You are on page 1of 252

G R O U P T H E O RY

AND
PHYSICS
02/10/2017, 14:55
Fisica Teorica – 2012/2013
N O TA

Queste note sono pensate come supporto didattico per il corso di Fisica Teorica
(Parte A) del Corso di Laurea Magistrale in Fisica dell’Università degli Studi di
Padova. E’ vietata la loro distribuzione non autorizzata.
CONTENTS

i introduction 1
1 symmetries in physics 3
1.1 Introduction 4
1.2 Symmetries and Groups 4
1.3 Examples of Symmetries 5
1.3.1 Symmetry of Geometric Objects 5
1.3.2 Spacetime Symmetries 6
1.3.3 Gauge Symmetries 7
1.3.4 Approximate Symmetries 9
1.4 Symmetry Breaking 10
2 principle of least action 13
2.1 Euler–Lagrange Equations 13
2.2 Examples 16
2.2.1 Free relativistic particle 16
2.2.2 Electromagnetic field and its coupling to a charged parti-
cle 18
2.3 Noether’s Theorem 19

ii finite groups 23
3 groups: basic elements 25
3.1 Definitions 25
3.2 Maps between groups 27
3.3 Finite groups 29
3.3.1 The Symmetric Group Sn 31
4 subgroups, quotients and products 37
4.1 Basic definitions 37
4.2 Quotients 38
4.3 Products 42
4.4 Classifying finite groups. 44
5 representations 47
5.1 Basic definitions 48
5.2 Equivalent and reducible representations 51

i
ii Contents

5.3 Unitary representations 54


5.4 Comments 57
6 properties of irreducible representations 61
6.1 The Great Orthogonality Theorem 63
7 properties of irreducible representations 67
7.1 Conjugacy classes 67
7.2 Characters and their properties 68
7.3 Character tables 71
8 irreducible representations of Sn 77
8.1 Young Tableaux and conjugacy classes 78
8.2 Young Tableaux and irreducible representations of Sn 80

iii elements of topology and differential geometry 87


9 homotopy groups 89
9.1 Homotopy 90
9.2 Homotopy groups 92
9.3 Universal covering 95
10 differentiable manifolds 97
10.1 Introduction 97
10.1.1 Differentiable manifolds 97
10.1.2 Properties 99
10.1.3 Examples 100
10.2 Tangent and cotangent space 102
10.3 Flows and exponential map 106
10.4 Lie brackets 108

iv lie groups 111


11 lie groups 113
11.1 Compact and non-compact groups 114
11.2 Invariant measure 114
11.3 Groups of matrices 115
11.3.1 Linear constraints 116
11.3.2 Quadratic constraints 117
12 lie algebras 123
12.1 Algebra as tangent space 124
12.1.1 Algebras for groups of matrices 126
12.2 From the algebra g to the group G 128
12.2.1 Surjectivity 128
Contents iii

12.2.2 Injectivity and global properties 129


12.2.3 The algebra structure and the Baker–Campbell–Hausdorff
formula 130
12.2.4 Lie algebras 131
13 lie algebras – properties 135
13.1 Subalgebras, ideals 137
13.2 Simplicity, compactness and connectedness 138
13.3 Cartan’s criteria 139
13.4 Casimir operator 141
14 algebra representations 143
14.1 Morphisms and representations 143
14.2 Matrix representations 144
14.3 Differential representations 145
14.4 Bosonic representations 147
14.4.1 u(1)C ' gl(1, C) 148
14.4.2 u(2)C ' gl(2, C) 149
14.5 Fermionic representations 150
15 representations of su(2) and so(3) 153
15.1 Rotation group and its algebra 153
15.2 Isomorphism su(2) ' so(3) and homomorphism SU(2) 7→ SO(3) 154
15.3 SO(3) topology 157
15.4 Irreducible representations of SU(2) and SO(3) 159
15.5 Irreducible representations of su(2) 160
15.6 Matrix representations 163
16 su(n) irreps and products 165
16.1 Products of irreducible representations 165
16.1.1 Tensor products 165
16.1.2 Clebsch–Gordan decomposition 166
16.1.3 Decomposition in irrepses of a subgroup 167
16.2 SU(N) irrepses and Young Tableaux 167
16.2.1 Fundamental, anti-fundamental and adjoint representa-
tions 167
16.2.2 Irrepses and Young Tableaux 170
16.2.3 Products using Young Tableaux 176
16.2.4 Again on SU(2) 178
17 poincar é group 183
17.1 Topological properties 184
iv Contents

17.1.1 Topology of SL(2,C) and SO+ (3,1) 188


17.2 The so+ (3, 1) algebra 189
18 poincar é irrepses 193
18.1 Unitary representations 195
18.2 Finite dimensional representations 199
18.3 Reducibility of spinor representations 201
18.4 Casimir operators 204
18.4.1 Casimir for the Lorentz group 204
18.4.2 Casimir for the Poincaré group 205
19 symmetries of molecules and solids 209
19.1 Classic vibrations of molecules 209
19.1.1 Symmetries 211
19.1.2 Diatomic molecule 213
19.1.3 Diatomic molecule in 2 dimensions 215
19.1.4 Triatomic molecule 217
a solutions 219
Part I

INTRODUCTION
1
SYMMETRIES IN PHYSICS

2012 has been a remarkable year for physics. The discovery of the Higgs boson-
like particle at the Large Hadron Collider in Geneva provides a spectacular
confirmation of our understanding of the fundamental interactions governing
our Universe, embodied in the Standard Model of elementary particles. This
discovery marks the pinnacle of the journey started with the unification of elec-
tricity and magnetism in terms of a single symmetry principle and completed
with the further unification of the weak nuclear force, via the electro-weak sym-
metry breaking mechanism.
Symmetry principles guide theoretical physics since its birth. Geometric
symmetries can explain many physical properties of materials and character-
ize diffraction patterns found in crystallography. Dynamical symmetries are
widely used in Nuclear Physics to classify the energy of rotational and vibra-
tional modes of nuclei. Symmetry properties of spacetime itself and symme-
try transformations relating different observers are at the base of Special and
General Relativity. Broken symmetries are used to explain a wide variety of
physical phenomena, from Superconductivity to electro-weak interactions.
In this lectures we want to explore the concept of symmetry in physics. In par-
ticular we want to give a theoretical description of this concept, by construct-
ing mathematical models of the symmetries appearing in physical systems. We
are mainly going to focus on the applications of symmetry principles in the
context of Quantum Mechanics, but we are also going to discuss spacetime
symmetries, which are at the base of Relativity and Quantum Field Theory,
and internal symmetries (global and local gauge symmetries), which guide our
current understanding of the physics of fundamental interactions.
The objective is to provide students with instruments and techniques that
are useful to extract information on the physics of a definite system by using

3
4 symmetries in physics

symmetry arguments. This is not a course in group theory and/or differen-


tial geometry and as such we do not seek completeness nor rigor at the level
of a course on these topics. We will introduce only the ingredients that are
necessary for physical applications and try to provide as many examples and
exercises derived from physics as possible.

1.1 introduction

Man is naturally attracted by symmetric objects. Actually, the word symmetry


derives from συν − µéτρoν and its adjective is usually associated to proportion-
ate objects, displaying equal measures. In everyday life we often identify the
word symmetry with bilateral symmetry. This is a symmetry under reflection
with respect to a plane, i.e. invariance under exchange of its left and right sides.
The fact that such a symmetry is so common when we look at men and ani-
mals has to do with the existence of 2 preferred axis, a vertical one determined
by gravity and a horizontal one, determined by the direction of motion. Their
combination singles out a preferred plane in space and explains the consequent
symmetry of our bodies with respect to such a plane. Plants do not move and
in fact they usually display a rotational symmetry around the vertical axis. As
we know, these are not exact symmetries, but rather approximate symmetries. Still,
we are able to somehow predict their existence because of the preferred plane
explained above and their existence explains our appearance.
How do we recognize symmetries? An object is symmetric if we can act on
it with an operation that does not change it with respect to our observation.
We can therefore argue that a symmetry transformation is a map that leaves an
“object” invariant. In general, if we apply a certain transformation rule to a
physical system, we put the system in a new and different state. If we are not
able to distinguish it from the previous one, we say that the transformation we
used is a symmetry.

1.2 symmetries and groups

We will use a great part of these lectures to discuss group theory. We will do
so because symmetry transformations are described by groups.
Consider three different states of a physical system: A, B and C. If state A
is similar to state B and state B is similar to state C, then state A is similar
to state C. Hence, the product of two symmetry transformations is again a
1.3 examples of symmetries 5

symmetry transformation. Also, each state is mapped onto itself by a trivial


transformation that leaves everything the same (the identity) and the inverse of
a symmetry transformation is still a symmetry. Thus, the set of all symmetry
transformations that characterize the symmetry of a system, are elements of a
group.

id S1 id S2 id
A B C
S1 ◦ S2

Figure 1.: Example of symmetry transformations between three different physical


states A, B and C.

For example, reflections with respect to a plane form a group that contains
two elements: the reflection operation and the identity. Rotations in three-
dimensional space, on the other hand, form a group that has an infinite number
of elements. Groups with a finite (or numerable) number of elements are called
discrete groups, while groups that depend on continuous parameters are called
continuous groups.
Symmetries of a system imply relations between observable quantities, which
may be obeyed with great precision, independently of the nature of the forces
acting in the system. For example, the energies of several different states of
the Hydrogen atom are exactly equal, as a consequence of the rotational invari-
ance of the system. In other instances symmetries of physical systems are only
approximately realized or broken. Still, an approximate or a broken symmetry
can teach us as many important lessons as an exact one, like in the case of
superconductivity or of the Higgs particle.

1.3 examples of symmetries

1.3.1 Symmetry of Geometric Objects

The first obvious example is given by symmetries of geometric objects. A


generic geometric object may display symmetry under rotations and/or reflec-
tions and these transformations may generate a discrete or a continuous group.
6 symmetries in physics

For instance, we know that an equilateral triangle is mapped to itself whenever


we perform a rotation of 120 degrees around its center, or a reflection around
one of three axes connecting the center to one of its vertices. A snowflake looks
like itself when we perform a 60 degrees rotation around its center or if we per-
form a mirror reflection. A circle, on the other hand, is mapped to itself by any
rotation around its center. This means that in this case we have a continuous
set of symmetry transformations (or isometries).

Figure 2.: Symmetry axes of geometric objects.

Geometric symmetries affect physical properties of molecules and materials,


like the appearance of conducting and isolating bands in crystals.

1.3.2 Spacetime Symmetries

As mentioned in the introduction, very important symmetry laws in physics are


the transformations of spacetime coordinates that map intertial frames among
them. These laws constrain both kinematic and dynamic properties of physical
systems and can be derived from general properties of spacetime. In Newto-
nian mechanics, such transformations change the coordinate system (t,~r ) by

• spatial translations:

~r → ~r +~a; (1.1)

• time translations:

t → t + b; (1.2)

• constant rotations:

~r → R~r; (1.3)
1.3 examples of symmetries 7

• relative motion at constant velocity ~v (Galilean boosts):

~r → ~r − ~v t. (1.4)

The first two symmetries derive from space and time homogeneity, while the
third one is related to space isotropy. Altogether they form a 10-dimensional
continuous group called the Galilei group, which constrains newtonian mechan-
ics. In Special Relativity Galilean boosts are replaced by Lorentz boosts and
together with the other transformations form the Poincaré Group, which we
will discuss in detail in chapter 17.
This type of symmetries is extremely important because they severely restrict
the form that physical laws can take. Also, as we will soon discuss, continuous
symmetries give rise to conserved quantities via Noether’s theorem and in the
case of spacetime symmetries we can obtain energy-momentum and angular
momentum conservation laws.

1.3.3 Gauge Symmetries

A special kind of symmetry transformations we encountered when studying


electrodynamics are gauge symmetries. They are coordinate-dependent trans-
formations that leave the equations of motion invariant. In fact, Maxwell’s
equations can be written as
4π ν
∂µ F µν = − j , eµνρσ ∂ν Fρσ = 0, (1.5)
c
where the electric and magnetic fields are described by the rank 2 antisymmet-
ric tensor Fµν . The second equation in (1.5) can be solved in terms of a potential
Aµ , by

Fµν = ∂µ Aν − ∂ν Aµ . (1.6)

However this solution is not unique because gauge transformations by an arbi-


trary function Λ( x ) change the Aµ potential as

A µ ( x ) → A µ ( x ) + ∂ µ Λ ( x ), (1.7)

but leave Fµν invariant and hence do not change the value of the electric and
magnetic fields. As we will now see, this function is actually specifying an
element of a group of symmetries of the Schrödinger equation for a charged
particle.
8 symmetries in physics

An electrically charged particle is described by a complex wave function


ψ(t, ~x ) whose Schrödinger equation remains valid if we perform a local rotation
in the complex plane

ψ( x ) → eiΛ( x) ψ( x ), (1.8)

provided the wavefunction is minimally coupled to the Aµ potential. This


means that each derivative acting on ψ has to be replaced by a covariant one

∂ µ → Dµ = ∂ µ − i A µ . (1.9)

In fact, if we do so, the Schrödinger equation is mapped to itself under gauge


transformations because

Dµ ψ 7→ eiΛ Dµ ψ. (1.10)

The set of all possible phases

U = exp(iΛ( x )) (1.11)

forms a group of 1 by 1 unitary matrices called U(1):

U†U = 1 ⇔ U ∈ U(1). (1.12)

Since a symmetry group of this type has an infinite number of elements,


whose value may differ at each spacetime event, the couplings and the dy-
namics of particles charged with respect to it are extremely restricted. Gauge
symmetries are therefore extremely powerful symmetry principles.
Fundamental interactions are nowadays described by generalizations of the
quantum field theory describing electrodynamics, where the internal symme-
try group U(1) has been replaced by larger ones. The Standard Model of ele-
mentary particles is based on the gauge group SU(3) × SU(2) × U(1). Great
unification theories use larger groups like SU(5) and SO(10) and String Theory
makes use of even larger groups like E8 × E8 .

Gravitational interactions, on the other hand, are described by General Rela-


tivity. Although this theory is rather different from ordinary gauge theories, it
still shows a local invariance that determines its dynamics.
Physics should be described by the same laws in any reference frame. For
this reason we should be able to describe them in a way that preserves their
form under general coordinate transformations:

x µ → ξ µ ( x ). (1.13)
1.3 examples of symmetries 9

This general covariance is at the base of Einstein’s theory of Relativity and


the role of “gauge field” for such transformations is played by the metric field
gµν ( x ).

1.3.4 Approximate Symmetries

We already discussed the fact that often approximate symmetries are useful to
constrain the theoretical description of physical systems. These are symmetries
valid in a certain regime of energy or length and may arise dynamically.
An example is given by crystals. An infinite crystal is invariant under trans-
lations for which the displacement is an integral multiple of the distance be-
tween two adjacent atoms. Real crystal, however, have a definite size and their
surface perturbs the symmetry under translations. Nevertheless, such pertur-
bations have little effects on the properties at the interior if the crystal contains
a sufficiently large number of atoms.
Another example of a symmetry that is only approximately realized is the
isospin symmetry. One of the baryons, called ∆+ , decays into a nucleon and a
pion in two ways:
∆+ → n + π + , or ∆+ → p+ + π 0 . (1.14)
Although the masses of the neutron and the proton are very close and the same
is true for the pions, so that the only distinctive element is the electric charge
distribution, the second process occurs twice as much as the first. The explana-
tion for this discrepancy comes from a symmetry argument. In fact nucleons
and pions sit in different representations of the isospin symmetry group. This
information is actually relevant for any symmetry principle. The way physical
quantities transform under symmetry transformations is not determined just
by the symmetry group, but also by its representation. For instance, we know
that we cannot sum mass and velocity as they are quantities that transform
differently under the group of rotations in 3-dimensional space. The mass is
an invariant quantity. It is a scalar under rotations and hence transforms in
the trivial representation. The velocity transforms as a vector and hence the
group of rotation is represented in its fundamental representation. Coming
back to the example above, the two decays mode differ because nucleons form
a doublet of the isospin symmetry, pions a triplet and ∆ particles a quadruplet
and hence they are constrained by the rules governing products of representa-
tions and their decomposition in irreducible components. We stress that this
argument is valid because particles inside the same representation have nearly
10 symmetries in physics

equal masses, but different electric charges. Symmetry arguments are therefore
valid in the approximation of equal masses, which we know is violated only by
few percents.
We conclude this discussion of approximate symmetries by recalling that
we have a very simple example of a discrete symmetry that is omnipresent in
physics: the reflection symmetry Z2 : x 7→ − x. Any degree of freedom can
be approximately described by the 1-dimensional harmonic oscillator when
looking at a minimum of its potential

1
V ( x ) ' V ( x0 ) + m2 δx2 + O(δx3 ) (1.15)
2

and obviously this is invariant for δx 7→ −δx.

1.4 symmetry breaking

Consider now a situation where we have a symmetry principle that is valid in


a certain energy regime, but which is (spontaneously) broken in another. We
can still have essential information on the description of such physical system
in the regime where the symmetry is broken by knowing that it was valid in
the other.
Let us consider for example a system described by 2 complex degrees of
freedom zi , i = 1, 2, whose potential is given by

 2
V ( z i ) = − µ2 Z † Z + λ Z † Z , (1.16)

 
z1
where λ > 0 and Z = . It is easy to check that such a potential is
z2
invariant under unitary transformations of the Z doublet

Z 7→ UZ, where U † U = 12 (U ∈ U(2)) (1.17)

and therefore the resulting physics should keep trace of this fact. We can also
see that the potential (1.16) has two critical points, satisfying
 
∂i Z = −µ2 + 2λ| Z |2 zi∗ = 0, (1.18)
1.4 symmetry breaking 11

at zi = 0, or at | Z |2 = µ2 /(2λ). Actually, zi = 0 is a maximum and hence an


unstable point, while | Z |2 = µ2 /(2λ) is a minimum. Physics at the maximum
is still invariant under the whole symmetry group because
    
a b 0 0
= (1.19)
c d 0 0
for any value of a, b, c and d and hence for any U ∈U(2).

Figure 3.: The “mexican hat” potential (1.16). The maximum has U(2) symmetry, while
the family of minima have only a residual U(1).

On the other hand, the family of minima is invariant only with respect to a
subset of such matrices, forming
p a U(1) subgroup. In fact, chosen for instance
the point z1 = 0 and z2 = σ = µ2 /(2λ) we have that
    
a b 0 bσ
= ⇒ b = 0, d = 1. (1.20)
c d σ dσ

In addition we have to remember that U † U = 1 and hence


 ∗
a c∗ | a |2 + | c |2 c ∗
     
a 0 1 0
= = , (1.21)
0 1 c 1 c 1 0 1

which implies c = 0 and a = eiα . We therefore have a spontaneous breaking of the


symmetry group

U(2) → U(1). (1.22)

This means that if we study the system around the stable minima, we explic-
itly see only a U(1) invariance, but we still have constraints on the possible
12 symmetries in physics

interaction terms due to the original larger U(2) symmetry group. For instance,
expanding the potential (1.16) around the minima we see that the symmetry
breaking terms are not generic, but fixed by the original U(2)-invariant poten-
tial.
2
THE PRINCIPLE OF LEAST ACTION AND NOETHER’S
THEOREM

In this chapter we are going to derive the Euler–Lagrange equations for con-
tinuous systems and examine the relation between conservation laws and the
existence of continuous symmetries. After a brief general discussion we will
concentrate on relativistic systems.

2.1 euler–lagrange equations

The principle of least action is a very elegant formalism that can be used to
derive the equations of motion of a physical system, starting from a functional,
the action, which is manifestly invariant under a definite set of symmetries.
Imagine we throw an object in the Earth’s gravitational field. We know that in
the presence of gravity the trajectory will be a parabola connecting two extrema
A and B. What is the distinctive feature of such a trajectory? We could imagine
an infinite number of different paths connecting A and B, but if we sum the
difference between kinetic and potential energy at each of the points of these
paths the result will be a number that is always greater than the one obtained
for the route followed by an actual object. In fact, while K ( A), K ( B), U ( A)
and U ( B) are the same for any trajectory, their value along the path changes
and crucially depends on the chosen trajectory. If we simplify the problem
and consider simply an object in vertical motion, whose height is described by
the function y(t), the law of motion is the one associated to an extremum (not
necessarily a minimum) of the action functional
" #
dy(t) 2
Z tB  
1
S[y] = m − mg y(t) dt. (2.1)
tA 2 dt

13
14 principle of least action

The result is the same as the one we would obtain by directly solving the
equation of motion
ÿ(t) = − g, (2.2)
with boundary conditions specified by y(t A ) and y(t B ). In fact, computing the
extrema of (2.1)
δS[y] S[y + cδy] − S[y]
= lim =0 (2.3)
δy c →0 c
one reproduces (2.2).
This line of reasoning works in general for any classical system. We can
recover its equations of motion as the result of a variational problem for an
action functional constructed in terms of a Lagrangian L given by the difference
of kinetic and potential energy of the system. For a generic classical physical
system with N degrees of freedom q I (t), where I = 1, . . . , N, the state at a time
t is determined by {q I (t), q̇ I (t)} and the action giving the evolution of such
system from state A at time t A to state B at time t B is
Z tB
S[q] = L(t, q I , q̇ I ) dt. (2.4)
tA

The evolution of the system is then described by the functions q I (t) extremizing
the action S. This means that if we change path
q0I (t) = q I (t) + δq I (t), (2.5)
with the assumption that the extrema are fixed δq I (t A ) = δq I (t B ) = 0, we
require that
δS = S[q0 ] − S[q] = 0. (2.6)
This leads to (assuming synchronous variations δt = 0)
Z tB Z tB 
δL δL

δS = δL dt = δq I + δq̇ I dt
tA tA δq I δq̇ I
Z tB 
δL δL d

= δq I + δq I dt
tA δq I δq̇ I dt
Z tB  (2.7)
δL δL δL
    
d d
= δq I + δq I − δq I dt
tA δq I dt δq̇ I dt δq̇ I
Z tB B
δL δL δL
   
d
= − δq I dt + δq I .
tA δq I dt δq̇ I δq̇ I A
2.1 euler–lagrange equations 15

The last term is vanishing because extrema are fixed and, since we consider
arbitrary variations δq I , the integral vanishes if and only if
δL δL
 
d
− = 0. (2.8)
δq I dt δq̇ I
These are the Euler–Lagrange equations, which must be satisfied by physical
trajectories.
These equations have a vast application in physics. Although not all systems
necessarily admit a Lagrangian description, whenever we have such a descrip-
tion we can derive its equations of motion by using (2.8). In addition, even
if we restricted our discussion to systems with a finite number of degrees of
freedom, we can extend it and repeat the derivation above for continuous sys-
tems: N → ∞. Field theories are prominent examples of this type. For instance,
the electromagnetic field can have different values at different points of space
and at different times Aµ = Aµ (t,~r ). We can therefore think of it as a vari-
able whose index I runs over an infinite number of possible values labelling
different points in space:

q I (t) −→ q~r (t) = q(t,~r ). (2.9)

Consider now a system that requires several fields φa ( x ), a = 1, . . . , n, to


specify it. The index a may label components of the same field (like in the
case of the electromagnetic potential Aµ ) or it may refer to different fields. The
Lagrangian should depend on φa (t,~r ), φ̇a (t,~r ), but also on ∂~r φa (t,~r ) and for
relativistic systems we can use the covariant notation

L = L φa ( x ) , ∂ µ φa ( x ) . (2.10)

It is also obvious that now we need to replace the sum over I by an integration
on the ~r variable
Z
∑→ d3~r (2.11)
I

and define a Lagrangian density


Z tB Z B
S= L dt = L d4 x, (2.12)
tA A

with A and B two events in spacetime. We note that while t depends on the
observer A and B do not. In the following we will often omit the extrema to
consider general equations of motion.
16 principle of least action

The Euler–Lagrange equations for fields follow as before (sum over repeated
indices a understood):
Z B Z B 
4 δL δL
δS = δL d x = δφa + δ∂µ φa d4 x
A A δφa δ∂µ φa
Z B 
δL δL
= δφa + ∂µ δφa d4 x
A δφa δ∂µ φa
Z B      (2.13)
δL δL δL 4
= δφa + ∂µ δφa − ∂µ δφa d x
A δφa δ∂µ φa δ∂µ φa
Z B    B
δL δL 4 δL
= − ∂µ δφa d x + δφa .
A δφa δ∂µ φa δ∂µ φa A

Once again the last term vanishes because we assume that the extrema are fixed
and therefore δS vanishes if and only if
 
δL δL
− ∂µ = 0. (2.14)
δφa δ∂µ φa

We observe that S is a scalar (L is a scalar density) and that provided the


fields φa have definite transformation properties under the action of (groups of)
symmetries the Euler–Lagrange equations are covariant.

2.2 examples

We now discuss some significant examples of actions for classical relativistic


particles immersed in generic electromagnetic fields and compute their equa-
tions of motion.

2.2.1 Free relativistic particle

The worldline of a relativistic particle can be parameterized by

x µ = x µ ( τ ), (2.15)

where τ is its proper time. Physical worldlines are selected by solving the
relativistic equations of motion, which, as we will now show, can be derived
using the principle of least action. In the case of a free particle such action must
2.2 examples 17

be a scalar function constructed solely from worldline data and the simplest
possibility is to use proper time τ for such purpose. An invariant definition of
the proper time is given by the square root of the line element
3
ds2 = dx µ ⊗ dx ν ηµν = −(dx0 )2 + ∑ (dxi )2 . (2.16)
i =1

In order to determine the action, in the non-relativistic limit it should repro-


duce the kinetic energy of a free particle. When β  1
1 v2
 
c
dτ = dt ' cdt 1 − +... . (2.17)
γ 2 c2
This implies that if we multiply proper time by mc we get an action
Z
S0 [ x µ ] = −mc dσ, (2.18)

which is a functional in terms of all possible worldlines x µ (τ ). The non-


relativistic limit of (2.18) gives
1 v2
 
1 2
Z Z Z
2
S0 ' −mc cdt 1 − = −mc dt + mv dt (2.19)
2 c2 2
and the first term does not depend on the position nor on the velocity and
therefore does not contribute to the equations of motion.
The equations of motion then follow from the variation of S0 with respect to
all possible paths x µ (τ ) connecting the events A and B:
Z B Z B
δ q 
δS = −mc δdτ = −mc µ − dx ρ dx
ρ δx µ
A A δx
Z B 
1 1 
= −mc √ −2dx µ dδxµ
A 2 −ds2
Z B Z B
(2.20)
dx µ µ
= mc dδxµ = mc u dδxµ
A dτ A
Z B  Z B
= d mcuµ δxµ − δxµ d(mcuµ ).
A A
Once more the last term is vanishing because δxµ ( A) = δxµ ( B) = 0 and there-
fore the equation of motion is
dpµ
= 0, (2.21)

18 principle of least action

with pµ = mcuµ , which is the correct equation of motion for a free relativistic
particle.

2.2.2 Electromagnetic field and its coupling to a charged particle

In an analogous fashion, we can compute free Maxwell’s equations from an


action involving only the electromagnetic tensor Fµν = ∂µ Aν − ∂ν Aµ :

1
Z
S1 = − d4 x Fµν F µν . (2.22)
16πc
The Euler–Lagrange equations for this case are
 
δL δL
− ∂ν =0 (2.23)
δAµ δ∂ν Aµ

and, using (2.22), we get that



δ Fρσ F ρσ δFρσ ρσ
=2 F = 2 ( F νµ − F µν ) = −4 F µν (2.24)
δ∂ν Aµ δ∂ν Aµ

and hence

∂ν F µν = 0. (2.25)

There is another quadratic tensor constructed from F, which may define a


Lagrangian density:

L = eµνρσ Fµν Fρσ . (2.26)

However, we can see that this is a total derivative and therefore it does not
contribute to the equations of motion:

L = eµνρσ Fµν Fρσ = 4 eµνρσ ∂µ Aν ∂ρ Aσ = 4 ∂µ eµνρσ Aν ∂ρ Aσ .



(2.27)

Its contribution may become relevant when multiplied by a function of other


fields or in the case that boundary effects have to be taken into account, but we
will not consider it for the time being.R
µ
If we introduce the current jµ = cq dxq δ4 ( x − xq ), we can deduce Maxwell’s
equations in the presence of matter from the interaction term
1
Z
S2 = d4 x j µ A µ . (2.28)
c2
2.3 noether’s theorem 19

We first prove that such an action is gauge invariant provided the continuity
equations ∂µ jµ = 0 are valid. Gauge transformations map Aµ → Aµ + ∂µ Λ.
Hence, the action S2 changes accordingly
1 1 1
Z Z Z
δgauge S2 = d4 xjµ ∂µ Λ = S2 + d4 x ∂ µ ( j µ Λ ) − d4 x ∂µ jµ Λ. (2.29)
c2 c2 c2
Discarding boundary terms, the action is gauge invariant if ∂µ jµ = 0. Finally,
we can put together S0 + S1 + S2 and compute both the equations of motion for
a field in the presence of charges and for the motion of the charges in this field.
1
j δAµ d4 x we get
δS1 δS2 R µ
We already computed δA µ
. If we add δA µ
= c 2

4π ν
∂µ F µν = − j . (2.30)
c
Using the explicit expression of the 4-current for a single charged particle in S2
we can also write
q
Z
S2 = dx µ Aµ (2.31)
c
and varying with respect to δx µ (up to boundary terms)
dpµ q
Z Z 
δx (S0 + S2 ) = − δxµ dτ + dδx µ Aµ + dx µ δx Aµ
dτ c
dpµ q
Z Z 
= − δxµ dτ + d δx µ Aµ
dτ c
(2.32)
q q
Z Z
− δx µ dAµ + dx ν ∂µ Aν δx µ
c c
dpµ q  dx ν
Z Z
= − δxµ dτ + δx µ ∂µ Aν − ∂ν Aµ dτ
dτ c dτ
leading to the equations of motion
dpµ q
= F µν uν , (2.33)
dτ c
which is the relativistic form of Lorentz force.

2.3 noether’s theorem

The action formalism is also extremely useful in deriving conservation laws


by looking for symmetry transformations that leave it invariant. The formal
20 principle of least action

representation of this idea is given by Noether’s theorem. This theorem shows


how in classical mechanics and in (classical and quantum) field theories the
existence of a group of continuous symmetries implies the existence of conserved
quantities.
Consider a system with N degrees of freedom, described by a Lagrangian
independent on time L(q I , q̇ I ) (obviously we can extend the reasoning to the
time-dependent case) and assume that there is a symmetry transformation

q I → q̂ I , q̇ I → q̂˙ I , (2.34)

which leaves the Lagrangian invariant without using the equations of motion. For-
mally

δL(q, q̇) = L(q̂, q̂˙ ) − L(q, q̇) = 0. (2.35)

Since
δL δL δL
    
d d
δL(q, q̇) = − δq I + δq I (2.36)
δq I dt δq̇ I dt δq̇ I

is vanishing at each instant because of the invariance of the Lagrangian under


the symmetry transformation we are considering, we see that

δL
Q= δq I (2.37)
δq̇ I

must be conserved along the motion, where the first term in (2.36) is vanish-
ing because of the Euler–Lagrange equations. Note that in this derivation the
assumption that the symmetry is continuous is crucial, otherwise the infinites-
imal variation of L does not make sense.
Also in Lagrangian field theory, the invariance of the action gets translated
in the existence of conserved quantities, but now this is expressed by a local
conservation of a current. In fact, following the same steps performed above
    
δL δL δL
δL = − ∂µ δφ + ∂µ δφ , (2.38)
δφ δ∂µ φ δ∂µ φ

and therefore we have


δL
∂µ jµ = 0, for jµ = δφ. (2.39)
δ∂µ φ
2.3 noether’s theorem 21

Charge conservation follows by defining


Z
Q= d3 x j0 ( x ). (2.40)

In fact
d
Z Z
Q= d3 x ∂0 j0 ( x ) = − d3 x ∂i ji ( x ) = 0, (2.41)
dt

where the last equality follows from the vanishing of ji at the boundary.

exercises

1. Compute the equations of motion following from the Lagrangian

1 1
L = − ∂µ Ai ∂µ Ai − ∂µ Bi ∂µ Bi − m2 ( A1 B2 − B1 A2 ),
2 2
where Ai and Bi are real fields and i = 1, 2.

2. Compute the equations of motion for the two real scalar fields φi , i = 1, 2,
following from

1 1
L = − δij Dµ φi D µ φ j + Cijk φi φ j φk ,
2 6

where Dµ φi ≡ ∂µ φi + Aµ eik φk , with e12 = −e21 = 1, and Aµ is an external


field. Cijk are constant coefficients.

3. Compute the equations of motion for

1
L = −∂µ Φ† ∂µ Φ + Φ† ηΦ − (Φ† Φ)2 ,
2
 
φ1 ( x )
 φ2 ( x ) 
where Φ is a complex field with 4 components Φ( x ) =  φ3 ( x )  and

φ4 ( x )
η = diag{−1, −1, 1, 1}. Compute also the extrema of the potential.
Part II

FINITE GROUPS
3
GROUPS: BASIC ELEMENTS

We have seen in our introductory lecture why symmetries are important to


physics and why symmetries are related to groups. In this chapter we begin to
introduce some basic elements of abstract Group Theory, which allow us to for-
malize the concept of symmetry. Our aim is to apply group theory to situations
relevant to physics, but the power of abstract algebra is to see the structures un-
derlying concrete cases and generalize them. We therefore concentrate on the
abstract concepts here, without following an axiomatic approach, but rather il-
lustrating them with simple examples. We will then explore the consequences
of these algebraic structures for applications in physics in some later lectures.

3.1 definitions

Definition 3.1. A Group (G, ◦) is a set G with a composition law, called multi-
plication,
◦ : G×G → G
( g1 , g2 ) 7 → g1 ◦ g2
satisfying the following properties:

i) associativity: a ◦ (b ◦ c) = ( a ◦ b) ◦ c, ∀ a, b, c ∈ G;

ii) existence of neutral element (or identity) 1: a ◦ 1 = 1 ◦ a = a, ∀ a ∈ G;

iii) existence of the inverse: ∀ a ∈ G, ∃ a−1 | a−1 ◦ a = a ◦ a−1 = 1 .

The closure property ensures that the composition law does not generate ele-
ments outside G. Associativity implies that the computation of an n-fold prod-
uct does not depend on how the elements are grouped together.

25
26 groups: basic elements

The remaining assumptions can be replaced by the existence of a neutral el-


ement and an inverse for left multiplication only (or right). The other relations
follow. For instance, if we assume the existence of a neutral element for left-
multiplication, 1 ◦ a = a, and the existence of the left inverse a−1 ◦ a = 1, we
can prove that a−1 is also an inverse under right multiplication:
a ◦ a−1 = 1 ◦ a ◦ a−1 = ( a−1 )−1 ◦ a−1 ◦ a ◦ a−1 = ( a−1 )−1 ◦ a−1 = 1, (3.1)
where we used the neutral element from the left and that a−1 should have an
inverse from the left, called ( a−1 )−1 . Using this result, we can also prove that
the identity is an identity also when we compose it from the right:
a ◦ 1 = a ◦ ( a−1 ◦ a) = ( a ◦ a−1 ) ◦ a = 1 ◦ a = a. (3.2)
We can also prove that
Theorem 3.1. A group has a unique identity element and each element has a unique
inverse.
Proof. If 1 and I are both identities, then I ◦ 1 = 1, but also I ◦ 1 = I and hence
I = 1.
If the element a has two inverses, namely a−1 and b, from b ◦ a = 1 we get that
b ◦ a ◦ a−1 = (b ◦ a) ◦ a−1 = 1 ◦ a−1 = a−1 , but also b ◦ a ◦ a−1 = b ◦ ( a ◦ a−1 ) =
b ◦ 1 = b.
The neutral element is its own inverse, but this may not be the only one. We
may have other elements that are their own inverse. For instance in (R∗ , ·) we
have that 1 · 1 = 1, but also (−1) · (−1) = 1.
We should also note that in general a group may have a multiplication law
that is not commutative. In fact we distinguish Abelian and non-Abelian groups
Definition 3.2. A group G is said to be Abelian (commutative) if the composi-
tion law is commutative: a ◦ b = b ◦ a, ∀ a, b ∈ G.
Conversely
Definition 3.3. A group G is said to be Non-Abelian (non-commutative) if at
least two elements a, b ∈ G satisfy a ◦ b 6= b ◦ a.
Example 3.1. A simple example of an Abelian group is (R, +), where the neutral
element is the element 0. Also (Z, +) has a group structure, while (N, +)
does not admit an inverse. If we use standard multiplication (R∗+ , ·) has a
group structure, while (N, ·) does not. In fact if n ∈ N, the inverse under
/ N.
multiplication is 1/n ∈
3.2 maps between groups 27

All these examples have an infinite number of elements, but there are also
groups with a finite number of elements. We will call
Definition 3.4. Order of a group G: #( G ) = number of elements in G.
Groups with a finite order are called finite groups. An example of a finite
group is
Definition 3.5. Cyclic group of order n: Zn ≡ { gk , 1 ≤ k ≤ n| gn = 1}.
A simple way to construct this group is by means of the n-th root of unity
g = e2πi/n . Let {ζ 1 , . . . , ζ n } be the n distinct roots of the equation zn = 1, so
that ζ k = gk . This set forms a group under multiplication, with composition
law ζ k ◦ ζ m = ζ m ◦ ζ k = e2πi/n(k+m) = ζ (k+m mod n) . From this composition
we can also derive an alternative definition of Zn as (Z, +) with equivalence
relation g + n ∼ g, ∀ g ∈ Z.
From the examples above we can also distinguish discrete groups like Zn or
(Z, +) and continuous groups, like (R, +).
Example 3.2. Other examples we will in the following use are the sets of real
and complex square matrices Mat(n, R) and Mat(n, C). These are semigroups
(not all their elements have an inverse) using matrix multiplication.
The sets of real or complex square invertible matrices form a group, instead.
They are denoted by GL(n, R) and GL(n, C), respectively, where GL stands for
general linear.
We also note that while Zn , (Z, +) and (R, +) are Abelian groups, GL(n, R)
and GL(n, C) are non-Abelian.

3.2 maps between groups

If we have two groups with the same number of elements and analogous prop-
erties, how do we distinguish them? When can we consider the two groups to
be the same? In order to answer to these questions we need to discuss maps
between groups.
As we explained above, a group is defined by a set, but also by a multiplica-
tion. The map between the elements of the two groups may or may not respect
such multiplication law.
Definition 3.6. A Homomorphism is a map φ : G1 → G2 such that the composi-
tion in G1 is mapped into the composition in G2 :
φ ( g1 ◦ 1 g2 ) = φ ( g1 ) ◦ 2 φ ( g2 ) .
28 groups: basic elements

Homomorphisms map the identities of the two groups among themselves


φ(1G1 ) = 1G2 . Also, the map of the inverse of an element is the inverse element
of the map of the original one: φ( a−1 ) = [φ( a)]−1 . In fact, if we respect the
group structure:

φ( a) = φ(1 ◦G1 a) = φ(1) ◦G2 φ( a), (3.3)

and

1G2 = φ(1G1 ) = φ( a−1 ◦G1 a) = φ( a−1 ) ◦G2 φ( a). (3.4)

However, homomorphic maps may loose some of the structure of the original
group and relate groups with different properties. For instance we will see that
homomorphisms can map non-Abelian groups to Abelian ones.
If we want to construct a map that fully respects the structure of a group we
need to require that it is also 1 to 1:
Definition 3.7. A homomorphism φ : G1 → G2 is called isomorphism if it is
surjective and invertible:

∃ φ−1 : G2 → G1 | φ−1 (φ( g)) = g, ∀ g ∈ G1 .


If the map φ is an isomorphism each element of G1 is uniquely mapped into
an element of G2 and therefore the two groups are essentially the same, up to
a relabeling of their elements. Symbolically G2 ' G1 .
Finally, it is useful to introduce a name for the maps from a group to itself:
Definition 3.8. An isomorphism between a group and itself is an automorphism.
Example 3.3. The first example is a homomorphism between the non-Abelian
group of invertible matrices and the set of reals (without zero), which form an
Abelian group under multiplication. The map is given by the determinant:

ϕ: GL(n, R) −→ R∗
(3.5)
M 7−→ det( M).
This map preserves the group multiplication thanks to the properties of the
determinant: ϕ( AB) = det( AB) = det( A) det( B) = ϕ( A) ϕ( B).
Example 3.4. The second example is a homomorphism between a continuous
and a discrete group:

ϕ: R∗ −→ {1, −1}
(3.6)
a 7−→ |aa| .
3.3 finite groups 29

ab a b
The map ϕ satisfies ϕ( ab) = | ab|
= | a| |b|
= ϕ( A) ϕ( B), as expected for a
homomorphism.
Example 3.5. Finally, we present an isomorphism between two continuous groups:

exp : (R, +) −→ (R∗+ , ·). (3.7)


The exponential map correctly satisfies exp( a + b) = exp( a) · exp(b). In addi-
tion the identity of the real numbers with addition as the group composition
law 1(R,+) = 0 is mapped to the identity of the real numbers with multiplica-
tion as composition law: 1(R∗+ ,·) = 1. The inverse map is the logarithm.

3.3 finite groups

Finite groups have properties that are not shared by infinite (or continuous)
groups. For example, if an element g of a finite group is multiplied by itself
enough times, the identity of the group is recovered. In fact, multiplying any
element of a finite group by itself more than #G times must lead to a recurrence
of the product, because there are at most #G distinct elements. Take g ∈ G,
where G is a finite group with #G = n. We can easily prove that there must be
a m ≤ n such that gm = 1. Consider the set { g, g2 , . . . , gn+1 }, which has one
element more than those of G. Since each of the elements of this set is in G by
closure, at least two of them must coincide. This means that ∃ p = q + m such
that g p = gq . We then deduce that g p = gq gm = gq and hence gm = 1, together
with m ≤ n.
We then call
Definition 3.9. The order of an element g ∈ G is the smallest integer n such
that gn = 1.
We can also prove that multiplying a group by any of its element we recover
all its elements.
Theorem 3.2. Rearrangement theorem.
g G = G, ∀ g ∈ G.
Proof. g G contains #G elements and hence they are all distinct and therefore
it is identified with G, or at least two of them are equal. If G = { g1 , . . . , gn }
and g = g j , two elements of gG are equal if g j gk = g j gm , but this implies that
gk = gm and therefore gG = G.
30 groups: basic elements

An important application of this theorem is given by the fact that we can


represent the composition law for a finite group via a multiplicative table. This
is a square array whose raws and columns are labelled by the elements of the
group and whose entries correspond to the products. The element in the i-th
raw and j-th column corresponds to the element gi ◦ g j , where gi and g j are the
elements labeling that raw and column.
A simple instance is the table for the group Z2 . Since the requirement that
Z2 is a group implies that 1 ◦ 1 = 1 and a ◦ 1 = 1 ◦ a = a, we only have to fix
a ◦ a, which is determined by the rearrangement theorem to be 1:

◦ 1 a
1 1 a (3.8)
a a 1

Note that the table is symmetric about the diagonal, which means that the
group is Abelian.
A different and efficient way to present a discrete group is via its generators
and their relations.

Definition 3.10. A group presentation hS| Ri contains a set of generators S and a


set of relations R between elements constructed in terms of the generators in S.

The group defined by such presentation is the result of all the inequivalent
(with respect to relations in R) words constructed by using the generators in S
(and their inverse) as letters. Note that for the same group we may have several
presentations!
Example 3.6. Cyclic groups have the following presentation:

Zn = h g | g n = 1 i. (3.9)

We have only one generator and the inequivalent words are g, g2 , . . . , gn = 1.


Using group presentations we can introduce another group we will exten-
sively use during our lectures, the Dihedral group. Its definition is given by

Definition 3.11. Dihedral group: Dn = hρ, σ | ρn = 1 = σ2 , σρ = ρ−1 σ i

and from such presentation we can deduce all the elements of the group. We
see from the relations in the presentation that both ρ and σ have finite order
(n and 2). We can also use the last relation in the presentation to move all
ρ’s appearing in the definition of a group element to the left of any σ. We
3.3 finite groups 31

therefore conclude that inequivalent words are ρ A σ B , with A = 0, 1, . . . , n − 1


and B = 0, 1. This means that #Dn = 2n.
This group has a simple geometric origin: it is the group of symmetries of
a regular polygon with n sides. ρ is the group element related to a rotation of
2π/n radians and σ is the reflection around a vertical axis.
We conclude this lecture by introducing two more groups that will be widely
used in this course: the Symmetric group and the Alternating group.

3.3.1 The Symmetric Group Sn

The symmetric or permutation group of n elements (Sn ) has a special role in


group theory because of a theorem (Cayley’s theorem) that shows that any
group of order n is isomorphic to some subgroup of Sn . It also has an obvious
important relevance in physics whenever one discusses identical particles.
Definition 3.12. The Permutation group of n elements (Sn ) is the group of bijec-
tive maps from the set X = {1, 2, . . . , n} to itself.
Note that its definition depends only on the order of the elements and not
on the labels.

1 2 3 2

2 3 2 1

3 1 1 3

Figure 4.: Two identical elements of the permutation group of three objects.

First of all we can see that such maps admit a group structure. The compo-
sition of two maps f , g : X → X defined as f ◦ g : x 7→ f ( g( x )) is obviously
again a map f ◦ g : X → X and hence the group is closed under multiplication
◦. This product is also clearly associative and the identity is the map 1 : x 7→ x,
∀ x ∈ X. If we restrict our attention to bijective maps then there is an inverse
map for each one of them and hence we have an inverse element for each of
the group elements.
Note that if we have two sets X and Y that have the same number of elements,
the associated permutation groups are isomorphic.
32 groups: basic elements

We can also determine the order of Sn rather easily. Since permutations are
maps σ : X → X, we just have to check how many inequivalent ways we have
to map X to itself. The first element can be mapped into each of the n elements
of X. Hence σ (1) can have n different values. At this point σ (2) can take at
most n − 1 different values, because we have to exclude σ(1), and so on. This
means that

#Sn = n! . (3.10)

There are several different ways to describe these maps, but, instead of giving
additional abstract definitions, we will introduce them by examples.
Example 3.7. Consider S3 , the group of permutations of 3 elements. This group
has order 3! = 6. One way to describe these elements is to use matrices describ-
ing how each element is mapped:
 
1 2 3
 . (3.11)
σ (1) σ (2) σ (3)

In this notation the top line represents the initial order of the objects and the
lower line represents the final order, after the permutation. The 6 inequivalent
maps are then
     
1 2 3 1 2 3 1 2 3
1=  , p1 =   , p2 =  ,
1 2 3 2 3 1 3 1 2
      (3.12)
1 2 3 1 2 3 1 2 3
d1 =   , d2 =   , d3 =  .
1 3 2 3 2 1 2 1 3

An alternative representation is given by cycles. This is actually more concise


and much more efficient for large n. A permutation that shuffles k < n objects
into themselves, leaving the others untouched is called a k-cycle. For instance,
we see that d1 permutes 2 and 3, but leaves 1 untouched, so we represent it by
(23), or by (23)(1) if we want to keep track of the fixed element. The permuta-
tion p1 on the other hand maps 1 → 2 → 3 → 1. We therefore represent it by
the 3-cycle (123). The whole group is then represented by:

1 = (1)(2)(3), p1 = (123), p2 = (132),


(3.13)
d1 = (1)(23), d2 = (2)(13), d3 = (3)(12).
3.3 finite groups 33

The composition is trivial and follows by applying the maps in series. For
instance:

1 → 3 → 1
(123) ◦ (13) : 2 → 2 → 3 = (23). (3.14)
3 → 1 → 2

Another important property of Sn (which we will not prove) is that every


permutation can be uniquely resolved into cycles which operate on mutually esclusive
sets. These are called disjoint cycles, like in d1 = (1)(23). Another general
property is that each element can be obtained as a product of 2-cycles (in this
case they may not act on mutually esclusive sets). Also, the decomposition in
2-cycles is not generically unique. For instance,

1 → 3 → 2
(123) = (32) ◦ (31) = 2 → 2 → 3 (3.15)
3 → 1 → 1

but also (13) ◦ (12) = (123). For each element however there is a minimum
number of 2-cycles needed to describe it and one can also see that if such
number is even all the decompositions of that group element will be given
in terms of an even number of 2-cycles and if such number is odd all the
decompositions will be given in terms of an odd number of 2-cycles.
We therefore define

Definition 3.13. The parity of a permutation g ∈ Sn is π ( g) = (−1)n( g) , where


n( g) is the number of 2-cycles needed in an arbitrary decomposition of g.

We can then split our symmetric group in terms of the even and odd elements
with respect to parity. It is interesting to note that the subset of elements with
even parity define also a group:

Definition 3.14. The Alternating group An is defined by the subset of elements


of Sn with even parity and form a group under the multiplication rules of Sn .

In the case of S3 we see that A3 = {1, p1 , p2 }. Note also that in this case A3
is Abelian, while S3 is not.
34 groups: basic elements

exercises

1. Consider the following sets of elements and composition laws. Determine


whether they are groups and, if not, identify which group property is
violated:
• Q+ , with a ◦ b = a/b;
• Even integers with respect to ordinary addition;
• Z with a ◦ b = a − b;
• Q∗ , with a ◦ b = ab;
• Z∗n ≡ { a ∈ Zn | gcd( a, n) = 1}, with a ◦ b = ab. (Tricky. Try first
with Z5∗ = {1, 2, 3, 4})

2. Show that a group is Abelian if and only if ( ab)−1 = a−1 b−1 .

3. Write the multiplicative table for a group of order 3 and show that it is
unique.

4. Determine all groups of order 4 (up to isomorphisms) and discuss if they


are Abelian or not.

5. The group of quaternions is the group generated by the elements {1, −1, I, J, K }
satisfying the composition laws

I 2 = J 2 = K2 = −1, I J = K, (−1)2 = 1.

Compute its multiplication table.

6. Consider the following couples of groups:

Z4 , S3 ,
Z6 , D3 ,
S3 , D3 .

Which lines contains isomorphic groups?

7. Show that the inverse of a k-cycle is the same k-cycle written in reverse
order.

8. Compute (34) ◦ (623174) ◦ (34), (12) ◦ (13)(45) ◦ (12) and (43) ◦ (13)(452) ◦
(34). What do we learn from this?
3.3 finite groups 35

9. A3 is isomorphic to another group we already know, which one?

10. A presentation for A4 is A4 = h a, b| a2 = 1 = b3 = ( ab)3 i. Find two


permutations in S4 representing the elements a and b of the given presen-
tation.
4
SUBGROUPS, QUOTIENTS AND PRODUCTS.

Now that we have introduced the basic notions of group theory, we would like
to understand when a subset of the elements of a group form themselves a
group under the same composition law, how to take products and quotients of
groups.

4.1 basic definitions

Definition 4.1. A subset of a group H ⊂ G is a subgroup H < G if and only if


it is closed under the composition law of G (and contains the inverse of each
element):

H<G ⇔ h1 ◦G h2−1 ∈ H, ∀h1 , h2 ∈ H. (4.1)

For finite groups there is no need of the requirement of having the inverse,
because an = 1 for some n for any a ∈ H.
According to this definition, the unit element {1} forms a subgroup of G as
well as the full G. These are improper subgroups.
In physics subgroups will be extremely useful, especially in the context of
symmetry breaking. We often encounter systems whose Lagrangian (or Hamil-
tonian) is invariant under a certain symmetry group, which is broken by the
addition of additional terms to a subgroup H < G.
Example 4.1. Dn contains a subgroup isomorphic to Zn . This is obtained as the
result of the repeated application of the generator ρ in the presentation (3.11).
Example 4.2. Zn < Sn . If we name ρ = (12 . . . n), we see that ρn = 1 and ρk 6= 1
for any k 6= n.

37
38 subgroups, quotients and products

Example
 4.3. Dn < Sn : ρ has been identified above. σ = (1 n)(2 n − 1) . . . =
1 2 ... n−1 n
.
n n−1 ... 2 1
Once again all these examples let us sense that all finite groups can be real-
ized as subgroups of Sn for n sufficiently large (which is the content of Cayley’s
theorem). If we only had finite groups we could then forget all the theoretical
constructions we introduced so far and directly study Sn . However it is still
useful to use finite groups as a laboratory to develop the necessary intuition to
study the properties of continuous groups.
A special subgroup in every group is

Definition 4.2. The Center of a group is the subgroup ZG < G defined as

ZG = { a ∈ G | a ◦ g = g ◦ a ∀ g ∈ G }.

This subgroup contains all the elements of a group that commute with all the
others. By definition, the center of a group is Abelian and it is straightforward
to see that if G is Abelian than the ZG is the whole group.
Example 4.4. Consider GL(2,R), i.e. the group of 2 by 2 invertible matrices. The
center is given by all the matrices proportional to the identity: s1, with s ∈ R∗ .

4.2 quotients

When the symmetry group G of a given physical system is broken to a sub-


group H, some of the elements of the original invariance will have a non-trivial
action. In detail, while H will still leave the system invariant, elements of G
that are not in H will not. It is therefore interesting to ask what is the structure
emerging for the trasformations under G when we identify all the transfor-
mations in H < G. For this reason we now discuss the quotients of groups.
Quotients essentially emerge by identifying elements in a group via equiva-
lence relations. In order to define them we therefore need first to recall the
properties of

Definition 4.3. An equivalence relation ∼ is a binary relation ∼ on a set if and


only if it is

1. reflexive: a ∼ a;

2. symmetric: a ∼ b ⇒ b ∼ a;
4.2 quotients 39

3. transitive: a ∼ b and b ∼ c ⇒ a ∼ c.

Once we have an equivalence relation we can split the elements according to


their properties under such a relation. In particular, we will group together all
the equivalent elements in so-called equivalence classes:

Definition 4.4. An equivalence class for an element a ∈ A is the set of all ele-
ments in A equivalent to a:

[ a ] ≡ { b ∈ A | b ∼ a }.

It can be seen that equivalence relations split a set into disjoint subsets given
by equivalence classes. This splitting is going to be crucial in the definition of
group quotients.
If we now want to identify elements in G that are related by elements in H,
we need to define appropriate equivalence classes by multiplying each element
of G by all the elements in H. Since for a generic group the product is not
commutative, the result will depend on the order. Given H < G we therefore
define distinct equivalence classes using left and right multiplication.

Definition 4.5. Right cosets are the sets obtained by multiplying H < G by
elements g ∈ G from the right: Hg. Analogously, Left coset ≡ gH, with H < G
and g ∈ G.

Note that in general gH and Hg are not in H unless also g ∈ H and they are
not even subgroups of G.
Example 4.5. Consider H = {1, d1 } < S3 . This group is isomorphic to Z2 . Right
cosets are obtained by multiplying H from the right with the elements of S3 .
The group structure shows that we have only 3 distinct cosets:

H1 = Hd1 = {1, d1 }, Hd2 = H p1 = {d2 , p1 }, Hd3 = H p2 = {d3 , p2 }. (4.2)

Similarly, left cosets are

1H = d1 H = {1, d1 }, d2 H = p2 H = { d2 , p2 }, d3 H = p1 H = {d3 , p1 }. (4.3)

As expected, left and right cosets can be different and indeed in this case they
are different. We can also note that the intersection between left cosets (or
between right cosets) is empty and that their union is the original group S3 .
In fact, this is a consequence of a general theorem:
40 subgroups, quotients and products

Theorem 4.1. Two cosets of a subgroup have the same elements or they do not share
any element in common.

Proof. Let Hg1 and Hg2 be two different (right) cosets. If there is at least one
element in common, this means that for hi , h j ∈ H, we have hi g1 = h j g2 . From
this assumption we derive that g2 ( g1 )−1 = (h j )−1 hi ∈ H. Using the rearrange-
ment theorem, we then conclude that Hg2 ( g1 )−1 = H and hence Hg2 = Hg1 ,
so all their elements are in the same.

We can finally discuss quotients of groups. Identifying elements in the same


coset we get

Definition 4.6. The Right quotient G/H of a group G by its subgroup H is the
set of all equivalence classes of elements of G obtained by right multiplications
with H:
g1 ∼ g2 if ∃ h ∈ H | g1 = g2 h.

The left quotient H \ G is obviously defined by the relation g1 ∼ g2 if ∃ h ∈


H | g1 = hg2 . From these definitions we see that quotients are collections of
equivalence classes defined by cosets.
In general G/H 6= H \ G and they do not form a group with respect to the
group multiplication in G, unless H satisfies a special condition: being self-
conjugate. This is a special condition under which the group multiplication in
G can be transferred without inconsistencies to G/H.
First of all we need to define what we mean by conjugation. This is an
operation by which we relate different elements of a group. In detail:

Definition 4.7. a, b ∈ G are conjugate if ∃ g ∈ G | a = gbg−1 .

This definition implies that the elements a and b give the same element when
multiplied by g, the first from the right and the second from the left. We then
see that we can use such a definition to establish a relation between left and
right cosets. Actually, we can extend this definition to whole subgroups of a
group:

Definition 4.8. N < G is Normal (and we write N C G) if it is selfconjugate:

gNg−1 = N, ∀ g ∈ G.

This definition implies that the conjugate of any element in N is also an ele-
ment in N for any element in G defining the conjugation: gng−1 ∈ N, ∀g ∈
4.2 quotients 41

G and ∀n ∈ N. The straightforward consequence is that right and left cosets de-
fined in terms of N coincide: gN = Ng. Since left and right cosets are the same,
also left and right quotients coincide. This eventually leads to the fact that G/N
quotients inherit the property of being a group under the multiplication of the
original group G:
Theorem 4.2. If N C G then G/N is a group.
Proof. The elements of the quotient are defined as the equivalence classes given
by the cosets. Hence all the elements in gN for a given g ∈ G have to be
identified. The product in G/N is the product in G, where again we identify
the elements which are part of the same coset. We then have to verify that all
group properties are respected for G/N. Closure follows from the composition
in G:
g1 N ◦ g2 N = g1 ( g2 N ( g2 )−1 ) g2 N = g1 g2 N ( g2 )−1 g2 N = g1 g2 N,
where we first used the fact that N is normal and then the fact that N is a
group (N ◦ N = N). The associative property is straightforward. The existence
of an inverse for each element follows from the proof of closure above. Given
an element gN, its inverse is defined by g−1 N.
Example 4.6. We first give an instance of a subgroup that is not normal. Con-
sider Z2 < Dn , defined by the σ generator in the presentation (3.11). From
the equivalence relations given in the presentation we can see that ρσρ−1 =
σρ−2 ∈/ Z2 . As expected, this implies that the product of elements in the quo-
tient using the multiplication in Dn does not close consistently. In fact, Dn /Z2
implies the identification ρk ∼ ρk σ, but if we multiply elements in the same
equivalence class, we get inconsistent results:
ρ a ρb = ρ a+b 6= (ρ a σ )(ρb σ ) = ρ a−b . (4.4)
Example 4.7. An example of a normal subgroup is given by Zn < Dn , defined
by ρ. Since we identify all the elements ρk ∼ 1 and σρk ∼ σ, the quotient is
given by two elements, which define the group Z2 = 1, σ. The multiplication
of elements in the same equivalence class now gives consistent results. For
instance, we see that
σρ a σρb = σ2 ρb− a = ρb−a ∼ 1 = σσ. (4.5)
Example 4.8. Although for n = 3 D3 ' S3 , in general Zn < Sn is not normal
for n > 3. This can be seen right away by checking that (12)(12345)(12) =
(21345) 6= (12345) a for any a.
42 subgroups, quotients and products

4.3 products

In Physics we can imagine systems where we have symmetries of different


nature that are simultaneously present. Both should be represented by groups,
with independent actions. We could also have instances where one symmetry
interferes with the way the other acts. In this case the two actions will not be
independent and the resulting group will be a non-trivial combination of the
original ones.
For these reasons we now discuss group products.

Definition 4.9. Given two groups G and K their (direct) product is

G × K = {( g, k), g ∈ G, k ∈ K },

with composition

( g1 , k 1 ) ◦ ( g2 , k 2 ) = ( g1 ◦ g2 , k 1 ◦ k 2 ) .

For finite groups we can deduce immediately the order of the product as

#( G × K ) = #G #K. (4.6)

In general, there are two obvious normal subgroups: G × 1K and 1G × K. The


quotients with respect to these normal subgroups are K and G, respectively.
Using the presentations of the two groups forming the product, we can define
the product group via a presentation that uses all their generators and relations
with an additional one, which we now explain. If we call S1 = { gi } the set of
generators of G and S2 = {k i } the set of generators of K, the generators of
G × K are S = S1 ∪ S2 . Moreover, the fact that we have a direct product G × K
is equivalent to the fact that all the elements gi ∈ G × 1K and k j ∈ 1G × K (we
loose the identity factors in the following) should commute: gi k j = k j gi . We
can then write

G × K = hS1 ∪ S2 | R1 ∪ R2 ∪ { gkg−1 k−1 = 1, ∀ g ∈ G, k ∈ K }i. (4.7)

This means that we need to take all the words constructed from the genera-
tors of G and K with their equivalence relations plus the commutation relation
between elements of G and K.
As we discussed in the introduction to this section, we may also need to
specify groups that are products where the action of one of the factors depends
4.3 products 43

on which of the elements of the other factor is taken in the product. For this
reason we can introduce a new definition, where the last relation in (4.7) is
replaced by a new one.

Definition 4.10. Given two groups G and K their semidirect product is defined
by

G n K ≡ hS1 ∪ S2 | R1 ∪ R2 ∪ { gφg (k ) g−1 k−1 = 1, ∀ g ∈ G, k ∈ K }i,

where φg : K → K is a g-dependent automorphism.

Note that the symbol defining the product contains an asymmetry, which is
reflected in the different ways the two groups enter in the definition: G n K =
K o G. In fact, for semidirect products elements in G × 1K and in 1G × K do
not necessarily commute.
From the definition we also see that K is a normal subgroup, i.e. K C ( G n K ),
while G is not, i.e. G < ( G n K ).
Example 4.9. Dn ' Z2 n Zn , where Z2 = hσ | σ2 = 1i, Zn = hρ | ρn = 1i and
φσ (ρ) = ρ−1 , so that σρ = ρ−1 σ. Zn is a normal subgroup, Z2 is not.
Example 4.10. The group of orthogonal matrices in N dimensions O(N) is de-
fined as the set of matrices O ∈ GL(N, R), which satisfy O T O = 1 N = OO T .
Since det (O T O) = (det O)2 = det 1N = 1, the matrices of O(N) satisfy det
O = ±1. Those with positive determinant form a subgroup, called special or-
thogonal group SO(N) < O(N), which obviously contains the identity. The
quotient with respect to this subgroup generates two classes of equivalence,
the one of matrices with positive determinant (which contains the identity)
and the set of matrices with negative determinant. This is actually a Z2 group
under the original multiplication. In fact we leave as an exercise to the reader
to check that generically O(N ) ' SO( N ) o Z2 , which, for odd N reduces to a
direct product.
Example 4.11. The Euclidean group E( N ) ' ISO( N ). It is defined as the combi-
nation of rotations SO(N) and translations in R N :

E( N ) = {(O, a) | O T O = 1 N , det O = 1, a ∈ R N }. (4.8)

Given v ∈ R N the action is v → Ov + a. The composition gives

(O2 , a2 ) ◦ (O1 , a1 ) = (O2 O1 , O2 a1 + a2 ), (4.9)


44 subgroups, quotients and products

(O1 ,a1 ) (O2 ,a2 )


in fact v −→ O1 v + a1 −→ O2 O1 v + O2 a1 + a2 . We have two subgroups
(O, 0) and (1, a), but

( A, 0)(1, a)( A−1 , 0) = (1, Aa) = φ A−1 ( a). (4.10)

Hence

ISO( N ) ' SO( N ) n R N . (4.11)

4.4 classifying finite groups.

Using direct and semidirect products we can construct groups from other
groups, but obviously these are not really new, because they can be classified
as products. So, how do we distinguish genuinely new groups?

Definition 4.11. A group is called simple if it does not admit proper normal
subgroups.

Definition 4.12. A group is called semisimple if it does not admit nontrivial


normal Abelian subgroups.

Non-simple groups can be obtained as (semi)direct products.


The classification of all finite groups (simple ones) required computers. Most
of them are in infinite families (like Zn with prime n, An , etc...) There are some
finite simple exceptional groups (for instance the Monster Group has about 8
·1053 elements)
The enormous theorem shows that there are a few infinite families and a finite
number of exceptional cases (sporadic groups).

exercises

1. Prove Lagrange’s theorem, which states that the order of a subgroup H


of a finite group G is a divisor of the order of G, i.e. #H divides #G.

2. Compute the center of Dn .

3. Find subgroups of order 2,3 and 4 of S4 .

4. Show that for Abelian groups left and right cosets coincide.
4.4 classifying finite groups. 45

5. Consider the following permutations in S8 :


 
1 2 3 4 5 6 7 8
a= ,
3 6 1 5 4 2 8 7
 
1 2 3 4 5 6 7 8
b= ,
2 3 4 5 6 7 1 8
 
1 2 3 4 5 6 7 8
c= .
2 4 6 1 3 5 7 8
a) Are these conjugate to each other?
b) a, b, c generate a group G < S8 . Is G < A8 ?
c) Take now H < G generated from a e b. Is H C G?
(Attention! #G = 168 and #H = 56. Find meaningful products to solve
the exercise)

6. Find all proper subgroups of the group of quaternions Q and discuss the
right and left quotients of Q with respect to them.

7. Is Z2 × Z2 isomorphic to Z4 ? And what about Z2 × Z3 and Z6 ?

8. Take A = Z2 and B = Zk . Discuss all possible products C = A n B for


k = 2, 3, 4 and check if the resulting group is isomorphic to Zn , Dn , An or
Sn .
5
R E P R E S E N TAT I O N S

So far we discussed groups using their abstract definition. To really start ap-
plying them (especially to physics), we must be able to represent them in a way
that could be used for practical calculations. For this reason we need to map
abstract groups into groups of mathematical objects we know how to deal with,
like matrices or differential operators. We therefore introduce in this lecture the
concept of representations of a group.
If a group codifies the symmetries of a physical system, i.e. the relations
between the transformations we can apply to a system without changing the
physics, we want to be able to represent this action on the system.
Beware that the existence of a symmetry does not mean that nothing changes
when applying it. For instance the discrete group D3 rotates or reflects a trian-
gle, leaving the image fixed, but not the vertices:

σ ρ
refl. rot.

Figure 5.: Example of some of the actions of D3 on a triangle.

In the same way, when we discuss a physical system, its invariance under a
certain symmetry does not imply that all its quantities remain unchanged. On
the other hand, the existence of a symmetry transformation tells us how the
variables describing a certain physical state change. For instance, Special Rela-
tivity dictates invariance of physics under Lorentz transformations. This does
not mean that two inertial frames, related by a Lorentz transformation, share

47
48 representations

exactly the same value for all their physical quantities, but rather that their
transformation is such that physical laws do not change in form. For instance,
the value of pµ from an inertial system to another changes by pµ → Λµ ν pν , but
dpµ
free particles are described by the same law dσ in both systems. In fact, Λ pro-
vides a concrete representation of how the group of Lorentz symmetries acts
on the various components of the momentum, which are grouped in a multiplet
of the symmetry group. In a similar fashion, when we consider a quantum-
mechanical system with a certain symmetry, we know that there must exist a
unitary operator commuting with the Hamiltonian, which gives a representa-
tion to the action of such symmetry. The wavefunctions, however, do not share
the symmetry of the Hamiltonian and they will transform in a way determined
by such operator and, in particular, by the representation of the symmetry
group provided by the unitary operator mentioned above. This, in turn, will
provide a way to classify the eigenfunctions of the Hamiltonian, as we will
discuss shortly.

5.1 basic definitions

As understood from our introduction, there may be different ways to repre-


sent a group. However, we will be mainly interested in linear representations.
These are representations by means of matrices. We will therefore dedicate a
considerable amount of time to the discussion of groups of matrices (both finite
and continuous) and to their properties.
Definition 5.1. A linear representation of a group G is a homomorphism (iso-
morphism) D : G → GL( N, C).
In a linear representation, each element of the group is represented by a
N × N complex matrix D ( g), whose composition law respects the composition
law in G:
D ( g1 ◦ g2 ) = D ( g1 ) D ( g2 ) ,
D (1) = 1 N , (5.1)
D ( g−1 ) = [ D ( g)]−1 .

From a different point of view, we can describe the linear representation D as


a map relating elements of a linear vector space D : G × V → V, which we
generically take to be a complex space V = C N . N fixes the size of the matrices
by which we represent the group G and therefore we call
5.1 basic definitions 49

Definition 5.2. The dimension of the vector space on which the matrices D ( g)
act is the dimension of the representation.
Since we are representing an abstract group G by a group of matrices, the
mappings between the two groups may not necessarily preserve all the infor-
mation of the original group G. For this reason we distinguish between faithful
representations, obtained by means of isomorphisms, and unfaithful representa-
tions, obtained by homomorphisms. In this last case, we will have that different
elements of the group are represented in the same way. If more than one el-
ement is represented by the same matrix, the group element connecting them
must be represented by the identity. This further implies that more than one
group element must be represented by the identity matrix and therefore that
the kernel of the map is going to be non-trivial. A faithful representation then
is such if D ( g) = D ( g0 ) ⇔ g = g0 .
It is also easy to convince ourselves that the elements in the kernel of the
representation are going to be a normal subgroup of the original group G:
H = ker D C G. In fact, if D ( g) = D ( g0 ) = 1, we also have that D ( g ◦ g0 ) =
D ( g) D ( g0 ) = 1. Obviously D (1) = 1, the inverses are again represented by
the identity and finally D ( g ◦ h ◦ g−1 ) = D ( g)1D ( g)−1 = 1 for any g ∈ G and
h ∈ H. We can then conclude that if our representation is unfaithful for G,
the same matrices are providing a representation of G/H that is faithful! We
therefore conclude that:
Theorem 5.1. All non-trivial representations of a simple group are faithful.
Once we choose a basis for V, {ei } with i = 1, . . . , N, the matrix action is
defined as
D ( g)ei = e j D ji ( g), (5.2)
where sum over repeated indices is understood. This is needed to preserve the
composition rule
D ( g ◦ g̃) = D ( g) D ( g̃). (5.3)
In fact:
D ( g ◦ g̃)ei = e j D ji ( g ◦ g̃) = D ( g)[ D ( g̃)ei ] =
(5.4)
D ( g)ek Dki ( g̃) = e j D jk ( g) Dki ( g̃)
and hence
D ji ( g ◦ g̃) = D jk ( g) Dki ( g̃), (5.5)
50 representations

which provides the standard matrix composition law. Be careful with the index
position when passing to an explicit representation.
We can now ask ourselves whether we can always represent a group in such
a way and also what kind of matrix representations are really useful. A repre-
sentation that always exists is
Definition 5.3. The trivial representation is defined by

D ( g) = 1, ∀ g ∈ G.

Another representation that is easy to construct is


Definition 5.4. The regular representation D R acts on a basis of vectors | gi ∈ C#G
as
D R ( g)| g0 i = | g̃ = g ◦ g0 i.
We directly used g as an index for such a representation, because we are
assuming that the linear space on which it acts has one basis vector for each
element of G. The matrix representation can then be given explicitly by the
formula:

DgR1 g2 ( g3 ) = δg1 ,g3 ◦ g2 . (5.6)

Example 5.1. Take Z2 = {1, σ }, V = span{|1i, |σ i} and then


   
R 1 0 R 0 1
D (1) = , D (σ) = . (5.7)
0 1 1 0

So that D R (1)|1i = |1i, D R (σ )|1i = |σ i, D R (1)|σ i = |σ i and finally D R (σ )|σi =


|1i.
Representations of G induce in an obvious way representations of any H < G.
We just need to restrict the maps D to the elements h ∈ H.
Example 5.2. The representations of D3 induce representations of Z3 by D (ρ)
and also of Z2 < D3 by using D (σ ).
Since representations are specified by matrices describing each of the group
elements, we may ask ourselves whether we really need to give explicitly each
of them or not. Actually, it turns out that it is often sufficient to provide the
explicit representation of a meaningful subset. For instence, once we have a
presentation for a group, we can provide a representation for the same group
by choosing D (s) for each s ∈ S in a way that satisfies the presentation relations:
D (r ) = 1 ∀r ∈ R.
5.2 equivalent and reducible representations 51

Example 5.3. Zn = hρ | ρn = 1i can be represented trivially by D (ρ) = 1. Then


D (ρn ) = D (ρ) . . . D (ρ) = 1 = D (1). However, we can also provide a faithful
representation by choosing D (ρ) = e2πi/n .
Example 5.4. D3 can be represented by its geometric action on a plane. This can
be represented by two matrices acting on the x and y coordinates giving the lo-
cation of the vertices of an equilateral triangle. The first one, D2 (ρ), represents
a rotation by 2/3π around the origin and the second one, D2 (σ ), is a reflection
on the x axis:
 √   
1 −√1 3 −1 0
D2 (ρ) = , D2 (σ ) = . (5.8)
2 − 3 −1 0 1

It is easy to check that the relations in the presentation are identically satisfied:
[ D2 (ρ)]3 = [ D (σ)]2 = 12 , (5.9)
 √ 
1 1 3
D (ρ) D (σ) = √ = D (σ)[ D (ρ)]2 . (5.10)
2 3 −1
In general, we may have an infinite number of different representations for
the same group. To our advantage, we can often classify them, and this is the
subject of the next part of this lecture. In fact, knowing all the representations
of a given group allows us to specify all its (linear) actions on a set of quantities
and it is therefore of primary importance for applications to physics.

5.2 equivalent and reducible representations

We already identified the conditions under which two groups are the same. In
this section we would like to identify the conditions under which two represen-
tations are the same.
Definition 5.5. Two representations D1 and D2 of the same group G are equiv-
alent if we can relate them by a change of basis, i.e.

∃S | D1 ( g) = S−1 D2 ( g)S, ∀ g ∈ G.

The matrix S is called a similarity matrix and its application to V changes the
basis of the representation space.
From the definition above, we see that S preserves the trace

trD1 = tr(S−1 D2 S) = trD2 (5.11)


52 representations

as well as the determinant


detD1 = det(S−1 D2 S) = det(S)−1 det( D2 )det(S) = trD2 . (5.12)
Hence, if detD1 6= detD2 or trD1 6= trD2 the two representations D1 and D2
cannot be equivalent.
Once we have constructed a map D that provides a representation for G, we
can easily construct other representations of the same group by considering
D = [ D T ]−1 and D ∗ . We should stress that, although we can construct them
in terms of D, they are not necessarily equivalent representations! When they
are equivalent, if we have a basis such that D ( g) = D ∗ ( g) then we call D a real
representation and if D ( g) = S−1 D ∗ ( g)S, but we do not have a basis where the
two coincide, then D is pseudoreal.
This brief discussion already shows that for a given group we can have more
representations. We could now ask which is the most efficient way to find them
and classify them. For instance, starting from two non-trivial representations
D1 and D2 we can construct a new representation D1 ⊗ D2 acting on V1 ⊗ V2 ,
but we could also construct another one by
 
D1 ( g) 0
D1 ( g) ⊕ D2 ( g) = , ∀ g ∈ G. (5.13)
0 D2 ( g)
It is obviously not useful to study all equivalent representations. We just need
one element in each equivalence class. How do we decide whether it is possible
to reduce a representation to more fundamental ones? First we try to separate
all possible different cases and then give criteria to understand which is the
one we face.
Definition 5.6. A representation is reducible (or totally reducible) if we can write
it as direct sum of other representations up to a change of basis.
Alternatively, D : G × V → V is reducible is V = V1 ⊕ V2 with V1 and V2 both
invariant under the action of D: D ( g)V1 ⊆ V1 and D ( g)V2 ⊆ V2 , ∀ g ∈ G.
Definition 5.7. A representation is decomposable (or reducible) if it is equivalent
to a block triangular representation.
Alternatively, D : G × V → V is decomposable if there is a V1 ⊂ V invariant
under the action of D:
 
D1 ( g) B( g)
D ( g) = , ∀ g ∈ G. (5.14)
0 D2 ( g)
This representation maps (v1 0) vectors to themselves.
5.2 equivalent and reducible representations 53

Definition 5.8. A representation is irreducible if it is non-decomposable.

Following the same arguments as above, a representation is irreducible if


there are no invariant subspaces under its action.

Example 5.5. S3 ' D3 . We already saw the 2-dimensional representation D2 in


equation (5.8). Using the fact that S3 is the group of permutations of 3 objects,
we can now construct a 3-dimensional representation, where the basis of the
vector space on which it acts is given by the three elements acted upon by the
permutations. A simple computation gives

   
1 1
D (1) =  1 , D ( d1 ) =  1 ,
1 1
   
1 1
D ( d2 ) =  1 , D ( d3 ) =  1 , (5.15)
1 1
   
1 1
D ( p1 ) =  1 , D ( p2 ) =  1 .
1 1

1 ρ ρ2 σ σρ σρ2

Figure 6.: Graphic representation of the 3-dimensional representation of equation (5.15)


as the action on the vertices of a triangle: (1, 0, 0) is the red one, (0, 1, 0) is
the blu one and (0, 0, 1) is the green one.

It is straightforward to check that such a representation is reducible. We can


immediately see that the three vertices are always mapped among themselves
54 representations

 
1

and therefore e3 = 1/ 3  1  is invariant. Also the subspace generated by its
1
orthogonal complement
   
1 1
1 1
e1 = √  − 1  , e2 = √  1  (5.16)
2 0 6 −2
is invariant. This becomes obvious if we use as basis for the vector space
on which these matrices act the set {e1 , e2 , e3 }. In this new basis the same
representation matrices become
 √   
−√1/2 3/2 0 −1 0 0
D (ρ = p1 ) =  − 3/2 −1/2 0  , D (σ = d3 ) =  0 1 0  . (5.17)
0 0 1 0 0 1
These are block-diagonal matrices and so will be all the matrices obtained by
their multiplications. We therefore see that this representation is the direct sum
of the 2-dimensional representation D2 and of the trivial representation.
Before proceeding let us stress the importance of the field used for the vector
space V in order to decide whether a given representation is reducible or not.
If we consider SO(2), the group of rotations on a plane, a generic element is
represented by
 
cos θ sin θ
D (θ ) = . (5.18)
− sin θ cos θ
This representation is irreducible if we act on R2 , but becomes reducible on C2 :

eiθ
 
0
D̃ (θ ) = S−1 D (θ )S = , (5.19)
0 e−iθ
 
1 1
where S = √1 .
2 i −i

5.3 unitary representations

If V is a complex vector space with a positive definite norm then V is a pre-


Hilbertian space (Hilbert if complete). It is then obvious the importance of
discussing unitary representations of groups.
5.3 unitary representations 55

Definition 5.9. A unitary representation of a group G is a representation D such


that ∀ g ∈ G and x, y ∈ V

h x |yi = h D ( g) x | D ( g)yi ⇔ D ( g)† D ( g) = 1.

Hence D ( g−1 ) = [ D ( g)]−1 = [ D ( g)]† .

A very interesting fact comes from the following theorem, which proves that
for finite groups unitary representations are all we need.

Theorem 5.2. Any representation of a finite group is equivalent to a unitary represen-


tation.

Proof. Take D ( g) a non-unitary representation of g ∈ G. Define

M≡ ∑ D † ( g ) D ( g ). (5.20)
g∈ G

This satisfies

D † ( g̃) MD ( g̃) = ∑ g∈G D † ( g̃) D † ( g) D ( g) D ( g̃)


(5.21)
= ∑ g∈G D † ( g ◦ g̃) D ( g ◦ g̃) = ∑ g0 D † ( g0 ) D ( g0 ) = M,

where we used the rearrangement theorem in the second line. We also have that
M = M† and hence M is a matrix with positive real eigenvalues (v† Mv > 0).
We can diagonalize it by means of a unitary matrix U

M = U † mU, with m = diag{m1 , . . . , m N } (5.22)


√ √
and we can define s = diag{ m1 , . . . , m N }. In this way

M = S2 = U † sUU † sU = U † s2 U = U † mU, (5.23)

where S = U † sU and S† = S. We can now prove that SD ( g)S−1 is unitary for


any g ∈ G:

(SDS−1 )† (SDS−1 ) = (S−1 )† D † S† SDS−1


(5.24)
= S−1 D † MDS−1 = S−1 MS−1 = S−1 SSS−1 = 1.
56 representations

Although we proved this theorem for finite groups, we could prove an anal-
ogous one for continuous compact groups. For continuous compact groups, how-
ever, we need to sum over infinite elements. This means that we can still per-
form the same proof provided we replace the sum with an integral over the
group, with an appropriate invariant measure (called Haar measure). Only the
existence of such a measure allows us to repeat the demonstration with
Z
M≡ dµ g D † ( g) D ( g), (5.25)

otherwise we cannot use the rearrangement theorem.


For finite groups we can also prove that

Theorem 5.3. Every unitary decomposable representation is reducible.

Proof. If D is decomposable we know that there is a subspace of the total vector


space such that D ( g)V ⊆ V, ∀ g ∈ G. To show that it is reducible we should
also show that D ( g)V ⊥ ⊆ V ⊥ , ∀ g ∈ G. This follows because for any |vi ∈ V
and |wi ∈ V ⊥ we have that hw|vi = 0, but also

D ( g−1 )|vi = |v0 i ∈ V, ∀ g ∈ G. (5.26)

These facts imply that D ( g)|wi = |w0 i ∈ V ⊥ , because

hw0 |vi = hw| D † ( g)|vi = hw| D ( g−1 )|vi = 0. (5.27)

This means that for finite groups and for continuous compact groups we can
always construct representations that are unitary and block diagonal. Hence we
can constrain ourselves to studying the irreducible representations composing
them.
We stress, however, that if the group is not finite nor compact, we may have
representations that are decomposable, but not reducible, like
 
1 x
D(x) = , (5.28)
0 1

representing (R, +), in fact D ( x ) D (y) = D ( x + y).


5.4 comments 57

5.4 comments

Unitary representations play a fundamental role in QM. If we prepare a system


in a state |1i, the probability to find it in a state |2i is

|h2|1i|2
P12 = .
|| |1i ||2 || |2i ||2
The change of state given by the action of the symmetry group G (for instance
the rotation group or Sn for n identical particles) is represented by a unitary
operator and therefore its action on the Hilbert space is such that the transition
probabilities are not affected. We see then that knowledge of group theory
implies knowledge of QM systems.
For instance, energy levels are given by eigenstates of the Hamiltonian: H |i =
E|i. If G is a group of transformations leaving H invariant, there is a uni-
tary representation of this group that commutes with H and hence UH | Ei =
EU | Ei = H (U | Ei). This means that the unitary operator U maps eigenstates
with eigenvalue E into states with the same eigenvalue and that therefore
the portion of the Hilbert space containing the eigenvectors with energy E,
H E ⊂ H, is invariant. If such a space is invariant under the action of U, this op-
erator must be in an irreducible representation of the symmetry group. Hence
the level degeneracy, which is equivalent to the dimension of H E , must also be
the dimension of an irreducible representation.

exercises

1. Given two matrices A ∈ Mat( p × q, R) and B ∈ Mat(r × s, R), their direct


sum is the matrix

A i ≤ p and j ≤ q,
 ij


A⊕B = B(i− p)( j−q) i > p and j > q,


0 otherwise.

Their direct product is defined by

( A ⊗ B)(ik)( jl ) = Aij Bkl .

Show that if the dimensions are the same (i.e. p = r and q = s) then
( A ⊕ B)(C ⊕ D ) = AC ⊕ BD and ( A ⊗ B)(C ⊗ D ) = AC ⊗ BD.
58 representations

If A ∈ GL(n, R) and B ∈ GL(m, R), show that

tr( A ⊗ B) = (trA)(trB).

Consider now Pauli matrices:


     
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = .
1 0 i 0 0 −1
Compute

γ0 = −i σ1 ⊗ 1, γ1 = σ2 ⊗ σ1 , γ2 = σ2 ⊗ σ2 , γ3 = σ2 ⊗ σ3 .

Verify that σi σj = ieijk σk + δij 1 and, using what you just learned, compute
γ5 = −iγ0 γ1 γ2 γ3 .

2. Consider again exercise 8 of chapter 4. Construct the regular representa-


tion for all cases with k = 2 and show their decomposition in irreducible
representations.

3. Verify that the matrices


   √ 
1 0 1 −1 3
D (1) = , D ( a) = √
0 1 2 3 1

form a representation for the group Z2 . Is this representation irreducible?


If not, determine the 1-dimensional representations which form the direct
sum of this representation.

4. The group A4 < S4 is generated by all possible products of the following


elements of S4 : a = (12)(34) and b = (123). A linear representation for
these permutations is given by

−1 0 0 0
   
0 1 0 0
 0 1 0 0   0 0 1 0 
D ( a) = 
 0 0 −1 0  ,
 D (b) = 
 1 0 0
.
0 
0 0 0 1 0 0 0 e2πi/3

a) Discuss whether this representation is irreducible or not.


b) Knowing that A4 admits 3 irreducible 1-dimensional representations
and 1 irreducible 3-dimensional representation, provide the explicit
form of the matrices D ( a) and D (b) for such representation.
5.4 comments 59

c) Discuss the quotient of A4 with H = {1, a, b2 ab, bab2 }.

5. In this exercise we want to discuss a representation of the group of per-


mutations of two objects S2 = Z2 = {1, ρ}, when the two objects are a
couple of quantum-mechanical particles. The state of each particle is an
element of a Hilbert space H and the state of the couple is an element of
H ⊗ H. The action of the S2 group on vectors of H ⊗ H is

D (ρ)[ψ ⊗ χ] = χ ⊗ ψ, ∀ψ, χ ∈ H.

If the two particles are identical, only states satisfying D (ρ)Ψ = Ψ or


D (ρ)Ψ = −Ψ are physical. Let us analyze these states as representations
of the group S2 .
Once constructed the operators P± = ( I ± D (ρ))/2, where I is the identity
on H ⊗ H
2 = P , P + P = I and that D ( ρ ) P = ± P ;
a) show that P± ± + − ± ±
b) find the invariant subspaces under the action of D (ρ);
c) if H is finite-dimensional, compute the dimension of P± (H ⊗ H).

6. Consider a particle in the potential well



0 if x ∈ [−1, 1] and y ∈ [−1, 1],
V ( x, y) =
∞ for other values of x, y.

The Hamiltonian is
1 h 2 i
Ĥ = p̂ x + p̂2y + V ( x, y) .
2m
The group D4 acts on the wavefunction Ψ( x, y) as (U ( g)Ψ)( x, y) ≡ Ψ( D ( g)( x, y)),
where g ∈ D4 and D ( g) is the 2-dimensional representation of D4 .
a) Show that Ĥ commutes with the representation U.
b) For n positive integer and k n = nπ/2 define the functions

cos(k n z) if n is odd,
f n (z) =
sin(k n z) if n is even.

Show that Ψn,m ( x, y) = f n ( x ) f m (y) inside the square (and zero out-
side) is an eigenfunction of Ĥ and compute its eigenvalue. Discuss
their degeneracy.
60 representations

c) Compute the action of U on the eigenfunctions Ψ2,1 and Ψ1,2 . Do


their linear combinations span an invariant subspace? If so, is the
induced representation on this subspace irreducible?
d) Compute the action of U on Ψ1,1 . Is this generating an invariant
subspace? If so, is the induced representation on this subspace irre-
ducible?
6
P R O P E R T I E S O F I R R E D U C I B L E R E P R E S E N TAT I O N S

Reducible representations of a group can be put in a block diagonal form by


means of similarity transformations and in each block one has irreducible rep-
resentations of the same group. We can then use irreducible representations as
the fundamental building blocks of more general representations. It is therefore
important to find criteria telling us when a given representation is reducible or
not and how to compare irreducible representations. We will see various ways
to do so, but all of them are based on a fundamental theorem that we are going
to discuss in this lecture: the Great Orthogonality Theorem.
In order to prove this theorem we need first a few technical results collected
in Schur’s lemmas and their corollaries. The idea behind these lemmas is very
simple: a matrix commuting with all those of a given irreducible representation
of a group must be proportional to the identity and irreducible representations
of different dimensions cannot be connected by a non-trivial similarity trans-
formation.
Theorem 6.1. Schur’s lemma: Given D1 and D2 two irreducible representations
of G (D1 : G × V1 → V1 and D2 : G × V2 → V2 ), if T is a linear map (a matrix)
T : V2 → V1 such that

D1 ( g) T = TD2 ( g), ∀ g ∈ G,

then we have one of these two cases:


1. T = 0;

2. T is a bijection and D1 and D2 are equivalent.


Proof. We first show that ker T is an invariant subspace of V2 , namely that
D2 (ker T ) ⊆ ker T. A vector v is in the kernel of the map T, v ∈ ker T, iff

61
62 properties of irreducible representations

Tv = 0. If we act on a null vector with the irreducible representation D1 we are


bound to obtain again a null vector. However, by using the main assumption
of the theorem, we see that

0 = D1 ( Tv) = T ( D2 v)

and therefore D2 v ∈ ker T. Since D2 is an irreducible representation we con-


clude that either ker T = V2 or that ker T = {0}. If ker T = V2 , T maps every
element of V2 to zero and therefore T = 0, closing our proof. Otherwise, we
proved that the map T is injective. We can now prove that it is bijective by
showing that im T ⊆ V1 is an invariant subspace for D1 . Take a vector w in the
image of the map T: w ∈ im T ⇔ w = Tv. Using again the main assumption of
the theorem, we see that the action of D1 on this vector gives again a vector in
the image of T:
D1 w = D1 ( Tv) = T ( D2 v) ∈ im T.
Since also D1 is irreducible we conclude that either im T = V1 or im T = {0}.
If im T = {0} we would have again that T = 0, but we already excluded that,
because we assumed that the map was injective. We therefore conclude that
im T = V1 . This means that our map T satisfies both im T = V1 and ker T = {0},
from which we conclude that dim V2 = dim V1 and that T is an invertible
square matrix. Summarizing, if V1 6= V2 then T = 0. If V1 ' V2 , then either
T = 0 or D1 and D2 are equivalent and T is the similarity transformation.

A direct consequence of this theorem is the following corollary:

Theorem 6.2. An invertible matrix T commuting with all the matrices of an irre-
ducible representation is proportional to the identity:

[ D ( g), T ] = 0, ∀g ∈ G ⇒ T = c 1.

Proof. This is just a special case of the previous theorem where D = D1 = D2


and we have TD e = D T, e for Te = T − c 1 (if a matrix T commutes with D also
T
e commutes with it). In this case, however, we just need to show that ker T e
is an invariant subspace for D. This follows in the same way as before using
the assumption: 0 = D ( Tv e ) = T e( Dv). The consequence is once again that
either ker T = {0} or ker T = V. Now, however, T is invertible and therefore
e e
e 6= {0} because there is at least one eigenvector of T with eigenvalue λ,
ker T
which means that Tv = λv for this vector. Choosing c = λ, we conclude that
e = 0. Hence ker T
Tv e = V and then T e = T − c 1 = 0.
6.1 the great orthogonality theorem 63

We also have another corollary:


Theorem 6.3. Irreducible representations on C for an Abelian group are 1-dimensional.
Proof. If G is Abelian we must have that [ D ( g), D ( g0 )] = 0, ∀ g, g0 ∈ G. This
implies that D ( g) = c g 1, again ∀ g ∈ G, but this would be reducible, unless it
is 1-dimensional.

Let us stress that this corollary uses the fact that the vector space on which
the representation acts is V = C N . As we have seen before we can have irre-
ducible 2-dimensional representations for SO(2) over the reals.

6.1 the great orthogonality theorem

Now that we established these technical results, we can finally prove the great
orthogonality theorem between matrix elements.
In order to introduce an orthogonality criterion, we need first to introduce
an inner product between matrix elements. This is easily identified if we real-
ize that matrix elements can be viewed as complex functions defined over the
group G:
Dij : G → C.
They are therefore part of the vector space C[ G ] of complex functions defined
over G and we can use as inner product the ordinary inner product for a com-
plex vector space. Note that this vector space has complex dimension equal to
the order of the group, because C[ G ] ' C#G , where each factor C is the set of
all possible values on which one can map each element g ∈ G. We can then
define a scalar product on this space by summing the values of the functions
f , h ∈ C[ G ] (vectors in this space) over their components, labelled by the group
elements:
1
( f , h) ≡
#G ∑ f ( g ) h ( g ). (6.1)
g∈ G

Using this scalar product we can finally introduce the concept of orthogonality
between matrix elements.
Theorem 6.4. Great Orthogonality Theorem. Given D1 and D2 irreducible repre-
sentations of dimensions N1 and N2 of a finite group G. We have that either D1 '
/ D2
and then
( Dij1 , Dkl
2
) = 0,
64 properties of irreducible representations

or D1 ' D2 (which means that N = N1 = N1 and that there is a basis where


D = D1 = D2 ) and then
1
( Dij , Dkl ) = δik δjl .
N
Proof. The proof is tedious, but straightforward. We apply Schur’s lemmas,
building an appropriate map interpolating between D1 and D2 . We start by
defining a basis for the N2 × N1 matrices:
(kl )
Eij ≡ δik δjl ,

where kl label different matrices and ij are the row and column indices. We
then build T such that D1 T = TD2 . There are actually N1 N2 matrices that play
this role and they are given by

T ( E(kl ) ) = ∑ D1 ( g) E(kl ) D2† ( g). (6.2)


g∈ G

We can indeed check that


D1 (h) T ( E(kl ) ) = ∑ g D1 (h ◦ g) E(kl ) D2 ( g−1 ) = ∑ g̃ D1 ( g̃) E(kl ) D2 ( g̃−1 ◦ h)
(6.3)
= ∑ g̃ D1 ( g̃) E(kl ) D2 ( g̃−1 ) D2 (h) = T ( E(kl ) ) D2 (h),
where we used the rearrangement theorem ( g̃ = h ◦ g) as well as the fact that for
a finite group all representations are equivalent to unitary ones (which means
that we can assume D2 ( g−1 ) = D2−1 ( g) = D2† ( g)). Because of Schur’s lemmas
we conclude that if the two representations have different dimensions N1 6= N2
the interpolating matrix T must vanish, which means that

Tij ( E(kl ) ) = ∑ Dim


1 k l 2
( g)δm δn D jn ( g) = #G ( D2jl , Dik
1
) = 0. (6.4)
g

If, on the other hand, N1 = N2 = N, then we have that either the two represen-
tations are inequivalent, and then again T = 0, or they are equivalent D1 ' D2 ,
in which case we can assume they coincide. If D1 = D2 = D then we can repeat
the same argument, using the corollary of Schur’s lemmas to argue that

T ( E(kl ) ) = c(kl ) 1. (6.5)

This means that Tij ( E(kl ) ) = c(kl ) δij , whose trace determines the constants c(kl ) :

Tii ( E(kl ) ) = Nc(kl ) . (6.6)


6.1 the great orthogonality theorem 65

For a finite group we can always choose for D a unitary representation (D † D =


1) and therefore

Tii ( E(kl ) ) = ∑ Dim ( g)δmk δnl Dni† ( g) = ∑ δkl = #G δlk . (6.7)


g g

We are then able to completely fix the proportionality coefficient


#G kl
c(kl ) = δ (6.8)
N
and close our proof.

We can also combine the two conditions in a unique condition on two generic
irreducible representations a and b saying that

δ ab
b
( Dija , Dkl )= δ δ . (6.9)
N ik jl
From this theorem we also get an important limit on the number of inequiv-
alent irreducible representations for a finite group. In fact we said we can con-
sider the matrix elements of a given representation as elements in the vector
space C[ G ]. Since dim C[ G ] = #G we can have at most #G mutually orthogonal
vectors in this space. On the other hand, for each representation a of dimension
Na we can have at most Na2 different matrix elements, from which we deduce
that the total of orthogonal elements for a given representation is Na2 . Summing
over al irreducible representations we obtain the inequality

∑ Na2 ≤ #G. (6.10)


a

We will actually see in the next lecture that such inequality should be an exact
equality.

exercises

1. Show that the matrices


D (1) = D ( a) = D (b) = 12 ,
 √ 
1 −√1 − 3
D (c) = D (d) = D (e) = ,
2 − 3 1
66 properties of irreducible representations

provide a representation for the 6 group elements of D3 . Use Schur’s


lemma to determine if this representation is reducible or not and if re-
ducible show which irreducible representations compose the one given.

2. Verify the Great Orthogonality Theorem for the following irreducible rep-
resentation of S3 :
 √   
1 −√1 3 −1 0
D (ρ) = , D (σ) = .
2 − 3 −1 0 1

3. Verify the Great Orthogonality Theorem for the following representation


of Z3 :
 √ 
1 −√1 3
D (ρ) = .
2 − 3 −1
7
P R O P E R T I E S O F I R R E D U C I B L E R E P R E S E N TAT I O N S

We saw that the Great Orthogonality Theorem (GOT) gives us a chance to


distinguish inequivalent representations, by providing a statement about the
orthogonality of matrix elements corresponding to different irreducible repre-
sentations of a given group. Unfortunately, the procedure needed to prove that
two representations are inequivalent by means of the GOT can be rather te-
dious. In this lecture we are going to see how we can obtain analogous results
(derived from the GOT) by introducing the concept of character for a representa-
tion. Characters, as we will see, allow us to obtain a lot of useful information on
the irreducible representations of a given group as well as on their properties.
The character of a representation is a particular function in C[ G ], which has
different values only on certain classes of elements of G. For this reason, in or-
der to introduce it in an efficient way, we first need to consider further relations
between the elements of G.

7.1 conjugacy classes

Recall that if g1 = gg2 g−1 for some g ∈ G, then g1 and g2 are conjugate. Also
recall that conjugation is an equivalence relation. We can therefore split a given
group in disjoint subsets defined by equivalence classes under conjugation as
follows:
Definition 7.1. All the elements of a group G that are conjugate to g ∈ G define
the class of conjugation of g.
We can immediately see that the identity is always alone in its conjugacy
class:
1 = g ◦ 1 ◦ g −1 , ∀ g ∈ G. (7.1)

67
68 properties of irreducible representations

We can also see that all the equivalence classes in an Abelian group have a
single element:

g̃ ◦ g ◦ g̃−1 = g ◦ g̃ ◦ g̃−1 = g, ∀ g, g̃ ∈ G. (7.2)

This implies that an Abelian group has #G conjugacy classes.


For a generic non-Abelian group the number of conjugacy classes is less than
the number of elements and their union will give back the full set of group
elements.
Example 7.1. As an example we discuss here the conjugacy classes of D3 ' S3 .
One equivalence class is given by the identity, as expected: C1 = {1}. The other
elements are either of the form ρ a , for a = 1, 2, with inverse ρ− a , or of the form
σρb , where b = 0, 1, 2, whose inverse is again σρb (in fact σρb σρb = σσρ−b ρ a =
1). The conjugacy class of ρ is then determined by

ρ a ρρ− a = ρ, σρ a ρ a σρ a = σ2 ρ− a−1 ρ a = ρ−1 = ρ2 . (7.3)

Since conjugation is an equivalence relation, if ρ2 is conjugate to ρ, we also


have that ρ is conjugate to ρ2 . The conjugacy class of σρ on the other hand is
determined by

ρ a σρρ− a = σρ1−2a , σρ a σρσρ a = σρ2a−1 . (7.4)

The group elements split therefore into three different conjugacy classes

C1 = {1}, C2 = {ρ, ρ2 }, C3 = {σ, σρ, σρ2 }. (7.5)

Once we group the elements of G in conjugacy classes we can study them


collectively. In particular we can study the functions defined on these classes.
Definition 7.2. A function that is constant on each conjugacy class is called
central:
f ( g) = f ( g̃g g̃−1 ), ∀ g̃ ∈ G.
As we will now see, the character of a representation is a special central
function.

7.2 characters and their properties

Definition 7.3. The character of a representation is the function χ D : G → C


defined as
χ D ( g) = tr D ( g), for g ∈ G.
7.2 characters and their properties 69

We can see immediately that the character is a central function:

χ D ( g g̃g−1 ) = tr( D ( g g̃g−1 )) = tr D ( g) D ( g̃) D ( g−1 )



(7.6)
= tr D ( g) D ( g̃)[ D ( g)]−1 = trD ( g) = χ D ( g̃),


where we used the cyclic property of the trace. Because of this fact we can see
that the value of χ D is the same on all the elements of a conjugacy class. Using
the cyclic property of the trace we can also conclude that all equivalent repre-
sentations have the same character. If two representations are equivalent then
D1 = SD2 S−1 and therefore χ D1 ( g) = trD1 ( g) = tr(SD2 ( g)S−1 ) = tr( D2 ( g)) =
χ D2 ( g).
As we did for the single matrix elements of a representation, we can intro-
duce the concept of orthogonal characters, as follows from their interpretation
as elements of the vector space C[ G ], where we introduced the inner product
(6.1). Actually, it is an immediate consequence of the GOT that

Theorem 7.1. Characters associated to inequivalent irreducible representations are


orthogonal:
(χ Da , χ Db ) = δab .

Proof. It follows straightforwardly from the definition:

δ ab δ
( Diia , Dkk
b
)= δ δ = δ ab ii = δ ab . (7.7)
N ik ik N

However we can say more. For finite groups all decomposable represen-
tations are also reducible. This means that they are always direct sums of
irreducible representations. We then see that

Theorem 7.2. Representations with the same set of characters are equivalent.

Proof. If D is reducible then D = ∑ a c a Da = Da ⊕ Da ⊕ . . . ⊕ Da ⊕ . . . From this


| {z }
ca
we get that

(χ D , χ a ) = c a (7.8)

and hence D is fully determined by χ D .


70 properties of irreducible representations

Since we just saw that characters fully specify a representation we can limit
our study to the irreducible ones. The first thing we should determine then
is how many inequivalent irreducible representations we can construct for a
given group. This is determined by a theorem in the following way:

Theorem 7.3. The number of inequivalent irreducible representations of a finite group


is equal to the number of conjugacy classes.

Proof. The proof has three steps.


First we show that matrix elements are a good basis for C[ G ].
Given a representation D a , the matrix elements Dija are orthogonal among
them and they are elements of C[ G ]. In order to form a basis of C[ G ], we must
show that they span C[ G ], i.e. that any arbitrary function f ∈ C[ G ] can be
expanded in terms of matrix elements of irreducible representations of G. For
any given g ∈ G we can write

f ( g) = f g = ∑0 f g0 δg0 g = ∑0 f g0 DgR0 1 ( g), (7.9)


g g

where the first equality is really just a definition of an element of C#G and
the last one follows from the definition of regular representation. The regular
representation is reducible and using previous results we can always write
D R = ∑ a c a Da , which implies

f ( g) = ∑ fija Dija ( g). (7.10)


a,i,j

Hence any element of C[ G ] can be written by using Dija as a basis.


The second step is the proof that if f ∈ C[ G ] is a central function then we
can expand it using characters as a basis:

f ( g) = ∑ f a ij Dija ( g) = f ( g̃g g̃−1 ) = ∑ f a ij Dija ( g̃g g̃−1 )


a ( g̃ ) D a ( g ) D a ( g̃−1 )
= ∑ f a ij Dik km mj (7.11)
a
a ( g̃ ) D a ( g ) D ( g̃ )
= ∑ f a ij Dik km jm
7.3 character tables 71

and being central f ( g) = f ( g̃g g̃−1 ) = 1


#G ∑ g̃ f ( g̃g g̃−1 ), which means that
1 a
a ( g̃ ) D ( g̃ ) D a ( g )
f ( g) = #G ∑ g̃ ∑ a f a ij Dik jm km
a , D a ) D a ( g)
= ∑ a f a ij ( D jm ik km
δij δkm a (7.12)
= ∑ a f a ij Na Dkm ( g )
f a ii
= ∑a Na χ D a ( g ).

Hence characters form a basis for central functions.


The third step is simply a correspondence between characters and central
functions. Since inequivalent irreducible representations have orthogonal char-
acters, they are in 1 to 1 correspondence with the basis of characters, which
are also a basis for central functions. This also means that characters are in 1
to 1 correspondence with conjugacy classes and that in turn these are in 1 to 1
correspondence with irreducible representations.

7.3 character tables

As we said before, characters are central functions and therefore they have the
same value on each element of the same conjugacy class. This means that when
we evaluate the inner product between two characters we can actually restrict
the sum to elements of different conjugacy classes:
1
(χ a , χb ) =
#G ∑ nc χa ( gc )χb ( gc ) = δab , (7.13)
c

where c runs on the different conjugacy classes and nc is the number of ele-
ments in the c conjugacy class. Using this rewriting we can now produce some
interesting definitions and relations that will completely specify irreducible rep-
resentations of a given group. The first thing we can introduce is the matrix of
characters. For any group we have the same number of conjugacy classes and
irreducible representations and this means that to fully specify all irreducible
representations of a given group we need a square matrix encoding the value
of the character of a given irreducible representation on an element of a given
conjugacy class, for each one of them.
Definition 7.4. The matrix of characters of a representation is
r
na
Uab ≡ χ ( g a ).
#G b
72 properties of irreducible representations

This matrix has some very interesting properties that restrict its possible
entries.

Theorem 7.4. The character matrix is unitary

Proof.
r 2
nb

Uab Ubc = Ūba Ubc = ∑ #G
χ a ( gb )χc ( gb ) = δac . (7.14)
b

This property means that different rows or different columns are orthogonal
and therefore we can fix their entries. Using it in the opposite way we can also
say more and actually prove a theorem that allows us to fix the number and
dimensions of the irreducible representations of a finite group.

Theorem 7.5.
∑ Na2 = #G.
a

Proof. We use unitarity of the character matrix, only in a way different from
(7.14):

n a nc

Uab Ubc = Uab U cb = ∑ χb ( gc )χb ( ga ) = δac . (7.15)
b
#G

If we specify this formula for ga = gc = 1, we can also conclude that

∑ χb (1)χb (1) = #G, (7.16)


b

where we used that n a = nc = 1, and since now χ a (1) = tr(1 Na ) = Na we


conclude.

Another noteworthy relation follows from what we just proved and tells us
when a representation is irreducible:

Theorem 7.6. If ∑c nc χ D ( gc )χ D ( gc ) = #G then D is irreducible.


† , for the
Proof. It follows from the orthogonality relation given above Uab Ubc
choice gc = ga .
7.3 character tables 73

Often we do not want to keep track of the normalization factors in defining


the unitary character matrix and therefore we introduce

Definition 7.5. The character table is the matrix collecting all the values of the
characters for each irreducible representations and for every conjugacy class.

We conclude this lecture by computing the character table for a couple of


examples.
Example 7.2. The first example is given by the group Z3 .
Each Zn is Abelian and therefore we have n conjugacy classes and n irre-
ducible representations. As we already discussed, all these representations
must be 1-dimensional and in fact they fulfill the condition ∑na=1 Na2 = n = 12 +
2πi
. . . + 12 . All these 1-dimensional representations can be defined by ω k = e n k ,
where k = 1, . . . , n.
For Z3 the conjugacy classes are then C1 = {1}, C2 = {ρ}, C3 = {ρ2 } and
the representations are generated by D1 (ρ) = 1, D10 (ρ) = e2πi/3 = ω and
D1” (ρ) = ω 2 . We eventually deduce the character table, which is

C1 C2 C3
1 1 1 1
(7.17)
10 1 ω ω2
1” 1 ω2 ω

Example 7.3. We now discuss D3 ' S3 .


We already know that the conjugacy classes are C1 = {1}, C2 = {ρ, ρ2 } and
C3 = {σ, σρ, σρ2 }. Since irreducible representations must be in 1 to 1 corre-
spondence with the conjugacy classes, we conclude that we have 3 irreducible
representations and from theorem 7.5 we can deduce that their dimension is 1,
1 and 2. In fact

#G = 6 = 12 + 12 + 22 . (7.18)

The characters of the trivial representation are 1 and that of the dimension 2
representation can be computed from their explicit expression, we gave previ-
ously. The rest follows by orthogonality

C1 C2 C3
1 1 1 1
(7.19)
10 1 1 −1
2 2 −1 0
74 properties of irreducible representations

exercises

1. Consider the group of geometric symmetries of a square (D4 ). Split its


elements in conjugacy classes, find all its subgroups and discuss which
ones are normal. Compute the quotient of D4 with its normal subgroups
and identify the resulting group.
2. Find all irreducible representations of D4 .
3. Take the representations of D4 generated by
   
0 1 1 0
D1 (ρ) = , D1 (σ) = ,
−1 0 0 −1
   
83 130 43 66
D2 (ρ) = , D2 (σ) = .
−53 −83 −28 −43
Using characters, discuss if they are equivalent.
4. Prove that given two representations D1 and D2 , χ D1 ⊗ D2 = χ D1 · χ D2 .
5. Prove that for unitary representations χ D ( g−1 ) = (χ D ( g))∗ .
6. Consider once more the group of quaternions Q.
a) Provide a matrix representation for each of its irreducible represen-
tations.
Hint: note that Q can be constructed as a discrete
 subset of the SU(2)
cos θ eiα sin θ eiβ
matrices U (θ, α, β) =
− sin θ e−iβ cos θ e−iα
b) Give its character table.
7. The dicyclic group Γ( Q6 ) is defined by Γ( Q6 ) = h a, b| a6 = 1, b2 = a3 , aba =
bi. Its conjugacy classes are

C1 = {1}, C2 = { a, a5 }, C3 = { a2 , a4 },

C4 = { a3 }, C5 = {b, a2 b, a4 b}, C6 = { ab, a3 b, a5 b}.


a) Are the following subgroups normal?
Z2 = {1, a3 }, Z3 = {1, a2 , a4 }, Z4 = {1, b, b2 , b3 }.

Hint: use the information contained in the conjugacy classes.


7.3 character tables 75

b) Compute Γ( Q6 )/Z3 , where Z3 is the one considered in the previous


question.
c) Discuss which of the following representations are equivalent:
 iπ/3   
e 0 0 1
D1 ( a) = , D1 (b) = ,
0 e−iπ/3 −1 0

ei2π/3
   
0 0 1
D2 ( a) = , D2 (b) = ,
0 e−i2π/3 1 0

e−i2π/3
   
0 0 −1
D3 ( a) = , D3 (b) = ,
0 e2iπ/3 −1 0
   
cos π3 i sin π3 0 −1
D4 ( a) = , D4 (b) = .
i sin π3 cos π3 1 0
8
I R R E D U C I B L E R E P R E S E N TAT I O N S O F S n

All finite groups are subgroups of some group of permutations Sn . Also, ir-
reducible representations of any H < G can be obtained by restricting the
representations of G to the elements of H. Combining these two facts we see
that we can obtain the representations of any finite group by studying and
building the irreducible representations of an appropriate permutation group.
It is therefore extremely useful and interesting to try to provide a classification
of the irreducible representations of Sn .
In order to do so, we will use a graphical description that is very efficient:
Young Tableaux. This type of diagrams not only is useful to discuss and classify
both conjugacy classes as well as irreducible representations of permutation
groups, but, as we will see in a subsequent lecture, this graphical description
will also be very useful for the construction and classification of representations
of continuous groups.
Definition 8.1. A standard Young Tableau of order k is a diagram with k boxes,
where the ( j + 1)th row contains a number of boxes that is less or equal to that
of the jth row and the ( j + 1)th column contains a number of boxes that is less
or equal to that of the jth column.
For instance, good Young Tableaux are

, , ,

while the following are not good Young Tableaux:

, , , .

77
78 irreducible representations of s n

8.1 young tableaux and conjugacy classes

In this first part we discuss the relation between Young Tableaux and conjugacy
classes. We will see that there is a 1-to-1 correspondence between regular Young
Tableaux and conjugacy classes. In order to establish such a correspondence we
first need to classify the conjugacy classes of Sn .
Recall that any permutation associated to an irreducible j-cycle can be written
as a product of 2-cycles. For instance

(12 . . . m) = (1 m)(1 m − 1) . . . (13)(12).

We also established that the inverse of a given j-cycle is a j-cycle with the same
elements written in the opposite order, i.e.

(12 . . . m)−1 = (1 m . . . 2),

and this in turn can be written as the product of the same 2-cycles defining
the original j-cycle, but in the opposite order, (1m . . . 2) = (12)(13) . . . (1 m −
1)(1 m). This implies that we can constrain our analysis of the conjugacy classes
of Sn using 2-cycles.
It is easy to see that the conjugates of a j-cycle by 2-cycles is again a j-cycle,
where the elements of the 2-cycle in the j-cycle are swapped:

(12) · ( a1b . . . c2d . . .) · (12) = ( a2b . . . c1d . . .). (8.1)

An analogous demonstration works for composite cycles (like, for instance,


(12)(345) ∈ S5 ).
Since conjugation by 2-cycles can change any element in any cycle, we see
that all the cycles with the same length must be in the same conjugacy class.
We can then put all conjugacy classes in 1-to-1 correspondence with Young
Tableaux where each column represents one of the cycles defining the conju-
gacy class. For instance, the cycle (1352)(4) in S5 is composed by a 4-cycle and
a 1-cycle. The corresponding representation via a Young Tableaux is therefore

. The same thing is true for the cycle (361)(24)(5) ∈ S6 , which is an element

in the conjugacy class represented by .


Example 8.1. Conjugacy classes of S3 . We already know that we have 3 conjugacy
classes C1 = {1 = (1)(2)(3)}, C2 = { p1 = (123), p2 = (132)} and C3 = {d1 =
(1)(23), d2 = (2)(13), d3 = (3)(12)}. As expected, they can be distinguished
8.1 young tableaux and conjugacy classes 79

by the length of the cycles defining them. In terms of Young Tableaux we can
represent them as

C1 ∼ , C2 ∼ , C3 ∼ .

From this analysis we can deduce a general lesson: we can find all the conju-
gacy classes of Sn by considering all Young Tableaux with n boxes arranged in
all possible admissible ways. We also learn that the number of classes depends
on the number of ways we can partition n in terms of integer numbers. For
instance, for S3 , we split n = 3 in 3 possible ways
3 = 1 + 1 + 1 = 2 + 1 = 3, (8.2)
corresponding to its three conjugacy classes and to its 3 inequivalent Young
Tableaux.
The shape of the tableau also tells us how many elements are present in
each conjugacy class. This is determined by the number of different ways
one can arrange the n elements one is permuting in the boxes of the tableau.
For instance, the Young Tableau corresponds to a triple 1-cycle and since
(1)(2)(3) = (1)(3)(2) = (2)(1)(3) = . . . = (3)(2)(1) we see that there is only
one element in this class:
1 2 3 = 1 3 2 = 2 1 3 = ... = 3 2 1 .

The Young Tableau on the other hand has three different elements in its
class
1 3 = (12)(3), 2 3 = (23)(1), 1 3 = (13)(2). (8.3)
2 1 2
We see then that computing the number of elements in the conjugacy class
defined by the Young Tableau under consideration follows by a combinatorial
analysis. In detail, if we have k j j-cycles, the corresponding Tableau represents
n!
nr. of permutations = k
. (8.4)
∏nj=1 j j k j !
Example 8.2. The Tableau has n = 3, k1 = 3, k2 = k3 = 0 and hence it
contains 1 element. The Tableau has n = 3, k1 = k2 = 1 and k3 = 0, hence it

contains 3 elements. The Tableau has n = 3, k1 = k2 = 0 and k3 = 1, hence it


contains 2 elements.
80 irreducible representations of s n

8.2 young tableaux and irreducible representations of Sn

We established that Young Tableaux are in 1-to-1 correspondence with conju-


gacy classes of the permutation group. On the other hand, we also know that
irreducible representations are in 1 to 1 correspondence with conjugacy classes.
This means that we can use Young Tableaux of rank n to represent irreducible
representations of the group of permutations Sn , too.
Each Young Tableau of rank k is associated to an irreducible representation
of dimension

n!
d= , (8.5)
∏ik=1 hi

where hi is the hook number associated to each single box in the diagram. The
hook number is the sum of all the boxes to the right and below a given box
plus one (see the picture 7 for an example).

9 6 3 1
7 4 1
1 5 2
4 1
2
4 1

Figure 7.: Example of hook number computation.

The proof of this formula is very long, tedious and not particularly illumi-
nating. For this reason, we will not see here a full proof of this formula, but
we will discuss a significative example that is also useful to understand how to
construct all the irreducible representations of a given permutation group.
The starting point is given by the regular representation of Sn , which is
reducible and decomposes in the sum of all the irreducible representations
of a given permutation group. For the sake of simplicity we will focus our
analysis to the group S3 , but we will mention the consequences of consider-
ing general n whenever possible. The regular representation of S3 is a map
8.2 young tableaux and irreducible representations of Sn 81

D R : S3 × C6 → C6 , where the natural basis for the vector space C6 is given by


vectors associated to the elements of S3 :
| gi = {|1i, |(12)i, |(13)i, |(23)i, |(123)i, |(132)i} (8.6)
and the explicit action is D R ( g)| g̃i = | g ◦ g̃i. We can then decompose the
matrices D R ( g) by looking at its action on the subspaces associated to each
Young Tableau as follows from the theorem:
Theorem 8.1. Given a Young Tableau T, the vectors |vi ≡ a T s T |1i generate invariant
subspaces under the action of the regular representation.
Here s T and a T are the operators that symmetrize all the elements in the rows
of a given Young Tableau and anti-symmetrize the elements in the columns as
follows. Given a Tableau T, we can fill its boxes with numbers from 1 to n in all
possible ways and define HT , which contains all permutations of the numbers
in the rows, and VT , which contains all the permutations of the numbers in the
columns. The symmetrizing and anti-symmetrizing operators are then defined
by
sT ≡ ∑ D R ( σ ), aT = ∑ π ( σ ) D R ( σ ), (8.7)
σ ∈ HT σ ∈VT

where π is the parity of the element σ ∈ Sn , defined in Def. 3.13. For in-
1 3
stance, the tableau 2 corresponds to the operators s T = D R (1) + D R (13),
a T = D R (1) − D R (12).

We will now explicitly show that for S3 the regular representation admits
invariant subspaces associated to the vector spaces V , V and V , defined

by the corresponding Young Tableaux.


Take first V . We can fill the corresponding diagram in 3! different ways:

1 2 3, 1 3 2, 2 1 3, 2 3 1, 3 1 2, 3 2 1 (8.8)
and we can see that they are related by all the permutations in S3 . Conse-
quently s T is given by the sum of all the elements in S3 and a T = D R (1). The
corresponding vector, generating V , is
i = D R (1) D R (1) + D R (12) + D R (13) + D R (23)

|v

+ D R (123) + D R (132) |1i



(8.9)

= |1i + |(12)i + |(13)i + |(23)i + |(123)i + |(132)i.


82 irreducible representations of s n

This means that V is a 1-dimensional subspace and, since it is generated


by the sum of all the vectors corresponding to the different elements of S3 ,
the action of the regular representation on |v i is trivial: each element is
mapped to a different element in S3 and the sum of all the elements is mapped
to itself, because of the rearrangement theorem

D R ( g)|v i = |v i, ∀ g ∈ S3 . (8.10)

The second invariant subspace V is again 1-dimensional and is spanned by

the vector

D R (1) − D R (12) − D R (13) − D R (23)



|v i =

+ D R (123) + D R (132) D R (1)|1i



(8.11)

= |1i − |(12)i − |(13)i − |(23)i + |(123)i + |(132)i.

In this case the regular representation reduces to a 1-dimensional representa-


tion determined by the parity of the element which is representing. For in-
stance:

D R (12)|v i = |(12)i − |1i − |(123)i − |(132)i + |(23)i + |(13)i = −|v i. (8.12)

We are left with the space V . In this case s T and a T are not uniquely
determined, but they depend on the arrangement of the elements of Sn in the
1 2
tableau. For instance, for 3 we have s T = D R (1) + D R (12) and a T = D R (1) −
1 3
D R (13), while for we have s T = D R (1) + D R (13) and a T = D R (1) − D R (12).
2

The corresponding vectors are

|v 1 2 i = |1i + |12i − |13i − |123i, (8.13)


3

|v 1 3 i = |1i + |13i − |12i − |132i. (8.14)


2
8.2 young tableaux and irreducible representations of Sn 83

These vectors however are not invariant under the action of the elements of S3 ,
but rather generate two orthogonal invariant subspaces of dimension 2. In fact,
the action of the S3 elements on the first gives

D R (1)|v 1 2 i = | v 1 2 i,
3 3

D R (12)|v 1 2 i = |(12)i + |1i − |(132)i − |(23)i ≡ |wi,


3

D R (13)|v 1 2 i = |(13)i + |(123)i − |1i − |(12)i = −|v 1 2 , i (8.15)


3 3

D R (23)|v 1 2 i = |(23)i + |(132)i − |(123)i − |(13)i = −|wi + |v 1 2 i,


3 3

D R (123)|v 1 2 i = |(123)i + |(13)i − |(23)i − |(132)i = |wi − |v 1 2 i,


3 3

D R (132)|v 1 2 i = |(132)i + |(23)i − |(12)i − |1i = −|wi.


3

By consistency, the application of the same representation matrices to |wi re-


mains also in the same 2-dimensional subspace, generated by |wi and |v 1 2 i
3
and hence, we can reconstruct the 2-dimensional representation of S3 by look-
ing at D R restricted to this space.
The same computation on the second element gives, on the other hand:

D R (1)|v 1 3 i = | v 1 3 i,
2 2

D R (12)|v 1 3 i = |(12)i + |(132)i − |1i − |(13)i = −|v 1 3 i,


2 2

D R (13)|v 1 3 i = |(13)i + |1i − |(123)i − |(23)i ≡ |ui, (8.16)


2

D R (23)|v 1 3 i = |(23)i + |(123)i − |(132)i − |(12)i = −|ui + |v 1 3 i,


2 2

D R (123)|v 1 3 i = |(123)i + |(23)i − |(13)i − |1i = −|ui,


2

D R (132)|v 1 3 i = |(132)i + |(12)i − |(23)i − |(123)i = −|v 1 3 i + |ui.


2 2
84 irreducible representations of sn

Also in this case, by consistency, the application of D R to |ui remains in the


same 2-dimensional subspace and its restriction to this subspace gives a 2-
dimensional irreducible representation that is equivalent to the previous one.

We conclude that we have two orthogonal invariant subspaces of C6 corre-


sponding to the two Young Tableaux we started from and by a more accurate
computation we can also check that the vectors |wi and |ui can be associated
to linear combinations of vectors corresponding to other Young Tableaux of the
same type. We have then determined 4 invariant subspaces of dimensions 1, 1,
2 and 2 and therefore we exhausted the rewriting of the original basis of vectors
generating C6 in terms of new ones associated to the same invariant subspaces.
We conclude that

D R = D1 + D10 + D2 + D2 , (8.17)

where the 2-dimensional representations are obviously equivalent, because we


know that there is only one such irreducible representation for S3 . This is
actually a result that applies in general for the decomposition of the regular
representation of Sn , where each irreducible representation appears a number
of times equal to its dimension. The important lesson we learn from this is that
among all possible arrangements in the tableaux , only 2 of them generate
orthogonal subspaces and these are in correspondence with the tableaux where
the numbers are always growing in any row and column of the diagram. By
repeating this decomposition for a generic Sn we can then deduce the general
formula for the dimension of its irreducible representations, but also construct
them explicitly.

The general formula for the dimension of the irreducible representations of


Sn associated to a given Young Tableau can also be obtained by induction, con-
sidering the Tableaux of Sn−1 < Sn . In fact, each Young Tableau of Sn contains
only a subset of the diagrams associated to Sn−1 and hence its dimension is
the sum of the dimensions of the corresponding diagrams. In this way we can
determine the resulting dimension by an iterative procedure, like in figure 8.
8.2 young tableaux and irreducible representations of Sn 85

Figure 8.: Dimension of Sn representations determined by induction from the dimen-


sion of the representations of Sn−1 .

exercises

1. Find the character table of S4 . To do so, we can use that the charac-
ters of the symmetric group are always integers and that the character
of the product of two representations is the product of the characters,
i.e. χ Da ⊗ Db = χ a · χb .
a) Draw all Young Tableaux with 4 boxes and then determine the con-
jugacy classes, the number of their elements and the dimension of
the corresponding irreducible representations.
b) Fill the column related to χ(1) and the raw associated to the trivial
representation.
c) By using the results of this chapter, fill the raw χ1 A , where 1 A refers
to the 1-dimensional alternating representation defined by the parity
of the elements of S4 .
d) Using χ Da ⊗ Db = χ a · χb deduce that 2 ⊗ 1 A ' 2. What does this
imply for the conjugacy classes containing elements constructed by
odd numbers of 2-cycles?
e) Use orthogonality of the characters of irreducible representations to
complete χ2 .
86 irreducible representations of s n

f) Argue that for the 3-dimensional representations we must have 30 =


3 ⊗ 1A.
g) Complete the table using again that (χ a , χb ) = δab for irreducible
representations.
Part III

ELEMENTS OF TOPOLOGY AND DIFFERENTIAL


G E O M E T RY
9
HOMOTOPY GROUPS

So far we discussed general properties of groups and some of their applica-


tions to finite groups. On the other hand, in physics we have to deal very
often with symmetry transformations that are continuous (e.g. rotations) and
hence we must expand our discussion to include symmetry groups with an
infinite number of elements. In order to do so, we need to introduce and re-
call some concepts in topology and differential geometry that are relevant for
understanding their properties. In fact, continuous groups are also topological
(and differential) spaces where each point is identified with an element of the
group.
The first important task is to classify and distinguish them according to their
topological and differential properties. In this lecture we will discuss one partic-
ular topological property that allows to separate spaces (and hence continuous
groups) in equivalence classes: homotopy. Actually, homotopy is not only rel-
evant in the context of group theory, but it has more interesting applications
both in condensed matter physics (where it allows to distinguish materials in
terms of their defects) and in theoretical high energy physics (both in gauge the-
ories and in higher dimensional field theories). For this reason we will give a
general introduction to this concept, with some final remarks on its application
to group theory.
First we need to introduce maps between topological spaces:

Definition 9.1. Let X and Y be two topological spaces. The map φ : X → Y is a


homeomorphism if φ is a continuous bijection and φ−1 is continuous. Two spaces
related by such a map are homeomorphic to each other.

The existence of homeomorphic maps between two spaces implies that we


can continuously deform them into each other. In this sense we can consider

89
90 homotopy groups

them to be topologically equivalent and they will be characterized by the same


topological invariants (like the number of connected components, ...). We can
actually group all spaces with the same topological invariants in equivalence
classes, whose equivalence relations are homeomorphisms. A complete charac-
terization in terms of invariants is rather difficult, but one important thing we
know is that
Theorem 9.1. Topological spaces having different topological invariants are not home-
omorphic.

9.1 homotopy

One way to construct invariants is the use of homotopy. This is a weaker notion
than that coming from homeomorphisms. In fact, in the case of homotopic
spaces we consider continuous functions between them that need not have an
inverse.
Homotopy also creates equivalence classes of continuous maps.
Definition 9.2. Let X and Y be two topological spaces. Consider two contin-
uous maps f 0 , f 1 : X → Y, then f 0 is homotopic to f 1 ( f 0 can be continuously
deformed into f 1 ):

f0 ∼ f1 (9.1)

if there is a continuous function

F : X × [0, 1] → Y (9.2)

such that

F ( x, 0) = f 0 ( x ), and F ( x, 1) = f 1 ( x ). (9.3)

Definition 9.3. F is called a homotopy.


Homotopies define homotopy classes for functions, where two functions lie
in the same class if they are homotopic.
We can also apply the same idea to topological spaces and define homotopic
equivalent spaces.
Definition 9.4. Two topological spaces X and Y are homotopic equivalent if
there are continuous maps f : X → Y and g : Y → X such that

f ◦ g ∼ 1Y and g ◦ f ∼ 1X . (9.4)
9.1 homotopy 91

Note that we did not require that g was the inverse of the function f , but
only that their combination is homotopically equivalent to the identity.
Example 9.1. Here is an example of homotopic spaces. Let X = Rn r {0},
Y = Sn−1 . We can show that these two spaces are homotopic, which means
that we can continuously deform one into the other. The proof is simple if one
consider the maps

~x
f : Rn r {0} 3 ~x 7→ ∈ S n −1 , (9.5)
|~x |

g : Sn−1 3 ~n 7→ ~n ∈ Rn r {0} where |~n| = 1. (9.6)

We see that f ◦ g = 1Y , but g is not the inverse of f , because g ◦ f 6= 1X .


However, we can construct a continuous map F : X × [0, 1] → X that shows
that g ◦ f is homotopically equivalent to the identity map in Y. This is built by
defining F such that F (~x, 0) = 1X (~x ) = ~x, and F (~x, 1) = g ◦ f (~x ) = |~~xx| :

~x
F (~x, t) = (1 − t)~x + t . (9.7)
|~x |

We see then that f ◦ g ∼ 1Y .

S1

Figure 9.: The space Rn r {0} can be squeezed to a sphere Sn−1 .

Example 9.2. We now discuss an interesting example of non-trivial homotopic


maps between X = S1 , the unit circle described by an angle θ, and Y =U(1),
represented by the unimodular complex numbers eiϕ . Consider the maps f α :
X → Y defined by

f α (θ ) = ei(nθ +α) . (9.8)


92 homotopy groups

These maps generate homotopy classes for different α and fixed integer n. Ex-
plicitly, for a fixed n, f α0 ∼ f α1 by means of the continuous family of functions

F (θ, t) = exp(i [nθ + (1 − t)α0 + tα1 ]), (9.9)

so that F (θ, 0) = f α0 (θ ) and F (θ, 1) = f α1 (θ ). An important point, which will


become of primary importance later, is that different integers n and m define
different homotopy classes:

einθ ∼ imθ
/e , if n 6= m. (9.10)

In fact, a map generalizing (9.9) and interpolating between the two would be

exp(i n(1 − t) θ + i t m θ ), (9.11)

but this is not continuous around θ = 0, 2π for t ∈]0, 1[ and generic m and n.
We actually have a multivalued function.
We can then differentiate equivalence classes in terms of the integer defining
the maps f α . This integer is called winding number, because we can see that it
represents how many times the circle is covered when θ goes from 0 to 2π.

n=0 n=1

Figure 10.: Different homotopy classes of functions covering the circle.

We then see that we cannot unwind in a continuous fashion functions with


different winding numbers (see figure 15).

9.2 homotopy groups

Introducing loops, like those mentioned above, in a topological space is useful


to classify the same spaces in terms of their connectedness properties.
9.2 homotopy groups 93

Definition 9.5. A continuous map α : [0, 1] → X, where X is a topological


space, with α(0) = α(1) = x0 ∈ X is called a loop with base point x0 .

Once again we can continuously deform loops among themselves by means


of homotopies. A homotopy between two loops α and β can be constructed by
defining

H : [0, 1] × [0, 1] → X (9.12)

such that

H (s, 0) = α(s), H (s, 1) = β(s), ∀s ∈ [0, 1], (9.13)

H (0, t) = H (1, t) = x0 , ∀t ∈ [0, 1]. (9.14)

 
H s, 41

 
H s, 12

x0

Figure 11.: Example of homotopy shrinking a loop with base point x0 .

Using the homotopy H we can then define equivalence classes of loops.


There are two kinds of loops, those containing a hole in them, which cannot be
shrunk to a point, and those that do not contain a hole and can be shrunk. We
can then use loops to classify topological spaces:

Definition 9.6. If any loop in X can be continuously shrunk to a point then X


is simply connected.
94 homotopy groups

β
x0

Figure 12.: Space X with a hole. Loop α cannot be shrunk to a point. Loop β can.

The interesting point for us is that the equivalence classes defined by loop
homotopies have a group structure.
Definition 9.7. The fundamental group or first homotopy group of a topological
space X at a point x0 , Π1 ( X, x0 ) is the group of equivalence classes of homo-
topic loops on X with base point x0 .
We can check the group properties rather easily by composing loops having
the same base point. In fact we define the product of loops by

 α(2s), for s ∈ [0, 1/2],
β◦α = (9.15)
 β(2s − 1), for s ∈ [1/2, 1].

This means that we first follow the loop α, from x0 to x0 and then the loop β
from the same point to x0 . The result is a special loop with basepoint x0 , which
passes in x0 for s = 1/2. The inverse is trivially defined by

α −1 ( s ) = α (1 − s ), (9.16)

and the identity map is 1(s) = x0 , ∀s. It must be noted that α−1 ◦ α 6= 1, but
obviously it is homotopic to the identity map α−1 ◦ α ∼ 1.
It is very interesting that in the case of arcwise connected topological spaces,
the fundamental group does not depend on the base point x0 anymore:
Theorem 9.2. If X is an arcwise connected topological space and x0 , x1 ∈ X, then

Π1 ( X, x0 ) ' Π1 ( X, x1 ) (9.17)

and we generically talk about Π1 ( X ).


This is understood by looking at figure 14, where the arcwise connectedness
assures the existence of the maps β and β−1 connecting x0 with x1 .
9.3 universal covering 95

α −1 α ◦ α −1
α

x0 x0

Figure 13.: A loop α times its inverse α−1 is homotopic to the identity.

α
β

x0 x1
β −1

Figure 14.: Loops with base-point x0 can be extended to loops with base-point x1 if the
space is arcwise connected.

Example 9.3. From our previous discussion we immediately see that Π1 (U(1)) '
Π1 (S1 ) = Z, because inequivalent loops are classified by n ∈ Z. Combining
loops in the equivalence classes n and m we obtain a loop in the equivalence
class defined by n + m.
Example 9.4. Π1 (Sn≥2 ) = 0. In this case all loops can be shrunk to a point.

9.3 universal covering

Whenever Π1 ( X ) = 0 we see that X is simply connected.


If, on the other hand X is not simply connected, then we can define its univer-
sal covering X,e which is simply connected and which is locally homeomorphic
to X. Actually, it can be proved that X e is unique up to homeomorphisms.
In the special case that X = G is a group, it can also be proved that G e is a
group, which is homomorphic to G (i.e. there is a map ϕ : G → G, called the
e
covering map, respecting the products). Also, the kernel of the homomorphism
ϕ is Π1 ( G ), so that Π1 ( G ) / G
e and therefore

e 1 ( G ) = G.
G/Π (9.18)

For instance, R is the universal covering group of U(1) ' R/Z.


96 homotopy groups

In the following lectures we will also need that Π1 (SO(n)) = Z2 , for n > 2.
We will define then the spin group

^
Spin(n) ' SO ( n ). (9.19)

exercises

1. Consider the disk D2 = {~x ∈ R2 ||~x | ≤ 1} where we make two holes at


points ~x1 and ~x2 . Show that the first homotopy group Π1 ( D2 r {~x1 } r
{~x2 }) is non Abelian.
10
DIFFERENTIABLE MANIFOLDS

10.1 introduction

In the rest of the course we want to analyze infinite-dimensional continuous


groups. In order to do so, we need to introduce some notions of derivability on
these spaces and for this reason we will discuss now differentiable manifolds.
These are one of the most fundamental concepts in mathematics and physics.
We already learned in previous courses how to perform derivatives and in-
tegrals of functions defined on the n-dimensional Euclidean space Rn . In this
lecture we will learn how to extend these concepts to spaces that cannot be
identified directly with Rn , but which look like Rn if considered in appropriate
regions. Generally, we can think of a differentiable manifold M as a smooth
surface that is locally homeomorphic to Rn . We can then associate to each point
of M some coordinates in Rn and require that the transition maps from a set of
coordinates to another be regular. With this approach we can analyze functions
on such spaces by converting them (locally) to functions in a Euclidean space.

10.1.1 Differentiable manifolds

We will introduce the concept of differentiable manifold by introducing one by


one the ingredients needed for its definition. The starting point is obviously a
topological space M, but then we need to add some structure to it. The first
step is to parameterize locally this space by coordinates in Rn .
Definition 10.1. Given a topological space M, a chart (Uα , ϕα ) is a homeomor-
phism ϕα from an open set Uα ⊂ M to an open set Rα ⊂ Rn :

ϕα : M ⊃ Uα → Rα ⊂ Rn .

97
98 differentiable manifolds

ϕ1

Rn

U1 M

ϕ1 ◦ ϕ2−1

U2 ϕ2 ◦ ϕ1−1
Rn

ϕ2

Figure 15.: Example of a manifold with arrows exemplifying its local maps ϕ1 , ϕ2 to
Rn and the maps ϕ1 ◦ ϕ2−1 , ϕ2 ◦ ϕ1−1 giving the coordinate changes in the
overlapping region.

Then we would like that different sets of coordinates on overlapping charts


could be exchanged.

Definition 10.2. Two charts (U1 , ϕ1 ) and (U2 , ϕ2 ) are compatible if ϕ1 ◦ ϕ2−1 ∈
C ∞ and ϕ2 ◦ ϕ1−1 ∈ C ∞ (when U1 ∩ U2 6= ∅).

Finally, we want to cover the whole topological space M with charts, so that
we can give a coordinate in Rn to all its points.

Definition 10.3. The set of (compatible) charts covering M is called atlas.

This may not be a unique procedure and in this case we will say that two
atlases are compatible if all their charts are compatible.
We are now in a position that lets us introduce the concept of differentiable
manifold.

Definition 10.4. A smooth differentiable manifold M satisfies:

1. M is a topological space;

2. there is a family of charts {Uα , ϕα }α covering M, i.e. ∪α Uα = M;


10.1 introduction 99

3. for any couple Uα , Uβ ⊂ M such that Uα ∩ Uβ 6= ∅, ϕ1 ◦ ϕ2−1 ∈ C ∞ and


ϕ2 ◦ ϕ1−1 ∈ C ∞ .
If the last assumption is not satisfied we have a manifold that is not smooth.
In conclusion a manifold is constructed by smoothly sewing together regions
which are homeomorphic to a Euclidean space. A crucial point it that the
dimensionality of the Euclidean spaces being used must be the same in every
patch.
Definition 10.5. The dimension of a manifold M is given by the dimension of
the vector space Rm to which it is locally homeomorphic.
It should be noted that this definition does not rely on an embedding of the
manifold in some higher-dimensional Euclidean space. While any n-dimensional
manifold can be embedded in R2n by Whitney’s embedding theorem, it is im-
portant to recognize the existence of a manifold independently of any embed-
ding.

10.1.2 Properties

Now that we have introduced the notion of a manifold, we can start looking at
its properties and at the possibility of deforming different manifolds onto each
other. Since we are now interested also in the differential structure and not
just to the topological structure, we better introduce a new definition of map
between topological spaces that preserves also the differentiable properties.
Definition 10.6. A map f : M1 → M2 continuous and differentiable with in-
verse continuous and differentiable is a diffeomorphism.
Two manifolds related by a diffeomorphism are called diffeomorphic.
Other properties of manifolds should be decided looking at the underlying
topological space, but, thanks to their atlases, we can often analyze them by
looking at their image in Rn (with some care). This is the case for their com-
pactness or orientability.
Definition 10.7. A manifold is compact if for any atlas {Uα }α=1,...,∞ there is a
finite sub-collection {Uα }α=1,...,n covering the whole manifold.
For instance, in R we can always have a single open subset covering the
whole manifold, R itself, but if we choose the atlas given by {Uj =] j, j + 2[} j∈Z
there is no finite sub-collection covering the whole manifold.
100 differentiable manifolds

An interesting concept, although with limited applications in our course, is


the orientability of a manifold. For a sphere it is obvious that one can dis-
tinguish an interior and an exterior, but this is not always the case for any
manifold. Mathematically, we can use the determinant of the Jacobian of the
coordinate change to decide whether a manifold is orientable or not.
Definition 10.8. A manifold is orientable if there is an atlas such that det ( ϕα ◦
ϕ− 1
β ) > 0 for any couple of charts (Uα , ϕα ), (Uβ , ϕ β ) with non-trivial intersec-
tion Uα ∩ Uβ 6= ∅.
It is interesting to know that compact orientable manifolds have been classi-
fied up to dimension 3.

10.1.3 Examples

We now give some examples to illustrate the concepts introduced so far. First
of all let us introduce some obvious examples of manifolds: Euclidean spaces.
Example 10.1. Rn is a manifold with an obvious atlas given by a unique chart
with homeomorphism the identity map.
On such manifolds it is easier to see why the concept of compatible charts is
important.
Example 10.2. Let M = R. We can construct charts covering the whole manifold
by using the maps ϕ1 : R → R and ϕ2 : R → R such that ϕ1 ( x ) = x and
ϕ2 ( x ) = x3 . These are not compatible, because ϕ1 ◦ ϕ2−1 = x1/3 , which is not
C ∞ at x = 0.
It is very easy to produce examples of manifolds that are not simply Rn .
For instance, the n-sphere Sn , defined as the locus of all points at some fixed
distance from the origin of Rn+1 , is a manifold. The n-torus T n , resulting from
taking an n-dimensional cube and identifying points on opposite sides, is a
manifold. And we can easily continue.
It is therefore also interesting to understand what is not a manifold and for
this reason we now give a few simple examples.

Example 10.3. , , are not manifolds. In the intersections


they are not locally R. If we remove the points at the intersections of the first
two examples in the picture we get 3 or 4 disconnected components, while in
R we always get 2 disconnected components if we remove a point. A similar
argument applies to the third picture.
10.1 introduction 101

When introducing the notion of manifold, we insisted on charts, atlases and


their compatibility. The reason is that for a generic manifold one cannot cover
the whole space by a single chart and we therefore need the discussion pre-
sented above. This is already clear for one of the simplest manifolds one can
think of: the circle.
Example 10.4. The circle S1 ≡ {( x, y) ∈ R2 | x2 + y2 = 1} is a compact 1-
dimensional manifold and therefore we can locally construct homeomorphisms
from open subsets of S1 to open sets of R. It is easy to see that the minimal atlas
is composed of two charts. The first one associates points of a maximal open
subset of the circle to points in the open set ]0, 2π [⊂ R via the homeomorphism
ϕ1 : S1 \ {(1, 0)} →]0, 2π [⊂ R, whose inverse gives the embedding of the circle
in R2 :

ϕ1−1 : θ 7→ (cos θ, sin θ ). (10.1)

The second one does the same, but in a way that covers the point we excluded
in the previous map. This means that now we map the open set U2 = S1 \
{(−1, 0)} to ] − π, π [⊂ R, so that also the homeomorphism ϕ2 : U2 →] − π, π [
has an inverse map that has the same form as before:

ϕ2−1 : θ 7→ (cos θ, sin θ ). (10.2)

The two maps are represented in figure 16.

y y
ϕ1
2π π

θ x x
(-1,0)
(1,0) θ

0 ϕ2 −π

Figure 16.: Atlas for S1 containing two charts.

With this procedure we are assigning two different coordinates sets to two
different patches of the manifold. We should then require that the two corre-
sponding charts are compatible in the overlapping regions. The points that are
not covered by one of the two charts are at y = 0 in the R2 embedding. We
102 differentiable manifolds

then have two regions of overlap: one given by all the points with y > 0 and
one with all the points with y < 0.
When y > 0 the coordinate changes are trivial because ϕ1 ◦ ϕ2−1 (θ ) = θ and
ϕ2 ◦ ϕ1−1 (θ ) = θ. This effectively means that we do not really need to change
the coordinate assigned to the points of S1 in this region when we pass from
one chart to the other. However, when y < 0 we have ϕ1 ◦ ϕ2−1 (θ ) = 2π + θ and
ϕ2 ◦ ϕ1−1 (θ ) = θ − 2π, which means that we have assigned different coordinates
to the same points and we need to perform a coordinate change when we go
from one chart to the other. It is straightforward to see that both change of
coordinates are C ∞ .
Example 10.5. The sphere Sn = {~r ∈ Rn+1 | |~r | = 1} also needs two charts,
constructed generalizing the previous example. They come from the so-called
stereographic projection. Using once again the embedding of the sphere in a
higher-dimensional space, we can define the two charts as follows. From a pole,
calling ~x ∈ Rn the local coordinates we can use

1 − |~x |2
 
2~x
ϕ1−1 (~x ) = , ,
1 + |~x |2 1 + |~x |2

which covers the whole Sn except the point {0, 0, . . . , 1}, and

−1 + |~x |2
 
2~x
ϕ2−1 (~x ) = , ,
1 + |~x |2 1 + |~x |2

which covers the whole Sn except the point {0, 0, . . . , −1}.

10.2 tangent and cotangent space

Useful structures on manifolds are vector fields and tangent spaces and we
now discuss how to define and construct them.
We are already familiar with the concept of tangent vectors to curves in Eu-
clidean spaces. We know that they can be constructed by evaluating derivatives
of the functions defining such curves. When we deal with a generic topological
space M that is also a differentiable manifold, we can construct the tangent
space to one of the points of M by generalizing this approach, describing the
vectors tangent to the curves passing through this point via its image in Rn .
A curve on M is a function

γ : R ⊃ I → M. (10.3)
10.2 tangent and cotangent space 103

xi ( p)

p γ(t) xi

Figure 17.: We can define tangent vectors to a manifold by looking at tangent vectors
to curves defined on the same manifold using their image in Rn .

When its image has a non-trivial intersection with the open subset of M that
defines a chart x : M ⊃ U → Rn , we can associate to the curve an image in Rn
with coordinates xi (γ(t)), i = 1, . . . , n (see fig. 17 for a representation of this).
The vector tangent to the curve γ at the point p = γ(t∗ ) ∈ M can be defined
via the tangent vector to its image:

d i
≡ V i ( p ).

x (γ(t)) (10.4)
dt t=t∗

The functions V i are maps from points of M to Rn , associating an element of


the vector space Rn to each of the points of γ( I ) ⊂ M. Unfortunately, this
definition of tangent vector heavily depends on the homeomorphism defining
the chart. For two overlapping charts we may have different values for the
functions V i ( p) defining the vector tangent to the same point of the same curve.
In order to have an intrinsic definition of tangent vectors, independent of the
coordinates, it is useful to introduce real functions defined on a manifold:

f : M → R. (10.5)

Once again, using coordinates defined by an atlas of M, we can write their


value in each chart as f ( ϕ− 1 i i
α ( x )), or, with an abuse of notation, as f ( x ). The
variation of the function f along the curve γ is (in terms of local coordinates)
d ∂ d
f (γ(t)) = i f ( ϕ−1 ( x )) xi (γ(t)), (10.6)
dt ∂x dt
where we evaluated f as a function of Rn via f ◦ ϕ−1 : Rn → R. This allows us
to define the differential operator on the set of real functions on a manifold M

V ≡ Vi , (10.7)
∂xi
104 differentiable manifolds

where V i = ddt xi (γ(t)), so that the directional derivative of the test function f
along the curve γ can be obtained by application of the differential operator V:

d
V[ f ] ≡ f (γ(t)). (10.8)
dt
Using these differential operators we can eventually give an intrinsic definition
of vectors tangent to a curve.

Definition 10.9. The differential operator V ≡ V i ∂x



i computed in p ∈ M is the
vector tangent to M at p in the direction of the curve γ.

Now V is defined in an intrinsic way. In fact, if the same point is covered by


e xe), i.e. p ∈ U ∩ U,
two different charts (U, x ) and (U, e we have

xj
e i ( p) ∂e xi e j
∂e
V ( p ) = V i ( p ) ∂i = V
e i ( p) e
∂i = V ∂j ⇒ V i ( p) = V ( p). (10.9)
∂xi ∂x j
We can also generalize this concept in order to obtain a differential operator
defined on the whole manifold and not just on a curve. This is called a vector
field.

Definition 10.10. A vector field V on M is a linear operator from C ∞ ( M) to


C ∞ ( M), which, in each chart, is defined via the directional derivative V =
V i ( x ) ∂i .

V (x)

Figure 18.: Integral curves following from the definition of the vector field V on the
manifold M.

We can imagine a vector field as a set of differential operators defined at


each point of the manifold, whose action is to take differentiable functions and
produce directional derivatives at the point where they are evaluated.
10.2 tangent and cotangent space 105

The action of these vector fields is linear on the functions


V [ f + g] = V [ f ] + V [ g] (10.10)
and there is no obstruction to adding them or scaling them by a real number
in order to obtain new vector fields:
( a V + b W )[ f ] = a V [ f ] + b W [ f ], for a, b ∈ R. (10.11)
This implies that directional derivatives form a vector space and we can use
this vector space to introduce the notion of tangent space to a manifold at a
point p in a coordinate-independent way.

Tp M

p
M

Figure 19.: The tangent space Tp M to a point p ∈ M can be defined by the equivalence
classes of directional derivatives along all the curves passing through p.

The tangent space to a point p ∈ M, whose symbol is Tp M, is the vector


space of linearly independent directional derivatives evaluated at that point.
Since we associate these derivatives to curves passing through p we can also
define T p M as the collection of curves generating different directional deriva-
tive operators.
Definition 10.11. The tangent space to the manifold M at the point p (in short Tp M)
is the vector space of dimension N = dim M, defined by the equivalence classes

{ γ : R → M | γ (0) = p ∈ M }
Tp M ≡ . (10.12)
{γ1 ∼ γ2 ⇔ ddt γ1 (0) = ddt γ2 (0)}
The collection of all the tangent vector spaces at each of the points of M is
then the tangent space of the manifold.
Definition 10.12. The Tangent Space to a manifold M is defined as
[
TM = Tp M. (10.13)
p∈ M
106 differentiable manifolds

Although we will not make much use of it in this course, starting from Tp M
we can define its dual space Tp∗ M, called the cotangent space.

Definition 10.13. The cotangent space to a manifold M at the point p ∈ M (in


short Tp∗ M) is defined as the dual space of Tp M via the duality operation
 
i ∂
dx , j = δji , (10.14)
∂x

so that the set of differentials dxi define a basis on this space.




10.3 flows and exponential map

A vector field on a manifold describes quite naturally a flow in M. These flows


are the curves that that gave origin to the vector field V and which can be
reconstructed by using the exponential map we are going to introduce in the
following.

Definition 10.14. Given a vector field V on a manifold M, the curve x (t, x0 )


solving the differential equation

d i
x (t, x0 ) = V i ( x (t, x0 )), and xi (0, x0 ) = x0i , (10.15)
dt
where x0 ∈ M and xi are the local coordinates, is called the integral curve or
flow generated by the vector field V, passing through x0 at the time t = 0.

For simplicity, we use x to indicate both points in M as well as their local


coordinates in Rn .

Theorem 10.1. Let V ∈ TM be a vector field on the manifold M. For any point
x0 ∈ M there is an integral curve of V, a flow x : R × M → M such that x (t, x0 ) is
a solution of the differential equation (10.15).

It is interesting to note that flows satisfy the properties required to be a


group. In fact, for fixed t, xt ( x0 ) ≡ x (t, x0 ) is a diffeomorphism xt : M → M
and represents a 1-parameter Abelian group:

1. the product closes (and is Abelian): xt ◦ xs = xs ◦ xt = xt+s (or, explicitly


xt ( xs ( x0 )) = xs ( xt ( x0 )) = xt+s ( x0 ));

2. there is an identity: x0 = 1;
10.3 flows and exponential map 107

3. each element has an inverse: xt−1 = x−t .


Also, flows can be expressed in terms of the exponential map of the vector
fields that generate them:
Definition 10.15. The exponential map of a vector field V (denoted by exp(tV ))
is defined as the family of diffeomorphisms mapping the point x0 ∈ M to
x (t, x0 ):
exp(tV )[ x0 ] = x (t, x0 ). (10.16)
We can verify this by the analysis of the expansion of the coordinate functions
defining the image of the flow around x0 :
∞  n
tn

d
xi (t, x0 ) = ∑ n
x i
( t, x 0 )
n =0
dt t=0 n!
 
d i

= exp t [ x (t, x0 )] (10.17)
dt t =0

= exp(tV )[ x0i ],
where in the last step we used the definition (10.15).
We give here only a simple example of the construction of the tangent space
to a manifold and the reconstruction of the curves on the manifold via the
exponential map.
Example 10.6. Take M = R. The tangent space Tx0 R can be constructed by
considering the diffeomorphisms γ( x0 , t) = x0 + t. In fact
d f (γ( x0 , t)) − f (γ( x0 , t))
f (γ( x0 , t)) = lim
dt t →0 t
(10.18)
f ( x0 + t ) − f ( x0 ) d
= lim = f ( x0 ).
t →0 t dx
This means that ddt γ( x0 , t) = 1 = V x and the vector field is V = dx
d
.
We can get back the integral curve that originated V by the exponential map
of V:
 
d
exp(tV )[ x0 ] = exp t [ x0 ]
dx
  (10.19)
d
= 1+t + . . . x = x0 + t = γ ( x0 , t ).
dx x = x0
108 differentiable manifolds

10.4 lie brackets

00
W ( p) V ( p )
p0 [V, W ]
p
V ( p) p0 W ( p0 )

Figure 20.: Geometrically the Lie bracket [V, W ] describes the non-commutativity of
the flows generated by the vector fields V and W.

The last concept we need to introduce is the Lie bracket. This is an operation
that associates a new vector field to the combination of two vector fields. It
expresses how a vector field changes along the flow generated by another vector
field. In detail, we have seen that there is a direct relation between a vector field
and its integral curves by means of the exponential map. If, starting from the
same point p ∈ M, we compose the result of integral curves coming from
different vector fields, the result will depend on the order. At the linearized
level, this should be represented once again by a vector associated to another
vector field and this is represented by the Lie brackets.
Definition 10.16. The Lie brackets are the commutator of two vector fields on
M:
[V, W ] = V i ∂i W j ∂ j − W i ∂i V j ∂ j =
 

= V i ∂i W j − W i ∂i V j ∂ j

(10.20)
= Z j ∂ j ≡ Z.
We do not discuss further this new ingredient at this stage, but we will come
back to it when we will discuss the relation between a Lie group and its tangent
space, defined by a Lie algebra.

exercises

1. The maps φi : R → R defined as φ1 ( x ) = x and φ2 ( x ) = x3 are homeo-


morphisms from R to R.
10.4 lie brackets 109

Are they diffeomorphisms?


If we consider φi as homeomorphisms in two different charts covering
Mi = R (the manifolds covered by such charts), are M1 and M2 diffeo-
morphic?

2. On the real projective space RPn we can introduce homogeneous coordi-


nates
[ x ] = [ x0 : . . . : xn ] ∈ RPn .
Define also the open sets Ui = {[ x0 : . . . : xn ]| xi 6= 0} and the charts

φi−1 : Rn → RPn
( x0 , . . . , x̂i , . . . , xn ) 7→ [ x0 : . . . : |{z}
1 : . . . : x n ],
i

where x̂i signals the fact that we remove the coordinate xi from the set.
Compute the transition functions. Recall the stereographic projection of
the spheres and build an homeomorphism between RP1 and S1 . (One
can proceed in a similar way for CP1 and S2 or for HP1 , the projective
space on quaternions, and S4 ).

3. Show that if a smooth differentiable manifold M ⊂ Rn of dimension k is


defined by the zeros of f : U ⊂ M → Rn−k (or U ∩ M = f −1 (0)) we have
that
Tp M = ker [d f ( p)].
Use this fact to find the tangent space to the unit sphere of dimension k.
Part IV

LIE GROUPS
11
LIE GROUPS.

In the first part of these lectures we mainly discussed discrete and finite groups.
We now want to consider the theory of continuous groups, where the group
elements depend continuously on a set of parameters.
The theory of continuous groups is much more complicated than the the-
ory of discrete groups. In this case the elements cannot be enumerated and
in addition to algebraic concepts we need topological notions like continuity,
connectedness and compactness and this is why we reviewed in the previous
lectures some of these notions. However, we will not discuss all possible types
of continuous group, but rather focus on a particular class of continuous groups
relevant for physical applications: Lie groups.
Definition 11.1. A Lie group is a smooth differentiable manifold with a group
structure such that the following group operations are differentiable (smooth):
• the composition: ◦ : G × G → G;
• the inversion map: G → G given by g 7→ g−1 .
The dimension of a Lie group G is determined by the dimension of G as a
manifold.
A Lie group is a special kind of differentiable manifold, where each point
is an element of a group and where different points may be related by group
operations.
Remember that our goal is not to give a full mathematical treatment of Lie
groups and their associated Lie algebras, but to show the results that are most
relevant for physical applications. For this reason we will often mention theo-
rems without demonstrations, just to inform the student of their existence, and
provide examples that are instructive to understand more general mathematical
results.

113
114 lie groups

11.1 compact and non-compact groups

Before discussing examples of Lie groups, we present some general statements


about their general properties that are going to be useful in the following.
Lie groups are also differentiable manifolds and this implies that we can
associate topological properties to Lie groups by means of the same properties
for their underlying manifold. For instance, we can separate compact and
non-compact groups using the corresponding definition 10.7 for differentiable
manifolds.
Since a generic element of a Lie group is determined by the values of dim
G parameters defined via local charts, we call compact groups the ones having
a compact parameter domain and non-compact those having a non-compact
parameter domain. For instance, the rotation group in 3 dimensions, SO(3), is
compact and its elements are defined via angles. The group of translations on
the other hand is non-compact.

11.2 invariant measure

Most of the theorems we proved in the first part of this course relied on carrying
out sums over the group elements, often in conjuction with the Rearrangement
Theorem 3.2. For instance, thanks to the rearrangement theorem, we often
replaced the sum over the group elements of a function of the elements of
the group, with the sum of the values of the same function evaluated on the
elements multiplied by the same g0 from the left

∑ f ( g0 g) = ∑ f ( h = g0 g) = ∑ f ( g ), (11.1)
g∈ G h∈ G g∈ G

or from the right

∑ f ( gg0 ) = ∑ f ( g ), (11.2)
g∈ G g∈ G

or summing over the inverses

∑ f ( g −1 ) = ∑ f ( g ). (11.3)
g∈ G g∈ G

When we deal with continuous groups, discrete sums have to be replaced


with continuous ones
Z
dµ g f ( g) (11.4)
G
11.3 groups of matrices 115

and in order to do so, we need to introduce a measure on the group dµ g ,


possibly in a way that the integral gives finite results for continuous functions
f . Obviously, different choices of measures may lead to different results, even
if we sum over the same elements. For this reason we would like to have
a measure that remains invariant when the group elements in the sum are
rearranged. This means that we need an invariant measure over G, namely a
measure dµ g satisfying

dµ gg0 = dµ g0 g = dµ g−1 = dµ g . (11.5)

Unfortunately such a measure does not always exist:

Theorem 11.1. If G is a compact group there is a measure satisfying dµ gg0 = dµ g0 g =


dµ g−1 = dµ g and it is unique up to a normalization factor.
This measure is called the Haar measure.

This theorem implies that only for compact groups we can safely extend our
proofs of orthogonality for the matrix elements of their representations, for the
equivalence of the group representations to unitary ones, etc. One needs more
care when discussing non-compact groups. For instance, one may still have
left-invariant or right-invariant measures, but their integral on the group may
diverge.

11.3 groups of matrices

The groups of invertible matrices play a special role among Lie groups for two
reasons:

• Invertible matrices provide linear representations of groups;

• Groups of matrices can be systematically classified using constraints.

We start by repeating a definition that we already encountered, but which


provides the starting point of our discussion.

Definition 11.2. The set of N × N invertible matrices with entries in F = R or


C forms a group called General Linear group, or GL(N, F).

If A, B ∈ GL(n, C) it is obvious that AB ∈ GL(n, C) and also ( AB)−1 =


B −1 A −1∈ GL(n, C), where the composition is given by the usual matrix prod-
uct. Also, we have an obvious identity element, which is the n × n unit matrix
116 lie groups

1n . The only constraint needed to have a group is that the matrices in GL(n, C)
should have non-vanishing determinant, so that we can invert them. This does
not constrain single entries of the matrix and therefore

dim GL(n, R) = n2 (11.6)

and

dim GL(n, C) = 2n2 . (11.7)

Recall that we discussed real manifolds and therefore we are giving here their
real dimension.
If we restrict ourselves to the group of matrices with unit determinant we
can define
Definition 11.3. The Special Linear group

SL(n, F) = { M ∈ GL(n, F) | detM = 1}.

The proof that this is still a group is straightforward because det ( AB) =
det A · det B, det 1n = 1 and det ( A−1 ) = 1/det A.
Since one of the entries of the matrices in SL(n, F) is fixed in terms of the
others because of the constraint on the determinant we have that

dim SL(n, R) = n2 − 1 (11.8)

and

dim SL(n, C) = 2(n2 − 1). (11.9)

Starting from these groups we will now see a series of other groups defined
as subsets satisfying linear or quadratic constraint. The matrices resulting from
these constraints provide the so-called fundamental representation of a group.
As we will see in the following, this is not the only irreducible representation
existing for such groups, but it is fundamental in the sense that the others could
be obtained by decomposing its products (in general a Lie group has an infinite
number of irreducible representations).

11.3.1 Linear constraints

Definition 11.4. The upper triangular matrices form a group

UT( p, q) = { M ∈ GL( p + q, F) | Mia = 0, ∀ a ≤ p and i ≥ p.}


11.3 groups of matrices 117

The form of a generic matrix in UT(p, q) is


!
A p× p B p×q
m= . (11.10)
Oq× p Cq×q

Composition of matrices of this form close on matrices of the same form. The
identity is part of UT( p, q) and the inverse is also in the group

A− 1
− A− 1 −1 !
−1 p× p p× p B p×q Cq×q
m = . (11.11)
Oq × p Cq−×1q

This definition can be generalized straightforwardly to UT(p, q, r) = UT( p, q +


r ) ∩ UT( p + q, r ) and this can be repeated, up to UT(1, . . . , 1). In this case the
group gets a new name:

Definition 11.5. The group of n × n strictly triangular matrices is called Solvable


group:
Solv(n) = UT(1, 1, . . . , 1).

Definition 11.6. The subgroup of matrices of Solv(n) having all diagonal ele-
ments equal to 1 is called Nilpotent group, Nil(n).

11.3.2 Quadratic constraints

We can also restrict invertible matrices by requiring that they preserve another
matrix (a metric), that is not necessarily part of the group. We then introduce a
quadratic constraint, where matrices M ∈ GL(n, F) satisfy also a constraint of
the form M T η M = η, or M† η M = η, or M∗ η M = η, for η a matrix that may be
singular or non-singular, antisymmetric or symmetric and, if symmetric, posi-
tive definite or indefinite. All these distinct cases will generate distinct groups,
provided that the constraint allows for group composition to be closed and for
the inverse to satisfy the same constraint (M = 1 always satisfies quadratic
constraints like those mentioned above).

Compact groups
Definition 11.7. The Unitary group

U(n) ≡ { M ∈ GL(n, C) | M† 1n M = 1n }.
118 lie groups

This group has n2 real constraints because ( M† M )† = M† M and hence


dim U(n) = n2 . (11.12)
Definition 11.8. The Special Unitary group
SU(n) ≡ { M ∈ U (n) | detM = 1}.
Obviously SU(n) < U(n), but, contrary to naive expectations, U(n) is not
simply the product of SU(n) with a phase. In fact the map
SU(n) × U(1) −→ U(n)
(11.13)
( M, eiα ) 7→ e = eiα M
M

is a homomorphism, but not an isomorphism. If det( M ) = 1, det( M e ) = eiαn ∈


C, but obviously this map is not 1 to 1. If we choose M = e2πi/n 1n , det( M ) = 1,
but we have n such matrices that we can map to the same element in U(n):
 
gm = e2πi m/n 1, e−2πi m/n 7→ 1n , (11.14)

where m = 1, . . . , n. Since the elements gm generate a Zn < SU(n)×U(1), we


have that
U(n) ' (SU(n) × U(1))/Zn . (11.15)

If we restrict ourselves to real matrices, the same quadratic constraint gives


rise to another notable group.
Definition 11.9. The Orthogonal group
O(n) ≡ { M ∈ GL(n, R) | M T M = 1n }.
We now have a real symmetric constraint ( M T M ) T = M T M. We thus have
n(n + 1)/2 constraints on n2 real entries leaving
n ( n − 1)
dim O(n) = . (11.16)
2
We already saw that the constraint defining O(n) implies that its matrices have
determinant equal to 1 or to −1. Those with unit determinant include the
identity and form a subgroup:
Definition 11.10. The Special Orthogonal group
SO(n) ≡ { M ∈ GL(n, R) | M T M = 1n , detM = 1}.
11.3 groups of matrices 119

Non-compact groups
If we consider the same type of constraints as those given above, but now for
an indefinite metric of the form

1p
!
η= (11.17)
−1q

we get some non-compact groups:

Definition 11.11.

U( p, q) ≡ { M ∈ GL( p + q, C) | M† η M = η }.

Definition 11.12.

O( p, q) ≡ { M ∈ GL( p + q, R) | M T η M = η }.

It is very easy to see that the set of parameters labeling the elements of these
groups becomes non-compact. For instance,
!
cosh ψ sinh ψ
= m ∈ O(1, 1), ∀ψ ∈ R. (11.18)
sinh ψ cosh ψ

Another non-compact group important for physical applications is the

Definition 11.13. Symplectic group

Sp(2n, F) = { M ∈ GL(2n, F) | M T ΩM = Ω},

On 1n
!
where Ω = .
−1n On

Other groups may be obtained using singular metrics, like

1
 

η= −1q ,
Or

generating U( p, q, r ), O(p, q, r ), etc.


120 lie groups

We will not discuss all possibilities here. We close this lecture mentioning
the fact that some of these groups with different definitions are actually iso-
morphic.
 For  instance, a matrix M ∈ Sp(2,R) has unit determinant. In detail
a b
M= ∈ Sp(2, R) if and only if
c d
     
a c 0 1 a b 0 1
= (11.19)
b d −1 0 c d −1 0
and this gives ad − bc = 1. But this is the definition of a matrix in the special
linear group of 2-dimensional real matrices and therefore

Sp(2, R) ' SL(2, R). (11.20)

exercises

1. Take the fundamental representation of SO(2) and show that Zn < SO(2).
The matrices you obtained provide a representation of Zn . Is it reducible?
2. Show that Dn < O(2) by using appropriate matrices in the fundamental
representation of O(2).
3. Consider the matrices in the fundamental representation of SO(3) and
show that S4 < SO(3).
Hint: look for a transformation corresponding to a 3-cycle and one corresponding
to a 4-cycle.
4. Let (qi , pi ), i = 1, . . . , n, be the coordinates and momenta of a classical me-
chanical system with n degrees of freedom. Poisson brackets are defined
as
n  
∂ f ∂g ∂g ∂ f
{ f , g} = ∑ i ∂p
− .
i =1 ∂q i ∂qi ∂pi
a) Show that {qi , q j } = { pi , p j } = 0 e {qi , p j } = δji .
b) Show that if we redefine the coordinates with a linear transformation
   
Q q
=S ,
P p

the relations { Qi , Q j } = { Pi , Pj } = 0 e { Qi , Pj } = δji are still true if


and only if S ∈ Sp(2n, R).
11.3 groups of matrices 121

5. Write the matrices R1 , R2 ∈ SO(3) corresponding to a rotation of π/2


around the z axis and of π/4 around the x axis respectively. Compute
R3 = R2 R1 and show that the corresponding element gives a rotation of
an angle φ around an axis determined by ~n = (n x , ny , nz ). Show that tr
( R3 ) = 1 + 2 cos φ.

6. Using the results of the previous exercise show that conjugate elements
of SO(3) are characterized by the angle specifying the rotation.

7. This is just for fun. Use what you just learned to find the conjugacy classes
of S4 .

8. Find the center of O(n) and of SO(n).

9. Find the maximal Abelian subgroup of U(2).

10. Given M ∈ Sp(2n, R), show that M−1 ∈ Sp(2n, R) and also M T ∈
Sp(2n, R), i.e. prove that if M T ΩM = Ω, then M−T ΩM−1 = Ω as well as
MΩM T = Ω.

11. Discuss the isomorphism between Sp(2, R) and SU(1,1).


 
1 i
Hint: Use the map µ( g) = τgτ −1 , where τ = √1 and check that it
2 i 1
defines an isomorphism.

12. Consider again the Lagrangian density

1
L = −∂µ Φ† ∂µ Φ + Φ† ηΦ − (Φ† Φ)2 ,
2
 
φ1 ( x )
 φ2 ( x ) 
where Φ is a complex scalar field with 4 components Φ( x ) =  
 φ3 ( x ) 
φ4 ( x )
e η = diag{−1, −1, 1, 1}.
a) What is the group of linear transformations Φ( x ) → SΦ( x ), with
S ∈ Mat(4, C), preserving the kinetic term and the potential, respec-
tively?
122 lie groups

 
c1
 c2 
b) For which constant values Φ0 =   c3  the Euler–Lagrange equa-

c4
tions are satisfied and what is the residual symmetry group?
(Otherwise, which matrices S satisfy Φ0 → SΦ0 = Φ0 )
12
LIE ALGEBRAS.

As already done for finite groups, also for Lie groups we would like to have a
rapid way to understand their structure and their possible relations. Given their
double nature of groups and manifolds, in order to specify an isomorphism we
need to have their equivalence as topological spaces as well as groups. We
then need to understand when we can deform two groups onto each other in a
continuous way, so that the composition laws are equivalent up to a change of
variables. This last requirement may be very complicated because composition
laws are not necessarily linear. Also, we would like to compare only a finite
number of elements if possible, rather than the infinite number that is defining
a continuous group. For these reasons we now introduce the notion of Lie
algebra, which is going to help us in this tasks.
A Lie algebra is the linearization of the corresponding group around the
identity element. It contains the information that is needed to reconstruct the
group in a neighborhood of the neutral element. Actually, the first useful thing
to note is that linearizing the group G at any of its elements gives the same
result. In fact, since G is a manifold endowed with a group structure, we can
map any element g to any other g0 using the composition of the first with
h = g0 g−1 and in the same way we can map any open neighborhood of g, Ug ,
to an open neighborhood of g0 , Ug0 , by multiplying points of Ug with h from
the left. Since the composition map on a Lie group is a continuous function,
we can do the same for the curves in Ug and eventually for the directional
derivatives defining the tangent space in g. The result is that we can focus on
the tangent space to the group G at the identity without losing in generality,
but simplifying some of the calculations because of the properties of the neutral
element.

123
124 lie algebras

Tg G g 0 ◦ g −1

g g0
Tg0 G Ug0
Ug

G
Figure 21.: Tangent spaces at different points on a group manifold can be related by
group multiplication.

12.1 algebra as tangent space

The general procedure to construct the tangent space requires the identification
of curves M (t) ∈ G, passing through the element around which we want to
construct the tangent space, which is the identity. Since we do not want to enter
in a mathematical discussion that would be useless for our purpose, we will
focus on groups of matrices, though obvious generalizations apply to generic
Lie groups. We therefore choose M (t) as a 1-dimensional subgroup of the
group of matrices G, so that

M (0) = 1 (12.1)

and

M ( t1 ) M ( t2 ) = M ( t1 + t2 ), (12.2)

which implies

M (t)−1 = M (−t). (12.3)

Locally this group is isomorphic to R. This means that we can define in an


obvious way its derivatives, following the construction explained in chapter 10.
For each of the inequivalent curves M(t) we define the corresponding tangent
vector to the identity as
d
g ≡ T1 G 3 X = M (0). (12.4)
dt
12.1 algebra as tangent space 125

The subgroup defined by M (t) can actually be seen as an integral curve of a


vector field V as follows from
d M (t + δt) − M (t) M (δt) − M(0)
M(t) = lim = M(t) lim
dt δt→0 δt δt→0 δt
d
= M(T ) M(0) = M (t) X, (12.5)
dt
and hence
d
M(t) = V (t)[ M (0)] = M(t) X, V (0) = X. (12.6)
dt
From the general procedure of section 10.3 we also know that we can recon-
struct the 1-parameter group M (t) < G via the exponential map built from
X ∈ g:

M(t) = exp(t X )[ M(0) = 1]. (12.7)

We then provided explicit maps between a group and its algebra in both direc-
tions:
d
dt
−→
G g
←−
exp
When G = GL(n, C), both maps are easy to compute. In particular, for M ∈
GL(n, C), we can easily obtain the tangent space at the identity by considering
the expansion

M = 1+em+..., e  1, (12.8)

where the condition det M 6= 0 is trivially satisfied for e  1 for any

m ∈ Mat(n, C), (12.9)

the space of square complex matrices. The inverse operation gives back the
original group close to the identity via the exponential map. By exponentiation
we can indeed obtain n2 different 1-parameter subgroups of GL(n, C):

tk mk
M(t) = exp(tm) = ∑ k!
, (12.10)
k =0
126 lie algebras

where mk is defined by the usual matrix product. This series is always conver-
gent in the space of matrices and we can easily check that M(t) is in fact in
GL(n, C). Its inverse is simply M(t)−1 = M(−t) = exp(−tm). We therefore
conclude that
Definition 12.1. The Lie algebra of the general linear group is

gl(n, F) = Mat(n, F).

We are obviously interested in understanding if and how we can reconstruct


the group G and all its properties from its algebra g = T1 G, but first we pause
and analyze some other specific instances of groups of matrices.

12.1.1 Algebras for groups of matrices

Also for the Special Linear group we can obtain its algebra by expanding the
constraint defining its fundamental representation around the identity:

1 = det M = det(1 + e A + . . .) = 1 + e trA + . . . . (12.11)

Definition 12.2.
sl(n, F) = { A ∈ Mat(n, F) | trA = 0}.
In the case of Unitary groups an analogous expansion gives

(1 + e A + . . . ) † (1 + e A + . . . ) = 1 + e ( A + A † ) + . . . , (12.12)

for e  1 and real (any phase can be reabsorbed in the definition of A).
Definition 12.3.
u(n) = { A ∈ Mat(n, C) | A† = − A}.
Obviously dim U(n) = dim u(n) = n2 . This is easy to check for the algebra,
where the n diagonal entries are purely imaginary and the n(n − 1)/2 entries
above the diagonal are complex conjugate of those below the diagonal. We then
n ( n −1)
have only 2 × 2 + n = n2 real degrees of freedom.
Since we obtained the conditions that matrices in the algebra should satisfy
by looking at the linear expansion around the identity, we may ask whether
their exponential, which goes beyond linear level, is really in the group. This
is straightforward to verify because of matrix properties:

[exp(tA)]† exp(tA) = exp(tA† ) exp(tA) = exp(−tA) exp(tA) = 1, (12.13)


12.1 algebra as tangent space 127

where we used the definition of the exponential of a matrix in the first equality
and the fact that A ∈ u(n) in the second equality.
Note also that in Quantum Mechanics the relation between elements of the
Unitary group and the algebra is usually done by considering the exponen-
tial of Hermitian operators, which are related to our group generators by an
imaginary unit.
For the orthogonal group and the special orthogonal group we get that

(1 + e A + . . . ) T (1 + e A + . . . ) = 1 + e ( A + A T ) + . . . , (12.14)

which means that matrices in their algebras are antisymmetric and therefore
with n(n − 1)/2 degrees of freedom. In addition, for the special orthogonal
group we should add the constraint giving unit determinant matrices, but this
is given by the vanishing of the trace, which is guaranteed by their antisymme-
try.

Definition 12.4.

o(n) ' so(n) = { A ∈ Mat(n, R) | A T = − A}.

Note that, as expected, the algebras described above have the structure of a
vector space, but differently from the corresponding groups, they do not close
under matrix multiplication. For instance, if A, B ∈ so(n), AB ∈
/ so(n), because

( AB)T = B T A T = (− B)(− A) = BA 6= − AB. (12.15)

Using the same trick we can also establish that:

Definition 12.5.
n o
sp(2n, R) = A ∈ Mat(2n, R) | A T Ω = −ΩA ,

1
 
0
where Ω = .
−1 0

Since Ω T = −Ω, we can also rewrite the constraint as (ΩA) T = ΩA, and
this means that the number of parameters defining the matrices A is equal to
that of 2n × 2n symmetric matrices. Hence

dimR sp(2n, R) = n(2n + 1). (12.16)


128 lie algebras

12.2 from the algebra g to the group G

Given g ≡ T1 G, we have a description of G in a neighborhood of the identity


as follows from:

Theorem 12.1. The exponential map is bijective in a neighborhood of the identity.

However, using g 3 X 7→ exp X ∈ G can we reconstruct all the information


needed to specify G? In particular, we are interested in the following questions:

• Is the exponential map surjective?

• Is the exponential map injective?

• If we have isomorphic algebras (in a sense that we will specify soon), are
the corresponding groups isomorphic?

We will see that unfortunately the answer to all these questions is NO. However,
the exponential map becomes surjective if the group is connected and compact and
it becomes injective if G is simply connected. Hence we conclude that using the
exponential map we can fully reconstruct a group that is simply connected to
the identity and compact.

12.2.1 Surjectivity

The counterexample has been found by Cartan analyzing the relation between
SL(2, R) and sl(2, R). If A is a matrix in sl(2, R), it has vanishing trace and can
be parameterized by three real parameters a, b, c:
 
a b+c
A= . (12.17)
b − c −a

The generic matrix in SL(2, R) that can be obtained from an element of the
algebra by exponentiation is going to be exp( A), for an arbitrary choice of
a, b, c. We will now show that such elements, however, do not cover the whole
group. The reason is that they satisfy a constraint on the trace that some of the
elements in the group do not fulfill. In fact, any matrix A of the form (12.17)
can be diagonalized
 
−1 λ
SAS = , (12.18)
−λ
12.2 from the algebra g to the group G 129


by means of the similarity matrix S. Here ±λ = ± a2 + b2 − c2 are the eigen-
values of the matrix A. They are real if a2 + b2 ≥ c2 and purely imaginary
otherwise. Using once more the properties of the exponential map for matri-
ces, we see that

tr(exp( A)) = tr(S exp( A)S−1 ) = tr(exp(SAS−1 )) = eλ + e−λ . (12.19)

This implies that

tr[exp( A)] ≥ −2. (12.20)

In fact, if λ is real tr(exp A) = 2 cosh λ ≥ 2 and if λ is complex, λ = iµ and


tr(exp A) = 2 cos µ ∈ [−2, 2]. On the other hand we can easily construct a
matrix in SL(2, R) that does not fulfill this constraint. This is
 
−x
X= , for x ∈ R∗+ . (12.21)
− 1x

We have that det X = 1 and therefore X ∈ SL(2, R), but also


 
1
trX = − x + ≤ −2, (12.22)
x

which does not fulfill the constraint (12.20) for x 6= 1.


Note that, although the exponential map is clearly not surjective, we can still
obtain the element X by multiplying elements obtained from the exponential
of the algebra:
     
0 1 1 0
X = exp π exp log x . (12.23)
−1 0 0 −1

This, on the other hand, is a consequence of the fact that SL(2, R) is connected
and cannot be extended to other groups (for instance in O(n) we have two
disconnected components and we will never get matrices with determinant −1
by exponentiating matrices in the algebra, nor by taking their products).

12.2.2 Injectivity and global properties

We already saw that so(n) = o(n), but we can also construct identical algebras
that generate both compact and non-compact groups. This is the case of o(2) '
u(1) ' R. All these algebras have a unique element, which we can choose to
130 lie algebras

 
0 1
be X = for o(2), X = i for u(1) and X = 1 for R. The group
−1 0
elements are obtained by taking exp(θX ). For U(1) and SO(2) we get that for
any θ = c + 2π Z we get the same element of the group, while for (R+ , ·) the
map is injective.

12.2.3 The algebra structure and the Baker–Campbell–Hausdorff formula

As long as we consider g = T1 G just as a vector space we only recover the


notion of G as a differentiable manifold, but not its group structure. However,
once we introduce an additional operation (the commutator), we can relate
properties of g with those of G.
In order to see how this structure emerges we analyze the simple case of
SO(n). Starting from A, B ∈ so(n) we know that exp(tA), exp(tB) ∈ SO(n),
for any t. We also know that the closure of the product operation within the
group ensures that also exp(tA) · exp(tB) ∈ SO(n). How can we characterize
this from the point of view of the algebra? Take the expansion of the product
for small t:
t2 t2
  
exp(tA) · exp(tB) = 1 + tA + A2 + . . . 1 + tB + B2 + . . .
2 2

t2 2
= 1 + t( A + B) + ( A + B2 + 2AB) + O(t3 )
2

t2 t2
= 1 + t( A + B) + ( A + B)2 + [ A, B] + O(t3 )
2 2
t2
 
= exp t( A + B) + [ A, B] + . . . , (12.24)
2
where ( A + B)2 = A2 + B2 + AB + BA and

[ A, B] ≡ AB − BA. (12.25)

We see that in order to determine the new element of the group from the prod-
uct of exponentials of elements of the algebra we need to know also their com-
mutator, no matter how close we are to the identity.
Note also that

[ A, B] T = ( AB)T − ( BA)T = B T A T − A T B T = BA − AB = −[ A, B] (12.26)


12.2 from the algebra g to the group G 131

and therefore [ A, B] ∈ so(n), so the commutator associates a new element of the


algebra to two other elements of the algebra. Hence we can use the commutator
as a product operation that closes in the algebra.
We conclude that close to the identity, the structure of a (matrix) group is
determined by the vector space given by T1 G with the additional information
of the commutators of the elements of the algebra.
We could continue order by order, but the result is just a particular case of a
general formula we are not going to prove in detail:

Theorem 12.2. Baker–Campbell–Hausdorff formula: e X eY = e Z , where


Z
Z= X+ dt ψ(exp( ad X ) exp(t adY ))Y,

u log u
where ψ(u) = u −1 = 1 + (u − 1) + 61 (u − 1)2 + . . . is regular in u = 1 and

ad X Y ≡ [ X, Y ]. (12.27)

In the following we will also need the

Theorem 12.3. Hadamard Lemma

e A Be− A = exp( ad A ) B.

This is rather easy to check explicitly for matrices:

A2 A2
   
1+A+ +... B 1− A+ +...
2 2
1 2 
= B + AB − BA + A B − 2ABA + BA2 + . . . (12.28)
2
= ( B + [ A, B] + [ A, [ A, B]] + . . .) = exp([ A, ·]) B.

12.2.4 Lie algebras

Definition 12.6. A Lie algebra is a vector space g on a field F with a map

[·, ·] : g × g → g

which is
132 lie algebras

1. bilinear: ∀ x, y, z ∈ g and a, b ∈ F

[ ax + by, z] = a[ x, z] + b[y, z],

[z, ax + by] = a[z, x ] + b[z, y];

2. antisymmetric
[ x, y] = −[y, x ], ∀ x, y ∈ g;

3. such that the map [ x, ·] : g → g satisfies the Leibnitz rule

[ x, [y, z]] = [[ x, y], z] + [y, [ x, z]], ∀ x, y, z ∈ g.

The last condition also goes under another name

Definition 12.7. Jacobi identity. Given any X, Y, Z ∈ g:

[ X, [Y, Z ]] + [Y, [ Z, X ]] + [ Z, [ X, Y ]] = 0.

The map we just introduced is simply the commutator for matrices, but in
the general case it is an operation that we can define starting from Lie brackets.
In full generality we can define g from G as its tangent space at the identity T1 G.
This is a vector space on which we can introduce an operation associating a new
vector of the tangent space [ X, Y ] to any other couple of vectors X, Y ∈ T1 G by
taking the generator of the vector field that is obtained from the Lie brackets of
the vector fields generated by the other two. Operatively, we introduce exp(tX )
and exp(tY ), compute the Lie bracket [exp(tX ), exp(tY )] and expand for t → 0.
The element we obtain is [ X, Y ].

exercises

1. Matrix properties and exponential of a matrix.


a) Assume that A and B are generic N × N complex matrices. Check

that tr [ A, B] = 0, eU AU = Ue A U † , det e A = etrA .
b) Compute eαM and eαN , where α ∈ C and
   
0 0 1 0 0 1
M= 0 0 0 , N =  0 0 0 .
1 0 0 −1 0 0
12.2 from the algebra g to the group G 133

2. The Heisenberg algebra, generating the Heisenberg group, can be repre-


sented by the following matrices:
     
0 1 0 0 0 0 0 0 1
tx =  0 0 0 , tp =  0 0 1 , th̄ =  0 0 0 .
0 0 0 0 0 0 0 0 0

Compute their commutators, compute their exponentials Ux , U p e Uh̄ and


verify the Baker–Campbell–Hausdorff formula for the product Ux U p .

3. Build the fundamental representation of SU(1,1) and derive su(1, 1) =


T1 SU(1,1) by taking derivatives of appropriate 1-parameter subgroups.
13
P R O P E RT I E S O F L I E A L G E B R A S .

Lie algebras are first of all vector spaces. As such, for any Lie algebra g we
can introduce a basis of n = dim g = dim G vectors {t I } I =1,...,n , so that any
element of g can be written as their linear combination. Lie algebras, however,
are also closed under the product defined by the map in 12.6 (which we will
call commutator in the following). Because of its vector space nature, all the
information contained in the commutator of two generic elements is fixed by
that of the basis vectors:
[t I , t J ] = f I J K tK . (13.1)
Definition 13.1. The coefficients f I J K in (13.1) are called structure constants.
Using the linearity properties of the commutator, we can always express the
result of the commutator of any couple of elements of the algebra X, Y ∈ g in
terms of the structure constants. In detail, we can expand the generic vectors
X and Y as X = x I t I , Y = y J t J , where x I , y J ∈ F, the field on which the algebra
is defined, and do the same for the vector Z, associated to the commutator:
[ X, Y ] = x I y J [t I , t J ] = x I y J f I J K tK = Z ∈ g. (13.2)
From the relation (13.1), defining the structure constants, we can also deduce
some of their properties. First of all, they must be antisymmetric in the first
two indices
fI J K = − fJI K. (13.3)
Then, expanding the Jacobi identity 12.7 in terms of the basis vectors, we can
also deduce that the structure constants must satisfy
1 L 
0 = f [ I J L f K] L M = f I J f KL M + f KI L f JL M + f JK L f IL M . (13.4)
3

135
136 lie algebras – properties

Example 13.1. The algebras u(1) ' R are generated by a unique element t1 and
all their structure constants are vanishing.
Example 13.2. The algebra of the Heisenberg group is defined by 3 generators,
with a unique non-trivial commutator:

[ t1 , t2 ] = t3 . (13.5)

The only non-vanishing structure constants are f 12 3 = − f 21 3 = 1.


Example 13.3. The algebra of angular momenta so(3) has 3 generators t I , I =
1, 2, 3, satisfying the commutator relations:

[t I , t J ] = e I JK tK . (13.6)

From this introduction, we see that a Lie algebra is specified by the dimen-
sion of the vector space and by the commutator relations. Hence a classification
of inequivalent structure constants allows to classify inequivalent algebras. If
we want to compute two algebras, though, we should also remember that we
can always choose different basis for the same algebra. This means that we can
always change the values of the structure constants by taking linear combina-
tions of the generators:

t̃ I = ∑ c I K tK . (13.7)
K

Since this is just a redefinition of the basis of vectors generating the vector space
g, it does not change the algebra. For this reason, if two algebras g and e g are
connected by a change of basis like (13.7), we say that they are isomorphic (we
will give a more appropriate definition in 14.2). One should be careful on the
choice of the coefficients c I K , though. The algebra g is defined a vector space
on a specific field F, which is usually taken to be C or R. We should then
only take linear combinations where the coefficients c I K are in the same field,
otherwise we are effectively changing the nature and dimension of the algebra.
Later we will see that sometimes establishing an isomorphism between dif-
ferent real Lie algebras using complex change of coordinates can be useful to
study their representations. For this reason we introduce another useful con-
cept: the complexification of a real Lie algebra.

Definition 13.2. If gR = span(t I ) is a real vector space, its complexification gC


is the real span of the basis {t I , i t I } ∈ gC .
Note that dimR gC = 2 dimR gR .
13.1 subalgebras, ideals 137

The process of complexification of an algebra is useful to connect algebras


that are not isomorphic. For instance,

so(3) = {span(t1 , t2 , t3 ) , [t1 , t2 ] = t3 , [t2 , t3 ] = t1 , [t3 , t1 ] = t2 } (13.8)

and

so(2, 1) = {span(t̃1 , t̃2 , t̃3 ) , [t̃1 , t̃2 ] = t̃3 , [t̃2 , t̃3 ] = t̃1 , [t̃3 , t̃1 ] = −t̃2 } (13.9)

are not isomorphic as real algebras, but we can map them onto each other if
we allow the use of complex coefficients:

t̃1 = it1 , t̃2 = t2 , t̃3 = it3 . (13.10)

We then say that so(3) and so(2, 1) are different real forms of the same complex
algebra so(3)C ' so(2, 1)C ' sl(2, C).
Definition 13.3. A real form gR of the complex algebra gC is obtained by con-
sidering a complex basis {t I } ∈ gC , I = 1, . . . , dimC gC and restricting the
vector space to the one generated by their real combinations.

13.1 subalgebras, ideals

Given an algebra g, we can always consider some subspace that is again a vector
space and we can ask whether the result is also an algebra.
Definition 13.4. A subspace h ⊂ g is a subalgebra of g if it is closed with respect
to the bracket operation in g, i.e.

∀ x, y ∈ h, [ x, y] ∈ h.

An invariant subalgebra is called ideal.


Definition 13.5. An ideal n ⊂ g is a subspace of an algebra that is stable under
multiplication, i.e.
[n, g] ∈ n.
It is clear that such a definition derives from the linearization of the condition
for a subgroup to be normal. In the same fashion, we can distinguish Abelian
and non-Abelian algebras by their commutators.
Definition 13.6. An algebra a is Abelian if and only if all its structure constants
are vanishing.
138 lie algebras – properties

13.2 simplicity, compactness and connectedness

We can now analyze how various group properties are translated into prop-
erties of their algebras. For instance, in 4.11 and 4.12 we introduced simple
and semi-simple groups. The corresponding definitions for the algebras are
the following:
Definition 13.7. A Lie algebra g is simple is it does not contain any ideal.
Definition 13.8. A Lie algebra g is semi-simple if it does not admit any Abelian
ideal.
Note that while it is obvious that
G simple ⇒ g simple, (13.11)

G semi-simple ⇒ g semi-simple, (13.12)


the inverse is not guaranteed! A simple counterexample is given by the su(2)
and so(3) algebras. As we will see in detail in section 15.2, although the alge-
bras are isomorphic, the corresponding groups have different properties: SO(3)
is simple, while SU(2) has an invariant Abelian subgroup1 : Z2 (which is also
not a Lie group).
By extension of meaning, we will also call compact the Lie algebra of a com-
pact Lie group.
Putting together these definitions one can come to an interesting result that
explains several properties of many algebras and algebra representations that
are useful in Physics.
Theorem 13.1. Every semi-simple complex Lie algebra admits a unique compact real
form.
Let us stress that since algebras come from groups by linearizing around the
identity, we cannot deduce global properties of G starting from g. In particular,
if G is not connected and K < G is the part connected to the identity
T1 G = g = k = T1 K. (13.13)
In addition, if a group is not simply connected and G
e is its universal covering,
we have that
g=e
g. (13.14)
1 Pay attention, because some texts define semi-simple groups as those that do not have invariant
Abelian continuous groups, and in this sense also SU(2) would be semi-simple.
13.3 cartan’s criteria 139

This is what happens for instance for u(1) = R, so(3) = su(2), so(1, 3) =
sl(2, C), where R = U ](1), SU(2) = SO^ (3), SL(2, C) = SO^(1, 3), but obviously
R 6= U(1), etc.
Summarizing, for any group G there is an algebra g that is obtained as its
tangent space at the identity g = T1 G, while for any algebra g there is a unique
simply connected group G that has g as its tangent space at the identity T1 G.
All the other connected groups K that have g as algebra are of the form

K/H = G, (13.15)

where H C K (maybe discrete).


For non-connected groups, G is the component connected to the identity.

13.3 cartan’s criteria

We can understand even more properties of algebras and the relations with
properties of the groups that generated them by introducing a new object: the
Cartan–Killing form.

Definition 13.9. The Cartan–Killing form is defined by

( X, Y ) = (Y, X ) = tr ( ad X adY (·)) ,

where X, Y ∈ g.

If we recall formula (12.27), we see that ad X is an operator acting on the


elements of the vector space spanned by the basis of generators of the algebra.
Effectively, we can define it as a matrix operator linearly mapping elements of
g into other elements of g:

ad X (t I ) = t J D ( ad X ) J I , (13.16)

where

D ( ad X ) J I ≡ x L f LI J . (13.17)

In this way ad X ( adY (·)) can be interpreted as a product of matrices, whose


trace is easy to compute. If we then expand X and Y on the basis of the algebra
X = x I t I , Y = y I t I , we can rewrite the definition 13.9 as

( X, Y ) = k I J x I y J , (13.18)
140 lie algebras – properties

where
k I J = k J I = f I M L f JL M . (13.19)
Note also that
k I J = ( t I , t J ). (13.20)
Using this form one can prove the following very useful theorem:
Theorem 13.2. Cartan’s criteria:
1. A Lie algebra is semi-simple if and only if det k 6= 0;
2. g semi-simple, real and compact if and only if k is negative definite.
Example 13.4. The algebra so(3) has non-vanishing structure constants f 12 3 =
f 31 2 = f 23 1 = 1 and therefore k11 = f 12 3 f 13 2 + f 13 2 f 12 3 , etc... so that
 
−2
k= −2 . (13.21)
−2
We see immediately that det k = −8 6= 0 and that it is negative definite. In fact
so(3) is semi-simple and compact.
Example 13.5. In heis(3) the only non-vanishing structure constants are f 12 3 =
− f 21 3 = 1. The Cartan–Killing matrix is then
 
0
k= 0 , (13.22)
0
which is obviously degenerate. The algebra is not semi-simple. It is straight-
forward to see that a = span(t3 ) is a 1-dimensional vector space that is also an
Abelian subalgebra of g:
[a, g] = 0 ⊂ a ⊂ g. (13.23)
Example 13.6. For so(2, 1) the structure constants are f 12 = f 13 = f 23 1 = 1,
3 2

which gives
 
2
k= −2 . (13.24)
2
This is non-degenerate, but it is not definite. The algebra is semi-simple, but
not compact.
13.4 casimir operator 141

13.4 casimir operator

When k is non-degenerate, we can construct its inverse k−1 , which we denote


with upper indices. In this case there is another interesting object that we can
introduce: the quadratic Casimir operator.

Definition 13.10. Quadratic Casimir operator

C2 ≡ k I J t I t J .

This operator2 plays an important role in the analysis of representations and


has the property that

Theorem 13.3.
[C2 , X ] = 0, ∀ X ∈ g.

Proof. The theorem follows if we prove that C2 commutes with all the basis
vectors t I .

[C2 , tk ] = k I J [t I t J , tK ] = k I J t I [t J , tK ] + k I J [t I , tK ]t J

= k I J f JK L t I t L + k J I f IK L t L t J (13.25)

= k I J f IK L t J t L + t L t J = k I J k LP f IKP (t J t L + t L t J ),


where we have introduced the completely antisymmetric tensor

f I JL ≡ f I J K k KL . (13.26)

This is completely antisymmetric, because f I JK = f [ I J ]K by definition, but also


f I JK = f JKI , as follows from

([t I , t J ], tK ) = tr(t I t J tK − t J t I tK ) = tr(t I t J tK − t I tK t J ) = (t I , [t J , tK ]), (13.27)

where we used the definition of commutator as well as the cyclic property of


the trace. The conclusion is that we are contracting a symmetric tensor with an
antisymmetric one and hence the result is zero.
2 Formally C2 is a combination that does not live in the algebra g, but in the so-called universal
enveloping algebra, defined as the associative algebra of polynomials in the elements in g. Since
we will be interested mainly in matrix representations, also C2 will be a matrix obtained from the
product of the matrices representing the generators.
142 lie algebras – properties

exercises

1. Verify that (13.4) follows from the Jacobi Identity for three vectors defin-
ing a basis of the algebra.
2. Compute the Cartan–Killing form for su(1, 1) and su(2) and show that
they are different real forms of sl(2, C).
3. The algebra su(3) in the fundamental representation is determined by
Gell-Mann matrices
     
0 1 0 0 −i 0 1 0 0
λ1 =  1 0 0  , λ2 =  i 0 0  , λ3 =  0 −1 0  ,
0 0 0 0 0 0 0 0 0
     
0 0 1 0 0 −i 0 0 0
λ4 =  0 0 0  , λ5 =  0 0 0  , λ6 =  0 0 1  ,
1 0 0 i 0 0 0 1 0
   
0 0 0 1 0 0
1
λ7 =  0 0 − i  , λ8 = √  0 1 0  ,
0 i 0 3 0 0 −2
via Ta = 2i λ a so that Ta† = − Ta and tr( Ta Tb ) = − 12 δab .
a) Compute f abc .
b) Show that su(3) has a proper su(2) ⊕ u(1) subalgebra.
4. Let G = SO(2,2) and its algebra g = so(2, 2).
a) Provide dim g matrices for the basis of elements of g in the funda-
mental representation.
b) Show that H = SO(2) × SO(2) < G and that the two elements a1 , a2 ∈
g corresponding to the generators of the two SO(2) commute.
c) Are there other elements of g that are not in the span of a1 and a2
and that commute with them?
5. Let A = Sp(4, R) and B = SO(4).
a) Give their dimension and define their algebras in terms of matrices.
b) Discuss C = A ∩ B by analyzing a and b.
c) Compute the Cartan–Killing metric of the corresponding algebra c
and discuss if C can be simple, semi-simple and/or compact.
14
R E P R E S E N TAT I O N S O F L I E A L G E B R A S

We have discussed at length representations of finite and of continuous groups.


We would like now to discuss how to provide representations of Lie algebras
and how these representations may be connected to those of the underlying
group.
The first kind of representation we are going to discuss is provided by ma-
trices, but obviously in physics we are often interested in other types of repre-
sentations, either by differential operators or by quantum bosonic or fermionic
operators. For this reason we will also briefly discuss these instances, mainly
providing examples for some particularly interesting cases.

14.1 morphisms and representations

When we want to provide a representation of an algebra g we need to map its


elements to some matrices or operators, which fulfill the same commutation
relations that define g. More in general, we may be interested in maps between
algebras and see when we can map an algebra into another, preserving the
composition operation.

Definition 14.1. An algebra morphism between the algebras g1 and g2 is a linear


map φ : g1 → g2 preserving the bracket operation

[φ(t1 ), φ(t2 )]g2 = φ ([t1 , t2 ]g1 ) , ∀t1 , t2 ∈ g1 .

Definition 14.2. An invertible morphism is an isomorphism.

Since Lie algebras are first of all vector spaces, invertible linear maps respect-
ing the structure of the commutators are changes of basis.

143
144 algebra representations

As expected, we get isomorphic algebras from isomorphic groups, but not


the opposite:

G1 ' G2 ⇒ g1 ' g2 , (14.1)

but

g1 ' g2 ⇒
/ G1 ' G2 . (14.2)

Representations of an algebra are obtained when we consider morphisms of


an abstract algebra to an algebra on which we know how to perform explicit
computations.

14.2 matrix representations

Definition 14.3. A matrix representation of an algebra g is a morphism

D : g → gl (n, C).

Definition 14.4. We call the representation D : g → gl (n, C) faithful if it is an


injective morphism.
When dealing with matrices it is easy to relate group representations to al-
gebra representations. We already saw in chapter 11 how to provide a matrix
representation (the fundamental representation) to many groups and in chapter
12 how this generates a matrix representation for the algebra, by differentiating
1-parameter subgroups of the original group. The two representations can also
be related in the opposite direction by the exponential map:

exp [ D (t1 )] = D (exp[t1 ]), (14.3)

where t1 ∈ g, exp(t1 ) ∈ G and exp [ D (t1 )] is the standard matrix exponentia-


tion.
An important point for our interest in physical applications is the following
theorem:
Theorem 14.1. (Ado’s theorem) Every finite-dimensional Lie algebra admits a faith-
ful matrix representation.
This implies that we can always represent finite-dimensional Lie algebras
using matrices and therefore we can always map other types of representations
to matrix representations.
14.3 differential representations 145

As in the case of groups, we may have many different irreducible represen-


tations for the same algebras. Most of them depend on the algebra one is
studying, but some can be constructed in the same way for any algebra. One
such representation is the adjoint representation. We already saw in section
12.2.3 the definition of the ad X operator, for X ∈ g. This is an operator acting
on the vector space defined by the algebra g itself. We now show that it actually
provides a representation of the algebra (though not always faithful!).
Definition 14.5. The adjoint representation of an algebra g is provided by
adt I (·) ≡ [t I , ·], t I ∈ g.
This is a representation of dimension dim g because the vector space on
which it acts is g itself and it is indeed a representation because
[ ad X , adY ] = ad[X,Y ] , (14.4)
as follows straightforwardly from the Jacobi identity.
It is also interesting to note that we can lift this definition to a representation
of the corresponding group by defining
Ad g (t I ) = gt I g−1 , ∀ g ∈ G, t I ∈ g. (14.5)
This is also a representation of dimension dim g and its tangent space is pre-
cisely given by the adjoint representation of the algebra:

d d −1 d −1

Ad g(t) (t I )
= g(t)
t I g (0) + g (0) t I g(t) =
dt t =0 dt t =0 dt t =0 (14.6)
= [ X, t I ] = ad X (t I ),

where g(0) = 1 and X = ddt g(t) ∈ g.

t =0
The adjoint representation of a Lie algebra has some resemblance with the
regular representation of a finite group. In that case group elements were
used as a basis of the vector space on which the representation was acting
and in this case we use the Lie algebra itself as the vector space on which the
representation acts.

14.3 differential representations

We can often provide representations for Lie algebras in terms of differential


operators, both in classical and quantum mechanics (with obvious limitations
146 algebra representations

on the spaces of functions on which they act). In the following we will not
be careful about the definition of a proper space of functions on which these
operators act (be it L2 (R) or C ∞ or tempered distributions or else), but always
assume that the operations are legitimate. We are only interested in showing
how some of the most common groups and algebras in physics can be repre-
sented in this way.
We already met the first example, which is the group of translations, isomor-
phic to (R, +). A representation for its 1-dimensional algebra is given by the
derivative of real functions:

d
D ( t1 ) = . (14.7)
dx

A representation for the group follows from the exponential map


 
d
D (exp( a t1 )) = exp a (14.8)
dx

and its action on test functions realizes the group action:


 
d
exp a f ( x ) = f ( x + a ). (14.9)
dx

A group that is extremely useful in physics is the group of rotations in n-


dimensional Euclidean space. The corresponding algebra is so(n) and is given
by the generators of such rotations. In general we can construct infinitesimal
generators of rotations by means of the differential operators

∂ ∂
Tij ≡ xi j
− xj i . (14.10)
∂x ∂x

The antisymmetric double index ij = − ji is useful to create a 1 to 1 correspon-


dence with the n(n − 1)/2 generators of the group and therefore provide a
faithful representation. The commutator relations are

[ Tij , Tkl ] = −δik Tjl + δjk Til − δjl Tik + δil Tjk , (14.11)

as it can be easily verified by computing it on a test function f (~x ) and recalling


j
that ∂i x j = δi and noting that the terms with two derivatives acting on the
function f (~x ) disappear because of antisymmetrizations.
14.4 bosonic representations 147

In the simplest case, namely so(2), we have only one such generator T =
x ∂y − y ∂ x and this can be simplified using polar coordinates x = r cos θ, y =
r sin θ, leading to

∂ ∂x (r, θ ) ∂y(r, θ )
T= = ∂x + ∂y = −r sin θ ∂ x + r cos θ ∂y . (14.12)
∂θ ∂θ ∂θ
For so(3) we get the expected commutation rules if we define

Tij = −eijk lk , (14.13)

so that

[li , l j ] = eijk lk . (14.14)

The same type of operators could be easily introduced for so( p, q), defining

Tij ≡ ηik x k ∂ j − η jk x k ∂i , (14.15)

where
1p
!
η= . (14.16)
−1q

The commutator of the generators (14.15) can also be computed rather easily
and gives

[ Tij , Tkl ] = η jk Til + ηil Tjk − ηik Tjl − η jl Tik . (14.17)

14.4 bosonic representations

The realization of Lie algebras using bosonic operators is very useful in Physics
in the resolution of problems like the harmonic oscillator in Quantum Mechan-
ics and in Quantum models of rotational and vibrational models of nuclei and
molecules.
Assume that we have n bosonic operators bα , α = 1, . . . , n, satisfying the
following commutation relations

[bα , b†β ] = δαβ , [bα , bβ ] = [bα† , b†β ] = 0. (14.18)

We can think of these operators as creation and destruction operators for quan-
tum states, starting from a vacuum satisfying bα |0i = 0.
148 algebra representations

By means of the defining relations (14.18) we can easily prove that

Tαβ ≡ bα† bβ (14.19)

provide a representation of the gl(n, C) generators, satisfying

[ Tαβ , Tγδ ] = δβγ Tαδ − δδα Tγβ . (14.20)

This is clear if one notes that they are in 1-to-1 correspondence with the matrices
representing the same group. For n = 2 for instance
   
1 0 0 1
T11 = , T12 = ,
0 0 0 0
    (14.21)
0 0 0 0
T21 = , T22 =
1 0 0 1
satisfy exactly the same commutation relations and therefore we can establish
an isomorphism between the two algebras. Note also that these generators
can be further split into those of sl(n, C), whose matrix representation has
vanishing trace, plus the identity matrix, that commutes with all the others,
providing a “linear Casimir” C1 = ∑i Tii .
Recall that sl(n, C) is the complexification of su(n) and gl(n, C) is the com-
plexification of u(n), which are their real sections. This implies that we can
relate their representations by taking appropriate real linear combinations.
We now provide some examples that hopefully will clarify the role of these
symmetry generators in Physics.

14.4.1 u(1)C ' gl(1, C)

Consider the 1-dimensional harmonic oscillator, whose Hamiltonian is


1 2 1 2
H= p̂ + x̂ . (14.22)
2 2
Using the operator
1
b ≡ √ ( x̂ + i p̂) (14.23)
2
we can rewrite (14.22) as
1 1
H = b† b + = T11 + , (14.24)
2 2
14.4 bosonic representations 149

where, according to the discussion above, T11 is the generator of gl(1, C). This
algebra is obviously Abelian and hence all its irreducible representations are
1-dimensional and labeled by the value of the action of the linear Casimir T11
on the corresponding invariant subspaces of the Hilbert space. A basis of states
generating these invariant spaces is given by

1
| N i = √ ( b † ) N |0i. (14.25)
N!
Since the Hamiltonian is given in terms of the linear Casimir, we can immedi-
ately deduce its eigenvalues and their degeneracies, computing its values on
the invariant spaces of the Hilbert space:
 
1
H|Ni = N + | N i. (14.26)
2

14.4.2 u(2)C ' gl(2, C)

We now move to the 2-dimensional harmonic oscillator. In this case the Hamil-
tonian is
1 2 1
H= ( p̂ + p̂2y ) + ( x̂2 + ŷ2 ) = b1† b1 + b2† b2 + 1, (14.27)
2 x 2
where
1 1
b1 ≡ √ ( x̂ + i p̂ x ), b2 ≡ √ (ŷ + i p̂y ). (14.28)
2 2

Using bα and bα† operators we can construct the 4 generators of the gl(2, C)
algebra:

T11 = b1† b1 , T12 = b1† b2 , T21 = b2† b1 , and T22 = b2† b2 . (14.29)

A basis for the states of the Hilbert space is

1 †
|Ni = b . . . bᆠN |0i, (14.30)
N α1
where N is some normalization factor, and they form an irreducible and com-
pletely symmetric representation of dimension N + 1 (We will come back on
this in section 16.2.3). Once again we see that the Hamiltonian is a function of
150 algebra representations

the linear Casimir N b = b† b1 + b† b2 . This implies that its eigenvalues and their
1 2
degeneracies are also determined by such Casimir.
Note that usually one speaks of the symmetry generators of the harmonic
oscillator as generators of su(n) rather than sl(n, C). The reason is that we
are really interested in unitary transformations that preserve the orthonormal-
ity of the state vectors and this poses obvious restrictions on the actual linear
combinations of the Tij operators that generate admissible symmetries.

14.5 fermionic representations

In certain instances creation and annihilation operators have to be taken with


the opposite statistic, i.e. they have to be (complex) fermionic operators, satis-
fying

{ aα , a†β } = aα a†β + a†β aα = δαβ , (14.31)

and

{ aα , a β } = 0 = { a†α , a†β }. (14.32)

Also in this instance

Tαβ = a†α a β (14.33)

generate gl(n, C), but now the states forming a basis of the Hilbert space on
which they act are in fully antisymmetric representations

1 †
|Ni = a . . . a†α N |0i. (14.34)
N α1
The check is straightforward also in this case, but pay attention to the commu-
tators:

[ Tαβ , Tγδ ] = a†α a β a†γ aδ − a†γ aδ a†α a β

= a†α { a β , a†γ } aδ − a†γ { aδ , a†α } a β − a†α a†γ a β aδ + a†γ a†α aδ a β (14.35)

= δβγ Tαδ − δδα Tγβ − { a†α , a†γ } a β aδ + a†γ a†α { aδ , a β }

= δβγ Tαδ − δδα Tγβ . (14.36)


14.5 fermionic representations 151

exercises

1. Prove that the Adjoint representation Dad (tK ) I J = f K J I for t I ∈ g is indeed


a representation of g, i.e. verify that

[ Dad (t I ), Dad (t J )] = f I J K Dad (tK ).

2. What about ( TI ) J K = − f I J K ?
Explicitly: consider the matrices TI , with non-zero elements as given above and
check the commutator [ TI , TJ ].

3. What is the algebra defined by the operators Li = − 2i x α (σi )α β ∂ β acting


on functions f ( x1 , x2 )?

4. Let
D ( Jµν ) = −ηµρ x ρ ∂ν + ηνρ x ρ ∂µ , D ( Pµ ) = ∂µ ,
where µ, ν, ρ = 0, 1, 2, 3. Find the structure constants of the resulting alge-
bra.

5. Casimir operators for gl(n, C).


a) Show that the matrices ( Tij ) ab = δia δjb , form a basis of gl(n, C) and
compute their structure constants.
Again: Tij are matrices with row and column indices a and b.
b) Compute the Cartan–Killing form in this basis and verify the algebra
properties using Cartan’s criteria.
c) Show that the elements of the enveloping algebra

Cr ≡ ∑ Ti1 i2 Ti2 i3 . . . Tir i1


1≤i1 ,i2 ,...,ir ≤n

commute with all the elements Tij and therefore they are Casimir
operators of degree r.
Try this at least for r = 1, 2, 3.

6. Important! Recall the definition of so(n) as given in (14.10). Show that


Tij , where i, j = 1, . . . , n − 1 generate a so(n − 1) ⊂ so(n).
Focus now on the n = 4 case. In this case we can prove that so(4) =
su(2) ⊕ su(2).
152 algebra representations

(1) (2)
Split the so(4) generators in Li = { T23 , T31 , T12 } and Li = { T14 , T24 , T34 }.
( A) ( B)
Compute [ Li , L j ].
 
(1) (2)
Define Li± ≡ 12 Li ± Li and close the argument.

7. An exceptional group. (very long)


Let Jab , with a, b = 1, . . . , 7, be the 21 generators of so(7). Introduce the
totally antisymmetric tensor φabc , whose non-zero entries follow from

φ123 = φ516 = φ624 = φ435 = φ471 = φ673 = φ572 = 1.

Show that the generators surviving the projection ∑ j,k φijk Jjk = 0 define a
subalgebra of so(7) (i.e. consider the vector space generated by the linear
combinations of the Jij generators orthogonal to the 7 vectors defined by
∑ j,k φijk Jjk ). Compute its structure constants and the Cartan–Killing form.
Is this one of the algebras we studied so far (so(n), su(n), sp(n))?
15
R E P R E S E N TAT I O N S O F S U ( 2 ) A N D S O ( 3 )

15.1 rotation group and its algebra

The relevance of the group associated to the invariance of a physical system un-
der spatial rotations in 3 dimensions is obvious. This group can be determined
as the set of transformations preserving the norm of 3-dimensional vectors

~ → RV,
V ~ |V ~ T R T RV
~ |2 → V ~ |2
~ = |V (15.1)

and is isomorphic to SO(3), defined in 11.10. Its algebra, so(3), is defined by


the commutators

[li , l j ] = eijk lk , (15.2)

where li , i = 1, 2, 3 form a basis of the so(3) vector space. In the fundamental


representation, the generators li are represented by 3 × 3 antisymmetric real
matrices, fulfilling commutation relations isomorphic to (15.2):

     
0 0 1 0 −1
D3 ( l1 ) =  0 − 1  , D3 ( l2 ) =  0  , D3 ( l3 ) =  1 0 .
1 0 −1 0 0
(15.3)

Using an explicit index notation for the matrix elements

D3 (li ) jk = −eijk . (15.4)

153
154 representations of su(2) and so(3)

15.2 isomorphism su(2) ' so(3) and homomorphism su(2) 7→ so(3)

The algebra of angular momenta, as already mentioned in previous lectures,


is isomorphic to the su(2) algebra. We can easily show this by explicitly con-
structing the latter and relating it to the one provided in (15.2).
The starting point is the fundamental matrix representation of SU(2). A 2 ×
2 matrix U provides such a representation if it satisfies U † U = 1 and det U = 1
(see definition 11.8). This means that we can parameterize U in terms of two
complex numbers z, w ∈ C as follows:
 
z w
U= , with |z|2 + |w|2 = 1. (15.5)
−w∗ z∗

As a byproduct, this parameterization shows that we can identify the manifold


underlying the SU(2) Lie group with the 3-sphere:

SU(2) ' S3 . (15.6)

In fact, rewriting z = x4 + i x3 and w = x2 + i x1 , for xi ∈ R, we have that the


constraint in (15.5) becomes

∑(xi )2 = 1, (15.7)
i

which is the defining equation of S3 as a hyper-surface embedded in R4 .


Starting from general U matrices as in (15.5), by differentiation we obtain
the su(2) algebra in the fundamental representation. This is the space of 2-
dimensional matrices A satisfying A† = − A and tr A = 0. Pauli matrices can
be used to provide a basis for such a space:
     
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 = . (15.8)
1 0 i 0 0 −1

In fact these matrices satisfy

σi† = σi and σi σj = i eijk σk + δij 12 (15.9)

and we can see that the combinations


i
D2 (li ) = − σi (15.10)
2
15.2 isomorphism su(2) ' so(3) and homomorphism su(2) 7→ so(3) 155

are anti-hermitian D2 (li )† = − D2 (li ) and have vanishing trace. We then con-
clude that D2 (li ) is a 2-dimensional representation of su(2) whose commutators
define the commutators of the algebra. It is also easy to see that the resulting
structure constants are isomorphic to the ones of so(3):
[ D2 (li ), D2 (l j )] = eijk D2 (lk ). (15.11)
We have then established that
su(2) ' so(3). (15.12)
We can now prove that by taking the exponential map of this 2-dimensional
representation of su(2) we can generate back the 2-dimensional representation
of SU(2) that covers the whole group.
A generic element of the algebra can be given in terms of the D2 representa-
tion by A = − 2i θ i σi ∈ su(2). Its exponential gives
∞  n
i i 1
U = exp( A) = ∑ − θ σi
n =0
2 n!
(15.13)
i i 1 1 i j
= 1 − θ σi − θ θ σi σj + . . .
2 2 4
and it can be simplified using that
1 i j  1
θ i θ j σi σj = θ θ σi , σj = θ i θ j δij 12 ,

(15.14)
2 2
which implies
  
i i 1 1 i 1
U = 12 − θ σi − θ i θ j δij 12 + − θ k σk − θ i θ j δij + . . .
2 8 3! 2 4
! !
1 |~θ |2 i k 1 |~θ |2
= 12 1− + . . . − θ σk 1 − +...
2 4 2 3! 4

|~θ | θi |~θ |
= 12 cos −i σi sin .
2 |~θ | 2
(15.15)
Matching this expression with the one for U given in (15.5), we see that

|~θ | ~θ |~θ |
x4 = cos and ~x = − sin , (15.16)
2 |~θ | 2
156 representations of su(2) and so(3)

so that ∑i ( xi )2 = 1 and we cover the whole S3 .


Note that su(2) also has a 3-dimensional irreducible representation: the ad-
joint representation. This acts on the su(2) elements as
 
i
ad− i ~θ ·~σ [σa ] = − ~θ ·~σ, σa = θ c ecab σb = σb (−θ c ecba ) = σb ~θ · D3 (~l )ba , (15.17)
2 2

where D3 (~l ) provides its matrix representation. It is straightforward to check


that this representation coincides with the fundamental representation of so(3).
As presented in (14.5), the adjoint representation of an algebra induces the
adjoint representation for the corresponding group:

AdU [σa ] = Uσa U † = σb D3 (U )ba . (15.18)

Explicitly
   
i i~  
exp − ~θ ·~σ σa exp θ ·~σ = exp ad− i ~θ ·~σ [σa ] =
2 2 2

= σa + ad(− i ~θ ·~σ) σa + 21 ad(− i ~θ ·~σ) ad(− i ~θ ·~σ) σa + . . .


2 2 2
   (15.19)
= σa + σb ~θ · D3 ~l + σc 21 ~θ · D3 ~l ~θ · D3 ~l +...
ba cb ba
  
= σb exp ~θ · D3 ~l .
ba

Since SU(2) is locally parameterized by the coordinates θ i , each element of SU(2)


corresponds to a distinct value of θ i . By using this fact we can now build a map
that relates the 2-dimensional and the 3-dimensional representations of SU(2)
and in turn elements of SU(2) with elements of SO(3):

−θ 3 θ2
 
0
θ3 θ1 − i θ2
  
i
exp − 7→ exp  θ 3 0 −θ 1  . (15.20)
2 θ + i θ2
1 −θ 3
−θ 2 θ1 0

This map shows two important things:

• the adjoint representation of SU(2) is not faithful;

• SU(2) and SO(3) are homomorphic, but not isomorphic.


15.3 so(3) topology 157

In fact we can see that the map in (15.20) is not a bijection, because it is not
invertible. Let us consider the simplified case where θ 1 = θ 2 = 0. The map
relates elements of SU(2) and SO(3) as

cos θ 3 − sin θ 3
!  
3 /2
e−iθ
3 /2 7→  sin θ 3 cos θ 3  (15.21)
eiθ 1

and this means that different elements in SU(2) may have the same adjoint
representation and therefore be mapped to the same element in SO(3). A simple
instance is obtained by looking at the elements associated to the values θ3 =
0 and θ3 = 2π. While in SU(2) these two elements are distinct and the 2-
dimensional faithful representation shows this explicitly, their 3-dimensional
representation is the same. In fact in SO(3) there is a unique element associated
to such values of the parameters. Explicitly, at θ 3 = 0 we have

D2 (0, 0, 0) = 12 7→ D3 (0, 0, 0) = 13 , (15.22)

while at θ 3 = 2π we have

D2 (0, 0, 2π ) = −12 7→ D3 (0, 0, 2π ) = 13 . (15.23)

We conclude that the map (15.20) is not a bijection, but a 2 to 1 homomorphism,


whose kernel is a subgroup of SU(2) given by Z2 = {12 , −12 }.
As usual, we can restore an isomorphism by identifying group elements in
the kernel:

SO(3) ' SU(2)/Z2 . (15.24)

15.3 so(3) topology

From (15.24) we see that topologically SO(3) is equivalent to the 3-sphere where
points are identified by a Z2 projection.
158 representations of su(2) and so(3)

x4 < 0 x4 > 0
Z2

Figure 22.: Z2 map identifying points on the 3-sphere. Contractible and non-
contractible cycles.

The 3-sphere can be described by its embedding in R4 as the set of points


satisfying ∑i ( xi )2 = 1. Thus, for eachp value of | x4 | the coordinates x1 , x2
and x3 span two 2-spheres of radius 1 − | x4 |2 . While each of these points
corresponds to a different element in SU(2), only half of them are distinct in
SO(3). For instance, the map (15.20) shows that the identity matrix 12 appears
when x1 = x2 = x3 = 0 and x4 = 1 and that −12 corresponds to the choice
x1 = x2 = x3 = 0 and x4 = −1. On the other hand, both points are representing
the identity of SO(3). We then conclude that the action of Z2 identifies points
of the 2-sphere with x4 > 0 to those of the 2-sphere with x4 < 0 (see figure 22
for a description of this fact). The resulting manifold is SO(3) and it is multiply
connected, with first homotopy group

π1 (SO(3)) = Z2 . (15.25)

This implies that we have 2 classes of equivalence for loops in SO(3): the class
of contractible loops and the class of loops that cannot be contracted to a point.
A pictorial explanation of this fact is the following. The SO(3) manifold is the
set of rotations in 3 dimensions. Any element g ∈ SO(3) can be described by a
rotation of an angle between −π and π around an axis, oriented in a specific
direction in R3 . Obviously rotations by π must be identified with rotations by
−π. This means that we can describe SO(3) as a closed ball in 3 dimensions
of π radius, where opposite points on the surface must be identified. We can
now easily see that a loop that starts from a point on the boundary and runs
in the interior up to the opposite point on the surface of the ball is a closed
loop in SO(3) that can never be contracted to a point. On the other hand if we
have a loop that starts from a point on the boundary, runs in the interior up
15.4 irreducible representations of su(2) and so(3) 159

to another point on the boundary, different from the first, and does the same
thing on the opposite side, we can continuously deform it to a closed loop that
can be contracted to a point by moving different points on the boundary on top
of each other (see fig. 23).

Figure 23.: Contraction of a loop that passes twice through the boundary of the ball
representing SO(3). Red and green dots are points on the boundary, which
are identified. First we move the red dot and its image on top of the green
ones and then we can contract the loop.

This is just a special instance of what happens for any SO(n) group. All
SO(n) spaces are not simply connected and their universal coverings, which
are simply connected, are the Spin groups Spin(n). We conclude that

Spin(3) ' SU(2). (15.26)

15.4 irreducible representations of su(2) and so(3)

What we just presented shows that SU(2) and SO(3) representations are related
and that SU(2) representations provide representations for the universal cover-
ing group of SO(3).
In Quantum Mechanics we are interested in unitary representations for the
group elements and hermitian representations for the algebra. The homomor-
phism between SU(2) and SO(3) shows that irreducible representations of SU(2)
are identical to irreducible representations of SO(3) up to a sign. In fact, if we
still want to insist and think of D2 as a representation of SO(3), we can think
of it as a projective representation, whose composition law depends on a phase
(which is invisible in the ray defining a state):

D2 ( g1 ) D2 ( g2 ) = exp(iφ( g1 , g2 )) D2 ( g1 ◦ g2 ), g1 , g2 ∈ SO(3). (15.27)


160 representations of su(2) and so(3)

What we can infer from this is that SO(3) admits genuinely projective represen-
tations and that its covering group automatically takes them into account. In
fact, for all practical applications we can actually study the representations of
the latter. For this reason we will now discuss representations of SU(2) and of
its algebra rather than those of SO(3).
In general, a series of theorems by Bargmann1 shows that

Theorem 15.1. A simply connected group without continuous Abelian normal sub-
groups does not admit genuinely projective irreducible representations.

This means that whenever the group we want to represent is not simply
connected, we can use its covering group.

15.5 irreducible representations of su(2)

In order to study representations of su(2) it is useful to realize that su(2)C '


sl(2, C). In fact it is easier to construct representations for this algebra and
then deduce the form of the su(2) representations by taking appropriate linear
combinations sections.
We start by showing that su(2)C ' sl(2, C), indeed. If li , i = 1, 2, 3, are the
su(2) generators, we can construct a complex basis for its complexification by
defining the linear combinations

J± ≡ i (l1 ± i l2 ), J3 ≡ i l3 . (15.28)

These generators satisfy

[ J3 , J± ] = ± J± , [ J+ , J− ] = 2 J3 (15.29)

and the isomorphism of this algebra with sl(2, C) is obvious in the 2-dimensional
representation
1
     
2 0 0 1 0 0
D2 ( J3 ) = , D2 ( J+ ) = , D2 ( J− ) = , (15.30)
0 − 12 0 0 1 0

which is equivalent to the fundamental representation of sl(2, C), i.e. 2 × 2


matrices with vanishing trace.
1 Again, we admit finite Abelian subgroups and therefore we do not use the word semi-simple,
consistently with our definition.
15.5 irreducible representations of su(2) 161

In order to build a generic finite-dimensional irreducible representation, we


need a basis of vectors |mi spanning the corresponding invariant space V. We
can always choose a basis in which one of the generators is diagonal. It is cus-
tomary to use J3 for this purpose, so that the index m denotes its eigenvalues:

D ( J3 )|mi = m|mi. (15.31)

We then note that J± |mi are still eigenstates of J3 with eigenvalues increased or
reduced by 1 with respect to |mi:

D ( J3 ) D ( J± )|mi = [ D ( J3 ), D ( J± )]|mi + D ( J± ) D ( J3 )|mi


= ± D ( J± )|mi + m D ( J± )|mi (15.32)
= (m ± 1) D ( J± )|mi.

If the representation is finite-dimensional there must be a j such that D ( J+ )| ji =


0 and a j− such that D ( J− )| j− i = 0. The basis of the invariant subspace is
then constructed acting with D ( J− ) on | ji until we reach the state j− . In fact
D ( J− )| ji ∼ | j − 1i, D ( J− )| j − 1i ∼ | j − 2i, etc., up to the moment that we find
a j− such that D ( J− )| j− i = 0.
Another important point that fixes the representation is that tr D ( J3 ) = 0,
which follows from
1
tr D ( J3 ) = tr ([ D ( J+ ), D ( J− )]) = 0, (15.33)
2
where we used (15.32) in the first equality and the cyclic property of the trace
in the second. We then conclude that

j + ( j − 1) + ( j − 2) + . . . + j− = 0, (15.34)

which implies that j− = − j and that the vector space has dimension 2j + 1,
with basis

{| ji, | j − 1i, . . . , | − ji}. (15.35)

Since the dimension of the vector space V must be an integer number, we also
conclude that

j ∈ N/2. (15.36)
162 representations of su(2) and so(3)

The number j is called the spin of the representation and the vector space on
which it acts is given by

V = span(| j, mi), where m = j, j − 1, . . . , − j. (15.37)

It is interesting to note that we can always find the spin of the representation
of sl(2, C) by using an operator that commutes with the entire set of matrices
providing the representation. By means of Schur’s lemmas we know that on
the common invariant space V this must be proportional to the identity and
therefore have a unique value when acting on any of the vectors of V. An
operator that has this property is the quadratic Casimir. In the sl(2, C) basis
defined by J3 , J± satisfying (15.29), the Cartan–Killing form is
 
4
k= 4  (15.38)
2

and the quadratic Casimir is


 
1 1 1
C2 = J32 + J+ J− + J− J+ . (15.39)
2 2 2

Its representation D (C2 ) must be diagonal on V, i.e.

D (C2 )| j, mi = λ( j) | j, mi (15.40)

for any m. Since D ( J− )| j, − ji = 0, we can simplify the expression of C2 using


the fact that J− J+ = J+ J− − 2J3 and therefore

1  1
D (C2 )| j, − ji = D ( J3 )2 − D ( J3 ) + D ( J+ ) D ( J− ) | j, − ji = j( j + 1)| j, − ji.
2 2
(15.41)

We conclude that the Casimir operator on the 2j + 1-dimensional representation


of sl(2, C) has always the same value on V, which is

1
λ( j) = j ( j + 1). (15.42)
2
Irreducible representations of sl(2, C) are then distinguished by the eigenvalues
of D (C2 ) and the basis of vectors in each of the invariant spaces is labelled by
the eigenvalues of D ( J3 ).
15.6 matrix representations 163

15.6 matrix representations

The procedure we just outlined let us also find a (2j + 1)-dimensional matrix
representation of su(2) explicitly. We start from the basis of normalized vec-
tors providing a basis for the invariant space of the corresponding irreducible
representation of sl(2, C) and then take the linear combinations leading to the
appropriate real form.
Define | j, ji as a norm one state:

|| D2j+1 ( J− )| j, ji ||2 = h j, j| D2j+1 ( J+ ) D2j+1 ( J− )| j, ji

= h j, j|[ D2j+1 ( J+ ), D2j+1 ( J− )]| j, ji (15.43)


= h j, j|2 D2j+1 ( J3 )| j, ji = 2j.

We can then define


1
| j, j − 1i ≡ p D2j+1 ( J− )| j, ji. (15.44)
2j

We can then proceed in the same way for the other states and eventually define
the form of D2j+1 (la ) as

D2j+1 (l3 ) = −i D2j+1 ( J3 ), (15.45)

i 
D2j+1 (l1 ) = − D2j+1 ( J+ ) + D2j+1 ( J− ) (15.46)
2
and
1 
D2j+1 (l2 ) = − D2j+1 ( J+ ) − D2j+1 ( J− ) (15.47)
2
Note that when j ∈ N we also have a representation for so(3).
Example 15.1. The j = 1/2 representation has highest state | 21 , 12 i. We start by

D2 ( J3 )| 21 , ± 12 i = ± 21 | 12 , ± 21 i. (15.48)

We then have that

| 12 , − 12 i ≡ D2 ( J− )| 12 , 12 i (15.49)
164 representations of su(2) and so(3)

and

D2 ( J− )| 12 , − 12 i = 0. (15.50)

Analogously

D2 ( J+ )| 12 , 12 i = 0, (15.51)

and
D2 ( J+ )| 21 , − 12 i = D2 ( J+ ) D2 ( J− )| 12 , 21 i

= [ D2 ( J+ ), D2 ( J− )]| 12 , 12 i (15.52)

= 2D2 ( J3 )| 21 , 12 i = | 21 , 21 i.

If the generic vector in Vj= 1 has two components and | 12 , 21 i is the upper com-
2
ponent and | 12 , − 12 i is the lower component, the corresponding representation
2 is given by
 1     
2 0 0 1 0 0
D2 ( J3 ) = , D (
2 +J ) = , D (
2 −J ) = , (15.53)
0 − 12 0 0 1 0

which in fact is the same as


i
D2 (li ) = − σi . (15.54)
2

exercises

1. Build the 8-dimensional adjoint representation of su(3). Take the 3 ma-


trices related to su(2) ⊂ su(3) and show that they provide a reducible
representation of su(2).
2. Build the irreducible representation of dimension 5 for so(3).
a) Start from the basis of eigenstates of its complexified algebra |2j +
1 = 5, mi, with m = −2, . . . , 2 and build D5 ( J3 ).
b) Using the properties of J± and their action on the same states build
D5 ( J± ) and D5 ( J1 ), D5 ( J2 ).
c) Build the corresponding representation for the quadratic Casimir op-
erator D5 (C2 ).
16
I R R E D U C I B L E R E P R E S E N TAT I O N S O F S U ( N ) A N D T H E I R
PRODUCTS

There are many groups relevant to physics, but a special role is played by SU(N)
groups. These groups appear naturally in Quantum Mechanics, for instance in
the analysis of the harmonic oscillator and of rotational symmetries, but they
are also at the core of the so-called Standard Model of Particle Physics. This
model is in fact a gauge theory coupled to matter, with gauge group SU (3) ×
SU (2) × U (1) and fundamental particles are sitting in specific representations
of these groups. In addition, there are approximate global symmetries that also
play a role in the analysis of fundamental processes, which are again described
by groups isomorphic to SU(N).
We are not just interested in constructing representations of SU(N) groups,
but also in understanding their products, because it is often useful to under-
stand how a composite system transforms once the transformation properties
of its components are known. Thus, after a brief general discussion, we present
in detail SU( N ) irrepses and their products using Young Tableaux. We stress
that, although some details of the techniques we use here are specific to SU(N)
groups, most of the discussion can be repeated and extended to other groups.

16.1 products of irreducible representations

16.1.1 Tensor products

A method that allows to build all irreducible representations (irrepses) of a


group starting from its fundamental representation is that of considering their
products and their decompositions according to their invariant subspaces.

165
166 su(n) irreps and products

The product of irrepses of a given group or algebra provides a new represen-


tation for the same group or algebra. The resulting representation, however, is
in general reducible, unless one of the two irrepses has dimension 1. In order
to obtain a reduction of the product according to its invariant subspaces it is
useful to work at the algebra level and then exponentiate the fragments result-
ing from the decomposition. For SU(N) groups we are sure to cover the whole
group by doing so, because SU(N) is compact, simply connected and has no
continuous Abelian normal subgroups.
Definition 16.1. Let Da be a representation of a group (algebra) acting on the
vector space Va and Db a representation acting on Vb , their product Da ⊗ Db acts
on Va ⊗ Vb as

( Da ⊗ Db )( g)[v a ⊗ vb ] = Da ( g)[v a ] ⊗ Db ( g)[vb ],

for g ∈ G (or g ∈ g) and v a,b ∈ Va,b .

16.1.2 Clebsch–Gordan decomposition

The result of a product of representations is not irreducible in general. When it


is also reducible (and not just decomposable), we can decompose it in irrepses.
This is always the case for compact Lie groups.
The decomposition of the product takes the name of Clebsch–Gordan decom-
position:
Definition 16.2. Clebsch–Gordan decomposition

D ⊗ D0 = ⊕ j Dj .

If G is finite or if it is a compact Lie group and its irrepses have been classified,
we can use the multiplicity m a of each irreducible fragment to write

D ⊗ D 0 = ⊕ a m a Da . (16.1)

Obviously we have that their characters are also related by

χ D⊗ D0 ( g) = χ D ( g) χ D0 ( g) = ∑ m a χ Da ( g ) (16.2)
a

and their dimensions (take the previous formula for g = 1) by

ND⊗ D0 = ND · ND0 = ∑ ma Na . (16.3)


a
16.2 su(n) irrepses and young tableaux 167

This decomposition is in 1 to 1 correspondence with the decomposition of


the vector space on which the same linear representations act. In detail, if
| a, mi ∈ Va and |b, ni ∈ Vb , with m = 1, . . . , dim Va and n = 1, . . . , dim Vb are
their bases, we can decompose the vectors in the product space as follows:

| a, m; b, ni ≡ | a, mi ⊗ |b, ni = ∑ |ci , pih p, ci |a, m; b, ni, (16.4)


c,i,p

where c is a label for the invariant subspace for the related irreps, i runs over
the number of times the same representation is repeated and p is the index
labeling the basis vectors. The coefficients Cci ,p| a,m,b,n = h p, ci | a, m; b, ni are
called the Clebsch–Gordan coefficients.

16.1.3 Decomposition in irrepses of a subgroup

Although it is not going to be addressed here, it is useful to note that there


is another case of interest where a given representation can be decomposed
in irreducible fragments. This happens in the case of a group H < G, where
irrepses of H are constructed by restricting those of G to the elements in H.
In this case D (h) is not in general irreducible, even if D ( g) is irreducible. For
instance, the dimension 2 irreps of the group D3 induces a representation of
Z2 < D3 which is reducible, because Z2 is Abelian.

16.2 su(n) irrepses and young tableaux

Without giving too many technical details, we now discuss irrepses of the Spe-
cial Unitary group SU(N).

16.2.1 Fundamental, anti-fundamental and adjoint representations

The fundamental representation of SU(N) is defined by matrices U ∈ GL( N, C)


that fulfill U † U = 1 N and det U = 1. These linear operators act on a vector
space V = C N , mapping vectors v = vi ei ∈ V (where ei is a basis for V) as

vi 7 → U i j v j , (16.5)
168 su(n) irreps and products

where i, j = 1, . . . , N label rows and columns of the matrix U. It is customary


to label irreducible representations with their dimension and therefore we have
that

Fundamental Representation ⇔ N. (16.6)

Once we proved that the matrix U ( g) provides a good representation of the


element g ∈ SU( N ), we see that also U ∗ ( g) provides a representation for the
same element:

(U ∗ )† U ∗ = (U † U )∗ = 1∗N = 1 N (16.7)

and

det U ∗ = (det U )∗ = 1. (16.8)

For N > 2 this is an inequivalent representation called anti-fundamental, which


we can represent by its action on w ∈ C N as

w∗ ı̄ 7→ U ∗ ı̄ ̄ w∗ ̄ . (16.9)

Once more, it is customary to refer also to this representation using its dimen-
sion, though in this case we also use a bar to distinguish it from the fundamen-
tal representation and stress that it can be obtained from the former by complex
conjugation:

Anti-Fundamental Representation ⇔ N. (16.10)

Using the scalar product in C N we also see that

w† v = w∗ı̄ vi δiı̄ = wi∗ vi (16.11)

and we can characterize the N representation by the action on covariant vectors,


whose index has been lowered by means of δiı̄ :

wi∗ 7→ U ∗ i j w∗j . (16.12)

We now move to the adjoint representation. The product of the two irrepses
N ⊗ N can be described by the set of operators acting on the vector space
defined by rank 2 tensors as

vi ⊗ w∗j ≡ Mi j 7→ U i k U ∗ j l Mk l . (16.13)
16.2 su(n) irrepses and young tableaux 169

As expected, this is not an irreducible representation. In fact the scalar product


of the elements is invariant (i.e. it transform in the trivial representation)

w† v 7→ wk∗ U ∗ i k U i l vl = w† U † U v = w† v (16.14)

and thus we can decompose the tensor Mi j into its trace and its traceless parts:

1 i ◦ ◦
Mi j = δj tr M + M i j , with M k k = 0. (16.15)
N
In summary

N ⊗ N = 1 ⊕ ( N 2 − 1 ), (16.16)

where the first one is the trivial representation and the second one is the adjoint.
We recognize that this is really the adjoint representation by the fact that the
dimension is the correct one (dimR SU(N) = N 2 − 1) and by the fact that the

action on M is precisely the one expected by the Adjoint representation, which
takes vectors in the algebra and acts on them by elements of the group from
the left and inverse elements from the right:
◦ ◦ ◦
AdU ( M ) = U M U −1 = U M U † . (16.17)

As we said in the introduction, we would like to see how all irrepses can
be generated by taking products of the fundamental representation. We start
by giving a simple example, where we consider rank 2 tensors built from 2
fundamental representations. These tensors transform as

Mij ≡ vi ⊗ w j 7→ U i k U j l Mkl . (16.18)

They are obviously always decomposable in the symmetric and antisymmetric


parts

Mij = Aij + Sij = − A ji + S ji , (16.19)

but what is interesting to note is that these fragments define invariant sub-
spaces under the action of SU(N):

Aij 0 = U i k U j l Akl = −U i k U j l Alk = −U j l U i k Alk = − A ji0 , (16.20)

Sij 0 = U i k U j l Skl = U i k U j l Slk = U j l U i k Slk = S ji0 . (16.21)


170 su(n) irreps and products

We conclude that
N(N − 1) N(N + 1)
N⊗N = ⊕ . (16.22)
2 2
Using this tensorial method we can extract all irreducible components in a
product, but it is generically too elaborate and complicated. However, a much
more effective method can be put in place by using Young Tableaux.

16.2.2 Irrepses and Young Tableaux

Tensorial representations can be decomposed in invariant parts using their


properties under permutations of their indices. Since conjugacy classes and
irreducible representations of the group of permutations are effectively repre-
sented by Young Tableaux, we can apply this technique to decompose products
of SU(N) irrepses. Each irreps is obtained as a product of k fundamental ir-
repses, projected on the irreducible part that is selected by the elements of the
permutation group represented by a specific Young tableau.
The fundamental representation acts on a single vector and can be repre-
sented by the unique element of S1 :

N ⇔ . (16.23)

Other irrepses follow from its k-fold product, with symmetrization and anti-
symmetrization rules determined by the Young tableau we use to represent it.
For instance the completely symmetric representation is

(N + k − 1)!
··· ⇔ T i1 ...ik = T (i1 ...ik ) ⇔ . (16.24)
| {z } (N − 1)!k!
k

and the fully antisymmetric one is








 N!
k ⇔ T i1 ...ik = T [i1 ...ik ] ⇔ . (16.25)
· · ·
 (N − k)!k!



There is an obvious limit on the number of boxes one can pile in vertical,
determined by the fact that in C N we have at most N different indices. This
16.2 su(n) irrepses and young tableaux 171

case is unique and it is represented by the totally antisymmetric tensor, which


transforms trivially:

ei1 ...i N 7→ U i1 k1 . . . U i N k N ek1 ...k N = det U ei1 ...i N = ei1 ...i N . (16.26)

Using this same tensor we can also construct an object transforming in the anti-
fundamental representation, starting from N − 1 vectors in the fundamental
representation. In detail, we can define

wi∗1 ≡ ei1 ...i N vi2 . . . vi N , (16.27)

which transforms as

wi∗1 7→ U ∗ i1 k1 ek1 ...k N U ∗ i2 k2 . . . U ∗ i N k N U i2 l2 . . . U i N l N vl2 . . . vl N . (16.28)

Using U † U = 1 the right hand side is the same as

Ui∗1 k1 wk∗1 ≡ Ui∗1 k1 ek1 ...k N vk2 . . . vk N . (16.29)

We therefore see that N − 1 vertical boxes correspond to the N representation.


Other tensors will be related to mixed symmetry diagrams with k boxes,
whose permutations can be applied to rank k tensors. In detail, when consider-
ing such diagrams we construct irrepses by defining tensors whose symmetry
properties are determined by the projector

P = aT sT , (16.30)

where a T and s T are the antisymmetric and symmetric permutation operators


defined in Theorem 8.1. Obviously the result may depend on the order of the
indices, but it is easy to see that because of the symmetry properties of the
operator we acted with, we can always exchange indices in the same row and
remove diagrams where in one of its columns we have the same index repeated.
For instance, SU(2) tensors have indices with i = 1, 2 and applied to a tensor
with 2 indices has only 3 inequivalent combinations because

1 2 = 2 1. (16.31)

The other tableau with two boxes, applied to the same tensor gives the
remaining combination

1 =−2 , 1 = 0, 2 = 0. (16.32)
2 1 1 2
172 su(n) irreps and products

The number of inequivalent possibilities is then represented by standard tableaux.


A standard tableau is a tableau where the index numbers do not decrease when
going from left to right in rows and always increase from top to bottom in
columns. Non-standard tableaux give tensors that, by symmetrization or anti-
symmetrization either vanish or are not independent of the standard tableaux.
By using some easy combinatorics, we can argue that the dimension of each
of the irrepses can be computed using the formula that gives the number of
standard tableaux associated to a given configuration:

N−i+j
D= ∏ h(i, j)
, (16.33)
boxes i,j

where i labels the row, j labels the column and h(i, j) is the corresponding hook
number.

Example 16.1. The SU(4) Young Tableau has dimension D = 140. For
N = 4 the numerator is given by the product of the coefficients in

4 5 6 7
3 4 5 (16.34)
2 3

and the denominator is given by the products of the coefficients in

6 5 3 1
4 3 1 . (16.35)
2 1

Example 16.2. We apply this decomposition technique to rank 3 tensors. A


generic rank 3 tensor has 3 indices without any definite symmetry property.
Once we have a basis for the vector space VN where the fundamental represen-
tation acts, we can construct a basis of the vector space where rank 3 tensors
are defined, namely V⊗3 N = VN ⊗ VN ⊗ VN , by taking their products, so that

T = T ijk ei ⊗ e j ⊗ ek . (16.36)

In order to split T into its irreducible fragments we build the projector operators
associated to the different representations of S3 . There are a total of 6 possible
16.2 su(n) irrepses and young tableaux 173

states and 4 different projectors: P , P , P1 2 and P 1 3 , associated to the


3 2

different standard tableaux. Explicitly:


P = 1 + (12) + (13) + (23) + (123) + (132), (16.37)

P = 1 − (12) − (13) − (23) + (123) + (132), (16.38)

P1 2 = 1 + (12) − (13) − (123), (16.39)


3

P1 3 = 1 + (13) − (12) − (132). (16.40)


2

Their action on the basis of the vector space where rank 3 tensors are defined,
splits V⊗3 N into invariant subspaces whose basis elements are linear combina-
tions of the original ones (we omit the tensor product symbol ⊗ for the sake of
having more compact expressions):
1 
Sijk = ei e j e k + ei e k e j + e j e k ei + e j ei e k + e k ei e j + e k e j ei ,
6
1 
Aijk = ei e j e k − ei e k e j + e j e k ei − e j ei e k + e k ei e j − e k e j ei ,
6
1 
M1,1 ijk = ei e j e k + e j ei e k − e k ei e j − e k e j ei , (16.41)
2
1 
M1,2 ijk = √ ei e j e k − 2 ei e k e j − 2 e j e k ei + e j ei e k + e k ei e j + e k e j ei ,
2 3

1 
M2,1 ijk = e e e − e j ei e k − e j e k ei + e k e j ei ,
2 i j k
1 
M2,2 ijk = √ ei e j e k + 2 ei e k e j − 2 e j e k ei − e j ei e k + e k ei e j + e k e j ei .
2 3
The labels S, A and M stand for symmetric, antisymmetric and mixed sym-
metry, which are the symmetry properties of the corresponding basis. The
projectors in (16.37)–(16.40) select these invariant spaces: P projects into S,
P projects into the 1-dimensional space spanned by A and finally, P 1 2 and
3

P1 3 project into the 2-dimensional spaces spanned by M1 and M2 , respectively.


2
174 su(n) irreps and products

A generic tensor can be decomposed by considering all the fragments that re-
main when contracting its T ijk components with the various basis elements.
For instance, the fully symmetric and fully antisymmetric combinations are
1  ijk 
T ijk Sijk = T + T ikj + T kij + T jik + T jki + T kji ei ⊗ e j ⊗ ek ,
6
1  ijk 
T ijk Aijk = T − T ikj + T kij − T jik + T jki − T kji ei ⊗ e j ⊗ ek , (16.42)
6
while the mixed symmetry combinations give
1  ijk 
T ijk M1,1 ijk = T + T jik − T jki − T kji ei ⊗ e j ⊗ ek ,
2
1  
T ijk M1,2 ijk = √ T ijk − 2T ikj − 2T kij + T jik + T jki + T kji ei ⊗ e j ⊗ ek ,
2 3

1  ijk 
T ijk M2,1 ijk = T − T jik − T kij + T kji ei ⊗ e j ⊗ ek , (16.43)
2
1  
T ijk M2,2 ijk = √ T ijk + 2T ikj − 2T kij − T jik + T jki + T kji ei ⊗ e j ⊗ ek .
2 3
One thing, however, that needs to be stressed here is that while M1 and M2
belong to different invariant subspaces of the S3 basis space, they give tensors
associated to the same representation of the permutation group on their indices
and therefore they provide equivalent representations of SU(N). In fact they all
fulfill the same conditions
ijk jki kij
TMa + TMa + TMa = 0, a = 1, 2, (16.44)

which remove the fully symmetric TS and fully antisymmetric TA representa-


tions, and have the symmetry property
ijk kji ijk jik
TM1 = − TM1 , TM2 = − TM2 . (16.45)

Clearly the fact that the antisymmetry is on the first two indices or in the first
and third, or, for that matter, in the second and third, gives obviously the same
SU(N) representation.
Example 16.3. The same kind of decomposition is also useful to analyze invari-
ant subspaces of the Hilbert space of k identical particles transforming under
16.2 su(n) irrepses and young tableaux 175

the action of the fundamental representation of the SU(N) group. The Hilbert
space for each of these particles is H N = C N and the total Hilbert space is
H = (H N )k , which we now decompose using Sk irrepses.
For the sake of simplicity we focus on the decomposition of the 8-dimensional
Hilbert space of k = 3 identical particles, transforming as a doublet of SU(2),
i.e. N = 2. We denote the basis of the 2-dimensional space on which the fun-
damental representation of SU(2) acts by |+i and |−i. The generic element of
the 8-dimensional space H = H2 ⊗ H2 ⊗ H2 is then a linear combination of the
states labelled by | ± ±±i = |±i ⊗ |±i ⊗ |±i, where each of the signs refers
to the corresponding sign in the choice of the basis elements for each of the
2-dimensional Hilbert spaces of the 3 identical particles.
Since k = 3, the SU(2) irrepses acting on H are in correspondence with the

Young Tableaux with 3 boxes: , and . According to (16.33), their


dimensions are 4, 0 and 2, respectively. The decomposition of H into invariant
subspaces follows now from the application of the projectors associated to these
Young Tableaux as in the previous example.
The fully symmetric tensor is unique in S3 (dS3 = 1) and the application of
this operator to H selects 4 different invariant states, classified according to
their total spin:
1
| + ++i, √ (| + +−i + | + −+i + | − ++i) ,
3
(16.46)
1
√ (| + −−i + | − +−i + | − −+i) , | − −−i.
3
The totally antisymmetric tensor has no states.
The mixed tensor has dS3 = 2 (there are 2 independent permutations) and
has dimension 2, because we have 2 different states in each invariant space.
The corresponding states in H are (again according to their total spin)
1
√ (| + +−i − | − ++i) , (16.47)
2

1
(| + +−i + | − ++i − 2| + −+i) , (16.48)
2
1
√ (| − −+i − | + −−i) , (16.49)
2

1
(| − −+i + | + −−i − 2| − +−i) . (16.50)
2
176 su(n) irreps and products

Note that the states with the same spin transform in one of the irreducible
representations ψ Ma of S3 . In conclusion dim H = 8 = d ·D +d ·
D = 1 · 4 + 2 · 2.

As a general rule one can prove that

Theorem 16.1. SU(N) irrepses are in 1 to 1 correspondence with Young Tableaux


with a number of rows lower than N.

The dimension of the irreps is determined by the formula (16.33) and the
number of times it appears is determined by the dimension as a representation
of the corresponding Symmetric group.
This classifies completely irreducible representations. We can now move on
and analyze how to explicitly compute products of representations for the same
group.

16.2.3 Products using Young Tableaux

The analysis given in the previous section for the construction of the irreducible
representations of SU(N) suggests that Young Tableaux should be used effec-
tively to compute products of representations. In this case we do not discuss
the derivation of the rules that determine how to construct this product and
obtain the Clebsch–Gordan decomposition, but rather explain the result and
provide some examples that can convince us of the effectiveness of the adopted
technique.

Theorem 16.2. The Clebsch–Gordan decomposition for products of representations in


SU(N) can be represented by means of Young Tableaux using the following rule:

1. Label the second tableau in the product with the letters a, b, c, . . . in the 1st, 2nd,
3rd, etc. row:

a a a a
b b b (16.51)
c c c
...

2. Attach all the boxes with an a to the first tableau in all possible different ways, but
such that there are never two a in the same column and such that the resulting
diagram is still a legitimate Young tableau.
16.2 su(n) irrepses and young tableaux 177

3. Repeat the procedure for the boxes with b, c etc...


4. Read all tableaux in the outcome from right to left and from top to bottom and
delete all those associated to strings that have more b than a to the left of any
letter or more c than b etc.
Example 16.4. In SU(2) the fundamental and antifundamental representations
coincide
2∼2⇔ . (16.52)
The product gives

⊗ = ⊕ ⇔ 2 × 2 = 1 + 3. (16.53)

Since corresponds to the trivial representation, the only irreducible represen-


tations that can be obtained are those with k boxes in a row, whose dimension
is k + 1. This is easily seen in the product of two 3-dimensional representations:

⊗ a a = a + + · , (16.54)
a a + = +
a a a
which can be represented by
3 × 3 = 5 + 3 + 1. (16.55)
Example 16.5. In SU(3) the fundamental representation 3 and the anti-fundamental
representation 3 are represented by two different tableaux:

3⇔ and 3⇔ . (16.56)

Using the first 3 rules in 16.2, their product gives


 
⊗ a = a + b = a b +
a + b +
a . (16.57)
b a b a
b
Using the fourth rule in 16.2, the 4 resulting tableaux correspond to the strings
ba, ab, ba and ab respectively. This means that the first and third tableau have
to be discarded and we are left with

⊗ = + ·, (16.58)
178 su(n) irreps and products

which corresponds to

3 × 3 = 8 + 1, (16.59)

in agreement with the result discussed in (16.16).


We can also compute the product of two fundamentals, which is the same
in terms of Young Tableaux as the one for SU(2), but corresponds to different
dimensionality of the corresponding irreducible representations:

⊗ = ⊕ ⇔ 3 ⊗ 3 = 6 ⊕ 3. (16.60)

Another interesting lesson we can take home from these examples is that
Young Tableaux corresponding to conjugate irrepses have complementary shapes.
This means that complex conjugation of a given representation is represented
by the complementary tableau, rotated by 180 degrees.

8= = = 8̄ = 8

6= = = 6̄

Figure 24.: Conjugate representations in SU(3). The adjoint representation 8 is real and
its conjugate Young Tableaux has the same shape as the original one. On
the other hand, the representation 6 is complex and its complex conjugate
is represented by a different tableau.

16.2.4 Again on SU(2)

We conclude this lecture, by reminding that you already encountered the Clebsch–
Gordan decomposition for SU(2) in physical problems related to spin and an-
gular momentum. In that case it is also useful to it may be useful to obtain
the exact multiplicative coefficients in the decomposition and this was done as
follows.
16.2 su(n) irrepses and young tableaux 179

Let D j1 and D j2 two irrepses of SU(2)

~~ ~
 
D j1 ( g) ⊗ D j2 ( g) = eψ· J(1) ⊗ evecψ· J(2) = 1 + ψ
~ ~J(1) ⊗ 1 + 1 ⊗ ~J(2) + . . . (16.61)

We can define the product representation of the algebra generators as

~J = ~J(1) ⊗ 1 + 1 ⊗ ~J(2) . (16.62)

The spaces on which the irrepses are defined, namely V1 and V2 , admit a or-
thonormal basis of eigenstates of J(31) and J(32) :

{| j1 , m1 i} ∈ V1 , {| j2 , m2 i} ∈ V2 . (16.63)

Obviously J3 is diagonal in this basis:

D ( J3 )(| j1 , m1 i ⊗ | j2 , m2 i) = (m1 + m2 )(| j1 , m1 i ⊗ | j2 , m2 i). (16.64)

We can then discuss J± . Now the states are mixed by the operators and, starting
from the highest spin state, whose spin is j1 + j2 , one generates 2 states with
spin j1 + j2 − 1:

| j1 , j1 i ⊗ | j2 , j2 − 1i (16.65)

and

| j1 , j1 − 1i ⊗ | j2 , j2 i. (16.66)

This implies that we have a non-trivial degeneracy in the interval [−| j1 −


j2 |, | j1 − j2 |]. In particular we have 2 min( j1 , j2 ) + 1 states.
The natural basis of the product is given in terms of the basis vectors by
unitary matrices whose coefficients are Clebsch–Gordan coefficients:

| j, mi = ∑ Cj,m,j1 ,m1 ,j2 ,m2 | j1 , m1 i ⊗ | j2 , m2 i. (16.67)


j1 ,m1 ,j2 ,m2

Example 16.6. We show that

D1 ⊗ D1/2 = D3/2 ⊕ D1/2 . (16.68)

The spin 3/2 state is unique

| 32 , 32 i = |1, 1i ⊕ | 12 , 12 i. (16.69)
180 su(n) irreps and products

We apply J− to this:
√ √
D ( J− )| 23 , 32 i = 3| 32 , 21 i = 2(|1, 0i ⊗ | 12 , 21 i) + |1, 1i ⊗ | 12 , − 21 i. (16.70)

We deduce that
r
2 1
| 32 , 12 i = (|1, 0i ⊗ | 12 , 21 i) + √ (|1, 1i ⊗ | 12 , − 21 i). (16.71)
3 3

Applying once more D ( J− ) we get


r
1 2
| 32 , − 12 i = √ (|1, −1i ⊗ | 12 , 12 i) + (|1, 0i ⊗ | 12 , − 12 i), (16.72)
3 3

| 32 , − 32 i = |1, −1i ⊗ | 12 , − 21 i. (16.73)

In addition, to complete the expected degeneracy, we should also consider the


states orthogonal to | 32 , mi:
r
1 1 1 1 1 2
| 2 , 2 i = √ (|1, 0i ⊗ | 2 , 2 i) − (|1, 1i ⊗ | 12 , − 12 i), (16.74)
3 3
r
1 1 2 1
|2, −2i = (|1, −1i ⊗ | 12 , 12 i) − √ (|1, 0i ⊗ | 12 , − 12 i), (16.75)
3 3

where obviously | 21 , − 21 i = D ( J− )| 12 , 12 i.

exercises

1. Product of two fundamental representations of SO(4).


a) The first step requires to build the tensor Mij = V i ⊗ W j where i, j =
1, 2, 3, 4. If V and W transform in the fundamental representation
V i 7→ Ri j V j and W i 7→ Ri j W j , we have that Mij 7→ Ri k R j l Mkl . Build
the corresponding matrix representation for the product (this is a 16
× 16 matrix).
b) We now show that this representation is not irreducible.
Show that the subspaces generated by symmetric matrices Sij = S ji
and by antisymmetric matrices Aij = − A ji are invariant subspaces
of dimensions 10 and 6.
16.2 su(n) irrepses and young tableaux 181

c) Show that the space of symmetric tensors can be further reduced


into the sum of two irreducible representations of dimensions 9 and
1.
d) Show that also the space of antisymmetric tensors can be further
reduced as follows:
ij 1
A+ = 2 eijkl Akl
+;
ij
A− = − 12 eijkl Akl
−.

2. We can repeat the analysis of the previous problem by using the isomor-
phism so(4) ' su(2) L ⊕ su(2) R established in an exercise of previous
chapter. (We use here L, R to distinguish the two copies of su(2) in corre-
spondence with Li+ and Li− ). Use the symbol (m, n) = Dm ( g L ) ⊗ Dn ( gR )
for the representation of the elements ( g L , gR ) ∈ SU(2) L × SU(2) R by
means of matrices of dimensions m and n.
Show that the fundamental representation of SO(4) is a couple of dou-
blets of SU(2) and compute their products using the properties of SU(2)
products.
(1)
a) Compute Li on (2, 2) and note that it contains a triplet and a scalar
(the fourth component).
b) Compute (2, 2) ⊗ (2, 2) and decompose the result in its irreducible
parts.
c) Relate this result to the one obtained in the previous exercise.

3. Describe the decomposition of (H2 )5 in irreducible representations of S5


and SU(2).

4. Describe the decomposition of (H3 )4 in irreducible representations of S4


and SU(3).

5. Using Young Tableaux describe the 6 representations of SU(4) of dimen-


sions lower or equal to 10 and consider their products.
17
T H E P O I N C A R É G R O U P

We already discussed some direct applications of group theory to Physics,


mainly focusing on Quantum Mechanics. So far, however, we concentrated
on simple instances where the only spacetime symmetry was related to the
group of rotations or to its (discrete) subgroups. On the other hand, we are ob-
viously interested in descriptions of systems where one of the basic principles
of modern Physics is taken into account: the principle of Relativity. By this
principle, observers in different inertial frames must have equivalent physical
descriptions of the same phenomena.
Inertial frames measure the same spacetime interval between any two events

ds2 = −(dx0 )2 + (dx1 )2 + (dx2 )2 + (dx3 )2 = dx µ ηµν dx ν , (17.1)


where
−1
 
 1 
η=  (17.2)
 1 
1
is the Minkowski metric. Coordinate transformations relating different inertial
frames should then preserve ds2 and have the form
x µ0 = Λµ ν x ν + aµ , where a ∈ R4 , Λ T ηΛ = η. (17.3)
The point of interest for us is that coordinate changes of the form (17.3) form
a group: the Poincaré group. This fact tells us that all relevant physical quan-
tities in a relativistic theory must transform in definite representations of this
group and that therefore we can classify quantum relativistic states by quantum
numbers that specify irreducible representations of the Poincaré group.

183
184 poincar é group

Definition 17.1. The Poincaré group is the set


n o
(Λ, a) ∈ Mat(4, R) × R4 | Λ T ηΛ = η

with composition law

( Λ2 , a2 ) ◦ ( Λ1 , a1 ) = ( Λ2 Λ1 , Λ2 a1 + a2 ),

where matrix product and the action of matrices on vectors is understood.

We leave to the reader the check that this composition law satisfies all group
axioms.
The Poincaré group contains several interesting subgroups. The first and
most interesting one is the subset of elements of the form (Λ, 0), which is called
the Lorentz group. Other interesting subgroups are the group of translations,
whose matrices have the special form (14 , a), and the group of rotations, whose
matrices are of the form ( R, 0), with R0 0 = 1, R0 i = Ri 0 = 0 and Ri j Rk j = δik .
Any rotation matrix can actually be parameterized by the unit norm vector n̂
fixing the axis of rotation and by the rotation angle θ:

Ri j = δji cos θ + ei jk n̂k sin θ + n̂i n̂ j (1 − cos θ ). (17.4)

17.1 topological properties

Before proceeding with the analysis of the representations of the Poincaré


group and of its Lorentz subgroup, we analyze its topological properties. We
actually focus on the Lorentz subgroup first, because some interesting facts are
already clear when we analyze elements of this subgroup.
As explained above, the Lorentz group is specified by 4 by 4 matrices Λ that
preserve the mixed signature metric η. These matrices provide the fundamental
representation of the Lorentz group, which is trivially isomorphic to

O(3, 1) = {Λ ∈ Mat(4, R) | Λ T ηΛ = η },

which describes “rotations” in a 4-dimensional non-Euclidean space with 3


isotropic dimensions. Just like any O(N) group has two disconnected compo-
nents, characterized by the determinant of the matrices representing its ele-
ments, also O(3,1) has two disconnected components. In detail:

−1 = det η = det(Λ T ηΛ) = det η (det Λ)2 (17.5)


17.1 topological properties 185

implies

det Λ = ±1. (17.6)

The full Lorentz group has therefore at least 2 disconnected components, cor-
responding to the set of matrices having determinant equal to plus or minus
1.
The subset of matrices with unit determinant contains also the identity ma-
trix and actually forms a subgroup of the original one called

SO(3, 1) = {Λ ∈ O(3, 1) | det Λ = 1}. (17.7)

This is also analogous to the SO(N) < O(N) groups. However, we now prove
that SO(3,1) is not connected.
If we focus on the constraint that characterizes the matrices of the group
SO(3,1) we see that the 00 component gives
 2  2
−1 = η00 = Λµ 0 Λν 0 ηµν = − Λ0 0 + ∑ Λi 0 , (17.8)
i

which implies
 2  2
Λ0 0 = 1 + ∑ Λi 0 ≥ 1. (17.9)
i

Using this inequality we can separate elements of SO(3,1) in two disconnected


components. One component, connected to the identity, is represented by ma-
trices with

Λ0 0 ≥ 1, (17.10)

while the other component has

Λ0 0 ≤ −1. (17.11)

Obviously we cannot change continuously the value of Λ0 0 to go from one


component to the other.
The same conclusion can also be drawn for matrices with det Λ = −1. We
then see that the Lorentz group has 4 disconnected components, specified by
the value of det Λ and by the sign of Λ0 0 for the matrices Λ providing its fun-
damental representation. We distinguish these sets by calling proper all trans-
formations with positive determinant and orthochronous all transformations
preserving the direction of time.
186 poincar é group

Definition 17.2. The set of matrices Λ ∈ O(3, 1) with det Λ = 1 and Λ0 0 ≥ 1


forms a subgroup called the proper and orthochronous Lorentz group:

SO+ (3, 1).

The other components are subsets but not subgroups. One can cover them by
taking the product of elements in SO+ (3, 1) with appropriate elements defining
a discrete subgroup of SO(3,1). In detail, the subset with det Λ = −1 and
Λ0 0 ≥ 1 contains the space parity operator
 
1
 −1 
P= , (17.12)
 −1 
−1

which satisfies P 2 = 1, while the subset with det Λ = −1 and Λ0 0 ≤ 1 contains


the temporal inversion operator

−1
 
 1 
T = , (17.13)
 1 
1

which also satisfies T 2 = 1. Finally, the subset with det Λ = 1 and Λ0 0 ≤ −1


contains the combination of these two discrete operators: −14 = P T = T P .
We therefore see that a generic O(3,1) transformation is the combination of a
SO+ (3,1) transformation with those of the Z2 × Z2 group generated by P and
T.
There are many more considerations that are interesting from a physical
point of view, like the fact that boosts in the same direction form a subgroup
while boosts in different directions close only up to rotations, that physics is
invariant only under SO+ (3,1) etc., but these are for the courses introducing
special relativity.
In the following we focus on the component connected to the identity and
we now prove that it is homomorphic to the simply connected group SL(2,C).
We will then discuss the physical implications of this fact.
Theorem 17.1. There is a 2 to 1 homomorphism between SL(2, C) and SO+ (3,1).
Proof. The fundamental representation of SO+ (3,1) acts linearly on spacetime
coordinates: x 7→ Λx. In order to obtain an action of SL(2,C), we think of
17.1 topological properties 187

the coordinates x µ as parameters specifying a generic 2 by 2 hermitian matrix


A† ( x ) = A( x ). A generic 2 by 2 hermitian matrix is specified by 4 real parame-
ters, which determine its expansion in terms of the basis1

σµ = (12 , σi ). (17.14)

In fact, σµ† = σµ and any 2 by 2 hermitian matrix can be written as A( x ) = x µ σµ ,


where x µ are 4 real parameters. We can actually extract the x µ parameters from
A( x ) by taking
1
xµ = tr ( A( x )σ̂µ ) , (17.15)
2
where σ̂µ = (12 , σi ), and tr (σµ σ̂ν ) = δµν . Note that

x0 + x3 x1 − i x2
 
det A( x ) = det = − x µ ηµν x ν . (17.16)
x1 + i x2 x0 − x3

We can then map hermitian matrices into themselves by


e = SAS† ,
A 7→ A (17.17)

so that Ae† = A,
e which means that A e = xeµ σµ , with xeµ = ΛS µ ν x ν . This in turn
implies that we have a map from SL(2, C) to SO+ (3, 1). We can see that the
map is linear because
 
xeµ = 12 tr A e( xe)σ̂µ = 1 tr SA( x )S† σ̂µ

2
(17.18)
= 21 tr Sσν S† σ̂µ x ν = ΛS µ ν x ν .


If, in addition, we require that S ∈ SL(2, C) (and hence that det S = 1), then
e = det (SAS† ) = det S(det S)∗ det A = − x µ x ν ηµν . (17.19)
− xeµ xeν ηµν = det A

This implies that

x µ x ν ηµν = (ΛS x )µ (ΛS x )ν ηµν (17.20)

and hence that

ΛS ∈ O(3, 1). (17.21)


1 If you are interested in a proof that (17.14) provides an orthonormal basis, check the exercises at
the end of this chapter.
188 poincar é group

In addition, SL(2, C) is connected, and the same will be true by the image of
the continuous map given by ΛS . We therefore conclude that

S 7→ ΛS (17.22)

maps elements of SL(2, C) into elements of SO+ (3,1). Finally, since S and −S
are mapped to the same element ΛS we see that the map is 2 to 1.
The fact that such map is a homomorphism follows by the proof that

S 1 S 2 7 → Λ S1 S2 = Λ S1 Λ S2 . (17.23)

This is tedious, but easy to obtain by using the explicit expression for ΛS :

Λ S1 S2 µ ν 1
tr S1 S2 σν S2† S1† σ̂µ = 21 tr S2 σν S2† S1† σ̂µ S1 =
 
= 2
(17.24)
1 † b S† σ̂ µ S
  a
= 2 S2 σν S2 a 1 1 b ,

where we used the cyclic property of the trace, and using that (see the exercises
for a proof of this identity)

(σρ ) a b (σ̂ρ )c d = 2 δad δcb . (17.25)

Inserting this identity in the expression above:

Λ S1 S2 µ ν 1
S2 σν S2† c (2 δad δcb ) S1† σ̂µ S1 a
 
= 4 d b
1 c ( σ̂ ρ ) d ( σ ) b
S2 σν S2† S1† σ̂µ S1 b a
 
= 4 d c ρ a (17.26)

1 = Λ S1 ρ Λ S2 ν .
1 † ρ 1 tr σ S† σ̂ µ S
  µ ρ
= 2 tr S2 σν S2 σ̂ 2 ρ 1

From this theorem we also have a hint of the fact that SL(2,C) provides the
universal covering of SO+ (3,1) and that

SO+ (3, 1) ' SL(2, C)/Z2 . (17.27)

17.1.1 Topology of SL(2,C) and SO+ (3,1)

Given that SL(2,C) and SO+ (3, 1) are homomorphic, they are also related as
topological spaces.
17.2 the so+ (3, 1) algebra 189

We can actually determine their topologies by studying them as spaces of


matrices. SL(2,C) is the space of 2 by 2 complex matrices with unit determinant
and SO+ (3,1) is the same space, but where couples of matrices are identified
by means of the map (17.22). A generic complex matrix S can be written as the
product of a unitary matrix and the exponential of a hermitian one

S = u eh , u† u = 1, h† = h. (17.28)

Since we are interested in SL(2,C) matrices, we compute the determinant of


(17.28), which is

det S = (det u)(exp(tr h)), (17.29)

where det u = eiα ∈ U(1) and exp(tr h) ≥ 0. The request of having a unit
determinant for S forces

det S = 1 ⇔ det u = 1, tr h = 0. (17.30)

Thus any SL(2,C) matrix can be written as the product of a matrix u ∈ SU(2)
' S3 and the exponential of a traceless, hermitian, 2 by 2 matrix h, which is
parameterized by three real numbers. We conclude that topologically

SL(2, C) ' R3 × S3 . (17.31)

Also, since the 2 to 1 map (17.22) identifies matrices S with opposite signs and
only u can change the overall sign, we argue that

SO+ (3, 1) ' R3 × S3 /Z2 . (17.32)

17.2 the so+ (3, 1) algebra

As is often the case, also for the SO+ (3, 1) group, it is convenient to determine
representations of the algebra rather than those of the group. For this reason
we now discuss the structure of the algebra generators and their fundamental
representation. The group SO+ (3, 1) can be thought of as a group of matrices.
We can then follow the same derivation of the algebra as in section 12.2, by
expanding matrices Λ ∈ SO(3,1) close to the identity. In detail, we can linearize
the condition Λ T ηΛ = η by choosing Λµ ν = δν + ω µ ν , assuming a small ω:
µ

= Λρ µ Λσ ν ηρσ = (δµ + ω ρ µ ) (δνσ + ω σ ν ) ηρσ


ρ
ηµν
(17.33)
= ηµν + ηµσ ω σ ν + ηνσ ω σ µ + . . .
190 poincar é group

The so(3, 1) algebra is then determined by the set of matrices ω fulfilling the
condition above, namely
n o
so(3, 1) = ω ∈ Mat(4, R) | ηω = −(ηω ) T . (17.34)

This also tells us immediately that

dim so(3, 1) = 6, (17.35)

because this is the number of independent conditions in a 4 by 4 antisymmetric


matrix (which is the set of ηω matrices, otherwise, if one considers ω, we would
have 3 symmetric and 3 antisymmetric matrices).
A basis for its fundamental representation is then given by
   
0 0
 0   0 1 
D ( l1 ) =   , D ( l2 ) =  ,
 0 −1   0 
1 0 −1 0
  (17.36)
0
 0 −1 
D ( l3 ) =  ,
 1 0 
0

which represent generators of so(3) ⊂ so(3, 1), and by


   
0 1 0 1
 1 0   0 
D (k1 ) =   , D (k2 ) =  ,
 0   1 0 
0 0
  (17.37)
0 1
 0 
D (k3 ) =  ,
 0 
1 0

which make the algebra non-compact and whose exponential generates boosts:

 
cosh ψ sinh ψ
 sinh ψ cosh ψ 
exp(ψD (k1 )) =  , (17.38)
 1 
1
17.2 the so+ (3, 1) algebra 191

where ψ is the rapidity tanh ψ = − β. Note that since the group SO+ (3,1) is
non-compact, we cannot generate all its elements by exponentiation. However,
being SO+ (3,1) connected, we can always cover the whole underlying manifold
by taking products of boosts and rotations:
   
Λ = exp ψ ~ · D (~k) exp ~θ · D (~l ) . (17.39)

A description of the same algebra, independent on the representation, is


given by the commutators. Let us define
J0i ≡ k i , Jij ≡ eijk lk , (17.40)
whose fundamental representation, provided in (17.36)–(17.37), is given by
β β
D ( Jµν )α β = −ηµα δν + ηνα δµ . (17.41)
From the commutators of the faithful representation given above, we can now
define the algebra so(3, 1) by means of its structure constants, as follows from
the brackets
[ Jµν , J ρσ ] = 4 δ[µ [ρ Jν] σ] , (17.42)
where indices have been raised by means of Minkowski metric η and square
brackets on the indices denote antisymmetrizations with weight one. Alterna-
tively
[li , l j ] = eijk lk , [li , k j ] = eijk k k , [k i , k j ] = −eijk lk . (17.43)
As for the group of rotations, it is going to be extremely useful to complexify
this algebra in order to study its representations.
Theorem 17.2. so(3, 1)C ' sl(2, C) ⊕ sl(2, C).
Proof. Define
1
li± ≡ ( l + i k i ). (17.44)
2 i
These fulfill
[li± , l j± ] = eijk lk± , [li+ , l j− ] = 0, (17.45)

which show that we have two copies of su(2)C ' sl(2, C).
This simple proof shows that also in this case we can build representations
of so(3, 1) by means of representations of sl(2, C), just like we did for su(2) '
so(3).
192 poincar é group

exercises

1. Check that (17.4) satisfies R T ηR = η.

2. Show that the space of 2 by 2 hermitean matrices is spanned by real linear


combinations of the σµ , µ = 0, 1, 2, 3, matrices, where σ0 = 12 and σi are
the Pauli matrices. For this purpose we introduce the inner product

h A| Bi = Tr ( A† B).

a) Show that hσµ |σν i = N δµν , for some N ∈ R.


b) We can therefore think of σµ as orthogonal vectors in R4 , whose
components are (σµ )ν . Show that

∑(σρ )∗µ (σρ )ν = N δµν ,


ρ

which leads to
∑(σµ )∗ α β (σµ )γ δ = N δαγ δβδ
µ

and finally to

∑(σi )α β (σi )γ δ = −δα δγδ + 2 δαδ δγ .


β β

i
18
I R R E D U C I B L E R E P R E S E N TAT I O N S O F T H E P O I N C A R É
GROUP

Now that we have determined the structure of the Poincaré and Lorentz groups
and of their algebras we can discuss their representations. As we will see in
the following, we will be interested in two different kinds of representations
for these groups, because we will need to employ different strategies in the
description of states and operators of a relativistic quantum mechanical theory.
Let me anticipate that we will need unitary infinite-dimensional representa-
tions to describe physical states, while we will make use of non-unitary, but
finite-dimensional representations for operators.
We restrict our discussion to elements of the Poincaré group that are con-
nected with the identity, i.e.

ISO+ (3, 1) = SO+ (3, 1) n R3,1 , (18.1)

where ISO stands for inhomogeneous special orthogonal transformations. We


know how the corresponding algebra can be represented in terms of differential
operators. Using anti-hermitian operators:

D ( Jµν ) = −ηµρ x ρ ∂ν + ηνρ x ρ ∂µ , (18.2)

D ( Pµ ) = ∂µ . (18.3)

The corresponding commutators are those of the Lorentz algebra (17.42) for Jµν
and

[ Pµ , Pν ] = 0, [ Jµν , Pρ ] = 2 ηρ[µ Pν] , (18.4)

193
194 poincar é irrepses

for the others. The last commutator tells us that translation generators trans-
form in the fundamental representation of the Lorentz rotations (and in fact
they carry a vector index µ):

[ Jµν , Pρ ] = Pσ D ( Jµν )σ ρ , (18.5)

with D ( Jµν )ρ σ = −ηµρ δνσ + ηνρ δµσ , as in (17.42).


Starting from the basis of generators of the Poincaré algebra we can con-
struct by exponentiation a representation for the Poincaré group acting on an
appropriate vector space, which we want to be the Hilbert space of a quantum
relativistic theory. In particular, if we want the Poincaré group to be the set of
spacetime symmetries that leave physics invariant, we are interested in a rep-
resentation of its elements by unitary operators. However, as noted before, the
Poincaré group is not compact and therefore unitary irreducible representations
are infinite dimensional. This implies that the vector space on which they act is
infinite dimensional and a basis of states for this space has labels that vary over
an infinite number of values. Although this may sound scary, it is a rather rea-
sonable fact that one should expect, given that among the quantum numbers
characterizing a generic physical state there is the momentum pµ , which has
continuous values.
However, these are not going to be the only representations of the Poincaré
group we are interested in. Today we know that quantum relativistic theories
are field theories. Fields (continuous functions of spacetime coordinates x µ
at the classical level) contain creation and destruction operators for quantum
states. In order to be able to work with them and in order to have a sensible
classical limit it is therefore useful to introduce also finite-dimensional representa-
tions of the Poincaré group, which are necessarily non unitary. For instance, the
electromagnetic field is described by the vector potential Aµ ( x ). Under Lorentz
transformations Aµ 7→ Λµ ν Aν and Λ is generically not unitary, as can be seen
immediately from the matrix representation given in the previous chapter.
Summarizing, we will use unitary representations for the Poincaré group ac-
tion on the Hilbert space of physical states of quantum field theories and we
will use finite-dimensional representations for the operators acting on the same
Hilbert space. In this lecture we therefore first analyze in detail the construction
of unitary representations and then discuss finite-dimensional ones. We con-
clude by discussing the role of Casimir operators to determine the appropriate
quantum numbers defining these irreducible representations.
18.1 unitary representations 195

18.1 unitary representations

For the sake of simplicity, we are going to consider single particle states, trans-
forming under the action of irreducible representations of the Poincaré group.
The classification of these representations has been worked out first by Wigner,
but one can look for a more detailed discussion in Weinberg’s book ”Quantum
theory of Fields: I”.
As mentioned in the introduction, an irreducible unitary representation of
the Poincaré group must be infinite dimensional. The states providing a ba-
sis for the vector space on which this representation acts form an infinite-
dimensional multiplet and transform onto each other when acted upon by
Poincaré generators. As is always the case, we can decide that a set of com-
muting operators are diagonal on this basis and start labeling the correspond-
ing states by using their eigenvalues. Recall that irreducible representations of
abelian groups are always 1-dimensional.
In the Poincaré algebra, Pµ operators commute among themselves and we
can therefore classify quantum states by using their eigenvalues

P µ | p µ , σ i = i p µ | p µ , σ i, (18.6)

where σ are other possible quantum numbers. As expected, their exponential


provides a unitary representation for translations

U (1, a) = exp( aµ Pµ ) (18.7)

and the action on this basis of states is also diagonal:


µ
U (1, a)| p, σ i = eiaµ p | p, σ i. (18.8)

In order to complete the basis of states of the vector space on which a unitary
representation of the Poincaré group acts, we need to provide a representation
for the proper Lorentz group elements. The corresponding operators U (Λ) ≡
U (Λ, 0) should map Pµ eigenstates into states with momentum Λp, because Pµ
transform in the fundamental representation of the Lorentz group and so must
transform its eigenstates. Actually, using Hadamard’s lemma 12.3 and (18.5),
we can see this explicitly by taking the action of
 
1
U (Λ) = exp ωµν J µν (18.9)
2
196 poincar é irrepses

on the momentum operators


 
− 12 ωµν J µν ρ 1 ωµν J µν 1
e P e 2 = P exp − ωµν D ( J µν )
σ ρ
, (18.10)
2 σ

which also implies that

U † ( Λ ) P ρ U ( Λ ) = U ( Λ −1 ) P ρ U ( Λ ) = P σ ( Λ −1 ) σ ρ = Λ ρ σ P σ , (18.11)

where we used that Λµ ρ ηµν Λν σ = ηρσ and hence (Λ−1 )ρ σ = ηρα (Λ−1 )α β η βσ =
Λσ ρ . This in turn implies that an eigenstate of Pµ with momentum pµ trans-
forms into an eigenstate with momentum Λµ ν pν , when acted upon by U (Λ):

Pµ (U (Λ)| p, σ i) = U (Λ)U † (Λ) Pµ U (Λ)| p, σi = U (Λ)Λµ ν Pν | p, σi


(18.12)
= i (Λµ ν pν ) U (Λ)| p, σi.

Does this imply that the only result of the action of a Lorentz transformation
on a state | p, σ i is to map it to

|Λp, σi = U (Λ)| p, σi? (18.13)

Actually, we can see that this cannot be correct, because we did not take into
account the possibility that there may different states whose momenta are in-
variant under the action of Λ and which therefore may mix between then when
acted upon by U (Λ). In order to solve this issue we introduce the so-called little
group and show that irreducible representations of Poincaré group are related
to irreducible representations of this group.

Definition 18.1. The little group of a vector v ∈ V (where V is the space on


which the representation D acts) is the subgroup Gv < G that leaves invariant
v:
IsoG (v) = { g ∈ G | D ( g)v = v}.

Since we are interested in the possible degeneracy between momentum eigen-


states acted upon by Lorentz transformations, we are interested in the little
groups of generic 4-vectors pµ with respect to G = SO+ (3, 1). We can prove
that irreducible representations of the Poincaré group must then be determined
in terms of irreducible representations of the little groups associated to physical
momenta pµ .
18.1 unitary representations 197

Theorem 18.1. Unitary irreducible representations of the Poincaré group are fixed by
irreducible representations of Iso (k) for fixed k.

Proof. Start from a fixed momentum k, whose little group Iso (k ) is determined
by elements Λ ∈ Iso (k) fulfilling

Λk = k. (18.14)

All eigenstates of the momentum operator with eigenvalue k are then mapped
onto each other by the action of Lorentz transformations

U (Λ)|k, σ i = Dσ0 σ (Λ)|k, σ0 i, (18.15)

where D provides an irreducible representation of Iso (k ), with σ = 1, . . . , dim D,


which we want to be unitary in order to preserve orthonormality of states

h p, σ| p0 σ0 i = δ3 (~p − ~p0 )δσσ0 . (18.16)

Also D may not be finite dimensional if Iso (k) is not compact. Fortunately this
will not be necessary if one is not interested in non-physical cases.
We can now determine the result of the action of a unitary Lorentz transfor-
mation on a generic state of momentum p. First we find a Lorentz transforma-
tion relating p to k

pµ = Λ p µ ν kν , (18.17)

which we assume does not mix states with the same momentum k:

| p, σi ≡ U (Λ p )|k, σi. (18.18)

Then we compute the action of a generic Lorentz transformation on | p, σi and


note that any such transformation can be rewritten as the product of a boost
mapping k to Λp (which again does not mix states with the same momentum
k) and a Lorentz transformation in the little group of k. In fact, if we compute
U (Λ)| p, σi, we see that we can rewrite it as

U (Λ)| p, σi = U (ΛΛ p )U † (ΛΛ p )U (Λ)| p, σi

= U (ΛΛ p )U † (ΛΛ p )U (Λ)U (Λ p )|k, σi (18.19)

= U ( Λ Λ p )U ( Λ − 1
Λ p ΛΛ p )| k, σ i,
198 poincar é irrepses

where

Λ− 1 −1
Λ p ΛΛ p k = ΛΛ p Λp = k (18.20)

and therefore

Λ− 1
Λ p ΛΛ p ∈ Iso ( k ). (18.21)

We can use this result to see that the generic unitary representation of a Lorentz
transformation acts on states | p, σ i mixing them according to the little group of
the associated momentum k

U (Λ)| p, σ i = U (ΛΛ p ) Dσ0 σ (Λ̃)|k, σ0 i = Dσ0 σ (Λ̃)|Λp, σ0 i (18.22)

where Λ̃ = Λ− 1
Λp ΛΛ p ∈ Iso ( k ). This provides a representation for Λ on each
state | p, σ i.

While the final expression may not be particularly elegant, we reduced the
problem of classifying representations of the Poincaré group to the problem of
classifying representations of the little groups associated to physical momenta,
which can be differentiated by the value of their invariant scalar combinations.
Since the physical interpretation of pµ does not allow for the case p0 < 0 and
since p2 > 0 implies tachyonic states, we can distinguish three physical orbits1 :
• kµ = 0. Iso (k) = SO(3, 1) and its orbit is given by a unique state. It has
the property that all generators of the Poincaré group are represented
by 1-dimensional null matrices and, consequently, all group elements are
represented by identity matrices. This state has momentum and angular
momentum equal to zero and is invariant under Poincaré transformations.
We interpret it as the vacuum.
• k2 < 0. There is a frame where kµ = (k, 0, 0, 0). In this case the little group
consists of space rotations Iso (k) = SO(3), whose covering group is SU(2).
These states describe massive particle and are classified by well known
spin representations D2j+1 and σ = ( j, j3 ). However, since [l3 , Pi ] = e3ij Pj ,
j3 should not be directly interpreted as the eigenvalue of the state with
respect to l3 , but rather with respect W3 , where W µ is the Pauli–Lubanski
operator we will introduce later, when discussing Casimir operators for
the Poincaré group.
1 Also, since we are interested in proper and orthochronous transformations, we should distinguish
p0 > 0 from p0 < 0. However, once more, physics tells us we should only concentrate on p0 > 0.
18.2 finite dimensional representations 199

• k2 = 0. There is a frame where k = (1, 1, 0, 0), or k = (1, −1, 0, 0). These


are massless particles. At the infinitesimal level the elements of so(3, 1)
that leave invariant k follow from Λk = (1 + ω )k = k and hence ωk = 0:
 
0 0 a b
 0 0 a b 
θ 1 l1 + a ( k 2 − l3 ) + b ( k 3 + l2 ) =  . (18.23)
 a −a 0 − θ1 
b −b θ1 0

The resulting group is ISO(2) ' E2 , the Euclidean group in 2 dimen-


sions, where l1 is the rotation generator and k2 − l3 and k3 + l2 commute
and transform onto each other by the action of exp(αl1 ). This is a non-
compact group and most of its irreducible representations are infinite
dimensional. The corresponding quantum numbers are, however, not ob-
served in nature. We conclude that physical states are classified only by
the representations of l1 ∈ u(1), which are labeled by an integer number,
called helicity.

18.2 finite dimensional representations

Since we are interested also in finite-dimensional non unitary representations


of the Lorentz group, because they are directly relevant to determine the fields
of quantum field theories, we now discuss how to find them in a constructive
manner. We proceed in the same way as for the su(2) algebra, by considering
first the complexification of the Lorentz algebra and then taking the appropriate
real section. We therefore first build representations for so(3, 1)C ' sl(2, C) ⊕
sl(2, C) and then consider the appropriate linear combinations. We know that
irrepses of each of the two sl(2, C) factors are characterized by a half-integer
label j ∈ N/2, called spin, and have dimension 2j + 1.

Theorem 18.2. Finite-dimensional representations of so(3, 1) are characterized by two


semi-integer numbers ( j1 , j2 ), with j1 , j2 ∈ N/2.

We therefore proceed by effectively generalizing the spinorial representations


of 3-dimensional rotations to 4 dimensions, taking into account the Lorentzian
signature. For 3 dimensional rotations we introduced Pauli matrices σi , i =
1, 2, 3, satisfying

{σi , σj } = 2 δij , (18.24)


200 poincar é irrepses

where δij is the Euclidean metric in 3 dimensions, and we saw that by means of
them we could provide an irreducible 2-dimensional spinorial representation
for the so(3)C algebra by using
 
1 1 i
DS ( Jij ) = − σij ≡ − [σi , σj ] = eijk − σk = eijk DS (lk ) (18.25)
2 4 2
and then, starting from that, build all the other representations.
For so(3, 1) we introduce 4 by 4 matrices γµ , µ = 0, 1, 2, 3, with analogous
anti-commutation properties:
Definition 18.2. The Clifford algebra Cl (3, 1) is the algebra generated by γµ ,
µ = 0, 1, 2, 3, satisfying
{γµ , γν } = 2 ηµν .
The minimal matrix representation one can give to γµ is 4 by 4 (see the ex-
ercises) and there are obviously various legitimate equivalent choices for their
expression. One of them is the following:

γ0 = i σ1 ⊗ 12 , γi = σ2 ⊗ σ i . (18.26)

We can now provide the spinorial representation of the Lorentz generators by


considering
 1
DS Jµν = − γµν , (18.27)
2
where
1
γµν ≡ [ γµ , γν ] . (18.28)
2
We leave to the reader the check of the commutation relations2 (17.42).

 18.3. The spinorial representation of the Lorentz group is provided by


Definition
DS Jµν = − 21 γµν .
States transforming under the action of such a representation of the Lorentz
group are called spinors.
The group representation is obtained by exponentiation:
      
1 1 1
DS exp ωµν J µν
= exp ωµν DS ( J ) = exp − ωµν γ
µν µν
. (18.29)
2 2 4
ρσ
2 It may be useful to realize that [γµν , γρσ ] = −8 δµν .
18.3 reducibility of spinor representations 201

From the discussion above, this is actually a representation of the covering


group of the Lorentz group
   
1
Spin(3, 1) = exp − ωµν γ µν
| ωµν = −ωνµ (18.30)
4

and again, just like between SU(2) and SO(3) we have a 2 to 1 map, we also
have a 2 to 1 map between Spin(3, 1) and SO+ (3, 1). We can actually identify

Spin(3, 1) ' SL(2, C). (18.31)

Differently from what happened for SU(2) we can now see that this representa-
tion is not irreducible, though.

18.3 reducibility of spinor representations

Starting from the 4 γ-matrices defining Cl (3, 1) we can always introduce a new
matrix3

γ5 ≡ i γ0 γ1 γ2 γ3 , (18.32)

satisfying

{γ5 , γµ } = 0 (18.33)

and

(γ5 )2 = 1. (18.34)

We can then introduce two projectors

1
P± ≡ (1 ± γ5 ) , (18.35)
2
which split the vector space on which the spinorial representation acts in two
complementary subspaces. This is actually straightforward to see using the
representation (18.26) for the γ-matrices, which gives

12 O2
   
P+ = , P− = . (18.36)
O2 12
3 The index 5 is not related to spacetime properties and therefore γ5 = γ5 .
202 poincar é irrepses

We now show that the subspaces im( P± ) are invariant with respect to the action
of the spinorial representation of the Lorentz group DS .
Take an element in any of the two subspaces: η± ∈ Im( P± ). By definition
P± η± = η± and P∓ η± = 0. Hence

γ5 η± = ± η± . (18.37)

If we apply a Lorentz transformation to this spinor, we get a new spinor


 
1
ηe± = exp − ωµν γ µν
η± , (18.38)
4

which is again in the same subspace because


   
1 1
γ5 ηe± = γ5 exp − ωµν γµν η± = exp − ωµν γµν γ5 η± = ±ηe± . (18.39)
4 4

We will call spinors in im( P+ ) spinors with positive chirality and those in
im( P− ) spinors with negative chirality.
Actually im( P± ) are vector spaces where irreducible representations of the
sl(2, C) ⊕ sl(2, C) algebra act. In fact we can write

1 1 1
ωµν J µν = ω0i J 0i + ωij J ij = −ω0i ki + eijk ωij lk = li+ ωi+ + li− ωi− , (18.40)
2 2 2
where
1 1 
li± ≡ ( l ± i k i ), ωi± ≡ eijk ω jk + 2i ω0i . (18.41)
2 i 2
In the Weyl basis (18.26)
   
ωi+ − 2i σi
1
− ωµν γµν =   . (18.42)
 
4

ωi− − 2i σi

We then see that Vs = V + ⊕ V − on which we have a split action of li+ and li− .
We also conclude that spinors with positive chirality are objects in the ( 21 , 0)
representation and those with negative chirality in the (0, 12 ). Alternatively,
using the dimensions of the irrepses, we denote them by (2, 1) and (1, 2).
18.3 reducibility of spinor representations 203

Finally, we can show that Spin(3, 1) ' SL(2, C) and that we have a 2 to 1
map between the Spin group and SO+ (3, 1). A generic element of Spin(3, 1) is
represented by4
   
1 S 0
exp − ωµν γµν = , (18.43)
4 0 ( S −1 ) †
where
 
i
S = exp − √ ωi+ σi (18.44)
2 2
 
and S ∈ SL(2, C), with (S−1 )† = exp − 2i ωi− σi , because (ωi+ )∗ = ωi− . As
for SU(2), we can build the adjoint representation on γµ x ν and produce the
2 to 1 map between SL(2, C) and SO+ (3, 1). This map translates elements of
SL(2, C) into ΛS matrices giving the action x µ 7→ ΛS ν x ν and again S and −S
µ

are mapped to the same element ΛS .

The generic representation, labeled by ( j1 , j2 ), is built from the tensor product


of these representations and the elements of the corresponding vector space on
which they act can be interpreted as tensors with 2j1 + 2j2 indices
... ⊗ ... ⇔ Sα1 ...α2j ,α̇1 ...α̇2j2 , (18.45)
| {z } | {z } 1

2j1 2j2

whose transformation properties are obvious.


Before closing this discussion, it is interesting to point out that each of the ten-
sors Sα1 ...α2j ,α̇1 ...α̇2j also contain pieces with well defined transformation prop-
1 2
erties under the group of spatial rotations. In fact, SO(3) < SO(3,1) and the
Lorentz algebra so(3, 1) contains the so(3) rotation generators li as special lin-
ear combinations of its generators: li = li+ + li− . This immediately implies that
the ( j1 , j2 ) irreps of so(3, 1) decomposes into irrepses of so(3) according to the
branching of the product j1 ⊗ j2 . For instance, the vector transforming in the
fundamental representation of SO(3,1) V µ corresponds to an object in the ( 21 , 12 ).
In fact
1 1
⊗ = 0 ⊕ 1, (18.46)
2 2
where the scalar under SO(3) is V 0 and the vector is V i .
4 We can now also come back to the issue of covering the whole group from the algebra. The
element S = − exp(γ0 γ1 + γ1 γ2 ) = −12 − γ0 γ1 − γ1 γ2 is in Spin(3, 1) ' SL(2, C), but it can never
be obtained as the exponential of any combination of the form (18.43).
204 poincar é irrepses

18.4 casimir operators

During these lectures we stressed several times that an effective way of classi-
fying group representations is given by Casimir operators. Casimir operators
commute with all the generators of a given algebra and therefore allow us to
distinguish irreducible representations by their eigenvalues. This is true also
for the Lorentz and Poincaré group and for this reason we now revisit the dis-
cussion of their irreducible representations in the light of the existence of such
Casimir operators.

18.4.1 Casimir for the Lorentz group

The Lorentz algebra has two quadratic Casimir operators

1
C1 = Jµν J µν (18.47)
2
and
1 µνρσ
C2 = e Jµν Jρσ . (18.48)
4
The existence of these two quadratic Casimir operators is obviously related
to the fact that its complexification is isomorphic to the direct sum of two
other algebras and therefore we should be able to use their quadratic Casimir
operators to classify irrepses:

so(3, 1)C = sl(2, C) ⊕ sl(2, C). (18.49)

Recalling that J0i = k i , Jij = eijk lk and li± = √1 (li ± i k i ) we have that
2

1
2
C1 = − J0i + J J = −|~k|2 + |~l |2 = (~l + )2 + (~l − )2 (18.50)
2 ij ij
and
 
C2 = eijk J0i Jjk = 2~k · ~l = −i (~l + )2 − (~l − )2 . (18.51)

We then conclude that we can use two labels ( j1 , j2 ) to differentiate the irre-
ducible representations.
18.4 casimir operators 205

18.4.2 Casimir for the Poincaré group

In the case of the Poincaré group [C1 , Pµ ] 6= 0 and also [C2 , Pµ ] 6= 0. We then
need to find new Casimir operators. The first one is simply the operator defin-
ing the mass of the state:

P2 | m i = − m2 | m i. (18.52)

In fact [ P2 , Pµ ] = 0 trivially and [ P2 , Jµν ] = 0 because P2 is a Lorentz scalar.


This allows us to distinguish between the two orbits of massive and massless
states.
The second invariant must reflect the discussion in section 18.1, namely that
physical states are classified according to their transformation properties under
the little group. The correct operator is called the Pauli–Lubanski operator and is
built using the square of the quadratic operator
1 µνρσ
Wµ = e Jνρ Pσ . (18.53)
2
Before proving that W 2 is a good Casimir operator, we note that
1 ijk
W0 = e eijk l l Pk = ~l · ~
P, (18.54)
2
1 ijk  i
Wi = −eijk J0i Pk − e Jjk P0 = −~k ∧ ~
P + P0 ~l , (18.55)
2
which tells us that W 0 is proportional to the helicity
~l · ~P
η= , (18.56)
|~P|
which defines the spin in the direction of the motion, and that W i becomes
proportional to the rotation generators l i whenever we consider the rest frame
of massive particles. We also see that
1 µνρσ
W µ Pµ = e Jνρ Pσ Pµ = 0 (18.57)
2
and that W µ commutes with the momentum operators Pν :
1 1
[W µ , P ν ] = 2 eµαβγ [ Jαβ Pγ , Pν ] = 2 eµαβγ [ Jαβ , Pν ] Pγ
(18.58)
1
= 2 eµαβγ 2 δαν Pβ Pγ = eµνβγ Pβ Pγ = 0.
206 poincar é irrepses

This means that W µ are in fact combinations of the Lorentz generators that
leave the momentum invariant and therefore generate the little group.
We then easily prove that W 2 is a good Casimir operator, because W µ is a
vector under Lorentz transformations

[ Jµν , Wρ ] = 2 ηρ[µ Wν] , (18.59)

and
1 2 2
W2 = J P + Pµ Jµν J νρ Pρ (18.60)
2
is therefore a scalar:

[ Jµν , W 2 ] = 0. (18.61)

When acting on massive particle states, whose rest frame momentum is pµ =


(m, 0, 0, 0), it reduces to

W 2 = −m2 (−|~k|2 + |~l |2 ) − m2 |~k|2 = −m2 |~l |2 , (18.62)

which means that irreducible representations are classified by the so(3) Casimir,
as expected.
When acting on massless particle states we have once more the problem with
the infinite quantum numbers not visible in nature. Choose a frame where
pµ = ( p, 0, 0, p) and decompose W µ accordingly
µ µ
W µ = η Pµ + N1 n1 + N2 n2 + Ws sµ , (18.63)

where

n1 = (0, 1, 0, 0), n2 = (0, 0, 1, 0) (18.64)

define the directions orthogonal to pµ and

sµ = (s0 , 0, 0, s3 ), with s0 6 = | s3 |. (18.65)

Since W µ Pµ = 0 and W µ Pµ ∼ Ws we have that Ws = 0. We then have that


W 2 = N12 + N22 , but there are no physical states with such continuous quantum
numbers and therefore we set them to zero, so that W 2 = 0. As expected, we
are left with η to classify the remaining states. In fact, since W 2 = 0, the helicity
determines fully W because

W 0 = η |~
P| ~ = η ~P.
and W (18.66)
18.4 casimir operators 207

exercises

1. Prove that the minimal representation for γ-matrices satisfying {γµ , γν } =


2η µν 1 with η = diag{−1, +1, +1, +1} is 4 by 4.
a) Compute the trace of γ0 and of γi using the definition of the Clifford
algebra.
b) Show that the only admissible eigenvalues for iγ0 and γi are ±1.
c) Show that 2 by 2 matrices are not enough to provide a faithful repre-
sentation.

2. In relativistic theories we often use two different representations for the


gamma matrices. One representation is called the Weyl representation
and was given in (18.26). The other representation is called the Dirac
representation and is related to the previous one by a similarity transfor-
mation:
12 12
 
µ µ −1 1
γD = SγW S , where S= √ .
2 −12 12

Build it explicitly.

3. Show that the 16 matrices Γn given by all possible products of different γ


matrices are independent (Γ0 = 1, Γ1 = γ0 , Γ5 = γ0 γ1 = −γ1 γ0 , etc...).
a) Compute (Γn )2 .
b) Show that tr Γn = 0, but for n = 0.
c) Show that Γm Γn is equal to one of the other Γ p up to a sign, and
p = 0 only when n = m.
d) Assuming that ∑n an Γn = 0 and taking traces of this relation after
multiplication with Γm one finally shows that all an coefficients must
be vanishing.

4. Compute ( 21 , 21 ) ⊗ [( 21 , 0) ⊕ (0, 12 )].

5. What kind of tensor corresponds to the (1, 0) ⊕ (0, 1) representation of


the Lorentz group?

6. Fields described by tensors with definite symmetry properties like gµν ( x ) =


gνµ ( x ), with gµ µ = 0, and Bµνρ ( x ) = B[µνρ] ( x ) transform under the action
of finite-dimensional representations of the Lorentz group. What are the
208 poincar é irrepses

labels of sl (2, C) ⊕ sl (2, C) describing them? Which spin states they con-
tain with respect to SO(3) < SO(1,3)?

7. A Dirac spinor ψ transforms under the action of the proper and


 orthochronous

Lorentz group in the spinorial representation D(ω ) = exp − 14 ωµν γµν ,
i.e. ψ → D(ω )ψ.
Using [γρ , γµν ] = 4η ρ[µ γν] show that D −1 γµ D = Λµ ν γν (prove it at the
infinitesimal level).
Using also 㵆 = γ0 γµ γ0 show that ψψ ≡ i ψ† γ0 ψ is a scalar quantity.

8. Prove that [ Jµν , Wρ ] = 2 ηρ[µ Wν] .


19
SYMMETRIES OF MOLECULES AND SOLIDS.

In this lecture we discuss some applications of the theory of group represen-


tations to Physics. In particular we will try to see what are the consequences
of the existence of symmetry properties of a physical systems and how to use
them to compute relevant physical quantities.
We will focus on systems consisting of identical particles (atoms) in an or-
dered configuration in space. The symmetries one has for these configurations
are described by discrete groups and are also subgroups of the groups of trans-
lations, rotations and reflections in space: the 3-dimensional Euclidean group

E3 = O(3) n R3 . (19.1)

Let ~x I ∈ R3 denote the position of the constituents of the molecule

M = {~x I , I = 1, . . . , N }. (19.2)

Definition 19.1. The symmetry group of a molecule M is the subgroup of E3 that


leaves M invariant:
G = { g ∈ E3 | D ( g ) M = M } .

19.1 classic vibrations of molecules

Imagine a molecule where the atoms are in a fixed position. The electronic
levels can be obtained as a function of their configuration and their energy de-
pends on the relative positions of the nuclei. These levels will change even if we
move the nuclei by a little amount. Fix then the average positions h~x I i of the N

209
210 symmetries of molecules and solids

units forming a molecule and call q A , where A = 1, . . . , 3N, the displacements,


weighted with their mass:
 
q3I −2
m I (~x I − h~x I i) =  q3I −1  . (19.3)
q3I

The potential energy of the system depends on the values of q A . Although the
resulting function is strictly valid only for fixed q A , we can still use V (q A ) to
compute the molecule vibrations, because q̇ A is going to be much smaller than
the average speed of the electrons.
The total energy of the system is the sum

H = K + V, (19.4)
1
where K = 2m |~p I |2 and V = V (~x I ).
If we use as coordinates the displacements q A , we can simplify the potential
and expand it around the equilibrium positions:
1 1
K=
2 ∑ q̇2A , V' F q qB ,
2 AB A
(19.5)
A

where we subtracted the constant part V (0) and we stopped at second order in
the expansion of the potential. The equations of motion of this system are

q̈ A + FAB q B = 0. (19.6)

Whenever FAB = ω 2A δAB we have the obvious solution

q A = c A cos(ω A t + φ A ). (19.7)

However, in general F is not diagonal. We assume it is real, symmetric and


positive definite, as a consequence of the expansion around a local minimum
configuration. In this case

F = S T ∆S, (19.8)

where
 
λ1
∆=
 .. ,

λ A = ω 2A > 0 (19.9)
.
λ3N
19.1 classic vibrations of molecules 211

and the columns of S are given by the 3N eigenvectors c( A) , so that the general
solution can be expressed in terms of the so-called Normal Modes

∑ cA
( B)
qA = α B cos(ω B t + φB ), (19.10)
B

where α and φ are integration constants, fixed by initial conditions.

Definition 19.2. The coordinates Q = Sq are called Normal Coordinates.

Using normal coordinates, the Hamiltonian simplifies to

1 1
H= PA PA + ω 2A Q A Q A . (19.11)
2 2
We now understand that the problem of studying molecule vibrations is equiv-
alent to the analysis of its normal modes. This problem becomes purely group
theoretical, once we use the symmetry properties of the molecule.

19.1.1 Symmetries

The ensemble of discrete points providing the equilibrium positions of a molecule


are invariant (or better, are mapped among themselves in a way that the final
result is indistinguishable from the original configuration) under the action of
the symmetry group G. For instance, a molecule described by a regular poly-
gon in 2 dimensions has symmetry group G = Dn .
The same group, however, also has a non-trivial action on the displacements.
In fact, we have to combine the vertex permutations with the rotations of the
relative reference frames. In figure 25 the two configurations are obtained by
reflection.

y3 y3
3 3
x3 x3

y1 y2 y2 y1

1 x1 2 x2 2 x2 1 x1

Figure 25.: Caption


212 symmetries of molecules and solids

Since the atom 2 is displaced (by the pink arrow in the picture), the final
configuration is obtained from the original one by permuting the vertices 1 and
2, but also acting on the displacement q2 , which is mapped by means of the
same element of D3 by a reflection around the y axis.
The generic action of G is therefore realized on the q A coordinates by the real
and unitary matrix

D ( g ) = p ( g ) ⊗ R ( g ), (19.12)

where p acts by permuting the I indices and R is a rotation acting on ( x I , y I , z I ),


in the same fashion for any value of I.
Since the matrix D ( g) comes from a product of different representations of
the same group, it is often reducible

D ( g) = ∑ n a D a ( g ), (19.13)
a

where a labels different irreducible fragments. Since D should represent a


symmetry of the system, whenever

q 7→ D ( g)q, (19.14)

K and V should be left invariant. Invariance of the kinetic term is straightfor-


ward because D is real and unitary. Invariance of the potential gives

D T ( g) FD ( g) = F, (19.15)

or, equivalently

FD ( g) = D ( g) F, ∀ g ∈ G. (19.16)

If we decompose D in irreducible fragments we get

F mn D n ( g) = D m ( g) F mn , (19.17)

where m and n are the dimensions of the irreducible blocks. Using Schur’s
lemma we can deduce that F mn = 0 whenever m 6= n and is proportional to
the identity on the same irreducible fragments. We conclude that F is block
diagonal on invariant subspaces of the space of displacements and therefore
we can reduce the computation of normal modes to the analysis of irreducible
representations of discrete groups.
19.1 classic vibrations of molecules 213

19.1.2 Diatomic molecule

Let us start by considering a diatomic molecule, like H2 , O2 or N2 . Assume that


displacements are allowed only in 1 dimension, we can fix the reference frame
so that

h x1 i = −d/2, h x2 i = d/2, (19.18)

where d is the average distance between the two nuclei.

Figure 26.: Caption

We call q1 and q2 the displacements, so that the distance between the two
nuclei is

δ = |(−d/2 + q1 ) − (d/2 + q2 )| = |d + q2 − q1 |. (19.19)

Following Hooke’s law, for small displacements

1 1
V= k ( δ − d )2 = k ( q2 − q1 )2 , (19.20)
2 2
so that
 
1 −1
F=k . (19.21)
−1 1

The symmetry group is the permutation group of two objects S2 ' Z2 , repre-
sented on (q1 , q2 ) by
   
1 0 0 1
D2 ( 1 ) = , D2 ( σ ) = . (19.22)
0 1 1 0

This representation is clearly reducible, because Z2 is abelian and its irreducible


representations are all 1-dimensional. The characters of the representation
given above are χ2 (1) = 2 and χ2 (σ ) = 0 and they must be the sum of the char-
acters of Z2 irrepses. Since χ1 (1) = χ1 (σ ) = 1 and χ10 (1) = 1, χ10 (σ ) = −1,
we have that D2 = D1 ⊕ D10 . As we proved above, F must be diagonal in this
basis and hence

tr( D ( g) F ) = λ1 χ1 ( g) + λ2 χ10 ( g). (19.23)


214 symmetries of molecules and solids

We then deduce that


  
1 −1
tr( D (1) F ) = tr k = 2k = λ1 + λ2 , (19.24)
−1 1

  
−1 1
tr( D (σ ) F ) = tr k = −2k = λ1 − λ2 , (19.25)
1 −1
which implies that
λ1 = 0, λ2 = 2k. (19.26)
This is exactly what we should get from the split of the displacements into
the center of mass displacement X = √1 (q1 + q2 ) and the displacement with
2
respect to the center of mass x = √1 ( q2 − q1 ). The Hamiltonian becomes
2
1  2  1
H= Ẋ + ẋ2 + (2k ) x2 . (19.27)
2 2
We see that the null eigenvalue is associated to the center of mass. In fact, its
motion is that of a translation and hence has no vibrational energy. On√the
other hand, x describes precisely a vibrational mode with frequency ω = 2k.
We can actually understand what is described by the normal modes by pro-
jecting the displacement vector into the subspace left invariant by the a-th irre-
ducible representation. This can be done by means of the projector
Na
Pa ≡
#G ∑ χ a ( g ) ∗ D ( g ). (19.28)
g

If one evaluates this projector as a matrix one sees that it is proportional to the
identity only in the block corresponding to the a irreducible representation:
a
( Pa )ij = Na ( Dkk , Dij ) = ∑ δab δij , (19.29)
b
where b runs over the various irrepses contained in D. For the case at hand we
have that
   
1 1 1 1 1 1
P1 = = √ √ (1, 1) (19.30)
2 1 1 2 1 2
and
   
1 1 −1 1 1 1
P10 = = √ √ (1, −1) . (19.31)
2 −1 1 2 −1 2

Their projection on (q1 , q2 ) gives QCM = q1 + q2 and 2 x = q2 − q1 .
19.1 classic vibrations of molecules 215

19.1.3 Diatomic molecule in 2 dimensions

If we now consider the same diatomic system in a 2-dimensional space, we


have 4 coordinates for the displacements and the symmetry group is G = D2 =
Z2 × Z2 , i.e. the set of reflections around the two axes.

y
y1 y2
x1 x2
x
1 2

Figure 27.: Caption

The symmetry group G has to be realized as a set of permutations of the


two centers and as the reflections of the x1 , y1 and x2 , y2 axes. We call ρ the
generator associated to the reflection symmetry around the x axis and σ the
generator of the reflection symmetry around the y axis. The representation of
the group elements as permutations of the index I = 1, 2 is

   
1 0 0 1
p (1) = p ( ρ ) = , p(σ ) = p(σρ) = . (19.32)
0 1 1 0

This realizes the fact that the equilibrium positions of the nuclei is not affected
by a reflection around the x axis.
Considering the embedding in O(2), the same elements are represented as

  

1 0 −1 0
R (1) = , R(σ) = ,
0 1 0 1
    (19.33)
1 0 −1 0
R(ρ) = , R(σρ) = .
0 −1 0 −1
216 symmetries of molecules and solids

We can now deduce the form of the representation matrices for G = D2 on


the displacements q A , A = 1, . . . , 4. They are

−1
   
1
1 1 
D4 ( 1 ) = 
  
, D4 ( σ ) = 
 −1
,
 1  
1 1
−1
   
1
 −1   −1 
D4 ( ρ ) =  , D4 (ρσ) = 
 −1
.
 1  
−1 −1
(19.34)

We can now show that D4 = D1 + D10 + D100 + D1000 , by computing χ4 (1) = 4,


χ4 (σ ) = χ4 (ρ) = χ4 (ρσ ) and using the character table of Z2 × Z2 :

{1} {σ} {ρ} {ρσ}


1 1 1 1 1
10 1 −1 1 −1 (19.35)
100 1 1 −1 −1
1000 1 −1 −1 1

In this basis

−1
 
1
 0 
F = k
 −1
. (19.36)
1 
0

We therefore deduce that

tr( D4 (1) F ) = 2 k = λ1 + λ10 + λ100 + λ1000 , (19.37)

tr( D4 (ρ) F ) = 2 k = λ1 + λ10 − λ100 − λ1000 , (19.38)

tr( D4 (σ ) F ) = 2 k = λ1 − λ10 + λ100 − λ1000 , (19.39)

tr( D4 (ρσ) F ) = 2 k = λ1 − λ10 − λ100 + λ1000 . (19.40)


19.1 classic vibrations of molecules 217

We solve this by λ1 = 2k and λ10 = λ100 = λ1000 = 0. The corresponding


projectors are

−1
 
1
1  0 
P1 =  , (19.41)
2  −1 1 
0

which correctly gives the vibration mode in x, with frequency 2k,
 
1 1
1  0 
P10 =  , (19.42)
2  1 1 
0
which gives the zero mode related to translations in x,
 
0
1  1 1 
P100 =  , (19.43)
2  0 
1 1
which gives the zero mode related to translations in y, and
 
0
1  1 −1 
P1000 =  , (19.44)
2  0 
−1 1
which produces anti-clockwise rotations around the origin.

19.1.4 Triatomic molecule

If we have three nuclei at the vertices of an equilateral triangle, the symmetry


group is D3 ' S3 , whose 3-dimensional representation as vertex permutations
should be multiplied by its 2-dimensional representation as rotations on the
plane. The resulting representation is reducible as

D6 = D1 = D1 0 + D2 + D2 . (19.45)

The corresponding
√ subspaces give the non-trivial dilatation mode, with fre-
quency ω = 3k, the zero energy, rotational mode, the 2-dimensional space of
218 symmetries of molecules and solids

translations in x and y (again with √


zero energy) and the 2-dimensional space
of vibrations, with frequency ω = 3k/2. We leave this straightforward, but
tedious analysis to the reader.
A
SOLUTIONS

Here are some possible very rough outlines for the solutions of the exercises
provided in the text.

chapter 2

1. The lagrangian can also be rewritten as


1 1
L = − ∂µ Ai ∂µ Ai − ∂µ Bi ∂µ Bi − m2 eij Ai Bj , (A.1)
2 2
where e12 = −e21 = 1. The equations of motion follow from the straigth-
forward application of the Euler–Lagrange equations:

 Ai − m2 eij Bj = 0, (A.2)

 Bi + m2 eij A j = 0. (A.3)

2. Euler–Lagrange equations follow from


δL δL
− ∂µ = 0. (A.4)
δφi δ∂µ φi

One has to pay attention to the fact that Dµ φi contains a piece without
δDµ φ a δDν φ a µ
derivatives, so that δφi
= Aµ e ai and δ∂µ φi
= δia δν . We then have

δL δDµ φk µ j
 
1
δφi
= −δkj δφi
D φ + 6 Cijk + Cjik + Cjki φ j φk
(A.5)
= −δkj Aµ eki D µ φ j + 12 C(ijk) φ j φk .

219
220 solutions

When varying the last term in the Lagrangian one has to take into account
the fact that only the fully symmetrized terms in the C tensor survive,
because they are contracted with three identical fields. Also
δL δDν φ a ν j
= − δaj D φ = −δij D µ φ j . (A.6)
δ∂µ φi δ∂µ φi

We then get that


1
δij Dµ ( D µ φ j ) = C φ j φk , (A.7)
2 (ijk)
where

Dµ ( D µ φ j ) = (∂µ δ jk + Aµ e jk )[(∂µ δka + Aµ eka )φ a ]. (A.8)

3. Euler–Lagrange equations:

Φ + ηΦ − (Φ† Φ)Φ = 0. (A.9)

When Φ is constant we have ηΦ − (Φ† Φ)Φ = 0, which can be solved only


for Φ = 0 or when (Φ† Φ) = 1, which implies c1 = c2 = 0, |c3 |2 + |c4 |2 =
1.

chapter 3

1. One by one:
• Q+ with a ◦ b = a/b is not a group because the associative property
fails. Let a, b, c ∈ Q+ , we have that
a
a a
( a ◦ b) ◦ c = ◦c = b = . (A.10)
b c bc
On the other hand
 
b a ca
a ◦ (b ◦ c) = a ◦ = = . (A.11)
c b b
c

• Yes. The sum of two even integers 2m and 2n, where m, n ∈ Z, is


2(m + n), which is even. Addition is associative. The unit is zero,
which is an even integer, and the inverse of 2n is −2n.
solutions 221

• Z with a ◦ b = a − b is not a group because we do not have a neutral


element, nor multiplication is associative. For instance
( n − m ) − p 6 = n − ( m − p ). (A.12)
Also, while 0 could be a good identity from the right:
n − 0 = n, (A.13)
it is not a good identity from the left
0 − n = −n 6= n (A.14)
and we know that any identity from the right in a group must also
be an identity from the left.
m
• Yes. A generic element of Q∗ can be written as , where m and n
n
are coprime integers. Closure:
m  k  mk
◦ = , (A.15)
n p np
x
which reduces to the form with x and y coprimes, after simplifi-
y
cations. The product is obviously associative. Neutral element is the
m n
number 1. The inverse of is .
n m
• Associativity is clear and 1 is the obvious neutral element under
multiplication. The existence of the inverse is tricky. Let us try first
with Z5∗ . Multiplying 2 with the other numbers
2 · 2 = 4, 2 · 3 = 6 ∼ 1, 2 · 4 = 8 ∼ 3 (A.16)
we see that 2−1 = 3 and 3−1 = 2. Multiplying 4 with the other
numbers
4 · 2 = 8 ∼ 3, 4 · 3 = 12 ∼ 2, 4 · 4 = 16 ∼ 1 (A.17)
and therefore 4−1 = 4. So there is always an inverse.
In general, take a ∈ Zn , we need to prove that there is always an
element b ∈ Zn such that ab = 1 mod n. First we show that for a
given a the map
Z∗n → Z∗n
(A.18)
b 7→ ab mod n
222 solutions

is injective. If it was not, we should have two elements x, y ∈ Z∗n


so that ax mod n = ay mod n. This, however, implies that a( x −
y) mod n = 0 and therefore that x − y can be divided by n, but this
is true in Z∗n only if x = y. Since it is injective and maps all elements
of Z∗n into elements of Z∗n it must also be surjective. This means that
there is an element b ∈ Z∗n for any a ∈ Z∗n such that ab mod n = 1.

2. For any group

( ab)−1 = b−1 a−1 . (A.19)

If G is Abelian any two elements commute and therefore we also have


that b−1 a−1 = a−1 b−1 , which implies ( ab)−1 = a−1 b−1 . On the other
hand if

( ab)−1 = a−1 b−1 (A.20)

for all a, b ∈ G we can multiply from the right by ba getting

( ab)−1 ba = 1 (A.21)

and multiplying from the left by ab we get

ba = ab. (A.22)

3. Assume we have 3 elements {1, a, b}. We can fill immediately the lines
and columns where the identity appears

1 a b
1 1 a b
(A.23)
a a
b b

The product ab cannot be equal to a or b otherwise we would imply that


either b = 1 or a = 1. Hence ab = 1. At this point the rearrangement
theorem fixes the other entries, which must be different in each row and
in each column
1 a b
1 1 a b
(A.24)
a a b 1
b b 1 a
solutions 223

4. Take G = {1, a, b, c}. We start as usual by filling rows and columns asso-
ciated to the identity:

1 a b c
1 1 a b c
a a (A.25)
b b
c c

We then consider the product aa. This cannot be equal to a otherwise


it would imply a = 1. We then have two possibilities, either aa = 1, or
aa = b (the case aa = c is just a relabeling of the elements). In the first
case, the table is

1 a b c
1 1 a b c
a a 1 (A.26)
b b
c c

and from the rearrangement theorem we immediately fill the rest of the
row:

1 a b c
1 1 a b c
a a 1 c b (A.27)
b b
c c

Also, ba 6= b, otherwise a = 1 and from the rearrangement theorem we


get

1 a b c
1 1 a b c
a a 1 c b (A.28)
b b c
c c b
224 solutions

The remaining entries could be completed in two possible ways:

1 a b c 1 a b c
1 1 a b c 1 1 a b c
a a 1 c b a a 1 c b (A.29)
b b c 1 a b b c a 1
c c b a 1 c c b 1 a

These are two distinct groups. We now come back to the other option
above, aa = b. By the rearrangement theorem we complete immediately
the second row and the second column
1 a b c
1 1 a b c
a a b c 1 (A.30)
b b c
c c 1

Actually, at this point also the remaining entries are fixed:

1 a b c
1 1 a b c
a a b c 1 (A.31)
b b c 1 a
c c 1 a b

This table is identical to the second in (A.29) if we relabel a 7→ b and b 7→ a.


We then have only two possible groups. In the first one all the elements
have order 2 and in fact it is Z2 × Z2 , where the first is generated by a
and the second by b. In the second we have an element of order 4, it is
Z4 , generated by b: b2 = a, b3 = ab = ba = c, b4 = a2 = 1.
5. Let us compute the multiplication table. From the data we have some
multiplications, but not all. First of all we introduce − I = (−1) I, − J =
(−1) J and −K = (−1)K and we show that they cannot be any of the
elements we already have. It is tedious, but straightforward. For instance,
− I cannot be I or −1 otherwise we would have that I = 1 or −1 = 1.
Then we prove that − I 6= 1, otherwise I = (− I ) I = (−1) I 2 = (−1)2 = 1.
Finally we prove that − I 6= K, otherwise − IK = −1 and hence IK = 1 =
KI, which implies J = KI J = K2 = −1. The same goes though for the
other elements.
solutions 225

So we assume that we have at least 8 elements ±1, ± I, ± J, ±K, whose


products are 1g = g1 = g, for any g, I 2 = J 2 = K2 = −1, (−1)2 = 1,
(−1) I = − I, (−1) J = − J, (−1)K = −K and I J = K.
The rest of the table is computed by using the properties above. For
instance, we have that I (−1) = I I I = (−1) I = − I and the same for J
and K. Also − I I = (−1)2 = 1 and hence also I (− I ) = 1 (and the same
for J and K). From I J = K we also get JK = I by multiplying by I from
the left and −K from the right. In the same fashion we also get KI = J
by multiplying by K from the left and − J from the right. From I J = K
we also get K J = − I by multiplying by J from the right and we can do
similar manipulations to get IK = − J and J I = −K. The rest follows.

1 I J K −1 − I − J − K
1 1 I J K −1 − I − J − K
I I −1 K −J −I 1 −K J
J J −K −1 I −J K 1 −I
K K J −I −1 −K − J I 1 (A.32)
−1 −1 −I −J −K 1 I J K
−I −I 1 −K J I −1 K −J
−J −J K 1 −I J − K −1 I
−K −K −J I 1 K J − I −1

6. Case by case:
• Z4 and S3 have different order, hence they cannot be isomorphic.
#(Z4 ) = 4, #(S3 ) = 3! = 6.
• Z6 and S3 have the same order, but the first is Abelian, while the
second is not. We conclude that they cannot be isomorphic.
• S3 and D3 are isomorphic. In general one can easily prove the
isomorphism by mapping generators between them and showing
that they respect the relations in the presentations. In this case we
can associate a 3-cycle permutation to ρ and a 2-cycle to σ, so that
σρ = ρ−1 σ. An instance that works is σ = (12)(3) ρ = (123).

7. A generic p-cycle of Sn (p ≤ n) has the form

a = ( i1 , i2 , . . . , i p ), (A.33)
226 solutions

where i j = 1, . . . , n, for j = 1, . . . , p, and must all be different between


themselves. The same p-cycle reversed is

b = ( i p , i p −1 , . . . , i 1 ). (A.34)

Their product gives the identity

ab = (i1 , i2 , . . . , i p ) ◦ (i p , i p−1 , . . . , i2 , i1 ) = 1 (A.35)

as follows from



 i p 7 → i p −1 7 → i p
..


. (A.36)


 i2 7 → i1 7 → i2

i 7 → i 7 → i

1 k 1

The same argument can be applied to the generic cycle, which is a disjoint
product of cycles.

8. Explicitly:

1 7 → 1 7 → 7 7 → 7


2 7 → 2 7 → 3 7 → 4




3 7 → 4 7 → 6 7 → 6



(34) ◦ (623174) ◦ (34) = 4 7→ 3 7→ 1 7→ 1 = (624173), (A.37)

5 7→ 5 7→ 5 7→ 5





6 7 → 6 7 → 2 7 → 2




7 7→ 7 7→ 4 7→ 3



 1 7→ 2 7→ 2 7→ 1

2 7 → 1 7 → 3 7 → 3



(12) ◦ (2)(13)(45) ◦ (12) = 3 7→ 3 7→ 1 7→ 2 = (1)(23)(45), (A.38)

4 7→ 4 7→ 5 7→ 5





5 7→ 5 7→ 4 7→ 4

solutions 227



 1 7→ 1 7→ 3 7→ 4

2 7 → 2 7 → 4 7 → 3



(34) ◦ (13)(452) ◦ (34) = 3 7→ 4 7→ 5 7→ 5 = (14)(352). (A.39)

4 7→ 3 7→ 1 7→ 1





5 7→ 5 7→ 2 7→ 2

We see that the result of the application of a 2-cycle (and its inverse) to a generic
permutation p, exchanges in p the position of the two numbers defining such
2-cycle:

( ab) ◦ (. . . a . . . b . . .) ◦ ( ab) = (. . . b . . . a . . .). (A.40)

9. A3 is given by all the elements in S3 with positive parity. These are 1,


(123) = (13) ◦ (12) and (132) = (12) ◦ (13). The only group of order 3 is
Z3 = hρ | ρ3 = 1i and indeed the two are isomorphic as follows from the
identification ρ = (123).

10. We should identify two permutations of S4 with a and b satisfying a2 =


1 = b3 and ( ab)3 = 1. The order of a generic permutation is the smallest
common multiple between the order of the disjoint cycles composing it.
This means that a should be given by a 2-cycle or by the product of 2-
cycles, while b should be a 3-cycle. All 3-cycles are in A4 , because they
arise as the product of connected 2-cycles. On the other hand single 2-
cycles are not in A4 , while the product of two 2-cycles is in A4 . Hence a
shold be one among

(12)(34), (13)(24), (14)(23), (A.41)

and b one among

(1)(234), (1)(324), (2)(413), (2)(431),


(A.42)
(3)(412), (3)(421), (4)(123), (4)(132).

We just need to check which couple satisfy the relation ( ab)3 = 1. An


instance is

a = (12)(34), b = (4)(123). (A.43)


228 solutions

chapter 4

1. Cosets have all their elements in common or they are distinct (see The-
orem 4.1). This means that every element of the group must appear in
exactly one distinct coset. Thus, since each coset has the same number
of elements, the number of distinct cosets multiplied by the number of
elements in the cosets is equal to the order of the group.
2. The center of a group G is C = { g ∈ G | gh = hg, ∀h ∈ G }. Let now
Dn = hρ, σ | ρn = σ2 = 1, ρσ = σρn−1 i. From this presentation we
see that any h ∈ Dn can be written as h = σ a ρb , where a = 0, 1 and
b = 1, ..., n. In order to check what is the center we can then see for which
values of a and b the element h commutes with the generators of Dn . The
commutator with σ gives
σ a ρb σ = σ a+1 ρn−b = σσ a ρb ⇔ b = n − b ⇔ b = n/2, (A.44)
which is satisfied only for even n, or when b = 0. The commutator with
ρ gives
 a b −1
σ ρ 6= σ a ρb ρ, a = 1, n 6= 2
ρσ a ρb = b + 1 (A.45)
ρ , a=0
We then conclude that
CDn = {1, ρn/2 }, (A.46)
when n is even, and
C Dn = { 1 } , (A.47)
when n is odd. Explicitly:
σ a ρb ρn/2 = ρn−n/2 σ a ρb = ρn/2 σ a ρb . (A.48)

3. We have 9 subgroups of order 2: {1, (12)}, {1, (13)}, {1, (14)}, {1, (23)},
{1, (24)}, {1, (34)}, {1, (12)(34)}, {1, (13)(24)}, {1, (14)(23)}. 4 subgroups
of order 3: {1, (123), (132)}, {1, (124), (142)}, {1, (134), (143)}, {1, (234),
(243)}. 7 subgroups of order 4, 3 of them are isomorphic to Z4 , namely {1,
(1234), (13)(24), (1432)}, {1, (1243), (14)(23), (1342)}, {1, (1324), (12)(34),
(1423)}, and 4 are isomorphic to Z2 × Z2 : {1, (12)(34), (13)(24), (14)(23)},
{1, (12), (34), (12)(34)}, {1, (13), (24), (13)(24)}, {1, (14), (23), (14)(23)}.
Other subgroups of higher order are S3 , D4 and A4 .
solutions 229

4. Any two elements of an Abelian group G commute, i.e. ab = ba, ∀ a, b ∈ G.


This implies that gH = Hg for any H < G and hence that left and right
cosets are identical.

5. As a first step we write the same permutations using disjoint cycles:

a = (13)(26)(45)(78), b = (1234567), c = (124)(365). (A.49)

a) From this expression it is obvious that they sit in different conju-


gacy classes because the disjoint cycles describing them have differ-
ent length.
b) Rewriting also b and c as product of 2-cycles, namely

b = (12)(23)(34)(45)(56)(67), c = (12)(24)(36)(56), (A.50)

we conclude that a, b, c ∈ A8 .
c) H is generated by a e b, while the elements of G come from products
of a, b and c. It is straightforward to see that a and c commute

ca = ac = (1, 6, 4, 3, 2, 5)(78), (A.51)

while we can move any b to the left of any c obtaining a power of b:

cb = b2 c = (1, 3, 5, 7, 2, 4, 6). (A.52)

We deduce that any g ∈ G can be written as the product g = cα h,


with h ∈ H and α = 0, 1, 2. From this we have that n ∈ H implies
gng−1 ∈ H. In detail

gng−1 = cα hnh−1 c3−α = cα ñc3−α , (A.53)

where ñ ∈ H. Using what we just found we can then move cα to the


right and we are left with an element in H.

6. The identity must be in any subgroup. We can then take the other ele-
ments one by one and see if they form a subgroup. An obvious subgroup
is H0 = {1, −1}. For any of the other elements the set {1, Ji } does not
close and, for any given i, needs also −1 = Ji Ji and − Ji = (−1) Ji . The
result is a subgroup Hi = {±1, ± Ji }, where i = 1, 2, 3. Starting from
{1, − Ji } we end up with the same subgroups Hi . Adding a third element
230 solutions

to H0 we end up again with Hi and if we add another element to Hi we re-


cover the full Q. From the multiplication table found before one sees that
all such subgroups are normal and hence left and right quotients coincide.
In detail Q/H0 = {1 ∼ −1, J1 ∼ − J1 , J2 ∼ − J2 , J3 ∼ − J3 } ∼ Z2 × Z2 and
Q/H1 = {1 ∼ −1 ∼ J1 ∼ − J1 , J2 ∼ − J2 ∼ J3 ∼ − J3 } ∼ Z2 (and similar
results follow for Q/H2 and Q/H3 ).

7. Z2 × Z2 is not isomorphic to Z4 bacause the latter has an element of


order 4, while all the elements in the first have at most order 2. On the
other hand Z2 × Z3 ' Z6 . In fact if a generates Z2 and b generates Z3 ,
i.e. a2 = 1 and b3 = 1, we can construct an isomorphism between the two
groups by mapping φ : ( a, b) 7→ ρ, where ρ is the Z6 generator. Explicitly:

φ( a, b) = ρ, φ( a2 , b2 ) = φ(1, b2 ) = ρ2 ,

φ( a3 , b3 ) = φ( a, 1) = ρ3 , φ( a4 , b4 ) = φ(1, b) = ρ4 , (A.54)

φ( a5 , b5 ) = φ( a, b2 ) = ρ5 , φ( a6 , b6 ) = φ(1, 1) = 1.

8. We have A = h a | a2 = 1i and B = hb | bk = 1i. The presentation for the


semidirect product A n B is C = h a, b | a2 = bk = 1, aφ(b) a = bi, where φ
is a B automorphism.
Whenever φ(b) = b we have the direct products, which are Z2 × Z2 ' D2 ,
Z2 × Z3 = Z6 and Z2 × Z4 , for k = 2, 3, 4. When k = 3 also φ(b) = b2 is
a good automorphism, and the result is the group D3 = S3 . When k = 4
also φ(b) = b3 is an automorphism and the result is the group D4 .
Other choices for φ(b) do not give an automorphism. We can never map
φ(b) = 1, because any homomorphism already has φ(1) = 1 and we want
φ to be 1 to 1. When k = 4 we also cannot choose φ(b) = b2 , because it is
not surjective: φ(bn ) = b2n will never cover b and b3 .

chapter 5

1. Using an explicit matrix form

A O
 
A⊕B = . (A.55)
O B
solutions 231

We then have that when A and C have the sane dimensions and so do B
and D

A O C O AC O
    
( A ⊕ B)(C ⊕ D ) = = = AC ⊕ BD.
O B O D O BD
(A.56)

For the direct product, using double indices as in the text, we have that

( A ⊗ B)(ij)(kl ) (C ⊗ D )(kl )(mn) = Aik Bjl Ckm Dln


(A.57)
= ( AC )im ( BD ) jn = ( AC ⊗ BD )(ij)(mn) .

From the same formulae we also have that

tr( A ⊗ B) = ( A ⊗ B)(ij)(ij) = Aii Bjj = (tr A)(tr B). (A.58)

We see that
−i −i
   
−i  −i
γ0 =  γ1 = 
  
, , (A.59)
 −i   i 
−i i

−1 −i
   
1 i 
γ2 =  γ3 = 
  
, . (A.60)
 1   i 
−1 −i

The product

γ5 = −i γ0 γ1 γ2 γ3 (A.61)

follows by taking the products in each factor of the tensor product defin-
ing the matrices:

γ5 = −i (−iσ1 σ2 σ2 σ2 ) ⊗ (1σ1 σ2 σ3 ) = σ3 ⊗ 1. (A.62)

2. When k = 2 we had only one case Z2 × Z2 . Its elements are 1, a, b, ab = ba,


where a and b are the generators of the two Z2 factors. We want to
232 solutions

represent these elements on a vector space of dimension #(Z2 × Z2 ) = 4,


whose basis is

| gi = {|1i, | ai, |bi, | abi} . (A.63)

The regular representation is then given by matrices D R ( g) such that


D R ( g)| g̃i = | g g̃i. We then have
 
1
1 
D R (1) = 14 , D R ( a) = 

, (A.64)
 1 
1

   
1 1
 1 1
D R (b) =  D R ( ab) = 
  
,  . (A.65)
 1   1 
1 1

Since the group is abelian all its irreducible representations are 1-dimensional.
This implies that we should reduce this 4-dimensional representation into
the sum of 4 1-dimensional representations. In other words, we can put
all its elements in a diagonal form. This is done by computing the eigen-
values and eigenspaces of the generators a and b. From this computation
we get that
    
  
1 1 1 1
1  1  1  −1  1  −1  1  1 
√  , √  , √  , √   (A.66)
4 1  4  −1  4 1  4  −1 
1 1 −1 −1

are a basis for the 4 invariant subspaces of C4 so that


   
1 1
−1 −1
D R ( a) =  D R (b) = 
   
,  . (A.67)
 1   −1 
−1 1

3. D ( a) D ( a) = 12 and therefore they represent Z2 . This representation must


be reducible because all irreducible representations of abelian groups are
solutions 233

1-dimensional. Since D (1) and D ( a) commute, we can diagonalize them


simultaneously and the diagonal form of a is simply
 
1 0
.
0 −1

This had to be expected from the trace, which is invariant. trD ( a) = 0


implies that the two 1-dimensional representations of a that appear in the
2-dimensional one provided in the text must have opposite signs.

4. a) The representation given in the text for the generators of the group
is block diagonal, with the first block given by a 3 by 3 matrix and
the second one 1-dimensional. Since all the other elements of A4 are
given by products of these matrices, they all have the same block-
diagonal structure.
b) From the decomposition of the matrices D ( a) and D (b) we already
have two representations, a 3-dimensional one and a 1-dimensional
one. We also know that we always have the trivial representation
D ( a) = D (b) = 1, which has also dimension 1. The last one is obvi-
ously the complex conjugate of the 1-dimensional complex represen-
tation D ( a) = 1, D (b) = e2πi/3 , namely D ( a) = 1, D (b) = e4πi/3 .
c) The elements of H have order 2, hence it is isomorphic to Z2 ×
Z2 . This is a normal subgroup of A4 . Its elements are (12)(34),
(13)(24), (14)(23) and the identity, which we know are closed under
conjugation in S4 and hence also in A4 . (Note that in general conjugacy
classes of Sn and An differ, because elements with the same length of cycles
may be connected by odd cycles, which are not in An . However the only
result may be to split a conjugacy class in multiple conjugacy classes and
hence the argument here goes through) The quotient gives Z3 , generated
by b.

5. Let H = C N , then (ψ ⊗ χ) ∈ C N ⊗ C N and therefore we can represent


2
it by elements of C N , namely N × N matrices, M = ψχ T . The opera-
tion generated by ρ is the transposition D (ρ)[ M ] = M T . Obviously any
representation D is such that D2 = 1, being a representation of S2 .
Let us now verify the properties given in the text. From the action

1
P± (ψ ⊗ χ) = (ψ ⊗ χ ± χ ⊗ ψ) (A.68)
2
234 solutions

we get

h i
1
P± P± (ψ ⊗ χ) = P± 2 (ψ ⊗ χ ± χ ⊗ ψ)
1 (A.69)
= 4 [(ψ ⊗ χ ± χ ⊗ ψ) ± (χ ⊗ ψ ± ψ ⊗ χ)]
1
= 2 (ψ ⊗ χ ± χ ⊗ ψ) = P± (ψ ⊗ χ).

2 (1 + D + 1 − D ) = 1 and P+ P− =
1
It is then trivial that P+ + P− =
P− P+ = 0.

We also see that DP± = D 1±2 D = 21 ( D ± D2 ) = 21 ( D ± 1) = ± P± .


This means that the spaces H± ≡ P± (H ⊗ H) are invariant under the
action of D. Obviously one could further reduce the representation to
1-dimensional invariant subspaces, because S2 is Abelian. Note also that
the invariant subspaces H+ and H− are each other’s complement because
P+ P− = 0. The dimension of P+ (H ⊗ H) equals that of N × N symmetric
N ( N +1)
matrices, and therefore is 2 . The dimension of P− (H ⊗ H) equals
N ( N −1)
that of N × N antisymmetric matrices: 2 .

6. To be filled in...

chapter 6

1. The group D3 has presentation D3 = hρ, σ | ρ3 = σ2 = 1, ρσ = σρ2 i. Note


that D (c)2 = D (d)2 = D (e)2 = D (1). We therefore have a representation
of D3 where c, d, e are identified with σ, σρ and σρ2 , while a and b are
identified with ρ and ρ2 .

The representation is reducible. In fact the matrix D (c) commutes with


all matrices of the representation provided in the text, but it is not pro-
portional to the unit matrix.
 
1 0
By diagonalization: D (c) = .
0 −1
solutions 235

2. We can explicitly check (summing in the order the elements 1, ρ, ρ2 , σ, σρ


and σρ2 ):

 
1 1 1 1 1 1
( D11 , D11 ) = 1+ + +1+ + = , (A.70)
6 4 4 4 4 2
1√ 1√ 1√ 1√
 
1
( D11 , D12 ) = 0− 3+ 3+0− 3+ 3 = 0,(A.71)
6 4 4 4 4
1√ 1√ 1√ 1√
 
1
( D11 , D21 ) = 0− 3+ 3+0+ 3− 3 = 0,(A.72)
6 4 4 4 4
 
1 1 1 1 1
( D11 , D22 ) = 1− − −1+ + = 0, (A.73)
6 4 4 4 4
 
1 3 3 3 3 1
( D12 , D12 ) = 0+ + +0+ + = , (A.74)
6 4 4 4 4 2
 
1 3 3 3 3
( D12 , D21 ) = 0+ + +0− − = 0, (A.75)
6 4 4 4 4
1√ 1√ 1√ 1√
 
1
( D12 , D22 ) = 0+ 3− 3+0− 3+ 3 = 0,(A.76)
6 4 4 4 4
 
1 3 3 3 3 1
( D21 , D21 ) = 0+ + +0+ + = , (A.77)
6 4 4 4 4 2
1√ 1√ 1√ 1√
 
1
( D21 , D22 ) = 0+ 3− 3+0+ 3− 3 = 0,(A.78)
6 4 4 4 4
 
1 1 1 1 1 1
( D22 , D22 ) = 1+ + +1+ + = . (A.79)
6 4 4 4 4 2

They agree with the GOT, which, for S3 requires

1
( Dij , Dkl ) = δ δ (A.80)
2 ik jl

for any irreducible representation.


236 solutions

3. Again, we can check:


 
1 1 1 1
( D11 , D11 ) = 1+ + = , (A.81)
3 4 4 2
√ √ !
1 3 3
( D11 , D12 ) = 0− + = 0, (A.82)
3 4 4
√ √ !
1 3 3
( D11 , D21 ) = 0+ − = 0, (A.83)
3 4 4
 
1 1 1 1
( D11 , D22 ) = 1+ + = , (A.84)
3 4 4 2
 
1 3 3 1
( D12 , D12 ) = 0+ + = , (A.85)
3 4 4 2
 
1 3 3 1
( D12 , D21 ) = 0− − =− , (A.86)
3 4 4 2
√ √ !
1 3 3
( D12 , D22 ) = 0− + = 0, (A.87)
3 4 4
 
1 1 1 1
( D21 , D21 ) = 1+ + = , (A.88)
3 4 4 2
√ √ !
1 3 3
( D21 , D22 ) = 0+ − = 0, (A.89)
3 4 4
 
1 1 1 1
( D22 , D22 ) = 1+ + = . (A.90)
3 4 4 2

In this case the theorem fails because of (A.84) and (A.86).

chapter 7

1. A presentation for D4 is hρ, σ | ρ4 = σ2 = 1, σρ = ρ3 σ i. Its elements are

{ρ, ρ2 , ρ3 , ρ4 = 1 = σ2 , σ, σρ = ρ3 σ, σρ2 = ρ2 σ, σρ3 = ρσ}. (A.91)


solutions 237

We can then write the generic element as σm ρn , with m = 0, 1 and n =


0, . . . , 3. Conjugacy classes are determined by taking all the elements of
the form
σm ρn gρ−n σm (A.92)
for a given g ∈ D4 and m = 0, 1, n = 0, . . . , 3.
One obvious conjugacy class is
C1 = {1}, (A.93)
because the identity commutes with all other elements.
When g = ρ p we get
σρn ρ p ρ−n σ = σρ p σ = σ2 ρ− p = ρ− p . (A.94)
This means that p = 1 is conjugate to p = 4 − 1 = 3 and p = 2 to
p = 4 − 2 = 2:
C2 = {ρ, ρ3 }, C3 = {ρ2 }. (A.95)
For g = σρn we get that
σρm σρn ρ−m σ = σρm σ2 ρm−n = σρ2m−n . (A.96)
This means that when n = 0 we have conjugate elements σρ2m , for m =
0, 1, 2, 3:
C4 = {σ, σρ2 }. (A.97)
When n = 1 we have conjugate elements σρ2m−1 for m = 0, 1, 2, 3:
C5 = {σρ, σρ3 }. (A.98)

Obvious subgroups are Z4 = hρ | ρ4 = 1i, which is normal, and Z2 =


{1, σ}, which is not normal. Other Z2 subgroups are {1, σρ} (which
is not normal, because σ (σρ)σ = σρ3 is not in the subgroup), {1, σρ3 }
(again not normal, because σ(σρ3 )σ = σρ is not in the subgroup), {1, σρ2 }
(ρ(σρ2 )ρ3 = σ is not in the subgroup) and Z2 = {1, ρ2 }, which is normal
(it is given by the union of conjugacy classes).
We also have Z2 × Z2 generated by {1, σρ, σρ3 , ρ2 } or {1, σ, ρ2 , σρ2 }. These
are again union of conjugacy classes and hence normal.
Quotients: D4 /Z4 = Z2 = {1, σ }, D4 /Z2 = Z2 × Z2 = {1 ∼ ρ2 , σ ∼
σρ2 , ρ ∼ ρ3 , σρ ∼ σρ3 }, D4 /Z2 × Z2 = Z2 .
238 solutions

2. From the results of previous exercise we have 5 conjugacy classes and


hence 5 irreducible representations. We know from the lectures that there
is a 2-dimensional faithful representation
   
0 1 −1 0
D2 (ρ) = , D2 (σ ) = . (A.99)
−1 0 0 1

We also know that we always have the trivial representation D1 (ρ) =


D1 (σ ) = 1. We are then left with 3 more irreducible representations, so
that

∑ Na2 = 1 + 22 + 1 + 1 + 1 = 8 = #G. (A.100)

These follow from the determinant of the 2-dimensional representation:


D10 (ρ) = 1, D10 (σ ) = −1, while the other two are determined by the
fact that D (σ )2 = 1, which means that D (σ ) = ±1 and D (ρ) D (σ) =
D (σ ) D (ρ)3 , which means that also D (ρ)2 = 1. We find D1” (ρ) = −1,
D1” (σ ) = 1 e D1000 (ρ) = −1, D1000 (σ ) = −1.

3. The two representations are equivalent because

(χ D1 , χ D2 ) = 1. (A.101)

The explicit change of basis is given by the unitary matrix


 
11 −3
U = eiπ/3 . (A.102)
−7 2

4.

χ D1 ⊗ D2 ( g) = tr( D1 ⊗ D2 )( g) = ( D1 ⊗ D2 )( g)(ij)(ij) = D1 ( g)ii D2 ( g) jj

= trD1 ( g) trD2 ( g) = χ D1 · χ D2 .
(A.103)

5. If D is unitary D † D = 1. Then

χ D ( g−1 ) = tr D ( g−1 ) = tr D −1 ( g) = tr D † ( g) = tr D ∗ ( g) = (χ D ( g))∗ .


(A.104)
solutions 239

6. a) Irreducible representations are in 1 to 1 correspondence with conju-


gacy classes. One obvious conjugacy class is given by C1 = {1}. It
is straightforward to see that also −1 commutes with all the other
elements and hence C2 {−1}. By using ( J i )−1 = − J i and − J j J i J j =
2δij J i − J i (without summation over i), we also deduce that C2+i =
{± J i }. Hence we have 5 conjugacy classes.
Since #Q = 8 = ∑ a Na2 and a = 1, . . . , 5, we deduce that 8 = 1 + 1 +
1 + 1 + 22 . The first 1-dimensional representation is the trivial one
D1 (±1) = D1 (± J i ) = 1. All other representations must preserve the
relations in the presentation of the group and therefore D ( J i ) D ( J j ) =
δij D (−1) + eijk D ( J k ). This implies that D ( J 1 )2 D ( J 2 )2 = D (−1)2 =
D (1) = 1 and D ( J 1 ) D ( J 2 ) D ( J 1 ) = D ( J 2 ), which gives D ( J 1 )2 = 1
and therefore D ( J 1 ) = ±1 and D ( J 2 ) = ±1. The three possibilities
are then
D (±1) = D (± J 1 ) = 1, D (± J 2 ) = D (± J 3 ) = −1, (A.105)

D (±1) = D (± J 2 ) = 1, D (± J 1 ) = D (± J 3 ) = −1, (A.106)

D (±1) = D (± J 3 ) = 1, D (± J 2 ) = D (± J 1 ) = −1. (A.107)

The 2-dimensional representation follows from the suggested matrix,


by choosing
   
1 0 −1 0
D (1) = U (0, 0, 0) = , D (−1) = U (π, 0, 0) = ,
0 1 0 −1
(A.108)

   
1
π  i 0 1
 π  −i 0
D ( J ) = U 0, , 0 = , D (− J ) = U 0, − , 0 = ,
2 0 −i 2 0 i
(A.109)

    0 
2
π  0 1 2
 π −1
D( J ) = U , 0, 0 = , D (− J ) = U − , 0, 0 = ,
2 −1 0 2 1 0
240 solutions

(A.110)

   
π π 0 i π π 0 −i
D( J 3 ) = U , 0, = , D (− J 3 ) = U , 0, − = .
2 2 i 0 2 2 −i 0
(A.111)

b) The character table is straightforward from the traces of the matrices


given above:
C1 C2 C3 C4 C5
1 1 1 1 1 1
10 1 1 1 −1 −1
(A.112)
100 1 1 −1 1 −1
1000 1 1 −1 −1 1
2 2 −2 0 0 0

7. a) A group N < G is a normal subroup iff gNg−1 ∈ N, ∀ g ∈ G. From


the conjugacy classes we have that ga3 g−1 = a3 and hence Z2 is
normal. Also Z3 is given by the union of conjugacy classes and
hence is normal. Z4 is not normal: gbg−1 gives a2 b and e a4 b, which
are not part of the group
b) Since Z3 is normal, left and right quotients coincide. In this case the
quotient is
Γ( Q6 )/Z3 = Z4 = {1 ∼ a2 ∼ a4 , b ∼ a2 b ∼ a4 b, b2 = a3 ∼ a5 ∼ a,

b3 = a3 b ∼ a5 b ∼ ab}. (A.113)
c) All elements of the form an b are off-diagonal and hence χ1 ( an b) =
χ2 ( an b) = χ3 ( an b) = χ4 ( an b) = 0. We then need only to com-
pute the characters of an element for each of the conjugacy classes
C2 , C3 and C4 . We have that χ1 ( a) = χ4 ( a) and χ2 ( a) = χ3 ( a),
χ1 (b2 = a3 ) = χ4 (b2 = a3 ) = −2 and χ2 (b2 = a3 ) = χ3 (b2 = a3 ) = 2,
χ1 ( a2 ) = χ4 ( a2 ) = 2 cos(2π/3) and χ2 ( a2 ) = χ3 ( a2 ) = 2 cos(4π/3).
This implies that the characters of D1 and D4 are identical and there-
fore they are equivalent representations. The same is true for D2 and
D3 . We can also check that (χ1 , χ2 ) = 0:
1
( χ1 , χ2 ) = 12 (1 · 2 · 2 + 2 · 2 cos π/3 · 2 cos(2/3π )
(A.114)
2 · 4 cos(2/3π )2 + (−2)(2) = 0.

+
solutions 241

chapter 8

1. a) The tableaux are:

• which corresponds to the conjugacy class C1 of the


identity, with a unique element. The corresponding representa-
tion has dimension 1;

• , representing the conjugacy class C2 of 4-cycles. It contains


6 elements and the corresponding representation has dimension
10 ;

• , representing double 2-cycles. There are 3 such elements


in C3 . The dimension of the corresponding irreps is 2;

• , representing the 8 3-cycles in C4 . The dimension of the


corresponding irreps is 3;

• , representing 6 2-cycles in C5 . The dimension of the


corresponding irreps is 30 .
We check easily that #G = 4! = 24 = 1 + 1 + 22 + 32 + 32 .
b) The column C1 and the first row, given by χ(1), are straightforward
to fill:
C1 C2 C3 C4 C5
1 1 1 1 1 1
10 1
2 2
3 3
30 3
242 solutions

c) The alternating group selects elements that can be written with an


even number of 2-cycles. 4-cycles come from multiplication of 3 2-
cycles (for instance (1234) = (12)(23)(34)) and therefore elements
in C2 are odd. The same is true for C5 , which contains single 2-
cycles. The other classes contain elements which appear from an
even number of 2-cycles.
We then fill the table accordingly:

C1 C2 C3 C4 C5
1 1 1 1 1 1
10 1 −1 1 1 −1
2 2
3 3
30 3

d) We can show that (χ2⊗10 , χ2⊗10 ) = 1 and it must be an irreducible


representation of dimension 2. Since there is only one such represen-
tation it must be equivalent to the one we already have. The result
follows from χ2⊗10 ( g) = χ2 ( g)χ10 ( g):

( χ2⊗10 , χ2⊗10 ) = 1
12 ∑ g (χ2⊗10 )∗ ( g)(χ2⊗10 )( g))
(A.115)
1 1
= 12 ∑ g |χ2 ( g)|2 · |χ10 ( g)|2 = 12 ∑ g |χ2 ( g)|2 = (χ2 , χ2 ),

because χ10 ( g) = ±1, for any g ∈ S4 .


If 2 ⊗ 10 ' 2, then χ2 χ10 = χ2 and therefore it must be vanishing for
C2 and for C5 :
C1 C2 C3 C4 C5
1 1 1 1 1 1
10 1 −1 1 1 −1
2 2 0 0
3 3
30 3

e) Orthonormality tells us that (χ1 , χ2 ) = 0 = 24 1


(χ2 (1) + 2 χ2 (C3 ) +
8 χ2 (C5 )) and (χ2 , χ2 ) = 1 = 24 (χ2 (1) + 3 χ2 (C3 )2 + 8 χ2 (C4 )2 ). If
1 2
solutions 243

only solutions with integer numbers can be accepted we fix χ2 (C4 ) =


−1 and χ2 (1) = χ2 (C3 ) = 2:

C1 C2 C3 C4 C5
1 1 1 1 1 1
10 1 −1 1 1 −1
2 2 0 2 −1 0
3 3
30 3

f) By the same line of reasoning as the one above 3 ⊗ 10 is an irreducible


representation and therefore it must be either equivalent to 3 or to
30 . If we assume that it is equivalent to 3 then we should again
conclude that we have a 0 in correspondence with the characters of
the elements in C2 and C5 , but then it would be impossible to fulfill
at the same time the condition following from (χ3 , χ3 ) = 1 and those
from (χ3 , χ1 ) = (χ3 , χ2 ) = 0. We then conclude that 3 ⊗ 10 ' 30 .
g) Writing down the system of orthonormality equations:

C1 C2 C3 C4 C5
1 1 1 1 1 1
10 1 −1 1 1 −1
2 2 0 2 −1 0
3 3 1 −1 0 −1
30 3 −1 −1 0 1

We could exchange 3 and 30 obtaining again a consistent matrix.


244 solutions

You might also like