03.02.2015 Views

GAUSS-BONNET THEOREM AND CROFTON TYPE FORMULAS IN ...

GAUSS-BONNET THEOREM AND CROFTON TYPE FORMULAS IN ...

GAUSS-BONNET THEOREM AND CROFTON TYPE FORMULAS IN ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>AND</strong> <strong>CROFTON</strong> <strong>TYPE</strong><br />

<strong>FORMULAS</strong> <strong>IN</strong> COMPLEX SPACE FORMS<br />

JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

Abstract. We give an expression, in terms of the so-called Hermitian<br />

intrinsic volumes, for the measure of the set of complex<br />

r-planes intersecting a regular domain in any complex space form.<br />

Moreover, we obtain two different expressions for the Gauss-Bonnet-<br />

Chern formula in complex space forms. One of them expresses the<br />

Gauss curvature integral in terms of the Euler characteristic and<br />

some Hermitian intrinsic volumes. The other one, which is shorter,<br />

involves the measure of complex hyperplanes meeting the domain.<br />

As a tool, we obtain new variation formulas in integral geometry<br />

of complex space forms. Finally, we express the average over the<br />

complex r-planes of the total Gauss curvature of the intersection<br />

domain.<br />

Prepublicació Núm 01, gener 2009.<br />

Departament de Matemàtiques.<br />

http://www.uab.cat/matematiques<br />

1. Introduction<br />

In the space of constant sectional curvature k, M n (k), Santaló [San04]<br />

gave the expression of the measure of the set of r-planes meeting a regular<br />

domain in terms of the mean curvature integrals. Suppose that L r<br />

denotes the space of r-dimensional planes in M n (k) and dL r a density<br />

of L r invariant under the isometry group of M n (k). If Ω ⊂ M n (k) is a<br />

regular domain and r = 2l, then<br />

(1)<br />

∫<br />

L 2l<br />

χ(Ω ∩ L 2l )dL 2l = c 0 vol(Ω) +<br />

l∑<br />

c i k l−i M 2i−1 (∂Ω),<br />

where c i are known and depend only on n, r and i, while M j (∂Ω)<br />

denotes the mean curvature integral defined by<br />

( ) −1 ∫ n − 1<br />

(2) M j (∂Ω) =<br />

σ j (II)dx<br />

j<br />

∂Ω<br />

1991 Mathematics Subject Classification. Primary 53C65; Secondary 52A22,<br />

53C55.<br />

Key words and phrases. Complex space forms, Gauss-Bonnet formula, Crofton<br />

formulas, Valuation.<br />

Work partially supported by FEDER/MEC grant # MTM2006-04353 and by<br />

the Departament d’Innovació Universitari i de Recerca from the Generalitat of<br />

Catalonia and the ESF.<br />

1<br />

i=1


2 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

where σ j (II) is the j-th symmetric elementary function of the eigenvalues<br />

of the second fundamental form. An analogous formula holds in<br />

the case of odd-dimensional planes.<br />

In the proof of formula (1), Santaló used the Gauss-Bonnet-Chern<br />

theorem in M n (k), that for n even states<br />

(3)<br />

M n−1 (∂Ω)= O n<br />

2 χ(Ω)−kc n−3M n−3 (∂Ω)−...−k n−2<br />

2 c1 M 1 (∂Ω)−k n 2 vol(Ω),<br />

and the reproductive property of the mean curvature integrals in M n (k),<br />

∫<br />

(4)<br />

(∂Ω ∩ L r )dL r = cM i (∂Ω),<br />

M (r)<br />

i<br />

L r<br />

where M (r)<br />

i (∂Ω∩L r ) denotes the i-th mean curvature integral of ∂Ω∩L r<br />

as a hypersurface in L r and c is known and depends only on n, r and<br />

i.<br />

In this paper we generalize formula (1) and (3) to the space of constant<br />

holomorphic curvature 4ɛ, CK n (ɛ). The role of L r in (1) will be<br />

played by L C r , the space of complex r-planes (totally geodesic complex<br />

submanifolds). For simplicity, we shall restrict to compact domains<br />

Ω ⊂ CK n (ɛ) with smooth boundary, and call them regular domains.<br />

The method used by Santaló cannot be applied in this situation.<br />

Mean curvature integrals are defined in a complex space form as in (2),<br />

but in the standard Hermitian space C n , it was not certain whether the<br />

reproductive property remains true for the mean curvature integrals<br />

when we integrate it over L C r . Moreover, in the complex projective<br />

space and in the complex hyperbolic space, an explicit Gauss-Bonnet<br />

formula is not known. Instead, we use variational arguments, as well<br />

as some facts about the theory of valuations, which we briefly describe<br />

next.<br />

Definition 1.1. Let K(V ) denote the family of non-empty compact<br />

convex subsets of a finite dimensional real vector space V of dimension<br />

n. A scalar valued functional φ : K(V ) → C is called a valuation if<br />

whenever A, B, A ∪ B ∈ K(V ).<br />

φ(A ∪ B) = φ(A) + φ(B) − φ(A ∩ B)<br />

The space of invariant valuations under the full group of isometries of<br />

R n was studied by Hadwiger [Had57], who proved that the dimension<br />

of this space is n + 1. Mean curvature integrals, volume and Euler<br />

characteristic form a basis of this space.<br />

On the standard Hermitian space C n with its isometry group IU(n) =<br />

C n ⋊ U(n), Alesker [Ale03] proved that there are more linearly independent<br />

valuations than on R n , the dimension of this space is ( )<br />

n+2<br />

2 .<br />

Both the Euler characteristic, and the measure of complex planes intersecting<br />

the domain belong to this space.


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 3<br />

Bernig and Fu [BF08], consider several valuation bases on C n . Here<br />

we will use the basis whose elements {µ k,q } {k,q} are called Hermitian<br />

intrinsic volumes. These valuations where first introduced by Park<br />

in complex space forms (cf. [Par02]). In section 2 we recall their<br />

definition.<br />

The main results of this paper can be stated as follows.<br />

Theorem 1.2. Let Ω be a regular domain in CK n (ɛ). Then, for r =<br />

1, ..., n − 1<br />

∫<br />

( ) −1 n − 1<br />

χ(Ω ∩ L r )dL r = vol(G C n−1,r) (ɛ r (r + 1)vol(Ω)+<br />

r<br />

(5)<br />

L C r<br />

+<br />

∑n−1<br />

k=n−r<br />

+<br />

ɛ k−(n−r) ω 2n−2k<br />

( n<br />

k) −1<br />

· ((k + r − n + 1)µ 2k,k +<br />

∑k−1<br />

q=max{0,2k−n}<br />

( )<br />

1 2k − 2q<br />

µ<br />

4 k−q 2k,q ))<br />

k − q<br />

where dL r denotes an invariant measure in the space of complex r-<br />

planes L C r . Moreover,<br />

(6)<br />

O 2n−1 χ(Ω) = 2n(n + 1)ɛ n vol(Ω) +<br />

⎛<br />

⎞<br />

∑n−1<br />

ɛ c O 2n−2c−1<br />

∑c−1<br />

( )<br />

+ ) ⎝<br />

1 2c − 2q<br />

µ<br />

4 c−q 2c,q +(c + 1)µ 2c,c<br />

⎠<br />

c − q<br />

c=0<br />

( n−1<br />

c<br />

q=max{0,2c−n}<br />

where ω i denotes the volume of the euclidean unit ball and O i the volume<br />

of the euclidean unit sphere.<br />

Formula (5) is a generalization of (1) in the sense that the measure<br />

of the complex r-planes meeting a regular domain is expressed as a<br />

linear combination of the volume and some other valuations related to<br />

mean curvature integrals. This answers a question posed by Naveira<br />

in [Nav05]. In case r = 1, formula (5) was already proved by different<br />

methods in [Aba09]. For 2r ≥ n, and ɛ = 0, formula (5) was proved in<br />

[BF08].<br />

Formula (6) generalizes to complex space forms the Gauss-Bonnet<br />

theorem (3). In complex dimensions n = 2, 3, formula (6) was obtained<br />

in [Par02].<br />

Combining expressions (5) and (6) we obtain<br />

∫<br />

M 2n−1 (∂Ω) = O 2n−1 χ(Ω) − 2nɛ χ(Ω ∩ L n−1 )dL n−1 −<br />

k=1<br />

L C n−1<br />

∑n−1<br />

( ) −1 n − 1<br />

− ɛ k O 2n−2k−1 µ 2k,k − 2nɛ n vol(Ω).<br />

k<br />

This expression is similar to the following one for real space forms.


4 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

Theorem ([Sol06]). Let S be a hypersurface bounding a compact domain<br />

Q in a real space form with sectional curvature k and dimension<br />

n. Then<br />

∫<br />

2(n − 1)<br />

M n−1 (S) = O n−1 χ(Q) − k χ(Q ∩ L n−2 )dL n−2 .<br />

O n−2 L n−2<br />

The main idea for the proof of Theorem 1.2 is to take variation along<br />

a vector field in CK n (ɛ), in both sides of equalities (5) and (6), and to<br />

compare them.<br />

In order to obtain a first expression of the variation of ∫ χ(∂Ω ∩<br />

L C r<br />

L r )dL r along a vector field in CK n (ɛ), we proceed as in [Sol06] (see<br />

Section 3.1). In C n , the variation of the Hermitian intrinsic volumes<br />

was obtained by Bernig and Fu [BF08]. Here we use the same method<br />

to find the generalization for ɛ ≠ 0 (see Section 3.2).<br />

Using formula (5), we prove in Section 6 that the total Gauss curvature<br />

does not satisfy the reproductive property and we get in C n the<br />

following expression:<br />

∫<br />

( ) −1 ( ) n − 1 n<br />

M 2r−1 (∂Ω ∩ L r )dL r = 2rω2rvol(G 2 C n−1,r)<br />

−1·<br />

r r<br />

L C r<br />

(7)<br />

⎛<br />

· ⎝<br />

∑n−r<br />

q=max{0,n−2r}<br />

( )<br />

1 2n − 2r − 2q<br />

4 n−r−q n − r − q<br />

⎞<br />

µ 2n−2r,q<br />

⎠ .<br />

In [Aba09], it is proved that the reproductive property (4) is not<br />

satisfied by the mean curvature integral either.<br />

Acknowledgments<br />

We wish to thank Andreas Bernig for illuminating discussions during<br />

the preparation of this work.<br />

2. Hermitian intrinsic volumes on CK n (ɛ)<br />

Let CK n (ɛ) be a (simply connected) complex space form with constant<br />

holomorphic curvature 4ɛ. We denote by T ′ CK n (ɛ) the unit tangent<br />

bundle of CK n (ɛ).<br />

Definition 2.1. Let Ω be a regular domain in CK n (ɛ). The unit (inner)<br />

normal bundle of ∂Ω is defined as<br />

N(Ω) ={(p, v) ∈ T ′ CK n (ɛ) : p ∈ ∂Ω, v inner normal to T ′ p∂Ω}.<br />

Given a 4n − 1 form ω in T ′ CK n (ɛ) we may consider, for every<br />

regular domain Ω the integral over N(Ω) of the form ω. The resulting<br />

functional is called a smooth valuation.<br />

Let z ∈ CK n (ɛ), e 1 ∈ T ′ CK n (ɛ) and let {z; e 1 , ..., e n } be a moving<br />

frame defined on an open subset U ⊂ T ′ CK n (ɛ). We denote by


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 5<br />

{ω 1 , ω 2 , ..., ω n } the 1-forms on T ′ CK n (ɛ) defined as the dual basis of<br />

{e 1 , ..., e n }, and by {ω ij } the connection forms of CK n (ɛ). That is, if<br />

( , ) denotes the Hermitian product on CK n (ɛ) and ∇ the Levi-Civita<br />

connection, then<br />

ω j = (dz, e j ) and ω jk = (∇e j , e k ) where j, k ∈ {1, ..., n}.<br />

Thus, these forms are C-valued. We denote<br />

(8)<br />

ω j = α j + iβ j ,<br />

ω jk = α jk + iβ jk .<br />

Remark 2.2. Forms α 1 , β 1 and β 11 are global forms in T ′ CK n (ɛ). We<br />

denote by α, β, γ the forms α 1 , β 1 , β 11 , respectively. Note that α<br />

coincides with the contact form of the unit tangent bundle.<br />

Lemma 2.3. Let Ω ⊂ CK n (ɛ) be a regular domain. Forms α and dα<br />

vanish at N(Ω) ⊂ T ′ CK n (ɛ).<br />

Proof. Let V ∈ T (p,v) N(Ω). Then, α(V ) (p,v) = 〈dπ(V ), v〉 = 0 where<br />

π : T ′ CK n (ɛ) → CK n (ɛ) is the canonical projection.<br />

To prove that dα vanishes at the unit normal bundle, we consider<br />

the inclusion of the unit normal bundle to the unit tangent bundle<br />

i : N(Ω) → T ′ (Ω) and we use that exterior differential commutes with<br />

the inclusion map to obtain<br />

(9)<br />

dα| N(Ω) = (i ∗ ◦ d)(α) = (d ◦ i ∗ )(α) = d(i ∗ α) = d(0) = 0.<br />

Consider the following 2-forms in T ′ CK n (ɛ)<br />

n∑<br />

θ 0 = −Im((∇e 1 , ∇e 1 )) = −Im( ω 1i ⊗ ω 1i )<br />

i=1<br />

θ 1 = −Im((dz, ∇e 1 ) − (∇e 1 , dz))<br />

n∑<br />

n∑<br />

= −Im( ω i ⊗ ω 1i − ω 1i ⊗ ω i )<br />

i=1<br />

i=1<br />

θ 2 = −Im((dz, dz)) = −Im(<br />

n∑<br />

ω i ⊗ ω i )<br />

i=1<br />

θ s = Re((dz, ∇e 1 ) − (∇e 1 , dz))<br />

n∑<br />

n∑<br />

= Re( ω i ⊗ ω 1i − ω 1i ⊗ ω i ).<br />

i=1<br />

This forms coincide with the invariant 2-forms, θ 0 , θ 1 , θ 2 and θ s<br />

defined in T ′ C n by Bernig and Fu [BF08, p.14]. Note that, θ s is the<br />

symplectic form of T CK n (ɛ).<br />

Remark 2.4. Park [Par02] defined invariant 2-forms in T ′ CK n (ɛ) similar<br />

to the ones in (9).<br />

i=1<br />


6 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

The forms β k,q and γ k,q defined in T ′ C n in [BF08], can be extended<br />

to T ′ CK n (ɛ) from (9).<br />

Definition 2.5. For positive integers k, q ∈ N with max{0, k − n} ≤<br />

q ≤ k 2 < n, it is defined in Ω2n−1 (T ′ CK n (ɛ))<br />

where<br />

β k,q := c n,k,q β ∧ θ n−k+q<br />

0 ∧ θ k−2q−1<br />

1 ∧ θ q 2, k ≠ 2q<br />

γ k,q := c n,k,q<br />

2 γ ∧ θn−k+q−1 0 ∧ θ k−2q<br />

1 ∧ θ2, q n ≠ k − q<br />

1<br />

c n,k,q :=<br />

q!(n − k + q)!(k − 2q)!ω 2n−k<br />

and ω 2n−k denotes the volume of the (2n − k)-dimensional euclidean<br />

ball.<br />

Given a regular domain Ω ⊂ CK n (ɛ), we define (for max{0, k − n} ≤<br />

q ≤ k < n) 2<br />

∫<br />

∫<br />

Bk,q(Ω) Ω := β k,q (k ≠ 2q), Γ Ω k,q(Ω) :=<br />

(n ≠ k−q).<br />

N(Ω)<br />

γ k,q<br />

N(Ω)<br />

In C n , it is satisfied B Ω k,q (Ω) = ΓΩ k,q (Ω) since dβ k,q = dγ k,q . Next we<br />

recall the exterior derivative of θ 0 , θ 1 and θ 2 , which can be found in<br />

[BF08] when ɛ = 0, or in [Par02] for general ɛ.<br />

Lemma 2.6 ([Par02]). In T ′ CK n (ɛ) it is satisfied<br />

dα = −θ s , dθ 0 = −ɛ(α ∧ θ 1 + β ∧ θ s ),<br />

dβ = θ 1 , dθ 1 = 0,<br />

dγ = 2θ 0 − 2ɛθ 2 − 2ɛα ∧ β, dθ 2 = 0.<br />

Next, we give the relation among {B Ω k,q (Ω)} and {ΓΩ k,q (Ω)} in CKn (ɛ)<br />

which generalizes the relation in C n .<br />

Proposition 2.7. In CK n (ɛ), for any positive integers k, q such that<br />

max{0, k − n} < q < k/2 < n it is satisfied<br />

c n,k,q<br />

Γ Ω k,q(Ω) = Bk,q(Ω) Ω − ɛ B<br />

c<br />

k+2,q+1(Ω).<br />

Ω<br />

n,k+2,q+1<br />

Proof. We denote by I the ideal generated by α, dα and the exact forms<br />

in N(Ω). If λ, ρ are (2n − 1)-forms in N(Ω) equal modulo I, then by<br />

Lemma 2.3<br />

∫ ∫<br />

λ = ρ.<br />

Thus, it is enough to prove<br />

N(Ω)<br />

c n,k,q<br />

N(Ω)<br />

(10) γ k,q ≡ β k,q − ɛ β k+2,q+1 mod I.<br />

c n,k+2,q+1


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 7<br />

Consider the form η = (θ s − β ∧ γ) ∧ θ n−k+q−1<br />

0 θ k−2q−1<br />

1 θ q 2. As dη is<br />

exact, we have dη ≡ 0 mod I. On the other hand, by Lemma 2.6 it<br />

follows that modulo I<br />

dη ≡ −γθ n−k+q−1<br />

0 θ k−2q<br />

1 θ q 2+2βθ n−k+q<br />

0 θ k−2q−1<br />

1 θ q 2−2ɛβθ n−k+q−1<br />

0 θ k−2q−1<br />

1 θ q+1<br />

2 .<br />

Using the definition of γ k,q and β k,q we obtain the relation in (10).<br />

Remark 2.8. In complex dimensions n = 2, 3, the previous relations<br />

were found in [Par02].<br />

In view of the previous equalities, we define (for max{0, k − n} ≤<br />

q ≤ k < n) 2<br />

{ B<br />

Ω<br />

(11) µ k,q (Ω) := k,q (Ω) if k ≠ 2q<br />

Γ Ω 2q,q(Ω) if k = 2q.<br />

Remark 2.9. In [BF08], it is proved that valuations {µ k,q , vol} where<br />

k ∈ {0, ..., n − 1} and q ∈ {max{0, k − n}, ..., ⌊k/2⌋} form a basis of<br />

invariant continuous valuations on C n .<br />

3. Variation formulas<br />

3.1. Variation of Hermitian intrinsic volumes. In order to study<br />

the variation on CK n (ɛ) of the Hermitian intrinsic volumes, we follow<br />

the method used by Bernig and Fu [BF08] in Corollary 2.6. First, we<br />

recall the definition of the Rumin operator, introduced in [Rum94], and<br />

the definition of the Reeb vector field in a contact manifold.<br />

Definition 3.1. Given µ ∈ Ω 2n−1 (T ′ CK n (ɛ)), let α be the contact form<br />

of T ′ CK n (ɛ), and let α ∧ ξ ∈ Ω 2n−1 (T ′ CK n (ɛ)) be the unique form such<br />

that d(µ + α ∧ ξ) is multiple of α (cf. [Rum94]). Then the Rumin<br />

operator D is defined as<br />

Dµ := d(µ + α ∧ ξ).<br />

Definition 3.2. Let M be a contact manifold and let α be the contact<br />

form. The Reeb vector field T is the unique vector field over M such<br />

that<br />

{<br />

iT α = 1,<br />

(12)<br />

L T α = 0.<br />

If the contact manifold is the unit tangent bundle of a Riemannian<br />

manifold, then the Reeb vector field is the geodesic flow (cf. [Bla76, p.<br />

17]).<br />

Lemma 3.3. In T ′ CK n (ɛ), it is satisfied<br />

i T α = 1, i T θ 1 = γ,<br />

i T θ 2 = β, i T β = i T γ = i T θ 0 = 0.<br />


8 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

Proof. The first equality comes directly from the definition (cf. (12)).<br />

As T is the geodesic flow, we have α i (T ) = β i (T ) = 0 and α 1i = β 1i = 0,<br />

i ∈ {2, ..., n}. By Definition in (9), we obtain the result.<br />

□<br />

Given a smooth valuation µ, and a vector field X with flow ϕ t , we<br />

are interested in computing<br />

δ X µ(Ω) := d dt∣ µ(ϕ t (Ω)).<br />

t=0<br />

This can be done by means of the following result stated in [BF08].<br />

Lemma 3.4 (Lemma 2.5 [BF08]). Suppose Ω ⊂ R n is a regular domain,<br />

N is the inner normal field to ∂Ω, X is a smooth vector field on<br />

C n and µ is a smooth valuation given by a (2n − 1)-form ρ. Then<br />

∫<br />

δ X µ(Ω) = 〈X, N〉 i T (Dρ)<br />

N(Ω)<br />

where T is the Reeb vector field on T ′ R n and Dρ is the Rumin operator<br />

of ρ.<br />

Although this result is stated and proved in C n , the given proof is<br />

also valid in an arbitrary Riemannian manifold.<br />

From Lemma 2.6, we obtain the exterior differential of the forms β k,q<br />

and γ k,q .<br />

Lemma 3.5. In CK n (ɛ)<br />

dβ k,q = c n,k,q (θ n−k+q<br />

0 ∧θ k−2q<br />

1 ∧θ q 2−ɛ(n−k+q)α∧β∧θ n−k+q−1<br />

0 ∧θ k−2q<br />

1 ∧θ q 2)<br />

and<br />

dγ k,q =c n,k,q (θ n−k+q<br />

0 ∧ θ k−2q<br />

1 ∧ θ q 2 − ɛθ n−k+q−1<br />

0 ∧ θ k−2q<br />

1 ∧ θ q+1<br />

2<br />

− ɛα ∧ β ∧ θ n−k+q−1<br />

0 ∧ θ k−2q<br />

1 ∧ θ q 2<br />

(n − k + q − 1)<br />

− ɛ α ∧ γ ∧ θ n−k+q−2<br />

0 ∧ θ k−2q+1<br />

1 ∧ θ q 2<br />

2<br />

(n − k + q − 1)<br />

− ɛ β ∧ γ ∧ θ s ∧ θ n−k+q−2<br />

0 ∧ θ k−2q<br />

1 ∧ θ q<br />

2<br />

2).<br />

Notation 3.6. Let Ω be a regular domain in CK n (ɛ) and let N be a<br />

normal field to ∂Ω. Let X be a smooth vector field on CK n (ɛ). We<br />

denote<br />

∫<br />

∫<br />

˜B k,q := ˜B k,q (Ω) = 〈X, N〉β k,q , ˜Γk,q := ˜Γ k,q (Ω) = 〈X, N〉γ k,q .<br />

∂Ω<br />

The variation of the valuations {µ k,q } on C n is given in [BF08, Proposition<br />

4.6]. The next proposition gives this variation in the previous<br />

notation.<br />

∂Ω


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 9<br />

Proposition 3.7 (Proposition 4.6 [BF08]).<br />

δ X µ k,q = 2c n,k,q (c −1<br />

n,k−1,q (k − 2q)2˜Γk−1,q − c −1<br />

n,k−1,q−1 (n + q − k)q˜Γ k−1,q−1<br />

+c −1<br />

n,k−1,q−1 (n + q − k + 1 2 )q ˜B k−1,q−1 −c −1<br />

n,k−1,q (k − 2q)(k − 2q − 1)˜B k−1,q ).<br />

We extend this result to CK n (ɛ) as follows.<br />

Proposition 3.8. Let X ∈ X(CK n (ɛ)) and let Ω ⊂ CK n (ɛ) be a regular<br />

domain. Then<br />

δ X B Ω k,q (Ω)= 2c n,k,q(c −1<br />

n,k−1,q (k − 2q)2˜Γk−1,q −c −1<br />

n,k−1,q−1 (n + q − k)q˜Γ k−1,q−1<br />

+ c −1<br />

n,k−1,q−1 (n + q − k + 1 2 )q ˜B k−1,q−1 − c −1<br />

n,k−1,q (k − 2q)(k − 2q − 1) ˜B k−1,q<br />

+ɛ(c −1<br />

n,k+1,q+1 (k − 2q)(k − 2q − 1) ˜B k+1,q+1 −c −1<br />

n,k+1,q (n − k + q)(q + 1 2 ) ˜B k+1,q ))<br />

and<br />

δ X Γ Ω 2q,q(Ω) =2c n,2q,q<br />

(<br />

− c −1<br />

n,2q−1,q−1 (n − q)q˜Γ 2q−1,q−1<br />

+ c −1<br />

n,2q−1,q−1 (n − q + 1 2 )q ˜B 2q−1,q−1<br />

+ ɛ(−c −1<br />

n,2q+1,q ((n − q)(2q + 3 2 ) − 1 2 (q + 1)) ˜B 2q+1,q<br />

+ c −1<br />

n,2q+1,q (n − q − 1)(q + 1)˜Γ 2q+1,q<br />

− ɛc −1<br />

n,2q+3,q+1 (n − q − 1)(q + 3 2 ) ˜B 2q+3,q+1<br />

) ) .<br />

Proof. The proof is based on that of Proposition 4.6 in [BF08].<br />

Consider first δ X Bk,q Ω (Ω). Lemma 3.4 provides an expression for the<br />

variation of a valuation given by a smooth form. By this lemma and<br />

Lemma 2.3, it is enough to find the Rumin operator of the form β k,q<br />

modulo α, dα since both forms vanish over the unit normal fiber bundle.<br />

In the case ɛ = 0 (cf. Proposition 4.6 in [BF08]) the Rumin differential<br />

of β k,q is given by Dβ k,q = d(β k,q + α ∧ ξ) where ξ is an invariant form<br />

fulfilling<br />

(13)<br />

ξ ≡ c n,k,q βγθ n+q−k−1<br />

0 θ k−2q−2<br />

1 θ q−1<br />

2<br />

∧ ( )<br />

(n + q − k)qθ1 2 − (k − 2q)(k − 2q − 1)θ 0 θ 2<br />

mod (α, dα).<br />

For general ɛ we take a form ξ ɛ such that ξ ɛ (p,v) ≡ ξ (p ′ ,v ′ ) when we identify<br />

T (p,v) T ′ CK n (ɛ) and T (p ′ ,v ′ )T ′ C n , for every (p, v) ∈ T ′ CK n (ɛ), (p ′ , v ′ ) ∈<br />

T ′ C n . Then, it is clear from Lemma 3.5 that d(β k,q +α∧ξ ɛ ) ≡ 0 modulo<br />

α, since the term multiple of ɛ in the expression of dβ k,q is also multiple<br />

of α.<br />

By Lemma 2.6, the exterior differential ξ for any ɛ is<br />

dξ ≡ c n,k,q θ n+q−k−1<br />

0 θ k−2q−2<br />

1 θ q−1<br />

2 ((n − k + q)qθ1 2 − (k − 2q)(k − 2q − 1)θ 0 θ 2 )<br />

∧ (γθ 1 − 2βθ 0 + 2ɛβθ 2 ) mod α


10 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

and the contraction of dβ k,q with respect to the field T , by Lemma 3.3,<br />

is<br />

i T dβ k,q ≡ c n,k,q θ n+q−k−1<br />

0 θ k−2q−1<br />

1 θ q−1<br />

2<br />

∧ ((k − 2q)γθ 0 θ 2 + qβθ 0 θ 1 − ɛ(n − k + q)βθ 1 θ 2 ) mod α.<br />

By substituting the last expressions in i T Dβ k,q ≡ i T dβ k,q − dξ (mod<br />

α, dα), we get the result.<br />

To compute δ X Γ 2q,q , note that dγ 2q,q has 3 terms which are not multiple<br />

of α (cf. Lemma 3.5). Then the Rumin differential is given by<br />

Dγ 2q,q = d(γ 2q,q + α ∧ (ξ + ξ 2 + ξ 3 )) where ξ corresponds, as above, to<br />

the form in (13),<br />

and<br />

ξ 2 ≡ c n,2q+2,q+1 (n − q − 1)(q + 1)βγθ n−q−2<br />

0 θ q 2 mod (α, dα)<br />

n − q − 1<br />

ξ 3 ≡ c n,2q,q βγθ n−q−2<br />

0 θ q 2 mod (α, dα).<br />

2<br />

Indeed, as in Proposition 4.6 in [BF08], ξ cancels the first term of dγ 2q,q<br />

modulo α and ξ 2 the second one. The third term is canceled by ξ 3 since<br />

it is the same multiple of dα = −θ s .<br />

□<br />

3.2. Variation of the measure of complex r-planes intersecting<br />

a regular domain. We denote by L C r , r ∈ {1, ..., n − 1} the space<br />

of complex r-planes in CK n (ɛ). Complex r-planes are totally geodesic<br />

submanifolds of complex dimension r isometric to CK r (ɛ) (cf. [Gol99,<br />

Lemma 2.2.4]). Moreover, Santaló [San52] proved the following properties<br />

of this space (as usual, J denotes the complex structure).<br />

Lemma 3.9 ([San52]). L C r , is a homogeneous space and<br />

where<br />

L C r ∼ = U ɛ (n)/U ɛ (r) × U(n − r)<br />

⎧<br />

⎨ C n ⋉ U(n), if ɛ = 0,<br />

U ɛ = U(1 + n), if ɛ > 0,<br />

⎩<br />

U(1, n), if ɛ < 0.<br />

Let {g; g 1 , Jg 1 , ..., g n , Jg n } be a local orthonormal frame such that<br />

{g 1 , Jg 1 , ..., g r , Jg r } generate the tangent space of a complex r-plane at<br />

g. The density of L C r is given by<br />

∣ ∣∣∣∣∣∣∣<br />

n∧ ∧<br />

(14) dL r = ∣ ω i ∧ ω i ω ij ∧ ω ij<br />

∣<br />

i=r+1<br />

i=1,...,r<br />

j=r+1,...,n<br />

where {ω i , ω ij } {i,j} are defined as in (8).


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 11<br />

On ∂Ω there is a canonical vector field given by JN, and a distribution<br />

D = 〈N, JN〉 ⊥ . At every point x ∈ ∂Ω, D x is the maximal<br />

complex linear subspace of T x CK(ɛ) contained in T x ∂Ω. We shall consider<br />

the bundle G C n,r(T ∂Ω) whose fiber at every point x ∈ ∂Ω is the<br />

Grassmanian G C n,r(T x ∂Ω) of r-dimensional complex subspaces of D x ;<br />

i.e., G C n,r(T ∂Ω) = {(x, l)|x ∈ ∂Ω, l is a J-invariant r-dimensional linear<br />

subspace of T x ∂Ω}.<br />

Proposition 3.10. Suppose Ω ⊂ CK n (ɛ) is a regular domain, X is<br />

a smooth vector field on CK n (ɛ), φ t is the flow associated to X and<br />

Ω t = φ t (Ω), then<br />

∫<br />

∫<br />

( ∫ )<br />

d<br />

dt∣ χ(Ω t ∩L r )dL r = 〈∂φ/∂t, N〉<br />

σ 2r (II| V )dV dx<br />

t=0 ∂Ω<br />

G C n,r(T x∂Ω)<br />

L C r<br />

where N is the inner normal field and σ 2r (II| V ) denotes the 2r-th symmetric<br />

elementary function of II restricted to V ∈ G C n,r(T x ∂Ω).<br />

Proof. This proof is based on the one in [Sol06, Theorem 4]. For every<br />

l ∈ G C n,r(T x ∂Ω), we make the parallel translation l t of l along φ t (x).<br />

Recall that parallel translation preserves the complex structure (cf.<br />

[O’N83, p. 326]). Then we project l t onto D φt(x), obtaining a complex<br />

r-plane l ′ t (at least for small values of t). We define<br />

As<br />

γ : G C n,r(T ∂Ω) × (−ɛ, ɛ) −→ L C r<br />

((x, l), t) ↦→ exp φt(x) l ′ t.<br />

γ ∗ (dL r ) = ι ∂t (γ ∗ (dL r ))dt = γ ∗ t (ι dγ∂t dL r )dt<br />

where γ t = γ(·, t), in the same way as in [Sol06], using the co-area<br />

formula and taking derivatives with respect to t, we get<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L r )dL r = signφ signK(L r )γ0(ι ∗ dφ∂t dL r ).<br />

t=0 G C n,r (T ∂Ω)<br />

L C r<br />

Suppose that {g; g 1 , Jg 1 , ..., g n , Jg n } is a local orthonormal frame<br />

defined on G C n,r(T ∂Ω 0 ) × (−ɛ, ɛ) such that g((x, l), t) = φ(x, t), γ =<br />

〈g, g 1 , Jg 1 , ..., g r , Jg r 〉 ∩ CK n (ɛ) and Jg n ((x, l), t) = N t (N t denotes the<br />

normal vector to ∂Ω t at φ t (x)). Consider the curve L r (t) given by the<br />

parallel translation of L r along the geodesic given by N, the inner normal<br />

vector to ∂Ω 0 . If P denotes the tangent vector to L r (t), then for<br />

t = 0<br />

ω i (P ) = 〈dg(P ), g i 〉 = 〈 d dt g(L r(t)), g i 〉 = 0,<br />

ω n (P ) = 〈dg(P ), N〉 = 1,<br />

ω kj (P ) = 〈dg k (P ), g j 〉 = 〈 d dt g k(L r (t)), g j 〉 = 0


12 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

where i ∈ {r + 1, r + 1, ..., n − 1, n − 1}, j ∈ {r + 1, r + 1, ..., n, n}, and<br />

k ∈ {1, 1, ..., r, r}. By (14) and last equations we get<br />

dL r = |ω n |ι P dL r<br />

since ι P dL r = |ω n (P )| · | ∧ n<br />

h=r+1 ω h ∧ ω h<br />

∧<br />

ωij |. Thus,<br />

and<br />

So,<br />

ι dγ∂t dL r = |ω n (dγ∂t)|ι P dL r + |ω n |ι dγ∂t ι P dL r<br />

ω n (dγ∂t) = 〈dg(dγ∂t), N〉 = 〈 ∂φ<br />

∂t , N〉,<br />

γ ∗ 0(ω n )(v) = 〈dg(dγ 0 (v)), N〉 = 0 ∀v ∈ T (p,l) G C n,r(T ∂Ω 0 ).<br />

γ0(ι ∗ dγ∂t dL r ) = |〈 ∂φ<br />

∂t , N〉|γ∗ 0(ι P dL r ).<br />

Finally, using that γ0(ι ∗ P dL r ) = |K(l)|dL C r[x]<br />

dx, we get the result. □<br />

Remark 3.11. In real space forms it is known<br />

∫<br />

G n−1,r<br />

σ r (II| V )dV = vol(G(n − 1, r))σ r (II),<br />

but in complex space forms we do not know how to evaluate directly<br />

the analogous integral over the complex Grassmanian.<br />

If r = n − 1, then this integral can be easily evaluated in CK n (ɛ).<br />

Corollary 3.12. Let Ω ⊂ CK n (ɛ) be a regular domain, let X be a<br />

smooth vector field on CK n (ɛ), let φ t be the flow associated to X and<br />

let Ω t = φ t (Ω). Then<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L n−1 )dL n−1 = ω 2n−1 ˜B1,0 (Ω).<br />

t=0<br />

L C n−1<br />

Proof. Since there is only one complex hyperplane contained in the<br />

tangent space of a hypersurface, using the following relations, we obtain<br />

the result directly from Proposition 3.10 .<br />

∫<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L n−1 )dL n−1 = 〈∂φ/∂t, N〉 σ 2n−2 (II| V )dV dx<br />

t=0 L C n−1<br />

∂Q 0 G C n−1,n−1<br />

∫<br />

= 〈∂φ/∂t, N〉σ 2n−2 (II| D )dx<br />

∂Q 0<br />

=<br />

∫∂Q 0<br />

〈∂φ/∂t, N〉 β ∧ θn−1 0<br />

(n − 1)!<br />

= c−1 n,1,0<br />

(n − 1)! ˜B 1,0 = ω 2n−1 ˜B1,0 .<br />


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 13<br />

4. Crofton type formulas<br />

4.1. In the standard Hermitian space.<br />

Theorem 4.1. Let Ω ⊂ C n , let X be a smooth vector field over C n ,<br />

let φ t be the flow associated to X and let Ω t = φ t (Ω). Then<br />

∫<br />

( )<br />

d<br />

−1 ( ) n − 1 n<br />

dt∣ χ(Ω t ∩ L r )dL r = vol(G C n−1,r)ω 2r+1 (r + 1)<br />

−1·<br />

t=0 L r r<br />

C r<br />

⎛<br />

⎞<br />

n−r−1<br />

∑<br />

( )<br />

· ⎝<br />

2q − 1 1<br />

(15)<br />

q 4 ˜B q−1 2n−2r−1,n−r−q (Ω) ⎠<br />

and<br />

∫<br />

L C r<br />

(16)<br />

q=max{0,n−2r}<br />

( ) −1 ( n − 1 n<br />

χ(Ω ∩ L r )dL r = vol(G C n−1,r)ω 2r<br />

r r<br />

⎛<br />

· ⎝<br />

∑n−r<br />

q=max{0,n−2r}<br />

) −1·<br />

1<br />

4 n−r−q ( 2n − 2r − 2q<br />

n − r − q<br />

⎞<br />

)<br />

µ 2n−2r,q (Ω) ⎠ .<br />

Proof. In order to simplify the following computations, we consider<br />

(17) B ′ Ω<br />

k,q (Ω) = c −1<br />

n,k,q BΩ k,q(Ω),<br />

and<br />

(18) ˜B′ k,q = c −1<br />

n,k,q ˜B k,q ,<br />

Γ ′ Ω<br />

k,q (Ω) = 2c −1<br />

n,k,q ΓΩ k,q(Ω)<br />

˜Γ′ k,q = 2c −1<br />

n,k,q ˜Γ k,q .<br />

The expression ∫ χ(L<br />

L C r ∩ Ω)dL r is a valuation on C n with degree<br />

r<br />

of homogeneity 2n − 2r. Thus, it can be expressed as a linear combination<br />

of the elements of a basis of valuations with the same degree of<br />

homogeneity. Then, by Remark 2.9 and (11), we have<br />

∫<br />

n−r−1<br />

∑<br />

(19) χ(Ω ∩ L r )dL r =<br />

C q B 2n−2r,q ′ + DΓ ′ 2n−2r,n−r<br />

L C r<br />

q=max{0,n−2r}<br />

for certain constants C q , D which we wish to determine. This will be<br />

done by comparing the variation of both sides of this equality. From<br />

here on we assume 2r < n. The case 2r ≥ n can be treated in the same<br />

way (cf. Remark 4.2).<br />

By Proposition 3.8, the variation of the right hand side of (19) is a<br />

linear combination of the following type<br />

(20)<br />

n−r−1<br />

∑<br />

q=n−2r−1<br />

c q ˜B′ 2n−2r−1,q +<br />

n−r−1<br />

∑<br />

q=n−2r<br />

d q ˜Γ′ 2n−2r−1,q<br />

where the coefficients c q and d q can be expressed in terms of a linear<br />

combination with known coefficients of the variables C q and D, that<br />

still remain unknown.


14 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

The variation of the left hand side of (19), by Proposition 3.10 is<br />

∫<br />

∫<br />

∫<br />

d<br />

(21)<br />

dt∣ χ(Ω t ∩ L r )dL r = 〈∂φ/∂t, N〉 σ 2r (II| V )dV dx.<br />

t=0<br />

L C r<br />

∂Ω<br />

G C n−1,r<br />

After fixing a frame, the integral ∫ σ<br />

G C 2r (II| V )dV is a polynomial<br />

n−1,r<br />

of the second fundamental form II restricted to the distribution D =<br />

〈N, JN〉 ⊥ . When pulling-back the form γ k,q from N(Ω) to ∂Ω, one<br />

gets a polynomial expression P k,q of degree k in the coefficients of II.<br />

Each of the monomials of P k,q is a minor k × k of II containing the<br />

entry II(JN, JN). For q ≠ q ′ , all the monomials of P k,q are different<br />

from those of P k,q ′. Therefore, every linear combination of {P k,q } q<br />

must contain the variable II(JN, JN). Since JN /∈ D, comparing<br />

the expressions of (20) and (21), it follows that d q = 0 for all q ∈<br />

{n − 2r, . . . , n − r − 1}.<br />

As c q and d q depend on C q and D, we will obtain the value of c q<br />

once we know the value of C q and D. We will get their value from the<br />

equalities {d q = 0}. Note that this gives r equations, since q runs from<br />

n − 2r to n − r − 1 in (20). As for the unknowns, we need to find r<br />

constants C q plus the constant D in (19).<br />

We will get an extra equation by taking II| D = Id and equating (21)<br />

to (20). Then, for any pair (n, r) we have a compatible linear system<br />

since constants in (19) exist. Next we find the solution, which turns<br />

out to be unique.<br />

Let us relate explicitly the coefficients {c q } and {d q } in (20) with<br />

C q and D in (19). To simplify the range of the subscripts, we denote<br />

d n−r−a with a = 1, . . . , r and c n−r−a with a = 1, . . . , r + 1.<br />

Coefficient d n−r−1 .From the variation of B ′ k,q in Cn (Proposition 3.7),<br />

the coefficient of ˜Γ ′ 2n−2r−1,n−r−1 comes from the variation of B ′ 2n−2r,n−r−1<br />

and Γ ′ 2n−2r,n−r. Then,<br />

(22)<br />

d n−r−1 = −2r(n − r)D + (2n − 2r − 2(n − r − 1)) 2 C n−r−1<br />

= 4C n−r−1 − 2r(n − r)D.<br />

Coefficient d n−r−a , a = 2, ..., r. The coefficient of ˜Γ ′ 2n−2r−1,n−r−a<br />

comes from the variation of B 2n−2r,n−r−a ′ and B 2n−2r,n−r−a+1. ′ Then,<br />

(23)<br />

d n−r−a = (2n − 2r − 2(n − r − a)) 2 C n−r−a<br />

− (2r + n − r − a + 1 − n)(n − r − a + 1)C n−r−a+1<br />

= 4a 2 C n−r−a − (r − a + 1)(n − r − a + 1)C n−r−a+1 .<br />

Coefficient c n−r−1 . The coefficient of ˜B ′ 2n−2r−1,n−r−1 comes from the<br />

variation of B ′ 2n−2r,n−r−1 and Γ ′ 2n−2r,n−r. Then,<br />

(24)<br />

c n−r−1 = 4(r + 1/2)(n − r)D − 4C n−r−1<br />

= 2(2r + 1)(n − r)D − 4C n−r−1 .


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 15<br />

Coefficient c n−r−a , a = 2, ..., r − 2. The coefficient of ˜B ′ 2n−2r−1,n−r−a<br />

comes from the variation of B ′ 2n−2r,n−r−a and B ′ 2n−2r,n−r−a+1. Then,<br />

c n−r−a = −2(2a)(2a − 1)C n−r−a + 2(r − a + 3/2)(n − r − a + 1)C n−r−a+1<br />

(25)<br />

= −4a(2a − 1)C n−r−a + (2r − 2a + 3)(n − r − a + 1)C n−r−a+1 .<br />

Coefficient c n−2r−1 . The coefficient of ˜B′ 2n−2r−1,n−2r−1 comes from<br />

the variation of B 2n−2r,n−2r. ′ Then,<br />

(26)<br />

c n−2r−1 = (2r − 2(r + 1) + 3)(n − r − (r + 1) + 1)C n−2r<br />

= (n − 2r)C n−2r .<br />

Now, we solve the linear system given by {d n−r−a = 0} where a ∈<br />

{1, ..., r}. From equations (22) and (23) we have that the system is<br />

given by:<br />

{<br />

r(n − r)D = 2Cn−r−1<br />

4a 2 C n−r−a = (n − r − a + 1)(r − a + 1)C n−r−a+1 .<br />

Thus,<br />

(27)<br />

(n − r − a + 1) · ... · (n − r)(r − a + 1)... · r<br />

C n−r−a = D<br />

2 · 4 a−1 a 2 (a − 1) 2 · ... · 1 2<br />

(n − r)!r!<br />

=<br />

2 2a−1 (n − r − a)!(r − a)!a!a! D<br />

= D ( )( n − r r<br />

.<br />

2 2a−1 a a)<br />

To obtain the value of D, we calculate ∫ σ<br />

G C 2r (p)dV and β 2n−2r−1,n−r−a<br />

′ n−1,r<br />

with II| D (p) = Id. On the one hand, we have<br />

∫<br />

σ 2r (p)(Id| V )dV = vol(G C n−1,r).<br />

G C n−1,r<br />

On the other hand, if II| D = Id, then the connection forms satisfy<br />

α 1i = ω i and β 1i = ω i . Thus, θ 1 = 2θ 2 and θ 0 = θ 2 and we obtain<br />

So, the equation<br />

must be satisfied.<br />

β ′ 2n−2r−1,n−r−a(p) = (β ∧ θ r−a+1<br />

0 ∧ θ 2a−2<br />

1 ∧ θ n−r−a<br />

2 )(p)<br />

= (β ∧ θ n−1<br />

2 )(p) = 2 2a−2 (n − 1)!.<br />

∑r+1<br />

vol(G C n−1,r) = c n−r−a 2 2a−2 (n − 1)!<br />

a=1


16 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

Equations in (24), (25) and (26) give us the relation among c n−r−a ,<br />

D and C n−r−a . Using these relations we have<br />

vol(G C n−1,r)<br />

= (2(2r + 1)(n − r)D − 4C n−r−1 )<br />

(n − 1)!<br />

r∑<br />

+ 2 2a−2 ((2r − 2a + 3)(n − r + a + 1)C n−r−a+1 − 4a(2a − 1)C n−r−a )<br />

a=2<br />

+ 2 2r (n − 2r)C n−2r<br />

= 2(2r + 1)(n − r)D + 4C n−r−1 ((2r − 1)(n − r − 1) − 1)<br />

∑<br />

(−2 2a−2 4a(2a − 1) + 2 2a (2r − 2a + 1)(n − r − a))C n−r−a<br />

r−1<br />

+<br />

a=2<br />

+ C n−2r (2 2r (n − 2r) − 2 2r−2 4r(2r − 1))<br />

r∑<br />

= 2(2r + 1)(n − r)D + 2 2a C n−r−a ((2r − 2a + 1)(n − r − a) − a(2a − 1))<br />

a=1<br />

(<br />

)<br />

(27)<br />

r∑ (2r − 2a + 1)(n − r − a) − a(2a − 1)<br />

= D 2(2r + 1)(n − r) + 2(n − r)!r!<br />

a!a!<br />

(n − r − a)!(r − a)!<br />

a=1<br />

(<br />

)<br />

n! − r!(n − r)!(2r + 1)<br />

= D 2(2r + 1)(n − r) + 2(n − r)!r!<br />

r!r!(n − r)!(n − r − 1)!<br />

2 n!<br />

= D<br />

r!(n − r − 1)! .<br />

Thus,<br />

D = vol(GC n−1,r)<br />

2 n!<br />

C n−r−a = vol(GC n−1,r)<br />

4 a n!<br />

( ) −1 n − 1<br />

,<br />

r<br />

( ) −1 ( )( n − 1 n − r r<br />

r a a)<br />

and, for 2r < n, we have<br />

∫<br />

r∑<br />

χ(Ω ∩ L r )dL r = C n−r−a B 2n−2r,n−r−a ′ + DΓ ′ 2n−2r,n−r<br />

L C r<br />

a=1<br />

= vol(GC n−1,r ) ( ) (<br />

n − 1<br />

−1 r∑ ( )( )<br />

n − r r<br />

2<br />

2 n! r<br />

a a)<br />

−2a+1 B 2n−2r,n−r−a+ ′ Γ ′ 2n−2r,n−r<br />

a=1<br />

and<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L r )dL r = (2(2r + 1)(n − r)D − 4C n−r−1 )B 2n−2r−1,n−r−1<br />

′<br />

t=0 L C r<br />

r∑<br />

+ ((2r − 2a + 3)(n − r + a + 1)C n−r−a+1 −4a(2a − 1)C n−r−a )B 2n−2r−1,n−r−a<br />

′<br />

a=2<br />

+(n − 2r)C n−2r B 2n−2r−1,n−2r−1<br />

′<br />

= vol(GC n−1,r ) ( ) (<br />

n − 1<br />

−1 ∑r+1<br />

( n − r<br />

n! r<br />

a<br />

a=1<br />

)( r + 1<br />

a<br />

) )<br />

a<br />

4 ˜B ′ a−1 2n−2r−1,n−r−a .


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 17<br />

Finally, we use the relation in (17) and (11) to obtain the result.<br />

Remark 4.2. If 2r ≥ n, then formula (15) follows directly from the<br />

relations among the different bases of valuations on C n given in [BF08]<br />

and the following relation in [Ale03]<br />

∫<br />

χ(Ω ∩ L r )dL r = 1 M 2r−3 (∂Ω ∩ L r )dL r = cU 2(n−r),n−r<br />

L O C r<br />

2r−1<br />

∫L C r<br />

for a certain constant c coming from the different normalizations in<br />

dL r .<br />

4.2. In CK n (ɛ).<br />

Corollary 4.3. Let Ω ⊂ CK n (ɛ), let X be a smooth vector field over<br />

CK n (ɛ), let φ t be the flow associated to X and let Ω t = φ t (Ω). Then<br />

∫<br />

( )<br />

d<br />

−1 ( ) n − 1 n<br />

dt∣ χ(Ω t ∩L r )dL r = vol(G C n−1,r)ω 2r+1 (r + 1)<br />

−1·<br />

t=0 L r r<br />

C r<br />

⎛<br />

⎞<br />

n−r−1<br />

∑<br />

( )<br />

· ⎝<br />

2q − 1 1<br />

(28)<br />

q 4 ˜B q−1 2n−2r−1,n−r−q (Ω) ⎠ .<br />

q=max{0,n−2r−1}<br />

Proof. Comparing equation (15) and Proposition 3.10 in case ɛ = 0<br />

shows that<br />

∫ ( ∫ )<br />

〈X, N〉 σ 2r (II|V )dV dx<br />

∂Ω<br />

G C n,r<br />

equals the right hand side of equation above. By taking a field X that<br />

vanishes outside an arbitrarily small neigborhood of any x ∈ ∂Ω, we<br />

deduce the following equality between forms<br />

( ∫ )<br />

σ 2r (II|V )dV dx = ω 2r+1<br />

)( n<br />

G C n,r<br />

r)·<br />

·<br />

n−r−1<br />

∑<br />

q=max{0,n−2r−1}<br />

( 2q − 1<br />

q<br />

( n−1<br />

r<br />

)<br />

cn,2n−2r−1,n−r−q<br />

β ∧ θ r−q+1<br />

4 q−1 0 ∧ θ 2q−2<br />

1 ∧ θ n−r−q<br />

2<br />

This equation extends obviously to CK n (ɛ) without change.<br />

using Proposition 3.10 gives the result.<br />

Theorem 4.4. Let Ω be a regular domain in CK n (ɛ). Then<br />

∫<br />

( ) −1 n − 1<br />

χ(Ω ∩ L r )dL r = vol(G C n−1,r) (ɛ r (r + 1)vol(Ω)+<br />

r<br />

L C r<br />

+<br />

∑n−1<br />

k=n−r<br />

+<br />

ɛ k−(n−r) ω 2n−2k<br />

( n<br />

k) −1<br />

· ((k + r − n + 1)µ 2k,k +<br />

∑k−1<br />

q=max{0,2k−n}<br />

( )<br />

1 2k − 2q<br />

µ<br />

4 k−q 2k,q )).<br />

k − q<br />

□<br />

Then,<br />


18 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

Proof. First, we write the formula to be proved in terms of {B k,q ′ } and<br />

{Γ ′ k,q } defined in (17):<br />

∫<br />

χ(Ω ∩ L r )dL r = vol(GC n−1,r ) ( ) n − 1 −1<br />

(ɛ r (r + 1)n!vol(Ω)+<br />

n! r<br />

+<br />

L C r<br />

(29)<br />

n−1<br />

∑<br />

k=n−r<br />

ɛ k−(n−r) ⎛<br />

⎝<br />

∑k−1<br />

q=max{0,2k−n}<br />

( )<br />

1 n − k k<br />

4 k−q B 2k,q<br />

k − q)( ′ q<br />

+ k + r − n + 1<br />

2<br />

Γ ′ 2k,k<br />

We calculate the variation of both sides of the equation (29) to prove<br />

that they coincide. The variation of the left hand side is given in<br />

Corollary 4.3. To calculate the variation of the right hand side we use<br />

Proposition 3.8 which we recall in the following table. We denote by<br />

(δ X B k,q ′ , ˜B r,s) ′ the coefficient of ˜B r,s in the expression of δ X B k,q ′ .<br />

⎧<br />

2q(n + q − k + 1/2), if r = k − 1, s = q − 1<br />

⎪⎨<br />

(δ X B<br />

k,q ′ , ˜B<br />

−2(k − 2q)(k − 2q − 1), if r = k − 1, s = q<br />

r,s)=<br />

′ 2ɛ(k − 2q)(k − 2q − 1), if r = k + 1, s = q + 1<br />

−2ɛ(n − k + q)(q + 1/2), if r = k + 1, s = q<br />

⎪⎩<br />

0, otherwise.<br />

⎧<br />

(δ X B<br />

k,q ′ , ˜Γ<br />

⎨ (k − 2q) 2 , if r = k − 1, s = q<br />

′<br />

r,s)= −(n + q − k)q, if r = k − 1, s = q − 1<br />

⎩<br />

0, otherwise.<br />

⎧<br />

4q(n − q + 1/2), if r = 2q − 1, s = q − 1<br />

⎪⎨<br />

(δ X Γ ′ 2q,q , ˜B r,s)=<br />

′ 4ɛ((q + 1)/2 − (n − q)(2q + 3/2)), if r = 2q + 1, s = q<br />

4ɛ ⎪⎩<br />

2 (n − q − 1)(q + 3/2), if r = 2q + 3, s = q + 1<br />

0, otherwise.<br />

⎧<br />

(δ X Γ ′ 2q,q , ˜Γ<br />

⎨ −2(n − q)q, if r = 2q − 1, s = q − 1<br />

′<br />

r,s)= 2ɛ(n − q − 1)(q + 1), if r = 2q + 1, s = q<br />

⎩<br />

0, otherwise.<br />

With these expressions, one can check that the variation of the right<br />

hand side of (29) coincides with the right hand side of (28).<br />

Finally, we take a deformation Ω t of Ω such that Ω t converges to a<br />

point. Since both sides of (29) vanish in the limit, the result follows. □<br />

5. Gauss-Bonnet theorem<br />

Theorem 5.1. Let Ω be a regular domain in CK n (ɛ). Then<br />

O 2n−1 χ(Ω) = 2n(n + 1)ɛ n vol(Ω) +<br />

⎛<br />

⎞<br />

∑n−1<br />

O 2n−2c−1 ɛ c ∑c−1<br />

( )<br />

+ ) ⎝<br />

1 2c − 2q<br />

(30)<br />

µ<br />

4 c−q 2c,q +(c + 1)µ 2c,c<br />

⎠.<br />

c − q<br />

c=0<br />

( n−1<br />

c<br />

q=max{0,2c−n}<br />

⎞<br />

⎠).


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 19<br />

Proof. We proceed analogously to the proof of Theorem 4.4. In fact,<br />

the same computations of the previous proof show (in case r = n) that<br />

the right hand side of (30) has null variation.<br />

If ɛ = 0, then we get the known Gauss-Bonnet formula in C n . Taking<br />

some smooth deformation of Ω to get a domain contained in a ball<br />

of arbitrarily small radius, we have that the equality is true for all<br />

ɛ ∈ R.<br />

□<br />

Theorem 5.2. Let Ω be a regular domain in CK n (ɛ). Then<br />

∫<br />

M 2n−1 (∂Ω) = O 2n−1 χ(Ω) − 2nɛ χ(∂Ω ∩ L n−1 )dL n−1 −<br />

k=1<br />

L C n−1<br />

∑n−1<br />

( ) −1 n − 1<br />

− ɛ k O 2n−2k−1 µ 2k,k − 2nɛ n vol(Ω).<br />

k<br />

Proof. First, we use the following relation between M 2n−1 (∂Ω) and<br />

µ 0,0 (Ω)<br />

∫<br />

∫<br />

γ ∧ θ0<br />

n−1<br />

M 2n−1 (∂Ω) = σ 2n−1 (II x )dx =<br />

∂Ω<br />

∂Ω (n − 1)!<br />

∫<br />

= 2c−1 n,0,0 c n,0,0<br />

γ ∧ θ0 n−1 = 2c−1 n,0,0<br />

(n − 1)! 2 ∂Ω<br />

(n − 1)! µ 0,0(Ω)<br />

(31)<br />

= 2nω 2n µ 0,0 (Ω) = O 2n−1 µ 0,0 (Ω).<br />

Now, from Theorems 4.4 and 5.1, it follows that<br />

⎛<br />

⎞<br />

n−1<br />

∑ ɛ c n−1<br />

c! ∑<br />

( )<br />

χ(Ω) = ⎝<br />

1 2c − 2q<br />

π c<br />

4 c−q µ 2c,q + (c + 1)µ 2k,k<br />

⎠<br />

c − q<br />

c=0<br />

+ ɛn (n + 1)!<br />

π n<br />

n−1<br />

∑<br />

= µ 0,0 +<br />

c=1<br />

q=max{0,2c−n}<br />

vol(Ω)<br />

⎛<br />

ɛ c c!<br />

π c<br />

⎝<br />

+ ɛn (n + 1)!<br />

π n vol(Ω)<br />

= µ 0,0 + ɛ n! n−1<br />

∑<br />

π n<br />

+ ɛ n! n−1<br />

∑<br />

π n<br />

c=1<br />

c=1<br />

= µ 0,0 + ɛ n! ∫<br />

π n<br />

( ɛ n (n + 1)!<br />

+<br />

π n<br />

n−1<br />

∑<br />

q=max{0,2c−n}<br />

⎛<br />

ɛ c−1 c!π n−c<br />

⎝<br />

n!<br />

ɛ c−1 c!π n−c<br />

n!<br />

⎞<br />

( )<br />

1 2c − 2q<br />

4 c−q µ 2c,q + (c + 1)µ 2c,c<br />

⎠<br />

c − q<br />

n−1<br />

∑<br />

q=max{0,2c−n}<br />

µ 2c,c + ɛn (n + 1)!<br />

π n vol(Ω)<br />

n−1<br />

∑<br />

χ(∂Ω ∩ L n−1 )dL n−1 +<br />

L n−1<br />

− ɛn n!n<br />

π n )<br />

vol(Ω).<br />

⎞<br />

( )<br />

1 2c − 2q<br />

4 c−q µ 2c,q + cµ 2c,c<br />

⎠<br />

c − q<br />

c=1<br />

ɛ c c!<br />

π c µ 2c,c


20 JUDIT ABARDIA, EDUARDO GALLEGO, GIL SOLANES<br />

Remark 5.3. For n = 2 and n = 3, the Gauss-Bonnet-Chern formula<br />

in CK n (ɛ) given in Theorem 5.1 was already stated in [Par02].<br />

6. Total Gauss curvature integral<br />

Theorem 6.1. Let Ω be a regular domain in C n . Then<br />

∫<br />

( ) −1 ( ) n − 1 n<br />

M 2r−1 (∂Ω∩L r )dL r = 2rω2rvol(G 2 C n−1,r)<br />

−1·<br />

L r r<br />

C r<br />

⎛<br />

⎞<br />

∑n−r<br />

( )<br />

· ⎝<br />

1 2n − 2r − 2q<br />

µ<br />

4 n−r−q 2n−2r,q<br />

⎠ .<br />

n − r − q<br />

q=max{0,n−2r}<br />

Proof. On the one hand, by Gauss-Bonnet formula in C n , we have<br />

∫<br />

∫<br />

χ(Ω ∩ L r )dL r = µ 0,0 (Ω ∩ L r )dL r .<br />

L C r<br />

On the other hand, by Theorem 4.4 in C n , we have<br />

∫<br />

( ) −1 ( ) −1<br />

n − 1 n<br />

χ(∂Ω ∩ L r )dL r = vol(G C n−1,r) ω 2r<br />

L r n − r<br />

C r<br />

⎛<br />

⎞<br />

∑n−r<br />

( )<br />

· ⎝<br />

1 2n − 2r − 2q<br />

µ<br />

4 n−r−q 2n−2r,q<br />

⎠ .<br />

n − r − q<br />

q=max{0,n−2r}<br />

If we equate both expressions and we use the relation (31) between<br />

the total Gauss curvature and the valuation µ 0,0 , we obtain the result.<br />

□<br />

Remark 6.2. The previous Theorem is not necessarily true in CK n (ε)<br />

for ɛ ≠ 0. Indeed, Howard’s transfer principle can not be used here since<br />

Theorem 6.1 is not local: it does not apply to general hypersurfaces,<br />

but only to the closed embedded ones.<br />

L C r<br />

□<br />

[Aba09]<br />

[Ale03]<br />

[BF08]<br />

[Bla76]<br />

[Gol99]<br />

References<br />

J. Abardia. Average of mean curvature integral in complex space forms.<br />

Preprints de la Universitat Autònoma de Barcelona, 2009.<br />

S. Alesker. Hard Lefschetz theorem for valuations, complex integral geometry,<br />

and unitarily invariant valuations. J. Differential Geom., 63(1):63–<br />

95, 2003.<br />

A. Bernig and J. H. Fu. Hermitian integral geometry. arXiv:0801.0711v5,<br />

2008.<br />

David E. Blair. Contact manifolds in Riemannian geometry. Springer-<br />

Verlag, Berlin, 1976. Lecture Notes in Mathematics, Vol. 509.<br />

W. M. Goldman. Complex hyperbolic geometry. Oxford Mathematical<br />

Monographs. The Clarendon Press Oxford University Press, New York,<br />

1999. Oxford Science Publications.


<strong>GAUSS</strong>-<strong>BONNET</strong> <strong>THEOREM</strong> <strong>IN</strong> COMPLEX SPACE FORMS 21<br />

[Had57] H. Hadwiger. Vorlesungen über Inhalt, Oberfläche und Isoperimetrie.<br />

Springer-Verlag, Berlin, 1957.<br />

[Nav05] A. M. Naveira. Two problems in real and complex integral geometry. In<br />

Complex, contact and symmetric manifolds, volume 234 of Progr. Math.,<br />

pages 209–220. Birkhäuser Boston, Boston, MA, 2005.<br />

[O’N83] B. O’Neill. Semi-Riemannian geometry, volume 103 of Pure and Applied<br />

Mathematics. Academic Press Inc. [Harcourt Brace Jovanovich Publishers],<br />

New York, 1983. With applications to relativity.<br />

[Par02] Heunggi Park. Kinematic formulas for the real subspaces of complex space<br />

forms of dimension 2 and 3. PhD-thesis. University of Georgia, 2002.<br />

[Rum94] M. Rumin. Formes différentielles sur les variétés de contact. J. Differential<br />

Geom., 39(2):281–330, 1994.<br />

[San52] L. A. Santaló. Integral geometry in Hermitian spaces. Amer. J. Math.,<br />

74:423–434, 1952.<br />

[San04] L. A. Santaló. Integral geometry and geometric probability. Cambridge<br />

Mathematical Library. Cambridge University Press, Cambridge, second<br />

edition, 2004. With a foreword by Mark Kac.<br />

[Sol06] G. Solanes. Integral geometry and the Gauss-Bonnet theorem in constant<br />

curvature spaces. Trans. Amer. Math. Soc., 358(3):1105–1115 (electronic),<br />

2006.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!