This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
dip)) A (— sin (p)d(p = sin 3 ip cos2 9dip A d9. Finally, we compute rlt : M —• N. Proposition 7.1.19 Let <j>Q,<j>\ : M -* N be two homotopic smooth maps. Then they induce identical maps in cohomology <% = $* : H*(N) -» : S —• M is a smooth map. We denote by Cjt(M) the set of ^-dimensional cycles in M. Definition 7.3.2 (a)Two cycles (S 0 , #o), (Si, 0i) 6 Cfc(M) are said to be c o b o r d a n t (we write this (So, 4>o) ~ c (Si, 4>i)) if there exists a compact, oriented manifold with boundary E and a smooth map <£ : E —• M such that (al) 3E = (—S0) U Si where —S0 denotes the oriented manifold So with the opposite orientation and U denotes the disjoint union. n t/) such that s u p p l e {\x{\ < 1/2 ; 9Xk9"1) = p(-Xi, • • •, -X"*) for all A"i,...,X f c e fl. Set G G in the above definition induces a K-algebra morphism tp* •. r{G) The elements of ker ip* are called universal The following result is immediate. £ G . kx and aN(k0,p0)
Js
2
r2lT
sin3 ip cos2 9dip A d9 = / sin3 ipdip • / J [0,7r]x[0,27r] it Jo Jo
cos2 9d9
100
Calculus on manifolds A.'K
= — = volume of the unit ball B 3 C I 3 . As we will see in the next subsection the above equality is no accident. Example 3.4.10 (Invariant integration on compact Lie groups) Let G be a compact Lie group. Fix once and for all an orientation on £Q- Consider u> G det £,*G a positively oriented volume element. By left translation we can extend a; to a leftinvariant volume form on G which we continue to denote by u>. This defines an orientation and in particular, by integration, we get a positive scalar c=
I oi. JG
Set dg = \u> so that / dg = 1.
(3.4.4)
JG
dg is the unique left-invariant n-form (n = dimG) on G satisfying (3.4.4) (assuming a fixed orientation on G). We claim dg is also right invariant. Assume for simplicity that G is connected. To prove this consider the modular function G 3 / I H A ( / I ) 6 1 defined by R*h(dg) = A(h)dg. A(h) is a constant because R*hdg is a left invariant form so it has to be a scalar multiple oi dg. Since (R^^)* = {RhiRh,)* = R^R*^ w e deduce A{hxh2) = A(/n)A(/i 2 )
V/H, h2 € G.
Hence h i-> A/i is a smooth morphism G - > ( R \ {()},•)• Since G is connected A(G) C M+ and since G is compact A(G) is bounded. If there exists x G G such that A(z) / 1 then either A(x) > 1 or A(a; _1 ) > 1 and in particular we would deduce the set (A(xn))nez is unbounded. Thus A = 1 which establishes the right invariance of dg. The invariant measure dg provides a very simple way of producing invariant objects on G. More precisely, if T is tensor field on G, then for each x G G define
T[ = [ ({Lg).T)xdg. JG
Then x i-¥ Tx defines a smooth tensor field on G. We claim T is left invariant. Indeed, for any ft G G we have
(L^T1 = J{Lh).{{Lg).T)dg = fG((Lhg)tT)dg
101
Integration on manifolds "=fts f (Lu)*Td{h-lu)
= Te
(d(/i- 1 «)) = Ll-,du = du).
JG
If we average once more on the right we get a tensor jG{{R9)tTl)xdg
x H+
which is both left and right invariant. Exercise 3.4.3 Let G be a Lie group. For any X € £ G denote by ad{X) the linear map £(j —> £ G defined by
£G^Y^[X,Y]eSlG(a) If u> denotes a left invariant volume form prove that VX £ £Q LXUJ =
trad(X)u}.
(b) Prove that if G is a compact Lie group then trad(X) §3.4.3
= 0 for any X € £ Q .
Stokes formula
The Stokes' formula is the higher dimensional version of the fundamental theorem of calculus (Leibniz-Newton formula)
f"df = f(b)-f(a) Jo.
where / : [a, b] —• R is a smooth function and df = f'(i)dt. In fact, the higher dimensional formula will follow from the simplest 1-dimensional situation. We will spend most of the time finding the correct formulation of the general version and this requires the concept of manifold with boundary. The standard example is the upper half-space Hl = {{x1,...,xn) e l " ; ! 1 >0}. Definition 3.4.11 A smooth manifold with boundary is a closed subset M of a smooth manifold M (dim M = n) such that (a) M° = int M + 0. (b) For each point p G dM = M\M° there exist local coordinates (x 1 , ...,xn) defined on an open neighborhood 91 of p in M such that
(6i) M° n m = {{x\..., xn) • xl > o}. 1 (62) dMnm = {x = o}.
dM is called the boundary of M. orientable if M° is orientable.
A manifold with boundary (M, dM) is called
Example 3.4.12 A closed interval / = [a,b] is a smooth 1-dimensional manifold with boundary dl = {a, b}.
102
Calculus on manifolds
Example 3.4.13 The closed unit ball B3 C R 3 is an orientable manifold with boundary dB3 = S2. Proposition 3.4.14 Let (M, dM) be a smooth manifold with boundary. Then dM is also a smooth manifold of dimension dimdM = d i m M — 1. Moreover if M is orientable then so is its boundary. The proof is left to the reader as an exercise. Important convention Let (M, dM) be an orientable manifold with boundary. There is a (non-canonical) way to associate to an orientation on M° an orientation on the boundary. This will be the only way in which we will orient boundaries throughout this book. If we do not pay attention to this convention then our results may be off by a sign. We now proceed to described this induced orientation on dM. For any p £ dM choose local coordinates (a;1, ...,xn) as in Definition 3.4.11. Then the induced orientation of TpdM is defined by edx2 A---AdxnedetTpdM,
e = ±1
1
where e is chosen so that for x > 0 (i.e. inside M ) the form ei-dx1)
A dx2 A • • • A dxn
is positively oriented, dx1 is usually called an inner conormal since x1 increases as we go towards the interior of M. —dx1 is then the outer conormal for analogous reasons. The rule by which we get the induced orientation on the boundary can be rephrased as {outer conormal}A{induced orientation on boundary} = {orientation in the interior}. We may call this rule "outer (co)normal first" for obvious reasons. Example 3.4.15 The canonical orientation on Sn C M n+1 coincides with the induced orientation of S" + 1 as the boundary of the unit ball Bn+1. Exercise 3.4.4 Consider the hyperplane Hi C K" defined by the equation {xl = 0}. Prove that the induced orientation of Hi as the boundary of the half-space H"' 1 = {x' > 0} is given by the (n — l)-form (—lj'dx 1 A • • • A dx1 A • • • dxn where as usual the hat indicates a missing term. Theorem 3.4.16 (Stokes formula)Let M be an oriented n-dimensional manifold with boundary dM and u € fln_1(M) a compactly supported form. Then / JM°
du = /
u.
JdM
In the above formula d denotes the exterior derivative and dM has the induced orientation.
103
Integration on manifolds
Proof Via partitions of unity the verification is reduced to the following two situations. Case 1. w is a compactly supported (n — 1) form in W1. We have to show / du> = 0. JW It suffices to consider only the special case u) = f(x)dx2
A • • • A dxn
where f(x) is a compactly supported smooth function. The general case is a linear combination of these special situations. We compute 1 / du= f ^/rdx JW JW dxl
A f ^-rdx1)
= f
l
1
Jw~ \m dx
A • • • A dxn
dx 2 A - • • A dx n = 0
)
since L &dxl=/(o°'x2' •••• xn) ~ / ( _ o ° ' x 2 ' •••• xn) and / has compact support. Case 2. u is a compactly supported (n — 1) form on H " . Let = ] T fi^dx1
w
A • • • A dx1 A • • • A dxn.
i
Then ^=(E(-l)
i + 1
^)^1A--Ado;".
One verifies as in Case! 11 that that /
-J-dx1 A • • • A dxn = 0 for i ^ 1.
For i = 1 we have /
l^^
1
JH^ OX1
A • • • A dx" = / f / ° ° l ^ 1 | dx 2 A • • • A dx" yR"-i Vyo a x 1 J
= f ^ (/(oo, a;2, ..., x") - / ( 0 , x 2 ,..., xn)) dx2 A • • • A dx" = ~ /
^ / ( 0 , x 2 , ...,x")dx 2 A • • • A dxn = | a H n w.
The last equality follows from the fact that the induced orientation on 3 H " is given by —dx2 A • • • A dx n . This concludes the proof of the Stokes formula. •
104
Calculus on manifolds
Remark 3.4.17 Stokes formula illustrates an interesting global phenomenon. It shows that the integral JM duj is independent of the behavior of ui inside M. It only depends on the behavior of w on the boundary. Example 3.4.18 / xdy/\dz=
dx Ady Adz = vol (B 3 ) — — .
/
Js2
3
JB*
Remark 3.4.19 The above considerations extend easily to more singular situations. For example, when M is the cube [0, l ] n its topological boundary is no longer a smooth manifold. However, its singularities are inessential as far as integration is concerned. The Stokes formula continues to hold
f du= f OJ Vw6 Q:n-l(In).
Ji"
Jdi
The boundary is smooth outside a set of measure zero and is given the induced orientation: " outer (co)normal first". The above equality can be used to give an explanation for the terminology "exterior derivative" we use to call d. Indeed if w 6 n n _ 1 (M n ) and Ih = [0, h] then we deduce du) L = 0 = lim hTn [ h->0
w.
(3.4.5)
Jdl"
When n = 1 this is the usual definition of the derivative. Example 3.4.20 We now have sufficient technical background to describe an example of vector bundle which admits no flat connections thus answering the question raised at the end of Section 3.3.3. Let M be the unit sphere in M3. As in Chapter 1, we have a distinguished cover of M, {[/„, Us} where U„ = S2 \ {south pole} and U, = S2 \ {north pole}. Each of the two sets can be identified with the complex plane C and we get coordinates zn on Un and zs on Us. On the overlap U, D U„ the two sets of coordinates are related by _ 1 Consider the complex line bundle L obtained by gluing two trivial line bundles Ln —> U„ and Ls —> Us via the transition map gsn :UnnUs^
{zn ± 0} -»• C*, gSn{Zn) = zn.
A connection on L is a collection of two complex valued forms ujn G n 1 (f/„) ® C, u>s G O^iUs) ® C satisfying a gluing relation on the overlap (see Example 3.3.6) dz
Vn(z) = — + W,(z), z
{Z = Zn).
Integration on manifolds
105
If the connection is flat then duin = 0 on Un and dujs = 0 on u>s. Let E+ be the equator equipped with the induced orientation as the boundary of the northern hemisphere and E~ the equator with the opposite orientation (as the boundary of the southern hemisphere). The orientation of E+ coincides with the orientation given by the form dQ where zn = exp(i#). We deduce from the Stokes formula (which works for complex valued forms as well) that /
u)n = 0
JE+
/
u>s = -
JE-
us = 0. JE+
On the other hand over the equator we have uin — u>s = — = ido
z
from which we deduce 0= /
JE+
un - ws = [
JE+
id0 = 27ri !!!
Thus there exist no flat connections on the line bundle L and at fault is the gluing cocycle defining L. In a future chapter we will quantify the measure in which the gluing data obstruct the existence of flat connections. §3.4.4
Representations and characters of compact Lie groups
The invariant integration on compact Lie groups is a very powerful tool with many uses. Undoubtedly, one of the most spectacular application is Hermann Weyl's computation of the characters of representations of compact semi-simple Lie groups. The invariant integration occupies a central place in his solution to this problem. We devote this subsection to the description of the most elementary aspects of the representation theory of Lie groups. Let G be a Lie group. Recall that a (linear) representation of G is a left action on a (finite dimensional) vector space V
GxV^V
(g,v)^T(g)veV
such that the map T(g) is linear for any g. One also says V has a structure of Gmodule. If V is a real (resp. complex) vector space then it is said to be a real (resp. complex) G-module. Example 3.4.21 Let V - C n . Then G = GL(n, C) acts linearly on V in the tautological manner. Moreover V , V®k, A.mV and SeV are complex G-modules.
106
Calculus on manifolds
Definition 3.4.22 A morphism of G-modules V\ and V2 is a linear map L : V\ —> V2 such that for any g € G the diagram below is commutative i.e. T2{g)L = LT\(g)
The space of morphisms of G-modules is denoted by Hom<3(Vi, V2). The collection of isomorphisms classes of complex G-modules is denoted by G-Mod. If V is a G-module than an invariant subspace (or submodule) is a subspace U C V such that T(g)(U) C U, \/g G G. A G-module is said to be irreducible if it has no invariant subspaces other than {0} and V itself. Proposition 3.4.23 The direct sum "©" and the tensor product "®" define a structure of semi-ring with 1 on G-Mod. 0 is represented by the null representation {0} while 1 is represented by the trivial module G —> Aut (C), g i-+ 1. The proof of this proposition is left to the reader. Example 3.4.24 Let Ti : G -+ Aut ([/,) (i = 1, 2) be two complex G-modules. Then U{ is a G-module by
(s,u*).->i7GrVHence Horn {U\, U2) is also a G-module. Explicitly, the action of g € G is given by (g,L)^T2(g)LTl(g-1). We see that H o r n e d , [72) can be identified with the linear subspace in Horn (Ui, t/ 2 ) consisting of the linear maps U\ —¥ U2 unchanged by the above action of G. Proposition 3.4.25 (Weyl's unitary trick) Let G be a compact Lie group and V a complex G-module. Then there exists a hermitian metric h on V which is Ginvariant i.e. h(gv1,gv2) = h{v1,v2) Vui, v2 G V. Proof
Let h be an arbitrary hermitian metric on V. Define its G-average by h(u, v)=
h(gu, gv)dg JG
where dg denotes the normalized bi-invariant measure on G. One can now check easily that h is G-invariant. • In the sequel, G will always denote a compact Lie group.
Integration on manifolds
107
Proposition 3.4.26 Let V be a complex G-module and h a G-invariant hermitian metric. If U is an invariant subspace of V then so is Ux, where "_l_" denotes the orthogonal complement with respect to h. Proof Since h is G-invariant it follows that Vg e G, T(g) is an unitary operator, T(g)T*(g) = \ v . Hence, T*(g) = T " 1 ^ ) = TO?" 1 ), V € G. If x e UL then for all u G [/ and V5 € G MTGOz.u) = M(i.r*(ff)u) = h{x,T{g-l)u)
= 0.
Thus T{g)x e C1- so that [71- is G-invariant. D Corollary 3.4.27 Every G-module V can be decomposed as a direct sum of irreducible ones. If we denoted by Irr(G) the collection of isomorphism classes of irreducible G-modules then we can rephrase the above corollary by saying that Irr(G) generates the semigroup (G-Mod, ©). To gain a little more insight we need to use the following remarkable trick due to Isaac Schur. Lemma 3.4.28 (Schur lemma)Let Vi, V2 be two irreducible complex G-modules. Then
dimcHomG(y1,^2) = | j ' £ ~ £ * Proof Let L G HomG(Vi, V2). Then kerL C VY is an invariant subspace of V\. Similarly, Range (L) C V2 is an invariant subspace of V2. Thus, either kerL = 0 or k e r i = Vi. The first situation forces range L 7^ 0 and since V2 is irreducible Range L = V2. Hence L has to be in isomorphism of G-modules. We deduce that if V\ and V2 are not isomorphic as G-modules H o r n e d , V2) = {0}Assume now that Vx = V2 and S : V\ —> V2 is an isomorphism of G-modules. According to the previous discussion, any other nontrivial G-morphism L : V\ —> V2 has to be an isomorphism. Consider the automorphism T = S^L : V\ —>• Vx. Since V\ is a complex vector space T admits at least one (non-zero) eigenvalue A. The map Alv^ — T is an endomorphism of G-modules and ker (XlVl — T) / 0. Invoking again the above discussion we deduce T = A l ^ i.e. L = XS. This shows dimHom G (V 1 ,V 2 ) = 1. • Schur's lemma is powerful enough to completely characterize S ^ M o d .
108
Calculus on manifolds
Example 3.4.29 The irreducible (complex)representations of S 1 .) Let V be a complex irreducible 5 1 -module {eie, v) i-»- T9v
where In particular this implies that each Tg is an 5 1 -automorphism since it obviously commutes with the action of this group. Hence Tg = \(8)ly. In particular dim V = 1 since any 1-dimensional subspace of V is S'-invariant. We have thus obtained a smooth map A : S 1 ->• C* such that \{e'9 • eiT) =
\{eie)\{eie).
Hence A : S 1 —» C* is a group morphism. As in the discussion of the modular function we deduce that |A| = 1. Thus A looks like an exponential {verify!) A(e i S )=exp(ia0). Moreover, when exp(27ria) = 1 so that a £ Z . Conversely, for any integer n e Z w e have a representation 5x^Aut(C)
{eie,z)^einez.
The exponentials exp(m#) are called the characters of the representations pn. Exercise 3.4.5 Describe the irreducible representations of T"-the n-dimensional torus. Definition 3.4.30 (a) Let V be a complex G-module, g i-> T(g) e Aut (V). character of V is the smooth function Xv--G->C
Xv{g) =
The
trT(g).
(b) A class function is a continuous function / : G —> C such that f{hgh-l)
= f(g)
Vg,heG.
(The character of a representation is an example of class function). Theorem 3.4.31 Let G be a compact Lie group, U\, U2 complex G-modules and Xut their characters. Then the following hold. ( a ) x ^ e c / 2 = XUi + Xu2, XUi<su2 = XUi • XUi-
(b)xc/ i (l) = dim[/ i . ( c ) Xu" = Xt/;-^ 6 c o m p l e x conjugate of xu{-
Integration on manifolds
109
(d) f Xui{g)dg = AimUf JG
where Up denotes the space of G-invariant elements of Ui U? = {xeUi;x
= Ti(g)x V 5 e G}.
e / XuAd) • Xu2{9)d9 = dim H o r n e d , ^ i ) . Jc P r o o f (a) and (b) are left to the reader. To prove (c) fix an invariant hermitian metric on U = Ui. Thus each T(g) is an unitary operator on U. The action of G on U* is given by T f [g~ l ). Since T(g) is unitary we have Tf(gl) = T(g). This proves (c). P r o o f of ( d ) . Consider P:U
->U
Pu=
f
T(g)udg.
Note that PT(h) = T(h)P, Vft 6 G i.e. P e RomG{U, U). We now compute T(h)Pu
= JGT(hg)udg I T(j)udj
=
jj^uRUd-f
= Pu.
JG
Thus, each Pu is G-invariant. Conversely, if x G U is G-invariant then Px = / T(g)xdg = / xdg = x i.e. UG = Range P . Note also that P is a projector i.e. P2 = P. Indeed P2u = J T(g)Pudg = f Pudg = Pu. Hence P is a projection onto UG and in particular dime UG = t r P = fGtrT{g)dg
=
jGXu{g)dg.
P r o o f of (e). J XUi • Xu2d9 = J Xc/i • Xv;dg = J Xui®u;dg = / Xnom(u2,Ui)
= dime (Horn {U2, Ux)f
= dim c Hom G (i7 2 , U{)
since HoniG coincides with the space of G-invariant morphisms. •
Calculus on manifolds
no Corollary 3.4.32 Let U, V be irreducible G-modules. Then {Xu, Xv) = I Xv Xvd9 = <W = JG
Proof Follows from Theorem 3.4.31 using Schur lemma. D Corollary 3.4.33 Let U, V be two G-modules. Then U = V if and only if xu = XvProof
Decompose U and V as direct sums of irreducible G-modules U = ®?{mlUt)
V = ®[{njVj).
Hence xu = Y,miXVi and xv = Y,njXvr The equivalence "representation" •*=>• "characters" stated by this corollary now follows immediately from Schur lemma and the previous corollary. D Thus, the problem of describing the representations of a compact Lie group boils down to describing the characters of its irreducible representations. This problem was completely solved by Hermann Weyl. Its solution requires a lot more work and goes beyond the scope of this book. We will spend the remaining part of this subsection analyzing the equality (d) in Theorem 3.4.31. Describing the invariants of a group action was a very fashionable problem in the second half of the nineteenth century. Formula (d) mentioned above is a truly remarkable result. It allows (in principle) to compute the maximum number of linearly independent invariant elements. Let V be a complex G-module and denote by xv its character. The complex exterior algebra A*V* is a complex G-module, as the space of complex multi-linear skew-symmetric maps V x • • • x V -> C. Denote by bck(y) the complex dimension of the space of G-invariant elements in A^V*. One has the equality bck(V) = /
XMv-dg.
These facts can be presented coherently by considering the Z-graded vector space
3e(V) = ©*ALV. Its Poincare polynomial is
^W*) = Y,tkK(v) = I/xw-dg. To obtain a more concentrated formulation of the above equality we need to recall some elementary facts of linear algebra.
Integration on manifolds
111
For each endomorphism A of V denote by ak (A) the coefficient of tk in the characteristic polynomial at{A) =det(lv + tA). Explicitly,
°k(.A) =
det a
J2
( >«v)
(n = dimV).
l
Equivalently ak(A) =
tr(AkA)
k
where A M is the endomorphism of A V induced by A. If g e G acts on V by T(g) then g acts on A*V* by A ^ T ' ^ " 1 ) = A*T(ff). (We implicitly assumed that each T(g) is unitary with respect to some G-invariant metric on V). Hence XAJV = ak{f{g)). (3.4.6) We conclude that P3{v)(t)
= ^^(TigVdg
= fGdet(lv
+ tf(g))dg.
(3.4.7)
Consider now the following situation. Let V be a c o m p l e x V-module. Denote by A*V the space of R-multi-linear, skew-symmetric maps V x • • • x V ->• K. A* V* is a real G-module. We complexify it. A*V ® C is the space of M-multi-linear, skew-symmetric maps V x ••• x V - > C and as such it is a complex G-module. The real dimension of the subspace 3k of G invariant elements in AkV* will be denoted by brk(V) so that the Poincare polynomial of TT{V)=®k3kis On the other hand bTk(V) is equal to the complex dimension of AkV* ® C. Using the results of Subsection 2.2.5 we deduce A*TV
= 0
(ei+,-=*AiV ® A ' F ) .
(3.4.8)
k
Each of the above summands is a G-invariant subspace. Using (3.4.6) and (3.4.8) we deduce P
x(v) = E / E it
JG
^(TGrtK^cfo,))fdfl
i+j=k
f d e t ( l v + tT(s)) d e t ( l F + tT( f f ))d 5 = / | d e t ( l v + *T(ff))|2dff.
(3.4.9)
We will have the chance to use this result in computing topological invariants of manifolds with a "high degree of symmetry" like e.g. the grassmannians.
112 §3.4.5
Calculus on manifolds Fibered calculus
In the previous section we have described the calculus associated to objects defined on a single manifold. The aim of this subsection is to discuss what happens when we deal with an entire family of objects parameterized by some smooth manifold. We will discuss only the fibered version of integration. The exterior derivative also has a fibered version but its true meaning can only be grasped by referring to Leray's spectral sequence of a fibration and so we will not deal with it. The interested reader can learn more about this operation from [32]. Assume now that instead of a single manifold F we have an entire (smooth) family of them (Fj,)(,€S. In more rigorous terms this means we are given a smooth fiber bundle p : E —> B with standard fiber F. On the total space E we will always work with split coordinates (xl; yi) where (x1) are local coordinates on the standard fiber F and (yi) are local coordinates on the base B (the parameter space). The model situation is the bundle
E = Rk xRm^Rm
=B
{x,y)&y.
We will first define a fibered version of integration. This requires a fibered version of orientability. Definition 3.4.34 Let p : E -» B be a smooth bundle with standard fiber F. The bundle is said to be orientable if (a) F is oriented; (b) there exists an open cover (Ua) and trivializations
p-\Ua)
*AFxUa
such that the gluing maps ipp o V" 1 : F x Ua0 -> F x Ua0
{Ua/3 = Uan Up)
are orientation preserving in the sense that for each y € Uap, the diffeomorphism
is orientation preserving. Exercise 3.4.6 If the the base B of an orientable bundle p : E —• B is orientable, then so is the total space E (as an abstract smooth manifold). Important convention Let p : E —> B be an orientable bundle with oriented basis B. The natural orientation of the total space E is defined as follows.
Integration on manifolds
113
If E = F x B then the orientation of the tangent space Ty,b)E is given by QF x LJB where up S det T/F (resp. wg € detT(,B) defines the orientation oiT/F (resp. TbB). The general case reduces to this one since any bundle is locally a product and the gluing maps are orientation preserving. This convention can be briefly described as orientation total space = orientation fiber A orientation base. The natural orientation can thus be called the fiber-first orientation. In the sequel all orientable bundles will be given the fiber-first orientation. Let p : E —> B be an orientable fiber bundle with standard fiber F. Proposition 3.4.35 The integration along fibers is an operation P
* = /,,„ : ^ P i (£) -> 0£{B
(r = dimF).
uniquely defined by its action on forms supported on domains D of split coordinates D ^ l
r
If <j = fdx1 A dyJ (/ G C?(W+m)) r JE/B
x E
m
4 l
m
(x; y) M- y.
then
f 0 , \I\^r \ (/ R r fdx1) dyJ , | / | = r
The proof goes exactly as in the non-parametric case (i.e. when B is a point). One shows using partitions of unity that these local definitions can be patched together to produce a well defined map
/ • « ; * ( £ ) ^:P7(B). I E/B
The details are left to the reader. Proposition 3.4.36 Let p : E —> B be an orientable bundle with an r-dimensional standard fiber F. Then for any w € tt*cpt(E) and 77 e w*p((B) such that degw+degr/ = dim E we have / dEto = {-l)rdB I u. JE/B
JE/B
If B is oriented then
Js">*P'W = fB(fE/B<>>)*V-
(F»bini)
114
Calculus on manifolds
The last equality implies immediately p„(u> Ap*rf) = ptLj A T / .
Proof
It suffices to consider only the model case p:£ = l ' x I 1
and ui = fdx
m
4 l
m
= B
{x;y)Ay
J
A dy . Then
dEuJ = Y, J~idxi
A dx
Ad J
V + (- 1 )' 7 ' E J~jdxI
'
Ad
^
A
V-
The above integrals are defined to be zero if the corresponding forms do not have degree r. The Stokes formula shows that the first integral is always zero. Hence
/
JE/B
djw=(-l)l^ dyi
/
£ d x ' ) *dyiAdyJ j
WR' J
= (-iydB
[
JE/B
u.
The second equality is left to the reader as a practice exercise. • Definition 3.4.37 A d-bundle is a collection (E, dE,p, B) consisting of the following. (i) A smooth manifold E with boundary dE. (ii) A smooth map p : E —>• B such that the restrictions p : Int E —> B and p : dE —• B are smooth bundles. The standard fiber of p : IntE —¥ B is called the interior fiber. One can think of a 9-bundle as a smooth family of manifolds with boundary. Example 3.4.38 The projection p : [0,1] x M -> M
(t;m) >-> m
defines a 9-bundle. The interior fiber is the open interval (0,1). The fiber of p : d(I x M) —> M is the disjoint union of two points. Standard Models
A d-bundle is obtained by gluing two types of local models.
Interior models I ' x C - t Km Boundary models H ^ x l " 1 - ) Mm where Kr+ = {{x\---,xr)e&r
; x1 > 0 } .
Integration on manifolds
115
Remark 3.4.39 Let p : (E,dE) - > B b e a d-bundle. If p : Int E -> B is orientable and the basis B is oriented as well, then on dE one can define two orientations, (i) The fiber-first orientation as the total space of an oriented bundle dE —> B. (ii) The induced orientation as the boundary of E. Exercise 3.4.7 Prove that the above orientations on dE coincide. Theorem 3.4.40 Let p : (E, dE) —>• B be an orientable 9-bundle with an r-dimensional interior fiber. Then for any u> 6 Q.*cpt(E) we have / u> = dsu — (—Wd.B I JdE/B JE/B JE/B
u
(Homotopy formula).
The last equality can be formulated as
I JdE/B
= l
dE- (~iydB [
JE/B
.
JE/B
This is "the mother of all homotopy formulae". It will play a crucial part in Chapter 7 when we embark on the study of DeRham cohomology. Exercise 3.4.8 Prove the above theorem.
Chapter 4 Riemannian Geometry Now we can finally put to work the abstract notions discussed in the previous chapters. Loosely speaking the Riemannian geometry studies the properties of surfaces (manifolds) "made of canvas". These are manifolds with an extra structure arising naturally in many instances. In particular we will study the problem formulated in Chapter 1: why a plane (flat) canvas disk cannot be wrapped in an one-to-one fashinon around the unit sphere in R3. Answering this requires the notion of Riemann curvature which will be the central theme of this chapter.
4.1 §4.1.1
Metric properties Definitions and examples
To motivate our definition we will first try to formulate rigorously what do we mean by a "canvas surface". A "canvas surface" can be deformed in many ways but with some limitations: it cannot be stretched as a rubber surface and this is because the fibers of the canvas are flexible but not elastic. Alternatively, this means that the only operations we can perform are those which do not change the lengths of curves on the surface. Thus one can think of "canvas surfaces" as those surfaces on which any "reasonable" curve has a well defined length. Adapting a more constructive point of view one can say that such surfaces are endowed with a clear procedure of measuring lengths of piecewise smooth curves. Classical vector analysis describes one method of measuring lengths of curves in M3. If 7 : [0,1] ->• K3 is such a curve then its length is given by
length (7) =
Jo
f\iit)\dt
where |7(£)| is the euclidian length of the tangent vector j(t). We want to do the same thing on an abstract manifold and we are clearly faced with one problem: how do we make sense of |7(£)|? Obviously, this problem can be 116
117
Metric properties
solved if we assume that there is a procedure of measuring lengths of tangent vectors at any point on our manifold. The simplest way to do achieve this is to assume that each tangent space is endowed with an inner product (which can vary from point to point in a smooth way). Definition 4.1.1 (a) A Riemann manifold is a pair (M, g) consisting of a smooth manifold M and a metric g on the tangent bundle i.e. a smooth, symmetric positive definite (0,2) tensor field on M. g is called a Riemann metric on M. (a) Two Riemann manifolds (M;, &) (i — 1,2) are said to be isometric if there exists a diffeomorphism cj>: M\ -> M2 such that <£*g2 = 9iIf (M, g) is a Riemann manifold then for any x € M the restriction gx : TXM xTxM —>• R is an inner product on the tangent space TXM. We will frequently use the alternative notation (•, -)x = gx(-, •). The length of a tangent vector v € TXM is defined as usual as
K=VM) 1 / 2 If 7 : [a, b] —> M is a piecewise smooth curve then we define its length by
1(7) = fm)\,(t)dt. If we choose local coordinates (a;1,..., xn) onM then we get a local description of g as g ~ gijdx'dx3,
g{j = g{-g~, j - ) •
Proposition 4.1.2 Let M be a smooth manifold and denote by SHM the set of Riemann metrics on M. Then 9 1 ^ is a non-empty convex cone in the linear space of symmetric (0,2) tensors. Proof The only thing we have to prove is that SH^ is non-empty. We will use again partitions of unity. Cover M by coordinate neighborhoods (Ua)aeA- Let (xla) be a collection of local coordinates on Ua. Using these local coordinates we can construct by hand the metric ga on Ua by ga = {dx\)2 + ••• +
[dxlf.
Now pick a partition of unity B C C™(M) subordinated to the cover {Ua)ae^ there exits a map <j>: B —> A such that V/3 £ B supp/5 C U^p). Then define
i.e.
The reader can check easily that g is well defined and is indeed a Riemann metric on M. D
118
Riemannian
Figure 4.1: The unit sphere and an ellipsoid look
geometry
"different".
Example 4.1.3 (Euclidian space) The space H" has a natural Riemann metric g0 = (dx1)2
+ ••• +
(dxn)2.
The geometry of (Rn,g0) is the classical Euclidian geometry. Example 4.1.4 (Induced metrics on submanifolds)Let (M,g) be a Riemann manifold and S C M a submanifold. If i: S —» M denotes the natural inclusion then we obtain by pull back a metric on S 9s = i*9 = 9\s • For example, any invertible symmetric nx n matrix defines a quadratic hypersurface in Kn by •HA = {xeRn; (Ax,x) = l} where (•, •) denotes the euclidian inner product on W. %A has a natural metric induced by the Euclidian metric on M". For example when A = In then 'Hin is the unit sphere in H" the induced metric is called the round metric of 5 n _ 1 .
Remark 4.1.5 On any manifold there exist many Riemann metrics and there is no natural way of selecting one of them. One can visualize a Riemann structure as denning a "shape" of the manifold. For example the unit sphere x2 + y2 + z2 = 1 is diffeomorphic to the ellipsoid y? + f? + W = ^ ^ut they look "different " (see Figure 4.1). However, appearances may be deceiving. In Figure 4.2 it is illustrated the deformation of a sheet of paper to a half cylinder. They look different but the metric structures are the same since we have not changed the lengths of curves on our sheet. The conclusion to be drawn from these two examples is that we have to be very careful when we use the attribute "different".
119
Metric properties
Figure 4.2: A plane sheet and a half cylinder are "not so different". Example 4.1.6 (The hyperbolic plane) The Poincare model of the hyperbolic plane is the Riemann manifold (D, g) where D is the unit open disk in the plane R2 and the metric g is given by 9= i -2 j ( ^ 2 + dy2). 1 — xl — yi Exercise 4.1.1 Let S) denote the upper half-plane
Sj = {(u,v) e l 2 ; v> 0} endowed with the metric h = ^-{du2
+ dv2).
Show that the Cayley transform z = x + iyi-)w
.z + i = —l =u+ w 2 — 1
establishes an isometry (D, ) = (Sj, h). Example 4.1.7 (Left invariant metrics on Lie groups) Consider a Lie group G and denote by £ its Lie algebra. Then any inner product (-, •) on £,Q induces a Riemann metric h = (•, -)g on G defined by
hg(x,Y) = (x,Y)9 = <(V0.*>(V'W vs e G, x,y eT9G where (Lg-i), : TgG —> T\G is the differential at g G G of the left translation map Lg-i. One checks easily that g i-> (•, -)g is a smooth tensor field and is left invariant i.e. L*gh = h MgeG. If G is also compact one then can use the averaging technique of Subsection 3.4.2 to produce metrics which are both left and right invariant.
120
Riemannian
geometry
Exercise 4.1.2 Let M = Gk,n be the grassmannian of complex k-planes in C". For each such A;-plane V denote by Py the orthogonal projection onto V. Thus Pv = Pv = Pv and Range (Pv) = V so that Py can be viewed as a complex, selfadjoint n x n matrix. Denote by Sn the linear space of such matrices and by h0 the inner product h0{A, B) = tr (AB*) \/A, B G Sn. (a) Prove that the map Pk : GkiH —» Sn defined by V >-» Pv is an embedding i.e. it is a one-to-one immersion. (b) Let k = 1 so that Glt„ = a P n _ 1 . Describe the induced metric P{h0 on OP"" 1 using the homogeneous coordinates introduced in Section 1.2.2. §4.1.2
The Levi-Civita connection
To continue our study of Riemann manifolds we will try to follow a close parallel with classical euclidian geometry. The first question one may ask is whether there is a notion of "straight line" on a Riemann manifold. In the euclidian space M3 there are at least two ways to define a line segment. (i) A line segment is the shortest path connecting two given points. (ii) A line segment is a smooth path 7 : [0,1] —> M3 satisfying 7(0=0.
(4.1.1)
Since we have not said anything about the calculus of variations (which deals precisely with problems of type (i) ) we will use the second interpretation as our starting point. We will see however that both points of view yield the same conclusion. Let us first reformulate (4.1.1). As we know the tangent bundle of K3 is equipped with a natural trivialization and as such it has a natural trivial connection V° defined by V?9,-=0 Vi,j ( $ = £ ; , V i
=
Vaj)
i.e. all the Christoffel symbols vanish. Moreover if go denotes the euclidian metric then
(v°ff0) (a,-, 9*) = v°i53k - so(v?a,-, ak) - 9o(dv v%) = o i.e. the connection is compatible with the metric. Condition (4.1.1) can be rephrased as V°(t)7W = 0 (4.1.2) so that the problem of defining "lines" in a Riemann manifold reduces to choosing a "natural" connection on the tangent bundle. Of course we would like this connection to be compatible with the metric but even so, there are infinitely many connections to choose from. The following fundamental result will solve this dilemma.
121
Metric properties
Proposition 4.1.8 Consider a Riemann manifold (M, g). Then there exists a unique symmetric connection V on TM compatible with the metric g i.e. T(V) = 0
Vg = 0.
V is usually called the Levi-Civita connection associated to the metric g. Proof We first prove that there exists at most one connection with the required properties. We will achieve this by producing an explicit description of it. Let V be a connection with the desired properties i.e. Vg = 0 and For any X,Y,Z
VXY
- VYX
= [X, Y] VZ, Y G Vect (M).
£ Vect (M) we have Zg(X, Y) = g(VzX, Y) + g(X,
VZY)
since Vg = 0. Using the symmetry of the connection we compute Zg(X, Y) - Yg(Z, X) + Xg(Y, X) = g(VzX, Y) - g(VYZ, +g(X, VZY)
- g(Z, VYX)
= g([Z, n X) + g([X, Y],Z)+
+ g(Y,
X) + g(VxY,
Z)
VXZ)
g([Z, X], Y) + 2g(VxZ,
Y).
We conclude g(VxZ,
Y) = l-{Xg{Y, Z) - Yg(Z, X) + Zg(X,
-g([X,Y],Z)
Y)
+ g([Y, Z],X) - g([Z,X],Y)}.
(4.1.3)
The above equality establishes the uniqueness of V. Using local coordinates {xl,...,xn) on M we deduce from (4.1.3) (with X = foTi'Y=
d^'
Z
=
a!~)
that
g{Vidj, dk) = guYlj = - (digjk - dkg{j +
djgik).
If (g'e) denotes the inverse of (g^) we deduce
4 = \gu (di93k - dk9i] + dm).
(4.1.4)
Existence It boils down to showing that (4.1.3) indeed defines a connection with the required properties. The routine details are left to the reader. • We can now define the notion of "straight line" on a Riemann manifold.
122
Riemannian
geometry
Definition 4.1.9 A geodesic on a Riemann manifold (M, g) is a smooth curve 7 : (a, b) —> M satisfying V W ,7(*) = 0 (4.1.5) where V is the Levi-Civita connection. Using local coordinates (x1, ...,xn) with respect to which the Christoffel symbols are (Tkj) and j(t) = (x1^), ...,xn(t)) we can reformulate the geodesic equation as a second order nonlinear system of ordinary differential equations. Set £ = -y(i) = xldi. Then V±Mt) = £% + x'VjLdi dt
w
dt
= xkdk + rk]lxiijdk
(r£- = rj4)
so that the geodesic equation is equivalent to xk + rkjxixj
= 0
VJfc = 1,..., n.
(4.1.6)
k
Since the coefficients r^- = T j(x) depend smoothly upon x we can use the classical Banach-Picard theorem on existence in initial value problems (see e.g. [4]). We deduce the following local existence result. Proposition 4.1.10 Let (M, g) be a Riemann manifold. For any compact subset K C TM there exists e > 0 such that for any (a;, X) 6 K there exists a unique geodesic 7 = jx<x • (—s,e) —>• M such that 7(0) = x, 7(0) = X. One can think of a geodesic as defining a path in the tangent bundle t i-t (7(i),7(i)). The above proposition shows that the geodesies define a local flow $ on TM by $ t (z,A-) = (7(i),7(t)) 7 = 7*,*. Definition 4.1.11 The local flow defined above is called the geodesic flow of the Riemann manifold (M,g). When the geodesic glow is a global flow i.e. any j X t X is defined at each moment of time t for any (x, X) 6 TM then the Riemann manifold is called geodesically c o m p l e t e . The geodesic flow has some remarkable properties. C o n s e r v a t i o n of energy: If j(t) is a geodesic then the length of j(t) is independent of time. Indeed ^|7(«)|2 = ^(7(*),7(*)) = 2g{Vi(thj(t))
= 0.
Thus if we consider the sphere bundles ST{M) = {XeTM
; \X\ = r}
we deduce that Sr(M) are invariant subsets of the geodesic flow.
123
Metric properties Exercise 4.1.3 Describe the infinitesimal generator of the geodesic flow. Example 4.1.12 Let G be a connected Lie group and £ G its Lie algebra. X 6 £ G defines an endomorphism ad (X) of £ G by ad{X)Y
Any
= [X,Y\.
The Jacobi identity implies that ad([X,Y])
=
[ad{X),ad(Y)}
where the bracket in the right hand side is the usual commutator of two endomorphisms. Assume there exists an inner product (•, •) on £ G such that for any X 6 £ G , ad{X) is skew-adjoint i.e. {[X,Y],Z)
= ~(Y,[X,Z})
(4.1.7)
We can now extend this inner product to a left invariant metric h on G. We want to describe its geodesies. First, we have to determine the associated Levi-Civita connection. Using (4.1.3) we get h{VxZ,
Y) = \{Xh{Y,
Z) - Y(Z, X) + Zh(X,
Y)
-h([X, Y],Z) + h([Y, Z], X) - h([Z, X], Y)}. If we take X, Y, Z e £ G i-e. these vector fields are left invariant then h(Y, Z) = constant, h(Z,X) — const., h(X,Y) = const, so that the first three terms in the above formula vanish. We obtain the following equality (at l e G ) {VXZ,Y)
= i{-([X,Y],Z) + <[Y,Z\,X)-
([Z,X],Y)}.
Using the skew-symmetry of ad(X) and ad(Z) we deduce (VXZ,Y)
=
~([X,Z},Y)
so that at 1 € G VxZ=l-[X,Z]
VX,Ze£G.
(4.1.8)
This formula correctly defines a connection since any X £ Vect (G) can be written as a linear combination X = YJ(*iXi
aiEC^iG)
XiEZo.
124
Riemannian
geometry
Now, if j(t) is a geodesic we can write j(t) = Y,liXi so that
0 = V W 7(t) = E 7 ^ +
\Y,wAxi>Xi]i,j
i
Since psT^X,] = —[Xj,Xi\ we deduce 7* = 0 i.e. 7W = £ 7 i ( 0 L Y i = X. This means 7 is an integral curve of the left invariant vector field X so that the geodesies through the origin with initial direction X e TiG are 7*(i) =
exp{tX).
Exercise 4.1.4 Let G be a Lie group and h a bi-invariant metric on G. Prove that its restriction to £Q satisfies (4.1.7). In particular, on any compact Lie groups there exist metrics satisfying (4.1.7). Definition 4.1.13 Let £ be a finite dimensional real Lie algebra. pairing or form is the bilinear map K : £ x £ —> K, K(X, y) = —tr (ad(x)ad(y))
The Killing
x, y G £ .
The Lie algebra £ is said to be semisimple if the Killing pairing is a duality. A Lie group G is called semisimple if its Lie algebra is semisimple. Exercise 4.1.5 Prove that SO(n) and SU(n) and SL{n, M) are semisimple Lie groups but U(n) is not. Exercise 4.1.6 Let G be a compact Lie group. Prove that the Killing form is positive definite and satisfies (4.1.7). N o t e The converse of the above exercise is also true i.e. any semisimple Lie group with positive definite Killing form is compact. This is known as Weyl's theorem. Its proof, which will be given later in the book, requires substantially more work. Exercise 4.1.7 Show that the parallel transport of X along exp(iF) is ( i exp(fy))*(^exp(|y))*^-
Example 4.1.14 (Geodesies on flat tori and SU(2)) The n-dimensional torus Tn = S1 x • • • x S1 is an abelian, compact Lie group. If (9l, •••,6n) are the natural angular coordinates on T™ then the flat metric is defined by g
= {de1)2 + •••
[denf.
125
Metric properties
g is a bi-invariant metric on on Tn and obviously its restriction to the origin satisfies the skew-symmetry condition (4.1.7) since the bracket is 0. The geodesies through 1 will be the exponentials 701
an{t)
t^(eta»t,...,ete-')
ak€R.
If the numbers ak are linearly dependent over the rationals Q then obviously 7Q1,...,Q„(t) is a closed curve. On the contrary, when the a's are linearly independent over Q then a classical result of Kronecker (see e.g. [33]) states that the image of 7Ql,...lQn is dense in Tn\\\ (see also Section 7.4 to come) The special unitary group SU(2) can also be identified with the group of unit quaternions {a + bi + cj + dk; a2 + b2 + c2 + d2 = 1} so that SU(2) is diffeomorphic with the unit sphere S3 C K4. The round metric on S 3 is bi-invariant with respect to left and right (unit) quaternionic multiplication (verify this) and its restriction to (1,0,0,0) satisfies (4.1.7). The geodesies of this metric are the 1-parameter subgroups of S3 and we let the reader verify that these are in fact the great circles of S3 i.e. the circles of maximal diameter on S3. Thus, all geodesies on S3 are closed. §4.1.3
The exponential map and normal coordinates
We have already seen that there are many differences between the classical euclidian geometry and the the general riemannian geometry in the large. In particular we have seen examples in which one of the basic axioms of euclidian geometry no longer holds: two distinct geodesic (read lines) may intersect in more than one point. The global topology of the manifold is responsible for this "failure". Locally however, things are not "as bad". Local riemannian geometry is similar in many respects with the euclidian geometry. For example, locally, all of the classical incidence axioms hold. In this section we will define using the metric some special collections of local coordinates in which things are very close to being euclidian. Let (M, g) be a Riemann manifold and U an open coordinate neighborhood with coordinates (x 1 , ...,xn). We will try to find a local change in coordinates (a;*) >-> (yj) in which the expression of the metric is as close as possible to the euclidian metric So = Sijdtfdyi. Let q g U be the point with coordinates (0,...,0). Via a linear change in coordinates we may as well assume that 9iM) -
S
'j-
We can formulate this by saying that (gi3) is euclidian up to order zero. We would like to "spread" the above equality to an entire neighborhood of q. To achieve this we try to find local coordinates (yJ) near q such that in these new
126
Riemannian
geometry
coordinates the metric is euclidian up to order one i.e.
We now describe a geometric way of producing such coordinates using the geodesic flow. Denote as usual the geodesic from q with initial direction X G TqM by 7,,x(i). Note the following simple fact Vs > 0 jq,sx(t)
= -yq,x(st).
Hence, there exists a small neighborhood V of 0 € TqM such that for any X EV the geodesic 7,,x(*) is defined for all \t\ < 1. We define the exponential map at q by exp, : V C TqM -+ M
X^
7,,x(l)-
TqM is an euclidian space and we can define D ? (r) C TqM - the open "disk" of radius r centered at the origin. We have the following result. Proposition 4.1.15 Let (M,g) and q € M as above. Then there exists r > 0 such that the exponential map exp g : D,(r) ->• M is a diffeomorphism onto. The supremum of all radii r with this property is denoted by PM{q)Definition 4.1.16 PM{I) is called the injectivity r a d i u s of M at q. The infimum pM =
'mipM(q)
is called the injectivity radius of M. The proof relies on the following key fact. Lemma 4.1.17 The Frechet differential at 0 6 TqM of the exponential map Do exp, : TqM -> TexPq{a)M = TqM is the identity TqM ->• TqM. Proof of the lemma. Consider X G TqM. It defines a line in TqM by t >-* tX which is mapped via the exponential map to the geodesic 7,,x(i)- By definition (D0expq)X
= j9,x(0)=X.
O
Proposition 4.1.15 follows immediately from the above lemma using the inverse function theorem.
Metric properties
127
Now choose an orthonormal frame (ei,...,e n ) of TqM and denote by ( x 1 , . . . ^ " ) the resulting cartesian coordinates. For 0 < r < PM{Q) any point p 6 exp g (D g (r)) can be uniquely written as V = exp^x^i) so the collection (x 1 ,..., x") provides a coordinatization of the open set exp g (D ? (r)) c M. The coordinates thus obtained are called normal coordinates at q , the open set exp 9 (D,(r)) is called a normal neighborhood and will be denoted by B r (g) for reasons that will become apparent a little later. Proposition 4.1.18 Let (x1) be normal coordinates at q € M and denote by gy the expression of the metric tensor in these coordinates. Then we have dgigij(
€ Kn the curve x* = mlt is the geodesic
rjt(x(t)K"m* = 0. In particular T)k(Q)m?mk = 0 Vmj 6 » " from which we deduce the lemma. • The result in the above lemma can be formulated as ff(V^--,—I) \
=0
Vt.j.fc
a*J OX* OX* J
so that V ^ — - = 0 at q Vt,j. 8xJ OTC
Using Vp = 0 we deduce
g^) = (^g,)l,= o. *
(4.1.9)
128
Riemannian
geometry
The reader may ask whether we can go one step further and find local coordinates which produce a second order contact with the euclidian metric. At this step we are in for a big surprise. This thing is in general not possible and in fact there is a geometric way of measuring the "second order distance" between an arbitrary metric and the euclidian metric. This is where the curvature of the Levi-Civita connection comes in and we will devote an entire section to this subject. §4.1.4
The minimizing property of geodesies
We defined geodesies via a 2nd order equation imitating the 2nd order equation defining lines in an euclidian space. As we have already mentioned, this is not the unique way of extending the notion of Euclidian straight line to arbitrary Riemann manifolds. One may try to look for curves of minimal length joining two given points. We will prove that the geodesies defined as in the previous subsection do just that, at least locally. We begin with a technical result which is interesting in its own. Let (M, g) be a Riemann manifold. Lemma 4.1.20 For each } E M there exists 0 < r < PM(Q) and e > 0 such that Vm G B r (g) we have e < pM{m) and B e (m) 3 B r (g). In particular, any two points of Gr(q) can be joined by a unique geodesic of length < e. We must warn the reader that the above result does not guarantee that the postulated connecting geodesic lies entirely in B r (g). This is a different ball game. Proof of the lemma Using the smooth dependence upon initial data in ordinary differential equations we deduce that there exists an open neighborhood V of (q, 0) € TM such that exp m X is well defined for all (m, X) G V. We get a smooth map F:V
^ M xM
(m,X)^{m,
exp m X).
We compute the differential of F at (q, 0). First, using normal coordinates (x1) near q we get coordinates (x*; ~XP) near (q, 0) G TM. The partial derivatives of F at (q, 0) are
Thus the matrix defining D^i0)F has the form id *
0 id
and in particular it is nonsingular. It follows from the implicit function theorem that F maps some neighborhood V of (9, 0) G TM diffeomorphically onto some neighborhood U of (q, q) G M x M. We
129
Metric properties
can choose V to have the form {(m, X) ; \X\m < e, m 6 Bs(q)} for some sufficiently small e and 5. Choose 0 < r < min(e,pAf (g)) such that Vm1,m2€Br(g)
(m 1 ,m 2 ) € [/.
In particular, it follows that for any m G Br() exp m : D £ (m) C T m M -» M is a diffeomorphism onto and B £ (m) = exp m (D E (m)) D B r (g). Clearly, for any m the curve t H-» exp m (£X) is a geodesic of length < e joining m to exp m (X). It is the unique geodesic with this property since F : V —> U is injective. D We can now formulate the main result of this subsection. Theorem 4.1.21 Let q, r and e as in the previous lemma and consider 7 : [0,1] —• M the unique geodesic of length < e joining two points of B r (g). If tv : [0,1] —>• M is a piecewise smooth path with the same endpoints as 7 then
f W)\dt < f \u{t)\dt
Jo
Jo
with equality if and only if w([0,1]) = 7([0,1]). Thus 7 is the shortest path joining its endpoints. The proof relies on two lemmata. Let m € M be an arbitrary point and assume 0 < R < pM{m). Lemma 4.1.22 (Gauss) In Bfi(m) the geodesies through m are orthogonal to the hypersurfaces E, = exp,(5 4 (0)) = {exp m (X) ; |X| = &} 0 < 6 < R. Proof Let t i-t X(t), 0 < t < 1 denote an arbitrary smooth curve in TmM such that |X(t)| m = 1 i.e. X(t) is a curve on the unit sphere Si(0) C TmM. We have to prove that the curves t >-¥ exp m (5X(i)) are orthogonal to the radial geodesic s i-» exp m (sX(0)), 0 < s < R. Consider the parameterized surface f(s, t) = expm(sX(t)) Set
(a, t) e [0, R) x [0,1].
130
Riemannian
geometry
Define -£ similarly. | £ and | £ are not vector fields in the usual sense since the surface / may have double points and at such point /» may associate two tangent vectors. One way around this problem is to pullback everything (metric, connection etc.) to the rectangle [0, R] x [0,1] via / and everything will be correctly defined. We will denote the pulled-back objects by the same symbols as the originals. It is convenient to view these "vector fields" as the tangent vectors to the "coordinate lattice" f(t = const.), f(s = const.). We have to show
,df df.
, d d,
w/
.
where (•, •} denotes the inner product defined by g. The first equality is tautological. Using the metric compatibility of the (pulled-back) Levi-Civita connection we compute d , d d, , _ d d. ,d _ d, os os ot 9* ds ot os o> ot Since the curves f(t = const.) are geodesies we deduce 9' OS
On the other hand, since | Jj, JU = 0 we deduce (using the symmetry of the LeviCivita connection) d_
W
d_ _ } {
d_
%dt ~ ds'
d_
mds'~
2 _ld_ d_ 2dt ds
0
since | ^ = \X(t)\ = 1. We conclude that the quantity (Jj, ^ ) is independent of s. For s = 0 we have / ( 0 , t) = exp m (0) so that | £ = 0 and therefore
as needed. • Now consider any continuous, piecewise smooth curve w: [a,6]->BR(m)\{m}. Each uj(t) can be uniquely expressed in the form cj(t)=exPm(p(t)X(t))
\X(t)\ = l
0<\p(t)\
L e m m a 4.1.23 The length of the curve u>(t) is > \p(b) — p(a)\. The equality holds iff X(t) = const and p(t) > 0. In other words, the shortest path joining two concentrical shells Tig is a radial geodesic.
131
Metric properties Proof
Let f{p,t)
= expm(pX(t))
so that w{t) =
f(p{t),t).
Since the vectors §^ and | £ are mutually orthogonal and since = \X(t)\ = 1 we get 2
H2 = | / f + | ; >|P|2The equality holds if and only if §£ = 0 i.e. X = 0. Thus / |w|dt> /" | p | d t > | / 9 ( 6 ) - p ( a ) | . Ja
Ja
Equality holds if and only if p(t) is monotone and X(i) is constant. This completes the proof of the lemma. • The proof of Theorem 4.1.21 is now immediate. Let m0,mi e B r (g) and 7 : [0,1] —¥ M a geodesic of length < e such that 7(1) = rrij, i = 0,1. We can write j(t) = expmo(tX)
l £ D
£
( 4
Set .R = |X|. Consider any other piecewise smooth path u> : [a, b] —> M joining m 0 to rri!. For any 8 > 0 this path must contain a portion joining the shell T,j(m0) to the shell E R (m 0 ) and lying between them. By the previous lemma the length of this segment will be > R — 6. Letting 5 —> 0 we deduce l(u>)>R=l{j). If ui([a, b}) does not coincide with 7([0,1]) then we obtain a strict inequality. • Any Riemann manifold has a natural structure of metric space. More precisely we set d(p, q) = inf {l(w) ui : [0,1] —> M piecewise smooth path joining p to q} A piecewise smooth path w connecting two points p, q such that 1(LJ) = d(p, q) is said to be minimal. From Theorem 4.1.21 we deduce immediately the following consequence.
132
Riemannian
geometry
Corollary 4.1.24 The image of any minimal path coincides with the image of a geodesic. In other words, any minimal path can be reparameterized such that it satisfies the geodesic equation. Exercise 4.1.8 Prove the above corollary. Theorem 4.1.21 also implies that any two nearby points can be joined by a unique minimal geodesic. In particular we have the following consequence. Corollary 4.1.25 Let q € M. Then for all r > 0 sufficiently small exp,(D r (0))(= Br(q)) = {peM;
d{p,q) < r}
(4.1.10)
Corollary 4.1.26 For any q € M we have the equality PM(Q) = sup{r ; r satisfies (4.1.10)}. Proof The same argument as in the proof of Theorem 4.1.21 shows that Vr < PM(Q) the radial geodesies expJtX) are minimal. • Definition 4.1.27 A subset U C M is said to be convex if any two points in U can be joined by a unique minimal geodesic which lies entirely inside U. Proposition 4.1.28 For any q € M there exists 0 < R < LM(q) such that for any r < R the ball B r (g) is convex. Proof Choose a small e > 0 (0 < 2e < PM(Q)) a n d 0 < R < e such any two points m 0 , m-i in BR(q) can be joined by a unique minimal geodesic 7 m o , m i (t) (0 < t < 1) of length < e (not necessarily contained in BR(q)). We will prove that Vm0, m1 6 B fl (g) the map t >-> d{q, 7 m 0 j m i (t)) is convex and thus it achieves its maxima at the endpoints t = 0,1. Note that
d(q,j{t))
X{t)eTqM
with r{t) =
d{q,ym(hrni{t))-
It suffices to show J^(r 2 ) > 0 for t G [0,1] if d(q,m0) and d(q,mi) are sufficiently small. At this moment it is convenient to use normal coordinates (x1) near q. The geodesic 7 mo>mi takes the form (x ! (i)) and we have r2 = (x 1 ) 2 + • • • + (x") 2 . We compute easily ^ ( r 2 ) = 2r 2 (x 1 + --- + x") + |x| 2
(4.1.11)
133
Metric properties where x(t) = X) x ' e i e TqM. 7 satisfies the equation x i + r;.fc(x)xJ'x* = 0
Since 1^(0) = 0 (normal coordinates) we deduce that there exists a constant C > 0 (depending only on the magnitude of the second derivatives of the metric at q) such that
|r;*(x)| < c\x\. Using the geodesic equation we obtain x* > - C | x | | x | 2 . We substitute the above inequality in (4.1.11) to get ^(r2)>2|x|2(l-nC|x|3).
(4.1.12)
If we choose from the very beginning
then because along the geodesic \x\ < R + e the right-hand-side of (4.1.12) is nonnegative. This establishes the convexity of 11-> r2(t) and concludes the proof of the proposition. • In the last result of this subsection we return to the concept of geodesic completeness. We will see that this can be described in terms of the metric space structure alone. Theorem 4.1.29 (Hopf-Rinow)Let M be a Riemann manifold and q 6 M. The following assertions are equivalent: (a) exp„ is defined on all of TqM. (b) The closed and bounded (with respect to the metric structure) sets of M are compact. (c) M is complete as a metric space. (d) M is geodesically complete. (e) There exists a sequence of compact sets Kn C M, Kn C Kn+X and (Jn Kn = M such that if pn $ Kn then d(q,pn) —» 00. Moreover, on a (geodesically) complete manifold any two points can be joined by a minimal geodesic. Remark 4.1.30 On a complete manifold there could exist points (sufficiently far apart) which can be joined by more than one minimal geodesic. Think for example of a manifold where there exist closed geodesic (e.g the tori Tn).
134
Riemannian
geometry
Exercise 4.1.9 Prove the Hopf-Rinow theorem. Exercise 4.1.10 Let (M, g) be a Riemann manifold and let (Ua) an open cover consisting of bounded geodesically convex open sets. Set da = (diameter (Ua))2. Denote by ga the metric on Ua defined by ga = d~*g so that the diameter of Ua in the new metric is 1. Using a partition of unity (ipi) subordinated to this cover we can form a new metric 9 = Y, ViMi)
(SUPP W
c
U
«{i))-
i
Prove that g is a complete Riemann metric. Hence, on any manifold there exist complete Riemann metrics. §4.1.5
Calculus on Riemann manifolds
The classical vector analysis extends nicely to Riemann manifolds. We devote this subsection to describing this more general "vector analysis". Let (M,g) be an oriented Riemann manifold. We now have two structures at our disposal: a Riemann metric and an orientation and we will use both of them to construct a plethora of operations on tensors. First, using the metric we can construct by duality an isomorphism C : Vect (M) —>• &}{M) (called lowering the indices). Example 4.1.31 Let M = R 3 with the euclidian metric. A vector field V on M has the form ox
ay
az
Then W = CV = Pdx + Qdy + Rdz. If we think of V as a field of forces in the space then W is the infinitesimal work of V. On a Riemann manifold there is an equivalent way of describing the exterior derivative. Proposition 4.1.32 Let e : C^iTM
® A*T*M) ->
Cx'(A*+1T*M)
denote the exterior multiplication operator e{a®P)=a/\p,
VQ G Qr{M), /3 €
Qk(M).
Then d = eo V where d denotes the exterior derivative and V is the connection induced on AkT*M by the Levi-Civita connection.
135
Metric properties Proof We will use a strategy useful in many other situations. normal coordinates will payoff. Denote temporarily by D the equality d = D is a local statement and it suffices to prove neighborhood. Choose (x 1 ) normal coordinates at an arbitrary di = ^ T . Note that
Our discussion about operator e o V. The it in any coordinate point p € M and set
i k
Let u) e £l (M). Near p it can be written as u) — ^2 ujjdx1 i
where as usual, for any ordered multi-index / : (1 < ix < • • • < ik < n) we set dx1 = dx' 1 A • • • A dxlk. In normal coordinates a t p we have (Vj3,) | p = 0 from which we get the equalities (VidxO | p (dk) = -(dx.J(Vidk))
| p = 0.
Thus a t p Dw = ^ dx" A V i (w / dx / ) = J2
dx
'
A
(djUidx1 + w/Vi(dx f )) = ^ dx* A 3jW/ = dw.
i
i
Since the point p was chosen arbitrarily this completes the proof of Proposition 4.1.32. D
Exercise 4.1.11 Show that for any fc-form u> on the Riemann manifold (M,g) the exterior derivative dui can be expressed by
Mx°,Xi,---,,xk) = ^2(-i)i(vXiu)(x0,---,xi,---,xk) for all Xo, • • • ,Xk € Vect (M). (V denotes the Levi-Civita connection.) The Riemann metric defines by duality a metric in any tensor bundle TJ~(M) which we continue to denote by g. Thus, given two tensor fields 1\, T2 of the same type (r, s) we can form their pointwise scalar product MBp^
g(T, S)p = ^ ( ^ ( p ) , T 2 (p)).
In particular, any such tensor has a pointwise norm
M3p^\T\g,p
=
{T,T)y\
Using the orientation we can construct (using the results in subsection 2.2.4) a natural volume form on M which we denote by dvg and we call it the metric volume .
136
-Riemannian geometry
This is the positively oriented volume form of pointwise norm = 1. If (x 1 ,..., i") are local coordinates such that dx1 A • • • A dxn is positively oriented then dvg = \f\g\dxl A • • • A dx11 where \g\ = det(^y). In particular we can integrate (compactly supported) functions on M by
/
f=f[
fdvg
V/ G C0°°(M).
We have the following (not so surprising) result. Proposition 4.1.33 Vxdvg = 0
Proof
V i e Vect (M).
We have to show that for any p G M (Vxdvg)(eu...,en)
=0
(4.1.13)
where basis of TPM. Choose (x1) normal coordinates near p. Set d{ = -g~, Sij — g{di,dk) and e* = d^ \p. Since the expression in (4.1.13) is linear in X we may as well assume X = dk for some k = 1,..., n. We compute (Vxdv9)(ei,...,
en) = X{dvg{du ..., 3n)) | p - £ dv9(eu ..., (Vxd{) \p,..., dn).
(4.1.14)
z
We consider each term separately. Note first that dvg(d1, ...,dn) = (det(gy)) 1 / 2 so that X(det(gij))1^2 | p = 9/t(det(gy)) 1 / 2 | p is a linear combination of products in which each product has a factor of the form dkgij \p- Such a factor is zero since we are working in normal coordinates. Thus the first term in(4.1.14) is zero. The other terms are zero as well since in normal coordinates p
Vxdt = V^di = 0. Proposition 4.1.33 is proved. • . Once we have an orientation we also have the Hodge *-operator * : Qk(M) -+
Sln'k(M)
uniquely determined by a A *j3 = (a, p)dvg
Vaj3eQk{M).
In particular *1 = dvg.
(4.1.15)
Metric properties
137
Example 4.1.34 To any vector field F = Pdx + Qdy + Rdz on K3 we associated its infinitesimal work WF = C{F) = Pdx + Qdy + Rdz. The infinitesimal energy flux of F is the 2-form §F = *WF = Pdy Adz + Qdz Adx + Rdx A dy. The exterior derivative of WF is the infinitesimal flux of the vector field curl F dWF = {dyR - dzQ)dy Adz + (dzP - dxR)dz Adx + (dxQ - dyP)dx A dy = $cur\F = *WcurlF-
The divergence of F is the scalar defined as div F = *d*WF = * d $ F = * {(dxP + dyQ + dzR)dx AdyA dz} = dxP + dyQ + dzR. If / is a function on R3 then we compute easily *d*df = d2j + %f + d2j = Af. Definition 4.1.35 (a) For any smooth function / on M we denote by g r a d / the vector field dual to the 1-form df. In other words (grad / , X) = df{X) = X • f
V i e Vect (M).
(b) For any X € Vect (M) we denote by div X the smooth function defined by the equality Lxdvg = (div X)dvg. Proposition 4.1.36 Let A" be a vector field on M and denote by a the 1-form dual to X. Then (a) d i v X = t r ( V X ) where we view V X as an element of C°°(End (TM)) via the identifications V X 6 Q}{TM) =* CX(T*M
® TM) ^ C°°(End (TM)).
(b) d i v X = *d * a. (c) If (x1,..., xn) are local coordinates such that dx1 A • • • A dxn is positively oriented then
divx = ^Ld i (^jx') y\g\ where X = X^d*. The proof will rely on the following technical result which is interesting in its own.
138
Riemannian
geometry
Lemma 4.1.37 Denote by S the operator S = *d* : nk(M)
->
tf-^M).
Let a be a (k — l)-form and fi a &-form such that at least one of them is compactly supported. Then I {da,P)dvg = {-iy-» J {a,5P)dvg JM
JM
where vn£ = nk + n + 1. Definition 4.1.38 Define d* : Q,k{M) -> Qk~l{M)
by
,
d* = (-l)"»-*5 = ( - l ) ' » ' » * d * . Proof of the lemma
We have
/ (da, p)dv„ = f da A */3 = [ d(a A *j3) + (-1)* / a Ad* /}.
JM
JM
JM
JM
The first integral in the right-hand-side vanishes by the Stokes formula since a A *j3 has compact support. Since d * (3 € fin-*+1(M) and * 2 = (_i)(»-*+i)(*-i) 0 n
fi"-*+1(Af)
we deduce f (da,P)dvq = (~l)k+^-k+l^-k^
JM JM
f
aA*6
JM
This establishes the assertion in the lemma since (n-
k + l ) ( f c - 1) + k = j/„ i t (mod2). D
Proof of the proposition Set 0 = di>9 and let (Xi,...,Xn) frame of TM in a neighborhood of some point. Then
be a local moving
(Lxsi)(xu..., xn) = x{n(xu..., xn)) - Y, n(*i> -. [*> xl -. *»)•
(4-L16)
i
Since VQ. = 0 we get X • (fl(Xi,..., X n )) = £ to{Xu - , VxXi,...,
Xn).
i
Using the above equality in (4.1.16) we deduce from VXY {LxQ)(Xlt....
Xn) = Y,n{Xlt...,
— [X,Y] = VyX
VXiX,..., Xn).
that (4.1.17)
i
Over the neighborhood where the local moving frame is defined we can find smooth functions / / such that VXiX
= f?XJ=>tr(VX)
= fi.
139
Metric properties
Part (a) of the proposition follows after we substitute the above equality in (4.1.17). Proof of (b) From (a) we deduce that for any / € C™(M) we have
Lx(fuj) = (xm + fti(vx)n. On the other hand LxifSl)
= (ixd + dix){fQ)
=
dix{fQ).
Hence
{(Xf) + ftv(vx)}dv9 = d(ixfn). Since the form fQ is compactly supported we deduce from Stokes formula
[ d{ixfU) = 0. We have thus proved that for any compactly supported function / we have the equality - J ftr(VX)dvg JM
= [ Xfdv9 JM
= J (grad f,X)dvg=
= /
df{X)dvg
JM
f
{df,a)dvg.
Using Lemma 4.1.37 we deduce - J ftr{VX)dvg
= - [ f5advg
V/ € C0°°(M).
This completes the proof of (b). Proof of (c) We use the equality Lxiy/lgldx1
A • • • A dxn) = div {X^^dx1
A • • • A dxn).
The desired formula follows derivating in the left-hand-side. One uses the fact that Lx is an even s-derivation and the equalities Lxdxl
= diX'dx1
(no summation)
proved in Subsection 3.1.3. • Exercise 4.1.12 Let {M,g) be a Riemann manifold and X € Vect (M). Show that the following conditions on X are equivalent. (a) Lxg = 0. (b) g{VYX, Z) + g{Y, VZX) = 0 for all Y, Z e Vect (M). (A vector field X satisfying the above equivalent conditions is called a Killing vector field).
140
Riemannian
geometry
Exercise 4.1.13 Consider a Killing vector field X on the Riemann manifold (M, g) and denote by 77 the 1-form dual to X. Show that Srj — 0 i.e. in other words div (X) = 0. Definition 4.1.39 Let (M, g) be an oriented Riemann manifold (possibly with boundary). For any /c-forms a, @ define {a,/3) = {a,0)M=
J {a,/3)dvg=
JM
f a A */3
JM
whenever the integrals in the right-hand-side are finite. Let (M, g) be an oriented Riemann manifold with boundary dM. By definition, M is a closed subset of a boundary-less manifold M of the same dimension. Along dM we have a vector bundle decomposition
(TM)\9M=T(8M)®n where n = (TdM)1- is the orthogonal complement of TdM in (TM) \QM- Since both M and dM are oriented manifolds it follows that tt is a trivial line bundle. Indeed, over the boundary det TM = det {TdM) ® n so that n^detTM®det(TdM)*. In particular n admits nowhere vanishing sections and each such section defines an orientation in the fibers of n. An outer normal is a nowhere vanishing section a of n such that for each x £ dM and any positively oriented wx e det TxdM ax A w j i s a positively oriented element of det TXM. Since n carries a fiber metric we can select a unique outer normal of pointwise length = 1. This will be called the unit outer normal and will be denoted by v. Using partitions of unity we can extend v to a vector field defined on M. Proposition 4.1.40 (Integration by parts)Let (M, g) be a compact, oriented Riemann manifold with boundary, a G nk~l{M) and fi € ttk{M). Then / (da, P)dvg = J
(a A */?) \9M + [ (a, d*0)dvg
JdM
JM
= /
JdM
a\9M Ai(ip/3)\dM
JM
+ JM
[a,d*(3)dvg
where * denotes the Hodge ^-operator on dM with the induced metric g and orientation.
Metric properties
141
Using the (•, •) notation of Definition 4.1.39 we can rephrase the above equality as (da, /3)M = (a, ip/3)dM + (a, d*P)M. Proof of the proposition
As in the proof of Lemma 4.1.37 we have
(da, P)dvg = daA*P
= d(a A */?) + ( - l ) f c a Ad* p.
The first part of the proposition follows from Stokes formula arguing precisely as in Lemma 4.1.37. To prove the second part we have to check that (a A *P) \9M= a \9M Ai(ipP) \
m
.
This is a local (even a pointwise) assertion so we may as well assume M = H " = {(a;1,...,a;") G Mn ; x1 > 0} and the metric is the euclidian metric. Note that v = —d\. Let I be an ordered (k — l)-index and J an ordered k- index. Denote by Jc the ordered (n — fc)-index complementary to J so that (as sets) J U Jc = {1, ...,n}. By linearity, it suffices to consider only the cases a = dx1, P = dxJ. We have *dx] = ejdxJC
(ej = ±1)
(4.1.18)
and ivax
- j _dxj,
l e J
where J ' = J \ {1}. Note that if 1 £ J then 1 € Jc so that (a A *p) \BM= 0 = a \aM Ai(ipP) \gM so the only nontrivial situation left to be discussed is 1 €
= -*{dxJ')
= ~e'jdxJC
(e'j = ±1).
(4.1.19)
We have to compare the two signs ej and e'j. in (4.1.18) and (4.1.19). ej is the signature of the permutation JUJ C of {1,..., n} obtained by writing the two increasing multi-indices one after the other, first J and then Jc. Similarly, since the boundary dM has the orientation — dx2 A- • -Adxn, e'j is (—l)x(the signature of the permutation J ' U J c o f {2,...,n}). Obviously sign (JUJ c ) = sign (J'QJ c ) so that ej = —e'j. The proposition now follows from (4.1.18) and (4.1.19). • Corollary 4.1.41 (Gauss)Let {M,g) be a compact, oriented Riemann manifold with boundary and X a vector field on M. Then I div (X)dv9 = f JM
where9 9 = 5 | a M .
JdM
(X, u)dv9a
142 Proof
Riemannian
geometry
Denote by a the 1-form dual to X. We have / div (X)dv„ = / 1 A *d * adv„
JM
JM
= / (l,*d*a)dv9
= — /
JM
=
(l,d*a)dvg
JM
L„awdv*>= Ljx^dvn»- ° JoM
J oM
Remark 4.1.42 The compactness assumption on M can be replaced with an integrability condition on the forms a, 0 so that the previous results hold for noncompact manifolds as well provided all the integrals are finite. Definition 4.1.43 Let (M, g) be an oriented Riemann manifold. The geometric Laplacian is the linear operator A M : Cco(M) -> C°°(M) defined by A M = d*df = — *d*df
= - d i v (grad / ) .
A smooth function / on M satisfying the equation A M / = 0 is called harmonic. Using Proposition 4.1.36 we deduce that in local coordinates (a;1,..., xn) the geometric laplacian takes the form
AM = —)=di
(y/Wtdj)
where (g ,J ) denotes as usual the matrix inverse to (
A0 = -(a? + ... + a2) which differs from the physicists Laplacian by a sign. Corollary 4.1.44 (Green)Let (M,g) as in Proposition 4.1.40 and f,g€ Then (/, A M S ) M = (df, dg)M
C°°(M).
- (/, -~Z)BM-
{f, A M 5 ) M - ( A M / , S ) M = (^p, g)dM - (/,
^>aw-
Proof The first equality follows immediately from the integration by parts formula (Proposition 4.1.40) with a = f and /? = dg. The second identity is now obvious. •
143
Metric properties
Exercise 4.1.14 (a) Prove that the only harmonic functions on a compact oriented Riemann manifold M are the constant ones. (b) If u, f e C°°{M) are such that A M w = / show that
/ = o.
JM
Exercise 4.1.15 Denote by (w 1 ,... ,un) the coordinates on the round sphere Sn ' M n+1 obtained via the stereographic projection from the south pole. (a) Show that the round metric go on Sn is given in these coordinates by 00 =
1+ r
{(du1)2 + ••• +
(dunY}
where r 2 = (u 1 ) 2 H h (un)2. (b) Show that the n-dimensional "area" of Sn is r , 27r("+1>/2 n = n dv„90 Js r(a±i) where T is Euler's gamma function TOO
T(s) = / *s-1e~*1dt. Jo Hint: Need to know the "doubling formula"
7r1/2r(2s) = 2 2s - 1 r(s)r(s +1/2) and (see [31]) -dr = Vo (1 + r )" ' ~ Jo 2
(r(n/2)) 2 2r(n)
Exercise 4.1.16 Consider the Killing form on su(2) (the Lie algebra of SU(2)) defined by
(X,Y) =
-trX-Y.
(a) Show that the Killing form defines a bi-invariant metric on SU(2) and then compute the volume of the group with respect to this metric. SU(2) is given the orientation defined by ex A e2 A e$ £ A 3 SM(2) where e* € su{2) are the Pauli matrices ex
[ i
01
0 -i
[ 0 1 1 -1 0
e2 =
e3 =
[Oil i 0
(b) Show that the trilinear form on su(2) defined by
B(X,Y,Z)
= ([X,Y},Z) 3
is skew-symmetric. In particular B e A su(2)*. (c) B has a unique extension as a left-invariant 3-form on SU(2) known as the Cartan form on SU{2)) which we continue to denote by B. Compute !su^2) B.
144
Riemannian
geometry
Hint: Use the natural diffeomorphism SU(2) = S 3 and the computations in the previous exercise.
4.2
The Riemann curvature
A Riemann metric on a manifold has roughly speaking the effect of "giving a shape" to the manifold. Thus a very short (in diameter) manifold is different from a very long one. A large (in volume) manifold is different from a small one. However there is a lot more information encoded in the Riemann manifold than just its size. To recover it we need to look deeper in the structure and go beyond the first order approximations we have used so far. The Riemann curvature tensor achieves just that. It is an object which is very rich in informations about the "shape" of a manifold and loosely speaking provides a second order approximation to the geometry of the manifold. As Riemann himself observed we do not need to go beyond this order of approximation to recover all the informations. In this section we introduce the reader to the Riemann curvature tensor and its associates. We will describe some special examples and we will conclude with the Gauss-Bonnet theorem which shows that the local object which is the Riemann curvature has global effects. §4.2.1
Definitions and properties
Let (M, g) be a Riemann manifold and denote by V the Levi-Civita connection. Definition 4.2.1 The Riemann curvature denoted by R = R(g) is defined as R{g) = F(V) where F(V) is the curvature of the Levi-Civita connection. The Riemann curvature is a tensor R e fi2(End (TM)) explicitly defined by R{X, Y)Z = [ V x , Vy]Z -
V[XtY]Z.
In local coordinates (a;1, ...,xn) we have the description
4 ^ =^,3^. In terms of the Christoffel symbols we have r>i _ % -pi
n
ijk
— ujL
ik
a r ' x r ' r m _ r^
u L
k
ij "+" L mjL ik
L
rm
mkL ij •
Lowering the indices we get a new tensor Rijkt = 9imRfke = {R{dk, dt)dj, d{).
The Riemann
145
curvature
Theorem 4.2.2 (The symmetries of the curvature tensor) The Riemann curvature tensor R satisfies the following identities {X, Y, Z,U,V G Vect (M)). (a) g(R(X, Y)U, V) = -g(R{(Y, X), U, V). (b) g(R(X,Y)U,V) = -g{R{X,Y)V,U). (c) (The 1st Bianchi identity) R{X, Y)Z + R{Z, X)Y + R(Y, Z)X = 0. (d) g{R{X, Y)U, V) = g(R(U, V)X,
Y).
(e) (The 2nd Bianchi identity) (VXR)(Y,Z)
+ (VYR)(Z,X)
+ (VZR){X,Y)
= 0
In local coordinates the above identities take the form Rijkl = —Rjikt — —RijikRijkt — Rklij)kt + R\jk + R'ktj = 0-
R
(ViR)Le + {ViR)U + (VtR)L = 0. Proof (a) It follows immediately from the definition of R as an End(TM)-valued skew-symmetric bilinear map (X, Y) i-> R(X,Y). (b) We have to show the symmetric bilinear form Q(U, V) = g(R(X, Y)U, V) + g(R(X, Y)V, U) is trivial. Thus it suffices to check Q(U, U) = 0. We may as well assume that [X, Y] = 0 since (locally) X, Y can be written as linear combinations (over C°°(M)) of commuting vector fields. (E.g. X = Xldi). Then Q(U, U) =
ff((VxVy
- VYVX)U,
U).
We compute Y(Xg(U,U))
=
2Yg(VxU,U)
= 2(VyVXU, U) + 2g{VxU,
VYU)
and similarly X(Yg(U, U)) = 2g(VxVYU,
U) + 2g(VxU,
VyU).
Subtracting the two equalities we deduce (b). (c) As before, we can assume the vector fields X, Y, Z pairwise commute. The 1st Bianchi identity is then equivalent to VXVYZ
- VY\7XZ
+ VZVXY
- VXVZY
+ VYVZX
- VZVYX
= 0.
The identity now follows from the symmetry of the connection: V x ^ = VYX (d) We will use the following algebraic lemma ([44], Chapter 5).
etc.
146
Riemannian
geometry
Lemma 4.2.3 Let R : E x E x E x E ->• 1R be a quadrilinear map on a real vector space E. Define S(Xi, X2, X3, X4) = R{X\, X2, X3, X 4 ) + R(X2, X3, Xi, X4) +
R(X3,Xi,X2,X4).
If R satisfies the symmetry conditions R{X\, Xi-, X3, X4) = —R(X2, X\, X3, Xi) R(Xi, X2, X3, X^ — —R(Xi,X2,
X$, X3)
then R(Xi, X2, X3, X4) — R(X3, X^ X\, X2) =
7s \S(Xi, X2, X3, Xi) — S(X2, X3, Xi, X\) — S(X3, Xi, X\, X2) + S(Xi, X3, X\, X2)} •
The proof of the lemma is a straightforward (but tedious ) computation which is left to the reader. The Riemann curvature R = g{R{Xx, X2)X3, Xi) satisfies the symmetries required in the lemma and moreover, the 1st Bianchi identity shows the associated form 5 is identically zero. This concludes the proof of (d). (e) This is the Bianchi identity we established for any linear connection (see Exercise 3.3.3). • The Riemann curvature tensor is the source of many important invariants associated to a Riemann manifold. We begin by presenting the simplest ones. Definition 4.2.4 Let (M, g) be a Riemann manifold with curvature tensor R. Any two vector fields X, Y on M define an endomorphism of TM by U M- R(U, X)Y. The Ricci curvature is the trace of this endomorphism i.e. Ric {X, Y) = tr {U
H->
R(U,
We view it as a (0,2)-tensor (X, Y) H* Ric {X, Y) e
X)Y). CX(M).
If (x1,..., xn) are local coordinates on M and the curvature R has the local expression R = (Rekij) then the Ricci curvature has the local description Ric=(i*y) = £ ^ « I
The symmetries of the Riemann curvature imply that Ric is a symmetric (0,2)-tensor (as the metric).
The Riemann
147
curvature
Definition 4.2.5 The scalar curvature s of a Riemann manifold is the trace of the Ricci tensor i.e. in local coordinates s = gijRio = gijR^j where as usual (g*i) is the inverse matrix of (;J).
X,Y
Let (M, g) be a Riemann manifold and p e M. e TPM set KP{X,Y)-
For any linearly independent
| X A y |
where \X A Y\ denotes the Gramm determinant
\XAY\
(X,X) (Y,X)
(X,Y) (Y,Y)
and is non-zero since X and Y are linearly independent. (\X A Y\ll2 measures the area of the parallelogram in TPM spanned by X and Y.) Exercise 4.2.1 Let X, Y,Z,W 6 TPM such that span(X,Y)= span(Z,W) is a 2dimensional subspace of TPM prove that KP(X, Y) = KP(Z, W). According to the above exercise the quantity KP{X, Y) depends only upon the 2plane in TPM generated by X and Y. Thus Kp is in fact a function on Gr2(p) the grassmannian of 2-dimensional subspaces of TPM. Definition 4.2.6 The function Kp : Gr2{p) —> M defined above is called the sectional curvature of M at p. Exercise 4.2.2 Prove that Gr2(M) = disjoint union of Gr2{p)
p GM
can be organized as a smooth fiber bundle over M with standard fiber Gr2^n(R), n = d i m M such that if M is a Riemann manifold Gr2{M) 3 (p;7r) i-> Kp(n) is a smooth map. §4.2.2
Examples
Example 4.2.7 Consider again the situation discussed in Example 4.1.12. Thus G is a Lie group and (•, •) is a metric on £ G satisfying (od{X)Y, Z) = -(Y,
ad(X)Z)
148
Riemannian
geometry
In other words, (•, •) is the restriction of a bi-invariant metric m on G. We have shown that the Levi-Civita connection of this metric is
VXY = ±[X,Y] vx,re£ G . We can now easily compute the curvature R(X, Y)Z = i {[X, [Y, Z}} - [Y, [X, Z}}} - ±[[X, Y], Z] (Jacobi identity) = ±[[X, Y], Z] + hjY, [X, Z\\ - ~{Y, [X, Z}} - ±][X, Y], Z\ =
-\[lX,Y},Z}.
We deduce (R(X,Y)Z,W)
= -\{[[X,Y),Z],W)
= -\{{X,Y],ad{Z)W)
=
=
\(ad(Z)[X,Y},W)
-\{[X,Y],[Z,W}).
Now let 7r € Gr2(g) for some g £ G. If (X, Y) is an orthonormal basis of TT (viewed as left invariant vector fields on G) then the sectional curvature along 7r is Kg(rr) =
^{[X,Y},{X,Y})>0
Denote the Killing form by K(X,Y) — —tr (ad(X)ad(Y)). To compute the Ricci curvature we pick an orthonormal basis E\,...,En of £ Q . For any X = X'Ei,Y = YjEj e £ G we have Ric(X,Y)
=
±ti(Z^[[X,Z\,Y])
= \Yt([[X,Ei],Y],Ei)) =
—ZiadWfrEhEi)
i
i
= - £ < [ X , Et\, [Y, Ei]) = \ Y,{ad{X)Eu i
ad(Y)Et)
i
= -^(ad(Y)ad{X)Ei,Ei)
= -^tr
(ad(Y)ad(X))
i
=
\K{X,Y).
In particular, on a compact semisimple Lie group the Ricci curvature is a symmetric positive definite (0,2)-tensor (in fact it is a scalar multiple of the Killing metric.)
The Riemann
curvature
149
We can now easily compute the scalar curvature. Using the same notations as above we get s
= \ E
Ric
($> Ei) = \ £ <Ei> Ei)-
In particular, if G is a compact semisimple group and the metric is given by the Killing form then the scalar curvature is S(K) = - dimG. 4
Remark 4.2.8 Many problems in topology lead to a slightly more general situation than the one discussed in the above example namely to metrics on Lie groups which are only left invariant. Although the results are not as "crisp" as in the bi-invariant case many nice things do happen. For details we refer to [57]. Example 4.2.9 Let M be a 2-dimensional Riemann manifold (surface) and consider local coordinates on M, (x 1 ,^ 2 ). Due to the symmetries of R Rijkl = —Rijlk = Rklij
we deduce that the only nontrivial component of the Riemann tensor is R = #1212The sectional curvature is simply a function on M K = y-R12n
= rs(s)
where \g\ = d e t a -
in this case K is known as the total curvature or the Gauss curvature of the surface. If in particular M is oriented and the form dx1 A dx2 defines the orientation, we can construct a 2-form e
(ff) = —Kdvg 9 = — s(g)dv g9 = v; 27r"""' ~47r"
"
j^Ruudx1
A dx2
2TTJ|5|
e(g) is called the Euler form associated to the metric g. We want to emphasize that this form is defined only when M is oriented. We can rewrite this using the pfaffian construction of Subsection 2.2.4. The curvature R is a 2-form with coefficients in the bundle of skew-symmetric endomorphisms of TM so we can write R = A
A =
0 #2112
i?1212 0
150
Riemannian
geometry
Assume for simplicity that (x 1 ,^ 2 ) are normal coordinates at a point q € M. Thus at q, \g\ = 1 since d\, d2 is an orthonormal basis of TqM. Hence at q, dvg = dx1 A dx2 and ^-g(R(di,d2)d2,d1)dx1Adx2.
e{g) = Z7T
Hence we can write
*!
^Pfa{-R).
The Euler form has a very nice interpretation in terms of holonomy. Assume as before that (x1 ,x2) are normal coordinates at q and consider the square St = [0, \/i\ x [0, y/i\ in the (xx,x2) plane. Denote the (counterclockwise) holonomy along dSt by %t- This is an orthogonal transformation of TqM and with respect to the orthogonal basis (d\, 82) of TqM it has a matrix description as _ _ cos9{t) l ~ [ sin0(i)
—sm0{t) cos 9(t)
'
The result in Subsection 3.3.4 can be rephrased as sin 9{t) = -tg(R{dud2)d2,
di) + 0{t2)
so that #1212 = (0). Hence #1212 is simply the infinitesimal angle measuring the infinitesimal rotation suffered by d\ along St- We can think of the Euler form as a "density" of holonomy since it measures the holonomy per elementary parallelogram. §4.2.3
Cartan's moving frame method
This method was introduced by Elie Cartan at the beginning of this century. Cartan's insight was that the local properties of a manifold equipped with a geometric structure can be very well understood if one knows how the frames of the tangent bundle (compatible with the geometric structure) vary from one point of the manifold to another. We will begin our discussion with the model case of K n . Example 4.2.10 Consider Xa = X^-^, a = 1, ...,n an orthonormal moving frame on K" where (x1, ...,xn) are the usual cartesian coordinates. Denote by (0°) the dual coframe i.e. the moving frame of T*K™ defined by
e°(x0) = si. The 1-forms measure the infinitesimal displacement of a point P with respect to the frame (X Q ). Note that the TM-valued 1-form 9 = 9aXa is the differential of the identity map id : M™ —> R" expressed using the given moving frame.
The Riemann
curvature
151
Introduce the 1-forms wjj defined by dX& = w%Xa
(4.2.1)
where we set
We can form the matrix valued 1-form (u>f) which measures the infinitesimal rotation suffered by the moving frame (Xa) following the infinitesimal displacement x i-> x + dx. In particular u) = (u/g) is a skew-symmetric matrix since 0 = d(Xa,Xf,)
= (uXa,Xp)
+
(Xa,uXi,).
Since 9 = did we deduce d& = 0 and we can rewrite this as dOa = 90 A ul or dO = -w A 9.
(4.2.2)
Using d/'Xp = 0 in (4.2.1) we deduce dw'p = — w" A UJJ or equivalently dw = - w A w. (4.2.3) The equations (4.2.2)-(4.2.3) are called the structural equations of the Euclidian space. The significance of these structural equations will become evident in a little while. We now try to perform the same computations on an arbitrary Riemann manifold M. We choose a local orthonormal moving frame (Xa) and construct similarly its dual coframe (9°). Unfortunately, there is no natural way to define dXa to produce the forms w2 entering the structural equations. We will find them using a different (dual) search strategy. Proposition 4.2.11 (E. Cartan)There exists a collection of 1-forms a/g uniquely defined by the requirements (a)wjj = -wg. {b)d9a = 0? A 0$, Va. Proof Uniqueness Suppose a/g satisfy the conditions (a)&(b) above. Then there exist functions / ^ and gpy such that
<4 = ftJ1
152
Riemannian
geometry
Then the condition (a) is equivalent to
(al) fa = - / £ while (b) gives
(bl) fa ~ f°p = 9$, The above two relations uniquely determine the / ' s in terms of the g's via a cyclic permutation of the indices a, j3, 7 #7 = ^ Existence
7
+
«&-*&)
(4-2.4)
Consider the functions g^ defined by dea = \ g ^ ^
{g% =
-galfi).
Next define CJ2 = fS^O1 where the / ' s are given by (4.2.4). We let the reader check that the forms wjj satisfy both (a) and (b). • The reader may now ask why go through all this trouble. What have we gained by constructing the forms w and after all, what is their significance? To answer these questions consider the Levi-Civita connection V. Define u>0 by VX^ = u}pXa. Hence VXlX0 = w|(X 7 )X t t . Since V is compatible with the Riemann metric we deduce in standard manner that (a) C$ = -<%.
The differential of 9" can be computed in terms of the Levi-Civita connection (see Subsection 4.1.5) and we have d9a(X0,X7)
= X09a{X7)
- X^iXp)
- 9a(VXpX7)
(use 9a{Xp) = 6% = const) = -9a(v67{X0)Xs)
+ +
9a(VXiX0) 9a{^(X7)Xs)
= tf(X7) - &°{XP) = (P A tfXXf,, X7). Thus the w's satisfy both conditions (a) and (b) of Proposition 4.2.11 so that we must have u>$ = w$.
In other words, the matrix valued 1-form (ujf) is the 1-form associated to the LeviCivita connection in this local moving frame. In particular, the 2-form Q = (did + u A u )
The Riemann
153
curvature
is the Riemannian curvature (see the computation in Example 3.3.11). The structural equations of a Riemann manifold take the form d9 = -u> A 9 du) + w A u) = fl. Comparing these with the euclidian structural equations we deduce another interpretation of the Riemann curvature: it measures "the distance" between the given Riemann metric and the euclidian one". We refer to [70] for more details on this aspect of the Riemann tensor. The technique of orthonormal frames is extremely versatile in concrete computations. Example 4.2.12 We will use the moving frame method to compute the curvature of the hyperbolic plane i.e. the upper half space H + = {(!,»);
y>0}
endowed with the metric g = —(dx2 + dy2). {ydx, ydy) is an orthonormal moving frame and (6X = -dx, 0y = -dy) is its dual coframe. We compute easily d0x = d(-dx) y
= ~dx Ady = (-dx) A 6V y2 y
dOy = d(-dy) = 0 = (--dx) y
A 0*.
y
Thus the connection 1-form in this local moving frame is ' 0 1.
.v
•
_ i
y
0
dx.
The Riemann curvature is
0
Q = du> =
i
i y
--V o
dy A dx :
0 1
-1 0
y2
The Gauss curvature is 1
K = y-Mto(dx,dy)dy,dx) = y4{-
=-1.
154
Riemannian
§4.2.4
geometry
The geometry of submanifolds
We now want to apply Cartan's method of moving frames to discuss the the local geometry of submanifolds of a Riemann manifold. Let (M, g) be a Riemann manifold of dimension m and S a ^-dimensional submanifold in M. The restriction of g to S induces a Riemann metric on S. We want to analyze the relationship between the Riemann geometry of M (assumed to be known) and the geometry of S with the induced metric. Denote by V M (resp. V s ) the Levi-Civita connection on (M, g) (resp on (S,g|s). The metric g produces an orthogonal splitting of vector bundles (TM)\S=TS@NS. Ns is called the normal bundle of S <-> M. Thus, a section of (TM) \s, that is a vector field X of M along S, splits into two components: a tangential component XT and a normal component X". Now choose a local orthonormal moving frame (Xi, ...,Xk;Xk+i, •••,Xm) such that the first k vectors (Xi,...,Xk) are tangent to S. Denote the dual coframe by (6a)i k. Denote by (nf), (1 < a, f5 < m) the connection 1-forms associated to V M in this frame and let crfi, (1 < a, ft < k) be the connection 1-forms of V 5 . We will analyze the structural equations of M restricted to S <-> M. d0a = 0" A [fp 1 < a, 0 < m.
(4.2.5)
We distinguish two situations. A . 1 < a < k. Since 9P | s = 0 for /? > k the equality (4.2.5) yields
^Q = E ^ A ^ ,
I$ = -I4
l
0=1
The uniqueness part of Proposition 4.2.11 implies that along S aap=Hap
l
This can be equivalently rephrased as VSXY = ( V ^ F ) T B.
VX, Y e Vect (5).
k < a < m. We deduce 0=i
At this point we want to use the following elementary result.
(4.2.6)
The Riemann
155
curvature
Exercise 4.2.3 (Cartan's Lemma) Let V be a d-dimensional real vector space and consider p linearly independent elements u^, ...,ui p € AlV, p < d. If #1, ..., 8p 6 AlV are such that
then there exist scalars Aij, 1 < i,j < p such that Atj = Ajt and v i=i
Using Cartan's Lemma we can find smooth functions fpy, A > f c , 1 < (3, 7 < fc satisfying //37
=
A/3-
^ = /^7Now form We can view 01 as a symmetric bilinear map Vect (5) x Vect (5) -> C°°(Ars). If[/,VeVect(S)
then
U = Ul3Xl3=efi{U)Xp
\3
V = VXJ=8"<(V)XJ
l<-y
W. n = EIE fe /^7(*o) ^ 1 *A A>A: (_ 0
\ 1
1
)
= E(E/W'W A>k \ £
/
The last term is precisely the normal component of VyU so that we have established (v£ft/)" = W(U, V) = *tt(V, £/) = ( v £ v ) " .
(4.2.7)
9t is called the 2nd fundamental form of S <-» M. (The first fundamental form is the induced metric). There is an alternative way of looking at 9t. Choose U, V G Vect(S), N G C°°(NS). We have (g(-, •) = <-,•»
<9i(y, v), iv) = <(v£f v)", JV) = «
v, TV>
156
Riemannian =
geometry
V%{V,N)-{V,V}fN) =
-(V,(V^N)T).
We have thus established -(V, (V2?JV)T> = (Vl(U, V), N) = («rt(V, U), N) = -(U, (vjf JV)T>.
(4.2.8)
The 2nd fundamental form can be used to determine a relationship between the curvature of M and that of S. More precisely we have the following celebrated result. T H E O R E M A E G R E G I U M (Gauss) Let RM (resp. Rs) denote the Riemann curvature of (M, g) (resp. (S, g | s ) . Then for any X, Y,Z,Te Vect (5) we have (RM{X,Y)Z,T)
=
(RS(X,Y)Z,T)
+(w(x, z), ^(y; T)) - {m(x, r), m(Y, z)). Proof
(4.2.9)
Note that \>%Y = VsxY +
Vl(X,Y).
We have RM{X, Y)X = [V^, V f )Z -
V$X]Z
= V^f ( V y Z + *rt(y, Z)) - V f ( V | Z + OT(X, Z)) - Vf x , y ] Z - 9t([X, Y], Z). Take the inner product with T of both sides above. Since 9fl(-, •) is iVs-valued we deduce using (4.2.6)-(4.2.8) (RM(X, Y)Z, T) = ( V ^ V y Z , T) + (V$f*tt(Y, Z), T)
-(vf
V^Z.T)
-
= ([v|, v y ]z, T) - <«n(y, z),m(x, T » +(9fi(A-,z),«n(y;T)) - (vfx,y]z,T). This is precisely the equality (4.2.9). • . The above result is especially interesting when 5 is a transversally oriented hypersurface i.e. a codimension 1 submanifold such that the normal bundle Ns is trivial. (Locally, all hypersurfaces are transversally oriented since Ns is locally trivial by definition). Pick n an orthonormal frame of Ns (i.e. a length 1 section of Ns) and pick an orthonormal moving frame (Xi,..., X m _i) of TS- Then (Xi,..., Xm-i, n) is an orthonormal moving frame of (TM) \s and the second fundamental form is completely described by
mfi(x,Y) = {m(x,Y),n).
The Riemann
157
curvature
?&n is a bona-fide symmetric bilinear form and moreover according to (4.2.8) we have 9t*(X,Y) = - < V £ n , Y ) =
-(VPn,X).
Gauss formula becomes in this case (RS(X,Y)Z,T)
= {RM(X,Y)Z,T)
-
OT«(X,Z) mg{Y,Z)
*fln(X,T) 9fla(y,T)
Let us further specialize and assume M = K m . Then (RS{X,Y)Z,T)
=
Ol«(y,T)
(4.2.10)
*Hn(F,^)
In particular, the sectional curvature along the plane spanned by X, Y is {Rs(X,Y)Y,X)=Vln(X,X)-Vlii(Y,Y)-\Vln(X,Y)\2. This is a truly remarkable result. On the right-hand-side we have an extrinsic term (it depends on the "space surrounding S ") while in the left-hand-side we have a purely intrinsic term (which is defined entirely in terms of the internal geometry of 5). Historically, the extrinsic term was discovered first (by Gauss) and very much to Gauss surprise (?!?) one does not need to look outside S to compute it. This marked the beginning of a new era in geometry. It changed dramatically the way people looked at manifolds and thus it fully deserves the name of The Golden (egregium) Theorem of Geometrv. Example 4.2.13 (Quadrics in M3) Let A be a selfadjoint, invertible —> linear operator with at least one positive eigenvalue. This implies the quadric QA = {ueR3
; {Au,u) = 1}
is nonempty and smooth (use implicit function theorem to check this). Let MO € QAThen TU0QA = {x e R 3 ; (Au0,x) = 0} = {AUQ)2-. QA is a transversally oriented hypersurface in R3 since the map QA B U >-» Au defines a nowhere vanishing section of the normal bundle. Set n = rfiAu. Consider °
\Au\
(eo,ei,e2) an orthonormal frame of R 3 such that e0 = n(uo). Denote the cartesian coordinates in R3 with respect to this frame by (a: 0 ,! 1 ,^ 2 ) and set <% = ^ - . Extend (ei, e2) to a local moving frame of TQA near MoThe second fundamental form of QA at u0 is ^h{di,dj)
=
(difi,dj)\
158
Riemannian
geometry
We compute dtti = di ( 4 ^ 7 ) = 9i((Au, Au)-V2)Au \\Au\J _
(diAu, Au) -Au\Au\V2
Hence (
yin(di,dj)\uo 1
+
~Adti \Au\
" mdiAu-
1 {AdiU, ej) |„ 0 \Au0\ e
(diu,Aej)\V0=
iy
Aej).
(4.2.11)
We can now compute the Gaussian curvature at uQ. Kun =
1 \Au0\*
(i4e!,ei) (Ae2,e1)
(Aeue2) (Ae2,e2)
In particular, when A = r 2I so that QA is the round sphere of radius r we deduce K
" = ^ Vlul
Thus, the round sphere has constant positive curvature. We can now explain rigorously why we cannot wrap a plane canvas around the sphere. Notice that when we deform a plane canvas the only thing that changes is the extrinsic geometry while the intrinsic geometry is not changed since the lengths of the "fibers" stays the same. Thus, any intrinsic quantity is invariant under "bending". In particular, no matter how we deform the plane canvas we will always get a surface with Gauss curvature 0 which cannot be wrapped on a surface of constant positive curvature! Gauss himself called the total curvature a "bending invariant". Example 4.2.14 (Gauss)Let E be a transversally oriented, compact surface in M3. (E.g. a connected sum of g tori). Note that the Whitney sum N^ © T E is the trivial bundle R | . We orient N% such that orientation JVS A orientation T E = orientation M3. Let n be the unit section of N-^ defining the above orientation. We obtain in this way a map <J5 : E -> S2 = {u e K3 ; |u| = 1} E 3 i 4 n(x) € S2. & is called the G a u s s map of E •->• S2. It really depends on how E is embedded in M3 so it is an extrinsic object. Denote by 91^ the second fundamental form of E •-> M3 and let (xl,x2) be normal coordinates at q 6 E such that orientationTgE = Si A 62-
The Riemann
159
curvature
Consider en the Euler form on E with the metric induced by the euclidian metric in M3. Then, taking into account our orientation conventions, we have 2irex(d1,d2) = Rnn «a(9i,0i)
Wn(di,di)
(4.2.12)
Now notice that din = -«tta(Qj, dx)di - Vlgidi, d2)d2. We can think of n,di \q and d2 \q as defining a (positively oriented) K3. The last equality can be rephrased by saying that the derivative of the Gauss map (8>t : T,E —> Ts(,)5 acts according to $ ^ -9ta(Qj, 9i)9i - 0 ^ ( 9 i ; 9 2 )9 2 . In particular, we deduce <25, preserves (reverses) orientations •<=>• i?i2i2 > 0 (< 0)
(4.2.13)
because the orientability issue is decided by the sign of the determinant 1 0 0 -Vflnidudi) 0 -*fln(d2,di)
0 -mn(dl,d2) -QM^.ft)
At q, di _L 9 2 so that (a ; n, 8,-fi) = gna(ft, 9001^(9,, 9X) + 5 ^ ^ , ^ S t ^ a , - , 9 2 ). We can rephrase this coherently as an equality of matrices {d\fi, d\n) (d2n, dxn) Vln(di,di) ^(92,90
(din, d2n) (d2n, d2n)
Qflfl(9i,92) Vln(d2,d2)
*>Mdi,ai) OTa^.ft)
^ln(9i,9 2 ) 5rin(9 2 ,9 2 )
Hence <
yin(dudl) ^n(9i,92)
VU(dud2) ^-(92,92)
( 9 ^ , d\n) (din, d2n)
(din, d2n) (d2n, d2n)
(4.2.14)
If we denote by dv0 the metric volume form on S2 induced by the restriction of the euclidian metric on M3 we see that (4.2.12) and (4.2.14) put together yield 27:\es(di,d2)\
= \dv0(din,d2n)\
=
\dv0{&,{di),&,{d2))\.
160
Riemannian
Using (4.2.13) we get
e = & Edv =
* h * ° h&^£s2-
geometry
(4 2 15)
--
This is one form of the celebrated Gauss-Bonnet theorem . We will have more to say about it in the next subsection. Note that the last equality offers yet another interpretation of the Gauss curvature. Prom this point of view the curvature is a "distortion factor". The Gauss map "stretches" an infinitesimal parallelogram to some infinitesimal region on the unit sphere. The Gauss curvature describes by what factor the area of this parallelogram was changed. §4.2.5
The Gauss-Bonnet theorem
We conclude this chapter with one of the most beautiful results in geometry. Its meaning reaches deep inside the structure of a manifold and can be viewed as the origin of many fertile ideas. Recall one of the questions we formulated at the beginning of our study: explain unambiguously why a sphere is "different" from a torus. This may sound like forcing our way in through an open door since everybody can "see" they are different. Unfortunately this is not a conclusive explanation since we can see only 3-dimensional things and possibly there are many ways to deform a surface outside our tight 3D Universe. The elements of Riemann geometry we discussed so far will allow us to produce an invariant powerful enough to distinguish a sphere from a torus. But it will do more than that. Theorem 4.2.15 (Gauss-Bonnet: preliminary version)Let S be a compact oriented surface without boundary. If po a n d
Jse90(S) = Jse91(S). Hence the quantity fs £g(S) is independent of the Riemann metric g so that it really depends only on the topology of S!!! The idea behind the proof is very natural. Denote by gt the metric gt = go + t{gi — So)- We will show
! / / * = ov* e [ o,i}. It is convenient to consider a more general problem. Definition 4.2.16 Let M be a compact oriented manifold. For any Riemann metric g on E define <£() = / JM
s{g)dvg
The Riemann
161
curvature
where s(g) denotes the scalar curvature of {M,g). Einstein functional.
<£(g) is called the Hilbert-
We have the following remarkable result. Lemma 4.2.17 Let M be a compact oriented manifold without boundary and gl = {g\j) be a 1-parameter family of Riemann metrics on M depending smoothly upon t e R. Then
-£<*(<) = - / (Ric - ^ ( f f V . f l ' M t y at
JM
I
Vt.
In the above formula (•, -)t denotes the inner product induced by gl on the space of symmetric (0,2)-tensors while the dot denotes the i-derivative. The (0, 2)-tensor -s{x)gij{x) is called the Einstein tensor of (M, g). Definition 4.2.18 A Riemann manifold (M,g) is said to be Einstein if the metric g satisfies Einstein's equation Ric 9 =
^s(g)g.
Remark 4.2.19 The attribute Einstein usually refers to a larger class of Riemann manifolds satisfying the condition Ric9(:r) =
\(x)g(x)
for some smooth function A € Cca(M). We refer to [12] for a comprehensive presentation of this subject. In this book we reserve the name Einstein to the special case described in the above theorem. Exercise 4.2.4 Consider a 3-dimensional Riemann manifold {M,g). Show that Rijkt = £ik9ji - £ie9jk + £ji9ik — £jk9u + 7}{9ie9jk — 9ik9ji)In particular, this shows that on a Riemann 3-manifold the full Riemann tensor is completely determined by the Einstein tensor. Exercise 4.2.5 (Schouten-Struik, [67])Prove that the scalar curvature of an Einstein manifold of dimension > 3 is constant.
162
Riemannian
geometry
Hint Use the 2nd Bianchi identity. Notice that when (S, g) is a compact oriented Riemann surface two very nice things happen. (i) (S, g) is Einstein. (Recall that only #1212 is nontrivial). (ii)C(p) = 2 / s £ 8 . Theorem 4.2.15 is thus an immediate consequence of Lemma 4.2.17. Proof of the lemma We will produce a very explicit description of the integrand (*(»')*¥ = = (-(s('))*V + s (+s5(9 ' )t)(j ( -t d v ) ^(s(fl')dty
(4.2.16)
of ^<£((). We will adopt a "get down in the mud and just do it" strategy reminiscent to the good old days of the tensor calculus frenzy. By this we mean that we will work in a nicely chosen collection of local coordinates and keep track of the zillion indices we will encounter. As we will see, "life in the mud" isn't as bad as it may seem to be. We will study the integrand (4.2.16) at t = 0. The general case is entirely analogous. For typographical reasons we will be forced to introduce new notations. Thus, g will denote (gl) for t = 0, while gt will be denoted simply by g. A hat over a quantity means we think of that quantity at t = 0 while a dot means differentiation at t = 0. Let q be an arbitrary point on S and denote by (a;1,..., xn) a collection of g-normal coordinates at q. Denote by V the Levi-Civita connection of g and let rj fc denote its Christoffel symbols in the coordinates (x'). Many nice things happen at q and we list a few of them which will be used later. g^ = g,3 = Sij
dk9ij = o.
Vidj = 0 f} fc = 0.
(4.2.17) (4.2.18)
If a = a,dx' is a 1-form then at q 5tja = Yldiai.
{5 = *d*)
(4.2.19)
i
In particular, for any smooth function u (AMiSu)(q)
= -J29i2u. i
Set h = (hi,) = (5) = (<)„). his a, symmetric (0,2)-tensor. Its ^-trace is the scalar tTgh =
giihij=tiC-1(h)
(4.2.20)
The Riemann curvature
163
where C is the lowering the indices operator defined by g. In particular a t q
%x~ah = Y,h«-
(4-2-21)
The curvature of g is given by R-ikj
=
~^Hjk
=
d>^ij — dfiik + ^mk^Tj ~ r m j T ^ .
The Ricci tensor is T> "ij
T>k o pA: o -pk . pfc T^m pA: r m — "ikj — °ki ij " °jL ik + l mkL ij ~ l mjl ik-
Finally the scalar curvature is s = tr 9 Rtj = gl3Rij
= j* (^r* - art + r^r^ - r^.rs). Derivating s at t = 0 and then evaluating a t q we obtain
s = gij (&f£ - djt 1k) + s» (fci* - dji*) = g^Rij + Yl{dktki-ditkk).
(4.2.22)
i
The term ,J can be computed by derivating the equality g'kgjk = S'k at t = 0. We get so that g'j = - / l y ,
(4.2.23)
To evaluate the derivatives T's we use the known formula? Fv = \
+ djgik)
which upon derivation at t = 0 yield II? = \ {di9ik ~ dkgij + dj9ik) + ±gkm (dihjk - dkh{j + djhik) = ^ (dihim ~ dmhij + djhim). We substitute (4.2.23) -(4.2.24) in (4.2.22) and we get a t q s = - X) hijRij + - Y,(dkdihik ij
i,k
- dk2hu + dkdihik)
(4.2.24)
164
Riemannian ~ o Jl(di2hkk
-
geometry
didkhik)
= - X) hijRij - X di2hkk + X didkhik i,j
i,k
i,k
= -(Ric,)j + AMjtii)g
+ Y,didkhik.
(4.2.25)
i,k
To get a coordinate free description of the last term note that a t q {Vkh)(di,dm)
= dkhim.
The total covariant derivative V/i is a (0,3)-tensor. Using the g-trace we can construct a (O.l)-tensor trj(V/t) = t r ( £ r ! v / i ) where C^1 is the raising the indices operator defined by g. In the local coordinates (x') we have Using (4.2.17) and (4.2.19) we deduce that the last term in (4.2.25) can be rewritten (at q ) as 5trg(Vh) = Strg{Vg). We have thus established that s = -(Ric,g)g
+ AM,gtrgg
+ 6trg(Vg).
(4.2.26)
The second term of the integrand (4.2.16) is a lot easier to compute. dvg = ±yj\g\dxx
• ••dxn
so that
dv9 =
±\\g\-^jt\g\dxl.-.dx\
A t q the metric is Euclidian, gy = 6ij and jt\g\
= Y,9u = \g\-tTS(g) =
\g\{g,g)g\g\.
Hence £(ff) = /
{~s(9)9-Ric{g),g)gdvg
+ / AM^gttg g +
5tig(Vg)dvg.
JM
The Green formula shows the last two terms vanish and the proof of the Lemma is concluded. D
The Riemann
curvature
165
Definition 4.2.20 Let S be a compact, oriented surface without boundary. define its Euler characteristic as the number
We
where g is an arbitrary Riemann metric on S. The number l
9(S) =
-(2-X(S))
is called the genus of the surface. Remark 4.2.21 According to the theorem we have just proved the Euler characteristic is independent of the metric used to define it. Hence, the Euler characteristic is a topological invariant of the surface. The reason for this terminology will become apparent when we discuss DeRham cohomology, a Z-graded vector space naturally associated to a surface whose Euler characteristic coincides with the number defined above. So far we have no idea whether x(S) is even an integer. Proposition 4.2.22 X (S
Proof
2
) = 2 and
X (T
2
) = 0.
To compute x{S2) we use the round metric 50 for which K = 1 so that x(5 2 ) = ^ /
s 2
^ o = ^ a r e a 9 0 ( 5 2 ) = 2.
To compute the Euler characteristic of the torus we think of it as an abelian Lie group with a bi-invariant metric. Since the Lie bracket is trivial we deduce from the computations in Subsection 4.2.2 that its curvature is zero. This concludes the proof of the proposition. D Proposition 4.2.23 If 5; (i=l,2) are two compact oriented surfaces without boundary then x(S1#S2)=X(5i) + x(52)-2. Thus upon iteration we get
x(5i#---#5 t ) = E x ( S i ) - 2 ( * - l ) i=l
for any compact oriented surfaces S i , . . . , 5 * . In terms of genera, the last equality can be rephrased as i=l
166
Riemannian
f
geometry
f
Figure 4.3: Generating a hot-dog-shaped
surface
In the proof of this proposition we will use special metrics on connected sums of surfaces which require a preliminary analytical discussion. Consider / : (—4,4) —> (0, oo) a smooth, even function such that (i) f{x) = l for \x\ < 2. (ii) f(x) = yjl-(x + 3) 2 for x € [-4, -3.5]. (iii) f(x) = yjl-{x3) 2 for x € [3.5,4]. (iv) / is non-decreasing on [-4,0]. One such function is graphed in Figure 4.3 Denote by Sf the surface inside R3 obtained by rotating the graph of / about the x-axis. Because of properties (i)-(iv) 5 / is a smooth surface diffeomorphic to S2. (One such diffeomorphism can be explicitly constructed projecting along radii starting at the origin). We denote by g the metric on Sf induced by the euclidian metric in R 3 . Since Sf is diffeomorphic to a sphere f Kgdvg =
2-KX(S2)=4ir.
Set
sf = sfn {±x > o} sf = sf n {±x > i} Since / is even we deduce
j
Kgdvg = '-[ K.gdvg 2Jsf
!
2%.
(4.2.27)
The Riemann curvature
167
Figure 4.4: Special metric on a connected sum On the other hand, on the neck C = {\x\ < 2} the metric g is locally euclidian g = dx2 + d62 so that over this region Kg = 0. Hence f Kgdvg = 0.
(4.2.28)
Proof of the Proposition 4.2.23 Let A C Si (i = 1,2) be a local coordinate neighborhood diffeomorphic with a disk in the plane. Pick a metric & on Si such that (Di,gi) is isometric with Sj!" and (1)2,52) is isometric to Sj. The connected sum 5 i # 5 2 is obtained by chopping off the regions Sj from D\ and SJ1 from Di and (isometrically) identifying the remaining cylinders 5 * n {|x| < 1} = C and call O the overlap created by gluing (see Figure 4.4). Denote the metric thus obtained on S\#S2 by g. We can now compute
+
hfs2K*dv°>-hLK°dv°
168
Riemannian (4
geometry
= 7 ) x(5 1 )+x(5 2 )-2.
The proposition is proved. • Corollary 4.2.24 (Gauss-Bonnet)Let E 9 denote the connected sum of p-tori. (By definition E 0 = S 2 . Then X(E 9 ) = 2 - 2 5 and g(Eg) = g. In particular a sphere is not diffeomorphic to a torus. Remark 4.2.25 It is a classical result that the only compact oriented surfaces are the connected sums of g-tori (see [52]) so that the genus of a compact oriented surface is a complete topological invariant.
Chapter 5 Elements of the calculus of variations This is a very exciting subject lieing at the frontier between mathematics and physics. The limited space we will devote to this subject will hardly do it justice and we will barely touch its physical significance. We recommend to anyone looking for an intellectual feast the Chapter 16 in vol.2 of [27] which in our opinion is the most eloquent argument for the raison d'etre of the calculus of variations.
5.1 §5.1.1
The least action principle l-dimensional Euler-Lagrange equations
From a very "dry" point of view the fundamental problem of the calculus of variations can be easily formulated as follows. Consider a smooth manifold M and denote L : t x TM —> R be a smooth function called the lagrangian. Fix two points po, p\ 6 M. For any piecewise smooth path 7 : [0,1] —> M connecting these points we define its action by 5(7) = 5 L ( 7 ) = / 1 L ( t , 7 ( t ) , 7 ( t ) ) d * . Jo
In the calculus of variations one is interested in those paths as above with minimal action. Example 5.1.1 Given U : K3 —> R a smooth function called the potential we can form the lagrangian L(q, q) : K3 x R 3 ^ TR 3 -» R as L = Q — U = kinetic energy — potential energy = -m|<j| 2 — U(q). (m is the mass). The action of a path (trajectory) 7 : [0,1] —> R3 is a quantity called the newtonian action. 169
170
Calculus of variations
Example 5.1.2 To any Riemann manifold (M, g) one can naturally associate two lagrangians L\, L2 : TM -» M defined by Li(q,v)=gq{v,v)1'2
(veTqM)
L2{q,v) =
-g(v,v).
We see that the action defined by L\ coincides with the length of a path. The action defined by L2 is called the energy of a path. Before we present the main result of this subsection we need to introduce a bit of notation. Tangent bundles are very peculiar manifolds. Any collection (ql,...,qn) of local coordinates on a smooth manifold M automatically induces local coordinates on TM. Any point in TM can be described by a pair (q, v) where q £ M, v £ TqM. Furthermore, v has a decomposition v = v>di
(* = £ ) .
We set q' = v' so that The collection (q1,... ,qn; q1,... ,qn) defines local coordinates on TM. These are said to be holonomic local coordinates on TM. This will be the only type of local coordinates we will ever use. Theorem 5.1.3 (The least action principle) Let L : RxTM —> K be a lagrangian and p0, pi € M two fixed points. If 7 : [0,1] —> M is a smooth path such that (1)7(1) = Pi, i = 0 , l .
(ii) S.L(7) < £x(7) for any smooth path 7 : [0,1] —> M joining p0 to pi. Then 7 satisfies the Euler-Lagrange equations d d
r
,
L(
.,
7 7) =
5*F *' '
dL
^-
More precisely, if (q';^) are holonomic local coordinates on TM such that "/(t) = {ql{t)) and 7 = {
JtW{t,qi'^)
=
dq~k{t'9i'4J)
fc = 1
'-'n=
dimM
-
Definition 5.1.4 A path 7 : [0,1] —>• M satisfying the Euler-Lagrange equations with respect to some lagrangian L is said to be an extremal of L.
171
The least action principle
To get a better feeling of these equations consider the special case discussed in Example 5.1.1 L = \m\q\2 - U(q). The Euler-Lagrange equations become mq = -VU(q).
(5.1.1)
These are precisely Newton's equation of the motion of a particle of mass m in the force field —VU generated by the potential U. In the proof of the least action principle we will use the notion of variation of a path. Definition 5.1.5 Let 7 : [0,1] —> M be a smooth path. A variation of 7 is a smooth map a = as(t) : (-e,e) x [0,1] ->• M such that «o(t) = 7(i). If moreover, as(i) = Pi Vs (i=0,l) then we say that a is a variation rel endpoints. Proof of Theorem 5.1.3
Let as be a variation of 7 rel endpoints. Then SL(CHO) < SL(as)
Vs
so that -H s =o SL(as)
= 0.
Assume for simplicity that the image of 7 is entirely contained in some open coordinate neighborhood U with coordinates (q1,..., qn). Then for very small \s\ we can write a s (t) = (9 I (s,i)) and
_ | = (,f ( s ,*)).
Following the tradition, we set da
9q'
d
das
dq3 d
5a is a vector field along 7 called infinitesimal variation (see Figure 5.1). In fact, the pair (5a;6a) € T{TM) is a vector field along t H-> {j{t),j{t)) £ TM. Note that 6a = jt5a and at endpoints 6a = 0. Exercise 5.1.1 Prove that if t >-> X(t) e T^t)M is a smooth vector field along 7 such that X[t) = 0 for t = 0,1 then there exists at lest one variation rel endpoints a such that 6a = X.
172
Calculus of variations
Figure 5.1: Deforming a path rel endpoints (Hint: Use the exponential map of some Riemann metric on M.) We compute (at s = 0)
Jo dql
Jo dqi
s
Integrating by parts in the second term we deduce
The last equality holds for a n y variation a. From Exercise 5.1.1 deduce it holds for any vector field Sa'di along 7. At this point we use the following classical result of analysis. " If f(i) is a continuous function on [0,1] such that f f{t)g(t)dt Jo
= 0 V 5 eC 0 °°(0,l)
then / is identically zero." Using this result in (5.1.2) we deduce the desired conclusion. • Remark 5.1.6 (a) In the proof of the least action principle we used a simplifying assumption i.e. the image of 7 lies in a coordinate neighborhood. This is true locally and for the above arguments to work it suffices to choose only a special type of variations, localized on small intervals of [0,1]. In terms of infinitesimal variations
173
The least action principle
this means we need to look only at vector fields along 7 supported in local coordinate neighborhoods. We leave the reader fill in the details. (b) The Euler-Lagrange equations were described using holonomic local coordinates. The minimizers of the action (if any) are objects independent of any choice of local coordinates so that the Euler-Lagrange equations have to be independent of such choices. We check this directly. If (xl) is another collection of local coordinates on M and (xl\£i) are the coordinates induced on TM then we have the transition rules rM .
so that d _dxi__d_ 'dq~i ~~ a ^ & r J
+
d V .k d dqkdqi9 dxi
_d__dx^_d__ dxj d dqi ~ dqi dxi ~~ dqi dxi Then
aL_a^aL dq* ~ dq{ dxi d (dL\
ay .kdL dqkdq'9
_ d (dxi
dxi ' dL\
Jt yWj ~ dl {dq* OH J d2xi .kdL dxi_d_ (dL\ ~~ dqkdqiq 'd±i + 'dq7dl \d&) ' We now see that the Euler-Lagrange equations in the ^-variables imply the EulerLagrange in the x-variable i.e. these equations are independent of coordinates. The aesthetically conscious reader may object to the way we chose to present the Euler-Lagrange equations. These are intrinsic equations we formulated in a coordinate dependent fashion. Is there any way of writing these equation so that the intrinsic nature is visible "on the nose" ? If the lagrangian L satisfies certain nondegeneracy conditions there are two ways of achieving this goal. One method is to consider a natural nonlinear connection V L on TM as in [59] . The Euler-Lagrange equations for an extremal y(t) can then be rewritten as a "geodesies equation" V^7The example below will illustrate this approach on a very special case when L is the lagrangian L^ defined in Example 5.1.2 in which the extremals are precisely the geodesies on a Riemann manifold. Another far reaching method of globalizing the formulation of the Euler-Lagrange equation is through the Legendre transform which again requires a nondegeneracy
174
Calculus of variations
condition on the lagrangian. Via the Legendre transform the Euler-Lagrange equations become a system of first order equations on the cotangent bundle T*M known as the Hamilton equations. These equations have the advantage that can be formulated on manifolds more general than the cotangent bundles namely on symplectic manifolds. These are manifolds carrying a closed 2-form whose restriction to each tangent space defines a symplectic duality (see §2.2.4.) Much like the geodesies equations on a Riemann manifold the Hamilton equations carry many informations about the structure of symplectic manifolds and are currently the focus of very intense research. For more details and examples we refer to the monographs [5] or [21]. E x a m p l e 5.1.7 Let [M,g) be a Riemann manifold. We will compute the EulerLagrange equations for the lagrangians L\, L2 in Example 5.1.2. L2(q,q) =
^gaiqWq3
so that dL2
W
., = 9jkq
dL2
W
=
1 dgij .t., qq
2W -
The Euler-Lagrange equations are
^+w^
= ^q3-
(5L3)
Since gkmgjm = 5™ w e get
^""(tN§0«'^°-
5
< -">
When we derivate with respect to t the equality gikQ' = gjkq3
we deduce km^gj^-i-i
9n
qq
_
1
km f%*
w -2
9n
_, dgjk\
Ai
w+wiqq
We substitute this equality in (5.1.4) and we get
*-+WiM?-|?)«=°-
<5i5»
The coefficient of qlq* in (5.1.5) is none other than the Christoffel symbol r™ so this equation is precisely the geodesic equation.
175
The least action principle Consider now the lagrangian L\{q,q) = {gijCfqi)1!2- Note that the action
[Pl L{q,q)dt •>po
is independent of the parameterization t >-» q(t) since it computes the length of the path. Thus, when we express the Euler-Lagrange equations for a minimizer 70 of this action we may as well assume it is parameterized by arclength i.e.
I-Vol = 1 . The Euler-Lagrange equations for L\ are d dt
gkjgj
\l9itfq?
_
% W
2\l9irfq>'
Along the extremal we have gutftf = 1 (arclength parameterization) so that the previous equations can be rewritten as
Jt^q)
= 2Wqq-
We recognize here the equation (5.1.3) which, as we have seen, is the geodesic equation in disguise. This fact almost explains why the geodesies are the shortest paths between two nearby points. §5.1.2
Noether's conservation principle
This subsection is intended to offer the reader a glimpse at a fascinating subject touching both physics and geometry. We first need to introduce a bit of traditional terminology commonly used by physicists. Consider a smooth manifold M. The tangent bundle TM is usually referred to as the space of states or the lagrangian phase space. A point in TM is said to be a state. A lagrangian L : R x TM —> K associates to each state several meaningful quantities. (a) The generalized momenta: pi = jpj. (b) The energy: H = p,g' — L. (c) The generalized force: F = f t . This terminology can be justified by looking at the lagrangian of a classical particle in a potential force field F = —VU
L = lm\q\2-U(q). The momenta associated to this lagrangian are the usual kinetic momenta of the Newtonian mechanics Pi = mql
176
Calculus of variations
while H is simply the total energy H=1-m\q\2
+ U(q).
It will be convenient to think of an extremal for an arbitrary lagrangian L(t, q, q) as describing the motion of a particle under the influence of the generalized force. Proposition 5.1.8 (Conservation of energy) Let -y(t) be an extremal of a time independent lagrangian L = L(q, q). Then the energy is conserved along 7 i.e.
Proof
By direct computation we get ! * ( 7 , 7 ) = ! ( * * * - L)
dt \dqi)q = { l ( i ) - | }
dqit} ?
= °
dqiq
dqiQ
( ^ Euler-Lagrange).
•
At the beginning of this century (1918) Emmy Noether discovered that many of the conservation laws of the classical mechanics had a geometric origin: they were, most of them, reflecting a symmetry of the lagrangian!!! This became a driving principle in the search for conservation laws and in fact, conservation became synonymous with symmetry. It eased the leap from classical to quantum mechanics and one can say it is a very important building block of quantum physics in general. In the few instances of conservation laws where the symmetry was not apparent the conservation was always "blamed" on a "hidden symmetry". What is this Noether principle? To answer this question we need to make some simple observations. Let X be a vector field on a smooth manifold M defining a global flow $ ' . This flow induces a flow \l>' on the tangent bundle TM defined by *'(z,i>) = ($'(*).*'.(«))• One can think of \P' as defining an action of the group ffi on TM and thus defining a symmetry of M. Example 5.1.9 Let M be the unit round sphere S2 C ffi3. The rotations about the 2-axis define a 1-parameter group of isometries of S2 generated by Jg (9 is the longitude on S2).
177
The least action principle
Definition 5.1.10 Let L be a lagrangian on TM and X a vector field on M. The lagrangian L is said to be X-invariant if L o $ ' = L Vt. Denote by X G Vect (TM) the infinitesimal generator of * ' and by Lx the Lie derivative on TM along A". We see that L is X-invariant iff CXL = 0. We describe this derivative using the local coordinates (ql,(f).
Set (ql(t), q>(t)) =
jt\t=0q-(t)=XkSl To compute Jj | (= o ${t)-£j
we use the definition of the Lie derivative on M d
q
-7
d
T
-Jt ^ 'X
dqk
=
d
t-i
Lx{q
\
]
W
^ dqk J dqqj ~
-4{
dXi d dq1 dqi
since dqj/dqi = 0 on TM. Hence dqi
dXi d dqk dip'
Corollary 5.1.11 L is X-invariant iff vi3L
x
.kdXi
0L
+q
w ww~
Theorem 5.1.12 (E. Noether) If the lagrangian L is X-invariant then the quantity
* = *£-*» is conserved along the extremals of L. Proof
Consider an extremal 7 = 7(9'(i)) of L. We compute d
o /
M
d
\ v i , u^dL\
dX> .k0L
k = (Euler - Lagrange) -7ri-q -7rrr + X{—~=0 oqk dq1 oq*
vld
(dL\
by Corollarv 5.1.11.
•
The classical conservation of momentum law is a special consequence of Noether's theorem.
178
Calculus of variations z
r
J
—
/
—/
—^22
y
*
Figure 5.2: A surface of revolution Corollary 5.1.13 Consider a lagrangian L = L(t,q,q) component of the force is zero) then ^ of the momentum is conserved).
on K n . If f^ = 0 (the z-th
= 0 along any extremal (the i-th component
To prove this it suffices to take X — A in Noether's conservation law.. The conservation of momentum has an interesting application in the study of geodesies. Example 5.1.14 (Geodesies on surfaces of revolution) Consider a surface of revolution S in E 3 obtained by rotating about the 2-axis the curve y = f(z) situated in the yz plane. If we use cylindrical coordinates (r, 8, z) we can describe S as r = f(z). In these coordinates the euclidian metric in M3 has the form ds2 = dr2 + dz2 + r2d92. We can choose (z, 9) as local coordinates on S then the induced metric has the form 9s = {1 + (f'(z))2}dz2
+ f2(z)d92
= A(z)dz2 + r2d92. (r =
f(z))
The lagrangian defining the geodesies on 5 is L=
2 2 2 \(A z + r 9 )
We see that L is independent of 9: ^ = 0 so that the generalized momentum d L
2n
—- = r 9 39 is conserved along the geodesies.
179
The variational theory of geodesies
This fact can be given a nice geometric interpretation. Consider a geodesic 7 = (z(t),8(t)) and compute the angle
<7,a/w>
=
\i\-\d/d6\
i!j r| 7 |
i.e. rcos> = r 2 # | 7 | _ 1 . The conservation of energy implies that |7| 2 = 2L = H is constant along the geodesies. We deduce the following classical result. Theorem 5.1.15 (Clairaut)On a a surface of revolution the quantity r cos <> / is constant along any geodesic. cf> € (—7r, n) is the oriented angle the geodesic makes with the parallels z = const. Exercise 5.1.2 Describe the geodesies on the round sphere S2 and on the cylinder {x2 + y2 = 1} C t t 3 .
5.2
The variational theory of geodesies
We have seen that the paths of minimal length between two points on a Riemann manifold are necessarily geodesies. Conversely, given a geodesic joining two points qo, qi it may happen it is not a minimal path. This should be compared with the situation in calculus when a critical point of a function / may not be a minimum or a maximum. To decide this issue one has to look at the second derivative. This is precisely what we intend to do in the case of geodesies. This situation is a bit more complicated since the action functional
S = \jWdt is not defined on a finite dimensional manifold. It is a function defined on the "space of all paths" joining the two given points. With some extra effort this space can be organized as an infinite dimensional manifold. We will not attempt to formalize these prescriptions but rather follow the ad-hoc, intuitive approach of [55]. §5.2.1
Variational formulae
Let M be a connected Riemann manifold and consider p,q G M. Denote by fiM = fiM(M) the space of all continuous, piecewise smooth paths 7 : [0,1] —> M connecting p to q. An infinitesimal variation of a path 7 e fiM is a continuous, piecewise smooth vector field V along 7 such that ^(0) = 0 and V(l) = 0. The space of infinitesimal variations of 7 is an infinite dimensional linear space denoted by T =T n
180
Calculus of variations
Definition 5.2.1 Let 7 G QP:q. A variation of 7 is a continuous map a = as(t) : ( - e , e ) x [0,1] ->• M such that (i) Vs € (-e,e), as e Qp>,. (ii) There exists a partition 0 = ta < t\ • • • < tk-i < tk = 1 of [0,1] such that the restriction of a to each (—£,£•) x (£;_!,£;) is a smooth map. Every variation a of 7 defines an infinitesimal r
variation
def das
Exercise 5.2.1 Given 7 e T T construct a variation a such that 5a = V. Consider now the energy functional
E-.nM->R
1 r1.
E(1) = -Jg I^WI2(ii-
Fix 7 6 QPig and let a be a variation of 7. The velocity j(t) has a finite number of discontinuities so that the quantity A t 7 = lira (j(t + h)-
j(t - h))
ft-»0+
is nonzero only for finitely many t's. Theorem 5.2.2 (The first variation formula) E.(Sa)
=' 4- l»=o E{a.) = - £<(<*")(*). ^ 7 ) - l\&a, V±i)dt. (5.2.1) as ( Vo <" V denotes the Levi-Civita connection. (Note that the right-hand-side depends on a only through 5a so it is really a linear function on T7.) Proof
Set as = ^-.
We derivate under the integral sign using the equality — |d s | 2 = 2 ( V B Ads,ds) us >
and we get
d r1 — \s=oE(as)= / (VjLas,as)\s=0dt. as Jo a" Since the vector fields Jj and J^ commute we have V j _ ^ = V a . f j - Let 0 = to < t2 < • • • < tk = 1 be a partition of [0,1] as in Definition 5.2.1. Since as = 7 for s = 0 we conclude
^(H = E£ i (Vafo, 7).
181
The variational theory of geodesies We use the equality — (<5a,7) = (Va_5a,j)
+ (5a,V
±j)
to integrate by parts and we obtain
E.(5a) = J2(6a,i) |*;_x " £ f {6<*,V±i)dt. i=iy«i-i
i=i
*'
This is precisely equality (5.2.1). D Definition 5.2.3 A path 7 € Qp,q is critical if
£,(10 =0 w e r 7 . Corollary 5.2.4 A path 7 G fiP), is critical if and only if it is a geodesic. Exercise 5.2.2 Prove the above corollary. Note that a priori a critical path may have a discontinuous first derivative. The above corollary shows that this is not the case; the criticality also implies smoothness. This is a manifestation of a more general analytical phenomenon called elliptic regularity. We will have more to say about it in Chapter 9. The map E, : T 7 —> E, 5a H-» E„(5a) is called the first derivative of E at 7 e f2p>?. We want to define a second derivative of E in order to address the issue raised at the beginning of this section. We will imitate the finite dimensional case which we now briefly analyze. Let / : X —> R be a smooth function on the finite dimensional smooth manifold X. If x0 is a critical point of / i.e. df(xo) = 0 then we can define the hessian at x0 / „ : TX0X x TXoX -> 1 as follows. Given Vi, V2 6 TX0X consider a smooth map (si, S2) •-> a(si, s 2 ) € X such that dc*. a(0,0) = zo and — (0,0) = VI, i = 1,2. (5.2.2) Now set h,{VuV2)-
dsids2
-|(o,o).
Note that since x 0 is a critical point of / the hessian /**(Vi, V2) is independent of the function a satisfying (5.2.2). We now return to our energy functional E : i1p^q —¥ M. Let 7 G Qpq be a critical path. Consider a 2-parameter variation of 7 aSUS2{t)
: U x [0,1] ->• M.
U is the tiny square (—e, e) x (—e,e) C R2 and a0,o = 7- a is continuous and has second derivatives everywhere except maybe on finitely many "coordinate planes Sj = const or t = const. Set ^ a = ~ |(0,o), 2 = 1,2. Note that 5ia 6 T 7 .
182
Calculus of variations
Exercise 5.2.3 Given Vi,V2 € T 7 construct a 2-parameter variation a such that Vi = Sta. We can now define the hessian of E at 7 by E„(6ia,8,a)
=
d g
^
|(o.o) -
Theorem 5.2.5 (The second variation formula) £„(o"ia, 52a) = - ^ ( f c a , A(<5lQ> - ^ V 2 a , V 2 J^i<* - #(7,5 i a )j)dt
(5.2.3)
where i? denotes the Riemann curvature. In particular, E„ is a bilinear functional onT7. Proof
According to the first variation formula we have 9E
v-^.,.
rllr
. da.
_
9a. ,
Hence _d2E ds fl - / {V_a_5 2a,Va_j)dtds st 7o
i
r1 da / {52a,V_a_Va_—)dt. Bt Bs
Jo
i
ot
(5.2.4)
Since 7 is a geodesic A t 7 = 0 and V.a.7 = 0. Using the commutativity of Jj with ^ we deduce
Finally, the definition of the curvature implies V A V I = V 3 V_a_ + R(Sxa, 7).
aj!
at
at dax
Putting all the above together we deduce immediately the equality (5.2.3). • Corollary 5.2.6 Ett(Vi, V2) = E„{V2, Vi) W i , V2 e TT
183
The variational theory of geodesies §5.2.2
Jacobi fields
In this subsection we will put to work the elements of calculus of variations presented so far. Let (M,g) be a Riemann manifold and p,q € M. Definition 5.2.7 Let 7 € fip,9 be a geodesic. A geodesic variation of 7 is a smooth map a3(t) : ( - e , e ) x [0,1] -> M such that a0 = 7 and 11-> a,(t) is a geodesic for all s. We set as usual 5a = | ^ | s = 0 . Proposition 5.2.8 Let 7 € fiPi, be a geodesic and (as) a geodesic variation of 7. Then the infinitesimal variation 5a satisfies the Jacobi equation V2t5a = R(i,6a)j
(V t = V J L ) .
Proof
^.-v.(v.|)-v.(v.|) Definition 5.2.9 A smooth vector field J along a geodesic 7 is called a Jacobi field if it satisfies the Jacobi equation
Vp =
R(j,J)j.
Exercise 5.2.4 Show that if J is a Jacobi field along a geodesic 7 then there exists a geodesic variation as of 7 such that J = 5a. Exercise 5.2.5 Let 7 € fiPi, and J a vector field along 7. (a) Prove that J is a Jacobi field if and only if
E„(J,V} = 0 W 6 T
r
(b) Show that J € T 7 (i.e. J is a vector field along 7 vanishing at endpoints) is a Jacobi field if any only if S*,(J, W) = 0 for all vector fields W along 7. (It is important to emphasize the point that not all vector fields W along 7 belong to T 7 .)
Exercise 5.2.6 Let 7 6 fiPi9 be a geodesic. Define 3 P to be the space of Jacobi fields V along 7 such that V(p) = 0. Show that d i m 3 P = d i m M and moreover, the evaluation map
ev, :ZP^TPM is a linear isomorphism.
V^VtV{p)
Calculus of variations
Figure 5.3: The poles are conjugate along meridians. Definition 5.2.10 Let j(t) be a geodesic. Two points 7(^1) and 7^2) on 7 are said to be conjugate along 7 if there exists a nontrivial Jacobi field J along 7 such that
Jfc) = 0, i = l,2. Example 5.2.11 Consider 7 : [0, 2n] -> S2 a meridian on the round sphere connection the poles. One can verify easily (using Clairaut's theorem) that 7 is a geodesic. The counterclockwise rotation by an angle 9 about the 2-axis will produce a new meridian, hence a new geodesic jg. Thus (79) is a geodesic variation of 7 with fixed endpoints. £7 is a Jacobi field vanishing at the poles. We conclude that the poles are conjugate along any meridian (see Figure 5.3). Definition 5.2.12 Let 7 e Qpq be a geodesic. 7 is said to be nondegenerate if q is not conjugated to p along 7. The following result (partially) explains the geometric significance of conjugate points. Theorem 5.2.13 Let 7 G Qp,g be a nondegenerate, minimal geodesic. Then p is conjugate with no point on 7 other than itself. In particular, a geodesic segment containing conjugate points cannot be minimal ! Proof We argue by contradiction. Let pi = 7(^1) be a point on 7 conjugate with p. Denote by Jt a Jacobi field along 71[0 tt] such that J 0 = 0 and Jtl = 0. Define V e T-, by f Jt , t e [o,t!] \ 0 , t>h We will prove that Vt is a Jacobi field along 7 which contradicts the nondegeneracy of 7. In the sequel I denotes the length.
185
The variational theory of geodesies Step 1 E„(U,U)>0
V£/GT 7 .
(5.2.5)
Indeed, let as denote a variation of 7 such that 5a = U. One computes easily that j^E{a.*)
=
2E„{U,V).
Since 7 is minimal for any small s we have l(as2) > l{a0) so that
E(as>) > \ (jf1 |ds,|dt)2 = i[(a s2 ) 2 > |l(«o)2 = |l(T)2 =
E(aa).
Hence — | J=0 £ ( « . » ) > 0. This proves (5.2.5). Step 2 Ett(V, V) = 0. This follows immediately from the second variation formula and the fact that the nontrivial portion of V is a Jacobi field. Step 3 E„{U,V) = 0 V l / € T 7 . From (5.2.5) and Step 2 we deduce 0 = E„{V, V) < £U(V + TU,V
+ TU)
= fu(r)
Vr.
Thus, T = 0 is a global minimum of /c/(r) so that fu(0) = 0. Step 3 follows from the above equality using the bilinearity and the symmetry of E„. The final conclusion (that V is a Jacobi field) follows from Exercise 5.2.5. • Exercise 5.2.7 Let 7 : M -> M be a geodesic. Prove that the set {t € R ; 7(£) is conjugate to 7(0)} is discrete. Definition 5.2.14 Let 7 G f2p>, be a geodesic. We define its index, denoted by ind (7), as the cardinality of the set C 7 = {t e (0,1) ; is conjugate to 7(0)} which by Exercise 5.2.7 is finite.
186
Calculus of variations
Theorem 5.2.13 can be reformulated as follows: the index of a nondegenerate minimal geodesic is zero. The index of a geodesic obviously depends on the curvature of the manifold. Often, this dependence is very powerful. Theorem 5.2.15 Let M be a Riemann manifold with non-positive sectional curvature i.e. {R(X, Y)Y, X) < 0 \/X, Y € TXM Vz e M. (5.2.6) Then for any p,q e M and any geodesic 7 G n P i „ ind(7) = 0. Proof It suffices to show that for any geodesic 7 : [0,1] —• M the point 7(1) is not conjugated to 7(0). Let Jt be a Jacobi field along 7 vanishing at the endpoints. Thus
V2tJ =
R(j,J)j
so that
[\v2J,J)dt
Jo
= f\R{j,J)j,J)dt Jo
= - f1 {R( J, 7)7, J)dt. Jo
We integrate by parts the left-hand-side of the above equality and we deduce
(v,J,J)\l - f \vtj\2dt = Jo
[\R{J,J)J,J)dt. Jo
Since J{T) = 0 for r = 0,1 we deduce using (5.2.6)
f \VtJ?dt
Jo
This implies V ( J = 0 which coupled with the condition J(0) = 0 implies J = 0. The proof is complete. • The notion of conjugacy is intimately related to the behavior of the exponential map. Definition 5.2.16 Let / : X —> Y be a smooth map between the smooth manifolds x and Y. (a) A point i G X i s said to be critical if r a n k ( D / x : TXX -»• Tf{x)Y)
< minfdimT.X,
dimTf(x)Y}.
(b) A point y G Y is called a critical value for / if / _1 (2/) contains at least one critical point of / . (c) y € Y is a regular value if it is not a critical value.
187
The variational theory of geodesies
Theorem 5.2.17 Let (M, g) be a connected, complete, Riemann manifold and g0 € M. A point q G M is conjugated to go along some geodesic if and only if it is a critical value for the exponential map exp go : TqoM -> M. Proof Let q = exp w t ( « € TqoM). Assume first that q is a critical value for exp9o and v is a critical point. Then Dvexpqo(X) = 0 for some X € Tv(TqoM). Let v(s) be a path in TqoM such that v(0) = v and v(0) = X. The map (s, t) *-¥ expqo(tv(s)) is a geodesic variation of the radial geodesic -yv : 1i-+ exp 9o (tv). Hence the vector field W = — | s = 0 exp go (to(s)) is a Jacobi field along 7„. Obviously W(0) = 0 and moreover d_
W{1) = — | s = 0 expw(t>(a)) = Dv expJX)
= 0.
ds
On the other hand this is a nontrivial field since VtW
d_ = V 5 |,=o ^ exp go (to(s)) = V,w(a) | s = 0 / 0. dt
This proves
Tv(TqoM)
(5.2.7)
The existence of such a Jacobi field follows from Exercise 5.2.6. As in that exercise denote by 3 , 0 the space of Jacobi fields J along j v such that J(qo) = 0. The map T,(rwM)->3w
X^tJx
is a linear isomorphism. Thus, a Jacobi field along 7„ vanishing at both q0 and q must have the form Jx where X € Tv(TqoM) satisfies Dvexpqo(X) = 0. Since v is not a critical point this means X — 0 so that Jx = 0. • Corollary 5.2.18 On a complete Riemann manifold M with non-positive sectional curvature the exponential map exp g has no critical values for any q € M. We will see in the next chapter that this corollary has a lot to say about the topology of M. Consider now the following experiment. Stretch the round two-dimensional sphere of radius 1 until it becomes "very long". A possible shape one can obtain may look like in Figure 5.4. The long tube is very similar to a piece of cylinder so that the total
188
Calculus of variations
Figure 5.4: Lengthening a sphere. (= scalar) curvature is very close to zero, in other words is very small. The lesson to learn from this intuitive experiment is that the price we have to pay for lengthening the sphere is decreasing the curvature. Equivalently, a highly curved surface cannot have a large diameter. Our next result offers a more quantitative description of this phenomenon. Theorem 5.2.19 (Myers)Let M be an n-dimensional complete Riemann manifold. If for all X € Vect (M) Ric(X,X)>^-=i!|X|2 then every geodesic of length > irr has conjugate points and thus is not minimal. Hence diam (M) = sup{dist (p, q) ; p,q € M} < irr and in particular the Hopf-Rinow theorem implies that M must be compact. Proof Fix a minimal geodesic 7 : [0, £} —> M of length t and let e,(i) be an orthonormal basis of vector fields along 7 such that en(t) = j(t). Set q0 = 7(0) and qi = "/(£)• Since 7 is minimal we deduce £«(V,V)>0 Set W{ = sm(irt/£)ei.
WeT7.
Then
E„(Wi7 W{) = - j\Wi, Jo = j
sin2(irt/e) [fit2
VtW, +
R{Wi,i)j)dt
- (R(ei,j)-y,ei))
dt.
189
The variational theory of geodesies We sum over i = 1 , . . . , n — 1 and we obtain J2 E„{Wh Wi) = / sin2 Trt/e {[n - 1)TT 2 /£ 2 - Ric (7,7)) dt > 0. i=l
''0
If I > 7ir then (n - 1)TT 2 /£ 2 - Ric (7,7)
< 0
so that
£fi„(Wi)Wi)<0. Hence for at least for some W{ we have Ett(Wi,Wi) minimality of 7. The proof is complete. •
< 0. This contradicts the
Corollary 5.2.20 A semisimple Lie group G with positive definite Killing pairing is compact. Proof The Killing form defines in this case a bi-invariant Riemann metric on G. Its geodesies through the origin 1 € G are the 1-parameter subgroups exp(tX) which are defined for all t € R. Hence, by Hopf-Rinow theorem G has to be complete. On the other hand we have computed the Ricci curvature of the Killing metric and we found Ric (X, Y) = ^K{X, Y) VX, Y e £ G . The corollary now follows from Myers' theorem. • Exercise 5.2.8 Let M be a Riemann manifold and q £ M. For the unitary vectors X, Y £ TqM consider the family of geodesies ~/s{t)=expqt{X
+ sY).
Denote by Wt = Sjs the associated Jacobi field along 70(t). Form /(£) = \Wt\2. Prove the following. (a) Wt = Dtx exp g (y) = Frechet derivative of D H expq(v) (b) f(t) = t 2 - \(R(Y,X)X, Y)qt* + 0(f). (c) Denote by x 1 a collection of normal coordinates at q. Prove that 9u(x) =
6u-±Rkijixix?+0{3l).
detpy(x) = 1 - ^ x V
+0(3).
(e) Let D r («) = {x e TqM ; |x| < r } .
190
Calculus of variations
Prove that if the Ricci curvature is negative definite at q then volo (D r (q)) < vols (exp„(D r (g)) for all r sufficiently small, volo denotes the euclidian volume in TgM while vol9 denotes the volume on the Riemann manifold M. Remark 5.2.21 The interdependence "curvature-topology" on a Riemann manifold has deep reaching ramifications which stimulate the curiosity of many researchers. We refer to [23] or [55] and the extensive references therein for a presentation of some of the most attractive results in this direction.
Chapter 6 The fundamental group and covering spaces In the previous chapters we almost exclusively studied local properties of manifolds. This study is interesting only if some additional structure is present since otherwise all manifolds are locally alike. We noticed an interesting phenomenon: the global "shape" (topology) of a manifold restricts the types of structures that can exist on a manifold. For example, the Gauss-Bonnet theorem implies that on a connected sum of two tori there cannot exist metrics with curvature everywhere positive because the integral of the curvature is a negative universal constant. We used Gauss-Bonnet theorem in the opposite direction and we deduced-the intuitively obvious fact- that a sphere is not diffeomorphic to a torus because they have distinct genera. The Gauss-Bonnet theorem involves a heavy analytical machinery which obscures the intuition. Notice that S2 has a remarkable property which distinguishes it from T 2 : on the sphere any closed curve can be shrunk to a point while on the torus there exist at least two "independent" unshrinkable curves (see Figure 6.1). In particular this means the sphere is not diffeomorphic to a torus. This chapter will set the above observations on a rigorous foundation.
Figure 6.1: Looking for unshrinkable loops.
191
192
6.1
The Fundamental group and covering spaces
The fundamental group
§6.1.1
Basic notions
In the sequel all topological spaces will be locally path connected spaces. Definition 6.1.1 (a)Let X and Y be two topological spaces and / o , / i : X —¥ Y continuous maps, /o and / i are said to be homotopic if there exists a continuous map F:[0,l]xX^Y (t,x)^Ft{x) such that Fi = fa for i = 0,1. We write this as f0 ^h / i (b) Two topological spaces X, Y are said to be homotopically equivalent if there exist maps / : X —> Y and g : Y —> X such that / o j ~ f t 1Y and g o / ~ h lx. We write this X ~^ Y. (c) A topological space is said to be contractible if it is homotopically equivalent to a point. Example 6.1.2 The unit disk in the plane is contractible. The annulus {1 < \z\ < 2} is homotopically equivalent with the unit circle. Definition 6.1.3 (a) Let X be a topological space and x0 £ X. A loop based at x0 is a continuous map 7 : [0,1] ->• X such that 7(0) = 7(1) = x0. The space of loops in X based at x0 is denoted by Q(X,x0). (b) Two loops 70,71 : I -¥ X based at x0 are said to be homotopic rel x0 if there exists a continuous map r:[0,l]x/->A-
(i,s)^r((s)
such that ri(s)=7i(a)
t = 0,l
and
{s^rt(s))eQ{x,x0) Vie [0,1]. We write this as 70 —xo 7iNote that a loop is more than a closed curve; it is a closed curve + a description of a motion of a point around the closed curve. Example 6.1.4 The two loops 71,72 : / -> C, jk(t) different though they have the same image.
= exp(2knt),
k = 1,2 are
The fundamental
193
group
Definition 6.1.5 (a) Let 71,72 be two loops based at x0 G X. The product of 71 and 72 is the loop 7l *7 2 (s)
7i *l2(s)
={
^2s) > °^sl/2
|72(2s_i)
, 1/2 < s < 1 •
The inverse of a based loop 7 is the based loop 7 " defined by 7-(s)
=7(l-s).
(c) The identity loop is the constant loop e l 0 (s) = xa. Intuitively, the product of two loops 71 and 72 is the loop obtained by first following 71 (twice faster) and then 72 (twice faster). The following result is left to the reader as an exercise. Lemma 6.1.6 Let a 0 c^l0 OJI, Po—XoPi based loops.. Then (a) a0*p0~xoa1 * fa.
an
d 7o—x07i be three pairs of homotopic
(b) a0 * QQ
(c) a0 * cxo~xoaa. (d) (a0 * Po) * 7o-x 0 Q o * (Po * 7o)Hence the product operation descends to an operation "•" on tt(X, Xo)/—x0~ t n e set of homotopy classes of based loops. Moreover the induced operation is associative, it has a unit and each element has an inverse. Hence (Q,(X,xo)/^xo, •) is a group. Definition 6.1.7 The group (£1{X, Zo)/—xa, •) is called the f u n d a m e n t a l g r o u p (or the Poincare group) of the topological space X and is denoted by TTI(X,X0). The image of a based loop 7 in ^ ( X , XQ) is denoted by [7]. The elements of ni(X, x0) are the "unshrinkable loops" discussed at the beginning of this chapter. The fundamental group ni(X,x0) "sees" only the connected component of X which contains a;0- To get more information about X one should study all the groups {ni(X,x)}xeXProposition 6.1.8 Let X and Y be two topological spaces x 0 € X and yo £ Y. Then any continuous map / : X —>• Y such that f(x0) = y0 induces a morphism of groups / , ITTip^Zo) ->7Ti(Y,2/ 0 )
satisfying the following functoriality properties. ( a ) ( l x ) , = l^(x,io)-
194
The Fundamental group and covering spaces
Figure 6.2: Connecting base points. (b)If
(X,x0)A(Y,y0)^(Z,z0) are continuous maps (such that f{x0) = y0 and g{yo) = ZQ) then (g o /)„ = gt o /„. (c) Let /o, / i : (X, x0) —> (Y, y0) be two base-point-preserving continuous maps. Assume f0 is nomotopic to f\ rel xo i.e. there exists a continuous map F : I x X —¥ Y, (t,x) i-> Ft(x) such that Fi(x) = fi(x) for i = 0,1 and Ft(x0) = yo- Then (/o). = ( / i ) . -
Proof
Let 7 G 0.(X, x0) Then 7(7) G fl(Y, j/ 0 ) and one can check immediately that 7^xoV => / ( 7 ) - W /(V)-
Hence the correspondence
n{X,x0)By^f{y)eQ{Y,yo) descends to a map / : ni(X,x0) —• ^ ( Y , j/ 0 )- This is clearly a group morphism. (a) and (b) are now obvious. We prove (c). Let / o , / i : {X,x0) -¥ {Y,yQ) be two continuous maps and Ft a homotopy rel x0 connecting them. For any 7 G fi(X, x0) A = /o(7)-M/i(7)=AThe above homotopy is realized by Bt = Ft(j).
•
A priori, the fundamental group of a topological space X may change as the base point varies and it almost certainly does if X has several connected components. However, if X is connected (and thus path connected since it is locally so) all fundamental groups ITI{X,x), x G X are isomorphic. Proposition 6.1.9 Let X be a connected topological space. Any continuous path a : [0,1] —> X joining x 0 to X\ induces an isomorphism a , :7ri(X,xo)
-^-K\(X,xx)
defined by a»([7]) = \pT * 7 * a] (see Figure 6.2).
The fundamental
195
group
Exercise 6.1.1 Prove Proposition 6.1.9. Thus, the fundamental group of a connected space X is independent of the base point modulo some isomorphism. We will write ni(X, pt) to underscore this weak dependence on the base point. Corollary 6.1.10 Two homotopically equivalent connected spaces have isomorphic fundamental groups. Example 6.1.11 (a) 7Ti(M",pi) ~ ^i{pt,pt) (b) 7Ti(annulus) ~ x^S1).
= {1}.
Definition 6.1.12 A connected space X such that iri(X,pt) simply connected.
= {1} is said to be
Exercise 6.1.2 Prove that the spheres of dimension > 2 are simply connected. Exercise 6.1.3 Let G be a connected Lie group. Define a new operation "* " on
n(G,l)by (o*/?)(a) = a(s)-/J(s) where • denotes the group multiplication. (a) Prove that a* ft ~i a * f3. (b) Prove that 7Ti(G, 1) is abelian. Exercise 6.1.4 Let E —> X be a rank r complex vector bundle over the smooth manifold X and let V be a flat connection on E i.e. F ( V ) = 0. Pick x0 € X and identify the fiber Exo with C . For any continuous, piecewise smooth 7 e il(X,xo) denote by T 7 = T 7 (V) the parallel transport along 7 so that T 7 £ GL (r, C). (a) Prove that a~Xo/3 =>• Ta =Tp. (b) Tpm =TaoTp. Thus, any fiat connection induces a group morphism T:7r1(X,a;o)^GL(7-,C)
J^T~\
(This morphism (representation) is called the monodromy of the connection). Example 6.1.13 We want to compute the fundamental group of the complex projective space CP". More precisely, we want to show it is simply connected. We will establish this by induction. For n = 1, CP 1 = S2 and by Exercise 6.1.2 the sphere S 2 is simply connected. We next assume CPfc is simply connected for k < n and prove the same is true for n. Notice first that the natural embedding C fc+1 "—> C" + 1 induces an embedding CP* ^-» CP". More precisely, in terms of homogeneous coordinates this embedding is given by [z0 : ... : zk] .-> [zo : ... : zk : 0 : . . . : 0] € C P n .
196
The Fundamental group and covering spaces
Choose as base point p t = [1 : 0 : . . . : 0] e OP" and let 7 € fi(OPn,pt). We may assume 7 avoids the point P = [0 : . . . : 0 : 1] since we can homotop it out of any neighborhood of P. We now use a classical construction of projective geometry. We project 7 from P to the "hyperplane" H = Q P " - 1 ^ OP". More precisely, if C = [-zo : • • • : zn] e OP" we denote by n(Q the intersection of the line P £ vvith the hyperplane %. In homogeneous coordinates "•(0 = [zo(l - zn) : ...: z„_i(l - zn) : 0] (= [z0 : . . . : z n _i : 0] when zn ^ 1). Clearly 7r is continuous. For t € [0,1] define ^(C) = [2o(l - tzn) : ... : z„_i(l - tzn) : (1 - t)zn]. Geometrically, 7rt flows the point C, along the line P(, until it reaches the hyperplane %. Note that irt(() = C Vt and V£ e W. Clearly 7r( is a homotopy rel p t connecting 7 = 71-0(7) to a loop 71 in % = O P " - 1 based at p t . Our induction hypothesis shows that 7x can be shrunk to p t inside % . This proves QP" is simply connected. §6.1.2
Of categories and functors
The considerations in the previous subsection can be very elegantly presented using the language of categories and functors. This brief subsection is a minimal introduction to this language. The interested reader can learn more about it from the monograph [51]. A c a t e g o r y is a triple C=(Ob, Horn, o) where (i) O b is a collection of elements called the objects of the category, (ii) Horn is a family of sets Horn (X, Y), one for each pair of objects X and Y. The elements of Horn (X, Y) are called the morphisms (or arrows) from X to Y. (iii) o is a collection of maps o : Horn (X, Y) x Horn (Y, Z) -» Horn (X, Z)
(f,9)^9°fwhich satisfies the following conditions. (Cl) For any object X there exists a unique element lx 6 Horn (X, X) such that f°lx
=f
9°lx=9
V/eHom(X,y)V5GHom(Z,X).
(C2) V/ € Horn {X, y ) j g Horn (Y, Z) h £ Horn (Z, W) ho(gof)
=
(hog)of.
The fundamental
group
197
R e m a r k 6.1.14 We note in passing that we were deliberately sloppy when we used the term "collection" when we introduced Ob. It may happen O b is not a set and one has to deal with thorny foundational issues which are beyond the scope of this book E x a m p l e 6.1.15 Top is the category of topological spaces. The objects are topological spaces and the morphisms are the continuous maps. (Top, *) is the category of marked topological spaces, the objects are pairs (X, *) where X is a topological space and * is a distinguished point of X. The morphisms
(A»4(y,*) are the continuous maps / : X —>Y such that /(*) = Jfr. pVect is the category of vector spaces over the field F. The morphisms are the F-linear maps. G r is the category of groups, while A b denotes the category of abelian groups. The morphisms are the obvious ones. iiMod denotes the category of left .R-modules, where R is some ring. Definition 6.1.16 Let €.\ and
con-
T : O b ( d ) x Horn (Ci) - • O b (C 2 ) x Horn (C 2 )
(X,f)»{F{X),F{f)) such that if X -4 Y then T(X) (i) T(lx) = 1HX)
^W F{Y) (resp.^(X) ^W ?{Y))
and
(ii) Ha) o HI) = Hg ° /) (resP. ?{f) ° r{9) = T{9 O /)). E x a m p l e 6.1.17 Let V be a real vector space. Then right tensoring with V defines a covariant functor ®V : ]g> Vect —> jgVect defined by U^U^V
L
%v U2 ® V).
{U1AU2)^{U1®V
On the other hand, taking the dual * : jjVect —¥ RVect V
H->
V
(U A V)
H>
(V* %• U")
defines a contravariant functor. The fundamental group construction of the previous is a covariant functor 7T! : (Top, *) ->• Gr. In Chapter 7 we will introduce other functors very important in geometry.
198
6.2 §6.2.1
The Fundamental group and covering spaces
Covering spaces Definitions and examples
As in the previous section we will assume that all topological spaces are locally path connected. Definition 6.2.1 (a) A continuous map ir : X —> Y is said to be a covering map if for any y £Y there exists an open neighborhood U such that 7r _1 ([/) is a disjoint union of open sets V, C X each of which is mapped homeomorphically onto U by n. Such an U is said to be an evenly covered neighborhood. The sets V, are called the sheets over U. (b) Let Y be a topological space. A covering space of Y is a topological space X together with a covering map 7r : X —> Y. If 7r : X —> Y is a covering map then for any y £ Y the set 7r_1(y) is called the fiber over y. Example 6.2.2 Let D be a discrete set. Then for any topological space X the product X x D is a covering of X with covering projection ir(x, d) = x. This type of covering space is said to be trivial. Exercise 6.2.1 Show that a fibration with standard fiber a discrete space is a covering. Example 6.2.3 The exponential map exp : t -> S 1 , t H exp(27ri<) is a covering map. However its restriction to (0, oo) is no longer a a covering map. (Prove this!). Exercise 6.2.2 Let (M, g) and (M, g) be two Riemann manifolds of the same dimension such that (M, g) is complete. Let <j> : M —> M a surjective local isometry i.e. $ is smooth and \v\g = \D(j>{v)\-g VveTM. Prove that 4> is a covering map. The above exercise has a particularly nice consequence. Theorem 6.2.4 (Cartan-Hadamard) Let (M, g) be a complete Riemann manifold with non-positive sectional curvature. Then for every point q € M the exponential map exp, : T„M -+ M is a covering map.
Covering spaces
199
Proof Consider the pull-back h = exp*(p). h is a symmetric non-negative definite (0,2)-tensor field on TqM. It is in fact a metric (i.e. it is positive definite) since the map exp. has no critical points due to the non-positivity of the sectional curvature. The lines 11-» tv through the origin oiTqM are geodesies of h and they are defined for all t € ffi. By Hopf-Rinow theorem we deduce that (TqM, h) is complete. The theorem now follows from Exercise 6.2.2. • Exercise 6.2.3 Let G and G two Lie groups of the same dimension and 4>: G —> G a smooth, surjective group morphism. Prove that 0 is a covering map. In particular, this explains why exp : R —> S1 is a covering map. Exercise 6.2.4 Identify S 3 C K4 with the group of unit quaternions
S3 = { ? e i ; M = i}. The linear space R3 can be identified with the space of purely imaginary quaternions
l 3 = a m l = {xi + yj + zk}. (a) Prove that Vq G S 3 qxq~l e J m H . (b) Prove that for any q G S3 the linear map x H-> qxq'1
T , : 3 m i - > 3mH
is an isometry so that Tq G S O ( 3 ) . Moreover the map S3Bq^Tqe
50(3)
is a group morphism. (c) Prove the above group morphism is a covering map. Example 6.2.5 Let M be a smooth manifold. A Riemann metric on M induces a metric on the determinant line bundle det TM. The sphere bundle of det TM (with respect to this metric) is a covering space of M called the orientation cover of M. Definition 6.2.6 Let X\ ^ Y and X2 ^ Y be two covering spaces of Y. A morphism of covering spaces is a continuous map F : X\ -¥ X2 such that 7r2 ° F = 7Ti i.e. the diagram below is commutative.
Y
200
T i e Fundamental group and covering spaces
If F is also a homeomorphism we say F is an isomorphism of covering spaces. Finally, if X A Y is a covering space then its automorphisms are called deck transformations. The deck transformations form a group denoted by Deck (X, 7r). Exercise 6.2.5 Show that Deck(Re-^? S 1 ) =* Z. Exercise 6.2.6 (a) Prove that the natural projection Sn —> is a covering map. (b) Denote by r i the tautological (real) line bundle over KP n . Using a metric on this line bundle form the associated sphere bundle S^Tik) —> MP". Prove that this defines a covering space isomorphic with the one described at part (a). §6.2.2
Unique lifting property
Definition 6.2.7 Let X A Y be a covering space and F : Z - > Y a continuous map. A lift of / is a continuous map F : Z —> X such that IT O F = f i.e. the diagram below is commutative. X
Y Proposition 6.2.8 (Unique P a t h Lifting)Let X A Y be a covering map, 7 : [0,1] —»• Y a path in Y and XQ a point in the fiber over j/o = 7(0), X0 € 7r_1(y0)- Then there exists at most one lift of 7, T : [0,1] —• Y such that T(0) = Xsy. Proof We argue by contradiction. Assume there exist two such lifts, I Y r ^ : [0,1] ->• X. Set
5 = {te[o,i] ; r!(t) = r2(t)}. S / 0 since 0 € 5. Obviously S is closed so it suffices to prove that it is also open. We will prove that there exists ro > 0 such that [0, r 0 ] C S. The general situation is entirely similar. Pick a small open neighborhood U of x0 such that n restricts to a homeomorphism onto w{U). There exists r0 > 0 such that 7t([0,ro] C U, i = 1,2. Since 7roFi = 7ror 2 we deduce r \ |[o,r0]= ^2 l[o,r0]- The proposition is proved. • Theorem 6.2.9 Let X 4 V be a covering space and / : Z - > F a continuous map, where Z is a connected space. Fix z0 € Z and XQ a point in 7T-1(?/o)i where y0 = f(z0). Then there exists at most one lift F : Z —>• X of / such that F(z0) = x0.
201
Covering' spaces
Proof For each z G Z let az be a continuous path connecting z0 to z. If Fi,F2 are two lifts of / such that Fi(z 0 ) = F2(z0) = xo then for any z G Z the paths i \ = F,(a z ) and T2 = F2{az) are two lifts of 7 = f(az) starting at the same point. From Proposition 6.2.8 we deduce that Tx = T2 i.e. Fi(z) = F2(z) for any z G Z. • §6.2.3
Homotopy lifting property
Theorem 6.2.10 (Homotopy lifting property)Let X 4 7 be a covering space, / : Z - > F a continuous map and F : Z —¥ X a lift of / . If ft: [0,1] x Z->Y
(i,z)>->ft ( (z)
is a homotopy of / (fto(z) = f{z)) then there exists a unique lift of ft H : [0,1] x Z -> X
(t, z) >-> Ht{z)
such that #0(2) = F(z). Proof For each z G Z we can find an open neighborhood C/z of z G Z and a partition 0 = t0 < £1 < • • • < tn = 1 (depending on z) such that ft maps [£;_i, ij] x Uz into an evenly covered neighborhood of ft^^z). Following this partition can now succesively lift ft \jxu, to a continuous map H = Hz : I x Uz —> X such that #„(£) = F(C), VC G Uz. By unique lifting property the liftings on / x UZl and / x UZ2 agree on / x (UZ1 n C/Z2) for any Zi, Z2 € Z and hence we can glue all these local lifts together to obtain the desired lift H on I x Z. • Corollary 6.2.11 (Path lifting property)Let X A Y be a covering space and 7 : [0,1] —• y a continuous path starting at y0 G Y. Then for every xo 6 7r_1 (j'o) there exists a unique lift r : [0,1] —¥ X of 7 starting at a;0. Proof 7(i).
Use the previous theorem with / : {pt} —»• F , / ( p t ) = 7(0) and ft((pt) = •
Corollary 6.2.12 Let X A y be a covering space, y0 G Y and 70,71 G fi(y, y0) If 70 and 7! are homotopic rel yo then any lifts r 0 , Ti which start at the same point also end at the same point i.e. F 0 (l) = T ^ l ) . Proof Lift the homotopy connecting 70 to 71 to a homotopy in X. By unique lifting property this lift connects r 0 to IY We thus get a continuous path r ( ( l ) inside the fiber 7r~1(y0) which connects r 0 ( l ) to T ^ l ) . Since the fibers are discrete this path must be constant. • Let X -^ y be a covering space and y0 G Y. Then for every x G ir~l{ya) and any 7(5) G fi'y, yo) denote by Tx(s) the unique lift of 7 starting at x. Set d
x-7=
4rx(i).
202
The Fundamental group and covering spaces
By Corollary 6.2.12 if w G Q,(Y,y0) is homotopic to 7 rel yo then x • 7 = x • u>.
Hence x • 7 depends only upon the equivalence class [7] € ^i(Y,y0).
Clearly
*-([7]-M) = ( * - [ 7 ] ) - M and X ' CyQ
X
so that the correspondence Tr~l(y0) x 7Ti(Y,2/o) 3 (a;,7) ^
x
• 7 e x • 7 G 7r_1(y0)
defines a right action of ^ ( F , 2/0) on the fiber 7r_1(yo)- This action is called the monodromy . The map x H-> X • 7 is called the monodromy along 7. Note that when Y is simply connected the monodromy is trivial. The map 7r induces a group morphism 7rt :7r x (X,a; 0 )->7ri(r,j/o) ^o G 7r_1(j/0)Proposition 6.2.13 TT» is injective. Proof Indeed, let 7 G tt(X, xo) such that 7r(7) is trivial in 7rj (F, yo)- The homotopy connecting n(j) to cOT lifts to a homotopy connecting 7 to the unique lift of eyo at xo which is txo. • §6.2.4
On the existence of lifts
Theorem 6.2.14 Let X A Y be a covering space, x0 € X, y0 = ir(x0) G Y, / : 2 4 V a continuous map and z0 € Z such that /(zo) = 2/o- Assume the spaces Y and .Z are connected (and thus path connected). / admits a lift F : Z —> X such that F(z0) = x0 if and only if /.MZ.^CTr.fa^.zo))-
(6.2.1)
Proof Necessity. If F is such a lift then using the functoriality of the fundamental group construction we deduce / , = 7r« o Ft. This implies the inclusion (6.2.1). Sufficiency For any z € Z choose a path j z from zo t o z- Then az = f(ryz) is a path from y0 to y = f(z). Denote by Az the unique lift of az starting at x0 and set F(z) = AZ{1). We claim that F is a well defined map. Indeed, let uz be another path in Z connecting z0 to z. Set \ z = }{LJZ) and denote by Az its unique lift in X starting at XQ. We have to show that A z (l) = Az{\). Construct the loop based at z0 /}z = uiz*~(z.
203
Covering spaces
f(/3z) is a loop in Y based at y^. From (6.2.1) we deduce that the lift Bz of /(/9Z) at x0 € X is a closed path (i.e. the monodromy along /(/3 2 ) is trivial). We now have A,(l) = B , ( l / 2 ) = A,(0) =
Az(l).
This proves F is a well defined map. We still have to show this map is also continuous. Pick z € Z. Since / is continuous, for every arbitrarily small, evenly covered neighborhood U of f(z) € Y there exists a path connected neighborhood V of z S Z such that f(V) C U. For any ( £ V pick a path a = a^ in V connecting z to £. Let UJ denote the path w = j z * a^ (go from zo to z along yz and then from z to C along
V7G7V
Then (Y,p) is an universal covering space of Y.
•
Exercise 6.2.7 Finish the proof of the above theorem.
204
The Fundamental group and covering spaces
Example 6.2.19 K ^? 5 1 is the universal cover of Sl. More generally, Exp : Mn -> (*i, - - -, *«) •—> (exp(27rit 1 ),..., exp(27rztn)) is the universal cover of Tn. The natural projection p : Sn -> I P " is the universal cover of MP". Example 6.2.20 Let (M, g) be a complete Riemann manifold with non-positive sectional curvature. By Cartan-Hadamard theorem the exponential map exp 9 : TqM —> M is a covering map. Thus the universal cover of such a manifold is a linear space of the same dimension. In particular the universal covering space is contractible!!! We now have another explanation why Exp : Kn —• Tn is a universal covering space of the torus: the sectional curvature of the (flat) torus is zero. Exercise 6.2.8 Let (M,g) be a complete Riemann manifold and p : M —>• M its universal covering space. (a) Prove that M has a natural structure of smooth manifold such that p is a local diffeomorphism. (b) Prove that the pullback p*g defines a complete Riemann metric on M locally isometric with g. Example 6.2.21 Let (M,g) be a complete Riemann manifold such that Ric {X, X) > const.\X\2g
(6.2.2)
where const denotes a strictly positive constant. By Myers theorem M is compact. Using the previous exercise we deduce that the universal cover M is a complete Riemann manifold locally isometric with (M, g). Hence the inequality (6.2.2) continues to hold on the covering M. Myers theorem implies again that the universal cover M is compact!! In particular, the universal cover of a semisimple, compact Lie group is compact!!! §6.2.5
The universal cover and the fundamental group
Theorem 6.2.22 Let i ^ X b e the universal cover of a space X. Then ni{X,pt) Proof
^Deck(X^X).
Fix (,o € X and set x0 = p(£o)- There exists a bijection Ev:Deck(X) ->p_1(i0)
205
Covering spaces given by the evaluation
Ev(F)=F(£o). _1
For any £ 6 7r (xo) let 7^ be a path connecting £0 to £. Any two such paths are homotopic rel endpoints since X is simply connected (check this). Their projections on the base X determine identical elements in it\{X, xo). We thus have a natural map tfiDeckW-^psT.Xo)
F ^
p{-yF{io)).
$ is clearly a group morphism. (Think monodromy!). The injectivity and the surjectivity of \£ are consequences of the lifting properties of the universal cover. Q Corollary 6.2.23 If the space X has a compact universal cover then iri(X,pt) finite.
is
Proof Indeed the fibers of the universal cover have to be both discrete and compact. Hence they must be finite. The map E v in the above proof maps the fibers bijectively onto Deck (X). O Corollary 6.2.24 (Weyl) The fundamental group of a compact semisimple group is finite. Indeed from we deduce from Example 6.2.21 that the universal cover of such a group is compact. Example 6.2.25 From Example 6.2.5 we deduce that 7ri(5x) = (Z, + ) . Exercise 6.2.9 (a)Prove that 7r1(MP",pt) S Z 2 , Vn > 2. (b) Prove that T T ^ T ^ S Z " . Exercise 6.2.10 Show that the natural inclusion U(n — 1) •-+ U(n) induces an isomorphism between the fundamental groups. Conclude that the map det : U{n) -> S 1 induces an isomorphism
Chapter 7 Cohomology 7.1 §7.1.1
D e R h a m cohomology Speculations around the Poincare lemma
To start off, consider the following partial differential equation in the plane: given two smooth functions P and Q find a smooth function u such that £
= '•
%">•
<"•*>
As it is, the formulation is still fuzzy since we have not specified the domains of the functions u, P and Q. As it will turn out, this aspect has an incredible relevance in geometry. Equation (7.1.1) has another interesting feature: it is overdetermined i.e. it imposes too many conditions on too few unknowns. It is therefore quite natural to impose some additional restrictions on the data P, Q just like the zero determinant condition when solving overdetermined linear systems. To see what restrictions one should add it is convenient to introduce the 1-form a = Pdx + Qdy. (7.1.1) can be rewritten as du = a.
(7.1.2)
If (7.1.2) has at least one solution u then 0 = d2u = da so that a necessary condition for existence is da = 0 i.e. dP__dQ^ dy dx' 206
(7.1.3)
DeRham
207
cohomology
A form satisfying (7.1.3) is said to be closed. Thus, if the equation du = a has a solution then a is necessarily closed. Is the converse also true? Let us introduce a bit more terminology. A form a such that the equation (7.1.2) has a solution is said to be exact. The motivation for this terminology comes from the fact that sometimes du is called the exact differential of u. We thus have an inclusion of vector spaces {exact forms} C {closed forms}. Is it true the opposite inclusion also holds? Amazingly, the answer to this question depends on the domain on which we study (7.1.2). The Poincare lemma comes to raise our hopes. It says that this is always true at least locally. Lemma 7.1.1 (Poincare lemma) Let C be an open convex set in R" and a € Qfc(C). Then the equation du = a (7.1.4) has a solution w € Q* -1 (C) if and only if a is closed, da = 0. Proof Necessity is clear. We prove sufficiency. We may as well assume 0 £ C. Consider the radial vector field on C d
r- = xli — OXi
and denote by <3>( the flow it generates. $ t is the linear flow $ t (x) = e'x, i G B " . Note that since C is convex the flow lines of $ s , which are the straight lines through the origin, intersect C along connected segments originating at 0. We begin with an a priori study of (7.1.4). Let u satisfy du = a. Using the homotopy formula Lf — dif + ifd we get dL/u = d(dif + ifd)u = difCt i.e. d{LfU — ifOt) = 0. This suggests looking for solutions of LfU = if a.
(7.1.5)
208
Cohomology
If u is a solution of this equation then Lfdu = dLfU — difCc = LfOt — igda. = Lfa. Hence u also satisfies Lf(du — a) = 0. 1
Set u) = du — a = Y,i i^idx . Using the computations in Subsection 3.1.3 we deduce Lfdx' = dx% so that LfOj = y^XLftu^dx1 = 0. / We deduce that Lftjj = 0 and consequently the coefficients uij are constants along the the flow lines of $ ( which all converge at 0. Thus u>\ = ci = constant. Each monomial cidx1 is exact i.e. there exits r\i 6 Q,k~l(C) such that dr]j = Cjdx1. For example, when I =\ <2 < •••
u= J
mifct)dt.
(7.1.6)
J —CO
Here the convexity assumption on C enters essentially since it implies that *,(C)CC
Vi<0
so that if the above integral is convergent then u is a form on C. If we write
then u{x) = Y,(f
rfj{x)dt\ dx1.
(7.1.7)
DeRham
209
cohomology
We have to check two things. A. The integral in (7.1.7) is well denned. To see this we first write ajdxJ
a = ^ \j\=k
and set A(x) = max v
'
laj(r:r)|. v
J;0
"
Then $*(va) = e ' v f ^2
aJ(etx)dxJ
so that 1^/(^)1 < C*e*|a:tA(x) Vt < 0. This proves the integral in (7.1.7) converges. B . u defined by (7.1.6) is a solution of (7.1.5). Indeed, on differential forms LfU = lim - (
lixa-U'f
S
&{ipa)dt-
s->0 s \
[
J-oo
= lim f f
$*t{ifa)dt) I
J-oo
$*t(i?a)dt - f
s->0 \J-oo
$*t{i?a)dt) J
J-oo
= lim / $*t{ifa)dt = %{ifa)
= ifO.-
Poincare lemma is proved. • Local solvability does not in any way implies global solvability. Something happens when one tries to go from local to global. Example 7.1.2 Consider the form d9 on K2 \ {0} where (r, 6) denote the polar coordinates in the punctured plane. To write it in cartesian coordinates (x, y) we use the equality V tan0 = x so that d Jfl__ (1 + tan, 22 m 0)d0 = — y\2djx _ +, v x x
(-i^y2\M ,.
-ydx + xdy
-ydx + xdy x2 + y2
•• a .
210
Cohomology
Obviously, da = d?9 = 0 on I 2 \ {0} so that a is closed on the punctured plane. Can we find a smooth function u on I 2 \ {0} such that du — a? We know that we can always do this locally. However, we cannot achieve this globally. Indeed, if this was possible then
f du= f 1 a= f l d6 = 2-K.
Js1
Js
Js
On the other hand, using polar coordinates u — u(r, 9) we get
f du=f 1 %d9 Js1 Js 89 = JQ ^ d 0 = u(l,2ir)-u(l,O) = O. Hence on M2 \ {0} {exact forms} / {closed forms}. We see what a dramatic difference a point can make: K2 \ {point} is structurally very different from K2. The artifice in the previous example simply increases the mystery. It is still not clear what makes it impossible to patch-up local solutions. The next subsection describes two ways to deal with this issue. §7.1.2
Cech vs. D e R h a m
Let us try to analyze what prevents the "spreading" of local solvability of (7.1.4) to global solvability. We will stay in the low degree range. Cech approach Consider w a closed 1-form on a smooth manifold. To solve the equation du = u, u 6 C°°(M) we first cover M by geodesically convex open sets (Ua) (with respect to some fixed Riemann metric). By Poincare lemma we can solve du = u on each open set Ua so that we can find a smooth function fa € C°°(Ua) such that dfa — w. We get a global solution if and only if
fa/) = fa - h = °
on each U
"P = UanU0^
For fixed a the solutions of the equation du = uo on Ua differ only by additive constants (i.e. closed 0-forms). The quantities /Q/? satisfy dfap = 0 on the (connected) overlaps Ua/3 so they are constants. Clearly they satisfy the conditions /Q/3 + hi + ha = 0 on (7Q/37 7^ 0.
(7.1.8)
DeRham
211
cohomology
On each Ua we have as we have seen several choices of solution. Altering a choice is tantamount to adding a constant fa —> fa + ca. The quantities fap change according to faB —> faB +Ca— Cp.
Thus the global solvability issue leads to the following situation. Pick a collection of local solutions fa. The equation du = u> is globally solvable if we can alter each fa by a constant ca such that faB = {cp - ca)
Va, /? such that Ua3 ± 0.
(7.1.9)
We can start the alteration at some open set Ua and work our way up from one such open set to its neighbors, always trying to implement (7.1.9). It may happen that in the process we might have to return to an open set whose solution was already altered. Now we are in trouble. (Try this on S1 and w = dQ.) After several attempts one can point the finger to the culprit: the global topology of the manifold may force us to always return to some already altered local solution. Notice that we replaced the partial differential equation du = ui with a system of linear equations (7.1.9), where the constants fap are subject to the constraints (7.1.8). This is no computational progress since the combinatorics of this system makes it impossible solve in most cases. The above considerations extend to higher degree and one can imagine the complexity increases considerably. This is however the approach Cech adopted in order to study the topology of manifolds and although it may seem computationally hopeless, its theoretical insights are invaluable. D e R h a m Approach This time we postpone asking why the global solvability is not always possible. Instead, for each smooth manifold M one considers the Z-graded vector spaces B*(M) = ®k>0Bk(M),
(B°(M)
d
=f {0})
Z*(M} =
®k>0Zk(M)
where Bk(M)
= {du; 6 Qk(M) ; u) € Q,k~l(M)} = exact k - forms
and Zk(M = {rie nk{M) k
; dq = 0} = closed k - forms.
k
Clearly B C Z . Form the quotients Hk{M) =
Zk{M)/Bk{M).
Intuitively, this space consists of those closed fc-forms w for which the equation du =
UJ
has no global solution u € fl*-1(M). Thus if we can somehow describe these spaces we may get an idea " who" is responsible for the global nonsolvability.
212
Cohomology
Definition 7.1.3 For any smooth manifold M the vector space Hk(M) fc-th DeRham cohomology group.
is called the
Clearly Hk{M) = 0 for k > dimM. Example 7.1.4 The Poincare lemma shows that Hk(Rn) sion in Example 7.1.2 shows that H^M2 \ {0}) # 0.
= 0 for k > 0. The discus-
Proposition 7.1.5 For any smooth manifold M dim H°(M) = number of connected components of M. Proof
Indeed H°(M) = Z°(M) = {/ € C^iM)
; df = 0}
Thus H°(M) coincides with the linear space of locally constant functions. These are constant on the connected components of M. O Already H°(M) contains an important topological information. Obviously the groups Hk are diffeomorphism invariants and its is reasonably to suspect the higher cohomology groups may contain more topological information. Thus, to any manifold we can now associate the graded vector space H*{M)"=f
($Hk{M).
A priori, the spaces Hk may be infinite dimensional. The Poincare polynomial, denoted by PM{t) is defined by P M (t) = ^ i f c d i m F A : ( M ) k>0
every time the right-hand-side expression makes sense. The number dim Hk (M) is usually denoted by 6^(M) and is called the k-th Betti number of M. Hence PM(t) = ] > > ( M ) t * . k
Exercise 7.1.1 Show that Psi(t) =
l+t.
We will spend the remaining of this chapter trying to understand what is that these groups do and which (if any) is the connection between the two approaches outlined above.
DeRham §7.1.3
cohomology
213
Very little homological algebra
At this point it is important to isolate the common algebraic skeleton on which both DeRham and Cech approaches are built. This requires a little terminology from homological algebra. In the sequel all rings will be assumed commutative with 1. Definition 7.1.6 (a) Let R be a ring, C = © „ 6 z C n and D = © n 6 z A i two Z-graded left .R-modules. A degree fc-morphism > : C -» D is an fl-module morphism such that 4>{Cn)cDn+k VneZ. (b) Let C = ®nez b e a Z-graded fl-module. A boundary (resp. coboundary) operator is a degree -1 (resp. a degree 1) endomorphism d : C —» C such that d2 = 0. A chain (resp. cochain) complex over R is a pair (C, d) where C is a Z-graded i?-module and d is a boundary operator (resp. a coboundary operator). In this book we will be interested mainly in cochain complexes so in the remaining part of this subsection we will stick to this situation. In this case cochain complexes are usually described as (C = © n 6 z C " , d ) . Moreover we will consider only the case C" = 0 for n < 0. Traditionally, a cochain complex is represented as a long sequence of -R-modules and morphisms of .R-modules > Cn~l
(C, d) :
d
^z\Cn
such that range [dn-i) C ker (dn) i.e. dndn-i
-^
Cn+l -> • • •
= 0.
Definition 7.1.7 Let >
£„_!
d^cn
_d^
cn+1
_^
be a cochain complex. Set Zn(C) = kerd„
Bn{C) = range
(dn^).
The elements of Zn(C) are called cocycles, while the elements of Bn(C) are called coboundaries. Two cocycles c, d £ Zn(C) are said to be cohomologous if c — d G Bn(C). The quotient module Hn{C)
d
=
Zn{C)/Bn(C)
is called the n-th cohomology group (module) of C. It can be identified with the set of equivalence classes of cohomologous cocycles. A cochain complex complex C is said to be acyclic if Hn(C) = 0 for all n > 0.
214
Cohomology For a cochain complex C one usually writes H'(C) =
@Hn{C). n>0
Example 7.1.8 ( D e R h a m complex) Let M be an m-dimensional smooth manifold. Then the sequence
o -> n°(M) -4 9} -4 • • • 4 nm(M) -> o (where d is the exterior derivative) is a cochain complex called the DeRham complex. Its cohomology is the DeRham cohomology of the manifold. Example 7.1.9 Let (g, [•, •]) be a real Lie algebra. Define d : A*g* -> A*+1fl* by
(<M(*o, Xlt...,Xk)=
£
( - W K . **]. *o, • • •, Xit..., Xit..., X,)
0
where as usual, the hat indicates a missing argument. According to the computations in Example 3.2.6 d is a coboundary operator so that (A*g*, d) is a cochain complex. Its cohomology is called the Lie algebra cohomology and is denoted by H*(g). Exercise 7.1.2 (a) Let g be a real Lie algebra. Show that H1(g)^(g/[g,g}T where [g, g] = span {[X, Y] ; X,Y e g}. (b) Compute Hl{gl{n, K)) where gUn, R) denotes the Lie algebra of nxn real matrices with the bracket given by the commutator. (c) (Whitehead) Let g be a semisimple Lie algebra (i.e. nondegenerate Killing pairing). Prove that H1^) = {0}. (Hint: Prove that [fl^]"1" = 0 where _L denotes the orthogonal complement with respect to the Killing pairing.) Proposition 7.1.10 Let (C,d):
• C"-1 ^ C " - ^ > C n + 1 -> • • •
be a cochain complex of /^-modules. Assume moreover that C is also a Z-graded .R-algebra i.e. there exists an associative multiplication such that Cn-CmcCm+n
Vm,n.
If d is a quasi-derivation i.e. d(x • y) = ±(dx) • y ± x • (dy)
\/x, y G C
then H*(C) inherits a structure of Z-graded .R-algebra.
DeRham
cohomology
215
A cochain complex as in the above proposition is called a differential graded algebra or D G A . Proof It suffices to show Z*{C) • Z*{C) C Z*{C) and B*{C) • B*{C) C B*{C). If dx = dy = 0 then d(xy) = ±(dx)y ± x{dy) = 0. If x = dx' and y = dy' then since d2 = 0 we deduce xy = ±(dx'dy'). D Corollary 7.1.11 The DeRham cohomology of a smooth manifold has an K-algebra structure induced by the exterior multiplication of differential forms. Definition 7.1.12 Let (A,d) and {B,S) be two cochain complexes. (a) A cochain map is a degree 0 morphism <j> : A —¥ B such that <j> o d = 5 o cj> i.e. the diagram below is commutative for any n.
An
_*!_
^n+1
Bn
_i2_^
Bn+1
(b) Two cochain maps 4>,ip : A —> B are said to be cochain nomotopic and we write this
such that tp o
B and ip : B -¥ A k
1B.
Example 7.1.13 The commutation rules in Subsection 3.2.1, namely [Lx,d] = 0 and [ i x i ^ s = Lx show that for each vector field X on a smooth manifold M the Lie derivative along X, Lx : fl*(M) —> £l*(M) is a cochain map homotopic with the trivial map (= 0). The interior derivative ix is the cochain homotopy achieving this. Proposition 7.1.14 (a) Any cochain map <j> : (A, d) -> (B,5) induces a degree zero morphism in cohomology 0 , : H*{A) -»• f T ( 5 ) . (b) If the cochain maps cj),ip : A —• B are cochain homotopic then they induce identical morphisms in cohomology, >» = i/>,. (c) (1A)» = 1H'(,A) and if
(A ) d)4(A , ,d')4(A",d") are cochain m a p s t h e n (ip o <[>)t = %j)t o <^>„.
216 Proof
Cohomology (a) It boils down to checking the inclusions
and 4>{Bn{A)) c
Bn{B).
These follow immediately from the definition of a cochain map. (b) We have to show that >(cocycle) — ^(cocycle) = coboundary. Let da = 0. Then 4>{a) — ip(a) = ±5(x{a) ± x{da) =
dA
0
dB
An+1
•
An
• dA
0
An~l
• dA
t l * Bn+1
Wn V'n+l.
+\
Qn
dB
dA
0
dc
B
n
ll>n
C
dB
dp
Bn
C
dB
dc
n
(7.1.10)
n
- 0
DeRham cohomology
217
in which the rows are exact. Then there exists a long exact sequence > Hn-\C)
V
Hn{A) h Hn(B) % Hn{C) % Hn+1(A)
->••••.
(7.1.11)
We will not include a proof of this proposition. We believe this is one proof in homological algebra the reader should try it on its own. We will just indicate the construction of the connecting maps dn. This construction and in fact the entire proof relies on a simple technique called diagram chasing. Start with x e Hn(C). x can be represented by some cocycle c e Zn(C). Since tp„ is surjective there exists b G Bn such that c = i>n{b). From the commutativity of the diagram (7.1.10) we deduce 0 = dc^n{b) = ipn+idB(b) i.e. dB(b) € keri/'n+i = range>n+i- In other words, there exists a € An+1 such that >n+i(a) = dBb. We claim o is a cocycle. Indeed
= 0.
Since
o^(c)
=
we can
This is not entirely rigorous since a depends on various choices. We let the reader check that the correspondence Zn(C) 3 c >-> a e Zn+1(A) above induces a well defined map in cohomology, dn : Hn{C) —>• Hn+1(A} and moreover, the sequence (7.1.11) is exact. Exercise 7.1.3 (Abstract Morse inequalities) Let C = (&n>0Cn be a cochain complex over the field F . Assume each of the vector spaces Cn is finite dimensional. Form the Poincare series Y,tn
Pc{t) = 71>0
and J^tndimFHn(C).
PH'(C)(t) = n>0
Prove that there exists a formal series R(i) (E Z[[t]] with non-negative coefficients such that Pc(t) = PH.{c)(t) + (1 +t)R(t). In particular, whenever it makes sense, the graded spaces C* and H* have identical Euler characteristics X(C*) = P C ( - 1 ) = P H . ( C ) ( - 1 ) =
X(H*(C)).
218
Cohomology
Exercise 7.1.4 (Additivity of Euler characteristic) Let 0-+yl->B->C->0 be a short exact sequence of cochain complexes over the field F. Prove that if at least two of the cohomology modules H*(A), H*{B) and H*(C) have finite dimension over F then the same is true about the third one and moreover x(H*(B))=x(H*(A})
+
x(H*(C)).
Exercise 7.1.5 (Finite dimensional Hodge theory) Let (©„>oV™, dn) be a cochain complex over the reals such that dim Vn < oo for all n. Assume each Vn is an euclidian space and denote by d*n : Vn+1 —> V" the adjoint of dn. We can now form the Laplacians A„ : Vn -)• Vn An = d*ndn + <*„_!<_!. (a) Prove that Anc = 0 iff dnc = 0 and d'.jC = 0. (b) Let c € Zn(C). Prove that there exists a unique c G Zn(C) such that \c\ = min{|c'| ; c-c' e Bn(C)}
cohomologous to c
where | • | denotes the euclidian norm in V". (c) Prove that c determined in part (b) satisfies A n c = 0. Deduce from all the above the Hodge theorem Hn(V) = kevAn. Exercise 7.1.6 Let V be a real vector space and v0 G V. Define dk : AkV —• Ak+1V by LJ t-> Vo A u).
(a) Prove that . . . " V A*v 4 Ak+1V %' ••• is a cochain complex (known as Koszul complex). (b) Use the finite dimensional Hodge theory described in previous exercise to prove that the Koszul complex is acyclic if v0 ^ 0 i.e. Hk{A*V,d) = 0 V/fc>0 §7.1.4
Functorial properties of D e R h a m cohomology
Let M and N be two smooth manifolds and <j>: M —> N a smooth map. The pullback >* : Q*{N) -*•
is a cochain map i.e. 4>*dN = dM>*.
Q*(M)
DeRham
219
cohomology
Thus <j>* induces a morphism in cohomology which we continue to denote by >*;
(£, m) i-> $ ( (m)
such that i for i = 0,1. (b) A smooth map cj>: M —• N is said to be a (smooth) homotopy equivalence if there exists a smooth map ip : N —>• M such that cj>otp ~ s h 1^ and ip o
H*(M).
Proof According to the general results in homological algebra it suffices to show the pullbacks #J,
-¥Q*~1{M)
such that
^!M - $S(w) = ±x(<M ± dxw. At this point, our discussion on the fibered calculus of Subsection 3.4.5 will pay off. The projection $ : / x A f - t t f defines an oriented 9-bundle (with standard fiber I). For any u> G Q*(N) we have the equality
220
Cohomology
JldIxM)/M J(dIxM)/M
We now use the homotopy formula of Theorem 3.4.40 of Subsection 3.4.5. We get J(dIxM)/M
J(IxM)/M
= f
J(IxM)/M
&(dNU>)-dU[
J(IxM)/M
*».
J(IxM)/M
7(/xM)/M
is the sought for cochain homotopy. • Corollary 7.1.20 Two homotopically equivalent spaces have isomorphic cohomology rings. Consider now a smooth manifold and U, V two open subsets such that M = ULiV. Denote by ly (resp. iy) the inclusions U *-» M (resp. V •-» M). These induce the restriction maps i\j : Q*(M)
->• Q*(U)
u)^u\u
and ^ : fi*(M) -> fi*(V) u H u |
F
.
We get a cochain map r :fi*(M) -»•«*(£/) ©fi*(V)
There exists another cochain map
5 •. n*(u) efl'(v) -> n*(u nv) {w,rf) H-> -w|t/nv +v\unv • Lemma 7.1.21 The short Mayer-Vietoris sequence
o -+ n*(M) A n*(i7) © n*(v) 4o*(c/nv)->o is exact. Proof Obviously r is injective. The proof of the equality Range r = ker 6 can be safely left to the reader. The surjectivity of 5 requires a little more effort. {U, V} is an open cover of M so we can find a partition of unity {
DeRham
221
cohomology
subordinated to this cover i.e. supp ipu C U, supp ipv C V 0 <
¥>t/ + W = 1-
Note that for any u> 6 Q*([/ n V) supp ^ w C supp ipv
cV
and thus, upon extending ipuui by 0 outside V we can view it as a form on M. Restricting to U we get a form ipvu> £ Q*(C/). Similarly, ipvuj £ fT(K). Note that 5(-
Hk+\M)
-> • • •
called the long Mayer-Vietoris sequence. The connecting morphisms d can be explicitly described using the prescriptions following Proposition 7.1.16 in the previous subsection. Start with ui € Qk(U D V) such that dui = 0. Writing as before
we deduce d(
Thus we can find r\ € Ct
on U n V.
(M) such that r] \u= d(ipvtu)
T)\v=
d(-ipuu).
Then duj =
TJ.
The reader can prove directly that the above definition is independent of the various choices. The Mayer-Vietoris sequence has the following functorial property.
222
Cohomology
P r o p o s i t i o n 7.1.23 Let <j> : M —> TV be a smooth map and {U, V} an open cover of N. Then U' = >_1(C/), V = ^> _1 (^) f ° r m a n ° P e n cover of M and moreover, the diagram below is commutative. Hk{N)
Hk{U)®Hk{V)
Hk{UCiV)
Hk+\N)
Hk{M)
Hk{U')®Hk{V)
Hk{U' n V)
Hk+1(M)
Exercise 7.1.7 Prove the above proposition. §7.1.5
Some simple e x a m p l e s
The Mayer-Vietoris theorem established in the previous subsection is a very powerful tool for computing the cohomology of manifolds. In principle, it allows one to recover the cohomology of a manifold decomposed into simpler parts, knowing the cohomologies of its constituents. In this subsection we will illustrate this principle on some simple examples. E x a m p l e 7.1.24 ( T h e cohomology of spheres)The cohomology of S 1 can be easily computed using the definition of DeRham cohomology. We have /f°(S' 1 ) = R since S 1 is connected. The closed 1-forms on S 1 have the form const.dO so that H^S^^M. Thus Psi(t) = l + t. To compute the cohomology of higher dimensional spheres we use Mayer-Vietoris theorem. The (n + l)-dimensional sphere Sn+1 can be covered by two open sets Us = Sn+1 \ {north pole} and Un = Sn+1 \ {south pole}. Each is diffeomorphic to M n+1 . Note that the overlap Un D Us is homotopically equivalent with the "Equator" 5". The Mayer-Vietoris sequence gives Hk{Un) 0 Hk{Us) -> Hk(Un (~l Us) -> Hk+1(Sn+1)
-> Hk+\Un)
® Hk+\US)
^ 0.
The last isomorphism is precisely Poincare lemma. For k > 0 the group on the left is also trivial so that we have the isomorphisms Hk(Sn)^Hk(UnDUs)^Hk+1{Sn+1)
k>0.
Denote by Pn{i) the Poincare polynomial of Sn and set Qn(t) = Pn{t) — Pn(0) = P„{t) — 1. We can rewrite the above equality as Qn+i(t) = tQn(t)
n>0.
DeRham
223
cohomology
Since Q\{t) = t we deduce Qn(t) = tn i.e. Psn{t) =
l+tn.
Example 7.1.25 Let {U, V} be an open cover of the smooth manifold M. We assume that all the Betti numbers of U, V and U C\V are finite. Using the Mayer-Vietoris short exact sequence and Exercise 7.1.4 in Subsection 7.1.3 we deduce that all the Betti numbers of M are finite and moreover we have x(M) = x(U) + x(V)-X(UnV).
(7.1.12)
This resembles very much the classical inclusion-exclusion principle in combinatorics. We will use this simple observation to prove that the Betti numbers of a connected sum of p-tori is finite and then compute its Euler characteristic. Let E be a surface with finite Betti numbers. From the decomposition E = (E \ disk) U disk we deduce (using again Exercise 7.1.4) x(E) = X ( S \ disk) + x(disk) -
X
((E \ disk) n (disk)).
Since (E \ disk) n disk is homotopically a circle and x(disk) = 1 we deduce X (E)
= x(E \ disk) + 1 - xiS1) = x ( S \ disk) + 1.
If now Ej and £2 are two surfaces with finite Betti numbers then E X # E 2 = (E x \ disk) U (E 2 \ disk) where the two holed surfaces intersect over an entire annulus, which is homotopically a circle. Thus X(Ei#E2)
= X (Ei \ disk) + X (E 2 \ disk) -
l
X(S
)
= x(2i) + X(E2) - 2. This equality is identical with the one proved in Proposition 4.2.23 of Subsection 4.2.5. We can decompose a torus as a union of two cylinders. The intersection of these cylinders is the disjoint union of two annuli so homotopically, this overlap is a disjoint union of two circles. In particular, the Euler characteristic of the intersection is zero. Hence x(torus) = 2x(cylinder) = 2x(circle) = 0. We conclude as in Proposition 4.2.23 that X (connected sum of g tori) = 2 — 2g. This is a pleasant, surprising connection with the Gauss-Bonnet theorem. And the story is not over.
224 §7.1.6
Cohomology The Mayer-Vietoris principle
We describe in this subsection a "patching" technique which is extremely versatile in establishing general homological results about arbitrary manifolds building up from elementary ones. Definition 7.1.26 A smooth manifold M is said to be of finite type if it can be covered by finitely many open sets U\,..., Um such that any nonempty intersection f/ij n • • • n Uik (k > 1) is diffeomorphic to R d l m M . Such a cover is said to be a good cover. Example 7.1.27 All compact manifolds are of finite type. To see this it suffices to cover such a manifold by finitely many open sets which are geodesically convex with respect to some Riemann metric. If M is a finite type manifold and U C M is a closed subset homeomorphic with the closed unit ball in R d l m M them M \ U is a finite type non-compact manifold. (It suffices to see that M" \ closed ball is of finite type). The connected sums and the direct products of finite type manifolds are finite type manifolds. Proposition 7.1.28 Let p : E —> B be a smooth vector bundle. If the base B is of finite type then so is the total space E. In the proof of this proposition we will use the following fundamental result. Lemma 7.1.29 Let p : E —> B be a smooth vector bundle such that B is diffeomorphic to K". Then p : E ->• B is a trivial bundle !!! Proof of Proposition 7.1.28. Denote by F the standard fiber of E. F is a vector space. Let (C/j) be a MV-cover of B. For each ordered multi-index I: {i\ < • • • < i^} denote by Uj the multiple overlap U^ D • • • D Uik. Using the previous lemma we deduce that each Ej = E \Vl is a product FxUi and thus it is diffeomorphic with some vector space. Hence (£*) is a MV-cover. D. Exercise 7.1.8 Prove Lemma 7.1.29. Hint Assume E is a vector bundle over the unit open ball B e l " . Use V-parallel transport along f= —x'-^ where V is a connection on E. We organize the family of n-dimensional, finite type smooth manifolds as a category 9Jln. The morphisms of this category are the embeddings (i.e. 1-1 immersions) Afi -> M 2 (Mi G mn).
DeRham
225
cohomology
Definition 7.1.30 Let R be a commutative ring with 1. A contravariant MayerVietoris functor (or MV-functor for brevity) is a contravariant functor from the category 9Jt n to the category of Z-graded .R-modules T = ®n^Tn
-> G r a d R-MoA
M .->
®Pi{M)
with the following property. If {U, V} is a MV-cover of M G SD?„ i.e. U, V, UC\ V G 97t„ then there exist morphisms of .R-modules dn:Fl{Ur\V)^Tn+1{M) such that the sequence below is exact.
—>^"(M) 4 J^{U)® J^{V) - 4 ^ ( u n v ) ^
J^+1(M)
->••••
where r* is defined by r* = F(iv) ® T{iv) while 8 is defined by 5(x ®y)=
T{iunv)(y)
-
T(tunv)(x).
(The maps i. denote natural embeddings.) Moreover, if TV G 97tn is an open submanifold of TV and {U, V) is a MV cover of M such that {U f~l TV, V D TV} is a MV-cover of TV then the diagram below is commutative.
r(f/nV) —^— ^"+1(M)
^"(c/nvnTV) —^-
^"+1(TV)
The vertical arrows are the morphisms ^ ( J . ) induced by inclusions. The covariant MV-functors are defined in the dual way, by reversing the orientation of all the arrows in the above definition. Definition 7.1.31 Let T, Q be two contravariant MV-functors 97tn -> G r a d ijMod. A c o r r e s p o n d e n c e between these functors is a collection of .R-module morphisms $M = ©
-»• © e n ( M )
226
Cohomology
(one collection for each M e 97ln) such that for any embedding Mx A M 2 the diagram bellow is commutative ^ " ( M 2 ) -?&L
^»(Mi)
0M,
en(M2)
g(v).
^(MO
and for any M G !J7tn and any MV cover {U, V} of M the diagram below is commutative. F((/(17) ~^- r'+^M) <}>UC\V
gn{unv)
<$>M
- ^ — gn+1(M)
The correspondence is said to be a natural equivalence if all the morphisms <J>M are isomorphisms. Theorem 7.1.32 (Mayer-Vietoris principle)Let T, Q be two (contravariant) MayerVietoris functors on 2DIn and <j>: T —> Q & correspondence. If <j)kw : Fk{Rn)
-> 0*(R")
is an isomorphism for any k € Z then 0 is a natural equivalence. Proof
The family of finite type manifolds SDTn has a natural filtration
tml c m2n c • • • c an; c • • • where 9JTn is the collection of all smooth manifolds which admit a good cover consisting of at most r open sets. We will prove the theorem using an induction over r. The theorem is clearly true for r = 1 by hypothesis. Assume 4>kM is an isomorphism for all M € SD^ -1 . Let M € WlTn and consider a good cover {Uu..., Ur} of M. Then {U =
U1U---UUr-1,UT}
is a MV-cover of M. We thus get a commutative diagram .^"(E/) © ^ " ( t / r ) — • J ^ ( * y n t / r )
Gn{U) ® Gn{Ur)
gn{unur)
^"+1(M)
— ^" +1 (c/)©a n+1 (c/r
e n+1 (M) — ^ i + 1 (c/)©e n + 1 (^-)
DeRham
227
cohomology
The vertical arrows are defined by the correspondence
/-l
B-2 —>• B - i — * Bo
h " B\
h
*- Bi
If fi is an isomorphism for any i ^ 0 then so is / 0 . Exercise 7.1.9 Prove the five lemma. The five lemma applied to our situation shows that the morphisms 4>M must be isomorphisms. D Remark 7.1.34 (a) The Mayer-Vietoris principle is true for covariant MV-functors as well. The proof is obtained by reversing the orientation of the horizontal arrows in the above proof. (b) The Mayer-Vietoris principle can be refined a little bit. Assume that !F and Q are functors from £DT„ to the category of Z-graded .R-algebras and
The Kiinneth formula
We learned in principle how to compute the cohomology of an "union of manifolds". We will now use the Mayer-Vietoris principle to compute the cohomology of products of manifolds. Theorem 7.1.36 (Kiinneth formula) Let M G 971m and N € SDTn. Then there exists a natural isomorphism of graded M-algebras H'{M xN)
= H'(M)
® H*{N) = 0
(®p+q=nH>>(M) ®
In particular, we deduce PMxN{t)
= PU{t) • PN(t).
H"{N)).
228 Proof
Cohomology We construct two functors JF,S:S!7tm->GradRAlg T : M •->• © ^ ( M ) = 0 r>0
{®p+q=rH"{M)
® H"(N)}
r>0
© r > 0 £(M) = © r > 0 e r ( M x N).
G:M^
) V/ : M -> M 2 $ ( / ) = 0 ( / x l w ) * U'(M2xiV)
V/ : Mi ^
M2.
r>0
We let the reader check the following elementary fact. Exercise 7.1.10 T and Q are contravariant MV-functors. For M € WlM define 4>M • T{M) -> Q(N) by def
<^M(W ® 7?) = W X TJ = 7T^o; A n*Nr] (w € H*(M),T] e
H*(M))
where 7rM (resp. TTN) are the canonical projections M xN —• M (resp. M x TV —» TV). The operation x:H*(M)®H*(N)^H*{MxN)
(w ® r?) >->• W X 77
is called the cross product. The Kiinneth formula is a consequence of the following lemma. Lemma 7.1.37 (a) cj) is a correspondence of MV-functors. (b) >]]jm is an isomorphism. Proof of the lemma The only nontrivial thing to prove is that for any MV-cover {U, V} of M G 93tm the diagram below is commutative. ®p+q=rH"(U
f)V)®
H"{N) - i - ® p + 9 = r i F + 1 ( M ) ® #«(/V)
^(/nv
Hr({U xN)n(Vx
<£M
N))
Hr+1{M
xN)
We briefly recall the construction of the connecting morphisms d and &'. One considers a partition of unity {tpu,
DeRham
229
cohomology
and ipv — TMVV form a partition of unity subordinated to the cover {U x N, V x N} of M x N. If w ® 7? € # * ( [ / n V)
where ^ \u= — d((pvw)
& = d(ipuuj).
On the other hand 0tmv(w® 77) = w x 77 and ^(w x 77) = w x 77. This proves (a). To establish (b) note that the inclusion j : TV M- E m x N
ii4(0,i)
is a homotopy equivalence with ir^ a homotopy inverse. Hence, by the homotopy invariance of the DeRham cohomology we deduce S(R m ) =* H*{N). Using the Poincare lemma and the above isomorphism we can identify the morphism <%» with % - . D Example 7.1.38 Consider the n-dimensional torus, Tn. By writing it as a direct product of n circles we deduce from Kunneth formula that Pm{t) = {Psl(t)}n
= (l + t)n.
Thus &*(Tn)=Q,
dimH*(Tn)
= 2n
and
xcn = 0. One can easily describe a basis of H*(Tn). Choose angular coordinates (01,... ,6n) on Tn. For each ordered multi-index I : 1 < i\ < • • • < j * < n we have a closed, nonexact form dO1. These monomials are linearly independent (over K) and there are 2" of them. Thus, they form a basis of H*{Tn). In fact, one can read the multiplicative structure using this basis. We have an isomorphism of K-algebras H*{Tn) S* A*R".
230
Cohomology
Exercise 7.1.11 Let M e 9Jtm and N € QJI„. Show that for any w* e rjj e H*(N) (i,j=0,l) the following equality holds. (wo x no) A (Wl x Vl) = {-l)de^de^(tj0
H*(M),
A W l ) x (% A tfc).
Exercise 7.1.12 (Leray-Hirsch)Let p : £ -> M be smooth bundle with standard fiber F. We assume the following: (a) Both M and F are of finite type. (b) There exists cohomology classes e j . , . . . , e r € H*(E) such that their restrictions to any fiber generate the cohomology algebra of that fiber. The projection p induces a H* (M)-modu\e structure on H*(E) by U-TI = P*LJAT1
VW € H*(M)r) e # * ( £ ) .
Show that H*(E) is a free i?*(M)-module with generators e i , . . . , e r .
7.2 §7.2.1
The Poincare duality Cohomology with compact supports
Let M be a smooth n-dimensional manifold. Denote by f2^,t(M) the space of smooth compactly supported &-forms. Then
0^fi° p t (M)4--.4fi; t (M)->0 is a cochain complex. Its cohomology is denoted by H*t(M) and is called the DeRham cohomology with compact supports. Note that when M is compact this cohomology coincides with the usual DeRham cohomology. Although it looks very similar to the usual DeRham cohomology, there are many important differences. The most visible one is that if <j> : M —> N and u> G fl*p((Af) is a smooth map then the pull-back <jfu> may not have compact support so this new construction is no longer a contravariant functor from the category of smooth manifolds and smooth maps to the category of graded vector spaces. On the other hand if dim M = dim N and cj> is an embedding we can identify M with an open subset of N and then any 77 e fi*p((M) can be extended with 0 outside M C N. This extension by zero defines a push-forward map >, •. n ^ ( M ) -»• n ; t ( A 0 . One can verify easily that >, is a cochain map so that it induces a morphism
4>.: #; t (M) -> #; ( (w). In terms of our category SDT„ we see that H*pt is a covariant functor from the category SDT„ to the category of graded real vector spaces. As we will see, it is a rather nice functor.
Poincare duality
231
Theorem 7.2.1 H^t is a covariant MV-functor and moreover
<((Kn) = { E0
, k < n , k= n
The last assertion of this theorem is usually called the Poincare lemma for compact supports. We first prove the Poincare lemma for compact supports. The crucial step is the following technical result we borrowed from [10]. L e m m a 7.2.2 Let E A B be a rank r real vector bundle, orientable in the sense described in Subsection 3.4.5. Denote by p„ the integration-along-fibers map
p. : ^t(E)
-> U£{B).
Then there exists a smooth bilinear map -> n ^ T " 1 ^ )
m : n%t(E) x Q^iE) such that p'p.a A 0 - a A p'p,/3 = {-l)rd{m(a,
Proof
/?)) - m{da, /?) + ( - l ) d e g a m ( a , dp).
Consider the 9-bundle n:£ •K
= I x(E®
E) ^ E
: (t; v0, vi) i-» (t- v0 + t(vi - v0)).
Note that d£ = ({0} x(E®
E)) U ({1} x {E © E)).
Define 7r*: E ® E -> E as the composition E © E 9* {<} x £ © E ^ I x (E © E) A £\ Observe that d7r = 7r|a £ = ( - T T ^ U T T 1 .
For (a,/?) € fi^(S) x fi^(£) define a © /? e 0 ^ ( 2 © £ ) by a © p = (7r°)*a A (vr1)*^.
232
Cohomology
(Verify that a 0 /? has indeed compact support). For a G Q'cpt(E) and /? G ^ p ( ( i ? ) we have the equalities A P = ^ ( a © /?) G n j + T r ( E )
p'p.a
aAp*p» / S = ^ ( a O / 3 ) G Q - + r r ( £ ) . Hence £>(a, /?) = p*p,a A /? - a A p*p,/? = irKa O /?) - TT°(Q O /?) = /
a©/?.
We now use the fibered Stokes formula to get D(a,(3)=[ JE/E
d£T*(aOP)
+ (-iydE
T is the natural projection £ = Ix(E®E)->E®E. {a, 13)= /
Je/E
T>0/3).
[
T{aQ0).
JS/E
The lemma holds with •
Proof of the Poincare lemma for compact supports such that 0< S<1 f 5(x)dx = 1.
Consider 5 G CQ°(K")
Define the (closed) compactly supported n-form r = 6(x)dx1 A • • • A dxn. We want to use Lemma 7.2.2 in which E is the rank n bundle over a point i.e. E = {pt} x R " A {pt}- The integration along fibers is simply the integration map. p . : J2^(R") -> K
{
0 , degw < n / K » ( j , degw = n
If now o> is a closed, compactly supported form on R n we have o> = (p*ptr) A u). Using Lemma 7.2.2 we deduce U - T A p*pmw = (—l) n dm(T, w). Thus any closed, compactly supported form u on ffin is cohomologous to r Ap*p,w. The latter is always zero if degw < n. When degw = n we deduce that w is cohomologous to (/ US)T. This completes the proof of the Poincare lemma. •
233
Poincare duality
To finish the proof of Theorem 7.2.1 we must construct a Mayer-Vietoris sequence. Let M be a smooth manifold decomposed as an union of two open sets M = UUV. The sequence of inclusions
unv^u,v^M induces a short sequence
o -> wcpt(u n v) -4 tr^u) © %t(v) 4 orcpt{M) -+ o where i(w) = (w,w) j{u,ri) =
fi-w.
The hat denotes the extension by zero outside the support. This sequence is called the Mayer-Vietoris short sequence for compact supports. Lemma 7.2.3 The above Mayer-Vietoris sequence is exact. Proof i is obviously injective. Clearly, Range (i) = ker(j). We have to prove that j is surjective. Let (
VUT, e n*cpt(u)
oyy).
In particular we have
n = 3{-{pvn,'pvn)Hence j is surjective.
•
We get a long exact sequence called the long Mayer-Vietoris sequence for compact supports. • • • - » • Hkcpt(U HV)^
Hkcpt(U) © Hkcpt{V)
- 4 Hkcpt(M)
4 Hkpf
-> • • •
The connecting homomorphism can be explicitly described as follows. If UJ 6 Q,k t{M) is a closed form then d((puL)) = d(—ipvu) on U Pi V. We set Sui = d(ipuu). The reader can check immediately that the cohomology class of <5w is independent of all the choices made. If <j> : N <-» M is a morphism of 9Jt„ then for any MV-cover of {U, V} of M {<^_1([/),<)!i-1(V)} is an MV-cover of N. Moreover, we almost tautologically get a commutative diagram
H^{M)-^H^{UnV) This proves H*pt is a covariant Mayer-Vietoris sequence. •
234
Cohomology
Remark 7.2.4 To be perfectly honest (from a categorial point of view) we should have considered the chain complex
y3 Q = (J) fl^ fc<0 fc<0
and correspondingly the associated homology H
* =
H
^t-
This makes the sequence 0 <- Q.~n i
<- £l~l <- Cl° <- 0
a chain complex and its homology H* is a bona-fide covariant Mayer-Vietoris functor since the connecting morphism 5 goes in the right direction H* —¥ H*~x. However, the simplicity of the original notation is worth the small formal ambiguity so we stick to our upper indices. From the proof of the Mayer-Vietoris principle we deduce the following. Corollary 7.2.5 For any M e 9Jtn and any k < n we have dim H^,t(M) §7.2.2
< oo.
The Poincare duality
Definition 7.2.6 Denote by 9Dt+ the category of n-dimensional, finite type, oriented manifolds. The morphisms are the embeddings of such manifold. The MV functors on 971^" are defined exactly as for 9Jl„. Given M £ 9Jt+ there is a natural pairing ( , )K : Q.k{M) x Q.nc;tk{M) -> R defined by (LJ,rj)K=
/
uAi|.
JM
which is called the Kronecker pairing . We can extend this pairing to any (w, rj) € Q* x n^ as 0 , deg u + deg rj / n \u,T))K | / M W A T ? , degw + degr/ = n This pairing induces maps V = Vk : Qk(M) -> {QP-tk{M)Y (V(u),T])
= <w,r/)K.
235
Poincare duality
( , ) denotes the natural duality between a vector space and its dual, V* x V —> R. We continue to denote by T>(OJ) the restriction to the space of closed, compactly supported (n — /e)-forms, Z^k. If moreover w is closed, this functional on Z^k k vanishes on the subspace B™~t of exact, compactly supported (n — fc)-forms. Indeed, if 7? = dr/, i 6 n ^ * _ 1 ( M ) then (X>(w), TJ> = I uAdn'
St es
=
JM
± f dwAV'
= Q.
JM
Thus, if w is closed V(u>) defines an element of (H^tk{M))*. If moreover u> is exact, a computation as above shows that V(u>) = 0 € {H^~tk{M))*. Hence V descends to a map in cohomology V : Hk(M) -> (H£k{M))* which is the same as saying that the Kronecker pairing descends to a pairing in cohomology. Theorem 7.2.7 (Poincare duality)The Kronecker pairing in cohomology is a duality for all M € 2rt+. Proof
The functor 971+ —>• Graded Vector Spaces
defined by
M->©ff*(M)=®(fl£*(M))' k
k
is a contravariant MV-functor. (The exactness of the Mayer-Vietoris sequence is preserved by transposition. This is where the fact that all the cohomology groups are finite dimensional vector spaces plays a very important role). For purely formal reasons (which will become apparent in a little while) we define the connecting morphism of the functor Hk (Hk(U D V) -4 Hk+l(U U V)) to be (-l) f c 5 ( , where H^tk~l{UnV) A H^k(UUV) denotes the connecting morphism in the DeRham cohomology with compact supports and <$f denotes its transpose. The Poincare lemma for compact supports can be rephrased
The Kronecker pairing induces linear maps VM : Hk{M) ->
Hk{M).
Lemma 7.2.8 Qk Vk is a correspondence of MV functors.
236
Cohomology
Proof We have to check two facts. FACT A. Let M <-t N be a morphism in SDT+. Then the diagram below is commutative. Hk(N) - ^ Hk(M)
Hk{N)
-*--
Hk{M)
F A C T B . If {U, V} is a i W - c o v e r of M e 97t+ then the diagram bellow is commutative Hk(UHV) -^ Hk+1(M) VM
T>unv
ff'([/ny)l-^ffw(M) Proof of FACT A. Let w € Hk(N). For any rj e H^k{M) natural duality between a vector space and its dual) (if °VN{u),rj) =
/
= (^tyVN{u),ri)
w A <j>*r] =
JN
/
((•,•) denotes the
= (VN{u>),
w \M AT? =
JM<-+N
/
JM
= (VM(
Choose rj e
fi^fc_1(M)
_ f d(-ifivuj) \ d(tpuu)
on U on V
such t h a t ofr? = 0. We have (Viudu), rj) = / Out Arj JM
=
6ujArj+dujArj
—
9UAIJ
7(7 Jv ./m-iv = — / dfojyo;) A n + / dipu<J) Ar) + dUpvw) A rj. Ju Jv Junv Note t h a t t h e first two integrals vanish. Indeed, over U we have the equality (d(ipvw) A rj = d (tpvv A rj)
237
Poincare duality
and the vanishing now follows from Stokes formula. The second term is dealt with in a similar fashion. As for the last term we have / (dipvuj) A ri = f d(pv A UJ A r] = {-l)desu f Junv Junv Junv = (-l)k
[
Junv
(-l)k(VunVLj,
UJA5V=
w A (d
SV)
{{-l)HtVmVU,Tl).
=
This concludes the proof of Fact B. The Poincare duality now follows from the Mayer-Vietoris principle. •
Remark 7.2.9 Using the Poincare duality we can associate to any smooth map / : M —> N between compact oriented manifolds of dimensions m and respectively n a natural push-forward morphism / . : H*{M) -> H*+q{N)
(q = dim TV - d i m M = n - m)
defined by the composition H*(M)
V
-H {Hm-(M)Y
{f
-7 (Hm-(N)y
V
A
H*+n~m{N)
where (/*)' denotes the transpose of the pullback morphism. Corollary 7.2.10 If M e SDT+ then Hkcpt{M) = Proof
Since Hkpt(M)
{Hn~k(M)Y.
is finite dimensional the transpose ^M •• (HHk(M)Y*
-v
(Hk(M)Y
is an isomorphism. On the other hand, for any finite dimensional vector space there exists a natural isomorphism V" =* v. a Corollary 7.2.11 Let M be a compact oriented n-dimensional manifold. Then the pairing Hk(M)
x Hn-k(M)
->• K (to, r)) .-> J w A 77 JM
is a duality. In particular bk(M) — bn-k(M),
\fk.
238
Cohomology
If M is connected H°(M) ^ Hn(M) ^ R so that Hn{M) is generated by any volume form defining the orientation. The symmetry of Betti numbers can be translated in the language of Poincare polynomials as tnPM(\)
= PM(t).
(7.2.1)
Example 7.2.12 Let E 9 denote the connected sum of g tori. We have shown that X (E 9 )
= b0 - h + b2 = 2 - 2g.
Since E s is connected, the Poincare duality implies b2 = b0 = 1. Hence b\ = 2g i.e. PSg{t) = l + 2gt + t2. Consider now a compact oriented smooth manifold such that dim M = 2k. Kronecker pairing induces a non-degenerate bilinear form 3 : Hk(M)
x Hk(M)
-*• M
3(LJ,
The
rf) = / u A t). JM
3 is called the cohomological intersection form of M. When k is even (so that n is divisible by 4) 3 is a symmetric form. Its signature is called the signature of M and is denoted by cr(M). When k is odd 3 is skew-symmetric (i.e. it is a symplectic form). In particular, we deduce the following result. Corollary 7.2.13 For any compact manifold M £ ^Xfltk+2 bik+x(M) is even.
tne
middle Betti number
Exercise 7.2.1 (a) Let P 6 Z[t] be an odd degree polynomial with non-negative integer coefficients such that P(0) = 1. Show that if P satisfies the symmetry condition (7.2.1) there exists a compact, connected, oriented manifold M such that Pu(t) = P(t). (b) Let P e Z[i] be a polynomial of degree 2k with non-negative integer coefficients. Assume P(0) = 1 and P satisfies (7.2.1). If the coefficient of tk is even then there exists a compact connected manifold M € StTtJt s u c n * n a t P~M(t) = P{t)Hint: Describe the Poincare polynomial of a connect sum in terms of the polynomials of its constituents. Combine this fact with the Kiinneth formula. Remark 7.2.14 The result in the above exercise is sharp. Using his intersection theorem F. Hirzebruch showed that there exist no smooth manifolds M of dimension 12 or 20 with Poincare polynomials 1 + t6 + t12 and respectively l+t10+t20. Note that in each of these cases both middle Betti numbers are odd. For details we refer to J. P. Serre, "Travaux de Hirzebruch sur la topologie des varietes", Seminaire Bourbaki 1953/54,n° 88.
Intersection
239
theory
Figure 7.1: A cobordism in R3
7.3 §7.3.1
Intersection t h e o r y Cycles a n d t h e i r duals
Definition 7.3.1 Let M G SDT+. A fc-dimensional cycle in M is a pair (S,
(a2)*ta=&, i = 0,l. (b) A cycle (S, >) € Cfc(M) is said to be d e g e n e r a t e if <j> is homotopic to the constant map S -> M. We write (S, 0) ~ c 0. We denote by ©^(M) the set of degenerate cycles. Exercise 7.3.1 Let (S 0 , <j>0) ~ c {Si, 4>\). Prove that ( - S 0 U Si, fa LI <£i) ~ c 0. We denote by 3 * (M) the free abelian group generated by Cjt(M) and by 2$k(M) C 3*(-W) the subgroup generated by cycles cobordant to degenerated ones. We can form the quotient group Sjk{M). For any cycle {S,<j>) G £/t(M) we denote by [S,<j>] its image in Sj^{M).
240
Cohomology
Figure 7.2: The dual of a point is Dirac's
distribution
Exercise 7.3.2 (a) Prove that [So, 4>o] + [Si, 4>x] = [So uSufau
4>i] V(5i,
gCk(M)
and -[S,4>] =
[-S,4>}€Sik(M).
(b) Prove that[5, 4>] = 0 in fik(M) if and only if there exists a degenerate cycle (S'fi) such that {S,
Vs
Stokes formula shows that this map is well defined i.e. it is independent of the closed form representing a cohomology class. Indeed, if w is exact, i.e. u> = du' then
f 'dJ = / # V = 0.
Js
Js
In other words, each cycle defines an element in (Hk(M))* which can be identified via the Poincare duality with H™~tk(M). Thus there exists 5S G H™~tk(M) such that [ LJA5S= JM
[
Hk{M).
JS
5s is called the Poincare dual of (S, >). There exist many closed forms r) e QJl~k{M) representing 5s- When there is no risk of confusion, we continue denote any such representative by 6S.
Intersection
241
theory
Example 7.3.3 let M = K" and S is a point {pt} C Mn. pt is canonically a 0-cycle. Its Poincare dual is a compactly supported n-form w such that for any constant A (i.e. closed 0-form)
f Xu = [ A = A i.e.
1.
Ar'
Thus (5P( can be represented by any compactly supported n-form with integral 1. In particular we can choose representatives with arbitrarily small supports. Their "profiles" look like in Figure 7.2. "At limit" they approach Dirac's delta distribution. Example 7.3.4 Consider an n-dimensional, compact, oriented manifold M. denote by [M] the cycle (M, 1 M ). Then <5[M] = l e H°{M).
We
For any differential form UJ we set (for typographical reasons) |w| = deg a;. Example 7.3.5 Consider the compact manifolds M 6 SDt+ and TV £ 93T+. To any cycle (S,4>) £ CP(M) and (T,tp) € Cq{N) we can associate the cycle (SxT,
W
x <5T
(7.3.1)
where uxvd=f
TT*MU A Tr^n
V(w, n) e 0*(M)
x Q*(TV).
By Kiinneth formula we have H*{M x N) = H*(M) ® #*(TV) and the cohomology of M x TV is generated (as a linear space) by the cross products w x r\. Pick (w, r/) e ft* (M) x ft* (TV) such that deg u> + deg n = (m + n) - (p + 9). Then, using Exercise 7.1.11 /
(«,)Afex{
T
) = (-lf-*l/
JMxN
(ujA5s)x(riA5T). JMxN
The above integral should be understood in the generalized sense of Kronecker pairing. The only time when this pairing does not vanish is when |w| = p and |n| = q. In this case the last term equals
(~iy^(JsuA6s)
(/ r ,A^) = ( - i ) ^ ( / s f U )
= (_!)?(m-p) J JSxT
This establishes equality (7.3.1).
(^
X
^ V o ; A „).
(fTrv
242
Cohomology
Example 7.3.6 Consider a compact manifold M G SDTj. Fix a basis (a;*) of H*(M) such that each a;* is homogeneous of degree \tVi\ = dt and denote by u>1 the basis of H*(M) dual to (a;,) with respect to the Kronecker pairing i.e. < W > J ) B = (-1)I U 'I-" U 'I< W J V>« =
4
I n M x M there exists a remarkable cycle, the diagonal A : M —• M x M
IH(I,I).
We claim that the Poincare dual of this cycle is <5A = *M d= E l - 1 ) ' " " 1 ^
X
"i-
(7-3.2)
i
Indeed, for any homogeneous forms a,/? 6 fl*(M) such that |a| + |/?| = n we have (w x r?) A D M = E l - 1 ) ' " ' 1 /
/ JMxM
= E(-l)|lJ,|(-l)l/JHui| /
J MxM
i
=
(a x /?) A (a/ x
Wi )
JMxM
i
(a A u>{) x (/? A «,-)
B-l)W(-l)WVl(/M«A^)(/M/JAWi)
The i-th summand is nontrivial only when |/?| = jcj11 and |or| = |w,-|. Using the equality |w'| + |<J'| 2 = 0(mod 2) we deduce /
(ax0)ABM = E ( 7
JMxM
aAf/)(7
~? \JM Q CJI
= E( ' i
/?Aw 4 )
I \JM
J
? W
>«(/ ' 0K-
From the equalities
we conclude
/ A>x/3)= / aA/J=/ ( E ^ ( ^ , a ) . ) A ( E ( A ^ ) » = / E( W '' a )«(^' a ; j)'' < J < A '^
=
E(W,'a)«(^'a'j)"(W»wJ}''
= ^(-l)KI(»-MI)(a, W i )-(-l) | w , | ( "- | u i l ) <)8,w i >« =
Equality (7.3.2) is proved.
E(Q'W')K(/?'W0K-
Intersection
243
theory
Proposition 7.3.7 Let M 6 tXfl* and (S i ; fa) G €k(M) (a) If (So, <M ~c (Si,
(i = 0,1) - two cycles in M.
Proof (a) Consider a compact manifold E with boundary 9E = So U Si and a smooth map $ : E —> M such that $ \aiz=fa,U fa. For any closed fc-form w € fifc(M) we have 0 = / $*(dw) = / d$*w StO*l 'ofces
/
=
/
I
A*
I
A*
(J = / fau - / 0„W.
(b) is left to the reader, (c) is obvious. To prove (d) consider E = [0,1] x S 0 and $ : [0,1] x S 0 -> M, $(t, x) = 0 o (x) V(t, x) e E. Note that 5E = (-S 0 ) U S 0 so that 5-s 0 + SSo = 5_Sous0 =
fl£*(M).
This is usually called the homological Poincare duality. §7.3.2
Intersection theory
Consider M 6 fXJl* and S a fc-dimensional compact oriented submanifold of M. We denote by i : S <-¥ M inclusion map so that (S, i) is a fc-cycle. Definition 7.3.8 A smooth map <j> : T -» M from an {n — fc)-dimensional, oriented manifold T is said to be transversal to S (and we write this S ffl>) if (a) <j>~1(S) is a finite subset of T; (b) for every x G
(direct sum).
In this case, for each x € 0 - 1 ( S ) we define the local intersection number at x to be (or= orientation) x{
°'
>
f 1 , or {T^x)S) A or{faTxT) = or(T 0 ( x ) M) \ - 1 , o r ( T , ( l ) S ) A o r ( ^ T x T ) = -or(T« x ) Af)
Finally, we define the intersection n u m b e r of S with T to be S-T=
Y,
i*(S,T).
244
Cohomology
Figure 7.3: The intersection number of the two cycles on T2 is 1 Our next result offers a different description of the intersection number indicating how one can drop the transversality assumption from the original definition. Proposition 7.3.9 Let M G 971+. Consider S <-> M a compact, oriented, kdimensional submanifold and (T,
(7.3.3)
JM
where 5. denotes the Poincare dual of •. The proof of the proposition relies on a couple of technical lemmata of independent interest. Lemma 7.3.10 (Localization lemma) Let M € 971+ and (S,
{?Aw=
4>*LO.
Intersection
theory
245
Hence, Sg represents the Poincare dual of S in H"ptk(M)
and moreover, supp 5$ C Af.
a Definition 7.3.11 Let M e 9DT+ and S <-» M a compact, A:-dimensional, oriented submanifold of M. A local transversal at p 6 S is an embedding 4>:Bc
Rn~k -» M
(B = open ball centered at 0)
such that 5 rfi0 and 0 _ 1 (S) = {0}. Lemma 7.3.12 Let M e 90?^, S <-> M a compact, fc-dimensional, oriented submanifold of M and (B, 4>) a local transversal at p € S. Then for any sufficiently "thin" closed neighborhood TV of S C M we have S •(£>*) = / / # • Proof Using the transversality S (\\(j>, the implicit function theorem and eventually restricting <> / to a smaller ball, we deduce that (for some sufficiently "thin" neighborhood TV of 5) there exist local coordinates (a; 1 ,..., xn) defined on some neighborhood U of p G M diffeomorphic with the cube {|i*'| < 1, Vt} such that (i)5 n U = {xk+1 = -.- = xn = 0},p=(0,...,0) (ii) The orientation of S n U is defined by dxl A • • • A dxk. (iii) The map >: B C K?"* „_M —>• M is expressed in these coordinates as x 1 = 0 , . . . ,xk = O , ^ 1 = y\ ... ,xn =
yn-k.
{rv)Mr\U = {\xi\ < 1/2; j = l , . . . , n } . Let e = ± 1 such that edx1 A . . . A dxn defines the orientation of TM. words
e= 1
S-(B,
k
For each £ = (re ,..., x ) e S n £/ denote by P ? the (n - fc)-"plane" P€ = { ( e ; i * + 1 , . . . , i " ) ; | ^ ' | < i
j>k}.
We orient each P^ using the (n - fe}-form dxk+1 A • • • A ds" and set
"(0 = I 4-
In other
246
Cohomology
Equivalently
v(t) = IBfBWS where ^ : B —• M is defined by
My\---,yn-k)
(Z;y\---,yn-k)-
=
To any function
i
we associate the fc-form uv = tpdx1 A---Adxk
= cpd£ € Sl%t(S n f/).
Extend the functions a; 1 ,... ,xk G C°°(Ur\/f) to smooth compactly supported functions x{ e C^{M) ->• [0,1] The form ui^ is then the restriction to U n S of the closed compactly supported form u)v = ipix1,..., We have
xk)dxl
A • • • A dxk.
/>A ^=L u * A ^ = h u * = J * ^-
(?-3-4)
The integral over U can be evaluated using the Fubini theorem. Write 6% = fdxk+1
A---Adxn
+Q
where Q is an (n —fc)-formnot containing the monomial dxk+1 A • • • A dxn. Then J uov A 6$ = f f(pdxl A • • • A dxn = e / fip\dxx A • • • A dxn\ F
=nie[
Jsnu
(\dxl A • • • A dxn\ = Lebesgue density)
y»(0 f/"
\JPS
f\dxk^A---Adxn\)m J
= e Js
Intersection
theory
247
If in the above equality we use the assumption (iii) we recognize in the left-hand-side the integral JB
The local transversal lemma is proved. • Proof of Proposition 7.3.9
Let (j>-l(S) = { p i , . . . , p m } .
The transversality assumption implies that each pt has an open neighborhood B, diffeomorphic to an open ball such that
S-T = £ S •(£;,&)• i
The compact set K =
does not intersect S so that we can find a "thin", closed neighborhood" N of S '-* M such that i f n A r = 0. Then (f)*5-g is compactly supported in U-Dj
From the local transversal lemma we get j D tfSg = (-1)*<»-*>S • (Bit4>) =
iPi(S,T).
The Proposition is proved. • Equality (7.3.3) has a remarkable feature. Its right-hand-side is an integer but is defined only for cycles S, T such that 5 is embedded and S fflT, while the left-handside makes sense for any cycles of complementary dimensions but a priori it may not be an integer. In any event, we have a remarkable consequence. Corollary 7.3.13 Let ( $ , & ) e £k(M) i = 0,1. If
and ( ( T ^ i ) € C„_fc(M) where M e Ort+,
(a) So ~ c S i , T 0 ~ c T i ,
(b) the cycles Si are embedded and (c) Si !\\TZ
Then So • T0 = S\ • Ti.
248
Cohomology
Definition 7.3.14 The homological intersection pairing is the Z-bilinear map 3 = ?M • &k(M) x Sjn.k(M)
-> M
(M e 9Jt+) defined by 0r(5,T)= /
5SA5T.
J M
We have proved that in some special instances 3(S,T) that when M is compact this is always the case.
6 Z. We want to prove
Theorem 7.3.15 Let M e 2K+ be compact manifold. Then for any (S,T) S)k{M) x Sjn-k(M) the intersection number 3{S,T) is always an integer.
e
The theorem will follow from two lemmata. The first one will show that it suffices to consider only the situation when one of the two cycles is embedded. The second one will show that if one of the cycles is embedded then the second cycle can be deformed so that it intersects the former transversally. (This is called a general position result). Lemma 7.3.16 (Diagonal trick)Let M, S and T as in Theorem 7.3.15. Then 3(S,T)
= (-1)""* /
6SxTA5A
JMxM
where A is the diagonal cycle A : M —>• M x M, x i-¥ (x,x). (It is here where the compactness of M is essential, since otherwise A would not be a cycle). Proof
We will use the equality 7.3.1 SsxT = ( - 1 ) " - Ss x ST-
Then (-1)"-* = /
5SxT A SA = /
JMXM
(Ss x ST) A 6A JMXM
= [ A*(5 S x 6T) = / 8s A ST. JM
JM
•
Lemma 7.3.17 (Moving lemma) Let S € C*(M) and T €
Intersection §7.3.3
249
theory
The topological degree
Consider two compact, connected, oriented smooth manifolds M, N having the same dimension n. Any smooth map F : M —• TV canonically defines an n-dimensional cycle i n M x J V IV : M -> M x N x^{x,F(x)). Tp is called the graph of F. Any point y G N defines an n-dimensional cycle M x {y}. Since N is connected all these cycles are cobordant so that the integer Tp • {M x {y}) is independent of y. Definition 7.3.18 The topological degree of the map F is defined by
degF^(Mx{y})-rF. Note that the intersections of TF with M x {y} correspond to the solutions of the equation F(x) = y. Thus the topological degree counts these solutions (with sign). Proposition 7.3.19 Let F : M ->• N as above. Then for any n-form w G Q.n(N)
f F*u) = degF [ u). JM
JN
Remark 7.3.20 The map F induces a morphism U = Hn(N)
^Hn{M)^R
which can be identified with a real number. The above proposition guarantees this number is an integer. Note that if w G Q.n(N) is exact then
Proof of the proposition
/ u> = / F*u = 0.
JN
JM
Thus, to prove the proposition it suffices to check it for any particular form which generates Hn(N). Our candidate will be the Poincare dual 5y of a point y G TV. We have while equality (7.3.1) gives &Mx{y] = SM X Sy = 1 X Sy. We can then compute the degree of F using Theorem 7.3.15 degF = d e g F f 5y = [ JN
(1 x <$„) A &rF
JMxN
"
= I VF(l x 6y) - f F*6y. D JM
JM
250
Cohomology
Corollary 7.3.21 ( G a u s s - B o n n e t ) Consider a connected sum of g-tori E = E , embedded in K3 and let © s : E -> S 2 be its Gauss map. Then deg©s=x(S) = 2-2P. This corollary follows immediately from the considerations at the end of Subsection 4.2.4. Exercise 7.3.3 Consider the compact, connected manifolds M0,Mi,N € SDT^ and the smooth maps Ft : Mi -> N, i = 0,1. Show that if F0 is cobordant to Fx then deg F0 = deg Fi. In particular, homotopic maps have the same degree. Exercise 7.3.4 Let A be a nonsingular n x n real matrix. It defines a smooth map
Prove that deg FA = sign det A. H i n t : Use the polar decomposition A = PO (where P is a positive symmetric matrix and O is an orthogonal one) to deform A inside GL(n, tt) to a diagonal matrix. Exercise 7.3.5 Let M —» N be a smooth map (M, TV are smooth, compact oriented of dimension n). Assume y €. N is a. regular value of F i.e. for all x G F~l(y) the derivative DXF : TXM -+ TVN is invertible. For x £ F~1(y) define deg(F,x)
=
1 , DXF preserves orientations — 1 , otherwise
Prove that degF=
Y,
deg(F,x).
F(x)=y
Exercise 7.3.6 Let M denote a compact oriented manifold and consider a smooth map
F:M^M. Regard H*(M) as a superspace with the obvious Z2-grading H*(M) = Heven(M)
© H°dd(M)
and define the Lefschetz number X(F) of F as the supertrace of the pull back F* : H*{M) -+ H*{M). Prove that X(F) = A • r > . and deduce from this the Lefschetz fixed point theorem: X(F) ^ 0 => F has a fixed point.
Intersection §7.3.4
251
theory
Thorn isomorphism theorem
Let p : E —> B be an orientable fiber bundle with standard fiber F and compact, oriented basis B. Let d i m B = m and d i m F = r. In this subsection we will extensively use the techniques of fibered calculus described in Subsection 3.4.5. The integration along fibers
/
=
p,:Wcpt{E)^W{B)
J E/B
satisfies p,dE
{-\)rdBp*
=
so that it induces a map in cohomology P, : H*cpt{E) -+
H*~r(B).
This induced map in cohomology is sometimes called the Gysin map. Exercise 7.3.7 Consider a smooth map / : M -¥ N between compact, oriented manifolds M, N of dimensions m and respectively n. Denote by if the embedding of M in M x N as the graph of / M 3 x i-> (x, f(x)) g M x N. The natural projection M x N -+ N allows us to regard M x N as a trivial fiber bundle over N. Show that the push-forward map / . : H*(M) ->• H*+n~m(N) defined in Remark 7.2.9 can be equivalently defined by / , = 7T» O
(if)t
where 7r, denotes the integration along fibers while (i/). : H*cpt(M) -> H*cpt(N) is the natural morphism defined by the embedding if. Let us return to the fiber bundle p : E —• B. Any smooth section o:B-+
E
defines an embedded cycle in E of dimension m = dim B. Denote by 5a its Poincare dual in Hlpt(E). Using the properties of the integration along fibers we deduce that for any u) e Q,m(B) we have
/ <5CTAp*w= f ( f
JE
JB \JE/B
5a)to. J
252
Cohomology
On the other hand by Poincare duality we get
[ 6aAp*Lj = (-l)rm
JE
fp'wAi
JE
= ( - i r / era; = (-1)™ f {P°T" = (-i)rm / <. JB
JB
JB
Hence
'A = I
Sa = ("I)™ € fi°(B).
JE/B
Proposition 7.3.22 Let p : E —> B a bundle as above. If it admits at least one section then the Gysin map p. : H*cpt(E) ->
H*-r(B)
is surjective. Proof
Denote by Ta the map r„ : H'(B)
->
fl^CE)
w H-> (-l) r m 5 C T A p*w = p*w A <J„.
Then ra is a right inverse for p„. Indeed u
= ( - l ) ™ p , { „ A u = (-l) r m p«(5 f f A p ' u ) =p.(T ff w).n
The map p* is not injective in general. If for example (5, <j>) is a fc-cycle in F then it defines a cycle in any fiber 7T_1(&) an consequently in E. Denote by 5s its Poincare dual in H™tr-k{E). Then for any w € fim~*(B) we have / (p.fe) A CJ = / <SS A p*u> = ± f fp'uj VD
= [ (p o 0)*<j = 0 JS
since p o ^ is constant. Hence pt5s = 0. Hence if F carries nontrivial cycles kerp, may not be trivial. The simplest example of standard fiber with only trivial cycles is a vector space. Definition 7.3.23 Let p : E —> B be an orientable vector bundle over the compact oriented manifold B (dimB = m, rank(£') = r). The Thorn class of E, denoted by TE is the Poincare dual of the cycle defined by the zero section Co : B —> E, b^0€Eb. Note that TB € Hrcvt{E).
Intersection
253
theory
Theorem 7.3.24 (Thom isomorphism)Let p : E -» B as in the above definition. Then the map T : H*(B)
-» tf^YX-E) W H T J A p*w
is an isomorphism called the T h o m isomorphism. Its inverse is the Gysin map (-l)""p. : H*cpt(E) -»• Proof
ff-'(S).
We have already established that
P.T =
Mr
To prove the reverse equality rp, = (—l) rm we will use Lemma 7.2.2 of Subsection 7.2.1. For p e Sl*q,t(E) we have (p*P,r B ) A 0 - TE A (p*pt(3) = ( - l ) r d ( m ( r E , P)) where m(rE,P)
G Q^, ( (£). Since p*p,rE = (—l) rm we deduce (-1)""/? = (TE Ap*(p,/?) + exact form
i.e. ( - l H = TEop,(^)
inff;t(£).Q
Exercise 7.3.8 Show that rE = C.l where C* : ff*(M) -> H'c+tAXmM{E) is the pushforward map defined by a section £ : M —> E. §7.3.5
Gauss-Bonnet revisited
We now examine a very special type of vector bundle: the tangent bundle of a compact, oriented, smooth manifold M. Note first the following fact. Exercise 7.3.9 Prove that M is orientable if and only if TM is orientable as a bundle. Definition 7.3.25 Let £ - > M b e a real orientable vector bundle over the compact, oriented n-dimensional smooth manifold. Denote by TE € H*pt(E) the Thom class of E. The Euler class of E is defined by e(E) = QTE G Hn(M) where ( 0 : M —> E denotes the zero section. e(TM) and is denoted by e(M).
is called the Euler class of M
254
Cohomology
Note that the sections of TM are precisely the vector fields on M. Moreover, any such section a : M —> TM tautologically defines an n-dimensional cycle in TM and in fact any two such cycles are homotopic - try an affine homotopy along the fibers of TM. Any two sections 0o,(?i : M —>• TM determine cycles of complementary dimension and thus the intersection number a0 • ax is a well defined integer, independent of the two sections. It is a number reflecting the topological structure of the manifold. Proposition 7.3.26 Let aQ, <7i : M -» TM be two sections of TM. Then / e(M) =<70 -o\JM
In particular, if dim M is odd then / e(M) = 0. JM
Proof The section
= [
JTM
I Sa JTM
A 5a
rM A SCo = f
JM
=
is the
/ TM A TM JTM
QTM
= f e(M). JM
If dim M is odd then
J e(M) = <70 • ox = -<7i • 0O = - J e(M). D JM
JM
Theorem 7.3.27 Let M be a compact oriented n-dimensional manifold and denote by e(M) its Euler class. Then / e(M)=x(M) = ^ ( - l ) ^ ( M ) .
In the proof we will use an equivalent description of x{M). Lemma 7.3.28 Denote by A the diagonal cycle in M x M. Then X(M)
= A • A.
Proof of the lemma Consider a basis (WJ) of H*{M) consisting of homogeneous elements. We denote by (a/) the dual basis i.e. (uji,ojj)K
=
5ij.
Intersection
255
theory
According to (7.3.2) we have
<5A = £ ( - l ) | w V
x
<*•
i
Similarly, if we start first with the basis (wl) then its dual basis is (_l)kl-l^l w . so we also have (taking into account that \u>j\ + |OJJ| • |w,| = n\u>j\ (mod 2))
5A = 'Ei-iyMuj
x J.
i
Using Exercise 7.1.11 we deduce
A A
• =/
(EC-1)1"'1^ X W *)
A
(E(-1)n|w,lwJ x ^ I
( E ( - l ) | w i | ( - l ) n • k K - 1 ) 1 ^ 1 ^ 1 ^ A W j j x (Wi x UJJ)
= /
= E ( - i ) k l ( - i ) " • KK-i^-'-^'k,^). (Wi)^)K. In the last expression we now use the duality equations (^ V)K
= (-l)^H^'
and the congruence \J\ + n\u)i\ + \tJi\2 + \wi\ • H^l = | ^ | + n\ui\ + \ui\2 + \ui\(n - |WJ|) H |U/| (mod 2) to conclude that A.A = £(-1)I^=X(M).
D
(J 1
Proof of theorem 7.3.27 The tangent bundle o f M x M restricts to the diagonal A as a rank 2n vector bundle. If we choose a Riemann metric on M x M then we get an orthogonal splitting T{M x M ) | A = 7 V A e T A . The diagonal map M -+ M x M identifies M with A so that TA = TM. have the following remarkable result. Lemma 7.3.29 NA =* T M .
We now
256 Proof
Cohomology Use the isomorphisms T(M x M) | A ^ TA © NA ^ TM © NA
and T ( M x M) | A = T M © TM.
•
From this lemma we immediately deduce the equality of Thorn classes TNA = TM.
(7.3.5)
At this point we want to invoke a technical result whose proof is left to the reader as an exercise in Riemann geometry. Lemma 7.3.30 Denote by exp the exponential map of a Riemann metric g on M x M. Regard A as a submanifold in N& via the embedding given by the zero section. Then there exists an open neighborhood TV of A C TVA C T(M X M) such that exp |AT: TV -> M x M is an embedding. Let A/" be a neighborhood of A C TVA as in the above lemma and set 9 t = exp(TV). We can view A as a submanifold of both 9 t and TV. Denote by <5/ (resp. <5^f) the Poincare dual of A in 91 (resp. TV). SA is the Thorn class of N& —> A which in view of (7.3.5) means =
<^A
T
NA
= TA =
TM.
We get A •A =
/
sf-
JA
= /
=
/
6% = f C0VA JA
JA
QrM = /
e(M).
IM
Hence / e(M)=X(M).
D
If n is a connected sum of g tori then we can rephrase the Gauss-Bonnet theorem as follows. Corollary 7.3.31 For any Riemann metric ft on a connected sum of -tori E 9 we have —e{h) = —shdvh Z7T
47T
= e(E 9 ) in ff*(£9).
Intersection
257
theory
The remarkable feature of the Gauss-Bonnet theorem is that once we choose a metric we can explicitly describe a representative of the Euler class in terms of the Riemann curvature. The same is true for any compact oriented even dimensional Riemann manifold. In this generality the result is known as Gauss-Bonnet-Chern and we will have more to say about it in the next chapter. We now have a new interpretation of the Euler characteristic of a compact oriented manifold M. Given a smooth vector field X on M, its "graph" in TM
Tx = {(x,X(x))eTxM;
xeM}
is an n-dimensionat submanifold of TM. The Euler characteristic is then the intersection number X (M) = Tx • M where we regard M as a submanifold in TM via the embedding given by the zero section. In other words, the Euler characteristic counts (with sign) the zeroes of the vector fields on M. For example if x{M) ^ 0 this means that any vector field on M must have a zero ! We have thus proved the following result. Corollary 7.3.32 If x(Af) / 0 then the tangent bundle TM is nontrivial. The equality x{S2n) = 2 is particularly relevant in the vector field problem discussed in Subsection 2.1.4. Using the notations of that subsection we can write v{S2n) = 0. We have thus solved "half" the vector field problem. Exercise 7.3.10 Let X be a vector field over the compact oriented manifold M. A point i G M i s said to be a non-degenerate zero of X if X(0) = 0 and
det
(§K^ a
for some local coordinates (xl) near XQ such that the orientation of T*0M is given by dx1 A • • • A dx". Prove that the local intersection number of Tx with M at x0 is given by *x 0 (rx,M) =sign det
dX£ dxi '
h=x
°
(This is sometimes called the local index of X at x0 and is denoted by i(X, x0). From the above exercise we deduce the following celebrated result.
258
Cohomology
Corollary 7.3.33 (Poincare-Hopf)If X is a vector field along a compact, oriented manifold M with only non-degenerated zeros xi,...,Xk then X(M)
1
=
£i(X,xj).
3
Exercise 7.3.11 Let X be a vector field on M" and having a non-degenerate zero at the origin. (a) prove that for all r > 0 sufficiently small X has no zeros on Sr — {\x\ = r}. (b) Consider Fr : ST -> S"" 1 defined by
Prove that i(X, 0) = degF r for all r > 0 sufficiently small. Hint: Deform X to a linear vector field.
7.4
S y m m e t r y and topology
The symmetry properties of a manifold have a great impact on its global (topological) structure. We devote this section to presenting some aspects of this avenue. §7.4.1
Symmetric spaces
Definition 7.4.1 A homogeneous space is a smooth manifold M acted transitively by a Lie group G called the symmetry group. Recall that a left action G x M —>• M
(g, m) >-> g • m
is called transitive if for any m £ M the map
tym:G3gi-^g-meM is surjective. For any point a; of a homogeneous space M we define the isotropy group at x by 3x = {geG
; g-x
= x}.
Lemma 7.4.2 Let M be a homogeneous space with symmetry group G and x,y € M. Then closed subgroup of G; (b) 3X =* 3y.
259
Symmetry and topology Proof that
(a) is immediate. To prove (b), choose g e G such that y = g-x. Then note 3y = g3xg~1-
•
Remark 7.4.3 It is worth mentioning some fundamental results in the theory of Lie groups which will shed a new light on the considerations of this section. Their proofs can be found in the monograph [73]. FACT 1. Any closed subgroup of a Lie group is also a Lie group. In particular, the isotropy groups 3X of a homogeneous space are all Lie groups. They are smooth submanifolds of the symmetry group. FACT 2. Let G be a closed group and H a closed subgroup. Then the space of left cosets G/H = {g • H ; g € G} can be given a smooth structure such that the map
Gx(G/H)^G/H
(gi,g2H)^(gig2)-H
is smooth. G/H becomes a homogeneous space with symmetry group G. All isotropy groups are isomorphic to H. FACT 3. If M is a homogeneous space with symmetry group G and x G M then M is equivariantly diffeomorphic to G/3X i.e. there exists a diffeomorphism 4>: M -> G/3X such that
M.
(c) a : M x M —> M is a smooth map (mi,m 2 ) !-• omi{mi) such that (cl) Vm 6 M am : M —> M is an isometry and am(m) = m. (c2) amoam = 1 M . (c3) Dam\TtnM= -lTmM(c4) Ggm = gamg~l (d) v : M x G —> G, (m, g) i-> i,„g is a smooth map such that (dl) Vm 6 M, im : G -> G is a homomorphism of G. (d2) im o im = 1G(d3) i 9 m = g\mg~l, V(m, g) e M xG. (e) amga^-{x)
= amgam{x)
= ^(g)
• x, Vm,x € M, g &G.
260
Cohomology
Remark 7.4.5 This may not be the most elegant definition of a symmetric space and certainly it is not the minimal one. Optimizing it will require a substantial amount of work and we refer to [34] for an extensive presentation of this subject. This definition has one advantage. It lists all the properties we need to establish the topological results of this section. The next exercise offers the reader a feeling of what symmetric spaces are all about. In particular, it describes the geometric significance of the family of involutions am. Exercise 7.4.1 Let connection and by R (a) Prove that VR = (b) Fix m 6 M and
(M,h) be a symmetric space. Denote by V the Levi-Civita the Riemann curvature tensor. 0. let j{t) be a geodesic of M such that 7(0) = m. Show that
°"m7(*) = 7 H ) Example 7.4.6 Perhaps the most popular example of symmetric space is the round sphere Sn C M™+1. The symmetry group is SO(n + 1) - the group of orientation preserving "rotations" of K" +1 . For each m e Sn+1 we denote by am the orthogonal reflection through the 1-dimensional space determined by the radius Om. We then set im(T) =CTmTo-"1,V T e S O ( n + l ) . We let the reader check that a and i satisfy all the required axioms. Example 7.4.7 Let G be a connected Lie group and m a bi-invariant Riemann metric on G. The direct product G x G acts on G by {91,92) •h =
g1hg21.
This action is clearly transitive and since m is bi-invariant its is also isometric. Define a : G x G -¥ G agh = gh~lg~x and i:Gx(GxG)->GxG
ig{gug2)
= (sSiS~\552P" 1 )-
We leave the reader check that these data define a symmetric space structure on (G, m). The symmetry group is G x G. Example 7.4.8 Consider the complex grassmannian M = Gk{n,C). S G M can be identified with a rank k, selfadjoint idempotent P : C " ->
Any element
P2.
P = Ps is the orthogonal projection onto the fc-dimensional subspace S C C". Denote by Symn the linear space of selfadjoint n xn complex matrices. The map M 3 S >-> Ps € Symn
261
Symmetry and topology is an embedding of M in Symn.
The linear space Symn has a natural metric
go{A,B)=9\etr{AB*). It defines by restriction a Riemann metric go on M. The unitary group U(n) acts on Symn by A^>T*A
= TAT*
TeU{n)AeSymn.
Note that U(n) * M = M and g0 is J7(n)-invariant. Thus £/(n) acts transitively and isometrically on M. For each subspace S € M define Rs = Ps-
PS± = 2P S - 1.
Rs is the orthogonal reflection through S-1. Note that Rs e U(n) and Rs = 2. The map A>-^ RS*A is an involution of Symn. family of involutions
It descends to an involution of M. We thus get an entire
a:MxM^M,
{PSl,Ps2)
•-> RSl * Ps2-
define i : M x t/(n) ->• [/(n),
isT =
flsTi?S-
We leave the reader check that the above collection of data defines a symmetric space structure on Gk(n,C). Exercise 7.4.2 Fill in the details left out in the above example. §7.4.2
Symmetry and cohomology
Definition 7.4.9 Let M be a homogeneous space with symmetry group G. A differential form u € Q*(M) is said to be (left) invariant if £*gw = ui \/g e G where we denoted by
tg: n*(M) -> n*(M) the pullback defined by the left action by g: rrn-^r g • m. Proposition 7.4.10 Let M be a compact homogeneous space with compact, connected symmetry group G. Then any cohomology class of M can be represented by a (not necessarily unique) invariant form.
262
Cohomology
Proof Denote by dfi9 the normalized bi-invariant volume form on G. For any form u € fi*(M) we define its G-average by uJ =
tgUidfig.
aJ is an invariant form on M. The proposition is a consequence of the following result. Lemma 7.4.11 If a; is a closed form on M then ZJ is closed and cohomologous to u. Proof of the lemma The form u> is obviously closed so we only need to prove it is cohomologous to UJ. Consider a bi-invariant Riemann metric m on G. Since G is connected the exponential map exp : £ G -> G X >-*• exp(iX) is surjective. Choose r > 0 sufficiently small such that exp : Dr = {\X\m
= r ; X G £ G } -> G
is an embedding. Set Br = expD T . We can select finitely many gi,..-,gm that
€ G such
771
G={jBj
(B^gjBr).
Now pick a partition of unity (a.j) C C°°(G) subordinated to the cover (Bj) i.e. 0 < OLJ < 1,
suppQj C Bj,
^2ai
=
!•
Set Oj = /
ajd/j,g.
For any j = 1 , . . . , m define Tj : ft*(M) -»• fi'(M) by Tj-w = /
aj{g)e*gujdfj,g.
JG
Note that a; = J ^ 3}w and dTjUi = Tjdtv. 3
Each Tj is thus a cochain map. It induces a morphism in cohomology which we continue to denote by Tj. The proof of the lemma will be completed in several steps. Step 1. £g = id on H*{M) for all g G G. Let X e £ G such that g = expX. / : / x M -)• M
/ t (m) = exp(tX) • m =
e*exp(tx)m.
Define
263
Symmetry and topology / is a homotopy connecting 1 M with I*. This concludes Step 1. Step 2 Tj = Ojlfl-.(Af).
For t € [0,1] consider 4>j,t • Bj —> G defined as the composition B.
9
4 Br •%' DT f 4 Dtr *% Btr 4 G.
Define Tj>t : Q*(M) -> fi*(M) by 2},(w = / a , - ( f l ) ^
(t) wd/i,
=T^>.
(7.4.1)
We claim that T j 0 is cochain homotopic to T ^ . To verify this claim set t = es, - c o < s < 0 and p s = exp(e s exp" 1 (g)) Vp £ B r . Then t/sw = TjJC,u> = J
a>j(gjg)e*g.gau; = J
a{gjg)tgfg.u}dng.
For each g 6 BT the map ys(m)
= g,(m)
defines a local flow on M. We denote by Xg its infinitesimal generator. Then d r — {Usu) = JB = ]
aj{gjg)(dMiXg
aj(gjg)LXgtgstg.u>dng +
ix9dM)(tgatg.u})diig.
Consequently TjfiU ~ T J M W = U-ooU -U0LO
= - 1 ^ [JB
aj{gjg){diXg
+ iXsd){tgstgju)diJ,g)
ds.
(An argument entirely similar to the one we used in the proof of the Poincare lemma shows that the above improper integral is pointwise convergent). From the above formula we immediately read a cochain homotopy x : fi*(M) —> fl*_1(M) connecting U-oo to U0. More precisely {xH}
\XEM= ~ j ^ yjB aj{gjg) {ixSgfgv}
Now notice that Tj.ou = [JB aj(g)diig\
= a/'.cj
U d(J.g) ds.
264
Cohomology
while Tj,iu> =
TjUj.
Taking into account Step 1 we deduce T, 0 = a,id. Step 2 is completed. The lemma and hence the proposition follow from 1H-(M) = YLa.jlH'(M) = Y^Ti j
= G
~
avera
ge-
D
i
The proposition we have just proved has a greater impact when M is a symmetric space. Proposition 7.4.12 Let (M, h) be an, oriented symmetric space with symmetry group G. Then the following are true. (a) Every invariant form on M is closed. (b) If moreover M is compact then the only invariant form cohomologous to zero is the trivial one. Proof (a) Consider an invariant fc-form u. Fix m £ M and set <2i = O^UJ. We claim <2> is invariant. Indeed, Vp € G Z*gU = < * / > =
(0Vnff)*O>
Since Z)
on M.
In particular diu = {-l)kduj
on M.
The (k + l)-forms dui = a^dw and du) are both invariant and as above we deduce du =
{-l)k+lduj.
The last two inequalities imply du) = 0. (b) Let u) be an invariant form cohomologous to zero i.e. u = da. Denote by * the Hodge *-operator corresponding to the invariant metric h. Since G acts by isometries
Symmetry and topology
265
7] = *u> is also invariant so t h a t drj = 0. We can now integrate ( M is compact) a n d use Stokes theorem t o get / u A *w = / daA = ± / a A dr) = 0.
•/M
JM
VM
This forces w = 0. • Form Proposition 7.4.10 and the above theorem we deduce the following celebrated result of Elie Cartan ([17]) Corollary 7.4.13 (Cartan)Let (M, h) be a compact, oriented symmetric space with compact, connected symmetry group G. Then the cohomology algebra H*{M) of M is isomorphic with the graded algebra Q.*inv{M) of invariant forms on M. In the coming subsections we will apply this result to the symmetric spaces discussed in the previous subsection: the Lie groups and the complex grassmannians. §7.4.3
T h e cohomology of compact Lie groups
Consider a compact, connected Lie group G and denote by £ G its Lie algebra. According to Proposition 7.4.10, in computing its cohomology it suffices to restrict our considerations to the subcomplex consisting of left invariant forms. This can be identified with the exterior algebra A*£ G . We deduce Corollary 7.4.14 H*(G) = H*(£G) £ A •„„£<; where A* nt) £ G denote the algebra of bi-invariant forms on G, while H*(£,Q) denotes the Lie algebra cohomology introduced in Example 7.1.9. Using the Exercise 7.1.2 we deduce the following consequence. Corollary 7.4.15 If G is a compact semisimple Lie group then HX(G) = 0. Proposition 7.4.16 Let G be a compact semisimple Lie group. Then H2(G) = 0. Proof A closed bi-invariant 2-form u on G is uniquely defined by its restriction to £ G and satisfies the following conditions. duj = 0 ^=> u>([X0,Xx},X2)
-CJ([X0,X2},X1)
+ LJ([XUX2],X0)
= 0
and (right-invariance) (LXow)(XuX2)=Q
VX0 G
£G<=*U([X0,X1},X2)-CJ([X0,X2},X1)=0.
Thus u{[X0,X1],X2)
= 0 V*0,XltX2€ £G.
On the other hand since ^(HQ) = 0 we deduce (see Exercise 7.1.2) £ G = [ £ G , £ G ] so that the last equality can be rephrased as u(X,Y)
=0 VX,ye£G. D
266
Cohomology
Definition 7.4.17 A Lie algebra is called simple if it has no nontrivial ideals. A Lie group is called simple if its Lie algebra is simple. Exercise 7.4.3 Prove that SU(n) and SO(m) are simple. Proposition 7.4.18 Let G be a compact, simple Lie group. Then H3(G) Moreover, H3(G) is generated by the Cartan form a(X,Y,Z)
=
= R.
K([X,Y],Z)
where n denotes the Killing pairing. The proof of the proposition is contained in the following sequence of exercises. Exercise 7.4.4 Prove that a simple Lie algebra is necessarily semi-simple. Exercise 7.4.5 Let w be a closed, bi-invariant 3-form on a Lie group G. Then u(X,y,
[Z,T]) = u([X, Y], Z,T)
VX, Y,Z,Te
£G.
Exercise 7.4.6 Let w be a closed, bi-invariant 3-form on a compact, semisimple Lie group. (a) Prove that for any X £ £ G there exists a unique left invariant TJX € O1 (G) such that (ixuj)(Y,Z) = r,x([Y,Z}). Moreover, the correspondence X i-> rjx is linear. HintrUse H\G) = H2{G) = 0. (b) Denote by A the linear operator £G —> £ G defined by K(AX,Y)=T,X(Y).
Prove that A is selfadjoint with respect to the Killing metric. (c) Prove that the eigenspaces of A are ideals of £ G . Use this to prove Proposition 7.4.18. Exercise 7.4.7 Compute / JSU(2)
a and /
a
JSO(3)
where a denotes the Cartan form. (These groups are oriented by their Cartan forms.) Hint: Use the computation in the Exercise 4.1.16 and the double cover SU(2) —> 5 0 ( 3 ) described in the Subsection 6.2.1. Pay very much attention to the various constants.
Symmetry §7.4.4
267
and topology Invariant forms on grassmannians and Weyl's integral formula
We will use the results of Subsection 7.4.2 to compute the Poincare polynomial of the complex grassmannian Gk(n, C). Set £ = n — k. Gk{n,C) is a symmetric space with symmetry group U(n). It is a complex manifold so that it is orientable (cf. Exercise 3.4.2). Alternatively, the orientability of Gk{n, C) is a consequence of the following fact. Exercise 7.4.8 If M is a homogeneous space with connected isotropy groups then M is orientable. We have to describe the [/(n)-invariant forms on Gk(n, C). These forms are completely determined by their values at a particular point in the grassmannian. We choose this point to correspond to the subspace So determined by the canonical inclusion C* ^ Cn. The isotropy of So is the group H = U(k) x U(£). H acts linearly on the tangent space V0 = Ts0Gk{n, C). If UJ is an [/(n)-invariant form then its restriction to V0 is an //-invariant skew-symmetric, multilinear map VQ
x • • • x Vo -»• M.
Conversely, any if-invariant element of A*V0* extends via the transitive action of U(n) to an invariant form on Gk{n, C). Denote by A*nv the space of .ff-invariant elements of A*V0*. We have thus established the following result. Proposition 7.4.19 There exists an isomorphism of graded M-algebras: H*{Gk(n,C))
= AJni).
We want to determine the Poincare polynomial of the complexified graded space, A* TO ®C
PkAt) = Et3
dim
c AL, ® c = PGt(n,C)(0-
3
Denote the action of H on Vo by H B h^Th
e Aut(Vo).
Using the equality (3.4.9) of Subsection 3.4.4 we deduce PhM
= / I d e t ( V 0 + tTh)\2dh
(7.4.2)
where dh denotes the normalized bi-invariant volume form on H. At this point the above formula may look hopelessly complicated. Fortunately, it can be dramatically simplified using a truly remarkable idea of H.Weyl.
268
Cohomology
Note first that the function
H3h^
is a class function i.e.
xTl
where T* = {diag(e I ( \...,e 1 9 *) € [ / ( £ ) } and similarly T ' = {diag (e 1 * 11 ,..., e1*') <E fJW} Each h € U(k) x U(£) is conjugated to diagonal unitary matrix i.e. there exists g e H such that ghg~l g T. We can rephrase this fact in terms of the adjoint action of H on itself Ad: HxH
^H
{g,h)^> Adg{h) = ghg'1.
The class functions are constant along the orbits of the adjoint action and each such orbit intersects the maximal torus T. In other words, a class function is completely determined by its restriction to the maximal torus. Hence, at least theoretically, we should be able to describe the integral in (7.4.2) as an integral over T. This is achieved in a very explicit manner by the next result. Define A n : Tn ->• C by
A„(0i,...A)=
II (eWi-ewO- (i = V=l) l
Proposition 7.4.20 (Weyl's integration formula) Consider a class function tp on the group G = U(ki) x • • • x U(ks). Then
fa ^g)d9
=
fci-^fc! L V^U '''' <s)|A*l(il)|2 ''' lA*-(*«)l2(fti A - ' - A dt°-
Above, dg denotes the normalized bi-invariant volume form on G. For each j = 1 , . . . , s we denoted by t = ti the collection of angular coordinates on Tk = Th> while dt denotes the normalized bi-invariant volume on Tk
Symmetry and topology
269
The remainder of this subsection is devoted to the proof of this proposition. The reader may skip this part at the first lecture and go directly to Subsection 7.4.5 where this formula is used to produce an explicit description of the Poincare polynomial of a complex grassmannian. Proof of Proposition 7.4.20 We will consider only the case s = 1. The general situation is entirely similar. Thus G = U(k) and T = Tk. Denote the angular coordinates on T by (01,..., 9k). Given a class function ip : G —> C form the complex valued form iov =
q:TxG/T^G
Note that if gxT = g{T then g\tgxx = gitg^1 so q is well defined. Pick a real metric m on £ G which is Ad-invariant i.e. Ad;m = m
WgeG.
The natural choice m(X, Y) = - 9 t e tr (XY*)
u{k) = £ G
will do the trick. The Lie algebra £<j splits orthogonally as £G
=
£T®
£G/T-
{£G/T
= £r)
The tangent space to 1 • T € G/T can be identified with £G/TFix x € G/T. Any g eG defines a linear map Lg : TXG/T -* TgxG/T. Moreover, if gx = hx = y then Lg and Lh differ by an element in the stabilizer of a; € G/T. This stabilizer is isomorphic to T and in particular it is connected. Hence, if a; € det TXG/T then Lgtj € det TyG/T and Lhoj € det TyG/T define the same orientation of TyG/T. In other words, an orientation in one of the tangent spaces of G/T "spreads" via the action of G to an orientation of the entire manifold. Thus, we can orient G/T by fixing an orientation on £G/T- We fix an orientation on HQ and orient £? using the form d6l A • • • A dOk. The orientation on HG/T will be determined by the condition (or= orientation) or(£ G ) = o r ( £ r ) A o r ( £ G / r ) The proof Weyl's integration formula will be carried out in two steps. Step 1 / (J = 77 / JG
k\
q*LJ V w .
JTXG/T
Step 2 For any class function tp on G we have
/ JTxG/T
q'uv=
[
270
Cohomology
S t e p 1. We use the equality /
q*u = deg q \ u
JTxG/T
JG
so it suffices to compute the degree of q. Denote by N(T) the normalizer of T in G i.e. N(T) = {geG;
gTg'1 C T}
and then form the Weyl group 2IT =
N(T)/T.
Lemma 7.4.21 2U = (S^-the group of permutations of k symbols. Proof This is a pompous rephrasing of the classical statement in linear algebra that two unitary matrices are similar iff they have the same spectrum (multiplicities included). The adjoint action of N(T) on T = diagonal unitary matrices simply permutes the entries of a diagonal unitary matrix. This action descends to an action on the quotient ?2XS so that 2H C &kConversely, any permutation of the entries of a diagonal matrix can be achieved by a conjugation. Geometrically this corresponds to a reordering of an orthonormal basis. • Lemma 7.4.22 Let a 1 , . . . , ak) £ Mfc such that 1, f^,..., ^ are linearly independent over Q. Set r = (exp(ia 1 ),... ,exp(ia* : )) G Tk. Then the sequence {Tn)n&2, is dense in Tk. (T is said to be a generator of Tk.) For the sake of clarity, we defer the proof of this lemma to the end of this subsection. Lemma 7.4.23 Let r € Tk C G be a generator of Tk. consists of |23J| = A;! points.
Then ^ ( r ) C T x G/T
Proof q{s, gT) = T<==> gsg-1 = r ^=> grg'1 = n
l
In particular, gr g~
n
= s £T,Vn
seT.
n
£ Z. Since (r ) is dense in T we deduce
gTg-lcT^geN{T). Hence 9 _ 1 (r) = {{g-lT9,gT) _1 T
and thus ? ( )
nas tne
eTxG/T;g&
N(T)}
same cardinality as the Weyl group 2B. •
271
Symmetry and topology The metric m on £ G / r extends to a G-invariant metric on G/T. left-invariant volume form on G/T, dv\n- Let v0 = /
JG/T G/T
It defines a
dv m
and set d\i = — dvxnLemma 7.4.24 q'dg = \Ak(t)\2dt A d/j,. In particular, any generator r of Th C G is a regular value of q since Afc(
hs = q(toexp(sX),g0exp(sY)T)
e G.
We want to describe —h^h, ds
6 TXG = £ G .
Using the Taylor expansions exp(sX) = 1 + sX + G(s 2 ) and exp( S y) = 1 + s F + 0(s2 we deduce V % = <7ot0-1(l + ay)*o(l + sX)(l = 1 + s (got^Ytogo1
- sY)g^
+ goXg^1 - goYgo1) +
+ 0(s2) 0(s2).
Hence Ar„ : TX0{T x G/T) * £r © £ G / T -> £ r © £ G / r = £ G can be written as DXQq{X ®Y)=
Ad 90 (Ad ( -. - id)Y + Ad 90 X
or in block form DXoq = Ad9,
0
Ad,-i - 1 P
Ad 9 is an m-orthogonal endomorphism of £ G so that det Ad 9 = ± 1 . On the other hand, since G = U(k) is connected det Ad 9 = det Adi = 1. Hence detDSog = d e t ( A d 1 - i - l i » o / r ) . We can identify £ G / r with { X e w ( f c ) ; Xjj = 0 j =
l,...,k}.
272
Cohomology
Given t = diag (exp(W 1 ),... , exp(i0*)) 6 T ' c U(k) we can explicitly compute the eigenvalues of Ad 4 -i acting on £G/T- They are {exp(-i(0i - 0,)) ; 1 < i± j < k}. Consequently = \/\k(t)\2.
detDXoq = det{Adt-i-l)
n
Lemma 7.4.24 shows that q is an orientation preserving map. Using Lemma 7.4.23 and Exercise 7.3.5 we deduce degg = |2B| = k\. Step 1 is completed. Step 2 follows immediately from Lemma 7.4.24. Weyl's integration formula is proved. • Proof of Lemma 2 We follow Weyl's original approach ([75],[76]) in a modern presentation. Let X = C(T, C) denote the Banach space of continuous complex valued functions on T. We will prove that
If U C T is an open subset and / is a continuous, non-negative function supported in U (/ ^ 0) then for very large n
This means that f(r') / 0 i.e. T> £ U for some j . To prove the equality (7.4.3) consider the continuous linear functionals Ln, L : X ->C
M/) = ^ £ / ( ^ ) • 3=0
and
L(f) = f fdt. JT
We have to prove that lim Ln(f)
= L(f)
/€X.
(7.4.4)
It suffices to establish (7.4.4) for any / G <S where 5 is a subset of X spanning a dense subspace. We let S be the subset consisting of the trigonometric monomials e c ( 0 \ . . . , 6k) = exp(iCi^1) • • • exp(iC^ fc )
C = (Ci, • • •, Ct) 6 %"•
273
Symmetry and topology
The Weierstrass approximation theorem guarantees that this <S spans a dense subspace. We compute easily L
^) = ^TlS^ (a) = ^ T l e c ( a ) - l •
Since 1, ^ r Q i , . . . , ^ak all C € Zk. Hence "
are linearly independent over Q we deduce that e^(a) / 1 for lim LJeA
= 0 = / e<-d£ = L(eA-
Lemma 7.4.22 is proved. • §7.4.5
The Poincare polynomial of a complex grassmannian
After this rather long detour we can continue our search for the Poincare polynomial ofGt(n,C). Let So denote the canonical subspace C* <-» C". The tangent space of Gk{n, C) at So can be identified with the linear space <£ of complex linear maps Cfc —> Ce, t — n — k. The isotropy group at So is H = U(k) x U{tj. Exercise 7.4.9 Prove the isotropy group H acts on <£ = {L : Ck -* Ce} by {T,S)-L
= SLT*
VL€<E,TeU{h),S
&U(l).
Consider the maximal torus Tk x Te C H formed by the diagonal unitary matrices. We will denote the elements of Tk by g = ( e i , . . . , ej.), eQ = exp(ir a ) and the elements of Tl by e = ( e i , . . . , ei), ej = exp(i^). The normalized measure on Tk is denoted by dr while the normalized measure on Te is denoted by d6. The element (e, e) 6 Tk x Te viewed as a linear operator on (£ has eigenvalues {eaej ; 1 < a < k
l<j<£}.
Using the Weyl integration formula we deduce that the Poincare polynomial of Gk {n, C) is PkM
=
= ~
^JTkxTtmi+t^j\2\AMr))\2\^(e(0))\2drAde fTkxTtn\ea+tej\2\AMT))\2\Me((>))\2dT
Ad6.
We definitely need to analyze the integrand in the above formula. Set hA*) = I ] > a + <e,)Afc(e)A£(e)
274
Cohomology
so that P
w W = TiH fu
JkAVhMdr
A d$.
(7.4.5)
We will study in great detail the formal expression Jk,t{t\ x'> V) = I[(x<* + *%•)• The Weyl group 2U = ©* x &e acts on the variables (x; y) by separately permuting the T-components and the ^/-components. If (p, ip) € 2U then JkAt;a(x)'>ip(y))
=
JkA^^y)-
Thus we can write Jj.^ as a sum
^MW = Eid<3d(a:)^(l/) where Qd(x) and i?d(y) are symmetric polynomials in x and respectively y. To understand the nature of these polynomials we need to introduce a very useful class of symmetric polynomials, namely the Schur polynomials. This will require a short trip in the beautiful subject of symmetric polynomials. An extensive presentation of this topic is contained in the monograph [50]. A partition is a compactly supported, decreasing function A:{1,2,...}^{0,1,2,...}. We will describe a partition by an ordered finite collection (Ai, A2,..., An) where Ai > • • • > An > A n+1 = 0. The length of a partition A is the number L(A) = max{n ; A„ ^ 0}. The weight of a partition A is the number |A| =
£ V n>\
Traditionally, one can visualize a partition using Young diagrams. A Young diagram is an array of boxes arranged in left justified rows (see Figure 7.4). Given a partition (Ai > • • • > An) its Young diagram will have Ai boxes on the first row A2 boxes on the second row etc. Any partition A has a conjugate A defined by An = # 0 ' > 0 ; A,- > n } . The Young diagram of A is the transpose of the Young diagram of A (see Figure 7.4). A strict partition is a partition which is strictly decreasing on its support. Denote by Vn the set of partitions of length < n and by "P* the set of strict partitions A of length n - 1 < L(X) < n. Clearly V*CVn. Denote by 6 = 6n G V*n the partition (n-l,n-2,...,l,0,...).
275
Symmetry and topology
Figure 7.4: The conjugate of (6,5,5,3,1) is (5,4,4,3,3,1) Remark 7.4.25 The correspondence VnB\>^\
+ 6n£V*
is a bijection. To any A e V* we can associate a skew-symmetric polynomial a A (:ri,...,x n ) = d e t ( a ^ ) =
Y,
e
(°)xa(i)-
a 6 e„ Note that asn is the Vandermonde determinant as(xu ...,xn)=
det(x" _ 1 _ I ) = Y[(xt - x,) =
An(x).
For each A G Vn we have A + 5 € "P* so that a^+s is well defined and nontrivial. Note that a\+s vanishes when xt = Xj so that the polynomial a\+s(x) is divisible by each of the differences (x( — Xj) and consequently is divisible by as- Hence s ( x )
=
^
)
is a well defined polynomial. It is a symmetric polynomial since each of the quantities a\+s and as is skew-symmetric in its arguments. S\(x) is called the Schur polynomial corresponding to the partition A. Note that each Schur polynomial 5A is homogeneous of degree |A|. We have the following remarkable result. Lemma 7.4.26 \i(t)=
£
trMS-x(x)S-x(y)
where Vk,i = {A ; Ai < fc L(A) < I}. For each A € T>k,i we denoted by A the complementary partition A = (k - Ac, k - X1-1, . . . , k - Ai).
Cohomology
276
6
1 0
7
j
4 4 2
1
0
*"
5 3
3 1
|||
Figure 7.5: The compiementarj of (6, 4,4, 2,1,0) £ P7,6 is (7,6, 5, 3,3,1) G P 6 , 7 Geometrically, the partitions in Vk,i are precisely those partitions whose Young diagrams fit inside &£xk rectangle. If A is such a partition then the Young diagram of the complementary of A is (up to a 180° rotation) the complementary of the diagram of A in the I x k rectangle (see Figure 7.5). For a proof of Lemma 7.4.26 we refer to [50] Section 1.4, Example 5. The true essence of the Schur polynomials is however representation theoretic and a reader with a little more representation theoretic background may want to consult [49], Chapter VI (Sec 6.4, Thm .V) for a very exciting presentation of the Schur polynomials and the various identities they satisfy, including the one in Lemma 7.4.26. Using (7.4.5), Lemma 7.4.26 and the definition of the Schur polynomial we can describe the Poincare polynomial of Gk(n, C) as
™=mL
TkxT<
£ *M,
^+ik
(e)aJ+S/(e)
dr A (
(7.4.6)
A€7>M
The integrand in (7.4.6) is a linear combination of trigonometric monomials e\l • • • e\ • e i ' " ' ' ei' where the r's and s' are nonnegative integers. Note that if A, /j, € Vk,i are distinct partitions then the terms a-x+6(e) and an+s{e) have no monomials in common. Hence jTkaX+M)aa+6(£)
dr = 0.
Similarly J iaj+5(e)a1J+s{e)de
= 0 if A / /z.
On the other hand a simple computation shows that
fTk\a-x+M2dr and JT<
*\+6\
= k\
277
Symmetry and topology
+1
•C 1
R l+\
,k+l
R
Figure 7.6: The rectangles Rt+i and Rk+1 are "framed" inside Rke+l In other words, the terms /
l
Nl/2
°A+«(£)aA+i(e)
KM)
form an orthonormal system in the space of trigonometric (Fourier) polynomials endowed with the L2 inner product. We deduce immediately from (7.4.6) that (7.4.7) ASPM
The map PH3AH>A6
V,k
is a bijection so that
PkA*)=
£
*2|A| =-fytW-
(7-4-8)
AeP,,t
Computing the Betti numbers i.e. the number of partitions in Vk/ with a given weight is a very complicated combinatorial problem and currently there are no exact general formulas. We will achieve the next best thing and rewrite the Poincare polynomial as a "fake" rational function. Denote by b^^w) the number of partitions A 6 V^k with weight |A| = w. Hence
Alternatively, bk,e(w) i s the number of Young diagrams of weight w which fit inside a k x £ rectangle. Lemma 7.4.27 h+i,i+i(w) = bk,t+1(w) + h+l>e(w - £ - 1)
278
Cohomology rectangle Rk+1 inside the (k + 1) x (1 + l)-rectangle
Proof Look at the (k + l)x£ Rke+l (see Figure 7.6). Then
bk+i,e+i{w) = # {diagrams of weight w which fit inside Rk+l} = #{diagrams which fit inside Rk+1} +#{diagrams which do not fit inside
Rk+1}.
bk+i,e = #{diagrams which fit inside
Rk+1}.
On the other hand
If a diagram does not fit inside Rk+1 this means that its first line consists of £ + 1 boxes. When we drop this line we get a diagram of weight w — £ — 1 which fits inside the k x (£ + 1) rectangle Rt+i of Figure 7.6. Thus, the second contribution to bk+u+i(w) is h,e+l(w - £ - 1). • The result in the above lemma can be reformulated as pk+u+1(t)
= Pk+U(t)
+ t2(m)PM+iW-
Because the roles of k and £ are symmetric (cf. (7.4.8)) we also have Pk+u+l(t)
= Pk4+1(t)
+
t2(k+Vpk+l,e(t)-
These two equality together yield P w + 1 ( l - ^ + 1 ) = P f c + 1 ,,(l-* 2 ( * + 1 ) ). Let m = k + £ + 1 and set Qd,m{t) = Pd,m-d{t) = PQ im C) W- The ^ a s t equality can be rephrased as Qk+l,m(t)
= Qk,m • i _ f2(k+l)
so that 1 _ f2(rn-k)
Qk+l,m{t) =
1
_
t2(fc+1)
j — fUm-k+l)
j " 3 ^
J _ ^2(n-l)
1 _*4
•
Now we can check easily that bl<m-i(w) = 1 i.e. Qi,mW = l + *2 + t 4 + --- + t 2 ( m - 1 ) Hence
(1 _ t2(m-k+l)j = <
9*,">(f) =
^(m.CjW
l-t2m 1-i2 . . . ( ! _ f2m)
Q _ t2\ . . . n _ f2*\
( 1 - t2) • • • ( 1 - t 2 m )
_ 2
~ (1 - 1 ) • • • (1 - t 2fc )(i - 1 2 ) • • • (i -1 2 (™-*))'
279
Cech cohomology
Remark 7.4.28 (a) The invariant theoretic approach in computing the cohomology of Gk{n, C) was used successfully for the first time by C. Ehresmann[24]. His method was then extended to arbitrary compact, oriented symmetric spaces by H. Iwamoto [38]. However, we followed a different avenue which did not require Cartan's maximal weight theory. (b) We borrowed the idea of using the Weyl's integration formula from Weyl's classical monograph [74]. In turn, Weyl attributes this line of attack to R.Brauer. Our strategy is however quite different from Weyl's. Weyl uses an equality similar to (7.4.5) to produce an upper estimate for the Betti numbers (of U(n) in his case) and then produces by hand sufficiently many invariant forms. The upper estimate is then used to established that these are the only ones. We refer also to [70] for an explicit description of the invariant forms on grassmannians. Exercise 7.4.10 Show that the cohomology algebra of QP" is isomorphic to the truncated ring of polynomials R[x}/{xn+1) where x is a formal variable of degree 2 while (xn+1) denotes the ideal generated by xn+l. Hint: Describe A| nv CP n explicitly.
7.5
Cech cohomology
In this last section we return to the problem formulated in the beginning of this chapter: what is the relationship between the Cech and the DeRham approach. We will see that these are essentially two equivalent facets of the same phenomenon. Understanding this equivalence requires the introduction of a new and very versatile concept, namely that of a sheaf. This is done in the first part of the section. The second part is a fast paced introduction to Cech cohomology. A concise yet very clear presentation of these topics can be found in [36]. For a very detailed presentation of this subject we refer to [29]. §7.5.1
Sheaves and presheaves
Consider a topological space X. The collection Ox of its open sets can be organized as a category. The morphisms are the inclusions U °-» V. A presheaf of abelian groups on X is a contravariant functor S : Ox —> A b . In other words, S associates to each open set an abelian group Sa and to each inclusion U °-> V a group morphism ry : Sy —>• Su such that if U •-• V M- W then r w — rv ° rw- If £ G <Sv then for any [ / ^ F w e set x\uaH 4(x)
G 5,
280
Cohomology
If / € Su then we define dom / = U. The presheaves of rings, modules, vector spaces are defined in an obvious fashion. E x a m p l e 7.5.1 Let X be a topological space. For each open set U C X denote by C(U) the space of continuous functions U —» R. The assignment U !-• C(U) defines a presheaf of R-algebras on X. The maps ry are determined by the restrictions \u- C(V) -> C(U). If X is a smooth manifold we get another presheaf U (->• C°°(U), of smooth function. More generally the differential forms of degree k can be organized in a presheaf fi*(»). If E is a smooth vector bundle then the E-valued differential forms of degree k can be organized as a presheaf of vector spaces
U^QkE{U) = nk(E\u). If G is an abelian group equipped with the discrete topology then the G-valued continuous functions C(U, G) determine a presheaf called the constant G-presheaf which is denoted by G_x. Definition 7.5.2 A presheaf S on a topological space X is said to be a sheaf if the following hold. (a) If (Ua) is an open cover of the open set U and f,geSu satisfy / \ua= 9 \ua Va then f = g. (b) If (Ua) is an open cover of the open set U and / „ e Sua satisfy fa \uanu0= fp krw f l then there exists / 6 Su such that f\ua=
VC/a n Up / 0
fa, Va.
E x a m p l e 7.5.3 All the presheaves discussed in Example 7.5.1 are sheaves. E x a m p l e 7.5.4 Consider the presheaf S over R defined by S(U) = continuous, bounded functions / : U —> R. We let the reader verify this is not a sheaf since the condition (b) is violated. The reason behind this violation is that in the definition of this presheaf we included a global condition namely the boundedness assumption. Definition 7.5.5 Let X be a topological space and R a commutative ring with 1. (a) A space of g e r m s over X ( "espace etale " in the french literature) is a topological space £ together with a continuous map ir : £ —>• X such that (al) 7r is a local homeomorphism (that is each point e G £ has a neighborhood U such that ir\u is a homeomorphism onto the open subset TT{U) C X). (a2)The stalk £x = 7r_1(a;) is an i?-module Vi 6 X. (a3) The module operations (u, v) H* (ru + sv) (r,s 6 R, u,v € -n~l{x)) depend continuously upon u, v. (b) A section of a space of germs 7r : £ -> X defined over U C X is a continuous function s : U —• £ such that s(x) G £x Mx e X. The spaces of sections defined over U will be denoted by £{U).
281
Cech cohomology
Example 7.5.6 (The space of germs associated to a presheaf) Let 5 be a presheaf of abelian groups over a topological space X. For each x G X define an equivalence relation ~ x on U3x
by / ~x 9 <==*• 3 open U B x such t h a t / \v= g \u . The equivalence class of / G \Ju3x $u is denoted by [f}x and is called the germ of / at x. Set Sx = {[/]*; d o m / 3 z } . and
s=\_\sx. xEX
There exists a natural projection n : S —> X which maps [f]x to x. The "fibers" of this map are TT~1(X) = <SX - the germs at i € X . Any f € Su defines a subset f(U) = {[/]„ ; u G [/} C 5 We can define a topology in S by indicating a basis of neighborhoods. A basis of open neighborhoods of [f]x G <S is given by the collection {g{U) ;UBx,g£Su
[g]x = [/],}.
(We leave the reader check that this collection of sets satisfies the axioms of a basis of neighborhoods as discussed e.g. in [43].) With this choice of topology each / G Sy defines a continuous section of 7r over U [/] : U B u M> [/]„ G Su. Note that each fiber Sx has a well defined structure of abelian group [f]x + [g]x = [(/ + g) \v]x U 3 x is open and U C d o m / n domg. (Check this addition is independent of the various choices.). Since n : f{U) —> U is a homeomorphism it follows that IT : S —> X is a space of germs. It is called the space of germs associated to the presheaf 5 . If the space of germs associated to a sheaf <S is a covering space we say that S is a sheaf of locally constant functions (valued in some discrete abelian group). When the covering is trivial i.e. it is isomorphic to a product X x { discrete set} then the sheaf is really the constant sheaf associated to a discrete abelian group.
282
Cohomology
Example 7.5.7 (The sheaf associated to a space of germs) Consider a space of germs £ A X over the topological space X. For each open subset U C X denote by £{U) the space of continuous sections U —> £. The correspondence U H-> £(U) clearly a sheaf. £ is called the sheaf associated to the space of germs. Proposition 7.5.8 (a) Let £ A X he a space of germs. Then £. = £. (b) Let S be a presheaf over the topological space X. S is a sheaf if and only if S = S. Definition 7.5.9 If S is a presheaf over the topological space X then <S = <S is called the sheaf associated to S. Exercise 7.5.1 Prove the above proposition. Definition 7.5.10 (a) Let A morphism between the (pre)sheaves of abelian groups (modules etc.) S and S over the topological space X is a collection of morphisms of abelian groups (modules etc.) hv : Su —> Su, one for each open set U C X, such that when V C U hv ° r^ — fy o hy. The Ty denote the restriction morphisms of S while the fy denote the restriction morphisms of S. A morphism h is said to be injective if each hy is injective. (b) Let S be a presheaf over the topological space X. A sub-presheaf of S is a pair (T, i) where T is a presheaf over X and i : T —> S is an injective morphism. i is called the canonical inclusion of the sub-presheaf. Let h : S —>• T be a morphism of presheaves. The correspondence U i-> ker hy C Su defines a presheaf called the kernel of the morphism h. It is a sub-presheaf of <S. Proposition 7.5.11 Let h : S -4 T be a morphism of presheaves. If both <S and T are sheaves then so is the kernel of h. The proof of this proposition is left to the reader as an exercise. Definition 7.5.12 (a)Let £* ^V X (i = 0,1) be two spaces of germs over the same topological space X. A morphism of spaces of germs is a continuous map h : £Q —» £\ such that (al) iti o h = 7To i.e. /i(7rj1(a;)) C 7rf 1(a;) Vz e X. (a2) For any x e X the induced map hx : 7r^"1(a;) -> 7rf1(x) is a morphism of abelian groups (modules etc.). The morphism is called injective if each hx is injective. (b) Let £ A X be a space of germs. A subspace of germs is a pair (T, j) where T is a space of germs over X and j : T —> £ is an injective morphism. Proposition 7.5.13 (a) Let h : £Q —> £\ he a morphism between two spaces of germs over X. Then h(£o) -^ X is a space of germs over X called the image of h and denoted by Im h. It is a subspace of £ x .
283
Cech cohomology Exercise 7.5.2 Prove the above proposition.
Lemma 7.5.14 Consider two sheaves S and T and let h : S —> T be a morphism. Then h induces a morphism between the associated spaces of germs h : S —> T• The definition of h should be obvious. If / € Sv and x 6 U then h([f]x) = [h(f)]x where h(f) is now an element of T(U). We leave the reader check that h is independent of the various choices and it is a continuous map S —> T with respect to the topologies described in Example 7.5.6. The sheaf associated to the space of germs Im/i is a subsheaf of T called the image of h and is denoted by Im h. Exercise 7.5.3 Consider a morphism of sheaves (over X) h : S —• T. Let U C X be an open set. Show that a section g ETu belongs to (Im h)u if an only if for every x € X there exists an open neighborhood Vx C U such that g \vx— h(f) for some
/eSv„. Definition 7.5.15 (a) A sequence of sheaves and morphisms of sheaves • • • —> Sn -4 Sn+i
-4 >S„+2 —• • • •
is said to be exact if Im hn = ker hn+i, Vn. (b) Consider a sheaf S over the space X. A resolution of S is a long exact sequence 0 -> S A S0 4 5 i 4 • • • -> Sn % Sn+1 -> • • •. Exercise 7.5.4 Consider a short exact sequence of sheaves 0 -> 5_i -> S0 ->• Si -> 0. For each open set U define S(U) = S0(U)/S-i(U). (a) Prove that U *-* S(U) is a presheaf. (b) Prove that Si = S = the sheaf associated to the presheaf S. Example 7.5.16 (The D e R h a m resolution)Let M be a smooth ra-dimensional manifold. Using the Poincare lemma and the Exercise 7.5.3 we deduce immediately that the sequence
o^R
M
^ n°M A nlM A • • • A o,nM ^ o
is a resolution of the constant sheaf RM. Q.kM denotes the sheaf of fc-forms on M while d denotes the exterior differentiation.
284 §7.5.2
Cohomology Cech cohomology
Let U \-> S(U) be pre-sheaf of abelian groups over a topological space X. Consider an open cover i l = (t/ a )as^t of X. A g-simplex (q € Z+) is an ordered (q + l)-uple a = (a 0 , Q i , . . . , aq) € Aq+1 such that
Ua=ff]Uai^0. o The set of all g-simplices is denoted by ii( ? ). Their union
U-"fo> is denoted by Af(il) and is called the nerve of the cover. Define
C<(S,U) = II
S
° (S* = S(Ua)).
The elements of C(S,ii) are called Cech g-cochains (subordinated to the cover il). In other words, a g-cochain c associates to each g-simplex a an element (c, a) 6 S„For each g-simplex a = (a0,... ,aq) we define its j - t h boundary as the (q — 1)simplex &o = &qo = (a0,..., otj,..., aq) e il(,-i) where as usual a hat indicates a missing entry. Exercise 7.5.5 Prove that 9J_ X ^ = d^Zld'q for j > i. We can now define an operator
5:Co-1{S,il)->C'l(S,iX) which assigns to each (q — l)-cochain c a g-cochain 6c whose value on a g-simplex a is given by
(5c,a) =
f:(-iy(c,cVa)\u,.
Using the Exercise 7.5.5 above one deduces immediately the following result. L e m m a 7.5.17 52 = 0 so that
o <-> c°(s,u) 4 c 1 ^,!!) -4 • • • 4 cq(s,ii) 4 • • • is a cochain complex. The cohomology of this cochain is called the Cech cohomology of the cover il with coefficients in the pre-sheaf S.
285
Cech cohomology
Example 7.5.18 Let i l and S as above. A 0-cochain is is a correspondence which associates to each open set Ua € i l an element cQ € S(Ua)- It is a cocycle if for any 1-simplex (a, /?) of the nerve we have C/3 -
CQ =
0.
A 1-cochain associates to each 1-simplex (a, (5) an element Cal3 G
S{Uap).
This correspondence is a cocycle if for any 2-simplex (a, /?, 7) we have C/?7 — C a 7 + Ca/} = 0.
For example if X is a smooth manifold and i l is a good cover then we can associate to each closed 1-form w e Q1(M) a Cech 1-cocycle valued in Rx as follows. First, select for each Ua a solution fa G C°°(Ua) of d/„ = oi. Since d(/ Q — /^) = 0 on f a| a we deduce there exist constants cap such that fa — Jp = cap. Obviously this is a cocycle and it is easy to see that its cohomology class is independent of the initial selection of local solutions fa. Moreover, if UJ is exact this cocycle is a coboundary. In other words we have a natural map
H1(X)^H1(Af(iX)Mx)We will see later this is an isomorphism. Definition 7.5.19 Consider two open covers i l = (Ua)aeA and 2J = ( V ^ g s of the same topological space X. We say 2J is finer than i l and we write this i l -< 2J if there exists a map Q : B —> A such that
v0cum
V(3eB.
The map Q is said to be a refinement map. Proposition 7.5.20 Consider two open covers i l = {Ua)aeA and 93 = (VJj)/jgB of the same topological space X such that i l -< 93. Fix a sheaf of abelian groups S. Then the following are true. (a) Any refining map Q induces a cochain map
0.:©c%u,s)->©c(aj,s). (b) If r : A —• B is another refinement map then Q, is cochain homotopic to r„. In particular, any relation i l X 23 defines a unique morphism in cohomology t^l : H*(il,S)
-> H*(V3,S).
(c) If i l X 23 -< 223 then 213 ! 2D 23 i l = 2 3 °*U-
J
286
Cohomology
Proof
(a) We define Q. : C(Xl) -s- C?(2J) by S(Va) B
VC e C ( i l ) a €
VM)
q
where by definition g(a) € A is the g-simplex {Q{PQ), . . . , Q{Pq))- The fact that £>, is a cochain map follows immediately from the obvious equality Q o &q = 8% o p.
(b) We define ft, : 2J(,-i) ->• it( g ) by hj-(A), • • • ,/?,-i) = (e(A>),..., e{Pj),r{Pj), • •-.»•(/?,_!)). The reader should check that fcj(
Vc e C(ii)
Va e 2J(,-i).
We will show that S o X ,(c) + X,+i ° *(*) = g.(c) - r.(c)
Vc e C ' ( i l ) .
Let a = (/?0, • • •, Pq) € 9J(?) and set e(ff) = (Ao, • • •, A,), r(a) = ( p 0 , . . . , nq) £ il ( ,) so that ftj(cr) = (A0, ...,\j,fij,...,
nq).
Then q
9+1
= £(-i) J '(£(-inc,9* + 1 ft>)>k) ^E(- 1 )ME(- 1 ) fc ( c '( A o,-.-,A,-,... > A j)ft ,...,/i,))iv
\fc=0
+ ZX-l) m (c, (A°' • • •' K Hi • • •. A*. • • •. M«)> k :
i(-l)ME(-l)*
\fc=0
Cech cohomology
287 i
+ Y,
(- 1 ) m ( c >( A o,---,A.,-,A i j,---,Wb---,M«)>l
t=j+i i
+ Z X - ! ) J {(c> (•*<>> • • •. ^ j - i . Mj, • • • > M«)> k +(c, (A0, • •., Aj, / x j + 1 ) . . . , //,)) | v „} • The last term is a telescopic sum which is equal to (c, (A 0 ,..., A,) \v„ - ( c , (/x0, • • •, M,)) k = (e.c,CT>- (r.c, a). If we change the order of summation in the first two term we recover the term
(Sxc,a). (c) is left to the reader as an exercise. • We now have a collection of groups {H*{ii, S) ; i l - open cover of X} and maps {»g:.ff%U,S); i l ^ S T } such that ,i i l = i d
ti
and
lU = J 0 J o l i l whenever il -< 2J -< 2U. We can thus define the inductive limit = / lim/f*(il,«S).
H'(X,S)
if* (X, 5) is called the Cech cohomology of the space X with coefficients in the pre-sheaf S. Let us briefly recall the definition of the direct limit. One defines an equivalence relation on il by i T ( i l ) B f ~ g G # • ( » ) <=> 30B V U, Of : i f f / = z g V We denote the equivalence class of / by / . Then lim/T(it)= *
(Uff*CU,$))/. \il
At limit we get maps *iI:/T(.U,lS)->ir(X,<S).
288
Cohomology
Example 7.5.21 Let S be a sheaf over the space X. For any open cover i t = (UQ) a 0-cycle subordinated to i t is a collection of sections fa 6 S(Ua) such that every time Ua n Up ^ 0 fa\uafi=
fp\ua0
•
According to the properties of a sheaf, such a collection defines a unique global section / € S(X). Hence H°(X,S) = S(X). Proposition 7.5.22 Any morphism of pre-sheaves h : So —> <Si over X induces a morphism in cohomology
h,:H*(X,S0)^H%X,S1). Sketch of proof
Let i t be an open cover of X. Define
ht:C'(iX,S0)^C(il,S1) by (h,c, a) = hu((c, a}) Vc 6 C(il, S0) a e il ( g). The reader can check easily that ht is a cochain map so it induces a map in cohomology
h^:H*(ii,S0)^H*(ii,S1) which commutes with the refinements i^/. The proposition follows by passing to direct limits. • Theorem 7.5.23 Let 0 -> 5_i 4 S 0 ^ <5i -> 0 be an exact sequence of sheaves over a paracompact space X. Then there exists a natural long exact sequence • H"{X, S-i) 4 H"(X, So) ^ H"(X, Si) 4 Hq+1(X, 5_i) -> • • •. Sketch of proof For each open set U C X define S(U) = S0{U)/S-i(U). Then the correspondence U H-> S(U) defines a pre-sheaf on X. Its associated sheaf is isomorphic with Sx (see Exercise 7.5.4). Thus for each open cover i t we have a short exact sequence 0 -> C»CU,5_i) A C 9 (it,«S 0 ) 4 C(.U, 5) -*• 0. We thus get a long exact sequence in cohomology • / / ' ( i t , 5_i) -»• / / ' ( i t , So) -> / / ' ( i t , S) -> / / « + 1 ( i i , A") -> • • •. Passing to direct limits we get a long exact sequence • H"{X, S-i) -> H"{X, So) -» / / ' ( X , 5) -> / / ' + 1 ( X , 5_i) - • • • • . To conclude the proof of the proposition we invoke the following technical result. Its proof can be found in [68].
Cecil cohomology
289
Lemma 7.5.24 If two pre-sheaves <S, S' over a paracompact topological space X have isomorphic associated sheaves then H*(X,S)^H*{X,S'). Definition 7.5.25 A sheaf S is said to be fine if for any locally finite open cover iX = (Ua)a£A there exist morphisms ha : S —¥ S with the following properties. (a) For any a & A there exists a closed set Ca such that Ca C Ua and ha(Sx) = 0 for x £ Ca where Sx denotes the stalk of S at x G X. The set Ca is called a support of ha and and we write this supp ha C Ca( b ) £ a ha = ls- This sum is well defined since the cover i l is locally finite. Example 7.5.26 Let X be a smooth manifold. Using partitions of unity we deduce that the sheaf £lkx of smooth fc-forms is fine. More generally, if E is a smooth vector bundle over X then the space Q.kE of E-valued fc-forms is fine. Proposition 7.5.27 Let 5 be a fine sheaf over a paracompact space X. H"{X,S) £ 0 f o r g > 1.
Then
Proof Because X is paracompact any open cover admits a locally finite refinement. Thus it suffices to show that for each locally finite open cover il — {Ua)aeA the cohomology groups Hq(il, S) are trivial for q > 1. We will achieve this by showing that the identity map C(il, S) —¥ C(il,S) is cochain homotopic with the trivial map. We thus need to produce a map Xq:C{ii,S)-^C''-1{iX,S) such that X«+1<5» + 5q-\*
= id.
(7.5.1)
Consider the morphisms ha : S ~> S associated to the cover i l postulated in the definition of a fine sheaf. For each a € A, a € iX(,-i) and / € C'(il, S) we construct (ta{f),o) € S(Ua) as follows. Consider the open cover of IJ„ {V = UanU„,W
= UtT\Ca}
(supp/iQcCa).
Note that haf \vnw= 0 and according to the axioms of a sheaf ha(f \v) can be extended by zero to a section (ta(f), a) € S(Ua). Now, for every / e C(ii, S) define q l by X feC<~ (il,S) q
(x V),°) = Y,(W),v)a
The above sum is well defined since the cover i l is locally finite. We leave the reader check that x9 satisfies (7.5.1). •
290
Cohomology
Definition 7.5.28 Let S be a sheaf over a space X. A fine resolution is a resolution
o -> s ^ s0 4 Si -4 • • • such that each of the sheaves <S, is fine. T h e o r e m 7.5.29 ( A b s t r a c t D e R h a m t h e o r e m ) Let 0 ^ S ->• S0 $ Sx $ • • • be a fine resolution of the sheaf <S over the p a r a c o m p a c t space X. Then 0 -> SQ{X) 4 Si{X) 4 • • • is a cochain complex and there exists a natural isomorphism Hq{X) =*
H"(Sq{X)).
Proof The first statement in the theorem can be safely left to the reader. For q > 1 denote by Zq the kernel of the sheaf morphism dq. We set for uniformity ZQ = S. We get a short exact sequence of sheaves 0 ->• Zq -> Sq -> Zq+1 - » 0 q>0.
(7.5.2)
We use the associated long exact sequence in which Hk(X, Sq) = 0 for k > 1 since Sq is a fine sheaf. This yields the isomorphisms Hk-\X,
Zq+1) £ Hk(X, Zq)
k>2.
We deduce inductively that Hm{X,Z0)^H\X,Zm_l)
m>\.
Using again the long sequence associated to (7.5.2) we get an exact sequence H°(X, V i ) ^ ' H°(X, Zm) -». H\X,
Zm_x) -> 0.
We apply the computation in Example 7.5.21 we get H\X,
2 m _ 0 3 Zm(X)K~x
(«S m _iP0).
This is precisely the content of the theorem. D Corollary 7.5.30 Let M be a smooth manifold. Then H*(M,RM)^H*(M).
(7.5.3)
291
Cecil cohomology Proof
M is paracompact. We conclude using the fine resolution 0 -+ 1
M
M - ^ M - ^ ^ M ->••••
D
Exercise 7.5.6 Describe explicitly the isomorphisms H1{M)^H1{M,RM) and H2(M)^H2{M,RM). Remark 7.5.31 The above corollary has a surprising implication. Obviously the Cech cohomology is a topological invariant and thus, so must be the DeRham cohomology which is denned in terms of a smooth structure. Hence if two smooth manifolds are homeomorphic they must have isomorphic DeRham groups! Such exotic situations do exist. In a celebrated paper [54], John Milnor has constructed a family of nondiffeomorphic manifolds all homeomorphic to the sphere S7. More recently, the work of Simon Donaldson in gauge theory was used by Michael Freedman to construct a smooth manifold homeomorphic to M4 but not diffeomorphic with M4 equipped with the natural smooth structure. (This is possible only for 4-dimensional vector spaces!) These three mathematicians, J. Milnor, S. Donaldson and M. Freedman were awarded Fields medals for their contributions. Theorem 7.5.32 Let M be a smooth manifold and iX = (Ua)aeA M i.e. Ua ^ l d i m M . Then
a
good cover of
Ht(ii,mM)^H*{M).
Proof Let Zk denote the sheaf of closed fc-forms on M. Using the Poincare lemma we deduce Zk{U„) = dQ."-1^). We thus have a short exact sequence
o -> cq{n, zk) -> c(ii, n*-1) 4 c(u, zk) -> o. Using the associated long exact sequence and the fact that Q,k~1 is a fine sheaf we deduce as in the proof of the abstract DeRham theorem that H^il,RM)^H"-1(U,Z1)^--s* H\il,
Z"-1) s H°(U, Z")/d,H°{iX, W1)
£ Z"{M)ldQ.q-x{M)
£ H*{M).
•
292
Cohomology
Remark 7.5.33 The above result is a special case of a theorem of Leray: if S is a sheaf on a paracompact space X and i l is an open cover such that H"(Ua,S)
=0
\/q>laeU{k)
then H*(il,S) = H*(X,S). For a proof we refer to [29]. When G_M is a constant sheaf (where G is an arbitrary abelian group) we have a Poincare lemma (see [25], Chapter IX, Thm. 5.1) H"(Rn,G)
=0
q>\.
Hence for any good cover iX H'(M,GM)=H*(U,GM).
Example 7.5.34 Let M be a smooth manifold and i l = (Ua)a^A a good cover of M. A 1-cocycle of R M is a collection of real numbers /a/g - one for each pair (a,/3) € A2 such that Ua0 ^ 0 satisfying fa0 + hi + fja = 0 whenever Uap7 ^ 0. The collection is a coboundary if there exist the constants fa such that fap = fp — fa. This is precisely the situation encountered in Subsection 7.1.2. The abstract DeRham theorem explains why the Cech approach is equivalent with the DeRham approach. Remark 7.5.35 Often in concrete applications one finds it is convenient to work with skew-symmetric Cech cochains. A cochain c G C(S, il) is skew-symmetric if for any q simplex a = (ao, • • •, o-q) and for all ip € &q+i we have (c, (a0,...,
a,)) = e{tp)(c, ( a ^ o ) , . . . , Q^,)))-
One can then define a "skew-symmetric" Cech cohomology following the same strategy. (The various intervening formulae will be more elaborate.) The resulting cohomology coincides with the cohomology described in this subsection. For a proof of this fact we refer to [68].
Chapter 8 Characteristic classes We now have sufficient background to approach a problem formulated in Chapter 2: find a way to measure the "extent of nontriviality" of a given vector bundle. This is essentially a topological issue but, as we will see, in the context of smooth manifolds there are powerful differential geometric methods which will solve a large part of this problem. Ultimately, only topological techniques yield the best results.
8.1 §8.1.1
Chern-Weil theory Connections in principal G-bundles
In this subsection we will describe how to take into account the possible symmetries of a vector bundle when describing a connection. All the Lie groups we will consider will be assumed to be matrix Lie groups i.e. Lie subgroups of a general linear group GL(n, K). This restriction is neither severe, nor necessary. It is not severe since, according to a nontrivial result (Peter-Weyl theorem), any compact Lie group is isomorphic with a matrix Lie group and these groups are sufficient for the applications in geometry. It is not necessary since all the results of this subsection are true for any Lie group. We stick with this assumption since most proofs are easier to "swallow" in this context. For a matrix Lie group G, its Lie algebra JJ is a Lie algebra of matrices in which the bracket is the usual commutator. Let M be a smooth manifold. Recall that a principal G-bundle P over M can be defined by an open cover (Ua) of M and a gluing cocycle 9ap • Uap —> G.
The Lie group G operates on its Lie algebra g via the adjoint action Ad:G->
GL{g),
g .-*• Ad(g) e GL{g)
where Ad{g)X = gXg-1 293
VX e g.
294
Characteristic classes
We denote by Ad(P) the vector bundle with standard fiber g associated to P via the adjoint representation. In other words, Ad(P) is the vector bundle defined by the open cover (Ua) and gluing cocycle Ad(gaP) : UaP -> GL{g). The bracket operation in the fibers of Ad(P) induces a bilinear map [•,•] : Qk(Ad{P))
x tf{Ad(P)) -»
Qk+e{Ad(P))
defined by [w* ® X, rf ® Y] = (w* A T)1) ® [X, Y] k
(8.1.1)
£
for all tj 6 fi*(M), ?7 g 0 £ (M) and X, F e n°(ylrf(P)). Exercise 8.1.1 Prove that for any u>, rj, > € fi*(Ad(F)) the following hold. K ^ ] = -(-l)H-M[,,w].
(8.1.2)
[[u,,7?],^] = [[w, fl,77] + (-1)M-I«[ W , fa, *}].
(8.1.3)
In other words, fT(yW(F)), [, ]) is a super Lie algebra. Definition 8.1.1 (a) A connection on the principal bundle P is a collection Aa € ^(Uc) <8> 0 satisfying the transition rule A
f>ix) = 9«p{x)dgap{x) + g-p(x)AQ{x)ga/3(x)
Vz € Uaf).
(b) The curvature of the above connection is defined by the collection Fa € fl2([/Q)® g where Fa = dAa +
-[Aa,Aa].
Proposition 8.1.2 (a) The set 21(F) of connections on F is an affine space modeled by Q^AdiP)). (b) The collection (Fa) defines a global j4d(F)-valued 2-form. (c) (The Bianchi identity) dFa + [Aa,Fa] = Q VQ. Proof (a) If (Aa, {Ba) £ 21(F) then their difference Ca = Aagluing rules Cp = g/3aCagj3a so that it defines an element of Q1(Ad(P)). (b) We need to check that the forms Fa satisfy the gluing rules F0 =
g-lFag
(8.1.4) Ba satisfies the
295
Chern-Weil theory where g = ga0 = g^.
We have
= d(g-ldg + g-xAag)
+ -\g~ldg + g~lAag, g~ldg +
g~lAag}.
Set m = g~*dg. Using 8.1.2 we get Fp = +d(g~1Aag)
dm-\--[m,m]
+ [vj,g-lAag\
+ -{g-lAag,g-lAag\.
(8.1.5)
We will check two things. A. The Maurer-Cartan structural equations. dm + -[m,m] = 0. B. d{g~lAag)
+ [m, g~lAag] =
g-x{dAa)g.
Proof of A. Let us first introduce a new operation. Let gl{n, K) denote the associative algebra of K-valued n x n matrices. There exists a natural operation A : Slk{Ua) ® gl(n, K) x Qe(UQ) ® gl(n, K) -> Slk+e{Ua) ® p/(n, K) uniquely defined by (<J* ®A)A
{rf ®B) = (u>k A rf) ® (A • B)
(8.1.6)
where uk € ft*(£4), ?/ € ft£(£/a) and A, B G p/(n,K) (see also Example 3.3.11). The space gl(n7 K) is naturally a Lie algebra with respect to the commutator of two matrices. This structure induces a bracket [, ] : Slk(Ua) ® gl(n, K) x tf(Ua) ® gj.(n, K) -> ttk+e{Ua) ® gi(n, K) defined as in (8.1.1). A very simple computation yields the following identity. uA7]=-[ui,r]}
^u),r]&Q1(Ua)®gi{n,K).
(8.1.7)
Assume the Lie group lies inside GL(n, K) so that its Lie algebra g lies inside gl(n, K). We can think of the map gap as a matrix valued map so that we have dm = d{g~ldg) = (dg'1) Adg=
-(g'1
• dg • g~l)dg = -{g~ldg)
A {g~xdg)
296
Characteristic classes = -G7As7
=
— -[m, UJ\.
Proof of B. We compute d{g-lAQg) = -g'1
= {dg-lAa)
• g + g~\dAa)g
• dg • g-1 A Aa • g + g-l{dAa)g = -w A g~lAag + g~lAa
+
+ ( s " M a S ) A g~xdg g-1{dAa)g
Azu +
-\w, g-xAag] + -[w, gxAag] +
2l
(8.1.2)
[nj,g ^cgj
g^A^g
g'^dA^g
x
+g
{dAa)g.
Part (b) of the proposition now follows from A, B and (8.1.5). (c) First, we let the reader check the following identity d[w)T/] = [«L;>^] + (-l)M[w)d»7]
(8-1-8)
where w, rj 6 Q,*(Ua)
= [Fa,Aa] - \[\Aa,Aa], The proposition is proved.
Aa]
(8
(8
= 2 ) [dAa,Aa]
= 3 ) [Fa,AQ].
•
Exercise 8.1.2 Let ua G iik(Ua) ® g satisfy the gluing rules w/3 = gpawag0a
on
Ua0.
In other words, the collection uia defines a global /c-form w e £lk(Ad(P)). the collection dtoa + [Aa,ua\
Prove that
defines a global .<4d(P)-valued (k + l)-form on M which we denote by d^uj. Thus, the Bianchi identity can be rewritten as dAF{A) = 0 for any A € 2l(P).
Chern-Weil theory §8.1.2
297
G-vector bundles
Definition 8.1.3 Let G be a Lie group and E —¥ M a vector bundle with standard fiber a vector space V. A G structure on E is defined by the following collection of data. (a) A representation p : G —> GL(V). (b) A principal G-bundle P over M such that E is associated to P via p. In other words, there exists an open cover (Ua) of M and a gluing cocycle 9a/3 '• Uap —> G
such that the vector bundle E can be defined by the cocycle P(gap) •• Ua0 ->
GL(V).
We denote a G-structure by the pair (P,p). Two G-structures (P*, p,) on E (i = 1,2) are said to be isomorphic if pi = p2 and the principal G-bundles Pi are isomorphic. Example 8.1.4 Let E —>• M be a rank r real vector bundle over a smooth manifold M. A metric on E allows us to talk about orthonormal moving frames. They are easily produced from arbitrary ones via the Gramm orthonormalization technique. In particular, two different orthonormal local trivializations are related by a transition map valued in the orthogonal group 0(r) so that a metric on a bundle allows one to replace an arbitrary collection of gluing data by an equivalent (cohomologous) one with transitions in 0(r). In other words, a metric on a bundle induces an 0(r) structure. The representation p is in this case the natural injection 0(r) «-» GL(r, M). Conversely, an 0(r) structure on a rank r real vector bundle is tantamount to choosing a metric on that bundle. Similarly, a Hermitian metric on a rank k complex vector bundle defines an U(k)structure on that bundle. Let E — (P, p, V) be a G-vector bundle. Assume P is defined by an open cover (Ua) and gluing maps 9ag • Uap —> G. 1
If the collection {Aa 6 0 ([/ a ) ® g} defines a connection on the principal bundle P then the collection p,{Aa) defines a connection on E. Above, p, : g —> End (V) denotes the derivative of p at 1 e G. A connection of E obtained in this manner is said to be compatible with the G-structure. Note that if F(Aa) is the curvature of the connection on P then the collection p„(F(Aa)) coincides with the curvature F(p„(Aa)) of the connection p„(Aa). For example, a connection compatible with some metric on a vector bundle is compatible with the orthogonal/unitary structure of that bundle. The curvature of such a connection is skew-symmetric which shows the infinitesimal holonomy is an infinitesimal orthogonal/unitary transformation of a given fiber.
298 §8.1.3
Characteristic classes Invariant polynomials
Let V be a vector space over K = R, C. Consider the symmetric power Sk(V)
C (V*)®*
which consists of symmetric, multilinear maps
->K.
Note that any
=
••-,»).
This follows immediately via the polarization formula ip{vi
'---Vk)
=
hdt1---dtkPv{tiVi+---+tkVk)-
If dim V = n, then fixing a basis of V we can identify Sh(V*) with the space of degree k homogeneous polynomials in n variables. Assume now that A is a K-algebra with 1. Starting with ip € S*(V*) we can produce a K-multilinear map ip = ipA : (A®V)
x • • • x (A®V)
-* A
uniquely determined by v(«i ®vi,...,ak®vk)
= tp{vu ..., vk)aia2 • • • a^ £ A.
If moreover the algebra A is c o m m u t a t i v e then (pA is uniquely determined by the polynomial Pv{x)
=
x e A ® V.
This is no longer the case if A is not commutative. Example 8.1.5 Let V = gl(n,C). For each matrix T € V we denote by ck(T) the coefficient of A* in the characteristic polynomial
cx(T) = det ( l - ^ T ) = Y, cdT)\k,
(i = >/=!).
ck{T) is a degree k homogeneous polynomial in the entries of T. For example c l ( T) = - ^ t r r ,
c n (T) =
(^)
n
detT.
Via polarization, ck(T) defines an element of Sk(gl(n, C)*).
299
Chern-Wei! theory
If A is a commutative C-algebra with 1 then A ® gUp., C) can be identified with the space gl{n, A) of n x n matrices with entries in A- For each T € gl{n, A)
det 1
( -Si T ) e - A W
and C).(T) continues to be the coefficient of \k in the above polynomial. Consider now a matrix Lie group G. The adjoint action of G on its Lie algebra g induces an action on Sk(g*) still denoted by Ad. We denote by Ik(G) the >ld-invariant elements of Sk(g*). It consists of those ip € Sk(g*) such that
/*(G) = 0/*(G) *>o and
r*(G) = n J*(G). fc>0
The elements of I*(G) are usually called invariant polynomials. 7**(G) can be identified (as vector space) with the space of Ad-invariant formal power series with variables from g*. E x a m p l e 8.1.6 Let G = GL(n,C)
so that g = g/(n,C). The map
p/(n, C) 3 Jf i-> tr exp(X) defines an element of I**(GL(n, C)). To see this we use the "Taylor expansion"
exp(X) = £ - L r * /t>o K -
which yields trexppO = £ ± t r X \ fc>0 Kk
For each fc, t r X * G I (GL(n,C))
since
tr( S X f f - 1 )*=trflXV 1 =trX*. Proposition 8.1.7 Let y> G Ik(G).
...,Xk)
Then for any
X,X1,...,Xk€g
+ --- +
The proposition follows immediately from the equality ft\t=ov(etxXle-t\...,etxXke-tx)
= 0. Q
(8.1.9)
300 §8.1.4
Characteristic classes The Chern-Weil theory
Let G be a matrix Lie group with Lie algebra g and P —> M a principal G-bundle over the smooth manifold M. Assume P is defined by an open cover (Ua) and a gluing cocycle gap '• Uag -> G.
Pick A G 21(F) defined by the collection Aa e ^(Ua) defined by the collection Fa = dAa +-[Aa,
® g. Its curvature is then
Aa].
Given <j> e /*(G) we can define as in the previous section (with A =
£leven(Ua),
V = o) p
Because <j> is yld-invariant and Fp = gpaFag^ P^(F Q ) = P^Fp)
we deduce on Uap
so that the locally defined forms P^{Fa) patch-up to a global 2/c-form on M which we denote by >(F(A)). Theorem 8.1.8 (Chern-Weil) (a) The form <j>{F{A)) is closed VA e 21(F). (b) If A0, A1 £ 21(F) then the forms ){F(A0)) and 4>{F{A1)) are cohomologous. In other words, (f>(F(A)) defines a cohomology class in H2k(M) which is independent of the connection A £ 21(F). Proof
We use the Bianchi identity dFa = — [Aa,Fa\. The Leibniz' rule yields d
= <j>{dFa, Fa,...,Fa)
= -<j>{[Aa, Fa], Fa,...,Fa)
+ --- +
>(Fa, ...,Fa,
[Aa, Fa})
(8
= 9 ) 0.
(b) Let A1 e 21(F) (i = 0,1) be defined by the collections
A^en^U^^g. Set Ca = A1a- A°a and for 0 < t < 1 we define Ala e H 1 ^ ) ® g by Ala = A°a + tCa. The collection (A*a) defines a connection A* e 21(F) and t H ^ ' e 21(F) is an (affine) path connecting A0 to A1. Note that C = (CQ) = A*. We denote by F* = (F£) the curvature of A f . A simple computation yields Fla = F° + i(dC Q + [A°a, Ca]) + y [C a , C J .
(8.1.10)
301
Chern-Weil theory Hence Fla = dCa + [A°a, Ca] + t[Ca, Ca] = dCa + K , Ca]. Consequently dt 0 ( O - 4>(K) = Si {*&>
n,-,**a)
+ -
+
= Si {ttdCa, F<,..., F<) + • • • + 4>{Fl ..., F<, dCa)}dt + jf * {<MK, c„], F%,-,K)
+ --- + *(Fl • • • > * t K . ca])} dt.
Because the algebra Ueven(Ua) is commutative we deduce for all a e &k and any wi,..., wfc € Q.even(Ua) ® fl. Hence ^ i ) -
We claim that
F<) -
^ d>(dCa +
[Ata,Ca},Fta,...,Fta)
Hence ^(Fi) - 4>(F°) = d /"' fc0(Co, J*,..., F^)di ./o = d Z"1 fc^.Pi,... ,Fl)dt HI dT^AlAl). (8.1.11) ./o Since C/j = gpaCag^ and F/j = g/3aFagaa o n ^a/9 w e conclude from the Ad-invariance of <j> that the collection T(Ala, A°) defines a global (2k — l)-form on M which we denote by T(Al,A°) and we name it the (/^transgression from A° to A1. We have thus established the transgression formula HF(A1)) - 4>(F(A0)) = dT+(A\A0). (8.1.12) The Chern-Weil theorem is proved.
•
302
Characteristic classes
Example 8.1.9 Consider a matrix Lie group G with Lie algebra g and denote by Po the trivial principal G-bundle over G, P0 = G x G. Denote by w the tautological 1-form w = g~xdg g Q1(G)
* = ¥^ ( / 0 1 ( * 2 " 1 )*" l r f i ) • ,
lVfc_,
^w'I07'07]''"'[vM)
fc 22fc-1A:!(fc - 1)' ofc
We thus have a natural map r : 7*(G) -> Hodd(G) called transgression. The elements in the range of r are called transgressive. When G is compact and connected then a nontrivial result due to the combined efforts of H. Hopf, C. Chevalley, H. Cartan, A. Weil and L. Koszul states that the cohomology of G is generated as an K-algebra by the transgressive elements. We refer to [18] for a beautiful survey of this subject. Exercise 8.1.3 Let G = SU(2). The Killing form K is a degree 2 Ad-invariant polynomial on su(2). Describe rK g fi3(G) and then compute
L
rK
G
Compare this result with the similar computations in Subsection 7.4.3. Let us now analyze the essentials of the Chern-Weil construction. Input: (a) A principal G-bundle P over a smooth manifold M (defined by an open cover (Ua) and gluing cocycle gap : Uap —>• G). (b) A connection A € 2l(P) defined by the collection Aa
gfi1(f/a)®0
303
Chern-Weil theory satisfying the transition rules
4s = 9a0dgaj3 + ga(jAagai3 on Uap. (c) <£ € Ik(G). Output: A closed form (f>(F(A)) e Q,2k(M) whose cohomology class is independent of the connection A. We denote this cohomology class by <j>(P). Thus, the principal bundle P defines a map, called the C h e r n - W e i l correspondence crop : I*{G) -> H*(M) 4>^<j>{P). One can check easily the map crop is a morphism of K-algebras. Definition 8.1.10 Let M and N be two smooth manifolds and F : M —> N a smooth map. If P is a principal G-bundle over N defined by an open cover (Ua) and gluing cocycle gap : Uap —• G then the pullback of P by F is the principal bundle F*(P) over M defined by the open cover F _ 1 ([/ t t ) and gluing cocycle F-l(Ua())
4 Ua0 9^ G.
The pullback of a connection on P is defined similarly. The following result should be obvious. Proposition 8.1.11 (a) If P is a trivial G-bundle over the smooth manifold M then <j>{P) = 0 € H*{M) for any > £ I*(G). (b) Let M —> N be a smooth map between the smooth manifolds M and N. Then for every principal G-bundle over N and any (f> e P{G) we have # f ( P ) ) = F*(«(P)). Equivalently, this means the diagram below is commutative. /*(G) C-Hf iT(AT)
\
IF-
cro F . ( P ) >v • H*(M) Denote by ^P G the collection of smooth principal G-bundles (over smooth manifolds). For each P G V($G we denote by Bp the base of P. Finally, we denote by T a contravariant functor from the category of smooth manifolds (and smooth maps) to the category of abelian groups.
304
Characteristic classes
Definition 8.1.12 An ^-valued G-characteristic class is a correspondence «p G 3 P n . c(P) e
HBp)
such that (a) c(P) = 0 if P is trivial and (b) J r (F)(c(P)) = c(F*(P)) for any smooth map F : M -¥ AT and any principal G-bundle P -> AT. Hence, the Chern-Weil construction is just a method of producing G-characteristic classes valued in the DeRham cohomology. Remark 8.1.13 (a) We see that each characteristic class provides a way of measuring the nontriviality of a principal G-bundle. (b) A very legitimate question arises. Do there exist characteristic classes (in the DeRham cohomology) not obtainable via the Chern-Weil construction? The answer is negative but the proof requires an elaborate topological technology which is beyond the reach of this course. The interested reader can find the details in the monograph [58] which is the ultimate reference on the subject of characteristic classes. (b) There exist characteristic classes valued in contravariant functors other then the DeRham cohomology. E.g., for each abelian group A the Cech cohomology with coefficients in the constant sheaf A defines a contravariant functor H*(—,A) and using topological techniques one can produce H*{—, j4)-valued characteristic classes. For details we refer to [58] or the classical [71].
8.2
Important examples
We devote this section to the description of some of the most important examples of characteristic classes. In the process we will describe the invariants of some commonly encountered Lie groups. §8.2.1
The invariants of the torus T"
The n-dimensional torus T" = (7(1) x • • • x U(l) is an abelian group so that the adjoint action on its Lie algebra t n is trivial. Hence F(Tn)
= S*((t n )*).
In practice one uses a more explicit description. This is obtained as follows. Pick angular coordinates 0 < 8' < 2n (1 < i < n) and set ~^-.d0]. 27ri The x'jS form a basis of (t™)* and now we can identify Xj1 =
/•(rjanin,..,^.
Important §8.2.2
305
examples C h e r n classes
Let E be a rank r complex vector bundle over the smooth manifold M. We have seen that a Hermitian metric on E induces an f/(r)-structure (P, p) where p is the tautological representation
p:U{r)^GL{r,C)Exercise 8.2.1 Prove that different Hermitian metrics on E define isomorphic {/(restructures. Thus, we can identify such a bundle with a principal U(r) bundle in a tautological way. A connection on the principal bundle is then equivalent with a linear connection V on E compatible with a Hermitian metric ( , ) i.e. V x (Aw, v) = A{(V x u, v) + (u,
Vxv)}
VA e C, u, v € C°°(E), X € Vect (M). The characteristic classes of E are by definition the characteristic classes of the tautological principal [/(r)-bundle. To describe these characteristic classes we need to elucidate the structure of the ring of invariants I*(U(r)). I*(U(r)) consists of symmetric, r-linear maps cf>: u(r) x • • • x u(r) -¥ K invariant under the adjoint action
u(r) 9 1 4 TXT'1
6 u(r), T € U(r).
It is convenient to identify such a map with its polynomial form P*{X) =
(X,...,X).
The Lie algebra u(r) consists of r x r complex skew-adjoint matrices. Classical results of linear algebra show that for any X € u(r) there exists T € U(r) such that TXT~l is diagonal TXT*1 = idiag(Xi,..., A r ). The set of diagonal matrices in u(r) is called the C a r t a n a l g e b r a of u(r) and is denoted by C„(r). It is a (maximal) abelian Lie subalgebra of u(r). Consider the stabilizer 9a(r) = {T€ U(r) ; TXT'1 = X VX € C„ (r) } and the normalizer 9lp ( p ) = {T € U(r) ; TC^T'1
C C„ (r) }.
306
Characteristic classes
5t/( r ) is a normal subgroup of ^Sluir) so we can form the quotient
W{U{r))dM mu{T)/$u(r) called the Weyl group of U(r). As in §7.4.4 we see that it is isomorphic with the symmetric group <5r because two diagonal skew-Hermitian matrices are unitarily equivalent if and only if they have the same eigenvalues (including multiplicities).We see that P^ is j4d-invariant if and only if its restriction to the Cartan algebra is invariant under the action of the Weyl group. The Cartan algebra is the Lie algebra of the (maximal) torus Tn (consisting of diagonal unitary matrices) and as in the previous subsection we introduce the variables X; = 3
-6Bi2?ri
3
The restriction of P$ to Ca(r) is a polynomial in the variables x\,...,xT. The Weyl group <5r permutes these variables so that P^ is j4d-invariant if and only if P$(xi,..., xr) is a symmetric polynomial in its variables. According to the fundamental theorem of symmetric polynomials the ring of these polynomials is generated (as an M-algebra) by the elementary ones Ci
=
E j Xj
Thus r(U(r))=R[c1,c2,...,cT}. In terms of matrices X G u{r) we have J2ck(X)tk
= det ( l - ^X)
G
r{U{r))[t).
The above polynomial is known as the universal Chern polynomial and its coefficients are called the universal Chern classes. Returning to our rank r vector bundle E we get the Chern classes ck(E) = ck(F(V))
G H2k(M)
and the Chern polynomial ct{E)
= det ( l - ^ T F ( V ) ) .
V denotes a connection compatible with a Hermitian metric ( , ) on E while F ( V ) denotes its curvature.
307
Important examples
Remark 8.2.1 The Chern classes produced via the Chern-Weil method capture only a part of what topologists usually refer to characteristic classes of complex bundles. To give the reader a feeling of what the Chern-Weil construction is unable to capture we will sketch a different definition of the 1st Chern class of a complex line bundle. The following facts are essentially due to Kodaira and Spencer [45]; see also [32] for a nice presentation. Let L —> M be a smooth complex Hermitian line bundle over the smooth manifold M. Upon choosing a good open cover (Ua) of M we can describe L by a collection of smooth maps zap : Uap —> f/(l) = S 1 satisfying the cocycle condition zaf}Zf}-,z-,a = 1 Va,/?,7.
(8.2.1)
If we denote by C°°(-,5 1 ) the sheaf of multiplicative groups of smooth S^-valued functions we see that the family of complex line bundles on M can be identified with the Cech group HX{M, C°°(-, S1)). This group is called the smooth Picard group of M. The group multiplication is precisely the tensor product of two line bundles. We will denote it by Pic°° (M). If we write zQ/3 = exp(27ri#Q/g) (8pa = — 9ap £ C°°(Ua0,R)) we deduce from (8.2.1) we deduce that V[/Q/37 ^ 0 #a/3 + #/37 + $ 7 a
=
napj
€ Z.
It is not difficult to see that Vf/Q/37j ^ 0 riffjs — nQ7<5 + naps — n a ^ 7 = 0. In other words nQ/g7 defines a Cech 2-cocycle of the constant sheaf Z. On a more formal level we can capture the above cocycle starting from the exact sequence of sheaves 0 - • Z -> C°°(-, R) exp^''") C°°(-, S 1 ) -> 0. The middle sheaf is a fine sheaf so its cohomology vanishes in positive dimensions. The long exact sequence in cohomology then gives 0 ->• Pic 00 (M) A H2(M, Z) ->• 0. The cocycle (na/37) represents precisely the class S(L). The class S(L), L e Pic 00 (M) is called the topological 1st Chern class and is denoted by cf(L). This terminology is motivated by the following result of Kodaira and Spencer, [45]: "The image of cf(L) in the DeRham cohomology via the natural morphism H*{M,Z)
-> H*{M,R) <* H*DR(M)
308
Characteristic classes
coincides with the 1st Chern class obtained via the Chern-Weil procedure.". The Chern-Weil construction misses precisely the torsion elements in H2(M,Z). For example if a line bundle admits a flat connection then its Chern class is trivial. This may not be the case with the topological one. The line bundle may not be topologically trivial. §8.2.3
Pontryagin classes
Let E be a rank r real vector bundle over the smooth manifold M. An Euclidean metric on E induces an 0(r) structure (P, p). The representation p is the tautological one p : 0(r) -> GL(r, M). Exercise 8.2.2 Prove that two metrics on E induce isomorphic 0(r)-structures. Hence, as in the complex case, we can naturally identify the rank r real vector bundles with principal <9(r)-bundles. A connection on the principal bundle can be viewed as a metric compatible connection in the associated vector bundle. To describe the various characteristic classes we need to understand the ring of invariants /•(O(r)). As usual we will identify the elements of lk(0(r)) with the degree k, Ad-invariant polynomials on the Lie algebra o(r) consisting of skewsymmetric r xr real matrices. Fix P e lk(0(r)). Set m = [r/2] and denote by J the 2 x 2 matrix J =
0 1
-1 0
Consider the Cartan algebra CO(T)
{A X J®--
© A m J 6 o(r) ; Aj 6 K} Am J © 0 S o(r) ; A^ 6 » }
r = 2m r = 2m + 1
Ce(r) is the Lie algebra of the (maximal) torus T
Re, © • • • ® Rem € 0(r) Re, © • • • © Rem © 1]K € °(r)
:2m > r = 2m + l
where for each 6 € [0,2ir] we denoted by Re the 2 x 2 rotation cos 0 — sin 9 sin 9 cos 9
Re =
As in Subsection 8.2.1 we introduce the variables XA = 3
——dBi.
2n
}
309
Important examples
Using standard results concerning the normal Jordan form of a skew-symmetric matrix we deduce that for every X € o(r) there exists T £ 0(r) such that TXT~l € C2(r)Consequently, any Ad-invariant polynomial on o(r) is uniquely defined by its restriction to the Cartan algebra. Following the approach in the complex case, we consider &>(r) = {T e 0(r) ; TXT'1 % )
= I V I e C„(r)}
= {T € O(r) ; T C ^ T - 1 C C B(r) }.
ffo(r) is a normal subgroup in 9 t ( 0 ( r ) ) so we can form the Weyl group 2»(0(r)) = Stow/SoMExercise 8.2.3 Prove that 2U(0(r)) is the subgroup of GL(m,R) involutions <Jij : \X\, . . • , X{, . . . , Xj, . . . , Xm)
generated by the
I V {Xij . . . , Xjj . . . , Xj, . . . , £ m J H-^ ^Xi, . . . , — Xj, • • • , ^ m j -
Sj : ( X i , . . . , £ j , . . . , XmJ fc
The restriction of P G 7 (0(r)) to C2(r) is a degree k homogeneous polynomial in the variables x\,...,xm invariant under the action of the Weyl group. Using the above exercise we deduce that P must be a symmetric polynomial P = P(xi,..., xm) separately even in each variable. Invoking once again the fundamental theorem of symmetric polynomials we conclude that P must be a polynomial in the elementary symmetric ones Pi = Y,j A ?2
=
Pm
—
Z/i<j xi
x
j
2 2 l' ' ' xm
x
Hence /*(0(r)) = K[p 1 ,... ) p [ r / 2 ]]. In terms of X e o(r) we have PtiX) = Y.PA*)t2i
= det ( l - ±X)
€
I'lO(r))[t\.
The above polynomial is called the universal Pontryagin polynomial while its coefficients Pj(X) are called the universal Pontryagin classes. The Pontryagin classes pi(E),.. .,pm(E) of our real vector bundle E are then defined by the equality
ME) = I > W
= det (l - ^ F ( V ) ) 6 H'{M)[t]
where V denotes a connection compatible with some (real) metric on E, while -F(V) denotes its curvature. Note that Pj(E) 6 Hij(M).
310 §8.2.4
Characteristic classes The Euler class
Let E be a rank r, real oriented vector bundle. A metric on E induces an 0(r)structure but the existence of an orientation implies the existence of a finer structure, namely an SO(r)-symmetry. The groups 0(r) and SO(r) share the same Lie algebra so(r) = o(r). The inclusion i : SO(r) ^
0{r)
induces a morphism of K-algebras i* : l'{0{r))
-+
I*{SO{r)).
Because so{r) = o(r) one deduces immediately that i* is injective. Lemma 8.2.2 When r is odd then i* : l*(0(r))
-» I*(SO(r))
is an isomorphism.
Exercise 8.2.4 Prove the above lemma. The situation is different when r is even, r = 2m. To describe the ring of invariants I*(SO(2m)) we need to study in greater detail the Cartan algebra
c2(2m) = {Ai J e • • • e \mJ e o(2m)} and the corresponding Weyl group action. The Weyl group 2F(50(2m)), defined as usual as the quotient W(SO(2m)
= 9ft so(2m) /3r S o(2m)
is the subgroup of GL(Co(2m)) generated by the involutions Oij : ( A i , . . . , A{,..., A j , . . . , Xm) — i > ( A i , . . . , \j,...,
A;,..., Amj
and where e"i,..., £m = ± 1 and E\ • • • em = 1. (Check this!) Set as usual Xi = — A;/27r. The Pontryagin invariants Pj(xu...,xm)=
(xii---xij)2
J2 l
continue to be 2U(50(2m)) invariants. There is however a new invariant A ( z i , . . . , z m ) = Y[XJ. i
In terms of
x = A1Je---eAmJecS0(2m)
Important
311
examples
we can write
where Pf denotes the pfaffian of the skewsymmetric matrix X viewed as a linear map M2m —> tt2m, when R 2m is endowed with the canonical orientation. Note that Pm = A 2 . Proposition 8.2.3 7*(50(2m)) 3 R[Zlt Z 2 , . . . , Zm; Y}/(Y2 - Zm)
(Zj = Pj,
Y = A)
where (Y2 — Zm) denotes the ideal generated by the polynomial Y2 — Zm. Sketch of proof We follow an approach used by H. Weyl in describing the invariants of the alternate group ([74], Sec. II.2). The isomorphism will be established in two steps. Step 1 I*(SO(2m)) is generated (as an M-algebra) by p i , . . . , p m , A. Step 2 The kernel of the morphism R[Zl,...,Zm,;Y}
A/*(50(2m))
denned by Zj >->• p3•„ Y i-> A is the ideal (Y2 - Zm). Proof of Step 1 Note that W{SO(2m)) has index 2 as a subgroup in 9B(0(2m)). Thus 2U(SO(2m)) is a normal subgroup and <8 = 2U(0(2m))/2n(SO(2m)) S Z 2 . & = {1, c} acts on I*{SO{2m))
by
(eF)(a; 1 ,a;2,...,x m ) = F{-xx,x2,
• • • ,xm) = ••• - F(xux2,
and moreover r ( 0 ( 2 m ) ) = ker(l - e). For each F € J*(SO(2m)) we have F+ d=f (1 + e)F € ker(l - e) so that F+ =
P(Pl,...,Pm).
On the other hand, F~ "^ (1 -
t)F
is separately odd in each of its variables. Indeed, F~{-x1,x2,---,xm)
= eF-(x1,...,zm)
•. •,
-xm)
312
Characteristic classes = e(l - e)F(xu
...,xm)
= -(l-
e)F(xu
...,xm)
= -F
{xu ...,
xm).
Hence, F~ vanishes when any of its variables vanishes so that F~ is divisible by their product A = xi • • • xm F~ = A-G. Since eF~ = -F~ and eA = - A we deduce eG = G i.e. G € 7*(0(2m)). Consequently, G can be written as G =
Q(p1,...,pm)
so that F~ =
A-Q(Pl,...,Pm).
Step 1 follows from F = ±(F+ + F-) = l-{P{Pl,... ,Pm) + A • Q(pu ..
.,Pm)).
Proof of Step 2 From the equality det X = Pf(X)2
VX 6 so{2m)
we deduce
(r 2 -Z m )Cker> so that we only need to establish the opposite inclusion. Let P = P{ZU Z2, • • •, Zm\ Y) € kerip. Consider P as a polynomial in Y with coefficients in R[Zi,..., Zm\. Divide P by the quadratic polynomial (in Y) Y2 — Zm. The remainder is linear R = A(Z1,...,Zm)Y
+
B(Z1,...,Zm).
Since Y2 — Zm, P 6 kerV' we deduce R € ker^. Thus A(pi, • • • ,p m )A + B(pu .. .,pm) = 0. Applying the morphism e we get -A{pu
... ,p m )A + B(Pl,...
Hence A = B = 0 so that P is divisible by Y2 - Zm.
,pm) = 0. •
Let E be a rank 2m, real, oriented vector bundle over the smooth manifold M. As in the previous subsection we deduce that we can use a metric to naturally identify E with a principal SO(2m)-bundle and in fact, this principal bundle is independent of the metric. Finally, choose a connection V compatible with some metric on E.
Important
313
examples
Definition 8.2.4 (a) The universal Euler class is defined by e = e(X) = _ L - P / ( _ A : )
e
7m(50(2m)).
(b) The Euler class of E, denoted by e(E) G H2m(M) represented by the 2m-form
is the cohomology class
e(V) = ( ^ V ( - n V ) ) (According to Chern-Weil theorem this cohomology class is independent of the metric and the connection.) Example 8.2.5 Let (S, g) be a compact, oriented, Riemann surface and denote by V s the Levi-Civita connection. The the Euler form £
(d) = -jzs{g)dvg 4ir~
9
coincides with the Euler form e(V ) obtained via the Chern-Weil construction. Remark 8.2.6 Let £ be a rank 2m, real, oriented vector bundle over the smooth, compact, oriented manifold M. We now have two apparently conflicting notions of Euler classes. A topological Euler class etop(E) 6 H2m(M) defined as the pullback of the Thorn class via an arbitrary section of E. A geometric Euler class egeom{E) e H2m(M) defined via the Chern-Weil construction. The most general version of the Gauss-Bonnet theorem, which will be established later in this chapter, will show that these two notions coincide! §8.2.5
Universal classes
In each of the situations discussed so far we discussed characteristic classes for vector bundles with a given rank. In this section we show how one can coherently present these facts all at once, irrespective of rank. The algebraic machinery which will achieve this end is called inverse limit. We begin by first describing a special example of inverse limit. A projective sequence of rings is a a sequence of rings {Rn}n>o together with a sequence of ring morphisms Rn t^ Rn+i. The inverse limit of a projective system (Rn, 4>n) is the subring
limRn C n # n n>0
consisting of the sequences (x\,x2, • • •) such that 4>„(xn+i) = xn, Vn > 0.
314
Characteristic classes
E x a m p l e 8.2.7 Let Rn = ^ [ [ X ^ . . .,Xn]] be the ring of formal power series in n variables with coefficients in the commutative ring with unit 9t. (RQ = 9t.) Denote by
FX(XX) • • •
Fn{Xn),...)
which defines an element in Dtfpfi, X2, •..]] denoted by Fi(Xi)F2(X2) •• •. When F\ = F2 = • • • Fn = • • • = F the corresponding elements is denoted by (F)°°. Exercise 8.2.5 (a) Let SK^a;]]1' denote the set of formal power series F 6 5H[[a;]] such that F(0) = 1. Prove that (fH[[x]]k, •) is an abelian group, (b) Prove that VF, G € ^{[xtf {F • G)°° = (F)°° • (G)°°
Similarly, given Gn E ^[[^jji {ji ^ 1) such that Fn(0} — 0 we can from the sequence of sums (0, Gi(Xi), GiiXJ
+ G2(X2),...,
G1(X1) + • • • e
Gn(Xn),...)
which defines an element in •SH[[X1,X2,.. •]] denoted by Gi(Xi) + G2(X2) -i . When G\ = G2 = • • • = Fn — • • • = F we denote the corresponding element (G)^. In dealing with characteristic classes of vector bundles one naturally encounters the increasing sequences £7(1) - » V{2) M- • • • (8.2.2) (in the complex case) and (in the real case) 0(1) ^ 0(2) - > • • • .
(8.2.3)
We will discuss these two situations separately. The complex case The sequence in (8.2.2) induces a projective sequence of rings R <- I"(U(l))
<- r*(U(2))
i
.
(8.2.4)
We know that I"(U(n)) = © n [fo]]= the ring of symmetric formal power series in n variables with coefficients in R. Set ©<»[[*,•]]= l i m © " .
315
Important examples
Given a rank n vector bundle E over a smooth manifold M and
(n > A;) the elementary symmetric polynomial in V"^ l
Then the sequence
(0,...,o,cik\c{r >,...)
defines an element in <5°°[[:rj]] denoted by c^ which we call the universal A;-th Chern class. Formally, we can write l
We can present the above arguments in a more concise form as follows. Consider the function F(x) = (1 + tx) € 9t[[a;]] where 9 t = R[i\. Then (F)°° defines an element in
{F)°°{xux2,...)
We can perform the above trick with any F £ M.[[x]] such that -F(O) = 1. We get a semigroup morphism
m[x}]\-)3F^(Fre(eea[[xj}],-). (see Exercise 8.2.5) One very important example is T
1
1
i
oo
r>
The coefficients B^ are known as the Bernoulli numbers. The product
316
Characteristic classes
defines an element in <5°°[[:r.,]] called the universal Todd class and denoted by Td. Using the fundamental theorem of symmetric polynomials we can write Td = 1 + Tdi + Td 2 + •• • where Td n G ©°°[[XJ]] is an universal symmetric, homogeneous "polynomial" of degree n, hence expressible as a combination of the elementary symmetric "polynomials" Ci,c2, — By an universal "polynomial" we understand element in the inverse limit M[x\,X2, • • •] = limR[a;i,a:2,...]. An universal "polynomial" P is said to be homogeneous of degree d if it can be represented as a sequence
P=(PUP2,...) where Pm is a homogeneous polynomial of degree d in m variables and Pm+l(xli
x
2, • • • > xmi 0 ) = Pm{X\,
...
Xm).
For example Tdi = -Cl,
Td 2 = —{c\ + c 2 ), Td 3 = — c l C 2 etc.
Analogously, any function G 6 K[[x]] such that G(0) = 0 defines an element &co[[xj]}
(GU = G(Xl) + G(x2) + •••€
We have two examples in mind. First, consider G(x) = xk. We get the symmetric function sk = (a;*)oo = (G)oo = x\ + x\ + • • • called the universal k-th power sum. We will denote these by s^. Next, consider G(x) = e 1 - 1 = V
=-.
m>l
We have
(C- - 1)„ = E i(^m)oc = E i»™ 6 e°°[fe']] m>l m-
m=l
m
-
Given a complex vector bundle E we define ch(jE;) = r a n k £ + (ex - !)«,(£). ch(E) is called the Chern character of the bundle E. If V is a connection on E compatible with some Hermitian metric then we can express the Chern character of E as c h ( £ ) = tr (eF
Important examples
317 00
1
= rank(£) + £ - i t r ( . F ( V r * ) k=i
K
-
where A is the bilinear map ft* (End (£)) x ft* (End (£)) -> Q i+J '(End {E)) defined in Example 3.3.11. Proposition 8.2.9 Consider two complex vector bundles E\,E2 over the same manifold M. Then ch{E1 © E2) = chiEi) + ch{E2) and ch(Ex ® E2) = ch(E1) • ch(E2) € H*{M).
Proof Consider a connection V 1 on Ei compatible with some Hermitian metric ht, i = 1,2. Then V 1 ©V 2 is a connection on E\®E2 compatible with the metric h\®h2 and moreover F(V 1 ffiV 2 ) = F ( V 1 ) © F ( V 2 ) . Hence exp(F(V 1 © V 2 )) = exp(F(V 1 )) © exp(F(V 2 )) from which we deduce the first equality. As for the second equality consider the connection V on E\ ® E2 uniquely defined by the Leibniz rule V(si ® s2) = (V x si) ® s2 + si ® (V 2 s 2 ),
Si
6 C00^)
where the operation
®: n*(Fi) x n £ (F 2 ) -» n*+/(Je1 ® F 2 ) is defined by (w* ® si) ® (7/'s2) = (w* A rf) ® (si ® s 2 ) Si S C°°(Fi), w* € fi*(M) and r/ e fi'(M). We compute the curvature of V using the equality F(V) = (d v ) 2 . If s, € then F ( V ) ( S l ® s2) = d v {(V 1 si) ® s 2 + S l ® (V 2 s 2 )} = {{F{Vl)Sl)
(8.2.5) Cx{Ei)
® s 2 - ( V ^ i ) ® (V 2 s 2 ) + ( V ^ ) ® (V 2 s 2 ) + s x ® (F(V 2 ))} = F(V1)®lfi2 + lBl®F(V2).
The second equality in the proposition is a consequence the following technical lemma.
318
Characteristic classes
Lemma 8.2.10 Let A (resp. B) be a skew-adjoint, n x n (resp. m x m) complex matrix. Then tr (exp(A
diag^eK)
so that tr ( 4 ® l C m + l c » ® B) = YleXieN The proposition is proved.
=
tr
( e/1 ) • t r ( e B )-
n
D
Exercise 8.2.6 ( N e w t o n ' s formulae) Consider the symmetric polynomials ck=
]T
xh---xik€R[xi,...,xn}
l
and (r > 0) 1)
•
•
• 1 • " •71J-
3
Set
n
- £.,'£).
/(*) = .7=1
(a) Show that 1 /'(*) = - £ •s.r- . /(*) r
(b) Prove the Newton formulae •iySr.
jC^- = 0
VI < r < n.
(c) Deduce from the above formulae the following identities between universal symmetric polynomials. Si=ci,
s 2 = c 1 - 2 c 2 , s 3 = Ci - 3cic 2 + 3c 3 .
Important
319
examples
The real case
The sequence (8.2.3) induces a projective system J**(0(l))<-I**(0(2))<
.
We have proved that J**(0(n)) = ©' n/2 '[[a^]]= the ring of even, symmetric power series in [n/2] variables. The inverse limit of this system is 7* * (O(oo) = l i m / " ( 0 ( r ) ) = ©<»[[:#] ^ lim © " [ [ : # ] . As in the complex case, any element of this ring is called a universal characteristic class. In fact, for any real vector bundle E and any 4> £ ©""M3^]] there is a well defined characteristic class <j>{E) which can be expressed exactly as in the complex case, using metric compatible connections. If F e lR[[z]]l> then (F(a;2))°° defines an element of ©°°[[a$]. In topology, the most commonly encountered situations are the following.
The universal characteristic class (F(x))°° is denoted by A and is called the A — genus. We can write A = 1 + Aj + A 2 + • • • where A^ are universal, symmetric, even, homogeneous "polynomials" and as such they can be described using universal Pontryagin classes
E
2 X
jl
2 X
' ' ' jm
•
1<J1 < - * »
The first couple of terms are Ai = ~ ,
A2 =
27,32.5(-4p2
+ 7p?) etc.
B . Consider
F(x) = v
^ _ = i + lx
tanh^/z
3
+
JL3? + ... = i +
45
^
'
X2(-i)^l^Bkxk. (2k)\
The universal class (F(x2))°° is denoted by L and is called the L-genus. As before we can write L = 1 + L! + L 2 + • • • where the L^'s are universal, symmetric, even, homogeneous "polynomials". They can be expressed in terms of the universal Pontryagin classes. The first few terms are Li = - P i , L 2 = —(7p 2 - p j ) etc. o 45
320
8.3
Characteristic classes
Computing characteristic classes
The theory of characteristic classes is as useful as one's ability to compute them. In this section we will describe some methods of doing this. Most concrete applications require the aplication of a combination of techniques from topology, differential and algebraic geometry and Lie group theory that go beyond the scope of this book. We will discuss in some detail a few invariant theoretic methods and we will present one topological result more precisely the Gauss-BonnetChern theorem. §8.3.1
Reductions
In applications, the symmetries of a vector bundle are implicitly described through topological properties. For example, if a rank r complex vector bundle E splits as a Whitney sum E = Ei ®E2 with rankjBj = Ti then E, which has a natural U(r)- symmetry, can be given a finer structure of U(ri) x C/(r2) vector bundle. More generally assume that a given rank r complex vector bundle E admits a Gstructure (P, p), where P is a principal G bundle and p : G —> U(r) is a representation of G. Then we can perform two types of Chern-Weil constructions: using the U(r) structure and using the G structure and in particular we obtain two collections of characteristic classes associated to E. One natural question is whether there is any relationship between them. In terms of the Whitney splitting E = Ei © E2 above, the problem takes a more concrete form: compute the Chern classes of E in terms of the Chern classes of E\ and E2- Our next definition formalizes the above situations. Definition 8.3.1 Let ip : H —> G be a smooth morphism of (matrix) Lie groups. (a) If P is a principal //-bundle over the smooth manifold M defined by the open cover ({/„) and gluing cocycle hap '• Uap —» H
then the principal G-bundle defined by the gluing cocycle ga0 =
is said to be the (^-associate of P and is denoted by (p(P). (b) A principal G-bundle Q over M is said to be (^-reducible if there exists a principal //-bundle P -» M such that Q =
->
r(H).
identities.
Computing characteristic classes
321
Proposition 8.3.2 Let P be a principal G-bundle which can be reduced to a principal ^-bundle Q. Then for every rj G ker
Proof
Denote by ip„ the differential of tp at 1 e H
Pick a connection [Aa) on Q and denote by (Fa) its curvature. Then the collection
v(
ck(E®F)=
Yl
(8.3.1) Equivalently, this means
ct(E)-Cj(F).
i+j=k
To check this, pick a Hermitian metric g on E and a Hermitian metric h on F. g®h is a Hermitian metric on E © F. Hence, E © F has an U(r + s) structure reducible to an U(r) x U(s) structure. The Lie algebra of U(r) x U(s) is the direct sum u(r) © u(s). Any element X in this algebra has a block decomposition XE X =
Xr 0
0 X.,
= xr®xs
where XT (respectively Xs) is an r x r (resp. sxs) complex, skew-adjoint matrix. Let i denote the natural inclusion u(r) ©«(s) <-> u(r + s) and denote by 4 " ' e I*(U{v))[t] the Chern polynomial.
322
Characteristic classes
We have
i*(c[r+s))(xr e x.) = det (i r + , - ^Xr e x.) =
det(l
r
-^X
r
).det(l
s
-^
s
)
The equality (8.3.1) now follows using the argument in the proof of Proposition 8.3.2. Exercise 8.3.1 Let E and F be two complex vector bundles over the same manifold M. Show that Td{E ®F) = Td(E) • Td(F). Exercise 8.3.2 Let E and F be two real vector bundles over the same manifold M. Prove that pt{E®F)=Pt{E)-Pt{F) (8.3.2) where pt denotes the Pontryagin polynomials. Exercise 8.3.3 Let E and F be two real vector bundles over the same smooth manifold M. Show that L{E®F) = L(E)-L(F) A{E®F)
=
A{E)-A{F).
Example 8.3.4 The natural inclusion R" •-+ C induces an embedding i : 0(n) °-> U(n). (An orthogonal map T : R n —»• R n extends by complexification to an unitary map T(£ : C n —> C n ). This is mirrored at the Lie algebra level by an inclusion o(n) ^-¥ u(n) and we obtain a morphism : P(U(n))
i
-+
P{0(n))
We claim that j*(c2fc+i) = 0
and
t'Coa) = (-I)*P*. Indeed, for X G o(n) we have ,
1
x 2/fc+l «+l
•(c2,+1)(X)= (- — ) \
Z7T1/
£ l<:,<-
^PO-.-'WA-)
Computing characteristic classes
323
where Xj(X) are the eigenvalues of X over C. Since X is in effect a real skewsymmetric matrix we have
XiiX) = \^X)
=-\j(X).
Consequently i*(c2k+1)(X)
= i*(c2k+l(X))=
(-—) \
= (-1)2fc+1 (-i) 2 f c + 1 \
Z7T1/
£
Z7T1/
s
\h(X)-..\i2k+l(X)
l
* . w • • • w*)=-»•(<*+!)(*)•
l
The equality i*(c2k) = {—l)hpk is proved similarly. From the above example we deduce immediately the following consequence. Proposition 8.3.5 If E -> M is a real vector bundle and E®C is its complexification then = (-l)kc2k(E®C), A = 1,2,.... (8.3.3) Pk(E) In a more concentrated form this means pt{E)=p_t(E)
=
cit{E®C).
Exercise 8.3.4 Let E —• M be a complex vector bundle of rank r. (a) Show that ck(E*) = (-l)kck{E)
i. e. ct(E') =
c.t(E).
(b) One can also regard E as a real, oriented vector bundle E^. Prove that
E(-i)^ 2 V(%) = ct(£;)-c_((s) k
i.e. p i t ( £ t t ) = ct{E) • c_ ( (£). and moreover Cr(E)=e(Em).
Exercise 8.3.5 (a) The natural morphisms I**(U(r)) -> l"(0(r)) induce a morphism $oo : I"(U(oo)) -+ / " ( O ( o o ) ) .
described above
324
Characteristic classes
As we already know, for any F e K[[a;]]b, (-F)00 is an element of P*{U(oo)) and if moreover F is even then (F)°° can be regarded as an element of I**(0(oo)). Show that for every F € ffi[[i]]1' $0O((F)co)
= (F • F-)°° G 7"(0(oo))
where F~(x) = F{-x). (b) Let £ be a real vector bundle. Deduce from part (a) that Td(E®C) =
A(E)2.
E x a m p l e 8.3.6 Consider the inclusion i: SO{2k) x SO{2l)) ^ SO{2k + 21). This induces a ring morphism i* : r{SO(2k
+ 21) -> r(SO{2k)
x SO(2£)).
Note that r{SO{2k)
x 50(2^)) ^ /*(S0{2k)) ® R /*(S0(2£).
Denote by e'"' the Euler class in /* (SO (2;/)). We want to prove that
Let X = Xk © Xt € so(2fc) © so(2£). Modulo a conjugation by (S,T) 6 50(2fc) x SO(2£) we may assume that Xk = A i J e - - - © A f c J and A^ = jUiJffi •••® mJ where as usual J denotes the 2 x 2 matrix
JJ = [[°1 - 01 " We have S(e^)(X)
= (-ij
Ax • • • A* • W • • • W = eW(Xk)
= eW®eM(X f c ®^).
•
e^(Xe)
325
Computing characteristic classes The above example has an interesting consequence.
Proposition 8.3.7 Let E and F be two real, oriented vector bundle over the same manifold M and having even ranks. Then e(E 0 F) = e(E) • e(F) where as usual the • denotes the A-multiplication in
Heven(M).
Example 8.3.8 Let £ be a rank 2k real, oriented vector bundle over the smooth manifold M. We claim that if E admits a nowhere vanishing section f then e(E) = 0. To see this, fix an Euclidean metric on E so that E is now endowed with an S'0(2A;)-structure. Denote by L the real line subbundle of E generated by the section £. Clearly, L is a trivial line bundle and E splits as an orthogonal sum E = L © L^. The orientation on E and the orientation on L defined by. £ induce an orientation on L1- so that L1- has an SO(2k — restructure. What we have just said shows that the SO(2k) structure of E can be reduced to an 5 0 ( 1 ) x SO(2k - 1) =* SO(2k - l)-structure. Denote by i* the inclusion induced morphism
r{so{2k)) -> r(so(2k -1)). Since i*(e(fc)) = 0 we deduce from the Proposition 8.3.2 that e(E) = 0. The result proved in the above example can be reformulated more suggestively as follows. Corollary 8.3.9 Let E be a real oriented vector bundle of even rank over the smooth manifold M. If e(E) ^ 0 then any section of E must vanish somewhere on M! §8.3.2
The Gauss-Bonnet-Chern theorem
If E —> M is a real oriented vector bundle over a smooth, compact, oriented manifold M then there are two apparently conflicting notions of Euler class naturally associated toE. The topological Euler class e«op(-E) = QTE where TE is the Thorn class of E and £0 : M —>• E is the zero section. The geometric Euler class „ (p^-f egeam(*)-{
(£yPf(-FW) Q
if rank (E) is even if rank (S) is odd
where V is a connection on E compatible with some metric and 2r = r a n k ( £ ) . The next result, which generalizes the Gauss-Bonnet theorem, will show that these two notions of Euler class coincide.
326
Characteristic classes
Theorem 8.3.10 (Gauss-Bonnet-Chern)Let E A M be a real, oriented vector bundle over the compact oriented manifold M. Then etop(E) =
egeom{E).
Proof We will distinguish two cases. A rank (E) is odd. Consider the automorphism of E \:E-+E
«4-«Vue£.
Since the fibers of E are odd dimensional we deduce that i reverses the orientation in the fibers. In particular, this implies KXTE
= -7T,T£ =
7T,(-TE)
where IT, denotes the integration along fibers. Since 7r» is an isomorphism (Thom isomorphism theorem) we deduce i*TE = -TEHence etop(E) = -C0*iVB.
(8.3.4)
On the other hand notice that Coi* = CoIndeed Coi* = (Ko)* = (-Co)* = ( 0 * (Co = -Co)The equality etop = e 9eom now follows from (8.3.4). B . rank (E) = 2k. We will use a variation of the original argument due to Chern ([19]). Let V denote a connection on E compatible with a metric g. The strategy of proof is very simple. We will explicitly construct a closed form u> 6 Q!^t(E) such that (i) 7r„w = 1 e n°(M). (ii) Qu, = e(V) = (2w)-kPf(-F(W)). The Thom isomorphism theorem coupled with (i) implies that u> represents the Thom class in H^t(E). (ii) simply states the sought for equality e top = egeom. Denote by S(E) the unit sphere bundle of E S(E) = {ueE;
\u\g = 1}.
S(E) is a compact manifold and dim S{E) = d i m M + 2 f c - l .
327
Computing characteristic classes
Denote by ir0 the natural projection S(E) —¥ M and by TT^E) —> S(E) the pullback of E to S(.E) via the map •KQ. TTQ{E) has an 50(2A:)-structure and moreover it admits a tautological, nowhere vanishing section T : S(EX) Be^eeEx
= K ( £ ) x ) e (x G M).
Thus according to Example 8.3.8 we must have ege0m«(E))
= 0G
H2k(S(E))
where e9eom(7TQ^) denotes the differential form e9eom(7roV) = J L p / ( - F ( j r ; v ) ) . Hence there must exist V 6 fl2A_1(S(.E)) such that dip =
egeam{ixlE).
The decisive step in the proof of Gauss-Bonnet-Chern theorem is contained in the following lemma. Lemma 8.3.11 There exists * = # ( V ) G
fi2*"1^^))
such that
d*(V) = egeom(Tr*0(E))
(8.3.5)
and /
W(V) = - 1 G
fi°(M).
(8.3.6)
JS(E)/M
The form \1/(V) is sometimes referred to as the global angular form of the pair (E, V). For the clarity of the exposition we will conclude the proof of the GaussBonnet-Chern theorem assuming Lemma 8.3.11 which will be proved later on. Denote by r : E —> M+ the norm function E Be
H->
|e|j
If we set E° = E \ {zero section} then we can identify £°^(0,oo)xS(£)
e^(\e\±e).
Consider the smooth cutoff function p = p(r) : [0, oo) -J- R
328
Characteristic classes
such that p(r) = - 1 for r € [0,1/4] and p(r) = 0 for r > 3/4. Finally define u, =
W (V)
= -p'(r)dr
A tf (V) - p(r)7r*(e(V)).
w is well defined since p'(r) = 0 near the zero section. Obviously w has compact support on E and satisfies the condition (ii) since CJw = - p ( 0 X o V e ( V ) = e(V). From the equality f
JE/M
p(r)7r*e(V) = 0
we deduce f
JE/M
ui = - [
JE/M
p'Mdr A * ( V )
=- r p'(r)dr-1 Jo
*( v ) JS(E)/M
*(V)(8^6)1.
= -(p(l)-p(0))/ JS(E)/M
To complete the program outlined at the beginning of the proof we need to show that u) is closed. duj = p'{r)dr A d*(V) - p'(r) A 7r*e(V) (8
= 5 ) p'(r)dr A {7rJe(V) - 7r*e(V)}.
The above form is identically zero since 7Tge(V) = 7r*e(V) on the support of p'. Thus ui is closed and the theorem is proved. • P r o o f of L e m m a 8.3.11 We denote by V the pullback of V to ir^E. The tautological section T : S(E) —>• ir^E can be used to produce an orthogonal splitting •KQE =
L®L±
where L is the line bundle spanned by T while LL is its orthogonal complement in •KQE with respect to the pullback metric g. Denote by P : 7r0*£ -> ir*0E the orthogonal projection onto LL. Using P we can produce a new metric compatible connection V on TTQE by
V = (trivial connection on L) © PVP. We have an equality of differential forms ^ e ( V ) = e(V) =
1 (2TT)
:P/(-F(V)).
Computing characteristic classes
329
Since the curvature of V splits as a direct sum F(V) = 0 © F'(V) where F'(V) denotes the curvature of V \LJ. we deduce P/(F(V)) = 0. We denote by V the connection V + t(V - V) so that V° = V and V1 = V. If F1 is the curvature of V we deduce from the transgression formula (8.1.11) that 7 r 0 *e(V)=e(V)-e(V)
=d{{^)kkLlpfv^>Ft>-->Ft)dt}We claim that the form
*(V) = ( ^ ) kkfolPf(V-V,F\...,
F*)dt
satisfies all the conditions in Lemma 8.3.11. By construction d-Jr(V) = 7r*e(V) so all that we need to prove is [
JS(E)IM
It suffices to show that for each fiber Ex of E we have / ^ ( V ) = -l. Along this fiber TTQE is naturally isomorphic with a trivial bundle TT;E\EX^(EXXEX^EX).
Moreover V restricts as the trivial connection. By choosing an orthonormal basis of Ex we can identify w^E \EI with the trivial bundle R2* over R2fc. The unit sphere S(EX) is identified with the unit sphere S2k~x C R2*. The splitting L © LL over S(E) restricts over S{EX) as the splitting R2* = v © TS210-1 where v denotes the normal bundle of S2*""1 •-* R2*. The connection V is then the direct sum between the trivial connection on v and the Levi-Civita connection on
TS2k-\
330
Characteristic classes
Fix a point p € 5 2 f c _ 1 and denote by ( x 1 , . . . ,x 2 * - 1 ) a collection of normal coordinates near p such that the basis ( ^ | p ) is positively oriented. Set d{ = ^ 7 for % = 1 , . . . , 2k — 1. Denote the unit outer normal vector field by d0. For a = 0 , 1 , . . . , 2fc — 1 set fa = da\p. The vectors fa form a positively oriented orthonormal basis of K2*. We will use Latin letters to denote indices running from 1 to 2k — 1 and Greek letters indices running from 0 to 2k — 1. Vida = ( V A ) " + (V { cU T where the superscript v indicates the normal component while the superscript r indicates the tangential component. Since at p 0 = Via,- =
(VidjY
we deduce Vidj = (Vidj)"
at p.
Hence
Vja,- = (Via„a0)ao = -(a,-, Vi&jaaRecall that V is the trivial connection in M2* and we have Vido\P= (-gfdo I | p = f, = di\p. Consequently Vidj = ~5jid0 2k x
If we denote by 9' the local frame of TS ~ equality as
at p.
to (3,) then we can rephrase the above
vdj = -(e1 + --- + e2k-1)®d0. On the other hand Vid0 = dt i.e.
va 0 = 91 ® ai + • • • + 0 2 * -1 ® a2t_!.
Since (x 1 , •'• •, x2k x) are normal coordinates with respect to the Levi-Civita connection V we deduce that Vda = 0, Va so that
A=(V-V)|„=
0 01 92 6>2*-
924-1
0 0
Computing characteristic classes
331
Denote by F° the curvature of V° = V at p. Then
F° = 0 e R where R denotes the Riemann curvature of V at p. The computations in Example 4.2.13 show that the second fundamental form of the embedding S2k-i
^
R
»
coincides with the induced Riemann metric (which is the first fundamental form). Using Teorema Egregium we get (R{di, dj)dk, dt) = SuSjk - 6ik5jeIn matrix format we have F° = 0 0 ( % ) l
where fi^ = 6 A (P. The curvature Fl at p of V ' = V + tA can be computed using the equation (8.1.10) of subsection 8.1.4 and we get Fl = F° + t2A A A = 0 © (1 We can now proceed to evaluate *
(V).
*(V) lp= O" kll = {^)k
t2)F°.
Pf{A (1
' " t2)F°'''''(1 " t2)F°)dt
k ^J\l - t2)*"1*) Pf(A, F°, F°, • • •, F°).
(8.3.7)
We need to evaluate the pfaffian in the right-hand-side of the above formula. Set for simplicity F = F°. Using the polarization formula and Exercise 2.2.19 in Subsection 2.2.4 we get "J (A, r, r, • • •, r J =
2__, £((y)Aaoai-t'o2cr3 ' ' ' ^
.
"£&2k
For i = 0,1 define & = {a e &a
; at = 0}.
We deduce 2kk\Pf(A,F,---,F) +{-l)k
= {-l)k
Y,
^e1
£ rt6°
^ ^ ( - ^ ' J A ^ 2 A ^ 3 A---A6C 7 "- 2 A ^ " " 1
f W ^ A f
A0" 3 A - - - A I T " - 2 A P - 1 .
332
Characteristic classes
For each a € G° we get a permutation 4>: (0-1,172, •••,CT2k-l)
€ ©2*-l
1
such that e(o-) = e(<£). Similarly, for a € (5 we get a permutation
such that e(cr) = e(<j>). Hence 2kk\Pf{A,F,---,F)
= 2{-l)k+1
*T, e ( ^ A
0*""1
= 2(-l)* +1 (2A;-l)!0 1 A---A0 2 * : - 1 2(-l)k+1dvolff,k-i.
=
Using the last equality in (8.3.7) we get
*(V) |„= ( ^ ) * * ( j f ( l - t 2 )* - 1 *) • 2(-l)*+1(2fc -
l)\dvols^
2irk
Using the Exercise 4.1.15 we get that f and consequently
/^.•w-ifc-.^GCd-fl"*)
-^-D.Gf'1-*1^)The above integral can be evaluated inductively using the substitution t = cos tp Ik = (/ o X (l - t 8 )*- 1 *) = [
\**
= (cos^) 2fc - 2 sin^|J /2 + (2* - 2) I"1*(cas
= (2k - 1)4_! - (2* - 2)4 so that Ik
_2fc-3
-2k=lIk-1-
One now sees immediately that / JS2k-l
Lemma 8.3.11 is proved.
•.
*(V) = - 1 .
333
Computing characteristic classes
Corollary 8.3.12 (Chern [Ch])Let {M,g) be a compact, oriented Riemann Manifold of dimension 2n. If R denotes the Riemann curvature then
The next exercises provide another description of the Euler class of a real oriented vector bundle E —>• M over the compact oriented manifold M in terms of the homological Poincare duality. Set r = rank E and let TE be a compactly supported form representing the Thom class. Exercise 8.3.6 Let $ : N —> M be a smooth map, where N is compact and oriented. We denote by $ * the bundle map $*E -> E induced by the puUback operation. Show that <3>*TE = ( $ # ) * T £ € Qr($*E) is compactly supported and represents the Thom class of $*£. In the next exercise we will also assume M is endowed with a Riemann structure. Exercise 8.3.7 Consider a nondegenerate smooth section s of E i.e. (i) Z = s _ 1 (0) is a codimension r smooth submanifold of M. (ii) There exists a connection V° on E such that V°xs / 0 along Z for any vector field X normal to Z (along Z). (a) (Adjunction formula) Let V be an arbitrary connection on E and denote by Nz the normal bundle of Z H M. Show that the adjunction map
a v : Nz
-+E\z
defined by X M- Vxs
X e C°°(7VZ)
is a bundle isomorphism. Conclude that Z is endowed with a natural orientation. (b) Show that there exists an open neighborhood 9 t of Z «-• Nz <-> TM such that exp : 91 -¥ exp(9t) is a diffeomorphism. (Compare with Lemma 7.3.30.) Deduce that the Poincare dual of Z in M can be identified (via the above diffeomorphism) with the Thom class of Nz. (c) Prove that the Euler class of E coincides with the Poincare dual of Z .
334
Characteristic classes
The part (c) of the above exercise generalizes the Poincare-Hopf theorem (see Corollary 7.3.33). In that case the section was a nondegenerate vector field and its zero set was a finite collection of points. The local index of each zero measures the difference between two orientations of the normal bundle of this finite collection of points: one is the orientation obtained if one tautologically identifies this normal bundle with the restriction of TM to this finite collection of points while and the other one is obtained via the adjunction map. In many instances one can explicitly describe a nondegenerate section and its zero set and thus one gets a description of the Euler class which is satisfactory for most topological applications. Example 8.3.13 Let r n denote the tautological line bundle over the complex projective space OP". According to Exercise 8.3.4 ci(r„) = e ( r „ ) when we view r n as a rank 2 oriented real vector bundle. Denote by [H] the (2n — 2)cycle defined by the natural inclusion i : OP"" 1 ^ OP",; [z0 : . . . *„_!] H-> [z0 : . . . : zn^
: 0] € OP".
We claim that the (homological) Poincare dual of Ci(r n ) is —\H]. We will achieve this by showing that Ci(r*) = —Ci(Tn) is the Poincare dual of [H]. Let P be a degree 1 homogeneous polynomial P G C[zo,.. .,zn]. For each complex line L <-> C" + 1 the polynomial P defines a complex linear map L —¥ C and thus an element of L* which we denote by P\i. We thus have a well defined map OP"9Z,H->P|LeL* = < | L and the reader can check easily that this is a smooth section of T£ which we denote by [P]. Consider the special case P 0 = zn. The zero set of the section [P0] is precisely the image of [H]. We let the reader keep track of all the orientation conventions in Exercise 8.3.7 and conclude that CX(T*) is indeed the Poincare dual of [H]. Exercise 8.3.8 Let Us = S2\ {north pole} and UN = S2\ {south pole} ^ C. The overlap UN n Us is diffeomorphic with the punctured plane C* = C \ {0}. For each map g : C* -> U(l) = S 1 denote by Lg the line bundle over S 2 defined by the gluing map
gNS-UNnUs^U(l),
gNS{z) = g{z).
Show that where degg denotes the degree of the smooth map |sicC'—* S1-
Chapter 9 Elliptic equations on manifolds Almost all the objects in differential geometry are defined by expressions involving partial derivatives. The curvature of a connection is the most eloquent example. One is often led to studying such objects with specific properties. For example, we inquired whether on a given vector bundle there exist flat connections. This situation can be dealt with topologically, using the Chern-Weil theory of characteristic classes. Very often, topological considerations alone are not sufficient and one has look into the microstructure of the problem. This is where analysis comes in and more specifically, one is led to the study of partial differential equations. Among them, the elliptic ones play a crucial role in modern geometry. This chapter is an introduction to this vast and dynamic subject which has numerous penetrating applications in geometry and topology.
9.1 §9.1.1
Partial differential operators: algebraic aspects Basic notions
We first need to introduce the concept of partial differential operator (p.d.o. for brevity) on a smooth manifold M. To understand what we are looking for we begin with the simplest of the situations, M = MN. Perhaps the best known partial differential operator is the Laplacian A : C°°{MN) -> C^QR")
Au = - £ . 9 i 2 u i
where as usual 9; = -g^-. This is a scalar operator in the sense it acts on scalar valued functions. Note that our definition of the Laplacian differs from the usual one by a sign. The Laplacian defined as above is sometimes called the geometric Laplacian. Next in line is the exterior derivative d:ttk(RN)^-Slk+l(RN). 335
336
Elliptic equations on manifolds
This is a vectorial operator in the sense it acts on vector valued functions. A degree k form u> on R " can be viewed as a collection of (\) smooth functions or equivalently, as a smooth function UJ : M.N —> M\k'. Thus d can be viewed as an operator
d : C°°{RNM"h
->
C O O (R J V ,K(*+0).
It is convenient to think of C°°(RN, R") as the space of smooth sections of the trivial bundle R" over RN For any smooth K=R, C-vector bundles E, F over a smooth manifold M we denote by Op(E, F) the space of K-linear operators C°°(£) ->
C^iF).
Op(E,E) is an associative K-algebra. The spaces C°°{E), C°°{F) are more than just K-vector spaces. They are modules over the ring of smooth functions C°°(M). The partial differential operators are elements of Op which interact in a special way with the above C°°(M)-module structures. First define PDO°{E,F) = Horn {E,F). Given T € P D O ° , u € C°°(E) and / e C°°{M) we have T(fu)
- f(Tu)
=0
or, in terms of commutators [T,f}u = T{fu)-f(Tu)
= 0.
(9.1.1)
Each / € C°°(M) defines a map ad(f):Op(E,F)^Op(E,F) by ad{f)T = Tof-foT=[T,f]
\/TeOp(E,F).
Above, / denotes the C°°(M)-module multiplication by / . equality (9.1.1) as P D O ° ( £ , F ) = {TeOp
{E,F) ; ad(f)T
We can rephrase the
= 0 V/ e C™{M)} = ; kerad.
Define PDO<m>(£;, F) = kero
Algebraic aspects
337
Remark 9.1.1 Note that we could have defined P D O ( m ) inductively as P D O ( m ) = {TeOp;
[T, f] G P D O ' 7 " - 1 ' V/ G C°°(M)}.
This point of view is especially useful in induction proofs. Example 9.1.2 Denote by R the trivial line bundle over RN. The sections of R are precisely the real functions on RN. We want to analyze P D O ( 1 ) = P D O ( 1 ) ( l , R ) . Let L e P D O ^ , u, f € C°°(R Ar ). Then
[L,f]u = where cr(/) G C°°(RN).
o(f)-u
On the other hand, for any / , g G C°°{RN)
o{fg)u = [L, fg}u = [L, f](gu) + f([L, g]u) = a{f)g • u + fa{g) • u. Hence o{jg) = o{f)g + fcr(g). In other words the map / i-> a(f) is a derivation of C°°(RN) and consequently (see Exercise 3.1.2) there exists a smooth vector field X on RN such that a(f)=X-f, V/eC°°(Rw). Let n = L(l) G C°°{RN).
Then for all u G C 0o (R /v ) we have
Lu = L(u • 1) = [L, u] • 1 + u • L(l) = X -u + fx-u.
Lemma 9.1.3 Any L G P D O ( m ) ( £ , F) is a local operator i.e. Vu G C°°{E) supp Lu C supp u.
Proof We argue by induction over m. For m — 0 the result is obvious. Let L G P D O ( m + 1 ) and u G C°°(£). For every / G C°°(M) we have L{fu) = [L,f]u + fLu. Since [L, / ] G P D O ( r a ) we deduce by induction s u p p l ( / w ) C supp u U supp / V / G C°°(M). For any open set O such that 0 3 suppu we can find / = 1 on suppu and / = 0 outside O so that fu = u). This concludes the proof of the lemma. • The above lemma shows that in order to analyze the action of a p.d.o. one can work in local coordinates. Thus understanding the structure of an arbitrary p.d.o. boils down to understanding the action of a p.d.o. in P D O ( m ) ( K p , Kq) where Kp and Kg are trivial K-vector bundles over RN. This is done in the exercises at the end of this subsection.
338
Elliptic equations on manifolds
Proposition 9.1.4 Let E, F, G be smooth K-vector bundles over the same manifold M.lIPe P D O ( m ) ( F , G ) and Q e P D O ( n ) ( £ , F ) then PoQ e P D O ( m + n ) ( £ , G ) . Proof We argue by induction over m + n. For m + n = 0 the result is obvious. In general, if / € C°°(M) [PoQ,f\
= [P,f]oQ
+
Po[Q,f\.
By induction, the operators in the right-hand-side have orders < m + n — 1. The proposition is proved. • Corollary 9.1.5 The operator L = Y,
°a(*)d Q : C°°(RN) -> C°°{RN)
{da = d?1 • • • d%N)
\a\<m
(a = (QI, . . . , a if) 6 Z J ,
\a\ = Y, OH) is a P-d.o. of order < m.
Proof According to the computation in Example 9.1.2 each partial derivative <% is a 1st order p.d.o. According to the above proposition multiple compositions of such operators are again p.d.o.'s. D Lemma 9.1.6 Let E, F —> M be two smooth vector bundles over the smooth manifold M. Then for any P 6 P D O (E, F) and any / , g € C°°(M) ad(f) • (ad(g)P) = ad(g) • (ad(f)P). Proof ad(f) • (ad(g)P) = [[P, g], / ] = [[P, / ] , g] + [P, [/, g] = [[P,f},g] = ad(g)-(ad(f)P).
a
From the above lemma we deduce that if P G P D O ( m ) then for any C {M) the bundle morphism ca
ad(h)ad{f2)---ad{fm)P does not change if we permute the / ' s . Proposition 9.1.7 Let P 6 PBO{m){E,F), that at a point x0 G M dfi{x0) = dgi(x0) £ T*oM
/ i ) f t e C°°{M) (i = l , . . . , m ) such
Vi = 1 , . . . , m.
Then {ad{fi)ad{f2)
• • • ad{fm)P)
| I 0 = {ad{gx)ad{g2) • • • ad{gm)P} \Xo .
Algebraic aspects
339
In the proof we will use the following technical result which we leave to the reader as an exercise. Lemma 9.1.8 For each XQ € M consider the ideals of C°°(M) m I 0 = {/ e C°°(M) • f(x0) = 0} 3 I 0 = {/ e C°°{M) ; f(x0) = 0, df(x0) = 0} Then 3xo =
m
lo
i.e. any function / which vanishes at x0 together with its derivatives can be written as / = X)9jhj 9j, hj £ m I 0 . i Exercise 9.1.1 Prove the above lemma. Proof
Let P 6 P D O ( m ) and ft, g{ € C°°(M) such that dfi(x0) = dgi{x0)
Vi = 1 , . . . , m.
Since ad(const.) = 0 we may assume (eventually altering the / ' s and the g's by additive constants) that fi(x0) = 9i(xQ)
Vi
We will show that {ad(f1)ad(f2)
• • • ad(fm)P}
| x o = {ad(9l)ad{f2)
• • • ad{fm)P}
\xo .
Iterating we get the desired conclusion. Let
and set Q = ad(f2) • • • ad(fm)P
G P D O ( 1 ) . We have to show that
{ad{
a
3,0j€mxo-
3
We have {adf»Q}U,=
l^adiajP^Q
(9.1.2)
340
Elliptic equations on manifolds
= £{[Q, «,-]&} L + £ K [ Q , £•]} \xo= 0. J i This proves the equality (9.1.2) and hence the proposition.
•
The proposition we have just proved has an interesting consequence. Given P G PDO ( m ) (£> F ) , z 0 6 M and £ e C°°(M) (i = 1 , . . . , m) the linear map i^ad{fi)
• • • ad(fm)P}
\xo: Exo - • FXo
depends only on the quantities £ = dfi(x0) £ T*oM. Hence for any f* G T*oM (i = 1 , . . . , m) the above expression unambiguously induces a linear map a(P)(Zu...,U)(P)--EX0^FX0 which moreover is symmetric in the variables &. Using the polarization trick of Chapter 8 we see this map is uniquely determined by the polynomial
a(P)(fl = am(P)(0 =
aiP)^^). m
If we denote by n : T*M -» M the natural projection then for each P G P D O ( m ) ( F , F) we have a well defined map am(P){-)eRom{n*E,TT*F) where n*E and n*F denote the pullbacks via n. Along the fibers of T*M am{P){^) looks like a degree m homogeneous "polynomial" with coefficients in Horn {Exo, Fxo). Proposition 9.1.9 Let P G P D O ( m ) ( £ , F ) and Q G PDO ( n ) (F,G). Then 0-m+n(Q ° P) = O-n(Q) °CT m (P).
Exercise 9.1.2 Prove the above proposition. Definition 9.1.10 A p.d.o. P G P D O ( m ) is said to have order m if am{P) ^ 0. In this case am{P) is called the (principal) symbol of P . The set of p.d.o.'s of order m will be denoted by P D O m . Definition 9.1.11 Let P G P D O m ( £ , F). P is said to be elliptic if for any x G M and any ZeT*M\ {0} om{P){0 : Ex-> Fx is a linear isomorphism.
Algebraic aspects
341
The following exercises provide a complete explicit description of p.d.o.'s on RN. Exercise 9.1.3 Consider the scalar p.d.o. on RN described in Corollary 9.1.5 a
a[x)da-
£= E \a\<m
Show that
om{pm= E M*)r= E o-W 1 •••&"• |a|<m
|a[<m
N
N
Exercise 9.1.4 Let C : C°°{M ) ->• C°°{®L ) be a p.d.o. of order m. Its principal symbol has the form ffm(i)(0= E "«{*)€• \a\<m
Show that a
L~ E
a{xW
|a]=m
is an p.d.o. of order < m — 1. Conclude that the only scalar p.d.o. on RN are those indicated in Corollary 9.1.5. Exercise 9.1.5 Let L € P D O ( m ) ( £ , F ) and u € C°°{E). Show the operator C°°{M) Bf^
[L,f}u € C°°(F)
belongs to P D O ( m ) ( l M , F). Exercise 9.1.6 Let Kp and Kq denote the trivial K-vector bundles over RN of rank p and respectively q. Show that any L e P D O ( m ) ( K p , l « ) has the form
L= E
Mx)d°
\a\<m
where for any a Aa € C°°(K N ,Hom (K P ,K ? )). Hint: Use the previous exercise to reduce the problem to the case p = 1. §9.1.2
Examples
At a first glance, the notions introduced so far may look too difficult to "swallow". To help the reader get a friendlier feeling towards them we included in this subsection a couple of classical examples which hopefully will ease this process. More specifically, we will compute the principal symbols of some p.d.o.'s we have been extensively using in this book. In the sequel • will denote the multiplication by a smooth scalar function.
342
Elliptic equations on manifolds
Example 9.1.12 (The euclidian Laplacian)This is the second order p.d.o. A : C°°(RN) -> C ^ l " )
A = - E
ft2
i
Let / G C ^ R " ) . Then (ft = £ - ) ad(/)(A) = - £ a d ( / ) (ft) 2 = - E W ( / ) ( $ ) ° ^ + ft o ad(/)(ft)}
= - E ( / * ' • $ + $(/»<•)} i
= ~ E ( / x i • #• + /*<*' • +/x- • di) i
= (A/)--2E/«'-ft. Hence ad(/) 2 (A) = ad(/)(A/-) - 2 E /.< • «*(/) (ft) i
2
= -2E(/ I 0 - = -2M/| 2 -. t
If we set £ = df in the above equality we deduce
* 2 (A)(o = -iei 2 •. In particular this shows A is an elliptic operator. Example 9.1.13 (Covariant derivatives)Consider a vector bundle E —¥ M over the smooth manifold M and V a connection on E. We can view V as a p.d.o. V : C°°{E) -J- C°°(T*M ® E). Its symbol can be read from a d ( / ) V , / € C°°(M). For any « G C°°(£) (ad(/)V)w = V(/w) - / ( V u ) =
df®u.
By setting £ = df we deduce ffi(V)(0=£® i.e. the symbol is the tensor multiplication by £. Example 9.1.14 (The exterior derivative)Let M be a smooth manifold. The exterior derivative d : n*{M) -*• Q*(M) is a first order p.d.o. To compute its symbol consider u G fl*(M) and / G C°°(M). Then (ad(f)d)u) = d(fuj) — fdu> = df Aw. If we set f = d/ we deduce <7i(d) = e(£) = the left exterior multiplication by £.
343
Algebraic aspects
Example 9.1.15 Consider an oriented, n-dimensional Riemann manifold (M, ). As in Chapter 4 we can produce an operator 5 = *d*:fi*(M) -> ft*(M) where * is the Hodge *-operator *:ft*(M)->-nn-*(M). 6 is a first order p.d.o. and moreover CJI(6) = *a1(d)* = *e(£) * .
This description can be further simplified. Fix £ G T*M and denote by £* € TjM its metric dual. For simplicity we assume |£| = |£*| = 1. Extend C1 = £ to an oriented orthonormal basis ( £ \ . . . ,£ n ) of TX*M and denote by £* the dual basis of TXM. Consider u £ AkT*M a monomial d£7 where / = ( » i , . . . , ijt) denotes as usual an ordered multi-index. Note that if 1 ^ / then 1 Cr1(£ )w
= 0.
(9.1.3)
If 1 € / e . g . 7 = ( 1 , . . . , A ; ) then *c(0(*w) = *(£J A £* +1 A • • • A O = ( - l ) ^ - * ) ^ - 1 ) ^ A • • • A f* = -(-l)"-»i(Dw
(9.1.4)
where i/n]fc = nk + n + 1 is the exponent introduced in Subsection 4.1.5 while i(£*) denotes the interior derivative along £*. Putting together (9.1.3) and (9.1.4) we deduce
(9-1.5)
Example 9.1.16 (The Hodge-DeRham operator) Let (M,g) be as in the above example. The Hodge-DeRham operator is d + d" :Q*{M)
-+ST{M)
k
where d* = (—l)""' 5. Hence d + d* is a first order p.d.o. and moreover
a(d +
(e(o - t(D)2 = -«mo+iiCHt))
= -lei2 •.
Notice that d + d* is elliptic (since the square of its symbol is invertible).
344
Elliptic equations on manifolds
Definition 9.1.17 Let E —)• M be a smooth vector bundle over the Riemann manifold (M,g). A second order p.d.o.
L:Cco(E)^C'x{E) is called a generalized Laplacian if o2(L){£) = — If ljNotice that all generalized Laplacians are elliptic operators. §9.1.3
Formal adjoints
For simplicity, all vector bundles in this subsection will be assumed complex, unless otherwise indicated. Let EX,E2 —» M be vector bundles over a smooth oriented manifold. Fix a Riemann metric g on M and Hermitian metrics (-, •)* on E{, 2 = 1,2. We denote by dvg = *1 the volume form on M defined by the metric g. Finally, C^(Ei) denotes the space of smooth, compactly supported sections of E{. Definition 9.1.18 Let P € P D O (EUE2). The operator Q e P D O {E2, Ex) is said to be a formal adjoint of P if Vu € C^iEJ and Vw e C^{E2) I (Pu, v)2dvg = / (u, Qv)xdvg. JM
JM
Lemma 9.1.19 Any P € P D O (E\,E2) Proof
admits at most one formal adjoint.
Let Qu Q2 be two formal adjoints of P. Then Wv e C$°(E2) I (u, (Qi - Q2)v)idvg = 0 VM e C o 0 0 ^ ) . JM
This implies (Qx - Q2)v = 0 V u £ C^{E2). If now v G C°°(E2) is not necessarily compactly supported then , choosing (a) C C™(M) a partition of unity, we conclude using the locality of Q = Qi — Q2 that Qv = J2 aQ(av) = 0. D The formal adjoint of a p.d.o. P g PT>0(Ei,E2) (whose existence is not yet guaranteed) is denoted by P*. It is worth emphasizing that P* depends on the choices of g and (•, •),. Proposition 9.1.20 (a) Let L0 6 P D O ( £ 0 , £ i ) and Lx G P D O ^ , ^ ) admit formal adjoints L\ € P D O (Ei+i,Ei) (i = 0,1) ( with respect to a metric g on the base and metric (•, •)., on Ej, j = 0,1, 2. Then L^LQ admits a formal adjoint and (LILQ)*
(b) If L e P D O ( m ) ( £ , 0 , £ i ) then L* e
= L*0L\.
VDO{m)(EuE0).
345
Algebraic aspects Proof
(a) For any u; € C^(Ei)
we have
/ (LiL0uo, u2}2dvg = / {L0u0,Llu0)ldvg=
/ (M 0 , L*0Llu2)0dvg.
(b) Let / e C°°{M). Then (ad(f)LY
= ( I o / - / o L ) ' = -[I',/] =
-ad(f)L\
Thus ad(f0)adUi)
• --adUmW
= ( - l ) m + 1 ( a d ( / o M / i ) • • -ad(/ m )L)* = 0. a
The above computation yields the following result. Corollary 9.1.21 If P e P D O ( m ) admits a formal adjoint then om{p*) =
{~\)mom{py
where the * in the right-hand-side denotes the conjugate transpose of a linear map. Let E be a Hermitian vector bundle over the oriented Riemann manifold Definition 9.1.22 A p.d.o. L e P D O (E,E) L = L*.
(M,g).
is said to be formally selfadjoint if
The above notion depends clearly on the various metrics Example 9.1.23 Using the integration by parts formula of Subsection 4.1.5 we deduce that the Hodge-DeRham operator d + d* :Q*(M)->Q*(M) on an oriented Riemann manifold (M, g) is formally selfadjoint with respect with the metrics induced by g in the various intervening bundles. In fact d* is the formal adjoint of d. Proposition 9.1.24 Let (£ i ,(-,-) i ) (i = 1,2) be two arbitrary Hermitian vector bundle over the oriented Riemann manifold (M,g). Then any L 6 P D O ( £ i , £ 2 ) admits at least (and hence exactly) one formal adjoint L*. Sketch of proof We prove this result only in the case when E\ and E2 are trivial vector bundles over K^. However we do not assume the Riemann metric over RN is the euclidian one. The general case can be reduced to this via partitions of unity and we leave the reader check it for him/her-self.
346
Elliptic equations on manifolds
Let Ei = C p and E2 = Cq. By choosing orthonormal moving frames we can assume the metrics on Et are the euclidian ones. According to the exercises at the end of Subsection 9.1.1 any L e P D O ( m ) ( £ i , £ 2 ) has the form
L= £
Aa(x)da
\a\<m
where Aa e Cco(RN, Mqxp(C)). transpose
Clearly, the formal adjoint of Aa is the conjugate
To prove the proposition it suffices to show that each -^
ePBO1
(EuEr)
admits a formal adjoint. It is convenient to consider the slightly more general situation. Lemma 9.1.25 Let X = X{£:
e Vect {RN) and denote by Vx the first order p.d.o.
Vxu = xi4-u u e c n c ) . OXi
Then Vx = -Vx-div9(X) where divg(X)
denotes the divergence of X with respect to the metric g.
Proof of the lemma Let u,v 6 C^(UN,CP)Choose R > 0 such that the euclidian ball BR of radius R centered at the origin contains the supports of both u and v. From the equality X • (u, v) = {Vxu, v) + {u, Vxv) we deduce JRN ( V X W , v)dvg = J = [
(Vxu,
X-(u,v)dvg-f
(u,Vxv)dvg. J
JBR
v)dvg
Set / = {u,v) e C£°(RN,C) and denote by a € Q,1^) respect to the Riemann metric g i.e. (a,P)g=(3{X)
(9.1.6)
BR
V/?€n'(M").
Equivalently a = gijX'dx3'.
the 1-form dual to X with
347
Algebraic aspects The equality (9.1.6) can be rewritten /
{Vxu,v)dvg=
JBR
df(X)dvg-
{u,Vxv)dvg *'J&
JBR
= J
(df, ct)gdvg - J^N (u,
Vxv)dvg.
The integration by parts formula of Subsection 4.1.5 yields /
(df, a)gdvg =
J BR
(df A *ga) \9BR + / JdBft
fd*advg.
JBR
Since / = 0 on a neighborhood of of 8BR we get / JBR
Since d*a = — divg(X) /
JBR
X-(u,v)dvg
(df,a)gdvg=
JBR
(u,v)d*advg.
(see Subsection 4.1.5) we deduce = —
JBR
(u,v)diVg(X)dvg=
/
JBR
(u, —
divg(X)v)dvg.
Putting all the above together we get J
(Vxu,v)dvg
= JmN(v; ( - V * -
dWg(X))v)dvg
i.e. V x = - V x - div 9 (X) = - V x - 4 = E di(^/\g\X').
(9.1.7)
\9\ i
The lemma and consequently the proposition is proved.
•
E x a m p l e 9.1.26 Let E be a rank r smooth vector bundle over the oriented Riemann manifold (M, g), d i m M = m. Let (•, •) denote a Hermitian metric on E and consider a connection V on E compatible with this metric. V defines a first order p.d.o. V : Cco(E) -+ C°°(r*M ® E). The metrics g and (•, •) induce a metric on T*M®E. We want to describe the formal adjoint of V with respect to these choices of metrics. As we have mentioned in the proof of the previous proposition this is an entirely local issue. So we fix Xo G M and denote by (x1,... ,xm) a collection of g-normal coordinates on a neighborhood U of x. Next, we pick a local synchronous frame of E near x0 i.e. a local orthonormal frame (eQ) such that a t XQ VjeQ = 0 Vi = 1 , . . . , m.
348
Elliptic equations on manifolds
The adjoint of the operator dxk® : C°°{E \v) -> C°°{TU
is the interior derivative (contraction) along the vector field g-dual to the 1-form dxk i.e. {dxk®Y = Ck d=f g^-iidj). Since V is a metric connection we deduce as in the proof of Lemma 9.1.25 that Vk* =
-Vk-dWg{dk).
Hence
V* = Y.v** ° c" = E ( ~ v * - div,(3t)) o Ck k
k
= -Y(Vk
+ dk(log(^\))oCk.
(9.1.8)
k
In particular, since dkg = 0 at x0 we get
v*u=-£vfc°c*. k
The covariant Laplacian is the second order p.d.o. A = A v : C°°(E) -> C°°{E),
A = V*V.
To justify the attribute Laplacian we will show that A is indeed a generalized Laplacian. Using (9.1.8) we deduce that over U (chosen as above) we have A
= - ( E ( V * + d"logy/\9\) ° Ck\ o hzdxJ = -
fe(V,
®VJ}
+ dk log yfc\) J o (g"i • Vj)
= - E{5 k J 'V, + dkgki + gkjdk log J\g~\} ° V,k,j
kj
\/\g\
The symbol of A can be read easily from the last equality. More precisely
= -\Z\29.
Algebraic aspects
349
In the following exercise we use the notations in the previous example. Exercise 9.1.7 (a) Show that
gkerke =
i cM^*) \J\g\
where T\t denote the Christoffel symbols of the Levi-Civita connection associated to the metric g. (b) Show that Ay = -Tr 9 (V T * M ® B V B ) where S7T~M®E is the connection on T*M ® E obtained by tensoring the Levi-Civita connection on T*M and the connection VE on E while Tr 3 : C^iT'M®2
® E) -> C°°{E)
denotes the double contraction by g TtgiSij ® u) = gijSijU. Let E and M as above. Proposition 9.1.27 Let L € P D 0 2 ( £ ) be a generalized Laplacian. Then there exists a unique metric connection V on E and SH = £H(£) € End (E) such that
The endomorphism 9t is known as the Weitzenbock remainder of the Laplacian L. Exercise 9.1.8 Prove the above proposition. Hint Try V denned by V / g r a d W M = t{(Agh)u
- (ad{h)L)u}
f,he
C°°{M), u € C*>(E).
Exercise 9.1.9 (General Green formula) Consider a compact Riemannian manifold (M, g) with boundary dM. Denote by n the unit outer normal along dM. Let E, F -* M be Hermitian vector bundles over M and suppose L e PDO* (E, F). Set go = g |dM> E0 = E \SM and F0 = F \QM- The Green formula states that there exists a sesquilinear map BL : C'X{E) x C°°{F) ->• C^idM) such that / (Lu,v)dv{g) JM
= [ JdM
BL(u,v)dv{g0)
+ f JM
(u,L*v)dv(g).
350
Elliptic equations on manifolds
Prove the following. (a) If L is a zeroth order operator (i.e. L is a bundle morphism) then Bi = 0. (b) If Lx € P D O (F, G) and L2 6 P D O {E, F) then BLIL2
(u, v) = BLl (L2u,
v) + BLi
(u,
L\v).
(c) BL.(v,u)
=
-BL(u,v).
(d) Suppose V is a Hermitian connection on E and X G Vect(M). Set L = Vx '• C°°(E) -> C°°(E). Then BL{u,v) = (u,v)g(X,n). (e) Let L = V : C°°(E) -+ C°°(T*M ® E). Then BL(u,v)
=
(u,inv)E.
where in denotes the contraction by n. (f) Denote by V the section of T*M \QM ff-dual to n. Suppose L is a first order p.d.o. and set J := OL{P)-
Then
BL{u,v)
=
{Ju,v)F.
(g) Using (a)-(f) show that for all u e C°°(E), v G C°°(F) and any z 0 e dM the quantity BL(u, v)(xo) depends only on the jets of u, v at xo of order at most k — 1.
9.2
Functional framework
The partial differential operators are linear operators in infinite dimensional spaces and this feature requires special care in dealing with them. Linear algebra alone is not sufficient. This is where functional analysis comes in. In this section we introduce a whole range of functional spaces which represent the suitable enviroment for p.d.o.'s to live in. The presentation assumes the reader is familiar with some basic principles of functional analysis. As a reference for these facts we recommend the excellent monograph [15] or the very comprehensive [79]. §9.2.1
Sobolev spaces in M^
Let D denote an open subset of RN. We denote by Ljoc(D) the space of locally integrable real functions on D i.e. Lebesgue measurable functions / : RN —> K such that Va G C?{RN) of G L 1 ^ ) Definition 9.2.1 Let / G L}0C(D). A function g G L}0C(D) is said to be the weak fc-th partial derivative of / if f g
JD
V^ G C?{D).
351
Functional framework Lemma 9.2.2 Any / G L}0C(D) admits at most one weak partial derivative. The proof of this lemma is left to the reader. Warning Not all locally integrable functions admit weak derivatives. Exercise 9.2.1 Let / G C°°{D). also its weak A:-th derivative.
Prove that the classical partial derivative dkf is
Exercise 9.2.2 Let H G L}oc{&) denote the Heaviside function, H(t) = 1 for t > 0, H(i) = 0 for t < 0. Prove that H is not weakly differentiate. Exercise 9.2.3 Let fuf2£ g\ + 52 weakly.
L}0C{D). If dkft = ft G L}oc weakly then dk{h + f2) =
The definition of weak derivative can be generalized to higher order derivatives as follows. Consider L : C°°{D) -> C°°(£>) a scalar p.d.o. and f,g G L\0C{D). Then we say that Lf = g weakly if / gcpdx = / fUipdx JD
V
JD
Above, L* denotes the formal adjoint of L with respect to the euclidian metric on
mN.
Exercise 9.2.4 Let f,g G C°°{D). Prove that Lf — g classically <=$• Lf = g weakly. Definition 9.2.3 Let k G Z+ and p G [l,oo]. The Sobolev space Lk'p(D) consists of all the functions / G LP(D) such that for any multi-index a satisfying \a\ < k the mixed partial derivative daf exists weakly and moreover daf G ^(D). For every / G Lk*(D) we set i/p
a
Pdx
||/lk P =||/lk P , 0 =(E / \d f\ \\a\
if p < oo while if p = oo ||/lk»=
£esssup|c><7|.
When k = 0 we write ||/|| p instead of ||/||o, p . Definition 9.2.4 Let k € Z+ and p G [1, oo]. Set Lk£(D) = {/ € L L ( ^ ) ; ¥>/ 6 L fc ' p (^) VV 6 C0°°(D)}.
352
Elliptic equations on manifolds
Exercise 9.2.5 Let f(t) = \t\a, t 6 I , a > 0. Show that for every p > 1 such that a > (P - l ) / p Theorem 9.2.5 Let k e Z+ and 1 < p < oo. Then (a) ( ^ ' " ( l " ) , || • ||fciP) is a Banach space. (b) If 1 < p < oo the subspace C£>(RN) is dense in Lk*{RN). (c) If 1 < p < oo the Sobolev space Lk'p(RN) is reflexive. The proof of this theorem relies on a collection of basic techniques frequently used in the study of partial differential equations. This is why we choose to cover the proof of this theorem in some detail. We will consider only the case p < oo leaving the p = oo situation to the reader. Proof Using the exercise 9.2.3 we deduce that Lk>p is a vector space. From the classical Minkowski inequality ( v
/ v
\ 1/P
(El^+^J
/ v
\ l/P
P
p
\ 1/J>
+(X>l ) •
<(£N )
we deduce that || • ||fc.p is a norm. To prove that Lk,p is complete we will use the well established fact that V is complete. Convention To simplify the notations, throughout this chapter all the extracted subsequences will be denoted by the same symbols as the sequences they originate from. Let (/„) C Lk'p(RN) be a Cauchy sequence i.e. lim ||/ m - / „ | k p = 0. m,n—•oo
'r
In particular, for each multi-index |a| < k the sequence (9 a /„) is Cauchy in L ^ R " ) and thus 3 % 3 ft, V\a\
/ M N a Q / n ^ = (-i)H/ MN /„av Since d Q / n —• ga and / n —» / in 77 and ip £ Lq(RN) where l/q = 1 — 1/p) we conclude J^N ga •
= lim(-l)H |
/„ • d°
f • r
353
FunctionaJ framework
Part (a) is proved. To prove that Cp°(]Rw) is dense we will use mollifiers. Their definition uses the operation of convolution. Given / , g e L1 (RN) define (/ * 9){x) = JRN f{x -
y)g{y)dy.
We leave the reader check that / * g is well defined (i.e. y H-> f(x — y)g{y) G L1 for almost all x). Exercise 9.2.6 If / e L1^) / * g e LP(mN) and
and g g 1 / ( 1 " ) then / * g is well defined. Moreover II/*SIIP
(9-2.1)
To define the mollifiers one usually starts with a function p 6 CQ°(K W ) such that p > 0,
suppp C {|:r| < 1}
and / „ pdx = 1. Next, for each 5 > 0 we define =
8-Np{x/6).
JRNPsdx
= l.
Pi{x)
Note that suppp,; C {\x\ < 6} and
The sequence {ps) is called a mollifying sequence. The next result describes the main use of this construction. Lemma 9.2.6 (a) For any / € L}oc(MN) the convolution ps * f is a smooth func£ion/(This explains the term mollifier. ps smoothes out the asperities), (b) If / € V>(RN) (1 < p < co) then Ps * / —> / as<5—>0.
Proof of the lemma Part (a) is left to the reader as an exercise in the differentiability of integrals with parameters. To establish part (b) we will use the fact that C^(RN) is dense in LP(RN). Fix s > 0 and choose g € C^(SLN) such that ll/-flllp<e/3.
354
Elliptic equations on manifolds
We have
\\pt * f -
<
/UP
UP*
* (/ -
)IIP
+ \\P6 * g - slip + lis - /||p.
Using the inequality (9.2.1) we deduce \\P6 * f ~ /||p < 2 | | / - g\\p + Up, * g - 5 || p .
(9.2.2)
We need to estimate \\ps * g — g\\p. Note that Ps*g (x) = JRN p(z)g(x - 8z)dz and
9
^ = ImN
p z x dz
(M) -
Hence \ps*g{x)
-g[x)\
< J^N p(z)\g{x - Sz) - g(x)\dz < 5sup \dg\.
Since s u p p p s * g C {x + z ; x G s u p p g ; z e suppp s } there exists a compact set K C RN such that supp (ps * g - g) C K
V5e(0,l).
We conclude that IIP* * 9 ~ g\\P < (JK 5 p (sup \dg\Ydx^
" = vol (Ky/'Ssup
\dg\.
If now we pick 6 such that vol(K) 1 / p Jsup|dp| < e / 3 we conclude from (9.2.2) that \\P6*f-f\\P<eThe lemma is proved.
D
The next auxiliary result describes another useful feature of the mollification technique especially versatile in as far as the study of partial differential equations is concerned. Lemma 9.2.7 Let / , g € L}0C(RN) such that 9*/ = g weakly. Then dk(ps*f) = Ps*gMore generally, if
Q
L = Y, «^ |a|<m
is a p.d.o with constant coefficients aQ 6 1 and Lf = g weakly then L(P6 * / ) = Ps * g classically.
355
Functional framework
Remark 9.2.8 The above lemma is a commutativity result. It shows that if L is a p.d.o. with constant coefficients then [L,
Ps*]f
= L{ps *f)-pt*
(Lf) = 0.
This fact has a fundamental importance in establishing regularity results for elliptic operators. Proof of the lemma It suffices to prove only the second part. We will write Lx to emphasize that L acts via derivatives with respect to the variables x = (a; 1 ,... ,xN). Note that L{P6 * f) = JRAL*Ps(x - V))f(y)dy. (9.2.3) Since
iript{x-y)
and
=
-wpt{x~y)
dC = -di we deduce from (9.2.3) that
L{ps *f) = JRN{L*yps{x - y))f{y)dy = J^N ps{x - y)g{y)dy = ps * g since Lf — g weakly and y >-» ps{x — y) € C%?(RN) Vs. The lemma is proved.
•
After this rather long detour we return to the proof of Theorem 9.2.5. Let / g Lk*(MN). We will construct /„ G C?(RN) such that /„ -» / in Lkj" using two basic techniques: truncation and mollification. Step 1: Truncation The essentials of this technique are contained in the following result. Lemma 9.2.9 Let / € Lk'p(UN). Consider for each fl>0a smooth function rjR € C^(RN) such that r)(x) = 1 for |x| < R, r]R(x) = 0 for |x| > R + 1 and \dqR[x)\ < 2 Vx. Then T]R • f G Lkj,(RN) VR > 0 and moreover VR
" / -* f as # —> oo.
Proof We consider only the case k = 1. The general situation can be proved by induction. We first prove that di(r]Rf) exists weakly and as expected dtivR • f) = {diT]R) -f + rin- dif. Let ip e C?(RN).
Since T]R
we have
JRN(diVR)
356
Elliptic equations on manifolds
This confirms our claim. Clearly (<%%)/ + r)Rdif e LP(RN) so that nRf € L l p ( R w ) . Note that d{r)R = 0 for | s | < R and \x\ > R + 1. In particular we deduce ($»te)/ -*• 0 a.e. Clearly |(^7?ij)/| < 2\f(x)\ so that by the dominated convergence theorem we conclude (dm) • f % 0 as R -> oo. Similarly ?7fl9i/ ->• 9 i / a.e. and |7?H9,/| < |<9j/| which implies rjudif —> 9 j / in ZA The lemma is proved.
•
According to the above lemma the space of compactly supported Lfc'p-functions is dense in Lk'p(RN). Hence it suffices to show that any such function can be arbitrarily well approximated in the Z^'P-norm by smooth, compactly supported functions. Let / € Lk'p(RN) and assume e s s s u p p / C {|a;| < R}. Step 2: Mollification ps * f L-^ f as 5 -> 0. Note that each p$ * / is a smooth, function supported in {\x\ < R + 5}. According to Lemma 9.2.7 da(Ps *f)=Ps* (d°f) V|a| < k. The desired conclusion now follows using Lemma 9.2.6. To conclude the proof of Theorem 9.2.5 we need to show that Lk'p is reflexive if 1 < p < oo. We will use the fact that LP is reflexive for p in this range. Note first that Lk'p(RN) can be viewed as a closed subspace of the direct product
II Lp(mN). \a\
via the map T : Lk'p(RN)
-> n
V(RN)
f H- ( d Q / ) H < *
|a|
which is continuous 1-1 and and has closed range. Indeed, if 9 a /n % fa then arguing as in the proof of completeness we deduce that fa = 9af0 weakly where f0 = l i m / n . Hence (/ Q ) = Tf0. We now conclude that Lk'p(MN) is reflexive as a closed subspace of a reflexive space. Theorem 9.2.5 is proved. •
357
Functional frameworJc Remark 9.2.10 (a) For p = 2 the spaces Lk,2(RN) inner product is given by (">v)'< =
are in fact Hilbert spaces. The
fBis['Eaf'u-ff>vdx\dx.
(b) If D C WLN is open then Lk'p(D) is a Banach space, reflexive if 1 < p < oo. However C^{D) is no longer dense in Lk'p{D). The closure of C^{D) in Lk'p{D) is denoted by LQ'P(D). Intuitively, Lkl'p{D) consists of the functions u G Lk'p(D) such that dju — = 0 ondD Vi = 0 , l , . . . , A - l where d/dv denotes the normal derivative along the boundary. The above statement should be taken with a grain of salt since at this point it is not clear how one can define u \gr> when u is defined only almost everywhere. We refer to [Adm] for a way around this issue. The larger space C'X{D) n Lk>p{D) is dense in Lk,p(D) provided the boundary of D is sufficiently regular. We refer again to [3] for details. Exercise 9.2.7 Prove that the following statements are equivalent. (a) u e Ll'"(RN). (b) There exists a constant C > 0 such that for all
kN
u
dxi
V» = 1,--.,JV
where p' = p/(p — I). (c) There exists C > 0 such that for all / s e l " ||A h u||L, < C\h\ where AftM(x) = u{x + h) — u(x). Exercise 9.2.8 Let / € Ll'p{RN)
and <j> € C°°(R) such that \d
const.
and (/) G L 1 * ^ ) and
= #{f) • dif.
W) l P
N
Exercise 9.2.9 Let / G L - (R ).
a\t\-j i m
~\
Show that | / | G Ll'"(MN) and d
if
-dj
ae
- o n {/ ^ °)
a.e. o n { / < 0 }
Hint: Show that fe = (e2 + f 2 ) 1 ' 2 converges to / in L1'" as e -+ 0.
358 §9.2.2
Elliptic equations on manifolds E m b e d d i n g t h e o r e m s : integrability p r o p e r t i e s
The embedding theorems describe various inclusions between the Sobolev spaces Lk'P(RN). Define the "strength" of the Sobolev space LktP(RN) as the quantity a(k,p) = aN(k,p)
= k-
N/p.
The "strength" is a measure of the size of a Sobolev size. Loosely speaking, the bigger the strength the more regular are the functions in that space and thus it consists of "fewer" functions. R e m a r k 9.2.11 The origin of the quantities aN(k,p) can be explained by using the notion of conformal weight. A function u : R " —> R can be thought as a dimensionless physical quantity. Its conformal weight is 0. Its partial derivatives <%u are physical quantities measured in meter-1 = variation per unit of distance and they have conformal weight —1. More generally, a mixed partial 9 Q u has conformal weight —\a\. The quantities |d Q | p have conformal weight — p\a\. The volume form dx is assigned conformal weight TV: the volume is measured in meterN. The integral of a quantity of conformal weight w is a quantity of conformal weight w + TV. For example
J \dfu\vdx has conformal weight TV — p\a\. In particular the quantity '•IV
( / { £ \dau\"}dx V M=* has conformal weight (TV — kp)/p = —a^(A:,p). Geometrically, the conformal weight is captured by the behavior under the rescalings x = \y. If it\(y) = u(Xy) then diUX = A(5;M) A
On an abstract manifold M of dimension TV the quantities of conformal weight w are the sections of the bundle of w/TV-densities \A\^N. For example a 1-density (measure) can be integrated and has weight TV. T h e o r e m 9.2.12 (Sobolev)If aiy(k,p) = a^(m,q)
< 0 and k > m
then L**(R») - 4
Lm'p{MN)
and the natural inclusion is continuous i.e. there exists C = C[N, k, m,p, q) > 0 such that l l / l k , < C | | / | | t , V/ e Lk'»(RN).
359
Functional framework
Proof We follow the approach of [63] which relies on the following elementary but ingenious lemma. 6 LN-1{RN'1).
Lemma 9.2.13 Let N > 2 and fu...,fN 1 < i < N define
for each i 6 l *
and
(x\...,xi,...,,xN)emN-1.
£i = Then
/(i) = /i(fi)/2(6)---/W(^)eL1(Rw) and moreover
n/iii
0 : I M U / ^ - D < C\\ \du\ ||x
where |dw|2 = \d\u\2 + • • • + |djv«|2-
Vu e C0°°(IR")
This result then extends by density to any
We have \u(x\...,xN)\<
fXl
\diu(x\...,xi~\t,xi+1,...,xN)\dt'^
g^i).
J—oo
Note that g{ € L ^ R * " 1 ) SO that
/ite) = *te) 1/(w - 1, eL w - 1 (K JV - 1 )Since we conclude from Lemma 9.2.13 that u(x) e LN^N~1)(RN)
II«HWAT-1)
< (niift(6)iii)
and
= ((niift«nij
•
Using the classical inequality geometric mean < arithmetic
mean
we conclude 1
hUw-Q
N
rimtf
< jf E llft«Hi < ^
N
E ll i<*«l IK-
360
Elliptic equations on manifolds
We have thus proved that Z , 1 , 1 ^ ) embeds continuously in Z / W ^ - 1 ) ^ ) . Now let 1 < p < oo such that aN(l,p) = 1 - N/p < 0 ( i.e. p < N). We have to show that L1'"(RN) embeds continuously in IP* (RN) where . P Let u €
CQ°(RN).
Np N-p
Set v = |v| r-1 v where r > 1 will be specified later. The inequality \V
,N
IMU/(N-I)<(^III^II implies 1/N
r 1
IHU/(iv-1)<^(nilN - ^l u
1
If g = p/(p — 1) is the conjugate exponent of p then using the Holder inequality we get Consequently (N
\
^
1
ll«ll^/( J y-i)<'-|Hlrtr -i)[nilft«llpj
•
Now choose r such that rN/(N — 1) = q{r — 1). This gives ,7V-1 and we get II«IIP-<
Millie"!!,)
This shows L1,p ^-> IP* if 1 < p < N. The general case Lk,P
^
Lm,g if aN(k,p)
= aN(m,q) < 0 fc > ro
follows easily by induction over £. We leave the reader fill in the details.
•
Theorem 9.2.14 (Rellich-Kondratchov)Let (k,p), (m, q) G Z + x [1, oo) such that k > m and 0 > a^(k,p) > a^(m,q). Then any bounded sequence (un) C Lk'p(RN) supported in a ball BR(0), R > 0 has a subsequence strongly convergent in Lm'q(RN).
Functional framewo.de
361
Proof We discuss only the case k = 1 so that the condition a ^ ( l , p ) < 0 imposes 1 < p < N. Let (u„) be a bounded sequence in L1,P(RN) such that esssuppw n C {|a;| < R} Vn. We have to show that for every 1 < q < p* = Np/(N — p) the sequence (u n ) contains a subsequence convergent in LP. The proof will be carried out in several steps. Step 1 We will prove that for every 0 < 5 < 1 the mollified sequence u„:s = ps * un admits a subsequence uniformly convergent on {\x\ < R + 1}. To prove this we will use the Arzela-Ascoli theorem and we will show there exists C = C{5) > 0 such that |u n ,i(x)|
- un,l{xz)\
Vn, V|x| < -R+ 1
< C\X1 - xl\
Vn
. V\xl\> \Xl\ < -R + 1-
Indeed \ps*u(x)\<5-N
f
<5~N
p(^-)\un(y)\dy
[
\un(y)\dy
C(N,p)6-N\\un\\p.-vol(Bs)^-1^'
<
< C(J)||w n || liP (by Sobolev embedding theorem). Similarly \un,s(xi) ~ unAxi)\
< J/
BR+1
\Ps(xi - y) - Ps{x2 - y)\ • \un(y)\dy
< C{5) • \xi -x2\
/
< C(8) • \xi -x2\
\un(y)\dy • |K||i,P.
Step 1 is completed. Step 2: Conclusion Using the diagonal procedure we can extract a subsequence of (un) (still denoted by (u n )) and a subsequence Sn \ 0 such that (gn = un,(Sn) (i) gn is uniformly convergent on BR. (ii) lim„ \\gn ~ «n|| l i P i E « = 0. We claim the subsequence («„) as above is convergent in Lq(BR) for all 1 < q < p. Indeed for all n, m \\Un ~ Um\\q,BR
< | | " n ~ gn\\q,BR
+ Hffn ~ 9m\\q,BR
+ \\9m ~
Um\\qtBR-
We now examine separately each of the three terms in the right-hand-side.
362
Elliptic equations on manifolds
A . The first and the third term. \\U» ~ 9n\\lBR
( Holder inequality) < (f
IBR
= J \un(x) IBR
~
gn{x)\d,X
qq \un(x) - gng{x)\ dx n(x)\dx\
' • vol ( 5 R ) ( r - 1 ) / r
(r=p*/q)
Hence the (sub)sequence (u„) is Cauchy in LQ(BR) pactness theorem is proved. • §9.2.3
and thus it converges. The com-
Embedding theorems: differentiability properties
A priori, the functions in the Sobolev spaces Lk'p are only measurable and are defined only almost everywhere. However, if the strength arf(k,p) is sufficiently large then the functions of Lk'v have a built-in regularity: each can be modified on a negligible set to become continuous and even differentiable. To formulate our next results we must introduce another important family of Banach spaces, namely the spaces of Holder continuous functions. Let a € (0,1). A function u : D C K^ -> R is said to be a-Holder continuous if [u]a d=f
sup
JT Q osc (u; BR{z)
r\D)
0
where for any set S C D we denoted by osc (w; S) the oscillation of u on S i.e. osc (u; S) = sup {\u(x) — u(y)\ ; x, y G S}. Set ||«IU,D=sup|M(a:)| xeD and define C°>a(D) = {u:D^R;
\\u\\0,a>D "=f I M U + [u]tt < oo}.
More generally, for every integer k > 0 define Ck'a(D) = { « £ Cm{D) • dpu € C°'a(D),
V|/?| < &}.
363
Functional framework Ck,a(D)
is a Banach space with respect to the norm IMI*,- =
\\90u\\oo,D
E
+ E
\0\
PNa.D-
|/3|=*
Define the strength of the Holder space Ck'a as the quantity a(k,a)
= k + a.
Theorem 9.2.15 (Morrey) Consider (m,p) € Z+ x [l,oo] and (k,a) G Z+ x (0,1) such that m> k and aN(m,p) = a(k,a) > 0. Then L m ' p (M w ) embeds continuously in Ck'a(MN). Proof We consider only the case k = 1 and (necessarily) m = 0. The proof relies on the following elementary observation. Lemma 9.2.16 Let u G C°°{BR) n L u ( B j t ) and set
" = vork)/B« u(a;)da; Then
In the above inequality cr^-i denotes the "area" of the (N — l)-dimensional round sphere S ^ " 1 C RN. Proof of the lemma I
\
IS
f^~Vl
d
N I
I
/•
1"!/,
u(x) — u(y) = — / —-u(x + rw)dr (o> = — ~). Jo or \x — y\ Integrating the above equality with respect to y we get r
vol (BR)(u(x)
- u) = - \
IB JBRR
r\x—y\ Q /•l*-y| d
dy
-?r-u(x + ru)dr. or
J JO
If we set \dTu(x + ru>)\ = 0 for \x + ru\ > R then vol (BR)\u(x)-u\
< /
dy
J\x-y\<2R
\dru(x + rui)\dr. Jo
If we use polar coordinates (p,ui) centered at x then dy = pN~1dpdu where p = \x — y\ and dw denotes the euclidian "area" form on the unit round sphere. We deduce roo
vol(BR)\u(x)~u\<
r
dr I N Jui /
r2R
\dru(x + ru>)\pN-1dp
364
Elliptic equations on manifolds (2R)N • N
_ (2R)N N [z
\dr(x + rw)\dudr
f°° t 1 / rN~ldr -—-\d r u(x + rLj)\du Jo Js"-! r " - 1 Rj^r \dTu(z)\ 1 BR\X_Z\N^< N ]JB R \X - z^-
x + ru)
The lemma is proved.
dr
(2fl) w f N N JBR
<
N
~
jBR\x^z\N-idV-
•
We want to make two simple observations. 1. In the above lemma we can replace the round ball BR centered at origin by any other ball centered at any other point. In the sequel for any R > 0 and Xo £ RN we set w*o,R =
uu
i
u
(y)dV-
\\ I
vo\(BR{x0))
JBR(XO)
2. The inequality (9.2.4) can be extended by density to any u £
Ll,1(BR).
We will complete the proof of Morrey's theorem in three steps. Step 1: L°°-estimates. We will show there exists C > 0 such that Vu € L1,P(KJV) n
IMU < ciMix,. For each x £ K^ denote by B(x) the unit ball centered at x and set ux = uXji. Using (9.2.4) we deduce
Hx)\<\ux\
+ cN f JB(X)
|du(
ff-i<*y
\x - y\"
1
Since ffff(l,p) > 0 we deduce that p > N so that its conjugate exponent q satisfies
"
p p-1 <
N N-l
In particular, the function y i-¥ \x - y\~(N~1'1 lies in L9(B(x))
and
/ i , ,„ „dy < C(N, q) y JB(X) \x-y\i(N-V ~ x H> where C(N, q) is an universal constant depending only on N and q. Using the Holder inequality in (9.2.5) we conclude | u ( z ) | < C | M | l i P Vrr.
365
Functional framework Step 2: Oscillation estimates.
We will show there exists C > 0 such that for all
Ma < CIMIi,. Indeed, from the inequality (9.2.4) we deduce that for any ball BR(x0) BR{x0) \du(y)\ i(x) - uXOiR\
J/l^"
\x - y\iiN-i)dy)
(Holder inequality) < C\\ \du\ \\p • U
and any x e
^q = P^P
~
^
V
< C\\ \du\ \\PR where l-{q-l)N-l)
, = 1
v = q Hence
1 P
\u{x) - uX0,R\ < CRa
N
1 , + - = 1 P P
Vx €
N = a. P
BR{x0)
and consequently osc(u;BR{x0})
Step 2 is completed. Step 3: Conclusion. Given u € L1'P(RJV) we can find (u„) e C^°(K^) such that « n —> u in L 1,p and almost everywhere. The estimates established at Step 1 and 2 are preserved as n —> oo. In fact these estimates actually show the sequence (un) converges in the C 0,a -norm to function v € C°' a (K' v ) which agrees almost everywhere with u. • Exercise 9.2.11 Let u 6 Ll'l(RN)
3C>0:
\ j r"
JBr(x)
satisfy a {q, v)-energy estimate i.e.
\du{y)\Qdy
where 0 < v < q and q > 1. Show that (up to a change on a negligible set) u is a-H61der continuous (a = 1 — v/q) and moreover
M a < C*2 where the constant Ci depends only on N, q, v and C\. Hint: Prove that JBT(X) \X - yl"-1
and then use the inequality (9.2.4).
~
366
Elliptic equations on manifolds
Remark 9.2.17 The result in the above exercise has a suggestive interpretation. If u satisfies the (q, ^)-energy estimate then although \du\ may not be bounded, on average, it "explodes" no worse that r Q _ 1 as r —> 0. Thus \u(x)-u(0)\
« C [X ta-ldt Jo
« C|x| Q .
The energy estimate is a very useful tool in the study of nonlinear elliptic equations. The Morrey embedding theorem can be complemented by a compactness result. Let (k,p) 6 Z+ x [1, oo] and (m, a) 6 Z+ x (0,1) such that OM{k,p) > a(m,a)
k > m.
Then a simple application of the Arzela-Ascoli theorem shows that any bounded sequence in Lk,p(RN) admits a subsequence which converges in the Cm'a norm on any bounded open subset of RN. The last results we want to discuss are the interpolation inequalities. They play an important part in applications but we chose not to include their proofs long but elementary proofs since they do not use any concept we will need later. The interested reader may consult [3] or [11] for details. Theorem 9.2.18 (Interpolation inequalities) For each R > 0 choose a smooth, cutoff function r]R e ^ ( K ^ ) such that r)R = 1 if |a;| < R T]R = 0 if |x| > R + 1 \dr)R{x)\ < 2 Vx € K^. Fix (m,p) 6 Z+ x [1, oo) and (k, a) 6 1+ x (0,1). (a) For every 0 < r < R + 1 there exists C = C(r,R,m,p) 0 < j < m, e > 0 and for all u 6 Lm-"{RN) \\VRu\\jtpr
< Ce\\vRu\\mf9s
such that for every
+Ce-*"-'">||! te u|| JI , flp .
(b) For every 0 < r < R + 1 there exists C = C(r,R,k,a) 0 < j < k, e > 0 and for all u 6 Ck'a(RN)
such that for every
The results in this and the previous section extend verbatim to slightly more general situations namely to functions / : UN -> H where H is a finite dimensional Hilbert space.
367
functional framework §9.2.4
Functional spaces on manifolds
The Sobolev and the Holder spaces can be defined over manifolds as well. To define these spaces we need two things: an oriented Riemann manifold (M,g) and a Kvector bundle TT : E —> M endowed with a metric h = {-, •) and a connection V = V E compatible with h. The metric g = (•, •) defines two important objects: (i) the Levi-Civita connection V s and (ii) a volume form dvg = * 1 . In particular, dvg defines a Borel measure on M. We denote by If(M, K) the space of K-valued p-integrable functions on (M, dvg) (modulo the equivalence relation of equality almost everywhere). Definition 9.2.19 Let p € [1, oo]. An L^-section of E is a Lebesgue measurable map tp : M —>• E (i.e. t/) _1 ([/) is Lebesgue measurable for any open subset U C E) such that: (i) 7r o ij)(x) = x for almost all x € M except possibly a negligible set. (ii) The function x H-» |^(x)| h belongs to L"(M,K). The space of .//-sections of E (modulo equality almost everywhere) is denoted by ^(E). We leave the reader check the following fact. Proposition 9.2.20 V{E)
I M U ••
is a Banach space with respect to the norm UuW{x)\vdvg(x))llf ifp
Note that if p, q € [1, oo] are conjugate, l/p+l/q defines a continuous pairing
= 1, then the metric h : ExE
—• KM
(;-):V{E)xL''(E)-¥L1{M,K) I (tp,(f>)dvg<
W\\,,B •
ULE
This follows immediately from the Cauchy inequality \h(il>(x),
®---®E*k)
368
Elliptic equations on manifolds
then for every Pi,---,Pk
€E [1,oo] such that 1 - 1/po = 1/pi + • • • + l/Pk
and Vipj £ Vi(Ej),
j = 1,..., k
\[ x(A,---,^k)dvs
For each m = 1,2,... define V m as the composition Vm
. c-oo^j v^
C
oo(T»M0^
v
^®
£
c°°(rM82®£) 4 • • •4
Cco{T*M®m®E)
where we used the symbol V to generically denote the connections in the tensor products T*M®J®E induced by V 9 and V £ The metrics g and h induce metrics in each of the tensor bundles T*M®m
We say that
® E).
(b) Define Lm'p(E) as the space of sections u G ^(E) such that Vj = 1 , . . . , m there exist Uj- e U^T'M®' ® £ ) such that V % = w7- weakly. We set
INU,P
= |M|ra,,B = £ ||V^||P.
A word of warning The Sobolev space Lm'p(E) introduced above depends on several choices: the metrics on M and E and the connection on E. When M is non-compact this dependence is very dramatic and has to be seriously taken into consideration. Example 9.2.22 Let (M, g) be the space R^ endowed with the euclidian metric. The trivial line bundle E = M.M is naturally equipped with the trivial metric and connection. Then I^QR^) = U{M, R). Denote by D the Levi-Civita connection. Then for every u 6 C°°(M) and m € Z + we have Dmu = Y, dx®a ® dau \a\—m
where for every multi-index a we denoted by dx®a the monomial dxai ® • • • ® dxafl.
369
functional framework The length of Dmu(x)
is
E |9°«WI2 |a|=m
• /
The space L m , p (K M ) coincides as a set with the Sobolev space Lk>p(RN). The norm || • || m JJ is equivalent with the norm || • || mw introduced in the previous sections. Proposition 9.2.23 (Lk'p(E), || • ||/fc,p,s) is a Banach space which is reflexive if 1 < p < oo. The proof of this result is left to the reader as an exercise. The Holder spaces can be defined on manifolds as well. If (M, g) is a Riemann manifold then g canonically defines a metric space structure on M (see Chapter 4) and in particular we can talk about the oscillation of a function u : M —»• K. On the other hand, defining the oscillation of a section of some bundle over M requires a little more work. Let (E, h, V) as before. We assume the injectivity radius pM of M is positive. Set Po = min-jXpAf}. If a;, y 6 M are two points such that dist 9 (x,y) < p0 then they can be joined by a unique minimal geodesic 7 I ) S starting at x and ending at y. We denote by TXtV : Ey —> Ex the V B -parallel transport along j X i V . For each ( 6 £ , and r] £ Ey we set by definition If - V\ = 1^ " Tx,yr,\x = |T? -
Ty^\y.
If u : M —> E is a section of E and S C M has the diameter < p0 we define osc (u ; 5) = sup{|w(:r) — u(y)\ ; x, y € S}. Finally set [u]a,E — sup{r~ Q osc (u ; BT{x)) ; 0 < r < p0, x G M } . For any fc > 0 define k
||«|U,a,B = E l|VJ«||oo,B + [Vm«]Q,T.M®™®E and set Ck'a{E) = {« e C fc (£) ; ||n||*,Q < oo}. Theorem 9.2.24 Let (M, 3) be a compact, TV-dimensional, oriented Riemann manifold and E a vector bundle over M equipped with a metric h and compatible connection V. Then the following are true.
370
Elliptic equations on manifolds
(a) The Sobolev space Lm-P(E) and the Holder spaces Ck'a(E) do not depend on the metrics g, h and on the connection V. More precisely, if gx is a different metric on M and V 1 is another connection on E compatible with some metric hi then Lm'p{E, g, h, V) = Lm'"(E, guhu
V1)
as sets of sections
and the identity map between these two spaces is a Banach space isomorphism. A similar statement is true for the Holder spaces. (b) If 1 < p < oo then C°°{E) is dense in Lk
= k0 - N/Po > h - N/p1 =
ko po
kl pl
then L - (E) embeds continuously in L ' (E).
aN(kt,pi)
If moreover
k0 > kx and k0 — N/pa > ki — N/pi then the embedding Lko,po(E) •-• Lhupl(E) is compact i.e. any bounded sequence of Lko'po(E) admits a subsequence convergent in the L* 1,pl -norm. (d) If (m,p) eZ+x [l,oo) and (k,a) € Z+ x (0,1) and m — N/p > k + a then Lm'p(E)
embeds continuously in Ck'a(E).
If moreover
m — N/p > k + a then the embedding is also compact. We developed all the tools needed to prove this theorem and we leave this task to the reader. The method can be briefly characterized by two phrases: partition of unity and interpolation inequalities. We will see them at work in the next section.
9.3
Elliptic partial differential operators: analytic aspects
This section represents the analytical heart of this chapter. We discuss two notions which play a pivotal role in the study of elliptic partial differential equations. More precisely we will introduce the notion of weak solution and a priori estimate. Consider the following simple example. Suppose we want to solve the partial differential equation Au + u = f eL2(S2,R) (9.3.1) where A denotes the Laplace-Beltrami operator on the round sphere. Riemann suggested one should consider the energy functional
371
Analytic aspects If u0 is a minimum of E i.e. E(u0) < E(u)
VM
then 0 = — 1 ( = 0 E(u0 + tv)= f {du0, dv)+uo-vdt Js2 Integrating by parts we get
f- vdvg
Vw.
(9.3.2)
/ (d*duo + UQ — f) • v dvg = 0 V« so that necessarily Au 0 + Mo = /• There are a few grey areas in this approach and Weierstrass was quick to point them out: what is the domain of E, «o may not exist and if it does it may not be C 2 so the integration by parts is ilegal etc. This avenue was abandoned until the dawns of this century when Hilbert reintroduced them into the spotlight and emphasized the need to deal with these issues. His important new point of view was that the approach suggested by Riemann does indeed produce a solution of (9.3.1) " provided if need be that the notion of solution be suitable extended ". The suitable notion of solution is precisely described in (9.3.2). Naturally, one asks when this extended notion of solution coincides with the classical one. Clearly, it suffices that u of (9.3.2) be at least C 2 so that everything boils down to a question of regularity. Riemann's idea was first rehabilitated in Weyl's faimous treatise [77] on Riemann surfaces. It took the effort of many talented people to materialize Hilbert's program formulated as his 19th and 20th problem in the famous list of 27 problems he presented at the Paris conference at the beginning of this century. We refer the reader to [1] for more details. This section takes up the issues raised in the above simple example. The key fact which will allow us to legitimize Riemann's argument is the ellipticity of the partial differential operator involved in this equation. §9.3.1
Elliptic estimates in WN
Let E = C r denote the trivial vector bundle over MN. Denote by (•, •) the natural Hermitian metric on E and by d the trivial connection. The norm of Lk'v{E) (defined using the euclidian volume) will be denoted by || • \\ktP. Consider an elliptic operator of order m L = £ Aa(x)d% : C°°{E) -> C°°(E). |a|<m
For each k e Z+ and R > 0 define
||L|| t , R =
E
sup \\dP(Aa(x))\\.
M<m,|/9|<M-m-|a| B«(°)
In this subsection we will establish the following fundamental result.
372
Elliptic equations on manifolds
Theorem 9.3.1 (a) Let (k,p) € Z+ x (l,oo) and R > 0. Then there exists C = C{\\L\\k+hR,k,p,N,R) > 0 such that Vu € C?(E\Bltm)
IMU+™, < C{\\Lu\\Kp + ||«||p).
(9.3.3)
(b) Let (k, a) € Z+x(0,1) and R > 0. Then there exists C = C(\\L\\k+1,R, k, a, N, R) > 0 such that Vu G C^(E\BR{0)) IMU+m,a < C{\\Lu\\k>a + ||U||0,Q).
(9.3.4)
The proof consists of two conceptually distinct parts. In the first part we establish the result under the supplementary assumption that L has constant coefficients. In the second part, the general result is deduced from the special case using perturbation techniques in which the interpolation inequalities play an important role. Throughout the proof we will use the same letter C to denote various constants C = C(\\L\\k+liR,k,p,N,R)>0 Step 1 We assume L has the form L=
£
Aa<%
\a\=m
where Aa are r x r complex matrices , independent of x 6 M^. We set
l|£|| = EII4.IIWe will prove the conclusions of the theorem hold in this special case. We will rely on a very deep analytical result whose proof goes beyond the scope of this book. For each / e Ll{$LN ,£) denote by /(£) its Fourier transform
Theorem 9.3.2 (Calderon-Zygmund) Let m : SN_1 and define m ( 0 : H* \ {0} -> C by
m{0=m
{w\}-
Then the following hold. (a) There exists ft € C°°(5 J V - 1 ,C) and c 6 C such that ( a i) Is"-1 QdvsN-i = 0.
—> C be a smooth function
Analytic
373
aspects
(a 2 ) For any u G C£°(lir) the limit (Tu) (x) = l h r J
-Mu(x-y)dy
\y\N
e\0jy>c
exists for almost every x G K^ and moreover cu(x) + Tu{x) = ,2ixsN/2
JRN exp(ix • 0™(f )«(£ K -
(b) For every 1 < p < oo there exists C = C{p, ||"i||oo) > 0 such that ||T«|| P < C||«|| p Vu G C%°(RN). (c) (Korn-Lichtenstein) For every 0 < a < 1 and any R > 0 there exists C — C(a, ||m|| 0 , aiJ R) > 0 such that Vu G Cl'a{BR) [Tu}a < C\\u\\0ta,BR.
For a proof of part (a) and (b) we refer to [72], Chap II §4.4. Part (c) is "elementary" and we suggest the reader to try and prove it. In any case a proof of this inequality can be found in [11], Part II.5. Let us now return to our problem. Assuming L has the above special form we will prove (9.3.3). The proof of (9.3.4) is entirely similar and is left to the reader. We discuss first the case k = 0. Let u G C™(E\BR). U can be viewed as a collection u{x) = {v}{x),...
,uT{x))
of smooth functions compactly supported in BR. Define fi(0 = (fi 1 (0.---.fi r (0)If we set v = Lu then for any multi-index (3 such that |/3| = m we have L&u = d^Lu = tfv because L has constant coefficients. We Fourier transform the above equality and we get ( - i ) m Y, AaZadPu((,) = ( - i ) * V » < 0 (9.3.5) \a\=m
Note that since L is elliptic the operator a ( L ) ( 0 = A(Z) = Y, \ot\=m
>U°
:Cr^U
374
Elliptic equations on manifolds
is invertible for any £ / 0. From (9.3.5) we deduce
d^lO = ^5(£M0 V£ + 0 where £?(£) = A(£)~l. Note that 5 ( £ ) is homogeneous of degree — m so that M(£) = £"£?(£) is homogeneous of degree 0. Thus we can find functions my(f) G C°°(R"\{0}) which are homogeneous of degree 0 such that
9^ i (0 = E m «(0« , '(0Using Theorem 9.3.2 (a) and (b) we deduce H^UIIP < C||«||p = C\\Lu\\p. This proves (9.3.3) when L has this special form. S t e p 2 The general case. L is now an arbitrary elliptic operator of order m. Let r > 0 sufficiently small (to be specified later). Cover BR by finitely many balls Br(x„) and consider nu £ C
HdVll < CV-1"1 V|/?| < m. Ifw6C 0 °°(Bfl) then
Set u„ = 77„u and
def
We rewrite the equality v„ = Luv as Lvuv = {Lv - V)uv + vv \a\=m
where ea>v(x) = Aa(xv) I K I I m j i < C{\\*A\p,Br(x,)
- Aa(x).
\0\<m
Using (9.3.3) we deduce
+ |H|p,B,(*.,) +
I ] lka,i/(a:)9 0 «||p,B 1 .(x„) + | | u | L - l , p , B , ( i „ ) ) \a\=m
Since |ea,i/(a:)| < Cr on B r (x v ) where C = C(||Z,||i,R)we deduce IKI|mj>,B,(z„) < C(r\\uv\\PlBr{x„)
+ ||«v||p,B,(x„) + IMIp,Br(:r„))-
Analytic
375
aspects
We can now specify r > 0 such that Cr < 1/2 in the above inequality. Hence IKI|m,p,B,.(x„) < C ' d K H j v M ^ ) + IKIIj>,Br(*„))-
We need to estimate ||iv||p- We use the equality vv = L(T]„U) = r]uLu + [L, r)v]u
in which [L, r)„] = ad(r)v)L is a p.d.o. of order m — 1 so that \\[L,r,M\P,Br(x») < Cr--<m-1>|M|m-i,p,Br(*„). Hence IH|p,Br(z„) < CiW^LuWj,^^)
+ r- (m - 1) ||u|| m _i, p , Br (x„))
so that \\uu\\m,P,BT(x„) < C{\\Lu\\PtBa + r-( m - 1 ) ||u|| m _ 1 , p , B J. We sum over v taking into account that the number of spheres Br(xv) we deduce IML,p,Bfi < E I K I U P < CRN{r-N\\Lu\\p,BR
is
0((R/r)N)
+r-(m+N-V\\u\\m^PtBR).
V
Note that r depends only on R, p, ||L||I,R so that IMkp.Bfl < C(II^M||p,BR + NU-l,p,B«) where C is as in the statement of Theorem 9.3.1. We still need to deal with the term ||M||m_1>PiBR in the above inequality. It is precisely at this point where the interpolation inequalities enter crucially. View u as a section of C^(E \B2R)- If we pick r? e C™(B2R) such that i j E l o n BR we deduce from the interpolation inequalities that there exists C > 0 such that IMIm-iA, < e\H\m,P,R + C e - ^ I M L u * . Hence ||«IU,P,B„ < C(||Lu||p,fljl +£||«|U,p,B« + e- (m - 1) ||u|| p , fl J. If now we choose e > 0 sufficiently small we deduce (9.3.3) with k = 0. To establish it for arbitrary k we argue by induction. Consider a multi-index 1,01 = k. If u € C£>{E\BR)
and Lu = v then
L{d^u) = d0Lu + [L, d^u = d0v + [L, d^]u. The crucial observation is that [L, d®] is a p.d.o. of order < m + k — 1. Indeed arn+k([L,dV})
= [am(L),ak(dp)}
= 0.
376
Elliptic equations on manifolds
Using (9.3.3) with k = 0 we deduce I I ^ I U P . B * < C{\\dPv\\9jtR + \\[L,d»HPiBlt + | | ^ U | | P , B J
< C{\\v\\kj,tBn
+ ||u|| m+fc _ 1}P , Bfi + |Mkp,B R ).
The term ||u||m+*-i,p,BR + ||w||/t,p,Bn can be estimated from above by elMlm+fc.p.B* + CHuHp.Bfi
using the interpolation inequalities as before. The inequality (9.3.3) is completely proved. The Holder case is entirely similar. It is left to the reader as an exercise. Theorem 9.3.1 is proved. Q Using the truncation technique and the interpolation inequalities we deduce the following consequence. Corollary 9.3.3 Let L as in Theorem 9.3.1 and fix 0 < r < R. Then for every k G Z+, 1 < p < oo and a G (0,1) there exists C = C(k,p, a, TV, ||Z-||*-n,fi, R,r) > 0 such that Vu € C°°{E) ||«IU+mj,,Br < C(||.Lu||,fciP,BR +
\\U\\P)BR)
and | | w | U + m , a , B r < C{\\Lu\\k^BR
+ ||«|to,n,BR).
Exercise 9.3.1 Prove the above corollary. §9.3.2
Elliptic regularity
In this subsection we continue to use the notations of the previous subsection. Definition 9.3.4 Let u, v : RN —> C be measurable functions. (a) u is a classical solution of the partial differential equation Lu = £
A»(z)d Q « (z) = v(x)
(9-3-6)
\H\<m
if there exists a 6 (0,1) such that v G C?£{RN), v e C%f(RN) and (9.3.6) holds everywhere. (b) u is said to be an LP strong solution of (9.3.6) if u e L™f(RN) and v e L?oc(RN) and (9.3.6) hold almost everywhere. (The partial derivatives of u should be understood in generalized sense.) (c) u is said to be an LP weak solution if u, v G L?oc(RN) and JRN (u, L*4>)dx = JRN (v, J>)dx V^ G
C?(E).
Analytic
377
aspects
Note the following obvious inclusion {LP weak solutions} D {LP strong solutions}. The principal result of this subsection will show that when L is elliptic then the above inclusion is an equality. Theorem 9.3.5 Let 1 < p < oo and L : C°°(E) -» C°°{E) an elliptic operator of order m. Then any I^-weak solution u of Lu = v e Vloc{E) is an LP strong solution i.e. u €
L^(E).
Remark 9.3.6 Loosely speaking the above theorem says that if a "clever" (i.e. elliptic) combination of mixed partial derivatives can be defined weakly then any mixed partial derivative (up to a certain order) can be weakly defined as well. The essential ingredient in the proof is the technique of mollification. For each 5 > 0 set us = ps * u G C°°(E) v6 = ps*v€ C°°{E). The decisive result in establishing the regularity of u is the following. Lemma 9.3.7 Let ws = Lus -vs£
C°°(E). Then for every <j> 6 C$>{E)
i.e. ws converges weakly to 0 in Lvloc. Roughly speaking this lemma says that [L, ps*} -> 0 as 5 -> 0. We first show how one can use Lemma 9.3.7 to prove u £ L^f(E). Note that us is a classical solution of Lus = vs + u>sUsing the elliptic estimates of Corollary 9.3.3 we deduce ||M«IU,p,B2r < CflK|| p , B!h . + lk||p,B 3 r + IKIIp,B3r)Since us —>• u and v& —> v in LP(E\B3T)
we deduce
IKIIj>,B3r. IMLflsr < C
Fix 0 < r < R.
378
Elliptic equations on manifolds
On the other hand, since u>s is weakly convergent in LP(E\B3T) bounded in this norm. Hence \Hm,V,Br < C.
we deduce it must be
In particular, because Lm'p(E\B2r) is reflexive we deduce that a subsequence of (us) converges weakly to some u € Lm'p(E |B 2 ,). Moreover, using Rellich-Kondratchov compactness theorem we deduce that on a subsequence us —• u strongly in
LP(E\BT).
Since us -» u in Lpoc (as mollifiers) we deduce u \gr= u € Lm'p(E \Br). This shows u e U^p because r is arbitrary. Theorem 9.3.5 is proved. D Proof of Lemma 9.3.7 Pick cj> e C™(E). Assume supp> C B = BR. We have to show lim I / (Lus,
V = £
Beix)%.
We have / {Lus,4>)dx=
JB
I (us(x),L*x
JB
= JB [JRN (Pt(x - y)u{y), L*J {x))dyj dx = E j
B
( / R * (Ps(x - y)u(y), B0(x)%4> (x))dy) dx
= E jB (/ M K <«(v). pfa - y)B0(x)dfy (x))dy"j dx. Similarly JB(vs{x),4>{x))dx = jBJ^N{v{y),ps{x = JBJ^^y)'L*yP^x
-
- y)<j>(x))dydi y)Hx))dydx
= E JB JRN <«(!/). Bp(y)dSMx - y)*(i))>dydx (switch the order of integration) = E j R N jBWv),
B/,(y)dS(Ps(x
- y)<j>(x)))dxdy
Analytic aspects (dxps(x -y)
379
= -dyps{x
- y))
= E l " 1 ) " ( / K N fBWv),BP{y)(d2p,(x
-
y)Mx))dxdy)
(integrate by parts in the interior integral) = E JRN J (u(v)> Bf>(y)Ps(x - y)d%
= E fB JR* (u^> Bp(y)p*(x - v)%
JB
= T,JJRN(u(y),Ps(x-y)
(Bp(x) -
B0(y))%
We will examine separately each term in the above sum.
L LN{U{V)' =
JB JRN =
PS{X
PS{X
JB{URN
- fB((fRN
~y) (B"(x) ~ B^y)) d*^x))dydx
~ V){u{y)'{Bfi{x) PS{X
~
B
~
ff(vMHx))dydx
y)u(y)dy)'B^x)d^(x))dx
Ps(x - y)B;(y)u(y)dy)
= f (us(x),B0(x)^(x))dx-
[
,d^{x))dx
{(B;u)s(x),d0xcf>(x})dx
JB
JB
where (Bpu)s denotes the mollification of B^u. As 5 —>• 0 us -> u and {B*0u)s ->• B*0u in Vloc. Hence Jim fg fRN(u(y),Pi(x = jB{u{x), Lemma 9.3.7 is proved.
- y) {B0(x) - B0(y))
Bp{x)di
^{x))dydx 8£0(x))dx = 0.
380
Elliptic equations on manifolds
Corollary 9.3.8 If u £ LP0C(E) is weak Z^-solution of Lu = v k
( 1 < p < oo) and v £ L £(E)
then u £ Lk+m'p(E)
and for every 0 < r < R
\\u\\m+k,p,Br < C(\\v\\ktP,BR + \\u\\PtBR) where as usual C = C(||L||* + m + i ? i j,R,r,...). Proof We already know that u G Lk*cm'p(E).
Pick a sequence un £ C^(E) k mp
un -> u strongly in L + '
such that
(E).
Then Lun -» Lu strongly in
Lk+m,p{E)
and ||«n||m+*,p,Br < C(||Un||o,p,B„ + t|i« n ||tj,,B i ,).
The desired estimate is obtained by letting n —> oo in the above inequality. • Corollary 9.3.9 (Weyl Lemma)If u G Lpoc(E) is a weak ZAsolution of Lu = v and v is smooth then u must be smooth. Proof Since v is smooth we deduce v G L^P Vfc G Z + . Hence u £ £f„t m ' p Vfc. Using Morrey embedding theorem we deduce that u £ C™f Vm > 0. • The results in this and the previous section are local and so extend to the more general case of p.d.o. on manifolds. They take a particularly nice form for operators on compact manifolds. Let (M, g) be a compact, oriented Riemann manifold and E, F —> M two metric vector bundles with compatible connections. Denote by L 6 PDOm(E, F) an elliptic operator of order m. Theorem 9.3.10 (a) Let u £ IP{E) and v £ Lk'p(F) f {v,4>)Fdv9 = f (u,L'^>)Edvg JM
then u G Lk+m'»(E)
(1 < p < oo) such that
V0 £
C?(F).
JM
and IM|*+mp,B < C(\Mk,P,F + | | « I U )
where C = C(L, k,p). (b) If u G Cm-a(E) and v £ Ck'a{F) (0 < a < 1) are such that Lu = v m+k a
then u £ C
' (E)
and ||^||fc+m,a,i? < C(||u||o,o, E + |M|jb,a,F)
where C = C(L, k, a).
Analytic
381
aspects
Remark 9.3.11 The regularity results and the a priori estimates we have established so far represent only the minimal information one needs to become an user of the elliptic theory. Regrettably we have mentioned nothing about two important topics: equations with non smooth coefficients and boundary value problems. For generalized Laplacians these topics are discussed in great detail in [28] and [61]. The boundary value problems for first order elliptic operators require a more delicate treatment. We refer to [13] for a very nice presentation of this subject. Exercise 9.3.2 (Kato's inequalities) Let {M,g) denote a compact oriented Riemann manifold without boundary. Consider a metric vector bundle E —> M equipped with a compatible connection V. (a) Show that for every u G Lr'2(E) the function x H-> \U(X)\ is in I}'2{M) and moreover \d\u{x)\\<\Vu{x)\ for almost all x G M. (b) Set Ag = V*V and denote by A M the scalar Laplacian. u e L?'2(E) we have A M ( M 2 ) = 2{ABu,u)E - 2|Vu| 2 .
Show that for all
Conclude that V0 G C°°(M) such that
JM
/
JM
{AEu,u)E
i.e. |W(X)|A M (|M(X)|) < {AEu{x),u(x))E §9.3.3
weakly.
A n application: prescribing the curvature of surfaces
In this subsection we will illustrate the power of the results we proved so far by showing how they can be successfully used to prove an important part of the celebrated uniformization theorem. In the process we will have the occasion to introduce the reader to some tricks frequently used in the study of nonlinear elliptic equations. We will consider a slightly more general situation than the one required by the uniformization theorem. Let (M, g) be a compact, connected, oriented Riemann manifold of dimension N. Denote by A = d*d : C'X{M) ->• Cco(M) the scalar Laplacian. We assume for simplicity that voi 9 (M) = / dvg = 1 so that the average of any integrable function tp is defined by Jp=
382
Elliptic equations on manifolds
We will study the following partial differential equation. Au + f(u) = s{x)
(9.3.7)
where / : M —• ]R and s € C°°(M) satisfy the following conditions. (Ci) / is smooth and strictly increasing. (C2) There exist a > 0 and b e M such that /(f) > a t + 6 V t € M . Set
F(u)=
Jo
fUf(t)dt.
We assume (C3) lim| i |_ >00 (F(t) — si) = 00 where s denotes the average of s /
s(x)dvg.
JM
Theorem 9.3.12 Let / and s(x) satisfy the conditions (Ci — C3) Then there exists a unique u 6 C^iM) such that Au (a;) + f(u(x))
= s(x) Vx € M.
The proof of this theorem will be carried out in two steps. Step 1: Existence of a weak solution A weak solution of (9.3.7) is a function u e Ll'2(M) such that f{u(x)) <E L2(M) and / {{du,dcj>) + f{u)cf>}dvg= J s(x)ct>(x)dvg{x) JM
\i
JM
Step 2: Regularity We show that a weak solution is in fact a classical solution. Proof of Step 1 We will use the direct method of the calculus of variations (outlined at the beginning of the current section). Consider the energy functional / : L 1>2 (M) -> K ' ( « ) = / {hdu\2 JM
+ F{u) -
g{x)u(x)}dvg{x).
Z
This functional is not quite well defined since there is no guarantee that F(u) 6 L1 (M) for all u e Ll'2{M). Leaving this issue aside for a moment we can perform a formal computation d la Riemann. Assume u is a minimizer of I i.e. I(u) < I{v)
Vv e
Ll'2{M).
Analytic aspects
383
Thus for all cj> £ 7> 2 (M) 7 (it) < I(u + t(j>) Vt € K. Hence t = 0 is a minimum of h^t) computation shows that
= I(u + t
K(°) = I i(du> d(t>) + f(u^
~ s(x)
JM
so that a minimizer of 7 is a weak solution provided we deal with the integrability issue raised at the beginning of this discussion. Any way, the lesson we learn from this formal computation is that minimizers of 7 are strong candidates for solutions of (9.3.7). We will circumvent the trouble with the possible non-integrability of F(u) by using a famous trick in elliptic partial differential equations called the maximum principle Lemma 9.3.13 Let h : M —> R be a continuous, strictly increasing function and u, v e Ll'2(M) such that (i) h(u), h{v) e L2{M). (ii) AM + h(u) > Ai> + h(v) weakly i.e. f {{du,dcj>) + h{u)(f>}dvg > f {dv,d<j)) + h(v)ct>}dvg V
Let
(u — v)~ = min{u — v, 0) = - { ( « — v) — \u — v\}. According to the Exercise 9.2.9 we have (u — v)~ € Ll'2(M)
and
Using
Clearly (d(u - v), d(u - v)~) = \d(u - v)~\2 and since h is nondecreasing {h(u) - h(v))(u - v)~ > 0 .
(9.3.8)
384
Elliptic equations on manifolds
Hence /
\d(u - v)~\2dvg <-
JM
f (h{u) - h(v))(u - v)~dvg < 0 JM
so that
\d(u-v)~\
=0.
Since M is connected this means (u — v)~ = c < 0. If c < 0 then (u — v) =
(M
— v)~ = c
so that w = u + c < v. Since /i is strictly increasing we conclude / h(u)dvg < /
h(v)dvg.
On the other hand, using (j> = 1 in 9.3.8 we deduce / h{u)dv„ > I
JM
JM
Hence c cannot be negative so that (u — v) The maximum principle is proved. •
h(v)dvg
•V
= 0 which is another way of saying
u>v.
Exercise 9.3.3 Assume h in the above lemma is only non-decreasing but u and v satisfy the supplementary condition u > v. Show the conclusion of Lemma 9.3.13 continues to hold. We now return to the equation (9.3.7). Note first that if u and v are two weak solutions of this equation then AM + f(u) > (<) At; + h(v) weakly so that by the maximum principle u > (<) v. This shows the equation (9.3.7) has at most one weak solution. To proceed further we need the following a priori estimate. Lemma 9.3.14 Let u be a weak solution of (9.3.7). If C is a positive constant such that /(C)>sups(x) M
then u{x) < C a.e. on M.
Analytic Proof
385
aspects The equality / ( C ) > sup s(x) implies A C + / ( C ) > s{x) = Au + / ( u ) weakly
so the conclusion follows from the maximum principle.
•
Fix C 0 > 0 such that /(Co) > sup s(x). Consider a strictly increasing C2-function / : 1 -»• 1 such that / ( « ) = / ( u ) for u < Co f(u)
is linear for M > Co + 1.
The condition (C2) implies there exist A, B > 0 such that \f(u)\
+ B.
(9.3.9)
Lemma 9.3.15 If u is a weak solution of AM + / ( « ) = s(x)
(9.3.10)
then u is also a weak solution of (9.3.7). Proof We deduce as in the proof of Lemma 9.3.14 that u < Co which is precisely the range where / coincides with / . D The above lemma shows that instead of looking for a weak solution of (9.3.7) we should try to find a weak solution of (9.3.10). We will use the direct method of the calculus of variations on a new functional I{u)=
/" 1 / {-\du(x)\2 JM
+ F{u{x)) -
s{x)u(x)}dvg{x)
2
where
u
F{u) = f
f(t)dt.
Jo The advantage we gain by using this new functional is clear. The inequality (9.3.9) shows F has at most quadratic growth so that F(u) 6 L}{M) for all u G L2(M). The existence of a minimizer is a consequence of the following fundamental principle of the calculus of variations. Proposition 9.3.16 Let X be a reflexive Banach space and J : X - > l a convex, weakly lower semi-continuous, coercive functional i.e. the level sets Jc = {x ; J{x) < c} are respectively convex, weakly closed and bounded in X. Then J admits a minimizer i.e. there exists Xo € X such that J{x0) < J(x)
Vx 6 X.
386 Proof
Elliptic equations on manifolds Note that inf J(x) = inf J(x) c
Consider xn € J such that lim J(xn) = inf J. n
v
'
c
Since X is reflexive and J is convex, weakly closed and bounded in the norm of X we deduce that Jc is weakly compact. Hence a (generalized) subsequence (xu) of xn converges weakly to some x0 £ Jc. Using the lower semi-continuity of J we deduce J(x0) < lim inf J(xl/) = inf J. Hence xo is a minimizer of J. The proposition is proved.
•.
The next result will conclude the proof of Step 1. Lemma 9.3.17 I is convex, weakly lower semi-continuous and coercive (with respect to the L 1,2 -norm). Proof The convexity is clear since F is convex on account that / is strictly increasing. The terms t i n - / \du\2dv„ « ! - > — / s(x)u(x\dvJx) 2 JM JM are clearly weakly lower semi-continuous. We need to show the functional u H-> /
F(u)dvg
JM
is also weakly lower semi-continuous. Since F is convex we can find a > 0, P £ R such that
F{u) If u n —• u strongly in Ll'2(M)
-au-P>0.
we deduce using the Fatou lemma that
/ {F{u) — au — P}dvg < lim inf / {F(u n ) — aun — f3}dvg. On the other hand lim / aun + Bdv„ = n JM
au + Bdv„ JM
which shows that I F(u)dvg < lim inf /
>M
F(un)dvg.
This means the functional LU2{M) 3u^
f JM
F(u)dvg
Analytic
387
aspects
is strongly lower semi-continuous. Thus the level sets {u;
f
F{u)dvg
JM
are both convex and strongly closed. Hahn-Banach separation principle can now be invoked to conclude these level sets are also weakly closed. We have thus established that / is convex and weakly lower semi-continuous. Remark 9.3.18 We see that the lower semi-continuity and the convexity conditions are very closely related. In some sense they are almost equivalent. We refer to [20] for a presentation of the direct method of the calculus of variations were the lower semi-continuity issue is studied in great detail. The coercivity will require a little more work. The key ingredient will be a Poincare inequality. We first need to introduce some more terminology. For any u e L2(M) we denoted by u its average. Now set u ^ x ) = u(x) — u. Note the average of uL is 0. This choice of notation is motivated by the fact that uL is perpendicular (with respect to the L 2 (M)-inner product) to the kernel of A which is the 1-dimensional space spanned by the constant functions. Lemma 9.3.19 (Poincare inequality)There exists C > 0 such that /
\du\2dvg >C J \uL\2dvg
VM G
Lia{M).
Proof We argue by contradiction. Assume that for any e > 0 there exists ue e L 1 ' 2 (M) such that
f tefdvg = l JM
and / \due\2dvg < e. JM
The above two conditions imply the family (u^) is bounded in L 1 , 2 (M). Since Ll'2{M) is reflexive we deduce that on a subsequence uf~±v The inclusion L1:1{M) subsequence
<-»• L2(M)
weakly in
l}'2(M).
is compact (Rellich-Kondratchov) so that on a
uf —> v strongly in
L2(M).
388
Elliptic equations on manifolds
This implies v / 0 since f \v\2dvg = \im[ fc
JM
\u±\2dv9 = l. JM
On the other hand du£ = duf —> 0 strongly in I?. We conclude / (dv,d
J/ M
E
JJM M
In particular / (di>, dv)dvg = 0 so that dv = 0. Since M is connected we deduce v = c = const, and moreover c = / u(a;)dw9(s) = lim / ufdvg = 0. e
JM
JM
This contradicts the fact that v / 0. The Poincare inequality is proved.
O
We can now establish the coercivity of I(u). Let K > 0 and u G L 1,2 (M) such that / JM JM
-\du\2 + F(u) - s(x)udvg < K. 2.
(9.3.11)
Since \du\2dvg = /
/
\duL\2dvg
JM
JM
and J \u\2 = \u\2 + f
JM
JM
\uL\2dvg
it suffices to show the quantities u, / \uLfdvg, JM
I
\duL\dvg
JM
are bounded. In view of the Poincare inequality the boundedness of /
\duL\2dvq
JM
implies the boundedness of HU^I^M SO that we should concentrate only on u and
IM« X I| 2 . The inequality (9.3.11) can be rewritten as / {^Idu-f + F{u) - s^u^dvg JM 2
-s-u<
Analytic
389
aspects
The Poincare and Cauchy inequalities imply C\\^\\l
- \\sL\\2 • ||u x || 2 + /
F{u)dvg
-S-U
Since volfl(M) = 1 and F is convex we have a Jensen inequality
so that C\\uL\\\ - l l ^ l h • Hu-'-IU + F(5) - 3 • u < /c. 2
(9.3.12)
1
Set P{t) = Ct - lls- ^* and let m = inf P(t). From the inequality (9.3.12) we deduce F(u) — s-u<
K — m.
Using condition (C3) we deduce that \u\ must be bounded. Feed this information back in (9.3.12). We conclude that POIir1"!^) must be bounded. This forces Hir 1 ^ to be bounded. Thus I is coercive and Lemma 9.3.17 is proved. Q Step 2: Regularity of the minimizer We will use a technique called b o o t s t r a p ping which blends the elliptic regularity theory and the Sobolev embedding theorems to gradually improve the regularity of the weak solution. Let u be the weak solution of (9.3.10). Then u(x) is a weak L2 solution of Aw = h(x) on M where h(x) = —f(u(x)) — s(x). Note that since the growth of / is at most linear f{u(x)) € L2(M). The elliptic regularity theory shows implies u € L2'2(M). Using Sobolev (or Morrey) embedding theorem we can considerably improve the integrability of u. We deduce that (i) either u € L"(M) if -N/q < 2 - N/2 < 0 (dimM = N) (ii) or u is Holder continuous if 2 — N/2 > 0 In any case this shows u(x) € Lqi(M) for some qx > 2 which implies h(x) G qi L (M). Using again elliptic regularity we deduce u G L 2 '' 1 and Sobolev inequality implies that u e Lq2 (M) for some q2 > q\ • After a finite number of steps we conclude that h(x) e Lq(M) for all q > 1. Elliptic regularity implies u e L 2 , '(M) for all q > 1. This implies h{x) £ L2'q(M) for all q > 1. Invoking elliptic regularity again we deduce that u € L*'q(M) for any q > 1. (At this point it is convenient to work with / rather than with / which was only C 2 ). Feed this back in h(x) and regularity theory improves the regularity of u two orders at a time. In view of Morrey embedding theorem the conclusion is clear: u G C°°(M). The proof of Theorem 9.3.12 is complete. D From the theorem we have just proved we deduce immediately the following consequence.
390
Elliptic equations on manifolds
Corollary 9.3.20 Let (M, g) be a compact, connected, oriented Riemann manifold and s(x) e Cco(M). Assume vol 9 (M) = 1. Then the following two conditions are equivalent. ( a ) $ = IM s(x)dvg > 0 (b) For every A > 0 there exists a unique u =.u\ € C°°(M) such that Au + Xe" = s{x).
(9.3.13)
Proof (a) => (b) follows from Theorem 9.3.12. (b) => (a) follows by multiplying (9.3.13) with v(x) = 1 and then integrating by parts so that s = X f eu(x)dvg(x) > 0 . • Although the above corollary may look like a purely academic result it has a very nice geometrical application. We will use it to prove a special case of the celebrated uniformization theorem. Definition 9.3.21 Let M be a smooth manifold. Two Riemann metrics g\ and 52 are said to be conformal if there exists / € Cco(M) such that gi = efg\. Exercise 9.3.4 Let (M,g) be an oriented Riemann manifold of dimension N and / 6 C°°(M). Denote by g the conformal metric g = e!g. If s(x) is the scalar curvature of g and s is the scalar curvature of g show that s(x) = e-!{s{x) + (N-
1)A,/ -
{N
~ ^
"
2)
\df(x)\2g}
where A 9 denotes the scalar Laplacian of the metric g while | • | 9 denotes the length measured in the metric g. Remark 9.3.22 A long standing problem in differential geometry which was only relatively recently solved is the Yamabe problem: "If (M,g) is a compact oriented Riemann manifold does there exist a metric conformal to g whose scalar curvature is constant?" In dimension 2 this problem is related to the uniformization problem of complex analysis. The Yamabe problem was solved in its complete generality due to the combined efforts of T. Aubin, [7, 8] and R. Schoen [66]. For a very beautiful account of its proof we recommend the excellent survey of J. Lee and T.Parker, [47]. One can formulate a more general question than the Yamabe problem. Given a compact oriented Riemann manifold (M, g) decide whether a smooth function s(x) on M is the scalar curvature of some metric on M conformal to g. This problem is known as the Kazdan-Warner problem. The case d i m M = 2 is completely solved in [42]. The higher dimensional situation d i m M > 2 is far more complicated both topologically and analytically.
Analytic
391
aspects
Theorem 9.3.23 (Uniformization Theorem)Let (E,g) be a compact, oriented Riemann manifold of dimension 2. Assume vol y (E) = 1. If x(E) < 0 (or equivalently if its genus is > 2) then there exists a unique metric g conformal to g such that s(g) = - 1 . Proof We look for g of the form g = eug. Using Exercise 9.3.4 we deduce that u should satisfy - l = e""{s(i) + A«} i.e. A« + e" =
s{x)
where s(x) is the scalar curvature of the metric g. The Gauss-Bonnet theorem implies that s = 4TTX(£) < 0
so that the existence of u is guaranteed by Corollary 9.3.20. The uniformization theorem is proved. Q On a manifold of dimension 2 the scalar curvature coincides up to a positive factor with the sectional curvature. The uniformization theorem implies that the compact oriented surfaces of negative Euler characteristic admit metrics of constant negative sectional curvature. Now, using the Cartan-Hadamard theorem we deduce the following topological consequence. Corollary 9.3.24 The universal cover of a compact, oriented surface of negative Euler characteristic is diffeomorphic to K2. In the following exercises (M, g) denotes a compact, oriented Riemann manifold without boundary. Exercise 9.3.5 Fix c > 0. Show that for every / £ L2(M) the equation Au + cu = f has an unique solution u 6 I/ 2,2 (M). Exercise 9.3.6 Consider a smooth function / : M x K —* M such that for every x € M the function u t-> /(ar, u) is increasing. Assume the equation A , u = /(x,w)
(9.3.14)
admits a pair of comparable sub/super-solutions i.e. there exist u0,U0 € LX'2{M) n L°°{M) such that U0{x) > u0{x) a.e. on M
392
Elliptic equations on manifolds
and Asf/0 > f(x, U0(x)) > f(x,u0{x))
> Agu0
weakly in
LU2(M).
Fix c > 0 and define (rj„)„>i C L 2 , 2 (M) inductively as the unique solution of the equation Aun(x)
+ CUn(x)
= CUn-^x)
+
f{x,Un_i(x))
(a) Show that u0(x) < u^x) < u2(x) < ••• < un(x) <••• < U0(x)
Vi 6 M.
(b) Show that un converges uniformly on M to a solution u 6 C°°(M) of (9.3.14) satisfying u0 < u < I/ 0 . (c) Prove that the above conclusions continue to hold even if the monotonicity assumption on / is dropped.
9.4
Elliptic operators on compact manifolds
The elliptic operators on compact manifolds behave in many respects as finite dimensional operators. It is the goal of this last section to present the reader some fundamental analytic facts which will transform the manipulation with such p.d.o. into a less painful task. What makes these operators so "friendly" is the existence of a priori estimates. These coupled with the Rellich-Kondratchov compactness theorem are the keys which will open many doors. §9.4.1
The Predholm theory
Throughout this section we assume the reader is familiar with some fundamental facts about unbounded linear operators. We refer to [15] Ch.II for a very concise presentation of these notions. An exhaustive presentation of this subject can be found in [41]. For the reader's convenience we describe the fundamental notions related to unbounded operators. Definition 9.4.1 (a) Let X, Y be two Hilbert spaces over K = R, C and T : D(T) C X -> Y a linear operator (not necessarily continuous) defined on the linear subspace D(T) C X. T is said to be densely defined if D(T) is dense in X. T is said to be closed if its graph T T = {{x,Tx) e D(T) x Y C X x Y} is a closed subspace in X x Y.
Elliptic operators on compact
manifolds
393
(b) Let T : D(T) C X —> Y be a closed, densely defined linear operator. The adjoint of T is the operator T* : D{T*) c Y -> X defined by its graph I V = {{y*,x*) G Y x X ; <x',i> = (y*,Tx)
Vx G £>(T).
where (•,•): Z x Z —>K denotes the inner product on a generic Hilbert space Z. (c) A closed, densely defined operator T : D(T) C X -+ X is said to be selfadjoint i f T = T*. R e m a r k 9.4.2 (a) In more concrete terms T : D(T) C X —> Y is closed if for any sequence (xn) C .D(T') such that (xn,Txn) —> (a;,?/) it follows that (i) x G D(T) and (ii) j/ = Tx. (b) If T : X —> Y is a closed operator then T is bounded (closed graph theorem). Also note that if T : D(T) C X —> Y is a closed, densely defined operator then kerT is a closed subspace of X (c) One can show that the adjoint of any closed, densely defined operator is a closed, densely defined operator. (d) The closed, densely defined operator T : D(T) C X —> X is selfadjoint if the following two conditions hold. (i) {Tx, y) = {x, Ty) for all x, y G D(T) and (ii) D(T) = {y G X ; 3C> 0 : \(Tx,y)\
< C\x\
Vx G D(T)}.
If only the condition (i) is satisfied the operator T is called s y m m e t r i c . Let (M, g) be a compact, oriented Riemann manifold, E, F two metric vector bundles with compatible connections and L G PDO* = VDOk{E,F) a fc-th order elliptic p.d.o. We will denote the various L2 norms by || • || and the L^-norms by
II • litDefinition 9.4.3 The analytical realization of L is the linear operator La : D(La) C L2(E) -»• L2{E) which acts by u *-* Lu for all u e D{La) =
Lk'2(E).
Proposition 9.4.4 (a) The analytical realization La of L is a closed, densely defined linear operator. (b) If U : C°°{F) -> C°°(E) is the formal adjoint of L then (L*) e =
(La)\
394
Elliptic equations on manifolds
P r o o f (a) Since C°°(E) C D(La) = Lk-2{E) is dense in L2(E) we deduce that La is densely defined. To prove La is also closed consider a sequence (u„) C Lk'2(E) such that un —> u strongly in L2(E) and Lun —> v strongly in L2(F). From the elliptic estimates we deduce I K - « m ||* < C{\\Lun - Lum\\ + \\un - um\\) -)• 0 as m, n -» oo. Hence (u n ) is a Cauchy sequence in Lk'2(E) clear that u = Lw. (b) From the equality / {Lu,v)dvg
V« e Lk'2(E),
= [ {u,L*v)dvg
JM
so that un —>• u in Lk'2(E).
v e
Now it is
Lk'2(F)
JM
we deduce D((LaY) and (La)* = (L*)a on Lk-2{F).
D D((L*)a) = Lk'2(F)
To prove that
D((LaY)
C D((i*)„) = L fc ' 2 (F)
we need to show that if v £ L2(F) is such that 3C > 0 k 2
/ {Lu, v)dvs < C\\u\\ \/u e L ' {E) \JM
then v 6 Lk'2(F).
Indeed, the above inequality shows that the functional «H> I
(Lu,v)dvg
JM
extends to a continuous linear functional on L2{E). Hence there exists 4> £ such that / {(L*Yu,v)dvg= / {u,4>)dvg Vu<ELk'2(E). JM
L2(E)
JM
In other words, v is a weak i 2 -solution of the elliptic equation L*v = (f>. Using elliptic regularity theory we deduce v G Lk,2(M). The proposition is proved. • Following the above result we will not make any notational distinction between an elliptic operator (on a compact manifold) and its analytical realization. Definition 9.4.5 (a) Let X and Y be two Hilbert space over K = R, C and T : D(T) C X -» Y a closed, densely defined linear operator. T is said to be semiF r e d h o l m if (i) dimkerT < oo and (ii) The range R ( T ) of T is closed.
Elliptic operators on compact
manifolds
395
(b) The operator T is called Predholm if both T and T* are semi-Fredholm. In this case the integer ind T =' d i m K ker T - d i m K ker T* is called the Fredholm index of T. Remark 9.4.6 The above terminology has its origin in the work of Ivar Fredholm at the beginning of this century. His result (later considerably generalized by F. Riesz) states that if K : H —> H is a compact operator from a Hilbert space to itself then 1H + K is a Fredholm operator of index 0. Consider again the elliptic operator L of Proposition 9.4.4. Theorem 9.4.7 The operator La : D{La) C L2(E) -> L2{F) is Fredholm. Proof
The Fredholm property is a consequence of the following compactness result.
Lemma 9.4.8 Any sequence (u„) C Lk'2{E) such that {||«n|| + ||Lu n ||} is bounded contains a subsequence strongly convergent in L2(E). Proof of the lemma
Using elliptic estimates we deduce that IKIlt < C{\\Lun\\ + IKH) < const.
Hence (un) is also bounded in Lk'2{E). On the other hand, since M is compact Lk'2(E) embeds compactly in L2(E). The lemma is proved. • We will first show that dim ker L < oo. In the proof we will rely on the classical result of F. Riesz which states that a Banach space is finite dimensional if and only if its bounded subsets are precompact (see [15], Chap. VI). Note first that according to Weyl's lemma kerL C Cco(E). Next, notice that kerL is a Banach space with respect to the L 2 -norm since according to Remark 9.4.2 (a) kerL is closed in L2(E). We will show that any sequence (u n ) C kerL which is also bounded in the L 2 -norm contains a subsequence convergent in L 2 . This follows immediately from Lemma 9.4.8 since ||u„|| + ||Lu n || = ||«n|| is bounded. To prove that the range R (T) is closed we will rely on the following very useful inequality, a special case of which we have seen at work in Subsection 9.3.3. Lemma 9.4.9 (Poincare inequality) There exists C > 0 such that
114 < C||L«|| kj2
2
for all u e L (E) which are L -orthogonal to kerL i.e. / (u,(j>)dvg = 0 V<£ G kerL.
396
Elliptic equations on manifolds
Proof We will argue by contradiction. Denote by X C Lk'2(E) the subspace consisting of sections L 2 -orthogonal to ker L. Assume that for any n > 0 there exists « , e X such that IKH = 1 and ||Lu n || < 1/n. Thus ||iw n || —> 0 and in particular ||u„|| 4- ||£u n || is bounded. Using Lemma 9.4.8 we deduce that a subsequence of (un) is convergent in L2(E) to some u. Note that ||w|| = 1. It is not difficult to see that in fact u G X. We get a sequence (un, Lun) C TL = the graph of L such that (un,Lun)
->• (u,0).
Since L is closed we deduce u e D(L) and Lu = 0. Hence u e ker LdX contradicts the condition ||w|| = 1. •
= {0}. This
We can now conclude the proof of Theorem 9.4.7. Consider a sequence (vn) C R(L) such that vn —>• v in L2{F). We want to show t / e R ( i ) . For each vn we can find an unique un E X = (kerL) x such that Lun = vn. Using the Poincare inequality we deduce \\u„ - um\\ < C\\vn
-vm\\.
When we couple this inequality with the elliptic estimates we get ||u„ - u m || fc < C(\\Lu„ - Lum\\ + \\un - um\\) < C\\vn - vm\\ ->• 0 as m, n -> CXD. Hence (u„) is a Cauchy sequence in Lk'2(E) so that un —> u in Lk,2(E). Clearly Lu = v so that » e R (L). We have so far proved that ker L is finite dimensional and R (L) is closed i.e. La is semi-Fredholm. Since {La)* = (L*)a and L* is also an elliptic operator we deduce {Lay is also semi-Fredholm. This completes the proof of Theorem 9.4.7. • Using the closed range theorem of functional analysis we deduce the following important consequence. Corollary 9.4.10 (Abstract Hodge decomposition)Any k-th order elliptic operator L : C°°(E) —» C^iF) over the compact manifold M defines natural orthogonal decompositions of L2(E) and L2{F). More precisely we have L 2 ( £ ) = k e r L e R ( L * ) and L2{F) = kerL* © R(L).
Elliptic operators on compact
397
manifolds
Corollary 9.4.11 If kerL* = 0 then for every v G L2(F) the partial differential equation Lu = v admits at least one weak L 2 -solution u €
L2(E).
The last corollary is really unusual. It states the equation Lu = v has a solution provided the dual equation L*v = 0 has no nontrivial solution. A nonexistence hypothesis implies an existence result! This partially explains the importance of the vanishing results in geometry i.e. the results to the effect that ker L* = 0. With an existence result in our hands presumably we are more capable of producing geometric objects. In the next chapter we will describe one powerful technique of producing vanishing theorems based on the so called Weitzenbock identities. Corollary 9.4.12 Over a compact manifold ker L = ker L'L Proof Then
ker L* = ker LU.
Clearly k e r i C kerL*L. Conversely, let tp € C°°(E) such that L'Lip = 0. | | L ^ | | 2 = f {Lil>LiP)dvg= J (L*Li>,iP)dvg = 0. JM
•
JM
The Fredholm property of an elliptic operator has very deep topological ramifications culminating with one of the most beautiful results in mathematics: the Atiyah-Singer index theorems. Unfortunately this would require a lot more extra work to include it here. However, in the remaining part of this subsection we will try to unveil some of the natural beauty of elliptic operators. We will show that the index of an elliptic operator has many of the attributes of a topological invariant. We stick to the notations used so far. Denote by Ellk(E, F) the space of elliptic operators C°°{E) —> C°°(F) of order k. By using the attribute space when referring to Ell* we implicitly suggested it carries some structure. It is. not a vector space, it is not an affine space it is not even a convex set. It is only a cone in the linear space PDO*" 1 '. But it carries a natural structure of metric space which we now proceed to describe. Let Lu L2 G El\k(E,F). We set 6(LU L2) = sup{||I 1 w - L2u\\ ; ||u||fc = 1}. Define d(Lu L2) = max{5(Lu L2), 6(L{, L*2)}. We let the reader check that (Ell*, d) is indeed a metric space. A continuous family of elliptic operators (L^) A6A (where A is a topological space) is then a continuous map A3AnI
A
e E1U.
In more intuitive terms this means that the coefficients of L\ and their derivatives up to order k depend continuously upon A.
398
Elliptic equations on manifolds
T h e o r e m 9.4.13 The index map ind:EUfc(£,F)->Z,
L H-> ind (L)
is continuous. The proof relies on a very simple algebraic trick which however requires some analytical foundation. Let X, Y be two Hilbert spaces. For any Fredholm operator L : D(L) C X —> Y denote by IL • ker L —> X (resp. by PL • X —> ker L) the natural inclusion ker L <-> X (resp. the orthogonal projection X —> kerL). If L; : D(Li) C X -¥ Y (i = 0,1) are two Fredholm operators define TlLo(Li) • D{LX) ® ker L*0 C X 0 ker L*0 -> Y © ker L 0 by ^L„(ii)(M, 0) = {Liu + iL-
U € D(Li),
4> S kerLJ.
In other words 72.£,0(L1) is given by the block decomposition 1lLo{L{) =
PL0
0
We will call 72.£,0(Li) is the regularization of Li at LQ. The operator LQ is called the pivot of the regularization. For simplicity, when L0 = L\ — L, we write
nL = nL{L). The result below lists the main properties of the regularization. L e m m a 9.4.14 (a) TZia(Li) is a Fredholm operator. (b) 1VLa{L{) = nL.(Lt). (c) TIL0 1S invertible (with bounded inverse). Exercise 9.4.1 Prove the above lemma. We strongly recommend the reader who feels less comfortable with basic arguments of functional analysis to try to provide the no-surprise proof of the above result. It is a very good "routine booster". Proof of Theorem 9.4.13 Let L0 6 Ellk{E, F). We have to find r > 0 such that VL G Ellk(E, F) satisfying d(L0, L) < r we have ind(L) = ind(Lo)We will achieve this end in two steps.
Elliptic operators on compact
399
manifolds
Step 1 We will find r > 0 such that VL satisfying d(L0,L) L at Lo is invertible (with bounded inverse).
< r the regularization of
Step 2 We will conclude that if d(L, L 0 ) < r where r > 0 is determined at Step 1 then ind (L) = ind (L 0 ). Step 1 Since %ia(L) is Fredholm it suffices to show that both 7£L 0 (L) and 1ZL*(L*) are injective if L is sufficiently close to L0. We will do this only for TZi0(L) since the remaining case is entirely similar. We argue by contradiction. Assume there exists a sequence (un,(f>n) C Lk'2(E) x kerLo a n d a sequence (L n ) C E1U(S,F) such that ||«n||* + ||0h|| = l
(9.4.1)
KLo(Ln){un,
(9.4.2)
d(L0,L„)
(9.4.3)
and From (9.4.1) we deduce that {<j>n) is a bounded sequence in the finite dimensional space kerLJ. Hence it contains a subsequence strongly convergent in L2 and in fact in any Sobolev norm. Set (/> = lim>„. Note that \\4>\\ = lim n \\(j>„\\. Using (9.4.2) we deduce i.e. the sequence (Lnun) is strongly convergent to —(/> in L2(F}. now gives \\L0un - Lnun\\ < 1/n
The condition (9.4.3)
i.e. limL n M n = limioMn = —<£ 6 L2(F). n
n
Since un _L kerLo (by (9.4.2)) we deduce from the Poincare inequality combined with the elliptic estimates that ||«n ~ "mil* < C\\L0un - L0um\\ ->• 0 a s m , n ^ o o . Hence the sequence un strongly converges in Lk'2 to some u. Moreover
"m|KIU = ll«lU
and
ll«IU + H\\ = 1-
Putting all the above together we conclude that there exists a pair (u, <j>) € Lk'2(E) x ker LJ such that
\\u\\k + U\\ = l LQU = — <j> and u ± k e r L .
(9.4.4)
400
Elliptic equations on manifolds
This contradicts the abstract Hodge decomposition which coupled with (9.4.4) implies u = 0 and cj> = 0. Step 1 is completed. S t e p 2 Let r > 0 as determined at Step 1 and L € Ellk(E, F). Hence
PL0
0
is invertible. We will use the invertibility of this operator to produce an injective operator ker L* ffi ker L0 •-> ker L ffi ker L*Q. This implies dim ker L* + dim ker L0 < dim ker L + dim ker L*0 i.e. ind(L0)
KLo{L)
A C
where T : Lk'2(E) n (kerL) 1 C (keri)- 1 -> (keri*)- 1 = Range (L) denotes the restriction of L to (ker L)L. T is a n invertible o p e r a t o r w i t h b o u n d e d inverse. Since HL0{L) is invertible, V» e V there exists a u n i q u e pair (4>, u) € (ker L)L(&U such that
nLo(L) ~4>
u
This means T(/> + Au = 0 and B<j) + Cu = v.
Elliptic operators on compact
401
manifolds
Thus we can view both <j> and u as (linear) functions of v, <j> = (/>(v) and u = u(v). We claim the map v H-> U = u(w) is injective. Indeed if w(w) = 0 for some v then T(j> = 0 and since T is injective ^ must be zero. From the equality v = B(j> + Cu we deduce v = 0. We have thus produced the promised injective map V <—> U. Theorem 9.4.13 is proved. • The theorem we have just proved has many topological consequences. We mention only one of them. Corollary 9.4.15 Let L0, Lx e Ellk{E,F) ind(Li).
if ak{LQ) = ak{L{) then ind(L 0 ) =
Proof For every t G [0,1] Lt = (1 — t)L0 + tL\ is a fc-th order elliptic operator depending continuously on t. (Look at the symbols). Thus ind(L ( ) is an integer depending continuously on t so it must be independent of t. • This corollary allows us to interpret the index as as a continuous map from the elliptic symbols to the integers. The analysis has vanished ! This is (almost) a purely algebraic-topologic object. There is one (major) difficulty. These symbols are "polynomials with coefficients in some spaces of endomorphism". The deformation invariance of the index provides a very powerful method for computing it by deforming a "complicated" situation to a "simpler" one. Unfortunately our deformation freedom is severely limited by the "polynomial" character of the symbols. There aren't many polynomials around. Two polynomial-like elliptic symbols may be homotopic in a larger classes of symbols (e.g. symbols which are only positively homogeneous along the fibers of cotangent bundle). At this point one should return to analysis and try to conceive some operators that behave very much like elliptic p.d.o. and have more general symbols. Such objects exist and are called pseudo-differential operators. We refer to [46] for a very efficient presentation of this subject. We will not follow this long but very rewarding path but we believe the reader who reached this point can complete this journey alone. Exercise 9.4.2 Let L £ Elljt(.E, F). A finite dimensional subspace V c L2(F) is called a stabilizer of L if the operator SLy
: Lk?{E) © V -4- L2{F)
SLy{u
®v) = Lu + v
is surjective. (a) Show that any subspace V C L2{F) containing k e r i * is a stabilizer of L. More generally, any finite dimensional subspace of L2(F) containing a stabilizer is itself a stabilizer. Conclude that if V is a stabilizer then ind L = dim ker SL,V — dim V.
402
Elliptic equations on manifolds
Exercise 9.4.3 Consider a compact manifold A and L : A -» Ellk(E, F) a continuous family of elliptic operators. (a) Show that the family L admits an uniform stabilizer i.e. there exists a finite dimensional subspace V C L2(F) such that V is a stabilizer of each operator L\ in the family L. (b)Show that if V is an uniform stabilizer of the family L then the family of subspaces ker SiXtv defines a vector bundle over A. (c) Show that if Vi and V2 are two uniform stabilizers of the family L then we have a natural isomorphism vector bundles ker SLyl © Y_2 = ker SL,v2 © } \ In particular, we have an isomorphism of line bundles det ker SLtVl ® det V? = det ker SLy2 ® det V2*. Thus the line bundle det ker SL,V ® det V* —¥ A is independent of the uniform stabilizer V. It is called the determinant line bundle of the family L and is denoted by detind(L). §9.4.2
Spectral theory
We mentioned at the beginning of this section that the elliptic operators on compact manifold behave very much like matrices. Perhaps nothing illustrates this feature better than their remarkable spectral properties. This is the subject we want to address in this subsection. Consider as usual a compact, oriented Riemann manifold (M, g) and a complex vector bundle E —» M endowed with a Hermitian metric (•,•) and compatible connection. Throughout this subsection L will denote a fc-th order, formally selfadjoint elliptic operator L : C°°(E) -> C°°(E}. Its analytical realization La : Lk'2{E) C L2(E) -+ L2{E) is a selfadjoint, elliptic operator so its spectrum is an unbounded closed subset of R. Note that for any A € R the operator A — La is the analytical realization of the elliptic p.d.o. A I B — L and in particular A — La is a Fredholm operator so that A e o(L) « = • ker(A - L) ± 0. Thus the spectrum of L consists only of eigenvalues of finite multiplicities. The main result of this subsection states that one can find an orthonormal basis of L2(E) which diagonalizes La. Theorem 9.4.16 Let L G Ellk(E) be a formally selfadjoint elliptic operator. Then the following are true.
403
Elliptic operators on compact manifolds
(a) The spectrum a(L) is real a(L) C M and for each A G a(L) the subspace ker(A — L) is finite dimensional and consists of s m o o t h sections. (b) o(L) is a closed, countable, discrete, unbounded set. (c) There exists an orthogonal decomposition
L2(E)= 0
ker(A-L).
Ae
(d) Denote by P\ the orthogonal projection onto ker(A — L). Then Lk'2(E) = D(La) = {i>e L\E)
; £A 2 ||P A >|| 2 < oo}. A
Part (c) of this theorem allows one to write 1=£PA A
and £ = EA^A. A
The first identity is true over the entire L2(E) while part (d) of the theorem shows the domain of validity of the second equality is precisely the domain of L. Proof (a) We only need to show that ker(A — L) consists of smooth sections. In view of Weyl's lemma this is certainly the case since A — L is an elliptic operator. (b)&(c) We first show a{L) is discrete. More precisely given A0 € c{L) we will find e > 0 such that ker(A - L) = 0, V|A - A 0 |,e, A ^ A0. Assume for simplicity A0 = 0. We will argue by contradiction. Thus there exist An -> 0 and un € C°°(E) such that Lun = \nun,
\\un\\ = 1.
Clearly un G R(L) = (kerL*)-1 = (keri)- 1 so that the Poincare inequality implies
1 = IKH < C\\Lun\\ = CXn -»• 0. Thus a(L) must be a discrete set. Now consider i 0 € K \ o{L). Thus t0 — L has a bounded inverse
T=(t0-L)-\ Obviously T is a selfadjoint operator. We claim that T is also a compact operator. Assume (vn) is a bounded sequence in L2{E). We have to show un = Tvn admits a subsequence which convergent in L2(E). Note that un is a solution of the partial differential equation (t0 - L)un = t0un - Lun = vn
404
Elliptic equations on manifolds
so that using elliptic estimates we deduce
IKIU < c(|K|| + |K||). Obviously un = Tvn is bounded in L2(E) so the above inequality implies ||un||jt is also bounded. The desired conclusion follows from the compactness of the embedding Lk'2{E) -» L2{E). Thus T is a compact, selfadjoint operator. We can now use the spectral theory of such well behaved operators as described for example in [15], Chap. 6. The spectrum of T is a closed, bounded, countable set with one accumulation point, /J, = 0. Any fj. 6 cr(T) \ {0} is an eigenvalue of T with finite multiplicity and since ker T = 0 L\E)=
©
ker(,x-T).
f£<7(T)\{0}
Using the equality L = to-
T"1
we deduce
||£AJW||2 = £ m t f | | 2 < ° o . A
A
Conversely, if
EA2||P^||2<(x, A
consider the sequence of smooth sections
<j>n = £
APA>
|A|
which converges in L2(E) to
* = £APAtf. A
On the other hand
converges in L (E) to $. Using the elliptic estimates we deduce \\i>n ~ 1pm\\k < C ' ( | | V ' n - V ' m | | + l l 0 n - 0 m | | ) - > O
Hence ?/> G Lk,2(E)
aS 71, Ul ->• CO.
as a L*'2-limit of smooth sections. The theorem is proved.
•
Elliptic operators on compact manifolds
405
E x a m p l e 9.4.17 Let M = S1, E = C M and L = - i ^ : C ^ S S Q -> C 0 0 (S 1 ,C). L is clearly a formally selfadjoint elliptic p.d.o. The eigenvalues and the eigenvectors of L are determined from the periodic boundary value problem du - i — = Au, u(0) = u(2ir) which implies u(0) = Cexp(iA0) and exp(2?rAi) = 1. Hence cr(L) = Z and ker(n — L) = span£{exp(in^)}. The orthogonal decomposition L 2 (5 1 ) = 0 k e r ( n - L ) n
is the usual Fourier decomposition of periodic functions. Note that u(9) = £ > n e x p ( i n 0 ) €
Ll'2{Sl)
n
if and only if £ ( l + n2)K|2
The following exercises provide a variational description of the eigenvalues of formally selfadjoint elliptic operator L 6 Ellk(E) which is bounded from below i.e. inf{ / (Lu, u)dvg ; u £ Lk'2{E),
\\u\\ = 1} > - c o .
JM
Exercise 9.4.4 Let V e Lk'2(E) be a finite dimensional invariant subspace of L. Show that (a) V consists only of smooth sections. (b) The quantity A(VX) = inf{ / (Lu, u)dvg ; u e Lk'2{E) n VL,
\\u\\ = 1}
is an eigenvalue of L. V1- denotes the orthogonal complement of V in
L2(E).
406
Elliptic equations on manifolds
Exercise 9.4.5 Set V0 = 0 and denote A0 = A(Vr0J-) According to the previous exercise A0 is an eigenvalue of L. Pick (f>0 an eigenvector corresponding to A0 such that H^oll = 1 and form VI = V0 © span{>0}. Set Ai = A(V11) and iterate the procedure. After m steps we have produced m + 1 vectors <j)0,
4>m,...}
2
is a Hilbert basis of L (E) and
Show
^JM
Exercise 9.4.6 Use the results in the above exercises to show that if L is a bounded from below, k-th order formally selfadjoint elliptic p.do. over an ./V-dimensional manifold then Am(L) = 0(mk/N) asm->oo and d(A) = dim © A
as A -> oo.
Remark 9.4.18 (a)When L is a formally selfadjoint generalized Laplacian then the result in the above exercise can be considerably sharpened. More precisely H.Weyl showed that lim A-"/ 2 d(A) = r a n k ( g ) - V O ' " ( M ) The very ingenious proof of this result relies on another famous p.d.o. namely the heat operator dt + L. For details we refer to [10]. (b) Assume L is the scalar Laplacian A on a compact Riemann manifold (M, g) of dimension M. Weyl's formula shows that the asymptotic behavior of the spectrum of A contains several geometric informations about M: we can read the dimension and the volume of M from it. If we think of M as the elastic membrane of a drum then the eigenvalues of A describe all the frequencies of the sounds the "drum" M can produce. Thus "we can hear" the dimension and the volume of a drum. This is a special case of a famous question raised by V.Kac in [39]: can one hear the shape of a drum? In more rigorous terms this question asks how much of the geometry of a Riemann manifold can be recovered from the spectrum of its Laplacian. This is what spectral geometry is all about. It has been established recently that the answer to Kac's original question is negative. We refer to [30] and the references therein for more details.
Elliptic operators on compact
407
manifolds
Exercise 9.4.7 Compute the spectrum of the scalar Laplacian on the torus T 2 equipped with the flat metric and then use this information to prove the above Weyl asymptotic formula in this special case. §9.4.3
Hodge theory
We now have enough theoretical background to discuss the celebrated Hodge theorem. It is convenient to work in a slightly more general context than Hodge's original theorem. Definition 9.4.19 Let {M,g) be an oriented Riemann manifold. An elliptic complex is a sequence of first order p.d.o.'s 0 - • C°°(£y % C°°{EX) $ • • • " V C°°(£ m ) -> 0 satisfying the following conditions. (i) {C°°(Ei),Di) is a cochain complex i.e. A A - i = 0, VI < i < m. (ii) For each (x, £) G T*M \ {0} the sequence of principal symbols
o -• (E0)x "(D-^'°
(Exyx -• • • • "
{Em
) x -• o
is exact. Example 9.4.20 The DeRham complex (Q*(M),d) is an elliptic complex. In this case the associated sequence of principal symbols is (e(£) = exterior multiplication byO 0 -> 1 e - ^ T*M e 4 ' • • • ^ det(T;M) -> 0 is the Koszul complex of Exercise 7.1.6 of Subsection 7.1.3 where it is shown to be exact. Hence the DeRham complex is elliptic. We will have the occasion to discuss another famous elliptic complex in the next chapter. Consider an elliptic complex (Cco(E.,D.)) over a compact oriented Riemann manifold (M, g). Denote its cohomology by H*(E.,D.). A priori these may be infinite dimensional spaces. We will see that the combination ellipticity + compactness prevents this from happening. Endow each Et with a metric and compatible connection. We can now talk about Sobolev spaces and formal adjoints D*. Form the operators At = D*D, + D^DU : C ° ° ( ^ ) -> C 0 0 ^ ) . We can now state and prove the celebrated Hodge theorem. Theorem 9.4.21 (Hodge)Assume M is compact. Then the following are true. (a) #'(£.,£>.) =* ker A; C C°°(Ei). (b) dim#'(£.,£>.) < o o V i
408
Elliptic equations on manifolds
(c) (Hodge decomposition) There exists an orthogonal decomposition L2(Ei) = ker A* © R ( A - i ) © R ( A ) where we view both A - i and A as bounded operators L1'2 —> L2. Proof
Set E = ®Ei, D = © A , D* = ©A* A = ©A ; , D = D + D*.
Thus D, D* and A are p.d.o.'s C°°{E) ->• C°°(E). D2 = (D*)2 = 0. We deduce
Since A A - i = 0 we deduce
A = D*D + DD* = (D + £>*)2 = £>2. We now invoke the following elementary algebraic fact which is a consequence of the exactness of the symbol sequence. (Look back to the finite dimensional Hodge theory in §7.1.3.) Exercise 9.4.8 The operators D and A are elliptic formally selfadjoint p.d.o. Note that according to Corollary 9.4.12 ker A = ker D so that we have an orthogonal decomposition L2 (E) = ker A © R (£>). (9.4.5) This is precisely part (c) of Hodge's theorem. For each i denote by Pt the orthogonal projection L2(Ei) —> ker A;. Set ? = { » £ C°°{Ei) ; DiU = 0} and W = A-i(C°°(^-i)) so that Hi(E.,D)=Zi/Bi. We claim that the map Pi : Z* —> ker A; descends to an isomorphism H'iE^D.)
^ ker Ai.
This will complete the proof of Hodge theorem. The above claim is a consequence of several simple facts. Fact 1 ker A, C Z'. This follows from the equality ker A = ker I). Fact 2 If u € Zi then u — Ptu 6 A - Indeed, using the decomposition (9.4.5) we have u = PiU + DiP
ijjeL1'2(E).
Elliptic operators on compact
409
manifolds
Since u - Ptu € C°°{Ei) we deduce from Weyl's lemma that V € C°°{E). Thus there exist v € C°°(Ei_i) and w € C 0 0 ^ ) such that %j) = v®w and u = PjU + A-iW + D*w. Applying Dt on both sides of this equality we get 0 = D{u = DtPiU + A A - i W + DiD*w = DiD*w. Since ker D? = ker A A* the above equalities imply u~PiU
= D^iv
e B\
We conclude that Pt descends to a linear map kerA; —> H'(E.,D.). Thus B' C R(.D) = (kerA) x and we deduce that no two distinct elements in kerAi are cohomologous since otherwise their difference would have been orthogonal to kerAj. Hence the induced linear map Pt : ker Aj —• H'{E., D.} is injective. Fact 1 shows it is also surjective so that kerAj *±B?(E.,D). Hodge theorem is proved.
•
Let us apply Hodge theorem to the DeRham complex on a compact oriented Riemann manifold
0 -» fi°(M) 4 fix(M) 4 • • • 4 fin(M) -> 0 (n = dimM). We know the formal adjoint of d : fi*(Af) -> £lk+l(M)
is
d* = (-1)""> * d* where !/„,* = nfc + n +1 and * denoted the Hodge *-operator defined by the Riemann metric g and the fixed orientation on M. Set A = dd* + d*d. Corollary 9.4.22 ( H o d g e ) Any smooth fc-form u € fi*(M) decomposes uniquely as
w = w0 + drj + dX rie nk~l(M), c e fi*+1(M) and u>0 € Q,k(M) is p-harmonic i.e. Aw0 = 0 <=>• du) = 0 and d*w = 0. If moreover <j is closed then C = 0 and this means any cohomology class [z] £ is uniquely represented by a harmonic fc-form.
Hk(M)
410
Elliptic equations on manifolds
Denote by H*(M, g) the space of j-harmonic fc-forms on M. The above corollary shows Hk(M,g)^Hk(M) for any metric g. Corollary 9.4.23 The Hodge *-operator defines a bijection *:Hk(M,g)^B.n-k(M,g). Proof
If w is -harmonic then so is *w since d*u> = ±*d*cv
=0
and *d * (*w) = ± * (duS) = 0. * is bijective since *2 = (—l)k(n~k\
D
Using the L 2 -inner product on fi*(M) we can identify Hn~k(M, g) with its dual and thus we can view * as an isomorphism H* A (H"-')«. On the other hand the Poincare duality described in Chapter 7 induces another isomorphism H* P4 (H"- s )* defined by {PD(u),ri}0=
/wAi) JM
where (•, -) 0 denotes the natural pairing between a vector space and its dual. Proposition 9.4.24 PD = * i.e. / <*w, ri)gdvg = [ uAri JM
Proof
Vu>eHk
7]e
Hnk.
JM
We have (*w, ri)gdvg = (r], *ui)gdvg - T] A *2w = (-l)k(-n~k)T) Aw = wAt).
•
k
Exercise 9.4.9 Let o>o G Q. (M) be harmonic k form and denote by Cuo its cohomology class. Show that J
\u0\gdvg < J
\iv\2gdvg Vw e Cuo
with equality if and only if u> — w0Exercise 9.4.10 Let G denote a compact connected Lie group equipped with a biinvariant Riemann metric h. Prove that a differential form on G is /i-harmonic if and only if it is bi-invariant.
Chapter 10 Dirac operators We devote this last chapter to a presentation of a very important class of first order elliptic operators which have numerous applications in modern geometry. We will first describe their general features and then we will spend the remaining part discussing some frequently encountered examples.
10.1 §10.1.1
The structure of Dirac operators Basic definitions and examples
Consider a Riemann manifold (M, g) and a smooth vector bundle E —> M. Definition 10.1.1 A Dirac operator is a first order p.d.o. D : C°°(£} -> C°°(£) such that D2 is a generalized Laplacian i.e.
a(I?){x,0
= -\tfglE,
VfoOeT'M.
The Dirac operator is said to be graded if E splits as E = E0 © Ex and D(C°°(Ei)) C C , ° c (£(j + i) mo d2). In other words, D has a block decomposition D =
0 B
A 0
Note that the Dirac operators are 1st order elliptic p.d.o. Example 10.1.2 (Hamilton-Floer) Denote by E the trivial vector bundle M2n over the circle S1. Thus C°°{E) can be identified with the space of smooth functions u : Sl -¥ K 2n .
411
412
Dirac operators
Let J : C°°(E) —> C°°(E) denote the endomorphism of E which has the block decomposition
[i r
o
with respect to the natural splitting JR2" = W1 © R". We define the Hamilton-Floer operator $:Cco{E)-+Cco{E) by Su = J^do Clearly, 5 2 =
—
WeC°°(E).
jjp is a generalized Laplacian.
E x a m p l e 10.1.3 ( H o d g e - D e R h a m ) L e t (M, g) be an oriented Riemann manifold. Then, according to the computations in the previous chapter, the Hodge-DeRham operator d + d* :ft*(M) - > f T ( M ) . is a Dirac operator. Let D : C'X{E) —¥ C°°{E) be a Dirac operator over the oriented Riemann manifold (M, g). Its symbol is an endomorphism a{D)
: n*E
-s- TT*E
where 7r : T*M —» M denotes the natural projection. Thus, for any x e M and any £ £ T*M, c(£) = a(D)(x,£) is an endomorphism of Ex depending linearly upon £. Since D2 is a generalized Laplacian C(02
= CT(D2)(X,0
=
-\^91ET.
To summarize, we see that each Dirac operator induces a bundle morphism c:T*M®E^E
(£,e)M.c(0e
such that c(£) 2 = —1£|2. From the equality
JEM
we conclude that {c(0,c(n)}
=-2g(Z,V)lEl
where for any linear operators A, B we denoted by {A, B} their anticommutator {A,B}"^
AB + BA.
413
The structure of Dirac operators
Definition 10.1.4 (a) A Clifford structure on a vector bundle E over a Riemann manifold (M, g) is a bundle morphism
c:
TM®E^E
such that
{c(0,c(!7)} = -2 fl foi/)l B . c is usually called the Clifford multiplication of the (Clifford) structure. A pair (vector bundle, Clifford structure) is called a Clifford bundle. (b) A Z 2 -grading on a Clifford bundle E -> M is a splitting E = E0 © Ex such that Va € fl1 (M) the Clifford multiplication by a is an odd endomorphism of the superspace C°°(£ 0 ) © C 0 0 ^ ) . Proposition 10.1.5 Let £ -> M be a smooth vector bundle over the Riemann manifold (M, g). Then the following conditions are equivalent. (a) There exists a Dirac operator D : C°°(E) -> C°°(E). (b) The bundle E admits a Clifford structure. Proof
We have just seen that (a)=>(b). To prove the reverse implication let
c : T*M
®E->E
be a Clifford multiplication. Then for every connection V : C°°(£) -> C°°{T*M ® E) the composition D = c o V : C°°(E) %• C°°(T*M ® E) A- C°°{E) is a first order p.d.o. with symbol c. Clearly D is a Dirac operator. • Example 10.1.6 Let (M,g) be a Riemann manifold. For each x € M and f e T*M define c(£) : A*T*M -> A*T*M by
c(£)w = (ee ~ ^ V where e ? denotes the (left) exterior multiplication by f while i^ denotes the interior differentiation along £* e TJJM - the metric dual of f. The Exercise 2.2.16 of Section 2.2.4 shows that c defines a Clifford multiplication on A*T*M. If V denotes the LeviCivita connection on A*T*M then the Dirac operator c o V is none other than the Hodge-DeRham operator. Exercise 10.1.1 Prove the last assertion in the above example. The above proposition reduces the problem of describing which vector bundles admit Dirac operators to an algebraic-topological one: find the bundles admitting a Clifford structure. In the following subsections we will address precisely this issue.
414 §10.1.2
Dirac operators Clifford algebras
The first thing we want to understand is the object called Clifford multiplication. Consider (V, g) a (real) finite dimensional, euclidian space. A Clifford multiplication is then a pair (E, p) where E is a K-vector space and p : V —> End (E) is an K-linear map such that {p(u),p(v)}
= -2g(u,v)lE
Vu,v€V.
If (e,) is an orthonormal basis of V then p is completely determined by the linear operators pi = p(e{) which satisfy the anticommutation rules {Pi,Pj} = -2<5yl£The collection (pi) generates an associative subalgebra in End (E) and it is natural to try to understand its structure. We will look at the following universal situation. Definition 10.1.7 Let V be a real, finite dimensional vector space and
a symmetric bilinear form. The Clifford algebra Cl(V, q) is the associative M-algebra with unit, generated by V subject to the relations {u, v} = uv + vu = —2q(u, v) • 1 VM, V 6 V. Proposition 10.1.8 The Clifford algebra Cl(V,q) exists and is uniquely defined by its universality property: for every linear map j : V —>• 21 such that 21 is an associative R-algebra with unit and {j(u), j(v)} = — 2q(u, v) • 1 there exists an unique morphism of algebras $ : Cl(V,q) —• 21 such that the diagram below is commutative. V -^
Cl(V,q)
i denotes the natural inclusion V '-» Cl(V, q). Sketch of proof Let A = ®k>oV®k (V®° = K) denote the free associative Malgebra with unit generated by V. Set Cl(V, q) =
A/1
where X is the ideal generated by {u<8)v + v®u + 2q(u, v)
415
The structure of Dime operators Exercise 10.1.2 Prove the universality property.
Remark 10.1.9 (a)When q = 0 then Cl(V, 0) is the exterior algebra A* V. (b) In the sequel the inclusion V ^-» Cl(V, q) will be thought of as being part of the definition of a Clifford algebra. This makes a Clifford algebra a structure richer than merely an abstract M-algebra: it is an algebra with a distinguished real subspace. Thus in thinking of automorphisms of this structure one should concentrate only on those automorphisms of K-algebras preserving the distinguished subspace. Corollary 10.1.10 Let (V;, qi) (i = 1,2) be two real, finite dimensional vector spaces endowed with quadratic forms $ : V -> R. Then any linear map T : Vi -> V2 such that qi{Tv) = q\{v), W G Vi induces an unique morphism of algebras T # : C7(Vi,gi) —> Cl{V2,q-2) such that 7#(Vi) C V2, where we view Vj as a linear subspace in C7(Vi,
Using the universality property of Cln we deduce that the map V -> Cl(V)
v ^ - v e
Cl(V)
extends to an automorphism of algebras a : Cl(V) —> Cl(V). Note that a is involutive i.e. a 2 = 1. Set Cl°(V) = ker(a - 1), Cll{V) = ker(a + 1). Note that Cl{V) = Cl°(V) 0 Cll(V)
and moreover
Cle{V) • Cli(V)
C C/ (£+7() mod 2 (V)
i.e. the automorphism a naturally defines a Z 2 -grading of Cl(V). In other words, the Clifford algebra Cl(V) is naturally a super-algebra.
416
Dirac operators
Let (Cl(V),+, *) denote the opposite algebra of Cl(V). Cl(V) coincides with Cl(V) as a vector space but its multiplication * is defined by x *y = y • x Vx, y G Cl(V) where "•" denotes the usual multiplication in Cl(V). Note that for any u,v £ V u-v + v-u = u*v + v*u so that using the universality property of Clifford algebras we conclude that the natural injection V <-¥ Cl(V) extends to a morphism of algebras Cl(V) —¥ Cl(V). This may as well be regarded as an antimorphism Cl(V) —> Cl(V) which we call the transposition map, x >-> x*. Note that ( u i • u2 • • • ur)
= uT • • • u\
VMJ
G V.
For x G Cl(V) we set x* = {a(x)f = a(a;1'). x* is called the adjoint of x. For each v eV define c(v) € End (A* V) by c(v)u — (e„ — iv)u> Vw € A*V where as usual e„ denotes the (left) exterior multiplication by v while iv denotes the interior derivative along the metric dual of v. Invoking again the Exercise 2.2.16 we deduce c(v)' = -\v\l so that by the universality property of the Clifford algebras the map c extends to a morphism of algebras c : Cl(V) —> End (A*V). Exercise 10.1.3 Prove that Mx G Cl(V) we have c{x*) = c{Xy where the * in the right-hand-side denotes the adjoint of c(x) viewed as a linear operator on the linear space A*V endowed with the metric induced by the metric on V. For each x G Cl(V) c(x)l is an element of A*V called the symbol of a;. The linear map Cl(V) 9 i H a(x) G A'V is called the symbol map. If (e,) is an orthonormal basis then o{eh
• • • eik) = eh A • • • A eh
Ve*..
This shows the symbol map is bijective since the ordered monomials {ei1---e,-Jt ; l
form a basis of Cl(V). The inverse of the symbol map is called the quantization map and is denoted by q : A'V —> CUV).
The structure ofDirac
417
operators
Exercise 10.1.4 Show that q(Aevm'oddV)
CrenloM(V).
=
Definition 10.1.11 (a) A K(=M-,C)-vector space E is said to be a K-ClifFord module if there exists a morphism of M-algebras p : Cl(V) -»•
EndK(E).
(b) A K-superspace E is said to be a K-Clifford s-module if there exists a morphism of s- algebras p : Cl(V) -> ErTd K (£). (c) Let £ b e a K-Clifford module, p : Cl(V) -> E n d K ( £ ) . E (or p) is said to be selfadjoint if there exists a metric on E (euclidian if K=R- or Hermitian if K=C) such that p{x*) = p(xy Vx e Cl{V). We now see that what we originally called a Clifford structure is precisely a Clifford module. Example 10.1.12 A*V is a selfadjoint, real Cl(V) super-module. In the following two subsections we intend to describe the complex Clifford modules. The real theory is far more elaborate. For more information we refer the reader to the excellent monograph [46]. §10.1.3
Clifford modules: the even case
In studying complex Clifford modules it is convenient to work with the complexified Clifford algebras Cln = Cl„ ® K C. The (complex) representation theory of G n depends on the parity of n so that we will discuss each case separately. The reader may want to refresh his/her memory of the considerations in Subsection 2.2.5. Let n = 2k and consider an n-dimensional euclidian space (V, g) . The decisive step in describing the complex Cl(V) modules is the following. Proposition 10.1.13 There exists a complex CZ(V)-module S = S(V) such that Cl(V) ^ E n d c ( § ) as C-algebras. (The above isomorphism is not natural; it depends on several auxiliary choices.)
418
Dirac operators
S(V) is known as the (even) complex spinor module. The reader familiar with the representation theory of associative algebras can immediately grasp the relevance of this proposition since Weddeburn's theorem completely describes the modules over the algebra of endomorphisms of a vector space. We will have to say more about that a little later. Proof of the proposition Consider a complex structure on V i.e. a skew-symmetric operator J : V —> V such that J 2 = — l y . Such a J exists since V is even dimensional. Let {ei, fu • • •', e/t, fk} be an orthonormal basis of V such that Je* = fr Vi. Extend J by complex linearity to V ®JJ C. We can now decompose V ® C into the eigenspaces of J V = Vlfi 0 V 0 ' 1 where V 1,0 = ker(i - J) and V-1 = ker(i + J ) . Alternatively, V1'0 = span c ( e j - - i/,-),
V 0 ' 1 = s p a n c ( e j + if,).
The metric on V defines a Hermitian metric on the complex vector space (V, J) h(u, v) = g(u, v) + ig(u, Jv) which allows us to identify (see Subsection 2.2.5)
(V* denotes the complex dual of the complex space (V, J)). Hermitian metric the collection fe = - / | ( e j - i / j ) ; 1 <j
is an orthonormal basis of V 1 ' 0 while
fa =-fi(ei + Vi) i l < i < fc } is an orthonormal basis of V 0 ' 1 . Set §2k = A'V 1 , 0 = \*>°V. Any morphism p : Cl{V) -)• End (Sat) is uniquely defined by its restriction to V®C
=
Vlfi®V0'1.
With respect to this
419
The structure of Dirac operators
We thus have to specify the action of each of the components V1'0 and V 0,1 . The elements w € V1'0 will act by exterior multiplication c(w)u> = V2e(w)u = y/2w Aw Vw£ A*'°V. The elements w e V0,1 can be identified with complex linear functionals on V 1,0 and as such they will act by interior differentiation c(w)u>i A • • • A W( = —V2i(w}(wi
A"-At»t)
t
= \/2^2(—iyg£(wj,w)wi
A • • • A u>j A • • • A iw/
3=1
where g£ denotes the extension of g to (V <8> C) x (V ® C) by complex linearity. To check that the above constructions do indeed define an action of Cl(V) we need to check that V » e V This boils down to verifying the anticommutation rules {<ei),c(fj)}
= 0, {cfe),c( e j )} = -2Stj = {c(/ i ),c(/ J )}.
We have
so that c(e;) = e(ei) - i(fi),
c(fj) = i(e(sj) + i{sj)).
The anticommutation rules follow as in the Exercise 2.2.16 using the equalities 9c(£i>£j) = hiThis shows S2k is naturally a C^V^-module. Note that dim(^S2*: = 2k so that d i m c E n d c ( § ) = 2" = dim c CZ(V). A little work (left to the reader) shows the Clifford multiplication map c : Cl(V) —> End(£(§2;t) is injective. This completes the proof of the proposition. D Using basic algebraic results about the representation theory of the algebra of endomorphisms of a vector space we can draw several useful consequences. (See [69] for a very nice presentation of these facts.) Corollary 10.1.14 There exists an unique (up to isomorphism) irreducible complex C/2fc-module and this is the complex spinor module §2*-
420
Dirac operators
Corollary 10.1.15 Any complex C^t-module has the form §2fc ® W, where W is an arbitrary complex vector space. The action of 0 2 * on §2jt ® W is defined by v • (s ® w) = c(v)s ® w. W is called the twisting space of the given Clifford module. Remark 10.1.16 Given a complex O 2 jrmodule E its twisting space can be recovered as the space of morphisms of Clifford modules W =
HomCljE2k,E).
Assume now that (V,g) is an oriented, 2A;-dimensional euclidian space. For any positively oriented orthonormal basis e\,..., e2jt we can form the element r = i*ei---eakeCZ(V). One can check easily this element is independent of the oriented basis and thus it is an element intrinsically induced by the orientation. It is called the chirality operator defined by the orientation. Note that T 2 = 1 and Tx = ( - l ) d e g a T Vx e Cl°(V) U
Cl^V).
Let § = S(V) denote the spinor module of Cl(V). The chirality operator defines an involutive endomorphism of S and thus defines a Z 2 -grading on S § = §+©§-
(S* = k e r ( ± l - T))
and hence a ^-grading of End(S). Since {«, T} = 0 Vw e V we deduce the Clifford multiplication by v is an odd endomorphism of S. This means any isomorphism Cl(V) = End (S(V)) is an isomorphism of ^ - g r a d e d algebras. Exercise 10.1.5 Let J be a complex structure on V. This produces two things: it defines an orientation on V and identifies S = S(V) = A*'°V. Prove that with respect to these data the chiral grading of S is g+/- ^
^even/oddfiy
The above considerations extend to arbitrary Clifford modules. The chirality operator introduces a Z2-grading in any complex Clifford module which we call the chiral grading. However this does not exhaust the family of Clifford s-modules. The family of s-modules can be completely described as {§®W ; W complex s - space} where ® denotes the s-tensor product. The modules endowed with the chiral grading form the subfamily in which the twisting s-space W is purely even.
421
The structure of Dirac operators
Example 10.1.17 Let (V,g) be a 2fc-dimensional, oriented, euclidian space. Then A™V ® C is naturally a Clifford module. Thus it has the form AfaV ® C ^ § ® W To find the twisting space we pick a complex structure on V whose induced orientation agrees with the given orientation of V. This complex structure produces an isomorphism A^V ® C ^ A*'V ® A ° ' V =* § ® (A*-V)* = § ® §*. This shows the twisting space is S*. On the other hand the chirality operator defines Z 2 gradings on both § and §* so that Am V ® C can be given two different s-structures: the chiral superstructure (in which the grading of §* is forgotten) and the grading as s-tensor product §®S*. Using Exercise 10.1.5 we deduce that the second grading is precisely the degree grading AfaV ® C = AevenV ® C 0 A°diV ® C. To understand the chiral grading we need to describe the action of the chiral operator on AmV ® C. This can be done via the Hodge *-operator. More precisely we have r • u = i*+P&'-1) * w
VweAT®C.
(10.1.1)
Exercise 10.1.6 Prove the equality 10.1.1. In order to formulate the final result of this subsection we need to extend the automorphism a and the anti-automorphism b to the complexified Clifford algebra Cl(V). a can be extended by complex linearity a(x ® z) = a(x) ® z V i e
Cl{V)
while ^ is extends according to (x ® z)1, = a;1, ® z. As in the real case set y* = a(jf) = a(yf
Vy G Cl(V).
Proposition 10.1.18 Let S(V) denote the spinor module of the 2fc-dimensional euclidian space {V,g). Then for every morphism of algebras p : Cl(V) -> E n d c ( S ( V ) ) there exists a Hermitian metric on S(V) such that
p(y') = p(yY
Vyeci(v).
Moreover, this metric is unique up to a multiplicative constant.
422
Dirac operators
Sketch of proof Choose an orthonormal basis { e i , . . . , e2k} of V and denote by G the group generated by these elements. G is finite and it consists of the monomials {eh •••eit;
l
2k}.
As a set, G generates Cl(V) as a complex vector space. Pick a Hermitian metric h on S(V) and for each g £ G denote by hg the pulledback metric hg(su s2) = h(p(g)si,p(g)s2) Vsx, s 2 € §. We can now form the averaged metric
l G l sec Each p(g) is an unitary operator with respect to this metric. We leave the reader check this is the metric we are after. The uniqueness follows from the irreducibility of S(V) using Schur's lemma. • Corollary 10.1.19 Let (V, g) as above and p : Cl(V) —> Endj^iJ) be a complex Clifford module. Then ; 25 admits at least one Hermitian metric with respect to which p is selfadjoint. Proof Decompose E as § ® W and p as A ® idw for some isomorphism of algebras A : C/(V) —• End(S). The sought for metric is a tensor product of the canonical metric on S and some metric on W. • §10.1.4
Clifford modules: the odd case
The odd dimensional situation can be deduced using the facts we have just established concerning the algebras CZ2fc- The bridge between these two situations is provided by the following general result. Lemma 10.1.20 Clm ^ Cl%%{. Proof Pick an orthonormal basis { e 0 , e i , . . . , em} in the standard euclidian space jjm+i These generate the algebra Clm+\. We view Clm as the Clifford algebra generated by { e i , . . . , e m } . Now define * : Clm -»• ClZ+x by V{x0 + x1)=x°
+
e0-x1
where x° 6 Cl^en and xl e C7£fd. We leave the reader check that this is indeed an isomorphism of algebras. D
423
The structure of Dirac operators
Proposition 10.1.21 Let (V,g) be a (2k + l)-dimensional euclidian space. Then there exist two complex, irreducible Cl(V) modules S+(V) and §~(V) such that Cl(V) ^ E n d c ( S + ) 0 E n d c ( S " )
as ungraded algebras
The direct sum S(V) = §y © S^ is called the (odd) spinor module. Proof Fix an orientation on V and a positively oriented orthonormal basis ei, e 2 , • •., e2k+i. Denote by S2fc+2 the spinor module of Cl(V © R) where V © M is given the direct sum euclidian metric and the orientation or(V © tt) = OT(V) A or(K). Choose an isomorphism p : Cl(V © K) -> End c (§2* + 2 )Then §2/t+2 becomes naturally a supermodule §2*+2 = §2fc+1 © Sjfc+i and we thus we get the isomorphisms of algebras Cl(V) ^ Cl(V © M)6"6" ^ Ende"en(S2A:+2) = End (§+^2) © End (§^+2).
•
The above characterization can be used to describe the complex (super) modules of C/2fc+i- We will not present the details since the applications we have in mind do not require these facts. For more details we refer to [46}. §10.1.5
A look ahead
In this heuristic section we interrupt a little bit the flow of arguments to provide the reader a sense of direction. The next step in our story is to glue all the pointwise data presented so far into smooth families (i.e. bundles). To produce a Dirac operator on an n-dimensional Riemann manifold (M, g) one needs several things. (a) A bundle of Clifford algebras C -> M such that Cx = Cln or Cln, V 1 6 M (b) A fiberwise injective morphism of vector bundles 1: T"M ^-> C such that \/x € M {i(u), i{v)}c = -2g(u, v) Vn, v 6 T*M. (c) A bundle of Clifford modules i.e. a vector bundle £ —> M together with a morphism c : C —¥ End (£) whose restrictions to the fibers are morphisms of algebras. (d) A connection on £. The above collection of data can be constructed from bundles associated to a common principal bundle. The symmetry group of this principal bundle has to be a Lie group with several additional features which we now proceed to describe.
424
Dirac operators
Let (V, g) denote the standard fiber of T*M and denote by Autv the group of automorphisms <j> of CUV) such that (/>(V) C V. The group Auty is defined similarly, using the complexified algebra Cl(V) instead of Cl(V). For brevity, we discuss only the real case. We need a Lie group G which admits a smooth morphism p : G —> Auty. Tautologically, p defines a representation p : G —» GL(V) which we assume is orthogonal. We also need a Clifford module c : Cl(V) —> End (E) and a representation p: G —> GL(E) such that for every v € V and any g € G the diagram below is commutative. E ^ E g\
s\
(10.1.2)
This commutativity can be given an invariant theoretic interpretation as follows. View the Clifford multiplication c : V -> End (E) as an element c £ F * ® £ ' ® & The group G acts on this tensor product and the above commutativity simply means that c is invariant under this action. In concrete applications E comes with a metric and we need to require that p is an orthogonal/unitary representation. To produce all the data (a)-(d) all we now need is a principal G-bundle P —• M such that the associated bundle P xp V is isomorphic with T*M. (This may not be always feasible due to possible topological obstructions). Any connection V on P induces by association metric connections V M on T*M and VE on the bundle of Clifford modules £ = P x^ E. (In practice one requires a little more namely that V M is precisely the Levi-Civita connection on T*M. This leads to significant simplifications in many instances.) With respect to these connections the Clifford multiplication is covariant constant i.e. VE{c{a)u)
= C ( V M Q ) + c(a)VEu
\/a e fi^Af) u € C°°{E).
This follows from the following elementary invariant theoretic result. Lemma 10.1.22 Let G be a Lie group and p : G —> Aut (E) a linear representation of G. Assume there exists eo € E such that p(g)e0 = e« Vj £ G. Consider an arbitrary principal G-bundle P —> X and an arbitrary connection V on P. Then eo canonically determines a section Ug on P xp E which is covariant constant with respect to the induced connection V B = p.(V) i.e. VEu0 = 0.
Exercise 10.1.7 Prove the above lemma.
425
The structure of Dirac operators
Apparently the chances that a Lie group G with the above properties exists are very slim. The very pleasant surprise is that all these (and even more) happen in most of the geometrically interesting situations. Example 10.1.23 Let (V, g) be an oriented euclidian space. Using the universality property of Clifford algebras we deduce that each g 6 SO (V) induces an automorphism of Cl(V) preserving V °-» Cl(V). Moreover it defines an orthogonal representation on the canonical Clifford module c : Cl(V) -» End (A*V) such that c(g • v){u) = g • (cfajfo- 1 • u>)) V € SO(V),
veV,ueA*V
i.e. SO(V) satisfies the equivariance property (10.1.2). If (M, g) is an oriented Riemann manifold we can now build our bundle of Clifford modules starting from the principal SO bundle of its oriented orthonormal coframes. As connections we can now pick the Levi-Civita connection and its associates. The corresponding Dirac operator is the Hodge-DeRham operator. The next two sections discuss two important examples of Lie groups with the above properties. These are the spin groups Spin(n) and its "complexification" Spin°(n). It turns out that all the groups one needs to build Dirac operators are these three classes: SO, Spin and Spin0. §10.1.6
Spin
Let (V, g) be a finite dimensional euclidian space. The group of automorphisms of the Clifford algebra Cl(V) contains a very rich subgroup consisting of the interior ones. These have the form (px : Cl(V) -> Cl(V)
u i-» (px(u) =x:u-x~1
Vu € Cl{V)
where x is some invertible element in Cl(V). The candidates for the Lie groups with the properties outlined in the previous section will be sought for amongst subgroups of interior automorphisms. It is thus natural to determine the subgroup {a; € Cl(V)* ; x • V • aT 1 C V} where Cl(V)* denotes the group of invertible elements. We will instead try to understand the Clifford group T(V) = {x€ CV(V) ; a(x) • V • x~l C V}
426
Dirac operators
where a : Cl(V) —> Cl(V) denotes the involutive automorphism of Cl(V) defining its Z 2 -grading. In general, the map px = {Cl(V) 3 m - t a{x)ux~x e Cl(V)} is not an automorphism of algebras but, as we will see by the end of this subsection, if x £ T(y) then px = ±ipx and hence a posteriori this alteration has no impact. Its impact is mainly on the aesthetics of the presentation which we borrowed from the elegant paper [6]. By construction T(V) comes equipped with a tautological representation p : T(V) -» GL(V)
p{x) : v ^ a(x) • v • x~\
Proposition 10.1.24 kei p = (K*, •) C Cl(V)*. Proof Clearly W c ker p. To establish the opposite inclusion choose an orthonormal basis (ej) of V and let x G ker p. x decomposes into even/odd components x = x0 + xi and the condition a(x)eix~1
= ei
translates into (x0 - xi)et = ei(x0 + Xi) Vi which is equivalent to the following two conditions [x0, et] = x0ei - eiX0 = Xie* + e,Xi = {xi, e,} = 0 Vi. In terms of the s-commutator the above two equalities can be written as one [e^ x]s = 0 Vi.
Since [•, x] is a superderivation of Cl{V) we conclude that [y,x]s = 0 V y e C / ( V ) . In particular, XQ lies in the center of Cl(V). elementary fact.
We let the reader check the following
Lemma 10.1.25 The center of the Clifford algebra is the field of scalars R C Cl{V). Note that since {xi,e;} = 0 then X\ should be a linear combination of elementary monomials ejl •••ej, none of which containing e^ as a factor. Since this should happen for every i this means X\ = 0 and this concludes the proof of the proposition. • Definition 10.1.26 The spinorial norm is the map N : Cl(V) -» Cl{V)
N(x) = x b x.
The structure ofDirac operators
427
Proposition 10.1.27 (a) N{T(V)) C M*. (b) The map N : T(V) -> M* is a group morphism. Proof Let x G r ( V ) . We first prove that xb G F(V). Since a{x)vx~l we deduce that a({a(x) • v • x~lY) = — a(x) • v • x'1 G V
G V, Vv G V
Using the fact that x >-» xb is an anti-automorphism we deduce Q( ( X ^
1
• i; • a ( x b ) ) G 1/
so that aiixY1) • v xb e V that is (a;1")-1 G T(V). Hence x* G T(^) Vx G r ( F ) . In particular, since a(T(V)) T{V), we deduce JV(r(V)) C T(V). For any v G V a(N(x))
• v • {Nix))-1
= a(x b x) • v • (x^x)" 1 = a(x b ) • {a{x) • v • x" 1 } •
C
(x^1-
On the other hand y = a(x)-v-x~l is an element in V which implies y* = a{y*) = —y. Hence a(N(x)) • v • {N{x))~l = - a ( x b ) • y* • (x^)" 1 = —a(x b ) • Q(X 1 > )~ 1 • D* • xb • ( x b ) _ 1 = — v* = v.
This means N{x) € kerp = R*. (b) If x, y G r(V) then JV(x • ?/) = (xy) k (zy) = y^x^xy = y*N{x)y = N{x)y>y = N(x)N(y).
a
Theorem 10.1.28 (a) For every x € r ( V ) the transformation p(x) of V is orthogonal. (b) There exists a short exact sequence of groups
i ->• R* -4 r(v) -4 o{v) -» I. (c) Every x G r ( V ) can be written (in a non-unique way) as a product x = v\- ••vic, Vj G V. In particular, every element of r ( V ) is Z2-homogeneous i.e. it is either purely even or purely odd. Proof
(a) Note that Vw G V we have N(v) = — \v\2. For every x G r ( V ) we get
N(p(x)(v))
= Nia^vx-1)
= 7V(a(x))iV(i;)iV(x-1) = ]V(a(x))7V(x)-1Af(u).
On the other hand x 2 = a(x 2 ) = a(x) 2 we deduce N(x)2 =
N(a(x))2
428
Dirac operators
so that N(p(x)(v)) = ±N(v). Since both N(v) and N(px(v)) are negative numbers we deduce that the only logical choice of signs in the above equality is +. Hence p{x) is an orthogonal transformation, (b) & (c) We only need to show p{T{V)} = 0(V). For x € V with \x\g = 1 we have a (a;) = —x = x"1. If we decompose v € V as Xx + u where A 6 l and u L x then we deduce p(x)v = —Xx + u. In other words, p(x) is the orthogonal reflection in the hyperplane through origin which is perpendicular to x. Since any orthogonal transformation of V is a composition of such reflections we deduce that for each T e 0(V) we can find V\,..., vk 6 V such that T =
p{vi)---p{vk).
Incidentally this also establishes (c). • Set T°(V) = T{V) n Cleven. Note that p{T°(V)) C SO{V) = {Te
0{V) ; d e t T = 1}.
Hence we have a short exact sequence
i ->• K* M- r°(v) A I. Definition 10.1.29 Set Pm(V) Spin(V)
= {xe
= {xe
T(V) ; \N(x)\ = 1}
T{V) ; N[x) = 1} = Pin(V) n r°(V).
The results we proved so far show that Spin(V) can be alternatively described by the following "friendlier" equality Spin(V)
= {v1---v2k;
k>0,Vi£
V,
\vt\ = 1, Vz = 1..
..2k}.
Proposition 10.1.30 There exist short exact sequences 1 -> Z 2 -> Pin(V) -> O(V) ->• 1 1 -> Z 2 -> 5pm(l/) ->• 5 0 ( ^ ) -^ 1. SpiniV) is a Lie group and p : Spin(V) —>• 50(V / ) is a covering map. It is connected if d i m F > 2 and simply connected if dimV > 3. In particular, Spin(V) is the universal cover of SO(V) when dim V > 3.
429
The structure of Dirac operators
Proof The exactness of the two sequences is left to the reader. It is also fairly easy to prove that p : Spin(V) —> SO(V) is a covering map using the following simple observations: (i) p is a group morphism ; (ii) p is continuous; (iii) ker p is discrete. This shows that Spin(V) can be naturally endowed with a smooth structure pulled back from SO(V) via p. Since SO(V) is connected if dim V > 2 the fact that Spin(V) is connected would follow if we showed that any points in the same fiber of p can be connected by arcs. It suffices to look at the fiber p - 1 ( l ) = {—1,1}Using Theorem 10.1.28(c) we see that Spin(V) = {v1---v2k;
k>l,Vj
€ V, |u,| = 1}.
Thus if u,v £ V are such that |u| = \v\ = 1, t i i s the path 7(£) = (u cos t + v sin t)(u cos t — v sin i), lies inside Spin(V)
0< t <
TT/2
and moreover 7(0)
= -1,
7 ( T / 2 ) = 1.
To prove that Spin{V) is simply connected if dim V > 3 it suffices to observe that each closed path in Spin(V) is homotopic to a "monomial loop" 7(*)="i(*) •••«»(*)
where Vj(i) are closed paths in the unit sphere SdlmV~1 C (V,g). If dimV > 3 then the unit sphere 5 d i m V ' _ 1 is simply connected which forces Spin(V) to be simply connected. • Corollary 10.1.31 ^(50(71)) = Z 2 if n > 3. Let Spin(n) = Spin(Rn) where E n denotes the standard euclidian space. Since the natural map p : Spin(n) —• SO(n) is a cover we deduce that its derivative at the "origin" induces an isomorphism of Lie algebras T = p* : spin(n) -=* so(n). We would like to spend some time discussing some often confusing aspects of this isomorphism. We can view Spin(V) as a submanifold of Cl(V) and as such we can identify its Lie algebra spin(V) with a linear subspace of Cl(V). The next result offers a more precise description.
430
Dirac operators
Proposition 10.1.32 Consider the quantization map q : A*V -» Cl(V).
Then
spin{V) = q ( A V ) . The Lie bracket is given by the commutator in Cl(V). Proof The group r(V) is a Lie group as a closed subgroup of the group of linear transformations of Cl(V). Since the elements of T(V) are either purely even or purely odd we deduce that the tangent space at 1 G r(V) can be identified with the subspace E = {x e Cren{V)
; xv - vx 6 V,
Vv € V}.
Fix i 6 £ and let e i , . . . , e„ be an orthonormal basis of V. We can decompose x as x = x0 + e\X\ where x0 € Cl(V)even and x\ € Cl(V)°dd are linear combinations of monomials involving only the vectors e%,... e„. Since [x0, ex] = 0 and {xi, e{\ = 0 we deduce
In particular this means xi e K © V G C7(V). Repeating the same argument with every vector e^ we deduce that i 6 R ® span{ei • ej ; 1 < i < j < dim V} = K © q(A2T0Thus T^iV)
Cleq(AV).
The tangent space to Spin(V) satisfies a further restriction obtained by differentiating the condition N(x) = 1. This gives spin{V) C { i £ l f f i q ( A V ) ; xb + x = 0} = q(A 2 V). Since dim spin(V) = dim SQ(V) = dimA 2 V we conclude the above inclusion is in fact an equality of vector spaces. Now consider two smooth paths x, y : (—e, e) —> SpiniV) such that x(0) = y(0) = 1. The Lie bracket of x(0) and y(0) is then found (using the Exercise 3.1.5) from the equality x^y^xity'yit)-1 = 1 + [i(0), y(0)}t2 + 0{t3) (as t ->• 0) where the above bracket is the commutator of x(0) and y(0) viewed as elements in the associative algebra CUV). • To get a more explicit picture of the isomorphism p» : spin{V) —• so(V)
431
The structure of Dirac operators
we fix an orientation on V and then choose a positively oriented orthonormal basis { e j , . . . ,e n } of V, (n = dim V). For every x € spin(V) the element p,{x) g spin(V) acts on V according to p„(x)v = x • v — v • x. If X — / , XijCiCj
then p*\X)ej
—
Z /
j
XjjCjj
\%ij —
^ji)'
i
Note the following often confusing fact. If we identify as usual so(V) = A2V by so(n) B A t-> U>A = ^2, g{Aei,ej)ei A e, = — ^2g(e,, Ae3)ei A ej i<j
i<j
then the Lie algebra isomorphism p, takes the form i
= 2 ^2 Xi3ei A e^ = 2a(x) € A2V. where a : C7(V) —• A*V is the symbol map, eiCj H-> ej A e^-. A word of warning If A e so_(V) has the matrix description i
with respect to an oriented orthonormal basis { e i , - - - , e „ } , (n = dimV) then the 2-form associated to A has the form ujA = ~ Yl aijet A e,so that Pi,1 {A) = -7;Y^aiJeieJ * i<3
-
4
-jY,a'Jeier i,3
The above negative sign is essential and in many concrete problems this can mean the difference between "Heaven" and "Hell". Any Clifford module <j>: Cl(V) —• Endjj (E) defines by restriction a representation $ : Spin{V) -»-
GLK(E).
The (complex) representation theory described in the previous sections can be used to determine the representations of Spin(V).
432
Dirac operators
E x a m p l e 10.1.33 ( T h e complex spinor r e p r e s e n t a t i o n s ) Consider a finite dimensional oriented euclidian space {V,g). Assume first that dimV is even. The orientation on V induces a Z 2 -grading on the spinor module § = § + © §~. Since Spin(V) C Cleven(V) we deduce that each of the spinor spaces S* is a representation space for Spin(V). They are in fact irreducible, nonisomorphic complex Spin(V)modules. They are called the positive/negative complex spin representations. Assume next that dim V is odd. The spinor module §{V) §(V) =
§+(V)®§-(V).
is not irreducible as a Cl(V) modules. Each of the modules S ± (V) is a representation space for Spin(V). They are irreducible but also isomorphic as Spin(V)-modules. If we pick an oriented orthonormal basis ei, • • •, e2n+i of V then the Clifford multiplication by a; = ei • • • e2„+i intertwines the ± components. This is a Spin(V) isomorphism since ui lies in the center of Cl(V). C o n v e n t i o n For each positive integer n we will denote by §„ the Spin(n) define by
module
§2* =* S+(R2k) 0 §-(R 2 *) (n = 2k) and § 2 * +1 = S+(K2A:+1) =Spin{ik+1)
§-(R M + 1 ) (n = 2k + 1).
§ n will be called the fundamental complex spinor module of Spin(n). Exercise 10.1.8 Let (V, g) be a 2fe-dimensional euclidian space and J : V —> V a complex structure compatible with the metric g. Thus we have an explicit isomorphism A:Cl(V) ->End(A*-V). Choose u E V such that \u\g = 1 and then for each t 6 M set q = q(t) = cost + u • vsint,
v = Jv.
Note that q € Spin(V) so that A(q) preserves the parities when acting on A*,0(V). Hence A(q) = A + (g)©A_(g) where A+/_() acts on hevenlodd'°V. Compute tr (A + / _(g)) and then conclude that hevm,/°dd,°y a r e non-isomorphic Spin(V)-modu\es. P r o p o s i t i o n 10.1.34 Let cj> : Cl(V) -> End (E) be a selfadjoint Clifford module. Then the induced representation of Spin(V) is orthogonal (unitary). Exercise 10.1.9 Prove the above proposition. Exercise 10.1.10 Prove the group Spin(V) Subsection 10.1.5.
satisfies all the conditions discussed in
433
The structure of Dirac operators §10.1.7
Spin0
The considerations in the previous case have a natural extension to the complexified Clifford algebra O n . The canonical involutive automorphism a : Cln -> Cln extends by complex linearity to an automorphism of Cln while the adjoint anti-automorphism * : Cln —> Cln extend to G „ according to the rule (v ® zf = v ® z. As in the real case set x* = a{xf and N(x) = x* • x. Let (V, g) be an euclidian space. The complex Clifford group TC(V) is defined by TC(V) = {xe Cl{V)*; a(x) • v • x~l € V Vv <E V}. We denote by pc the tautological representation pc:Tc{V)
-»GL(V,R).
As in the real case one can check that c P
(r<(v)) = o(F)
and kerp c = C*. The spinorial norm JV(:r) defines a homomorphism N : TC{V) -)• C*. Define Pm c (V) = { i £ r c ( V ) ; |AT(a;)| = 1}. We leave the reader check the following result. Proposition 10.1.35 There exists a short exact sequence 1 -> S 1 -> Pinc{V)
-> O(V) -» 1.
Corollary 10.1.36 There exists a natural isomorphism Pinc{V)
^ (Pin{V) x S 1 )/ ~
where "~" is the equivalence relation (x, z) ~ (-x, -z)
V(x, z) € Pin(V) x S 1 .
434 Proof
Dirac operators The inclusions Pin(V)
C Cl(V), S ' c C induce an inclusion
(Pin(V)
x S 1 ) / —> Cl{V).
The image of this morphism lies obviously in TC(V) n {|iV| = 1} so that (Pin(V) x S 1 ) / ~ can be viewed as a subgroup of Pinc(V). The sought for isomorphism now follows from the exact sequence 1 -> S 1 -> {Pin(V) x S1)/
> 0(V) -> 1. D
c
We define 5pm (F) as the inverse image of SO(V) via the morphism pc : Pinc(V)
-» O(V).
Arguing as in the above corollary we deduce Spinc{V) ^ (Spm(V) x S 1 ) / ~ = (Spin(V) Exercise 10.1.11 Prove Spinc(V) 10.1.5.
x S1)/^-
satisfies all the conditions outlined in Subsection
Assume now dim V is even. Then any any isomorphism Cl(V) * E n d c ( S ( F ) ) induces a complex unitary representation Spinc{V)
->• Aut (S(V))
called the complex spinorial representation of Spin0. It is not irreducible since (once we fix an orientation on V), End (§(!/)) n a s a natural superstructure and by definition Spin0 acts through even automorphism. As in the real case, §(V) splits into a direct sum of irreducible representations S ± ( F ) . Any complex structure J on V defines two things: (i) a canonical orientation on V and (ii) a natural subgroup U(V, J) = {T€
SO(V) ; [T, J] = 0} C SO{V).
Denote by ij : U(V, J) —¥ SO[V) the inclusion map. Proposition 10.1.37 There exists a natural group morphism ZJ : U{V, J) -»• Spin°{V) such that the diagram below is commutative. U(V, J) -^
Spin°{V)
The structure ofDirac operators
435
Proof Let w € U(V) and consider a path 7 : [0,1] -+ U(V) connecting 1 to w. Via the inclusion f (V) w- S O ( ^ ) we may regard 7 as a path in SO(V). As such, it admits a unique lift 7 : [0,1] —• Spin(V) such that 7(0) = 1. Using the double cover Sl -¥ S1 z >-» Z2 we can find a path 6(t) in S1 such that 5(0) = land(5 2 (t) = det7(i). Define f(u) to be the image of (7(1), 5(1)) in Spinc(V). We have to check that (i) £ is well defined and (ii) a is a smooth group morphism. To prove (i) we need to show that if 77: [0,1] —> U(V) is a different path connecting 1 to w and A : [0,1] -> S 1 is such that A(0) = 1 and A(i) = detTj(t)2 then Spinc(V).
W(1),A(1)) = ( 7 (1),*(1)) in
The elements 7(1) and jj(l) lie in the same fiber of the covering Spin(V) so that they differ by an element in ker p. Hence
A
SO(V)
7(1) = «7(1) e = ± l We can identify e as the holonomy of the covering Spin(V) —> SO(V) along the loop 7 * rj~ which goes from 1 to u) along 7 and back to 1 along rj"(t) = rj(l — t). The map det : U(V) —> S1 induces an isomorphism between the fundamental groups (see Exercise 6.2.10 of Subsection 6.2.5). Hence in describing the holonomy e it suffices to replace the loop 7 * rf C U(V) by any loop v(t) such that det v{t) = det(7 * r}~) = A(t) € Sl. Such a loop will be homotopic to 7 * rj~l in U(V) and thus in SO(V) as well. Select v(t) of the form v(t)ex = A(i)ei, u{t)ei = e{\/i>2 where (e*) is a complex, orthonormal basis of (V, J). Set /; = Je,. With respect to the real basis (ei, f\\ e2, $2, • • •) v(t) (viewed as an element of SO(V)) has the matrix description [ cos9(t) -sm6(t) •••' sin0(t) cos 6(t) ••• :
Id
where 8 : [0,1] —>• K is a continuous map such that A(t) = e i9(i) . The lift of v{t) to Spin(V) has the form
v(t) =
(cos-Y-e1f1sm-Y).
We see that the holonomy defined by v(i) is nontrivial if and only if the holonomy of the loop 11-> S(t) in the double cover S1 ^-> 5 1 is nontrivial. This means that 5(1) and A(l) differ by the same element of Z 2 as 7(1) and ^(1). This proves ^ is well defined. We leave the reader to check that f is indeed a smooth morphism of groups.
•
436 §10.1.8
Dirac operators Low dimensional e x a m p l e s
In low dimensions the objects discussed in the previous subsections can be given more suggestive interpretations. In this subsection we will describe some of these interpretations. \n= 11 The Clifford algebra Cl\ is isomorphic with the field of complex numbers C. The Z 2 -grading is SHeC © 3 m C . The group Spin(l) is isomorphic with Z 2 . \n = 2 | The Clifford algebra C72 is isomorphic with the algebra of quaternions H. This can be seen by choosing an orthonormal basis {e 1 ,e 2 } in R 2 . The isomorphism is given by 11—>• 1, ei I—> i, e 2 >->• j , eie 2 >-» k where i, j and k are the imaginary units in H. Note that Spin(2) = {o + 6k ; a, 6 e R a 2 + b2 = 1} ^ S1. The natural map Spin(l)
->• SO(2) = Sl takes the form ew H-> e2W.
\n = 3[ As an (ungraded) algebra Cl3 is isomorphic to the direct sum H © H. More relevant is the isomorphism Clfen = C72 = H given by 11-> 1, eie2 i-> i, e 2 e 3 n- j ,
e 3 ei 1-* k
where {ei,e 2 ,e 3 } is an orthonormal basis in R3. Under this identification the operation x i-» xb coincides with the conjugation in H x = a + bi + cj + dki-^x
= a — 6i — cj — dk.
In particular the spinorial norm coincides with the usual norm on H N(a + 6i + cj + dk) = a2 + b2 + c2 + d2. Thus any a; 6 C7f e n \ {0} is invertible and x'1 =
l xb JV(i) '
Moreover, a simple computation shows that
rfar1
C R 3 Vi e Clfen \ {0} so that
r°(R3)^H\{0}. Hence Spin(3) ^ { i £ l ; |z| = 1} =* SU(2).
437
The structure of Dirac operators
The natural map Spin(3) —• 5 0 ( 3 ) is precisely the map described in the Exercise 6.2.4 of Subsection 6.2.1. The isomorphism Spm(3) = SU(2) can be visualized by writing each q = a + 6i + cj 4- dk as q = u+jv u = a + M, v = (c — di) € C To a quaternion g = u + jv one associates the 2 x 2 complex matrix Sq =
u —v V
U
e su{2).
Note that S ? = S*, Vg G H. For each quaternion g G H we denote by Lq (resp. i?9) the left (resp. right) multiplication. The right multiplication by i defines a complex structure on H. Define T H -> C 2 by q = u + ju
H->
Tg =
:]
A simple computation shows that T(fl,«) = iTg i.e. T is a complex linear map. Moreover \/q G Spm(3) = S 3 the matrix Sq is in SU(2) and the diagram below is commutative.
e - ^ c2 H - ^ C2 In other words, the representation Spin{3) 9 g >->• L, G G L C ( H ) of Spin(3) is isomorphic with the tautological representation of SU(2) on C 2 . On the other hand the correspondences eie 2 i-> $ 0 $ € End (C2) © End (C2) e 2 e 3 M. Sj © Sj € End (C2) © End (C2) e 3 ei H* S k © S k G End (C2) © End (C 2 ) eie 2 e 3 i-)- i? = l c » © ( - l c 0 € End (C2) © End (C2) extend to an isomorphism of algebras C/3 —> End (C 2 )©End (C 2 ). This proves that the tautological representation of SU(2) is precisely the complex spinorial representation §3
438
Dirac operators
From the equalities [RhLq] = {RhRi}
=0
we deduce that Rj defines an isomorphism of Spin (3)-modules
This implies there exists a Spm(3)-invariant bilinear map P : § 3 x §3 ->• C.
This plays an important part in the formulation of the recently introduced SeibergWitten equations (see [78]). Exercise 10.1.12 The left multiplication by i introduces a different complex structure on EL Prove the representation Spin(3) 3 q H-> (R^ : H —> H) is (i) complex with respect to the above introduced complex structure on H and (ii) it is isomorphic with complex spinorial representation described by the left multiplication.
\n — 4 | Cl\ can be realized as the algebra of 2 x 2 matrices with entries in H. To describe this isomorphism we have to start from a natural embedding R4 <-» M 2 (H) given by the correspondence " 0 -x x 0 A simple computation shows that the conditions in the universality property of a Clifford algebra are satisfied and this correspondence extends to a bona-fide morphism of algebras C/4 —> M 2 (H). We let the reader check this morphism is also injective and a dimension count concludes it must also be surjective. Proposition 10.1.38 Spin(4) = SU{2) x SU(2) Proof We will use the description of Spin(4) as the universal (double-cover) of SO (A) so we will explicitly produce a smooth 2—1 group morphism SU(2)xSU(2) —> 5 0 ( 4 ) . Again we think of SU(2) as the group of unit quaternions. Thus each pair (qi, g2) £ SU(2) x SU(2) defines a real linear map Tqim : H -> H x i->- Tquq2x = qxxq2. Clearly |x| = 1^1 • \x\ • \q^\ = |T gii92 x| Va; G H so that each Tqim is an orthogonal transformation of H. Since SU{2) x SU(2) is connected all the operators TquQ2 belong
The structure ofDirac
439
operators
to the component of 0(4) containing 1 i.e. T defines an (obviously smooth) group morphism T : SU(2) x SU(2) ->• 5 0 ( 4 ) . Note that kerT = { 1 , - 1 } so that T is 2 — 1. In order to prove T is a double cover it suffices to show it is onto. This follows easily by noticing T is an immersion (verify this !) so that its range must contain an entire neighborhood of 1 6 5 0 ( 4 ) . Since the range of T is closed (verify this\) we conclude that T must be onto because the closure of the subgroup (algebraically) generated by an open set in a connected Lie group coincides with the group itself (see Subsection 1.2.3). • The above result shows that so(4) ^ spin(4) ^ SM(2) © S U ( 2 ) ^ so(3) © so(3). Exercise 10.1.13 Using the identification CU = M 2 (H) show that Spin(4) corresponds to the subgroup {diag (p, q);p,q€M,
\p\ = \q\ = 1} C M 2 (H).
Exercise 10.1.14 Let {e 1 ,e 2 ,e 3 ,e4} be an oriented orthonormal basis of M4. Let * denote the Hodge operator defined by the canonical metric and the above chosen orientation. Note that * : A2K4 -> A2M4 is involutive * 2 = id so that we can split A2 into the ± 1 eigenspaces of * A2K4 = A2FM4©A2_R4. (a) Show that A | = s p a n j ^ * , if,
nf}
where + 1 , Vi = ^ ( e i A e 2 ± e 3 Ae 4 ) . 1 , % = 7|(ei Ae3±e4Ae2) + 1 , % = -7s(ei A e 4 ± e 2 A e 3 ) . (b) Show that the above splitting of A2K4 corresponds to the splitting so(4) = so(3) © so(3) under the natural identification A 2 E 4 = so(4).
440
Dirac operators
To obtain an explicit realization of the complex spinorial representations §J we need to describe a concrete realization of the complexification CZ4. We start from the morphism of M-algebras B x = u + jf i-> Sx This extends by complexification to an isomorphism of C-algebras H®RC^M2(C). (Verify this\) We now use this isomorphism to achieve the identification. E n d H ( H © H)
Bx^Tx
=
0 0
6 End (C 2 © C 2
Note that the chirality operator F = —6x626364 is represented by the canonical involution
r i-> iC2 e (-ica)We deduce the §J representations of Spin(4) = SU(2) x SU(2) are given by §J : SU{2) x 51/(2) B (p, q) n . p € GL(2; C) §4" : 51/(2) x SU(2) 3(p,q)^qe
GL(2; C).
As for Spin(3) these representations can be given quaternionic descriptions. Exercise 10.1.15 The space K4 = H has a canonical complex structure defined by Ri which defines (following the prescriptions in §10.1.3) an isomorphism c : O4 —• End(A*C 2 ). Identify C 2 in the obvious way with AXC2 and with A e " e "C 2 via ex i-> 1 G A°C2, e2 >-> e.\ A e 2 . Show that under this identification we have c{x) = TX Vz € M4 = (M, Ri) =* C 2 where Tx is the odd endomorphism of C 2 © C 2 defined above. Exercise 10.1.16 Let V be a 4-dimensional oriented euclidian space. Denote by q : A*V —> Cl(V) the quantization map and fix an isomorphism A : Cl(V) -¥ End (§(f^)) of Z 2 -graded algebras. Show that for any 77 e A2+(V) the image Aoq(r]) e End (§(V)) is an endomorphism of the form T © 0 e End (§ + (^)) © End (§~(V)).
441
The structure of Dirac operators
Exercise 10.1.17 Denote by V a 4-dimensional oriented euclidian space. (a) Show that the representation SJ ® §4" of Spm(4) descends to a representation of 5 0 ( 4 ) and moreover S+(V) ® c S+(V) = (A°(V) © A2+(V)) ® M C as SO{4) representations. (b) Show that §+{V) ^ §+{V) as Spin{4) modules. (c)The above isomorphism defines an element <j> £ § + (V) ® £ § + (K). Show that via the correspondence at (a) the isomorphism <j> spans A°(V). §10.1.9
Dirac bundles
In this subsection we discuss a distinguished type of Clifford bundle which is both frequently encountered in applications and is rich in geometric informations. We will touch only the general aspects. The special characteristics of the most important concrete examples are studied in some detail in the following section. In the sequel all Clifford bundles will be assumed to be complex. Definition 10.1.39 Let E —> M be a Clifford bundle over the oriented Riemann manifold (M, g). A Dirac structure on E is a pair (h, V) consisting of a Hermitian metric h on E and a Clifford connection i.e. a connection V compatible with h such that (i) Va e fix(M) the Clifford multiplication by a is a skew-Hermitian endomorphism of E. (ii) Va eQ}(M),X e Vect {M),ue C°°(E) V j ( c ( a ) u ) = c{V^a)u
+ c(a)(Vx«)
where V M denotes the Levi-Civita connection on T*M. (This condition means the Clifford multiplication is covariant constant). A pair (Clifford bundle, Dirac structure) will be called a Dirac bundle. A Z^-grading on a Dirac bundle (E, h, V) is a Z 2 grading of the underlying Clifford structure E = E0 © E\ such that h = h0 © hi and V = V° © V 1 where ht and resp. V 1 is a metric (resp. a metric connection) on Ei.
The next result addresses the fundamental consistency question: do there exist Dirac bundles? Proposition 10.1.40 Let E —> M be a Clifford bundle over the oriented Riemann manifold (M,g). Then there exist Dirac structures on E. Proof Denote by V the (possible empty) family of Dirac structure on E. Note that if %, V{) €V{i = l,2) and / G C°°{M) then (/Ax + (1 - })h2, / V 1 + (1 - / ) V 2 ) e V.
442
Dirac operators
This elementary fact shows the existence of Dirac structures is essentially a local issue: local Dirac structures can be patched-up via partitions of unity. Thus it suffices to consider only the case when M is an open subset of K" and E is a trivial vector bundle. On the other hand we cannot assume the metric g is also trivial (i.e. euclidian) since the local obstructions given by the Riemann curvature cannot be removed. We will distinguish two cases. A. n = d i m M is even. The proof will be completed in three steps. Step 1. A special example. Fix a selfadjoint endomorphism c : C/„ -» End (§„) where § n denotes the fundamental spinor representation. We will continue to denote by c the restriction to Spin(n) w- Cln. Fix a global, oriented, orthonormal frame (e^) of TM and denote by (e J ) its dual coframe. Denote by u> = («y) the connection 1-form of the Levi-Civita connection on T*M with respect to this moving frame i.e. Ve J = we7' = Y^uij ® e'>
w G
fi(M) ® so(n).
i
Using the canonical isomorphism p, : spin(n) —> so(n) we define UJ = p~l{u) = - - ] T V ' V • ej G n ' ( M )
i<j
This defines a connection V on the trivial vector bundle SM by V S u = du - \ ] T U y ® c(e i ) • c{e])u 2
VM G C ° ° ( § M ) .
«<j
The considerations in §10.1.5 show that V 5 is indeed a Clifford connection so that SM is a Dirac bundle. Step 2. Constructing general Dirac bundles Fix a Dirac bundle (E, hE, V B ) over M. For any Hermitian vector bundle (W, hw) equipped with a Hermitian connection Vw we can construct the tensor product E
443
The structure of Dirac operators
the proof is concepually identical. The straightforward details are left to the reader.
• Denote by (E,h,V) a Dirac bundle over the oriented Riemann manifold (M, g). The exists a Dirac operator on E canonically associated to this structure D = c o V : C°°{E) 4 C^iT'M
® E) A
C°°(E).
A Dirac operator associated to a Dirac structure is said to be a geometric Dirac operator. Proposition 10.1.41 Any geometric Dirac operator is formally selfadjoint. Proof The assertion in the above proposition is local so we can work with local orthonormal moving frames. Fix xo G M and denote by (xl) a collection of normal coordinates near x0. Set e* = ^ - . Denote by (e1) the dual coframe of (e;). If (h, V) is a Dirac structure on the Clifford bundle E then at XQ the associated Dirac operator can be described as D = '£c{J)Vi (V, = V e j ). i
We deduce D* = (V i )*c(e i )*. Since the connection V is compatible with h and div(e;) | x o = 0 we deduce
while C(eT
= -c(e j )
since the Clifford multiplication is skew-Hermitian. Hence, at x0
i
i
Since the Clifford multiplication is covariant constant and (Vf^e1) 1^,,= 0 we conclude [Vi,c{J)]\x*=0 This concludes the proof of the proposition.
Vi. D
Let D be a geometric operator associated to the Dirac bundle (E,h,V). definition D2 is a generalized Laplacian and consequently D2 = V*V + 5H
By
444
Dirac operators
where V is a connection on E and 9 t is the Weitzenbock remainder — an endomorphism of E. For geometric Dirac operators, V = V (!!!) and this remainder can be given a very explicit description with remarkable geometric consequences. To formulate it we need a little foundational work. Let (E, h, V) be a Dirac bundle over the oriented Riemann manifold (M, g). Denote by Cliff (M) —> M the bundle of Clifford algebras generated by {T*M, g). The curvature F ( V ) of V is a section of A 2 T*M®End (E). Using the quantization map q : A'T'M -> Cliff (M) we get a section q(F) € C 7 i / / ( M )
F(V) = J2ei/\eP®Fij
and c(F) = Y.c{ei)c{(J)Fij
= i^c(e')c(^)f;.
Theorem 10.1.42 (Bochner-Weitzenbock) Let D be a the geometric Dirac operator associated to the Dirac bundle (E, h, V) over the oriented Riemann manifold (M,g). Then £>2 = V*V + c(F(V)). Proof Fix x € M and then choose an oriented, local orthonormal moving frame (ei) of TM near x such that
where V M denotes the Levi-Civita connection. Such a choice is always possible because the torsion of the Levi-Civita connection is zero. Finally denote by (e1) the dual coframe of (ej). Then
Since [V;, c(ej)] \x= 0 we deduce D21,= E c(ei)c(e>) V.V, = - £ V 2 + £ c(e i )c(^) V, V,-
Fundamental
445
examples
i
i<j
Y,c(ei)c(eJ)Fij(V).
= "£V? + i
i<j
We want to emphasize again the above equalities hold only at x. The theorem now follows by observing that
(v*v)u=-^v?ju. n Exercise 10.1.18 Let (E, h, V) be a Dirac bundle over the oriented Riemann manifold (M, g) with associated Dirac operator D. Consider a Hermitian bundle W —> M and a connection V w compatible with the Hermitian metric hw(a) Show that (E ® W, h
e € C°°(E), w € C°°(W).
We denote by Dw the corresponding geometric Dirac operator. (b) Denote by cw(F{Vw)) the endomorphism of E ® W defined by the sequence F ( V W ) € C°°(A 2 T*M®End(W^)) A C°°(C/(T*M)®End(W)) H% C°°(End(E®VF)). Show that D2W = V*V + c(F(V)) +
10.2
cw(F{Vw)).
Fundamental examples
This section is entirely devoted to the presentation of some fundamental examples of Dirac operators. More specifically we will discuss the Hodge-DeRham operator, the Dolbeault operator the spin and spin0 Dirac. We will provide more concrete descriptions of the Weitzenbock remainder presented in Subsection 10.1.9 and show some of its uses in establishing vanishing theorems. §10.2.1
The H o d g e - D e R h a m operator
Let (M, g) be an oriented Riemann manifold and set A'CT*M = A*T*M
446
Dirac operators
is a Dirac operator. In fact, we will prove this operator is a geometric Dirac operator. Continue to denote by g the Hermitian metric induced by the metric g on the complexification A^T'M. V s will denote the Levi-Civita and its associates. When we want to be more specific about which Levi-Civita connection we are using at a given moment we will indicate the bundle it acts on as a superscript. E.g., V T * M is the Levi-Civita connection on T'M. Proposition 10.2.1 The pair (g, V 9 ) defines a Dirac structure on the Clifford bundle A^T'M and d + a" is the associated Dirac operator. Proof In Subsection 4.1.5 we have proved that d can be alternatively described as the composition C°°(A*T*M) 4 C'X'iT* ® A'T'M)
4
C°°(A'T'M)
where e denotes the exterior multiplication map. Thus d + d* = eoV
+ V*oe*.
If Xi, • • •, Xn is a local orthonormal frame of TM and 8l, • • •, 9" is its dual coframe then for any ordered multi-index I we have £*(0') = £ * * / • i Thus, for any u e C°°(A*T*M) V* o e* = - Y^ Vxkixk
+ div
(Xk)iXku.
Fortunately, we have the freedom to choose the frame (Xk) in any manner we find convenient. Fix an arbitrary point XQ £ M and choose {Xk) such that at xo we have Xk = 7T-, where (xk) denotes a collection of normal coordinates near x0. With such a choice div (Xk) = 0 at x0 and thus d*uj\xo= - ] T i 9 t V a t u ; . This shows d + d* can be written as the composition C°°(A*T*M) 4 C^iTM
® A'T'M)
A- C°°(A*T*M).
where c denotes the usual Clifford multiplication on an exterior algebra. We leave the reader verify that the Levi-Civita connection on A'T'M is indeed a Clifford connection i.e. the Clifford multiplication is covariant constant. This shows that d + d* is a geometric Dirac operator. D
Fundamental examples
447
We want to spend some time elucidating the structure of the Weitzenbock remainder. First of all, we need a better description of the curvature of V s viewed as a connection on A*T*M. Denote by R the Riemann curvature tensor i.e. the curvature of the Levi-Civita connection V 9 : C°°(TM) -> C°°(r*M
Va G fix(M) X, Y, Z G Vect (M).
= -a{R(Y, Z)X)
Lemma 10.2.2 fl is the curvature of the Levi-Civita connection V r M . Proof
The Levi-Civita connection on T*M is determined by the equalities {VTz'Ma){X)
= Z • a{X) - a{VTzMX)
Va G fi^M), X, Z e Vect (M).
Derivating along F 6 Vect (M) we get (VyVza)(X) = y • (Vza)(X) - (Vza)(VyX) = y • Z • a(X) - y • a(VzX) Similar computations give Vz^ya
- Z • a(VyX) - a(VzVyX).
and V[yiZ] and we get
(RT"M(Y, Z)a){X)
= -a(R™(Y,
Z)X).
D
The Levi-Civita connection V 9 = V r * M extends as an even derivation to a connection on K*T*M. More precisely, for every X € Vect (M) and any a i , . . . , ak G Q}(M) we define Vgx(ai A - A « k ) = (V^aj) A • • • A a t + • • • +
ai
A •'• • A (V^a/t).
Lemma 10.2.3 The curvature of the Levi-Civita connection on A*T*M is defined by RA'T'M{X,
y ) ( a x A • • • A a t ) = (R(X, Y)ai)
A • • • A ak + • • • + at A • • • A (R(X,
for any vector fields X, Y and any 1-forms a i , . . . , at-
Y)ak)
448
Dirac operators
Exercise 10.2.1 Prove the above lemma. The Weitzenbock remainder of the Dirac operator d + d* is c(fl A * T * M ). To better understand its action we need to pick a local, oriented, orthonormal moving frame (ej) of TM. We denote by (e') its dual coframe. The Riemann curvature tensor can be expressed as R = £ j A ePRij i<3
where Rij is the skew-symmetric endomorphism Rij = R{eh ej) : TM -> TM. Thus C{RA-T-M)
=
Y.c{ei)c{e^'T-M{el,ej)
=1-J2c(ei)c(^)RA'T'M(ei,ej)
i<3
i,j
J
where c(e ) = e(e^) — i(e^). Note that since both V*V and (d + d*)2 preserve the Z-grading of AfcT'M = ©fcA^.T*M so does the Weitzenbock remainder and consequently it must split as A T M
C(R
' ' ) = e,> 0 **.
Since 91° = 0 so the first interesting case is 91 1 . To understand its form pick normal coordinates (x{) near x 0 € M. Set et = £. |Xo(= T^M and e' = dx* |XoG T*0M. At xo the Riemann curvature tensor R has the form R=Y,ekAeeR(ek,ee) k<e
where R(ek, et)ej = R)klei = Rijkt^i- Using Lemma 10.2.2 we get R{ek,ee)eJ
= Rijue1.
Using this in the expression of 9\} at x 0 we get
3
k,l i,j
We need to evaluate the Clifford actions in the above equality. c ( e V = eeAei-
Sie
and c(e*)c(eV = ek A el A el - 5itek - 8kee{ + 6ikee.
Fundamental
449
examples
Hence ^ ( E
a eJ
J ) = \ E
<*jRiM A e ' A e j - 5uek - foe* + 5ikel).
Using the first Bianchi identity we deduce that Y, RijUek A e£ A e* = - £
R
^
A ek A ee = 0 Vj.
Because of the skew-symmetry i2„M = —Rijtk we conclude that J ] ajRijktSkte' = 0. Hence
s R ' ( E y ) = ^ E ajRiM6^i = 9 E ^RijU^
i ~ 9E
i,i,e
*«c*)
a R
3 Hkiek
i,i,k
= J2 ajRijikek
=E
i,j,k
otiRjkek
jk
where i
denotes the Ricci tensor at x 0 . Hence m1 = Ric.
(10.2.1)
In the above equality Ric is regarded (via the metric duality) as a selfadjoint endomorphism of T*M. The identity (10.2.1) has a beautiful consequence. Theorem 10.2.4 (Bochner)Let (M, g) be a compact, connected, oriented Riemann manifold. (a) If the Ricci tensor is non-negative definite then b\{M) < d i m M . (b) If Ricci tensor is non-negative definite but is somewhere strictly positive definite then bx{M) = 0. (Recall that &i(M) denotes the first Betti number of M). The above result is truly remarkable. The condition on the Ricci tensor is purely local but with global consequences. We have proved a similar result using geodesies (see Myers Theorem, Subsection 5.2.2 , 6.2.4 and 6.2.5). Under more restrictive assumptions on the Ricci tensor (uniformly positive definite) one deduces a stronger
450
Dirac operators
conclusion namely that the fundamental group is finite. If the uniformity assumption is dropped then the conclusion of the Myers theorem no longer holds (think of the flat torus). In particular this result gives yet another explanation for the equality Hl{G) = 0 where G is a compact semisimple Lie group. Recall that in this case the Ricci curvature is j x {the Killing metric}. Proof (a) Let Ai denote the metric Laplacian Ai = dd* + d*d : QX(M) ->
Ql(M).
Hodge theory asserts that h(M)
= dimker Aj
so that in order to estimate the first Betti number we need to estimate the "number" of solutions of the elliptic equation AX77 = 0 77 e f t 1 . Using the Bochner-Weitzenbock theorem and the equality (10.2.1) we deduce A1?7 = V* V77 + Ric T? = 0 on M. Taking the L 2 -inner product by 77 and then integrating by parts we get f \Vrj\2dvg+
f (RicV,V)dvg
= 0.
'
(10.2.2)
Since Ric is non-negative definite we deduce Vr? = 0. Hence any harmonic 1-form must be covariant constant. In particular, since M is connected, the number of linearly independent harmonic 1-forms is no greater than the rank of T*M = dim M. (b) Using the equality (10.2.2) we deduce that any harmonic 1-form 77 must satisfy (Ric(x)7jI,77I)I = 0
Vi€tf.
If the Ricci tensor is positive at some x0 G M then r/(x0) = 0. Since 77 is also covariant constant and M is connected, we conclude that 77 = 0. • Remark 10.2.5 For a very nice survey of some beautiful applications of this technique we refer to [9],
Fundamental §10.2.2
451
examples
The Dolbeault operator
This subsection introduces the reader to the Dolbeault operator which plays a central role in complex geometry. Since we had almost no contact with this beautiful branch of geometry we will present only those aspects concerning the "Dirac nature" of these operators. To define this operator we need a little more differential geometric background. Definition 10.2.6 (a) Let E —> M be a smooth real vector bundle over the smooth manifold M. An almost complex structure on E is an endomorphism J : E —>• E such that J 2 = —If;. (b) An almost complex structure on a smooth manifold M is an almost complex structure J on the tangent bundle. An almost complex manifold is a pair (manifold, almost complex structure). Note that any almost complex manifold is necessarily even dimensional and orientable so from the start we know that not any manifold admits almost complex structure. In fact the existence of such a structure is determined by topological invariants finer than the dimension and orientability. Example 10.2.7 (a) Any complex manifold is almost complex. Indeed any such manifold is locally modeled by C n and the transition maps are holomorphic maps C n —> C n . The multiplication by i defines a real endomorphism on K2n = C n which induces the almost complex structure on TM. (b) For any manifold M the total space of its tangent bundle TM is an almost complex manifold. If (xl) are local coordinates on M and (X*) are the coordinates introduced in the fibers of TM {TM 3 X = X'-g^) then the almost complex structure on TM is determined by _d_ _d_ _d_ __9_ dxi ^ ~dX~i dXi ^ dx{' We let the reader check this is a well defined operator (independent of the choice of local coordinates). Let (M, J) be an almost complex manifold. Using the results of Subsection 2.2.5 we deduce that the complexified tangent bundle TM ® C splits as TM ® C = (TM)1'0 ® (TM)0-1. The complex bundle TM1'0 is isomorphic (over C) with (TM, J). By duality J induces an almost complex structure in the cotangent bundle T*M and similarly we get a decomposition T*M ® C = (T'M)1'0
©
(T'M)0'1.
452
Dirac operators
In turn this defines a decomposition A*CT'M = 0
A™T*M.
We set W'q{M) = C°°(A"'9T*M). Example 10.2.8 Let M be a complex manifold. If (z3 = x3 +\y3) are local holomorphic coordinates on M then (TM)lfi is generated (locally) by the complex tangent vectors 1
dzi ~ 2 \dxi lfi
while (T*M)
dyi)
is locally generated by the complex 1-forms dz3 = dx3 + \y3.
(TM)1'0 is also known as the holomorphic tangent space while (T'M)1'0 is called the holomorphic cotangent space. The space (T'M)0'1 is called the anti-holomorphic cotangent space and is locally generated by the complex 1-forms dz3 = dx3 — iy3. Note that any (p, g)-form r\ € Q,p,q(M) can be locally described as 77=
nA,BdzAAdzB
£
(7M, B eC°°(M,C))
\A\=p,\B\=q
where we use capital Latin letters A, B, C... each such index A dzA = dzAl
to denote ordered multi-indices and for
A • • • A dzA".
dzB is defined similarly. Exercise 10.2.2 Let (M, J) be an arbitrary smooth almost complex manifold. (a) Prove that dfi p '"(M) C W+2«-\M)
® fip+1'?(M) © n M + 1 ( M ) f i p - 1 , 9 + 2 ( M ) .
(b) Show that if M is a complex manifold then dW'q{M) C fip+1'«(M) 0 Q M + 1 ( M ) . (10.2.3) (The converse is also true and is known as the Newlander-Nirenberg theorem. Its proof is far from trivial. For details we refer to the original paper [62]). Definition 10.2.9 An almost complex structure on a smooth manifold is called integrable if it can derived from a holomorphic atlas, i.e. an atlas in which the transition maps are holomorphic.
Fundamental
453
examples
Thus, the Newlander-Nirenberg theorem mentioned above states that the condition (10.2.3) is necessary and sufficient for an almost complex structure to be integrable. Exercise 10.2.3 Assuming the Newlander-Nirenberg theorem prove that an almost complex structure J on the smooth manifold M is integrable if and only if the Nijenhuis tensor N e CX{T*M®2 ® TM) defined by N{X, Y) = [JX, JY] - [X, Y] - J[X, JY] - J{JX, Y] VX, Y € Vect (M) vanishes identically. In the sequel we will exclusively consider only complex manifolds. Let M be such a manifold. The above exercise shows that the exterior derivative d-Mkc{M)
^fi^+1(M)
splits as a direct sum 0, =
(&p+q=kd
where dM = {d : n M ( M ) -> 9P+1'"{M) ©
QM+1{M)}.
The component Qp'q —> Qp+l'q is denoted by d = dp'q while the other component QP,9 _>. QP,Q+I
is
denoted by d = W'q.
Example 10.2.10 In local holomorphic coordinates (zl) the action of the operator d is described by d I E VABdzA A dzB) )
\A,B
= Y, (-l)lA]~^dzA j,A,B
A dz' A dzB.
"*
It is not difficult to see that
g p,,+1 oa p " = o v P ) 9 . In other words for any 0 < p < dim£ M the sequence
0->(l f '°(I)4f!'''(M)A.is a cochain complex known as the p-th Dolbeault complex of the complex manifold M. Its cohomology groups are denoted by H^'q(M). Lemma 10.2.11 The Dolbeault complex is an elliptic complex.
454
Dirac operators
Proof The symbol of W'q is very similar to the symbol of the exterior derivative. For any x G M and any £ 6 T*X (7(9™) ( 0 : AMTX*M -> AP'«+1TI*M is (up to a multiplicative constant) the (left) exterior multiplication by £0'1 where £ 0,1 denotes the (0,1) component of £ viewed as an element of the complexified tangent space. More precisely
where J0 : T*M —> T*M denotes the canonical complex structure induced on T*M by the holomorphic charts. The sequence of symbols is the cochain complex 0 -» A"-0 ® A0-0 "W-jyp1*
A *'°
® A"-0 ® A0-1 "*-_»$»*
...
This complex is the (Z-graded) tensor product of the trivial complex
0 -> Ap'° 4 Ap'° -> 0 with the Koszul complex 0 _>. A0'0 (-!)!^' lA A 0 - 1 (--1Z^lA
....
Since f0'1 / 0 for any f / 0 the Koszul complex is exact (see Subsection 7.1.3) . This proves the Dolbeault complex is elliptic. • To study the Dirac nature of this complex we need to introduce a Hermitian metric h on TM. Its real part is a Riemann metric g on M and the canonical almost complex structure on TM is a skew-symmetric endomorphism with respect to this real metric. The associated 2-form Qh = —3m h is nondegenerate in the sense that fin (n = diiri(£ M) is a volume form on M. According to the results of Subsection 2.2.5 the orientation of M defined by ft" coincides with the orientation induced by the complex structure. We form the Dolbeault operator
9 + 9*:n p -*(M)->Q p '*(M). Proposition 10.2.12 %/2(d + d ) is a Dirac operator. Proof
We need to show that
(a(d)(0 - <7(6>)(0f = - ^ I 2 i d
V£ e T*M.
Fundamental
455
examples
Denote by J the canonical complex structure on T*M and set n = J£. Note that i ± j] and |£| = \r]\. Then
*(5)(0 = (-l) p ^K + ^) = ^ ( 0 + «(»?)) where as usual e(-) denotes the (left) exterior multiplication. The adjoint of
-
1
(-iyo(d)(o* = -(»(r) - Kfa*)) where £* (resp. r;*) denotes the metric dual of £ (resp. ?j). We deduce ( a ( 9 ) ( 0 - a ( 9 ) ( 0 ' ) 2 = \ ( e ( 0 + i«fa) - <«*) + «fa*)} 2 = J {(e(0 - * ( 0 ) + i(c(>7) + ^ * ) ) } 2 =' J { c ( 0 + icfo)} 2 Note that
c(o2 = -(e(?wn+*(r)e(0) = -iei2
and C(T?)2 = e(r/)i(r/*) + ifa*)e(J7) = \r,\2. On the other hand since £ X 77 we deduce as above that c(Ocfa) + c(i7)c(O = 0. (Verify this!) Hence
{ a(9)(0 - a(9)(0* }2 = - J(|£|2 + M2) = - ^ H 2 . • A natural question arises as to when the above operator is a geometric Dirac operator. Note first that the Clifford multiplication is certainly skew-adjoint since it is the symbol of a formally selfadjoint operator. Thus all we need to inquire is when the Clifford multiplication is covariant constant. Since c(0 =
^7T ( e { 0 + ie(??) ~ m
~ii(77,))
where 77 = J£ we deduce the Clifford multiplication is covariant constant if V 9 J = 0. Definition 10.2.13 Let M be a complex manifold and h a Hermitian metric on TM (viewed as a complex bundle). Then h is said to be a K a h l e r m e t r i c if V J = 0 where V denotes the Levi-Civita connection associated to the Riemann metric 9te h and J is the canonical almost complex structure on TM. A pair (complex manifold, Kahler metric) is called a Kahler manifold.
456
Dirac operators
Exercise 10.2.4 Let (M, J) be an almost complex manifold and h a Hermitian metric on TM. Let Q = - 3 m / i . Using the exercise 10.2.3 show that if
then the almost complex is integrable and the metric h is Kahler. Conversely, assuming that J is integrable and h is Kahler show that
We see that on a Kahler manifold the above Clifford multiplication is covariant constant. In fact, a more precise statement is true. Proposition 10.2.14 Let M be a complex manifold and h a Kahler metric on TM. Then the Levi-Civita induced connection on A°'*T*M is a Clifford connection with respect to the above Clifford multiplication and moreover the Dolbeault operator \/2((9 + d*) is the geometric Dirac operator associated to this connection. Exercise 10.2.5 Prove the above proposition. Example 10.2.15 Let (M,g) be an oriented 2-dimensional Riemann manifold (surface). The Hodge * operation defines an endomorphism *:TM
-> TM
2
satisfying * = —ITM i.e. an almost complex structure on M. Using the Exercise 10.2.4 we deduce this almost complex structure is integrable since, by dimensionality dQ = 0 where tt is the natural 2-form £1{X,Y) = g{*X,Y) X, Y e Vect (M). This complex structure is said to be canonically associated to the metric. Example 10.2.16 Perhaps the favorite example of Kahler manifold is the complex projective space C n . To describe this structure consider the tautological line bundle Li —} OP". It can be naturally viewed as a subbundle of the trivial bundle C" + 1 —> OP". Denote by ho the canonical Hermitian metric on C" + 1 and by V° the trivial connection. If we denote by P : C" + 1 —• Lx the orthogonal projection then V = P o V° | i t defines a connection on L\ compatible with h\ = h l^. Denote by UJ the 1st Chern form associated to this connection
and set hFS(X, Y)x = -ux(X,
JY) + iu{X, Y)
Vx € M, X, Y 6 TXM.
Then hps is a Hermitian metric on OP" (verify!) called the Fubini-Study metric. It is clearly a Kahler metric since (see the Exercise 10.2.4) dQh = —du> = — dc x (V) = 0.
457
Fundamental examples
Exercise 10.2.6 Describe hpS in projective coordinates and then prove that h is indeed a Hermitian metric (i.e. it is positive definite). Remark 10.2.17 Any complex submanifold of a Kahler manifold is obviously Kahler. In particular, any complex submanifold of QP n is automatically Kahler. A celebrated result of Chow states that any complex submanifold of OP" is automatically algebraic i.e. it can be defined as the zero set of a family of homogeneous polynomials. Thus all complex nonsingular algebraic varieties admit a natural Kahler structure. It is thus natural to ask whether there exist Kahler manifolds which are not algebraic. The answer is positive and a very thorough resolution of this problem is contained in the famous Kodaira embedding theorem which provides a simple necessary and sufficient condition for a compact complex manifold to be algebraic. For this result (and many more others) Kodaira was awarded the Fields medal in 1954. His proofs rely essentially on some vanishing results deduced from the Weitzenbock formulae for the Dolbeault operator d + d and its twisted versions. A very clear presentation of this subject can be found in the beautiful monograph [32]. §10.2.3
The spin Dirac operator
Like the Dolbeault operator, the spin Dirac operator exists only on manifolds with a bit of extra structure. We will first describe this new structure. Let (Mn,g) be an n-dimensional, oriented Riemann manifold. In other words, the tangent bundle TM admits an SO(n) structure so that it can be defined by an open cover (Ua) and transition maps ga/3 • Uaft ->•
SO{n)
satisfying the cocycle condition. The manifold is said to posses a spin structure if there exist smooth maps 9Q/3 : Uap ->
Spin{n)
satisfying the cocycle condition and such that P(9<*p) = gap V Q , j3
where p : Spin{n) -> SO(ri) denotes the canonical double cover. The collection gaf} as above is also called a lifting of the SO(n) structure. A pair (manifold, spin structure) is called a spin manifold. Not all manifolds admit spin structure. To understand what can go wrong let us start with a trivializing cover i l = (Ua) for TM with transition maps gap and such that all multiple intersection £/ajg...7 are contractible. In other words, i l is a good cover. Since each of the overlaps Uap is contractible each map gap : Uap —> SO(n) admits at least one lift Qccfs : Uap ->
Spin{n).
458
Dime operators
From the equality p(gapgp7g^a) = gapgpjgja
= 1 we deduce
e
ap-r = gapgi3jgja 6 kerp = Z 2 .
Thus any lift of the gluing data gap to Spin(n) produces a degree 2 Cech cochain of the trivial sheaf Z 2 namely the 2-cochain Ua/3-r •""* e<*P-y-
(«•)
Note that for any a, /?, 7, S such that Uap~,6 ^ 0 we have (•Pf6 — ia-yd + €aP8 ~ (-aPj
= 0 €
Z2.
In other words, e. defines a Cech 2-cocycle and thus defines an element in the Cech cohomology group H2(M, Z 2 ). It is not difficult to see this element is independent of the various choices: the cover At, the gluing data gap and the lifts gap. This element is intrinsic to the tangent bundle TM. It is called the second Stiefel-Whitney class of M and it is denoted by w2(M). We see that if w 2 (M) / 0 then M cannot admit a spin structure. In fact, the converse is also true. Proposition 10.2.18 An oriented Riemann manifold M admits a spin structure if and only if w 2 (M) = 0. Exercise 10.2.7 Prove the above result. Remark 10.2.19 The usefulness of the above proposition depends strongly on the ability of computing w2- This is a good news/bad news situation. The good news is that algebraic topology has produced very efficient tools for doing this. The bad news is we will not mention them since it would lead us far astray. See [46] and [58] for more details. Remark 10.2.20 The definition of isomorphism of spin-structures is rather subtle (see [56}). More precisely, two spin structures defined by the cocycles g.. and h„ are isomorphic if there exists a collection e a € Z 2 C Spin(n) such that the diagram below is commutative for all x e Uap Spin(n)
-^-
.(*) Spin(n)
-^-
Spin(n) ^/3oW
Spin(n)
The group Hl(M,Z2) acts on Spin(M) as follows. Take an element e G Hl{M,Z2) represented by a Cech cocycle, i.e. a collection of continuous maps eap\ : Uap —> Z a C Spin(n) satisfying the cocycle condition E-aP ' £ / ? 7 * ^ 7 «
•!•-
Fundamental
examples
459
Then the collection e„ • g„ is a Spin(ri) gluing cocycle defining a spin structure we denote by e • a. It is easy to check that the isomorphism class of e • a is independent of the various choice( i.e Cech representatives for e and a). Clearly the correspondence HX(M, Z 2 ) x Spin(M) defines a left action of H1(M,Z2)
3 (e, a) >-> e • a €
on Spin(M).
Spin{M)
This action is transitive and free.
Exercise 10.2.8 Prove the above proposition and the statement in the above remark. Exercise 10.2.9 Describe the only 2 spin structures on S 1 . E x a m p l e 10.2.21 (a) A simply connected Riemann manifold M of dimension > 5 admits spin structures if and only if every compact orientable surface embedded in M has trivial normal bundle. (b) A simply connected four-manifold M admits spin structures if and only if the normal bundle N-% of any embedded compact, orientable surface E has even Euler class i.e.
I e(JVE) is an even integer. (c) Any compact oriented surface admits spin structures. Any sphere 5™ admits a unique spin structure. The product of two spin manifolds is canonically a spin manifold. (d) w2(MPn) = 0 iff n = 3 (mod 4) while OP" admits spin structures iff n is odd. Let (M n , g) be a spin manifold. Assume the tangent bundle TM is defined by the open cover (Ua) and transition maps Qap • Ua0 -» SO(n). Moreover assume the spin structure is given by the lifts (jap : Uap ->
Spin(n).
As usual, we regard the collection gap as defining the principal SO(n) bundle of oriented frames of TM. We call this bundle Pso(M)- The collection gap defines a principal Spin(n) bundle which we denote by Pspin(M)- We can regard PSO(M) as a bundle associated to PsPm(M) via p : Spin(n)•—• SO(n). Using the unitary spinorial representation A„ : Spin(n) -> Aut (§ n ) we get a Hermitian vector bundle S(M) = Pspin(M) >
460
Dirac operators
called the complex spinor bundle. From Lemma 10.1.22 we deduce that §(M) is naturally a bundle of Cl(TM) modules. Then, the natural isomorphism Cl(TM) = Cl(T*M) (via the metric) induces on §(M) a structure of C7(T*M)-module. As it turns out §(M) has a natural Dirac structure whose associated Dirac operator is the spin Dirac operator on M. We will denote it by D. Since An is a unitary representation the spinorial bundle. §(M) comes with a natural metric with respect to which the Clifford multiplication is self-adjoint. All we now need is to describe a natural connection on §(M) with respect to which the Clifford multiplication is covariant constant. We start with the Levi-Civita connection V* which we can regard as a connection on the principal bundle Pso(M)- Alternatively, V 9 can be defined by a collection of so(n)-valued 1-forms ua £ £ll(Ua) ® so.{n) such that w
/3 = 9a0d9ap + 9Zf)Uaga0 on Uap.
Denote by r the canonical isomorphism of Lie algebras T : spin(n) —> so(n). Then the collection Qa = r _ 1 (w a ) defines a connection V on the principal bundle Pspin(M) and thus via the representation A n it defines a connection V = V on the spinor bundle §(M). The above construction can be better visualized if we work in local coordinates. Choose a local, oriented orthonormal frame (ej) of TM\uo and denote by (e-7) is dual coframe. The Levi-Civita connection has the form Vej = ek ® u>'kjei so that ua = ek ® (u}3ki) where for each k the collection (oj^ij is a skew-symmetric matrix. Using the concrete description of r given in Subsection 10.1.6 we deduce
Z
k
i<j
i,j,k
Lemma 10.2.22 V is a Clifford connection on §(M) so that (§(M), V) is a Dirac bundle called the bundle of pure spinors. Proof Use Lemma 10.1.22. D
461
Fundamental examples
We now want to understand the structure of the Weitzenbock remainder of the geometric Dirac operator D associated to the bundle of pure spinors. If
R = k<e Y.e" A e ^ « Ku = (RU) = (^«) is the Riemann curvature tensor (in the above trivializations over Ua) we deduce that the curvature of V is R = Y/ekAel®T-1{Rke) k
4
k<e tj
From this we obtain
° ijkl
In the above sum the terms corresponding to indices (i,j,k,£) such that i = j or A; = £ vanish due to the corresponding skew-symmetry of the Riemann tensor. Thus we can write
c(^(V)) =
-WY.^W^WWW)
Using the equalities c(eI)c(e-') + c(e-J)c(eI) = —2<5y we deduce that the monomial c(e')c(e J ) anti-commutes with c(e*)c(e^) if the two sets{i, j} and {k, £} have an unique element in common. Such pairs of monomials will have no contributions in the above sum due to the curvature symmetry Rijkl
=
Rklij-
Thus we can split the above sum into two parts c(F(V)) = - I £
flwC(e')c(e'>(ei)c(e')
i,j
+
£
Rijklc{ei)c{ej)c{ek)c{ei).
i,j,fc,£distinct
Using the first Bianchi identity we deduce that the second sum vanishes. The first sum is equal to
where s denotes the scalar curvature of M. We have thus proved the following result. Theorem 10.2.23 (Lichnerowicz)([48])
4
462
Dirac operators
A section ip G C°°(§(M)) such that T>2%p = 0 is called a harmonic spinor. Lichnerowicz theorem shows that a compact spin manifold with positive scalar curvature admits no harmonic spinors. Exercise 10.2.10 Consider an oriented 4-dimensional Riemann spin manifold (M, g) and W —» M a Hermitian vector bundle equipped with a Hermitian connection V. Form the twisted Dirac operator •Dw : C°° ((§ + (M) 0 S"(M)) ® W) ->• C°° ((S + (M) 0 §~(M)) ® w) defined in Exercise 10.1.18. T)w is Z 2 -graded and hence it has a block decomposition
where DWt+ : C°°(§ + (M) ® W) -> C°°(§-(M)
+
^,
+
= V*V + ^ + c(F + (V))
where V denotes the product connection of §(M) ® W while F + (V) = ^ ( F ( V ) + *F(V)) denotes the self-dual part of the curvature of the bundle (W, V). §10.2.4
The spin0 Dirac operator
Our last example of Dirac operator generalizes both the spin Dirac operator and the Dolbeault operator. The common ingredient behind both these examples is the notion of spin0 structure. We begin by introducing it to the reader. Let (Mn,g) be an oriented, n-dimensional Riemann manifold. As in the previous section we can regard the tangent bundle as associated to the principal bundle PSO(M) of oriented orthonormal frames. Assume PSO(M) is defined by a good open cover i t = (Ua) and transition maps ffa/3 : Ua0 -> SO(n). The manifold M is said to posses a spin0 structure if there exists a principal Spin°(n) -bundle Pspi-n? such that PSO(M) is associated to Pspin' via the natural morphism p° : Spin°{n) -> SO{n): PsO(M)
= Pspin"
Xpc
SO{n).
Equivalently, this means there exist smooth maps gap • Uap —> Spin°{n), the cocycle condition, such that pc{gap) = gap-
satisfying
Fundamental
463
examples
As for spin structures, there are obstructions to spin0 structures as well which clearly are less restrictive. Let us try to understand what can go wrong. We stick to the assumption that all the overlaps C/a/j...T are contractible. Since Spin°(n) = (Spin(n) x S1)/Z2, lifting the SO{n) structure {gap) reduces to finding smooth maps hap • Uap -> Spin{n) and Zap '• Uap —> S such that p{Kn) = 9aff and {eaPyXafiy)
= ' {Kphfrh^
, Za0Z0JZla)
€ { ( - 1 , - 1 ) , (1, 1)}.
(10.2.4)
If we set \a0 = z\0 : Ua@ -> S 1 we deduce from (10.2.4) that the collection (\al3) should satisfy the cocycle condition. In particular, it defines a principal S1 bundle over M, or equivalently, a complex line bundle C. This line bundle should be considered as part of the data defining a spin0 structure. The collection (£Qja7) is an old acquaintance: it is a Cech 2-cocycle representing the 2nd Stiefel-Whitney class. As in Subsection 8.2.2 we can represent the cocycle Xa/3 as Xa/) = exp(i0Q/3). The collection "a/37 = -^—{6a0 + @t3-t + #70)
defines a 2-cocycle of the constant sheaf Z representing the the topological 1st Chern class of C. The condition (10.2.4) shows that na01 = eai3-f (mod 2). To summarize, we see that the existence of a spin0 structure implies the existence of a complex line bundle C such that c\op(C) = w 2 (M) (mod 2). It is not difficult to prove the above condition is also sufficient. In fact one can be more precise. Denote by Spin°(M) the collection of isomorphism classes of spin0 structures on the manifold M. Any a G Spin°(M) is defined by a a lift (hap,za/3) as above. We denote by Ca the complex line bundle defined by the gluing data (zap). We have seen that c f O C ) = iua(M) (mod 2). Denote by CM C # 2 ( M , Z ) the "affine " subspace consisting of those cohomology classes satisfying the above congruence modulo 2. We thus have a map Spin°(M)
-> £ M ,
a ^
cf{Ca).
464
Dirac operators
Proposition 10.2.24 The above map is a surjection. Exercise 10.2.11 Complete the proof of the above proposition. The smooth Picard group Pic°°(M) acts on Spin°(M) Spinc(M) More precisely if a e Spinc{M)
by
x Pic°°(M) 3 (a, L) H-J- a ® L. is given by the cocycle
a = [hap, za0] : Ua/3 -> Spin (n) x S 1 / ~ and L is given by the S 1 cocycle Ca/3 : Uap —> 5 then a ® L is given by the cocycle
Note that C-a®L = Ca® I? so that ^ " ( a ®L)
= C\OP{CJ) + 2c[op(L).
Proposition 10.2.25 The above action of Pic°°(M) on Spinc(M) sitive.
is is free and tran-
Proof Consider two spin0 structures a1 and a2 defined by the good cover (Ua) and the gluing coycles [haP'Zaph c
c
Since p {hal) = p (hal) by a sign) that
* = 1,2.
= ga/3 we can assume (eventually modifying the maps hap h(l) - h{2) n a/3 — "a/3
This implies that r
0
_ r ( 2 )/ 7 ( 1 )
^a0 —
Z
a0/Za0
is an S'1-cocycle defining a complex line bundle L. Obviously a2 = a1 ® L. This shows the action of Pic°°(M) is transitive. We leave the reader verify this action is indeed free. The proposition is proved. • Given two spin0 structures 0\ and a2 we can define their "difference" o^/ci as the unique line bundle L such that a2 = ux ® L. This shows that the collection of spin0 structures is (non-canonically) isomorphic with H2(X, Z) = Pic°°. It is a sort
Fundamental
465
examples
of affme space modelled on H2(X, Z) in the sense that the "difference " between two spin0 structures is an element in H2(X, Z) but there is no distinguished origin of this space. A structure as above is usually called a H2(M, Z)-torsor. Without a sufficient background in algebraic topology the above results may look of very little help in detecting spin0 structures. This is not the case and to convince the reader we will list below (without proofs) some examples of spin0 manifolds. Example 10.2.26 (a) Any spin manifold admits a spin0 structure. (b) Any almost complex manifold has a natural spin0 structure. (c) (Hirzebruch-Hopf, [37]; see also [61]) Any oriented manifold of dimension < 4 admits a spin0 structure. Let us analyze the first two example above. If M is a spin manifold then the lift gap • Uap -> Spin(n) of the SO structure to a spin structure canonically defines a spin0 structure via the trivial morphism Spin(n) —> Spin°(n) Xj, S1,
J H > (g, 1) mod the Z 2 — action.
We see that in this case the associated complex line bundle is the trivial bundle. This is called the canonical spin0 structure of a spin manifold. Thus on a spin manifold the torsor of spm c -structures does in fact possess a "canonical origin" so in this case there is a canonical identification Spin°(M) =* Pic°° £ H2(M, Z). To any complex line bundle L defined by the 51-cocycle (zap) we can associate the spin0 structure defined by the gluing data {(ga0,zc,p)}Clearly, the line bundle associated to this structure is L2 = L®2. In particular this shows that on a spin manifold M for any a £ Spin°(M) there exists a square root
OJ2oiLa.
To understand why an almost complex manifold (necessarily of even dimension n = 2k) admits a canonical spin0 structure it suffices to recall the natural morphism U(k) —»• SO(2k) factors through a morphism f : U{k) ->• Spinc(2k). The t/(/s)-structure of TM, defined by the gluing data Kp • Ua0 -> U(k)
466
Dirac operators
induces a spinc structure defined by the gluing data £(hap). Its associated line bundle is given by the S^cocycle detc(hQ/3) : Ua0 -> S 1 and it is precisely the determinant line bundle detcTlfiM
= A*'°TM.
10
The dual of this line bundle, det£(T*M) ' = Ak'°T*M plays a special role in algebraic geometry. It usually denoted by 3C and it is called the canonical line bundle. Thus the line bundle associated to this spin0 structure is 3C_1 = %*'. From the considerations in Subsection 10.1.5 and 10.1.7 we see that many (complex) vector bundles associated to the principal Spin0 bundle of a spin0 manifold carry natural Clifford structures and in particular one can speak of Dirac operators. We want to discuss in some detail a very important special case. Assume that (M, g) is an oriented, n-dimensional Riemann manifold. Fix a € Spinc(M) (assuming there exist spin c structures). Denote by (gap) a collection of gluing data defining the SO structure Pso(M) on M with respect to some good open cover (I/ Q ). Moreover, we assume a is defined by the data hap •• Ua0 ->
Spinc(n).
Denote by A£ the fundamental complex spinorial representation (defined in §10.1.6). A„ : Spin°(n) -» Aut (§„). We obtain a complex bundle §„(M) = Pspin- X An S n which has a natural Clifford structure. This is called the bundle of complex spinors associated to a. Example 10.2.27 (a) Assume M is a spin manifold. We denote by aQ the spin0 structure corresponding to the fixed spin structure. The corresponding bundle of spinors § 0 (M) coincides with the bundle of pure spinors defined in the previous section. Moreover for any complex line bundle L we have S L d=f Sff ^ § 0 ® L. where a = <70 ® L. Note that in this case L2 = L„ so one can write §ff ^ So ® C]/2. (b) Assume M is an almost complex manifold. The bundle of complex spinors associated to the canonical spin0 structure a (such that Ca = 3C_1) is denoted by S^(M). Note that S C ( M ) =* A°'*T*M. D
Fundamental
467
examples
We will construct a natural family of Dirac structures on the bundle of complex spinors associated to a spirf structure. Consider for warm-up the special case when TM is trivial. Then we can assume ga@ = 1 and ha0 — (l>2a/?) : Uap -> Spin{n) x S 1 —> Spinc{n). The S 1 cocycle (z^n) defines the line bundle Ca. It this case something more happens. The collection (zap) is also an S1-cocycle defining a complex Hermitian line bundle L such that L2 = Ca. Traditionally, L is denoted by CXJ2 though the square root may not be uniquely defined. We can now regard §„(M) as a bundle associated to the trivial Spm(n)-bundle Pspin and as such there exists an isomorphism of complex Spin(n) vector bundles Sff(M)^S(M)®4/2. As in the exercise 10.1.18 of Subsection 10.1.9 we deduce that twisting the canonical connection on the bundle of pure spinors So(M) with any Hermitian connection on C]J2 we obtain a Clifford connection on §a(M). Notice that if the collection
{Lja€u{l)®Ql{Ua)} defines a connection on Ca i.e. dZ
°P
,
TT
up = —^^a/J + u)a over Uap then the collection W
a — X^o
defines a Hermitian connection on L = ClJ2. Moreover if F denotes the curvature of (u>.) then the curvature of (Q.) is given by F = \F.
(10.2.5)
Hence any connection on Ca defines in an unique way a Clifford connection on § ff (M). Assume now that TM is not necessarily trivial. We can however cover M by open sets (Ua) such that each TUa is trivial. If we pick from the start a connection on Ca this induces a Clifford connection on each §a(Ua). These can be glued back to a Clifford connection on §CT(A^) using partitions of unity. We let the reader check the connection obtained in this way is independent of the various choices. Example 10.2.28 Assume (M, g) is both complex and spin. Then a choice of a spin structure canonically selects a square root X - 1 ^ 2 of the line bundle X - 1 because
468
Dirac operators
% 1 is the line bundle associated to the spin0 structure determined by the complex structure on M. Then S c = So ® 3C"1/2Any Hermitian connection on % induces a connection of § £ . If M happens to be Kahler then the Levi-Civita connection induces a complex Hermitian connection % and thus a Clifford connection on SC{M) = A°-*T*M. Let u be a connection on C„. Denote by V" the Clifford connection it induces on § cr (M) and by D u the associated geometric Dirac operator. Since the Weitzenbock remainder of this Dirac operator is a local object so to determine its form we may as well assume §CT = § ® C]/2. Using the computation of Exercise 10.1.18 and the form of the Weitzenbock remainder for the spin operator we deduce D l = (V?V" + iS + ^ ( F (
w
))
where F denotes the curvature of the connection u> on Ca. Since F(u>) has the form F = ixtt,
(]£(12(¥)
we deduce cc„(F) = iq(^) € Cl{T*M) ® C. Hence ^
= (VTV<" + ^ + ^ q ( f i ) .
(10.2.6)
c
Exercise 10.2.12 Consider the canonical spin structure a° on a compact Kahler manifold (M,g, J). The Levi-Civita connection induces a Clifford connection on X~ and thus a connection on the associated bundle of spinors § C ( M ) = A°'*T*M. (a) Show that the associated spin0 Dirac operator coincides with the Dolbeault operator. (b) Assume L —> M is a Hermitian line bundle over M equipped with a Hermitian connection and denote by 1)L the corresponding spin0 Dirac operator on SaC&L =* S c ® L ^ A°'*T*M ® L. Prove that D L C o o (A 0 'T*M) C C*°°(A*'?T*M) where OL is the generalized Laplacian T>\. (c)(Kodaira) Assume L is a positive line bundle i.e. h(X,Y)
:= Cl{V){X,JY)
= ^-FL{V)(X,JY)
VX,Y
0,q
G Vect (M).
defines a Riemann metric on M. Denote by H (L) the kernel of the restriction of • i to A°tQT*M. Show that there exists n0 > 0 such that H°'q(Lm)
= 0 Vn > n 0) Vg > 1.
Bibliography [1] Mathematical Developments Arising From Hilbert Problems, Proceedings of Sym. in Pure Math., vol . 28, Amer. Math. Soc. Providence RI, 1976. [2] J.F. Adams: Vector fields on spheres, Ann.of Math 75 (1960), p.603-632. [3] R.A. Adams: Sobolev Spaces, Academic Press, New York, 1975. [4] V.I. Arnold: Ordinary Differential Equations, Springer Texts, SpringerVerlag, 1992. [5] V.I. Arnold: Mathematical Methods Of Classical Mechanics, Springer Texts, Springer Verlag, 1989. [6} M.F. Atiyah, R. Bott, A. Shapiro: Clifford modules, Topology 3, Suppl.l (1964), p. 3-38. [7] T. Aubin: Problemes isoperimetriques Geom. 11 (1976), p. 533-598.
et espaces de Sobolev, J. Diff.
[8] T. Aubin: Equations differentielles non lineaires et probleme de Yamabe concernant la courbure scalaire, J. Math. Pures et Appl., 55 (1976), p. 269-296. [9] P. Berard: From vanishing theorems to estimating theorems: the Bochner technique revisited, Bull. A.M.S. 19(1988), p.371-406. [10] N.Berline, E.Getzler, M. Vergne: Heat Kernels And Dirac Springer Verlag, 1992.
Aperators,
[11] L. Bers, F. John, M. Schechter: Partial Differential Equations, Lectures in Applied Mathematics, vol. 3A, Amer. Math. Soc, Providence RI, 1964. [12] A. Besse: Einstein Manifolds, Ergebnisse der Mathematik, Bd.lO, Springer Verlag, 1987. [13] B. Boos, K. Wojciechowski: Elliptic Boundary Problems For Dirac Operators, Birkhauser, 1993. 469
470
Dirac operators [14] R.Bott, L.Tu: Differential Forms In Algebraic Topology, Springer-Verlag, 1982. [15] H. Brezis: Analyse Fonctionelle. Theorie Et Applications, Masson, 1983. [16] T. Brocker, K. Janich -.Introduction To Differential Topology, Cambridge University Press, 1982. [17] E. Cartan: Sur les invariants integraux des certains espaces homogenes et les proprietes topologiques de ces espaces, Ann. Soc. Pol. Math, 8(1929), p.181-225. [18] H. Cartan: La transgression dans un group de Lie et dans un espace fibre principal, p.1268-1282, vol. Ill of H. Cartan's " Collected Works", Springer-Verlag, 1979. [19] S.S. Chern: A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemann manifolds, Ann. Math, 45(1946), p.747-752. [20] B. Dacorogna: Direct Methods In The Calculus Of Variations, Applied Mathematical Sciences, vol. 78, Springer-Verlag, 1989. [21] B. A. Dubrovin, A.T. Fomenko, S.P Novikov: Modern Geometry - Methods And Applications, Vol.1-3, Springer Verlag, 1984, 1985, 1990. [22] J. Dieudonne: Foundations Of Modern Analysis, Academic Press, 1963. [23] M. P DoCarmo: Riemannian Geometry, Birkhiiuser, 1992. [24] C.Ehresmann: Sur la topologie de certains espaces homogenes, Ann. of Math., 35(1934), p.396-443. [25] S. Eilenberg, N. Steenrod: Foundations Of Algebraic Topology, Princeton University Press, 1952. [26] W.Fulton, J.Harris: Representation
Theory, Springer-Verlag 1991.
[27] R.P Feynman, R.B.Leighton, M. Sands: Physics, Addison Wesley, 1964.
The Feynamn Lectures
On
[28] D. Gilbarg, N.S. Trudinger: Elliptic Partial Differential Equations Of Second Order, Springer Verlag, 1983. [29] R. Godement: Topologie Algebrique Et Theorie Des Faisceaux, Hermann et Cie, Paris, 1964. [30] C. Gordon. D. L. Webb, S. Wolpert: One cannot hear the shape of a drum, Bull. A.M.S., 27(1992), p.134-138.
Fundamental
examples
471
[31] I.S.Gradshteyn, I.M.Ryzhik: Tables Of Integrals, Series And Academic Press, 1994. [32] P. Griffiths, J. Harris: ley&Sons, 1978.
Principles Of Algebraic Geometry,
Products,
John Wi-
[33] G.H. Hardy, E.M Wright: An Introduction To The Theory Of Numbers, Oxford University Press, 1962. [34] S. Helgason: Groups And Geometric Analysis, Academic Press, 1984. [35] M.Hirsch: Differential Topology, Springer-Verlag, 1976. [36] F. Hirzebruch: Topological Methods In Algebraic Geometry, Springer Verlag, New York, 1966. [37] F. Hirzebruch, H. Hopf: Felder von Flachenelementen in 4-dimensionalen Manigfaltigkeiten Math. Ann. 136(1958). [38] H. Iwamoto: On integral invariants and Betti numbers of symmetric Riemannian manifolds, I, J. of the Math. Soc. of Japan, 1(1949), p.91-110. [39] M. Kac: Can one hear the shape of a drum?, Amer. Math. Monthly, 73(1966) Part II, p.1-23. [40] M. Karoubi: K-Theory, Springer-Verlag, 1978. [41] T. Kato: Perturbation Theory For Linear Operators, Springer Verlag, 1984. [42] J. Kazdan, F. Warner: Curvature functions for compact 2-manifolds, Ann. Math. 99(1974), p.14-47. [43] J.L. Kelley: General Topology, Van Nostrand, 1955. [44] S. Kobayashi, K. Nomizu: Foundations Of Differential Geometry, Interscience Publishers, New York, 1963. [45] K. Kodaira, D.C. Spencer: Groups of complex line bundles over compact Kahler varieties. Divisor class group on algebraic varieties, Proc. Nat. Acad. Sci. USA 39(1953), p. 868-877. [46] H. B. Lawson, M.-L. Michelson: Spin Geometry, Princeton University Press, 1989. [47] J.M. Lee, T.H. Parker: The Yamabe problem, Bull. A.M.S. 17 (1987), p. 37-91.
472
Dirac operators [48] A. Lichnerowicz: Laplacien sur une variete riemannienne et spineure, Atti Accad. Naz. dei Lincei, Rendicotti 33(1962), 187-191. [49] D.E. Littlewood: T i e Theory Of Group Characters And Matrix Representations Of Groups, Clarendon Press, Oxford, 1940. [50] I.G.Macdonald: Symmetric Functions And Hall Polynomials, Clarendon Press, Oxford, 1979. [51] S. MacLane: Categories For The Working Mathematician, Springer Verlag, 1971. [52] W.S.Massey: A Basic Course In Algebraic Topology, Springer Verlag, 1991. [53] W.S.Massey: Cross products of vectors in higher dimensional spaces, Amer. Math. Monthly, 93(1983), p.697-700.
euclidian
[54] J. Milnor: On manifolds homeomorphic to the 7-sphere, Ann. of Math. 64 (1956), p.399-405. [55] J. Milnor: Morse Theory, Princeton University Press, 1963. [56] J.W. Milnor: Spin structures 9(1963), 198-203.
on manifolds,
L'Enseignement Math.,
[57] J. Milnor: Curvatures of left invariant metrics on Lie groups, Adv. in Math., 21(1976), p.293-329. [58] J. Milnor, J.D. StashefF: Characteristic Classes, Ann. Math. Studies 74, Princeton University Press, Princeton, 1974. [59] R. Miron, M. Anastasiei: The Geometry Of Lagrange Spaces: Theory And Applications, Kluwer Academic Publishers, 1994. [60] J.W. Morgan: The Seiberg-Witten Equations And Applications To The Topology Of Smooth Manifolds, Mathematical Notes, Princeton University Press, 1996. [61] C. Morrey: Multiple Integrals In The Calculus Of Variations, Springer Verlag, 1966. [62] A. Newlander, L. Nirenberg: Complex analytic coordinates on almost complex manifolds, Ann. Math., 54(1954), p.391-404. [63] L. Nirenberg: On elliptic partial differential equations, Ann. Scuola Norm. Sup. Pisa, 13(1959), p.1-48.
Fundamental
473
examples
[64 I. R. Porteous: Topological Geometry, Cambridge University Press, 1981. M. M. Postnikov: Lectures In Geometry. Semester V: Lie Groups an Lie [65; Algebras, Mir Publishers, Moscow, 1986. R. Schoen: Conformal deformations of a Riemannian metric to constant [66; curvature, J. Diff. Geom. 20 (1984), p. 479-495. [67 J.A.Schouten, D.J.Struik: On some properties of general manifolds relating to Einstein's theory of gravitation, Amer. J. Math 43(1921), p. 213-216. J.P. Serre: Faisceaux algebriques coherences, Ann Math. 61(1955), p.197278. [69; I.R. Shafarevitch: Algebra I. Basic Notions, Encyclopedia of Mathematical Sciences, vol. 11, Springer Verlag, 1990. [70; M. Spivak: A Comprehensive Introduction To Differential Geometry, vol. 1-5, Publish or Perish, 1979. [71 N.E. Steenrod: Press, 1951.
The Topology Of Fibre bundles, Princeton University
[72; E.M. Stein: Singular Integrals And Differentiability Properties Of Functions, Princeton University Press, 1970. [73; F.W. Warner: Foundations Of Differentiable Manifolds And Lie Groups, Springer Verlag, 1983. [74; H.Weyl: The Classical Groups, Princeton University Press, 1946. [75; H.Weyl: Nachr. Akad. Wiss Gottingen, 1914, p. 235-236. [76; H.Weyl: Math. Ann., 77 (1916), p.313-315. [77; H.Weyl: Die Idee Der Riemannschen Flache B.G. Tuebner, Berlin 1913. [78; E. Witten: Monopoles and four-manifolds, Math.Res. Letters 1 (1994), p.769-796. [79] K. Yosida: Functional Analysis, Springer Verlag, 1974.
Index C°°(M), 7 G-module, 105 Lk'", 351 A M , 142 A*y, 35 hP'iV, 51 detind(L), 402 exp ? , 126 A-genus, 319
atlas, 6 maximal, 7 berezinian, 46 Bernoulli numbers, 315 Betti number, 212 Bianchi identity, 88, 145 bootstrap, 389 bundle, 21 canonical line -, 466 Clifford, 413 complex spinor, 460 cotangent, 53 Dirac, 441 fiber, 56 frame, 59 G-fiber, 57 Hopf, 60 line, 26 tautological, 27 principal, 59, 293 pullback of a, 303 tangent, 22 vector, 24, 26 rank of a, 24 ample, 27 automorphism, 26 base of a, 24 endomorphism, 26 map, 26 morphism, 26 pullback, 27 section, 25 tautological, 27 total space of a, 24
/M,93 /B/SI H3 d-bundle, 114 TTJCA-.IO), 193 PM{Q),
126
( T ( M ) , 238
{•,•}, 412 h{M), 212 d", 138
7,46 d i v Z , 137 g r a d / , 137 X, 466 * [ [ * ] ] \ 314 ind (7), 185 tr s , 42 V(V), 32 gl(n), 15, 69 o(n), 15, 69 sl(n), 15, 69 su(n), 15, 69 M(n), 15, 69 v, 140 L-genus, 319 antipodal points, 11 474
Fundamental
475
examples
trivial, 28 Cartan algebra, 305, 308, 310 Cartan's Lemma, 155 categories, 196 character, 108 characteristic class, 304 2nd Stiefel-Whitney, 458 universal, 315, 319 charts, 6 Chern classes, 306, 315 universal, 306 Chern polynomial, 306 universal, 306 Chern-Weil correspondence, 303 Christoffel symbols, 89 circle, 15 class function, 108 Clifford algebra, 414 connection, 441 module, 417 multiplication, 413, 414 s-module, 417 structure, 413 coboundary, 213 cochain homotopy, 215 cocycle, 213 cohomologous, 213 cocycle condition, 24 cohomology Cech, 287 Cech , 284 DeRham, 212 compact supports, 230 complex chain, 213 cochain, 213 acyclic, 213 DeRham, 214 Dolbeault, 453 Koszul, 454
complex structure, 51 configuration space, 3 connection, 76, 294 flat, 84 Levi-Civita, 121 symmetric, 89 connection 1-form, 78 conormal, 102 inner, 102 outer, 102 contraction, 69 convolution, 353 coordinates, 1 Cartesian system, 2 local, 6 polar, 2 coordinatization, 2, 3 covariant derivative see connection 76 covering space, 198 sheets of a, 198 universal, 203 critical point, 186 critical value, 186 curvature, 83 Gauss, 149 Ricci, 146 Riemann, 144 scalar, 147 sectional, 147 cycle, 239 degenerate, 239 transversal, 243 cycles cobordant, 239 deck transformation, 200 density bundle, 91 derivative, 4 Frechet, 4 det V, 35 diffeomorphism, 7 differential form, 54
476 closed, 207 exact, 207 Dolbeault operator see complex 454 duality, 42 Hodge, 48 metric, 43 natural, 43 symplectic, 43 Dynkin polynomials, 69 Einstein manifold, 161 Einstein's equation, 161 elliptic complex, 407 equation Euler-Lagrange, 170 Euler characteristic, 38, 165 Euler class, 253, 313 universal, 313 exponential map see manifold 126 exterior, 35 algebra, 35 derivative, 70 product, 35 extremal, 170 five lemma, 227 flow, 61 infinitesimal generator of a, 62 local, 61, 63 Frechet derivative, 4 Fredholm index, 395 function, 3 Ck, 4 Frechet differentiable, 3 harmonic, 142 smooth, 4, 6 functors, 196 contravariant, 197 covariant, 197 gauge transformation, 26 Gauss map, 158 genus, 165
Dirac operators geodesic, 122 index of a, 185 nondegenerate, 184 geodesic flow, 122 global angular form, 327 gluing condition, 25 grassmannian, 12, 27, 267 complex, 12 real, 12 Green formula, 349 group, 13, 28 action, 57 effective, 57 free, 57 orbit, 57 Clifford, 425 complex Clifford, 433 connected, 15 fundamental, 193 left translation in a, 14 Lie, 13, 28, 67, 105, 123, 147, 265, 293, 300 exponential map of a, 68 representation, 57 orthogonal, 13 right translation in a, 28 special orthogonal, 14 special unitary, 14 unitary, 14 Gysin map, 251 hessian, 181 Hilbert-Einstein functional, 161 Hodge *-operator, see operator holonomy, 87 homogeneous space, 258 homotopy, 219 immersion, 23 injectivity radius, 126 integral curve, 62 interior derivative, 69 intersection form, 238
Fundamental
examples
intersection number, 243 invariant integration, 100 invariant polynomials, 299 isometry, 117 isotropy, 258 Jacobi equation, 183 field, 183 Jacobi identity, 41 Kiinneth formula, 227 Killing pairing, 124 Killing pairing, 214, 266 Kronecker pairing, 234 lagrangian, 169 invariant, 177 Laplacian, 142, 344 covariant, 348 Lefschetz number, 250 Lie bracket, 64 derivative, 63 group, see group Lie algebra, 41, 66, 67 simple, 266 semisimple, 124, 214, 266 Mobius strip, 95 manifold with boundary, 101 almost complex, 451 complex, 12 connect sum, 16 connected sum, 10 direct product, 10 finite type, 224 Kahler, 455 orientable, 16, 95 orientation of a, 96 oriented, 95, 96 Riemann, 117 convex sets in a, 132
exponential map of a, 126 geodesically complete, 122 smooth, 6 spin, 457 manifolds, 3, 9 map, see function gluing, 12 cochain, 215 covering, 198 degree of a, 249 homotopic, 192 lift of a, 200 quantization, 430 transition, 6, 24 cohomologous, 26 matrix, 4 invertible, 4 Jacobian, 4 orthogonal, 13 skew-Hermitian, 14 skew-symmetric, 14 unitary, 14 maximum principle, 383 Mayer-Vietoris cover, 225 functors, 225 correspondence of, 225 long sequence, 221, 233 principle, 226 short sequence, 220, 233 metric, 49, 53 Hermitian, 49 conformal, 390 Kahler, 455 Riemann, 54 metric volume, 135 mollifier, 353 monodromy, 202 moving frame, 78 nerve, 284 Newton formulae, 318
478 normal coordinates, 127 operator, 3 Hodge *, 48 boundary, 213 chirality, 420 closed, 392 Fredholm, 395 selfadjoint, 393 semi-Fredholm, 394 symmetric, 393 coboundary, 213 Dirac, 411 Fredholm, 395 geometric Dirac, 443 Hodge-DeRham, 412 linear, 3 transpose of a, 32 partial differential, see p.d.o. orientation canonical, 50 orientation cover, 199 orientationsee vector space 45 outer normal, 140 p.d.o., 335, 336 elliptic, 340 formal adjoint, 344 formally selfadjoint, 345 symbol of a , 340 pairing, 42 Hodge, 48 parallel transport, 80 partition, 274 conjugate of a, 274 length of a, 274 weight of a, 274 partition of unity, 8 pfaffian, 46 Pin(V), 428 Poincare duality, 235, 243 inequality, 387, 395
Dirac operators lemma, 207 polynomial, 212 series, 37 Pontryagin classes universal, 309 Pontryagin polynomial universal, 309 presheaf, 279 projection, 9 stereographic, 9 projective plane, see space pullback, 55 quantization map, 416 r-density, 49 regular value, 186 Riemann metric, 117 Schur lemma, 107 polynomial, 275 sheaf, 280 fine, 289 resolution, 283 fine, 290 smooth, see function structure, 6 space Banach, 3 complex projective, 12 locally compact, 6 real projective, 10 simply connected, 195 Sobolev, 351 tangent, 18, 21 space of germs, 280 morphism, 282 section of, 280 stalk of, 280 sphere, 9 sphere bundle, 60 Spin(V), 428
Fundamental
479
examples
spinor, 418, 423 split coordinates, 112 Stokes formula, 102 structural equations, 151 submanifold, 8 2nd fundamental form, 155 codimension of a, 8 submersion, 23 super, 39 algebra, 39 commutative, 39 commutator, 39 derivation, 40 Lie algebra, 40 space, 39 tensor product, 41 supertrace, 42 surface, 15 doughnut-shaped, 15 symbol map, 416 symmetric space, 259 symmetry, 13 tensor, 30 product, 30 algebra, 32 contravariant, 32 covariant, 32 field, 54 left invariant, 67 right invariant, 67 product universality property of, 30 skew-symmetric, 34 exterior product of, 34 symmetric, 34 type (r,s), 32 theorem, 3 Banach fixed point, 5 Bochner vanishing, 449 Bochner-Weitzenbock, 444 Calderon-Zygmund, 372
Cartan, 265 Cartan-Hadamard, 198 Chern-Weil, 300 Clairaut, 179 Fubini, 113 Gauss-Bonnet, 160, 223, 250 Gauss-Bonnet-Chern, 326 Hodge, 407, 409 Hopf-Rinow, 133 implicit function, 5 inverse function, 4 Leray-Hirsch, 230 Mayer-Vietoris, 221 Morrey, 363 Myers, 188 Noether, 177 Poincare-Hopf, 258 Rellich-Kondratchov compactness, 360 Sobolev embedding, 358 uniformization, 391 Weyl, 205 Theorema egregium, 156 Thorn class, 252 Todd class universal, 316 torsion, 89 torsor, 465 torus, 9, 13 transform, 13 Cay ley, 13 Fourier, 372 trivializing cover, 23 unit outer normal, 140 variation, 171, 180 geodesic, 183 infinitesimal, 179 vector, 19 field, 25 Killing, 139 space, 20, 30 Z-graded, 37
480 determinant line of a, 35, 45 dual, 31 orientation of a, 45 oriented, 45 volume form in a, 45 tangent , 19 volume form, 96 weak derivative, 350 Weitzenbock remainder, 349, 447, 461 Weyl group, 270, 306, 309 integral formula, 268 lemma, 380 unitary trick, 106 Whitney sum, 53 Young diagram, 274
Dirac operators