A finite-dimensional (complex) representation of a Lie group is a Lie group homomorphism
where is a finite-dimensional complex vector space.
Infinite-dimensional representations are important, but subtle. In general one must equip with some topology and add topological conditions to all notions which follow. For instance, one might take to be a Hilbert space.
An example of such a representation arises naturally if you attempt to generalise the regular representation! One must take to be something like the space of square-integrable functions on the group, rather than the space of arbitrary functions, to get a pleasant theory.
If is a finite-dimensional representation of as above, then we can take its derivative:
mapping from the Lie algebra of to the space of endomorphisms of . Note that
is a Lie algebra with bracket
The map is a Lie algebra homomorphism. Often we write, abusively, instead of .
Note that choosing an isomorphism induces isomorphisms and .
A (complex) representation of a Lie algebra is a Lie algebra homomorphism
where is a complex vector space. That is,
is -linear;
.
Note that by Theorem 1.37 the differential of a Lie group representation is a Lie algebra representation.
Warning! It is not the case that, if is a Lie algebra representation, then
Indeed, in general need not be an element of the Lie algebra at all, and even if it is the displayed equation will not usually hold.
The notions of -homomorphism (or -homomorphism, or intertwiner), isomorphism, subrepresentation, and irreducible representation stay the same as for finite groups. For example, a -homomorphism from to is a linear map such that
for all and .
A -linear representation of is a complex representation of such that
for all .
If is a complex Lie group, then a holomorphic representation of is a complex representation whose derivative is -linear; equivalently, the map is holomorphic.
Let be a Lie group, be a Lie algebra.
If and are irreducible finite-dimensional representations of or , then
If , then any - or -homomorphism is scalar.
Any irreducible finite-dimensional representation of an abelian Lie group or Lie algebra is one-dimensional.
If is an irreducible finite-dimensional representation of (or ) and (or ) is the center of (or ) then there is a homomorphism (or ) such that
for all (or ) and . We call this the central character.
The proofs are all the same as in the finite group case! ∎
Let be a finite-dimensional representation of a Lie group . Let be its derivative.
If is invariant under , then is invariant under .
If is irreducible, then is irreducible.
If is unitary, that is, there is a basis for such that for all , then is skew-Hermitian, that is, for all (using the same basis for ).
Let be another finite-dimensional representation of . If , then .
If is connected, then the converses to these statements hold.
So, for connected Lie groups, we can test irreducibility and isomorphism at the level of Lie algebras.
For (1), we know that for any and . Taking the derivative at , it follows that as required. Part (2) follows from (1).
For (3), if is unitary, then after choosing a basis appropriately it is a Lie group homomorphism . The derived homomorphism therefore lands in the Lie algebra of .
For part (4), let be a -isomorphism, so that in particular,
for all and . Taking the derivative at gives
so that is a -isomorphism as required.
If is connected, then is generated by . Hence all proofs above can be reversed. For example, for (1), suppose that is preserved by . If and , then
as is preserved by and also closed. Since every can be written as a finite product of for , we see that is preserved by as required. ∎
We give a list of various constructions with representations of Lie groups, and the analogous constructions for their derivatives.
The standard representation of a linear Lie group comes from its action on :
The direct sum of representations , is with derivative
The determinant representation of is which sends to . We have
which follows from .
If is a representation of , the dual representation of is defined by
for a linear functional on and . It has derivative
Given a basis of , then the matrix of with respect to the dual basis is
which differentiates to
If and are representations of , then as before the tensor product representation is the representation on defined by
Then using the product rule one sees
The symmetric square and alternating square are also as for finite groups. If is a representation of then has a representation :
and
Similarly we have a representation on :
We can take tensor/symmetric/alternating products of more than one factor. Suppose are representations of .
We form the tensor product
It is generated by symbols subject to the multilinear relations, that is, linearity in each slot:
One has
The action of is as before: for , ,
The derivative is, for and ,
We also write
The th symmetric power is the space generated by symbols with linearity in each slot and any permutation of the vectors giving the same element. We have
where . Indeed, if is a basis for then a basis for is
from which finding the dimension is a simple counting problem.
As for higher tensor powers, the actions of and are
and
The th alternating power is the space generated by symbols with linearity in each slot and having the alternating property: for any permutation , we have
In particular, switching the places of two components reverses the sign, while if two of the vectors coincide (more generally, if they are linearly dependent).
We have
where . Indeed, if are a basis for then a basis for is
from which finding the dimension is a simple counting problem. In particular, is one-dimensional generated by .
The representation on is given again as above: for ,
while for
To give an example of how to justify the claims about derivatives, we do the case of tensor products. Suppose , are vector spaces acted on by . Let , , . We must compute
Expanding:
This gives
as required.
We defined tensor products (and so on) of representations of Lie groups and then differentiated them. We could also directly make these definitions with Lie algebras. For instance, if is a Lie algebra and is a representation of , we define the symmetric square representation on by
We can construct representations as vector spaces of functions on topological spaces with actions of . If acts on a set , then it also acts on the vector space of functions by . Usually this will be infinite dimensional, and so out of the scope of our course, but sometimes we can impose conditions allowing us to handle it. For example, acts on , and hence on the space of polynomial functions in variables. Imposing a further restriction — to homogeneous polynomials of some fixed degree — gives a finite-dimensional representation. The derivative must be calculated on a case-by-case basis. We will see examples of this on the problem sheet.
Let be a linear Lie group and be its Lie algebra. Then the adjoint representation of is the action on by conjugation. We usually write instead of , so that
for and . By Proposition 1.24 (2), is indeed in . Thus the map is a Lie group homomorphism:
An alternative definition, that works for general Lie groups, is to consider the conjugation by map and then take the derivative:
The derivative of , denoted by , is called the adjoint representation of the Lie algebra . Thus
Again, we write for , so we have
By Theorem 1.37 we have the formula
as elements of .
Let and be as above and let . Then
The map is a Lie algebra homomorphism, so that
Thus, for all ,
Since , taking the differential at , we get
The map is a Lie algebra homomorphism because it is the differential of a Lie group homomorphism.∎
This explains the origin of the Jacobi identity:
while
Equating these gives the Jacobi identity.
The first formula, , could have been used to define the adjoint representation for any Lie algebra, without reference to Lie groups.
Warning! It is very easy to misinterpret some of the formulas concerning the adjoint representation. For example, does not mean that
but (as already noted and proved) that
Similarly, does not mean that is equal to but rather is the identity
If is abelian, so is . If, moreover, is connected, then the converse holds.
Suppose that is abelian. Then, for all and ,
Taking the derivative at we see that . Thus is trivial. Differentiating, we see is trivial, so for all . Thus for all as required.
Conversely, suppose is connected and is abelian. Then is trivial and, since is connected, is trivial. Thus for all . Thus for all and all . Since generates , we see that is commutative. ∎
We now discuss the representation theory of the unitary group , which is isomorphic to , the circle group. Its Lie algebra is with trivial Lie bracket, which is isomorphic to .
All irreducible finite-dimensional representations of are one-dimensional. They are given by
for .
It follows from Schur’s lemma that all irreducible finite-dimensional representations of are one-dimensional, so are homomorphisms . Since is connected, such a homomorphism is determined by the derivative , which has the form for some . As in Example 1.52, this exponentiates to a map if and only if for some , giving the homomorphism . ∎
All representations of are unitary.
All representations of are completely reducible, that is, decompose into irreducible representations.
Exercise (see Maschke’s theorem below).
This follows from (1) as in the proof of Maschke’s theorem for finite groups. We sketch another proof. So take , a finite dimensional representation of . Consider its differential . Let . We may write for some strictly upper triangular matrix and diagonal matrix such that and commute. Then
On the other hand,
since and commute. It follows that and (why?), whence (why?).
∎
Complete irreducibility does not hold for representations of a general Lie group. For example, the standard representation of
does not decompose into a direct sum of two one-dimensional invariant subspaces (otherwise we would diagonalize , which is impossible). Furthermore, it is not unitary (as unitary matrices are diagonalizable).
Some of this actually generalizes substantially:
(Maschke’s theorem for compact Lie groups) Let be a compact Lie group.
Any finite-dimensional representation of is unitarizable.
(complete reducibility) Any finite-dimensional representation of is a direct sum of irreducible representations.
The second part is proved exactly as for finite groups using the averaging technique. Let be a finite-dimensional representation and let be a subrepresentation. Let be the -invariant Hermitian inner product on guaranteed by the first part. Then the orthogonal complement is also a subrepresentation, and . Iterating, we obtain that is a direct sum of irreducible representations.
The proof of the first part also uses the same idea as for finite groups. Take to be any Hermitian inner product on . Then define
This is also a Hermitian inner product, and
| (putting ) | ||||
| (since ) | ||||
The challenge here is to show that there is an appropriate notion of for which the step “” is valid — this goes by the name of ‘existence of Haar measure’. For you can do it by hand, see problem 18. ∎