Prehomogeneous vector space

Last updated

In mathematics, a prehomogeneous vector space (PVS) is a finite-dimensional vector space V together with a subgroup G of the general linear group GL(V) such that G has an open dense orbit in V. The term prehomogeneous vector space was introduced by Mikio Sato in 1970. These spaces have many applications in geometry, number theory and analysis, as well as representation theory. The irreducible PVS were classified first by Vinberg in his 1960 thesis in the special case when G is simple and later by Sato and Tatsuo Kimura in 1977 in the general case by means of a transformation known as "castling". They are subdivided into two types, according to whether the semisimple part of G acts prehomogeneously or not. If it doesn't then there is a homogeneous polynomial on V which is invariant under the semisimple part of G.

Contents

Setting

In the setting of Sato, G is an algebraic group and V is a rational representation of G which has a (nonempty) open orbit in the Zariski topology. However, PVS can also be studied from the point of view of Lie theory: for instance, in Knapp (2002), G is a complex Lie group and V is a holomorphic representation of G with an open dense orbit. The two approaches are essentially the same, and the theory has validity over the real numbers. We assume, for simplicity of notation, that the action of G on V is a faithful representation. We can then identify G with its image in GL(V), although in practice it is sometimes convenient to let G be a covering group.

Although prehomogeneous vector spaces do not necessarily decompose into direct sums of irreducibles, it is natural to study the irreducible PVS (i.e., when V is an irreducible representation of G). In this case, a theorem of Élie Cartan shows that

G ≤ GL(V)

is a reductive group, with a centre that is at most one-dimensional. This, together with the obvious dimensional restriction

dim G ≥ dim V,

is the key ingredient in the Sato–Kimura classification.

Castling

The classification of PVS is complicated by the following fact. Suppose m > n > 0 and V is an m-dimensional representation of G over a field F. Then:

(G × SL(n), VFn) is a PVS if and only if (G × SL(mn), V*Fmn) is a PVS.

The proof is to observe that both conditions are equivalent to there being an open dense orbit of the action of G on the Grassmannian of n-planes in V, because this is isomorphic to the Grassmannian of (mn)-planes in V*.

(In the case that G is reductive, the pair (G, V) is equivalent to the pair (G, V*) by an automorphism of G.)

This transformation of PVS is called castling. Given a PVS V, a new PVS can be obtained by tensoring V with F and castling. By repeating this process, and regrouping tensor products, many new examples can be obtained, which are said to be "castling-equivalent". Thus PVS can be grouped into castling equivalence classes. Sato and Kimura show that in each such class, there is essentially one PVS of minimal dimension, which they call "reduced", and they classify the reduced irreducible PVS.

Classification

The classification of irreducible reduced PVS (G, V) splits into two cases: those for which G is semisimple, and those for which it is reductive with one-dimensional centre. If G is semisimple, it is (perhaps a covering of) a subgroup of SL(V), and hence G × GL(1) acts prehomogenously on V, with one-dimensional centre. We exclude such trivial extensions of semisimple PVS from the PVS with one-dimensional center. In other words, in the case that G has one-dimensional center, we assume that the semisimple part does not act prehomogeneously; it follows that there is a relative invariant, i.e., a function invariant under the semisimple part of G, which is homogeneous of a certain degree d.

This makes it possible to restrict attention to semisimple G ≤ SL(V) and split the classification as follows:

  1. (G, V) is a PVS;
  2. (G, V) is not a PVS, but (G × GL(1), V) is.

However, it turns out that the classification is much shorter, if one allows not just products with GL(1), but also with SL(n) and GL(n). This is quite natural in terms of the castling transformation discussed previously. Thus we wish to classify irreducible reduced PVS in terms of semisimple G ≤ SL(V) and n ≥ 1 such that either:

  1. (G × SL(n), VFn) is a PVS;
  2. (G × SL(n), VFn) is not a PVS, but (G × GL(n), VFn) is.

In the latter case, there is a homogeneous polynomial which separates the G × GL(n) orbits into G × SL(n) orbits.

This has an interpretation in terms of the grassmannian Grn(V) of n-planes in V (at least for n ≤ dim V). In both cases G acts on Grn(V) with a dense open orbit U. In the first case the complement Grn(V) U has codimension 2; in the second case it is a divisor of some degree d, and the relative invariant is a homogeneous polynomial of degree nd.

In the following, the classification list will be presented over the complex numbers.

General examples

G V Type 1 Type 2 Type 2 isotropy group Degree
G ⊆ SL(m, C) Cmnm+1 n = mGm
SL(m, C) Cmm − 1 ≥ n ≥ 1*
SL(m, C) Λ2Cmm odd, n = 1, 2 m even, n = 1 Sp(m, C) m/2
SL(m, C) S2Cmn = 1 SO(m, C) m
SO(m, C) Cmm − 1 ≥ n ≥ 1*SO(n, C) × SO(mn, C) 2
Sp(2m, C) C2m2m − 1 ≥ n ≥ 1*, n odd 2m − 1 ≥ n ≥ 1*, n even Sp(n, C) × Sp(2mn, C) 1

* Strictly speaking, we must restrict to n ≤ (dim V)/2 to obtain a reduced example.

Irregular examples

Type 1

Spin(10, C) on C16

Type 2

Sp(2m, C) × SO(3, C) on C2mC3

Both of these examples are PVS only for n = 1.

Remaining examples

The remaining examples are all type 2. To avoid discussing the finite groups appearing, the lists present the Lie algebra of the isotropy group rather than the isotropy group itself.

GVnIsotropy algebra Degree
SL(2, C) S3C21 0 4
SL(6, C) Λ3C61 (3, C) × (3, C)4
SL(7, C) Λ3C71 C
2
7
SL(8, C) Λ3C81 (3, C)16
SL(3, C) S2C32 0 6
SL(5, C) Λ2C33, 4 (2, C), 0 5, 10
SL(6, C) Λ2C32 (2, C) × (2, C) × (2, C)6
SL(3, C) × SL(3, C) C3C32 (1, C) × (1, C)6
Sp(6, C) Λ3
0
C6
1 (3, C)4
Spin(7, C) C81, 2, 3 C
2
, (3, C) × (2, C), (2, C) × (3, C)
2, 2, 2
Spin(9, C) C161 (7, C) 2
Spin(10, C) C162, 3 C
2
× (2, C)
, (2, C) × (3, C)
2, 4
Spin(11, C) C321 (5, C) 4
Spin(12, C) C321 (6, C) 4
Spin(14, C) C641 C
2
× C
2
8
GC
2
C71, 2 (3, C), (2, C)2, 2
EC
6
C271, 2 C
4
, (8, C)
3, 6
EC
7
C561 C
6
4

Here Λ3
0
C6C14
denotes the space of 3-forms whose contraction with the given symplectic form is zero.

Proofs

Sato and Kimura establish this classification by producing a list of possible irreducible prehomogeneous (G, V), using the fact that G is reductive and the dimensional restriction. They then check whether each member of this list is prehomogeneous or not.

However, there is a general explanation why most of the pairs (G, V) in the classification are prehomogeneous, in terms of isotropy representations of generalized flag varieties. Indeed, in 1974, Richardson observed that if H is a semisimple Lie group with a parabolic subgroup  P, then the action of P on the nilradical of its Lie algebra has a dense open orbit. This shows in particular (and was noted independently by Vinberg in 1975) that the Levi factor G of P acts prehomogeneously on V := /[, ]. Almost all of the examples in the classification can be obtained by applying this construction with P a maximal parabolic subgroup of a simple Lie group H: these are classified by connected Dynkin diagrams with one distinguished node.

Applications

One reason that PVS are interesting is that they classify generic objects that arise in G-invariant situations. For example, if G = GL(7), then the above tables show that there are generic 3-forms under the action of G, and the stabilizer of such a 3-form is isomorphic to the exceptional Lie group G2.

Another example concerns the prehomogeneous vector spaces with a cubic relative invariant. By the Sato-Kimura classification, there are essentially four such examples, and they all come from complexified isotropy representations of hermitian symmetric spaces for a larger group H (i.e., G is the semisimple part of the stabilizer of a point, and V is the corresponding tangent representation).

In each case a generic point in V identifies it with the complexification of a Jordan algebra of 3 × 3 hermitian matrices (over the division algebras R, C, H and O respectively) and the cubic relative invariant is identified with a suitable determinant. The isotropy algebra of such a generic point, the Lie algebra of G and the Lie algebra of H give the complexifications of the first three rows of the Freudenthal magic square.

HGVIsotropy algebra Jordan algebra
Sp(6, C) SL(3, C) S2C3(3, C) J3(R)
SL(6, C) SL(3, C) × SL(3, C) C3C3(3, C) J3(C)
SO(12, C) SL(6, C) Λ2C6(6, C) J3(H)
EC
7
EC
6
C27C
4
J3(O)

Other Hermitian symmetric spaces yields prehomogeneous vector spaces whose generic points define Jordan algebras in a similar way.

HGVIsotropy algebra Jordan algebra
Sp(2n, C) SL(n, C) S2Cn(n, C) Jn(R)
SL(2n, C) SL(n, C) × SL(n, C) CnCn(n, C) Jn(C)
SO(4n, C) SL(2n, C) Λ2C2n(2n, C) Jn(H)
SO(m + 2, C) SO(m, C) Cm(m − 1, C) J(m − 1)

The Jordan algebra J(m 1) in the last row is the spin factor (which is the vector space Rm1R, with a Jordan algebra structure defined using the inner product on Rm1). It reduces to J2(R), J2(C), J2(H), J2(O) for m = 3, 4, 6 and 10 respectively.

The relation between hermitian symmetric spaces and Jordan algebras can be explained using Jordan triple systems.

Related Research Articles

<span class="mw-page-title-main">Lie algebra</span> Algebraic structure used in analysis

In mathematics, a Lie algebra is a vector space together with an operation called the Lie bracket, an alternating bilinear map , that satisfies the Jacobi identity. In other words, a Lie algebra is an algebra over a field for which the multiplication operation is alternating and satisfies the Jacobi identity. The Lie bracket of two vectors and is denoted . A Lie algebra is typically a non-associative algebra. However, every associative algebra gives rise to a Lie algebra, consisting of the same vector space with the commutator Lie bracket, .

<span class="mw-page-title-main">Lie group</span> Group that is also a differentiable manifold with group operations that are smooth

In mathematics, a Lie group is a group that is also a differentiable manifold, such that group multiplication and taking inverses are both differentiable.

<span class="mw-page-title-main">Representation of a Lie group</span> Group representation

In mathematics and theoretical physics, a representation of a Lie group is a linear action of a Lie group on a vector space. Equivalently, a representation is a smooth homomorphism of the group into the group of invertible operators on the vector space. Representations play an important role in the study of continuous symmetry. A great deal is known about such representations, a basic tool in their study being the use of the corresponding 'infinitesimal' representations of Lie algebras.

<span class="mw-page-title-main">Lie algebra representation</span>

In the mathematical field of representation theory, a Lie algebra representation or representation of a Lie algebra is a way of writing a Lie algebra as a set of matrices in such a way that the Lie bracket is given by the commutator. In the language of physics, one looks for a vector space together with a collection of operators on satisfying some fixed set of commutation relations, such as the relations satisfied by the angular momentum operators.

In mathematics, a generalized flag variety is a homogeneous space whose points are flags in a finite-dimensional vector space V over a field F. When F is the real or complex numbers, a generalized flag variety is a smooth or complex manifold, called a real or complexflag manifold. Flag varieties are naturally projective varieties.

<span class="mw-page-title-main">Linear algebraic group</span> Subgroup of the group of invertible n×n matrices

In mathematics, a linear algebraic group is a subgroup of the group of invertible matrices that is defined by polynomial equations. An example is the orthogonal group, defined by the relation where is the transpose of .

<span class="mw-page-title-main">Reductive group</span>

In mathematics, a reductive group is a type of linear algebraic group over a field. One definition is that a connected linear algebraic group G over a perfect field is reductive if it has a representation that has a finite kernel and is a direct sum of irreducible representations. Reductive groups include some of the most important groups in mathematics, such as the general linear group GL(n) of invertible matrices, the special orthogonal group SO(n), and the symplectic group Sp(2n). Simple algebraic groups and (more generally) semisimple algebraic groups are reductive.

<span class="mw-page-title-main">Cartan subalgebra</span> Nilpotent subalgebra of a Lie algebra

In mathematics, a Cartan subalgebra, often abbreviated as CSA, is a nilpotent subalgebra of a Lie algebra that is self-normalising. They were introduced by Élie Cartan in his doctoral thesis. It controls the representation theory of a semi-simple Lie algebra over a field of characteristic .

<span class="mw-page-title-main">Symmetric space</span> A (pseudo-)Riemannian manifold whose geodesics are reversible.

In mathematics, a symmetric space is a Riemannian manifold whose group of symmetries contains an inversion symmetry about every point. This can be studied with the tools of Riemannian geometry, leading to consequences in the theory of holonomy; or algebraically through Lie theory, which allowed Cartan to give a complete classification. Symmetric spaces commonly occur in differential geometry, representation theory and harmonic analysis.

<span class="mw-page-title-main">Representation theory of the Lorentz group</span> Representation of the symmetry group of spacetime in special relativity

The Lorentz group is a Lie group of symmetries of the spacetime of special relativity. This group can be realized as a collection of matrices, linear transformations, or unitary operators on some Hilbert space; it has a variety of representations. This group is significant because special relativity together with quantum mechanics are the two physical theories that are most thoroughly established, and the conjunction of these two theories is the study of the infinite-dimensional unitary representations of the Lorentz group. These have both historical importance in mainstream physics, as well as connections to more speculative present-day theories.

In mathematics, a Gelfand pair is a pair (G,K ) consisting of a group G and a subgroup K (called an Euler subgroup of G) that satisfies a certain property on restricted representations. The theory of Gelfand pairs is closely related to the topic of spherical functions in the classical theory of special functions, and to the theory of Riemannian symmetric spaces in differential geometry. Broadly speaking, the theory exists to abstract from these theories their content in terms of harmonic analysis and representation theory.

<span class="mw-page-title-main">Hermitian symmetric space</span> Manifold with inversion symmetry

In mathematics, a Hermitian symmetric space is a Hermitian manifold which at every point has an inversion symmetry preserving the Hermitian structure. First studied by Élie Cartan, they form a natural generalization of the notion of Riemannian symmetric space from real manifolds to complex manifolds.

In mathematics, the orbit method establishes a correspondence between irreducible unitary representations of a Lie group and its coadjoint orbits: orbits of the action of the group on the dual space of its Lie algebra. The theory was introduced by Kirillov for nilpotent groups and later extended by Bertram Kostant, Louis Auslander, Lajos Pukánszky and others to the case of solvable groups. Roger Howe found a version of the orbit method that applies to p-adic Lie groups. David Vogan proposed that the orbit method should serve as a unifying principle in the description of the unitary duals of real reductive Lie groups.

In mathematics, a discrete series representation is an irreducible unitary representation of a locally compact topological group G that is a subrepresentation of the left regular representation of G on L²(G). In the Plancherel measure, such representations have positive measure. The name comes from the fact that they are exactly the representations that occur discretely in the decomposition of the regular representation.

<span class="mw-page-title-main">Lattice (discrete subgroup)</span> Discrete subgroup in a locally compact topological group

In Lie theory and related areas of mathematics, a lattice in a locally compact group is a discrete subgroup with the property that the quotient space has finite invariant measure. In the special case of subgroups of Rn, this amounts to the usual geometric notion of a lattice as a periodic subset of points, and both the algebraic structure of lattices and the geometry of the space of all lattices are relatively well understood.

<span class="mw-page-title-main">Representation theory</span> Branch of mathematics that studies abstract algebraic structures

Representation theory is a branch of mathematics that studies abstract algebraic structures by representing their elements as linear transformations of vector spaces, and studies modules over these abstract algebraic structures. In essence, a representation makes an abstract algebraic object more concrete by describing its elements by matrices and their algebraic operations. The theory of matrices and linear operators is well-understood, so representations of more abstract objects in terms of familiar linear algebra objects helps glean properties and sometimes simplify calculations on more abstract theories.

<span class="mw-page-title-main">Complexification (Lie group)</span> Universal construction of a complex Lie group from a real Lie group

In mathematics, the complexification or universal complexification of a real Lie group is given by a continuous homomorphism of the group into a complex Lie group with the universal property that every continuous homomorphism of the original group into another complex Lie group extends compatibly to a complex analytic homomorphism between the complex Lie groups. The complexification, which always exists, is unique up to unique isomorphism. Its Lie algebra is a quotient of the complexification of the Lie algebra of the original group. They are isomorphic if the original group has a quotient by a discrete normal subgroup which is linear.

In mathematics, given a quiver Q with set of vertices Q0 and set of arrows Q1, a representation of Q assigns a vector space Vi to each vertex and a linear map V(α): V(s(α)) → V(t(α)) to each arrow α, where s(α), t(α) are, respectively, the starting and the ending vertices of α. Given an element dQ0, the set of representations of Q with dim Vi = d(i) for each i has a vector space structure.

This is a glossary of representation theory in mathematics.

<span class="mw-page-title-main">Representation theory of semisimple Lie algebras</span>

In mathematics, the representation theory of semisimple Lie algebras is one of the crowning achievements of the theory of Lie groups and Lie algebras. The theory was worked out mainly by E. Cartan and H. Weyl and because of that, the theory is also known as the Cartan–Weyl theory. The theory gives the structural description and classification of a finite-dimensional representation of a semisimple Lie algebra ; in particular, it gives a way to parametrize irreducible finite-dimensional representations of a semisimple Lie algebra, the result known as the theorem of the highest weight.

References