Homogeneous space

Source: Wikipedia, the free encyclopedia.
flat torus is homogeneous under its diffeomorphism, homeomorphism, and isometry groups
.

In

G-orbit
.

Formal definition

Let X be a non-empty set and G a group. Then X is called a G-space if it is equipped with an action of G on X.[1] Note that automatically G acts by automorphisms (bijections) on the set. If X in addition belongs to some category, then the elements of G are assumed to act as automorphisms in the same category. That is, the maps on X coming from elements of G preserve the structure associated with the category (for example, if X is an object in Diff then the action is required to be by diffeomorphisms). A homogeneous space is a G-space on which G acts transitively.

Succinctly, if X is an object of the category C, then the structure of a G-space is a homomorphism:

into the group of automorphisms of the object X in the category C. The pair (X, ρ) defines a homogeneous space provided ρ(G) is a transitive group of symmetries of the underlying set of X.

Examples

For example, if X is a topological space, then group elements are assumed to act as homeomorphisms on X. The structure of a G-space is a group homomorphism ρ : G → Homeo(X) into the homeomorphism group of X.

Similarly, if X is a differentiable manifold, then the group elements are diffeomorphisms. The structure of a G-space is a group homomorphism ρ : G → Diffeo(X) into the diffeomorphism group of X.

Riemannian symmetric spaces
are an important class of homogeneous spaces, and include many of the examples listed below.

Concrete examples include:

Examples of homogeneous spaces
space X group G stabilizer H
spherical space Sn−1 O(n) O(n − 1)
oriented Sn−1 SO(n) SO(n − 1)
projective space PRn−1 PO(n) PO(n − 1)
Euclidean space En E(n) O(n)
oriented En E+(n) SO(n)
hyperbolic space Hn O+(1, n) O(n)
oriented Hn SO+(1, n) SO(n)
anti-de Sitter space AdSn+1 O(2, n) O(1, n)
Grassmannian Gr(r, n) O(n) O(r) × O(nr)
affine space A(n, K) Aff(n, K) GL(n, K)
Isometry groups
  • Positive curvature:
    1. Sphere (orthogonal group): Sn−1 ≅ O(n) / O(n−1). This is true because of the following observations: First, Sn−1 is the set of vectors in Rn with norm 1. If we consider one of these vectors as a base vector, then any other vector can be constructed using an orthogonal transformation. If we consider the span of this vector as a one dimensional subspace of Rn, then the complement is an (n − 1)-dimensional vector space that is invariant under an orthogonal transformation from O(n − 1). This shows us why we can construct Sn−1 as a homogeneous space.
    2. Oriented sphere (
      special orthogonal group
      ): Sn−1 ≅ SO(n) / SO(n − 1)
    3. Projective space (projective orthogonal group): Pn−1 ≅ PO(n) / PO(n − 1)
  • Flat (zero curvature):
    1. Euclidean space (Euclidean group, point stabilizer is orthogonal group): An ≅ E(n) / O(n)
  • Negative curvature:
    1. Hyperbolic space (
      orthochronous Lorentz group, point stabilizer orthogonal group, corresponding to hyperboloid model
      ): Hn ≅ O+(1, n) / O(n)
    2. Oriented hyperbolic space: SO+(1, n) / SO(n)
    3. Anti-de Sitter space: AdSn+1 = O(2, n) / O(1, n)
Others

Geometry

From the point of view of the Erlangen program, one may understand that "all points are the same", in the geometry of X. This was true of essentially all geometries proposed before Riemannian geometry, in the middle of the nineteenth century.

Thus, for example, Euclidean space, affine space and projective space are all in natural ways homogeneous spaces for their respective symmetry groups. The same is true of the models found of non-Euclidean geometry of constant curvature, such as hyperbolic space.

A further classical example is the space of lines in projective space of three dimensions (equivalently, the space of two-dimensional subspaces of a four-dimensional

line geometry of Julius Plücker
.

Homogeneous spaces as coset spaces

In general, if X is a homogeneous space of G, and Ho is the

stabilizer of some marked point o in X (a choice of origin), the points of X correspond to the left cosets
G/Ho, and the marked point o corresponds to the coset of the identity. Conversely, given a coset space G/H, it is a homogeneous space for G with a distinguished point, namely the coset of the identity. Thus a homogeneous space can be thought of as a coset space without a choice of origin.

For example, if H is the identity subgroup {e}, then X is the G-torsor, which explains why G-torsors are often described intuitively as "G with forgotten identity".

In general, a different choice of origin o will lead to a quotient of G by a different subgroup Ho′ that is related to Ho by an inner automorphism of G. Specifically,

(1)

where g is any element of G for which go = o. Note that the inner automorphism (1) does not depend on which such g is selected; it depends only on g modulo Ho.

If the action of G on X is

smooth manifold and so X carries a unique smooth structure
compatible with the group action.

One can go further to

properly discontinuously
.

Example

For example, in the line geometry case, we can identify H as a 12-dimensional subgroup of the 16-dimensional general linear group, GL(4), defined by conditions on the matrix entries

h13 = h14 = h23 = h24 = 0,

by looking for the stabilizer of the subspace spanned by the first two standard basis vectors. That shows that X has dimension 4.

Since the homogeneous coordinates given by the minors are 6 in number, this means that the latter are not independent of each other. In fact, a single quadratic relation holds between the six minors, as was known to nineteenth-century geometers.

This example was the first known example of a Grassmannian, other than a projective space. There are many further homogeneous spaces of the classical linear groups in common use in mathematics.

Prehomogeneous vector spaces

The idea of a prehomogeneous vector space was introduced by Mikio Sato.

It is a finite-dimensional

group action of an algebraic group G, such that there is an orbit of G that is open for the Zariski topology
(and so, dense). An example is GL(1) acting on a one-dimensional space.

The definition is more restrictive than it initially appears: such spaces have remarkable properties, and there is a classification of irreducible prehomogeneous vector spaces, up to a transformation known as "castling".

Homogeneous spaces in physics

Given the Poincaré group G and its subgroup the Lorentz group H, the space of cosets G / H is the spacetime algebra Minkowski space.[2]

metrics for some cosmological models; for example, the three cases of the Friedmann–Lemaître–Robertson–Walker metric may be represented by subsets of the Bianchi I (flat), V (open), VII (flat or open) and IX (closed) types, while the Mixmaster universe represents an anisotropic example of a Bianchi IX cosmology.[3]

A homogeneous space of N dimensions admits a set of 1/2N(N + 1)

For three dimensions, this gives a total of six linearly independent Killing vector fields; homogeneous 3-spaces have the property that one may use linear combinations of these to find three everywhere non-vanishing Killing vector fields ξ(a)
i
,

where the object Cabc, the "structure constants", form a

flat isotropic universe, one possibility is Cabc = 0 (type I), but in the case of a closed FLRW universe, Cabc = εabc, where εabcis the Levi-Civita symbol
.

See also

Notes

  1. ^ We assume that the action is on the left. The distinction is only important in the description of X as a coset space.
  2. W. A. Benjamin
  3. ^ Steven Weinberg (1972), Gravitation and Cosmology, John Wiley and Sons

References