Open access peer-reviewed chapter

An Introduction to the Generalized Gauss-Bonnet-Chern Theorem

Written By

Paul Bracken

Reviewed: 06 June 2022 Published: 11 July 2022

DOI: 10.5772/intechopen.105716

From the Edited Volume

Manifolds III - Developments and Applications

Edited by Paul Bracken

Chapter metrics overview

111 Chapter Downloads

View Full Metrics

Abstract

This work studies the mathematical structures which are relevant to differentiable manifolds needed to prove the Gauss-Bonnet-Chern theorem. These structures include de Rham cohomology vector spaces of the manifold, characteristic classes such as the Euler class, pfaffians, and some fiber bundles with useful properties. The paper presents a unified approach that makes use of fiber bundles and leads to a non-computational proof of the Gauss-Bonnet-Chern Theorem. It is indicated how it can be generalized to manifolds with boundary.

Keywords

  • manifold
  • Euler characteristic
  • bundle
  • fiber
  • projection
  • universal

1. Introduction

One of the great achievements of differential geometry is the Gauss-Bonnet theorem. In its original form, the theorem is a statement about surfaces which connect their geometry in the sense of curvature to the underlying topology of the space, in the sense of the Euler characteristic [1, 2, 3, 4, 5]. The most elementary case of the theorem states that the sum of the angles of a triangle in the plane is π radians. If the surface is deformed, the Euler characteristic does not vary as it is a topological invariant, while the curvature at certain points does change [6, 7, 8]. The theorem states that the total integral of the curvature remains the same, no matter how the deformation is performed. If there is a sphere with a ding, its total curvature is 4π since its Euler characteristic is two. This is the case no matter how big or deep the actual deformation is. A torus has Euler characteristic zero, so its total curvature must also be zero. If the torus carries the usual Riemannian metric from its embedding in 3, then the inside has negative Gaussian curvature, and the outside has positive Gaussian curvature, so the total curvature is zero. It is not possible to specify a Riemannian metric on the torus which has everywhere positive or everywhere negative Gaussian curvature. Manifolds M have dimension n unless stated otherwise [9, 10, 11]. There are many applications of this theorem in both mathematics and mathematical physics such as in gravity [12, 13, 14], string theory [15] and even in the study of Ricci flow [16].

Although the curvature K is defined intrinsically in terms of the metric on the manifold M. It can also be defined for n=2 extrinsically when the metric on M is induced by an embedding M3. In fact, it ν:MS2 is the normal map and da is the volume element on S2, then K=νda so that

MK=Mνda=degνS2da=4πdegν.E1

Without bringing in differential geometric considerations, it is seen to be the case that degν=1/2χM, where χM is the Euler characteristic of M. Using this fact in (1), the Gauss-Bonnet theorem for a compact oriented surface M, the first version of the theorem is obtained for the case in which the metric on M arises by means of an embedding in 3

MK=2πχM.E2

It is the intention here to state and prove a general version of the theorem which applies to manifolds of even dimension, so a surface with n=2 is a special case. An intrinsic proof of the theorem was obtained by Chern 1944. The kind of argument outlined above was used by Hopf in developing the first generalization of the theorem. To outline the basic idea, consider a compact surface Mnn+1 when n is even. If dμg is the volume form on the manifold and dsn denotes the volume element Sn, then

MnKndμg=Mnνdsn=volSndegν=12volSnχMn.E3

This can be extended to any compact oriented Riemannian n-manifold Mng which has even dimension, where Kn in a coordinate system is given by

Kn=12n/2n!i1,,inj1,,jnRi1i2j1j2Rin1injn1jn1gεi1in1gεj1jn.E4

The g(4) is the square root of the determinant of the metric. With Kn given by (4), and μg the volume form on the manifold, we are then led to conjecture that

MnKndμg=12volSnχMn,E5

where M is a compact, oriented Riemannian manifold with n even.

It is the objective to look at and study some of the ensuing developments which have led to a much deeper understanding of the foundations which underlie this theorem. It will be seen that this development leads to a completely non-computational proof of this deep theorem.

Advertisement

2. Characteristic classes

When an oriented n-dimensional manifold Mdμg is compact and closed, with dμg is the volume form and μ the orientation of M, so every form has compact support, Stokes theorem leads to the important theorem. Let η be any n1-form then

M=Mη.E6

Therefore an n-form ω on M, which is not exact, even though it must be closed as all n-forms on M are zero, can be found simply by locating an ω such that

Mω0.E7

Such a form always exists, as it is known there is a form ω such that for v1,,vnMp, ωv1vn>0 if v1vn=μp. If c:01nMμ preserves orientation, cω on 01n is gdx1dxn for some g>0 on 01n, hence cω>0. This observation leads to this theorem. A smooth, oriented manifold is not smoothly contractible to a point. In fact, it is the shape of M not the size which determines whether or not every closed form on M is exact. More information about the shape of M can be obtained by analyzing more closely the extent to which closed forms are not exact. So how many non-exact n-forms are there on a compact oriented n-manifold If ω is not exact, the same holds for ω+ for η any n1-form η. Thus it is necessary to regard ω and ω+ as equivalent. This suggests an equivalence relation and directs one to think of this in terms of quotient spaces.

For each k, ZkM denotes all closed k-forms on M and it is a vector space. The space BkM of all exact k-forms is a subspace since d2=0. The quotient space is called the k-dimensional de Rham cohomology vector space of M and is defined to be

HkM=ZkM/BkM.E8

The theorem of de Rham states that the vector space is isomorphic to a vector space defined just in terms of the topology of M called the k-dimensional cohomology group of M with real coefficients.

An element of HkM is an equivalence class ω of a closed form ω such that closed forms ω1 and ω2 are equivalent if and only if the difference is exact. In terms of these vector spaces, the Poincaré lemma gives Hkn=0, the vector space consisting of just the zero vector if k>0, or HkM=0 if M is contractible and k>0. To compute H0M note B0M=0 as there are no non-zero exact 0-forms as there are no non-zero minus one forms. Thus H0M is the same as the vector space of all C functions f:M with df=0. If M is connected, this condition implies f is constant so H0M and its dimension is the number of components of M.

The de Rham cohomolgy vector spaces with compact support HckM are defined similarly to (8), that is, HckM=ZckM/BckM, where ZckM is the vector space of all closed k-forms with compact support and BckM all k-forms where η is a k-form with compact support. If M is compact HckM=HkM.

Theorem 2.1. (The Poincaré-Duality Theorem) If M is a connected, oriented n-manfold of finite type, then the map

Π:HkMHcnkME9

is an isomorphism for all k.

This theorem eventually motivates the introduction of the Euler characteristic for any smooth connected oriented manifold M. Consider then a smooth k-dimensional vector bundle ξ=π:EM over M. Orientations μ for M, and ν for ξ give an orientation μν for the n+k-manifold E, since E is locally a product. Let U1Ur be a cover of M by geodesically convex sets so small that each bundle ξ restricted to Ui is trivial. Then π1U1π1Ur turns out to be a nice cover for E, so it is a manifold of finite type. For the section and projection maps s,π, πs=I on M and sπ is smoothly homotopic to the identity of E, so the map π:HlMHlE is an isomorphism for all l. The reason for mentioning (6) and Theorem 2.1 is that it shows there is a unique class UHckE such that

π:μU=μνHcn+kE.E10

This class is called the Thom class of ξ.

A theorem states that if Mμ is a compact oriented, connected manifold ξ=π:EM an oriented k-plane bundle over M orientation ν, the Thom class U is the unique element of HckE such that for all pM, and jp:FpE the inclusion map, we have jpU=νp. This condition has the implication that Fpνpjpω=1, where U is the class of closed form ω.

The Thom class U of ξ=π:EM can now be used to determine an element of HkM. Let s:ME be any section. There is always one, any two are clearly homotopic. Define the Euler class χEHkM of ξ by

χξ=sU.E11

If ξ has a non-zero section s:ME and ωCckM represents U, a suitable multiple cs of s takes M to the complement of support ω, so in this case, χξ=csU=0.

The term Euler class is connected with the special case of the bundle TM which has sections which are vector fields on M. If X is a vector field on M having an isolated zero at some point p, Xp=0, but Xq0 for qp in a neighborhood of p. An index of X at p can be defined. Suppose X is a vector field on an open set Un with an isolated zero at 0U. Define fX:U0Sn1 by fXp=Xp/Xp. If i:Sn1U is ip=εp mapping Sn1 into U, then the map fXi:Sn1Sn1 has a certain degree independent of ε for small ε, since maps i1,i2:Sn1U correspinding to ε1,ε2 will be smoothly homotopic. This degree is called the index of X at 0. Consider a diffeomorphism h:UVn with h0=0, so hX is the vector field on V such that hXy=hXh1y. So 0 is an isolated zero of hX. It can be shown, if h:UVn is a diffeomorphism with h0=0 and X has an isolated zero at 0, the index of hX at 0 equals the index of X at 0.

As a consequence of this, an index of a vector field on a mainifold can be defined. If X is a vector field on M, with isolated zero at pM, choose a coordinate system xU such that xp=p and define the index of X at p to be the index of xX at 0.

Theorem 2.2. Let M be a compact, connected manifold with orientation μ, also an orientation for the tangent bundle ξ=π:TMM. Let X:MTM be a vector field with only a finite number of zeros and let σ be the sum of indices of X at these zeros. Then

χξ=σμH0M.E12
Advertisement

3. Pfaffians

An intrinisic expression along with one in a coordinate system for the function Kn on a compact, oriented Riemannian manifold of even dimension has been given already. Another more important way of expressing Kn involves the curvature form Ωji for a positively oriented orthonormal moving frame X1,,Xn on M. In terms of these forms, the n-form Kndμg, the one to be integrated, can be written down. A sum over permutations such as

πSnBXπ1Xπn,E13

can be written just as well as

j1,,jnεj1,,jnBXj1Xjn.E14

Suppose this is the n-fold wedge product

Ωi2i1Ωinin1.E15

Since the Ωji are 2-forms, using the definition of wedge product,

Ωi2i1Ωinin1X1Xn=2++2!2!2!1n!j1,jnεi1ipΩi2i1Xj1Xj2Ωinin1Xjn1Xjn=12n/2j1,jnεj1,,jnRXj1Xj2Xj2Xj1RXjn1XjnXjnXjn1=12n/2j1,,jnRi1i2j1j2Rin1injn1jn.E16

Comparing this to (4), it may be concluded that

Kn=12n/2n!i1,,inεi1,,in2n/2Ωi2i1Ωinin1X1Xn.E17

When (17) is multiplied by the volume form dμg, it becomes

Kndμg=1n!i1,,inΩi2i1Ωinin1.E18

By (18) the form on the right does depend on the choice of the positively oriented orthonormal frame, X1,,Xn. There is a direct way to get this algebraically.

Suppose A is an n×n matrix A=aij with n=2m even. Define the Pfaffian, PfA of A to be

PfA=12mm!i1,,inai1i2ain1in.E19

Note that εi1in does not change when any permutation of the pairs i2l1i2l. For any set S=h1k1hmkm of pairs of integers between 1 and n, let us define

εS=εh1k1hmkm.E20

It is not necessary to specify any ordering of pairs hiki in S. Also a permutation of the pairs i2l1i2l does not change the factor ai1i2ain1in. So for each P above define aS=ah1k1ahmkm. If P is the collection of all such S, we clearly have

PfA=12mSPεSaS.E21

Theorem 3.1. Let n=2m then for all n×n matrices A and B,

PfBtAB=detBPfA.E22

and Bt denotes the transpose. If BSOn then

PfB1AB=PfA.E23

Proof:

2mm!PfBtAB=i1inεi1inj1jnbj1i1aj1j2bj2i2bjm1im1ajm1jnbjnin=j1jni1inεj1jnbj1i1bjninaj1j2ajn1jn=j1,,jnεj1,,jndetBaj1j2ajn1jn=2mm!detBPfA.E24

This theorem was stated for matrices of real numbers, but PfA can be defined provided the entries of A are in some commutative algebra over .

Consider again a positively oriented orthogonal moving frame X1,,Xn on M, with curvature forms Ωji. For each pM, the direct sum A=Ω2MpΩ2Mp is a commutative algebra over under the operation . So one can consider PfΩp, where Ωp is an n×n matrix of connection 2-forms at p

PfΩp=12mm!i1,,inεi1,,inΩi1i1Ωinin1p.E25

If X=Xa is another positively oriented orthonormal moving frame then apOn and the corresponding curvature forms satisfy Ω=a1Ωa. Then Theorem 3.1 implies that

PfΩp=Pfa1pΩap=PfΩpE26

so the form i1,,inεi1,,inΩi2i1Ωinin1 is well defined.

Advertisement

4. Bundles of paticular importance

Projective n-space n can be defined as the set of all pairs pp for pSnn+1 or the set of line through 0 in n+1, since each lines intersects Sn through two anti-podal points. A Grassmannian manifold Gnn is the set of all n-dimensional subspaces of N with N>0. Over the Grassmannian manifold GnN, there is a natural n-dimensional bundle ζnn constructed as follows. The total space of the bundle EζnN is the subset of GnN×N consisting of all pairs

WwGnN×M,wW.E27

The projection map which takes EζnNGnN is πWw=W. The fibre π1W over W of GnN will be W itself or, more explicitly, Ww:wW. A vector space structure is defined on π1W by using the vector space structure on WN; if a is a scalar, then Ww1+Ww2=Ww1+w2 and aWw=Waw. Also ζnN satisfies the local triviality condition.

For M>N there is a natural map α:GnNGnM, as an n-dimensional subspace of N can be considered an n-dimensional subspace of M. There is clearly a map α¯:EζnNEζnM such that α¯α is a bundle map from ζnN to ζnM and thus ζnNαζnM.

In algebraic topology, one often considers the union G0 of the increasing sequence Gnn+1Gnn+1 with weak topology; that is, a set UGn=lGnn+l is open if and only if UGnn+l is open in Gnn+l for all l. There is a natural n-dimensional bundle ζn over Gn defined in a way similar to ζnN such that the following properties are maintained: i for every bundle ξ over a paracompact space X, there is a map f:XGn such that ξfζn. ii if f0,f1:XGn are maps of a paracompact space X into Gn with f0:ζnf1ζn then f1f0.

For this reason ζn is called the universaln-dimensional bundle and Gn, is called the classifying space for n-dimensional bundles since equivalence classes of n-dimensional bundles over X are classified by homotopy classes of maps of X into Gn. Now Gn is not a manifold so we continue to use the bundles ζnN, which are usually called universal bundles.

An orientation for a vector space V is an equivalence class of ordered bases for V where v1vnw1wn if and only if aij defined by wi=jajivj has detaij>0. There are only two such equivalence classes η and η. An oriented vector space is a pair Vη, where η is an orientation for V.

An orientation for a bundle ξ=π:EX is a collection η=ηx of orientations for the fibres π1x which satisfy an obvious compatibility requirement, while an oriented bundle is a pair ξη, where η is an orientation for ξ. Orientation η of ξ gives another η=η if X is connected. This is the only other one for ξ. Define ξ1ξ2μ1μ2 to be the sum ξ1ξ2 with the indicated orientation.

Suppose ξ=π:EM is a smooth oriented n-dimensional vector bundle over a smooth manifold M of any dimension. The Euler class χξHnM was defined by first defining the Thom class UξHcnE. It can be proved Uξ is the unique class whose restriction to each π1p is the generator νp=Hcnπ1p determined by the orientation. This result leads directly into the next theorem.

Theorem 4.1. Let ξ=π:EM be a smooth manifold where M is also a compact manifold. If E is the total space of fξ and f:EE is a bundle map covering f,

fUξ=UfξHcnE.E28

Proof: Note f has the property inverse of a compact set is compact, so f takes HcnE to HcnE. Let fξ be π:EM. If pM is any point, and ip:π'1pE is the inclusion map, then

ipfUξ=fipUξ.E29

Recall how fξ is defined then fipUξ must be the generator of Hcnπ'1p, since ifpUξ is the generator of Hcn(π1fp. This shows fUξ must be Ufξ.

The Euler class χξ was defined as sUξ for any section s of ξ. Suppose s=0 is the zero section, which is chosen. It can be shown that

fχξ=χfξHnM.E30

A consequence of (30) is important as it gives the following.

Theorem 4.2. If n is even, then

χQnN0,N>n.E31

Proof: Since SnN for N>n, we have a bundle map ff:TSnEζnN so

χTSn=fχζnN.E32

However, it is known to be the case that χTSn is χSn times the fundamental class of Sn and χSn=20.

Advertisement

5. A unique one-form constructed from the curvature

This is an important characteristic class which is important and plays a significant role. Consider principal bundles associated with a smooth oriented n-dimensional vector bundle ξ over a smooth manifold M. There is the principal bundle of frames Fξ of E. If ξ has a Riemannian metric , the bundle OE of orthonormal frames can be considered, which is a principal bundle with group On. Since only paracompact M are considered here, there is an Ehresmann connection ω on OE. Thus ω is a matrix of one-forms ωji on OE with values in on, the curvature form Ω= is a matrix of two-forms Ωji with values in on. A connection ω on OE is equivalent to a covariant derivative operator on E compatible with the metric, and for a general ξ over M, there will be many connections compatible with the metric. One can not be singled out by requiring a symmetric connection which only makes sense for the tangent bundle. As ξ is oriented, we can also consider the bundle SOE of positively oriented frames. If X is connected, it is simply one of the two components of OE, with group SOn and Lie algebra on. A connection ω on SOn again has values in on as does the matrix of two-forms Ω.

If we specialize to the case of a smooth oriented n-dimensional vector bundle ξ=π:EM over M, with n=2m even. If is a Riemannian metric for ξ and ω is a connection on the corresponding principal bundle ω¯:SOEM, consider the n-form which is defined on SOE

2mm!PfΩ=i1,,inεi1,,inΩi2i1Ωinin1.E33

The following is an invariant formulation of a previous theorem.

Theorem 5.1. There is a unique n-form Λ on M such that

ω¯Λ=i1,,inεi1,,inΩi2i1Ωinin1=2mm!PfΩ.E34

Proof: Let X1,,XnMpY1,,YnSOEu be tangent vectors such that πYiXi, and choose some uω1p. Then form Λ must satisfy

ΛX1Xn=2mm!PfΩY1YnE35

This suffices to give uniqueness. If it can be shown this Λ in (35) is well-defined, then existence can be established.

Suppose different tangent vectors Z1,,Zn are taken such that ω¯xZi=Xi. Since ω¯xYiZi=0, all YiZi are vertical. However, ΩYZ=0 if either Y or Z is vertical. Consequently,

PfY1Yn=PfΩZ1Y2Yn=PfΩZ1Z2Y3Yn=PfΩZ1Zn.E36

This means the definition of Λ does not depend on the Yi selected. Suppose a different u¯ω¯1p is chosen. Then u¯=RAu=uA for some ASOn, and so let Y¯iSOEπ be given by Y¯i=RAYi and

PfΩY¯1Y¯n=PfΩRAY1RAYn=PfRAΩY1Yn=PfA1ΩAY1Yn=PfΩY1Yn.E37

Theorem 5.2. The unique n-form Λ in (35) is closed, dΛ=0.

Proof: Suppose X1,,Xn+1Mp be given and choose uω¯1p) and Y1,,Yn+1SOEu with ω¯xYi=Xi and hYi the horizontal component of Yi. Then working out dΛ

dΛX1Xn+1=dΛω¯xY1ω¯xYn+1=dΛω¯xhY1ω¯xhYn+1=ω¯dΛhY1hYn=dω¯ΛhY1hYn+1=2mm!dPfΩhY1hYn+1=2mm!DPfΩY1Yn+1.E38

However, DΩ=0 by Bianchi’s identity and a consequence of this is that (38) vanishes.

This result applies automatically when ξ is the tangent bundle. The implication of this is that the n-form Λ determines a de Rham cohomolgy class ΛHnM of M. The form Λ itself depends on the oriented n-dimensional bundle ξ=π:EM over M as well as the choice of metric for ξ and connection ω on the corresponding bundle SOE.

Theorem 5.3. The cohomology class Λ is independent of both the metric and the connection ω.

Proof: Suppose two metrics , are given for ξ. Then the corresponding principal bundles SOE and SOE are equivalent. If f:SOESOE is a fiber preserving diffeomorphism which commutes with the action SOn and ω a connection on SOE. Then ω=fω is a connection on SOE. Corresponding curvature forms satisfy Ω=fPfΩ so PfΩ=fPfΩ. The corresponding forms Λ and Λ are in fact equal. It suffices to show any two connection differential forms ω0, ω1 on the same SOE generate forms Λ0, Λ1 whose difference is exact. If π:M×01M is the projection πpt=p, consider the bundle πSOξ over M×01. Induced connections are πω0 and πω1 on πSOξ. Let τ:M×0101 defined here as τpt=t and define a connection

ω=1τπω0+τπω1E39

on πSOξ with Ω the connection form. If it maps M to M×01 and is defined as itp=pt, then i0ω can be identified with ω0 and i1ω with ω1. By Theorems 5.1 and 5.2, which hold for manifolds with and without boundary, there is a closed n-form Λ on M×01 which pulls back to 2mm!PfΩ on the total space of πSOξ. A theorem states for any k-form ω on M×01, i1ωi0ω=dI. So if =0, this implies i1ωi0ω=d. Substituting the form Λ in place of ω into this, it follows that Λ1Λ0 is exact.

Thus every smooth oriented smaooth bundle ξ over M of even fibre dimension n determines a de Rham cohomology class Cξ=ΛHnM and Cξ=Cη if ξη. It may be asked how does the object Cξ behave with respect to f.

Theorem 5.4. Let ξ=π:ξM be a smooth oriented bundle over M with fibre dimension n even, let f:MM be a smooth map. Then

Cfξ=fCξHnM.E40

Proof: The total space of fξ is called E. Let f:E'E be the bundle map covering f. If is a metric on E, then f is a metric on E. There is an equivalence f¯:SOESOE covering f with ω¯' taking SOE to M and ω¯ mapping SOE to M.

If ω is a connection on SOE, then f¯ω is a connection on SOE. It is seen that the corresponding connection forms satisfy Ω=fΩ. Aa a result, we have

PfΩ=Pff¯Ω=f¯PfΩ.E41

For n-forms Λ on M given by Theorem 5.1, we then have

ω¯'fΛ=f¯ω¯Λ=2mm!f¯PfΩ=2mm!PfΩ.E42

This means fΛ must be the n-form Λ on M given in (31).

When ξ is a smooth oriented bundle of odd fibre dimension, the definition of C may be extended. It would be remarkable if it were the case that Cξ were always a constant multiple of χξ. To this end, the following theorem is needed.

Theorem 5.5. Let ξi=πi:EiM for i=1,2 be smooth oriented vector bundles over M of fibre dimension n1 and n2. If ni=2mi, then

Cξ1ξ2=m1+m2!m1!m2!Cξ1Cξ2.E43

If n1 or n2 is odd, this reduces to Cξ1ξ2=0.

Proof: Pick two metrics which are Riemannian for each ξi and set =12 on ξ1ξ2=π:EM. Let ω¯i:SOEiM and ω¯:SOEM be the corresponding principal bundles. Over M consider the product principal bundle Q=SOE1SOE2 with corresponding group SOn1×SOn2SOn1+n2 whose fiber over pM is the direct product ω¯11×ω¯21p, so this bundle is a subset of SOE.

Let ρi be the projection maps for Q which project this down onto either of its factors. If ωi are connections on SOEi, with curvature forms Ωi, then

ρ1ω1ρ2ω2=ρ1ω100ρ2ω2E44

is a connection on Q and the curvature form is

Ω=ρ1Ω1ρ2Ω2=ρ1Ω100ρ2Ω2E45

The connection ω¯ can be extended uniquely to a connection ω on SOE. The requirement ωσM=M determines ω at the new vertical vectors, hence ω is determined at all points of Q, and then at all points of SOE by the requirement ωRAY=AdA1ωY.

At any point eQ, the horizontal vectors for ω are the same as that for ω¯. At E, it holds that Ω¯=Ω for tangent vectors to Q which implies, using PfAB=PfAPfB, that

PfΩ=PfΩ¯=Pfρ1Ω1Pfρ2Ω2=ρ1PfΩ1ρ2Ω2.E46

Consequently, if Λi are the forms given by (34), then at e it must hold that on tangent vectors to Q

ω¯Λ=2m1+m2m1+m2!PfΩ=m1+m2!m1!m2!2m1m1!ρ1PfΩ12m2m2!ρ2PfΩ2=m1+m2!m1!m2!ρ1ω¯1Λ1ρ2ω¯2Λ2=m1+m2!m1!m2!ω¯1Λ1ω¯2Λ2.E47

This implies that

Λ=m1+m2!m1!m2!Λ1Λ2.E48

Corollary 5.1. If the oriented bundle ξ=π:EM has a nowhere zero section s, then

Cξ=0.E49

Proof: Let E1E be written

pMsp,E50

and let E2E be the orthogonal complement

pMspE51

with respect to some Riemannian metric on E. Then ξ1=π1E1:E1M is an oriented one-dimensional bundle. Consequently, ξ2=π2E2:E2M is also an oriented bundle since ξ is oriented. Clearly ξξ1ξ2. An application of the previous result shows that Cξ=0.

This theorem is almost enough to characterize χ as we can now show the statement which relates Cξ and the Euler class.

Corollary 5.2. If ξ=π:EM is a smooth vector bundle of fibre dimension n over a compact oriented manifold M, then the class CξHnM is a multiple of the Euler class χξ.

Proof: Suppose S is the sphere bundle S=eE:ee=1, which is constructed with respect to some Riemannian metric on E. Let π0:SX be the restriction πS. The bundle π0ξ has a nowhere zero section. Corrollary 5.1 and Theorem 5.4 then yield

π0Cξ=Cπ0ξ=0.E52

However, there is a theorem which states a class αHnM satisfies π0α=0 if and only if α is a multiple of χξ. It can now be inferred that Cξ is a multiple of the Euler class χξ.

Advertisement

6. The Gauss-Bonnet-Chern theorem

If Corollary 5.2 is applied to the tangent bundle of a compact oriented manifold M of dimension n which is even, the class CTMHnM is some multiple of the Euler class χTM. This fact is not so interesting because HnM is one-dimensional since it means CTM=0 if χTM=0. The corollary does lead to something interesting when applied to the universal bundle.

Theorem 6.1 For every even n, there is a constant βn such that

Cξ=βnχξ.E53

for all smooth oriented n-dimensional bundles ξ over compact oriented manifolds. In this sense, it is universal.

Proof: Begin with the bundles zetanN for N>n. Corollary 5.2 implies there are constants βn,N such that

CζnN=βn,NχζnNHnζnN.E54

If j:GnNGnM is the natural inclusion, then jζNNζnN. Equation (30) and Theorem 5.4 yield

CζnN=jCζnM=jχζnM.E55

Thus, (54), (55) give

βn,NχζnN=βn,MχζnN.E56

Since χζnN0 by Theorem 4.2, this implies that βn,N=βn,M for all M,N>1. This common number is called βn, and we have

CζnN=βnχζnN.E57

However it is known that any smooth oriented n-dimensional bundle ξ over a compact manifold M is equivalent to fζnN for some smooth map f:MGN, then

Cξ=CfζnN=fCζnN=βnfχζnN=βnχξ.E58

The constant βN is universal in nature and it may be asked whether it can be computed. Since it has this universality property, it suffices to compute this constant for a special case where the calculation is easier and in turn implies another application of the next theorem.

Theorem 6.2 For integer n=2m, the constant βn in Theorem 6.1 is

βn=n!2VSn=πm2nm!.E59

If M is a compact manifold of even dimension n=2m then

MKndμg=πm2nm!n!χM.E60

Proof: Let ξ be the tangent bundle TM of a compact oriented manifold of dimension n. Now (17) gives a formula for Kndμg, where Ωji are curvature forms for some positively oriented orthonormal moving frame. This implies that the form Λ in Theorem 5.1 for the bundle SOξ=SOTM is

Λ=n!Kndμg.E61

If κ is the fundamental class of M then

MKndμgκ=1n!MΛκ=1n!Cξ=βnn!χξ=βnn!χMκ.E62

Hence, equating the coefficients of κ on both sides,

MKndμg=βnn!χM.E63

Consider the case of a specific manifold M=Sn in (63), where Kn=1 so the left side of (63) reduces to VSn

VSn=βnn!χSn=2βnn!.E64

Since the volume VSn is known to be πm2n+1m!/n!, solve (64) for βn in terms of VSn ,

βn=n!2VSn=πm2nm!.E65

This value of βn can be put back into (60) and for the manifold M, it follows that

MKndμg=πm2nm!n!χM.E66
Advertisement

7. The theorem for manifolds with boundary

Theorem 5.5 played a large part in the proof of (64). It allowed us to state that if ξ=π:EM is an oriented n-dimensional vector bundle with sphere bundle π0:S0M, then π0Cξ=0 was a large part of the proof of (). If Λ is the n-form on M representing Cξ, then the n-form π0Λ on S is exact π0Λ=dΦ for some n1-form on S. Suppose ξ=TM and let X be a unit vector bundle on M which has an isolated singularity at pM Let Bε be a closed ball of radius ε around p and set Mε=M\intBε where int denotes the interior. Then XMε is a manifold with boundary, the image of Mε under the section X:M\pS. Consequently

MΛ=MpΛ=limε0MεΛ=limε0MεXπ0Λ=limε0X(Mεπ0Λ
=limε0XMεdΦ=limε0XMεΦ.E67

If indXp is the index of X at p

MΛ=indXpπ01pΦ=χMπ01pΦ.E68

Since n=2m we also have the Gauss-Bonnet-Chern Theorem 6.2

MΛ=Mn!Kndμg=πmm!2nχM,E69

we finally obtain

π01pΦ=πmm!2n.E70

Let MM be a compact orientable manifold with boundary with Euler characteristic χM=dimH0MdimH1M+. A compact oriented manifold M2 can be constructed, the double of M, by taking two disjoint copies of M and identifying corresponding points of M.

Theorem 7.1 The Euler characteristic of the manifold M2 is given by

χM2=2χMχM.E71

Proof: Let U and V be open neighborhoods of the two copies of M in M2 such that HkUHVHkM for all k and HkUVHkM for all k. So there is the sequence 0H0M2HkM2HkUHkVHkUVHk+1M2. When the sequence is exact, a theorem can be applied to obtain the result.

This is very interesting since it claims different things depending on whether the dimension n of M is even or odd. When n is odd χM2=0 hence χM=1/2χM which implies χM must be even. But when n is even, χM=0, so the previous theorem implies

2χM=χM2.E72

Corollary 7.1 Let M be a compact orientable manifold with boundary of even dimension n. Let X be a vector field on M with only finitely many zeros all in M\M such that X is outward pointing on M. The sum of indices of X is χM.

Proof: Modify X near M so it is an outward pointing unit normal ν on the boundary and so there are no new zeros. Then there is a vector field on M2 which looks like X on one copy of M and X on the other. Since n is even, the index X of an isolated zero is the same as the index of X at that zero. The Theorem of Poincaré-Hopf on the sum of indices of X gives twice the sum of the indices of X equals χM2=2χM by (72).

Theorem 7.2 Let M be a compact oriented Riemannian manifold with boundary of even dimension n=2m, tangent bundle π:TMM and associated sphere bundle π0=πS:SM. Let ω be a connection on the principal bundle ω¯:SOTMM, with curvature form Ω. Let Λ be the unique n-form on M with

ω¯Λ=εi1,,inΩi2i1Ωinin1=2mm!PfΩ,E73

and Φ an n1-form on S with π0Λ=dΦ. Let ν:MS be the outward pointing unit normal on M. Then

MKndμg=1n!MΛ=πmm!2nn!χM+1n!MνΦ.E74

Proof: Extend ν to a vector field X on M with only finitely many zeros p1,,pkM\M. Let Biε be the closed balls of radius ε centered at pi which are disjoint from each other and from M. Put Mε=M\i=1kintBiε. Now integrate the form Λ over M and use (70)

MΛ=limε0XMεΦ=νMΦ+i=1klimε0BiεΦ=MνΦ+πmm!2ni=1MindXpi=MνΦ+πmm!2nχM.E75

The last line makes use of Corollary 7.1.

Theorem 7.2 presents one way in which Theorem 6.2 can be generalized to the case of manifolds with boundary. At this point an interpretation for the first term in (75) is not easy to provide. It is required to obtain an explicit Φ such that π0Λ=dΦ. In fact such a Φ can be constructed.

References

  1. 1. Chern SS. A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemannian manifolds. Annals of Mathematics. 1944;45(2):747-752
  2. 2. Chern SS. On curvature and characteristic classes of a Riemannian manifold. Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg. 1955;20:117-126
  3. 3. Bishop RL, Goldberg SI. Some implications of the generalized gauss-bonnet theorem. Transactions of the American Mathematical Society. 1964;112(3):508-535
  4. 4. Bell D. The Gauss-Bonnet Theorem for vector bundles. Journal of Geometry. 2006;85:15-21
  5. 5. Peterson P. Riemannian Geometry. New York: Springer-Verlag; 1998
  6. 6. Goldberg SI. Curvature and Homology. New York: Academic Press; 1962
  7. 7. Spivak M. A Comprehensive Introduction to Differential Geometry. Houston: Publish or Perish; 1999
  8. 8. Gilkey P. Invariance Theory, the Heat Equation, and the Atiyah-Singer Index Theorem. 2nd ed. Boca Raton: CRC Press; 1995
  9. 9. Boothby W. An Introduction to Differentiable Manifolds and Riemannian Geometry. New York: Academic Press; 1975
  10. 10. Kobayashi S, Nomizu K. Foundations of Differential Geometry. New York: Interscience; 1963
  11. 11. Rosenberg S. The Laplacian on a Riemannian Manifold. Cambridge: Cambridge University Press; 1997
  12. 12. Glavan D, Lin C. Einstein-Gauss-Bonnet gravity in four-dimensional spacetime. Physical Review Letters. 2020;124:081301
  13. 13. Herrero-Valea M. The shape of scalar Gauss-Bonnet gravity. Journal of High Energy Physics. 2022;13(03):075
  14. 14. Liu Y, Pan Q, Wang B, Cai R-G. Dynamical perturbations and critical phenomena in Gauss-Bonnet AdS black holes. Physics Letters A. 2010;693:343-350
  15. 15. D’Hoker E, Phong DH. The geometry of string perturbation theory. Reviews of Modern Physics. 1988;60(4):917-1065
  16. 16. Chow B, Lu L, Ni L. Hamilton’s Ricci Flow. Providence, RI: AMS; 2006

Written By

Paul Bracken

Reviewed: 06 June 2022 Published: 11 July 2022