Subpage 1 of User:Mathsci edit

 
Bernhard Riemann (1826-1866)
 
Élie Cartan (1869-1951)

The classical approach of Gauss was the standard elementary approach[1][2][3][4] which predated the emergence of the concepts of Riemannian manifold initiated by Bernhard Riemann in the mid-nineteenth century and of connection developed by Tullio Levi-Civita, Élie Cartan and Hermann Weyl in the early twentieth century. The notion of connection, covariant derivative and parallel transport gave a more conceptual and uniform way of understanding curvature, which not only allowed generalisations to higher dimensional manifolds but also provided an important tool for defining new geometric invariants, called characteristic classes.[5] The approach using covariant derivatives and connections is nowadays the one adopted in more advanced textbooks.[6][7][8]

Connections on a surface can be defined in a variety of ways. The Riemannian connection or Levi-Civita connection[9] is perhaps most easily understood in terms of lifting vector fields, considered as first order differential operators acting on functions on the manifold, to differential operators on the frame bundle: in the case of an embedded surface, the lift is very simply described in terms of orthogonal projection. Indeed the vector bundles associated with the frame bundle are all sub-bundles of trivial bundles that extend to the ambient Euclidean space; a first order differential operator can always be applied to a section of a trivial bundle, in particular to a section of the original sub-bundle, although the resulting section might no longer be a section of the sub-bundle. This can be corrected by projecting orthogonally.

The Riemannian connection can also be characterized abstractly independently of an embedding. The equations of geodesics are easy to write in terms of the Riemannian connection, which can be locally expressed in terms of the Christoffel symbols. Along a curve in the surface, the connection defines a first order differential equation in the frame bundle. The monodromy of this equation defines parallel transport for the connection, a notion introduced in this context by Levi-Civita.[10] This gives an equivalent more geometric way of describing the connection in terms of lifting paths in the manifold to paths in the frame bundle. This formalised the classical theory of the "moving frame", favoured by French authors.[11] Lifts of loops about a point give rise to the holonomy group at that point. The Gaussian curvature at a point can be recovered from parallel transport around increasingly small loops at the point. Equivalently curvature can be calculated directly infinitesimally in terms of Lie brackets of lifted vector fields.

The approach of Cartan, using connection 1-forms on the frame bundle of M, gives a third way to understand the Riemannian connection, which is particularly easy to describe for an embedded surface. Thanks to a result of Kobayashi (1956), later generalized by Narasimhan & Ramanan (1961), the Riemannian connection on a surface embedded in Euclidean space E3 is just the pullback under the Gauss map of the Riemannian connection on S2.[12] Using the identification of S2 with the homogeneous space SO(3)/SO(2), the connection 1-form is just a component of the Maurer-Cartan 1-form on SO(3). In other words everything reduces to understanding the 2-sphere properly.[13]

Covariant derivative edit

 
A vector field on the torus

For a surface M embedded in E3 (or more generally a higher dimensional Euclidean space), there are several equivalent definitions of a vector field X on M:

The last condition means that the assignment f   Xf on C(M) satisfies the Leibniz rule

 

The space of all vector fields  (M) forms a module over C(M), closed under the Lie bracket

 

with a C(M)-valued inner product (X,Y), which encodes the Riemannian metric on M.

Since  (M) is a submodule of C(M, E3)=C(M)  E3, the operator X  I is defined on  (M), taking values in C(M, E3).

Let P be the smooth map from M into M3(R) such that P(p) is the orthogonal projection of E3 onto the tangent space at p.

Pointwise multiplication by P gives a C(M)-module map of C(M, E3) onto  (M) . The assignment

 

defines an operator   on  (M) called the covariant derivative, satisfying the following properties

  1.   is C(M)-linear in X
  2.   (Leibniz rule for derivation of a module)
  3.   (compatibility with the metric)
  4.   (symmetry property).

The first three properties state that   is an affine connection compatible with the metric, sometimes also called a hermitian or metric connection. The last symmetry property says that the torsion tensor

 

vanishes identically, so that the affine connection is torsion-free.

The assignment   is uniquely determined by these four conditions and is called the Riemannian connection or Levi-Civita connection.

Although the Riemannian connection was defined using an embedding in Euclidean space, this uniqueness property means that it is in fact an intrinsic invariant of the surface.

It existence can be proved directly for a general surface by noting that the four properties imply

 

so that   depends only on the metric and is unique. On the other hand if this is used as a definition of  , it is readily checked that the four properties above are satisfied.[14]

Equivalently, in local coordinates (x,y) with basis tangent vectors e1=  and e2 =  , the connection   can be expressed purely in terms of the metric using the Christoffel symbols:

 

If c(t) is a path in M, then the Euler equations for c to be a geodesic can be written more compactly as

 

Parallel transport edit

 
Parallel transport of a vector around a geodesic triangle on the sphere. The length of the transported vector and the angle it makes with each side remain constant.

Given a curve in the Euclidean plane and a vector at the starting point, the vector can be transported along the curve by requiring the moving vector to remain parallel to the original one and of the same length, i.e. it should remain constant along the curve. If the curve is closed, the vector will be unchanged when the starting point is reached again. This is well known not to be possible on a general surface, the sphere being the most familiar case. In fact it is not usually possible to identify simultaneously or "parallelize" all the tangent planes of such a surface: the only parallelizable closed surfaces are those homeomorphic to a torus. [15]

Parallel transport can always be defined along curves on a surface using only the metric on the surface. Thus tangent planes along a curve can be identified using the intrinsic geometry, even when the surface itself is not parallelizable.

Parallel transport along geodesics, the "straight lines" of the surface, is easy to define. A vector in the tangent plane is transported along a geodesic as the unique vector field with constant length and making a constant angle with the velocity vector of the geodesic.

For a general curve, its geodesic curvature measures how far the curve departs from being a geodesics; it is defined as the rate at which the curve's velocity vector rotates in the surface. In turn the geodesic curvature determines how vectors in the tangent planes along the curve should rotate during parallel transport.

A vector field v(t) along a unit speed curve c(t), with geodesic curvature kg(t), is said to be parallel along the curve if

  • it has constant length
  • the angle θ(t) that it makes with the velocity vector   satisfies
 

This yields the previous rule for parallel transport along a geodesic, because in that case kg = 0, so the angle θ(t) should remain constant.[16] The existence of parallel transport follows from standard existence theorems for ordinary differential equations.

Arnold has suggested[17][18] that since parallel transport on a geodesic segment is easy to describe, parallel transport on an arbitrary C1 curve could be constructed as a limit of parallel transport on an approximating family of piecewise geodesic curves.[19]

This equation shows once more that parallel transport depends only on the metric structure so is an intrinsic invariant of the surface; it is another way of writing the ordinary differential equation involving the geodesic curvature of c. Parallel transport can be extended immediately to piecewise C1 curves.

The covariant derivative can in turn be recovered from parallel transport.[20] In fact   can be calculated at a point p, by taking a curve c through p with tangent X, using parallel transport to view the restriction of Y to c as a function in the tangent space at p and then taking the derivative.

Frame bundle edit

Let M be a surface embedded in E3. The orientation on the surface means that an "outward pointing" normal unit vector n is defined at each point of the surface and hence a determinant can be defined on tangent vectors v and w at that point:

 

using the usual scalar triple product on E3 (itself a determinant).

An ordered basis or frame v, w in the tangent space is said to be oriented if det(v, w) is positive.

  • The tangent bundle of M consists of pairs (p, v) in M x E3 such that v lies in the tangent plane to M at p.
  • The circle bundle of M consists of pairs (p, v) with ||v|| = 1. It is identical to the frame bundle because for each unit tangent vector v there is a unique unique tangent vector w with det(v, w) = 1.

Since the group of rotations in the plane SO(2) acts simply transitively on oriented orthonormal frames in the plane, it follows that it also acts on the frame or circle bundles of M.[21] The definitions of the tangent bundle, the unit tangent bundle and the (oriented orthonormal) frame bundle E can be extended to arbitrary surfaces in the usual way.[22][23] There is a similar identification between the latter two which again become principal SO(2)-bundles. In other words:

The frame bundle is a principal bundle with structure group SO(2).

There is also a corresponding notion of parallel transport in the setting of frame bundles:[24][25]

Every continuously differentiable curve in M can be lifted to a curve in E in such a way that the tangent vector field of the lifted curve is the lift of the tangent vector field of the original curve.

This statement means that any frame on a curve can be parallely transported along the curve. This is precisely the idea of "moving frames". Since any unit tangent vector can be completed uniquely to an oriented frame, parallel transport of tangent vectors implies (and is equivalent to) parallel transport of frames. The lift of a geodesic in M turns out to be a geodesic in E for the Sasaki metric (see below).[26] Moreover the Gauss map of M into S2 induces a natural map between the associated frame bundles which is equivariant for the actions of SO(2).[27]

Cartan's idea of introducing the frame bundle as a central object was the natural culmination of the theory of moving frames, developed in France by Darboux and Goursat. It also echoed parallel developments in Albert Einstein's theory of relativity.[28] Objects appearing in the formulas of Gauss, such as the Christoffel symbols, can be given a natural geometric interpretation in this framework. Unlike the more intuitive normal bundle, easily visualised as a tubular neighbourhood of an embedded surface in E3, the frame bundle is an intrinsic invariant that can be defined independently of an embedding. When there is an embedding, it can also be visualised as a subbundle of the Euclidean frame bundle E3 x SO(3), itself a submanifold of E3 x M3(R).

Cartan-Ehresmann connection edit

.

The theory of connections according to Élie Cartan, and later Charles Ehresmann, revolves around:

All "natural" vector bundles associated with the manifold M, such as the tangent bundle, the cotangent bundle or the exterior bundles, can be constructed from the frame bundle using the representation theory of the structure group K = SO(2), a compact matrix group.

Cartan's definition of a connection can be understood as a way of lifting vector fields on M to vector fields on the frame bundle E invariant under the action of the structure group K. This "universal lift" then immediately induces lifts to vector bundles associated with E and hence allows the covariant derivative, and its generalisation to forms, to be recovered. If X is a vector field on M, the lift X* is required to have the following properties:[29]

  • X* is a vector field on E;
  • the map X   X* is C(M)-linear;
  • X* is K-invariant and induces the vector field X on C(M)   C(E).

Here K acts as a periodic flow on E, so the canonical generator A of its Lie algebra acts as the corresponding vector field, called the vertical vector field A*. It follows from the above conditions that, in the tangent space of an arbitrary point in E, the lifts X* span a two-dimensional subspace of horizontal vectors, forming a complementary subspace to the vertical vectors. This becomes an orthogonal complement for the canonical Riemannian metric on E defined by Shigeo Sasaki.[30][31]

Horizontal vector fields admit the following characterisation:

  • Every K-invariant horizontal vector field on E has the form X* for a unique vector field X on M.

If σ is a representation of K on a finite-dimensional vector space V, then the associated vector bundle E XK V over M has a C(M)-module of sections that can be identified with

 

the space of all smooth functions ξ : E → V which are K-equivariant in the sense that

 

for all x ∈ E and g ∈ K.

The identity representation of SO(2) on R2 corresponds to the tangent bundle of M.

The covariant derivative   is defined on an invariant section ξ by the formula

 


The connection on the frame bundle can also be described using K-invariant differential 1-forms on E.[32] [33]

The frame bundle E is a 3-manifold. The space of p-forms on E is denoted Λp(E).[34] It admits a natural action of the structure group K.

Given a connection on the principal bundle E corresponding to a lift X   X* of vector fields on M, there is a unique connection form ω in

 ,

the space of K-invariant 1-forms on E, such that[35]

 

for all vector fields X on M and

 

for the vector field A* on E corresponding to the canonical generator A of  .

Conversely the lift X* is uniquely characterised by the following properties:

  • X* is K-invariant and induces X on M;
  • ω(X*)=0.

Cartan structural equations edit

On the frame bundle E of a surface M there are three canonical 1-forms:

  • The connection form ω, invariant under the structure group K = SO(2)
  • Two tautologous 1-forms θ1 and θ2, transforming according to the basis vectors of the identity representation of K

If π: E   M is the nature projection, the 1-forms θ1 and θ2 are defined by

 

where Y is a vector field on E and e1, e2 are the tangent vectors to M of the orthonormal frame.

These 1-forms satisfy the following structural equations, due in this formulation to Cartan:[36]

 
(First structural equations)
 
(Second structural equation)

where K is the Gaussian curvature on M.

Holonomy and curvature edit

Parallel transport in the frame bundle can be used to show that the Gaussian curvature of a surface M measures the amount of rotation obtained by translating vectors around small curves in M.[37] Holonomy is exactly the phenomenon that occurs when a tangent vector (or orthonormal frame) is parallelly transported around a closed curve. The vector reached when the loop is closed will be a rotation of the original vector, i.e. it will correspond to an element of the rotaion group SO(2), in other words an angle modulo 2π. This is the holonomy of the loop, because the angle does not depend on the choice of starting vector.

 
Geometric interpretation of the Lie bracket of two vector fields

This geometric interpretation of curvature relies on a similar geometric of the Lie bracket of two vector fields on E. Let U1 and U2 be vector fields on E with corresponding local flows αt and βt.

  • Starting at a point A corresponding to x in E, travel   along the integral curve for U1 to the point B at  .
  • Travel from B by going   along the integral curve for U2 to the point C at  .
  • Travel from C by going   along the integral curve for U1 to the point D at  .
  • Travel from D by going   along the integral curve for U2 to the point E at  .

In general the end point E will differ from the starting point A. As s   0, the end point E will trace out a curve through A. The Lie bracket [U1,U2] at x is precisely the tangent vector to this curve at A.[38]

To apply this theory, introduce vector fields U1, U2 and V on the frame bundle E which are dual to the 1-forms θ1, θ2 and ω at each point. Thus

 

Moreover V is invariant under K and U1, U2 transform according to the identity representation of K.

The structural equations of Cartan imply the following Lie bracket relations:

 

The geometrical interpretation of the Lie bracket can be applied to the last of these equations. Since ω(Ui)=0, the flows αt and βt in E are lifts by parallel transport of their projections in M.

Informally the idea is as follows. The starting point A and end point E essentially differ by an element of SO(2), that is an angle of rotation. The area enclosed by the projected path in M is approximately  . So in the limit as s   0, the angle of rotation divided by this area tends to the coefficient of V, i.e. the curvature.

This reasoning is made precise in the following result.[39]

Let f be a diffeomorphism of an open disc in the plane into M and let Δ be a triangle in this disc. Then the holonomy angle of the loop formed by the image under f of the perimeter of the triangle is given by the integral of the Gauss curvature of the image under f of the inside of the triangle.

In symbols, the holonomy angle mod 2π is given by

 

where the integral is with respect to the area form on M.

This result implies the relation between Gaussian curvature because as the triangle shrinks in size to a point, the ratio of this angle to the area tends to the Gaussian curvature at the point. The result can be proved by a combination of Stokes's theorem and Cartan's structural equations and can in turn be used to obtain a generalisation of Gauss's theorem on geodesics triangles to more general triangles.[40]

One of the other standard approaches to curvature, through the covariant derivative  , identifies the difference

 

as a field of endomorphisms of the tangent bundle, the Riemann curvature tensor.[41][42] Since   is induced by the lifted vector field X* on E, the use of the vector fields Ui and V and their Lie brackets is more or less equivalent to this approach. The vertical vector field W=A* corresponding to the canonical generator A of   could also be added since it commutes with V and satisfies [W,U1] = U2 and [W,U2] = —U1.

Example: the 2-sphere edit

The differential geometry of the 2-sphere can be approached from three different points of view:

S2 can be identifed with the unit sphere in E3

 

Its tangent bundle T, unit tangent bundle U and oriented orthonormal frame bundle E are given by

 
 
 

The map sending (a,v) to (a, v, a x v) allows U and E to be identified.

Let

 

be the orthogonal projection onto the normal vector at a, so that

 

is the orthogonal projection onto the tangent space at a.

The group G = SO(3) acts by rotation on E3 leaving S2 invariant. The stabilizer subgroup K of the vector (0,0,1) in E3 may be identified with SO(2) and hence

S2 may be identified with SO(3)/SO(2).

This action extends to an action on T, U and E by making G act on each component. G acts transitively on S2 and simply transitively on U and E.

The action of SO(3) on E commutes with the action of SO(2) on E that rotates frames

 

Thus E becomes a principal bundle with structure group K. Taking the G-orbit of the point ((0,0,1),(1,0,0),(0,1,0)), the space E may be identified with G. Under this identification the actions of G and K on E become left and right translation. In other words:

The oriented orthonormal frame bundle of S2 may be identified with SO(3).

The Lie algebra   of SO(3) consists of all skew-symmetric real 3 x 3 matrices.[43] the adjoint action of G by conjugation on   reproduces the action of G on E3. The group SU(2) has a 3-dimensional Lie algebra consisting of complex skew-hermitian traceless 2 x 2 matrices, which is isomorphic to  . The adjoint action of SU(2) factors through its centre, the matrices ± I. Under these identifications, SU(2) is exhibited as a double cover of SO(3), so that SO(3) = SU(2) / ± I.[44] On the other hand SU(2) is diffeomorphic to the 3-sphere and under this identification the standard Riemannian metric on the 3-sphere becomes the essentially unique biinvariant Riemannian metric on SU(2). Under the quotient by ± I, SO(3) can be identified with the real projective space of dimension 3 and itself has an essentially unique biinvariant Riemannian metric. The geometric exponential map for this metric at I coincides with the usual exponential function on matrices and thus the geodesics through I are have the form exp Xt where X is a skew-symmetric matrix.

The actions of G on itself, and hence on C(G) by left and right translation induce infinitesimal actions of   on C(G) by vector fields

 

The right and left invariant vector fields are related by the formula

 

The vector fields λ(X) and ρ(X) commute with right and left translation and give all right and left invariant vector fields on G. Since C(S2) = C(G/K) can be identified with C(G)K, the function invariant under right translation by K, the operators λ(X) also induces vector fields Π(X) on S2.

Let A, B, C be the standard basis of   given by

 

Their Lie brackets [X,Y] = XYYX are given by

 

The vector fields λ(A), λ(B), λ(C) form a basis of the tangent space at each point of G. Let α, β, γ be the corresponding dual basis of right invariant 1-forms on G. The Lie bracket relations imply the Maurer-Cartan equations

 

These are also the corresponding components of the Maurer-Cartan form

 

a right invariant matrix-valued 1-form on G, which satisfies the relation

 

The inner product on   defined by

 

is invariant under the adjoint action. Let π be the orthogonal projection onto the subspace generated by A, i.e. onto  , the Lie algebra of K. For X in  , the lift of the vector field Π(X) from C(G/K) to C(G) is given by the formula

 

This lift is G-equivariant on vector fields of the form Π(X) and has a unique extension to more general vector fields on G / K.

The right invariant 1-form α is the connection form on G corresponding to this lift. Thus

  • vertical vector fields on G are those of the form f · λ(A) with f in C(G);
  • horizontal vector fields on G are those of the form f1 · λ(B) + f2 · λ(C) with fi in C(G).

The existence of the basis vector fields λ(A), λ(B), λ(C) shows that SO(3) is parallelizable. This is not true for SO(3)/SO(2) by the hairy ball theorem: S2 does not admit any nowhere vanishing vector fields.

Parallel transport in the frame bundle amounts to lifting a path from SO(3)/SO(2) to SO(3). It is accomplished using a matrix-valued ordinary differential equation called the transport equation as follows:[45]

  • the space of matrices F satisfying F2 = I, F = F* and Tr F = –1 can be identified with S2 by choosing a base point F0 and using the transitive action by conjugation of SO(3);
  • a path F(t) starting at F0 is lifted to the path g(t) in SO(3) satisfying   and g(0) = I.[46]

Given a path F(t), the ordinary differential equation  , with initial condition  , has a unique C1 solution g(t) with values in G, giving the lift by parallel transport of F.

Set A = ½ Ft F. Since F2 = I and F is symmetric, A is skew-symmetric.

The unique solution g(t) of the ordinary differential equation

 

with initial condition g(0) = I guaranteed by the Picard-Lindelöf theorem, must have gTg constant and therefore I, since

 

Moreover

 

since g-1Fg has derivative 0:

 

Thus the gauge transformation g trivialises the path F(t).

There is another kinematic way of understanding parallel transport and geodesic curvature in terms of "rolling without slipping or twisting". Although well known to differential geometers since the early part of the twentieth century, it has also been applied to problems in engineering and robotics.[47] Consider the 2-sphere as a rigid body in three dimensional space rolling without slipping or twisting on a horizontal plane. The point of contact will describe a curve in the plane and on the surface. At each point of contact the different tangent planes of the sphere can be identified with the horizontal plane itself and hence with one another.

  • The usual curvature of the planar curve is the the geodesic curvature of the curve traced on the sphere.
  • This identification of the tangent planes along the curve corresponds to parallel transport.

This is particularly easy to visualize for a sphere: it is exactly the way a marble can be rolled along a perfectly flat table top.

The roles of the plane and the sphere can be reversed to provide an alternative but equivalent point of view. The sphere is regarded as fixed and the plane has to roll without slipping or twisting along the given curve on the sphere.[48]

Embedded surfaces edit

When the surface M is embedded in E3 (or a higher dimensional Euclidean space), parallel transport can be described in yet another way. Parallel transport along a curve c(t), with t taking values in [0,1], starting from a tangent from a tangent vector v0 also amounts to finding a map v(t) from [0,1] to R3 such that

  • v(t) is a tangent vector to M at c(t) with v(0) = v0
  • the velocity vector   is normal to the surface at c(t)

This always has a unique solution, called the parallel transport of v0 along c.

Analytically, the existence of parallel transport can be deduced from the transport equation

gt g–1 = ½ FtF

on G = SO(3) with F(t)=2P(c(t)) – I.[49] The transformation g trivialises the tangent bundle along c. Given a tangent vector v0 in the tangent space to M at c(0), the vector v(t)=g(t)v0 lies in the tangent space to M at c(t) and satisfies the equation

 

It therefore is exactly the parallel transport of v along the curve c.[50] In this case the length of the vector v(t) is constant. More generally if another initial tangent vector u0 is taken instead of v0, the inner product (v(t),u(t)) is constant. The tangent spaces along the curve c(t) are thus canonically identified as inner product spaces by parallel transport so that parallel transport gives an isometry between the tangent planes.

Parallel transport is related to the covariant derivative  , because the condition on the velocity vector   may be rewritten in terms of the covariant derivative as[51][52]

 

the fundamental defining equation for parallel transport.


There is another kinematic way of understanding parallel transport and geodesic curvature in terms of "rolling without slipping or twisting". Although well known to differential geometers since the early part of the twentieth century, it has also been applied to problems in engineering and robotics.[53] Consider the surface as a rigid body in three dimensional space rolling without slipping or twisting on a horizontal plane. The point of contact will describe a curve in the plane and on the surface. At each point of contact the different tangent planes of the surface can be identified with the horizontal plane itself and hence with one another.

  • The usual curvature of the planar curve is the the geodesic curvature of the curve traced on the surface.
  • This identification of the tangent planes along the curve corresponds to parallel transport.

This is particularly easy to visualize for a sphere: it is exactly the way a marble can be rolled along a perfectly flat table top.

The roles of the plane and the surface can reversed to provide an alternative but equivalent point of view. The surface is regarded as fixed and the plane has to roll without slipping or twisting along the given curve in the surface.[54]

This geometric way of viewing parallel transport can also be directly expressed in the language of geometry.[55] The envelope of the tangent planes to M along a curve c is a surface with vanishing Gaussian curvature, which by Minding's theorem, must be locally isometric to the Euclidean plane. This identification allows parallel transport to be defined, because in the Euclidean plane all tangent planes are identified with the space itself.

Such a lift was constructed above for the special case of the 2-sphere S2 = SO(3) / SO(2).

In 1956 Kobayashi proved that:[56]

  • The invariant connection on the frame bundle of S2 corresponds to the Riemannian connection on S2.
  • Under the Gauss map, the connection it induces on the frame bundle of M corresponds to the Riemannian connection on M.

In this case the Sasaki metric is a multiple of the unique biinvariant metric on SO(3), inherited from the standard metric on its double cover SU(2) = S3.[57][58]

The pullback of the connection form on SO(3) under the Gauss map gives the connection form on the frame bundle of M corresponding to the Riemannian connection.

There is another simple way of constructing the connection form ω using the embedding of M in E3.[59]

The tangent vectors e1 and e2 of a frame on M define smooth functions from E with values in R3, so each gives a 3-vector of functions and in particular de1 is a 3-vector of 1-forms on E.

The connection form is given by

 

taking the usual scalar product on 3-vectors.

Gauss-Codazzi equations edit

When M is embedded in E3, two other 1-forms ψ and χ can be defined on the frame bundle E using the shape operator.[60][61][62] Indeed the Gauss map induces a K-equivariant map of E into SO(3), the frame bundle of S2 = SO(3)/SO(2). The form ω is the pullback of one of the three right invariant Maurer-Cartan forms on SO(3). The 1-forms ψ and χ are defined to be the pullbacks of the other two.

These 1-forms satisfy further structure equations:

 
(symmetry equation)
 
(Gauss equation)
 
(Codazzi equations)

The Gauss–Codazzi equations for χ, ψ and ω follow immediately from the Maurer-Cartan equations for the three right invariant 1-forms on SO(3).

Notes edit

  1. ^ Eisenhart 2004
  2. ^ Kreyszig 1991
  3. ^ Berger 2004
  4. ^ Wilson 2008
  5. ^ Kobayashi & Nomizu 1969, Chapter XII.
  6. ^ do Carmo 1976
  7. ^ O'Neill 1997
  8. ^ Singer & Thorpe 1967
  9. ^ Levi-Civita 1917
  10. ^ Levi-Civita 1917
  11. ^ Darboux & 1887,1889,1896
  12. ^ Kobayashi & Nomizu 1969
  13. ^ Ivey & Landsberg 2003 This approach, together with its higher dimensional generalisations, is discussed in great detail in Chapters 1 and 2.
  14. ^ Kobayashi & Nomizu 1963, p. 160
  15. ^ Berger 2004, p. 127
  16. ^ Berger 2004, p. 129
  17. ^ Arnold 1982, p. 301-306, Appendix I.
  18. ^ Berger 2004, p. 263-264
  19. ^ Arnold's method of approximation also applies to higher dimensional Riemannian manifolds, after having given an appropriate geometric description of parallel transport along a geodesic. Parallel transport can be shown to be a continuous function on the Sobolev space of paths of finite energy, introduced in Klingenberg (1982), by an extension of the analysis in Nelson (1969) of ordinary differential equations of the form   to the case where the potential A is only assumed to be piecewise continuous or even just square integrable.
  20. ^ do Carmo 1992, p. 56-57
  21. ^ Singer & Thorpe 1967
  22. ^ Kobayashi & Nomizu 1963
  23. ^ {harvnb|Singer|Thorpe|1967}}
  24. ^ Kobayashi & Nomizu 1963, p. 68-71
  25. ^ Singer & Thorpe 1967, p. 181-184
  26. ^ Sasaki 1958
  27. ^ Kobayashi 1956
  28. ^ Ivey & Landsberg 2003
  29. ^ Kobayashi & Nomizu 1963, p. 63-64
  30. ^ Sasaki 1958
  31. ^ Berger 2004, p. 727-728
  32. ^ Singer & Thorpe 1967
  33. ^ A general connection on a principal bundle E with structure group H is described by a 1-form on E with values in   invariant under the tensor product of the action of H on 1-forms and the adjoint action. For surfaces, H is Abelian and 1-dimensional, so the connection 1-form is essentially given by an invariant 1-form on E.
  34. ^ The space of p-forms can be identified with the space of alternating p-fold C(E)-multilinear maps on the module of vector fields. For further details see Helgason (1978), pages 19-21.
  35. ^ Kobayashi & Nomizu 1963
  36. ^ Singer & Thorpe 1967, p. 185-189
  37. ^ Singer & Thorpe 1967, p. 190-193
  38. ^ Singer & Thorpe 1967, p. 143
  39. ^ Singer et al.
  40. ^ Singer & Thorpe 1967, p. 195
  41. ^ Kobayashi & Nomizu 1963
  42. ^ do Carmo 1992
  43. ^ The Lie algebra of a closed connected subgroup G of a real or complex general linear group consists of all matrices X such that exp tX lies in G for all real t; see Adams (1983) or Varadarajan (1984).
  44. ^ Geometrically this double cover corresponds to a spin structure on S2.
  45. ^ This standard treatment of parallel transport can be found for example in Driver (1995, p. 25).
  46. ^ In mathematical physics, the solution of this differential equation is often expressed as a path-ordered exponential; see for example Nelson (1969).
  47. ^ Sharpe 1997, p. 375-388, Appendix B: Rolling without Slipping or Twisting
  48. ^ Berger 2004, p. 130
  49. ^ Kobayashi & Nomizu 1963, p. 69
  50. ^ This standard treatment of parallel transport can be found for example in Driver (1995, p. 25).
  51. ^ do Carmo 1992, p. 52
  52. ^ Kobayashi & Nomizu 1963
  53. ^ Sharpe 1997, p. 375-388, Appendix B: Rolling without Slipping or Twisting
  54. ^ Berger 2004, p. 130
  55. ^ do Carmo 1976, p. 244
  56. ^ Kobayashi 1956, Theorem II.
  57. ^ Klingenberg & Sasaki 1975
  58. ^ Arnold 1978, Appendix 2: Geodesics of left-invariant metrics on Lie groups and the hydrodynamics of ideal fluids.
  59. ^ Singer & Thorpe 1967, p. 221-223
  60. ^ O'Neill 1997, p. 256-257
  61. ^ Ivey & Landsberg 2003, Chapter 2.
  62. ^ Kobayashi & Nomizu 1969, Chapter VII.

References edit

  • Adams, J. Frank (1983), Lectures on Lie Groups, University of Chicago Press, ISBN 0226005305
  • Aleksandrov, A.D.; Zalgaller, V.A. (1967), Instrinsic Geometry of Surfaces, Translations of Mathematical Monographs, vol. 15, American Mathematical Society
  • Arnold, V.I (1989), Mathematical Methods of Classical Mechanics, Springer-Verlag, ISBN 0-387-96890-3
  • Berger, Marcel (2004), A Panoramic View of Riemannian Geometry, Springer-Verlag, ISBN 3-540-65317-1
  • Cartan, Élie (1983), Geometry of Riemannian Spaces, Math Sci Press, ISBN 0915692341; translated from 2nd edition of Leçons sur la géométrie des espaces de Riemann (1951) by James Glazebrook.
  • Cartan, Élie (2001), Riemannian Geometry in an Orthogonal Frame (from lectures delivered by É Cartan at the Sorbonne in 1926-27) (PDF), World Scientific, ISBN 9810247478; translated from Russian by V. V. Goldberg with a foreword by S. S. Chern.
  • Darboux, Gaston (1887,1889,1896), Leçons sur la théorie génerale des surfaces: Volume I, Volume II, Volume III, Volume IV, Gauthier-Villars {{citation}}: Check date values in: |year= (help); External link in |title= (help)
  • do Carmo, Manfredo P. (1976), Differential Geometry of Curves and Surfaces, Prentice-Hall, ISBN 0132125897
  • do Carmo, Manfredo P. (1992), Riemannian geometry, Birkhäuser, ISBN 0-8176-3490-8
  • Driver, Bruce K. (1995), A primer on Riemannian geometry and stochastic analysis on path spaces (PDF), Lectures given at the E.T.H., Zurich
  • Eisenhart, Luther P. (2004), A Treatise on the Differential Geometry of Curves and Surfaces, Dover, ISBN 0486438201 Full 1909 text (now out of copyright)
  • Eisenhart, Luther P. (1947), An Introduction to Differential Geometry with Use of the Tensor Calculus, Princeton Mathematical Series, vol. 3, Princeton Univerisity Press
  • Euler, Leonard (1760), "Recherches sur la courbure des surfaces", Memoires de l'academie des sciences de Berlin, 16 (published 1767): 119–143.
  • Euler, Leonard (1771), "De solidis quorum superficiem in planum explicare licet", Novi Commentarii academiae scientiarum Petropolitanae, 16 (published 1772) {{citation}}: Text "3-34" ignored (help).
  • Gauss, Carl Friedrich (1827), General Investigations of Curved Surfaces, New York: Raven Press (published 1965) {{citation}}: Cite has empty unknown parameter: |1= (help) translated by A.M.Hiltebeitel and J.C.Morehead; "Disquisitiones generales circa superficies curvas", Commentationes Societatis Regiae Scientiarum Gottingesis Recentiores Vol. VI (1827), pp. 99-146.
  • Gray, Alfred; Abbena, Elsa; Salamon, Simon (2006), Modern Differential Geometry of Curves And Surfaces With Mathematica, CRC Press, ISBN 1584884487
  • Han, Qing; Hong, Jia-Xing (2006), Isometric Embedding of Riemannian Manifolds in Euclidean Spaces, American Mathematical Society, ISBN 0821840711
  • Helgason, Sigurdur (1978), Differential Geometry, Lie groups and Symmetric Spaces, Academic Press, ISBN 0-12-338460-5
  • Hilbert, David; Cohn-Vossen, Stephan (1952), Geometry and the Imagination (2nd ed.), New York: Chelsea, ISBN 978-0-8284-1087-8.
  • Ivey, Thomas A.; Landsberg, J.M. (2003), Cartan for Beginners: Differential Geometry via Moving Frames and Exterior Systems, Graduate Studies in Mathematics, vol. 61, American Mathematical Society, ISBN 0-8218-3375-8
  • Jacobowitz, Howard (1972), "Local Isometric Embeddings of Surfaces into Euclidean Four Space", Indiana Univ. Math. J., 21: 249–254
  • Klingenberg, Wilhelm (1982), Riemannian geometry, de Gruyter studies in mathematics, vol. 1, de Gruyter, ISBN 3-11-008673-5
  • Kobayashi, Shochichi (1956), "Induced connections and imbedded Riemannian space", Nagoya Math. J., 10: 15–25
  • Kobayashi, Shoshichi; Nomizu, Katsumi (1963), Foundations of differential geometry, Vol. I, Wiley Interscience, ISBN 0470496487
  • Kobayashi, Shoshichi; Nomizu, Katsumi (1969), Foundations of differential geometry, Vol. II, Wiley Interscience, ISBN 0470496487
  • Kreyszig, Erwin (1991), Differential Geometry, Dover, ISBN 0486667219
  • Levi-Civita, Tullio (1917), "Nozione di parallelismo in una varieta qualunque", Rend. Circ. Mat. Palermo, 42: 173–205
  • Milnor, John W. (1963), Morse theory, Annals of Mathematics Studies, vol. 51, Princeton University Press, ISBN 0691080089 {{citation}}: Cite has empty unknown parameter: |1= (help)
  • Narasimhan, M.S.; Ramanan, S. (1961), "Existence of Universal Connections", Amer. J. Math., 83: 563–572
  • Nelson, Edward (1969), Topics in dynamics — I: Flows, Mathematical Notes, Princeton University Press
  • O'Neill, Barrett (1997), Elementary Differential Geometry, Academic Press, ISBN 0-12-526745-2
  • Poznjak, E.G. (1973), "Isometric imbedding of two-dimensional Riemannian metrics in Euclidean spaces", Russian Math. Surveys, 28: 47–77
  • Pressley, Andrew (2001), Elementary Differential Geometry, Springer Undergraduate Mathematics Series, Springer-Verlag, ISBN 1852331526
  • Sharpe, Richard W. (1997), Differential Geometry: Cartan's Generalization of Klein's Erlangen Program, Springer-Verlag, ISBN 0387947329
  • Singer, Isadore M.; Thorpe, John A. (1967), Lecture Notes on Elementary Topology and Geometry, Springer-Verlag, ISBN 0-387-90202-3
  • Sternberg, Shlomo (1964), Lectures on differential geometry, Prentice-Hall
  • Struik, Dirk Jan (1988), Lectures on classical differential geometry: Second Edition, Dover, ISBN 0486656098
  • Toponogov, Victor A. (2005), Differential Geometry of Curves and Surfaces: A Concise Guide, Springer-Verlag, ISBN 0817643842
  • Valiron, Georges (1986), The Classical Differential Geometry of Curves and Surfaces, Math Sci Press, ISBN0915692392 Full text of book
  • Varadarajan, V. S. (1984), Lie Groups, Lie Algebras, and Their Representations, Springer-Verlag, ISBN 0387909699
  • Wilson, Pelham (2008), Curved Space: From Classical Geometries to Elementary Differential Geometry, Cambridge University Press, ISBN 978-0-521-71390-0