<<

Contents

1 Hermitian symmetric spaces: examples and basic properties1 1.1 Almost complex manifolds...... 1 1.2 Hermitian manifolds...... 2 1.3 Symmetric spaces...... 3 1.4 Bergman metric...... 4 1.5 A higher-dimensional example: Siegel upper half-space...... 6 1.6 of Hermitian symmetric domains...... 7

2 Classification of symmetric Hermitian domains9 2.1 Preliminaries: Cartan involutions...... 9

1 Hermitian symmetric spaces: examples and basic properties

The next goal is to explain classification of Hermitian symmetric domains. We begin with some preliminary expla- nations of what these are, along with toy examples.

1.1 Almost complex manifolds

Definition 1.1. Let M be a smooth real manifold (of class C ∞). We say that it is an almost complex manifold if we are given a smooth tensor field J on M of type (1,1) such that for each vector field X on M holds J(J(X )) X . = − In other words, we ask that to each point p M we associate (in a smooth manner) an endomorphism of the real ∈ J : T M T M such that J 2 Id. p p → p p = − n Example 1.2. Let M be a complex manifold of dimension n, i.e. a ringed space (M,OM ) locally isomorphic to C with the sheaf of holomorphic functions. Then M may be viewed as a smooth real manifold of dimension 2n. The tangent spaces T M carry naturally an action of i C, and M is an almost complex manifold. N p ∈ Example 1.3. Let V be a real vector space. To specify an action of C on V means to pick an R-algebra representation 2 ρ : R[J]/(J 1) EndR(V ). + → Since the algebra R[J]/(J 2 1) is generated by J, everything amounts to picking an endomorphism J : V V satis- + → fying J 2 Id. Then this gives an action of the whole C by = − (x J y) v x v y J(v). + · = + Note that V becomes a C-vector space, so we conclude that dimR V must be an even number, otherwise V does not admit an almost complex structure. Vice versa, if dimR V 2n, then we can pick a basis (x , y ,...,x , y ) and define an action of C by thinking of = 1 1 n n these as of coordinates z x i y : k = k + k

J(x ) : y , k = k J(y ) : x . k = − k However, there is nothing canonical about this! N

Example 1.4. We see that a real manifold M does not admit an almost complex structure if dimR M is odd. If dimR M is even, then each tangent space Tp M carries an almost complex structure Jp , but they might not glue to a smooth vector field J on M such that J 2 Id. For instance, among the spheres S2n, only S2 and S6 admit = − an almost complex structure (S2 simply because it may be viewed as the Riemann sphere; for S6 it is actually not known whether it has a complex structure, it’s a famous open problem). See e.g. N. Steenrod, The Topology of Fibre Bundles, §41, and J. P.May, A Concise Course in Algebraic Topology, §24.4. N

1 1.2 Hermitian manifolds Definition 1.5. Let M be a smooth manifold. We say that it is Hermitian if

• M carries an almost complex structure J.

• M carries a Riemannian structure, i.e. a smooth tensor field g of type (2,0), such that

– g(X ,Y ) g(Y , X ) for all vector fields X ,Y on M, = – g : T M T M R is a positive definite nondegenerate bilinear form for each p M. p p × p → ∈ • The Riemannian structure is invariant under J, i.e. for all vector fields X ,Y on M holds

g(JX , JY ) g(X ,Y ). =

Example 1.6. Consider the complex plane C with real coordinates Z X i Y . The standard Riemannian metric = + is given by µ ∂ ∂ ¶ µ ∂ ∂ ¶ µ ∂ ∂ ¶ gp (p), (p) 0, gp (p), (p) 1, gp (p), (p) 1. ∂X ∂Y = ∂X ∂X = ∂Y ∂Y = We recall that usually this is written as “ds2 dx2 d y2”. The action by i is ∂ ∂ , ∂ ∂ , and so clearly = + ∂X 7→ ∂Y ∂Y 7→ − ∂X g(i X ,i Y ) g(X ,Y ), and C is Hermitian. N · · = Example 1.7. Consider the complex upper half-plane

H : {z C Imz 0}. = ∈ | > It carries naturally a complex structure, and real coordinates Z X i Y . We consider a Riemannian metric given 2 (d X )2 (dY )2 = + by ds + , i.e. = y2 µ ∂ ∂ ¶ µ ∂ ∂ ¶ 1 µ ∂ ∂ ¶ 1 gp (p), (p) 0, gp (p), (p) , gp (p), (p) , ∂X ∂Y = ∂X ∂X = y2 ∂Y ∂Y = y2

and we see immediately that g(i X ,i Y ) g(X ,Y ), so H is a Hermitian manifold. N · · = Example 1.8. Consider P1(C) C { } as the Riemann sphere S2 , R3, which comes with a natural complex and = ∪ ∞ → Riemannian structure. z

x 0 C y

Geometrically, the action of i rotates C counterclockwise, and this satisfies g(i X ,i Y ) g(X ,Y ). N · · =

2 1.3 Symmetric spaces

Definition 1.9. Let M be a connected complex manifold. Then its underlying smooth manifold M ∞ has a canon- ical almost complex structure. Assume that M ∞ also carries a Riemannian structure which is compatible with the almost complex structure, in the sense that M ∞ is a Hermitian manifold. We say that M is a Hermitian symmetric space if for each point p M there exists a symmetry s : M M ∈ p → that satisfies the following properties:

1) s is an involution, i.e. s2 Id, p p =

2) sp is holomorphic with respect to the complex structure on M,

3) s is an with respect to the specified Riemannian structure on M ∞, i.e. g(s (X ),s (Y )) g(X ,Y ), p p p =

4) sp has p as its isolated fixed point.

Example 1.10. As we observed, the complex plane C is naturally a Hermitian manifold. To see that it is also sym- metric, note that the group of plane translations acts transitively on C, and these actions are holomorphic isome- tries. Hence we may check the conditions 1)–4) for one point, for example 0. Then a symmetry is given by z z: 7→ − it is an involution, a holomorphic isometry, and p is its only fixed point. Similarly, we have Hermitian symmetric spaces C/Λ, where Λ ω Z ω Z is a lattice. N = 1 ⊕ 2 Example 1.11. Recall that the group

µa b¶ SL (R) : {γ M (R) detγ ad bc 1} 2 = = c d ∈ 2 | = − =

acts on H by Möbius transformations az b γ z : + . · = cz d + 2 (d X )2 (dY )2 The maps z γ z are holomorphic, and they are also isometries. Indeed, the Riemannian metric ds + 7→ · = y2 may be written in coordinates z x i y and z x i y as = + = − dz dz ds2 4 . = − (z z)2 − Using the formula d( z) det dz dz , we see that γ γ (cz d)2 (cz d)2 · = + = + d(γ z)d(γ z) dz dz 4 · · 4 . − (γ z γ z)2 = − (z z)2 · − · − µ py x/py¶ The action is transitive, since x i y i for any x i y H . Hence we again may consider one + = 0 1/py · + ∈ particular point, e.g. i H . Now z 1/z is an involutive holomorphic isometry, having i as the only fixed point. ∈ 7→ − We conclude that H is a Hermitian symmetric space. N

Example 1.12. Consider P1(C) with the natural Riemannian structure on the Riemann sphere

C { } , R3. ∪ ∞ → Any rotation of the sphere is a holomorphic isometry, and the group of rotations acts transitively. Rotation by π along the Z -axis (that is, z z on C and ) is an involution, leaving fixed the poles 0 and . So P1(C) is a 7→ − ∞ 7→ ∞ ∞ Hermitian symmetric space. N

3 1.4 Bergman metric Let Ω Cn be a bounded domain, i.e. a nonempty bounded open connected subset. It turns out that there is a ⊂ canonical way to assign a Riemannian metric to Ω, so that every biholomorphic map Ω Ω is an isometry with → respect to it. We sketch the construction following S.G. Krantz, Function Theory of Several Complex Variables, referred as [Krantz] below. As always, for functions f ,g : Ω C we have an inner product → Z f ,g : f (z)g(z)dz, 〈 〉 = Ω which is correctly defined on the space of square-integrable functions

L2(Ω) : {f : Ω C f , f }. = → | 〈 〉 < ∞ We have the subspace H(Ω) L2(Ω) of holomorphic square integrable functions, and in fact it is a closed subspace. ⊂ This means that H(Ω) is a Hilbert space, just as L2(Ω)—namely, it is complete with respect to the associated norm p f : f , f . k k = 〈 〉 Definition-theorem 1.13. There exists a unique map K : Ω Ω C, such that × → 1) for fixed w Ω the function z K (z,w) lies in H(Ω), ∈ 7→ 2)K (z,w) K (w,z), = 3) it satisfies the reproducing property Z f (z) K (z,w) f (w)dw = Ω for all f H(Ω). ∈ We call K (z,w) the Bergman kernel1.

The construction is roughly the following. Let (φk )k N be an orthonormal basis of H(Ω). We set ∈ X K (z,w) : φk (z)φk (w). = k

This converges, and in fact satisfies 1) [Krantz, Proposition 1.4.7]. It is clear that 2) is satisfied as well. Finally, we check 3): X Z f f ,φk φk K ( ,w) f (w)dw. = k 〈 〉 = Ω − The uniqueness of K (z,w), and hence the fact that it does not depend on the choice of an orthonormal basis of H(Ω), is due to 2) and 3). Indeed, suppose K 0(z,w) is another Bergman kernel. Then Z Z K (z,w) K (w,z) K 0(z,t)K (w,t)dt K (w,t)K 0(z,t)dt K 0(z,w) K 0(z,w). = = Ω = Ω = = Example 1.14. As one can imagine, explicitly calculating the Bergman kernel is difficult. It is possible in some easy cases, e.g. for the unit open disk D : {z C z 1}. = ∈ | | | < The Bergman kernel for D is given by 1 1 K (z,w) = π (1 z w)2 − · —see the calculation in [Krantz, Theorem 1.4.22]. N

1 STEFAN BERGMAN (1895–1977), a Polish-born American mathematician

4 Proposition 1.15. Let Ω Cn be a bounded domain and let K (z,w) be the corresponding Bergman kernel. Then ⊂ K (z,z) 0 for all z Ω. > ∈ Proof. We have from the construction of the Bergman kernel

X X 2 K (z,z) φk (z)φk (z) φk (z) , = k = k | |

and there are k such that φ (z) , 0, since there are nonzero functions in H(Ω). | k | ■ Definition-theorem 1.16. Let Ω Cn be a bounded domain. Then there exists a canonical Hermitian metric on Ω ⊂ given by 2 X i j ∂ g gi j dz dz , gi j (z) logK (z,z). = = ∂zi ∂z j This is called the Bergman metric on Ω.

(Checking that g (z) is positive definite for all z Ω requires some thought; see [Krantz, Chapter 1, Exercise 39].) i j ∈ Here is a truly remarkable property of the Bergman metric:

Theorem 1.17. Let Ω ,Ω Cn be two bounded domains, and let f : Ω Ω be a biholomorphic map. Then f is 1 2 ⊂ 1 → 2 an isometry with respect to the Bergman metric.

(This is [Krantz, Proposition 1.4.15].)

Example 1.18. For the open unit disk we have K (z,z) 1 1 , and a little calculation (e.g. in real coordinates D π (1 z 2)2 ³ ´ = ³−| | ´ x i y z, keeping in mind that ∂ 1 ∂ i ∂ and ∂ 1 ∂ i ∂ ) gives the Bergman metric + = ∂z = 2 ∂x − ∂y ∂z = 2 ∂x + ∂y

∂2 2 g(z) logK (z,z) . = ∂z ∂z = (1 z 2)2 − | | In fact, this is the so-called Poincaré metric. N We recall that H and D are two different models of the hyperbolic plane, and the upper half plane H corre- sponds to the open disk D via the Cayley transform

H D, → i z z − 7→ i z 1 z + i − z. 1 z ← + [

5 i

0

A little calculation shows that the canonical Bergman metric on D corresponds to the metric on H that we considered above (up to scaling the latter by an appropriate constant). Indeed, let z,z denote the coordinates on H , and let w,w be the coordinates on D.

µ i z ¶ 2i dz dw d − − , = i z = (i z)2 + + µ i z ¶ 2i dz dw d + , = i z = (i z)2 − −

2i dz 2i dz 2 − 2dw dw · (i z)2 · (i z)2 8 dz d z 8 dz d z 2 dz d z 2 + − 2 2 2 2 . (1 w w) = Ã i z i z! = ((i z)(i z) (i z)(i z)) = (2i (z z)) = − (z z) − 1 − + + − − − + − − − i z · i z + − Remark 1.19. Any Hermitian symmetric domain D may be embedded in some Cn as a bounded symmetric do- main Ω. This means that D has a unique Hermitian metric that maps to the Bergman metric on Ω.

1.5 A higher-dimensional example: Siegel upper half-space

We have the transitive action of SL2(R) on the upper half-plane H by Möbius transformations. We see that the stabilizer of i H identifies with ∈ µ a b¶ SO (R) { a,b R, a2 b2 1}, 2 = b a | ∈ + = − which is a , and we have

H SL2(R)/SO2(R). =

6 This has a higher-dimensional generalization, the Siegel upper half-space, consisting of symmetric complex n n matrices with positive definite imaginary part: ×

Hn : {Z X i Y X ,Y Mn(R), X > X , Y > Y , Y 0} = = + | ∈ = = > n (n 1)/2 (in particular, H1 H is the upper half-plane). We may see this as an open subset of C + by sending a matrix = n (n 1)/2 (zi j ) to the point (zi j )j i C + , so there is a natural complex structure. ≥ ∈ We have the µ ¶ µ ¶ µ ¶ AB 0 In 0 In Sp (R) : {γ M (R) γ> γ } 2n = = CD ∈ 2n | · I 0 · = I 0 − n − n

(in particular, SL2(R) Sp (R)), and Sp (R) acts transitively on Hn by “Möbius transformations” = 2 2n 1 γ Z (AZ B)(CZ D)− . · = + + We have a subgroup of unitary matrices

U {Z M (C) Z † Z ZZ † I }, n = ∈ n | = = n † where Z Z > denotes the conjugate transpose (in particular, SO2(R)  U1). This may be identified with a sub- = group of Sp2n(R): µ XY ¶ X iY , + 7→ YX − and this is a maximal compact subgroup of Sp2n(R). In fact,

Hn Sp (R)/Un. = 2n Now consider the set of matrices

† Dn : {Z Mn(C) Z > Z and In Z Z 0}, = ∈ | = − > n (n 1)/2 which may be identified with a bounded domain in C + (in particular, for n 1 it is just the open unit disk), = so we know that for Dn there is a canonical Hermitian metric—the Bergman metric. There is a “Cayley transform”, identifying Hn and Dn:

Hn Dn, → 1 Z (i I Z )(i I Z )− 7→ n − n + 1 i (In Z )(In Z )− Z . − + ←[ This allows us to conclude that Hn also carries an invariant Hermitian metric. Further, it is a Hermitian sym- metric space (and hence Dn). For this observe that the point i In Hn is the only fixed point of the involution µ ¶ ∈ 0 In − Sp2n(R). In 0 ∈

1.6 Isometries of Hermitian symmetric domains Theorem 1.20. Let (M,g) be a symmetric space, with group of orientation preserving isometries Isom(M,g). Denote by Isom(M,g)+ the connected component of the identity (in the analytic topology). For a point p M denote by ∈ K Isom(M,g)+ be the subgroup leaving p fixed. p ⊂

7 For example,

(M,g) isometries orientation-preserving p Kp H PGL2(R)  PSL2(R) o Z/2 PSL2(R) i SO2(R)/{ 1} ± P1(C)O (R) SO (R) 0 U SO (R) 3 3 1 = 2

Then Kp is compact, and we have an isomorphism of smooth manifolds

Isom(M,g)+/K M, p → γ K γ(p); · p 7→ in particular, the action of Isom(M,g)+ on M is transitive.

For a proof we refer to [Helgason, , Lie Groups, and Symmetric Spaces, II.4.3].

Theorem 1.21. Let (M,g) be a Hermitian symmetric space. Consider the following groups:

• Isom(M ∞,g) = isometries of the underlying real manifold M ∞.

• Hol(M) = M M as a complex manifold. →

• Isom(M,g) = Isom(M ∞,g) Hol(M) = holomorphic isometries. ∩ Then in fact Isom(M ∞,g)+ Hol(M)+ Isom(M,g)+, = = in particular Hol(M)+/Kp  M ∞.

— this is claimed in Milne’s notes.

Theorem 1.22. Let (M,g) be a Hermitian symmetric domain. Consider the real H : Hol(M)+ and its Lie = algebra h : Lie(H). There is a unique connected algebraic subgroup G of GL(h) such that inside GL(h) =

G(R)+ Hol(M)+. = Moreover, G(R)+ G(R) Hol(M). = ∩ Example 1.23. For the upper half-plane H every isometry is either

• holomorphic (and orientation-preserving), or

1 • differs from a holomorphic map by the orientation change z z− , which is antiholomorphic. 7→

We have G PGL , and PGL (R)+ PSL (R). N = 2 2 = 2 Theorem 1.24. Let M be a Hermitian symmetric domain. For each point p M there exists a unique homomor- ∈ phism u : U Hol(M) p 1 → (where U : {z C z 1} is the ), such that u (z) fixes p and acts on T M as multiplication by z. 1 = ∈ | | | = p p (A proof of this is sketched in Milne’s notes)

8 Example 1.25. Consider M H . Then Hol(M)  PSL2(R) : SL2(R)/{ 1}. We want morphisms that fix i, and we = = ± keep in mind that Stab R (i) SO (R). In fact, U may be viewed as SO (R) via SL2( ) = 2 1 2

h : U SL (R), 1 → 2 µ a b¶ z a ib . = + 7→ b a − 2 h(z) fixes the point i, and we compute that it acts on Ti H as multiplication by z :

µµ ¶ ¶¯ µ ¶¯ d a b ¯ d aX b ¯ f X ¯ f + ¯ d X b a · ¯ = d X bX a ¯ − X i X i = − +¯ = ¯ d aX b ¯ d ¯ + ¯ f (X )¯ = d X bX a ¯X i · d X ¯X i − + = = 2 2 ¯ a b d ¯ + 2 f (X )¯ = (a bi) · d X ¯X i − ¯ = d ¯ (a bi)2 f (X )¯ . = + · d X ¯X i = Since we want action by multiplication by z, we define

u : U PSL (R) : SL (R)/{ 1}, 1 → 2 = 2 ± z h( pz) (mod 1). 7→ ±

The choice of pz does not change the class modulo 1, and so u is well-defined and u(z) acts on Ti H by ± multiplication by z. N

2 Classification of symmetric Hermitian domains

2.1 Preliminaries: Cartan involutions Definition-theorem 2.1. Let G be a connected reductive algebraic group over R. Then there exists an θ : G G (as an algebraic group over R) with θ2 Id, such that the group → = G(θ) : {g G(C) g θ(g)} = ∈ | = is compact (by g we denote complex conjugation). Such θ is called a Cartan involution, and it is unique up to conjugation by an element of G(R).

Example 2.2. Let V be a finite dimensional real vector space. Consider the group GL(V ). A choice of some basis defines an isomorphism GL(V )  GLn, and on GLn we have the usual matrix transpose X X >. The map 7→ 1 θ : X (X >)− 7→ is a Cartan involution: (θ) 1 GL (R) {Z GL (C) Z (Z >)− } U n = ∈ n | = = n is compact. A different choice of basis of GL(V ) differs from θ by conjugation by an element of GLn(R). N Example 2.3. If G is a reductive algebraic group over R and G , GL(V ) is a faithful representation over R, then → there exists some basis for V for which G is stable under transpose X X > (exactly because G is assumed to be 1 7→ reductive!). The restriction of X (X >)− to G is a Cartan involution. In fact, all Cartan involutions arise this 7→ way. N

9 µa b¶ µ d b¶ Example 2.4. For instance, take SL GL . For a matrix M the inverse is given by 1 − , so we 2 ⊂ 2 = c d detM · c a − take a Cartan involution µa b¶ µ d c¶ θ : − . c d 7→ b a − µ 0 1¶ Note that it is the same as conjugation by : 1 0 − 1 µ 0 1¶ µa b¶ µ 0 1¶− µ d c¶ − . 1 0 · c d · 1 0 = b a − − − Now we see that µ a b¶ SL(θ) { a 2 b2 1} SU , 2 = b a | | | + | | = = 2 − which is a closed bounded subset in C2, hence a compact subgroup. N

10