You are on page 1of 153

Notes on Group Theory

Mario Trigiante
November 18, 2010

Contents
1 Abstract Groups
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Definition of an Abstract Group . . . . . . . . . . . . .
1.2.1 Some examples . . . . . . . . . . . . . . . . . .
1.2.2 Finite and Infinite Groups. . . . . . . . . . . . .
1.2.3 Generators and Defining Relations . . . . . . .
1.3 Subgroups . . . . . . . . . . . . . . . . . . . . . . . . .
1.4 Classes and Normal Subgroups . . . . . . . . . . . . .
1.4.1 Conjugation classes in Sn and Young Diagrams.
1.4.2 Conjugation Classes in Dn . . . . . . . . . . . .
1.4.3 Normal Subgroups . . . . . . . . . . . . . . . .
1.5 Homomorphisms, Isomorphisms and Automorphisms .
1.5.1 The Automorphism Group . . . . . . . . . . . .
1.6 Product of Groups . . . . . . . . . . . . . . . . . . . .
1.7 Matrix Groups . . . . . . . . . . . . . . . . . . . . . .
1.7.1 The Group GL(n, F) . . . . . . . . . . . . . . .
1.7.2 The Group SL(n, F) . . . . . . . . . . . . . . .
1.7.3 The Group O(n, F) . . . . . . . . . . . . . . . .
1.7.4 The Group O(p, q; F) . . . . . . . . . . . . . . .
1.7.5 The Unitary Group U(n) . . . . . . . . . . . . .
1.7.6 The Pseudo-Unitary Group U(p, q) . . . . . . .
1.7.7 The Symplectic Group Sp(2n, F) . . . . . . . .
1.7.8 The Unitary-Symplectic Group USp(2p, 2q) . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

2 Transformations and Representations


2.1 Linear Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1 Homomorphisms Between Linear vector Spaces . . . . . . . .
2.1.2 Inner Product on a Linear Vector Space over R . . . . . . . .
2.1.3 Hermitian, Positive Definite Product on a Linear Vector Space
2.1.4 Symplectic product . . . . . . . . . . . . . . . . . . . . . . . .
2.1.5 Example: Space of Hermitian matrices . . . . . . . . . . . . .
2.2 Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1 Real Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . .
3

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

5
5
7
8
9
10
14
17
17
20
21
22
25
27
27
27
28
28
31
33
34
35
36

. . . .
. . . .
. . . .
over C
. . . .
. . . .
. . . .
. . . .

39
39
42
44
45
46
47
48
49

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

CONTENTS
2.2.2 Complex Spaces With Hermitian Positive Definite Metric . . . . . . .
2.2.3 Symplectic Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3 Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1 Transformations on Mn . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2 Homogeneous-Linear Transformations on Mn and Linear Transformations on Vn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.3 Affine Transformations . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.4 The volume preserving group SL(n, F) . . . . . . . . . . . . . . . . .
2.3.5 (Pseudo-) Orthogonal Transformations . . . . . . . . . . . . . . . . .
2.3.6 Unitary Transformations . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.7 Homomorphism between SU(2) and SO(3) . . . . . . . . . . . . . . .
2.3.8 Symplectic Transformations . . . . . . . . . . . . . . . . . . . . . . .
2.3.9 Active Transformations in Different Bases . . . . . . . . . . . . . . .
2.4 Realization of an Abstract Group . . . . . . . . . . . . . . . . . . . . . . . .
2.4.1 Representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.2 The Regular Representation and the Group Algebra . . . . . . . . . .
2.5 Some Properties of Representations . . . . . . . . . . . . . . . . . . . . . . .
2.5.1 Unitary Representations and Quantum Mechanics . . . . . . . . . . .
2.6 Schurs Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.7 Great Orthogonality Theorem . . . . . . . . . . . . . . . . . . . . . . . . . .
2.8 Characters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.9 Operations with Representations . . . . . . . . . . . . . . . . . . . . . . . .
2.9.1 Direct Sum of Representations . . . . . . . . . . . . . . . . . . . . . .
2.9.2 Direct Product of Representations . . . . . . . . . . . . . . . . . . . .
2.9.3 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.10 Representations of Products of Groups . . . . . . . . . . . . . . . . . . . . .

3 Constructing Representations
3.1 Constructing Representations of Finite Groups . . . . . . . . . . . . .
3.1.1 Irreducible Representations of Sn and Young Tableaux . . . .
3.2 Irreducible Representations of GL(n, C) . . . . . . . . . . . . . . . . .
3.3 Product of Representations . . . . . . . . . . . . . . . . . . . . . . . .
3.4 Branching of GL(n, C) Representations With Respect to GL(n 1, C)
3.5 Representations of Subgroups of GL(n, C) . . . . . . . . . . . . . . .
3.5.1 Representations of U(n) . . . . . . . . . . . . . . . . . . . . .
3.5.2 Representations of (S)O(p, q) . . . . . . . . . . . . . . . . . .
4 References

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

50
51
51
53
57
60
61
61
65
65
67
68
68
70
78
83
86
89
92
95
101
101
103
105
111
113
113
116
121
134
135
137
142
144
153

Chapter 1
Abstract Groups
1.1

Introduction

The mathematical notion of group plays an important role in natural sciences as it provides
a quantitative description of Symmetry. By symmetry we mean the property of an object
(e.g. a crystal, a molecule or any other physical system) to look the same after undergoing
a certain transformation. A symmetry transformation, in other words, brings the object to
overlap with itself so that the final configuration of the system is virtually indistinguishable
from the initial one. The system is then said to be invariant with respect to its symmetry
transformations. For instance the symmetry transformations of a square are rotations about
its center by multiples of 90o and reflections in two mutually perpendicular lines, while those
of an equilateral triangle are rotations about its incenter by multiple of 120o and reflections
in the three altitudes. In three dimensional Euclidean space a spherical object is symmetric
with respect to any rotation about any direction through its center and to reflections in
any plane containing the center. The symmetry of the five regular polyhedrons, i.e. the
tetrahedron, the cube, the octahedron, the dodecahedron and the icosahedron can be totally
characterized by 3 distinct sets (or groups as we shall see) of transformations, consisting of
rotations and reflections, which are enough to capture the symmetry properties of various
three dimensional structures: For instance the molecule of methane CH4 has the shape of a
tetrahedron in which the Hydrogen atoms are located at the four vertices and the Carbon
atom at the center; The Pyrite (F e S2 ) is a mineral whose crystal may appear in cubic or
2
octahedral shapes; The ion B12 H12
has an icosahedral shape.
Given two symmetry transformations, A and B, of a system, we can consider the transformation resulting from first performing B and then applying A to the resulting configuration.
This transfromation is defined as the product A B of the two transformations A and B and
is clearly a symmetry transformation (take as A and B two clockwise rotations by 90o of a
cube about its center, the product A B will then be a clockwise rotations by 180o ). For
any transformation A which maps an initial configuration of the system into a final one, we
can define its inverse A1 which brings the system back to its initial configuration. If A is a
symmetry transformation, so is A1 (in the example of a square, if A is a 90o clockwise rotation, A1 is a 90o anticlockwise rotation, or, equivalently, a 270o clockwise rotation). We
5

CHAPTER 1. ABSTRACT GROUPS

can define the identity transformation I as the trivial transformation mapping every point in
the original configuration into itself. It clearly is a symmetry transformation of any object.
Moreover, of A is a transformation, then it can be easily verified that: A I = I A = A.
Finally, given three transformations A, B, C, if we first apply B C and then A to an initial
configuration or first C and then A B the resulting configuration is the same, namely the
product of transformations is associative: A (B C) = (A B) C. A set of elements endowed
with a product operation satisfying the above properties is called a group. Therefore the set
of all the symmetry transformations of an object, together with the product defined above,
is a group called the symmetry group of the system.
The symmetries of a system have consequences on its physical properties. Consider
two point-like electric charges, such as the electron e and the proton p+ in a Hydrogen

atom. The electric field E generated by p+ has spherical symmetry, namely symmetry with
respect to any rotation about the proton itself. This implies that the physical properties of
the system, such as the energy, do not change if we rotate the electron around the proton

keeping their distance fixed. Consider now an electric point-charge q in the field E generated
by other two point-charges q1 , q2 . The symmetry of the latter now is no longer spherical,
but rather cylindrical, since it is invariant with respect to an arbitrary rotation about only
one direction, which is the one connecting the two source-charges. As a consequence of this
all the physical properties of the system, like the energy, should not change if we rotate q
around the symmetry axis keeping its distance from it fixed.
As an other instance consider a crystal with its own symmetry group. Since the physical
properties of any system are left unchanged under the action of a symmetry transformation,
the refraction index along directions connected to one another by the action of elements of
the symmetry group is the same.
In quantum mechanics, all physical properties of a system, like a sub-atomic particle,
an atom or a molecule, are encoded in the notion of quantum state , represented by a
vector |i in an infinite-dimensional Hilbert space. The effect of a transformation A, which
may be a rotation, a reflection or a translation in our three-dimensional Euclidean space, or
may also include time evolution or Lorentz boosts if we consider the larger four-dimensional
Minkowski space-time, is to map a state into a different one:
|i |0 i A |i .

(1.1.1)

If A is a symmetry transformation of the system, its action will not alter the energy: E =
E0 . The presence of a symmetry implies the existence of more than one independent state for
a given value of energy, namely degeneracy of the energy levels. The amount of degeneracy
will depend on the maximal number of independent states related to a given one by the
action of the elements of the symmetry group. Such numbers, for different initial states, are
a mathematical feature of the group itself, i.e. the dimensions of its representations.
Understanding all the physical properties of a system in relation to its symmetries is one
of the main successes of group theory applied to natural sciences.

1.2. DEFINITION OF AN ABSTRACT GROUP

1.2

Definition of an Abstract Group

We have introduced the notion of group in relation to the symmetry transformations of


an object. Different objects may have symmetry transformations which, through acting in
a different way, close groups having the same structure. Such structure, which we shall
define more rigorously in a moment, is encoded in the definition of a corresponding abstract
group. We say that different symmetry groups have the same structure if they are different
realizations of a same abstract group. As an example consider the (proper) symmetry group
of a square, consisting of rotations about its center by integer multiples of /2 and the
symmetry group of a parallelepiped with square basis, consisting of /2rotations about its
axis. The two groups are represented by different elements, since their transformations act
on different objects, but, nevertheless share the same structure and are different realizations
of the same abstract group to be denoted by C4 . Let us start then by giving the definition
of an abstract group.
A set of objects G, on which a product operation is defined, is said to be a group if the
following axioms are satisfied:
i G is closed with respect to : For any couple of elements g1 , g2 in G there exist a
third element g3 such that g3 = g1 g2 . In other words is a map from G G into G;
ii Existence of unit element: There exist an element e in G so that, for any g G
we have: e g = g e = g;
ii Existence of the inverse element: For any element g of G there exist its inverse
g 1 in G so that: g g 1 = g 1 g = e;
iii Associativity: For any three elements g1 , g2 , g3 of G the following property holds:
g1 (g2 g3 ) = (g1 g2 ) g3 . The result of this triple product is denoted by g1 g2 g3 ;
Using the associative property of the product, it is easy to verify that: (g1 g2 )1 = g21 g11 .
Moreover the inversion is involutive, namely, for any g in G, g = (g 1 )1 .
The correspondence between couples of elements g1 , g2 of G and elements g3 = g1 g2 in
G through the product operation defines the structure of G.
In general, given two elements g1 , g2 of G, the product g1 g2 may be different from g2 g1 ,
namely the two elements may not commute. If any two elements of G commute the group
is said to be abelian.
A group G is said to be finite if it consists of finitely many elements: G {g1 , . . . , gn }.
The number n of its elements is called the order of G and is also denoted by |G|. The
structure of a finite group G is conveniently represented by a multiplication table in which
the first row and column contain the elements of the group in the same order g1 , . . . , gn while
the entry (i, j) is the element gi gj .
Exercise 1.1: Show that the four complex numbers 1, i, 1, i form a group of order
four with respect to ordinary multiplication and write the multiplication table.

1.2.1

CHAPTER 1. ABSTRACT GROUPS

Some examples

The cyclic group: An example of finite group is the cyclic group which is the group
whose elements are powers g k , k integer, of an element g called the generator of the group:
Cyclic group {. . . , g 2 , g 1 , g 0 , g, g 2 , . . .} .

(1.2.1)

Let us define what we mean by power of an element g: If k is a positive integer then


g k g g . . . g k-times (g 2 g g, g 3 g g g and so on, so that g k g k1 g); If k is
a negative integer g k (g 1 )k ; Finally we define g 0 e. We say that the cyclic group is
finite of period n, for some positive integer n if g n = e, so that the group will consist of the
following elements
Cyclic group of period n {e = g 0 , g, g 2 , . . . , g n1 } ,

(1.2.2)

where we omitted the negative powers since g 1 = g n1 ,g 2 = g 2 n2 = g n g n2 = e g n2 =


g n2 and so on. The multiplication table of this group reads:
e
g
..
.
g n1

e
e
g

g
g
g2

...
...
...

g n1
g n1
e

g n1

...

g n2

From the symmetry of the table we may conclude that a cyclic group is abelian.
Exercise 1.2: Consider the group Cn of rotations about a point in the plane by integer
multiples of 2 /n (C3 is the symmetry group of an equilateral triangle, C4 of a square etc...).
Show that the group Cn is cyclic of order n.
The permutation group: Given a set of n objects x1 , . . . , xn a permutation is a one-toone correspondence between each of them and other objects in the same set. If a permutation
S maps x1 into xi1 , x2 into xi2 up to xn into xin , it is usually represented by the short-hand
notation:




x1 x2 . . . x n
1 2 ... n
S
or simply
,
(1.2.3)
xi 1 xi 2 . . . x i n
i1 i2 . . . i n
the order of elements in the first row is not important, what matters is the mapping. We
shall also denote the number ik corresponding to k through the permutation S by S(k) and
say that S maps 1 into the number S(1), 2 into S(2) and so on. Consider now the set Sn
of all possible permutations of a set of n objects. We can define on Sn a product operation:
If a permutation S maps a generic element xk into the element xik = xS(k) and an other
permutation T maps xik = xS(k) into xjk = xT (S(k)) , the product T S is the permutation
mapping xk into xjk = xT (S(k)) :

 
 

xi 1 xi 2 . . . x i n
x1 x2 . . . x n
x1 x2 . . . xn
T S =

=
xj 1 xj 2 . . . x j n
xi 1 xi 2 . . . x i n
xj1 xj2 . . . xjn


x1
x2
...
xn
=
.
(1.2.4)
xT (S(1)) xT (S(2)) . . . xT (S(n))

1.2. DEFINITION OF AN ABSTRACT GROUP

For instance if

S

1 2 3 4
3 4 2 1


and

1 2 3 4
4 2 3 1


,

(1.2.5)

we have:



1 2 3 4
T S
.
(1.2.6)
3 1 2 4


1 2 3 4
Clearly this product is not commutative, since S T =
6= T S. We define
1 4 2 3
the identity permutation I as the trivial permutation mapping each object in the set into
itself:


1 2 ... n
I
.
(1.2.7)
1 2 ... n
From the definition of product one can easily verify for any permutation S that: S I =
I S = S.
Given a permutation S mapping a generic element xk into the element xik , we define its
inverse S 1 as the permutation mapping xik back into xk :




1 2 ... n
i1 i2 . . . in
1
S
S
.
(1.2.8)
i1 i2 . . . in
1 2 ... n
Clearly the following property holds: S S 1 = S 1 S = I. Finally one can verify that the
product of permutations is associative. The set of all permutations of n objects, endowed
with the product defined above, forms therefore a (non-abelian) group Sn of order n!.
Exercise 1.3: Show that the following set of permutations:






1 2 3 4
1 2 3 4
1 2 3 4
T1 =
, T2 =
, T2 =
,
1 2 3 4
2 1 4 3
4 3 2 1


1 2 3 4
,
(1.2.9)
T4 =
3 4 1 2
for a group and write the multiplication table. Is the group abelian?

1.2.2

Finite and Infinite Groups.

Let G be a finite group and g any of its elements. It follows that g k , for any integer k is still
an element of G. However, since G is finite, not all g k , for different values of k are different.
0
In particular we must have for some integers ` and `0 : g ` = g ` . Multiplying a number of
times both sides by g 1 and using the associative property of the product, we deduce that
g k = e for a certain k. Therefore, for any element g there exist an integer k, called the order
or period of g, such that: g k = e, and therefore g k1 is the inverse of g.

10

CHAPTER 1. ABSTRACT GROUPS

A group G is called infinite if it is not finite. An example of infinite group is the set of
rational positive numbers with respect to ordinary product. An other example of infinite
group is an infinite cyclic group, such as the group of integer multiples of a real number a,
. . . , 2 a, a, 0, a, 2 a, . . ., on which the product operation is defined as the ordinary sum
of real numbers.
A special instance of infinite groups are the Lie groups, which will be extensively studied
in the present course. The elements of a Lie group are functions of a number of continuous
parameters g = g(1 , . . . , n ) (for this reason they are called continuous groups) and their
structure is described by specifying the parameters of each g3 = g1 g2 as analytic functions
of the parameters of g1 and g2 . The number n of parameters is called dimension of the Lie
group. An instance of Lie group is the group SO(3) of rotations in the three dimensional
Euclidean space, each element R being uniquely defined by the three continuous parameters
such as the Euler angles: R = R[, , ]. This group, together with the spatial reflections,
forms the symmetry group O(3) of the sphere. A simpler example of Lie group is the group
SO(2) of planar rotations R[] about a point by an angle . This group, as opposed to SO(3),
is abelian and its structure is defined by the relation: R[1 ] R[2 ] = R[1 + 2 ].
The Group GL(n, R). An other example of Lie group is the group GL(n, R) of general
linear transformations on a ndimensional linear vector space. Its elements are n n nonsingular matrices M[aij ]1 depending on n2 continuous parameters aij which are it own entries:

a11 a12 . . . a1n


a21 a22 . . . a2n
M[aij ] (aij ) =
.
(1.2.10)
..
..
...
.
.
an1

an1

. . . ann

This is a group with respect to ordinary rows-times-columns matrix multiplication. The


identity element of the group is the n n identity matrix 1n . With any non-singular matrix
we can associate its inverse matrix and the matrix multiplication is associative. This group
is clearly not abelian since the product of matrices is not.
The structure of this Lie group is defined by expressing the entries of the product of two
matrices in terms of their entries:
M[cij ] = M[aij ] M[bij ] ,
then cij =

1.2.3

Pn

k=1

(1.2.11)

aik bkj .

Generators and Defining Relations

According to the very definition, a group G is totally defined once its elements are given,
together with its structure, namely its multiplication table. The group G may be very large, if
1

In the present chapter the elements of a generic matrix will be labeled by two lower indices (e.g. aij ).
In next chapter we shall consider transformations implemented by matrices and will give a meaning to
vectors with upper and lower indices, according to their transformation property. The elements of matrices
representing transformations will have one lower and one upper index (e.g. M i j or Ri j ) since their action
on any vector should give a vector of the same kind, i.e. with index in the same (upper or lower) position.

1.2. DEFINITION OF AN ABSTRACT GROUP

11

not infinite and writing the multiplication table may be arduous, if not impossible. Consider
the infinite cyclic group with generator g. Although infinite, defining it was relatively simple
since its generic element can be written as an integer power of the generator g. In general
we can generalize this characterization of a group by defining a minimal set of elements of a
group {g1 , g2 , . . . , g` }, called the generators of the group, such that any group element can
be written as a sequence of products of these generators. The generators are like letters of an
alphabet and the group elements words (in fact a generic group element is also referred to as
a word when it is represented as a sequence of generators). Clearly a group is abelian if and
only if its generators commute with one another: gi gj = gj gi , for any i, j = 1, . . . , `. One
implication is straightforward. Let us show that, if this property is true a group is abelian.
Take two elements g and g 0 of G. Each of them is represented by a word. Compute then the
product g g 0 . Using the commutativity property of the generators, we can permute them
freely in this product and compose the word representing g 0 first and that representing g
second, obtaining g g 0 = g 0 g. The group G is then abelian.
Finite cyclic groups, such as the rotation symmetry group Cn of regular polygons, have
just one generator. For Cn groups, the generator is a rotation r by 2 /n. The dihedral
groups Dn consists of rotations and reflections: D3 is the symmetry group of an equliateral
triangle. It has order 6 and consists of rotations by multiples of 2 /3 (which themselves
close a group C3 ), and a reflection in three directions containing the altitudes , , (see
Figure 1.1); D4 is the symmetry group of a square. It has order 8 and consists of rotations

Figure 1.1: Group D3


by multiples of /2 (which themselves close a group C4 ), and reflections , , , in the
directions along the diagonals or connecting the midpoints of opposite sides (see Figure 1.2).
While Cn have just one generator r, Dn have two generators: The rotation r by 2 /n and
a reflection . This means that any element of Dn can be expressed as products of r and
, as it can be seen for the D3 and D4 examples in Figures 1.1 and 1.2. The group Dn in

12

CHAPTER 1. ABSTRACT GROUPS

Figure 1.2: Group D4

general consists of rotations by multiples of 2 /n, which themselves close the group Cn , and
n reflections in directions at angles multiple of /n. All these n reflections can be obtained
by first acting by one of them , and then applying the rotation r a number of times. The
order of Dn is then 2 n:
Dn : {e, r, r2 , . . . , rn1 , , r , r2 , . . . , rn1 } .

(1.2.12)

If we are only told that Dn had just two generators r, , we would construct an infinite set of
words our of these two letters. We used the geometric properties of rotations and reflections
to conclude that a generic word can always be reduced to one of the 2 n in (1.2.12). What
geometry is telling us is that not all words correspond to distinct elements of the group.
For instance, we perform three consecutive rotations r by an angle 2 /3 we obtain the
identity rotation, i.e. r3 = e. Moreover if we perform twice a reflection we end up with the
identity transformation, namely 2 = e. As a consequence of this, in D3 , the words r3+k
and rk represent the same element. Similarly 2+k and k . These two properties of the D3
generators have the effect of cutting down the number of independent words, though are not
enough to reduce them to six. We need to use the property that r is itself a reflection
and thus (r )2 = e. From this relation, the previous two and the defining axioms of the
groups, we find r = r1 = r2 , which allows us to permute the order of the rs and the
s and reduce a generic word to the form rk p , with k = 0, . . . , 2 and p = 0, 1. Similarly,
for Dn , we can reduce the number of independent words by using the relations rn = e and
2 = e and the fact that r is a reflection, namely that (r )2 = e. From these relations
it follows that r = rn1 , using which we can reduce a generic word to the form rk p ,
with k = 0, . . . , n 1 and p = 0, 1, yielding the 2 n independent elements in (1.2.12). Using

1.2. DEFINITION OF AN ABSTRACT GROUP

13

the three relations rn = e, 2 = e and (r )2 = e we can construct the multiplication table


of the group, namely they are enough to totally characterize its structure.
We can, in general, characterize the structure of any group by a set of defining relations.
By defining relations we mean a minimal set of conditions in the form of Wk = e, Wk being
suitable words, namely sequences of products of generators, which are sufficient to deduce
the whole group structure. Any other relation satisfied by the elements of G can be deduced
from the defining ones using the defining axioms of a group. For instance a cyclic group of
period n and generator g is defined by the single relation g n = e. Consider now the trivial
group G = {e}. Also in this case the relation g n = e, for any integer n, is trivially satisfied
by all its elements. However g n = e, for n > 1 is not the defining relation, since the relations
g n = e, for n < 1 cannot be derived from it. In this case the defining relation is simply
g = e. Notice that, as opposed to a finite cyclic group the infinite cyclic group, also referred
to as C , has no defining relation: Different words correspond to different group elements.
Let us summarize our discussion so far. The task of defining the elements of a group is
considerably simplified by assigning its generators, in terms of which all the elements are
written as words. Then the structure of the group can be deduced from a set of defining
relations. The defining relations for the Cn and the Dn groups are:
Cn : r n = e ,
Dn : rn = e , 2 = e , (r )2 = e .

(1.2.13)

From the last three relations we find r = rn1 . If n = 1, the first relation implies
r = e, namely D1 consists of the only reflection , apart from e, and is trivially abelian. If
n = 2 we find r = r and the group is again abelian. The reader can convince himself
that these two, namely D1 and D2 are the only abelian dihedral groups. Using the relations
(1.2.13) we are able to construct the multiplication table of a Dn group. Let us give below
those of D3 and D4 :
D3
e
r
r2

=r
= r2
D4
e
r
r2
r3

=r
= r2
= r3

e
e
r
r2
r3

r
r2
r3

e
e
r
r2

r
r2
r
r
r2
r3
e
r3

r
r2

r
r
r2
e
2
r

r
r2
r2
r3
e
r
r2
r3

r2
r2
e
r
r
r2

r3
r3
e
r
r2
r
r2
r3

r
r2
e
r
r2

r
r2
r3
e
r
r2
r3

=r
r
r2

r2
e
r
=r
r
r2
r3

r3
e
r
r2

= r2
r2

r
r
r2
e
= r2
r2
r3

r
r2
r3
e
r

= r3
r3

r
r2
r
r2
r3
e

14

CHAPTER 1. ABSTRACT GROUPS

Notice that the above tables are not symmetric and thus the groups are not abelian.

1.3

Subgroups

A subset H of G is a subgroup of G if it is itself a group. Clearly any group G has as


subgroups the one consisting of the unit element only {e} and G itself. Actually {e} is the
only order one subgroup of G. Indeed if there were an other order one subgroup {g}, for
some g G, it should contain all the powers g ` of g up to its period k: g k = e. Since the
subgroup consists of the element g only, k = 1 and g = e.
Here we prove a necessary and sufficient condition for a subset H of G to be a subgroup:
Property 1.1: H is a subgroup of G if and only if for any two elements h1 , h2 of H, h1 h1
2
is in H.
One implication is simple. Let us now suppose that for any two elements h1 , h2 of H,
h1 h1
2 is in H and show that H is a subgroup of G, namely a group with respect to the
product operation induced from G. Notice that, from the hypothesis, it immediately follows
1
1
that e = h1 h1
1 is in H, for any h1 H. Moreover given a h1 H, being e H, h1 = eh1
is in H as well. As a consequence of this, given any two elements h1 , h2 of H, h1 h2 is still
in H, namely that H is closed with respect to the product on G. Indeed for any h1 , h2 H,
1 1
since h1
is by hypothesis in H.
2 is still in H, it follows that h1 h2 = h1 (h2 )
Remark: If G is infinite, in order for a subset H of G to be a subgroup, it is not sufficient
for it to be closed with respect to the product. As an example consider the group G of al real
numbers except 0, with respect to ordinary multiplication. Consider now a positive number
x > 0 and the subset H of all numbers of the form xn , with n positive integer. H is closed
with respect to ordinary product, however it is not a group since, for n > 0, xn H but
(xn )1 = xn is not in H. If G is finite, on the other hand, any subset of G closed with
respect to the product is a subgroup. Indeed, for any h H, h` is contained in H, for any `.
Being G finite there exist an integer k such that hk = e, which implies that e and h1 = hk1
are contained in H as well.
Clearly the only subgroup of order one consists of the identity e only.
Dihedral groups: The dihedral group Dn has Cn as a subgroup. Listing the elements of
Dn as in (1.2.12), the first n n diagonal block of its multiplication table is precisely the
multiplication table of its Cn subgroup.
Cyclic permutations: A cyclic permutation of n objects x1 , . . . , xn , denoted by the symbol (x1 x2 . . . xn ), or simply (1 2 . . . n) is a permutation mapping each element into the
subsequent up to xn which is mapped into x1 :


1 2 ... n 1 n
(1 2 . . . n)
.
(1.3.1)
2 3 ...
n
1

1.3. SUBGROUPS

15

In this notation (1 2 . . . n) = (2 3 . . . n 1 1) = . . .. A cyclic permutation of n objects is


also called a cycle of order n or a ncycle. We can perform a cyclic permutation of n objects
a number of times. Clearly if we perform it n times we are back to the initial sequence,
namely we obtain the identity permutation:


1 2 ... n 2 n 1 n
2
(1 2 . . . n)
,
3 4 ...
n
1
2
..
.
n
(1 2 . . . n) = I .
(1.3.2)
One can show that any permutation in Sn of n objects can be decomposed in product of
cyclic permutations acting on different subsets of the original set:


1 2 ... n 1 n
= (j1 . . . jk1 ) . . . (s1 . . . sk` ) .
(1.3.3)
i1 i2 . . . in1 in
For instance the reader can prove that (1 2 3)2 = (3 2 1), (1 2 3 4)3 = (4 3 2 1) and in
(1 2 . . . n)n1 = (n n 1 . . . 2 1). Here are other examples of this decomposition:


1 2 3 4
= (1 2 3 4)2 = (1 3) (2 4) ,
3 4 1 2


1 2 3 4 5 6
= (1 4) (3 5) (2 6) ,
4 6 5 1 3 2


1 2 3 4 5
= (1 3 2) (4) (5) (1 3 2) ,
3 1 2 4 5

general

(1.3.4)
(1.3.5)
(1.3.6)

where, for the sake of simplicity, we have suppressed in the last equation the 1cycles (4), (5),
since it is understood that all indices which do not enter the (n > 1) -cycles in the decomposition are mapped into themselves. Since each factor in (1.3.3) is itself a permutation
acting
P`
non trivially on different sets of objects, the factors commute and moreover i=1 ki = n.
These cycles define the structure of the permutation. For instance the structure of (1.3.6)
consists of one 3-cycle and 2 1-cycles. Suppose that in a generic permutation (1.3.3) the
1-cycles occur 1 times, the 2-cycles 2 times and so on up to the n-cycles which occur n
times. Since each cycle acts on different sets of objects, we must have:
1 + 2 2 + . . . + n n = n .

(1.3.7)

We shall characterize the structure of a permutation by means of the nplet of numbers


(1 , 2 , . . . , n ). For instance the structures of the permutations (1.3.4), (1.3.5) and (1.3.6)
are (0, 2, 0, 0), (0, 3, 0, 0) and (2, 0, 1, 0) respectively.
A 2cycle is also called transposition. Clearly the identity permutation I is the product
of 1cycles:
I = (1) (2) . . . (n) .

(1.3.8)

16

CHAPTER 1. ABSTRACT GROUPS

The powers of a ncycle form a subgroup of Sn which is clearly a cyclic group called group
of cyclic permutations.
A ncycle can be decomposed in products of transpositions:
(1 2 . . . n) = (1 n) (1 n 1) . . . (1 3) (1 2) = (1 2) (2 3) . . . (n 1 n) ,

(1.3.9)

so that any permutation can be expressed as product of transpositions. A permutation is


even or odd if it is decomposed into an even or odd number of transpositions respectively.
The reader can prove that the identity I is an even permutation and that the set of all
even permutations form a subgroup of Sn . There is an equal number ( n!2 ) of even and odd
permutations. As an example let us give the explicit composition of S3 :

I, (1 2 3), (3 2 1) even
S3 =
.
(1.3.10)
(1 2), (1 3), (2 3) odd
We leave as an exercise to the reader to show that we can choose as generators of the
permutation group Sn the set of n 1 adjacent transpositions: (1 2), (2 3), . . . , (n 1 n).
Property 1.2: The order of any subgroup H of G is a sub-multiple of the order n of G.
To show this let m be the order of a subgroup H of G. If m = n or m = 1, H would
coincide with G or with {e} respectively and the property would trivially hold. Suppose
1 < m < n. Then there is an element g1 G which is not in H. Consider the set H g1 ,
called left coset of G, consisting of all the elements of the form h g1 with h in H. This
set has no elements in common with H, since, if there were an element of the form h g1
contained in H, also h1 (h g1 ) = g1 would be in H, which contradicts our assumption
. The number of elements of H g1 equals the order of H, since different elements of H
correspond to different elements of H g1 . To show this suppose there existed two distinct
elements h1 , h2 H such that: h1 g1 = h2 g1 . Multiplying both sides of the equation to the
right by g11 this equality would become h1 = h2 which is against our assumption. The set
H H g1 has then 2 m elements. If it coincided with G the property would be proven since
this would imply that n = 2 m. Suppose H H g1 be strictly contained in G. Then there is
an element g2 G which is not contained in H H g1 . Consider now the left coset H g2 .
This set is disjoint from H and from H g1 , as it follows from the fact that if there existed
two elements h1 , h2 H such that h1 g1 = h2 g2 , then, multiplying by h1
2 both sides to
1
the left we would obtain g2 = (h2 h1 ) g1 which is an element of H g1 , contradicting our
hypothesis on g2 . If the set H H g1 H g2 is still strictly contained in G we iterate our
construction until we decompose G in the following direct sum:
G = H H g1 H g2 . . . H gk ,

(1.3.11)

where g1 , . . . , gk are distinct elements of G which are not in H and (H gi ) (H gj ) = .


Since the direct sum on the right hand side of the above equation has dimension k m, we
have proven that n = k m. The number k is called index of H in G. If n is prime G can
only have a subgroup of order one, which is {e}. Moreover if n > 1 is prime, G can only be

1.4. CLASSES AND NORMAL SUBGROUPS

17

cyclic, since otherwise it would contain the cyclic group generated by one of its elements as
a subgroup of order 1 < m < n which will be a sub-multiple of n.
Following the same procedure, we could have decomposed G into the following direct
sum:
G = H g1 H g2 H . . . gk H ,

(1.3.12)

where gi H is called right coset and is the set of elements of G of the form gi h, with h H.
We denote by G/H and HG the collections of the k right and left cosets respectively:
G/H {H, H g1 , H g2 , . . . , H gk } ,
HG {H, g1 H , g2 H , . . . , gk H} .

1.4

(1.3.13)

Classes and Normal Subgroups

We say that two elements g1 , g2 of G are conjugate to one another in G, and write g1 g2
iff there exist a third element g G such that:
g1 = g g2 g 1 ,

(1.4.1)

g1 is also called conjugate of g2 by g and we say that it is obtained through the adjoint action
of g on g2 . It is straightforward to prove that is an equivalence relation, namely that it is
reflexive, symmetric and transitive. The elements of G arrange in equivalence classes with
respect to (conjugation classes). Clearly e forms a single class and, if G is abelian, each
element forms a different class.

1.4.1

Conjugation classes in Sn and Young Diagrams.

Let us consider the effect of a conjugation of a permutation. Let S and T be the following
two permutations:




1
2
...
n
1
2
...
n
S =
, T =
, (1.4.2)
S(1) S(2) . . . S(n)
T (1) T (2) . . . T (n)
and let us compute T S T 1 :


=

S(1)
T (S(1))

S(2)
...
T (S(2)) . . .

T (1)
T (2)
...
T (S(1)) T (S(2)) . . .

 
1

S(1)

T (n)
,
T (S(n))
S(n)
T (S(n))

2
S(2)

...
n
. . . S(n)

 
T (1)

T (2)
2

...
...

T (n)
n


=

(1.4.3)

namely the conjugate of S by means of T is the permutation obtained by acting on the first
and second rows of S by T . For example:




1 2 3 4 5 6
1 2 3 4 5 6
S =
, T =
then:
4 6 5 1 3 2
2 1 3 5 4 6


2 1 3 5 4 6
0
1
S = T ST =
.
(1.4.4)
5 6 4 2 3 1

18

CHAPTER 1. ABSTRACT GROUPS

If we write S as the product of cycles acting on different subsets of objects, as in (1.3.3), its
conjugate by T is then obtained by replacing each element in the cycle by their transformed
by T . For instance:
S = (1 2 6) (3 5) , T = (1 6 4) then:
S 0 = T S T 1 = (2 4 6) (3 5) ,

(1.4.5)

since T maps 1 6, 6 4, 4 1, 3 3, 5 5. In this example both S and S 0


have the same permutation structure, namely decompose in the same kind of cycles. In
the notation introduced in the previous section their structure is (1 , 2 , 3 , 4 , 5 , 6 ) =
(1, 1, 1, 0, 0 0). It is not difficult to convince ourselves that conjugation in general does not
alter the structure of a permutation. Moreover two permutations with the same structure are
clearly conjugate to one another. We conclude that the conjugation classes of Sn are in one
to one correspondence with the possible permutation structures, i.e. are labeled by the nplets
(1 , 2 , . . . , n ). Consider for instance S3 , whose content is displayed in (1.3.10). In this
case the group consists in three classes: (3, 0, 0), (1, 1, 0), (0, 0, 1), the first consisting of
one element (i.e. the identity), the second by three elements and the third by two elements.
Consider as a further example the group S4 whose class content is summarized below:
S4 :
class
(4, 0, 0,
(2, 1, 0,
(0, 2, 0,
(1, 0, 1,
(0, 0, 0,

0)
0)
0)
0)
1)

elements
I = (1) (2) (3) (4)
(1 2), (1 3), (1 4), (2 3), (2 4), (3 4)
(1 2) (3 4), (1 3) (2 4), (1 4) (2 3)
(1 2 3), (3 2 1), (1 2 4), (4 2 1), (1 3 4), (4 3 1), (2 3 4), (4 3 2)
(1 2 3 4), (1 2 4 3), (1 3 4 2), (1 4 3 2), (1 4 2 3), (1 3 2 4)

(1.4.6)

We have 12 even elements in the classes (4, 0, 0, 0), (0, 2, 0, 0), (1, 0, 1, 0) and 12 odd in
the remaining classes.
The conjugation classes of Sn are then defined by the solutions to equation (1.3.7).
Let us determine the number of elements within a generic class (1 , 2 , . . . , n ) of Sn . The
problem can be restated as that of determining the number of distinct ways n objects can be
distributed in a number of boxes, of which 1 can contain just one element, 2 two elements
. . . n n elements. Each class (1 , 2 , . . . , n ) is therefore in one to one correspondence with
the distinct partitions of n objects and can be represented as follows:
1 1cycles

2 2cycles

n ncycles

z }| { z
}|
{
z
}|
{
() . . . () () . . . () . . . ( . . . ) . . . ( . . . ) ,

(1.4.7)

the objects being represented by the symbol. Given a class, the number of its elements can
be computed as the number of permutations which change the element within the class. The
total number of permutations we can perform on the n objects is n!. Out of these, permutations whose effect is to exchange cycles of the same order k will leave the element invariant.
There are 1 ! 2 ! . . . n ! such permutations. Moreover cyclic permutations within each cycle

1.4. CLASSES AND NORMAL SUBGROUPS

19

will not change the cycle and thus the element. For instance (1 2 3) = (2 3 1) = (3 1 2). For
each of the k kcycles there are k such permutations and thus there are 11 22 . . . nn more
permutations which leave a generic element of the class invariant. The total number of permutations which do not effect a single element of a given is therefore 11 1 ! 22 2 ! . . . nn n !
and thus the number of elements nc in the class is:
nc =

11

! 22

n!
.
2 ! . . . nn n !

(1.4.8)

Applying this formula to S4 we find:


n(4 0 0 0) =
n(2 1 0 0) =
n(0 2 0 0) =
n(1 0 1 0) =
n(0 0 0 1) =

4!
= 1,
4!
4!
= 6,
4
1 2! 21 1!
4!
= 3,
2
2 2!
4!
= 8,
11 1! 31 1!
4!
= 6,
1
4 1!
14

in accordance with table 1.4.6.


We can alternatively describe each partition of n, i.e. each class in Sn , by a new set of
indices (1 , . . . , n ), where the positive indices i are related to i as follows:
1 = 1 + 2 + . . . + n ,
2 = 2 + . . . + n ,
..
.
,
n = n .

(1.4.9)

Pn
By definition we have: 1 2 . . . n and, in virtue of eq. (1.3.7),
i=1 i = n.
Moreover eqs. (1.4.9) are readily inverted to yield: 1 = 1 2 , . . . , n1 = n1 n , n =
n . Each class (1 , . . . , n ) is represented graphically by a Young diagram which consists
of n boxes, distributed in a number of rows of which, moving downwards, the first contains
1 boxes, the second 2 boxes and so on: Let us apply this formalism to the S3 and S4 case
and write the Young diagrams corresponding to their classes:

S3 :

(1 , 2 , 3 )

(1 , 2 , 3 )

(3, 0, 0)

(3, 0, 0)

(1, 1, 0)

(2, 1, 0)

(0, 0, 1)

(1, 1, 1)

Young diagrams

(1.4.10)

20

CHAPTER 1. ABSTRACT GROUPS

Figure 1.3: Young diagram

S4 :

1.4.2

(1 , 2 , 3 , 4 )

(1 , 2 , 3 , 4 )

(4, 0, 0, 0)

(4, 0, 0, 0)

(2, 1, 0, 0)

(3, 1, 0, 0)

(0, 2, 0, 0)

(2, 2, 0, 0)

(1, 0, 1, 0)

(2, 1, 1, 0)

(0, 0, 0, 1)

(1, 1, 1, 1)

Young diagrams

(1.4.11)

Conjugation Classes in Dn

Let us use the defining relations (1.2.13) for Dn to construct its conjugation classes. From
the relation (r )2 = e we find r = r1 , which implies, upon multiplication to the left
by suitable powers of r that:
r2 = r r1 , r3 = r (r ) r1 . . .
r2k = r (r2k2 ) r1 , r2k+1 = r (r2k1 ) r1 .

(1.4.12)

1.4. CLASSES AND NORMAL SUBGROUPS

21

From the above relations we conclude that all reflections of the form r2k , for some positive
integer k, are in the same class as , while all reflections r2k+1 , are in the same class as
r . If n is odd, n = 2k + 1 for some k, and thus = r2k+1 is in the same class as r and
thus all reflections belong to the same class. If n is even, reflections group into two classes:
n odd :
[] = {r` } , ` = 0, 1, 2, n 1 ,
n even :
[] = {, r2 , . . . , rn2 } ,
[r ] = {r , r3 , . . . , rn1 } .

(1.4.13)

As far as the rotations are concerned, using the same relation r = r1 = rn1 we
find that r = rn1 , r2 = rn2 and so on, so that r is in the same class as rn1
(recall that 1 = ), r2 in the same class as rn2 and so on. If n = 2k, we have k 1
classes [r` ] = {r` , rn` }, ` = 1, . . . , k 1 and a single class [rk ] = {rk }. If n = 2k + 1, we
have k classes [r` ] = {r` , rn` }, ` = 1, . . . , k. In all cases we have the class [e] consisting of
the identity element only. Summarizing:
D2k+1 :

D2k

1.4.3

k + 2 classes
[e] , [r` ] = {r` , rn` } (` = 1, . . . , k)
[] = {, r , . . . , rn1 } ,
: k + 3 classes
[e] , [r` ] = {r` , rn` } (` = 1, . . . , k 1) , [rk ] ,
[] = {, r2 , . . . , rn2 } ,
[r ] = {r , r3 , . . . , rn1 } .

(1.4.14)

Normal Subgroups

Let H be a subgroup of G and, for a given element g of G let us define the set g H g 1 as
the set of all elements of the form g h g 1 , with h H. This set is actually a group itself.
To show this let us take any two of its elements g1 = g h1 g 1 , g2 = g h2 g 1 and show
that g1 g21 is in g H g 1 :
1
g1 g21 = (g h1 g 1 ) (g h2 g 1 )1 = (g h1 g 1 ) (g h1
2 g ) =
1
= g (h1 h1
g H g 1 ,
(1.4.15)
2 )g

since h1 h1
2 is in H.
H is said to be a normal or invariant subgroup of G iff, for any g G, g H g 1 = H,
i.e. for any h H, g h g 1 H. This implies that for normal subgroups the left and right
cosets coincide: g H = H g. Clearly H = {e} is a normal subgroup of G, since, for any
g G, g e g 1 = g g 1 = e H.

22

CHAPTER 1. ABSTRACT GROUPS

Property 1.3: If H is a normal subgroup of G, the set G/H of cosets H gi close a group
called the factor group.
To show this let us prove that we can consistently define a product in G/H. Given two
elements H gi , H gj of G/H, define (H gi ) (H gj ) as the set of all elements of the form
(h1 gi ) (h2 gj ), with h1 , h2 H. We can write this generic element in the following form:
h1 (gi h2 gi1 ) gi gj . Since H is normal, h3 = gi h2 gi1 is still in H, and thus a generic
element of (H gi ) (H gj ) can be written as (h1 h2 ) (gi gj ) H (gi gj ). Similarly a
generic element h (gi gj ) of H (gi gj ) can be written as (h gi ) (e gj ) (H gi ) (H gj ).
We conclude that:
(H gi ) (H gj ) = H (gi gj ) .

(1.4.16)

Clearly the identity element with respect to this coset-product is H = H e itself. Eq. (1.4.16)
also implies that the inverse of a generic coset H g is the coset H g 1 . Associativity of this
product is also implied by the same property of the product in G.
Remark: Notice that if H is not a normal subgroup of G, the element (h1 gi ) (h2 gj ),
for different h1 , h2 H would not belong to a same coset and thus the product on G/H
would not be properly defined.
One may define normal elements of G and elements g 0 G such that, for any g G, we
have g g 0 g 1 = g 0 . As a consequence a normal element of G commutes with all elements of
G. Indeed, since for any g G, g g 0 g 1 = g 0 , multiplying both sides by g to the right we
find: g g 0 = g 0 g. An example of normal element is the identity element e. The set C G
of all normal elements of G is called the center of G. It is a subgroup of G. Indeed consider
g 0 , g 00 C and show that g 0 g 00 1 C. Take any g G:
g (g 0 g 00 1 ) g 1 = (g g 0 g 1 ) (g g 00 1 g 1 ) = (g g 0 g 1 ) (g g 00 g 1 )1 = g 0 g 00 1 ,
which implies that g 0 g 00 1 C. Since any element of C commutes with all the elements of
G, it will also commute with all the elements of C itself, namely C is abelian. C is clearly
a normal subgroup of G.

1.5

Homomorphisms, Isomorphisms and Automorphisms

Given two groups G, G0 a mapping from G to G0 is a homomorphism iff, given any


two elements g1 , g2 in G: (g1 , g2 ) = (g1 ) (g2 ). In other words a homomorphism
preserves the product. The group G is said to be homomorphic to G0 . We shall denote a
homomorphism of G into G0 by G G0 . here are come properties of homomorphisms:
1. A first property of homomorphisms is that (e) = e0 , e0 being the identity element
of G0 . Indeed take an element g 0 G0 which is the image of some element g of G,
g 0 = (g), and consider the product g 0 (e) = (g) (e) = (g e) = (g) = g 0 e0 .
Multiplying both sides by g 0 1 to the left we find: (e) = e0 .
2. As a second property, for any g G, (g)1 = (g 1 ). Indeed (g) (g 1 ) =
(g g 1 ) = (e) = e0 .

1.5. HOMOMORPHISMS, ISOMORPHISMS AND AUTOMORPHISMS

23

3. If H is a subgroup of G, H 0 = (H) is a subgroup of G0 . Take two elements h01 =


(h1 ), h02 = (h2 ) of H 0 , image of h1 , h2 H. We have h01 h021 = (h1 ) (h1
2 ) =
1
1
0
(h1 h2 ) H , since h1 h2 is an element of H. As a consequence of this, since G
is in particular subgroup of itself, it follows that (G) is a subgroup of G0 .
4. If H is the largest subset of G whose image though is a subgroup H 0 = (H) of G0 ,
then H is a subgroup of G. Take two elements h1 , h2 of H and evaluate (h1 h1
2 ) =
1
0
0
(h1 ) (h2 ) . This element is in H , since (h1 ), (h2 ) are and H is a subgroup of
G0 . Now, by assumption, H consists of all the elements of G whose image through
is in H 0 . We then conclude that h1 h1
2 H and thus that H is a subgroup of G.
5. If H is a normal subgroup of G then H 0 = (H) is a normal subgroup of (G). Take
H normal subgroup of G, and consider a generic element g 0 = (g) of (G). We have
that g 0 H 0 g 0 1 = (g H g 1 ) = (H) = H 0 .
6. If H is the largest subset of G such that its image though coincides with a normal
subgroup H 0 = (H) of G0 , then H is a normal subgroup of G. Let now H 0 be a normal
subgroup of G0 and consider the subgroup g H g 1 of G, for a given g G. Its image
(g H g 1 ) = g 0 H 0 g 0 1 = H 0 . By assumption we conclude that g H g 1 = H,
namely that H is normal.
7. The set E G of all the elements of G which map through into e0 is a normal
subgroup of G. This follows form the previous property, applied to H 0 = {e0 } which is
a normal subgroup of G0 .
8. A homomorphism is one-to-one if and only if the only element of G which is mapped
into the identity element e0 of G0 is the identity e. Let us start assuming the homomorphism to be one-to-one, which means that (g1 ) = (g2 ), g1 , g2 G implies
g1 = g2 . Then suppose (g) = e0 . Since also (e) = e0 = (g) it follows by hypothesis that g = e. Let us now assume that the identity e of G is the only element which is mapped into the identity e0 of G0 and suppose that two elements
g1 , g2 G are mapped into the same element (g1 ) = (g2 ) G0 . This implies that
(g1 g21 ) = (g1 ) (g2 )1 = (g1 ) (g1 )1 = e0 . By assumption we must have
g1 g21 = e, that is g1 = g2 .
We shall restrict from now on to onto homomorphisms: (G) = G0 . If the onto homomorphism is also one-to-one it is called an isomorphism. The concept of isomorphism is
important since two isomorphic groups have the same order and the same structure. From
this point of view two isomorphic groups can be considered as a same one. For isomorphic
groups we shall use the short-hand notation G G0 .
Property 1.4: If G is homomorphic to G0 , the G/E is isomorphic to G0 .
Let us first define the correspondence between G/E and G0 , by showing that all the
elements of a coset E g G/E map into a single element g 0 = (g) of G0 . Indeed, for any
e E, we have (
e g) = (
e) (g) = e0 g 0 = g 0 . We can then define (E g) as g 0 = (g).

24

CHAPTER 1. ABSTRACT GROUPS

Let us now show that the correspondence is one-to-one, namely that if (E g1 ) = (E g2 ),


we must have E g1 = E g2 . Indeed let (E g1 ) = (g1 ) = g10 = g 0 = g20 = (g2 ) = (E g2 ).
Then g1 and g2 are two elements of G which map into g 0 . This implies that g1 g21 is in
E since (g1 g21 ) = (g1 ) (g2 )1 = g 0 g 0 1 = e0 . As a consequence of this we have
E g1 = E (g1 g21 ) g2 = E g2 . We have thus shown that the mapping : G/E G0
is one-to-one. We need now to show that it preserves the product: Let g10 = (E g1 ) and
g20 = (E g2 ). The product (E g1 )(E g2 ) is the coset E (g1 g2 ) and its image through is:
(E (g1 g2 )) = (g1 g2 ) = g10 g20 . This proves that : G/E G0 is also a homomorphism.
Property 1.5: Let H be a normal subgroup of G, then G is homomorphic to the factor
group G/H.
Let us define : G G/H as the mapping which associates with each element of G the
coset it belongs to:
:

g G (g) = H g G/H .

(1.5.1)

Clearly this mapping is not one to one. Let us show it is a homomorphism:


(g1 g2 ) = H (g1 g2 ) = H H (g1 g2 ) = (H g1 ) (g11 H g1 ) g2 =
= (H g1 ) (H g2 ) = (g1 ) (g2 ) ,
(1.5.2)
where we have used the property that H be normal. Finally is onto since any coset in
G/H can be written as H g and thus is the image of g through .
Therefore if we have a homomorphism G G0 , since G/E G0 we can reduce it to the
homomorphism G G/E.
A homomorphism of G onto itself is called an endomorphism while an isomorphism of G
onto itself is called an automorphism. A meromorphism is a one-to-one endomorphism of
G which is not necessarily onto. In particular a proper meromorphism is a meromorphism
which is not an automorphism, namely such that (G) ( G. If G admits a proper meromorphism it is necessarily infinite. Indeed is G were finite, being a meromorphism a one-to-one
correspondence, (G) would have the same order as G and thus would coincide with G.
Exercise 1.4: Consider the infinite set G of all the integer powers of a positive number
a: G = {ak }, k Z. Show that it is a group and show that the mapping : G G such
that (ak ) = a2 k is a proper meromorphism.
Example : Let us show that the groups S3 and D3 are isomorphic. Consider the following
mapping:
: D3 S3 ,
(e) = I , (r) = (1 2 3) , (r ) = (3 2 1) , () = (1 2) ,
(r ) = (1 3) , (r2 ) = (1.5.3)
(2 3) .
2

The reader can easily show that is a homomorphism. Being one-to-one and onto, it is an
isomorphism.

1.5. HOMOMORPHISMS, ISOMORPHISMS AND AUTOMORPHISMS

1.5.1

25

The Automorphism Group

Consider two automorphisms 1 , 2 of a group G. They are in particular functions of G


into itself and thus we can consider their composition 2 1 as the mapping of G into itself
such that associates with any g G the element 2 1 (g) = 2 (1 (g)). Let us show that
2 1 is still an automorphism. It is a homomorphism since for any g1 , g2 G:
2 1 (g1 g2 ) = 2 (1 (g1 g2 )) = 2 (1 (g1 ) 1 (g2 )) = 2 (1 (g1 )) 2 (1 (g2 )) =
= 2 1 (g1 ) 2 1 (g2 ) .
(1.5.4)
Now let us prove that 2 1 is onto. Consider g 00 G. Being 1 , 2 onto, there exist g 0 G
such that 2 (g 0 ) = g 00 and g G such that 1 (g) = g 0 . This implies that there exist g G
such that g 00 = 2 (g 0 ) = 2 (1 (g)) = 2 1 (g) and thus 2 1 is onto. Let us prove now
that it is one-to one. Consider g1 , g2 G such that 2 1 (g1 ) = 2 1 (g2 ). This means
that 2 (1 (g1 )) = 2 (1 (g2 )). Being 2 one-to-one we have that 1 (g1 ) = 1 (g2 ), and, being
1 one-to-one, it follows that g1 = g2 .
We can consider now the set of all automorphisms of G endowed with the product .
Let us show that it is a group, called the automorphism group AG of G. To this end we
define the identity automorphism 1 as the trivial function which maps each element of G
into itself: 1(g) = g, for any g G. From the definition we can convince ourselves that,
given any automorphism : 1 = 1 = . The function 1 is trivially an automorphism.
Since an automorphism is one-to-one and onto, it can be inverted. Its inverse 1 is then
such that 1 will map g G into g 0 G iff (g 0 ) = g. The composition of with its
inverse function is clearly the identity automorphism:
1 = 1 = 1 .

(1.5.5)

Let us show that 1 is still an automorphism. Let g10 = (g1 ) and g20 = (g2 ) be two
elements of G. By definition, 1 (g10 ) = g1 and 1 (g20 ) = g2 . Then 1 (g10 g20 ) is defined as
the only element of G which is mapped through into g10 g20 . This element is g1 g2 , and so
1 (g10 g20 ) = g1 g2 = 1 (g10 ) 1 (g20 ). Since 1 is, by construction, one-to-one and onto,
it is an isomorphism.
We still need to prove that the product is associative. Given a generic element g G
we find:
3 (2 1 )(g) = 3 (2 1 (g)) = 3 (2 (1 (g))) = 3 2 (1 (g)) =
= (3 2 ) 1 (g) .

(1.5.6)

Let us now consider, for a given h G the correspondence h :


h : g G h g h1 G .

(1.5.7)

Clearly e = 1. Let us show that h is an automorphism. It is a homomorphism since,


given two elements g1 , g2 G, we have h (g1 g2 ) = h (g1 g2 ) h1 = (h g1 h1 ) (h
g2 h1 ) = h (g1 ) h (g2 ). h is one to one: Indeed suppose h (g1 ) = h (g2 ), which means

26

CHAPTER 1. ABSTRACT GROUPS

h g1 h1 = h g2 h1 . Multiplying both sides by h1 to the left and by h to the right we


find g1 = g2 . Finally h is onto: For any g 0 G there exist g = h1 g 0 h G such that
g 0 = h (g).
h is called inner automorphism of G. The set of all the h , for all h G, is a group.
To show this consider h1 and h2 , with h1 , h2 G and let us show that
h2 h1 = h2 h1 ,

(1.5.8)

by computing this product on a generic g G:


1
1
h2 h1 (g) = h2 (h1 (g)) = h2 (h1 g h1
= h2 h1 (g) .
1 ) h2 = (h2 h1 ) g (h2 h1 )

From this it follows that h1 = h1 since h h1 = h1 h = e = 1. Given any two inner


automorphisms h1 and h2 the product of the first times the inverse of the second is still an
inner automorphism. Indeed h1 h1
= h1 h1
= h1
. We have then proved that the
2
2
2 h1
inner automorphisms close a subgroup IG of AG called the group of inner automorphisms.
Property 1.6: IG is a normal subgroup of AG .
Let Ag be an automorphism of G and h IG . Let us show that h 1 IG .
To this end we compute this function on a generic element g G:
h 1 (g) = h ( 1 (g)) = (h ( 1 (g))) = (h 1 (g) h1 ) =
= (h) g (h)1 = (h) (g) .
(1.5.9)
This shows that h 1 = (h) IG .
We can define a correspondence of G into its group of inner automorphisms IG which
maps a generic element g G into g . In virtue of equation (1.5.8) this is a homomorphism:
G IG . It is not however one-to-one. Indeed for any h in the center C of G, h = 1:
h (g) = h g h1 = g = 1(g) ,

(1.5.10)

where we have used the property that any element of C commutes with any other element
of G. In fact C contains all the elements of G which are mapped into the identity of IG . In
virtue of Property 1.4 we have that G/C IG .
Example 1.1: Consider an infinite cyclic group G = {ak } consisting of the integer powers
of a generator a. If is an endomorphisms of G, it will map a into an other element of the
group a` , for some integer ` (we shall suppose ` 6= 0). Being a homomorphism, it will map
ak into (a)k = a` k . Therefore the most general endomorphism of a infinite cyclic group
has the following action (ak ) = a` k , for some integer `. This endomorphism is one-to-one,
namely it is a meromorphism. Clearly (G) is a proper subgroup of G if ` 6= 1, as the
reader can show. The only automorphisms then correspond to the cases in which ` = 1.
AG is an order 2 group. Since G is abelian, it coincides with its center: G = C. As a
consequence IG G/C is an order one group consisting of the identity 1 alone.

1.6. PRODUCT OF GROUPS

1.6

27

Product of Groups

Given two groups G1 and G2 we define their product G1 G2 as a group consisting of couples
of elements (g1 , g2 ), g1 G1 and g2 G2 , on which the following product is defined:
(g1 , g2 ), (g10 , g20 ) G1 G2

(g1 , g2 ) (g10 , g20 ) (g1 g10 , g2 g20 ) G1 G2 . (1.6.1)

From the above definition, if we denote by e1 , e2 are the unit elements of G1 and G2 respectively, it follows that:
The identity element of G1 G2 is (e1 , e2 ) since, for any (g1 , g2 ) G1 G2 : (g1 , g2 )
(e1 , e2 ) = (g1 e1 , g2 e2 ) = (g1 , g2 ) = (e1 , e2 ) (g1 , g2 );
For any element (g1 , g2 ) G1 G2 , its inverse (g1 , g2 )1 is (g11 , g21 ): (g1 , g2 )
(g11 , g21 ) = (g1 g11 , g2 g21 ) = (e1 , e2 ) = (g11 , g21 ) (g1 , g2 );
The associative property of the product in G1 G2 follows from the same property of
the product in the two factors.
Consider the set of elements {(g1 , e2 )}, g1 G1 , of G1 G2 . This subset is a subgroup, as the
reader can easily verify, which corresponds to the product G1 {e2 }. The correspondence:
g1 G1

(g1 , e2 ) G1 {e2 } ,

(1.6.2)

is an isomorphism: G1 G1 {e2 } G1 G2 . Similarly we can define the subgroup


{e1 } G2 of G1 G1 and show that G2 {e1 } G2 G1 G2 , that is G2 is isomorphic to
{e1 } G2 . The two subgroups G1 {e2 } and {e1 } G2 commute, namely a generic element
of the former commutes with any element of the latter:
(g1 , e2 ) G1 {e2 } , (e1 , g2 ) {e1 } G2 :
(g1 , e2 ) (e1 , g2 ) = (g1 e2 , g2 e2 ) = (e1 g2 , e2 g2 ) = (e1 , g2 ) (g1 , e2 ) .

1.7

Matrix Groups

We end this section by defining the main continuous matrix groups that we shall consider in
this course. They represent the main instances of Lie groups. Let F denote either the real or
the complex numbers. A matrix groups over F is a set of square matrices with entries in F,
endowed with the ordinary rows-times-columns product, and satisfying certain properties.

1.7.1

The Group GL(n, F)

We have already defined GL(n, F) as the group of all nn non-singular matrices with entries
in F, that is GL(n, R) and GL(n, C) are the groups of all n n real and complex matrices
respectively. Since real numbers are in particular complex numbers, namely R C, real
matrices are particular instances of complex matrices and therefore GL(n, R) is a subgroup

28

CHAPTER 1. ABSTRACT GROUPS

of GL(n, C). Its elements are defined by restricting to the complex matrices in GL(n, C)
with real entries. GL(n, R) is said to be a real form of GL(n, C).
As we shall see in next chapter, this group described a generic change of basis in a
linear vector space, or a generic linear homogeneous transformation on a space of points. A
generic element GL(n, F) depends on its entries which are n2 continuous parameters in F:
M = M[aij ] = (aij ). If F = R, the parameters will be n2 real, if F = C, there will be n2
complex parameters, that is 2 n2 real.

1.7.2

The Group SL(n, F)

SL(n, F) is the set of n n matrices with entries in F with unit determinant:


SL(n, F) = {S = (Sij ) : det(S) = 1} .

(1.7.1)

Clearly SL(n, F) GL(n, F). Let us show that SL(n, F) is actually a subgroup of GL(n, F).
Consider S = (Sij ) and T = (Tij ) in SL(n, F), let us prove that S T1 is still in SL(n, F).
det(S)
= 1. This proves
By assumption det(S) = det(T) = 1. Therefore det(S T1 ) = det(T)
1
that S T is an element of SL(n, F). The group SL(n, F) describes, as we shall illustrate
in next chapter, special linear transformations, which are volume preserving homogeneous
linear transformations on an n-dimensional space over F. As for the GL groups, SL(n, R)
is a real form of SL(n, C). A generic element of SL(n, F) depends on its entries which are
n2 continuous parameters in F, subject to the condition det(S) = 1. In the SL(n, C) case,
det(S) = 1 is one condition on a complex number which implies two real conditions on
the 2 n2 real parameters, leaving a generic element of the group to depend on 2 n2 2 free
parameters. As for SL(n, R), det(S) = 1 is one real condition on n2 real parameters and
thus the number of real independent parameters defining a generic element of the group is
n2 1.

1.7.3

The Group O(n, F)

The group O(n, F) is defined as the group of n n orthogonal matrices with entries in F,
namely of matrices S = (Sij ) satisfying the property:
ST S = 1n ,

(1.7.2)

or, in components
n
X

Ski Skj = ij ,

(1.7.3)

k=1

where ST denotes the transposed matrix of S, ij is 1 for i = j and 0 for i 6= j, 1n = (ij )


is the n n identity matrix. Multiplying both sides of equation (1.7.2) by S1 to the right,
we find that the inverse of any orthogonal matrix coincides with its transposed: S1 = ST .
Moreover one also finds, using the property that S S1 = S1 S = 1n , that S ST = 1n .

1.7. MATRIX GROUPS

29

Computing the determinant of both sides of eq. (1.7.2) we deduce that det(S)2 = 1, that
is det(S) = 1. To show that O(n, F) is a group, let us show it is a subgroup of GL(n, F).
Consider two elements of S and T of O(n, F), let us show that S T1 = S TT satisfies eq.
(1.7.2):
(S T1 )T (S T1 ) = (S TT )T (S TT ) = (T ST ) (S TT ) = T (ST S) TT =
= T TT = 1n .
(1.7.4)
We define the group of special orthogonal matrices SO(n, F) as the subset of orthogonal
matrices, with entries in F, which have unit determinant: det(S)
T = 1. Clearly SO(n, F) is
a subgroup of SL(n, F). More precisely: SO(n, F) = O(n, F) SL(n, F). For the sake of
simplicity we shall denote simply by (S)O(n) the group (S)O(n, R).
Exercise 1.5: Prove that SO(n, F) is a subgroup of O(n, F), while the subset of orthogonal matrices with determinant 1 is not.
Exercise 1.6: Prove that SO(n, F) is a normal subgroup of O(n, F).
As we shall see in next chapter, special orthogonal transformations describe rotations
in a n-dimensional space of points (e.g. the Euclidean space) over F, while orthogonal
transformations with determinant 1 describe reflections in the same space.
Exercise 1.7:
Prove that the matrix

S[] =

cos() sin()
sin() cos()


,

(1.7.5)

is in SO(2). In fact S[] represents a generic element of this group and, as we shall see
it describes a rotation about a point by an angle in the Euclidean plane. The reader
can show that the structure of this group is described by the relation S[1 ] S[2 ] =
S[1 + 2 ], which represents the simple fact that the result of two consecutive rotations
on the plane about a point is itself a rotation about the same point by an angle which
is the sum of the angles defining the two rotations.
Show that

R =

1 0
0 1


,

(1.7.6)

is in O(2) but not in SO(2). As we shall see, this matrix represents a reflection in the
Y axis in the Euclidean plane: x x, y y.
Finally prove that the matrix

0 0 1
S = 0 1 0 ,
1 0 0
is in O(3) but not in SO(3).

(1.7.7)

30

CHAPTER 1. ABSTRACT GROUPS

Exercise 1.8: Prove that the matrix


1
S =

3
1
2
1
6

1
3
12
1
6

1
3

0 ,
26

(1.7.8)

is in SO(3).
Consider the defining condition (1.7.3) for the real group O(n, R) and take the equations
corresponding to j = i:
n
X

(Ski )2 = 1 .

(1.7.9)

k=1

Since Sij are real numbers, the left hand side is a sum of positive numbers which has to be
equal to one. The above equation implies then that each term in the sum be less or equal
than one, i.e. |Sij | 1. We shall call bounded a matrix whose entries, as functions of free
parameters, are all, in modulus, bounded from above. We conclude that any element of
O(n, R) is a bounded matrix. A group of matrices is called compact if all its elements are
bounded. For this reason O(n, R) is a compact group.
Using equation (1.7.3), we can count the number of free continuous parameters a generic
O(n, F) matrix depends on, i.e. the dimension of the group. This number would be given by
the n2 parameters in F of a generic matrix in GL(n, F), minus the number of independent
conditions in eq. (1.7.3). We see that conditions (1.7.3) are symmetric in i, j: If we interconditions
change the values of i and j the condition is the same. This leaves us with n(n+1)
2
on numbers in F. The number of free parameters will then be:
n2

n(n 1)
n(n + 1)
=
2
2

parameters in F,

(1.7.10)

O(n, R) will then be described by n(n1)


real parameters, O(n, C) by n(n1) real parameters.
2
Since the orthogonality condition (1.7.3) already implies that det(S) = 1, the condition
of unit determinant, defining the SO(n, F) subgroup, is just a restriction on the sign of the
determinant which does not lower the number of independent parameters.
Relation between O(n) and SO(n) Since SO(n) is a normal subgroup of O(n) we can
construct the factor group O(n)/SO(n). This is done noticing that any two orthogonal
matrices O1 , O2 with determinant 1 are connected by the left (or right) action of a SO(n)
element S: O2 = S O1 . Indeed it suffices to take S = O2 O1
1 . Then take any orthogonal
matrix O with determinant 1. We will take for instance O = diag(1, +1, +1, . . . , +1), so
that O2 = 1n . We can decompose O(n) in the following cosets:
O(n) = SO(n) + SO(n) O .

(1.7.11)

O(n)/SO(n) is an order 2 group, isomorphic to {1n , O}, which is a cyclic group of period 2,
also denoted by Z2 .

1.7. MATRIX GROUPS

1.7.4

31

The Group O(p, q; F)

A generalization of the concept of orthogonal group is the group O(p, q; F), p + q = n,


consisting of pseudo-orthogonal n n matrices with entries in F. A pseudo-orthogonal
matrix S = (Sij ) is defined by the following property:
ST p,q S = p,q ,

(1.7.12)

or, in components
p
X

Ski Skj

k=1

n
X

Ski Skj = (p,q )ij ,

(1.7.13)

k=p+1

where p,q ia s diagonal matrix with p (+1) and q (1) diagonal entries:
q

p,q

z }| { z }| {
= diag(+1, . . . , +1, 1, . . . , 1) .

(1.7.14)

We leave to the reader to show that this condition defines a group, subgroup of GL(n, F).
In particular we can easily derive the following relations:
S1 = p,q ST p,q , S p,q ST = p,q , det(S) = 1 .
If p = n and q = 0 or p = 0 and q = n, p,q
O(p, q; F) = O(n, F). Consider the real matrix I

0q,p
I =
1p

(1.7.15)

becomes proportional to 1n and therefore


which we write in blocks as follows:

1q
,
(1.7.16)
0p,q

where 0p,q is the pq matrix made of zeros. The matrix I is orthogonal: I I T = 1n . Moreover
I has the property that I p,q I T = q,p . Therefore if S is in O(p, q; F), S0 = I S I T is in
O(q, p; F), as the reader can easily verify. In other words we can write:
O(q, p; F) = I O(p, q; F) I T ,

(1.7.17)

namely the two groups are isomorphic, they have the same structure.
Now consider the pseudo-orthogonal group over the complex numbers: O(p, q; C) and
define the following complex matrix:
p

Cp,q

z }| { z }| {
= diag(+1, . . . , +1, i, . . . , i) .

(1.7.18)

Clearly Cp,q
is the matrix Cp,q
whose entries are the complex conjugate of the corresponding
1 1
entries of Cp,q . The reader can also verify that Cp,q Cp,q = Cp,q
Cp,q = p,q . If S is an orthogonal
0
1
complex matrix, S = Cp,q S Cp,q is in O(p, q; C). Indeed in the equation ST S = 1n we can
express S in terms of S0 :
1
1 0
1n = (Cp,q S0 T Cp,q
) (Cp,q
S Cp,q ) = Cp,q S0 T p,q S0 Cp,q ,

(1.7.19)

32

CHAPTER 1. ABSTRACT GROUPS

T
= Cp,q . Multiplying both sides of
where we have used the symmetry property of Cp,q : Cp,q
1
the above equation, to the left and to the right by Cp,q , we find

p,q = S0 T p,q S0 ,

(1.7.20)

namely that S0 O(p, q; C). We conclude that the orthogonal group over the complex
numbers and the pseudo-orthogonal one, for any partition p, q of n, are isomorphic through
the adjoint action of the complex matrix Cp,q :
1
O(n; C) = Cp,q
O(q, p; C) Cp,q .

(1.7.21)

Notice that we cannot say that O(n; R) and O(q, p; R) are isomorphic, since the adjoint
action of the matrix Cp,q would transform a real matrix into a complex one. To summarize
we can write the following isomorphisms:
O(q, p; F) O(p, q; F) ,
O(n; C) O(p, q; C) ,

(1.7.22)

Just as we did for the orthogonal group, we can define the group of special pseudo-orthogonal
transformations SO(q, p; F) as the subgroup of O(q, p; F) consisting of all the matrices with
unit determinant.
In what follows we shall use the short-hand notation for the groups over the real numbers:
O(p, q) O(p, q; R), SO(p, q) SO(p, q; R).
An example of pseudo-orthogonal group over the real numbers is the Lorentz group O(1, 3)
describing the most general transformations between two inertial reference frames.
Exercise 1.9: Prove that the matrix


cosh() sinh()
T[] =
,
(1.7.23)
sinh() cosh()
is in SO(1, 1). In fact T[] represents a generic element of this group.
Consider the defining condition (1.7.13) for the real group O(p, q; R) and take the equations corresponding to j = i:
p
X
k=1

(Ski )

n
X

(Ski ) =

k=p+1

+1 i = 1, . . . , p
.
1 i = p + 1, . . . , n

(1.7.24)

Being the left hand side a sum of terms with indefinite sign, the above condition no longer
implies that each term be bounded from above. In fact the entries of a O(p, q) matrix, for
p or q different from 0, are in general, in absolute value, not bounded from above, see the
example in eq. (1.7.23). The group O(p, q), for pq 6= 0, is then non-compact, as it contains
unbounded matrices.
The counting of free parameters for the pseudo-orthogonal group is the same as for the
orthogonal one.

1.7. MATRIX GROUPS

1.7.5

33

The Unitary Group U(n)

The unitary group U(n) is the subgroup of GL(n, C) consisting of n n unitary matrices.
A unitary matrix U is defined by the property:
U U = 1n ,

(1.7.25)

or, in components,
n
X

Uki
Ukj = ij ,

(1.7.26)

k=1

where the hermitian -conjugate U of U is defined as the complex conjugate of the transposed
of U: U (UT ) . The reader can easily verify that U1 = U and that therefore U U =
1n . Moreover, computing the determinant of both sides of (1.7.25) and using the property
that det(U ) = det(U) , we find
|det(U)|2 = det(U) det(U) = 1 ,

(1.7.27)

that is the determinant of a unitary matrix is just a phase: det(U) = ei . Finally, just as
we did for the orthogonal and pseudo-orthogonal groups, it is straightforward to check that
U(n) is indeed a subgroup of GL(n, C).
The group U(1) consists of complex numbers c (i.e. 1 1 matrices) with unit modulus,
i.e. c c = 1. In other words U(1) consists of phases c = ei .
Define now the group SU(n) of special unitary matrices, namely of unitary matrices with
unit determinant. The reader can show that it is a normal subgroup of U(n). Let now U be
a unitary matrix and let the phase ei be its determinant. We can write U as follows:
i

U = en S ,

(1.7.28)

where S SU(n) and e n U(1).


Remark: Equation (1.7.28) shows that we can always write a unitary matrix as the
i
product of a U(1) element times a SU(n) element. Similarly, given a phase e n in U(1) and
a SU(n) matrix S we can define a unique U(n) matrix by equation (1.7.28). In other words
we can define a correspondence:
i

(e n , S) U(1) SU(n)

e n S U(n) .

(1.7.29)

This correspondence is a homomorphism, however it is not an isomorphism since it is n-toi


2 i k
one, as the reader can verify by showing that all the elements of the form (e n (+2 k) , e n S),
k = 1, . . . , n, are mapped into the same element of U(n). One can say that U(1) SU(n) is
contained n-times inside U(n).
Unitary groups are important in quantum mechanics, since they describe transformations
on quantum states and operators representing physical observables.

34

CHAPTER 1. ABSTRACT GROUPS


Exercise 1.10: Prove that the matrix
1
S =
2

1n
1n

i 1n
i 1n


,

(1.7.30)

is in U(2n) and determine its U(1) and SU(2n) components. The above matrix is also called
Cayley matrix.
Exercise 1.11: Prove that the matrix:


cos() ei(+) sin() ei(+)
U[, , , ] =
,
(1.7.31)
sin() ei() cos() ei()
is in U(2) and determine its U(1) and SU(2) components. In fact one can show that the
most general element of U(2) can be written in the form (1.7.31) and thus that it depends on
four real parameters. Equation (1.7.31) then defines a parametrization of the generic U(2)
element, the parameters being , , , .
Consider now the defining condition (1.7.26) for i = j:
n
X

|Uki |2 = 1 .

(1.7.32)

k=1

The left hand side is the sum of positive terms which should be equal to one, implying that
each term is bounded from above by 1: |Uik | 1. We conclude that the modulus of each
entry of a unitary matrix is bounded from above and thus the corresponding group is said
to be compact.
Let us count the dimension of U(n). We start from the 2 n2 real parameters of a generic
n n complex matrix and subtract the number of independent real conditions implied by
(1.7.26). For each couple (i, j) with i 6= j we have a complex condition, i.e. 2 real conditions.
Since the couples (i, j) and (j, i) yield the same real conditions (one expresses the vanishing
of a complex number, the other the vanishing of its complex conjugate) we have in total
2n(n 1)/2 = n(n 1) real conditions coming from the i 6= j entries of eq. (1.7.26). The
diagonal entries i = j are real and yield n real conditions. In total we have n(n 1) + n = n2
real conditions. The number of free real parameters a U(n) depends on is:
2 n2 n2 = n2 .

(1.7.33)

The determinant of a unitary matrix U is a phase ei , depending on a continuous real


parameter . Setting it to 1 suppresses then one real parameter, leaving, for a generic
element of SU(n), n2 1 real independent parameters.

1.7.6

The Pseudo-Unitary Group U(p, q)

The notion of unitary group is extended to that of pseudo-unitary group U(p, q), p + q = n,
consisting of complex n n matrices satisfying the pseudo-unitary requirement:
U p,q U = p,q .

(1.7.34)

1.7. MATRIX GROUPS

35

If p = n or q = n the group reduces to U(n). The reader can easily verify that:
U1 = p,q U p,q , U p,q U = p,q , det(U) = ei .

(1.7.35)

Moreover, using the matrix I, we can prove the isomorphism U(p, q) U(q, p), just as we
did for the pseudo-orthogonal group.
Finally we define the subgroup
SU(p, q) = {S U(p, q)|det(S) = 1} .

(1.7.36)

Just as we did for the unitary case, one can write the isomorphism: U(p, q) U(1)SU(p, q).
The counting of free parameters for U(p, q) and SU(p, q) goes in the same way as for the
unitary group, yielding n2 real parameters for the former and n2 1.
Remark: Real (pseudo-) unitary matrices are (pseudo-) orthogonal. Indeed if M is a
real matrix M = MT and conditions (1.7.34) and (1.7.25) become (1.7.12) and (1.7.2)
respectively. Therefore O(p, q) is a subgroup of U(p, q), that is real (pseudo-) orthogonal
matrices are in particular (pseudo-) unitary.

1.7.7

The Symplectic Group Sp(2n, F)

The symplectic group Sp(2n, F) consists of 2n 2n matrices with entries in F satisfying the
condition:
ST S = ,

(1.7.37)

where

=

0n
1n

1n
0n


,

(1.7.38)

0n being the n n zero-matrix. In components equation (1.7.37) reads:


n
X

Ski k` S`j = ij .

(1.7.39)

k,`=1

A matrix satisfying condition (1.7.37) is said symplectic. The matrix has the following
properties: T = , 2 = 12n .
From (1.7.37) we immediately find: S1 = ST and S ST = , namely ST is
symplectic as well. We leave to the reader to show that condition (1.7.37) defines a subgroup
of GL(2n, F).
If we write a symplectic matrix S in terms of four n n blocks:


A B
S =
,
(1.7.40)
C D

36

CHAPTER 1. ABSTRACT GROUPS

equation (1.7.37) implies the following conditions:


AT C = CT A , AT D CT B = 1n , BT D = DT B .

(1.7.41)

Symplectic matrices are relevant in the Hamiltonian description of a system with n degrees of
freedom described by the generalized coordinates (q i ) = (q 1 , . . . , q n ). In fact they represent
the effect of canonical transformations on the vector (dq i , dpi ), i = 1, . . . , n (such symplectic
matrix representing a canonical transformation in general is not constant, but its entries will
depend on the phase space coordinates).
Since (ij ) is an antisymmetric matrix, the diagonal entries i = j of equation (1.7.39) are
identically zero and thus imply no condition. The only independent conditions come from
the (i, j), i 6= j, entries. The equations from the (i, j) and the (j, i) entries are the same,
leaving n(n 2)/2 independent conditions on numbers in F. The number of free parameters
is therefore:
(2n)2

2n(2n + 1)
2n(2n 1)
=
= n (2n + 1)
2
2

parameters in F ,

(1.7.42)

which means that Sp(2n, R) and Sp(2n, C) have n(2n+1) and 2n(2n+1) free real parameters
respectively.
Exercise 1.12: Show that the matrix
n

n
}|
{
z }| { z
x
S = diag(e , . . . , ex , ex , . . . , ex ) ,

(1.7.43)

with x real, is in Sp(2n, R). Notice that its diagonal entries, as x varies in R, are not bounded
from above. Since the group Sp(2n, R) has an unbounded element, it is non-compact.

1.7.8

The Unitary-Symplectic Group USp(2p, 2q)

The unitary-symplectic group USp(2p, 2q), with p + q = n is defined as the set of symplectic
2n 2n complex matrices U = (Uij ) which satisfy the pseudo-unitarity conditions:
UT U = ,
U 2p,2q U = 2p,2q .
where now the signs in the diagonal of 2p,2q are

1p 0p,q
0q,p 1q
2p,2q =
0p 0p,q
0q,p 0q

arranged as follows:

0p 0p,q
0q,p 0q
.
1p 0p,q
0q,p 1q

The group USp(2p, 2q) can then also be written as the following intersection:
\
USp(2p, 2q) = U(2p, 2q) Sp(2n; C) .

(1.7.44)

(1.7.45)

(1.7.46)

If either p or q are zero, the group USp(2n), being contained inside U(n) is compact.

1.7. MATRIX GROUPS

37

Summary Let us summarize below the dimensions and other features of the matrix groups
discussed above (p and q in the table are both considered non-vanishing):
Group
GL(n, C)
GL(n, R)
SL(n, C)
SL(n, R)
(S)O(n, C)
(S)O(n)
(S)O(p, q; C)
(S)O(p, q)
U(n)
U(p, q)
SU(n)
SU(p, q)
Sp(2n, C)
Sp(2n, R)
USp(2n)
USp(2p, 2q)

Real dimension
2n2
n2
2n2 2
n2 1
n(n 1)
1
n(n 1)
2
n(n 1)
1
n(n 1)
2
n2
n2
n2 1
n2 1
2n(2n + 1)
n(2n + 1)
n(2n + 1)
n(2n + 1)

Compact (c.)/non-compact (n.c.)


n.c.
n.c.
n.c.
n.c.
n.c.
c.
n.c.
n.c.
c.
n.c.
c
n.c.
n.c.
n.c.
c.
n.c.

38

CHAPTER 1. ABSTRACT GROUPS

Chapter 2
Transformations and Representations
2.1

Linear Vector Space

We define a linear vector space V over the real numbers R (or complex numbers C) a
collection of objects, to be called vectors and denoted by boldface symbols, on which two
operations are defined: a summation + between vectors and a product of a number times
a vector, such that:
1. (V, +) is an abelian group:
For any V1 and V2 , their sum is still a vector: V1 + V2 V ;

(closure)

There exist the null vector 0 such that, for any vector V: 0 + V = V + 0 = V;
(null vector)
For any vector V, there exist the opposite vector V such that, V + (V) =
V+V = 0; The sum W+(V) will be simply denoted by WV;
(opposite
vector)
For any three vectors V1 , V2 , V3 , we have: V1 + (V2 + V3 ) = (V1 + V2 ) + V3 ;
(associativity)
For any two vectors V1 , V2 , we have: V1 + V2 = V2 + V1 ;

(commutativity)

2. The following additional axioms are satisfied:


i For any a in R (or C) and V V : a V V ;
ii For any a, b in R (or C) and V V : a (b V) = (a b) V;
iii For any V V : 1 V = V 1 = V;
iv For any a in R (or C) and V1 , V2 V : a (V1 + V2 ) = (a V1 ) + (a V2 );
v For any a, b in R (or C) and V V : (a + b) V = (a V) + (b V)
Real (or complex) numbers are called scalars. For the sake of simplicity hereafter the symbol
in writing the product between a scalar and a vector will be omitted: a V a V.
39

40

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

A set of n vectors, e1 , . . . , en , are said to be linearly independent iff the only vanishing
combination of them yielding the null vector is the one with vanishing coefficients:
a1 e 1 + . . . + an e n = 0

a1 = a2 = . . . = an = 0 .

(2.1.1)

A vector space V is said to be ndimensional, and is denoted by Vn , if the maximal number


of linearly independent vectors is n. This means that, if we denote by (ei ) = (e1 , . . . , en ) a
set of n linearly independent vectors, called a basis of Vn , the vectors consisting of these n
plus any other vector en+1 are not linearly independent. As a consequence of this any vector
V can be expressed as a unique combination of the n vectors ei :
n
X
V =
V i ei = V 1 e1 + . . . + V n en .
(2.1.2)
i=1
i

The unique set of coefficients (V ) = (V 1 , . . . , V n ) associated with V are called the components of V in the basis (ei ). They are conventionally labeled by an upper index (notice
that V 2 is the second component of V and should not be confused with the square of some
not well defined quantity V !) for reasons to be clarified in the sequel. The components of V
are real or complex numbers depending on whether Vn is defined over R or C. For the time
being let us introduce Einsteins summation convention, which will be extensively used in
the present course: When in a formula a same index appears in an upper and lower position,
summation over that index is understood. For instance in the expression V i ei the index
i appears in an upper (as label of the vector components) and lower (as label of
basis
Pthe
n
i
vectors) position. Summation is then understood and we should read: V ei i=1 V i ei .
Notice that if we make a linear combination of two vectors V = V i ei and W = W i ei with
real (or complex) coefficients, the components of the resulting vector will be the same linear
combinations of the components V i and W i of two vectors:
a V + b W = a (V i ei ) + b (W i ei ) = (a V i ei ) + (b W i ei ) = (a V i + b W i ) ei , (2.1.3)
where the reader should realize that we have first used properties iv and ii and then v of the
scalar-vector product. It will be of considerable help to describe vectors as column vectors,
namely to identify each of them with the column vector whose entries are the components
relative to a given basis:
1
V

V2

V = V i ei
(2.1.4)
... .
Vn
We shall often denote the column vector representing V, in a given basis, by the same symbol
V: V (V i ). The advantage of such a description is that all vector operations can now
be translated into matrix operations. Consider for instance the problem of computing the
components of a linear combinations (2.1.3) of two vectors:
1
1

aV 1 + bW1
V
W
2
2
V2
2
+ b W. = a V +. b W = (a V i + b W i ) . (2.1.5)
aV + bW a
.
..
..
..

n
n
n
n
V
W
aV + bW

2.1. LINEAR VECTOR SPACE

41

Identifying a vector V with the column consisting of its components can be done only if the
basis is fixed once for all. In this notation the basis elements are represented by the following
column vectors:



1
0
0
0
1
0


e1 =
(2.1.6)
... , e2 = ... . . . en = ... .
0
0
1
If we are considering two different bases (ei ) and (e0i ), a same vector V = V i ei = V i 0 e0i ,
will be represented with respect to them by two different column vectors: V = (V i ) and
V0 = (V i 0 ). In this case we will keep in mind that the column vectors V and V0 represent
the same abstract vector V.

Figure 2.1: Vector V in V3 and point P in E3

Example 2.1: A familiar example of linear vector space over the real numbers is the
collection of vectors in our three dimensional Euclidean space E3 . These vectors are defined

as arrows connecting two points A and B in E3 ad denoted by AB. We know how to multiply
any such vector by a real number or how to sum two of them. The result is still a vector
in E3 , namely there exist a couple of points of E3 connected by it. In particular the null
vector is the vector connecting a point to itself. They close the linear vector space V3 . Given
a basis of three vectors e1 , e2 , e3 we can decompose, in a unique way, any vector V along
them, see Figure 2.1:
1
V
V = V 1 e1 + V 2 e2 + V 3 e3 V 2 .
(2.1.7)
3
V

42

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Vectors in V3 are then associated with couples of points of E3 . This means that, if we
arbitrarily fix a point O in E3 , called the origin, we can uniquely associate with any other

point P the vector r OP V3 , called the position vector of P with respect to O. The
origin O in E3 and a basis (ei ) of V3 define a reference frame (RF). The components of r
relative to (ei ) are the coordinates of P in the chosen RF, see Figure 2.1:

x

r = x e1 + y e2 + z e3 y .
(2.1.8)
z
The coordinates (x, y, z) will also be denoted by (xi ) = (x1 , x2 , x3 ). One of the defining
property of an Euclidean space is also the notion of distance between two points, namely
that of metric.

2.1.1

Homomorphisms Between Linear vector Spaces

Consider two vector spaces Vn and Wm over F (which can be either R or C) of dimension n
and m respectively. A mapping S:
S : V Vn W = S(V) Wn ,

(2.1.9)

is said to be a homomorphism between the two linear vector spaces if, for any V1 , V2 Vn
and , F:
S( V1 + V2 ) = S(V1 ) + S(V2 ) ,

(2.1.10)

namely if S preserves all the operations defined on Vn . A homomorphism is also called a


linear mapping on Vn with values in Wm . S is onto if S(Vn ) = Wm and is one-to-one if,
S(V1 ) = S(V2 ) implies V1 = V2 . The latter condition is equivalent to saying that the
zero vector 0 of Vn is the only vector which is mapped into the zero-vector of Wm . S is an
isomorphism if it is one-to-one and onto.
If the two spaces Vn and Wm have the same dimension, n = m, they are isomorphic.
Consider indeed a basis (ei ) for Vn and (fi ) for Wn . Then define a mapping S between the
two spaces such that for any i = 1, . . . , n, S(ei ) fi . The reader can verify that S is an
isomorphism. Vice-versa, if Vn and Wm are isomorphic, they have the same dimension. Let
us prove this property. Take indeed a basis (ei ) of Vn and consider the set of n vectors (fi )
in Wm where fi S(ei ). These vectors are linearly independent since, if i fi = 0, using eq.
(2.1.10) we have that S(i ei ) = 0 and thus, being S one-to-one, i ei = 0, which implies
i = 0 for any i = . . . , n. Suppose now (fi ) is not a maximal system of linearly independent
vectors in Wm . This means that there exist a vector f Wm such that f + i fi = 0
implies = i = 0. Since S is onto, there exist e Vn , such that S(e) = f . We have
then that S( e + i ei ) = 0, or , equivalently, e + i ei = 0 implies = i = 0, which
contradicts the assumption that (ei ) is a basis of Vn . A homomorphism of Vn onto Vn is
called an operator or endomorphism on Vn . The space of all homomorphisms between Vn
and Wm is denoted by Hom(Vn , Wm ). The space of all endomorphisms on Vn is denoted by

2.1. LINEAR VECTOR SPACE

43

End(Vn ) Hom(Vn , Vn ). If a homomorphism is onto and one-to-one it is called isomorphism


and its space denoted by Isom(Vn , Wm ), where m = n. Isomorphisms of a linear vector space
on itself are called automorphisms and span a space denoted by Aut(Vn ) Isom(Vn , Vn ).
Transformations on a linear vector space Vn will be described by automorphisms on Vn .
The action of a homomorphism S between Vn and Wm is represented by a m n matrix
MS as follows. Consider a basis (ei ) for Vn , i = 1, . . . , n. Consider the set of vectors
e0i S(ei ) of Wm . Since S is not necessarily one-to-one, (e0i ) are not necessarily linearly
independent. We can nevertheless expand each of them in components with respect to the
basis (fa ) of Wm , a = 1, . . . , m. Let us denote by MS a j the component of e0i along fa :
e0i S(ei ) = MS a j fa .

(2.1.11)

Let us compute S on a vector V = V i ei Vn :


S(V) = S(V i ei ) = V i e0i = (MS a i V i ) fa = W a fa .

(2.1.12)

The components W a of the vector W S(V) Wm are then expressed as follows:


W a = MS a i V i .
Representing these vectors in terms of column vectors:
1
1
W
V
.
.
W .. , V .. ,
Wm
Vn

(2.1.13)

(2.1.14)

relation (2.1.13) can be written in the following matrix notation:


W S(V) = MS V ,

(2.1.15)

where MS denotes the m n matrix MS (MS a j ) which, acting on the n-vector V yields
the m-vector W. For the sake of simplicity, the matrix MS , representing the action of
the homomorphism S, will also be denoted by the corresponding boldface letter S. If S
Aut(Vn ), MS (MS i j ) will be a n n invertible matrix, namely an element of GL(n, F)1 .
In fact there is a one to one correspondence between S Aut(Vn ) and GL(n, F).
Just as for the groups we can define the product S T of two endomorphisms S, T on a
linear vector space Vn as the endomorphism defined on a generic vector V as follows:
S T (V) S(T (V)) .

(2.1.16)

Let (ei ) be a basis of Vn , so that S(ei ) = MS j i ej and T (ei ) = MT j i ej . Then


S T (ei ) = S(T (ei )) = S(MT j i ej ) = MT j i MS k j ek = (MS MT )k i ek ,
1

(2.1.17)

Recall that we have denoted by GL(n, R) or GL(n, C) the group of all n n real or complex matrices
equipped with the ordinary row-times-column product.

44

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

so that the matrix MST , representing the action of S T , is the product MS MT of the
corresponding matrices in the same order: On a column vector V = (V i ):
S T (V) = MS MT V .

(2.1.18)

If S, T are two automorphisms, S T is still an automorphism. It is straightforward to prove


that Aut(Vn ), with this product, is a group. Moreover the correspondence:
S Aut(Vn ) MS = (MS i j ) GL(n, F) ,

(2.1.19)

is a homomorphism since, as shown above, MST = MS MT . It is onto: Any matrix M


defines an endomorphism S on Vn , such that MS = M. Finally it is one-to-one, since a
n n matrix M uniquely defines an endomorphism on Vn : Two endomorphisms represented
by the same matrix M will define the same correspondence between vectors, according to
(2.1.19), and thus coincide. The correspondence (2.1.19) is an isomorphism between Aut(Vn )
and GL(n, F):
Aut(Vn ) GL(n, F) ,

(2.1.20)

so that we can characterize GL(n, F) as the group of automorphisms, or linear transformations, on a n-dimensional linear vector space over F.

2.1.2

Inner Product on a Linear Vector Space over R

Let us return now to the general discussion of linear vector spaces and define on them some
new structures.
On a linear vector space Vn over R we define an inner or scalar product between vectors
as the mapping:
(, ) : Vn Vn R ,

(2.1.21)

which associates with any couple of vectors V and W a real number (V, W). This mapping
has to be symmetric and bilinear, i.e. linear in any of its arguments:
1. For any V and W in Vn : (V, W) = (W, V)

(symmetry);

2. For any V1 , V2 and W in Vn and b R: (a V1 + b V2 , W) = a (V1 , W) + b (V2 , W)


(linearity).
In virtue of its symmetry, linearity in the first argument implies linearity in the second
as well. If now we write two generic vectors in components with respect to a basis (ei ),
V = V i ei , W = W i ei , using the bilinear property of the inner product, we are able to
write:
(V, W) = (V i ei , W j ej ) = V i (ei , W j ej ) = V i (ei , ej ) W j = V i gij W j , (2.1.22)

2.1. LINEAR VECTOR SPACE

45

where the matrix g = (gij ) ((ei , ej )) is a symmetric n n real matrix called the metric.
In matrix notation the scalar product (2.1.22) can be written in terms of row-times-column
matrix products:
(V, W) = VT g W ,

(2.1.23)

where, in the right hand side, we have denoted by V and W the columns corresponding to
the two vectors, so that VT , namely the transposed of the column vector V, is actually a
row vector: VT (V 1 , V 2 , . . . , V n ). We define the norm squared of a vector V the inner
product of the vector with itself:
||V||2 (V, V) = V i gij W j = VT g V .

(2.1.24)

We have not required that g = (gij ) be a positive defined matrix so far, and thus ||V||2 is not
necessarily a positive number. We will just require g to be non-singular, namely: det(g) 6= 0.
As we shall show in the next sections, we can always find a basis (ei ) with respect to which
the matrix g has a diagonal form with diagonal entries 1:
q

g = p,q

z }| { z }| {
diag(+1, . . . , +1, 1, . . . , 1) ,

(2.1.25)

where p + q = n. The signature q p of g is a feature of the inner product. This basis is


called orthonormal.

2.1.3

Hermitian, Positive Definite Product on a Linear Vector


Space over C

Consider now a vector space Vn over C. A hermitian positive definite product is defined as
a mapping:
(, ) : Vn Vn C ,

(2.1.26)

which associates with any couple of vectors V and W a complex number (V, W) and which
satisfied the following properties:
1. For any V and W in Vn : (V, W) = (W, V)

(hermitian);

2. For any V1 , V2 and W in Vn and b R: (W, a V1 + b V2 ) = a (W, V1 ) + b (W, V2 )


(linearity);
3. For any V in Vn : kVk2 (V, V) 0, the inequality is saturated only if V = 0
(positive definiteness);
From the first two properties it follows that (a V1 + b V2 , W) = a (V1 , W) + b (V2 , W). If
now we write two generic vectors in components with respect to a basis (ei ), V = V i ei , W =
W i ei , using the above properties of the hermitian product, we are able to write:
(V, W) = (V i ei , W j ej ) = V i (ei , W j ej ) = V i (ei , ej ) W j = V i gij W j (, 2.1.27)

46

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

where the matrix g = (gij ) ((ei , ej )) is now hermitian, namely it satisfies the property:
gij = (gji ) . In matrix notation, if we denote by M the hermitian conjugate of a matrix M,
obtained by taking the complex conjugate of the transposed of M, we can write: g = g .
From the general theory of matrices we know that g, being hermitian, has real eigenvalues.
We can also rewrite (2.1.22) in the following form:
(V, W) = V g W ,

(2.1.28)

where V (V 1 , . . . , V n ). Hermitianity of the product in particular implies that, for any


V Vn , (V, V) = (V, V) R. The fact that the product is positive definite further
requires that gij be a positive definite matrix, namely have real positive eigenvalues.
A complex linear vector space equipped with a hermitian, positive definite scalar product
is also called a Hilbert space.

2.1.4

Symplectic product

Consider a 2 n-dimensional linear vector space V2 n on R. A symplectic product on V2n is


defined as a mapping:
(, ) : V2n V2n R ,

(2.1.29)

which associates with any couple of vectors V and W a real number (V, W) and which
satisfied the following properties:
1. For any V and W in Vn : (V, W) = (W, V)

(skew-symmetry);

2. For any V1 , V2 and W in Vn and b R: (W, a V1 + b V2 ) = a (W, V1 ) + b (W, V2 )


(linearity).
The above properties imply bi-linearity of the product. If now we write two generic vectors in
components with respect to a basis (ei ), V = V i ei , W = W i ei , using the bilinear property
of the inner product, we are able to write:
(V, W) = (V i ei , W j ej ) = V i (ei , W j ej ) = V i (ei , ej ) W j = V i ij W j , (2.1.30)
where = (ij ) ((ei , ej )) is a skew-symmetric 2n 2n real matrix called symplectic
matrix: ij = ji .
We can always choose a basis (ei ) such that, if we split its elements into the first n, ea ,
labeled by a and the last n, en+a , we have:
(ea , eb ) = (ea+n , eb+n ) = 0 , (ea , eb+n ) = (eb+n , ea ) = ab ,

(2.1.31)

where ab is equal to 1 if a = b, to 0 if a 6= b. In this basis the symplectic matrix reads:




0n 1n
=
,
(2.1.32)
1n 0n
where 0n is an n n matrix made of zeros, while 1n is the n n identity matrix. In this
basis, the symplectic product between two vectors (2.1.30) reads:
(V, W) = VT W = V a W n+a V n+a W a .

(2.1.33)

2.1. LINEAR VECTOR SPACE

2.1.5

47

Example: Space of Hermitian matrices

Let us show that the space of hermitian n n matrices, is a vector space over the real
numbers. A hermitian matrix H (Hij ) is a complex n n matrix satisfying the property
H = H

Hij = (Hji ) .

(2.1.34)

We have already come across an example of hermitian matrix: The hermitian metric g =
(gij ). The reader can easily verify that if H1 and H2 are two hermitian matrices, also any
linear combination of them, a H1 + b H2 , is. Therefore the space of hermitian matrices is a
vector space with respect to the ordinary sum of matrices and product times real numbers.
To determine its dimension we should find the maximal number of free real parameters a
generic hermitian matrix depends on. These will be identified with the components of a
generic hermitian matrix relative to some basis. We start from the 2 n2 real parameters of
a generic complex n n matrix. The condition (2.1.34), for i = j implies that the diagonal
entries of H are real numbers (Hii = (Hii ) ). The same condition for i 6= j implies that the
entries below the diagonal are the conjugate of the symmetric entries above the diagonal.
The independent entries of a generic H consist therefore of the n real diagonal components
and the real and imaginary parts of n(n 1)/2 complex entries above the diagonal. Their
number is n + 2 n(n 1)/2 = n2 . We conclude that the space of hermitian matrices has
real dimension n2 .
As an example consider the hermitian 2 2 matrices. The most general such matrices
has the following form:


a
b ic
H =
,
(2.1.35)
b + ic
d
a, b, c, d being real numbers. The space has dimension 4 and we can choose as a basis the
following set of hermitian matrices:
e1 = 12 , e2 = 1 , e3 = 2 , e4 = 3 ,
where i are the Pauli matrices:






0 1
0 i
1 0
1 =
, 2 =
, 3 =
.
1 0
i 0
0 1

(2.1.36)

(2.1.37)

The generic matrix (2.1.35) is expressed in the basis (2.1.36) as follows:


H =

a+d
ad
e1 + b e2 + c e3 +
e4 = H i ei .
2
2

(2.1.38)

We can also define a positive definite inner product over the space of 22 hermitian matrices:
(H1 , H2 )

1
Tr(H1 H2 ) .
2

(2.1.39)

48

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

The reader can verify that with respect to the above inner product the basis (2.1.36) is
ortho-normal: (ei , ej ) = ij . The components of H can be computed as H i = H j ij =
H j (ej , ei ) = (H, ei ).
Exercise 2.1: Show that traceless hermitian matrices (H = H, Tr(H) = 0) form a
3-dimensional vector space, of which the Pauli matrices (i ) form a basis.
Exercise 2.2: Consider now anti-hermitian matrices. These are complex n n matrices
A = (Aij ) satisfying the condition: A = A, or, in components, Aij = (Aji ) . Show that
anti-hermitian matrices form a linear vector space over the real numbers, just as hermitian
matrices do. Show that the dimension of the algebra is still n2 and that, for n = 2 we can
choose as a basis the following matrices: fi = i ei , i = 1, 2, 3, 4, ei being defined in (2.1.36).
Let us end this subsection by mentioning some properties of the Pauli matrices, which
the reader can easily verify:
i j = ij 12 + i

3
X

ijk k ,

(2.1.40)

k=1

i j k = i ijk 12 + ij k ik j + jk i ,

(2.1.41)

where ijk is different from zero only if i 6= j 6= k 6= i, 123 = 1 and is totally antisymmetric,
so that ijk = +1 or 1 depending on whether (i, j, k) is an even or odd permutation of
(1, 2, 3) respectively. The following property of ijk , used to derive eq. (2.1.41) from (2.1.40)
holds:
3
X

ijk `nk = i` jn in j` .

(2.1.42)

k=1

Exercise 2.3: Compute 1 2 3 .

2.2

Spaces

Let us consider a space of points Mn (like the three dimensional Euclidean space) with which
there is a linear vector space Vn associated in the same way as a three-dimensional vector
space V3 is associated with the Euclidean space E3 : To each couple of points A, B in Mn

there corresponds a unique vector V = AB of Vn connecting them. Mn is then said to be


a ndimensional real or complex space of points depending on whether Vn is a linear vector
space over R or C. The three-dimensional Euclidean space is a real space of points.
If we fix an origin O in Mn we can associate with each point P Mn a unique position

vector r = OP and a unique set of n coordinates (xi ) representing the components of r


relative to a basis (ei ) of Vn :
1
x

x2

r = xi e i
(2.2.1)
... .
xn

2.2. SPACES

49

The coordinates xi are real or complex numbers depending on whether Mn is a real or


complex space respectively. Given two points P and P 0 of Mn , with position vectors r and
r0 respectively, the relative position vector r of P 0 with respect to P is defined as follows:

x1
x2

(2.2.2)
r r0 r = (xi ) ei =
... ,
xn
where xi xi 0 xi . If the two points are infinitely close to one another, the relative position
vector will be denoted by dr and represents an infinitesimal shift and its components are
the differentials of the coordinates:
1
dx

dx2

(2.2.3)
r (dxi ) ei =
... .
dxn

2.2.1

Real Metric Spaces

If Vn is endowed with an inner product, we can use it to define a metric on Mn : The squared
distance d(P, P 0 )2 between two points P and P 0 is the squared norm of the relative position
vector:
d(P, P 0 )2 ||r||2 = (xi ) gij (xj ) = rT g r .

(2.2.4)

If the two points are infinitely close, their squared distance is also denoted by ds2 and reads:
ds2 d(P, P 0 )2 ||dr||2 = dxi gij dxj = drT g dr .

(2.2.5)

With respect to an orthonormal basis (2.1.25) this infinitesimal distance reads:


ds2 = (dx1 )2 + . . . + (dxp )2 (dxp+1 )2 . . . (dxn )2 .

(2.2.6)

The space Mn , equipped with this notion of distance, is a metric space, characterized by the
signature p q of g and will therefore be denoted by M (p,q) . If n = 3, q = 0 and p = n the
metric is positive definite (g is a positive definite matrix having all positive eigenvalues) and
the space M (3,0) is the familiar Euclidean space E3 . In this case, if (x, y, z) and (x0 , y 0 , z 0 )
are the coordinates of P and P 0 respectively, equation (2.2.5) reads:
d(P, P 0 )2 (x0 x)2 + (y 0 y)2 + (z 0 z)2 0 ,

(2.2.7)

which is the usual expression derived from Pythagoras theorem. Fro the above expression
it is clear that, being g positive definite, d(P, P 0 ) = 0 if and only if P = P 0 . In general
||V||2 0 and is zero if and only of V = 0. Since the norm squared of a vector of V3 ,
in this case, coincides with its length squared, i.e. the distance squared between its end

50

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

points, an orthonormal basis is a basis of mutually orthogonal vectors of unit length. We


recover in this case the familiar geometric definition of scalar product between two vectors:
(V, W) = ||V|| ||W|| cos(), being the angle between the two vectors.
If n = 4, q = 3 and p = 1 the corresponding space M (1,3) is called Minkowskis space.
In the special theory of realativity a point in M (1,3) represents a physical event which took
place at a point of coordinates x, y, z of our Euclidean space, at a time t. Its coordinates
are labeled by an index = i 1 = 0, . . . , 3, so that x0 is related to time, x0 = c t, c being
the speed of light, while the remaining three coordinates x1 , x2 , x3 are identified with the
spatial coordinates x, y, z of the event. The distance between two nearby events reads:
ds2 dx g dx = (dx0 )2 (dx1 )2 (dx2 )2 (dx3 )2 = c2 dt2 dx2 dy 2 dz 2 .
(2.2.8)
The whole special theory of relativity is based on the requirement that this distance be the
same whatever inertial RF we use for studying the events and derives from the assumption
that the speed of light c be the same in any inertial RF. ds2 can now be positive, negative or
null, since the metric g = (g ) has indefinite signature. From this it follows that non-null
vectors can have vanishing norm squared and, for the same reason, two distinct points can
have vanishing distance. A vector V = V e (V ) of V4 with positive or negative norm
squared are called time-like or space-like respectively, while vectors with vanishing norm
are called light-like. Similarly we talk about time-like, space-like or light-like distances if
r = (x ) is time-like, space-like or light-like.

2.2.2

Complex Spaces With Hermitian Positive Definite Metric

Consider a complex space Mn whose associated linear vector space Vn is endowed with a
hermitian, positive definite, inner product. This product on Vn induces a metric on Mn , so
that the squared distance d(P, P 0 )2 between two points P and P 0 is defined as the squared
norm of the relative position vector:
d(P, P 0 )2 krk2 = (xi ) gij (xj ) = r g r .

(2.2.9)

With respect to an ortho-normal basis (ei ) of Vn , gij = (ei , ej ) = ij , the above distance
reads:
n
X
d(P, P 0 )2
|xi |2 = |x1 |2 + . . . + |xn |2 .
(2.2.10)
i=1

This distance, being the sum of squares, vanishes only if each term in the sum is zero, namely
if the two points coincide. If P, P 0 are infinitely close, so that their relative position vector
is described by the infinitesimal shift vector dr = (dxi ) their distance d(P, P 0 ), also denoted
by ds2 , reads:
ds

0 2

= d(P, P ) = kdrk = dr dr =

n
X
i=1

where we have used the ortho-normal basis g = 12 .

|dxi |2 = |dx1 |2 + . . . + |dxn |2 , (2.2.11)

2.3. TRANSFORMATIONS

2.2.3

51

Symplectic Spaces

Next we consider spaces of points M2n whose geometric properties are encoded in a linear
vector space V2n over R on which a symplectic product is defined. Choosing a RF on M2n
with respect to which the symplectic matrix has the form (2.1.32) and, given two infinitesimal
displacements dr = (dxi ), dr0 = (dxi 0 ), we can compute their symplectic product:
n
X
(dr, dr ) = dx ij dx = dr dr =
(dxa dx0 a+n dxa+n dx0 a ) .
0

j0

(2.2.12)

a=1

An example of symplectic space is the phase space of a Hamiltonian system with n degrees
of freedom. Each point is this space represents a mechanical configuration of the system
and is described by n generalized coordinates q a and n conjugate momenta pa (for a point
particle n = 3 and q a are the three spatial coordinates x, y, z, while pa the components of
the momentum vector px , py , pz ): xa = q a , xn+a = pa . The infinitesimal displacement vector
r is then computed between two nearby configurations and the symplectic product (2.2.12)
between two infinitesimal displacements reads:
(dr, dr0 ) = dxi ij dxj 0 = dq a dp0a dpa dq 0 a ,

(2.2.13)

where now Einsteins summation convention was used for the index a. The relevance of this
symplectic structure for Hamiltonian systems relies in the fact that canonical transformations, including time evolution, namely the correspondence between the configurations at
a time t and at a time t0 > t, are represented on dr by symplectic transformations which
preserve the product (2.2.13).

2.3

Transformations

In the present section we discuss transformations on a point-field. By point-field we mean an


arbitrary collection of elements (which can be either finite or infinite) which we call points.
An example of point-fields are the metric spaces M (p,q) introduced in the previous section.
These are sets of infinite points, think about the three dimensional Euclidean space. We
will also consider a vector space as a point-field, where the points are now vectors. A group
itself, finite or infinite, can be regarded as a point-field. Our notion of point-field so far is
rather generic.
Let M be a point-field. A transformation S on M is a correspondence of M into itself,
which maps each point P of M into its image P 0 = S(P ):
S : P M P 0 = S(P ) M .

(2.3.1)

A transformation is then defined by assigning its value on any point of M . Given two
transformations S, T , we define the product transformation T S as the transformation
resulting from the consecutive action on M of S and T . Therefore if S maps P into P 0 = S(P )
and T maps P 0 into P 00 = T (P 0 ), T S will map P into P 00 = T (S(P )). We define the

52

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

identity transformation I as the transformation mapping each point into itself: I(P ) = P .
If we restrict to one-to-one, onto correspondences, these can be inverted in M . Given a
transformation S we define its inverse S 1 as the mapping between a point P 0 and the only
point P of which P 0 is the image through S:
S : P P0

S 1 : P 0 P .

(2.3.2)

Clearly the inverse of the inverse of a transformation is the transformation itself: (S 1 )1 =


S. By definition, the consecutive action of a transformation and its inverse on any point
will yield the point itself, namely the product of any transformation by its inverse is the
identity transformation: S S 1 = S 1 S = I. If we have three transformations the reader
can convince himself that their product is associative: R (T S) = (R T ) S. We conclude
that the set of all transformations of a point-field M , endowed with the product defined
above, is a group called group of transformations of the point-field, or symmetry group of
the point-field, and denoted by TM .
Example 2.2 For example consider a set of n objects: x1 , . . . , xn . Any one to one transformation on this point-field is described by a permutation. Indeed consider a transformation
S. Its image on x1 will be one of the points of the same set, call it xi1 : S(x1 ) = xi1 . Similarly
S(x2 ) = xi2 , which cannot coincide with xi1 being the correspondence one-to-one, and so
on up to S(xn ) = xin , where xin is different from
 xi1 , . . . , xin1
 . The action of S will then
x1 . . . x n
totally be described by the permutation: S
.
xi 1 . . . x i n
Example 2.3 We can consider as an example of point-field the Euclidean plane E2 or
space E3 . These are real metric spaces, in which the metric is positive definite. Among
all the transformations in TE2 or TE3 , of particular interest are those transformations which
leave distances between points invariants. These are congruences and consist in translations,
rotations about a point or an axis (in E3 ) and reflections. Rotations and reflections are also
called proper congruences. Congruences will preserve the shape and size of any collection of
points (e.g. a triangle, a rectangle etc...) in E2 or E3 . We may also consider more general
transformations which preserve the shape but not the size of a geometrical object. These are
called similitudes and include, besides translations, rotations and reflections, also dilations.
Example 2.4 Let us consider M to be the familiar Euclidean plane E2 . We can consider
a subset N of points in E2 , like a geometrical object (e.g. an equilateral triangle, a square,
or any polygon), and look for its symmetry group TN . As we have shown in last chapter, for
an equilateral triangle, TN is the non abelian group D3 , containing rotations and reflections,
for a square D4 and so on for all the regular polygons...In any case TN is a subgroup
of TE2 . A geometrical object of finite size can only have as symmetry transformations
rotations and reflections. Indeed a finite size object is always contained inside a circle
of finite radius. If a translation t or a dilation d were a symmetry of the object, their
consecutive actions tn and dn , for any n would be symmetries as well. But for n sufficiently

2.3. TRANSFORMATIONS

53

Figure 2.2: Symmetry of a decoration under a translation by its period a.

large the resulting transformation, being it a translation or a dilation, would bring all the
points of the object outside the circle which originally contained it and thus tn and dn
cannot be symmetries, in contrast with our assumption. It can be shown that the groups
Cn and Dn exhaust all possible proper congruences on the Euclidean plane. Only infinitely
extended objects with periodic shape, like a lattice, can have a translation as symmetry.
An example are certain ornaments or regular tilings, see Figure 2.2. Similar arguments
can be given for geometrical objects in the three dimensional Euclidean space E3 as well.
Proper congruences (i.e. rotations and reflections), act on En as linear homogeneous, metric
preserving, transformations and, as we shall see, are contained in the Lie group O(n). On E2
the group O(2) can be considered as the limit n of Dn , also denoted by D , and is the
symmetry group of a circle, containing rotations about a point by any angle and reflections
in any straight line containing that point. The group SO(2), as we shall see, is abelian and
contains just rotations about a point by any angle. It can be identified with the group C .

2.3.1

Transformations on Mn

As point-field we can consider the spaces of points Mn discussed in Section 2.2. With respect
to a RF, each point is described by its n coordinates (xi ). The action of a transformation S
mapping a point P of coordinates xi into a point P 0 of coordinates xi 0 is then totally defined
by expressing the latter as functions of the coordinates of P : xi 0 = fSi (x1 , x2 , . . . , xn ). The
functions fSi (x1 , x2 , . . . , xn ), which we shall denote for the sake of simplicity by fSi (xj ), can
depend, in general, non-linearly on their arguments. If we consider the effect of S on the
relative position vector dr = (dxi ) between two nearby points of coordinates xi and xi + dxi ,
it can be deduced by expressing the relative position between the two image points (which

54

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

are still infinitely close to one another):


dxi 0 = fSi (xj + dxj ) fSi (xj ) = dfSi =

fSi j xi 0 j
dx =
dx = MS i j dxj ,
xj
xj

(2.3.3)

or, in matrix notation:


dr0 = MS dr ,

(2.3.4)

Where we have used the definition of the total differential of a function. The transformation
S acts by means of a coordinate-dependent matrix MS (MS i j ), which is the Jacobian of
the transformation, on the column vector dr of coordinate differentials (dxi ). The identity
transformation I maps xi into xi 0 = xi , namely fIi (xk ) = xi and the corresponding action
on the coordinate differentials is represented by the n n identity matrix: MIi j = ji , or, in
matrix notation, MI = 1n .
Consider now two transformations S and T whose action on the coordinates is defined
by the functions fSi and fTi respectively:
S : xi xi 0 = fSi (xj ) ,
T : xi xi 0 = fTi (xj ) ,

(2.3.5)

and T x into x = fTi (xj 0 )


which means that, if S maps x into x =
effect of T S is represented by the composites of the functions fTi and fSj :
T S : xi xi 0 = fTi (fSj (xk )) .
i

i0

fSi (xj )

i0

i 00

fTi (fSj (xk ))


(2.3.6)

The action of T S on coordinate differentials is readily computed using the property of


derivation of composite functions:
dx

i 00

fTi fSj k
fTi
j0
dx =
dx = MT i j MS j k dxk ,
=
xj 0
xj 0 xk

(2.3.7)

or, in matrix notations:


dr00 = MT MS dr = MT S dr .

(2.3.8)

If the action of S is described by a set of functions fSi such that xi 0 = fSi (xj ), its inverse
S 1 will be described by the inverse functions fSi 1 = fS1 i , such that: fSi 1 (xj 0 ) = xi =
fS1 i (fSj (xk )). The Jacobian of this transformation is the inverse of the matrix MS : MS 1 =
M1
S . Indeed 1n = MI = MS 1 S = MS 1 MS .
So far we have described the effect of a transformation on Mn as a coordinate transformation: xi xi 0 = f i (xj ). Coordinates are like the names we give to the points of Mn
in a given RF. We can view a same transformation either as a change of our description of
a same point, namely a change of the coordinate system we use (passive description of a
transformation), or as a true mapping between different points in a fixed coordinate system
(active description of a transformation). The two points of view are complementary and
lead to the same relations between the final and initial coordinates. In the former xi 0 and
xi are different descriptions of a same point in space, that is the transformation affects the
RF we use, while in the latter xi 0 and xi are the coordinates, relative to a same RF, of
two different points which are mapped one into the other by the transformation (this is the
view-point we originally adopted when introducing the notion of transformation).

2.3. TRANSFORMATIONS

55

Figure 2.3: Active an passive description of a rotation in the plane.

Example 2.5 As a simple example let us consider a clockwise rotation by an angle on


a plane, about a point O, see Figure 2.3. The plane is the two-dimensional Euclidean space
E2 , associated with the linear vector space V2 , containing all the vectors on the plane. Using
the active description of the transformation, a point P with coordinates x, y is mapped in a
point P 0 whose coordinates x0 , y 0 are related to x, y as follows:
 0
x = f 1 (x, y) = cos() x + sin() y
(x, y)
.
(2.3.9)
y 0 = f 2 (x, y) = sin() x + cos() y
Notice that the coordinates of the new point P 0 are linear and homogeneous functions of
the coordinates of P through coefficients depending on the angle . We can indeed rewrite
(2.3.9) in the following matrix form:


cos() sin()
0
r = Mr , M =
.
(2.3.10)
sin() cos()
It can be easily verified that M is also the Jacobian of the transformation:
!
1
1
M=

f
x
f 2
x

f
y
f 2
y

(2.3.11)

56

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

and therefore defines the effect of the rotation on infinitesimal displacements as well: dr0 =
M dr. Let us now consider the same transformation from the passive point of view and
observe that the same eqs. (2.3.9) can be interpreted as relating the coordinates of a same
point P relative to a basis ei (x, y) to those (x0 , y 0 ) relative to a new basis e0i , obtained
from the original one through an counter-clockwise rotation by the same angle about the

origin. To see this let us write the position vector of P in the basis (ei ): OP = x e1 + y e2 .
Notice that, since we are changing the basis for the vectors, we should not identify the
position vector with the column containing the coordinates, but rather write it explicitly in

terms of the basis elements. To avoid confusion we denote by OP the position vector, which
is associated with the point P and by r the column vector consisting of the components

of OP in the basis (ei ). The same position vector has the following form in the new basis

OP = x0 e01 +y 0 e02 . Consider next the relation between the two bases by writing each element
of the new basis in components with respect to the old one:
e01 = cos() e1 + sin() e2 ,
e02 = sin() e1 + cos() e2 .
(2.3.12)

Substituting the above formulas in OP we find:

OP = x0 (cos() e1 + sin() e2 ) + y 0 ( sin() e1 + cos() e2 ) = x e1 + y e2 .(2.3.13)


Equating the components of e1 and e2 the relation between the old and the new coordinates
follows:
x = cos() x0 sin() y 0
y = sin() x0 + cos() y 0 .

(2.3.14)

Finally, inverting the above relations we obtain the transformation law (2.3.9). In both
descriptions the transformation is realized as the action of a same matrix M on the column
vector of coordinates r.
We can consider now a function F on Mn with values in R:
F : P Mn F (P ) R .

(2.3.15)

This function can be described as a function of the n coordinates (xi ) of a generic point
P Mn : F (P ) F (x1 , x2 , . . . , xn ) = F (xi ). The gradient F of F is defined as the
row vector consisting of the partial derivatives of F (x1 , x2 , . . . , xn ) with respect to each
coordinate:

 

F F
F
F
F
,
,...,

,
(2.3.16)
x1 x2
xn
xi
such that the differential dF of F , which expresses the difference in value of F between two
infinitely close points is expressed as the row-times-column product of F and dr:


 dx1
F
F
F ..
dF = F (xi + dxi ) F (xi ) =
dxi =
,...,
= F dr(2.3.17)
.
.
i
1
x
x
xn
n
dx

2.3. TRANSFORMATIONS

57

Consider now a transformation S mapping xi into xi 0 = fSi (xj ). From the passive view-point,
the two sets of coordinates are different descriptions of a same point. Since the value of F
depends only on the point in which it is evaluated, the expression of F as a function of the
two set of coordinates will in general be different: Its expression F 0 (xi 0 ) in terms of the xi 0
is obtained from F (xi ) by expressing xi as functions of the xi 0 , through the inverse of the
functions fSi :
F 0 (xi 0 ) = F (P ) = F (xi (xj 0 )) .

(2.3.18)

The components of the gradient transform as follows:


F
F 0
xj F
F

=
=
M 1 j i ,
i
i
0
i
0
j
x
x
x x
xj S

(2.3.19)

0 F 0 = F M1
S .

(2.3.20)

or, in matrix notation:

We have come across two kind of objects so far: Objects represented by column vectors,
whose components are labeled by an upper index, like the vectors in Vn V = (V i ) or an
infinitesimal displacement dr = (dxi ) in Mn , and objects like the gradient of a function
F
F = ( x
i ), which are represented by row-vectors whose components are labeled by a lower
index. Under a transformation the former transform as in (2.3.3) and are called contravariant
vectors, while the latter transform as in (2.3.19) and are called covariant vectors.

2.3.2

Homogeneous-Linear Transformations on Mn and Linear Transformations on Vn .

Let us generalize our example to generic homogeneous-linear transformations on Mn .


Passive picture: These transformations, in the passive picture, originate from a generic
change of basis of Vn : (ei ) (e0i ), keeping the origin O of the RF fixed. The position vector

OP of a same point P in the two bases read:

OP = xi ei = xi 0 e0i .
(2.3.21)
In the first basis the position vector is described by the column vector r = (xi ) consisting
of the coordinates relative to the initial RF, while in se second basis it is represented by
the column vector r0 = (xi 0 ), consisting of the new coordinates of P . Let us express each
element of the new basis in terms of the old one:
e0i = Dj i ej ,

(2.3.22)

where Dj i represents the j th component of the ith vector e0i with respect to the old basis.
Equation (2.3.22) can be easily written in matrix notation by arranging the basis elements
in row vectors and writing:
(e01 , e02 , . . . , e0n ) = (e1 , e2 , . . . , en ) D ,

(2.3.23)

58

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

where D is the matrix (Di j ), i and j being the row and column indices respectively. Since
the matrix D is non-singular:
both (ei ) and (e0i ) are systems of linearly independent vectors, 

cos() sin()
i
. Let us
det(D) 6= 0. In the example 2, the matrix D = (D j ) =
sin() cos()
substitute eq. (2.3.22) into (2.3.21):

OP = xi ei = xi 0 Dj i ej .

(2.3.24)

Equating each component with respect to the old basis we find a relation between the old
and new coordinates:
xi = D i j xj 0

r = D r0 .

(2.3.25)

Since D is non-singular, the above system of n linear equations in n unknowns xi 0 can be


solved by multiplying both sides to the left by D1 . We then find:
r0 = D1 r

xi 0 = f i (xj ) = D1 i j xj .

(2.3.26)

We can readily compute the Jacobian matrix of the above transformation: M = (M i j ) =


i0
f i
1 i
1
( x
) = ( x
j ) = D . The matrix D represents therefore the Jacobian matrix
j ) = (D
xj
xi
of the inverse transformation: D = ( x
j 0 ). In terms of M let us rewrite the transformation
properties of the basis elements and the coordinate differentials:
e0i = Dj i ej = M 1 j i ej =
i0

dx

xj
ej ,
xi 0

xi 0 j
= M j dx =
dx ,
xj
i

(2.3.27)
(2.3.28)

or, in matrix notation:


(e01 , e02 , . . . , e0n ) = (e1 , e2 , . . . , en ) M1 ,
dr0 = M dr ,

(2.3.29)
(2.3.30)

Comparing (2.3.29) with (2.3.19) we see that, under a linear-homogeneous transformation,


the basis elements transform as components of a covariant vector, and are indeed labeled by
a lower index.
Active picture: If we adopt on the other hand the active point of view, points and vectors
are transformed into different points and vectors respectively. In this representation, we will
derive the effect of a homogeneous linear transformation S on Mn from the action of a linear
transformation on the corresponding linear vector space Vn . A linear transformation over a
vector space Vn is an invertible linear function of Vn onto Vn , that is an automorphism on Vn ,
as defined in Section 2.1.1. As discussed in Section 2.1.1, with respect to a given basis (ei )
of Vn , the action of an automorphism S on Vn is completely described by a n n invertible
matrix MS = (MS i j ), whose generic entry MS i j represents the ith component of the vector

2.3. TRANSFORMATIONS

59

S(ej ), i, j = 1, . . . , n. The action of S on the column vector V (V i ), given in eq. (2.1.15),


is expressed as the following matrix product:
V V0 = S(V) = MS V ,

(2.3.31)

V i V 0 i = MS i j V j .

(2.3.32)

or, in components,

The transformation S on Vn induces on the space of points Mn a transformation, to be


denoted by the same letter S, which maps a point P described, with respect to a given

RF, by the vector OP = xi ei , into a point P 0 described, in the same RF by the vector
0

OP S(OP ) = xi 0 ei :
S 0

OP OP S(OP ) .

(2.3.33)

using eq. (2.3.32) we can relate the components of the transformed vector, i.e. the coordinates of the new point P 0 , to the coordinates of the original point P :
xi 0 = MS i j xj

r0 = MS r .

(2.3.34)

We have described a homogeneous linear transformation on Mn , in the active picture, as


originating from a linear transformation (an automorphism) on the corresponding linear
vector space Vn . Notice that in the two descriptions (active and passive) the relation between
the final and initial coordinates is the same, their interpretation is different: In the former
r = (xi ) and r0 = (xi 0 ) are the coordinates of a same point relative to different bases,
in the latter the two vectors define the coordinates of different points relative to the same
basis. In particular, the matrix MS implementing the action of S on the coordinates (2.3.34)
and on the components of a generic vector (2.3.32) is the constant Jacobian matrix of the
i0
, so that just as dxi , also the components of a generic vector V i ,
transformation MS i j x
xj
i0
relative to (ei ), transform as contravariant quantities: V i 0 = MS i j V j = x
V j . From now
xj
on we shall not make any difference between homogeneous linear transformations on Mn and
the corresponding linear transformations on the corresponding Vn : Whatever we say for the
former holds true for the latter and vice-versa.
We have learned that the action of a linear homogeneous transformation on the coordinate
vector is described by a constant n n, non-singular matrix M (M i j ), which coincides
with its Jacobian matrix. The same transformation can be viewed as a linear transformation,
an automorphism, on the corresponding linear vector space Vn and is represented on the
components of a generic vector by the action of the same constant matrix M. Let us
show that the set of linear homogeneous transformations is a subgroup of the group of
transformation TMn on Mn . Take two linear homogeneous transformations T, S, let us show
that T S 1 is still linear and homogeneous. Suppose S maps r into r0 = MS r, S 1 will map
1
0
r0 into r = M1
S r and thus is realized by the matrix MS 1 = MS . Suppose now T maps r
0
into r00 = MT r, then T S 1 will map r0 into r00 = MT r = (MT M1
S ) r . We conclude that

60

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

T S 1 is still a linear homogeneous transformation realized by the matrix MT M1


S , and
thus that linear homogeneous transformations close a subgroup of TMn . In Section 2.1.1 we
have shown that the group of automorphisms on Vn is isomorphic to GL(n, F), where F is
R or C depending on whether Vn is a real or complex linear vector space. We conclude that
we can identify GL(n, F) with the group of linear homogeneous transformations on Mn or,
equivalently the group of linear transformations on Vn and that GL(n, F) is a subgroup of
TMn .

2.3.3

Affine Transformations

A more general form of transformations of Mn and on Vn are the affine ones which include,
besides homogeneous linear transformations also translations. Affine transformations are not
homogeneous since the 0-vector in Vn is not mapped into itself or, equivalently, the origin
O of the RF in Mn is not invariant. Let us describe an affine transformation on Mn in the
passive view-point. Consider two generic coordinate systems in Mn : One with origin O and
basis (ei ), the other with origin O0 and basis (e0i ). Let the two bases be related as in eq.

(2.3.27). A point P is described by the vector OP = xi ei with respect to the former and

by O0 P = xi 0 e0i with respect to the latter RF. Let O0 O = xi0 e0i be the position vector of O0
relative to O. From the relation:
0

O P = OP + O0 O ,

(2.3.35)

we derive the following relation between the new and old coordinates of P
xi 0 e0i = xi ei + xi0 e0i = xi M j i e0i + xi0 e0i .

(2.3.36)

Equating the components of the vectors on the right and left hand side we find
xi 0 = xi M j i + xi0 .

(2.3.37)

Let us introduce the column vectors of coordinates r0 (xi 0 ), r (xi ), r0 (xi0 ) and the
square non-singular matrix M (M i j ), so that we can rewrite the coordinate transformation
as follows:
r0 = M r + r0 .

(2.3.38)

The affine transformation, to be denoted by the couple (M, r0 ), is described by the n n


matrix M and by the n-component vector r0 describing its translational component O O0 .
Since r0 is a constant vector, independent of the point P , the coordinate differentials dr =
(dxi ) transform only through the homogeneous linear component of the affine transformation,
according to (2.3.28), as the reader can verify by differentiating both sides of (2.3.38). Recall
that in the passive description adopted so far r0 and r are column vectors describing the same
point P in two different RFs. In the active description they describe two different points
P, P 0 with respect to the same RF.

2.3. TRANSFORMATIONS

61

On Vn we define an affine transformation as a correspondence between two column vectors


V and V0 of the form (2.3.38):
V0 = M V + a .

(2.3.39)

This transformation will be denoted by the couple (M, a), a being a column vector a (ai ).
If r0 , or a, describing the translation, are the zero-vector, the affine transformation reduces
to a homogeneous linear transformation. Affine transformations close a group, as the reader
can verify by applying two consecutive transformations on a same vector. Indeed the product
of two affine transformations, the identity element and the inverse have the following form:
(M1 , a1 ) (M2 , a2 ) = (M1 M2 , M1 a2 + a1 ) ,
I = (1n , 0) ,
1
(M, a)
= (M1 , M1 a) .

(2.3.40)

The homogeneous linear transformations on Mn close a subgroup of the affine transformation


group.
Of particular interest in physics are the affine transformations on the Minkowski spacetime M (1,3) , which preserve the invariant distance ds2 in (2.2.8). These transformations
consist in a combination of a Lorentz transformation which represents their homogeneous
linear component and of a translation, and close a group called the Poincare group.

2.3.4

The volume preserving group SL(n, F)

Consider now transformations on Mn , space of points over F, which preserve the volume
element dV = dn x = dx1 . . . dxn . Under a generic transformation S mapping xi xi 0 (xj )
the volume element transforms with the Jacobian, namely with the determinant of the
Jacobian matrix:
 i0
x
0
i
i
= det(MS ) .
(2.3.41)
dV dV = J(x ) dV , J(x ) det
xj
Volume preserving transformations are those transformations under which dV 0 = dV and
thus are described by a Jacobian matrix M with unit determinant: det(MS ) = 1.
For general transformations the Jacobian matrix depends on the point through the coordinates (xi ). For homogeneous linear transformations the Jacobian matrix is constant.
Volume preserving homogeneous linear transformations on Mn are therefore represented by
n n matrices with entries in F and unit determinant. These transformations therefore close
the group SL(n, F).

2.3.5

(Pseudo-) Orthogonal Transformations

Transformation property of a metric Consider a real metric space M (p,q) . On it the


notion of metric (i.e. of distance) is defined by means of a symmetric scalar product (, )

62

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

on the linear vector space Vn over R. Consider two infinitely close points and let dr = (dxi )
denote the components of the relative position vector. The squared distance ds2 between
the two points is given by eq. (2.2.5). Consider now a transformation S under which dr
transforms with the corresponding Jacobian matrix MS as in eq. (2.3.4). In the passive
description the points do not change and neither does their distance. We can thus write the
same distance squared with respect to dxi and dxi 0 as follows:
xi
xj
0
g
dx` 0 = dxk 0 gk`
dx` 0 ,
ij
xk 0
x` 0

(2.3.42)

 0
1
ds2 = drT g dr = dr0 T MT
dr = dr0 T g0 dr0 .
S g MS

(2.3.43)

ds2 = dxi gij dxj = dxk 0


or, in matrix notation:

With respect to the new coordinate differentials, the distance is expressed in terms of a new
metric g0 = (gij0 ) related to the old one g = (gij ) as follows:
0
gk`

xj
xi
1
gij ` 0 g0 MT

S g MS .
k
0
x
x

(2.3.44)

Equation (2.3.44) defines the transformation property of a metric with respect to a generic
transformation.
A transformation S is said to be metric preserving if ds2 has the same expression in
terms of the new and old coordinates. This means that gij0 = gij , that is:
gk` =

xj
xi
1
g
g = MT
g = MTS g MS .
ij
S g MS
xk 0
x` 0

(2.3.45)

Covariant components of a vector Consider now a generic transformation S : xi


xi 0 (xj ) on M (p,q) not necessarily homogeneous linear or metric preserving. The metric transforms as in (2.3.44) while the components V = (V i ) of a generic vector transform as coni0
travariant quantities: V i 0 = M i j V j = x
V j . Let us define the following quantities:
xj
Vi gij V j ,

(2.3.46)

called covariant components of the vector, obtained by lowering the upper index of the
contravariant components by means of the metric gij . Let us prove that these quantities
transform indeed as covariant quantities, namely as the components of a gradient (2.3.19):
  j0 
 k
x
x`
x
xk
xk
s
j
j
0
0
j0
Vi gij V Vi = gij V =
g
V
=
g
V
=
Vk ,
k`
kj
xi 0
xj 0
xs
xi 0
xi 0
`

j0

x x
`
where we have used the usual property of Jacobian matrices: x
j 0 xs = s .
i
If we write a vector as V = V ei , using the linearity and the symmetry properties of the
inner product we find:

(V, ei ) = V j (ej , ei ) = gij V j = Vi .

(2.3.47)

2.3. TRANSFORMATIONS

63

Consider an Euclidean space Mn = En in which the metric gij is positive definite, and
we can choose an ortho-normal basis in which gij = ij (p = n, q = 0 case). Recalling
the geometrical meaning of the scalar product on En , Vi are nothing but the orthogonal
projection of the vector V along the direction ei and coincide with V i .
The scalar product between two vectors can also be written in the following way:
(V, W) = V i gij W j = Vj W j = V i Wi .

(2.3.48)

It is natural to think of the covariant components of a vector as entries of a row vector, so


that the inner product (2.3.48) can be written in a nice matrix notation:
1
1
W
V
W2
V2

(V, W) = (V1 , V2 , . . . , Vn )
(2.3.49)
... = (W1 , W2 , . . . , Wn ) ... .
Wn
Vn
Notice that, under any transformation:
Vi W i = Vi0 W i 0 ,

(2.3.50)

that is the expression (2.3.48) of the scalar product in terms of the covariant-contravariant
components of the two vectors is the same before and after the transformation. Its expression
in terms of the covariant or contravariant components only is not, as previously explained.
(Pseudo)-Orthogonal transformations Consider now homogeneous linear transformations on M (p,q) . We have learned that we can always choose an orthonormal basis vectors of
Vn in the definition of the RF so that g = p,q . The constant real matrix MS representing
the action of a metric preserving homogeneous linear transformation S will then satisfy the
following property:
1
MT
= p,q MTS p,q MS = p,q .
S p,q MS

(2.3.51)

The matrix MS is then a pseudo-orthogonal matrix.


We conclude that metric preserving linear homogeneous transformations on a real space
(p,q)
M
close the a group O(p, q).
In an Euclidean space, in which p = n and p,q = 1n , it is easy to show that, in an
ortho-normal basis, if M is an orthogonal transformation on Vn (and thus MT is), for any
vectors V and W:
(V, M W) = (MT V, MT M W) = (MT V, W) ,

(2.3.52)

where we have used MT M = 1n .


As an example consider the space M (1,1) on which the metric reads:
ds2 = (dx0 )2 (dx1 )2 = drT g dr ,

(2.3.53)

64

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS




where the metric g =


1 0
. Consider the transformation represented by the following
0 1

constant matrix M:

M =

cosh(a) sinh(a)
sinh(a) cosh(a)


,

(2.3.54)

so that
dx0 0 = cosh(a) dx0 + sinh(a) dx1 ,
dx1 0 = sinh(a) dx0 + cosh(a) dx1 .

(2.3.55)

This transformation is a symmetry of the metric (2.3.53) as the reader can verify by substituting the expression of dx in terms of dx 0 and substituting them in the metric ds2 . One
indeed finds:
ds2 = (dx0 0 )2 (dx1 0 )2 ,

(2.3.56)

that is the metric in the new coordinate differential has the same expression as in the old
ones. The matrix M is in fact in SO(1, 1).
Consider now the action of linear transformations on Vn . Given two vectors, whose
components with respect to an orthonormal basis are described by V = (V i ) and W = (W i )
respectively, their scalar product reads:
(V, W) = VT p,q W .

(2.3.57)

The value of the scalar product depends only on the two vectors on which it is computed
and not on their description. In the passive representation of a linear transformation, the
value of the scalar product is the same if computed in terms of the components V and W of
the two vectors in the old basis or in terms of the components V0 = M V and W0 = M W
of the same vectors in the new basis:
(V, W) = VT p,q W = V0 T MT p,q M1 W0 .

(2.3.58)

For a generic matrix M the product has a different expression in terms of the old and
new components: (V, W) = VT p,q W 6= V0 T p,q W0 = (V0 , W0 ). The transformation
preserves the inner product iff the expression of the product in terms of the old components
(V, W) coincides with that in terms of the new ones:
(V0 , W0 ) = (V, W) ,

(2.3.59)

forany V and W, or, in other words:


(M V, M W) = (V, W) .

(2.3.60)

Using (2.3.58) the above condition amounts to requiring that:


MT p,q M = p,q ,

(2.3.61)

namely that M O(p, q). We conclude that the orthogonal (or pseudo-orthogonal) group is
the group of linear transformations on Vn preserving a symmetric inner product.

2.3. TRANSFORMATIONS

2.3.6

65

Unitary Transformations

Consider now a complex space of points Mn on whose complex linear vector space Vn a
hermitian inner product (, ) is defined. Consider a linear transformation on Vn , represented
by a constant complex matrix M = (M i j ), and two vectors with (complex) components
V = (V i ), W = (W i ) before and V0 = M V, W0 = M W after the transformation. The
transformation preserves the hermitian inner product if the latter has the same form when
expressed in terms of the old and new components, for any couple of vectors:
(V, W) = V i gij W i = V i 0 gij W i 0 = (V0 , W0 ) = (M V, M W) .

(2.3.62)

Substituting in the above equation the expression of the new components in terms of the old
ones and requiring the equality to hold for any (V i ) and (W i ) we find:
(M i k ) gij M j ` = gk` M g M = g .

(2.3.63)

If we choose an ortho-normal basis, gij = ij , and thus g = 1n , equation (2.3.63) becomes:


M M = 1n ,

(2.3.64)

that is M is a unitary matrix. The group U(n) consists of all the transformations which
leave a hermitian inner product invariant.
From eq. (2.3.62), written in a ortho-normal basis, it follows that, if M is a unitary
matrix (and thus M is), for any V and W in Vn :
(V, M W) = (M V, M M W) = (M V, W) .

2.3.7

(2.3.65)

Homomorphism between SU(2) and SO(3)

At this point we are ready to consider an important instance of homomorphism between Lie
groups: That between SU(2) and SO(3). We define a mapping between elements of the two
groups as follows. Consider an element (2 2 complex matrix) S = (S a b ), a, b = 1, 2, of
SU(2) and its adjoint action on the Pauli matrices defined in (2.1.37): S1 i S = S i S,
i, j = 1, 2, 3. Since the Pauli matrices form a basis for hermitian traceless matrices, resulting
matrix is still hermitian traceless:
(S i S) = S i S = S i S ,

Tr(S i S) = Tr(SS i ) = Tr(i ) = 0 . (2.3.66)

Therefore S i S can be expanded in the basis (i ). Let us denote by R[S]i j the components
along i of S i S:
S i S = R[S]i j j .

(2.3.67)

Since R[S] (R[S]i j ) is a 3 3 matrix, we have thus defined a correspondence which maps
a 2 2 matrix S of SU(2) into a 3 3 matrix R[S]. We want to show first that this
correspondence is a homomorphism, namely that R[S1 S2 ]i j = R[S1 ]i k R[S2 ]k j :
(S1 S2 ) i (S1 S2 ) = S2 (S1 i S1 ) S2 = R[S1 ]i k (S2 k S2 ) = R[S1 ]i k R[S2 ]k j j =
= (R[S1 ] R[S2 ])i j j .
(2.3.68)

66

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Let us prove now that the matrix R[S] is real by computing the hermitian-conjugate of both
sides of eq. (2.3.67) and using the property that the left hand side is hermitian:
S i S = (S i S) = (R[S]i j ) j .

(2.3.69)

Since the components associated with any vector (in this space vectors are hermitian matrices!) are unique, comparing (2.3.69) to (2.3.67) we find: (R[S]i j ) = R[S]i j . Using the
definition of inner product for hermitian 2 2 matrices we can write
R[S]i j =

1
Tr[(S i S) j ] ,
2

(2.3.70)

Finally let us show that the matrix R[S] is orthogonal. To this end we use the general
property of homomorphisms that: R[S1 ] = R[S]1 and write
R[S]1 i j = R[S ]i j =

1
1
Tr[(S i S ) j ] = Tr[(S j S) i ] = R[S]j i ,
2
2

(2.3.71)

where we have used the cyclic property of the trace. We conclude that R[S]1 = R[S]T ,
which means that R[S] O(3). Let us show that R[S] SO(3), namely that det(R[S]) = 1.
To show this let us use the property that 1 2 3 = i 12 . Then, from unitarity of S it follows
that:
12 = S S = i S 1 2 3 S = i (S 1 S) (S 2 S) (S 3 S) =
= i (R[S]1 i i ) (R[S]2 j j ) (R[S]3 k k ) = i R[S]1 i R[S]2 j R[S]3 k (i j k )(2.3.72)
.
Now use eq. (2.1.41) to rewrite i j k . Notice that the terms with the P
matrix do not coni
j
tribute because of the orthogonality property of R: R[S]k R[S]` ij = 3i=1 R[S]k i R[S]` i =
k` , which is zero if k 6= `. The only term in i j k which contributes to the summation is
i ijk 12 , and therefore we can rewrite eq. (2.3.72) as follows:
R[S]1 i R[S]2 j R[S]3 k ijk 12 = 12 .

(2.3.73)

We recognize in the sum R[S]1 i R[S]2 j R[S]3 k ijk the expression of the determinant of a
matrix in terms of its entries and therefore we conclude that:
det(R[S]) = 1 ,

(2.3.74)

namely that R[S] SO(3). We have thus defined a homomorphism between SU(2) and
SO(3):
R

S SU(2) R[S] SO(3) .

(2.3.75)

This homomorphism is two-to-one. Indeed, recalling the notation introduced in Section 1.5
of the first chapter, let us compute the normal subgroup E of SU(2) consisting of all the
elements which are mapped into the identity 13 of SO(3):
S E SU(2)

S i S = i

(2.3.76)

2.3. TRANSFORMATIONS

67

The last condition implies that S commutes with all the Pauli matrices, i.e. S i = i S, for
all i = 1, 2, 3. The reader can convince himself that this condition can only be satisfied by
S = 12 . From 1 = det(S) = 2 it follows that = 1. We conclude that E = {12 , 12 },
namely E is the cyclic group of period 2, also denoted by Z2 . Since i are also SU(2)
matrices, E = Z2 is also the center C of SU(2). From Property 1.4 of Section 1.5 it follows
that
SU(2)/E = SU(2)/Z2 SO(3) .

(2.3.77)

Exercise 2.4: Consider the most general SU(2) matrix given in Exercise 1.11 of the first
chapter:


cos() ei
sin() ei
S[, , ] =
,
(2.3.78)
sin() ei cos() ei
and let us construct the corresponding generic element R[, , ] of SO(3). Observe that if
and , the matrix in (2.3.78) changes sign: S S. However, according
to our previous discussion, the corresponding SO(3) matrix should remain unchanged, since
both S and S correspond to the same orthogonal matrix. Using (2.3.70) prove that:
cos(2 ) cos2 () cos(2 ) sin2 ()
cos2 () sin(2 ) + sin(2 ) sin2 ()
cos( ) sin(2 )

R[, , ]

cos2 () sin(2 ) + sin(2 ) sin2 ()


cos(2 ) cos2 () + cos(2 ) sin2 ()
sin( ) sin(2 )

cos( + ) sin(2 )
sin( + ) sin(2 )
(2.3.79)
.
cos(2 )

We see that R[, , ] = R[, , ]. R[, , ] describes the most general rotation of
a system of Cartesian axes in the three dimensional Euclidean space: e0i = R[, , ]i j ej . Its
parameters , , are related to the well known Euler angles (, , ) as follows: = 12 ,
= 12 ( + ) and = 12 ( ), with [0, ], , [0, 2 ].

2.3.8

Symplectic Transformations

Let M2n be a space of points on whose linear vector space V2n over F, a symplectic product
is defined. Consider a linear transformation over V2n , represented by the 2n 2n matrix
M = (M i j ) with entries in F, which leaves the product invariant. Just as for the symmetric
and hermitian inner product, this means that, for any two vectors V = (V i ) and W = (W i ):
(M V, M W) = (V, W) .

(2.3.80)

Using the definition (2.1.30), the above condition implies:


M i k ij M j ` = k`

MT M = .

(2.3.81)

Choosing a basis of V2n in which = (ij ) has the form (2.1.32), condition (2.3.81) implies
that M is a symplectic matrix. The group Sp(n, F) consists of all the transformations which
leave a skew-symmetric (i.e. symplectic) product invariant.

68

2.3.9

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Active Transformations in Different Bases

Consider a linear vector space Vn over F and two bases (ei ), (fi ), related by a transformation
matrix A = (Ai j ):
fi = Aj i ej .

(2.3.82)

Now consider an active transformation S mapping a generic vector of components V = (V i )


= (V i ) = MS V relative to the same
relative to (ei ) into a new vector of components V
basis. We wish to describe the action of S relative to the basis (fi ). Consider a vector
with components V i 0 relative to (fi ). Its components relative to (ei ) are V i = Ai j V j 0 .
This vector is mapped by means of S into an other one with components V i = MS i j V j =
MS i j Aj k V k 0 relative to (ei ). Now let us express the new vector back in the basis (fi ). The
corresponding components will be V i 0 = A1 i j V j = (A1 MS A)i j V j 0 . The transformation
S will then map a vector with components V0 = (V i 0 ) relative to (fi ) into a different vector
0 = (V i 0 ) relative to the same basis, where:
with components V

0 = A1 MS A V0 .
V
(2.3.83)
To summarize, if the active action of a linear transformation S in the basis (ei ) is represented
by a matrix MS , the action of the same transformation in the basis (fi ) related to the former
as in (2.3.82) is represented by the matrix A1 MS A, obtained from MS through the adjoint
action of the matrix A, also called similarity transformation. In other words, the effect of
a change of basis on the matrix representing an active transformation is described by a
similarity transformation.

2.4

Realization of an Abstract Group

Given an abstract group G a realization of G on a point-field M is a mapping T between


the elements of G and elements of TM , i.e. transformations on M :
T : g G T [g] TM ,

(2.4.1)

such that, given two elements g1 g2 G, their product corresponds to the product of the
transformations associated with each of them:
T [g1 g2 ] = T [g1 ] T [g2 ] .

(2.4.2)

In other words a realization is a homomorphism between an abstract group G and the


group TM of transformations on M . As we have learned in the first chapter, when discussing
about homomorphisms between groups, the following properties hold: T [e] = I and T [g 1 ] =
T [g]1 .
The realization is called faithful if it is one-to-one, namely if T [g1 ] = T [g2 ] implies g1 = g2 .
As shown in Section 1.5, this is equivalent to requiring that T [g] is the identity transformation
I only if g = e.

2.4. REALIZATION OF AN ABSTRACT GROUP

69

Let now T be a faithful realization of G, then the image T [G] of G through T is a


group of transformations which is isomorphic to G (G T [G]), namely they have the same
structure. With an abuse of notation we shall call T [G], and not just the mapping T , a
faithful realization of G. Given a transformation group G we can always define an abstract
group G of which G is the image through a faithful realization T : G = T [G] G. If we have
two transformation groups G and G0 which are isomorphic G G0 , and thus share the same
structure, we can view them as the faithful realizations of a same abstract group. Indeed the
abstract groups associated with G and G0 are themselves isomorphic, but, being abstract
groups made just of symbols, they can be identified. G and G0 , on the other hand, in spite
of being isomorphic, cannot be identified since they are represented in general by different
objects (transformations on a point-field). The notion of faithful realization allows us then to
capture the common structure of all isomorphic transformation groups into a single abstract
group. We can then view an abstract group as an abstraction of a transformation group in
which the single transformations are replaced by a symbol, and which has its same structure.
For instance we can associate with each linear transformation group (GL(n, F), O(n, F), U(n)
etc...) on a linear vector space its own abstract group to be denoted by the same symbol.
As we shall see such abstract groups can have various other faithful realizations which are
different from the transformation groups we started from and which we used to define them
(defining realizations).
Let us fist pose the question: Given an abstract group G, does it always admit a faithful
realization over some point-field? The answer is positive. Indeed we can take as point-field
the group G itself, take an element a G and consider the mapping T [a] of G into itself:
gG

T [a](g) a g G

(2.4.3)

T is called left translation and is one-to-one. Indeed suppose for some a G T [a] is the
identity mapping T [a] = I. This means that, for any g G, T [a](g) a g = g. Multiplying
both sides to the right by g 1 we find a = e. It is a homomorphism: Take a1 , a2 G, for
any g G, T [a1 a2 ](g) (a1 a2 ) g = a1 (a2 g) = T [a1 ](T [a2 ](g)) = (T [a1 ] T [a2 ])(g).
We conclude that T is a faithful realization of G.
Let us give now an example of a non-faithful realization of G. For any a G let us define
T [a] as the mapping of G into itself which consists in the conjugation with respect to a:
gG

T [a](g) = a (g) a g a1 G .

(2.4.4)

We have considered this realization in last chapter, a being an inner automorphism of G,


and we have shown that it is indeed a homomorphism. However it is not faithful in general
since all the elements of the center C of G are mapped into the identity transformation:
T [C] = {I}. The group of inner automorphisms IG of G is a faithful realization of G/C.
In general, given a realization T of an abstract group G, if we denote by E the normal
subgroup of G which is mapped into the identity transformation, T [E] = {I}, see Section
1.5, as proven in the same section, the factor group G/E is isomorphic to T [G], G/E T [G],
namely T is a faithful realization of G/E.
Given an abstract group G the mapping between any of its elements and the identity
transformation I on some point-field is clearly a non-faithful realization called the identity

70

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

representation and denoted by 1. In this case E = G and G/E consists of the identity coset
only.
Property 2.1: Any finite group G of order n is realized by a subgroup of the permutation
group Sn .
To show this let us define the permutation which realizes a generic group element g G.
Let g1 , g2 , . . . , gn the generators of G and consider the action on them of the left translation
T [g] by g. This transformation maps a generic element gi into T [g](gi ) = g gi which must
coincide with some other element gki . This correspondence is one-to-one, as we have already
shown. The effect of T [g] on the group elements is therefore to permute them and so we can
describe it by means of a permutation:

 

g1
g2
...
gn
g1 g2 . . . g n
g G T [g]
=
Sn .
g g1 g g2 . . . g gn
gk1 gk2 . . . gkn
(2.4.5)
If g, g 0 G, the realization of their product is the product of the permutations realizing
each of them. In formulas:


g1
g2
...
gn
0
=
T [g g ]
(g g 0 ) g1 (g g 0 ) g2 . . . (g g 0 ) gn

 

g 0 g1
g 0 g2
...
g 0 gn
g1
g2
...
gn
=

=
g (g 0 g1 ) g (g 0 g2 ) . . . g (g 0 gn )
g 0 g1 g 0 g2 . . . g 0 gn
= T [g] T [g 0 ] .
(2.4.6)
This shows that the correspondence (2.4.5) is a homomorphism and thus a realization. It is
faithful since the left translation is.

2.4.1

Representations

A realization of an abstract group G in terms of homogeneous linear transformations on a


space of points Mn , or, equivalently transformations on a linear vector space Vn , is called
n-dimensional representation. A representation D is therefore a mapping which associates
with each element g of the group G an invertible transformation (automorphism) D[g] on Vn ,
named the representation space or base space, whose action is described by a n n matrix
D[g] (D[g]i j ), i, j = 1, . . . , n:
D

g G

D[g] Aut(Vn ) ,

(2.4.7)

where

D[g] : V Vn D[g](V) = D[g] V Vn ,

(2.4.8)

2.4. REALIZATION OF AN ABSTRACT GROUP

71

or, in components,
D[g]

V i V 0 i = D[g]i j V j .

(2.4.9)

We shall characterize a representation D equivalently as the homomorphism D between an


abstract group G and GL(n, F) which associates with each g G an invertible matrix D[g]
(D[g]i j ) GL(n, F). Being D a homomorphism we must gave that, for any g1 , g2 G:
D[g1 g2 ] = D[g1 ] D[g2 ]

D[g1 g2 ] = D[g1 ] D[g2 ] .

(2.4.10)

Depending on whether the representation space is defined over the real or complex numbers,
i.e. if F equals R or C, the representation is said to be real or complex.
Example 2.6: Any cyclic group G of finite period n and generator g admits the following
faithful representation on C (which is a one dimensional vector space over C):
g D[g] = e

2i
n

g k D[g k ] = D[g]k = e

2ik
n

(2.4.11)

2ik

The group D[G] = {e n }, k = 1, . . . , n, representing the generic cyclic group G is denoted


by Zn . An example of G is the group Cn of rotations about a point on the plane by angles
multiple of 2/n. Each of these rotations are realized by a phase in Zn .
Given a representation D of a group G over a representation space Vn and a non singular
b = A1 D A, defined as:
matrix A, also the mapping D
b = A1 D[g] A ,
g D[g]

(2.4.12)

is a representation. As shown in Subsection 2.3.9, the adjoint action of a non singular matrix
A represents the effect on the matrix representing a transformation over Vn , of a change of
b can be viewed as representing the abstract elements of G by means of
basis. Therefore D
the same transformations as D but expressed in a different basis of Vn . The representations
b related by the adjoint action of a non-singular matrix, are called equivalent.
D and D,
We can consider representing a group G in terms of transformations acting on covariant
vectors (Vi ). We have learned that, if a column vector of contravariant components V = (V i )
transform through the left action of a matrix M = (M i j ), a covariant vector (Vi ), seen as a
row vector, transforms through the right action of M1 , as in the basis elements in (2.3.29),
or, as a column vector, through the left action of MT . Therefore if we have a representation
D of G by means of transformations on contravariant vectors:
g G D[g] : V i

D[g]i j V j ,

(2.4.13)

we can associate with it a representation D0 as transformations on covariant vectors:


g G D0 [g] = D[g]T : Vi

D0 [g]i j Vj = (D[g]T )i j Vj .

(2.4.14)

As previously mentioned, we can associate with each matrix group G defined in section 1.7,
an abstract group G, by associating with each matrix M in G a unique symbol in g G

72

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

such that M = D[g], and requiring G to have the same structure as G. By construction this
correspondence:
g G D[g] G ,

(2.4.15)

is a representation, which defines the abstract group G and thus is called defining representation or fundamental representation. Representations of Lie groups are often denoted by
the boldface of their dimension, so that the defining representations D in terms of n n
matrices are dented by n. The same group G will have infinite other representations. We
will denote the abstract group associated with the matrix group by the same symbol, so
that, for instance SU(n) will also denote the abstract group associated with the group of
n n special unitary matrices.
Example 2.7: The homomorphism (2.3.75) between SU(2) and SO(3) is an example of
representation in which a generic element S of SU(2) is represented by a real 33 orthogonal
matrix R[S]. As previously discussed, the subgroup E which is mapped into the identity
matrix 13 is Z2 = {12 , 12 }, so that R is a faithful representation of SU(2)/E = SU(2)/Z2 .
R is a representation of SU(2) over the Euclidean space E3 and is denoted by the boldface of
its dimension: R 3. This is clearly different from the defining representation 2 of SU(2),
which is complex.
Example 2.8: Given a representation D of a group G over a space Vn over F (which can
be either R or C), an example of non-faithful representation is the determinant of D, over
V1 = F:
g G

det(D[g]) F .

(2.4.16)

It clearly is a representation since det(D[g1 g2 ] = det(D[g1 ] D[g2 ]) = det(D[g1 ]) det(D[g2 ]).


Consider the representation of a group G in terms of unitary matrices, unitary representation, on contravariant vectors V i :
g G D[g] = (D[g]i j ) U(n) ,

(2.4.17)

The corresponding representation on covariant vectors D0 associates with each g G a


matrix D0 [g] = (D0 [g]i j ) = D[g]T . Using the unitarity of D[g], i.e. D[g]1 = D[g] we
find D[g]T = (D[g] )T = D[g] = ((D[g]i j ) ), that is the matrices D0 [g] are simply the
complex conjugate of the matrices D[g] and the representation D0 is also denoted by D
and called the representation conjugate to D. We learn that, if a vector V = (V i ) in a
complex linear vector space Vn transforms as a contravariant vector with respect to a group
of unitary transformations, its complex conjugate V = (V i ) transform as a covariant
vector: (V i ) = Vi . As far as orthogonal representations are concerned, since D[g]T = D[g],
covariant and contravariant vectors transform in the same representation: D0 = D.
Remark: The only unitary group whose defining representation n coincides with its
conjugate n is SU(2). Indeed the matrix D[g] = S[, , ] corresponding to a generic SU(2)

2.4. REALIZATION OF AN ABSTRACT GROUP

73

element g in the defining representation D 2 is given in eq. (2.3.78) as a function of the


three parameters defining g. The reader can verify that:
S[, , ] = 21 S[, , ] 2 ,

(2.4.18)

namely the defining representation is equivalent to its conjugate: 2 2. The matrix A


defining the equivalence is the unitary matrix 2 .
Consider a representation D of an abstract group G over a linear vector space Vn . A
subspace Vk of Vn , for some k < n is called invariant or stable if the action of the transformations in D[G], representing G, maps vectors in Vk into vectors in Vk . A representation
is said to be irreducible if the space Vn it acts on does not admit an invariant subspace
aside from the trivial one consisting of the zero-vector {0}. A representation which is not
irreducible is said to be reducible. Consider a reducible representation D over Vn and let Vk
be an invariant subspace. We can consider a basis (ei ) of Vn so that the first k elements ea ,
a = 1, . . . , k, span Vk , while the last n k elements e` , which are not in Vk , will be labeled
by an index ` = k + 1, . . . , n. The subspace Vk consist of column vectors with components
V a only along the ea vectors:
1
1
V
V
.
.
.
...
k  a
k  a
V
V
V

= V
V k+1 =

V
,
V

Vk .
n
k
`

V
0
V .
0.
..
..
n
V
0
(2.4.19)
The transformation D[g] representing a generic G-element g has to map a column vector of
the form V0 in (2.4.19) into a vector of the same form. It therefore, as a n n matrix, will
have the following characteristic block structure:


Dk [g] B[g]k,nk
g G D[g] =
,
(2.4.20)
0nk,k Ank [g]
The action of D[g] on a generic Vn vector V = (V i ) = (V ` , V a ) is:

0
V ` 0 = A[g]` `0 V `
0
V V = D[g] V
.
V a 0 = B[g]a ` V ` + Dk [g]a b V b

(2.4.21)

If V Vk , we have V ` = 0 and therefore also V0 will have components only in Vk : V ` 0 = 0,


V a 0 = Dk [g]a b V b . The k k block Dk [g], for different g in G, defines a k-dimensional
representation Dk of G over the smaller subspace Vk .
The representation D is called completely reducible if there exist a basis of Vn on which
the matrix representation D[g] of a generic g G has a block diagonal form, namely a form
(2.4.20) with B[g]k,nk = 0k,nk . In this case Vn can be written as the direct sum of two
invariant subspaces Vk and Vnk : Vn = Vk Vnk . The block Ank [g], for a generic g G,

74

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

defines a (n k)-dimensional representation Dnk of G over Vnk . We shall then say that
the original representation D decomposes into the direct sum of Dk and Dnk and write:
D Dk Dnk .

(2.4.22)

The representations Dk , Dnk may still be completely reducible, and thus can be further
decomposed into lower dimensional representations. We can iterate the
P above procedure
until we end up with irreducible representations: Dk1 , . . . , Dk` , where `i=1 ki = n. This
corresponds to finding a basis in which the matrix representation under D of a generic
element g G has the following block structure:

Dk1 [g]
0

0
0
Dk2 [g]
0
D[g] =
.
(2.4.23)
..
..
...
.
.
0

Dk` [g]

We say that the original representation D is completely reducible into the irreducible representations Dki and write:
D

`
M

Dki Dk1 Dk2 . . . Dk` .

(2.4.24)

i=1

Correspondingly the representation space Vn of D has been decomposed into the direct sum
of spaces Vki on which Dk1 [g] act:
Vn = Vk1 Vk2 . . . Vk` ,

(2.4.25)

and the basis (ei ) in which D[g], for any g G, has the form (2.4.23), is the union of bases
in each of the subspaces: (ei ) = (ea1 , ea2 , . . . , ea` ), where ai = 1, . . . , ki and (eai ) is a basis
of Vki .
To summarize, a representation D is completely reducible if it is equivalent to a representation which, for any group element, has the matrix block diagonal form (2.4.23).
Example 2.9: Consider the representation of the group SO(2) as rotations in the XY
plane of the three dimensional Euclidean space:
x x0 = x cos() + y sin() ,
y y 0 = x sin() + y cos() ,
z z0 = z .
The above transformation is implemented by a 3 3 real matrix Dz [g()],
corresponding abstract element g() of SO(2)

cos() sin()
g() SO(2) Dz [g()] = Rz [] = sin() cos()
0
0

representing the

0
0 .
1

(2.4.26)

2.4. REALIZATION OF AN ABSTRACT GROUP

75

The above representation is clearly completely reducible since there is a subspace of the
Euclidean space, namely the Z axis, which is left invariant and the orthogonal complement,
which is the XY plane, also defines an invariant subspace. The action along the Z axis is the
identity representation 1 while the action on the XY plane, defined by the first 2 2 block
of the above matrix, is the defining the (irreducible) representation 2 of SO(2). Denoting
the representation D by its dimension 3 we will then write:
3 1 2 .

(2.4.27)

Similarly we can define other two 3-dimensional representations of SO(2), Dx , Dy , describing


rotations about the X and the Y axes:

1
0
0
Dx [g()] = Rx [] = 0 cos() sin() ,
(2.4.28)
0 sin() cos()

cos() 0 sin()
,
Dy [g()] = Ry [] = 0
1
0
(2.4.29)
sin() 0 cos()
The three representations Dx , Dy , Dz are equivalent as the
1
that: Dz = A1
1 Dx A1 = A2 Dy A2 , where:

0 0 1
1
A1 = 0 1 0 , A2 = 0
1 0 0
0

reader can show by verifying

0 0
0 1 ,
1 0

(2.4.30)

and therefore can be identified with the same (reducible) representation 3 of SO(2).
Question: Are A1 and A2 is O(3)? Are they in SO(3)?
Example 2.10: The group D4 Let us now consider a representation of the symmetry
group D4 of a square, discussed in Section 1.2.3. Let the square lie on the XZ plane, be
centered in the origin, and have its sides parallel to the axes. We wish to represent the
action of the D4 elements in terms of matrices acting on the three-dimensional Euclidean
space. Since all elements of D4 are expressed as products of its two generators r, , it is
convenient to start representing the action of these two transformations. Recall that r is a
90o clockwise rotation about the Y axis. It is illustrated in Figure 2.4 and represented by
the following homogeneous linear transformation on the three coordinates:
x0 = z , y 0 = y , z 0 = x .
Writing r0 = D[r] r, we deduce the form of the

0
D[r] = 0
1

(2.4.31)

matrix associated with this transformation:

0 1
1 0 .
(2.4.32)
0 0

76

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Figure 2.4: Active and passive representation of a clockwise rotation r by 90o about the Y
axis

As expected, since this is a particular rotation about the Y axis, it coincides with Ry [] in
(2.4.29), for = 2 . As far as the reflection is concerned, the corresponding homogeneous
linear transformation reads:
x0 = z , y 0 = y , z 0 = x ,

(2.4.33)

from which it follows that:

0 0 1
D[] = 0 1 0 .
1 0 0

(2.4.34)

Notice that D[e] = D[ 2 ] = D[]2 = 13 . Since the elements of D4 are e, r, r2 , r3 , , =


(r ), = (r2 ), = (r3 ), using the homomorphism property of a representation, we
may deduce the matrix representation of all the D4 elements, which we list below:

1 0 0
0 0 1
1 0 0
D[e] = 0 1 0 , D[r] = 0 1 0 , D[r2 ] = D[r]2 = 0 1 0 ,
0 0 1
1 0 0
0 0 1

1 0 0
0 0 1
0 0 1
D[r3 ] = 0 1 0 , D[] = 0 1 0 , D[] = 0 1 0 ,
1 0 0
1 0 0
0 0 1

1 0 0
0 0 1
D[] = 0 1 0 , D[] = 0 1 0 .
(2.4.35)
1 0 0
0 0 1

2.4. REALIZATION OF AN ABSTRACT GROUP

77

Figure 2.5: Active representation of the reflection in the z + x = 0 plane

Let us notice that all rotations (e, r, r2 , r3 ) are represented by SO matrices. In fact the
rotations close the subgroup C4 of SO(2), and are all obtained by computing the SO(2)
matrix in (2.4.29) for = 0, 2 , , 32 . The reflections are still orthogonal, but have negative
determinant, and thus are only in O(2). Moreover the matrices in (2.4.35) have a characteristic block structure (2.4.23), provided we order the basis elements as (e2 , e1 , e3 ), namely
we perform a change of basis described by:

0 1 0
A = 1 0 0 .
(2.4.36)
0 0 1
The representation is therefore completely reducible in a representation D2 acting on the
subspace e1 , e3 , that is on the XZ plane, and a one-dimensional representation acting on
e2 , that is on the Y axis, which is the trivial or identity representation. Let us write, for the
sake of completeness, the D4 elements in the two dimensional representation:






1 0
0 1
1 0
2
D2 [e] =
, D2 [r] =
, D2 [r ] =
,
0 1
1 0
0 1






0 1
0 1
1 0
3
D2 [r ] =
, D2 [] =
, D2 [] =
,
1 0
1 0
0 1




0 1
1 0
D2 [] =
, D2 [] =
.
(2.4.37)
1 0
0 1
The identity representation of D4 is usually denoted by A1 , while the two dimensional one
by E. Another representation of D4 is the determinant of the above matrices in the D2

78

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

representation, usually denoted by the symbol A2 :


A2 : det(D2 [e]) = det(D2 [r]) = det(D2 [r2 ]) = det(D2 [r3 ]) = 1 ,
det(D2 []) = det(D2 []) = det(D2 []) = det(D2 []) = 1 .

(2.4.38)

The determinant representation, just like the trivial identity representation, is real onedimensional. Rotations are represented by the number +1, reflections by 1. It clearly is
non-faithful, so we are losing information about the group. For D4 the reader can verify that
we have two further one-dimensional representations, denoted by B1 and B2 respectively:
B1 : DB1 [e] = DB1 [r2 ] = DB1 [r ] = DB1 [r3 ] = 1 ,
DB1 [r] = DB1 [r3 ] = DB1 [] = DB1 [r2 ] = 1 ,
B2 : DB2 [e] = DB2 [r2 ] = DB2 [] = DB2 [r2 ] = 1 ,
DB2 [r] = DB2 [r3 ] = DB2 [r ] = DB2 [r3 ] = 1 .

(2.4.39)
(2.4.40)

Therefore for D4 we have found four one-dimensional representations A1 , A2 , B1 , B2 and


one two-dimensional one E.
Exercise 2.5: Using eq. (2.4.28), write the matrix representation of the rotations in Cn
about the X axis.
Example 2.11: Consider the group U(1). We can take as the defining representation the
one in which the elements of the group are described by a phase ei , (0, 2 ). The
elements of the abstract group U(1) are then functions g() of the continuous parameter
. We can define, for any integer n, a one-dimensional representation D(n) of U(1) over the
complex numbers as follows:
g() U(1)

D(n) [g()] ei n .

(2.4.41)

The reader can verify that this is indeed a representation.

2.4.2

The Regular Representation and the Group Algebra

Let us define, for a generic order-n finite group G its regular representation. We have seen in
the previous section that any group admits a faithful realization by means of left-translations
on the group itself. We can associate with this realization a matrix representation of G. Let
us denote by g0 , g2 , . . . , gn1 the elements of G, g0 = e. Given a generic group element g,
the action of T [g] on any element gi is the unique element gki = g gi and is described by
the permutation (2.4.5). Let us associate each element gi of G with an element egi of a
basis of a n-dimensional real vector space VR . Define the matrix DR [g] (DR [g]j i ) whose
entries are all zero except DR [g]ki i which is 1. The matrix DR [g], for any g G, defines a
linear transformation on the vector space VR , to be denoted by DR [g], which maps the basis
element egi into the element eggi = egki :
g

egi DR [g](egi ) eggi = egki = DR [g]j i egj .

(2.4.42)

2.4. REALIZATION OF AN ABSTRACT GROUP

79

All the basis elements egi are constructed by acting on the same element eg0 ee by the
corresponding transformation DR [gi ]: egi = egi e = DR [gi ](ee ). Writing a generic vector
V = V i egi as the column vector V = (V i ) of components V i , namely identifying the basis
elements with the following column vectors:


1
0
0
0
.
.
.
.
eg0 = ee =
(2.4.43)
. , , egn = . ,
0
0
0
1
the action of the linear transformation DR [g] on a generic vector V is represented as follows:
DR [g](V) = V i DR [g]j i egj DR [g] V .

(2.4.44)

Clearly DR [e]i j = ji and it is the only matrix with diagonal entries. To see this suppose
there existed a g G such that DR [g]i j has a diagonal entry, say DR [g]k k (no summation
over k). This means that g gk = gk , which implies g = e.
Let us show that the correspondence DR between elements g G and transformations
DR [g] Aut(VR ) is a homomorphism. Take two generic elements g, g 0 G:
DR [g g 0 ](egi ) = egg0 gi = DR [g](DR [g 0 ](egi )) = (DR [g] DR [g 0 ])(egi ) ,

(2.4.45)

which implies that DR [g g 0 ] = DR [g] DR [g 0 ]. This implies an analogous relation between the
corresponding matrices: DR [gg 0 ] = DR [g] DR [g 0 ]. The correspondence DR between elements
g G and the n n matrix DR [g] is then a n-dimensional representation called the regular
representation of G. This representation is clearly faithful and, as we shall see, completely
reducible. Notice now that DR is an orthogonal representation (i.e. real unitary). Indeed if
the left translation by g is defined by the permutation gi gki and thus the non-vanishing
entries of the matrix DR [g] are DR [g]ki i = 1, the action g 1 is represented by the inverse
permutation gki gi and thus the non-vanishing entries of the corresponding matrix are
DR [g 1 ]i ki = 1. Therefore DR [g 1 ] = DR [g]T . Being DR a homomorphism, we then have
DR [g 1 ] = DR [g]1 = DR [g]T , which proves the orthogonality of DR .
We can easily construct the matrices DR [g] directly from the multiplication table of the
group. To this end we write the multiplication table corresponding to the chosen ordering
(g0 , g2 , . . . , gn1 ) of the group elements. We then permute the columns of the table so that
the unit element e only appears along the diagonal (for instance in the multiplication table
of D3 , given in Section 1.2.3, we should permute the second with the third column). DR [g]
is then the matrix obtained from the resulting table by replacing the element g by 1 and all
the other elements by 0. In the D3 case we then find, for the elements e, r, the following
matrices:

0 0 1 0 0 0
0 0 0 1 0 0
1 0 0 0 0 0
0 0 0 0 0 1

0 1 0 0 0 0
0 0 0 0 1 0

DR [e] = 16 , DR [r] =
0 0 0 0 0 1 , DR [] = 1 0 0 0 0 0 .

0 0 0 1 0 0
0 0 1 0 0 0
0 0 0 0 1 0
0 1 0 0 0 0

80

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Exercise 2.6: Compute the regular representation of the generators r, of D4 .


Given two linear transformations DR [g], DR [g 0 ] on VR , representing the corresponding
elements of G, we can consider the linear function on VR defined by an arbitrary linear
combination of the two:
( DR [g] + DR [g 0 ])(V) DR [g 0 ](V) + DR [g 0 ](V) = ( DR [g] + DR [g 0 ]) V ,
where the matrix describing the action of the linear combination is given by the same combination of the matrices associated with each transformation. In general such linear combination is not invertible and thus should not be regarded as a transformation, i.e. an automorphism on VR , but just an endomorphism2 . The linear transformation DR [g] + DR [g 0 ] in
general does not correspond to any group element either. As opposed to the abstract group
elements of G, on which no sum or multiplication by numbers is defined, we can define the
same operations on the corresponding linear transformations on VR . Consider the space of
all the linear transformations DR [gi ] on VR , defining the regular representation and extend
it with all the linear combinations of these transformations. We end up with a space AR [G]
of endomorphisms on VR , which is closed with respect to linear combinations of its elements
and therefore is itself a linear vector space. By construction, a basis of this space consists of
the linear transformations DR [gi ] : VR VR and therefore AR [G] is n-dimensional. A generic
element R of AR [G] is a endomorphism on VRP
which is expressed as a linear combination of
the n basis elements DR [gi ]: R = Ri DR [gi ] = gG R(g) DR [g], where Ri R(gi ):
X
R(g) DR [g]} .
(2.4.46)
AR [G] = {R End(VR )| R = Ri DR [gi ] =
gG

Similarly we can define the vector space AR [G] consisting of linear combinations of the n n
matrices DR [gi ]:
X
AR [G] = {R n n matrix| R = Ri DR [gi ] =
R(g) DR [g]} .
(2.4.47)
gG

In contrast with a generic linear vector space, on AR [G] a product operation is defined
between the basis elements DR [gi ]. Indeed, in virtue of eq. (2.4.45), if the left-translation
by gi in G maps gj into gi gj = gkj , we have:
DR [gi ] DR [gj ] = DR [gi gj ] = DR [gkj ] = DR [gi ]k j DR [gk ] .

(2.4.48)

An analogous relation holds between the corresponding matrices:


DR [gi ] DR [gj ] = DR [gi ]k j DR [gk ] .

(2.4.49)

We can extend on the whole vector space AR [G] the product operation by requiring it to
be linear in both its arguments (bi-linear), namely that, for any S1 , S2 , S, R1 , R2 , R AR [G]:
( S1 + S2 ) R = S1 R + S2 R ,
S ( R1 + R2 ) = S R1 + S R2 ,
2

(2.4.50)

To show this consider the combination DR [g] DR [g], which maps all vectors V into the zero vector 0
and is represented by the zero-matrix 0. It clearly is not invertible.

2.4. REALIZATION OF AN ABSTRACT GROUP

81

, being arbitrary numbers. Therefore, if R = Ri DR [gi ] and S = S j DR [gj ], we define the


product of the two transformations as the following linear transformation R S:
R S = (Ri DR [gi ]) (S j DR [gj ]) = Ri DR [gi ] (S j DR [gj ]) = Ri S j DR [gi ] DR [gj ] =
= (Ri S j DR [gi ]k j ) DR [gk ] .
We can also express the components
of the productPusing the equivalent representation of
P
the elements of AR [G], R = gG R(g) DR [g], S = g0 G S(g 0 ) DR [g 0 ]:
RS =

XX

R(g) S(g 0 ) DR [g] DR [g 0 ] =

gG g 0 G

XX

R(g) S(g 0 ) DR [g g 0 ] =

gG g 0 G

!
=

gG g 00 =gg 0 G

R(g) S(g 1 g 00 ) DR [g 00 ] =

g 00 G

gG

R(g) S(g 1 g 00 )

DR [g 00 ] .

A linear vector space on which a bi-linear product is defined is called an algebra. The
space AR [G], equipped with the product is therefore an algebra, called the group algebra
associated with the group G. More specifically it is an associative algebra since the product
is associative, as it can be easily verified. Similarly also the space AR [G] is an associative
algebra. Equations (2.4.48) and (2.4.49) express the product of two basis elements as a
linear combination of the same basis elements and are called the structure equations of the
algebras AR [G] and AR [G]. They uniquely define the product of any two elements within
each algebra and are characterized by the constants DR [gi ]j k , called the structure constants of
the algebra. The spaces AR [G] and AR [G] have the same dimension n and the same structure
constants. They are isomorphic algebras since the correspondence Ri DR [gi ] Ri DR [gi ]
is an isomorphism between the two vector spaces which preserves the product operation.
Therefore we shall use either AR [G] and AR [G] to describe the same group algebra.
The product on AR [G] has an identity element which is the identity transformation DR [e]
itself: DR [gi ] DR [e] = DR [e] DR [gi ] = DR [gi ]. Moreover it is associative since the product
of transformations is, being represented by product of matrices.
Notice that the whole representation space VR of the regular representation can be generated by acting on the single vector ee by the algebra AR [G]. Indeed to generate a generic
vector V = V i egi VR it suffices to act on ee by the element R = V i DR [gi ] of AR [G]:
R(ee ) = V i DR [gi ](ee ) = V i egi = V.
The correspondence between R = V i DR [gi ] AR [G] and V = V i egi VR is an isomorphism between the linear vector spaces AR [G] and VR , the difference between the two
being that on the former a product operation is defined while on the latter it is not. We
can view AR [G] itself as the representation space of the regular representation in which
DR [gi ] AR [G] plays the role of egi VR . Indeed, just as DR [gi ] : egj DR [gi ](egj ) = egi gj ,
we can define the action of AR [G] on itself: DR [gi ] : DR [gj ] DR [gi ] DR [gj ] = DR [gi gj ].
Finally let us consider a generic representation D of G on a m-dimensional linear vector
space Vm . The elements gi G, i = 0, . . . , n = |G| 1, are represented by linear transformations D[gi ] Aut(Vm ). Just as we did for linear transformations on VR we can consider

82

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

linear combinations of automorphisms on Vm and define the linear vector space AD [G] with
basis elements D[gi ], consisting of endomorphisms on Vm of the form: R = Ri D[gi ].
AD [G] = {R End(Vm )| R = Ri D[gi ]} ,

(2.4.51)

On this space a bilinear product is defined from the product of two generic basis elements:
D[gi ] D[gj ] = D[gi gj ] = D[gkj ] = DR [gi ]k j D[gk ] ,

(2.4.52)

which promotes AD [G] to an algebra. Similarly we can define the matrix description of the
above algebra:
AD [G] = {R = Ri D[gi ]} ,

(2.4.53)

characterized by the same structure:


D[gi ] D[gj ] = DR [gi ]k j D[gk ] ,

(2.4.54)

and thus isomorphic to AD [G]. Notice that AD [G] has the same dimension and structure
as AR [G] and thus the two are isomorphic algebras. In fact AR [G] is a particular instance
of AD [G] when the representation D is the regular one: D = DR . Just as we did for
groups, we can capture the common structure of isomorphic algebras by defining an abstract
algebra A[G] consisting of symbols, uniquely characterized by its dimension and structure,
and say that two isomorphic algebras of endomorphisms are different representations of a
same abstract algebra. AR [G] and AD [G] are then different representations, over VR and
Vm respectively, of a same abstract algebra A[G].
Let us define A[G] as a n-dimensional vector space, n = |G|, whose basis elements are
in correspondence with the elements of the group G and will be, with an abuse of notation,
denoted by the same symbols: g0 , g1 , . . . , gn . A generic element of A A[G] is then expressed
as a linear combination of the gi :
A[G] {A =

n
X
i=1

Ai gi =

A(g) g} .

(2.4.55)

gG

Let a bi-linear product on A[G] be defined, which maps couples of elements of A[G] into
their product which is still an element of A[G]: For any A, B A[G], AB A[G]. Being bilinear, this product is totally defined by the product of the basis elements, which is expressed
in components relative to the same basis through the structure constants. The product of
the basis elements is in turn defined by the product of the corresponding elements in G,
namely by the structure of G itself: Suppose the left-translation by the element gi maps gj
into gkj , then
gi gj = gkj = DR [gi ]k j gk .

(2.4.56)

Notice that the last equality in the above formula would be meaningless if gi , gj , gk were
elements of G, since on G there is no notion of sum, and thus the summation over k on

2.5. SOME PROPERTIES OF REPRESENTATIONS

83

the right hand side would be undefined. The product between two elements A = Ai gi and
B = B j gj of A[G] reads:
A B = Ai B j gi gj = (Ai B j DR [gi ]k j ) gk .

(2.4.57)

The structure constants DR [gi ]k j in (2.4.56), defining the product on A[G], are the same
as those for AD [G] in (2.4.52). A[G] is an associative algebra called the abstract group
algebra associated with G. Given a representation D of G on Vm , we define a corresponding
representation of its algebra A[G] as the following mapping between elements of A[G] and
elements of AD [G], which are endomorphisms on Vm :
D

A = Ai gi D[A] Ai D[gi ] AD [G] ,

(2.4.58)

or, equivalently, the mapping D between element of A[G] and m m matrices in AD [G]:
D

A = Ai gi D[A] Ai D[gi ] AD [G] .

(2.4.59)

By construction D is a homomorphism between the two linear vector spaces which preserves
the product:
D[A B] = D[A] D[B] ,

(2.4.60)

and thus is a homomorphism between the two algebras. The same for D. Just as for groups,
we shall not make any difference between D and D, they define the same representation.
We have then learned that a representation of G induces a representation of its abstract
algebra A[G]. The reverse is also true: A representation D of A[G] induces a representation
of G. We just need to associate with a generic element gi of G the automorphism D[gi ]
corresponding to the basis element gi of A[G].
Just as it was the case for AR [G] and in general for AD [G], we can view A[G], which
as the representation space of the regular representation in which gi A[G] plays the
role of egi VR . Indeed, we can associate with each gi G an automorphism DR [gi ] on
A[G] whose action on an element A = Aj gj is represented by the left-multiplication by
the corresponding element gi of A[G]: A DR [gi ](A) gi A. In other words, just as
DR [gi ] : egj DR [gi ](egj ) = egi gj = DR [gi ]k j egk , we can define the action of DR [gi ] on A[G]
in a similar fashion DR [gi ] : gj DR [gi ](gj ) gi gj = DR [gi ]k j gk , so that it is still described
by the same n n matrix DR [gi ] (DR [gi ]k j ).
The abstract group algebra A[G] is then also the representation space of the regular
representation of the algebra itself. With any A = Ai gi A[G] we associate the endomorphism DR [A] = Ai DR [gi ] on A[G], whose action on a generic element B A[G] is the left
multiplication by A: B DR [A](B) = Ai DR [gi ](B) = Ai (gi B) = A B.

2.5

Some Properties of Representations

Let us prove some important properties of representations of finite groups:

84

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Property 2.2: Any representation of a finite group is equivalent to a unitary one.


Consider a finite group G of order n, which has a representation D over a complex vector
space Vn on which a hermitian, positive definite, scalar product (, )0 is defined. We want
to show that we can always define a hermitian scalar product (, ) with respect to which D
is a unitary representation, that is, for any g G and for any vectors V and W:
(D[g] V, D[g] W) = (V, W) .

(2.5.1)

Define the scalar product (, ) on two generic vectors V and W as follows:


(V, W)

def. 1 X
(D[g] V, D[g] W)0 .
=
n gG

(2.5.2)

We can easily show that the product defined above is hermitian. In particular it clearly is
linear with respect to the second argument. Moreover:
(W, V) =

1 X
1 X
(D[g] W, D[g] V)0 =
(D[g] V, D[g] W)0 = (V, W) .(2.5.3)
n gG
n gG

The positive definiteness of (, ) follows from the same property of (, )0 since (V, V), being
the sum of positive numbers (D[g] V, D[g] V)0 , is itself positive, and moreover it vanishes
only if each term in the sum vanish, which is the case only if V = 0.
Let us show that with respect to this product D is unitary:
1 X
(D[g g 0 ] W, D[g g 0 ] V)0 =
n gG
1 X
1 X
(D[g g 0 ] W, D[g g 0 ] V)0 =
(D[g 00 ] W, D[g 00 ] V)0 =
=
n gg0 G
n g00 G

(D[g] V, D[g] W) =

= (V, W) .

(2.5.4)

To complete the proof, we should show that D is equivalent to a unitary representation


relative to the original scalar product (, )0 . Let us define the metrics associated with the
two products: gij0 (ei , ej )0 and gij (ei , ej ), relative to a given basis (ei ). These two
matrices g0 (gij0 ) and g (gij ) are both hermitian positive definite. From the general
theory of matrices it follows that there exist a (in general not unitary) complex matrix
M such that g = M g0 M or, equivalently (M1 ) g M1 = g0 . Let us show that the
M D M1 is unitary with respect to g0 . Choose a generic g G:
representation D
g0 D[g]

D[g]
= (M1 ) D[g] (M g0 M) D[g] M1 = (M1 ) D[g] g D[g] M1 =
= (M1 ) g M1 = g0 ,
(2.5.5)
where we have used unitarity of D with respect to g, namely that D[g] g D[g] = g. This
completes the proof.

2.5. SOME PROPERTIES OF REPRESENTATIONS

85

Remark: If the representation D is defined over a real space Vn with a symmetric scalar
product (, )0 , by the same token, we can define a new symmetric product (, ) with respect to which D is real orthogonal and, following the same procedure as above, define a
representation equivalent to D which is orthogonal with respect to (, )0 . Therefore any real
representation of a finite group is equivalent to an orthogonal one.
Property 2.3: Any reducible representation of a finite group G is completely reducible.
Let D be a reducible representation of G over a complex vector space Vn equipped
with a hermitian scalar product. In virtue of the previous property, we can define a new
hermitian scalar product (, ) with respect to which D is unitary. By assumption there
exists a subspace Vk of Vn , for some k < n, which is stable with respect to the action of
the transformation D[g], for any g G, i.e. such that D[G](Vk ) Vk . Let us choose an
ortho-normal basis (ei ) with respect to (, ), and define with respect to (, ) the orthogonal
complement Vnk of Vk in Vn , as the space consisting of the vectors in Vn which are orthogonal
to all vectors in Vk :
Vnk {V Vn | (V, W) = 0 for any W Vk } .

(2.5.6)

Vnk is clearly a vector space, since the linear combination of any two vectors orthogonal to
all vectors in Vk has itself this property. Let us show that Vnk is stable with respect to the
action of G defined by D. To prove this, it suffices to show that, if V Vnk , then, for any
g G, D[g] V Vnk . This is the case if D[g] V is orthogonal to all vectors in Vk . Let us
then take a generic vector W in Vk :
(D[g] V, W) = (V, D[g] W) = (V, D[g 1 ] W) = 0 ,

(2.5.7)

where we have used eq. (2.3.65) and the fact that D[g 1 ] W is still in Vk by assumption. We
have thus shown that Vn can be decomposed into two disjoint subspaces Vk and Vnk each
of which is stable with respect to D[G]. As a consequence of this, if we choose a basis for Vn
whose first k elements are in Vk and last n k elements are in Vnk , in this basis the matrix
form of D[g], for any g G, is block diagonal. In other words D is completely reducible into
two representations Dk and Dnk acting on Vk and Vnk respectively:
D Dk Dnk .

(2.5.8)

This completes the proof.


Let us mention, without proving, that the above two properties also hold for compact
Lie groups (such as O(n) or U(n)):
Property 2.4: Any representation of a compact Lie group is equivalent to a unitary one.
This for instance does not apply to non-compact groups, like the Lorentz group, which
do not admit any unitary representations aside from the trivial one.
Property 2.5:
ducible.

Any reducible representation of a compact Lie group is completely re-

86

2.5.1

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Unitary Representations and Quantum Mechanics

In quantum mechanics all information about the physical state of a system is encoded in the
mathematical notion of quantum state. A quantum state is a vector in a Hilbert space H ,
which, as previously mentioned, is a complex (possibly infinite dimensional) linear vector
space equipped with a hermitian, positive definite, scalar product. It is denoted by the
symbol |i (instead of the boldface letter V that we have been using so far), while the scalar
product (|1 i , |2 i) between two states |1 i , |2 i is denoted by h1 |2 i. One usually works
in a ontho-normal basis, with respect to which, a state |i can be represented as a column
vector consisting of its components and h| its hermitian conjugate, namely the row vector
consisting of the complex conjugate components of |i: h| |i . This notation for the
vectors representing quantum states was originally introduced by Dirac who named |i a
ket vector and its hermitian conjugate h| a bra vector. Given two states |1 i , |2 i, the
quantity:
P (1 , 2 )

|h2 |1 i|2
,
k2 k2 k2 k2

(2.5.9)

where ki k2 hi |i i > 0, represents the probability of finding, upon measurement, a


system which is initially prepared in a state |1 i, in a state |2 i, characterized by a definite
value of the quantity we are measuring. P (1 , 2 ) is also called transition probability between
the states |1 i to |2 i. Clearly the result of a measurement, described by the probability
P (1 , 2 ), is not affected if the two states are multiplied by complex number: |i i
|i i. Physical states are then represented by vectors in H modulo multiplicative complex
numbers, which are unphysical. They are, in other words, described by rays in H , defined as
one-dimensional subspaces { |i} of H consisting of vectors differing from a given one by
a multiplicative number. If we normalize a quantum state to unit norm kk2 h|i = 1,
the multiplicative complex number is fixed up to a phase.
Of particular relevance in quantum mechanics are operators S, i.e. linear functions, on
H . The action of S on a vector |i is denoted by |S i and is represented by the action of
a (possibly infinite dimensional) matrix MS on the corresponding vector components:
|S i S(|i) = MS |i .

(2.5.10)

Since S is not necessarily a transformation, MS need not be invertible. Clearly the action
of S on bra vectors is represented by the right action of the hermitian conjugate matrix:
hS | |S i = hS | MS .

(2.5.11)

In quantum mechanics observables, i.e. measurable physical quantities as energy, momentum, angular momentum, spin etc..., generically denoted by O, are represented by hermitian
An operator O
is hermitian if its matrix, in an ortho-normal basis, is hermitian:
operators O.

MO = MO . If O is hermitian, then, for any couple of states |1 i , |2 i, we have:


2 i = h1 | M |2 i = (h1 | M ) |2 i = hO
1 |2 i ,
h1 |O

O
O

(2.5.12)

2.5. SOME PROPERTIES OF REPRESENTATIONS

87

and therefore, on a same state |i:


i = hO|i

i ,
h|O
= h|O

(2.5.13)

i is a real number. The number:


that is h|O
hOi

i
h|O
R,
kk2

(2.5.14)

represents the expectation value of the observable O on the state |i, the most likely value
that a measure of O will give on that state. With an abuse of notation, to simplify life, we
and for the corresponding matrix: M = O,

shall use the same symbol for the operator O


O
i = O
|i. A state | i with a definite value of the observable is an eigen-state
so that |O

of O, to the eigen-value :
| i = | i .
O

(2.5.15)

Since eigen-vectors to different eigen-values are orthogonal, that is h |0 i = 0 if 0 6= .


This represents the experimental fact that, if a system is prepared in a state with a definite
value of O, we can be sure that any further measure of O on the system will give the same
result, namely the probability P ( , 0 ) of measuring the system in a state with a different
value 0 of the same observable is zero.
Consider now a linear transformation on the space of states H , described by an invertible
operator S. It may represent the effect on the states of a change in the space-time RF, due
to rotations, reflections, Lorentz boosts etc... If a system is described in a state |i by an
observer in a RF, it will in general be described by a different state | 0 i = |S i with respect
to the transformed RF3 . What all observers should agree on is the transition probability P
between two states, defined in (2.5.9), namely we should have:
P (1 , 2 ) = P (10 , 20 ) = P (S 1 , S 2 ) .

(2.5.16)

This is the case only if the scalar product between any two states |1 i , |2 i is invariant
in modulus: |hS 2 |S 1 i| = |h2 |1 i|. A theorem by Wigner, which we are not going
to prove, states that the phases of the vectors representing physical states can always be
chosen so that a transformation preserving transition probabilities be either realized as a
unitary transformation, hS 2 |S 1 i = h2 |1 i, or as an anti-unitary one, hS 2 |S 1 i =
h2 |1 i (transformations involving time-reversal). We shall restrict ourselves to unitary
transformations only. Suppose we have an abstract group G of transformations (like the
rotation group or the Lorentz group), whose action on the states is described by a unitary
representation D, which means that, for any element g of the abstract group G:
|i |g i = D[g] |i ,
3

(2.5.17)

We are always going to describe transformations of quantum states from the active point of view: States
are mapped into different states.

88

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

and, moreover, for any couple of states |1 i , |2 i:


hg 2 |g 1 i = h2 |D[g] D[g] |1 i = h2 |1 i ,

(2.5.18)

which implies that D[g] D[g] = D[g] D[g] = 1, for any g G.


Of course the expectation value of an observable will in general change since, for instance,
the position of a particle depends on the RF in which it is measured:
gG

hOi hOi0 =

D[g] |i
h|D[g] O
.
kk2

(2.5.19)

From the above formula it is clear that we can either think of the transformations as acting

on the states, as in (2.5.17), or on the operator O:


O
0 = D[g] O
D[g] ,
O

(2.5.20)

the result on the expectation value of O is the same. For instance if O is the position vector
r = (xi ) of a particle and the transformation group is SO(3), a generic rotation g SO(3) on
r is represented by the action of the 3 3 real matrix R[g] (R[g]i j ) in (2.3.79): r0 = R[g] r.
This is the transformation property we expect for the expectation value of the position on any
state of the particle. Let D be the unitary representation of SO(3) on the space of quantum
states, and r (
xi ) the hermitian operator associated with r. The expected position hri of
a particle in a state |i is given by eq. (2.5.14) and should transform under g SO(3) as in
(2.5.19):
gG

hxi i hxi i0 =

h|D[g] xi D[g] |i
= R[g]i j hxj i ,
kk2

(2.5.21)

which implies that the operator r should transform as follows:


xi x0 i = D[g] xi D[g] = R[g]i j xj .

(2.5.22)

Let us emphasize that D and R are different representations of the same rotation group:
The former describes the effect of rotations on the vectors representing physical states, the
latter describes rotations on three dimensional space vectors.
We may have transformations on H which are not the effect of any change in the spacetime RF. These are called internal transformations and are still represented by unitary
operators on physical states. Internal transformations have an important role in the theory
of elementary particles and of fundamental interactions.
Consider a group G of transformations which do not involve time (this excludes for
instance Lorentz transformations other than rotations, i.e. Lorentz boosts), like rotations, or
the permutation of particles in a system of particles, or, in general, internal transformations.
If these transformations represent symmetries of the system, the energy of the system should
not be affected. The energy is described by the Hamiltonian H, which corresponds, in

2.6. SCHURS LEMMA

89

If G is a symmetry group then, in virtue


quantum mechanics, to a hermitian operator H.
of eq. (2.5.20), we should have that, for any g G:
D[g] H
D[g] = H
,
H

(2.5.23)

namely, using the unitarity property of D[g], the Hamiltonian operator should commute with
D[g]] = 0 for any g G. Suppose the system is in a state |E i of definite energy
D[g]: [H,
to the eigenvalue E:
E. For what we have previously said, |E i is the eigen-state of H
|E i = E |E i .
H

(2.5.24)

There may be more than one eigen-state to the same eigenvalue E. In this case one speaks
of degeneracy of the energy level E. Let HE denote the space spanned by all the states with
corresponding to the eigenvalue E. It
definite energy value E, namely the eigen-space of H
clearly is a subspace of H and its dimension measures the degeneracy of the energy level E.
Let us show that it is a representation space for G, namely that it is invariant with respect
to the action of G. Take a vector |E i and act on it by a generic transformation g G:
|E i |g E i = D[g] |E i. Let us show that the transformed state is still in HE :
|g E i = H
D[g] |E i = D[g] H
|E i = E |g E i .
H

(2.5.25)

The dimension of HE is therefore the dimension of a representation of G, which, if we exclude


the so-called accidental degeneracy, is also irreducible. In these general cases we can then
deduce an important physical property of a system, such as the degeneracy of its energy
levels, merely from group theoretical considerations.

2.6

Schurs Lemma

Consider two representations D1 and D2 of an abstract group G, acting on the linear vector
spaces Vn and Wm , respectively. Consider a homomorphism T Hom(Vn , Wm ), whose action
on the (column) n-vectors in Vn is represented by the m n matrix T (T a i ), as explained
in Section 2.1.1: T (V) = T V. We define an intertwiner between the representations D1
and D2 a homomorphism T Hom(Vn , Wm ) such that, for any g G, the corresponding
matrix T satisfies the following condition:
T D1 [g] = D2 [g] T .

(2.6.1)

These maps are also called G-homomorphisms and their space is denoted by HomG (Vn , Wm ).
Property 2.5 (Schurs Lemma): Let D1 and D2 be two irreducible representations of
G on the linear vector spaces Vn and Wm respectively, and let T be an intertwiner between
them:
i) If D1 and D2 are inequivalent, T 0 ( i.e. T 0);

90

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS


ii) If D1 and D2 are equivalent (so that Vn and Wm are isomorphic, i.e. m = n), then, for
some number c, T = c A, where the n n matrix A defines the equivalence between
the representations: D2 = A1 D1 A.

Let us prove property (i) first. Define E as the subspace of Vn spanned by the vectors which
are mapped by T into the zero-vector (also called the kernel of T or Ker T ):
E {V Vn |T (V) = T V = 0 Wm } .

(2.6.2)

It can be easily shown that E is a G-invariant subspace. Indeed consider V E and a generic
g G. Then also D1 [g] V is in E since T (D1 [g] V) = T D1 [g] V = D2 [g] T V = D2 [g] 0 = 0.
Since, by assumption, D1 is irreducible, either
(a)

E = {0} or

(b) E = Vn .

(2.6.3)

In case of (b), all vectors of Vn are mapped into the zero-vector of Wn , namely T = 0, i.e.
T 0, and (i) is proven. In case of (a), T is one-to-one. Let us show that the space
T (Vn ) Wm is G-invariant:
For any V Vn and g G : D2 [g] T (V) = D2 [g] T V = TD1 [g] V = T (D1 [g] V) T (Vn ) .
Since D2 is irreducible and T (Vn ) can not, by assumption, be {0}and therefore we must
have: T (Vn ) = Wm , namely T is onto. We conclude that T is an isomorphism and thus,
from eq. (2.6.1) we deduce that D1 and D2 are equivalent: D1 = T1 D2 T, contradicting
our assumption. This completes the proof of (i).
Let us prove now proposition (ii). By assumption D1 and D2 are equivalent: D1 =
A1 D2 A, for some non singular matrix A. Using this equivalence, eq. (2.6.1) can be
re-written as follows:
T D1 [g] = A D1 [g] A1 T

(A1 T) D1 [g] = D1 [g] (A1 T) ,

(2.6.4)

which means that the matrix A1 T intertwines D1 with D1 . Let us focus now on this n n
matrix A1 T which maps n-vectors of Vn into n-vectors of Vn . Under a change of basis
of Vn A1 T will transform by a similarity transformation. From general matrix theory, we
know that any matrix can be reduced, through a similarity transformation, into its Jordan
normal form. This normal form is a block diagonal matrix:

J1
0
..
,
J =
(2.6.5)
.
0
J`
where the blocks Jk , k = 1, . . . , `, have the following form:

k 1
0
.

k . .

.
Jk =
..
. 1
0

(2.6.6)

2.6. SCHURS LEMMA

91

From eq. (2.6.6) it follows that




1
1
0
0


Jk
... = k ... ,
0
0

(2.6.7)

from which it that A1 T must have at least one eigenvalue. Let us call it c and the
corresponding eigenvector Vc . Define the homomorphism T 0 End(Vn ) described by the
matrix:
T0 A1 T c 1n .

(2.6.8)

Since both matrices A1 T and c 1n intertwine D1 with D1 , also T0 does. Define E = Ker T 0 .
Since D1 is irreducible, also for E we have the options (2.6.3). Notice however that T 0 can
not be an isomorphism, i.e. E cannot be {0}, since there exist the vector Vc on which:
T 0 (Vc ) = T0 Vc = A1 T Vc c Vc = 0 .

(2.6.9)

We conclude that T 0 = 0, that is A1 T = c 1n . This concludes the proof of proposition (ii).


As a corollary of Schurs lemma, we have
Properties 2.6:
iii Let D be an irreducible n-dimensional representation of a group G. A matrix T which
commutes with all matrices D[g], for any g G, is proportional to the identity matrix
1n ;
iv Let D be a n-dimensional representation of a group G. If there exist a matrix T which
commutes with all matrices D[g], for any g G, and is not proportional to the identity
matrix 1n , then D is reducible;
Proposition (iii) is a particular case of Schurs lemma, in which D1 = D2 and thus A = 1n .
From proposition (ii) it follows that T = c 1n , for some number c.
Schurs lemma and its corollary, provide us with a criterion for telling if a representation
is reducible and, in some cases, to determine its irreducible components: Suppose we find
an operator T on Vn which commutes with all the matrices D[g] representing the action of
a group G on the same space. The matrix representation T of T will then have the form:

c1 1k1
0
..
,
T =
(2.6.10)
.
0
cs 1ks
where c1 , . . . , cs are different numbers and the corresponding eigen-spaces Vk1 , . . . , Vks of
T correspond to different representations Dk1 , . . . , Dks of G. If G is a symmetry group
of a quantum mechanical system, we know, from our discussion in Section 2.5.1, that the

92

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

commutes with the action D[G] of G on the states, i.e. it intertwines


hamiltonian operator H
D with itself. Its matrix representation on the states will then have the form (2.6.10),
where c1 , . . . , cs (s may be infinite!) are the energy levels of the system, and k1 , . . . ks their
degeneracy. If there is no accidental degeneracy, Dk1 , . . . , Dks are irreducible representations
of G. If Dki is not irreducible, this may indicate that there exist a larger symmetry group
G0 containing G, whose action on Vki is irreducible. On other words accidental degeneracies
may signal the existence of a larger symmetry of the system.
Property 2.7: All irreducible representations of an abelian group are one-dimensional.
Let G be an abelian group and D an irreducible n-dimensional representation of G.
Consider a generic element g G. The abelian property of G implies that, for any g 0 G,
D[g] D[g 0 ] = D[g 0 ] D[g], that is D[g] intertwines the representation D with itself. In virtue
of Schurs lemma, being D irreducible, we must have D[g] = d[g] 1n , for some number
d[g] associated with g. In an irreducible representation the matrix elements can all be
proportional to the identity matrix only if the representation is one-dimensional, namely if
n = 1. In Section 2.8 we shall give, for finite groups, an alternative proof of this property.

2.7

Great Orthogonality Theorem

We prove in this section an important property of unitary representations of finite groups.


Property 2.8 (Great Orthogonality Theorem): Let G be a finite group and D, D0
two irreducible unitary representations of dimensions n and m respectively. With respect to
some bases of the corresponding representation spaces Vn and Wm , each element g G is
then represented by the matrices D[g] (D[g]i j ) and D0 [g] (D0 [g]a b ), i, j = 1, . . . , n and
a, b = 1, . . . , m. The following properties hold:
i) If D and D0 are inequivalent,
X

D[g]i j

D0 [g]a b = 0 ,

(2.7.1)

gG

for any i, j = 1, . . . , n and a, b = 1, . . . , m;


ii) If D = D0 ,
X

D[g]i j

D0 [g]k ` =

gG

|G| k j
,
n i `

(2.7.2)

for any i, j, k, ` = 1, . . . , n, |G| denoting the order of G.


Let us start proving proposition (i). Under the hypotheses of the theorem, we take a generic
a rectangular m n matrix X (X a i ) and define the following matrix M (M a i ):
def. X 0
M =
D [g] X D[g]1 ,
(2.7.3)
gG

2.7. GREAT ORTHOGONALITY THEOREM

93

or, in components,
def. X 0 a b
=
D [g] b X j D[g]1 j i .

M ai

(2.7.4)

gG

Let us show that M is an intertwiner between D0 and D. Take a generic h G:


X
X
D0 [h] M =
(D0 [h] D0 [g]) X D[g]1 =
D0 [h g] X D[g]1 (D[h]1 D[h]) =
gG

gG

D0 [h g] X D[h g]1 D[h] =

gG

D0 [g] X D[g]1 D[h] = M D[h](2.7.5)


.

gG

In virtue of Schurs lemma, if D and D0 are inequivalent, M = 0. Let us construct the set of
n m matrices Mi a [(Mi a )b j ], where a, i label the matrix, while b, j are the matrix indices,
out of the following matrices Xi a [(Xi a )b j ] = [ji ab ]. Each of these matrices is zero:
X
Mi a =
D0 [g] Xi a D[g]1 = 0m,n ,
(2.7.6)
gG

or, writing the (b, j)-component of the above matrix:


X
X
X

0 = (Mi a )b j =
D0 [g]b c ki ac D[g]1 k j =
D0 [g]b a D[g]1 i j =
D[g]j i D0 [g]b a ,
gG

gG

gG

where in writing last equality we have used unitarity or the representation D[g]1 = D[g] .
Let us now prove proposition (ii). By assumption D = D0 . Schurs theorem implies that
M = c 1n :
X
D[g] X D[g]1 = c 1n .
(2.7.7)
M =
gG

Taking the trace of the left hand side of the above equation we find:
X
 X
 X
Tr (M) =
Tr D[g] X D[g]1 =
Tr D[g]1 D[g] X =
Tr (X) = |G| Tr (X) .
gG

gG

gG

Since the trace of the right hand side is c Tr (1n ) = n c, we find:


c =

|G|
Tr (X) .
n

(2.7.8)

Now insert this result in (2.7.7), written in components:


def. P
0
i
k
1 `
M ij =
j =
gG D [g] k X ` D[g]

|G|
n

Tr (X) i j =

|G|
n

X k ` `k ij .

(2.7.9)

The above equation can be rewritten in the following form:


X

X
`
gG

D [g] k D[g]

1 `

|G| ` i

k j
n

!
= 0.

(2.7.10)

94

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Since (X k ` ) is an arbitrary matrix, we conclude that the expression in brackets has to be


zero for any k, `, namely that:
X

D0 [g]i k D[g]j `

gG

|G| ` i
k j,
n

(2.7.11)

where we have again used unitarity of D.


Let us denote by D(s) , s = 1, 2, . . ., all unitary irreducible representations of G, so that
different values of s correspond to inequivalent representations. Let ns denote the dimension
of D(s) . We can express both properties (i) and (ii), namely the whole great orthogonality
theorem, by the following single compact formula:
 0
 |G|
0
` k i j ss ,
D(s) [g]i k D(s ) [g]j ` =
ns
gG

(2.7.12)

If we represent the set of three indices s, i, j of D(s) [g]i k by a single index I, so that D[g](I)
D(s) [g]i k , equation (2.7.12) is rewritten in the following form:
X

|G| II 0
,
nk

D[g](I) D[g](I ) =

gG

(2.7.13)

which expresses the orthogonality of a set of vectors D[g](I) , labeled by the index I. These
are vectors in the finite dimensional linear vector space HG of complex valued functions on
the group G. This space has dimension given by the order |G| of the finite group, since any
function f over G is completely defined by specifying its values on each of the |G| elements
of the group, namely by specifying its |G| components f (g). Given two functions f, f 0 on G
the following scalar product is defined:
(f, f 0 )

def. X
=
f (g) f 0 (g) .

(2.7.14)

gG

With respect to this product eq. (2.7.13) expresses the ortho-normality of the set of functions
D[g](I) : The great orthogonality theorem states the mutual orthogonality of the matrix
elements of unitary irreducible representations of G vectors in HG . (D[g](I) ) provide a basis
for the space HG , so that any function f HG can be written as follows:
f (g) =

X
I

(I)

cI D[g]

ns
XX
s

csi j D(s) [g]i j .

(2.7.15)

i,j=1

Remark: A great orthogonality theorem can be stated also for compact Lie groups, like
O(n) or U(n). In this case, however, the group is not finite set of elements, since its elements are function of continuous parameters. Clearly we need to replace the sum over the

2.8. CHARACTERS

95

group elements
over the group by suitably defining a (G-invariant) integration
R
P by an integral
measure dg: gG G dg. The great orthogonality theorem then states that:
Z

 0
 Vol(G)
0
` k i j ss .
(2.7.16)
dgD(s) [g]i k D(s ) [g]j ` =
ns
G
R
where the volume of G is defined as Vol(G) = G dg. Again
define a space of
R we can
0
0
complex valued functions on G, with inner product (f, f ) G dg f (g) f (g), with respect
to which the matrix elements of unitary irreducible representations form a basis. We shall not
elaborate further on this case. We just mention, as an example, the case of the group U(1),
whose elements are functions g() of a continuous angular parameter (0, 2 ). For any
integer s we have defined the representation D(s) [g()] = ei s . We can take
R as integration
R 2
measure on the group d, so that the volume of the group is Vol(U(1)) = G dg = 0 d =
2 . Then eq. (2.7.16) is easily verified in this simple case, since the indices i, j, k, ` all run
over a single value (ns 1 for any representation s):
Z

(s)

dgD [g] D

(s0 )

[g] =

dei (ss ) = 2 ss =

Vol(U(1)) ss0
.
ns

(2.7.17)

Clearly complex valued functions over U(1) are functions f () of (0, 2), i.e. functions
over the circle. They form a linear vector space HU(1) , which is infinite dimensional. From
functional analysis we know that ei s , s = 0, 1, 2, . . ., form an ortho-normal basis of
functions for this space. Any function f () can be expanded (the Fourier expansion) in this
basis. Are there other irreducible U(1)-representations? If there existed an other irreducible
U(1)-representation D, it should still be one dimensional D[] = (D[]) and should be,
according to our analysis, orthogonal to all the ei s , which can not be the case since we can
Fourier expand D[] in the basis ei s .

2.8

Characters

Given a finite dimensional representation D of a group G we define the character (g) of an


element g G in the representation D the quantity:
(g)

def.
= Tr(D[g]) = D[g]i i ,

(2.8.1)

where summation over i is understood. The set of all characters {(g)|g G} is called the
set of characters of D. From the cyclic property of the trace we have that Tr(A1 D[g] A) =
Tr(D[g]), which implies that:
a) Equivalent representations have the same characters;
b) All elements of a same conjugation class have the same characters.

96

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

If D(s) are unitary irreducible representations of a finite group G (whatever we say for a
finite group can be extended to compact Lie groups), from eq. (2.7.12), setting i = k and
j = ` and summing over i and j, we find
X
0
0
0
((s) , (s ) ) =
(s) (g) (s ) (g) = |G| ss .
(2.8.2)
gG

That is the characters (s) (g), as complex valued functions over G, and thus elements of
HG , are mutually orthogonal. But characters have the same value over the elements of a
same conjugation class. If we denote by C0 , C1 , C2 , . . . , Cp1 the conjugation classes of G,
containing d0 , d1 , . . . , dp1 elements respectively (we use the convention to identify C0 with
the class [e], so that d0 = 1), and by (s) (Ci ) the value of (s) on the elements in Ci , the
orthogonality relation (2.8.2) can be rewritten as follows:
p1
X

di (s) (Ci ) (s ) (Ci ) = |G| ss .

(2.8.3)

i=0

Let the number of unitary irreducible representations of G be denoted by pr , so that the


index s runs from 0 to pr 1, 0 denoting the identity representation. The characters should
then be thought of as pr complex valued functions over the set of p conjugation classes of G.
These functions span a p-dimensional complex vector space, to be denoted by HC . On this
space we can define a hermitian scalar product, such that, for any , HC :
X

(, )

(g) (g) =

p1
X

di (Ci ) (Ci ) .

(2.8.4)

i=0

gG

Eq. (2.8.3) can then be viewed as an orthogonality relation over this space. Since in a p
dimensional linear vector space there can be at most p mutually orthogonal (and thus linearly
independent) vectors, there can be no more than p independent functions (s) . Recall that s
labels the unitary irreducible representations of the group. We conclude that, for any finite
group G:
number of unitary irreducible representations number p of conjugation classes ,
that is pr p. We shall prove below a general property which states that the above two
numbers, for finite groups, are actually equal.
On the trivial (identity) representation, which we label with s = 0, (0) (g) = 1 for any
g G. Applying eq. (2.8.2) to the representations s 6= 0 and s0 = 0, we find:
X

(s)

(g) =

gG

p1
X

di (s) (Ci ) = 0 ,

(2.8.5)

i=0

that is the sum, over the group elements, of all characters of a given representation is zero.
Writing (2.8.3) for s0 = s we find:
p1
X
i=0

di (s) (Ci ) (s) (Ci ) = |G| .

(2.8.6)

2.8. CHARACTERS

97

The above property holds only for irreducible representations and thus gives a criterion for
the reducibility of a representation.
Suppose we have a reducible n-dimensional representation D of G which decomposes
into irreducible ones so that the representation D(s) occurs a number as of times in the
decomposition:
as times
}|
{
Mz
D(s) . . . D(s) .
D
pr 1

(2.8.7)

s=0

Ppr 1

We have s=0 ns as = n, where ns is the dimension of D(s) . The trace of the n n matrix
D[g], for any g G, is the sum of the traces of each of its blocks, which consist in as blocks
D(s) [g], for all irreducible representations D(s) . We then find that, for any conjugation class
Ci :
pr 1

(g) =

as (s) (g) .

(2.8.8)

s=0

As anticipated, let us state and prove the following property:


Property 2.9:
A finite group G has as many inequivalent irreducible unitary representations as the number of conjugation classes.
To prove this, let us show that (s) (g) is a basis for HC , namely that any other element
(g) of HC , which is linearly independent of the (s) or, equivalently, which is orthogonal
to all the (s) , must be zero4 . Consider (g) an element of HC . This means that is a
complex valued function on G and moreover, for any g, h G, (h1 g h) = (g), namely
the value of is a function of the conjugation classes only. Suppose that is orthogonal to
all the (s) , that is, for any irreducible representation s:
X
(g) (s) (g) = 0 .
(2.8.9)
gG

This implies that is orthogonal to the character of any representation of G, since the
scalar product is bi-linear and can be expanded as a linear combination of the characters
(s) . Let us show that, given a representation D of G over a vector space Vn , using we can
construct an intertwiner S,D between D and itself. Indeed let S,D be a linear operator on
Vn whose action on a generic vector V is represented as follows:
X
S,D (V) = S,D V
(g) D[g] V .
(2.8.10)
gG
4

Let V and W be two elements of a complex linear vector space Vn on which a hermitian scalar product
W (V,W)
(, ) is defined. If the two vectors are linearly independent the vector W
kVk2 V is non zero. The
V) = 0. Conversely the reader can easily show that if V and W are orthogonal,
reader can show that (W,
they are linearly independent.

98

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Let us show that, for any g G, S,D D[g] = D[g] S,D . For any V Vn and h G:
X
X
S,D D[h] V =
(g) D[g] D[h] V =
(g) D[g h] V =
gG

gG

(h g 0 h1 ) D[(h g 0 h1 ) h] V =

(g 0 ) D[h g 0 ] V =

g 0 G

hg 0 h1 G

= D[h] S,D V .

(2.8.11)

From the arbitrariness of V we conclude that S,D is a matrix which intertwines the representation D with itself. If D is one of the irreducible representations D(s) of G, by Schurs
lemma, S,D(s) = s 1, where 1 is the identity matrix over the representation space of D(s) ,
whose dimension we denote by ns . We can express s as follows:
1 X
1
Tr(S,D(s) ) =
(g) (s) (g) .
(2.8.12)
s =
ns
ns gG
Using the orthogonality assumption (2.8.9) we deduce that s = 0, namely that, for any
irreducible representation S,D(s) = 0. Notice that a generic representation D is completely
reducible in combinations of D(s) and thus the matrix S,D will in general have, in a suitable
basis, a block diagonal form, each block being equal to S,D(s) . It then follows that, for any
representation D, S,D = 0, that is:
X
(g) D[g] = 0 .
(2.8.13)
gG

Consider then the regular representation DR . The |G| |G| matrices DR [g], for different
g in G, have non vanishing entries in different positions (i, j) and thus are |G| linearly
independent matrices. Equation (2.8.13), applied to the regular representation, then implies
that, for any g G, (g) = 0, that is is the zero-vector in HG , and thus that (s) form a
basis of the same space. This shows that the number of irreducible representations (range
of s) actually equals the number p of conjugation classes.
As an example of the above property let us consider the group D4 . In Section 1.4.2, we
have found for D4 five conjugation classes: [e], [r` ] = {r, r3 }, [r2 ] = {r2 }, [] = {, }, [] =
{, }. In Section 2.4.1, we have found for D4 precisely 5 irreducible representations: E,
A1 , A2 , B1 , B2 . As we shall show below these are all the inequivalent irreducible unitary
representations of D4 .
Because of the above property, we will take the label s to run from 0 (corresponding to
the trivial representation) to p 1.
Consider the decomposition (2.8.7) of a generic representation into the pr = p unitary
irreducible representations. Using the orthogonality relation (2.8.2) we can single out the
multiplicities as in (2.8.7) by multiplying both sides times (s) (g) and summing over g G.
In formulas:
p1
p1
X
1 XX
1
1 X
0
(s)

(s0 )
(s)

(g) (g) =
as0 (g) (g) =
|G|
as0 ss = as .(2.8.14)
|G| gG
|G| s0 =0 gG
|G|
s0 =0

2.8. CHARACTERS

99

Let us now compute the quantity (g) (g), for a given g G, using eq. (2.8.8):
p1
X

(g) (g) =

as as0 (s) (g) (s ) (g) .

(2.8.15)

s,s0 =0

If we sum over g G and use the orthogonality property (2.8.2), we find:


X

(g) (g) =

p1
X

as as0

s,s0 =0

gG

(s)

(s0 )

(g)

(g) = |G|

p1
X

as as 0

s,s0 =0

gG

ss0

= |G|

p1
X

(as )2 ,

s=0

that is:
p1
X

di (Ci ) (Ci ) = |G|

p1
X

(as )2 .

(2.8.16)

s=0

i=0

Comparing eqs. (2.8.6) to (2.8.16) we can state the following


Property 2.10: The sum of the squared moduli |(Ci )|2 of the characters over the conjugation classes, weighted by the order di of each class, equals the order |G| of the group only
if the representation is irreducible, otherwise it is an integer multiple of it.
Let us apply this property to the 3-dimensional representation D of D4 given in (2.4.35).
Suppose we want to tell wether this representation is reducible or not. Let us compute the
set of characters for each group element:
{(e), (r), (r2 ), (r3 ), (), (), (), ()} = {3, 1, 1, 1, 1, 1, 1, 1} .
or, equivalently, for each class:
([e]) = 3 , ([r]) = 1 , ([r2 ]) = 1 , ([]) = 1 , ([]) = 1 .

(2.8.17)

Compute the sum of the squares of (g), over the group elements:
X
(g)2 = 32 + 12 + (1)2 + 12 + 12 + 12 + 12 + 12 = 16 = 2 8 =
gG

= 2 |D4 | > |D4 | .

(2.8.18)

According to our criterion the three dimensional representation of D4 is therefore reducible.


This was indeed shown by inspection of the explicit form of the matrices in this representation. We have seen that:
D E A1 .

(2.8.19)

Since P
the two representations E and A1 occur in the above decomposition just once, we
have 1s=0 a2s = 1 + 1 = 2 and this explains the factor 2 on the right hand side of (2.8.18),
according to the general property (2.8.16).

100

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Let us now consider the regular representation, whose dimension equals the order |G|
of the group. Since the only matrix DR [g] which has diagonal entries, and therefore a non
vanishing trace, is DR [e] = 1|G| , the only non vanishing character of this representation is
R (e) = |G|.
Exercise 2.7: Apply the criterion expressed by Property 2.10 and prove that the regular
representation is reducible.
Applying (2.8.14) to the regular representation, the only contribution to the sum in the
left hand side, comes from g = e. Since (s) (e) = ns we find:
as =

1
R (e) (s) (e) = ns .
|G|

(2.8.20)

This is telling us that each irreducible representation enters the decomposition


of the regular
Pp1
ns as = |G|
representation a number of times precisely equal to its dimension. From s=0
we finally find (Burnside Theorem):
p1
X

n2s = |G| ,

(2.8.21)

s=0

that is the sum of the squared dimensions of each irreducible representation of a finite group
equals the order of the group itself. In particular, for abelian groups we have as many
classes as elements, p = |G|, which implies that, for any s, ns = 1, that is all irreducible
representations are one-dimensional.
Let us come back to the D4 example. In this case the order of the group is |D4 | = 8.
If we sum the square of the dimensions of the irreducible representations E, A1 , A2 , B1 , B2
that we found, recalling that E is two -dimensional and all the other are one-dimensional,
we obtain:
4
X

n2s = 1 + 1 + 1 + 1 + 22 = 8 = |D4 | .

(2.8.22)

s=0

The above sum already saturates the order of the group. According to our analysis, there
can not be other irreducible representations for D4 , than those listed above.
Character Table. Characters are very useful since they allow to decompose representations of a finite group into their unitary irreducible components according to (2.8.7). To
this end we need to classify the set of characters {(s) (Ci )}i=1,...,p for each irreducible representation D(s) . The set of characters {(Ci )}i=0,...,p1 of a given representation D then
decompose in a unique way into the characters of its irreducible components, according to
eq. (2.8.8), allowing us to determine the multiplicities as , for any s, and thus the whole decomposition (2.8.7). The characters for each irreducible representation are usually collected
in the character table, which has the following form:

2.9. OPERATIONS WITH REPRESENTATIONS


G
D(0)
D(1)
D(2)
..
.

[e]
1
(1) ([e]) = n1
(2) ([e]) = n2
..
.

C1
1
(1) (C1 )
2) (C1 )
..
.

101
C2
1
(1) (C2 )
(2) (C2 )
..
.

For instance let us use the explicit form of the D4 irreducible representations (2.4.37),
(2.4.38), (2.4.39), (2.4.40), to construct its character table:
D4
A1
A2
B1
B2
E

[e] {r, r3 }
1
1
1
1
1
1
1
1
2
0

{r2 }
1
1
1
1
-2

{, }
1
1
1
1
0

{, }
1
1
1
1
0

The reader can verify, with reference to the above table, the orthogonality property (2.8.2).
Using this character table we can easily work out the decomposition of the regular representation of D4 , whose character set is:
R (e) = |D4 | = 8 , R (g 6= 2) = 0 .

(2.8.23)

The only combination of the characters in Table 2.8 which reproduce (2.8.23), according to
the general formula (2.8.8) correspond to the following decomposition:
DR 2 E A1 A2 B1 B2 ,

(2.8.24)

that is, as expected, all the irreducible representations occur in the decomposition of the
regular one with a multiplicity given by their dimension (for instance E is two-dimensional
and occurs twice in (2.8.24)).

2.9

Operations with Representations

Let us define some operations which allow to construct new representations of a given group
from given ones.

2.9.1

Direct Sum of Representations

Consider two representations D1 and D2 of a group G over the spaces Vn and Wm respectively.
If we denote by (ei ) and (fa ), i = 1, . . . , n and a = 1, . . . , m, bases of Vn and Wm , respectively
recall that the direct sum Vn Wm is defined as the (n+m)-dimensional vector space spanned
by the basis (e0I ) (e1 , . . . , en , f1 , . . . , fm ), I = 1, . . . , n + m. A vector V0 in Vn Wm is then

102

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

represented by the following column vector:

V 01
...

 i 
V 0n
V
0I 0

= V eI =
V 0 n+1 = W a ,
.
..
V 0 n+m

V0

(2.9.1)

where V = V i ei and W = W a fa are the components of V0 along the subspaces Vn and Wm


respectively. Under the action of a group element g G, the component vectors transform
as follows:
V D1 [g] V ,

W D2 [g] W ,

(2.9.2)

which implies that the whole vector V0 transforms as follows:


V

D1 [g]i j V j
D2 [g]a b W b


=

D1 [g]
0
0
D2 [g]

V0 = D[g] V0 .

(2.9.3)

It is straightforward to show that the correspondence D:


D

g G


D[g]

D1 [g]
0
0
D2 [g]


,

(2.9.4)

is a representation of G acting on the space Vn Wm , that is D[g] = (D[g]I J ) is a (n +


m) (n + m) matrix, called direct sum of the representations D1 D2 , and is denoted by the
symbol:
D = D1 D2 .

(2.9.5)

We may generalize the above construction to define the direct sum D = D1 D2 . . . Dk


of the representations D1 , D2 , . . . Dk of G acting on the vector spaces Vn1 , Vn2 , . . . , Vnk as
P
the representation acting on the n-dimensional space Vn1 Vn2 . . . Vnk , n = k`=i n` by
means of the matrices

D1 [g]
0

0
0
D2 [g]
0
D
g G D[g]
.
(2.9.6)
.
..
.
..
..
.
0

Dk [g]

By construction D is reducible and


P its character (g) is the sum of the characters i (g) of
each representation Di : (g) = ki=1 i (g)

2.9. OPERATIONS WITH REPRESENTATIONS

2.9.2

103

Direct Product of Representations

Consider two representations D1 and D2 of a group G over the spaces Vn and Wm respectively.
As usual let us denote by (ei ) and (fa ), i = 1, . . . , n and a = 1, . . . , m, bases of Vn and Wm ,
respectively. We define the direct product Vn Wm a (n m)-dimensional vector space spanned
by the basis (ei fa ), where the operation is defined on Vn Wn and is bilinear in its two
arguments:
(a V1 + b V2 ) W = a (V1 W) + b (V2 W) ,
V (a W1 + b W2 ) = a (V W1 ) + b (V W2 ) ,

(2.9.7)

for any a, b numbers, Vi , V Vn and Wi , W Wm . Using this property we can write


the direct product of two vectors V = V i ei Vn and W = W a fa Wm as the vector
VW =
whose components are the product of the components of the two vectors: V
i
a

V W ei fa . We can represent V = V W only by means of its components V i W a , which


can either be arranges in a n m matrix, or in a n m vector whose components are labeled
by the couple of indices I = (i, a): (V I ) = (V 1 W 1 , V 2 W 1 , . . . V n1 W m , V n W m ). We will
write V W = (V I ) (V i W a ). Under the action of an element g G, this product will
transform as follows:
V W (D1 D2 )[g] (V W) (D1 [g] V) (D2 [g] W) ,

(2.9.8)

or, in components:
V W = (V I ) (V i W a ) (V 0 I ) (V i 0 W a 0 ) = (D1 [g]i j V j D2 [g]a b W b ) =
= (D1 [g]i j D2 [g]a b V j , W b ) =
= ((D1 D2 )[g]ia jb V j W b ) =
= ((D1 D2 )[g]I J V J ) =
= (D1 D2 )[g] (V W) ,

(2.9.9)

The n m n m matrix (D1 D2 )[g] = ((D1 D2 )[g]ia jb ) (D1 [g]i j D2 [g]a b ), for any g G,
defines a representation of G called direct product (or Kronecker product ) of the representations D1 and D2 and denoted by the symbol D1 D2 . A generic element of Vn Wm can not
in general be written as the product of two vectors and has the form F = F ia ei fa (F I ).
It will transform, under the action of a g G, as the product of two vectors, according to
the product representation:
F F0 = (D1 D2 )[g] F ,

(2.9.10)

which means that each if its indices will transform under g according to the corresponding
representation:
F ia F ia 0 = (D1 D2 )[g]ia jb F jb = D1 [g]i j D2 [g]a b F jb .

(2.9.11)

104

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

The character (D1 D2 ) (g) of a product of two representations of a group g is obtained by


tracing the matrix ((D1 D2 )[g]ia jb ):

(D1 D2 )

(g) =

n X
m
X
i=1 a=1

ia

(D1 D2 )[g]

ia

n
m
X
X
i
=(
D1 [g] i )(
D2 [g]a a ) = (D1 ) (g) (D2 ) (g) .
i=1

a=1

As for the direct sum of representations, we can extend the notion of direct product to
an indefinite number of representations D1 , D2 , . . . Dk ofQG acting on the vector spaces
Vn1 , Vn2 , . . . , Vnk as the representation acting on the (n = k`=1 n` )-dimensional space Vn1
Vn2 . . . Vnk . A generic element of this product space F (F a1 a2 ... ak ) has components
labeled by k plets of indices (a1 a2 . . . ak ), a1 = 1, . . . , n1 , a2 = 1, . . . , n2 and so on, and
transforms as the product of components of vectors in each factor space V a1 V a2 . . . V ak :
F a1 a2 ... ak

F a1 a2 ... ak 0 = (D1 D2 Dk )[g]a1 a2 ... ak b1 b2 ... bk F b1 b2 ... bk


D1 [g]a1 b1 D2 [g]a2 b2 Dk [g]ak bk F b1 b2 ... bk .
(2.9.12)

The above transformation property defines the product representation D1 D2 Dk :


F F0 = (D1 D2 . . . Dk )[g] F ,

(2.9.13)

In general the product of two or more representations of a same group is reducible, as we


shall show in explicit examples.
So far we have considered objects with two or more indices, like F a1 a2 ... ak , which transform
as product of contravariant components of vectors (i.e. all the indices are upper indices).
We can generalize our analysis to objects of the form F (F a1 a2 ... ap ap+1 ap+2 ... ak ) whose
components have p-upper and (q = k p)-lower indices and transform as products of pcontravariant and q-covariant components V a1 V ap Wap+1 Wak . Take for instance our
original example of two representations D1 and D2 acting on two vector spaces Vn and
Wm . We have defined a quantity F = (F ia ), whose components transform as products
V i W a of contravariant components V i and W a . We could have defined a quantity F =
(F i a ) transforming as the product V i Wa of contravariant components V i and covariant
components Wa . Recall that, if W a transform in the representation D2 , the corresponding
covariant components Wa transform in the representation D02 whose matrices are related to
those of D2 as follows: D02 [g] (D20 [g]a b ) = D2 [g]T
Wa D20 [g]a b Wb = D2 [g]1 b a Wb .

(2.9.14)

The quantity F = (F i a ), under the action of a g G, will then transform as follows


F i a F 0 i a = D1 [g]i j D2 [g]1 b a F j b = (D1 D02 )[g]i a b j F j b ,

(2.9.15)

or, in compact form:


F F0 = (D1 D02 )[g] F .

(2.9.16)

2.9. OPERATIONS WITH REPRESENTATIONS

105

The above transformation law defines the representation D1 D02 : (D1 D02 )[g]i a b j
D1 [g]i j D2 [g]1 b a . The reader can verify that, if F = (F i a ) and G = (Gi a ) are two objects
transforming according to (2.9.15), any linear combination of them F + G ( F i a +
Gi a ) will transform in the same way. Just like objects with two upper indices, also objects
with one upper and one lower indices span a linear vector space, which is the representation
space of D1 D02 .
Consider now the more general situation, mentioned above, of objects of the form F
a1 a2 ... ap
(F
ap+1 ap+2 ... ak ). Suppose the contravariant indices a1 a2 . . . ap transform with the representations D1 , . . . , Dp respectively and the covariant ones ap+1 ap+2 . . . ak with the representations D0ap+1 , . . . , D0k , where a1 , . . . , ak are indices of different vector spaces of dimensions
n1 , . . . , nk . The quantities F will transform in the representation D1 Dp D0ap+1 D0ak :
F F0 = (D1 Dp D0p+1 D0k )[g] F ,

(2.9.17)

for any g G, where the above action is defined as follows:


F a1 ... ap ap+1 ... ak

F 0 a1 ... ap ap+1 ... ak =


= D1 [g]a1 b1 Dp [g]ap bp Dp+1 [g]1 bp+1 ap+1 Dk [g]1 bk ak F b1 ... bp bp+1 ... bk .
(2.9.18)

Just as for quantities with one covariant and one contravariant indices, quantities with qcovariant and p-contravariant indices form a linear vector space which is the representation
space of D1 Dp D0p+1 D0k .
The irreducible identity representation 1 is one-dimensional and all group elements are
represented by the number 1. The direct product any representation times 1 is therefore the
representation itself:
D 1 = 1 D = D.

2.9.3

(2.9.19)

Tensors

Of particular interest is the case in which all the indices of the quantity F (F a1 ... ap b1 ... bq )
refer to the same linear vector space Vn , so that a1 , . . . , ap , b1 , . . . , bq = 1, . . . , n. This
corresponds, in our previous analysis, to the case in which all the representations coincide
with a same one acting on Vn : D1 = D2 = Dk = D. The quantity F (F a1 ... ap b1 ... bq ),
transforming in the representation:
p

Dp D0 p

}|
{
z
}|
{ z
D D D0 D0 ,

(2.9.20)

is called a tensor of the representation D of G or simply a tensor (we have denoted by Dp


the p-fold product of the representation D). A tensor of this kind, with p-contravariant and
q-covariant indices, is called type (p, q) tensor, while the total number of indices k = p + q

106

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

defines the rank of the tensor. Under the action of a group element g G the tensor F
therefore transforms as follows:
F a1 ... ap ap+1 ... ak

F 0 a1 ... ap ap+1 ... ak =


= D[g]a1 b1 D[g]ap bp D[g]1 bp+1 ap+1 D[g]1 bk ak F b1 ... bp bp+1 ... bk .
(2.9.21)

All type (p, q)-tensors span a linear vector space, which we denote by V (p,q) , which is the
representation space of (2.9.20). Since tensors have been defined in relation to their transformation property (2.9.21) with respect to a certain group G, we talk about G-tensors.
We have already come across tensors. Consider the metric g = (gij ) defining a symmetric inner product on a real linear vector space. Under a generic linear transformation g in
G = GL(n, R), the metric transforms as in (2.3.44):
gij gij0 = D[g]1 k i D[g]1 ` j gk` .

(2.9.22)

Comparing (2.9.22) with (2.9.21) we conclude that g = (gij ) transforms as a type (0, 2)
tensor. The Kronecker delta ij is an example of type-(1, 1) invariant GL(n, C)-tensor.
Indeed if we transform it according to (2.9.21) we find:
ij i0 j = D[g]1 k i D[g]j ` k` = ij .

(2.9.23)

We can define the inverse metric g1 = (g ij ): g ij gjk = ki . This defining condition of the
inverse metric is invariant under a generic change of the RF, implemented by a generic
element of G = GL(n, R), provided g ij transforms as a type-(2, 0) tensor
0
g 0 ij gjk
= D[g]i s D[g]j t D[g]1 ` j D[g]1 r k g st g`r = D[g]i s D[g]1 r k t` g st g`r =

= D[g]i s D[g]1 r k rs = ji ,

(2.9.24)

where we have used the property D[g]j t D[g]1 ` j = t` . Other examples are the contravariant
components of vectors (V i ) or the covariant ones (Vi ), which are type (1, 0) and type (0, 1)
tensor respectively. As opposed to the Kronecker delta, the metric is not an invariant tensor
with respect to a generic linear transformation, as we have seen, but it is invariant only
with respect to transformations in G = O(r, s) (if the metric has r positive and s negative
eigenvalues). In this case we can go to an ortho-normal basis in which g = (r,s) and say that
(r,s) is a type (0, 2) invariant O(r, s)-tensor. We can define the following type-(0, n) tensor:

0 two or more indices have the same value


1 (i1 i2 . . . in ) even permutation of (1, 2, . . . , n) .
i1 i2 ...in =
(2.9.25)

1 (i1 i2 . . . in ) odd permutation of (1, 2, . . . , n)


This definition makes sense only if we restrict to a transformation group which leaves the tensor invariant, otherwise it should refer to a specific RF. Such group is the group G = SL(n, R)
of volume preserving transformations (i.e. of n n real matrices with unit determinant).
Indeed if we transform (2.9.25) under a generic g G we find:
i1 i2 ...in D[g]1j1 i1 D[g]1jn in j1 j2 ...jn = det(D[g]1 ) i1 i2 ...in = i1 i2 ...in , (2.9.26)

2.9. OPERATIONS WITH REPRESENTATIONS

107

where we have used the property that det(D[g]1 ) = 1. The tensor i1 i2 ...in is a type-(0, n)
invariant SL(n, R)-tensor. Similarly we can define a type-(n, 0) invariant SL(n, R)-tensor
i1 ...in whose components are defined by the same rules (2.9.25).
Example:
The defining representation of the (proper) Lorentz group SO(1, 3) has two
invariant tensors: A type-(0, 2) (1,3) = ( ) = diag(1, 1, 1, 1), which is the Minkowski
metric, and the type-(0, 4) tensor  .
Exercise: Find a type (2, 0) invariant Sp(2n, R) tensor.
Given a type-(p, q) tensor F a1 ...ap b1 ...bq and a type-(r, s) tensor Ga1 ...ar b1 ...bs we can construct a type-(p + r, q + s) tensor T as the product of the two:
T a1 ...ap+r b1 ...bq+s F a1 ...ap b1 ...bq Gap+1 ...ap+r bq+1 ...bq+s .

(2.9.27)

The reader can show that the quantity defined above transforms indeed as a type-(p+r, q +s)
tensor. We have seen examples of this general property when we originally constructed type(2, 0) tensors (V i V j ) as products of two contravariant components of vectors, namely of two
type-(1, 0) tensors (V i ). Since product of tensors are still tensors, we say that the space of
all tensors, of whatever rank and type, close an algebra named tensor algebra.
Let us define a further operation on tensors. Given a type-(p, q) tensor F a1 ...ap b1 ...bq , we
can consider only the components in which one upper and one lower
(say ap and bq )
Pn index
a1 ...ap1 a
have the same value and sum over them, namely we compute
b1 ...bq1 a =
a=1 F
F a1 ...ap1 a b1 ...bq1 a (summation over a in the right hand side is understood). The result is a
type-(p 1, q 1) tensor G:
Ga1 ...ap1 b1 ...bq1 F a1 ...ap1 a b1 ...bq1 a ,

(2.9.28)

Let us show that the above object is actually a type-(p 1, q 1) tensor:


Ga1 ...ap1 b1 ...bq1

D[g]a1 c1 D[g]a cp D[g]1 d1 b1 D[g]1 dq a F c1 ... cp d1 ... dq =


= D[g]a1 c1 D[g]ap1 cp1 D[g]1 d1 b1 D[g]1 dq1 bq1 cdpq F c1 ... cp d1 ... dq
= D[g]a1 c1 D[g]ap1 cp1 D[g]1 d1 b1 D[g]1 dq1 bq1 F c1 ... cp1 a d1 ... dq1 a
= D[g]a1 c1 D[g]ap1 cp1 D[g]1 d1 b1 D[g]1 dq1 bq1 Gc1 ...cp1 d1 ...dq1 .

We say that Ga1 ...ap1 b1 ...bq1 is obtained from F a1 ...ap b1 ...bq by contracting the index ap with
bq , or tracing over the indices ap and bq . We could have contracted any upper index ai
with any lower one bj of the same tensor F a1 ...ap b1 ...bq thus obtaining p q independent traces
represented by different type-(p 1, q 1) tensors. A type-(1, 1) tensor (F i j ) can be thought
of as a n n matrix.
j, we are computing the trace of
Pn In ithis case if we contract i with
i
a1 ...ap
the matrix F i i=1 F i . The type (p, q) tensor F
b1 ...bq could have been a product
of two or more lower rank tensors and the indices over which we are contracting, ap and
bq , could have belonged to different tensors. The above proof tells us that the result is still
a type-(p 1, q 1) tensor. As an example consider the metric (gij ) and a contravariant
vector V = (V k ). The product of the two (gij V k ) is a type-(1, 2) tensor. If we contract

108

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

the index k with the index j we end up with a type-(0, 1) tensor (gij V j ) which is nothing
but the vector Vi of covariant components of V. We say that we have lowered the index j
of the contravariant vectors by means of the metric tensor gij . In general, using the metric
or its inverse, we can construct, out of a type-(p, q) tensor either a type-(p 1, q + 1) or a
type-(p + 1, q 1) tensor as follows:
 a1 ...ap1
a1 ...ap1 a
G
b1 ...bq
b1 ...bq bq+1 gbq+1 a F
a1 ...ap
,
(2.9.29)

F
b1 ...bq
ap+1 b a1 ...ap
a1 ...ap ap+1
F
G
b1 ...bq1 b
b1 ...bq1 g
in the first case we have lowered the index bq+1 using the metric, in the second we have raised
the index ap+1 using the inverse metric.
From this discussion it is clear that the representation (2.9.20) of G = GL(n, C) with
respect to which a type-(p, q) transform is reducible. Indeed we can take subspaces V (p`,q`) ,
` = 1, . . . , min(q, p), of its representation space spanned by tensors of the form
a

p`+1
V (p`,q`) {F a1 ...ap` b1 ...bq` bq`+1
bqp } V (p,q) ,

(2.9.30)

which are invariant under the action of G, as the reader can prove. The components of a
generic type-(p, q) over the V (p`,q`) subspace are obtained by taking all possible contractions of ` upper with ` lower indices. For instance the component of a type-(1, 1) (F i j ) in
V (1,1) over the subspace of type-(0, 0) tensors (scalars) consists of the trace of the corresponding matrix F k k ji . As a second example consider a type-(2, 2) tensor F = (F ij k` ). It
will be the sum of a trace-less component F0 = (F0ij k` ) in the space V (2,2) , i.e. a tensor
whose traces are all vanishing, a component F1 in the space V (1,1) and a scalar component
F2 in V (0,0) :
F ij k` = F0ij k` + F1ij k` + F2ij k` ,

(2.9.31)

F1ij k` = H i k `j + M i ` kj + N j k `i + P j ` ki ,
F2ij k` = F ki `j + G kj `i .

(2.9.32)

where

The tensors H i k , M i k , N i k , P i k are all traceless ( H i i = M i i = N i i = P i i = 0) and can be


expressed in terms of the four different traces of F ij k` .
Exercise: Show that:
1
[(n + 2) F1ik jk F1ik kj F1ki jk F1ki kj ] ,
(n + 3)(n 1)
1
=
[F1ik jk + (n + 2) F1ik kj F1ki jk F1ki kj ] ,
(n + 3)(n 1)
1
=
[F1ik jk F1ik kj + (n + 2) F1ki jk F1ki kj ] ,
(n + 3)(n 1)
1
=
[F1ik jk F1ik kj F1ki jk + (n + 2) F1ki kj ] ,
(n + 3)(n 1)

H ij =
M ij
N ij
P ij

2.9. OPERATIONS WITH REPRESENTATIONS

109

1
[n F2ij ij F2ij ji ] ,
n(n2 1)
1
[F2ij ij + n F2ij ji ] ,
G =
2
n(n 1)
F =

(2.9.33)

Let us show now that the representation of GL(n, C) over type-(2, 0) tensors (and over
type-(0, 2) tensors) is reducible. We can decompose the space V (2,0) = {(F ij )} of type-(2, 0)
(2,0)
(2,0)
tensors, which is n2 -dimensional, into two lower dimensional subspaces VS
and VA
consisting of symmetric and anti-symmetric tensors of the same kind:
(2,0)

V (2,0) = VS
(2,0)

VS

(2,0)

VA

,
(2,0)

= {(F ij )| F ij = F ji } , VA

= {(F ij )| F ij = F ji } .

(2.9.34)

The dimensions of these two subspaces is the number of independent parameters symmetric
and anti-symmetric tensors depend on. Consider a n n symmetric matrix (F ij ). All its
entries below the diagonal are equal to the symmetric ones above the diagonal, so that the
independent entries consist in the n diagonal entries and the 21 n(n 1) entries above the
(2,0)
diagonal. The dimension of VS
is therefore n + 12 n(n 1) = 21 n(n + 1). Similarly the only
independent entries of an anti-symmetric matrix F ij are the entries above the diagonal, so
(2,0)
(2,0)
(2,0)
that the dimension of VA is 21 n(n1). The reader can show that dim(VA )+dim(VS ) =
dim(V (2,0) ) = n2 .
Let us show that these subspaces are invariant under GL(n, C). Consider a type-(2, 0)
tensor with definite symmetry property, i.e. F ij = F ji and let us act on it by means of a
generic linear transformation g GL(n, C):
F 0 ij = D[g]i ` D[g]j k F `k = D[g]i ` D[g]j k F k` = D[g]j ` D[g]i k F `k = F 0 ij (2.9.35)
,
which shows that symmetric (anti-symmetric) tensors are transformed into symmetric (anti(2,0)
(2,0)
symmetric) tensors and thus VS
and VA
are invariant subspaces. They correspond to
irreducible representations. A generic tensor (F ij ) in V (2,0) is decomposed into its compo(2,0)
(2,0)
nents in VS
and VA
as follows:
F ij = FAij + FSij ,
1 ij
(2,0)
FAij =
(F F ji ) VA ,
2
1 ij
(2,0)
FSij =
(F + F ji ) VS .
2

(2.9.36)

The same decomposition can be given for type-(0, 2) tensors. Let us show that if we have a
type-(0, 2) and a type-(2, 0) with opposite symmetry, their contraction (i.e. cross-contraction
over all their indices) is zero. To show this consider a symmetric FSij and an antisymmetric
FA ij tensors of the two types and compute their contraction:
FSij FA ij = FSji FA ij = FSji FA ji = FSij FA ij = 0 .

(2.9.37)

110

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

We would have obtained the same result had we considered and anti-symmetric tensor FAij
and a symmetric one FS ij of the opposite type.
(2,0)
Consider now (G = O(n))-tensors of type-(2, 0) and let us show that VS
is further
reducible. We use now the property that, in an ortho-normal basis, the type-(2, 0) tensor
(2,0)
g ij = ij is G-invariant. Using this tensor we can split VS
into two orthogonal subspaces:
(2,0)
(2,0)
VT
(trace tensors) of VT L (traceless tensors), defined as follows:
(2,0)

VT

(2,0)

= {F ij = F ij } , VT L

= {F ij | F ij = F ji , F ij ij = 0} ,

(2.9.38)
(2,0)

the former is clearly one dimensional while the latter has dimensions dim(VS ) 1 =
(2,0)
1
n(n 1) 1. Let us show that these two subspaces are O(n)-invariant. The proof for VT
2
(2,0)
is trivial. As for VT L , take a symmetric traceless tensor F ij . Under a O(n) transformation
it is mapped into a symmetric tensor F 0 ij , which we show below to be still traceless:
F 0 ij ij = D[g]i k D[g]j ` F k` ij = F ij ij = 0 ,

(2.9.39)

where we have used the orthogonality property of D[g]: D[g]i k D[g]j ` ij = k` . Notice that
anti-symmetric tensors can not be decomposed, as we did for symmetric ones, into a trace
and a traceless part, since they are all traceless to start with. This is sue to the fact that ij
is symmetric and the contraction of a symmetric and an anti-symmetric tensor is zero.
The space V (2,0) is then decomposed into three invariant subspaces:
(2,0)

V (2,0) = VT
ij

Correspondingly a tensor (F ) in V
F

ij

FAij
FTij
FTijL

(2,0)

(2,0)

(2,0)

VT L VA

(2.9.40)

splits into its components along the three subspaces:

= FTij + FTijL + FAij ,


1 ij
(2,0)
=
(F F ji ) VA ,
2
1
(2,0)
F ij VT ,
=
n
1
(2,0)
= FSij F ij VT L ,
n

(2.9.41)

where F F k` k` .
Example: Consider the group O(3) and type-(2, 0) O(3)-tensors (F ij ) which span a (32 =
9)-dimensional space V (2,0) . This representation is the two-fold product of the defining
representation 3 of the group: 3 3. This representation is reducible into one acting on the
antisymmetric tensors, of dimension 12 3(3 1) = 3, an other acting on symmetric traceless
tensors, of dimension 21 3(3 + 1) 1 = 5 and a last one acting on the trace tensors, of
dimension 1. Denoting these irreducible representations by the boldface of their dimensions
we can write the following decomposition:
3 3 1 5 3.

(2.9.42)

Notice that, since for O(n) D = D0 , that is covariant and contravariant vectors transform
in the same representation, O(n) makes no difference between upper and lower indices, thus
(F ij ), (F i j ) and (Fij ) transform in the same representation.

2.10. REPRESENTATIONS OF PRODUCTS OF GROUPS

2.10

111

Representations of Products of Groups

Consider a representation D1 of a group G on a linear vector space Vn and a representation


D2 of a different group G0 on a linear vector space Wm . Generic vectors V = V i ei Vn and
W = W a fa Wm in the two representation spaces transform under the action of elements
g G and g 0 G0 , respectively as follows:
V D1 [g] V , W D2 [g 0 ] W .

(2.10.1)

Let us define now the action of the couple of elements (g, g 0 ) G G0 on the vector
V W V i W a ei fa Vn Wm as the simultaneous, but independent, action of the two
group elements on the respective vectors V and W:
V W (V0 W0 ) = (D1 D2 )[(g, g 0 )] (V W) (D1 [g] V) (D2 [g] W) ,
(2.10.2)
or, in components:
V i W a V 0 i W 0 a = (D1 D2 )[(g, g 0 )]ia jb V j W b D1 [g]i j D2 [g 0 ]a b V j W b (2.10.3)
.
The correspondence between elements (g, g 0 ) in G G0 and n m n m matrices D1
D2 )[(g, g 0 )] ((D1 D2 )[(g, g 0 )]ia jb ) = (D1 [g]i j D2 [g 0 ]a b ) is a representation of the product of the two groups. Consider indeed the action of the product of two elements (g1 , g10 )
and (g2 , g20 ) of G G0 on a same vector V W. Recalling the definition of product on G G0
(i.e. that (g1 , g10 ) (g2 , g20 ) (g1 g2 , g10 g20 )), we can write:
(D1 D2 )[(g1 , g10 ) (g2 , g20 )](V W) =
=
=
=
=

(D1 D2 )[(g1 g2 , g10 g20 )](V W) =


(D1 [g1 g2 ] V) (D2 [g10 g20 ] W) =
(D1 [g1 ] D1 [g2 ] V) (D2 [g10 ] D2 [g20 ] W) =
(D1 D2 )[(g1 , g10 )] [(D1 [g2 ] V) (D2 [g20 ] W)] =
(D1 D2 )[(g1 , g10 )] (D1 D2 )[(g2 , g20 )] (V W) ,

for any V Vn and W Wn , which proves that


(D1 D2 )[(g1 , g10 ) (g2 , g20 )] = (D1 D2 )[(g1 , g10 )] (D1 D2 )[(g2 , g20 )] , (2.10.4)
namely that D1 D2 is a homomorphism on G G0 and thus a representation. It can be
shown that , if D1 and D2 are irreducible D1 D2 is an irreducible representation of G G0 .
In Section 2.9.2 we have denoted by the same symbol D1 D2 , the product of two
representations of a same group, defined on products of vectors by eq. (2.9.9). Let us
emphasize here the conceptual difference between the product of two representations of a
same group G and the product of representations of two different groups G, G0 , which is a
representation of G G0 . The former describes the simultaneous action of a single group
element g G on the two vectors in the tensor product VW, according to eq. (2.9.9). The
latter describes the independent action of two elements g, g 0 on the corresponding vectors in
the tensor product V W, according in eq. (2.10.2) . The former is generally reducible,
the latter is irreducible if the two representations are.

112

CHAPTER 2. TRANSFORMATIONS AND REPRESENTATIONS

Chapter 3
Constructing Representations
3.1

Constructing Representations of Finite Groups

Let us now address the problem of constructing the irreducible representations of a finite
group G. In Section 2.4.2 we have introduced the notion of regular representation and of
abstract group algebra A[G]. We have later shown that the regular representation of a
finite group is completely reducible into the sum of all the irreducible representations D(s) ,
s = 0, . . . , p 1, with multiplicities given by their dimensions ns :
DR

p1
M

ns D(s) .

(3.1.1)

s=0

This amounts to saying that the representation space of the regular representation can be
written as the direct sum of the representation spaces of each irreducible representation
in (3.1.1). As shown in Section 2.4.2, we can choose as representation space of DR either VR , with basis (egi )i=0,...,n1 , or, equivalently, the group algebra A[G] itself, with basis
(gi )i=0,...,n1 . With each element gi G we associate the automorphism DR [gi ] on A[G]
defined as the left action of the algebra element gi on A[G]. Corresponding to the decomposition (3.1.1) A[G] will decompose into the direct sum of the representation spaces B (s) of
each irreducible representation D(s) :
A[G]

p1
M

ns B (s) = B1 B2 . . . B` ,

(3.1.2)

s=0

where we have denoted by different symbols


(Bk )k=1,...,` the spaces of all the irreducible
P
representations in (3.1.2). Clearly ` = p1
n
s=0 s . Since each Bk is invariant with respect to
the action of G, for each B Bk we must have that DR [gi ](B) gi B Bk , or, in other
words, for any gi G:
DR [gi ](Bk ) = gi Bk = Bk .

(3.1.3)

From the bi-linearity of the product it follows that Bk is invariant with respect to the leftmultiplication by any element of A = Ai gi A[G], represented by the action of DR [A] =
113

114

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Ai DR [gi ]: A Bk = Ai (gi Bk ) = Bk . Since Bk is a subspace of A[G], it will be invariant


with respect to the left-multiplication by its own elements. This means that, if B1 , B2 Bk ,
B1 B2 Bk . The space Bk is then close with respect to the product induced by A[G], and
thus it is an algebra itself. We say that Bk is a subalgebra of A[G]. Moreover it is invariant
with respect to the left-product by any element of A[G]:
A A[G] : A Bk = Bk .

(3.1.4)

This property makes the subalgebra Bk a left-ideal of A[G]. We conclude that the space of any
representation in which DR decomposes is defined by a left-ideal of A[G]. The converse is also
true: Any left-ideal of A[G] defines the space of a representation of G in the decomposition
of DR , being it invariant with respect to the action of all the automorphisms DR [gi ]. Of
course, just as any representation of a finite group is completely reducible into smaller
representations, A[G] can always be written as the direct sum of the corresponding leftideals: If B A[G] is a left-ideal, its orthogonal complement B 0 , such that A[G] = B B 0 , is
still a left-ideal. The two left-ideals may be still decomposable into direct sums of left-ideals.
This occurs if and only if the corresponding representations or G are further reducible. This
implies that the left-ideals (Bk )k=1,...,` in (3.1.2), corresponding to irreducible representations,
are themselves not further reducible into smaller left-ideals. We say that they are minimal
left-ideals. We have reduced the problem of decomposing the regular representation into
irreducible representations to that of finding finding the minimal left-ideals in A[G].
Any element A A[G] can be written, in a unique way, as the sum of its components
within each left-ideal in (3.1.2):
A = A1 + A2 + . . . + A` ,

(3.1.5)

where Ak Bk . In particular we can decompose the identity element e = g0 :


e = e1 + e2 + . . . + e` .

(3.1.6)

Let us take a generic element A A[G] and use the property A e = A. Decomposing e
according to (3.1.6) we find:
A = A e = A (e1 + e2 + . . . + e` ) = A e1 + . . . + A e` .

(3.1.7)

Since Bk are left-ideals, Aek Bk , but, being the decomposition (3.1.5) unique, we conclude
that
A1 = A e1 , . . . , A` = A e` .

(3.1.8)

In particular, if A Bk , A ek = A and A ek0 = Ak0 = 0, if k 0 6= k. We conclude that


each component ek is the identity element of the corresponding ideal Bk . Moreover, taking
A = ek we find:
e2k ek ek = ek , ek ek0 = 0 , k 0 6= k .

(3.1.9)

3.1. CONSTRUCTING REPRESENTATIONS OF FINITE GROUPS

115

An element A of A[G] which equals its own square, A2 = A, is called idempotent. The identity
element ek of each minimal left-ideal Bk of A[G] is, according to eq. (3.1.9), idempotent.
Clearly the identity element e0 of any left-ideal B of A[G] is idempotent: (e0 )2 = e0 .
Moreover we can write any element of a left-ideal B as the product of an element of A[G]
times the identity element e0 of B. On the other hand, being B a left-ideal, the product of
any element of A[G] times e0 is in B. Therefore we can obtain any left-ideal through the
left-action of A[G] on its identity element: B = A[G] e0 . We say that the identity element
e0 B generates the left-ideal B.
Suppose we find an idempotent element a A[G], a2 = a. Let us show that the space
A[G] a {A a| A A[G]}, obtained through the left action of A[G] on a, is a left-ideal
of A[G] of which a is the identity element. Clearly this space is a left-ideal since, for any
A A[G], A (A[G] a) = A[G] a. Given a generic element A a A[G] a we find
(A a) a = A a2 = A a. This shows that a is the identity element of A[G] a.
There is therefore a one-to-one correspondence between left-ideals of A[G] and idempotent
elements of the same algebra.
Suppose now that the identity element e0 of a left-ideal B A[G] can in turn be written
as the sum of two idempotent elements: e0 = e1 + e2 , with e21 = e1 , e22 = e2 , e1 e2 = 0.
These elements will generate two disjoint subalgebras (B e1 , B e2 ), of B so that we can
write: B = B1 B2 = B e1 B e2 . Indeed any element B = B e0 = B e1 + B e2
of B can be written as the sum of a component B e1 in B e1 and a component B e2 in
B e2 . The two subalgebras are disjoint since, if there existed a common non-zero element
element B (B e1 ) (B e2 ) of the two algebras, being B in the fist subalgebra we can write
B = B e1 . On the other hand, being B an element of the second subalgebra we can also
write B = B e1 = (B e1 ) e2 = B (e1 e2 ) = 0. This proves that the subspaces B e1 , B e2
have just the zero-vector in common and thus are disjoint. Each of them is a left-ideal of
A[G], since B e1 = A[G] e0 e1 = A[G] e1 and similarly B e2 = A[G] e2 . The left-ideal B
is therefore not-minimal and the corresponding representation of G is not irreducible since
it decomposes into smaller representations on the two invariant subspaces.
On the other hand, if B is not minimal, it can be decomposed into the direct sum of
left-ideals and its identity element decomposes into the sum of the identity elements of each
subalgebra, which are themselves idempotent. We conclude that, a left-ideal B is minimal,
and thus defines an irreducible representation if and only if its idempotent identity element
does not decompose into the sum of two or more idempotent elements. An idempotent
element with this property is called primitive idempotent element and our discussion shows
that there is a one-to-one correspondence between minimal left-ideals of A[G] and primitive
idempotent elements. Being each left ideal Bk in (3.1.2) minimal, their idempotent identity
elements ek are primitive.
The problem of decomposing the regular representation into irreducible components can
thus be reduced to the problem of finding primitive idempotent elements of A[G].
For what we have said each minimal left-ideal Bk is generated by its own primitive
idempotent element: Bk = A[G] ek . From eq. (3.1.2) it follows that there are as many
primitive idempotent elements P
of A[G] as irreducible representations in the decomposition
of the regular one, namely ` = p1
s=0 ns .

116

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

To summarize, in order to find the irreducible representations of a finite group G, we need


to determine the primitive idempotent elements (ek )k=1,...,` of A[G]. For each such elements
ek the space of the corresponding irreducible representation is just A[G]ek . Actually we may
generalize our search to elements a which are only idempotent up to a factor, namely such
that a2 = a. To such an element there correspond a unique idempotent element a0 = a .

3.1.1

Irreducible Representations of Sn and Young Tableaux

Let us apply our analysis to the permutation group1 Sm , of order n = m!. From the
general theory we know that there are as many irreducible representations of Sm as conjugation classes. Each conjugation class, see Section 1.4.1, is defined by a partition ()
(1 , 2 , . . . , m ) of m:
1 + 2 + + m = m ,

1 2 m ,

(3.1.10)

and is represented graphically by a Young diagram. Young has identified, for the permutation
group, a complete set of primitive idempotent (up to a factor) elements of A[Sm ], called
Young elements. Let us start giving two of them. One is the symmetrizer :
S

P,

(3.1.11)

(P )P ,

(3.1.12)

P Sm

the other is the antisymmetrizer :


A

P Sm

where (P ) is +1 if the permutation P is even, 1 if it is odd.


Exercise 3.1: Show that (P S) = (P ) (S), and also that (P ) = (S) (S P ).
Let us show that these two elements of the algebra of Sm are idempotent up to a factor.
Let us start showing that, for any S Sm :
SS

= S ,

S A = (S) A .

(3.1.13)

We indeed have:
SS

X
P Sm

SA

X
P Sm

SP =

S0 = S ,

(3.1.14)

S 0 =SP Sm

(P )S P =

(S) (S 0 ) S 0 = (S) A .

(3.1.15)

S 0 =SP Sm

In this case of the permutation group we shall denote, as we have always done, the identity element by
I: e = I.

3.1. CONSTRUCTING REPRESENTATIONS OF FINITE GROUPS

117

Using the above properties we find:


X
X
S2 =
SS =
S = m! S ,
SSm

SSm

SSm

S A

(S) S A =

(S)2 A = m! A ,

SSm

SA =(

SSm

(S))A = (

SSm

m! m!

)A = 0.
2
2

(3.1.16)

The corresponding idempotent elements are e1 = Sm! and e2 = Am! . In virtue of eqs.
(3.1.14) and (3.1.15), the corresponding minimal left-ideals are one-dimensional since
B1 = A[Sm ] S = {S S | S Sm } = {S } ,
B2 = A[Sm ] A = {S A | S Sm } = {A } ,

(3.1.17)

that is they correspond to one-dimensional representations of Sm , the former being the


identity representation, the latter being called the antisymmetric representation. Notice that
S and A are clearly primitives since the corresponding left-ideals, being one-dimensional,
are minimal.
Example 3.1: Consider S2 = {I, (1 2)}. Its algebra is two dimensional, generated by basis
elements denoted by the same symbols as the corresponding group elements. A generic
element of A A[S2 ] reads:
A = A0 I + A1 (1 2) .

(3.1.18)

The symmetrizer and antisymmetrizer read:


S
A

= I + (1 2) ,
= I (1 2) .

(3.1.19)

Example 3.2: Consider S3 . Its elements are:


S3 = {g0 , g1 , g2 , g3 , g4 , g5 } = {I, (1 2), (1 3), (2 3), (1 2 3), (3 2 1)} .

(3.1.20)

Each of them corresponds to a basis element of the group algebra A[S3 ], denoted by the
same symbol. A generic element of A A[S3 ] has then the form:
A = A0 I + A1 (1 2) + A2 (1 3) + A3 (2 3) + A4 (1 2 3) + A5 (3 2 1) .

(3.1.21)

The symmetrizer and antisymmetrizer read:


S
A

= I + (1 2) + (1 3) + (2 3) + (1 2 3) + (3 2 1) ,
= I (1 2) (1 3) (2 3) + (1 2 3) + (3 2 1) .

(3.1.22)

118

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Each conjugation class () of Sm is described by a Young diagram and is associated with an


irreducible representation. A ns dimensional irreducible representation occurs ns times in
the decomposition of the regular representation and thus corresponds to ns different primitive
idempotent elements, one for each copy of the same representation. Young gave a rule to
construct these ns primitive idempotent elements corresponding to a given diagram. We
start defining a Young tableaux which is obtained by filling the boxes of a Young diagram
with different numbers from 1 to m in an arbitrary order. For example consider the Young
diagram corresponding2 to () = (4, 3, 2), and define the following tableaux:
1 4 6 7
2 3 8
9 5

(3.1.23)

The numbering of the boxes is arbitrary. Let us now denote by P the permutations of the
m numbers in the tableaux (in the above example m = 9), which leave the rows globally
invariant, namely whose effect is to permute the numbers within each row. Examples of
permutations P for the tableaux (3.1.23) are: (1 4 6 7), (1 4), (4 7), (2 3 8), (3 8), (9 5), . . ..
The permutation (1 2), for example, is not of this kind, since it acts on numbers belonging to
two different rows. Define, in correspondence with a given Young tableau, the symmetrizer :
X
P =
P A[Sm ] .
(3.1.24)
P

1 2
is
For instance, the symmetrizer corresponding to the tableau 3
P = I + (1 2) A[S3 ] .

(3.1.25)

Let Q generically denote those permutations which leave the columns globally invariant,
namely which act on numbers lying on a same column of a given tableaux, and define the
antisymmetrizer:
X
Q =
(Q) Q A[Sm ] .
(3.1.26)
Q

1 2
For instance, the antisymmetrizer corresponding to the tableau 3
is
Q = I (1 3) A[S3 ] .

(3.1.27)

Since two or more numbers can not lie on a row and a column at the same time, the P
and the Q permutations form disjoint sets. The P and the Q permutations are products of
2

We use the convention of writing only the first non vanishing i s, the remaining being zero, so that, for
instance, (4, 3, 2, 0, 0, 0, 0, 0, 0) (4, 3, 2).

3.1. CONSTRUCTING REPRESENTATIONS OF FINITE GROUPS

119

permutations on each row and column respectively and close two different subgroups of Sm .
The Young element corresponding to the given Young tableau is defined as:
Y() Q P .

(3.1.28)

1 2
the Young element is:
For instance, for the tableau 3
Y1 2

= [I (1 3)] [I + (1 2)] A[S3 ] .

(3.1.29)

Clearly the elements S and A are the Young elements corresponding to the diagrams
(m, 0, . . . , 0) and (1, 1, . . . , 1) respectively (all tableaux defined out of these two diagrams
correspond to the same Young elements: S and A respectively). We shall not prove in
general that Y() is idempotent up to a factor, we shall restrict ourselves to some examples.
In defining a Young tableau, we assigned numbers to boxes. Of course we can choose a
different numbering. This amounts to performing a permutation on the numbers in the
original tableau. Let S be a permutation in Sm , the effect of S on the Young element
corresponding to a tableau is:
Y

Y 0 = S Y S 1 ,

(3.1.30)

where YS is the Young element corresponding to the tableau obtained from the original one
by performing S on its labels. For instance:
Y 2 3 = (1 2 3) Y 1 2 (1 2 3)1 .
1

(3.1.31)

We need to identify those tableaux which define primitive idempotent elements, such that,
if Y and Y 0 are two such elements Y Y 0 = 0, namely one can not be obtained one from
the other by the left action of the algebra: Y 0 6= A Y for some A A[Sm ]. Each of
them will generate a different irreducible representation of Sm . For a given Young diagram,
such elements are defined by standard tableaux. A standard tableau is obtained by filling
the corresponding diagram starting from the first row, from left to right, with the numbers
1, 2, . . . , m, respecting the following condition: Each number on the ith line should be greater
than the number on the same column of the (i 1)th line. A general property, which we
shall not prove, states that, for a given Young diagram, any Young tableaux can be obtained
as (combinations of) left-products of group algebra elements on standard ones, and different
standard tableaux can not be obtained one from an other in this way. The tableau (3.1.23)
for instance is not standard since the 3 on the second row is smaller than the corresponding
number (4) on the first. Examples of standard tableaux of the same type are:
1 2 3 4
,
5 6 7
8 9

1 2 3 7
,
5 6 4
8 9

(3.1.32)

The standard tableaux for the diagram (2, 1) of S3 are Y = 13 2 and Y 0 = 12 3 . Standard
tableaux corresponding to the same Young diagrams can, on the other hand, be obtained

120

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

one from the other through a conjugation by a certain permutation. Therefore they generate
different copies of the same irreducible representation of Sm . It turns out that the number of
standard tableaux corresponding to a given diagram coincides with the dimension of the corresponding representation, which is consistent with the property that a given representation
enters the decomposition of the regular one a number of times equal to its dimension.
Let us discuss the S3 case in detail. The Young diagrams are three and are given in
and the last

Table 1.4.10. The first

diagrams have just one standard tableaux each

while the second diagram


have two standard tableaux. The standard tableaux and the
corresponding Young elements, are:
S

1
1 2 , Y0 1 3 ,
1 2 3 , A 2 , Y
3
2
3

(3.1.33)

where
S
A
Y
Y0

=
=
=
=

I + (1 2) + (1 3) + (2 3) + (1 2 3) + (3 2 1) ,
I (1 2) (1 3) (2 3) + (1 2 3) + (3 2 1) ,
[I (1 3)] [I + (1 2)] = I (1 3) + (1 2) (1 2 3) ,
[I (1 2)] [I + (1 3)] = I + (1 3) (1 2) (3 2 1) ,

(3.1.34)

Exercise 3.2: Verify that:


S 2 = 6 S , A 2 = 6 A , Y 2 = 3 Y , (Y 0 )2 = 3 Y 0 ,
S A = A S = 0 , S Y = Y S = 0 , S Y 0 = Y 0 S = 0,
A Y = Y A = 0 , A Y 0 = Y 0 A = 0 , Y Y 0 = Y 0 Y = 0 . (3.1.35)
Exercise 3.3: Express the identity element I as a linear combination of the S , A , Y , Y 0 .
Let us verify that Y generates a two dimensional representation of S3 , namely that
A[S3 ] Y is a two-dimensional space. To this end we act on Y to the left by the S3
elements, namely we compute gi Y :
I Y
(1 2 3) Y
(1 3) Y
(1 2) Y
(2 3) Y
(3 2 1) Y

=
=
=
=
=
=

Y = I (1 3) + (1 2) (1 2 3) ,
(1 3) (2 3) + (1 2 3) (3 2 1) ,
I + (1 3) (1 2) + (1 2 3) = Y ,
I + (1 2) (2 3) (3 2 1) = Y + (1 2 3) Y ,
(1 3) + (2 3) (1 2 3) + (3 2 1) = (1 2 3) Y ,
((1 2) (1 3)) Y = (1 2) Y = Y (1 2 3) Y .

(3.1.36)

We see that all the elements gi Y can be expressed as combinations of two basis elements
f1 Y and f2 (1 2 3) Y . This means that A[S3 ] Y is a two-dimensional space on which
the following two-dimensional representation D of S3 act:
D[gi ](fa ) = D[gi ]b a fb gi fa .

(3.1.37)

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

121

Where the matrix representation D[gi ] = (D[gi ]b a ) of each group element reads:




1 0
1 1
D[I] = 12 , D[(1 2)] =
, D[(1 3)] =
,
1 1
0 1






0 1
0 1
1 1
D[(2 3)] =
, D[(1 2 3)] =
, D[(3 2 1)] =
.
1 0
1 1
1 0
Exercise 3.4: Following the same procedure as above, show that the space A[S3 ] Y 0 is
generated by Y 0 , (3 2 1) Y 0 and thus is two-dimensional.
Exercise 3.5: Construct the two-dimensional matrix representation on the subalgebra
generated by Y 0 , in the basis (Y 0 , (3 2 1) Y 0 ).
Exercise 3.6: Prove that Y 3 2 = Y (3 2 1) Y 0 .
1

To summarize, we have shown for S3 that the regular representation decomposes as


follows:
DR

(3.1.38)

which implies that S3 has two one-dimensional representations and a two-dimensional irreducible representation. Correspondingly the algebra A[S3 ] decomposes into the following
direct sum of minimal left-ideals:
A[S3 ] = A[S3 ] S A[S3 ] A A[S3 ] Y A[S3 ] Y 0 ,

(3.1.39)

where the last two spaces are two dimensional and are generated by the Young elements
corresponding to the standard tableaux 13 2 and 12 3 . They define the two copies of the
representation in eq. (3.1.38).
Exercise 3.7: Construct for the group S4 the Young tableaux corresponding to the Young
diagrams in Table 1.4.11. What is the dimension of the irreducible representation corresponding to each diagram?

3.2

Irreducible Representations of GL(n, C)

Consider the linear space V (p,0) of type-(p, 0) tensors T = (T i1 i2 ip ), the indices ii , . . . , ip


running from 1 to n, where V (p,0) represents the product of p copies of a n-dimensional
complex linear vector space Vn : V (p,0) = Vn Vn Vn . Recall that under a GL(n, C)
transformation D = (Di j )3 the tensor transforms as follows:
T i1 i2 ip T 0 i1 i2 ip = Di1 j1 Di2 j2 . . . Dip jp T j1 j2 jp .

(3.2.1)

Symbolically we shall represent the action of GL(n, C) on the tensor by the operator Dp
D . . . D, so that the above transformation will read
T T0 = Dp T .
3

(3.2.2)

Here, for the sake of simplicity, we denote by the symbol D the matrix D[g] corresponding to a generic
element g GL(n, C).

122

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Let us define the action of a permutation S Sp on type-(p, 0) tensors represented by an


defined,
operator S on V (p,0) which maps a type-(p, 0) tensor T into a type-(p, 0) tensor ST
in components, as follows:
i1 ip T iS(1) iS(p) ,
(ST)

(3.2.3)

where S maps the label k into S(k). The correspondence between S and the operator S on
V (p,0) is a homomorphism. To show this we need to prove that the operator corresponding
to the product of two permutations coincides with the product of the operators associated
with each of the two: S[
P = S P , for any two S, P Sp . Consider the action of the
product:
(S[
P T)i1 ip = T iS(P (1)) iS(P (p)) .

(3.2.4)

Consider now the consecutive action of the two operators:


(S (P T))i1 ip = (P T)iS(1) iS(p) = T iS(P (1)) iS(P (p)) .

(3.2.5)

This proves that the mapping S Sp S Aut(V (p,0) ) is a homomorphism, namely a


representation of the permutation group over type-(p, 0) tensors. For instance:
d
\
((1
2) T)i1 i2 = T i2 i1 , ((1
2 3) T)i1 i2 i3 = T i2 i3 i1 .

(3.2.6)

d
d
d
d
d
Suppose we want to compute ((1
3) (1
2) T)i1 i2 i3 , we have ((1
3) (1
2) T)i1 i2 i3 = ((1
2) T)i3 i2 i1 =
\
T i2 i3 i1 = ((1
2 3) T)i1 i2 i3 . Notice that the effect of (1 2) on a tensor is to permute the fist and
the second indices, whatever they are. The effect of a permutation on a tensor is just to reshuffle its components, changing their labels. This representation induces a representation of
the whole algebra A[G] associated with the group Sp on the type-(p, 0) tensors: An algebra elP
P

ement A = SSp A(S) S act on tensors by means of the endomorphism A = SSp A(S) S.
Let us show that the action of a GL(n, C)-transformation and of a permutation on a tensor
commute, namely that, if T0 = Dp T,
S T0 = (S T)0 ,

(3.2.7)

that is S and Dp are commuting operators. Let us write the tensor on the left hand side
in components:
(S T0 )i1 ...ip = T 0 iS(1) ...iS(p) = DiS(1) jS(1) . . . DiS(p) jS(p) T jS(1) ...jS(p) =
= DiS(1) j . . . DiS(p) j
(S T)j1 ...jp .
S(1)

S(p)

(3.2.8)

Now we can just re-shuffle the D matrices on the right hand side and write:
(S T0 )i1 ...ip = Di1 j1 . . . Dip jp (S T)j1 ...jp = (S T)0 i1 ...ip ,

(3.2.9)

representing a generic element of


which proves eq. (3.2.7). Since the endomorphism A,
S Sp , it will also commute with the action of
the algebra A[Sp ], is a combination of S,

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

123

GL(n, C) on tensors. In particular the action of the primitive idempotent (up to a factor)
Young elements Y() commutes with Dp :
c() (Dp T) = Dp (Y
c() T) .
Y

(3.2.10)

c() is invariant under GL(n, C):


As a consequence of this the image of V (p,0) through Y
c() V (p,0) ) = Y
c() (Dp V (p,0) ) = Y
c() V (p,0) .
Dp (Y

(3.2.11)

c() T represent? Consider


Given a tensor T = (T i1 ...ip ), and a Young tableau (), what does Y
p = 2. In this case the Young elements are:
S

= Y 1 2 = I + (1 2) , A = Y 1 = I (1 2) .

(3.2.12)

Since S 2 = 2 S and A 2 = 2 A , the idempotent elements are e1 = 21 S and e2 =


that the identity I decomposes in the corresponding left-ideals as follows:
I = e1 + e2 =

1
1
S + A.
2
2

1
2

A , so

(3.2.13)

The same decomposition holds for the corresponding action on the tensors T = (T i1 i2 ):
1
1
T = I T = (S T) + (A T) = TS + TA .
2
2

(3.2.14)

We have decomposed T into two components TS = (TSi1 i2 ) and TA = (TAi1 i2 ):


1
1
1
d
(S T)i1 i2 = [(I + (1
2)) T]i1 i2 = (T i1 i2 + T i2 i1 ) ,
2
2
2
1
1
1
d
(A T)i1 i2 = [(I (1
2)) T]i1 i2 = (T i1 i2 T i2 i1 ) ,
=
2
2
2

TSi1 i2 =
TAi1 i2

(3.2.15)

corresponding to the symmetric and antisymmetric part of T. The images S V (p,0) and
A V (p,0) of V (p,0) , through the symmetrizer and antisymmetrizer respectively, represent the
subspaces of symmetric and antisymmetric type-(2, 0) tensors respectively and, as discussed
in the previous sections, they indeed define irreducible representations of GL(n, C). The
action of the two elements S , A correspond to two distinct symmetry operations which
can be applied to a type-(2, 0) tensor: The symmetrization and the antisymmetrization with
respect to its indices. In general the action of a Young element Y() on a tensor represent
a maximal set of symmetry operations which can be applied to the tensor. By maximal we
mean that any further symmetry operation on the resulting tensor would give either zero
or the tensor itself. For instance if we further symmetrize the tensor TS with respect to
c is idempotent
its indices we obtain the tensor itself (this is due to the fact that e1 = 12 S
and thus e1 TS = e21 T = TS ), while if we antisymmetrize it we get zero (this being related
to the property e2 e1 = 0 which implies e2 TS = e2 e1 T = 0). We conclude that e1

124

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

and e2 are then projection operators into the symmetric and antisymmetric part of a type(2, 0) tensor. Consider now a type (p, 0) tensor T = (T i1 ...ip ). What is a maximal set of
symmetry conditions that we can impose on the tensor? We can group the indices into
subsets containing 1 , 2 , . . . , p indices (take 1 2 . . . p ), so that 1 + . . . + p = p.
We can then symmetrize the tensor to be symmetric with respect to the indices within the
same group. This amounts to construct a tensor which is totally symmetric with respect
to the interchange of any couple of indices within each subset. Each component of this
new tensor are obtained by summing together all the components of the original one which
are obtained from one another through the action of permutations P of indices within the
each group. In other words if we define the Young tableau () = (1 , 2 , . . . , p ) whose
rows are filled with the labels of the indices belonging to each group, recalling that the
symmetrizer P is the sum of all the permutations of the objects within each group, the
c to the
symmetrization procedure described above amounts to applying the operator P
original tensor. For instance we can group the three indices of T = (T i1 i2 i3 ) into two
subsets (i1 , i3 ) and (i2 ) (in this example 1 = 2, 2 = 1, 3 = 0). Symmetrizing the
c i 1 i2 i3 = T i1 i 2 i3 + T i3 i 2 i1 ,
tensor with respect to i1 and i3 yields the tensor N i1 i2 i3 = (PT)
d is the symmetrizer corresponding tot he Young diagram 1 3 , which
c = I + (13)
where P
2

is symmetric in i1 and i2 . After having symmetrized the tensor with respect to the indices
within each subset, we can still impose antisymmetry conditions. Of course these conditions
can not involve indices within a same subset, since only the zero-tensor can be antisymmetric
and symmetric with respect to a same couple of indices. We can impose antisymmetry
conditions among indices belonging to different sets. Since the order of indices within each
set is irrelevant, we can place the labels of these indices in the first column of the Young
tableau (). After having performed this first antisymmetrization, we can choose other
indices within different subsets, place them in the second column of the Young tableau and
proceed antisymmetrize the tensor with respect to them. Antisymmetrizing a tensor with
respect to a subset of indices amounts to defining a new tensor whose components are the
sum of all components of the original tensors obtained one from another by a permutation of
the indices within this subset, with a plus sign if the permutation is even and a minus sign if
it is odd. We iterate this procedure until we can apply no further symmetry condition. This
c T, obtained from
antisymmetrization procedure then amounts to applying to the tensor P
b
the first symmetrization, the anitsymmetrizer operator Q corresponding to the same tableau.
In other words the resulting tensor is obtained from the original one by applying to it the
c() = Q
b P.
c Each Young tableau defines therefore a maximal set of symmetry
operator Y
operations which can be implemented to a tensor by the action of the corresponding Young
element. In our example, on the tensor N i1 i2 i3 , which is symmetric with respect to (i1 , i3 ), we
can still antisymmetrize with respect to the couple (i1 , i2 ) and obtain the tensor M i1 i3 ,i2 =
d
b PT)
c i1 i2 i3 = N i1 i2 i3 N i2 i1 i3 = T i1 i2 i3 + T i3 i2 i1 (T i2 i1 i3 + T i3 i1 i2 ), being Q
b = I (12).
(Q
i1 i2 i3
Notice that the antisymmetrization spoils the symmetry property of the tensor N
with
i1 i3 ,i2
i3 i1 ,i2
respect to the indices in each subset, so that M
6= M
. There is no further symmetry
condition that we can impose on the tensor. Such tensor is then called irreducible and is
denoted by M i1 i3 ,i2 , where we divide the subsets of indices by commas. Any tensor can be

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

125

written as the sum of irreducible components. Each component is obtained by acting on


the tensor by means of a primitive idempotent element of the algebra A[Sp ]. Indeed take
for example a generic tensor T = (T i1 i2 i3 ) and lets act on it by means of the Young element
Y 1 3 = Y 0:
2

d
d
c0 T)i1 i2 i3 = [(I (1
2)) (I + (1
3)) T]i1 i2 i3 = T i1 i2 i3 + T i3 i2 i1 T i2 i1 i3 T i3 i1 i2 = M i1 i3 ,i2 .
(Y
The reader can verify that M i1 i3 ,i2 = M i2 i3 ,i1 . The effect of the symmetrizer P = (I +(1 3))
in the above Young element is to symmetrize with respect to (i1 , i3 ), while the effect of the
antisymmetrizer Q = I (1 2) is to antisymmetrize with respect to (i1 , i2 ). In summary,
each irreducible tensor, denoted by the symbol M , corresponds to a Young tableau whose
first row is filled with the labels of the first set of 1 indices, the second row, with the labels
of the second set of 2 indices and so on, and is obtained by acting on a generic tensor by
means of the corresponding Young element, which can then be thought of, up to a coefficient,
as a projection operator. Consider for instance the tableau:
1 3 6 8 9
2 4 7 11
5 10
12

(3.2.16)

The corresponding irreducible tensor is denoted by:


M i1 i3 i6 i8 i9 ,i2 i4 i7 i11 ,i5 i10 ,i12 .

(3.2.17)

c=Q
bP
c projects any type-(12, 0)
The Young element corresponding to the above tableau Y
c symmetrizes
tensor into the irreducible component (3.2.17) as follows: The symmetrizer P
the tensor with respect to the indices within each row, while the effect of the antisymmetrizer
b is to antisymmetrize the resulting tensor with respect to the indices within each column.
Q
As a result, the tensor M is antisymmetric with respect to the indices within each column.
For instance the tensor in eq. (3.2.17) is antisymmetric in the subsets of indices (i1 , i2 , i5 , i12 ),
(i3 , i4 , i10 ), (i6 , i7 ) and (i8 , i11 ).
c() V (p,0) consists of irreducible tensors and
For a given Young tableau () the space Y
defines a representation of GL(n, C). Since the only process of reduction of a tensor which
is consistent with the free action of GL(n, C) on its indices and commutes with it is the
c() V (p,0)
symmetrization procedure described above, and since the tensors in each subspace Y
are subject to a maximal set of symmetrization conditions, defined by the corresponding
c() V (p,0) defines an irreducible representation of GL(n, C).
tableau (), Y
We conclude that each Young tableau defines an irreducible representation of GL(n, C).
The fundamental representation D acts on contravariant vectors (V i ) which are type(1, 0) tensors, it is associated with the Young diagram 1 which coincides with its diagram
. We shall identify D with the corresponding Young diagram and represent it by a single
box: D . The p-fold product of fundamental representations, acting on the type-(p, 0)

126

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

tensors in V (p,0) will be then denoted by:


p

z
}|
{
z
}|
{
D D =

(3.2.18)

The decomposition of this product into irreducible representations correspond to the decomc() V (p,0) , for all
position of its representation space V (p,0) into irreducible tensors within Y
Young tableaux (). This decomposition is effected as follows: Suppose, for a given tableau
2
(), Y()
= Y() . The corresponding idempotent element e() is e() = 1 Y() . The representation I on V (p,0) of the identity element I splits into the sum of the representations e()
of the primitive idempotent elements e() corresponding to all standard Young tableau:
X
X
1 c
I =
e() =
Y() .

Young tableaux ()

Young tableaux ()

Applying both sides on V (p,0) we obtain the following decomposition:




X
X

1
(p,0)
(p,0)
(p,0)
(p,0)
V
= I V
=
e() V
=
Y() V
.

Young tableaux ()

Young tableaux ()

(3.2.19)
The effect of e() is to project the tensors in V (p,0) into the irreducible component corresponding to the diagram (). Equation (3.2.19) then defines the decomposition of the product
(3.2.18) into irreducible representations. Let us apply this procedure to type- (3, 0) tensors
and decompose the 3-fold product of the fundamental representation of GL(, C).
In Section 3.1.1 we have found four standard Young tableaux, the corresponding Young
elements being:
S

= Y1 2 3 , A = Y1 , Y = Y1 2 , Y 0 = Y1 3 .
2
3

(3.2.20)

These define four primary idempotent elements: e1 = 61 S , e2 = 16 A , e3 = 13 Y , e4 = 13 Y 0 , .


The identity element decomposes as follows:
I =

1
1
1
1
S + A + Y + Y 0.
6
6
3
3

(3.2.21)

The space of type- (3, 0) tensors decompose correspondingly into the following subspaces
1 c (3,0) 1 c (3,0) 1 c (3,0) 1 c0 (3,0)
V (3,0) = I V (3,0) = S
V
+ AV
+ Y V
+ Y V
, (3.2.22)
6
6
3
3
each of which is acted on by a corresponding irreducible representation of GL(n, C). Let us
analyze their content. Consider a generic tensor T = (T i1 i2 i3 ) and project it into the above
subspaces by acting on it with the corresponding idempotent element ei :
T = I T =

4
X
i=1

e1 T =

1c
1c
1c
1 c0
ST+ A
T+ Y
T+ Y
T.
6
6
3
3

(3.2.23)

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

127

where:
(
e1 T)i1 i2 i3 =
=
(
e2 T)i1 i2 i3 =
=
(
e3 T)i1 i2 i3 =
=
(
e4 T)i1 i2 i3 =
=

1
1 c i1 i2 i3
= (T i1 i2 i3 + T i2 i3 i1 + T i3 i1 i2 + T i2 i1 i3 + T i3 i2 i1 + T i1 i3 i2 ) =
(S T)
6
6
(3.2.24)
M i1 i2 i3 ,
1 c i 1 i2 i3
1
= (T i1 i2 i3 + T i2 i3 i1 + T i3 i1 i2 T i2 i1 i3 T i3 i2 i1 T i1 i3 i2 ) =
(A T)
6
6
(3.2.25)
M i1 ,i2 ,i3 ,
1 c i1 i2 i3
1
= (T i1 i2 i3 T i3 i2 i1 + T i2 i1 i3 T i2 i3 i1 ) =
(Y T)
3
3
i1 i2 ,i3
,
(3.2.26)
M
1 c 0 i1 i2 i3
1 i1 i 2 i3
= (T
T i2 i1 i3 + T i3 i2 i 1 T i 3 i1 i2 ) =
(Y T)
3
3
M i1 i3 ,i2 ,
(3.2.27)
(3.2.28)

We will then write the corresponding decomposition of the 3-fold product of the fundamental
representation of GL(n, C), as follows:

1 2 3

1
2 1 2 1 3
3
2
3

. (3.2.29)

The above decomposition is also written representing each diagram by the corresponding
-indices:
(1) (1) (1) (3) (1, 1, 1) 2 (2, 1) .

(3.2.30)

Each Young diagram () defines a kind of tensor with maximal symmetry and thus corresponds to an irreducible representation D() of GL(n, C)4 : For instance
corresponds to
,
the representation with tensors of the kind M . Different Young tableaux corresponding to
the same diagram, define different copies of the same representation in the decomposition of
a generic type-(p, 0) tensor, they just differ from the choices of the indices of the T- tensor on
which the symmetry operations are applied and thus lead to the same GL(n, C)-irreducible
representation since, with respect to the action of this group, all indices
Pk are on an equal
footing. Consider a k-row Young diagram () (1 , . . . , k ), with i=1 k = p, p being
the total number of indices of the corresponding irreducible tensor. Of course we know that
k p, but we have a more fundamental bound on k directly related to the group: k n.
Indeed recall that the tensor corresponding to a given Young tableau is antisymmetric with
respect to the indices in each column. Therefore the length of a column can not exceed
n boxes, otherwise we will have a tensor which is antisymmetric in a number k of indices
which is greater than their range n. Such tensor would have vanishing components since two
or more of the k indices should have the same value and a tensor which is antisymmetric
4

We shall also denote the representation D() simply by ().

128

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

in two indices have vanishing entry when the indices have the same value. Therefore the
action of a Young element Y corresponding to a tableau with more than n rows annihilates
c T = 0. We shall represent a k-row Young diagram (1 , . . . , k ),
a GL(n, C)-tensor T: Y
defining a GL(n, C)-irreducible representation, also by the n-plet (1 , . . . , n ), where all i ,
with k < i < len are zero.
Let us now address the problem of finding the dimension of a GL(n, C)-irreducible representation described by a given Young diagram. We should write all independent components
of the corresponding irreducible tensor M . Consider for instance M i1 i2 ,i3 . We should keep in
mind that these irreducible tensors are antisymmetric with respect to indices which are in
the same position within different subsets, namely in the same columns of the corresponding
Young tableaux. For instance M i1 i2 ,i3 = M i3 i2 ,i1 . Therefore components of the kind M ij,i
are zero. We can convince ourselves that the independent components are:
M 11,2 =
M 11,3 =
M 12,2 =
M 13,3 =
M 22,3 =
M 23,3 =
M 12,3 =
M 13,2 =

1
3
1
3
1
3
1
3
1
3
1
3
1
3
1
3


2 T 112 T 121 T 211 ,

2 T 113 T 131 T 311 ,

T 122 + T 212 2 T 221 ,

T 133 + T 313 2 T 331 ,

2 T 223 T 232 T 322 ,

T 233 + T 32,3 2 T 332 ,

T 123 T 321 + T 213 T 231 ,

T 132 T 231 + T 312 T 321 .

(3.2.31)

Any other component is a combination of the ones listed above. Consider or instance M 21,3 .
This component coincide with the difference M 12,3 M 13,2 . The independent components
in (3.2.31) can be represented by the following symbols:
1 1
2

, 1 1 , 1 2 , 1 3 , 2 2 , 2 3 , 1 2 , 1 3 .
3

(3.2.32)

The above diagrams, called standard components, should not be confused with the Young
tableaux, since the boxes are filled with values of the indices and not with their labels! The
standard components corresponding to a given Young diagram represent the independent entries of the corresponding irreducible tensor and their number therefore gives the dimension
of the representation. For instance we have found 8 standard components for the representation
which is therefore eight dimensional. There is a simple recipe to construct the
standard components for a given Young diagram and thus to deduce the dimension of the
corresponding representation: Fill the Young diagram with values of the indices, so that,

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

129

in reading each row from left to right these values are non-decreasing (they may be equal),
and in reading each column from top to bottom, they are in a strictly increasing order. For
instance the components 3 1 , 2 1 are not standard. Using this rationale let us compute
2
3
the dimensions of certain irreducible representations.
Totally antisymmetric type-(p, 0) tensor. Consider the representation described by a
single column of p-boxes, namely by the diagram (1, 1, . . . , 1):

..
.

(3.2.33)

When constructing the standard components, the values in each box have to be all different
and arranged in a strictly increasing order from top to bottom. Each standard component is
then defined by a set of distinct p values i1 , . . . ip of the indices. The number of components
is
 thus
 given by the number of sets of p distinct numbers out of n total values, namely by
n
. We then find:
p
 
n!
n
.
dim(1, 1, . . . , 1) =
=
p
p! (n p)!

(3.2.34)

Totally symmetric type-(p, 0) tensor. The corresponding diagram (p) consists of p


boxes in a single row:

(3.2.35)

To construct the standard components we need to fill them with p out of n, not necessarily
distinct, values i1 , . . . , ip in a non-decreasing order:
i1 i2 . . . ip n .

(3.2.36)

This implies the following relation among the values:


i1 < i2 + 1 < . . . < ip + p 1 n + p 1 .

(3.2.37)

namely each choice of values i1 , . . . , ip corresponds to a single choice of strictly increasing


values i1 , i2 + 1, . . . , ip + p 1 out of n + p 1. The total number of possibilities is still given
by a binomial:


(n + p 1)!
n+p1
dim(p) =
=
.
(3.2.38)
p
p! (n 1)!

130

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

The Gun-diagram (2, 1):


. Let us consider now the diagram (2, 1). When constructing the standard components, we have to distinguish three cases:
i) Three distinct values i < j < k;
ii) Two distinct values i < j.
Clearly we cannot use a single values since there is a column of two boxes. Let us consider
the two cases separately:
 
n
i) We have
possible choices of 3 distinct numbers i < j < k out of n. For each
3
choice we can arrange them as follows:
i j
k

i k
j

(3.2.39)

 
n
We have therefore 2
diagrams for this case.
3
 
n
ii) We have
distinct couples of values i < j out of n possible ones. For each couple
2
we have the following diagrams:
i j
j

i i
j

 
n
for a total number of 2
standard components.
2
The dimension of the representation is then:
 
 
n (n2 1)
n
n
.
dim(2, 1) = 2
+2
=
3
2
3

(3.2.40)

(3.2.41)

For n = 3 we find the eight diagrams in eq. (3.2.31).


The representation (2, 1, . . . 1) of type-(p, 0) tensors. The diagram of this representation is:

..
.

(3.2.42)

and consists of p-boxes arranges in a first column of p 1-boxes and a second single box
column. This generalizes the gun diagram corresponding to the p = 3 case. In constructing
the corresponding standard components we can either arrange p distinct values i1 < i2 <

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

131

.. . <
 ip in the boxes or p 1 values i1 < i2 < . . . < ip1 . In the former case, for each of the
n
choices of i1 < i2 < . . . < ip out of n, we have the following p 1 possibilities:
p
i1
i3
..
.
ip

i2

i1
i2
, ..
.
ip

i3
...

i1
i2
..
.

ip
,

(3.2.43)

ip1

 


n
n
yielding (p1)
standard components. As for the latter case, for each of the
p
p1
choices of i1 < i2 < . . . < ip1 out of n, we have the following p 1 arrangements:
i1
i2
..
.

i1

i1
i2
..
.

i2
...

i1
i2
..
.

ip1
,

(3.2.44)

ip1
ip1
ip1


n
yielding (p 1)
standard components. The dimension of the representation is
p1
therefore:
 


(p 1)(n + 1)!
n
n
.(3.2.45)
dim(2, 1, . . . 1) = (p 1)
+ (p 1)
=
p
p1
p!(n p + 1)!
Notice that for p = 3 we find (3.2.41).
The box diagram (2, 2):

When filling the diagram with numbers out of n, we have the following choices:
i) Four distinct values i < j < k < `;
ii) Three distinct values i < j < k;
iii) Two distinct values i < j.
 
n
Case i). For each of the
choices of i < j < k < `, we have the following two
4
components:
i j ,
k `

i k .
j `

(3.2.46)

 
n
Case ii). For each of the
choices of i < j < k, we have the following three components:
3
i i ,
j k

i j , i j .
j k
k k

(3.2.47)

132

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

 
n
Case ii). For each of the
choices of i < j, we have the following component:
2
i i .
j j

(3.2.48)

Summing up the numbers found above we get:


 
   
n2 (n2 1)
n
n
n
.
dim(2, 2) = 2
+3
+
=
4
3
2
12
The rifle diagram (3, 1):
standard components:

(3.2.49)

. Let us list the possible cases with the corresponding

i) Four distinct values i < j < k < `. The independent components are
i j k ,
`

i k ` , i j ` .
j
k

(3.2.50)

ii) Three distinct values i < j < k. The independent components are
i i j ,
k

i i k ,
j

i k k , i j j ,
j
k

i j k , i j k .
j
k

(3.2.51)

iii) Two distinct values i < j. The independent components are


i i i ,
j

i j j ,
j

i i j .
j

(3.2.52)

We conclude that:
 
 
 
n (n + 2)(n2 1)
n
n
n
dim(3, 1) = 3
+6
+3
=
.
3
2
4
8

(3.2.53)

A general formula. let us give for completeness a general formula for computing the
dimension of a GL(n, C)-irreducible representation. Define the Cayley van der Monde determinant:
Y
D(x1 , x2 , . . . , xn )
(xi xj ) .
(3.2.54)
i<j

The dimension of the representation (1 , 2 , . . . , n ) is:


dim(1 , 2 , . . . , n )

D(`1 , `2 , . . . , `n1 , `n )
,
D(n 1, n 2, . . . , 1, 0)

(3.2.55)

3.2. IRREDUCIBLE REPRESENTATIONS OF GL(N, C)

133

where `i i + n i. The reader can verify that:


D(n 1, n 2, . . . , 1, 0) = (n 1)!(n 2)! . . . 2! .

(3.2.56)

Let us apply this formula to the rifle diagram (3, 1). In this case 1 = 3, 2 = 1 and all i ,
for 2 < i n = 0. The numbers (`i ) are:
`1 = 3 + n 1 = n + 2 , `2 = 1 + n 2 = n 1 , `i = n i (i = 3, . . . , n) .
The reader can easily verify that:
D(`1 , `2 , . . . , `n1 , `n ) =

1
(n + 2)!(n 1)!(n 3)!(n 4)! . . . 2!1! .
8

(3.2.57)

From (3.2.55) we then find:


(n + 2)!
1
(n + 2)!(n 1)!(n 3)!(n 4)! . . . 2!1!
=
= n (n + 2) (n2 1) .
8 (n 1)!(n 2)! . . . 2!
8 (n 2)!
8

dim(3, 1) =

Exercise 3.8: By by directly applying formula (3.2.55), prove that:


For n = 3, dim(7, 4, 2) = 42;
For n = 4, dim(2, 2, 2) = 10;
For n = 5, dim(2, 2, 2) = 50;
For n = 3, dim(6, 6) = 28;
Now we are able, for instance, to compute the dimension of the irreducible representations
appearing on the right hand side of the decomposition (3.2.30). Representing the irreducible
representations of GL(n, C) by the boldface of their dimensions, eq. (3.2.30) can also be
written as follows:
3 3 3 1 10 8 8 .

(3.2.58)

Exercise 3.9: Prove that type-(4, 0) GL(3, C)-tensors decompose as follows:

(3.2.59)

or,
3 3 3 3 151 3 152 2 6 30 ,
where 3 =

, 30 =

, 151 =

(totally symmetric tensor) and 152 =

(3.2.60)
.

Notice that the totally antisymmetric tensor is not present since the number of its indices
would exceed their range.

134

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

3.3

Product of Representations

Suppose we have two GL(n, C)-irreducible tensors and multiply them together. Suppose
one is of type-(p1 , 0) and the other of type-(p1 , 0). The product is again a tensor with as
many upper indices and the sum of the number of indices of the two factors, namely it is
a type-(p1 + p2 , 0) tensor. This tensor will in general be reducible. We have learned that
type-(p, 0) tensors decompose in irreducible representations which are in correspondence
with irreducible representations of Sp . Therefore the representations in the decomposition
of the product will correspond to irreducible representations of Sp1 +p2 which acts on each
component of the product by shuffling all the p1 + p2 indices. Let us learn how, using Young
diagrams, we can decompose the product of two irreducible representations into irreducible
ones. Suppose the representations of the two tensors be described by the Young diagrams
() and (0 ) of Sp1 and Sp2 respectively. In which representations of Sp1 +p2 does the product
() (0 ) decompose? Consider a simple case in which () = (3, 1) and (0 ) = (2, 1) of S4
and S3 respectively. There is a simple figurative rule to perform such decomposition. Fill
the first row of the left diagram with the letter a, the second with b the third with c and so
on. It is convenient to choose the left diagram, to be filled with letters, as the simplest (i.e.
smaller) of the two. In our example we will have:
a a .

(3.3.1)

Now construct all possible (viable) Young diagrams by attaching the boxes filled with a to
the right diagram, with the only condition that two a-boxes should not lie on a same column.
In our example we have the following diagrams:
a
a a

a a

(3.3.2)

Next we attach the b-boxes to each of the above diagrams to construct new viable Young
diagrams. In doing this, however, we should not just follow the rule previously given for
the a-boxes, but also the following restriction: In reading each row from right ro left and
each column from top to bottom, the fist symbol we encounter should appear in the sequence
at least as many times as the second, the second symbol we encounter at least as many
times as the third and so on. This last restriction defines sequences of symbols called lattice
permutations. For instance if, reading a row from right to left, we find the sequence baa,
the fist symbol we encounter is b and appears less times than the second a. This is not a
lattice permutation and the corresponding diagram should be discarded. Next we attach the
c-boxes following the same restrictions given for the a and the b-boxes and so on. In our
example we find:
a
a
a
b

a
b

a
b

b
a

a
b

a
b

3.4. BRANCHING OF GL(N, C) REPRESENTATIONS WITH RESPECT TO GL(N 1, C)135

a
a

(3.3.3)

These are the representations of S7 which appear in the decomposition of the product of the
two tensors. Suppose now the group we are considering is GL(4, C), so that n = 4. We can
compute, using the general formula (3.2.55), the dimensions of the representations involved
in the product and of those appearing in the decomposition:
= 45 ,

= 201 ,

= 1401 ,

= 224 ,

= 1402 ,

= 60 ,

= 36 ,

= 120 ,

= 202 ,
= 203 ,

(3.3.4)

The decomposition of the product (3.3.1) then reads:

or, equivalently:
45 201 224 120 1401 2 1402 202 60 36 203 .
We can apply the above procedure to compute the 3fold product
mental representations of GL(3, C):

= (

of the funda-

3.4

(3.3.5)

Branching of GL(n, C) Representations With Respect to GL(n 1, C)

Consider a contravariant vector V = V i ei = (V i ) of Vn , in the fundamental representation


D n of GL(n, C). It is a type-(1, 0) tensor whose components are listed in a column

136

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

vector:

V1
...

V =
V n1 .
Vn

(3.4.1)

The representation space Vn can be split into the direct sum of a subspace Vn1 spanned
by the vectors (ea ), a = 1, . . . , n 1, parametrized by the first n 1 components V a in
(3.4.1) and a one dimensional subspace V1 = {en }, parametrized by the nth component V n
in (3.4.1): Vn = Vn1 V1 . We can consider the subgroup of GL(n, C) which has, through
the representation D, a block diagonal action on the vector (3.4.1), in which the first block
is a (n 1) (n 1) matrix acting on V a and the last block is 1. With respect to this action
Vn1 and V1 are invariant subspaces and thus n is, by construction, no longer irreducible
with respect to this subgroup. The largest subgroup of GL(n, C) which has this action is
GL(n 1, C). The fundamental representation n of the former decomposes into the direct
sum of the fundamental representation n 1 (acting on Vn1 ) and the identity representation
1 of the latter (the identity representation acting on V1 ). In other words, with respect to
GL(n 1, C):
GL(n, C) GL(n 1, C) ,
n (n 1) 1

n ,

(3.4.2)

where n represents the nth component V n of the tensor V i on which the identity representation acts. The component V n is called a singlet with respect to GL(n 1, C). This
subgroup in other words does not see the nth component of the vector. A decomposition
of a representation of a group with respect to a subgroup is called branching law.
How does a generic irreducible GL(n, C)-tensor branch with respect to GL(n 1, C)?
. Its components split into
Consider for instance a tensor (M i1 i2 ,i3 ) in the representation
ab,c
an,b
an,n
ab,n
the following subsets: M , M , M
,M
(recall that (M i1 i2 ,i3 ) is antisymmetric in
i1 , i3 , so that i1 and i3 can not have the same value and that the only independent components
are the standard ones). As the index i splits into (a, n), the standard components of the
correspondent irreducible representation split as follows:
i j
k

a b
c

a n
c

a n
n

a b
n

(3.4.3)

Notice that the index n can appear only at the bottom-end of each column. Since GL(n1, C)
does not see indices with value n, the components ac b , ac n , na n , na b correspond to the
representations

of GL(n 1, C), so that the following branching rule holds:


GL(n, C) GL(n 1, C) ,

(3.4.4)

In general, given an irreducible GL(n, C)representation, its branching with respect to


GL(n 1, C) is obtained by grouping the corresponding standard components according

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)

137

to the number and position of the boxes with the value n, keeping in mind that this value
can appear at most once at the bottom-end of each column. As a second example consider
the representation (3, 1) of GL(n, C):
GL(n, C) GL(n 1, C) ,
n

(3.4.5)

If () (1 , . . . , n ) is an irreducible GL(n, C)representation (as usual, if p < n, we define


i = 0 for p < i n), generalizing the above argument, one can show that it branches in
the following sum of GL(n 1, C)representations (01 , . . . , 0n1 ):
GL(n, C) GL(n 1, C) ,
M
(1 , . . . , n )
(01 , . . . , 0n1 ) ,

(3.4.6)

where the direct sum is extended over values of the indices 0a satisfying the following inequalities:
1 01 2 02 . . . 0n1 n .

3.5

(3.4.7)

Representations of Subgroups of GL(n, C)

So far we have been considering type (p, 0) GL(n, C)-tensors only. We may extend our
analysis to more general mixed tensors having p contravariant and q covariant indices. In
this case a first decomposition into lower-rank traceless tensors is described by eq. (2.9.30).
Within each subspace V (p`,q`) , traceless tensors are then further reduced by applying the
decomposition described in the previous sections to the upper and lower indices separately,
namely by applying Young symmetrizers acting on two kinds of indices. In particular irreducible mixed tensor T = (T i1 ...ip j1 ...jq ) will in general transform in a representation D()
0
with respect to the upper indices and in a representation D0 ( ) with respect to the lower
0
ones. The tensor as a whole, will then be acted on by the representation D() D0 ( ) ,
simply denoted by {(), (0 )}, of the group GL(n, C). For instance an irreducible tensor in
the representation {(2, 1), (3, 2, 1)} will have the form:
M i1 i2 ,i3 j1 j2 j3 ,j4 j5 ,j6 .

(3.5.1)

Let us now address the issue of branching GL(n, C)- irreducible representations with respect
to some of its subgroups. The action of an element g GL(n, C) on a type-(p, 0) tensor is
defined by the transformation:
(D() [g])i1 ...ip j1 ...jp = (D[g]i1 j1 D[g]ip jp )() ,

(3.5.2)

138

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

where the subscript () means that the indices i1 . . . ip are to be symmetrized/antisymmetrized


according to the tableau (). For instance the action on a type- (2, 0) antisymmetric tensor
(T ij ), belonging to the representation () = (1, 1) = , is given by:
(D

( )

[g])ij k` =

1
(D[g]i k D[g]j ` D[g]i ` D[g]j k ) .
2

(3.5.3)

From eq. (3.5.2) we see that the entries of D() [g] are homogeneous polynomials of degree p in
the entries of D[g] = (D[g]i j ). The representation D() is reducible with respect to a subgroup
G GL(n, C) if some of the non-vanishing entries of D() [g], or linear combinations of them,
vanish for all g G. Consider the subgroup G = GL(n, R). Recall that, from a standard
theorem of algebra, if a complex polynomial vanishes for all real values of its variables, it will
vanish identically. Therefore if D() were a reducible GL(n, R)-representation, one or more
entries of D() [g] would vanish for any real values of the variables D[g]i j (i.e. g GL(n, R)),
implying that the same entries would be zero for any GL(n, C) matrix. We conclude that
irreducible GL(n, C)representations are irreducible also with respect to GL(n, R).
Consider now the subgroup SL(n, C). The only condition satisfied by the entries D[g]i j
when g belongs to this subgroup is det(D[g]) 1. This does not imply the vanishing on any
entry of D() [g] on the subgroup. To show this suppose that some entry, to be denoted by
(p)
Pk [g], of D() [g] vanish for g SL(n, C). A generic nn matrix M can always be written as
1
the product of the nth root of its determinant times a unimodular matrix: M = m n S, where
m = det(M) and det(S) = 1. Therefore, for any g GL(n, C), there exists a g 0 SL(n, C)
1
(p)
such that D[g] = m n D[g 0 ], where m = det(D[g]). The entries of Pk [g], of D() [g] that
vanish on SL(n, C), are homogeneous polynomials of degree p in the entries of D[g]. This
implies that they will vanish on a generic g GL(n, C), since, using their homogeneity, we
p
(p)
(p)
have Pk [g] = m n Pk [g 0 ] = 0, being g 0 SL(n, C). We conclude that no entry of D() [g] can
vanish on the subgroup SL(n, C) only and thus that if D() is also irreducible with respect to
SL(n, C). Using the same theorem of algebra mentioned earlier, we can prove that irreducible
GL(n, C)representations are irreducible not just for SL(n, C) also with respect to SL(n, R).
Similar arguments can be used to show that the same holds true for the U(n) and SU(n)
subgroups of GL(n, C): Irreducible representations of the latter are irreducible also with
respect to the formers. As we shall see this is no longer true for the subgroups (S)O(p, q)
and Sp(n, C) (n even) of GL(n, C) which will be dealt with separately.
Consider now a totally antisymmetric type (p, 0)-tensor T (T i1 ...ip ). It spans the
irreducible representation () = (1 , 2 , . . . , p ) = (1, 1, . . . , 1) referred to the contravariant
indices, to be denoted by the symbol Dp (recall that D is the representation acting on
contravariant Vn vectors). With it we can associate a totally antisymmetric quantity T =
( Ti1 ...inp ) transforming in the product of the representation det(D) times the representation
of totally antisymmetric type-(0, n p) tensors, to be denoted by D0 (np) (recall that
D0 DT is the representation acting on covariant vectors (Vi )) and defined as follows:

Ti1 ...inp

def.

1
i ...i j ...j T j1 ...jp .
p! 1 np 1 p

(3.5.4)

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)

139

The above quantity will also be denoted by the short-hand notation: T  T. The correspondence between T and T is called dualization. Similarly if F = (Fi1 ...ip ) is a type-(0, p)
tensor transforming in the representation D0p , we can associate with it a type-(n p, 0)
tensor F using the contravariant symbol i1 ...in :

F i1 ...inp

def.

1 i1 ...inp j1 ...jp

Fj1 ...jp .
p!

(3.5.5)

To show that T transforms in the representation det(D)D0 (np) let us take a g GL(n, C)
and use the following matrix property:
D[g]i1 j1 D[g]in jn i1 ...in = det(D[g]) j1 ...jn .

(3.5.6)

The above property can also be represented symbolically as follows:


Dn [g]  = det(D[g])  .

(3.5.7)

Contracting the first n p indices on both sides of (3.5.6) by D[g]1 j` i` and recalling that
D0 [g] D[g]T we find:
i1 ...inp j1 ...jp D[g]j1 k1 D[g]jp kp = det(D[g]) D0 [g]i1 `1 D0 [g]inp `np `1 ...`np k1 ...kp .
(3.5.8)
Using the symbolic notation introduced above we will also write:
 Dp [g] = det(D[g]) D0(np) [g]  .

(3.5.9)

As T transforms into T0 (T 0 i1 ...ip ) Dp [g] T, defined by:


T 0 i1 ...ip = D[g]i1 j1 D[g]ip jp T j1 ...jp ,

(3.5.10)

The quantity T transforms into T0 whose components are given by:

1
1
i1 ...inp j1 ...jp T 0 j1 ...jp = i1 ...inp j1 ...jp D[g]j1 k1 D[g]jp kp T k1 ...kp =
p!
p!
1
det(D[g]) D0 [g]i1 `1 D0 [g]inp `np `1 ...`np k1 ...kp T k1 ...kp
=
p!
= det(D[g]) D0 [g]i1 `1 D0 [g]inp `np T`1 ...`np ,
(3.5.11)

Ti01 ...inp =

where we have used eq. (3.5.8). The above derivation in our symbolic notation is straightforward:

T0 =  T0 =  Dp [g] T = det(D[g]) D0(np) [g]  T = det(D[g]) D0(np) [g] T .

Similarly the dual F  F of a totally antisymmetric type-(0, p) tensor F, as defined in eq.


(3.5.5), transforms in the det(D)1 D(np) . This is easily shown by using the property
 D0 p [g] = det D[g]1 D(np) [g] , deduced from eq. (3.5.6) by contracting the last p indices

140

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

with D[g]1 j` i` . Quantities like T or F, transforming in the product of a power of det(D)


times a tensor representation, are called tensor densities.
If we restrict to SL(n, C) transformations we have det(D[g]) 1, i.e. det(D) is the identity representation, and T is then still a tensor transforming in the D0(np) representation.
The mapping (3.5.4) between type-(p, 0) and type-(0, n p) totally antisymmetric tensors,
T and T  T, is a homomorphism
  (asthe reader
 can verify) between two linear vector
n
n
spaces with the same dimension
=
, i.e. it is an isomorphism. We say then
p
np
that the corresponding SL(n, C)-representations Dp and D0 (np) are dual to each other
and are in fact equivalent. In particular this implies that Dn is equivalent to the identity
representation 1. Denoting the one-dimensional identity representation 1 also by and the
Young diagrams defining irreducible representations D0 () on covariant indices by crossed
boxes, we can write:

np
p
..
..

n
.
(3.5.12)
..

.
From this it follows that: In a Young diagram describing a SL(n, C)-representation we can
remove all n-box columns, since they amount to multiplying the irreducible representation
obtained deleting that column by the identity representation. In other words, attaching to
a Young diagram any number of n-box columns, we obtain equivalent representations:
(1 , . . . , n ) (1 + k, . . . , n + k)

(1 , . . . , n ) .

(3.5.13)

In tensor language, if a SL(n, C)-tensor is antisymmetric in n indices, such indices can be


thought of as belonging to an - tensor multiplying a lower rank tensor. For instance the
SL(3, C)-tensor M i1 i2 ,i3 i4 ,i5 , which is antisymmetric in i1 , i3 , i5 can be written in the following
factorized form:
M i1 i2 ,i3 i4 ,i5 i1 i3 i5 M i2 ,i4 .

(3.5.14)

The tensors M i1 i2 ,i3 i4 ,i5 and M i2 ,i4 define two equivalent SL(3, C)-representations. The tensor
M i1 i2 ,i3 i4 ,i5 is antisymmetric in the subsets of indices (i1 , i3 , i5 ) and (i2 , i4 ). We can think of
dualizing with respect to each subset to obtain a dual tensor transforming in the dual
representation:

Mk

1
1
ki2 i4 i1 i3 i5 M i1 i2 ,i3 i4 ,i5 .
2
3!

(3.5.15)

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)

141

Since Mk transforms as a covariant vector in , we can write

(3.5.16)

or simply (2, 2, 1) (1), where the symbol indicates that the equivalence is defined through
the duality operation.
We can generalize the above correspondence and associate with any SL(n, C) contravariant tensor () (1 , . . . n ), an equivalent covariant one, in the representation (0 )
(01 , . . . 0n ), obtained by dualizing the indices in each column of tableau (). We will call the
diagram (0 ) dual to () and write:
(1 , . . . n )

(01 , . . . 0n ) .

(3.5.17)

Let us define the relation between i and 0i . The reader can verify that a Young diagram
(1 , . . . n ) has n column with n boxes, n1 n columns with n 1 boxes, and so on
up to 1 2 one-box columns. The dual diagram will then have 1 2 = 0n1 0n
columns with n 1 boxes, 2 3 = 0n2 0n1 columns with n 2 boxes and so on up
to n1 n = 01 02 one-box columns. There will be no n-box columns, which implies
0n = 0. We have thus deduced the following relations
0n = 0 ,
1 2 = 0n1 0n ,
2 3 = 0n2 0n1 ,
..
.
n1 n = 01 02 ,

(3.5.18)

which are solved by 0i = 1 ni+1 , i = 1, . . . , n, so that:


(1 , . . . n )

(01 , . . . 0n ) = (1 n , . . . , 1 2 , 0) .

(3.5.19)

There is a geometrical way of constructing the dual of a given Young diagram. One should
first inscribe the diagram () in the upper half of a rectangle with the horizontal side consisting of 1 boxes and the vertical one of n-boxes. The complement of the diagram in the
rectangle is the dual diagram (0 ) upsidedown. Consider, for example, the contravariant
representation (7, 6, 5, 2, 1) of GL(7, C) and inscribe it in a 7 7 square, as illustrated below:

The dual diagram is (0 ) = (7, 7, 6, 5, 2, 1, 0).


What we have said for SL(n, C) also holds for its real form SL(n, R).

(3.5.20)

142

3.5.1

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Representations of U(n)

In the previous chapter we have learned that the complex conjugate of a contravariant vector
transforms under a unitary representation as a covariant one: (V i ) Wi . This is because
the unitarity condition implies D0 = D. This property generalizes easily to type-(p, q) tensors
T (T i1 ...ip j1 ...jq ): The complex conjugate of a type-(p, q) tensor is a type-(q, p) tensor:
(T i1 ...ip j1 ...jq ) = F j1 ...jq i1 ...ip .

(3.5.21)

In particular complex conjugation defines a pairing between covariant and contravariant


tensors which is however not an equivalence: The complex conjugate of a type-(p, 0) tensor
()
in the representation D() is a type-(0, p) one in the representation D0 () = D . If we
consider the special unitary group SU(n), which is a subgroup of SL(n, C), the duality
operation represents a further pairing between covariant and contravariant tensors and the
()
type-(0, p) tensor in D will be the dual of a type-(n p, 0) tensor in the representation
()
(1)
0
D( ) = D . For instance, in the SU(3) case D D0 (1) is equivalent through duality to
D(1,1) . In diagrams:

(3.5.22)

1
ijk M j,k .
2

(3.5.23)

or, in tensor language:


(V i ) = Wi =
The SU(3) representation (2, 1)
case we can write:

is self-dual, namely it coincides with its dual. In this

(3.5.24)

or, in tensor components:


(M i1 i2 ,i3 ) = Mi1 i2 ,i3 =

1
i1 i3 j2 i2 j1 j3 M j1 j2 ,j3 .
2 3!

(3.5.25)

For SU(2) the following relation holds:


(3.5.26)

or, in tensor components:


(V i ) = Wi = ij M j .

(3.5.27)

In fact SU(2) is the only special unitary group for which the fundamental representation
2=
is equivalent to its conjugate, as we have already shown earlier. Let us now observe

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)

143

that any Young diagram of SU(2) is equivalent to a single-row diagram (1 ), since any doublebox column can be erased. The corresponding SU(2)-representation is usually labeled by
the half-integer j 21 and denoted by D(j) , so that:
1

D( 2 )

, D(1)

, D( 2 )

(3.5.28)

The dimension of a representation D(j) is:





2 + 1 1
(j)
=
= 1 + 1 = 2 j + 1 .
dim D
1

(3.5.29)

A tensor in D(j) is totally symmetric in 1 = 2 j indices, each taking the values 1, 2: M i1 ...i2j .
0
A mixed tensor will transform with respect to SU(2) in a representation D(j) D0 (j ) and
will have the form: M i1 ...i2j j1 ...j2j0 . We can apply the rule defied in Section 3.3 to decompose
the product of two SU(2)-irreducible representations D(j1 ) and D(j2 ) . Consider for instance
the case j1 = 25 and j2 = 1. We have:

=
( 52 )

( 72 )

D(1) D

( 25 )

( 32 )

=
.

We can generalize the above result and write the following decomposition:
M
D(j1 ) D(j2 ) =
D(k) , k = |j1 j2 |, |j1 j2 | + 1, . . . , j1 + j2 .

(3.5.30)

(3.5.31)

We have discussed in the previous chapter the relation between SU(2) and the rotation group
SO(3) in the three-dimensional Euclidean space: The latter is somewhat contained twice in
the former. This implies in particular that SU(2) has more representations than SO(3). In
fact the two share only those representations with integer j (for instance the fundamental
two-dimensional representation of SU(2) is not an SO(3)-representation). In this sense SU(2)
represents a generalization of the three-dimensional rotation group. In quantum mechanics
physical systems with definite angular momentum are described by states transforming in
a SO(3)-representation defined by an integer value of j, which measures the corresponding
quantized value of angular momentum in units of ~. SU(2)-representations with half-integer
j describe an other physical property of a particle called spin, no longer related to spatial
rotations (it rather describes a kind of rotation in some internal space). By total angular momentum we shall refer to both ordinary angular momentum and spin, and it thus
corresponds to a representation D(j) of SU(2).
If we have two interacting systems with total angular momenta j1 ~ and j2 ~ respectively,
the total system will be described by a state transforming in the product of the representations describing each system, namely in D(j1 ) D(j2 ) . The decomposition (3.5.31) then
provides the irreducible representations in which the total system transforms and is called
sum rule of angular momenta.

144

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Whatever we said for (S)U(n) also applies to representations of (S)U(p, n p).


Let us consider irreducible representations of SL(3, R), SU(3) and SU(2, 1). The columns
of Young diagrams for these groups can consist of at most two boxes. However couples
of antisymmetric covariant (contravariant) indices can be dualized to single contravariant
(covariant) indices. Consider for instance the irreducible tensor M i1 i2 i3 i4 ,i5 i6 i7 . It has mixed
symmetry. However if we dualize the antisymmetric couples of indices we end up with a
tensor defining an equivalent representation, which is totally symmetric in its upper and
lower indices separately:
1
j i i j i i j i i M i1 i2 i3 i4 ,i5 i6 i7 .
8 115 226 337
We can represent graphically this transformation as follows:
M 0 i4 j1 j2 j3

(3.5.32)

(3.5.33)

In general we can state the following property:


Property 3.1:
All irreducible tensors with respect to the groups SL(3, R), SU(3) and
SU(2, 1) can be represented as mixed tensors which are totally symmetric in each kind of
indices. The representation can be therefore totally characterized by a couple of integers
(p, q) representing the number of contravariant and covariant indices respectively.

3.5.2

Representations of (S)O(p, q)

While the only operations which commute with the action of the GL(n, C) group on a tensor
with only one kind of indices (covariant or contravariant) are permutations on the order of
its indices, so that irreducible tensors are tensors on which a maximal set of symmetrization/antisymmetrization operations have been applied, when we restrict ourselves to the
action of the (S)O(p, q) group, there exists a further operation on tensors which commutes
with it: The contraction of two indices of the same kind. We have shown in the GL(n, C)
case that contractions of upper with lower indices in a tensor yield lower rank tensors which
define invariant subspaces. As far as the group O(n) is concerned, we have the invariant
tensors ij (type- (2, 0)) and ij (type- (0, 2)). Given a type-(p, q) tensor T = (T i1 ...ip j1 ...jq )
the contraction of two upper indices amounts to multiplying T by ij and then contracting
i and j with the two contravariant indices i` and ik :
T [`k]i1 ...i`1 i`+1 ...ik1 ik+1 ...ip j1 ...jq T i1 ...i`1 ii`+1 ...ik1 jik+1 ...ip j1 ...jq ij =
n
X
=
T i1 ...i`1 ii`+1 ...ik1 iik+1 ...ip j1 ...jq .
i=1
[`k]

The quantity T is a type- (p 2, q) tensor, as we can easily show by transforming T under


g O(n) (let us denote D[g] simply by D):
T [`k]i1 ...ip2 j1 ...jq

T [`k]0 i1 ...ip2 j1 ...jq =

n
X
i=1

T 0i1 ...i`1 ii` ...ik2 iik ...ip2 j1 ...jq =

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)


=
=

n
X
i=1
i1

145

D1 j1 m1 D1 jq mq T k1 ...m...s...kp2 m1 ...mq =
Di1 k1 Di m Di s Dkp2
p2
i

k1

D1 j1 m1 D1 jq mq T [`k]k1 ...kp2 m1 ...mq ,


Dkp2
p2

Pn
i
i
where we have used the property of orthogonal matrices:
i=1 D m D s = ms . Similarly,
using the invariant tensor ij we can trace over two lower indices obtaining a type-(p, q 2)
tensor T[`k] :
T[`k] i1 ...ip j1 ...jq2

n
X

T i1 ...ip j1 ...j`1 jj` ...jk2 jjk ...jq2 .

(3.5.34)

j=1

There are p(p 1)/2 independent traces T[`k] and q(q 1)/2 traces T[`k] that can be constructed out of a type-(p, q) tensor T. These tensors can be further traced, so that the space
V (p,q) naturally splits into the direct sum of subspaces V (p2`,q2k) which are invariant under
O(n) and whose elements have the form:
`

V (p2`,q2k) = {F (p2`) (q2k)

k
z }| { z }|
{
(2) (2) (2) (2) } V (p,q) ,

(3.5.35)

where F (p2`) (q2k) denotes a traceless type (p2`, q 2k) tensor, (2) = ( ij ) and (2) = (ij ).
A type- (p, q) tensor can then be written, in a unique way, as the sum of its components on
each of these subspaces. Let us consider for instance a type-(3, 0) tensor T = (T ijk ). It will
be written as the sum of a traceless tensor of the same type T0 = (T0ijk ) and a component
in V (1,0) :
T ijk = T0ijk + T1ijk ,
T1ijk = H i jk + F j ik + Gk ij ,

(3.5.36)

where H i , F i , Gi are expressed in terms of the traces of T through the conditions:


T

[12] i

n
X

T kki = H i + F i + n Gi ,

k=1

T [13] i =

n
X

T kik = H i + n F i + Gi ,

k=1

T [23] i =

n
X

T ikk = n H i + F i + Gi ,

(3.5.37)

k=1

which yield:
1
[(n + 1) T [23] i T [13] i T [12] i ] ,
(n 1)(n + 2)
1
=
[T [23] i + (n + 1) T [13] i T [12] i ] ,
(n 1)(n + 2)
1
=
[T [23] i T [13] i + (n + 1) T [12] i ] ,
(n 1)(n + 2)

Hi =
Fi
Gi

(3.5.38)

146

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

The components T0 and T1 are mutually orthogonal in the sense that:


n X
n X
n
X

T0ijk T1ijk = 0 .

(3.5.39)

i=1 j=1 k=1

We have defined so far two kinds of traces or contractions: One between upper and lower
indices which commutes with GL(n, C) transformations and an other involving couples of
indices of the same kind (both upper or lower), which commutes with the action of the
orthogonal group only. When reducing a type-(p, q) tensor into irreducible components with
respect to O(n), we first need to restrict to components which are traceless with respect
to the first kind of contraction, then we further decompose such components into tensors
which are traceless with respect to the second kind of contraction as well. We shall call
traceless a tensor on which both kinds of contractions are zero. On these tensors we can
apply symetrization and anti-symmetrization operations defined by Young tableaux (Young
symmetrizers), as we did for the GL(n, C)-tensors, to derive the irreducible components. Of
course this procedure makes sense since any permutation of indices takes a traceless tensor
into an other traceless one.
Example 3.3: Let us apply this procedure to an example that we have already worked
out in the previous chapter. Consider a type-(2, 0) tensor F = (F ij ) and decompose it in a
traceless and trace part:
F ij = F0ij + F ij ,
(3.5.40)
Pn
P
ii
where ni=1 F0ii = 0. We easily find F = n1
i=1 F . Next we apply to the traceless part,
2
in the representation
= D , symmetrization and anti-symmetrization operators to
construct its irreducible components according to the decomposition:

F0ij

1 ij
1
(F0 + F0ji ) + (F0ij F0ji ) = FSij + FAij .
2
2

(3.5.41)

The O(n) irreducible components of F are a singlet F in the representation 1, the trace, FSij
in the representation n (n + 1)/2 1 and FAij in the representation n (n 1)/2.
Example 3.4: Let us consider a type-(2, 1) tensor T = (T ij k ). Let us first decompose it
into a traceless and trace part:
T ij k = T0ij k + F i kj + H j ki + Gk ij .

(3.5.42)

Exercise 3.10: Find F i , H i , Gi in terms of the traces of T.


The traces F i , H i , Gi can not be further reduced with respect to O(n) and thus define
three representations n (recall that for orthogonal representations upper and lower indices
). Consider now the traceless part T0ij k .
are on an equal footing since D = D and thus n = n

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)

147

In applying symmetrization and anti-symmetrization operations we should recall that with


respect to the orthogonal group upper and lower indices of the fundamental representation n
are on an equal footing and thus the problem is the same as that of decomposing a traceless
tensor T0ijk into its irreducible components. We find:
T0

(3.5.43)

where now the dimensions of each of the above O(n) irreducible representations is obtained
from the dimension of the corresponding GL(n, C) by subtracting a the dimensions of traces.
Take, for instance, the rank three symmetric GL(n, C) tensor M ijk P
transforming according
n
iij
. It contains a O(n)-invariant trace
transforming
to the representation
i=1 M
in the fundamental representation n. The dimension of the totally symmetric irreducible
O(n)-representation is therefore given by the dimension of the corresponding GL(n, C)representation minus n.PSimilarly, for the gun diagram, a GL(n, C)-tensor M ij,k contains
a O(n)-invariant trace ni=1 M ii,j in the representation n. We can therefore determine the
dimensions of the O(n)-irreducible representations in (3.5.43):
n(n + 1)(n + 2)
n,
] =
6
h
i
n(n2 1)
dim
n,
=
3
" #
n(n 1)(n 2)
dim
=
,
6

dim [

(3.5.44)
where we have used the fact that a totally anti-symmetric tensor has no trace part. We can
now write the decomposition of a type-(2, 1) tensor. It is nothing but the decomposition of
the product of three fundamentals n n n = :



 

n(n2 1)
n(n + 1)(n + 2)
n(n 1)(n 2)
nnn
n
2
n
6
6
3
3 n ,


where the 3 n part act on the traces F i , H i , Gi .


When we write the Young diagrams for the traceless part of a tensor, the following
important property, which we are not going to prove, should be taken into account:
Property 3.2: The Young element, defined by a diagram in which the sum of the lengths
of the first two columns exceeds n, when applied to a traceless tensor, gives zero.

148

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

In the decomposition of the traceless part of a tensor, the only irreducible components
which appear correspond therefore to admissible diagrams in which the total number of
boxes in the first two columns does not exceed n. For example
is not admissible for
O(3) and

is not admissible for O(2).

Admissible diagrams can be grouped in pairs of associate diagrams, Y and Y 0 , in which


the length a of the first column of Y does not exceed n2 boxes, a n/2, and the first column
of Y 0 consists of b = n a boxes and all the other columns coincide with those of Y . Let c be
the length on boxes of the second column of both diagrams. We have that n/2 a c and
therefore a + c n, namely Y is admissible. On the other hand a c so that n a n c
and thus (n a) + c n, namely Y 0 is admissible as well. Examples of associate diagrams
are, for O(5):
Y

Y0 =

Y0 =

(3.5.45)

If n is even, n = 2 , we may have diagrams Y which coincide with Y 0 . These are called
self-associate diagrams and their first column consists of boxes.
To construct admissible diagrams we can first construct Y and then its associate counterpart Y 0 . Since Y does not contain more P
than n/2 rows, it can be characterized by integers
(1 , . . . , ), 1 2 . . . and i=1 i = p, p being the rank of the tensor (total
number of indices), where n = 2 if n is even and n = 2 + 1 if n is odd. The integers i
are nothing but the lengths in boxes of each row of Y . To any solution to the equation:
1 + 2 + . . . + = p ,

(3.5.46)

there corresponds a diagram Y . Knowing Y we determine Y 0 .


The relevance of this construction is apparent is we consider the group SO(n) which, as
opposed to O(n), is unimodular, namely is a subgroup of SL(n, R). With respect to this group
the dualization operation maps tensors into tensors and defines a correspondence between
equivalent representations. Since tensors in Y 0 are obtained from those in Y by dualizing
over the indices in the first column, with respect to SO(n) associate diagrams are equivalent
representations. Self-associate diagrams yield two non-equivalent representations. Consider
for instance a totally antisymmetric rank tensor F = (F i1 ...i ) with respect to SO(2). Since
this group makes no difference between upper and lower indices, covariant and contravariant
tensors belong to the same representation space. In particular the dualization operation maps
rank- tensors into rank- tensors. Moreover F = F, namely the operation is involutive.
This means that the space of totally anti-symmetric rank- tensors splits into two eigenspaces of of equal dimension, to the eigenvalues 1. These sub-spaces are spanned by the
following components of F:
1
(F F) ,
F()
2

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)


F

() i1 ...i



1 i1 ...i j1 ...j
i1 ...i
F

Fj1 ...j .
!

149
(3.5.47)

The two eigen-spaces are invariant under the action of SO(n) and define two inequivalent
irreducible representations. F() are called self-dual and anti-self-dual components of F.
All that we said for O(n) applies to the more general O(p, q) groups. The only difference
is that traces are defined using the invariant metric p,q :
T [`k]i1 ...i`1 i`+1 ...ik1 ik+1 ...ip j1 ...jq T i1 ...i`1 ii`+1 ...ik1 jik+1 ...ip j1 ...jq (p,q )ij =
p
X
T i1 ...i`1 ii`+1 ...ik1 iik+1 ...ip j1 ...jq
=
i=1

n
X

T i1 ...i`1 ii`+1 ...ik1 iik+1 ...ip j1 ...jq ,

i=p+1

the same for the traces T[`k] of a covariant tensor.


Consider the group SO(3). In this case = 1 and thus all representations are defined
by a single integer 1 , which we shall also denote by j, and all diagrams Y are single-row.
This means that the representation space of irreducible SO(3)-representations D(1 ) = D(j)
is described by totally symmetric tensors. Consider a generic irreducible representation,
described
bythe single jbox row. The corresponding GL(3, R) representation has dimension

3+j1
= (j+2)(j+1)
and acts on totally symmetric rank-1 tensors F = (F i1 ...ij ). To
2
j
obtain the dimension of the SO(3) representation we have to subtract to this number the
contribution of the traces. A generic trace F[`k] is a rank j 2 totally antisymmetric tensor
and thus its vanishing implies j (j1)
conditions on the entries of F. We conclude that the
2
dimension of the generic SO(3)-irreducible representation (j) is:
dim D(j)

(j + 2)(j + 1) j (j 1)

= 2j + 1.
2
2

(3.5.48)

A vector V = (V i ) in V3 is acted on by the representation j = 1 of SO(3), of dimension 3.


The traceless product of the coordinates of a particle, xi xj 31 ij krk2 , is a rank-2 tensor,
corresponding to the representation j = 2 of dimension 5.
Relation between SO(3) and SU(2) representations. Now we can make our previous
statements about the relation between SO(3) and SU(2) representations more precise. Let
us consider the Pauli matrices i = [(i )a b ], i = 1, 2, 3, a, b = 1, 2, defined in (2.1.37). We can
lower the first index of these matrices using the ab tensor and define the three symmetric
matrices (i )ab ac (i )c b :
(1 )ab = i  1 = i 3 ,
(2 )ab = i  2 = 12 ,
(3 )ab = i  3 = i 1 ,

150

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Now consider a type-(0, 2) symmetric tensor F = (Fab ) in the representation D0 (j=1) = .


F is a complex 2 2 matrix on whose complex entries we can impose a linear condition
which commutes with the action of SU(2). This condition reduces the space spanned by
tensors of this type to a lower dimensional SU(2)-invariant sub-space on which the irreducible
representation acts. Recalling that (Fab ) is a type-(0, 2) tensor with two upper indices, this
condition reads:
Fab = ac bd (Fcd ) F =  F  .

(3.5.49)

The most general tensor satisfying the above reality condition has the following form:


f + ih
ie
F =
,
(3.5.50)
ie
f ih
where e, f, h are real parameters. Therefore tensors F satisfying (3.5.49) depend on 3 real
parameters and thus span a three-dimensional linear vector space over the real numbers (( i )
being a basis of such space), namely the corresponding representation is three dimensional
and real. Let us show that this representation is the fundamental 3 of SO(3). We notice
that Fab in (3.5.50) can be expanded in the symmetric matrices ( i )ab :
Fab = F i (i )ab = h (1 )ab + f (2 )ab + e (3 )ab ,

(3.5.51)

or, in matrix notation:


F = i F i  i .

(3.5.52)

Let now S = S[, , ] = (S a b ) be a generic SU(2)-transformation. The tensor F transforms


as follows:
0
Fab Fab
= S 1 c a S 1 d b Fcd ,

(3.5.53)

F F0 = ST F S1 .

(3.5.54)

F0 = i F i ST  i S1 .

(3.5.55)

or, in matrix notation:

Using (3.5.52) we can write:

Now we use the property (2.4.18) which, taking into account that  = i 2 , reads: S =  S .
From this it follows that  S = ST , where we have used the unitarity property of S, and
we can write:
F0 = i Fi  S i S1 = i Fi  (R1 i j j ) = (Rj i F i ) j = F 0 j j .

(3.5.56)

Where R = R[S] = (Ri j ) is the orthogonal transformation corresponding to the unitary


matrix S, as defined in Section 2.3.7, and we have used the orthogonality property R1 = RT ,

3.5. REPRESENTATIONS OF SUBGROUPS OF GL(N, C)

151

which, in components, reads R1 i j = RT i j = Rj i . We see that the components F i of


F transform in the fundamental 3 of SO(3). In other words the matrices (i )ab define a
change of basis between the two-times symmetric representation of SU(2) and the defining
representation of SO(3) so that the two representations coincide:
of SU(2)

of SO(3) .

(3.5.57)

From this we deduce that all the SU(2)-representations D(j) , defined by integer j, coincide
with the corresponding representations of SO(3).

152

CHAPTER 3. CONSTRUCTING REPRESENTATIONS

Chapter 4
References
Lie Groups, Lie Algebras & Some of Their Applications , Robert Gilmore, John Wiley
and Sons, 1974
Group Theory and its Application to Physical Problems , Morton Hamermesh, AddisonWesley Series in Physics.
Prof. t Hoofts Lecture Notes on Discrete Groups,
http://www.phys.uu.nl/ sahlmann/teaching/lecture notes/discrete groups lecture notes.pdf

Symmetry, Hermann Weyl, Princeton Science Library.


The Theory of Groups and Quantum Mechanics, Hermann Weyl, Dover.

153

You might also like