This blog post provides mostly some motivation, basic definitions and examples for group representations, up to Maschke’s theorem. Only familiarity with linear algebra and elementary group theory is required for understanding the main part of this post. However, there are some examples for readers with more background which can be safely ignored. The same is true for all categorical remarks.
This will be the first part of a series.
Introduction
The reason to care about groups is because they act on objects. Group actions arise in many different contexts and can provide insight (into the objects as well as into the groups which act on them).
The basic definition of a group action is an action on a set. A set can be thought of as an object without any structure except size (i.e. cardinality). For finite sets, this is more structure than it might seem: we can use counting and divisibility arguments which leads to results such as strong results on p-groups such as the Sylow theorems, the easy-to-prove but ubiquitous orbit-stabilizer theorem or Burnside’s lemma, which has nontrivial combinatorial implications.
Group actions are everywhere in pure mathematics and frequently the object which is acted upon has more structure than just being a set, e.g. a topological space. In these situations, the natural thing to investigate (or to require, if you’re writing a definition) is some compatibility between the group action and the structure on the object. In the example of a topological space, one would require the action to be continuous.
A common structure are vector spaces. Their usefulness seems to stem from the fact that they are both very well understood and still give one a lot of tools to work with: we have duals, tensor products, traces, determinants, eigenvalues etc. Thus, a technique that is used sometimes is “linearization” in which one tries to reduce a problem to linear algebra or at least gain insight by using linear-algebraic methods. Examples are the tangent space of a smooth manifold or variety (and for smooth maps the derivative) or the linearization of a nonlinear ODE.
Representations of groups can be thought of as a linearization of group theory, or more precisely as a linearization of group actions: One can be define them as actions of groups on vector spaces that respect the linear structure. They arise with an ubiquity comparable to (nonlinear, merely “set-theoretic”) group actions. When they do, they are arguably even more useful, since vector spaces have a lot more structure than sets, as described in the previous paragraph.
Group Representations
Definition 1.1 Let be a field. Then for a group
, a (
-linear) representation on a
-vector space
is a map
that satisfies:
where
is the neutral element.
Definition 1.2 In the above situation, the degree or dimension of the representation is the dimension of over
.
Note that the first two axioms state that a representation is a group action and the other two axioms state that for any fixed the map
is
-linear. This map is also invertible, since the axioms imply that
is an inverse. By the second axiom, we also get that the map
is a group homomorphism. So by using a currying argument, we have seen that every group representation gives rise to a group homomorphism
.
Conversely, given a group homomorphism , we can uncurry to get a representation on
by setting
Thus we have another characterization, which allows us to apply concepts defined for group homomorphisms like kernel/image etc, whereas the first characterization allows us to apply notions defined for group actions.
While we’re at it, we may as well add a rephrasing in categorical language of the last characterization and get the following
Lemma 1.3 Let be a field, then for a group
and a vector space
over
, the following data are equivalent:
- A representation of
on
as in definition 1.1
- A group homomorphism
- A covariant functor from
considered as a one-object category to the category
of
-vector spaces that sends the single object in
to
This is the direct analog of equivalent characterizations of group actions: we can also view them as homomorphisms to the group of permutations of a set or as functors from a group to the category of sets.
(Viewing a group as a category works like this: Let be a group, then define a category
with a single object
and set
. Composition
, i.e.
is just group multiplication. Associativity and having an identity follows from the group axioms.)
The last characterization allows one to apply constructions from category theory, such as composing representations with other functors, but we will not use it in this post except as an alternative descrition.
Remark 1.4 One can also consider representations of monoids or the case where is any ring and
is a module.
Remark 1.5 We have defined only left representations. Reversing the chirality in the definitions is straightforward and gives rise to the notion of right representations.
Now for some examples.
As a zeroth example, note that representations of the trivial group are just vector spaces.
Example 1.6 Let be a finite cyclic group of order
with a fixed generator
, then a homomorphism from
to any group
is determined by where it sends
and we can send
to precisely those elements
such that
, so a representation of
is just a linear automorphism that satisfies this equation.
For , one can take for example a rotation matrix
to define a two-dimensional representation of
.
For one can take
to get a one-dimensional representation of
. Both representations correspond to having
act by a counterclockwise rotation of
degrees. The former can be obtained from the latter by restricting scalars from
to
.
Example 1.7 Let , the additive group of
, then
acts on
via transvections:
Example 1.8 If is a
-set (a set equipped with an action from
), then we can consider a vector space
such that the basis elements
are indexed by
. We can then define the action on the basis elements by setting
. This permutation of the basis elements extends uniquely to a linear automorphism of
and we get a respresentation, called the permutation representation associated to the group action.
From a categorical standpoint, if we view -sets as functors
and representations as functors
, then this construction is just composing a group action with the free module functor
. (Thus this construction is also functorial with the notion of morphisms of representations to be defined later.)
Example 1.9 Let be any
vector space. Let
. Then the identity
defines a representation. This corresponds to the natural action of
on
that comes from the definition of
as
-linear automorphisms.
Example 1.10 Generalizing the last example, all classical matrix groups such as ,
,
,
etc. are defined as subgroups of some general linear group, so that the subgroup inclusion defines a representation.
Example 1.11 Let be a finite group and let
be a prime number. Suppose
is a normal abelian subgroup of exponent
. Then
is a
vector space and the conjugation action of
on
is
-linear (this is automatic: any group homomorphism between vector spaces over
is
-linear.), thus we obtain a
-linear representation.
Example 1.12 Let be a Galois extension. Then
acts by definition on
by
-linear field automorphisms. We can just forget the “field automorphism” part and consider
just as a
-vector space, then we get a representation
.
If and
are number fields with rings of integers
and
and
is a non-zero prime ideal in
, then then
acts on
, giving a
-linear representation.
Example 1.13 Let and let
be any vector space over
, then
acts on
where we tensor
copies of
. Then
acts on
by permuting the components of the tensors. Explicitly, if
is an elementary tensor, then we can define the result of the action of
on that to be
.
Example 1.14 Let be a finite-dimensional
-algebra, then the group of units
acts on
by conjugation and this action is
-linear, and thus we obtain a representation of the unit group on the underlying vector space of the algebra. (If
and
, then a suitable restriction of the domain and codomain of this representation gives a description of the Hopf fibration.)
Example 1.15 In the same spirit as the last example, let be an algebraic group over
. Let
be the Lie algebra of
. For any
, the conjugation map
is a smooth automorphism of
, so we can take the derivative at the identity and get a linear automorphism
. The map
is a representation, called the adjoint representation of
. (The same construction works verbatim for Lie groups.)
Example 1.16 Let be a smooth connected manifold and let
be a vector bundle with a flat connection. Let
be a base point and set
and
. If we take a smooth loop
based at
, parallel transport along that loop defines an automorphism of
.
The flatness condition implies that this automorphism depends only on the homotopy class of and by smooth approximation, every homotopy class of continuous loops may be represented by a smooth loop, thus we obtain the holonomy representation
. It turns out that this representation uniquely determines the flat bundle.
As is common practice with group actions, if is a representation, we also write just
for
or
for the map
. By further abuse of notation, we will also just call
a representation of
where the action is clear from the context.
Definition 1.17 If and
are
-linear representations of a group
for some field
, then a morphism of representations (also called intertwining operator) from
to
is a
-linear map
such that
. (i.e.
is
-equivariant.)
Note that if we consider representations as functors, then a morphism of representations is just a natural transformation. Indeed, for any , naturality with respect to
as a morphism is precisely the requirement that
for all
.
Example 1.17 In the situation of example 1.13, let , then we can define
by acting on each factor:
for an elementary tensor. Since we act in the same way in each component, this commutes with permutation of the factors, thus
defines a morphism of the representation of
given by permuting the factors in the tensor product.
Definition 1.18 If is a representation of
, then a subspace
that is
-invariant (i.e.
for all
) defines again a representation of
. These subspaces are called subrepresentations of
.
If is a subrepresentation of
, then the inclusion is a morphism of representations, which gives a (quite general) family of examples for morphisms.
Example 1.19 In the situation of example 1.7, consider the subspace of spanned by
, this is a subrepresentation because
.
Example 1.20 Given a morphism of representations, the kernel and the image are subrepresentations of the domain and codomain, respectively.
Example 1.21 In the situation of example 1.8, suppose that is a sub
-set, i.e. we have
for all
, then
is itself a
-set and if we apply the same construction to
, the resulting vector space is a subspace of
in a canonical way, and so also a subrepresentation. (This is a special case of the mentioned functoriality of this construction.)
If is finite, another subrepresentaion is given by the span of
.
We now come to the first substantial theorem about representations.
Theorem 1.22 (Maschke) Let be a finite group and suppose that the order
is invertible in
. Then if
is a finite-dimensional representation and
is a subrepresentation, then there exists another subrepresentation
such that
,.
Proof By linear algebra, we can find a -linear projection
, i.e. we have that
and
is the identity on
. We have that
, but of course,
will not be a morphism of representations in general. The idea is to “average”
to get another projection onto
that is a morphism of representations.
Set (Here we use that
is invertible in
). This will be
-linear again. This is a morphism of representations, as for
, we have
. Since
is a subrepresentation and
is the identity on
,
is also the identity on
(it is crucial that we divided by
for this step.) and the image is also contained in
, so
is still a projection onto
.
Therefore, the kernel is a complement of and as
is a morphism of representations, the kernel is a subrepresentation.
Example 1.23 To show that the assumptions in Maschke’s theorem are necessary, consider the transvection representation of the additive group of on
described in example 1.7 and 1.19. Here
acts via
. As described in example 1.19, the subspace
of vectors in
with second component
is a subrepresentation.
But this subrepresentation doesn’t have a complement that is also a subrepresentation: Indeed, if is any vector in
such that
, then
and
are linearly independent, as they are clearly not multiples of each other. Thus any subrepresentation that is not contained in
is the whole of
, so
doesn’t have a complement.
This serves as a counterexample in two different ways: if we take to be a finite field, it shows that the assumption that the order is invertible is necessary. If we take
to be an infinite field (say of characteristic
), then it shows that even in characteristic
, the conclusion doesn’t need to hold when the group is infinite.