Skip to main content

Section 4.1 Lie groups

In the first few sections we discussed at length finite groups and their representations. But in the beginning we also introduced examples of infinite and continuous groups. In this section we go back to the study of continuous groups, and focus on those that also satisfy a differentiability criterion, known as Lie groups. But in fact, all examples of continuous groups that we have seen previously are also examples of Lie groups.

Subsection 4.1.1 Continuous (topological) groups and Lie groups: abstract definition

Let us go back to discrete groups for a moment. What is a discrete group? It is a discrete set \(G\text{,}\) with a binary operation (that we call group multiplication) satisfying a bunch of axioms. In particular, the binary operation closes, so we can think of it as a mapping \(\alpha: G \times G \to G\text{,}\) given by \(\alpha: (a,b) \mapsto a b\text{,}\) where \(a,b \in G\) and on the right-hand-side we mean group multiplication of \(a\) and \(b\text{.}\) We can also define another mapping \(\beta: G \to G\text{,}\) which maps elements to their inverses: \(\beta: g \mapsto g^{-1}\) for \(g \in G \text{.}\) So we encode the structure of a group in terms of the mappings \(\alpha\) and \(\beta\text{.}\)

Now, what is a continuous group? The idea is to replace the discrete set \(G\) by a continuous space \(M\) of a certain dimension, which we call the dimension of the continuous group. More precisely, to define a continuous group we need to define the notion of continuity. The precise statement is that we add to the set \(M\) the extra structure of a topology (i.e. a notion of open sets, or “proximity”), so that it becomes a topological space. Then, we require that the mappings \(\alpha\) and \(\beta\) be continuous with respect to this topology. This defines the concept of a topological group, which is what we really mean by “continuous group”.

The idea of Lie groups is to impose the extra structure of differentiability. To define what it means to be “differentiable”, we introduce an extra structure on the set \(M\text{:}\) we require that \(M\) be a smooth (i.e. \(C^\infty\)) manifold. Then, we require that the mappings \(\alpha: M \times M \to M\) and \(\beta: M \to M\) be smooth maps.

Definition 4.1.1. Lie groups.

A Lie group is a real smooth (\(C^\infty\)) manifold \(M\) that is also a group, in which the group operations of multiplication \(\alpha: M \times M \to M\) and inversion \(\beta: M \to M\) are smooth maps.

This is a nice, abstract definition, but it requires an understanding of manifolds and differential geometry. Fortunately, as we will see in practice we will not need to know much about differential geometry, as the Lie groups that we will work with will all be realized as groups of matrices that depend continuously on a set of parameters. But let us not get ahead of ourselves: let us first work through a few simple examples.

Consider the set \(\mathbb{R}^*\text{,}\) with binary operation given by multiplication. It is easy to see that it satisfies the axioms of a group. Moreover, \(\mathbb{R}^*\) is a smooth manifold, and the group operations of multiplication and inversion are given by the mappings \(\alpha: (a,b) \mapsto a b\) and \(\beta: a \to \frac{1}{a}\text{,}\) for \(a,b \in \mathbb{R}^*\text{,}\) which are smooth maps. Thus \(\mathbb{R}^*\) is a one-dimensional Lie group.

Consider the one-dimensional circle \(S^1\text{,}\) which can be parameterized by an angle \(\theta \in [0,2 \pi)\text{.}\) This is a group under the operation given by addition of angles. Again, \(S^1\) is a smooth manifold, and the group operations, given by \(\alpha: (\theta_1, \theta_2) \mapsto \theta_1 + \theta_2 \text{ (mod }2 \pi \text{)}\) and \(\beta: \theta_1 \mapsto - \theta_1 \text{ (mod }2 \pi \text{)}\) are smooth maps. Thus \(S^1\) is a one-dimensional Lie group.

Let us now consider the set of matrices of the form

\begin{equation*} T = \begin{pmatrix} a \amp b \\ 0 \amp 1 \end{pmatrix}, \end{equation*}

with \(a \in \mathbb{R}_{>0}\) and \(b \in \mathbb{R}\text{.}\) This set forms a group under matrix multiplication. Its underlying manifold is given by \(\mathbb{R}_{>0} \times \mathbb{R}\text{,}\) which is a smooth manifold. Using matrix multiplication, we see that

\begin{equation*} T_1 T_2 = \begin{pmatrix} a_1 \amp b_1 \\ 0 \amp 1 \end{pmatrix} \begin{pmatrix} a_2 \amp b_2 \\ 0 \amp 1 \end{pmatrix} = \begin{pmatrix} a_1 a_2 \amp a_1 b_2 + b_1 \\ 0 \amp 1 \end{pmatrix}, \end{equation*}

and hence the group operation on \(\mathbb{R}_{>0} \times \mathbb{R}\) can be written as:

\begin{equation*} \alpha: ( (a_1, b_1), (a_2, b_2) ) \mapsto (a_1 a_2, a_1 b_2 + b_1), \end{equation*}

which is smooth. As for inverses, given a matrix

\begin{equation*} T = \begin{pmatrix} a \amp b \\ 0 \amp 1 \end{pmatrix}, \end{equation*}

its inverse is

\begin{equation*} T = \begin{pmatrix} \frac{1}{a} \amp - \frac{b}{a} \\ 0 \amp 1 \end{pmatrix}. \end{equation*}

Thus the group inversion on \(\mathbb{R}_{>0} \times \mathbb{R}\) can be written as:

\begin{equation*} \beta: (a, b) \mapsto \left( \frac{1}{a}, - \frac{b}{a} \right), \end{equation*}

which is also differentiable since \(a \in \mathbb{R}_{>0}\text{.}\) Therefore, this is a two-dimensional Lie group.

Now that we have endowed our groups with the extra structure of a topology, we can define what it means for a group to be compact.

Definition 4.1.5. Compact groups.

We say that a topological group is compact if its topology is compact.

Of the three examples that we saw above, only the circle group \(S^1\) is compact.

Subsection 4.1.2 Matrix Lie groups

The abstract definition of Lie groups as groups that are also smooth manifolds is nice, but in practice it is a bit too abstract for our purposes. All the Lie groups that we will study in this class have very explicit realizations as matrix groups. To study those, we need to introduce the mother of all Lie groups: the general linear group, which we have already encountered many times.

Consider the group \(GL(n,\mathbb{R})\) of real \(n \times n\) invertible matrices. Real invertible matrices have \(n^2\) real independent entries. One can show that this is a Lie group of dimension \(n^2\text{.}\) The details of the proof are beyond the scope of this class, but let us briefly sketch the argument. Let \(M_n(\mathbb{R})\) be the set of \(n \times n\) matrices. Then the subset of real invertible matrices \(GL(n,\mathbb{R})\) is obtained by requiring that the determinant is non-zero. But since the determinant is a polynomial map, one can show that the subset \(GL(n,\mathbb{R})\) forms an open set in \(M_n(\mathbb{R})\) (with respect to the Zarisky topology). Using this, one can conclude that \(GL(n,\mathbb{R})\) is a Lie group of dimension \(n^2\text{.}\)

Using a similar line of reasoning, one can show that \(GL(n,\mathbb{C})\) is also a Lie group, with real dimension \(2 n^2\) (since complex invertible matrices have \(n^2\) independent complex entries, that is, \(2 n^2\) independent real parameters).

With this example, we can construct many Lie groups, because of the following crucial theorem, known as the “closed subgroup theorem”:

We will not prove this theorem here, as it is beyond the scope of this class. But the key is that it implies that all our favourite continuous groups are Lie groups.

The group \(SL(n,\mathbb{R})\) of real matrices with determinant one is a closed subgroup of \(GL(n,\mathbb{R})\text{,}\) and hence a Lie group. Imposing that the determinant is equal to one removes one degree of freedom, and hence the dimension of \(SL(n,\mathbb{R})\) is \(n^2-1\text{.}\)

Similarly, the group \(SL(n,\mathbb{C})\) of complex-valued matrices with determinant one is a closed subgroup of \(GL(n,\mathbb{C})\text{,}\) and hence a Lie group. Again, imposing that the determinant is equal to one fixes one constraint, but in this case it reduces the number of complex degrees of freedom by one, and hence the number of real degrees of freedom by two. So the real dimension of \(SL(n,\mathbb{C})\) is \(2 (n^2-1)\text{.}\)

Note that \(SL(n,\mathbb{R})\) and \(SL(n,\mathbb{C})\) are non-compact.

The group \(O(n)\) of real orthogonal matrices is a closed subgroup of \(GL(n,\mathbb{R})\text{,}\) and hence a Lie group. The number of independent real parameters in an \(n \times n\) orthogonal matrix can be calculated to be \(\frac{1}{2} n (n-1)\text{,}\) which is the dimension of the Lie group.

The group \(SO(n)\) of special orthogonal matrices (with determinant one) is also a closed subgroup of \(GL(n,\mathbb{R})\text{,}\) and hence a Lie group. In this case, imposing that the determinant of an orthogonal matrix \(A\) is equal to one is a discrete condition, as any orthogonal matrix must satisfy \(\det A = \pm 1\) to start with. Thus it does not reduce the number of degrees of freedom, and hence the dimension of \(SO(n)\) is also \(\frac{1}{2} n(n-1)\text{.}\)

The orthogonal groups \(O(n)\) and special orthogonal groups \(SO(n)\) are compact Lie groups.

The group \(U(n)\) of unitary matrices is a closed subgroup of \(GL(n,\mathbb{C})\text{,}\) and hence a Lie group. The number of real independent parameters in an \(n \times n\) unitary matrix can be calculated to be \(n^2\text{,}\) which is the dimension of \(U(n)\text{.}\)

The special unitary group \(SU(n)\) is also a closed subgroup of \(GL(n,\mathbb{C})\text{,}\) and hence a Lie group. Imposing the condition that the determinant is equal to one reduces the number of real degrees of freedom by one. Indeed, any unitary matrix \(A\) has \(|\det A|^2 = 1\text{,}\) which means that its determinant can be written as \(\det A = e^{i \theta}\) for some real number \(\theta\text{.}\) Imposing that \(\det A =1 \) fixes this real parameter \(\theta\text{.}\) As a result, the real dimension of \(SU(n)\) is \(n^2 - 1\text{.}\)

Just as for orthogonal groups, both \(U(n)\) and \(SU(n)\) are compact Lie groups.

In this course we will focus on matrix Lie groups, such as the orthogonal and unitary groups, which are realized as closed subgroups of the general linear groups. This is why we will rarely need to go back to the formal abstract definition of Lie groups.

Subsection 4.1.3 Lie groups as groups of transformations of a target space

We first introduced Lie groups abstractly, as groups that are also smooth manifolds. We then studied an important class of Lie groups, known as matrix Lie groups, that are realized as closed subgroups of the general linear groups. But concretely, we often think of Lie groups in a third way, as groups of transformations of a target space. For instance, we often think of \(SO(n)\) as the group of rotations in \(\mathbb{R}^n\text{.}\) Let us make these statements a little more precise.

Suppose that we are interested in a group of continuous transformations of some space, which we call the target space \(T\text{:}\)

\begin{equation*} L: T \to T. \end{equation*}

If the target space has dimension \(d\text{,}\) and we use coordinates \(x_1,\ldots,x_d\) on \(T\text{,}\) we can write these transformations as:

\begin{equation*} x_i' = g_i(x_1,\ldots,x_d; a_1, \ldots, a_n), \qquad i=1,\ldots,d, \end{equation*}

where the \(a_1,\ldots,a_n\) are parameters for the continuous transformations. From this point of view, the abstract continuous group is the parameter space of the transformations, which is parameterized by the \(a_1, \ldots, a_n\text{.}\) In other words, to each choice of real numbers \(a_1, \ldots, a_n\) in a certain set \(M\text{,}\) we assign a corresponding continuous transformation of the target space as above.

From the point of view of transformations, the group operation consists in composition of transformations. How does this relate to the group operation on parameter space? Given two transformations of the target space, corresponding to two points in parameter space, composing the transformations gives a new transformation of the target space, which is assigned to a new point in parameter space. Thus, from the point of view of the parameter space, composing transformation of the target space gives a mapping \(\alpha: M \times M \to M \) on parameter space, which specifies the group operation. If \(M\) is a smooth manifold, we recover the structure of Lie group.

Remark 4.1.11.

It is important however to distinguish between the dimension of the target space (\(d\)), and the dimension of the Lie group, which is the number of parameters \(n\) on which the transformations depend. In other words, the dimension of the Lie group is the dimension of the parameter space \(M\text{,}\) which is the underlying smooth manifold of the Lie group.

All the Lie groups that we have encountered above can be understood as groups of transformations.

Let \(x \in \mathbb{R}\text{,}\) and consider the group of rescalings

\begin{equation*} x' = a x \end{equation*}

parameterized by a non-zero real number \(a \in \mathbb{R}^*\text{.}\) These transformations form an abelian group, with the operation being composition. The parameter space is \(\mathbb{R}^*\text{.}\) Composing two transformations with parameters \(a\) and \(b\text{,}\) we get:

\begin{equation*} x'' = a x' = a b x, \end{equation*}

and hence the group operation on parameter space is the mapping \(\alpha: \mathbb{R}^* \times \mathbb{R}^* \to \mathbb{R}^*\) given by

\begin{equation*} \alpha: (a,b) \mapsto a b. \end{equation*}

The underlying parameter space is thus the group \(\mathbb{R}^*\) with operation given by multiplication, which is a one-dimensional Lie group. It is Example 4.1.2.

As a variation of the previous example, we consider the group of transformations:

\begin{equation*} x' = a_1 x + b_1, \end{equation*}

with \(a_1 \in \mathbb{R}_{>0}\) and \(b_1 \in \mathbb{R}\text{.}\) These transformations form a non-abelian group. Its parameter space is two-dimensional, and given by \(M=\mathbb{R}_{>0} \times \mathbb{R}\text{.}\) Composing two transformations with parameters \((a_1,b_1)\) and \((a_2, b_2)\text{,}\) we get:

\begin{equation*} x'' = a_1 x' + b_1 = a_1 (a_2 x + b_2) + b_1 = a_1 a_2 x + (a_1 b_2 + b_1). \end{equation*}

So the group operation corresponds to the mapping \(\alpha: M \times M \to M\) on parameter space given by:

\begin{equation} \alpha: ((a_1, b_1), (a_2, b_2)) \mapsto (a_1 a_2, a_1 b_2 + b_1).\label{equation-mapping-2d}\tag{4.1.1} \end{equation}

This is nothing else than the group operation on the two-dimensional Lie group constructed in Example 4.1.4.

We also introduced matrix Lie groups above as groups of matrices. But it is also customary to think of them as groups of transformations of target space.

Let the target space be an \(n\)-dimensional real vector space \(V\text{.}\) The linear invertible transformations \(L: V \to V\) form the general linear group \(GL(V)\text{,}\) which we have already studied. If \(V = \mathbb{R}^n\text{,}\) then we get the general linear group \(GL(n,\mathbb{R})\text{.}\) Note that, while the dimension of the target space is \(n\text{,}\) the dimension of the Lie group (which is the dimension of the parameter space for \(n \times n\) real invertible matrices) is \(n^2\text{.}\)

If we choose the complex vector space \(V = \mathbb{C}^n\) for target space, then we get the group \(GL(n,\mathbb{C})\text{.}\) The real dimension of the target space here is \(2 n\text{,}\) while the dimension of the Lie group is \(2 n^2\text{.}\)

Remark 4.1.15.

When we think of Lie groups as groups of transformations, or as groups of matrices, the underlying manifold structure of the parameter space many not be obvious. We will see that in the following examples.

One can restrict to general linear transformations in \(\mathbb{R}^n\) that preserve length and angles. Those are given by orthogonal transformations, and form the Lie group \(O(n)\text{,}\) studied in Example 4.1.9. One can also think of orthogonal transformations as symmetries of a real \(n\)-dimensional sphere, since they preserve its defining equation

\begin{equation*} x_1^2 + x_2^2+ \ldots + x_{n+1}^2 = R^2. \end{equation*}

Note that while the dimension of the target space is \(n\text{,}\) the dimension of the Lie group \(O(n)\) is \(\frac{1}{2} n (n-1)\text{,}\) as orthogonal \(n \times n\) matrices depend on \(\frac{1}{2}n (n-1)\) real independent parameters.

We can also further restrict to special orthogonal transformations, which form the group \(SO(n)\text{.}\) These consists in all elements of \(O(n)\) that are continuously connected to the identity. Geometrically, it can be understood as the group of rotations in \(\mathbb{R}^n\text{.}\)

We note that the underlying manifold structure of \(SO(n)\) (that is, of its parameter space) is not obvious to see in general. But for \(SO(2)\) it is clear. Rotations in two dimensions are parameterized in terms of a single angle \(\theta \in [0,2 \pi)\text{.}\) Thus the parameter space is the circle \(S^1\text{,}\) with the group operation given by addition of angles. We recover the one-dimensional Lie group of Example 4.1.3! So instead of thinking of \(SO(2)\) as the group of \(2 \times 2\) special orthogonal matrices, or as the group of rotations in \(\mathbb{R}^2\text{,}\) we can also think of it as the circle \(S^1\) with group operation given by addition of angles. This also makes it clear that it is a compact group, since the circle is compact.

For \(SO(3)\text{,}\) things are already more complicated. Rotations in \(\mathbb{R}^3\) are parameterized by three angles. So its parameter space is three-dimensional, which is indeed the dimension of the Lie group. But if you remember how to parameterize rotations in three dimensions with angles, it is not so straightforward. It turns out that the manifold structure of the parameter space here is \(\mathbb{R} \mathbb{P}^3\text{,}\) the real projective three-dimensional space. This example already higlights the fact that finding the manifold structure of the underlying parameter space of a matrix Lie group is not in general straightforward.

The unitary group \(U(n)\text{,}\) studied in Example 4.1.10 can be realized as the group of transformations in \(\mathbb{C}^n\) that preserve the inner product. One could also think \(U(n)\) as the group of symmetries of a “complex \(n\)-dimensional sphere”.

As for the the orthogonal group, the underlying manifold structure of \(U(n)\) and \(SU(n)\) is not so obvious to see. But for \(SU(2)\) we can find the parameter space directly. Any \(2 \times 2\) special unitary matrix \(A\) can be parametrized as:

\begin{equation*} A = \begin{pmatrix} \alpha \amp - \beta^* \\ \beta \amp \alpha^* \end{pmatrix} \end{equation*}

for \(\alpha, \beta \in \mathbb{C}\) such that \(|\alpha|^2 + |\beta|^2 = 1\text{.}\) Writing \(\alpha = a + i b\) and \(\beta = c + i d\text{,}\) the equation \(|\alpha|^2 + |\beta|^2 = 1\) becomes

\begin{equation*} a^2 + b^2 + c^2 + d^2 = 1, \end{equation*}

which is the equation of a three-sphere \(S^3\) in \(\mathbb{R}^4\text{.}\) Thus, the parameter space (or underlying manifold structure) of \(SU(2)\) is the three-sphere \(S^3\text{,}\) which is a compact manifold.

Remark 4.1.18.

When we think of Lie groups as groups of transformations, or equivalently as subgroups of \(GL(n,\mathbb{C})\text{,}\) formally what we are doing is constructing a representation of the underlying abstract group. Indeed, to every group element (i.e. points in parameter space), we assign a unique matrix (or a transformation of the target space). So we are building an explicit group homomorphism between the parameter space and \(GL(n,\mathbb{C})\text{.}\) We call this representation the fundamental (or defining) representation of the matrix Lie group.

For instance, the interpretation of \(SO(2)\) as \(2 \times 2\) special orthogonal matrices is the fundamental representation of the abstract Lie group, which is isomorphic to the circle \(S^1\text{.}\) Similarly, the interpretation of \(SU(2)\) as \(2 \times 2\) special unitary matrices is the fundamental representation of the abstract Lie group, which is isomorphic to the three-sphere \(S^3\text{.}\)