Group Theory For Maths, Physics and Chemistry Students: Arjeh Cohen Rosane Ushirobira Jan Draisma October 18, 2007
Group Theory For Maths, Physics and Chemistry Students: Arjeh Cohen Rosane Ushirobira Jan Draisma October 18, 2007
students
Arjeh Cohen
Rosane Ushirobira
Jan Draisma
1 Introduction 5
1.1 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Basic notions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 Permutation Groups 17
2.1 Cosets and Lagrange . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Quotient groups and the homomorphism theorem . . . . . . . . . 18
2.3 Loyd’s puzzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.4 Action of groups on sets . . . . . . . . . . . . . . . . . . . . . . . 28
4 Representation Theory 51
4.1 Linear representations of groups . . . . . . . . . . . . . . . . . . 51
4.2 Decomposing Displacements . . . . . . . . . . . . . . . . . . . . . 59
5 Character Tables 63
5.1 Characters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 Orthogonality of irreducible characters . . . . . . . . . . . . . . . 64
5.3 Character tables . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4 Application: symmetry of vibrations . . . . . . . . . . . . . . . . 77
5.5 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3
4 CONTENTS
Chapter 1
Introduction
1.1 Symmetry
Group theory is an abstraction of symmetry
Symmetry is the notion that an object of study may look the same from different
points of view. For instance, the chair in Figure 1.1 looks the same as its
reflection in a mirror that would be placed in front of it, and our view on the
wheel depicted next to the chair doesn’t change if we rotate our point of view
over π/6 around the shaft.
But rather than changing viewpoint ourselves, we think of an object’s sym-
metry as transformations of space that map the object ‘into itself’.
What do we mean when we say that an object is symmetric? To answer this
question, consider once more the chair in Figure 1.1. In this picture we see a
plane V cutting the chair into two parts. Consider the transformation r of three-
dimensional space that maps each point p to the point p0 constructed as follows:
5
6 CHAPTER 1. INTRODUCTION
• when designing a chair, the requirement that the object be symmetric may
reduce drawings to half the chair;
Symmetry conditions
A few more observations can be made regarding the previous cases.
Groups
The transformations under which a given object is invariant, form a group.
Group theory was inspired by these types of group. However, as we shall see,
‘group’ is a more general concept. To get a feeling for groups, let us consider
some more examples.
Planar groups
The hexagon, as depicted in Figure 1.2, is a two-dimensional object, lying in
the plane. There are lots of transformations of the plane leaving it invariant.
For example, rotation r around 0 over 2π/6 is one of them, as is reflection s in
the vertical line l through the barycentre. Therefore, the hexagon has at least
8 CHAPTER 1. INTRODUCTION
In fact, these are all symmetries of the hexagon. Thus, each is a (multiple)
product of r and s. These two elements are said to generate the group of
symmetries.
Exercise 1.1.1. Verify that the elements listed are indeed distinct and that
r6 = s2 = e, and srs = r5 .
We find that the symmetry group of the pattern is infinite. However, when
ignoring the translational part of the symmetries, we are left with essentially
only two symmetries: e and r.
Exercise 1.1.3. Find frieze patterns that have essentially different symmetry
groups from the one of Figure 1.3.
A pattern with a more complicated symmetry group is the wallpaper pattern
of Figure 1.4, which Escher designed for ‘Ontmoeting’ (Dutch for ‘Encounter’).
It should be thought of as infinitely stretched out vertically, as well as hori-
zontally. The translations s and t over the vectors a (vertical from one nose
1.1. SYMMETRY 9
to the next one straight above it) and b (horizontal, from one nose to the next
one to its right) leave the pattern invariant, but this is not what makes the
picture so special. It has another symmetry, namely a glide-reflection g, which
is described as: first apply translation over 21 a, and then reflect in the vertical
line l equidistant to a left oriented and a neighbouring right oriented nose. Note
that g 2 = s.
Space groups
Having seen symmetries of some 2-dimensional figures, we go over to 3-dimen-
sional ones.
The cube in Figure 1.5 has lots of symmetries. Its symmetry group is gen-
erated by the rotations r3 , r4 , r40 whose axes are the coordinate axes, and the
reflection t through the horizontal plane dividing the cube in two equal parts.
Exercise 1.1.5. What is the size of the symmetry group of the cube?
Closely related is the methane molecule, depicted in Figure 1.6. Not all of
the symmetries of the cube are symmetries of the molecule; in fact, only those
that map the hydrogen atoms to other such atoms are; these atoms form a
regular tetrahedron.
Exercise 1.1.6. Which of the symmetries of the cube are symmetries of the
methane molecule?
Finally consider the cubic grid in Figure 1.7. Considering the black balls to
be natrium atoms, and the white ones to be chlorine atoms, it can be seen as the
structure of a salt-crystal. It has translational symmetry in three perpendicular
10 CHAPTER 1. INTRODUCTION
directions along vectors of the same size. To understand its further symmetry,
consider all those symmetries of the crystal that leave a given natrium atom
fixed. Such a symmetry must permute the 6 chlorine atoms closest to the
given natrium atom. These atoms form a regular octahedron, as depicted in
Figure 1.8, and therefore we may say that the symmetry group of salt crystals
is generated by that of the regular octahedron and three translations.
2. Prove that {e} and {e, s0 }, where e denotes the identity map on R, are
the only finite groups of isometries of R.
G = {ta | a ∈ Z} ∪ {t a2 | a ∈ Z}.
4. Conclude that there are only two ‘essentially different’ groups describing
the symmetries of discrete subsets of R having translational symmetry.
12 CHAPTER 1. INTRODUCTION
Exercise 1.1.8. Analyse the symmetry groups of discrete subsets of the strip
R × [−1, 1] in a manner similar to that of Exercise 1.1.7. Compare the result
with your answer to Exercise 1.1.3.
Exercise 1.2.2. Prove that a group G cannot have more than one identity.
Also, the notation g −1 for the inverse of g seems to indicate the uniqueness of
that inverse; prove this.
Let us give some examples of groups.
Example 1.2.3.
1. All sets of transformations found in §1.1 form groups; in each case, the
composition ◦ serves as operation ·. Some of them are Abelian; like the
the chair, and some aren’t, like those of the cube.
2. The real numbers R form a group with respect to addition +, the unit
element being 0. They do not form a group with respect to ordinary
multiplication, as 0 does not have an inverse. Leaving out 0 we do obtain
an Abelian group (R \ {0}, ·), the unit element being 1.
3. Also Q and Z form additive groups. Again Q \ {0} is a multiplicative
group. However, Z cannot be turned into a multiplicative group, the only
invertible elements being {±1}.
4. Let X be a set. The bijections of X form a group. Composition is again the
group operation. The group inverse coincides with the ‘functional’ inverse.
The group is often denoted by Sym(X). If X is finite, say |X| = n, then
|Sym(X)| = n!. If two sets X and Y have the same cardinality, then the
groups Sym(X) and Sym(Y ) are essentially the same. For finite sets of
cardinality n, the representative Sym({1, ..., n}) of this class of groups is
denoted by Sym(n) or Sn .
Its element are also called permutations. For X = {1, 2, 3}, we have
Sym(X) = {e, (1, 2), (2, 3), (1, 3), (1, 2, 3), (3, 2, 1)}.
g · g · ... · g.
| {z }
n
Exercise 1.2.6. Show that the notation for fractions which is usual for integers,
does not work here: hg could stand for g −1 h as well as for hg −1 , and the latter
two may differ.
With the methane molecule in Figure 1.6, some of the symmetries of the
cube form a smaller group, namely the symmetry group of the tetrahedron. We
found an example of a subgroup.
1. H is non-empty.
3. H is closed under taking inverses. This means that, for all h ∈ H, we have
h−1 ∈ H.
Exercise 1.2.8. Prove that a subgroup of a group contains the group’s identity.
Prove that for a subset H of G to be a subgroup of G it is necessary and sufficient
that H be non-empty and that gh−1 ∈ H for all g, h ∈ H.
Generation
The following lemma is needed to formalize the concept of a subgroup generated
by some elements.
(0,0) (1,0)
Permutation Groups
Definition 2.1.1. For subsets S and T of a group G, we write ST for the set
{st | s ∈ S, t ∈ T }.
Example 2.1.3.
17
18 CHAPTER 2. PERMUTATION GROUPS
4. A left coset need not be a right coset. Take for instance G = S3 and
H the cyclic subgroup h(12)i. Then (13)H = {(13), (123)} and H(13) =
{(13), (132)}.
It is easy to verify that left cosets are equivalence classes of the equivalence
relation
g ∼ h ⇔ g −1 h ∈ H,
and similarly for right cosets. So we have the following lemma.
Lemma 2.1.4. The left cosets of a subgroup H in the group G form a partition
of the group G. For g ∈ G, the map h 7→ gh is a bijection between H and the
left coset gH.
Exercise 2.1.5. Show that (13)h(12)i is not a right coset of h(12)i in S3 , and
that (12)h(123)i is a right coset of h(123)i in S3 .
{g ∈ G | φ(g) = e}
φ(G) = {φ(g) | g ∈ G} ⊆ H
and one checks that this space is invariant under M 7→ AM A∗ for any A ∈
GL(2, C). Also, for A ∈ SL(2, C), ϕ(A) preserves the metric due to the multi-
plicative properties of the determinant:
Exercise 2.2.12. Take the group Z/4Z, the subgroup h(1234)i of S4 and the
multiplicative subgroup generated by i in C. Verify that they are all isomorphic.
As we have seen before, it is not always the case that left cosets and right
cosets of a subgroup H are the same. When that happens, we say that H is
normal.
If H is normal, then we can define an operation on the set G/H of all left
cosets of H in H, as follows:
This operation turns G/H into a group, called the quotient group of G by H.
The map G 7→ G/H, g 7→ gH is a homomorphism G → G/H.
2.2. QUOTIENT GROUPS AND THE HOMOMORPHISM THEOREM 21
Exercise 2.2.15. Prove that Inn(G) (cf. Example 2.2.4) is a normal subgroup
of Aut(G) (cf. Definition 2.2.1).
Example 2.2.16.
A B C D
E F G H
I J K L
M N O
A B C D
E F G H
I J K L
M O N
A B C D A B C D A B C D
E F G H E F G H E F G H
I J K L I J K I J K
M O N M O N L M O N L
i.e., π is the product of the 2-cycle (2, 6) and the 3-cycle (3, 5, 1). These
are called disjoint, because each fixes the elements occurring in the other,
and this fact makes them commute. Any permutation can be written as a
product of disjoint cycles, and this factorization is unique up to changes
in the order. In this notation, it is usual not to write down the 1-cycles.
Leaving these out, it is not always clear to which Sn the permutation
belongs, but this abuse of notation turns out to be convenient rather than
confusing.
Permutation matrix notation: consider once again the permutation π ∈ S5 .
With it we associate the 5 × 5-matrix Aπ = (aij ) with
(
1 if π(j) = i,
aij =
0, otherwise.
1 2 3 4 5
1 2 3 4 5
Exercise 2.3.1. Show that the 2-cycles in Sn generate that whole group. Can
you give a smaller set of 2-cycles that generates the whole group? How many
do you need?
sgn π := det Aπ .
Exercise 2.3.6. Prove that the sign of a permutation equals (−1)N where N
is the number of crossings of the lines drawn in the diagram notation of the
permutation.
1 2 3 4
5 6 7 8
9 10 11 12
13 14 15 16
No Return Possible
Let us return to Sam Loyd’s 15-puzzle. It can be described in terms of states
and moves. The Figures 2.1 and 2.2 depict states of the puzzle. Formally, a
state can be described as a function
Here, s(p) is the square occupying position p in the state s and ∗ stands for the
empty square. The positions are numbered as in Figure 2.5.
For p ∈ {1, . . . , 16}, we denote by M (p) ⊆ S16 the set of possible moves in a
state s with s(p) = ∗. A move m ∈ M (p) should be interpreted as moving the
square from position j to position m(j), for j = 1, . . . , 16. For example
Moves can be combined to form more general transformations. The set of all
possible transformations starting in a state with the empty square in position
p is denoted by T (p), and it is defined as the set of all products (in S16 ) of the
form mk mk−1 . . . m1 , where
for all j ≥ 0. This reflects the condition that mj+1 should be a possible move,
after applying the first j moves to s. In particular, in taking k = 0, we see
that the identity is a possible transformation in any state. Also, if t ∈ T (p) and
u ∈ T (t(p)), then u ◦ t ∈ T (p). Finally, we have t−1 ∈ T (t(p)).
Moves and, more generally, transformations change the state of the puzzle.
Let s ∈ S with s(p) = ∗ and let t ∈ T (p). Then t changes the state s to the
2.3. LOYD’S PUZZLE 27
In words, the square at position q after applying t is the same as the square at
position t−1 (q) before that transformation.
Now that we captured Sam Loyd’s 15-puzzle in a comfortable notation, we
can formalize our question as follows: is there a t0 ∈ T (16) that transforms the
messed up state s1 of Figure 2.2 into the initial state s0 of Figure 2.1, i.e., that
satisfies
t0 · s1 = s0 ?
The answer is no, and we shall presently make clear why. In each state, order
the non-empty squares according to the path in Figure 2.6. That is, the j-th
non-empty square one encounters when following this path in a particular state
s is called the j-th square in s, for j = 1, . . . , 15. Now consider any position
p, any t ∈ T (p) and some s ∈ S with s(p) = ∗. Then we define a permutation
πt ∈ S15 by
πu◦t = πu ◦ πt (2.1)
6 7 8
10 9
conclusion that πm is an odd cycle for any move m, hence an even permutation.
As a consequence of this and equation (2.3), a sequence of such moves will also
be mapped to an even permutation by π. Therefore, such a transformation
cannot possibly be (14, 15).
You conclude that your friend must have cheated while you were away mak-
ing coffee, by lifting some of the squares out of the game and putting them back
in a different way.
Exercise 2.3.7. Show that the set {πt | t ∈ T (p) for some p} forms a subgroup
of S15 .
is a subgroup of A16 . Use this to give another argument why your friend must
have cheated.
If the action α is obvious from the context, we leave α out and write gm for
α(g, m). Also, for subsets S ⊂ G and T ⊂ M we write ST for {α(g, m) | g ∈
S, m ∈ T }. For {g}T we write gT and for S{m} we write Sm.
α(g, m) = φ(g)m.
Example 2.4.3.
3. Consider the group of all motions of the plane, generated by all trans-
lations (x, y) 7→ (x + α, y) for α ∈ R in x-direction and all inflations
(x, y) 7→ (x, βy) for β ∈ R∗ . This acts on the solutions of the differential
equation
d
( )2 (y) = −y.
dx
Example 2.4.4. Consider G = S4 and put a = {12, 34}, b = {13, 24}, c =
{14, 23}. Here 12 stands for {1, 2}, etc. There is a natural action (as above) on
the set of pairs, and similarly on the set {a, b, c} of all partitions of {1, 2, 3, 4}
into two parts of size two. Thus, we find a homomorphism S4 → Sym({a, b, c}).
It is surjective, and has kernel a group of order 4. (Write down its nontrivial
elements!)
Example 2.4.5. The group SL(2, C) acts on the complex projective line P1 (C)
as follows. Denote by π : C2 \ {0} → P1 (C) the map sending a pair of homoge-
neous coordinates to the corresponding point. The linear action of SL(2, C) on
C2 by left multiplication permutes the fibers of π, hence induces an action on
P1 (C). On the affine part of P1 (C) where the second homogeneous coordinate
is non-zero, we may normalize it to 1, and the action on the first coordinate x
is given by
a b ax + b
x= .
c d cx + d
Note, however, that this affine part is not invariant under the group.
α(e, g) = ege−1 = g
f : Gm → G/Gm , g · m 7→ gGm
|G| = |Gm||Gm |.
Remark 2.4.12. One must distinguish the elements of G from their actions on
M . More precisely, the map sending g ∈ G to the map α(g, ·) : M → M is a
homomorphism, but need not be an isomorphism.
We now want to make clear how the action of groups can help in classification
and counting problems, and we shall do so by means of the following problem.
2.4. ACTION OF GROUPS ON SETS 31
1 2 1 2 1
2
3
3 4 3 4 4 5
Here X
2 stands for the set of all unordered pairs from X. If the function f
takes the value 1 on such a pair {x, y}, this means that the pair is an edge in
the graph. Often, graphs are associated with a picture like the ones in Figure
2.8. The set of all graphs on X is denoted by Graphs(X).
The first two graphs in Figure 2.8 are essentially the same; to make them
exactly the same one has to relabel the vertices. Such a relabeling is in fact a
permutation of X. Formally, this leads to an action of Sym(X) on Graphs(X)
in the following way. First of all Sym(X) acts on X 2 by
X
This diagram tells us that e ∈ 2 is an edge in the graph f if and only if π · e
is an edge in the graph π · f . In formula,
π · f (x) = f (π −1 x).
32 CHAPTER 2. PERMUTATION GROUPS
Graphs that can be obtained from each other by the action of some π ∈
Sym(X) are called isomorphic. Now our original question can be rephrased
very elegantly as follows.
where v = |X|.
(1 · 26 + 6 · 24 + 8 · 22 + 6 · 22 + 3 · 24 )/24 = 11.
Symmetry Groups in
Euclidean Space
3.1 Motivation
In low dimensions, there are few finite groups of isometries and this enables us,
when given a two- or three-dimensional object, to determine its symmetry group
simply by enumerating all axes of rotation and investigating if the object admits
improper isometries such as inversion and reflections. As an example, suppose
that we want to describe the symmetry group G of the CH4 molecule depicted
in Figure 3.1. The hydrogen atoms are arranged in a regular tetrahedron with
the C-atom in the center.
As a start, let us compute |G|. Consider a fixed H-atom a. Its orbit Ga
consists of all four H-atoms, as can be seen using the rotations of order 3 around
axes through the C-atom and an H-atom. Applying Lagrange’s theorem we find
|G| = 4|Ga |. The stabilizer of a consists of symmetries of the remaining 4 atoms.
Consider one of the other H atoms, say b. Its orbit under the stabilizer Ga has
cardinality 3, and the stabilizer (Ga )b consists of the identity and the reflection
in the plane through a, b and the C-atom, so that we find
33
34 CHAPTER 3. SYMMETRY GROUPS IN EUCLIDEAN SPACE
C
H
H
H
Figure 3.1: The CH4 molecule.
1. There are 4 three-fold axes, namely those through an H-atom and the
center of the opposite side of the tetrahedron.
2. There are three two-fold axes, namely those through the centers of two
perpendicular sides.
Later on, we shall see that this information determines the orientation pre-
serving part of G uniquely; it is denoted by T . From the fact that G does
not contain the inversion, but does contain other improper isometries, we may
conclude that G is T ∪ i(W \ T ), a group that can be described by Figure 3.2.
Here we see a regular tetrahedron inside a cube. Half of the proper isometries
of the cube are isometries of the tetrahedron, and the improper isometries of
the tetrahedron may be obtained as follows: take any proper isometry of the
cube which is not a isometry of the tetrahedron, and compose it with the in-
version i. Note that T ∪ i(W \) and W are isomorphic as abstract groups, but
their realizations in terms of isometries are different in the sense that there is
no isometry mapping one to the other. This point will also be clarified later.
3.2. ISOMETRIES OF N -SPACE 35
Exercise 3.1.1. Consider the reflection in a plane through two H-atoms and
the C-atom. How can it be written as the composition of the inversion and a
proper isometry of the cube?
The finite symmetry group of the tetrahedron and the translations along
each of three mutually orthogonal edges of the cube in Figure 3.2 generate the
symmetry group of a crystal-like structure. For this reason, the group G is
called a crystallographic point group.
The set of all isometries is a subgroup of Sym(Rn ), which we denote by AO(n, R).
The following definition and lemmas will explain the name of this group.
Definition 3.2.2. For a fixed a ∈ Rn , the isometry ta : x 7→ x + a is called the
translation over a.
Lemma 3.2.3. For any g ∈ AO(n, R), there exists a unique pair (a, r) where
a ∈ Rn and r ∈ AO(n, R) fixes 0, such that g = ta r.
Proof. Set a := g(0). Then t−1 −1
a g(0) = 0, so ta g is an isometry fixing 0, and it
is clear that a = g(0) is the only value for which this is the case.
Lemma 3.2.4. Let r ∈ AO(n, R) be an isometry fixing 0. Then r is an orthog-
onal R-linear map.
Proof. Let x, y ∈ Rn , and compute
where we use that r is an isometry fixing 0. Hence, r leaves the inner product
invariant. This proves the second statement, if we take the first statement for
granted. For any z ∈ Rn , we have
(z, rx + ry) = (z, rx) + (z, ry) = (r−1 z, x + y) = (z, r(x + y)),
T (G) := {a ∈ Rn | ta ∈ G};
this is an additive subgroup of Rn , which we shall identify with the group of all
translations in G.
Using the linearity of isometries fixing 0, one can prove the following.
Proposition 3.2.6. For a subgroup G ⊆ AO(n, R), the map R : G → O(n, R)
defined by
g = ta R(g) for some a ∈ Rn
is well-defined, and a group homomorphism. Its kernel is T (G), and this set is
invariant under R(G).
Proof. Lemma 3.2.3 and 3.2.4 show that R is well-defined. Let r1 , r2 ∈ O(n, R)
and a1 , a2 ∈ Rn , and check that
U ∩ V = {p}.
Proof. As 0 is an isolated point, there is an > 0 such that kvk ≥ for all
v ∈ L \ {0}. But then kv − wk ≥ for all distinct v, w ∈ L. Hence, any sequence
(vn )n in L with vn − vn+1 converging to 0, is eventually constant.
k
X
π(L) = Zπ(vi ).
i=2
k
X
π(v) = ci π(vi )
i=2
38 CHAPTER 3. SYMMETRY GROUPS IN EUCLIDEAN SPACE
Pk
with ci ∈ Z. Then v− i=2 ci vk ∈ L is a scalar multiple of v1 , and by minimality
of kvk the scalar is an integer. Hence,
k
X
L= Zvi .
i=1
If G ⊆ AO(n, R) is a group that acts discretely, then T (G) does so, too. In
general, this does not hold for R(G), as the following example shows.
This phenomenon does not occur if we require T (G) to have the maximal
possible rank.
Proof. By Proposition 3.2.6, T (G) is invariant under R(G). Now apply Theorem
3.2.13.
3.2. ISOMETRIES OF N -SPACE 39
This corollary allows for the following approach to the classification of crys-
tallographic groups in O(n, R): first classify all finite subgroups of O(n, R), then
investigate for each of them whether it leaves invariant a lattice of rank n. Fi-
nally, for each pair (R, L) of a crystallographic point group R and a lattice L of
rank n left invariant by R, there may be several non-equivalent groups G with
T (G) = L and R(G) = R. Among them is always the semi-direct product L o R:
the subgroup of AO(n, R) generated by L and R.
In this setting, ‘classification’ means ‘up to equivalence under AO(n, R)’, i.e.,
the groups G and hGh−1 for some h ∈ AO(n, R) are considered the same. This
is a finer equivalence relation than isomorphism, as is shown in the following
example.
Example 3.2.16. The groups C2 and D1 (see Section 3.3) are isomorphic as
abstract groups, but not equivalent under AO(2, R). The same holds for the
groups W and T ∪ i(W \ T ) (see Section 3.4).
Exercise 3.2.17. Show that the symmetry group G of Figure 1.4 is not the
semi-direct product of R(G) and T (G).
In the next sections, we shall carry out this classification in the cases n = 2
and 3, up to the point of determining the crystallographic point groups. The
determinant of an orthogonal map turns out to be a useful tool there.
Definition 3.2.18. An isometry r ∈ O(n, R) is called proper or orientation
preserving if det r = 1. The set of all proper isometries in O(n, R) is denoted
by SO(n, R). An isometry g ∈ AO(n, R) is called proper if R(g) is proper.
Otherwise, it is called improper.
Exercise 3.2.19. Prove that, in any subgroup G ⊆ AO(n, R) containing im-
proper isometries, the proper ones form a subgroup of index 2.
Exercise 3.2.20. Consider the following map:
ra : v 7→ v − 2(v, a)a,
Exercise 3.2.22. Let G be a finite subgroup of the group GL(n, R) of all invert-
ible n × n-matrices. Show that G leaves invariant a positive definite symmetric
bilinear form, and that it is hence conjugate to a subgroup of O(n, R). Hint: let
(., .) denote the standard inner product, and define
1 X
hx, yi := (gx, gy)
|G|
g∈G
for x, y ∈ Rn . Show that h., .i is symmetric and positive definite. Let A be the
matrix with entries aij = hei , ej i, and B ∈ GL(n, R) such that B T B = A. Show
that BGB −1 ⊆ O(n, R).
with respect to the standard basis is an element of SO(2, R), called rotation
around 0 over φ. Conversely, any element of this group equals some rφ . Any
improper isometry in O(2, R) is a reflection in a line, see Exercise 3.2.20.
In fact, these are the only two classes, as we shall now prove. Let G be a
finite subgroup of O(2, R). There are two possibilities:
Exercise 3.3.3. Show that O(n, R) is generated by reflections. Hint: show first
that it is generated by elements that fix an (n − 2)-dimensional subspace of Rn
pointwise, and that O(2, R) is generated by reflections.
Using this classification, we can find which of the finite groups are crystal-
lographic point groups.
C1 , C2 , C3 , C4 , C6 ,
D1 , D2 , D3 , D4 , D6 .
Proof. For the first part, one only needs to construct invariant lattices for D4
and D6 (see 3.3.5), as the other groups are contained in one of these see.
For the second statement, suppose that G is a crystallographic point group.
According to Theorem 3.2.13, G must be finite. The lattice L is invariant under
the subgroup K of O(2, R) generated by G ∪ {i}, where i denotes the inversion.
As the inversion commutes with all elements of G, K is finite (either equal to
G or twice as large). Therefore, according to Theorem 3.3.2, K is conjugate to
Cn or Dn for some n; as K contains the inversion, n must be even. So either
n ≤ 6, in which case G is of one of the types mentioned in the theorem, or
n ≥ 8. Suppose that the latter were the case and choose a vector c in the
lattice of shortest possible length. Now r2π/n c ∈ L and c − r2π/n c ∈ L. A
little calculus shows that the latter vector is shorter than c because the rotation
angle too small (for n = 8, this can be seen from Figure 3.3), and we arrive at
a contradiction.
Note that we have only classified all crystallographic point groups, and not
the crystallographic groups. There are 17 of the latter, and only 10 of the
former; see [2].
42 CHAPTER 3. SYMMETRY GROUPS IN EUCLIDEAN SPACE
c-hc
hc
1. Fix an axis l and consider the proper rotations around l over the angles
2kπ/n for k = 0, . . . , n − 1. They form a group, denoted by Cn , due to its
analogy to the class denoted by Cn in the two-dimensional setting.
2. Start with the group Cn of rotations around l and choose a second axis m
perpendicular to l. Consider the group generated by Cn and the rotation
around m over π. It is twice as large as Cn , and denoted by Dn0 . Note that
it contains only proper isometries. The proper symmetries of an n-prism
form a group of type Dn0 . Note that D10 is conjugate to C2 .
3.4. THE FINITE SUBGROUPS OF O(3, R) 43
Cn for n ≥ 1,
Dn0 for n ≥ 2,
T, W, P.
Count the number N of pairs (g, p), with e 6= g ∈ G and p on the unit sphere,
that satisfy g(p) = p. On one hand, each g 6= e fixes exactly two anti-podal
points, hence this number is
N = 2(|G| − 1).
On the other hand, for each p ∈ S, the number of g 6= e fixing p equals |Gp | − 1.
Hence X
N= (|Gp | − 1).
p∈S
where the sum is taken over all orbits of G in S. Hence, dividing both sides by
|G|, we get
2 X |o|
2− = (1 − ).
|G| o
|G|
44 CHAPTER 3. SYMMETRY GROUPS IN EUCLIDEAN SPACE
As each s ∈ S has non-trivial stabilizer, each orbit in S has size at most |G|/2;
hence each ao is at least 2. Now Exercise 3.4.3 leads to Table 3.1. That is, it
provides the numerical data for it; it remains to check that this data determines
the group up to conjugation.
local b,g,o;
if s >= 2 then return false;
else g:=2/(2-s); o:=List(a,x->g/x);
Print("Orbit sizes:",o," |G|:",g,"\n");
b:=a[Length(a)];
while extend(Concatenation(a,[b]),s+(1-1/b)) do
b:=b+1;
od;
return true;
fi;
end;
Explain why the function call extend([2,2],1) results in an endless loop. On
the other hand, when called as extend([2,3],1/2+2/3), the function does halt,
and prints
Orbit sizes:[ 6/5, 4/5 ] |G|:12/5
Orbit sizes:[ 6, 4, 4 ] |G|:12
Orbit sizes:[ 12, 8, 6 ] |G|:24
Orbit sizes:[ 30, 20, 12 ] |G|:60
3.4. THE FINITE SUBGROUPS OF O(3, R) 45
We do not give all details of the proof, but restrict ourselves to the following
remarks.
1. Note that D10 and C2 describe the same class; for this reason the second
list starts with n = 2.
2. The last group in the list is made possible by the fact that a group of
type W has a subgroup of type T of index 2. This can be seen as follows:
consider the group of proper symmetries of a cube with center 0, and
look at the lines between anti-podal vertices of the cube; they are axes of
rotations of order 3 leaving the cube invariant; these proper rotations form
the symmetry group of a regular tetrahedron with center 0 and points on
those axes.
1. List all rotation axes and their orders. They determine the subgroup H
of G of all proper rotations.
4. Suppose that G contains improper rotations, but not the inversion. Then,
according to the list of Theorem 3.4.5, there is only one possibility for G,
unless H is of type Cn .
Exercise 3.4.6. Consider the molecule depicted in Figure 3.5. Identify the
class of its symmetry group in the list of Theorem 3.4.5.
Exercise 3.4.7. Consider the full symmetry group of the tetrahedron, including
its improper rotations. Which class in the list of Theorem 3.4.5 does it belong
to?
Exercise 3.4.8. Determine the symmetry group of the CH4 molecule depicted
in Figure 3.1.
H H
H
Figure 3.5: The NH3 - molecule.
C
H H
H H
C
H
Figure 3.6: The ethane molecule in staggered configuration.
48 CHAPTER 3. SYMMETRY GROUPS IN EUCLIDEAN SPACE
P
P
N N
N N
P
P
Figure 3.7: The ethane molecule in staggered configuration.
Cm , Cm ∪ iCm m = 1, 2, 3, 4, 6
0 0 0
Dm , Dm ∪ iDm m = 2, 3, 4, 6
Cm ∪ i(C2m \ Cm ) m = 1, 2, 3
0 0 0
Dm ∪ i(D2m \ Dm ) m = 2, 3
0
Cm ∪ i(Dm \ Cm ) m = 2, 3, 4, 6
T, W, T ∪ iT, W ∪ iW, T ∪ i(W \ T )
Representation Theory
The notion
We introduce the concept of a linear representation of G.
Definition 4.1.1. A (linear) representation of a group G on a vector space V
is a homomomorphism
ρ : G → GL(V ).
If ρ is injective, it is called faithful.
After a choice of basis of V , we can view a representation as a homomorphism
G → GL(n, K),
where GL(n, K) is the group of invertible matrices with entries in K; such a map
is called matrix representation of G.
If ρ : G → GL(V ) is a representation of G on V , then we often write gv
instead of ρ(g)v, if no confusion is possible. Also, G is said to act linearly on V ,
and V is called a G-module. The dimension of the representation is by definition
the dimension of V . Note that (g, v) 7→ ρ(g)v is indeed an action of G on V in
the sense of Definition 2.4.1.
Exercise 4.1.2. To experience the use of matrix representations of groups, try
to guess which orthogonal linear map is the product ρx ρy ρz . Here ρx stands
for rotation over π about the x-axis, and similarly for y and z. To check your
guess, represent the rotations by matrices and compute the product.
51
52 CHAPTER 4. REPRESENTATION THEORY
We know that (1, 2)(1, 2, 3) = (2, 3) and this can be seen from the matrices as
well:
1 0 0
M(1,2) M(1,2,3) = 0 0 1 .
0 1 0
Exercise 4.1.5. Prove that the correspondence given above between S3 and
GL(n, K) is a group homomorphism, that is, a matrix representation of S3 .
Example 4.1.6. Recall from 3.3 the dihedral group Dn of order 2n. It contains
the cyclic group Cn of order n with generator c as a subgroup of index 2 and has
a reflection a (of order 2) with aca = c−1 . The group Dn has a two-dimensional
real representation r : Dn → GL(2, R) determined by
k cos(2kπ/n) − sin(2kπ/n) 0 1
c 7→ , a 7→ .
sin(2kπ/n) cos(2kπ/n) 1 0
from which we conclude Ax1 = (ad − bc)−1 (dx1 − bx2 ) and Ax2 = (ad −
bc)−1 (−cx1 + ax2 ). Hence A has matrix
d −c
(ad − bc)−1
−b a
d2 c2
−cd
−2
(ad − bc) −2bd ad + bc −2ac .
b2 −ab a2
1 X
πρ(h) = ρ(g)π 0 ρ(g −1 h);
|G|
g∈G
56 CHAPTER 4. REPRESENTATION THEORY
1 X
ρ(hg)π 0 ρ((hg)−1 h)
|G|
g∈G
1 X
= ρ(h)ρ(g)π 0 ρ(g −1 h−1 h)
|G|
g∈G
1 X
= ρ(h) ρ(g)π 0 ρ(g −1 )
|G|
g∈G
= ρ(h)π,
as claimed. To verify that π is a projection, we first note that ππ 0 = π 0 ; indeed,
both are zero on W 0 and the identity on U . Compute
1 X
π2 = π ρ(g)π 0 ρ(g −1 )
|G|
g∈G
1 X
= ρ(g)ππ 0 ρ(g −1 )
|G|
g∈G
1 X
= ρ(g)π 0 ρ(g −1 )
|G|
g∈G
= π,
where the second equality follows from the fact that π commutes with each
ρ(g). This concludes the proof that π is a projection. This implies that V =
im(π)⊕ker(π), where both im(π) and ker(π) are sub-G-modules of V by Exercise
4.1.15. Now note that im(π) = im(π 0 ) = U , so that we may take W = ker(π)
to conclude the proof of the lemma.
Remark 4.1.20. Let ρ be a linear representation of G on a finite-dimensional
complex vector space V . Suppose that V is endowed with a Hermitian inner
product (., .) satisfying
(ρ(g)(v), ρ(g)(w)) = (v, w) ∀v, w ∈ V, g ∈ G.
Now if U is a G-invariant subspace of V , then the orthogonal complement U ⊥
of U in V is an invariant subspace of V complementary to U .
It is important note that, if G is finite, one can always build an invariant
Hermitian inner product (.|.) from an arbitrary one (., .) as was done in Exercise
3.2.22 for positive definite inner products:
X
(v|w) = (ρ(g)(v), ρ(g)(w)).
g∈G
Thus, the above argument gives an alternative proof of the lemma in the case
where V is a finite-dimensional complex vector space.
4.1. LINEAR REPRESENTATIONS OF GROUPS 57
Exercise 4.1.21. Show that in Exercises 4.1.10 and 4.1.18, there are invariant
subspaces that have no invariant complements.
For finite groups, irreducible representations are the building blocks of all
representations.
Proof. Proceed by induction. Suppose that the stament holds for all representa-
tions of dimension smaller than n, and let V be an n-dimensional representation.
If V is irreducible, then we are done. Otherwise there exists a sub-G-module U
with 0 ( U ( V . By Lemma 4.1.19, U has a invariant complementary sub-G-
module W . As both dim U and dim W are smaller than n, we may apply the
induction hypothesis, and find that U is the direct sum of irreducible sub-G-
modules U1 , . . . , Uk of U and W is the direct sum of irreducible sub-G-modules
W1 , . . . , Wl . But then
V = U1 ⊕ . . . ⊕ Uk ⊕ W1 ⊕ . . . ⊕ Wl .
Exercise 4.1.27. Let p be a prime, and consider the cyclic group Cp with
generator c. The map Cp → GL2 (Z/pZ) defined by
1 k
ck 7→
0 1
Exercise 4.1.30. Prove that the following converse of Schur’s lemma holds:
Let r : G → GL(V ) be a completely reducible representation of a group G. For
r to be irreducible, it suffices that any linear map A : V → V that commutes
with all r(g), g ∈ G, is a scalar.
3. Conclude that the converse of Schur’s lemma does not necessarily hold if
G is not completely reducible, cf. Exercise 4.1.30.
4.2. DECOMPOSING DISPLACEMENTS 59
1 2
1 2
3 3
(1,2,3)
1 2 1 2
Our goal can now be formalized as follows: try to write Γ as a direct sum of
irreducible submodules. Theorem 4.1.25 tells us that this is possible.
Here, we shall try to solve this problem ‘by hand’, without using too much
representation theory. In this way, the reader will appreciate the elegant char-
acter arguments which are to be given in Example 5.2.23.
As a first step, we can write each displacement in a unique way as a purely
‘radial’ displacement and a purely ‘tangental’ displacement (see Figure 4.4), and
the spaces Γrad and Γtan of radial and tangental displacements, respectively, are
three-dimensional submodules of Γ.
The submodule Γrad contains a one-dimensional submodule Γ1rad , consisting
of displacements with the ‘same’ velocity vector in all three directions, such as
the one depicted in Figure 4.5. To be precise: this space consists of all radial
displacements f for which (1, i)(f (1)) = f (i).
4.2. DECOMPOSING DISPLACEMENTS 61
3 3
1 2 1 2
1 2
3 3
1 2 1 2
The space Γ1rad must have an invariant complement in Γrad , and looking at
the pictures we find that the space Γ2rad of radial displacements spanned by the
two in Figure 4.6 is also invariant. Note that this space consists of all radial
displacements for which (1, 3)f (1) + (1, 2)f (1) + f (1) = 0.
One may wonder if the two-dimensional module Γ2rad can be decomposed
further, but this is not the case, as we shall see in Example 5.2.23. The rep-
resentation Γtan can be decomposed in a similar fashion. However, instead of
going into details rightaway, we will first treat more representation theory, and
come back to this application in Example 5.2.23.
Notes
This section is based on a similar (but more difficult) example in [10]. If your
interest is aroused by this example, you are encouraged to read the pages in
that book concerned with the displacemements of a tetrahedral molecule.
Chapter 5
Character Tables
5.1 Characters
An extremely useful notion regarding matrices is the trace. The trace of a matrix
A = (aij )n×n is tr(A) = a11 + · · · + ann .
Exercise 5.1.1. Prove that for two matrices A and B, one has tr(AB) =
tr(BA).
Exercise 5.1.2. Show that in the case of a permutation representation π : G →
Sn , the number tr(Mπ (g)) (see Example 4.1.8) is equal to the number of fixed
points of the permutation π(g).
Definition 5.1.3. Let r be a representation of a group G on a finite-dimensional
vector space V . The character of r is the function χr defined on G by
63
64 CHAPTER 5. CHARACTER TABLES
for χ, ψ ∈ H.
5.2. ORTHOGONALITY OF IRREDUCIBLE CHARACTERS 65
We have noticed earlier that characters are class functions. We divide the
proof of Theorem 5.2.1 into two parts: first, we will show the orthonormality
of the irreducible characters, and then their being a full basis of the space of
class functions. For our first purpose, we introduce the tensor product of two
representations.
Definition 5.2.3. Let V and W be vector spaces over K. Then V ⊗ W , called
the tensor product of V and W , is the quotient of the free vector space over K
with basis {v ⊗ w | v ∈ V, w ∈ W } by its subspace spanned by all elements of
the following forms (for v, v1 , v2 ∈ V , w, w1 , w2 ∈ W and λ ∈ K):
(v1 + v2 ) ⊗ w − v1 ⊗ w − v2 ⊗ w,
v ⊗ (w1 + w2 ) − v ⊗ w1 − v ⊗ w2 ,
(λv) ⊗ w − λ(v ⊗ w), and
v ⊗ (λw) − λ(v ⊗ w).
{vi ⊗ wj | i = 1, . . . , m, j = 1, . . . , n}
is a basis of V ⊗ W .
It is not hard to see that if U, V, and W are vector spaces over K, then
the spaces (U ⊗ V ) ⊗ W and U ⊗ (V ⊗ W ) are canonically isomorphic. We
will identify them, leave out parentheses, and write T d (V ) for the d-fold tensor
product of V with itself. The direct sum of all T d (V ) for d ∈ N (including 0,
for which T d (V ) is isomorphic to K) forms an associative algebra with respect
to the operator ⊗; this algebra is called the tensor algebra on V and denoted
by T (V ).
Definition 5.2.5. Let V be a vector space over K and let d be a natural number.
Then the quotient of T d (V ) by the subspace spanned by the set
{v1m1 · · · vnmn | m1 , . . . , mn ∈ N, m1 + . . . + mn = d}
On the other hand, by the definition of the tensor representation, the left-hand
side equals
X X X
(r(g)vk ) ⊗ (s(g)wl ) = ( rik vi ) ⊗ ( sjl wj ) = rik sjl vi ⊗ wj . (5.2)
i j i,j
and therefore
X X
χs⊗r (g) = aij,ij = rii sjj = χr (g)χs (g),
i,j i,j
∗
Exercise 5.2.11. Show that χr = χr .
One final lemma before proceeding with the proof of Theorem 5.2.1.
T (w ⊗ f )(v) = f (v) · w, v ∈ V, f ∈ V ∗ , w ∈ W,
Exercise 5.2.13. Prove that the map T defined in the proof of Lemma 5.2.12
is really an intertwining map.
1 X
P = t(g).
|G|
g∈G
= (χ | χr ).
s
68 CHAPTER 5. CHARACTER TABLES
so that im(P ) = HomG (V, W ). This proves the claim. We now distinguish two
cases.
First, suppose that r and s are not equivalent. Then, by Schur’s lemma, the
space HomG (V, W ) is zero-dimensional, so that P and, a fortiori, its trace are
0. In this case (χs | χr ) = 0.
Alternatively, suppose that r and s are equivalent. As we only wish to
compute the inner product of their characters, we may assume that they are
5.2. ORTHOGONALITY OF IRREDUCIBLE CHARACTERS 69
equal: W = V and s = r. In this case, the image of P is the linear span of the
identity map I on V . Taking any basis of Hom(V, V ) that starts with I, we see
that the matrix of P with respect to that basis has only non-zero elements on
the first row. The only contribution to the trace of P is therefore the element in
the upper left corner. Since P I = I, that element is 1. Hence (χr | χr ) = 1.
This proves that the characters are an orthonormal system in the space of
class functions on G.
Proof. We saw in the lemma that the regular representation ρ can be written
as a sum of all irreducible representations ri of G. Moreover, each of these
representations appears ni times in the decomposition. So
n1 times ns times
z }| { z }| {
ρ = r1 ⊕ . . . ⊕ r1 ⊕ . . . ⊕ rs ⊕ · · · ⊕ rs .
Ps Ps
Therefore |G| = χρ (e) = i=1 ni χri (e) = i=1 n2i .
equals
1
(f | χr )I.
n
Proof. First of all, note that H verifies r(g)H = Hr(g) for all g ∈ G. Hence, as
r is irreducible, H must be multiplication by a scalar λI. Its trace equals nλ.
On the other hand, it equals
1 X
f (g)χr (g) = (f | χr ).
|G|
g∈G
Consequently,
λ = (f | χr )/n.
5.2. ORTHOGONALITY OF IRREDUCIBLE CHARACTERS 71
Proof of Theorem 5.2.1, second part. In order to prove that the irreducible char-
acters form a basis of H, it suffices to prove that the orthoplement of their linear
span in H is zero. Hence, let f ∈ H be orthogonal to all irreducible characters.
According to Lemma 5.2.20, the linear map
1 X
H= f (g)r(g)
|G|
g∈G
Thus, for H to be zero, f must be identically zero. This concludes the proof.
Half of the following proposition is Exercise 4.1.24, and can be done without
character theory.
Proposition 5.2.22. The group G is Abelian if and only if all of its irreducible
representations are one dimensional.
and
χr2 ((1)) = 2, χr2 ((1, 2)) = 0, and χr2 ((1, 2, 3)) = −1.
Both χr1 and χr2 have squared norm 1, so they are irreducible. The represen-
tation of D3 on E is equivalent to r2 , as can be read of from the traces of the
matrices on page 4.2.
An element of Γ can be extended in a unique way to a linear map V → E, so
that we may identify Γ with the space HomK (V, E) of K-linear maps from V to
E. By Lemma 5.2.12, HomK (V, E) is isomorphic to V ∗ ⊗E; let r : D3 → V ∗ ⊗E
be the corresponding representation. Lemma 5.2.8 shows that χr = χs χr2 .
Hence,
χr ((1)) = 6, χr ((1, 2)) = 0, and χr ((1, 2, 3)) = 0.
First, we can decompose χr = (χr1 + χr2 )χr2 = χr1 χr2 + (χr2 )2 = χr2 +
(χr2 )2 . The character (χr2 )2 has squared norm 3, so it is reducible. We have
(χr1 |(χr2 )2 ) = 1 and also (χr2 |(χr2 )2 ) = 1, so that (χr2 )2 − χr1 − χr2 is the char-
acter of a third irreducible representation, say r3 , of dimension 1. Its character
equals
χr3 ((1)) = 1, χr3 ((12)) = −1, and χ(r3 ) (123) = 1,
that is, χr3 = sgn . By Theorem 5.2.21, the characters r1 , sgn , r2 exhaust all
irreducible characters of D3 . We conclude that the character χr of D3 on Γ
decomposes into χr1 + 2χr2 + χr3 .
To evaluate the right-hand side of this equation we count the number of per-
mutations having precisely k fixed points (0 ≤ k ≤ n). This number is clearly
equal to
n
F (n − k),
k
m
X m!
F (m) = (−1)i .
i=0
i!
74 CHAPTER 5. CHARACTER TABLES
and it is not hard to prove that the right-hand side is zero for n ≥ 3.
Example 5.3.3. Consider the cyclic group Cn with generator c, and let ζ ∈ C
be a primitive n-th root of unity. Define, for j = 0, . . . , n − 1, the map χj :
Cn → C by χj (ck ) := ζ jk . Then the χj are characters of Cn , and irreducible
as they are one-dimensional. By Theorem 5.2.21 there are no other characters,
so that the character table of Cn is as follows.
Cn e c c2 ... cn−1
1 1 1 ... 1
χ1 1 1 1 ... 1
χ2 1 ζ ζ2 ... ζ n−1
χ3 1 ζ2 ζ4 ... ζ 2(n−1)
· · · · ... ·
· · · · ... ·
· · · · ... ·
(n−1)2
χn 1 ζ n−1 ζ 2(n−1) ... ζ
Exercise 5.3.12. Prove the column orthogonality of the character table: for
g, h ∈ G, and χ1 , . . . , χs a complete set of irreducible characters of G,
s
X
χi (g)χi (h) = |CG (h)|δg,h .
i=1
7. Conclude from the orthogonality of the character table that there is one
more irreducible character. Compute this character, and compare it with
the character of the restriction of the standard representation of S5 to A5 .
Cl
Cl
P
Cl Cl
Cl
Figure 5.1: The PCl5 molecule
D3h E C3 3C2 σh S3 σv
1 2 3 1 2 3
A01 1 1 1 1 1 1
A02 1 1 -1 1 1 -1
E0 2 -1 0 2 -1 0
A001 1 1 1 -1 -1 -1
A002 1 1 -1 -1 -1 1
E 00 2 -1 0 -2 1 0
The character χiso is the sum of the characters of D3h on translations and
on rotations. The character of D3h on translations equals χ3 = E 0 + A002 ,
and the character χ4 on rotations turns out to be as follows.
D3h E C3 3C2 σh S3 σv
χ4 3 0 -1 -1 2 -1
We find that χ4 = A02 + E 00 . We conclude that
N N N Cl
Pd Pd
Cl Cl Cl N
The following exercises shows how spectroscopic techniques may or may not
be used to recognize molecules.
Exercise 5.4.2. The square planar coordination compound PdCl2 (NH3 )2 can
be prepared with two different configurations; see Figure 5.2. Is it possible
to discern the cis- and the trans-configurations using information concerning
Pd − Cl stretch vibrations in IR spectroscopy?
Exercise 5.4.4. Let T be the octahedron whose 6 vertices are joined by strings.
Determine the decomposition of the 18-dimensional space of vibrations of irre-
ducibles of the group of rotational symmetries of T (of order 24). Interpret
the result. Which spaces fuse into irreducibles for the group of all (proper or
improper) symmetries of T (of order 48)?
5.5 Notes
Most of the content of this chapter can be found in Serre’s book [9]. The last
section is entirely based on Theo Beelen’s lectures in the course for which these
notes were prepared.
Chapter 6
In the previous two chapters we got acquainted with the representation theory
of finite groups. In this chapter we will treat some of the representation theory
of compact Lie groups, without going into the details of defining a Lie group and
showing, for example, the existence of an invariant measure. A good reference,
on which in fact most of this chapter is based, is [1]. Some general, less detailed
remarks can be found in [9].
1. GL(n, R),
2. SL(n, R),
3. O(n, R),
4. SO(n, R),
The latter two groups are called the unitary group and the special unitary group.
A more intrinsic definition is the following: let (., .) be a Hermitian inner product
on a finite-dimensional complex vector space V . Then the corresponding unitary
81
82CHAPTER 6. SOME COMPACT LIE GROUPS AND THEIR REPRESENTATIONS
group is the group of all linear maps on V with (gv, gw) = (v, w) for all g ∈ G
and v, w ∈ V . Note that each of the real (complex) matrix groups above is
the zero set of some smooth map F from Mm (R) (Mm (C)) to some Euclidean
space Rd such that the rank of the Jacobian of F at every point of the zero
set of F is d. If, in this situation, the zero set of F is a group with respect to
matrix multiplication—as it is in the above cases—then it is automatically a
Lie group. Such Lie groups are called linear because they are given by a linear
representation. In the present chapter, this class of Lie groups suffices for our
needs.
Exercise 6.1.1. How can GL(n, R) be constructed from Mn+1 (R) via the above
construction? What about (R, +)?
Exercise 6.1.2. Show that O(n, R) ∼ = SO(n, R) o (Z/2Z). Here o means that
the subgroup at the left is a normal subgroup and the subgroup at the right is
a complementary subgroup in the sense that their product is the whole group
and their intersection is the trivial group. Such a product is called semi-direct.
If n is odd, then it is a direct product.
The theory of Lie groups is a blend of group theory and topology (surfaces,
continuity, differentiability, etc.), as might be expected. For example, one can
ask whether a given Lie group is (path-wise) connected, or whether it is compact.
A typical lemma from the theory is the statement that the connected component
of the identity element is a normal subgroup.
Proof. We will connect an arbitrary matrix g = (aij )ij in SO(n, R) to the iden-
tity via a continuous and piecewise smooth path in SO(n, R). The first step is
to connect it to a diagonal matrix; to this end, suppose first that there exists a
position (i, j) such that i > j and aij 6= 0. Then one can choose this position
such that, in addition, ai0 j 0 = 0 for all (i0 , j 0 ) with i0 > j 0 and either j 0 < j or
j 0 = j and i0 > i. Schematically, g has the following form:
T ∗ ∗ ∗ ∗
0
ajj ∗ aji ∗
0
g= ∗ ∗ ∗ ∗
,
0 aij ∗ aii ∗
0 0 ∗ ∗ ∗
6.1. SOME EXAMPLES OF LIE GROUPS 83
Exercise 6.1.6. Show that SU(2) is connected. Hint: find a continuous path
from any given element of SU(2) to some e(t) as in Exercise 6.1.3.
84CHAPTER 6. SOME COMPACT LIE GROUPS AND THEIR REPRESENTATIONS
3. left-invariant, i.e., for the action of G induced on C(G) by the left multi-
plication, we have for all h ∈ G and f ∈ C(G) that
Z Z
h · f (g)dg = f (g)dg,
G G
−1
where h · f is the function g 7→ f (h g) (h ∈ G).
4. and normalized, i.e., Z
1dg = 1.
G
A similar but more laborious computation yields that the invariant Haar mea-
sure for SU2 (C) is
1
sin2 θ sin ψdθdψdφ,
2π 2
where
− sin θ sin ψeiφ
cos θ − i sin θ cos ψ
sin θ sin ψe−iφ cos θ + i sin θ cos ψ
is a general element of SU2 (C).
The result of the following exercise will be used in proving that certain
irreducible representations of SU2 (C) exhaust all irreducible representations.
Exercise 6.2.3. Show that the map sending a class function f : SU(2) → C
to (f ◦ e) : R → C, where e is the one-parameter subgroup of SU(2) defined
in Exercise 6.1.3, is a linear bijection between the space H of continuous class
functions on SU(2) and the linear space of all even, 2π-periodic continuous
functions R → C. Moreover, the norm corresponding to the invariant inner
product (6.1) below corresponds, under this linear bijection, to the normalized
L2 inner product on the space of continuous even 2π-periodic functions on R.
One could define representations of compact Lie groups in Hilbert spaces,
but here we shall restrict our attention to finite-dimensional representations.
Also, we shall only consider complex representations.
Definition 6.2.4. A finite-dimensional complex representation of the Lie group
G is a smooth homomorphism G → GL(V ), where V is a finite-dimensional
complex vector space.
In fact, one can prove that continuous homomorphisms from a Lie group G
are automatically smooth.
For finite-dimensional complex representations of compact Lie groups, al-
most all theorems that we proved in the preceding chapters for finite groups
hold, be it that at some places a sum must be replaced by an integral. To give
an idea how things work, one can try the following exercise.
Exercise 6.2.5. Let r : G → GL(V ) be a finite-dimensional representation of
the compact Lie group G, and let W ⊂ V be an invariant subspace. Show that
W has an invariant complement in V .
We will study the representations of the groups SU(2) and SO(3, R), so it is
good to verify that these are indeed compact Lie groups.
Exercise 6.2.6. In this exercise we prove that U(n), SU(n), O(n, R) and
SO(n, R) are compact Lie groups. To this end, endow the space Mn (C) with
the Hermitian inner product given by
X
((aij )ij , (bij )ij ) := aij bij .
i,j
Thus, U(n) is a Lie group by the construction of the beginning of this chapter,
and compact as it is a closed and bounded set in Mn (C). Find similar arguments
for the other groups above. Show, on the contrary, that the group O(n, C) of
complex matrices g satisfying g T g = Iis not compact for n ≥ 2.
A final remark concerns the compact analogon of the inner product on the
space of complex-valued functions on a finite group: in the case of a compact
Lie group, one should consider the space of Hilbert space of continuous class
functions with Hermitian inner product defined by
Z
(f |h) := f (g)h(g)dg. (6.1)
G
Here, the irreducible characters form an orthonormal Hilbert basis in the space
of continuous class functions.
First applying A and then rd (st ), and using the fact that the two commute, we
find that also
X n
A(st Pn ) = cosk t · sinn−k t · cn Pk .
k
k
Taking t such that cos t sin t 6= 0, it follows from the fact that the Pk are linearly
independent that cn = ck for all k = 1, . . . , n, and hence that A = cn I.
We wish to prove that these are all irreducible characters. To this end we
shall use some Fourier theory, and the result of Exercise 6.2.3.
Theorem 6.3.2. The representations rd are, up to equivalence, all irreducible
representations of the group SU(2).
Proof. In view of Exercise 6.2.3, we can identify H with the space of even 2π-
periodic complex-valued functions on R. The matrix of rd (e(t)) with respect to
the basis P0 , . . . , Pd is diagonal, and its trace is
d
X
κd (t) = ei(d−2k)t .
k=0
88CHAPTER 6. SOME COMPACT LIE GROUPS AND THEIR REPRESENTATIONS
So L is the group
From now on, φ will denote the restriction of this homomorphism to SU(2). We
claim that the image is the group
e⊥
0 = {x1 e1 + x2 e2 + x3 e3 | xi ∈ R} ⊆ M.
Proof. Any such x is in fact a Hermitian matrix with trace 0. As such it has
two real eigenvalues c, −c, where c > 0, (note that x is non-zero). Let two
(necessarily orthogonal) eigencolumns of norm 1 corresponding to c and −c
be the first and second column of a matrix B, respectively. Then the usual
coordinate transformation rule gives
x = B −1 (cH)B.
Exercise 6.4.6. Show that ker φ = {I, −I}. Hint: the elements of this kernel
must commute with all of M , hence in particular with e1 , e2 , and e3 .
Finally we can apply lemma 6.4.1: the irreducible representations of SO(3)
correspond to those of SU(2) for which −I is mapped to the identity transfor-
mation. More precisely, we wish to know for which d we have rd (−I) = I. For
this to be the case, it is necessary and sufficient that
This holds if and only if d is even. Hence the irreducible representations are
Wd := P (V )2d .
hd = {p ∈ P (V )d | ∆p = 0}.
As
was noticed in Example 4.1.17, the dimension of P (V )d is dim P (V )d =
d+2
d .
We now want to prove that the dimension of hd is 2d + 1. A homogeneous
polynomial p of degree d can be written as a sum of homogeneous polynomials
pk in x2 and x3 with power of x1 ’s as coefficients:
x21 xd
p(x1 , x2 , x3 ) = p0 (x2 , x3 ) + x1 p1 (x2 , x3 ) + p2 (x2 , x3 ) + · · · + 1 pd (x2 , x3 )
2 d!
d
X xk 1
= pk (x2 , x3 )
k!
k=0
where the pk ’s have degree d − k. If one applies the Laplace operator to p (you
see why we took factorials):
d−2 k d
xk1
2
∂2
X x 1
X ∂
∆(p) = pk+2 (x2 , x3 ) + + pk (x2 , x3 )
k! k! ∂x22 ∂x23
k=0 k=0
6.4. THE 3-DIMENSIONAL ORTHOGONAL GROUP 91
This formula implies that p in hd depends only on the first two homogeneous
polynomials p0 and p1 . Therefore we conclude that the dimension of hd is
precisely the number of homogeneous polynomials in two variables of degree d
and d − 1. That is, dim hd = 2d + 1.
Of course, the degree of hd reminds us of the irreducible representations Wd
of SO(3) we met before. Indeed, we have the following.
To see that what we just stated is true, we should first check that hd is a
representation of SO(3). We state here the following fact without proof.
Fact: The action of the Laplace operator on P (V )d commutes with the
action of SO(3).
From this we conclude that hd is a representation of SO(3).
To prove that hd is irreducible, we claim that hd = Wd . Set p a harmonic
polynomial in hd , p(x1 , x2 , x3 ) = (x2 + ix3 )d . Let R(t) be the matrix of SO(3)
defined by
1 0 0
R(t) = 0 cos t − sin t
0 sin t cos t
Apply R(t) to p. That gives us R(t)p = exp(−idt)p. So p generates an
invariant subspace W of hd under the action of R(t).
Any element in SO(3) is conjugate to R(t). So in order to find the character
of Wd on R(t) it suffices to compute it on R(t). That is just the value of
the character χ of rd (the irreducible representation of SU(2)) at et/2 , i.e.,
P2d i(d−k)t 2d
k=0 e .
By the decomposition of hd in irreducible Ws ’s, we know that its character
is a linear combination of exp(ist). So what we saw above about the action of
R(t) on W leads us to conclude that hd = Wd .
92CHAPTER 6. SOME COMPACT LIE GROUPS AND THEIR REPRESENTATIONS
Bibliography
[6] Juergen Hinze, editor. The permutation group in physics and chemistry,
volume 12 of Lecture notes in chemistry. Springer Verlag, Berlin, 1979.
[7] H.H. Jaffe and Milton Orchin. Symmetry in chemistry. Wiley, London,
1965.
[8] A. Nussbaum. Applied group theory; for chemists, physicists and engineers.
Prentice-Hall electrical engineering series. Prentice-Hall, Englewood Cliffs,
1971.
[9] J.P. Serre. Linear representations of finite groups. Springer Verlag, New
York, 1977.
[10] S. Sternberg. Group theory and physics. Cambridge University Press, Cam-
bridge, 1994.
93