Invariant theory

Source: Wikipedia, the free encyclopedia.

Invariant theory is a branch of

polynomial functions that do not change, or are invariant, under the transformations from a given linear group. For example, if we consider the action of the special linear group SLn on the space of n by n matrices by left multiplication, then the determinant
is an invariant of this action because the determinant of A X equals the determinant of X, when A is in SLn.

Introduction

Let be a group, and a finite-dimensional vector space over a field (which in classical invariant theory was usually assumed to be the complex numbers). A representation of in is a group homomorphism , which induces a

group action
of on . If is the space of polynomial functions on , then the group action of on produces an action on by the following formula:

With this action it is natural to consider the subspace of all polynomial functions which are invariant under this group action, in other words the set of polynomials such that for all . This space of invariant polynomials is denoted .

First problem of invariant theory:[1] Is a finitely generated algebra over ?

For example, if and the space of square matrices, and the action of on is given by left multiplication, then is isomorphic to a polynomial algebra in one variable, generated by the determinant. In other words, in this case, every invariant polynomial is a linear combination of powers of the determinant polynomial. So in this case, is finitely generated over .

If the answer is yes, then the next question is to find a minimal basis, and ask whether the module of polynomial relations between the basis elements (known as the

syzygies
) is finitely generated over .

Invariant theory of finite groups has intimate connections with Galois theory. One of the first major results was the main theorem on the symmetric functions that described the invariants of the symmetric group acting on the polynomial ring ] by permutations of the variables. More generally, the Chevalley–Shephard–Todd theorem characterizes finite groups whose algebra of invariants is a polynomial ring. Modern research in invariant theory of finite groups emphasizes "effective" results, such as explicit bounds on the degrees of the generators. The case of positive characteristic, ideologically close to modular representation theory, is an area of active study, with links to algebraic topology.

Invariant theory of

semisimple Lie groups
has its roots in invariant theory.

standard monomials
.

Examples

Simple examples of invariant theory come from computing the invariant monomials from a group action. For example, consider the -action on sending

Then, since are the lowest degree monomials which are invariant, we have that

This example forms the basis for doing many computations.

The nineteenth-century origins

The theory of invariants came into existence about the middle of the nineteenth century somewhat like Minerva: a grown-up virgin, mailed in the shining armor of algebra, she sprang forth from Cayley's Jovian head.

Weyl (1939b, p.489)

Cayley first established invariant theory in his "On the Theory of Linear Transformations (1845)." In the opening of his paper, Cayley credits an 1841 paper of George Boole, "investigations were suggested to me by a very elegant paper on the same subject... by Mr Boole." (Boole's paper was Exposition of a General Theory of Linear Transformations, Cambridge Mathematical Journal.)[2]

Classically, the term "invariant theory" refers to the study of invariant

representations of Lie groups
are rooted in this area.

In greater detail, given a finite-dimensional

invariants of binary forms
where n = 2.

Other work included that of Felix Klein in computing the invariant rings of finite group actions on (the

du Val singularities
.

Like the Arabian phoenix rising out of its ashes, the theory of invariants, pronounced dead at the turn of the century, is once again at the forefront of mathematics.

Kung & Rota (1984, p.27)

The work of David Hilbert, proving that I(V) was finitely presented in many cases, almost put an end to classical invariant theory for several decades, though the classical epoch in the subject continued to the final publications of Alfred Young, more than 50 years later. Explicit calculations for particular purposes have been known in modern times (for example Shioda, with the binary octavics).

Hilbert's theorems

ring of invariants of G acting on the ring of polynomials R = S(V) is finitely generated. His proof used the Reynolds operator
ρ from R to RG with the properties

  • ρ(1) = 1
  • ρ(a + b) = ρ(a) + ρ(b)
  • ρ(ab) = a ρ(b) whenever a is an invariant.

Hilbert constructed the Reynolds operator explicitly using

Cayley's omega process Ω, though now it is more common to construct ρ indirectly as follows: for compact groups G, the Reynolds operator is given by taking the average over G, and non-compact reductive groups can be reduced to the case of compact groups using Weyl's unitarian trick
.

Given the Reynolds operator, Hilbert's theorem is proved as follows. The ring R is a polynomial ring so is graded by degrees, and the ideal I is defined to be the ideal generated by the homogeneous invariants of positive degrees. By Hilbert's basis theorem the ideal I is finitely generated (as an ideal). Hence, I is finitely generated by finitely many invariants of G (because if we are given any – possibly infinite – subset S that generates a finitely generated ideal I, then I is already generated by some finite subset of S). Let i1,...,in be a finite set of invariants of G generating I (as an ideal). The key idea is to show that these generate the ring RG of invariants. Suppose that x is some homogeneous invariant of degree d > 0. Then

x = a1i1 + ... + anin

for some aj in the ring R because x is in the ideal I. We can assume that aj is homogeneous of degree d − deg ij for every j (otherwise, we replace aj by its homogeneous component of degree d − deg ij; if we do this for every j, the equation x = a1i1 + ... + anin will remain valid). Now, applying the Reynolds operator to x = a1i1 + ... + anin gives

x = ρ(a1)i1 + ... + ρ(an)in

We are now going to show that x lies in the R-algebra generated by i1,...,in.

First, let us do this in the case when the elements ρ(ak) all have degree less than d. In this case, they are all in the R-algebra generated by i1,...,in (by our induction assumption). Therefore, x is also in this R-algebra (since x = ρ(a1)i1 + ... + ρ(an)in).

In the general case, we cannot be sure that the elements ρ(ak) all have degree less than d. But we can replace each ρ(ak) by its homogeneous component of degree d − deg ij. As a result, these modified ρ(ak) are still G-invariants (because every homogeneous component of a G-invariant is a G-invariant) and have degree less than d (since deg ik > 0). The equation x = ρ(a1)i1 + ... + ρ(an)in still holds for our modified ρ(ak), so we can again conclude that x lies in the R-algebra generated by i1,...,in.

Hence, by induction on the degree, all elements of RG are in the R-algebra generated by i1,...,in.

Geometric invariant theory

The modern formulation of

symbolic method of invariant theory
, an apparently heuristic combinatorial notation, has been rehabilitated.

One motivation was to construct moduli spaces in algebraic geometry as quotients of schemes parametrizing marked objects. In the 1970s and 1980s the theory developed interactions with symplectic geometry and equivariant topology, and was used to construct moduli spaces of objects in differential geometry, such as instantons and monopoles.

See also

References

External links