Download as pdf or txt
Download as pdf or txt
You are on page 1of 104

MPH-001

MATHEMATICAL
Indira Gandhi National
Open University
METHODS IN PHYSICS
School of Sciences

Block

5
GROUP THEORY
UNIT 16
Sets and Mappings 7
UNIT 17
Groups 21
UNIT 18
Matrix Representations 45
UNIT 19
Continuous Groups 69
Programme Design Committee
Prof. V.B. Bhatia, Retd. Prof. Enakshi Sharma Prof. G. Pushpa Chakrapani Prof. Suresh Garg, Retd.
University of Delhi, Delhi University of Delhi, South BRAOU School of Sciences,
Prof. Abhai Mansingh, Retd. Campus, Delhi Prof. Y.K. Vijay IGNOU, New Delhi
University of Delhi, Delhi Prof. H.S. Mani, Retd. University of Rajasthan, Prof. Vijayshri
Prof. Feroz Ahmed, Retd. IIT Kanpur Rajasthan School of Sciences,
University of Delhi, Delhi Prof. S. Annapoorni Prof. J. Nag, Retd. IGNOU, New Delhi
Prof Yashwant Singh, Retd. University of Delhi, Delhi Jadavpur University Prof. S.R. Jha
Banaras Hindu University, Prof. D. Choudhury Prof. Zulfequar, School of Sciences,
Varanasi University of Delhi, Delhi Jamia Milia Islamia, New Delhi IGNOU, New Delhi
Prof. Deepak Kumar Prof. T.R. Seshadri Dr. Om Pal Singh Prof. Shubha Gokhale
J.N.U., New Delhi University of Delhi, Delhi IGCAR, Kalpakkam, School of Sciences,
Tamil Nadu IGNOU, New Delhi
Prof. Vipin Srivastava Prof. S. Ghosh
Central University of J.N.U., New Delhi Prof. Prabhat Munshi Prof. Sanjay Gupta
Hyderabad, Hyderabad IIT Kanpur School of Sciences,
Prof. Neeraj Khare
IGNOU, New Delhi
Prof. G.S. Singh IIT Delhi, Delhi Prof. R.M. Mehra, Retd.
IIT Roorkee, Roorkee Dept. of Electronics, South Dr. Subhalakshmi Lamba
Prof. V.K. Tripathi
Campus, University of Delhi, School of Sciences,
Prof. A.K. Rastogi. IIT Delhi, Delhi
Delhi IGNOU, New Delhi
J.N.U., New Delhi Prof. Pankaj Sharan, Retd.
Prof. A. K. Ghatak Jamia Milia Islamia, Prof. S.K. Kulkarni Dr. M.B. Newmai
IIT Delhi, Delhi New Delhi Pune University/ School of Sciences,
IISER Pune, Pune IGNOU, New Delhi
Prof. Rupamajari Ghosh Prof. Kirti Ranjan
J.N.U., New Delhi University of Delhi, Delhi

Course Design Committee


Prof. V.B. Bhatia, Retd. Prof. G.S. Singh Prof. Neeraj Khare Prof. Vijayshri
University of Delhi, Delhi IIT Roorkee, Roorkee IIT Delhi, Delhi SOS, IGNOU, New Delhi
Prof. Abhai Mansingh, Retd. Prof. A.K. Rastogi Prof. Pankaj Sharan, Retd. Prof. S.R. Jha
University of Delhi, Delhi J.N.U., New Delhi Jamia Milia Islamia, New Delhi SOS, IGNOU, New Delhi
Prof. Feroz Ahmed, Retd. Prof. S. Annapoorni Prof. Kirti Ranjan Prof. Shubha Gokhale
University of Delhi, Delhi University of Delhi, Delhi University of Delhi, Delhi SOS, IGNOU, New Delhi
Prof Yashwant Singh, Retd. Prof. D. Choudhury Prof. R.M. Mehra, Retd. Prof. Sanjay Gupta
BHU, Varanasi University of Delhi, Delhi Dept. of Electronics, South SOS, IGNOU, New Delhi
Prof. Deepak Kumar Prof. T.R. Seshadri Campus, Delhi University, Dr. Subhalakshmi Lamba
J.N.U., New Delhi University of Delhi, Delhi Delhi SOS, IGNOU, New Delhi
Prof. Vipin Srivastava Prof. S. Ghosh Prof. Suresh Garg, Retd. Dr. M.B. Newmai
Central University of J.N.U., New Delhi SOS, IGNOU, New Delhi SOS, IGNOU, New Delhi
Hyderabad, Hyderabad

Block Preparation Team


Prof. Pankaj Sharan, Retd. (Units 16-19) Dr. M.B. Newmai (Units 16-19)
Jamia Milia Islamia, New Delhi School of Sciences, IGNOU, New Delhi
Course Coordinators: Prof. Sanjay Gupta, Dr. Subhalakshmi Lamba
Block Production Team
Sh. Rajiv Girdhar
AR (P), IGNOU
Acknowledgement: Shri Gopal Krishan Arora, EDP, SOS for CRC preparation.
October, 2023
© Indira Gandhi National Open University, 2023
ISBN:
Disclaimer: Any materials adapted from web-based resources in this module are being used for educational purposes
only and not for commercial purposes.
All rights reserved. No part of this work may be reproduced in any form, by mimeograph or any other means, without
permission in writing from the Copyright holder.
Further information on the Indira Gandhi National Open University courses may be obtained from the University’s office at
Maidan Garhi, New Delhi-110 068 or the official website of IGNOU at www.ignou.ac.in.
Printed and published on behalf of Indira Gandhi National Open University, New Delhi by Prof. Meenal Mishra, Director,
SOS, IGNOU.
Printed at
CONTENTS
Block and Unit Titles 1
Credit Page 2
Contents 3
BLOCK 5: GROUP THEORY 5
Unit 16 Sets and Mappings 7
16.1 Introduction 7
16.2 Sets 8
16.2.1 Subsets 9
16.2.2 Intersection, Union and Complement of Sets 10
16.2.3 Venn Diagrams 11
16.2.4 Identities 12
16.2.5 Cartesian Product 12
16.2.6 Relations 13
16.2.7 Equivalence Classes 14
16.3 Mappings 15
16.3.1 Definition 15
16.3.2 Examples of Mappings 16
16.3.3 Range and Domain of a Mapping 16
16.3.4 ‘Into’ ‘Onto’ and ‘One-to-one’ Mappings 17
16.3.5 Composition of Mappings 17
16.3.6 Direct and Inverse Image of a Mapping 17
16.3.7 Inverse Mapping 18
16.4 Summary 19
16.5 Terminal Questions 19
16.6 Solutions and Answers 19
Unit 17 Groups 21
17.1 Introduction 21
17.2 Introduction to Groups 22
17.2.1 Definition of a Group 23
17.3 Group Multiplication Tables 24
17.3.1 Commutative or Abelian Groups 25
17.4 Subgroups 25
17.5 Permutation Group or Symmetric Group Sn 26
17.5.1 Geometrical Illustration of S3 28
17.6 Transpositions and Even and Odd Permutations 30
17.7 Cosets of a Subgroup 31
17.8 Cyclic Subgroups 33
17.8.1 Cyclic Subgroups of the Permutation Group 33
17.9 Homomorphisms 34
17.9.1 Isomorphism 35
17.10 Conjugation 36
17.11 Invariant Subgroups 37
17.12 Summary 38
17.13 Terminal Questions 39
17.14 Solutions and Answers 39
Unit 18 Matrix Representations 45
18.1 Introduction 45
18.2 Matrix Representations 46
18.2.1 2  2 Matrix Representation of D3 46
18.2.2 3  3 Matrix Representation of S3 49
18.3 Equivalent Representations 50
18.4 Reducible and Irreducible Representations 51
18.5 Group Representation by Linear Operators 51
18.6 Example of Reducing a Representation of S3 53
18.7 Some Definitions and Results 57
18.7.1 Every Representation of a Finite Group can be made
Unitary by a Similarity Transformation 57
18.7.2 Schur’s Lemma 59
18.8 Summary 61
18.9 Terminal Questions 61
18.10 Solutions and Answers 61
Unit 19 Continuous Groups 69
19.1 Introduction 69
19.2 Continuous Groups: An Introduction 70
19.2.1 Active and Passive Transformations 70
19.2.2 Simplest Continuous Groups 71
19.2.3 Orthogonal Group O(3) 72
19.2.4 The Subgroup SO(3) or Rotation Group 73
19.2.5 Lie Groups or Continuous Groups 73
19.2.6 One Parameter Subgroups of SO(3) 74
19.3 Generators of One-parameter Subgroups 75
19.3.1 Commutator of Generators 77
19.3.2 Lie Algebra of Generators 78
19.4 Minkowski Space and the Lorentz Group 81
19.4.1 Minkowski Space 83
19.4.2 The Lorentz Group 84
19.5 SL(2,C) and the Lorentz Group L 85
19.5.1 Homomorphism SL(2,C)  L†
 is 2 : 1 89
19.5.2 Six Parameters of SL(2,C) 90
19.5.3 Boosts and Hermitian SL(2,C) Matrices 90
19.5.4 Rotations and SU(2) 92
19.5.5 Form of a General SL(2,C) Matrix 94
19.5.6 The Matrix  and other Useful Formulas 95
19.6 Summary 96
19.7 Terminal Questions 96
19.8 Solutions and Answers 97

4
BLOCK 5 : GROUP THEORY
In the intricate world of science, sets, matrices, and groups emerge as indispensable
mathematical tools, offering profound insights and applications across various realms of
physics. In quantum mechanics, the state of a quantum system is described by a complex
vector in a Hilbert space, which is essentially a set of vectors. The concept of sets is crucial
for understanding the possible states and their probabilities. Similarly, matrices may be
employed to represent transformations in physics, such as rotations, reflections, and
translations. For example, in solid-state physics, matrices describe the rotation of crystal
lattices and their symmetry properties.

In Unit 16 we will revise the concepts of sets, subsets, intersection and union of sets,
complement, Venn diagrams and present some identities and relations. We then define
Cartesian product, relations and equivalence classes. We also discuss mappings and related
definitions and concepts such as range, domain, into, onto and one-to-one mappings and
composition of mappings.

In Unit 17, we discuss groups and related concepts like group multiplication tables,
subgroups, permutation groups. Then we discuss transpositions, even and odd permutations,
cosets of a subgroup, cyclic subgroups and Homomorphisms. In the last two sections, we
take up conjugation and Invariant Subgroups.

In Unit 18 we will discuss matrix representation. We also discuss equivalent representations,


reducible and irreducible representations, group representation by linear operators and give
an example of reducing a representation of S3.

In Unit 19 we discuss continuous groups in physics which are related to symmetry


transformations. We introduce continuous groups and discuss some simple continuous
groups and study the generators of one-parameter subgroups, Minkowski space, and the
Lorentz group.

We strongly encourage you to engage in a thorough process of working through different


steps, practicing with solved examples, tackling SAQs, and independently solving terminal
questions. We wish you good luck.

5
6
Unit 16 Sets and Mappings

UNIT 16
SETS AND MAPPINGS
Structure
16.1 Introduction 16.3 Mappings
Expected Learning Outcomes Definition
16.2 Sets Examples of Mappings
Subsets Range and Domain of a Mapping
Intersection, Union and Complement ‘Into’, ‘Onto’, and ‘One-to-one’ Mappings
of Sets Composition of Mappings
Venn Diagrams Direct and Inverse Image of a Mapping
Identities Inverse Mapping
Cartesian Product 16.4 Summary
Relations 16.5 Terminal Questions
Equivalence Classes 16.6 Solutions and Answers

16.1 INTRODUCTION
In this unit, we introduce some ideas of modern mathematical language which
you may already have encountered. We use the word ‘modern’ in the same
sense as ‘modern’ physics, which is supposed to refer to more than a century
old ideas. Similarly, the word ‘modern mathematics’ refers to centuries old
ideas of many mathematicians.
In Sec. 16.2, we will revise the concepts of sets, subsets, intersection and
union of sets, complement, Venn diagrams and present some identities and
relations. We will also define Cartesian product, relations and equivalence
classes. Sec. 16.3, we discuss mappings and related definitions and concepts
such as range, domain, into, onto and one-to-one mappings and composition
of mappings. We also discuss direct and inverse image of a mapping and
inverse mapping. In the next unit, you will learn about groups.
Expected Learning Outcomes
After studying this unit, you should be able to:

 define sets, subsets, intersection and union of sets, complement of sets;


 draw Venn diagrams;
 apply identities pertaining to sets;
 define Cartesian product, relations and equivalence classes;
 solve problems based on set theory; 7
Block 5 Group Theory
 define mappings and give examples;
 define range and domain of a mapping, and ‘into’, ‘onto’ and ‘one-to-one’
mappings, and composition of mappings;
 define direct and inverse image of a mapping, and inverse mapping; and
 solve problems based on mappings.

16.2 SETS
Let us begin by reviewing the basic facts and notation about sets. As you may
know, a set is a collection of objects, which are distinct and distinguishable.
The objects which belong to a set are called its elements. If an object named
x belongs to a set S then we write
xS
which we read as “x belongs to S” or “x is in S”. If we want to express that
some object y does not belong to a set S, we write
yS
It is a general practice to write the elements, or description of elements, within
curly brackets { } when defining a set.
Note the words ‘distinct’ and ‘distinguishable’. The collection which contains
three integers 1, 2, 3 but with one of them repeated, for example
S = {1, 2, 2, 3}
is not a set.
The ‘objects’ of a set can be of any type. For example, sets themselves can be
objects, and therefore elements of some other set.
It is important to understand the conceptual difference between an element
and a set. Let x be an object, then a set containing this single object
X = {x}
is a set. And, a set containing the set X, is
Y = {X} = { {x} }
Here Y is another set and although,
x  X = {x} but x  {x }

We give here some notations of useful sets:

Set Notation Description

R the set of all real numbers

Z the set of all integers, positive, negative and zero

Z0 the set of all positive integers, not including zero

C the set of all complex numbers


8
Unit 16 Sets and Mappings
Very often a set is described by a common property that its elements share.
Let R  be the set of all real numbers which are positive. Then we write:

R   x x  R, x  0

Notice the use of curly bracket and the vertical bar in the notation above.
If there are only a finite number of elements, we can write their names within
curly brackets separated by commas. For example,
S  n n  Z0 , n  5  {1, 2, 3, 4, 5}

It is important to know that when a set is written like this the order in which
the elements of the set are written does not matter. It is the same set
irrespective of the order
S  {1, 2, 3, 4, 5}  {5, 2, 4, 3,1}  {3, 2,1, 4, 5}

It is useful to define a special set, denoted by  and called the empty set, which
has no elements.

Example 16.1

Let us give the definition and standard notation of some sets. Suppose that R
is the set of all real numbers. Then

(0,1)  { x  R 0  x  1}
(0,1]  { x  R 0  x  1}
[0,1)  { x  R 0  x  1}
(0,1]  { x  R 0  x  1}

A set like (0, 1) is called an open interval between 0 and 1, and [0, 1], the
closed interval. Sets of the type (0, 1] or [0, 1) are called semi open or semi
closed intervals. We can define such intervals between any two real numbers
x1 and x2 (with x1 < x2 ):

( x1, x 2 )   x  R x1  x  x 2 ,

and so on.

16.2.1 Subsets
If all the elements of a set T are also members of another set S then we say
that the set T is a subset of S and we write it as:

TS

and read it as “T is contained in S” or simply “T is a subset of S”. When


convenient, we can also write it as:

ST

which we read as “S contains T”.

Note that it is always true that a set is its own subset: S  S. Also, the empty
set is the subset of every set: If S is a set then   S. 9
Block 5 Group Theory
Using the names of sets defined above, we see that

Z0  Z  R  C, R  R

There is another symbol used in set theory: X  Y . It also read as “X is a


subset of Y”. Similar to the symbol x  y between two real numbers where x
could be less than or equal to y, the symbol X  Y also includes the
possibility of the two sets being the same or equal. When Y contains an
element that is not an element of X we say that X is a proper subset of Y. But
remember that unlike x  y where x is strictly less than y, the symbol
X  Y does not always mean that there is an element which is not an element
of X. therefore it is safest to treat X  Y and X  Y as equivalent statements,
and when a set X is a proper subset of the set Y, then that is to be explicitly
mentioned.

16.2.2 Intersection, Union and Complement of Sets


Given two sets A and B, the set which contains all elements common to both A
and B is called the intersection of the two sets. It is denoted by:
AB
It is clear from the definition that
AB=BA
It may happen that there is no element which is both in A and in B. In that
case, the intersection of the two sets is the empty set:
AB=
If A  B = , we say that A and B are disjoint or mutually exclusive.
If B is a subset of A then B  A = B.
Given two sets A and B, the set which contains elements which either belong
to A or to B is called the union of the two sets. It is denoted by A  B. It is like
joining the two sets without repeating the elements which may be common to
both. This process is symmetric:
AB=BA
By using the process of union or intersection we can keep obtaining new sets.
For example, if A, B and C are three sets we can form
(A  B)  C
We could also have constructed
A  (B  C)
but that is the same, since
A  (B  C) = (A  B)  C
Therefore, since the order of taking unions does not matter, we just write
ABC
Similarly, we can construct sets by repeated intersections:

10
A  (B  C) = (A  B)  C = A  B  C
Unit 16 Sets and Mappings
The process for union and intersections of sets can continue for any number
of times, evaluated in any order: if A1, A2,..., An are any sets,

A1  A2 ...  An

or
A1  A2 ...  An

are the union or intersections of these sets. We write these as


ni1 Ai  A1  A2  ...  An

Similarly,
ni1 Ai  A1  A2  ...  An

As a matter of fact these sets can even be infinite in number. We shall give
some examples later on.
Example 16.2
If A = {a, b, c, d}, B = {c, e, g, d, h}
then
A  B = {a, b, c, d, e, g, h}, and A  B = {c, d}

Example 16.3
(0, 0)  

(0, 1)  (1, 2)  

[0, 1]  [1, 2]  {1},


n 1 [0, 1  1/ n ]  [0, 1], n is an integer,


n 1 ( 1/ n,  1/ n )  {0}, n is an integer,

 n1 ( 1/ n,  1/ n )  ( 1,  1), n is an integer.

Complement
Let A and B be two sets. The set which contains all elements of A which do
not belong to B is called the complement of B with respect to A. We can
denote this set by
A – B = {x  Ax  B}

16.2.3 Venn Diagrams


In order to visualize the relations between sets we show sets as circles,
rectangles (or any closed shapes) and place them so that the common or
overlapping parts show the sets corresponding to intersections and the outer
boundary shows the union. In the following diagrams the sets A  B and
A  B are shown by the thick lines:
11
Block 5 Group Theory

B B

AB

AB
A A

Fig. 16.1: Venn diagrams.

16.2.4 Identities
The following results are well known identities related to sets. If A, B, C are
sets, then

AA=A

AA=A

A–A=

A  (B  C) = (A  B)  (A  C)

A  (B  C) = (A  B)  (A  C)

SAQ 1

Make the Venn diagrams for the following sets, starting with A, B and C:

a) A – B,
b) A  B  C

16.2.5 Cartesian Product

We first define the ordered pair of two objects a and b as the object (a, b). It
is called ordered because the position of a and b cannot be interchanged.

Give two sets X and Y we can define a set, called the Cartesian product of X
and Y and denoted by X  Y as the set whose elements are all ordered pairs
(x, y) where x  X and y  Y:

X  Y  ( x, y ) x  X , y  Y 

Example 16.4

The Cartesian plane (a two dimensional plane with orthogonal axes) is the
cartesian product of the set R of real numbers with itself. It is written as
R 2  R  R.

12
Unit 16 Sets and Mappings

16.2.6 Relations
A relation R defined on a set X is a subset of the Cartesian product of X with
itself:
R  X X
If the ordered pair (a, b)  R, we say that
a is related to b by R,
or say that,
a R b is true.
Note that if a is related to b by R, it may or may not be true that b is related to
a.

Example 16.5

Let R be the set of all real numbers. We define the relation R  R × R called
‘less than’ as the subset

R = {(a, b)a, b  R, a < b}.

An equivalence relation is a relation R on a set X which satisfies the


following three conditions:

1. x R x for every x  X, (we say R is reflexive)

2. if x R y then y R x (R is symmetric)
3. if x R y and y R z then x R z, (R is transitive)

SAQ 2

a) Let X = {x, y, z} be a set with three elements. Define

R = {(x, x), (y, y), (z, z), (x, y), (y, x)}.

Is R an equivalence relation?

b) In the definition of the equivalence relation given above, do requirements


2 (symmetry) and 3 (transitivity) not already imply 1 (reflexivity)? Why then
is there a need to assume 1 separately?

For small sets it is useful to write the pairs (x, y) for which x R y, in the form of
a matrix leaving other spaces blank. For example, we can write R of SAQ 2a
as

( x, x ) ( x, y ) 
 
 
R  ( y , x ) ( y , y ) 
 
 
 ( z, z )
13
Block 5 Group Theory
16.2.7 Equivalence Classes
Given an equivalence relation defined on a set X, we can define subsets of X
such that every member of a subset is related to every other member of the
same subset, and not related to any member of the other subsets.

Example 16.6

Let a relation R be the following subset of X  X where X = {x, y, z, u, v, w}

R = {(x, x), (y, y), (z, z), (u, u), (v, v), (w, w),

(x, z), (z, x), (y, u), (u, y), (x, w), (w, x)}

The equivalence classes are

X1  {(x, x), (x, z), (z, x), (x, w), (w, x), (z, z), (w, w)}

X 2  {(y, y), (u, u), (y, u), (u, y)}

X 3  {(v, v)}

The equivalence classes look even clearer in the following matrix form if we
arrange the elements properly:

 ( x, x ) ( x, z) ( x, w ) 
 ( z, x ) ( z, z) 
 
( w, x ) ( w, w ) 
 
 ( y, y ) ( y,u) 
 (u, y ) (u,u) 
 
 ( v, v )

Example 16.7

On the set of all integers Z, we define a relation P called ‘parity’ as follows:


mPn if (m – n) is divisible by 2. The two equivalence classes of Z are the
subsets of even or odd integers.

SAQ 3

Which of the following relations are equivalence relations? Identify the


equivalence classes, if so.

Z is the set of all integers positive, negative, and zero.

a) i, j  Z and iRj if and only if i j  0.

b) i, j  Z and iRj if and only if i – j is a multiple of 5.


14
Unit 16 Sets and Mappings

16.3 MAPPINGS
Let us begin by defining mapping.

16.3.1 Definition
A function or a mapping f from a set X into a set Y is a rule by which every
element of X is assigned a unique element of Y. It is important to note that
there is only one element of Y which is assigned to a given element of X. Of
course, there may be more than one element of X which are assigned the
same element of Y.
Example 16.8
Two simple examples of mappings are: the identity mapping I X of a set X
into itself maps every element x into itself I X ( x )  x. The constant mapping
maps every element of x into a same fixed element of Y.
Note the use of the technical word ‘into’.
We write the function or mapping as
f:XY
and the value assigned to the element x  X as
f(x)  Y
We also write
x  f(x)
to say the same thing. The assigned element f(x) is also called the image of x,
or value of the function at x. Quite often, when the sets X and Y are already
understood, and need not be specified, it is enough to write the mapping as
x  f (x ). For example, if it is already known and we are defining real
functions of a real variable, then writing x  x 2 is sufficient.

Remark about Notation


Very often we see statements like ‘f(x) is a function of x’, or ‘sin(x) is a periodic
function’. These statements have become standard with use. But they are not
strictly accurate. Actually, the function is f, and f(x) is the value which the
function f assigns to the variable x. Similarly, the function sin is periodic, and
its value at x is sin (x).
But what about the function sq.: R  R defined from x  R into R for which the
rule is ‘multiply x with itself, the number obtained is the value sq (x).
By common understanding the function sq is written as x 2 . Here the
distinction between the function and its values is no longer there. But this is for
convenience of writing for these monomial functions.
It is important to keep in mind the conceptual distinction between a function or
mapping f : X  Y and its values. The values belong to Y, while the function f
belongs to neither X nor Y. It is an entity in itself providing a relationship
between the two sets.
It is also important to realize that there are no double-valued or multi-valued
functions. The function written simply as x is only defined as a mapping

 : {0}  R    {0}  R   : x  x 15
Block 5 Group Theory
Here 
{0}  R   is the set including zero and all positive numbers. There is
another function  x which maps

  : {0}  R    {0}  R   : x   x

where {0}  R   is the set including zero and all negative real numbers.
Because both these functions satisfy  x  2  x , they are often considered
as a single function with two values. That is not correct.

Let us now consider a few examples of mappings.

16.3.2 Examples of Mappings


Functions or mappings do not have to be confined to the real or complex
variables although they are the most common in applications. Here are some
other examples:

1. Let X be a set and A, B etc. its subsets and P(X) as the set of all subsets
of X including itself and the empty set. Then there is a mapping defined on
P(X) which maps the subset A  X to its complement X – A.

2. Let X  Y be the cartesian product of two sets X and Y. The mapping pr1
and pr2 called projections to the first and second factor are defined as:

pr1  X  Y  X , pr1( x, y )  x

pr2  X  Y  Y , pr2 ( x, y )  y

16.3.3 Range and Domain of a Mapping


In a mapping f : X  Y from the set X into Y, each element of the set X is
mapped to some unique element, that is its image f(x)  Y. But it is not
necessary that every element of the Y is the image of some element of X. The
subset R  Y consisting of all those elements of Y which are the image of
some element of X is called the range of the mapping f:

R  y  Y y  f ( x ) for some x  X .

The way we have defined a mapping f : X  Y, all elements of the set X are
mapped by f. Sometimes it is more convenient to choose a subset D  X and we
chose to only map D into Y : f : D  Y. In that case, D is called the domain of
the mapping f.

Example 16.9

In the mapping of real numbers to real numbers R  R : x  x 2 the range is


the set of all positive numbers along with zero: range = R   {0}.

SAQ 4

What is the range of the function R  R given by x  e x ?


16
Unit 16 Sets and Mappings

16.3.4 ‘Into’ ‘Onto’ and ‘One-to-One’ Mappings


For a mapping f : X  Y, we generally say that the mapping f maps X into Y.

If the range of the mapping f : X  Y is equal to the whole set Y, we say that the
mapping is onto or surjective.

Also, in general, it is possible that more than one element of X get mapped to the
same element of Y. But if each element x  X is mapped to a separate element
of Y then we say the mapping is one-to-one or injective. For one-to-one
mapping, the statement f ( x1)  f ( x2 ) implies that x1  x 2 .

A mapping or a function which is both one-to-one and onto is called one-to-one


correspondence or a bijective mapping. The inverse mapping or function can
only be defined for a bijective mapping.

Example 16.10

The function x  x 2 from R  R is not one-to-one because x and – x both


map to the same value.

For function x  sinx, for a fixed x, an infinite number of elements x + 2n


where n is an integer, map to the same value sin x.

The function x  e x is one-to-one, but not onto because its range is only
positive real numbers. The function x  x 3 is one to one and onto.

16.3.5 Composition of Mappings


Let f : X  Y and g  Y  Z be two mappings. Then we can define a mapping
g o f which takes an element from X to Z by first applying f and then g:

(g o f) : X  Z, (g o f) (x) = g (f(x))

The composition of mappings is the same as the idea of a function of a function.

Example 16.11

Let f : R  R be defined as f : x  y = e  x , and g : R  R as g : y  z = y2


then (g o f) : r  z = e 2x .

16.3.6 Direct and Inverse Image of a Mapping


For a mapping f : X  Y and a subset A  X, the direct image or just image f (A)
is the subset of Y which consists of all the elements which are mapped by
elements of A:

f(A) = {f(x)x  A}.

For a mapping f : X  Y, the inverse image f 1(B) of a subset B  Y is the set


of all x  X such that f(x)  B. We write

f 1(B )  x  X f ( x )  B 17
Block 5 Group Theory
If the range R of f is not all of Y then for those B which are outside the range of f,
that is, B  R = , the inverse image is just the empty set .

Example 16.12
The concept of inverse image is used in defining continuity of a function
f : R  R by the traditional  ,  method. We say that a function f is continuous
at x  a if f(x) can be made to differ from f(a) by not more than  for a chosen
 as small as we like, provided x is chosen close enough to a. Or, we say
there exists a  such that f ( x )  fa   if x  a  .

Let T be the subset T  f (a)   , f (a)    and S = (a – , a + ). The


definition of continuity is then that for every choice of T there is an S such that
f 1(T )  S.

It is well to remember the following rules about direct and inverse images.

For f : X  Y and subsets of X and Y,

f ( A1  A2 )  f ( A1)  f ( A2 ) (16.1)

f ( A1  A2 )  f ( A1)  f ( A2 ) (16.2)

f ( A1  A2 )  f ( A1)  f ( A2 ) , (16.3)

and

f 1(B1  B2 )  f 1(B1)  f 1(B2 ) (16.4)

f 1(B1  B2 )  f 1(B1)  f 1(B2 ) (16.5)

f 1(B1  B2 )  f 1(B1)  f 1(B) (16.6)

Notice that for the inverse images all the three relations are equality, whereas for
the direct images, only union of forward images is equal to the forward image of
the union.

SAQ 5

Find examples to verify that Eqs. (16.2) and (16.3) may not be equalities.

16.3.7 Inverse Mapping


If the mapping f is one-to-one correspondence, that is, one-to-one and onto, or
bijective, then it is possible to define the inverse mapping, because there is
exactly one element x  X for every element y  Y. We can then define for each
y  f (x ) , the inverse mapping.

f 1 : Y  X, f 1( y )  x

It is clear that both f o f 1 and f 1 o f are the identity mappings of sets Y and X
18 respectively.
Unit 16 Sets and Mappings

Example 16.13
The simple mapping x  y  x 2 from real numbers R to real positive
numbers R  has no inverse because for every value of y there are two
inverse images. The mapping is not one-to-one.

16.4 SUMMARY
In this unit, you have learnt about
 sets, subsets, intersection, union and complement of sets, Venn
diagrams, identities, Cartesian product, relations and equivalence
classes; and
 mappings, definition, example of mappings, range and domain of a
mapping, ‘Into’, ‘Onto’, and ‘One-to-one’ mappings, composition of
mappings, direct and inverse image of a mapping and inverse
mapping.

16.5 TERMINAL QUESTIONS


1. Make the Venn diagram for the sets.

A  (B  C), A  (B  C) starting with A, B and C and visually verify the


de Morgan laws:
A  (B  C) = (A  B)  (A  C)
A  (B  C) = (A  B)  (A  C).
2. What is the range of the function R  R given by x  sin( x ).

3. Is the mapping f : R  R so that x  x 2  x a one-to-one mapping or an


onto mapping? What is the inverse image of the set {0} and the direct image
of the set [ 1, 1]? (Hint: It will help to draw a graph of the function.)

16.6 SOLUTIONS AND ANSWERS


Self-Assessment Questions
1. See Figs. 16.2a and b for (a) and (b):

B C

A–B
A
D=ABC
A

(a) (b)

Fig. 16.2: Solution of SAQ 1. 19


Block 5 Group Theory
2. a) Yes.

b) The condition 1 has to be satisfied by every element of X. Condition 2


may not hold for every pair. In SAQ 2a, z is not related to any other
element.

3. a) No. – 1R0 and 0R + 2, for example. But (1) 2 < 0. Transitive property
is not satisfied.
b) Yes. iRi because i  i  0  5. If i – j is a multiple of 5 then so is j – i. And
lastly, i  i  n  5 and j  k  m  5 then
i  k  i  j  j  k  (n  m )  5. There are five equivalence classes:

[0] = ...,  5, 0, 5, 10, ..., [1] = ...  4, 1, 6, ..., [2] = ...  3, 2, 7, ...,

[3] = ...  2, 3, 8, ... [4] = ... =  1, 4, 9, ...

4. R  , the set of all real numbers not including the zero.

5. Let X = {x, y, z} and Y = 1, 2, 3. Let mapping f be defined as

f(x) = 1, f(y) = 2, f(z) = 1.


We choose A1  { x, y } and A2  {y, z} . Then

f ( A1)  {1, 2}, f ( A2 )  {1, 2}, f ( A1  A2 )  {2}  f ( A1)  {1, 2}.

Also, in this example A1  A2  { x } so that f ( A1  A2 )  {1} while


f ( A1)  f ( A2 )  , so that f ( A1  A2 )  f ( A)  f ( A2 ).

Terminal Questions
1. See Figs. 16.3a and 4b.
f e
C
C
l k
B
B
n c
h a b c m d
D g h
i
f
ff i
g
e
ee d j
i
A A  (B  C)
A A  (B  C) a b

(a) (b)

Fig. 16.3: Solution of TQ 1.

a) (acdea) = A  B, (ghbig) = A  C;

b) (abijklmna) = A  B, (abcdefgha) = A  C

2. The closed interval [ 1, 1].

3. It is neither one-to-one, nor onto mapping.


f 1{0}  {1, 0}, f [1,1]  [1/ 4, 2].
20
Unit 17 Groups

UNIT 17
GROUPS
Structure
17.1 Introduction 17.7 Cosets of a Subgroup
Expected Learning Outcomes 17.8 Cyclic Subgroups
17.2 Introduction to Groups Cyclic Subgroups of the Permutation
Definition of a Group Group
17.3 Group Multiplication Tables 17.9 Homomorphisms
Commutative or Abelian Groups Isomorphism
17.4 Subgroups 17.10 Conjugation
17.5 Permutation Group or 17.11 Invariant Subgroups
Symmetric Group Sn 17.12 Summary
Geometrical Illustration of S3 17.13 Terminal Questions
17.6 Transpositions and Even and 17.14 Solutions and Answers
Odd Permutations

17.1 INTRODUCTION
Mathematical quantities like groups appear primarily because of symmetry
which is so essential a concept for laws of physics. In this unit, we discuss
groups (Sec. 17.2) and related concepts like group multiplication tables,
subgroups, permutation groups (Secs. 17.3 to 17.5). Then we discuss
transpositions, even and odd permutations (Sec. 17.6), cosets of a subgroup,
cyclic subgroups and Homomorphisms (Secs. 17.6 to 17.9). In the last two
sections, we take up conjugation (Sec. 17.10) and Invariant Subgroups
(Sec. 17.11). In the next unit, you will learn about matrix representations.

Expected Learning Outcomes


After studying this unit, you should be able to:

 define a group and give its examples;


 write multiplication tables of finite groups;
 identify commutative and abelian groups; 21
Block 5 Group Theory
 define subgroups and permutation groups;
 transpositions, even and odd permutations;
 determine coset of a subgroup and cyclic groups;
 define homomorphism, isomorphism and conjugation;
 determine if a subgroup is invariant; and
 solve problems related to groups.

17.2 INTRODUCTION TO GROUPS


Let us begin the discussion with a few examples of groups.

Example 17.1
You know from UG physics courses that if a rigid body is symmetric about an
axis, then rotation by an angle  about the axis does not change the state of
the body. A further rotation about the same axis by an angle  also leaves the
state unchanged. This shows that the rotation by  +  does not change the
state. Also inverse rotation by – , and no rotation (that is rotation by angle 0)
do not change the state. This ‘group’ of rotations by all angles is
mathematically also a group.

Example 17.2

As you know from your earlier school and UG courses in physics, the
equations of motion for a free particle in a frame of reference with coordinates
(x, y, z) are:
d 2x d 2y d 2z
 0,  0, 0
dt 2 dt 2 dt 2
Let x   x  v x t, y   y  v y t, z  z  v z t

where (v x , v y , v z ) are components of some velocity. Then

d 2x d 2y  d 2z
 0,  0, 0
dt 2 dt 2 dt 2
Furthermore, if
x   x   v x t, y   y   v y t, z  z  v z t,

then
d 2 x  d 2 y  d 2 z 
 0,  0, 0
dt 2 dt 2 dt 2
But
x   x  v x t, v x  v x  v x

and similarly for y  and z  . The velocities define a group of transformations


from one frame of reference to another. There is the identity (no change) and
inverse (the velocity changes sign). They also form a group.
22
Unit 17 Groups
This shows that the equations of motion are unchanged when we replace
x  x   x  etc. using parameters v x v x connecting them.

Let us define a binary operation before proceeding further.

By definition, a binary operation B on a set X is a rule by which any two


elements of the set are assigned a third element of the set. We can say that a
binary operation is a mapping from the Cartesian product X  X into X:

B : ( X  X )  X, B( x1, x 2 )  X, x1, x 2  X

We can denote the binary operation by other symbols also, like x1 o x2 when
x1, x 2  X . Very often when there is no confusion, we also simply write
x1 x2 .

Examples of Binary Operations

1. The sum a + b of two real (or complex) numbers is a binary operation on


the set of real (or complex) numbers.

2. Similarly, the multiplication of two real (or complex) numbers is a binary


operation.

3. Multiplication of two square matrices is a binary operation defined on the


set of all square matrices of the same size.

A binary operation x1 o x2 defined on a set X is called associative if for any


three elements x1, x2 , x3 .

( x1 o x2 ) o x3  x1 o ( x2 o x3 )

The binary operation is called commutative if for any two elements x1 , x 2 .

x1 o x 2  x 2 o x1

Let us now define a group.

17.2.1 Definition of a Group


A group G is a set with a binary operation which satisfies the following three
conditions:

1. The operation is associative:

(g1 o g2 ) o g3  g1 o (g2 o g3 ), g1, g2, g3 G

2. There exists a unique element e  G called identity such that for any
g  G,
eog g goe

3. For every g  G, there exists a unique element g 1  G (called the inverse


of g) such that

g o g 1  g 1  g  e 23
Block 5 Group Theory
Remark 1

It is important to realize that G is just a set. It becomes a group only when a


binary operation satisfying the conditions above is defined on it. The same set
G becomes a different group with a different binary operation.

Therefore we should, strictly speaking, call the pair (G, ) as the group. When
there is no confusion what the binary operation is, we can say ‘group G’
without writing fully ‘group (G, )’.

The binary operation g1 o g 2 associated with a group is often called the


group multiplication law, or group composition law, although it is not like
multiplication of real or complex numbers. Also, for convenience of writing the
symbol  is very often omitted, so that g1 o g 2 is written simply as g1 g 2 .

We now give some more examples of groups.

Example 17.3
a) The set R of all real numbers with the binary operation (+), the addition of
numbers, forms a group with 0 as the identity, and for any a  R the
inverse is  a.
b) The set R  of all positive real numbers a, b with multiplication as the
binary operation is a group. The identity is the number 1, and inverse of a
is 1/a.
c) The set G has just two elements {e, a}, the identity and one more element
a. The binary operation is defined by
e  e = e, e  a = a = a  e, a  a = e.
Note that the inverse of a is a itself. This simple group appears in many
places in physics. For example, a could be rotation about an axis by  and
e would be no rotation, (same as rotation by 2) and the group
multiplication as the successive application of these operations. As another
situation, a can be parity or space-inversion which transforms the point
(x, y, z) as a : (x, y, z)  (x, y, z). The identity in this case is
e : (x, y, z)  (x, y, z).

SAQ 1

Define the binary operation (), the subtraction a – b for a, b  R as a group


operation. Does this make R a group?

17.3 GROUP MULTIPLICATION TABLES


A group with only a finite number of elements is called a finite group, and the
number of elements of a finite group is called its order. A group which has an
infinite number of elements is called an infinite group or a group of infinite
order. In the above examples, the groups on sets R or R  are infinite groups.
When the order of a finite group is small, it is useful to write the group
24 composition in the form of a table or a matrix with rows corresponding to the
Unit 17 Groups
left factor, and columns corresponding to the right factor. We write the
multiplication tables of some simple finite groups with two elements {e, a},
three elements {e, a, b} and four elements (e, a, b, c} below:

e a b c
e a b
e a e e a b c
e e a b
e e a a a e c b
a a b e
a a e b b c e a
b b e a
c c b a e

(17.1)

Notice that each row (or column) of the matrix of products (that is the matrix
below the horizontal line and to the right of the vertical line) has all the
elements of the group but in a different order. This is so because in a group
table if a  b = c and also a  d = c then (multiplying on the left by a 1 in both
the equations) b  a 1 c  d. But that cannot happen because b and d belong
to different columns.

17.3.1 Commutative or Abelian Groups


In general, in a group multiplication the order of elements matters: g1  g 2 is
different from g 2  g1. When g1  g 2 = g 2  g1 we say that these two
elements commute with each other.

When all elements of a group commute with each other, we call such a group
a commutative or Abelian group. A group which is not abelian is called non-
Abelian.

All groups given in the examples so far are Abelian groups. It can be proved
that all finite groups of order less than equal to 5 are Abelian.

17.4 SUBGROUPS
A subgroup is a group within a group with the same binary operation.

Let G be a group and S a subset: S  G. If the set S has the following


properties:

i) it is closed under the binary operation of G, that is, if a, b  S, then


a  b  S,
ii) since the identity must always be in a group, it includes the identity e  S,
and
iii) if a  S then a 1  S

then S is a group within G and it is called a subgroup of G.


25
Block 5 Group Theory
Note that according to the definition, the identity by itself {e} and the group G
are also subgroups. These are called trivial subgroups of G. Other
subgroups, if they exist are called proper subgroups.
Example 17.4
Let GL (n, R) be the group of all n  n real matrices with non-zero determinant.
The group multiplication is the multiplication of matrices, which is associative.
As the identity matrix 1 and the inverse matrix exist in GL (n, R) (because
determinant of the matrices is non-zero), GL (n, R) is a group.

An orthogonal matrix A is one for which AA T  1. Taking determinants of this


equation, we get det( AAT )  (det A) 2  1. Therefore, these matrices have
non-zero determinant and belong to GL (n, R). The product of two orthogonal
matrices is also orthogonal. If A1A1T  1 and A2 AT2  1 then

( A1A2 ) ( A1A2 )T  ( A1A2 ) ( AT AT )  A1( A2 AT


2 1 2
)A1T  A11A1T  1

Moreover, the inverse of an orthogonal matrix A is A 1  AT which is again


an orthogonal matrix: ( AT ) ( AT )T  AT A  A 1A  1. Therefore the set O(n)
of all n  n orthogonal matrices forms a subgroup.

The group GL(n, R) is called the general linear group of real matrices of
dimension n. The subgroup O(n) is called the orthogonal group of dimension
n.

There is still a smaller group within O(n). Since (det A)2  1, the O(n) matrices
with determinant +1 form a subgroup of O(n) (and therefore also of GL(n, R)).
This group is called the special orthogonal group of dimension n, and
denoted by SO(n). The members of O(n) with determinant – 1 cannot form a
group because the product of two matrices of determinant – 1 is a matrix with
determinant + 1.

SAQ 2

Show that n  n complex matrices with non-zero determinant form a group


GL (n, C).

Example 17.5

The group of four elements {e, a, b, c} is defined in Eq. (17.1) by group


multiplication tables has the following three proper subgroups:
{e, a}, {e, b}, {e, c}

17.5 PERMUTATION GROUP OR SYMMETRIC


GROUP Sn
Different ways of arranging n distinct objects in a line is called a permutation. It
is customary to denote these objects by the natural numbers 1, 2, …, n. There
26
Unit 17 Groups
are n! ways of arranging these numbers. For example, if n = 3, then there are
3! = 6 ways as shown below:

(1, 2, 3), (1, 3, 2), (2, 3, 1), (2, 1, 3), (3, 1, 2), (3, 2, 1)

As all the arrangements are the same elements in different order, it is very
useful to see them as one-to-one correspondence of the set X = {1, 2, 3} into
itself. Therefore every permutation is a one-to-one onto mapping X  X. Let
us name these mappings in a standard way:

1 2 3 1 2 3 1 2 3
e , p1   , p2   ,
     
1 2 3 1 3 2 2 3 1

1 2 3 1 2 3 1 2 3
p3   , p   , p5   
  4    
2 1 3 3 1 2 3 2 1

Here the upper row is always (1, 2, 3) and the lower row tells us what these
elements are mapped into, e is obviously identity where 1  1, 2  2, 3  3.
In p4 , for example, 1  3, 2  1, 3  2.

We now define the product of any of these two permutations. For example,
p1  p4 can be understood as

1 2 3  1 2 3  1 2 3
p1  p 4     p
      5
1 3 2  3 1 2  3 2 1

This is so because p1 maps 1  1 and p4 maps 1  3. The result is 1  3


for the first column. Similarly, 2  3 from p1 and 3  2 from p4 giving 2  2
for the product. And for the third column, 3  2  1.
This permutation group of three objects is called symmetric group S3 . A
permutation group of n objects is called symmetric group Sn .

Remark 2

As a permutation is a mapping, the notation, for example, for

1 2 3
 
 
2 3 1

can also be written

2 1 3
 
 
3 2 1

These are equivalent ways of representing the same permutation. All that
matters is which number goes to which under the mapping.

SAQ 3
Find the inverse element of e, p1, p2 , p3 , p4 , p5 .
27
Block 5 Group Theory
17.5.1 Geometrical Illustration of S3

Consider an equilateral triangle with vertices numbered as 1, 2, 3 (see


Fig. 17.1). The first triangle denotes the unchanged position of the triangle and
corresponds to the identity e. The next p4 (where the numbers 1, 2, 3 are
replaced by 3, 1, 2) is obtained by a 2/3 rotation about an axis passing
through the centre of the triangle and perpendicular to the plane of the
triangle. A further 2/3 rotation (that is 4/3 rotation on e) gives p2 . Another
2/3 rotation will bring it back to e (as it should because p4 p2  e ).

e p4 p2 p3 p1 p5
3 2 1
2 1 3

1 2 3 1 2 3 2 1 1 33 2

Fig. 17.1: Illustration of S 3 .

What about p1,p3 and p5 ?

These are mirror reflections about the line passing through a vertex and
perpendicular to the side opposite. Thus p3 is obtained with reflection about
the line passing through the vertex 3 of e and interchanging 1 and 2 (as shown
by a dotted line in e). Similarly, p1 is obtained from e by mirroring about a line
passing through 1. And p5 is obtained by mirroring about a line passing
through 2.

This shows that the group S3 is related to the symmetries of a regular


triangle. This group is also called the dihedral group D3 .

SAQ 4

Is the group S3 commutative?

Example 17.6

A square, just like the equilateral triangle has symmetries (see Fig. 17.2.) We
denote by e the original square. Let R1 be a rotation by angle /2 about an
axis passing through the centre of the square and the normal to its plane. R1
keeps the square unchanged. So does R2 , a rotation by  as well as R3 , a
rotation by 3/2. The rotation by a further /2 angle, that is a rotation by 2
brings the square back to its original place. So, it is the same as the identity.

Apart from these, rotations by  about the four axes lying in the plane passing
through the centre (two parallel to the sides R x , R y and two diagonals
R , R ) as shown in the Fig. 17.2 also leave the square unchanged.
28
Unit 17 Groups

4 3

1 2

Fig. 17.2: Symmetries of a square.

We can write the mappings as : (a, b, c, d are 1, 2, 3, 4 in some arbitrary


order.):

d c  d c 
e  
a b a b
   

d c   c b d c  b a  d c  a d 
R1   , R 2   , R 3   ,
 a b d         
   a a b c d   a b b c 

d c   a b d c  c d 
Rx   , R y   
 a b d c  a b b a 
       

d c  b c  d c  d a
R   , R    
 a b a d   a b  c b 
      

As we will define later, e, R1, R2 , R3 form a cyclic group of order 4. And


R , R are just interchange of two diagonal elements with the other two
remaining fixed; operating twice gives identity, R R  e and R R  e.

The group multiplication can be read out from these. For example,

d c  a b b c  d c 
RR x   R    R  ,
a b d c  a d   a b
       

Therefore, RRx  R . Similarly, for other products.

SAQ 5

Make the multiplication table for this group of eight symmetries of the square:
D4  {e, R1, R2 , R3 , R x , R y , R  , R  }.
29
Block 5 Group Theory
17.6 TRANSPOSITIONS AND EVEN AND ODD
PERMUTATIONS
In the example of the permutation group S3 given in Sec. 17.5, notice that the
permutations p1, p3 and p5 are such that they are obtained by interchanging
just two elements, keeping the third fixed. Such permutations in which only a
pair of elements interchange their place, are called transpositions. The
product of such transpositions with themselves is equal to the identity. A
transposition with the identity, therefore, forms a subgroup of two elements.
Therefore there are these obvious subgroups in Sn .

It is also interesting that all permutations can be built up from products of


transpositions: thus, for example, if we are given p1, p3 and p5 , the three
transpositions in S3 , the remaining permutations can be written as

e  p1 o p1, p2  p2 o p3 , p4  p1  p5

Of course, there are many ways to write these products. For example, we
could also have written

e  p3 o p3 , p2  p3 o p5 , p4  p5  p3

But the important point is that all permutations can be divided into two
subsets: Those which can be written as a product of an even number of
transpositions, are called even permutations, and those which are a product
of an odd number of transpositions, are called odd permutations.

In the example above, e, p2 , p4 are even. (e can be considered as a product


of zero transpositions and is always counted as even). And the remaining
p1, p3 , p5 are odd because they are themselves one transposition.

We have defined even and odd permutations above.

The number of elements in a permutation group of n objects is n! and this


number is always an even number (because n  2). We now prove that exactly
half of the permutations are even and the other half are odd.

The proof is by induction, for a general n.

We first illustrate the line of argument for our example for n = 3.

In the group of permutations of three objects there are 6 elements, 3 even and
3 odd:

123, 231, 312 even

132, 321, 213 odd

Now we introduce a fourth element. This fourth object (4) can be put at the
end of the string of three objects, and then this last object can be interchanged
one by one with the objects before it. For example,
30
Unit 17 Groups
1234  1234 no interchange even remains even

1234  1243 one interchange 3, 4 even becomes odd

1234  1432 one interchange 2, 4 odd

1234  4123 one interchange 1, 4 odd

Similarly, for the other two even permutations. There will be one even, and
three odd. Thus, there are 12 permutations, 3 even and (3 + 3 + 3 = 9) odd.
On the other hand, putting 4 at the end of an odd permutation and then
interchanging with other objects will given 3 odd and 9 even: for example,
starting with the odd permutation 1324,

1324  1324 no interchange odd remains odd

1324  1342 one interchange 2, 4 odd becomes even

1324  1423 one interchange 3, 4 even

1324  4132 one interchange 1, 4 even

Similarly, for the other two odd permutations. Thus, we have generated 24 = 4!
Permutations by introducing 4, and there are 3+9 = 12 even and 12 odd
permutations of 4 objects. We started with equal number of even and odd
permutations for n = 3, and showed that there are an equal number of even
and odd permutations for n = 4.

For the general case, we assume that n!/2 permutations are even and n!/2
odd. Proceeding as above, by introducing (n + 1) as the last member in all the
n!/2 even permutations we will get n!/2 even and n!/2  n odd. Working on the
other n!/2 odd permutations we will get n!/2 odd and n!/2  n even
permutations. Thus there are n!/2 + n!/2  n = (n + 1)!/2 even permutations
and the same number of odd permutations. This shows that by assuming the
result to be true for n we can prove that it holds for (n + 1). But we have seen
that it holds for n  2 and n  3 . Therefore, it is true for all n.

17.7 COSETS OF A SUBGROUP


Given a subgroup H of a group G we can divide the group G into many
mutually disjoint subsets as follows. (We are dropping the symbol for binary
operation  for convenience of writing.)

We take an element a  G which does not belong to H, and then construct a


set

aH = {ah h  H}  G

This subset which is obtained by multiplying a to every element of H on the left


is called the left coset of H determined by a. This coset aH contains a = ae
because the identity e  H. 31
Block 5 Group Theory
aH is disjoint from H because if ah1  ah2 where both h1, h2  H, then
multiplying by h11 we get a  h2 h11  H which is a contradiction because a
was chosen to be not in H.

Next we choose an element b neither in H nor in aH. We can construct


another left coset bH. This coset is also disjoint from H as well as from aH,
because if ah1  bh2 it follows that b  ah1 h2 1  aH , which is not true.

In this way we can keep going till all the elements of the group G are covered
in some or other coset.

Each of the cosets has the same number of elements. There is a one-to-one
correspondence ah , g  H between elements of the coset aH and H. This is
so because ah1  ah2 implies h1  h2 if we multiply by a1.

It is important to know that although we have denoted the coset as aH


because we chose a as the starting point in the construction, all
elements in this coset are on the same footing, that is, they are equally
qualified to represent the coset.

Suppose a  aH, then a  ah1 for some h1  H. But a  ah11 , therefore any
element a  ah2  aH, can also be written as a  ah11h2  ah3  aH .
Therefore, as a set aH = aH. This means that we can choose any element in
a coset to define that coset.

If G is a finite group of order N, say, and the order of H is n, then it must be


true that N = nm where m is the number of cosets, and all cosets have the
name number n of elements.

This puts a restriction on what is the order of a subgroup H of a finite group G


because the order of H must divide the order of the group G.

Example 17.7

For a group of three elements {e, a, b}, we cannot have a  a = e because


then {e, a} will form a subgroup of order 2, and it does not divide 3. The only
choice is a  a = b. See the multiplication table given in Sec. 17.3 of this unit.

For the same reason, a group whose order is a prime number cannot have a
proper subgroup.

Just as we defined the left cosets, we can define right cosets Ha etc.:

Ha = {ha h  H}  G

Again all the right cosets have the same number of elements and they are
mutually disjoint. The whole group is divided into disjoint union of right cosets.
32
Unit 17 Groups
SAQ 6

Find the left and right cosets of the subgroup {e, a} of the group:

e a b c

e e a b c

a a e c b

b b c e a

c c b a e

17.8 CYCLIC SUBGROUPS


For a finite group G we start with an element a different from the identity. We
construct aa, aaa, aaaa, …  G. It is traditional to write aa as a2 , and aaa as
a3 , etc. At each stage, either an is equal to e or to a different group element
from e and a because if an = a then multiplying by a 1 we get an 1  e .

If the group is of finite order, that is if the group has only a finite number of
elements, then for some value of integer n, we stop at an = e. After that,
multiplying by a will only repeat the cycle e, a, a2 , …, an 1.

This set of elements {e, a, a2 , …, an 1 } forms a subgroup of G, because

ar as  ar s , where a r  s  a m if (r  s  n, r  s  n  m)

SAQ 7

Find the cycles of the group S3 .

17.8.1 Cyclic Subgroups of the Permutation Group


In a permutation like

1 2 3 4 5
p 
 
3 5 1 2 4

if we follow the mapping 1  3  1 we get a cycle (a transposition). We next


pick 2  5  4  2. Therefore, we can consider the given permutation as a
product of the two cycles.

1 3 2 5 4
   
   
3 1 5 4 1

It is not necessary to write both the rows because the permutation of a cycle
just means shifting the numbers one place to the right. The standard notation
is, therefore, chosen as 33
Block 5 Group Theory
1 3 2 5 4
   (13)    (254)
   
3 1 5 4 1

and

p = (13) (254)

Every permutation can be so broken. Some other examples are

1 2 3 4 5
p   (13) ( 4) (25)
 
3 5 1 4 2

Sometimes ‘cycles’ like (4) above are omitted when there is no confusion, and
the permutation of four objects, (13) (4) (25) is simply written as (13) (25), it
being understood that the numbers not mentioned are mapped to themselves.

Example 17.8

Let two permutations of S3 be p1  (13) and (23). The product is

p1p2  (13)  (23)  (13) (2)  (1) (23)

1 2 3 1 2 3
   
   
3 2 1 1 3 2

1 2 3
 
 
2 3 1

 (123 ).

Similarly, for 5 objects if p = (143) (25) and p = (45) (12) then

pp = (143) (25)  (45) (12) = (15) (432)

17.9 HOMOMORPHISMS
Given two groups G and G with binary operations  and , a mapping

 : G  G

is called a homomorphism if for any two elements, a, b  G, if (a) = a and


(b) = b then

a  b = (a  b)

We will omit the symbols for binary operations, it being understood that when
two elements of a group are written as a simple product, the binary operation
is that of that group. So we write

34 a b = (a) (b) = (ab), for every a, b  G


Unit 17 Groups
It is clear that the identity of G must be mapped to the identity of G, because

(e) (a) = (ea) = (a)

and multiplying this equation by (a) 1 on the right we get

(e)  (a) (a) 1  e

where e is the identity of G.

Also, for any a  G the homomorphism  maps a 1 to (a) 1 ;

(a)  (a) 1  (aa 1 )  (e)  e

which shows that (a 1 )  (a) 1.

But, apart from e, there can be several elements of G which can also map to
the same identity element e  G. Let K  G be the subset of elements of G
which map into the identity of G (Fig. 17.3).

K  e

G G

Fig. 17.3: Several elements of G can map to the same identity element.

When the group G is the same as G we call such a homomorphism as an


endomorphism.

These elements in K form a subgroup in G because if a, b  K then

(a)  e, (b)  e, then (ab)  e e  e, therefore, ab  K

17.9.1 Isomorphism
If the identity e  G is the only element which is mapped by a homomorphism
 into the identity e of G then the mapping  is one-to-one, because if a and b
are distinct elements of G and they map to the same element a = (a) = (b),
then since  is a homomorphism,
(ab 1)  (a)(b 1)  a(a) 1  e

But as e is the only element mapped to e, we must have ab 1  e, or, a = b.

Thus, a homomorphism for which only the identity maps to identity is


one-to-one. If it also happens to be onto, that is, the range of the mapping is
all of G then the homomorphism is called an isomorphism.
Two groups which are isomorphic are practically the same in their group
algebraic properties. 35
Block 5 Group Theory
When the isomorphism  maps the group G into itself:  : G  G, we call it an
automorphism.

Example 17.9

3  3 real matrices A are called orthogonal if

AT A  1

Taking determinant we see that (det A) 2  1 or det A =  1 because


det A  det AT . Therefore, such matrices always have an inverse. Moreover,
the product of the orthogonal matrices is also orthogonal:

AT A  1, BT B  1, then ( AB)T AB  BT ( AT A)B  BT B  1

The inverse of A by definition is A 1  AT .

Let O(3) be the set of all such matrices. They form a group under matrix
multiplication. This group O(3) is called the orthogonal group. Define a
mapping  of this group into the group of two elements, the numbers
Z 2  {1,  1} with multiplication law of ordinary numbers as

 : O(3)  Z2 : A  det A

This is a homomorphism because

det( AB )  det A det B

The subset SO(3)  O(3) which consists of matrices A  O(3) with


determinant 1 map to 1 Z 2 . They form a subgroup called the special
orthogonal group.

17.10 CONJUGATION
Two elements a and b of a group G are called conjugate of each other if there
is a g  G such that b  gag 1. The relationship is mutual because we can
also write a  (g 1 ) b(g 1 ) 1. Actually this conjugation is an equivalence
relation: (1) any a  G is conjugate to itself a  eae 1  eae, (2) if a is
conjugate to b then b is conjugate to a as seen above, and, (3) if a is
conjugate to b  g 1ag 11 and b is conjugate to c  g 2 bg 21 then a is conjugate
to c:

c  g 2 bg 21  g 2 g1ag 11g 21  (g 2 g1 ) a (g 2 g1 ) 1

We know (from Unit 16 on Sets and Mappings), that every equivalence


relation will make the set as a union of disjoint sets called equivalence
classes. An equivalence class is determined by a subset where every member
36
of the subset is related by the equivalence relation.
Unit 17 Groups
Note that every fixed element g  G defines a conjugation automorphism as

 g : G  G : a  gag 1

It is one-to-one because gag 1  gbg 1 implies a = b, and it is onto because


for any b  G there is an a  g 1bg such that  g (a )  b.

Example 17.10

Let us calculate the equivalence classes of the group S3 (or D3 ) . Recall that
there are six group elements {e, p2 , p4 , p1, p3 , p5 } with p 2 and p 4 being
inverses of each other, and p1, p3 , p5 being their own inverses. Since
geg 1  e, for any g, therefore, e always forms an equivalence class by itself
under conjugation.

g e p2 p4 p1 p3 p5

gp2g 1 p2 p2 p2 p4 p4 p4

gp1g 1 p1 p3 p5 p1 p5 p3

We need not go through gp4 g 1 because p 4 already appears in the


conjugate elements of p 2 . Similarly, gp1g 1 brings up the remaining three
elements of the group, so there is no need to explore further.

Therefore equivalence classes of conjugation for S3 are:

{e}, { p2 , p4 }, { p1, p3 , p5 }

SAQ 8
Calculate the equivalence classes of the group D4 .

17.11 INVARIANT SUBGROUPS


Just as we have conjugate elements of a given group element, we can think of
all the conjugate elements of a subset S of G. Just as we collect all the
conjugate elements of a single element a  G to make the equivalence class,
we can collect all the conjugate elements of every element of a  S  G. We
write symbolically

GSG 1  {gag 1 a  S, g  G }

This is particularly important if the subset S happens to be a subgroup H.

An invariant subgroup or a normal subgroup H of a group G is such that all


the conjugate elements of H are within H itself. In other words,

GHG1  H 37
Block 5 Group Theory
For an invariant subgroup, the left and right cosets are the same. This is so
because for any a outside H, the condition for invariant H means

ah1a 1  H

Therefore, there is h2  H such that

ah1a 1  h2 , or ah1  h2a

The last equality shows ah1  aH in the left coset is the same element as
h2 a  Ha of the right coset.

With this equality of left and right cosets of an invariant subgroup, we can
define a composition law for cosets. We choose the left cosets for writing the
composition law

(aH)(bH) = (ab)H

The first thing to check is that this definition is independent of the


repesentatives a and b chosen, because we know from Section 17.7 that if
a  aH is any other element then a is equally good to represent the coset
because aH = aH. Similarly if b  bH is any other element then b is equally
good to represent the coset because bH = bH.

Therefore, we must make sure that

(aH)(bH) = (ab)H

This is possible because if a   ah1 and b   bh2 for some elements


h1, h2  H, then a typical element on the right hand side is

abh  ah1bh2h  ab(b 1h1b) h2 h

But h3  (b 1h1b  H ) because H is an invariant subgroup, so that


a b h  abh3 h2 h  abH. In other words, abH and a b H are the same coset,
and the composition of these cosets is well defined.

The group of left or right cosets of an invariant subgroup H is called the factor
group denoted by G / H . If G is a finite group of order N, then we know that
the order n of the subgroup H (which is also equal to the number of elements
in each coset) divides N. Therefore, the order of the factor group of an
invariant subgroup is N / n.

17.12 SUMMARY
In this unit we have discussed the following concepts:

 definition of a group, group multiplication tables;

 commutative or abelian groups;

38  subgroups;
Unit 17 Groups
 permutation of group or symmetric group S n ;

 transpositions, even and odd permutations;

 cosets of a subgroup;

 cyclic groups and cyclic groups of the permutation group;

 homomorphism, isomorphism;

 conjugation; and

 invariant subgroups.

17.13 TERMINAL QUESTIONS


1. Let R  be all positive real numbers. Define a binary operation

ab
ab .
7

Does this make R  a group?

2. Show that n  n unitary matrices form a subgroup U(n) which contains a


subgroup SU(n) of unitary matrices with determinant  1.

3. Construct the multiplication table for the group of permutations of {1, 2, 3} .

4. List the inverses of all the elements of D4 .

5. List the proper subgroups of D4 .

6. Find all the proper subgroups of the permutation group S3  {e, p1,..., p5 }
(defined in Sec. 17.5) and their cosets.

17.14 SOLUTIONS AND ANSWERS


Self-Assessment Questions
1. No, because there is no identity: a – e = a means e = 0, but e – a  a for all
a  0 in R.

2. a) Product of two n  n complex matrices M1 and M 2 with non zero


determinants, M1  0, M 2  0 is another such a complex matrix
because the determinant of the product is the product of determinants:
M1M 2  M1 M 2  0.

b) The identity matrix of size n is an identity element for the product.

c) The inverse for such matrices exist because the determinant is not
zero. This group is called GL(n,C ) which stands for General Linear
group of n  n complex matrices. 39
Block 5 Group Theory
3. The inverse mapping of a permutation

1 2 3
p2   
 
2 3 1

is obtained by interchanging the upper and lower rows:

2 3 1  1 2 3
p 21    p
    4
1 2 3 3 1 2

Similarly for others,


g e p1 p2 p3 p4 p5

g 1 e p1 p2 p3 p4 p5

4. No. As seen from the multiplication table above, p3  p5  p5  p3 .

5. The multiplication table for group of eight symmetries of the square,


D4  {e, R1, R 2 , R3 , R x , R y , R  , R  }

e R1 R2 R3 Rx Ry R R

e e R1 R2 R3 Rx Ry R R

R1 R1 R2 R3 e R R Ry Rx

R2 R2 R3 e R1 Ry Rx R R

D4  R3 R3 e R1 R2 R R Rx Ry

Rx Rx R Ry R e R2 R3 R1

Ry Ry R Rx R R2 e R1 R3

R R Rx R Ry R1 R3 e R2

R R Ry R Rx R3 R1 R2 e

6. The left and right cosets of the subgroup {e, a} of the given group are:
H  {e, a} and bH  (b, c}  Hb .
7. All the proper subgroups of S3 are cyclic:

H  {e, p2 , p4 }, H1  (e, p1}, H3  {e, p3 }, H5  {e, p5 }.

8. Look at the multiplication table given in solution to SAQ 5. The inverse of the
elements can be calculated as follows:
R1R 2  e, R11  R 3 , R 31  R1

R 2 R 2  e, R 21  R 2

Rx Rx  e, Rx1  Rx
40
Unit 17 Groups

R y R y  e, R y1  R y

RR  e, R1  R

RR  e, R1  R
We know that eis always an equivalence class by itself because
geg 1  e for all group elements.

Next

g gR1g 1

e R1

R1 R1R1R11  R1

R2 R 2 R1R 21  R 3 R 3  R 2

R3 R 3 R1R 31  eR 31  R1

Rx R x R1R x1  R _ R x  R 3

Ry R y R1R y1  R  R y  R 3

R R  R1R 1  R x R   R 3

R R  R1R 1  R y R   R 3

This shows that there is an equivalence class containing the elements


R1,R2,R3 
Since R2,R3 are already in this equivalence class we need not check
gR2g 1 or gR3g 1 .

Next
g gR x g 1

e Rx

R1 R1R x R11  R  R 3  R y

R2 R 2 R x R 21  R y R 2  R x

R3 R 3 R x R 31  R  R1  R y

Rx R x R x R x1  R x

Ry R y R x R y1  R 2 R y  R x

R R  R x R 1  R1R   R y

R R  R x R 1  R 3 R   R 3
41
Block 5 Group Theory
Therefore R x , R y form an equivalence class. We need not look at
R y because it is already counted in this class.

Next
g gR  g 1

e R

R1 R1R  R11  R y R 3  R y

R2 R 2 R  R 21  R  R 2  R 

R3 R 3 R  R 31  R x R1  R 

Rx R x R  R x1  R _ R x  R 

Ry R y R  R y1  R1R y  R 

R R y R  R 1  R 

R R  R  R 1  R 2 R   R 

This gives us the equivalence class R , R .


The group D4 has the following 4 equivalence classes:
e, R1,R2, R3 , Rx ,Ry , R , R 

Terminal Questions
1. Yes. The identity is e = 7, and the inverse of a is 49/a.

2. A unitary matrix U is such that U †U  1. Therefore U †U  1 .

As

U †  (U * )T  U *  U *

therefore

*
U †U  U U  1

proving that its determinant U is equal to a complex number of modules


unity. Also as U †U  1, we have U †  U 1 . This shows that the product of two
unitary matrices is unitary:

(U1U2 )†  U2†U1†  U21U11  (U1U2 )1.

The identity is unitary, because its hermitian adjoint is itself, and it is its own
inverse. Thus as the inverse and identity exist, all unitary matrices form a
group.
42
Unit 17 Groups
The subset of unitary matrices that have determinant equal to 1 form a
subgroup because the product of such matrices also has determinant 1:

U1U 2  U1 U 2  1.

3. The multiplication table for the group of permutations of {1, 2, 3}

e p1 p2 p3 p4 p5

e e p1 p2 p3 p4 p5

p1 p1 e p3 p2 p5 p4

S3  p2 p2 p5 p4 p1 e p3

p3 p3 p4 p5 e p1 p2

p4 p4 p3 e p5 p2 p1

p5 p5 p2 p1 p4 p3 e

4. Inverses of all the elements of D 4 are:

g e R1 R2 R3 Rx Ry R R

g 1 e R3 R2 R1 Rx Ry R R

5. Proper subgroups of D4 are:

{e, R x }, {e, R y }, {e, R  }, {e, R  }, {e, R1, R 2 , R3 }

6. The structure of subgroups becomes clear if we rearrange our elements in


the table:

e p2 p4 p1 p3 p5

e e p2 p4 p1 p3 p5

p2 p2 p4 e p5 p1 p3

p4 p4 e p2 p3 p5 p1

p1 p1 p3 p5 e p2 p4

p3 p3 p5 p1 p4 e p2

p5 p5 p1 p3 p2 p4 e

H  {e, p2 , p4 } H1  {e, p1}, H3  {e, p3 }, H5  {e, p5 }


43
Block 5 Group Theory
Cosets:

The two cosets (2 = 6/3) of H are

H  {e, p2 , p4 }, p1H  { p1, p3 , p5 }  p3H  p5H  Hp1  Hp3  Hp5

The three left cosets of H1 are

H1{e, p1}, p2H1  p5H1  { p2 , p5 } p3H1  p4H1  { p3 , p4 }

The three right cosets are

H1 {e, p1}, H1 p2  H1 p3  { p2 , p3 } H1 p 4  H1 p51  { p 4 , p5 }

The three left cosets of H3 and H5 can be constructed similarly. Notice


that the left coset p2 H1 is not the same as the right coset H1 p2 .

44
Unit 18 Matrix Representations

UNIT 18
MATRIX
REPRESENTATIONS
Structure
18.1 Introduction 18.6 Example of Reducing a
Expected Learning Outcomes Representation of S 3
18.2 Matrix Representations 18.7 Some Definitions and Results
2  2 Matrix Representation of D 3 Every Representation of a Finite Group
3  3 Matrix Representation of S3 can be Made Unitary by a Similarity
18.3 Equivalent Representations Transformation
18.4 Reducible and Irreducible Schur’s Lemma
Representations 18.8 Summary
18.5 Group Representation by 18.9 Terminal Questions
Linear Operators 18.10 Solutions and Answers

18.1 INTRODUCTION
You have learnt in Unit 17 that the real and complex matrices with non-zero
determinant form groups under matrix multiplication. For n  n matrices these
are called the general linear group of real matrices GL (n, R) and the general
linear group of complex matrices GL (n, C), respectively. If G is a group, then
a homomorphism of this group into GL (n, R) or GL (n, C) (for some n) is
called a matrix representation of G.
In this unit we will discuss matrix representation (Sec. 18.2). We also discuss
equivalent representations, reducible and irreducible representations, group
representation by linear operators (Secs. 18.3 to 18.5) and give an example of
reducing a representation of S3 (Sec. 18.6). Finally, in Sec. 18.7, you will
learn a few more definitions and results, and Schur’s lemma.

Expected Learning Outcomes


After studying this unit, you should be able to:

 define and determine matrix representations of groups;


 define and determine equivalent representations;
 define and determine reducible and irreducible representations and give
some examples; 45
Block 5 Group Theory
 represent a group by operators;
 state and derive Schur’s lemma; and
 solve problems related to matrix representations.

18.2 MATRIX REPRESENTATIONS


If G is a group, then a homomorphism of this group into GL (n, R) or GL (n, C)
(for some n) is called a matrix representation of G. The integer n is called the
dimension of the representation.

In other words, a matrix representation of dimension n assigns to all group


elements g  G, n  n non-singular matrices D(g) so that if g 1 and g 2 are any
two elements of the group G, then

D( g 1 ) D( g 2 ) = D( g 1 g 2 )

We say it is a real or complex, (as the case may be) matrix representation.

A few examples will clarify the concept.

The simplest group G = {e, a} with a  a = e, is an abstract group because we


know nothing about the mathematical nature of e or a. Or, about the nature of
operation , except that it is an associative product, which satisfies the group
axioms. This group is isomorphic to the group S 2 of permutation of two
objects.

A matrix representation of such a group is a mapping D : G  M of the


group elements into a set M of real (or complex) matrices such that the group
law is obeyed by corresponding matrices:

D(e)D(e) = D(e  e) = D(e), D(e)D(a) = D(e  a) = D(a)

D(a)D(e) = D(a  e) = D(a), D(a)D(a) = D(a  a) = D(e)

A representation by one-dimensional real matrices (that is real numbers) of the


group G = {e, a} is

D(e) = 1, D(a) =  1

A representation by two-dimensional real matrices of the same group (not


unrelated to the above one-dimensional representation) is

 1 0 1 0 
D(e )   , D(a )   
 0 1  0  1
   

18.2.1 2  2 Matrix Representation of D3


A less trivial example is the matrix representation of the group D3 . We saw in
Unit 17 (Sec. 17.5.1) that it is the group of rotations and mirror reflections of an
equilateral triangle in a plane. It is also the group S 3 of six permutations of the
vertices of the triangle. We reproduce the figure here, and rename the group
elements (Fig. 18.1).
46
Unit 18 Matrix Representations

e p4  R1 p2  R2 p3  M3 p1  M1 p5  M 2
3 2 1 3 2 1

1 2 3 1 2 3 2 1 1 33 2

Fig. 18.1: Illustration of S 3 .

S3 D3 name Remark

E E No change

p4 R1 Anti-clockwise rotation by 2/3

p2 R2  R1 R1 Anti-clockwise rotation by 4/3

p3 M3 Mirror reflection about the line passing through the


centre and the top vertex

p1 M1 Mirror reflection about the line passing through the


centre and the lower left vertex

p2 M2 Mirror reflection about the line passing through centre


and lower right vertex

The group multiplication tables can be summarised (omitting the action of


identity) as follows:

R1 R2 M1 M2 M3

R1 R2 e M3 M1 M2

R2 e R1 M2 M3 M1

M1 M2 M3 e R1 R2

M2 M3 M1 R2 e R1

M3 M1 M2 R1 R2 e

Rotations in the plane can be represented by 2  2 matrices which change


coordinates (x, y) as a column vector. Taking the centre of the triangle as the
origin, the equilateral triangle of side of unit length looks like:
47
Block 5 Group Theory

1 : ( 1/ 2,  1/ 2 3 )

2 : ( 1/ 2,  1/ 2 3 )

3 : (0,  1/ 3 )
1 2

Fig. 18.2: Diagram for matrix representation of D3 .

A rotation R1 by 2/3 = 120 brings every point (x, y) to ( x~, y~) such that

 x~   x    1/ 2  3 / 2  x 
   D(R )      
 ~  1     
  1 / 2   y 
 
y y   3 / 2

Thus R1 is represented by the 2  2 matrix

  1/ 2  3 / 2
 
D(R1 )   
 3 / 2  1 / 2 

And R2  R1R1 is then represented by two applications of the rotation by


120, that is, by D(R2 )  D(R1 ) D(R1 ) :

  1/ 2  3 / 2    1/ 2  3 / 2
  
D(R 2 )    
 3 / 2  1/ 2    3 / 2  1/ 2 

  1/ 2 3 / 2
 
 
 3 / 2  1/ 2 

Mirror reflections about the line through the vertex 3 and centre is easy: it just
changes the x coordinate to its negative and does not change the y
coordinate:

 x~   x  1 0  x    x 
   D(M )         
 ~  3       
 
y y   0 1  y   y 

Therefore

1 0
D(M 3 )   
 
48 0 1
Unit 18 Matrix Representations
The mirror reflection in a line from the centre making an angle  with the x-
axis, changes coordinates (x, y) to ( x~, y~) by a matrix as follows (TQ 1):

 x~   cos(2) sin(2)   x 
   
 ~    
 y   sin(2)  cos(2)   y 

For our purposes, the mirror reflection M1 in a line from vertex 1 is at an angle
/6 = 30 with the x-axis, therefore

  1/ 2 3 / 2
 
D(M1 )   
 3 /2  1 / 2 

Similarly, the mirror reflection M 2 in a line from vertex 2 is at   /6 = 150,


with the x-axis. Therefore,

  1/ 2  3 / 2
 
D( M 2 )   
 3 / 2  1 / 2 

The identity of the group is always represented by the unit matrix:

1 0
D(e )   .
 
0 1

This completes the construction of a 2  2 representation of the group D3 .

SAQ 1

Show that in the x-y plane, a rotation about the origin in an anti-clockwise
direction by an angle  will move the points (x, y) to ( x~, y~) by the matrix

 cos   sin  
 
 
 sin  cos  

18.2.2 3  3 Matrix Representation of S3


Remember that the group D3 is the same as the permutation group S 3 of
three objects. Just as we constructed a two-dimensional matrix representation
in the previous section by seeing the effect of rotations and mirror mappings
on a two-dimensional plane, we can construct a 3  3 representation by
observing how a given permutation works on a column vector containing
x1, x 2 , x 3 . For example, if p 4 takes 123  312 then a matrix can do that on
( x1, x 2 , x 3 ) by

 x3   0 0 1  x1   x1 
       
       
 x1    1 0 0  x 2   D( p 4 )  x 2 
       
 x  0 0  x  x 
 2  1  3  3 49
Block 5 Group Theory
Similarly we can build the other permutations:

1 0 0 1 0 0 0 1 0
     
     
D(e )   0 1 0 , D( p1 )   0 0 1 , D( p 2 )   0 0 1
     
0  0  1 0 
 0 1  1 0  0

(18.1)

0 1 0 0 0 1 0 0 1
     
     
D( p 3 )   1 0 0 , D( p 4 )   1 0 0 , D( p 5 )   0 1 0
     
0  0  1 0 
 0 1  1 0  0

(18.2)
We can check that, for example,

1 0 0 0 1 0 0 1 0
     
     
D( p1 ) D( p 2 )   0 0 1 0 0 1   1 0 0   D( p 3 )
     
0 0  1 0   0 1 
 1  0 0

and similarly for other products. Try this as an exercise.

18.3 EQUIVALENT REPRESENTATIONS


Suppose we have a representation g  D(g) of group elements g  G into real
(or complex) matrices of dimension n. Let S be a non-singular matrix of the
same dimension, then the matrices D(g )  SD(g ) S 1 also form a
representation because

D(g1 ) D(g 2 )  SD(g1 ) S 1SD(g 2 ) S 1

 SD(g1) D(g2 ) S 1  SD(g1g2 ) S 1

 D(g1g 2 )

This representation g  D(g) is not really different from the first one g  D(G)
because they are related by the matrix S. (This kind of relation between two
matrices A  SA S 1 is called a similarity transformation).

Two representations like D and D above are called equivalent


representations. In fact, the relation of being equivalent is an equivalence
relation and representations belonging to the same class are practically the
same.
50
Unit 18 Matrix Representations

18.4 REDUCIBLE AND IRREDUCIBLE


REPRESENTATIONS
Let D1(g ) (of dimension n1 ) and D2(g ) (of dimension n 2 ) be two
representations of the group G. Then we can construct a representation D(g)
of dimension n1  n2 by placing the matrices for the same group element in a
block-diagonal form

 D1(g ) 0 
D(g )   
 
 0 D2 ( g ) 

Since block diagonal matrices multiply block-wise, it is a representation:

 D1(g 1 ) 0   D1(g 2 ) 0 
D(g 1 ) D(g 2 )     
   
 0 D2 ( g 1 )   0 D2 ( g 2 ) 

 D1(g 1g 2 ) 0 
 
 
 0 D2 (g 1g 2 ) 

 D(g1g 2 )

A representation like this may not look like a block-diagonal representation if


we use a non-singular (n1  n2 )  (n1  n2 ) matrix S and define the equivalent
representation

 D1(g ) 0 
D (g )  S D(g )S 1  S   S 1
 
 0 D2 ( g ) 

The natural question is: suppose we were given the representation D (g ) first.
How could we have found out that it was made up from two smaller size
representations?

We call a representation reducible if it is equivalent to a representation in a


block-diagonal form of matrix representations of smaller dimensions. A
representation which cannot be reduced to or brought into a block-diagonal
form by a similarity transformation is called irreducible.

The irreducible representations of any group are very important because they
are the basic units out of which all representations are constructed.

18.5 GROUP REPRESENTATION BY LINEAR


OPERATORS
You have learnt in Unit 7 that matrices are intimately related to linear
operators in a vector space V. Recall that if A is a linear operator A : V  V in
a finite dimensional real or complex vector space, then the action of the linear
operator is completely determined by its action on vectors of the basis.
Let e1, e 2 ,..., e n  be a basis in V. Since Ae i , i  1,... n is again in V, it can be
expanded in the basis, with coefficients Aij : 51
Block 5 Group Theory
Ae i   A ji e j
j

(Notice the order of the indices in A ji ). The n  n matrix A ji is associated


with the linear operator A. If B is another linear operator then

BAe i   A ji Be j   A ji Bk j e k
j j ,k

 
    Bk j A ji e 
 k (BA) k i e k
k  j  k

This shows that the product BA of operators B and A is associated with the
product of corresponding matrices.
We can therefore also look for the representation of a group G as g  A(g) by
operators in a vector space in place of matrices.
This has many advantages especially when the vector space is infinite
dimensional like the Hilbert space of quantum mechanics, where it is not easy
to deal with infinite dimensional matrices.
Remember that the matrix corresponding to a linear operator is dependent on
the basis chosen. If f1, f2 ,..., fn  is another basis, and we expand each f in
the basis of es, then

fi   S ji e j (18.3)
j

(Notice again, the order of the indices in S ji ). Then, we can find the matrix
A ji associated with A with respect to the basis fi 

A fi   Aji f j
j

  Aji Sk jek (18.4a)


j ,k

On the other hand,

A fi   S ji Ae j
j

  S ji Ak jek (18.4b)
j

Comparing Eqs. (18.4a and b), and equating coefficients of e k we get

 Aji Sk j  S ji Ak j
j j

or, as matrices
(SA) ki  ( AS) ki

which can be rewritten as

A  SAS 1, or A  S 1AS (18.5)


52
Unit 18 Matrix Representations
Therefore, the matrices associated with an operator in two different bases are
related by a similarity transformation.
This result is of great use in finding irreducible representations.
Suppose we can represent group elements as operators on a vector space
which has an orthonormal basis. (This happens regularly in quantum
mechanics.) If e1, e 2 ,..., e n  is an orthonormal basis,

e i , e j   ij

then the matrix associated with operator A in this basis can easily be
calculated by the ‘matrix elements’ as follows. From
Ae i   A ji e j
j

it follows on taking inner product with e k , that

e k , Ae i   A ji ek ,e j   A ji  k j  Ak i
j j

Now suppose we find that some m-dimensional subspace V1 of the vector


space V is invariant under all the group operators. This means that the group
operators A(g) acting on vectors of the subspace give vectors of the same
subspace:
A(g )v  V1 for every v  V1 and for every g  G

Then we choose an orthonormal basis f1, f2 ,..., fn  such that the first m
vectors of the basis span the invariant subspace V1 and the remaining (n  m)
vectors span the orthogonal subspace V2 . This means that every vector of V2
is orthogonal to every vector of V1 and vice versa. The space V2 is also an
invariant subspace because a vector in V2 cannot map by any operator A(g)
into V1 because that would imply that A(g ) 1  A(g ) 1 will map a vector V1
into V2 . Thus we find that the matrix elements in the new basis are zero
between vectors of V1 and V2 and non-zero only between vectors of the same
subspace. The structure of the matrix A(g) in the basis of fs is:

 A1(g ) 0 
 ab 
A(g ) i j  

, a, b  1,..., m; c, d  m  1,..., n
 A2 (g )cd 

 0

Here A1(g ) and A2 (g ) are block diagonal matrices in A(g ) and they are
of dimension m and (n – m), respectively. Since the original matrices A(g)
were not block diagonal in the e-basis, we have reduced the representation by
changing over to the f-basis and using the knowledge of some invariant
subspace.

18.6 EXAMPLE OF REDUCING A


REPRESENTATION OF S3
We have discussed the six element permutation group S 3 and its other
geometrical version D3 in Unit 17. We also constructed their matrix 53
Block 5 Group Theory
representations: a two-dimensional representation above in Sec. (18.2.1) and
a three-dimensional matrix representation in Sec. (18.2.2).

We can ask the question: Are these representations of the same group
reducible or irreducible representations?

We start with the 3  3 representation. We defined its group elements as


acting on vectors in the form of a column matrix (in a three-dimensional
space). The matrices

1 0 0 1 0 0 0 1 0
     
     
D(e )   0 1 0 , D( p1 )   0 0 1 , D( p 2 )   0 0 1
     
0  0  1 0 
 0 1  1 0  0

(18.6a)

0 1 0 0 0 1 0 0 1
     
     
D( p 3 )   1 0 0 , D( p 4 )   1 0 0 , D( p 5 )   0 1 0
     
0 1  0 0  1 0 
 0  1  0

(18.6b)
work on the column vectors like

 x1 
 
 
 x2 
 
x 
 3

and permute them in the column to some other order. We can think of the
column vector as a vector in the three-dimensional space with components
( x1, x 2 , x 3 ) that is mapped by the representative of the permutation group.
For example, ( x1, x 2 , x 3 ) is mapped by the permutation 123  312 to
( x3 , x1, x 2 ) , and so on.

It is clear that if we were to act these matrices on a column vector with the
equal entries

 1
 
 
 1
 
 1
 

then this vector will remain invariant under all the permutations. Thus, there is
a one-dimensional invariant subspace in the direction of (1, 1, 1).
Our aim is to use this fact and choose a new orthogonal basis in which the unit
54
vector in the direction of (1, 1, 1) is one of the three basis vectors.
Unit 18 Matrix Representations

f3

2
1

Fig. 18.3: Three-dimensional space for reducing S3 .

The three-dimensional space is shown in Fig. 18.3. The triangle which we had
used to define its symmetries in Sec. (18.2.1) here corresponds to the points
(1, 0, 0), (0, 1, 0) and (0, 0, 1) at the vertices as shown.

The centroid is shown with the little circle . The vector joining the origin and
the centroid is proportional to (1, 1, 1) and is the invariant under symmetries of
the equilateral triangle. The axis about which rotations by 120 take place is
normal to the triangle and is in the direction of (1, 1, 1). If we choose this as
the third axis, then under rotations this coordinate will remain unchanged and
will correspond to the invariant subspace. Thus, we choose, after normalizing,

 1
 
1  
f3   1
3 
 1
 

The choice of the other two vectors f1,f2 of the orthonormal basis is open. We
can choose any two orthogonal unit vectors in the plane perpendicular to f3 ,
and then calculate the matrix representation in f-basis. This will make the
3  3 representation into a block diagonal form with the f1,f2 space providing
the 2  2 representation of the permutation group S3 (or D3 ) and the
one-dimensional space of f3 providing the trivial representation which assigns
1 to every group element.

But the general choice of f1,f2 will not give us the 2  2 representation we
constructed in Sec. (18.2.1). For reference we write it down once again. The
relation of names of the group elements are as follows: M1, M 2 , M3 are the
mirror reflections p1, p2 , p3 , respectively, and R1  p4 , R2  p2 .

1 0
D(e )   ,
 
0 1 55
Block 5 Group Theory
  1/ 2  3 / 2   1/ 2 3 / 2
   
D(R1 )   , D(R 2 )   
 3 / 2  1/ 2   3 / 2  1 / 2 
 

1 0
D(M 3 )   ,
 
0 1

  1/ 2 3 / 2   1/ 2  3 / 2
   
D(M1 )   , D(M 2 )   
 3 /2  1/ 2   3 / 2  1 / 2 
 

The reason why any choice of f1,f2 will not give us the 2  2 matrices above
is that these matrices were made with assumption that the line joining the
vertices 1 and 2 is the x-axis, and the y-axis is normal to it so that the x-y-z
axes form a right-handed basis of unit vectors.

Therefore, our choice of f1 should be (1) parallel to the line joining 1 and 2 in
the diagram and (2) perpendicular to the vector f3 . So, the vector should be
chosen as:

  1
 
1  
f1   1
2 
0
 

The two vectors then determine the third orthogonal vector (in a right-handed
system) by the cross product as:

  1
 
1  
f2  f3  f1    1
6 
2
 

The original basis is:

 1 0 0
     
     
e1   0 , e 2   1 , e3  0
     
0 0  1
     

Therefore, the similarity matrix S (See Eqs. (18.3) and (18.5)) which connects
the f-basis to e-basis by fi   S ji e j is given by,

 a  ab b
 
 
S a  ab b , a 1 2, b 1 3
 
 0 b 
 2ab
56
Unit 18 Matrix Representations
We can now convert the 3  3 matrices D(e), D( p1),..., D( p5 ) into the reduced
matrices by using the similarity transformation (18.5):

D(g )  S 1D(g )S, g  e, p1,..., p5

As S is change of basis from one right-handed orthonormal basis to another in


a real vector space, the matrix S is orthogonal, that is, its determinant is unity
and its inverse can be obtained by its transpose:

 a a 0 
 
 
S 1  S T    ab  ab 2ab 
 
 b b 
 b

Thus, the reduced matrices can be obtained through S. As an example,

 a a 0 1 0 0   a  ab b
   
   
D ( p1 )  S 1D( p1 )S    ab  ab 2ab   0 0 1  a  ab b
   
 b b   0 0   0 b 
 b 1 2ab

 a a 0  a  ab b  a2 3a 2 b 0 
    
    
   ab  ab 2ab   0 2ab b    3a 2 b  3a 2 b 2 0 
    
 b  
 b b   a  ab b   0 0 2
3b 

  1/ 2 3 /2 0
 
   D(M 1 ) 0
 3 /2  1/ 2 0   
   0 
 
1

 0 0 1 

Recall that p1 is the same group element as M1 . The choice of the new basis
has brought the 3  3 representation into a block diagonal form of 2  2
representation and the trivial one-dimensional representation.

18.7 SOME DEFINITIONS AND RESULTS


In this section, we give some more definitions and relevant results.

18.7.1 Every Representation of a Finite Group can be


made Unitary by a Similarity Transformation

Let G be a group and e  g1, g 2 ,..., g N etc. its elements. Let D(g i ) be a
representation. In general we assume matrices D(g) to be complex; real
matrices are just a special case. We write Di  D(g i ) for convenience.
57
Block 5 Group Theory
We define a quantity

A  Di Di†
i

Then A is a positive definite Hermitian matrix, which can be diagonalized by a


unitary similarity transformation U, (U †  U 1) with real positive numbers on
the diagonal. That is

 1 
 
 
  
   U AU 1 (18.7)
  
 
 
  N 

Also, let

Di  U Di U 1  U Di U † (18.8)

then

  U AU 1  (U Di U † )(U Di†U † )   Di (Di )†


i i

Next we define matrix square roots of  by

 11/ 2 
 
 
 21/ 2 
1/ 2   
 
  
 
 N1/ 2 

 11/ 2 
 
 
 21/ 2 
1/ 2   
 
  
 
  N 
 1 / 2

Multiplying Eq. (18.7) on the right and on the left by 1 / 2 and using
1 / 2 1 / 2  1 , we get:

1  1/ 2Di (Di )† 1/ 2 (18.9)


i

We will use this identity presently. We define for any j

Dj  1/ 2Dj 1/ 2 (18.10)


58
Unit 18 Matrix Representations
then we claim that D j is unitary. As

( 1/ 2Dj 1/ 2 )†  1/ 2 (Dj )† 1/ 2

Dj (Dj )†  ( 1 / 2Dj 1 / 2 ) ( 1 / 2Dj 1 / 2 )†

 1 / 2 Dj (D j )† 1 / 2

 1 / 2 Dj (D j )† 1 / 2

  1 / 2D j Di (Di )† (D j )† 1 / 2


i

  1/ 2Dj Di (Dj Di )† 1/ 2


i

Since Di  U Di U 1  U D(g i )U 1, and since D(g)s form a representation,

D j Di  U D(g j ) D(g i )U 1  U D(g j g i ) U 1

As j is fixed and i goes over the group elements, g j g i goes over the same
group elements. Therefore if g j g i  g k , then using Eq. (18.9), we get:

D j (D j )†   1 / 2D j Di (D j Di )† 1 / 2


i

  1 / 2Dk (Dk )† 1 / 2


k

1
This proves the unitarity of the matrix representation D which is obtained by
two similarity transformations [see Eqs. (18.8) and (18.10)].
D(g i )  1/ 2U D(g i )U 11/ 2  V D(g i )V 1, V  1/ 2U.

18.7.2 Schur’s Lemma


Suppose D(g) is an irreducible n  n matrix representation of a finite group,
and there is a matrix M of the same size which commutes with every D(g), g 
G. Then

M D(g) = D(g) M, for every gG

Schur’s lemma tells us that such a matrix M can only be a constant times the
identity matrix.

The proof goes as follows. The logic of the proof is that if the matrix is not a
multiple of identity, the representation will be reducible.

We can always assume that the representation D(g) is a unitary


representation. Then taking adjoint, M D(g) = D(g)M implies

D(g )† M †  M † D(g )†

Multiplying this equation on the left and on the right by D(g) and D(g)†,
respectively, we get

M † D(g )  D(g ) M † 59
Block 5 Group Theory
Therefore, if M commutes with every D(g) then M†
also does so. And

therefore, the Hermitian matrices M1  M  M and M2  (M  M † ) i also do
so. Therefore, there is no harm in proving the theorem for M1 and M 2 which
are Hermitian.

If M1 is Hermitian, there is a unitary matrix U such that it is diagonalized by it.


Let

d1  U M1U 1, M1  U 1d1U

The commutation property is now

U 1d1U D(g )  D(g )U 1d1U

Or, multiplying by U and U 1 on the left and the right, and calling
D(g )  U D(g )U 1,

d1D(g )  D(g )d1 for every gG

Let the matrix elements of D(g) and d1 be labelled by a, b, c  1, . .., n. The


above equation then is for all a, c.

 (d1)ab D(g )bc  D(g )ab (d1 )bc


b b

Since d1 is diagonal, the off-diagonal elements are zero, and hence,

(d1)aa D(g )ac  D(g )ac (d1)cc

This means that

D(g )ac ((d1)cc  (d1)aa )  0

If for some values of a and c, it is true that (d1)cc  (d1)aa then the matrix
elements D(g )ac  0. Therefore, if we rearrange all those rows and columns
for which the diagonal elements of d1 are unequal, the representation
matrices D(g) (for all g  G) in this re-arrangement will be in block diagonal
form and reducible. But that is a contradiction, because the representation is
assumed to be irreducible. Therefore, all the diagonal elements of d1 are
equal, and the matrix is multiple of identity. The same result follows for M 2 ,
and therefore, for M which was to be proved.

We had assumed the representation D(g) to be unitary to begin with. But that
is not necessary because we can use a similarity transformation to bring D(g)
to unitary form using a similarity matrix S, and prove that the matrix SMS1 is
a multiple of identity SMS1  c1. But then M  cS 11S  c1 which was to be
proved.

Let us now summarise what we have studied in this unit.


60
Unit 18 Matrix Representations

18.8 SUMMARY
In this unit, we have discussed the following concepts:
 Matrix representations, and matrix representations of D3 and S3 ;

 Equivalent representations;

 Reducible and irreducible representations;

 Group representation by linear operators;


 Reducing a representation of S3 ;

 Similarity transformation for making every representation of a finite


group, unitary; and

 Schur’s lemma and its proof.

18.9 TERMINAL QUESTIONS


1. Show that the mirror reflection in a line passing through the origin and
making an angle  with the x-axis takes a point (x, y) in the plane to ( x~, y~)
by the matrix:

 cos(2) sin( 2) 


 
 
 sin(2)  cos(2) 

2. Complete and verify the multiplication table for the 3 × 3 representation of


S3 .
3. Check that for the other group elements M 2 , M3 , p2 , p4 the same
block-diagonal form is obtained by the similarity matrix S.

18.10 SOLUTIONS AND ANSWERS


Self-Assessment Questions
1. We use radial coordinates r, . Then x  r cos  and y  r sin . Under
rotation r does not change, and  becomes   . Therefore,

x~  r cos(  )  r [cos  cos   sin  sin ]  x cos   y sin 

y~  r sin(  )  r [sin  cos   cos  sin ]  x sin   y cos 

Thus,
 x~   cos   sin   x
    
 ~     
 y   sin  cos   y 
and the matrix representation is
 cos   sin  
 
 
 sin  cos  
61
Block 5 Group Theory
Terminal Questions
1. See Fig. 18.4.

y-axis

y-axis

x-axis

(x, y) : (x, y)


~, y~)
 (x, y) : ( x

x-
axis

Fig. 18.4: Diagram for solution of TQ 1.

If we make the line with angle  as the new x-axis, with an orthogonal
y-axis as shown, then mirror reflection of any point (x, y) in the line would
be to take this point to (x, y). Under the change of coordinates

 x    cos  sin    x 
   
    
 y     sin  cos    y 

The inverse relations are

 x   cos   sin    x  
   
    
 y   sin  cos    y  

As y   x sin  y cos ,

 x    cos  sin    x 
   
    
  y    sin   cos    y 

Therefore, we use the inverse relations for ( x~, y~) , which are the
coordinates with respect to the x-y axes of the point which has
coordinates ( x ,y ) with respect to x - y - axes

 x~   cos   sin    x  
    
 ~    
y   sin  cos     y  

 cos   sin    cos  sin   x


     
     
 sin  cos    sin   cos   y 
62
Unit 18 Matrix Representations
 cos(2) sin( 2)  x
   
   
 sin( 2)  cos(2)  y 

2. The group multiplication table of the permutation group S3 is

2 p1 p2 p3 p4 p5

e e p1 p2 p3 p4 p5

p1 p1 e p3 p2 p5 p4

S3  p2 p2 p5 p4 p1 e p3

p3 p3 p4 p5 e p1 p2

p4 p4 p3 e p5 p2 p1

p5 p5 p2 p1 p4 p3 e

The matrix representation is given by

1 0 0 1 0 0 0 1 0
     
     
D(e )   0 1 0 , D( p1)   0 0 1 , D( p2 )   0 0 1
     
0  0 0  1 0 
 0 1  1  0

0 1 0 0 0 1 0 0 1
     
     
D( p3 )   1 0 0 , D( p4 )   1 0 0 , D( p5 )   0 1 0
     
0  0 0  1 0 
 0 1  1  0

No need to verify the multiplication by the identify matrix D(e )  1

Multiplication by D( p1) :

1 0 0 1 0 0  1 0 0
     
     
D( p1) D( p1)   0 0 1 0 0 1   0 1 0   D(e ) .
     
0 0  0 0   0 1 
 1  1 0

1 0 0 0 1 0 0 1 0
     
     
D( p1) D( p2 )   0 0 1 0 0 1   1 0 0   D( p3 ) .
     
0 0  1 0   0 1 
 1  0 0
63
Block 5 Group Theory
1 0 0 0 1 0 0 1 0
     
     
D( p1) D( p3 )   0 0 1 1 0 0  0 0 1   D( p2 ) .
     
0 0  0 1   1 0 
 1  0 0

1 0 0 0 0 1  0 0 1
     
     
D( p1) D( p4 )   0 0 1 1 0 0  0 1 0   D( p5 ) .
     
0 0  0 0   1 0 
 1  1 0

1 0 0 0 0 1  0 0 1
     
     
D( p1) D( p5 )   0 0 1 0 1 0   1 0 0   D( p4 ) .
     
0 0  1 0   0 0 
 1  0 1

Multiplication by D( p2 ) :

1 1 0 1 0 0 0 0 1
     
     
D( p2 ) D( p1)   0 0 1 0 0 1   0 1 0   D( p5 ) .
     
0 0  0 0   1 0 
 0  1 0

0 1 0 0 1 0 0 0 1
     
     
D( p2 ) D( p2 )   0 0 1 0 0 1   1 0 0   D( p4 ) .
     
1 0  1 0   0 0 
 0  0 1

0 1 0 0 1 0  1 0 0
     
     
D( p2 ) D( p3 )   0 0 1 1 0 0  0 0 1   D( p1) .
     
1 0  0 1   0 0 
 0  0 1

0 1 0 0 0 1  1 0 0
     
     
D( p2 ) D( p4 )   0 0 1 1 0 0  0 1 0   D(e ) .
     
1 0  0 0   0 1 
 0  1 0

0 1 0 0 0 1  0 1 0
     
     
D( p2 ) D( p5 )   0 0 1 0 1 0   1 0 0   D( p3 ) .
     
1 0  1 0   0 1 
 0  0 0
64
Unit 18 Matrix Representations
Multiplication by D( p3 ) :

0 1 0 1 0 0 0 0 1
     
     
D( p3 ) D( p1)   1 0 0 0 0 1   1 0 0   D( p4 ) .
     
0 1  0 0   0 0 
 0  1 1

0 1 0 0 1 0 0 0 1
     
     
D( p3 ) D( p2 )   1 0 0 0 0 1   0 1 0   D( p5 ) .
     
0 1  1 0   1 0 
 0  0 0

0 1 0 0 1 0  1 0 0
     
     
D( p3 ) D( p3 )   1 0 0 1 0 0  0 1 0   D(e ) .
     
0 1  0 1   0 1 
 0  0 0

0 1 0 0 0 1  1 0 0
     
     
D( p3 ) D( p4 )   1 0 0 1 0 0  0 0 1   D( p1) .
     
0 1  0 0   0 0 
 0  1 1

0 1 0 0 0 1  0 1 0
     
     
D( p3 ) D( p5 )   1 0 0 0 1 0  0 0 1   D( p2 ) .
     
0 1  1 0   1 0 
 0  0 0

Multiplication by D( p4 ) :

0 0 1 1 0 0 0 1 0
     
     
D( p4 ) D( p1)   1 0 0 0 0 1   1 0 0   D( p3 ) .
     
0 0  0 0   0 1 
 1  1 0

0 0 1 0 1 0  1 0 0
     
     
D( p4 ) D( p2 )   1 0 0 0 0 1   0 1 0   D(e ) .
     
0 0  1 0   0 1 
 1  0 0

0 0 1 0 1 0 0 0 1
     
     
D( p 4 ) D( p3 )   1 0 0 1 0 0  0 1 0   D( p5 ) .
     
0 0  0 1   1 0 
 1  0 0 65
Block 5 Group Theory
0 0 1 0 0 1  0 1 0
     
     
D( p4 ) D( p4 )   1 0 0 1 0 0  0 0 1   D( p2 ) .
     
0 0  0 0   1 0 
 1  1 0

0 0 1 0 0 1  1 0 0
     
     
D( p4 ) D( p5 )   1 0 0 0 1 0  0 0 1   D( p1) .
     
0 0  1 0   0 0 
 1  0 1

Multiplication by D( p5 ) :

0 0 1 1 0 0 0 1 0
     
     
D( p5 ) D( p1)   0 1 0 0 0 1   0 0 1   D( p2 ) .
     
1 0  0 0   1 0 
 0  1 0

0 0 1 0 1 0  1 0 0
     
     
D( p5 ) D( p2 )   0 1 0 0 0 1   0 0 1   D( p1) .
     
1 0  1 0   0 0 
 0  0 1

0 0 1 0 1 0 0 0 1
     
     
D( p5 ) D( p3 )   0 1 0 1 0 0   1 0 0   D( p4 ) .
     
1 0  0 1   0 0 
 0  0 1

0 0 1 0 0 1  0 1 0
     
     
D( p5 ) D( p4 )   0 1 0 1 0 0   1 0 0   D( p3 ) .
     
1 0  0 0   0 1 
 0  1 0

0 0 1 0 0 1  1 0 0
     
     
D( p5 ) D( p5 )   0 1 0 0 1 0  0 1 0   D(e ) .
     
1 0  1 0   0 1 
 0  0 0

3. S is given by
 a  ab b
 
 
S a  ab b , a  1/ 2, b  a/ 3.
 
 0 b 
 2ab

66 D(g )  S 1D(g )S, g  e, p1,..., p5


Unit 18 Matrix Representations
As S is a change of basis from one right handed orthonormal basis to another
in a real vector space the matrix S is orthogonal, that is, its determinant is
unity and its inverse can be obtained by its transpose:
 a a 0 
 
 
S 1  S T    ab  ab 2ab 
 
 b b 
 b
The group has 6 elements e, p1, p2 , p3 , p4 , p5 . We do not have to calculate
for e because that is a unit matrix and it remains a unit matrix: S 1eS  e
which is already block diagonal.
We have already seen that
 a a 0  1 0 0  a  ab b
     

     
D( p1)  S D( p1)S    ab
1  ab 2ab  0 0 1  a  ab b
     
 b b  0 0   0 b 
 b  1  2ab

 a a 0   a  ab b  a2 3a 2 b 0 
     
     
   ab  ab 2ab   0 2ab b    3a 2 b  3a 2 b 2 0 
     
 b  
 b b   a
  ab b   0 0 3b 2 

  1/ 2 3 /2 0
 
   D( M 1 ) 0
 3 /2  1/ 2 0   
   0 
 
1

 0 0 1 

 a a 0  0 1 0  a  ab b
     
     
D ( p2 )  S 1D( p2 )S    ab  ab 2ab  0 0 1  a  ab b
     
 b b  1 0   0 b 
 b  0  2ab

  a2 3a 2 b 0    1/ 2 3 /2 0
   
     D(R 2 ) 0
   3a 2 b  3a 2 b 2 0    3 / 2  1/ 2 0   
     0 
 
1
  
 0 0 3b   0
2 0 1 

 a a 0  0 1 0  a  ab b
     
     
D ( p 3 )  S 1D( p 3 )S    ab  ab 2ab  1 0   a  ab b
     
 b b  0 0   0 b 
 b  0  2ab

  2a2 0 0   1 0 0
   
     D(M3 ) 0

 0 6a2b2 0  0 1 0 
    
  0 1
 
 0 0 3b2   0 0 1  67
Block 5 Group Theory
 a a 0  0 0 1  a  ab b
     
     
D ( p 4 )  S 1D( p 4 )S    ab  ab 2ab  1 0 0  a  ab b
     
 b b  0 0   0 b 
 b  1  2ab

  a2  3a 2 b 0    1/ 2  3 /2 0
   
     D(R1 ) 0
  3a 2 b  3a 2 b 2 0  3 /2  1/ 2 0   
     0 
 
1
  
 0 0 3b 2   0 0 1 

 a a 0  0 0 1  a  ab b
     
     
D ( p 5 )  S 1D( p 5 )S    ab  ab 2ab  0 1 0  a  ab b
     
 b b  1 0   0 b 
 b  0  2ab

 a2  3a 2 b 0   1/ 2  3 /2 0
   
     D(M 2 ) 0
   3a 2 b  3a 2 b 2 0    3 / 2  1/ 2 0   
     0 
 
1
  
 0 0 3b 2   0 0 1 

68
Unit 19 Continuous Groups

UNIT 19
CONTINUOUS
GROUPS
Structure
19.1 Introduction 19.4 Minkowski Space and the
Expected Learning Outcomes Lorentz Group
19.2 Continuous Groups: An Introduction Minkowski Space
Active and Passive Transformations The Lorentz Group
Simplest Continuous Groups †
19.5 SL(2,C) and the Lorentz Group L
Orthogonal Group O(3) †
Homomorphism SL(2,C)  L is 2:1
The Subgroup SO(3) or Rotation Group
Lie Groups or Continuous Groups Six Parameters of SL(2,C)
One Parameter Subgroups of SO(3) Boosts and Hermitian SL(2,C) Matrices
19.3 Generators of One-parameter Rotations and SU(2)
Subgroups Form of a General SL(2,C) Matrix
Commutator of Generators The Matrix  and Other Useful Formulas
Lie Algebra of Generators 19.6 Summary
19.7 Terminal Questions
19.8 Solutions and Answers

19.1 INTRODUCTION
You have so far discussed groups that are finite groups which contain a
number of discrete elements. There can be groups with infinitely many
elements, like the set of all integers with respect to addition. This group has
infinitely many elements, but the elements are denumerable.
Now think about the set of rotations through an angle, say , about a fixed
axis. The rotations form a group, but the group is labelled by the continuous
parameter . There are infinitely many elements in this group and they are not
denumerable either. This is an example of a continuous group.
In this unit we discuss continuous groups in physics which are related to
symmetry transformations. In Sec. 19.2 we introduce continuous groups and
discuss some simple continuous groups. In Sec. 19.3 we discuss about the
generators of one-parameter subgroups. In Sec. 19.4 we will talk about
Minkowski space and the Lorentz group. In Sec. 19.5 we discuss the group of
2  2 complex matrices SL(2,C) and its connection to the Lorentz group. 69
Block 5 Group Theory
Expected Learning Outcomes
After studying this unit, you should be able to:
 define and identify active and passive transformations;

 define continuous groups and give examples;

 determine the generators of one parameter subgroups;

 solve problems based on continuous groups and their generators;

 define Minkowski space and Lorentz group;

 solve problems related to SL (2,C); and

 define the matrix  and solve problems related to it.

19.2 CONTINUOUS GROUPS: AN


INTRODUCTION
In this section we will introduce you to continuous groups, which in physics are
transformation groups. So, first we explain transformations – active and
passive.

19.2.1 Active and Passive Transformations


All concepts of symmetry, and symmetry groups, of physics deal with
transformations. It is important to remember the distinction between active
and passive transformations.

When we think of a rigid body being rotated about an axis by a certain angle, it
is an example of an active transformation. The position coordinates of various
points of the body change after the rotation. This change of the position
coordinates is an active transformation of rotation.

But we can think of rotation in another way.

We can keep the body fixed, but we move the coordinate system by a rotation.
The coordinates of the points of the body in the old coordinate system are
transformed to the coordinates of the same points in the new coordinate
system. This is a passive transformation, where the rigid body remains
stationary and it is the coordinate system describing it which rotates.

In the active case, there is one coordinate system, and infinitely many different
rotations can be imparted to the body.

In the passive case, there are an infinite number of coordinate systems related
by rotations.

Mathematically, both are symmetries and they are equivalent, but the
parameters have opposite signs.

If a body is actively rotated by an angle /4 about the z-axis, a point of the
body on the xy-plane with coordinates (1, 0, 0) will acquire coordinates

70
 
1/ 2,1/ 2, 0 (see Fig. 19.1a).
Unit 19 Continuous Groups
On the other hand if the rigid body is kept fixed, and the coordinate system is
rotated by the same amount, the same point will have coordinates
 
1/ 2,  1/ 2, 0 in the new coordinate system (Fig. 19.1b).

y y

y

x

 x  1/ 2, y   1/ 2 

x  1/ 2, y   1/ 2 
(x = 1, y = 0) (x = 1, y = 0)
x 
Passive
x
Active

(a) (b)

Fig. 19.1: a) Active transformation; b) passive transformation.

Passive and active transformations are both used frequently in physics.

19.2.2 Simplest Continuous Groups


Let us now define continuous groups. Continuous groups or Lie groups are
groups such that each element of the group is specified by one or several (but
finite number of) real parameters. Such groups necessarily have an infinite
number of elements. The group law is the rule which tells us how to calculate
the parameters of the product element given two sets of parameters specifying
the two elements.

Given below are some of the simplest examples of continuous groups.

1. Let  be the set of all real numbers. Then x   specifies the group
element. The group law is the ordinary addition of real numbers. The
identity element is the number 0 and the inverse of x is  x.

2. Let   be the subset of non-zero positive real numbers. The group law is
multiplication of real numbers. The identity is the number 1, and the
inverse of x    is 1/x.

3. The group U(1): Let U(1) be the set of complex numbers of modulus unity,
that is, complex numbers z with z  1 . The group elements can be
pictured as belonging to the unit circle in the complex plane. The product
of two elements of U(1) is also of modulus unity z1z 2  z1 z 2  1 . The
71
Block 5 Group Theory
identity is the number 1 and the inverse of z is z* = 1/z. The group is so
named because complex numbers of U(1) can be considered as 1  1
unitary matrices.

19.2.3 Orthogonal Group O(3)


As a simple and important example, let us consider rotations in three-
dimensional space. Rotations have the property that (1) there is one special
point which remains fixed, and (2) distances between any pair of points of the
rotated body remain unchanged.

We can take the fixed point, O, as the origin of Cartesian coordinates. Let a
general point P of the body have the coordinates ( x1, x 2 , x 3 ) before the
rotation, and ( x1 , x 2 , x 3 ) after the rotation. Then

( x1 )2  ( x 2 )2  ( x 3 )2  ( x1)2  ( x 2 )2  ( x 3 )2

We can write it as a matrix equation:

 x1   x 1 
   
If X   x 2  , X    x 2 
   
x 3   x 3 
   

then

X T X   X T X

If the relationship between the coordinates X and X is linear, then we can


write
X = RX,

where R is a 3  3 real matrix. Written fully,

 x 1  R11 R12 R13   x1 


     
 2    2
 x   R 21 R 22 R 23  x 
     
 x 3  R 31 R 32 R 33  x 3 
   

The condition X T X   X T X implies (see Sec. 8.4 in Unit 8).

RT R  1 or RT  R 1 therefore RR T  1 as well.

Thus, R is a 3  3 real orthogonal matrix.

Taking the determinant, we see that

RT R  RT R   R 2  1, or R  1

We call O(3) the set of all 3  3 real orthogonal matrices. They form a group
because the product of two such matrices is again one such matrix:

72 (R1R 2 )T (R1R 2 )  R T RT R R  RT
2 1 1 2
R 1
2 2
Unit 19 Continuous Groups
The identity matrix trivially satisfies orthogonality condition, and the inverse
R 1  R T is also in the set because

(R 1)T R 1  (RT )T RT  RRT  1

The space inversion (or parity) matrix which reverses the sign of each
coordinate

 1 0 0
 
I s   0 1 0   1
 
 0 0  1

belongs to O(3).

The space inversion Is is actually not a rotation, because it cannot be


physically achieved by rotating a body. The actual rotation group is a
subgroup of O(3) which we discuss next.

19.2.4 The Subgroup SO(3) or Rotation Group

A general element of O(3) can have a determinant either 1 or 1. The


product of two matrices with determinant 1 is again a matrix with determinant
1. Therefore, the subset of O(3) which has matrices with determinant 1 is a
subgroup denoted by SO(3) called the special orthogonal group.

Thus, the group O(3) has two cosets, the subgroup SO(3) and the set of
matrices with determinant –1. Since the space inversion matrix has
determinant equal to –1, any O(3) matrix T with determinant equal to –1 can
be written as:

T = Is R

Note that if T  O(3) then both T and T 1 have the same determinant
because T 1  T T .

SAQ 1

Is SO(3) an invariant subgroup of O(3)? If yes, find the factor group.

19.2.5 Lie Groups or Continuous Groups


O(3) and SO(3) are more elaborate examples of continuous or Lie groups.
As mentioned above these groups have a continuous infinity of group
elements. Each group element is labeled by a number of real parameters.

In our case O(3) needs three real parameters to specify a group element. A
3  3 matrix R has 9 elements. The orthogonality condition R T R  1 is
equivalent to six conditions, ( R T R is a symmetric matrix with six elements
actually independent: three elements on the diagonal and three elements on
one side of the diagonal. Elements on one side of the diagonal are equal to 73
Block 5 Group Theory
corresponding elements on the other side in a symmetric matrix.) Therefore,
there are only three free parameters.

The three independent parameters can be chosen in infinitely many ways.

19.2.6 One Parameter Subgroups of SO(3)

A rotation in which not just a single point, but a whole line remains fixed
is called rotation about an axis.

These are especially simple rotations because if the axis is fixed, a single
parameter, the angle of rotation is enough to specify the rotation completely.

It was proved by Euler (in 1776) that every rotation can be seen as a rotation
about an axis by a definite angle.

What it means is this: although one can arrive at the same final position of a
rigid body in infinitely many possible ways through all kinds of rotations, the
final position can always be obtained from the initial position by a single
rotation about an axis by a certain angle.

P

Q

Fig. 19.2: Rotation about an axis.

Refer to Fig. 19.2. Let the fixed point be O and P and Q two points on the
body. Let the final positions of the two points be P and Q.

The axis can be found as follows: draw a plane perpendicular to POP and
bisecting the angle POP. Similarly, draw a plane perpendicular to QOQ and
bisecting the angle QOQ. If these planes are not coincident, then the
intersecting line of these planes is the axis. If the two planes are coincident,
then the axis is the intersecting line of planes POQ and POQ.

Rotations about a fixed axis can be labeled by giving the direction of the axis,
which requires two parameters, and the angle by which the rotation is made,
which is a number lying between 0 and 2 when the angle is measured in
radians. At angle 2 the rotation becomes equal to identity. One should avoid
the identity element to be given by two different parameters, therefore rotation
by 2 is equated to identity.
74
Unit 19 Continuous Groups
If we fix the axis, then there is only one parameter to specify the rotation.
These rotations about a fixed axis define a subgroup because two successive
rotations by angles 1 and  2 give a rotation by angle 1 +  2 . It is possible
that 1 +  2 exceeds 2. In that case the added angle is the excess over 2.

This one parameter group of rotations with a fixed axis is called the group
SO(2) as rotations in the two dimensional plane perpendicular to the axis. It is
a subgroup of SO(3).

SAQ 2

Find the matrices corresponding to rotations about the z-axis and show that
they are in one-one correspondence with orthogonal 2  2 matrices with
determinant 1.

19.3 GENERATORS OF ONE-PARAMETER


SUBGROUPS
Since rotations about a fixed axis add up, it is possible to build finite rotations
by successive application of a very large number of very small rotations. This
suggests that we should look at the structure of infinitesimal rotations.

An active rotation by an angle  about the 3-axis (in a right-handed Cartesian


coordinate system) is the SO(3) matrix:

cos   sin  0 1   2 / 2!...     3 / 3!... 0


   
 
R3 ()   sin  cos  0     3 / 3!... 1   2 / 2!... 0  ...
   
 0 0 1  0 0 1

where  > 0 means that the rotation is in the right-handed screw sense (that is
anti-clockwise in the plane parallel to 1-2 plane).

For infinitesimal  this matrix is:

0 1 0 1 0 0
   
2 
lim R 3 ()  1    1 0 
0  0 1 0
0 2! 
   
0 0 0 0 0 0

2 2
 1   J3  J  ...
2! 3

The matrix J 3 is:

0 1 0
 
J 3   1 0 0
 
0 0 0 75
Block 5 Group Theory
We can see that

J 32  1, J 33  J 3 , J 34  1, ...

therefore even for finite ,

2 2
R 3 ()  exp( J 3 )  1  J 3  J  ...
2 3

J 3 is called the generator of rotations about the 3-axis because all finite
angle rotations about 3-axis can be built out of it.

Similarly, an infinitesimal rotation about the 1-axis,

1 0 0  1 0 0 
   
R1()  0 cos   sin   0 1   2 / 2!...     / 3!...  ...
3

   
0 sin  cos   0    3 / 3!... 1   2 / 2!... 

can be written as:

2 2
R1()  1  J1  J  ...  exp (J1 )
2 1

with

0 0 0
 
J1  0 0  1
 
0 1 0 

And an infinitesimal rotation about the 2-axis,

 cos  0 sin  
 
R 2 ()   0 1 0 
 
 sin  0 cos 

gives, in a similar manner,

2 2
R 2 ()  1  J 2  J  ...  exp (J 2 )
2 2

with

0 0 1
 
J 2   0 0 0
 
 1 0 0
76
Unit 19 Continuous Groups

19.3.1 Commutator of Generators

We have discussed the generators of rotations about the three coordinate


axes. There will be a similar generator about every axis.

Rotations about axes in different directions do not commute.

We study the simple case of two infinitesimal rotations: R1  R1() about


coordinate axis 1, and R2  R2 () about axis 2,  and  being small.

Let us see how much mismatch is there when these two infinitesimal rotations
are applied one after the other in two different ways.

If R1, R2   R1R2  R2R1 were zero then the result of R1R2 and R2R1 would
have been the same acting on any vector X.

But, in general

R1R2 X  R2R1 X or, multiplying by inverses, R 21R11R 2 R1  1.

The failure of commuting can therefore be estimated by calculating

R 21R11R 2 R1  R 2 ( )R1( )R 2 ()R1()

Keeping quantities up to second order,

R 2 ()R1()  (1  J 2   2 J 22 / 2  ...) (1  J1   2 J12 / 2  ...)

 1  J 2  J1  J 2 J1   2 J 22 / 2   2 J12 / 2  ...

Similarly, changing signs of  and 

R 2 ( )R1( )  1  J 2  J1  J 2 J1   2 J 22 / 2   2 J12 / 2  ...

Therefore,

R2 ()R1()R2 ()R1()  (1  J 2  J1  J 2 J1

  2 J 22 / 2   2 J12 / 2  ...)

 (1  J 2  J1  J 2 J1   2 J 22 / 2   2 J12 / 2  ...)

Keeping again up to second order, we see that

R2 ()R1() R2 ()R1()  1  (J 2 J1  J1J 2 )  ...

This shows that the effect of applying infinitesimal rotation by  about axis 1,
followed by an infinitesimal rotation by  about axis 2, and then an inverse
infinitesimal rotation about axis 1 by –  followed by an infinitesimal rotation
about axis 2 by angle –  is equivalent to an infinitesimal rotation by an angle
 about an axis whose generator is
77
Block 5 Group Theory
0 0 1 0 0 0  0 0 0 0 0 1
      
J2J1  J1J2   0 0 0 0 0  1  0 0  1. 0
 0 0
      
 1 0 0 0 1 0  0 1 0   1 0 0

 0 1 0
 
  1 0 0
 
 0 0 0

 J 3

Thus, the resulting rotation axis is the negative 3-axis!

We have taken rotations about the coordinate axes, but one could take any
arbitrary directions and obtain similar relations between generators.

The relation above shows that the commutator of the generators of rotations
in 1- and 2-directions is the generator in the 3-direction:

J1, J 2   J1, J 2  J 2 , J1  J 3

We can show similarly that

J 2 , J 3   J1 and J 3 , J1  J 2

19.3.2 Lie Algebra of Generators

We calculated the generators of rotations by looking at the infinitesimal


rotations about the coordinate axes. We now discuss a general infinitesimal
rotation.

A general rotation that is a member of SO(3) is an orthogonal 3  3 matrix R


with determinant equal to one. The identity matrix corresponds to no rotation.
Therefore, an infinitesimal rotation is a matrix infinitesimally different from the
identity matrix. We write

R=1+M

where elements of M are all small quantities.

The condition of orthogonality requires that

1  RT R  (1  M )T (1  M )

 1  M T  M  second order of smallness

Therefore, in the limit of infinitesimal rotation,

M T  M  0, or M T  M.

The infinitesimal rotation matrix differs from the identity matrix by an


antisymmetric matrix.
78
Unit 19 Continuous Groups
We have already seen that the generators J1, J 2 , J 3 in the last section are
antisymmetric matrices.

There are as many generators as the number of possible rotation axes, which
are infinite in number. These generators are all possible antisymmetric
matrices.

But, the set of all generators, that is, the set of all anti-symmetric 3  3 real
matrices forms a real vector space (of dimension 3 as we see below) because
the sum of two antisymmetric matrices is also antisymmetric and if we multiply
an antisymmetric matrix by a real number it remains antisymmetric.

Moreover, the commutator of two antisymmetric matrices is also


antisymmetric. Let M and N be two antisymmetric matrices, then, using
MT  M and NT  N , we have

[M, N]T  (MN  NM)T  NT MT  MT NT  NM  MN  [M, N]

What is the dimension of this space?

A general 3  3 real matrix has 9 independent elements. If it is antisymmetric,


the diagonal elements are zero, and the remaining 6 elements are pairwise
negative to each other. Thus there are only three independent matrices; all
others can be written as linear combination of these three.

0 a b
 
M    a 0 c 
 
 b c 0 

0 1 0 0 0 1 0 0 0
     
 a  1 0 0  b  0
 0 0  c 0
 0  1
     
0 0 0  1 0 0 0 1 0 

 cJ1  bJ 2  aJ3

Therefore every generator of SO(3) is a member of the vector space of 3  3


antisymmetric matrices. Moreover this vector space has the commutator of
two elements again belonging to the same space. The commutator acts like a
bilinear product,
[M, N  P ]  [M, N ]  [M, P ], [aM , N ]  a[M, N ],

[M  N, P ]  [M, P ]  [N, P ], [M, aN ]  a[M, N ]

In addition, it is antisymmetric

[M, N] =  [N, M]

and satisfies the Jacobi identity:

[L, [M, N]] + [M, N], L]] + [N, L], M]] = 0 79


Block 5 Group Theory
Definition

Any vector space which has a bilinear relation defined on it is an algebra. A


vector space (like the one above) with an antisymmetric bilinear product
satisfying Jacobi identity is called a Lie algebra.

The bilinear ‘product’ here is the commutator of two matrices. Most


products encountered in physics are associative products, for example,
the matrix multiplication, (AB)C = A(BC). But the Jacobi identity shows
that the product in a Lie algebra is non-associative: [[M, N], P]  [M, [N, P]]
in general. The only other place a non-associative product is
encountered in elementary physics is the Poisson bracket in classical
mechanics. There too, there is a Lie algebra of phase space functions.

Since every element of the Lie algebra can be written as a linear combination
of its basis elements, it is enough to specify the commutator only for the basis
elements.

In a Lie algebra of dimension n with basis E1,E2,..., En we can write [E i , E j ]


which again belongs to the vector space of the Lie algebra as a linear
combination of the basis vectors:

[E i , E j ]   c ij k E k ,
k

where the n 2 (n  1) / 2 real numbers c ij k  c ji k are called the structure


constants of the Lie algebra.

SAQ 3

Why are there n 2 (n  1) / 2 structure constants for a Lie algebra of dimension


n?

Important note on notation:

The Lie algebra of SO(3) is a real algebra with real structure constants.

But when the rotation group is used as a symmetry in Quantum Mechanics,


the rotations are represented by unitary operators, and infinitesimal rotations
can be written as

Uˆ (R)  1  Tˆ  ...

where Tˆ is an infinitesimal operator. The condition for unitarity means

1  U(R )† U(R )  1  Tˆ †  Tˆ  ...

Therefore the generators Tˆ of rotation in Quantum Mechanics are


anti-hermitian operators (Tˆ †  Tˆ ).

The commutator of two anti-hermitian operators is again anti-hermitian, so the


Lie algebra is a real Lie algebra, with real structure constants as in the Lie
algebra of rotation matrices.
80
Unit 19 Continuous Groups
However, the generators of all symmetry transformations are
proportional to physical quantities or observables, which are
represented by hermitian operators.

It is a time honoured practice to redefine generators with a factor proportional


to imaginary unit and convert the anti-hermitian operator into a hermitian
operator.

Thus, usually the operator version of the generators are defined as:

U (1  J1  ...)  1  iJˆ1  ...

U (1  J 2  ...)  1  iJˆ 2  ...

U (1  J 3  ...)  1  iJˆ3  ...

with the result that

[Jˆ1, Jˆ2 ]  iJˆ3 , and so on.

This may give the impression that the structure constants are imaginary, but
that is not so. As is well known, the generators Ĵ are called components of
angular momentum.

We have given the generators starting from the active rotation matrices. In
many books, as well as in this unit, passive transformations are used as well.
There the relations given above will appear as:

U(1  J1  ...)  1  iJˆ1  ...

and so on. You must be careful about these signs. Of course, just as bases
can be chosen in infinitely many ways, the generators too can be chosen in
every possible way. But the basis with [J1, J 2 ]  iJ3 , and its cyclic versions,
are universally accepted norm for angular momentum.

19.4 MINKOWSKI SPACE AND THE LORENTZ


GROUP
We now discuss Minkowski space and the Lorentz group. For this, let us first
revise Lorentz transformation. A word about notation.
Notation: We use units such that time is measured in metres with the
help of the universal constant c which represents velocity of light in
vacuum. This, effectively makes c = 1 in relativity formulas, and
velocities become dimensionless numbers.

Recall the familiar Lorentz transformation from your UG courses:

 t    1 1  v 2 v 1 v 2 0 0  t 
   
     
 x   v 1 v 2 1 1 v 2 0 0 x
    
 y    y 
   0 0 1 0  
     
 z   z
   0 0 0 1    81
Block 5 Group Theory
It relates the same event P which is specified by (t, x, y, z) and (t , x , y , z ) ,
respectively in the two frames of reference S and S (which means we are
using passive transformation). The two frames are such that

1. their axes are parallel,


2. the x and x-axes are collinear,
3. S moves with velocity v with respect to S along the x-axis,
4. the clocks are set so that when the origins of the two frames coincide,
t  0  t .

This is written as the matrix equation:

x = Lx

where x is the single column matrix

x0   t 
   
   
 x1   x 
  
 2  y 
x   
   
x3   z 
   

and similarly for x. The matrix L is a 4  4 matrix.

The example given above is a special Lorentz transformation called a “boost”


in the 1-direction.

Definition

A Lorentz transformation is a real 4  4 matrix L which transforms x into


x = Lx such that

 ( x 0 )2  ( x 1 )2  ( x 2 )2  ( x 3 ) 2  ( x  0 )2  ( x 1 ) 2  ( x  2 ) 2  ( x  3 ) 2

This condition can be written more compactly using a matrix,

1 
 
 
 1 
 
 1 
 
 
 1

as

x T x  x T x 

Therefore, if x   Lx then the condition for L to be a Lorentz matrix is:

x T x   xT LT Lx  xT x
82
Unit 19 Continuous Groups
Since this is true for any arbitrary x, it follows that

LT L  

Thus, any real 4  4 matrix that satisfies the above equation is a Lorentz
transformation.

In the next two sub-sections we discuss Minkowski space and Lorentz group.

19.4.1 Minkowski Space

We can think of a 4-dimensional vector space (in analogy to the 3-dimensional


space) where ‘position vectors’ are:

 x0 
 
 
 x1 
x  
 2
x 
 
 x3 
 

and where the inner product or ‘dot product’ between two vectors x and y is
given by

x, y  x . y  x T y

This four-dimensional space is called the Minkowski space.

Just as in three-dimensions, rotations do not change the dot product of two


vectors, similarly, the Lorentz transformations do not change the above dot
product.

But unlike the three-dimensional space, the dot product of a vector with itself
is not necessarily positive. It can be positive, negative or even zero. We still
call

x . y  xT x  ( x 0 )2  ( x1)2  ( x 2 )2  ( x 3 )2

as the “length squared” of a vector.

The Minkowski space gets divided into three regions with respect to the origin.

It is conventional to imagine the x 0 -axis or the time axis as a vertical line and
the plane perpendicular to it carrying the spatial axes corresponding to
x 1, x 2 , x 3 .

The set of vectors corresponding to x . x = 0 are called light-like or null


vectors.

Null vectors form a double cone. This is the surface in Minkowski space
satisfying

( x 0 )2  ( x 1)2  ( x 2 )2  ( x 3 )2
83
Block 5 Group Theory
or

( x 0 )   ( x 1)2  ( x 2 )2  ( x 3 )2

with the vertex at the origin (0,0,0,0)T and vertex angle of 45. This cone is
called the light cone. The time axis is the axis of the cone. The positive side
of the cone (that is, the half-cone through which the positive half of the time
axis passes) is called the forward light cone and the negative side of the
cone the backward light cone.

All vectors with x.x  0 lie within the light cone and they are called time-like
vectors.

The remaining vectors with x.x  0 lie in the wedge-like region and they are
called space-like vectors.

19.4.2 The Lorentz Group


Let L be the set of all 4  4 Lorentz transformation matrices. These form a
group, called the Lorentz group with matrix multiplication as the group law.
The inverse of each matrix exists because det L  0. We know that the
determinant of any Lorentz matrix can be either +1 or 1. This follows from
LT L   and taking determinant on both sides to obtain (det L)2  1.

You can check that

L1  LT 

SAQ 4

a) Prove that

 L00  L10  L20  L30 


 
 
  L01 L11 L21 L31 
L1  LT    ,
 L L12 L22 L32 
 02 
 
 L L33 
 03 L13 L23

so that the inverse Lorentz matrix is just the transposed matrix with 0i and
i0 elements occurring with a minus sign.

b) Show that if L00  1 then not only

L00   L10
2  L2  L2
20 30

(which follows from the definition) but also

L00   L201  L202  L203


84
Unit 19 Continuous Groups

19.5 SL(2,C) AND THE LORENTZ GROUP L

The proper orthochronous Lorentz group L has a deep relation with the
group SL(2, C) of 2  2 complex matrices.

SL(2, C) stands for Special Linear Group of 2  2 complex matrices. The


only special thing about elements of SL(2, C) is that their determinant is equal
to one. The fundamental fact about these matrices is this:

For every A  SL(2, C) there is a proper orthochronous Lorentz


transformation L(A)  L such that

L( A1 ) L( A2 )  L( A1 A2 )

To see this, let us consider the set of all 2  2 hermitian matrices.

Let us show that all 2  2 hermitian matrices form a four-dimensional real


vector space.

The sum of two hermitian matrices is hermitian, as well as multiplication by a


real number.

A 2  2 hermitian matrix has the general form:

 1  2  i 3 
 
 
  2  i 3 4 

where the ’s are four real numbers.

We can choose the following four matrices as a basis in this vector space.
These four standard hermitian matrices are called Pauli matrices:

1 0 0 1
0    1   
   
0 1 1 0

0 i 1 0
2    3   
   
i 0 0  1

Any four linearly independent hermitian matrices would have been equally
good, but these are the ones traditionally used and they have the properties
listed below:

i) The squares of all the Pauli matrices are equal to the identity

02  1, 12  1, 22  1, 32  1

ii) Matrices i  1, 2, 3 have trace zero, and

iii) 1 2  2 1  i 3 and other similar relations are obtained by rotating


1,2,3 cyclically. 85
Block 5 Group Theory
These properties imply that

Tr ( )  2

where  is the Kronecker delta, equal to zero when    and 1 when the
two indices are the same.

Caution about notation: Although the placing of subscripts on the


constant Pauli matrices is like that of a covariant vector, it is not a
covariant 4-vector. We will always use Pauli matrices with a lower index.

With a given space-time point with coordinates x   we form a hermitian


matrix,

3  x0  x3 x 1  ix 2 
 
X   x   
 x 1  ix 2 
 0  x 0  x 3 

whose determinant is

det X  x 0   x 1   x 2   x 3    x . x.
2 2 2 2

Let A  SL(2, C) be an arbitrary 2  2 complex matrix with determinant 1 and


X as above. We define

Y  AX A†

where A† is the hermitian conjugate of A. The matrix Y is also hermitian and,


therefore, can be expanded in the basis of   matrices:

3
Y   y  
 0

Moreover,

y . y  y T y   detY   det X  xT x  x . x

because the determinant of A as well as A† is equal to one.

What is the relation between coordinates y  and x  ? Certainly y depends


linearly on the x. Therefore, these coordinates are linearly related. We write
3
y   L( A) x 
 0

L(A) has to be a Lorentz matrix because y . y = x . x.

To obtain how L(A) depends on A, we proceed as follows.

We know that the defining relation for the Lorentz transformation


corresponding to A  SL(2, C) is:

AX A †  A( x    )A †  y     (   x  )  ,   L( A)
86
Unit 19 Continuous Groups
It holds good for any point x. Therefore, we must have:

A  A †     

Multiplying on the left by  and taking the trace, we find that

1
   L( A)   Tr (  A  A † )
2

By applying two successive transformations, we get:

X  Y  A1 X A1† , y   L( A1 )  x 

and

Y  Z  A2Y A2† , z   L( A2 )  y 

We deduce that

Z  A2 ( A1X A1† ) A2†  ( A2 A1) X ( A2 A1)† , z   L( A2 A1) x 

so that

z   L( A2 )  L( A1 )   L( A2 A1 )  x 

Therefore,

LA2 LA1   LA2 A1 

We can work out the relationship between A and L(A) in the following detailed
and very useful form:

a b 
A , ad  bc  1
c d 
 


L00  a 2  b 2  c 2  d 2 2 
L01   ab * cd *

L02   ab * cd *


L03  a 2  b 2  c 2  d 2 2 
L10   a * c  b * d  L11   ad *  b * c 

L12   (ad * b * c ) L13   ac *  bd *

L20   (a* c  b * d ) L21   (a* d  b * c )

L22   (a * d  bc*) L23   ( a * c  bd *) 87


Block 5 Group Theory
L30  a  2
b 2
c 2
d 2
2
L31   ab *  cd * L32   ab *  cd *


L33  a 2  b 2  c 2  d 2 2 
Here the real and imaginary parts of a complex number z are denoted by (z )
and (z ) , respectively.

We just calculate two of the elements. You can calculate the remaining
L (A) yourself.

1 1
L00  Tr ( 0 A 0 A † )  Tr AA†
2 2

1  a b

a *

c *

 Tr
2  c d  b *
 d * 

1  a  b ca * db * 
2 2
 Tr
2  ac *  bd * c2  d2


1 2
2

a  b2  c2  d 2 
Similarly,

1
L12  Tr (1A 2 A † )
2

1  0 1 a b  0  i  a * c *
 Tr     
2  1 0  c
 d   i 0   b * d * 

1 c d    ib *  id * 
 Tr   
2 a b   ia * ic * 

1   icb * ida *  

 Tr
2    iad * ibc * 

 ad * b * c 

Proof of L( A)  L†

From the expressions above, we see that L00 ( A)  0 for any A. As L(A) is a
Lorentz transformation, it maps a time-like vector to a time-like vector. This
shows that the L(A) preserves the time direction.

To show that det L(A) = +1 we must prove that there is no A which can give
space inversion.
88
Unit 19 Continuous Groups
The formulas above show that for L( A)  I s , to be true, L00  L33  0 implies
that a 2  d 2  0 or a = 0 and d = 0.

As det A = ad – bc = 1, this means bc = –1.

But L11  L22  0 implies that (b * c )  0, and L12  L21  0 implies that
(b * c )  0, giving b * c  0. This is a contradiction because bc = –1.

19.5.1 Homomorphism SL(2,C)  L† is 2 : 1

The formula relating A  L(A) also shows that the correspondence is a 2 : 1


mapping, with A and – A giving the same L(A)  L† .

But, how do we know that there are just two, and not more A’s for a given
L(A)?

We try to find its kernel.

We ask in the following example: If L(A) = 1 (the 4  4 unit matrix), then what
should A be?

Example 19.1

Using the formulas above, show that the only possibilities are A = 1, or A = – 1
(the 2  2 unit matrices).

Solution : Given the formulas above for L (A) we find that if L(A) = 1, then /

i) L03  0, and L30  0 imply that L03  L30  0 which gives a 2  d 2 and
c2  b2.

ii) L00  1 and L33  1 imply that L11  L33 which gives b 2  c 2  0
meaning that b = 0 and c = 0.

iii) L01, L10, L02, L20, L13, L31, L23, L32 become zero because of b = 0, c = 0.

iv) L11  L22  (ad *)  1. And L12  (ad *)  0 as well as (a * d )  L21  0.
This means that ad* is a real number. Therefore, ad* = a*d = 1. The
determinant of an SL(2, c) matrix is equal to one, that is ad – bc = ad = 1.
Therefore, a = 1/d = 1/d* or d = d* is real. So is a = 1/d*.

v) a 2  d 2 , and therefore, d   a. But ad = 1, so that ad  a2  1. As a is


real, only d  a is allowed. Therefore, a2  1 and both a and d can be +1
or –1, but both must have the same sign.

vi) Therefore A can be 1 or A = – 1, proving the mapping SL(2, C)  L† to be


2 : 1.
89
Block 5 Group Theory
19.5.2 Six Parameters of SL(2,C)

A typical matrix in SL(2, C) is:

a b
A , ad  bc  1
 
c d

where a, b, c, d are complex numbers with one complex condition among


them. This means that there are eight real parameters with two real conditions
among them. Such a matrix is parametrized by six parameters equal to the
number of independent parameters in the Lorentz group (three for rotations
and three for boosts).

19.5.3 Boosts and Hermitian SL(2,C) Matrices


Every hermitian SL(2, C) matrix can be written in the standard form:

A  cosh(  / 2)  n.  sinh(  / 2),

because if A hermitian, it can be written as A   a with real a  , and


the determinant det A = 1 means that a02  a12  a22  a32   1. Such matrices
represent boosts in the direction of the unit vector n with velocity given by
tanh .

As an example
e / 2 0 
 
B3 ( )  cosh(  / 2)   3 sinh(  / 2)   
 0 e  / 2 

corresponds to
 cosh  0 0 sinh  
 
 
 0 1 0 0 
L(B3 ( ))   
 0 0 1 0 
 
 
 sinh  cosh  
 0 0

Note that if  > 0 then this is an active Lorentz boost in the direction of
coordinate 3-axis, with velocity v = tanh . This boost transforms the
momentum four vector (m, 0, 0, 0) of a particle of mass m at rest, to that of the
particle moving along the 3-axis with momentum 4-vector
p   ( p 0  m cosh , 0, 0, p 3  m sinh ). The speed of the particle is:
p
v  tanh .
p0

SAQ 5

Show that if A is hermitian, then L(A) is symmetric.


90
Unit 19 Continuous Groups
Let us take up another example.

Example 19.2

The SL(2, C) matrix Bp for a boost in p direction that transforms a four vector
(m, 0, 0, 0) representing the four-momentum of a particle of rest mass m in its
 
rest frame into p 0  p 2  m 2 , p , that is

 p0  m
   
   
 p1  0
   L(Bp )  
 2 0
p   
   
 p3  0
   

is given by

B p  cosh(  / 2)  n .  sinh(  / 2).

Show that we can write it as

m 0  p   
Bp 
2mp 0  m 

where n  p / p and cosh( )  p 0 / m, sinh( )  p / m.

Solution : We know that

p0
2 cosh2 ( / 2)  cosh   1  1
m

Therefore,

cosh   1 p0  m p0  m
cosh(  / 2)   
2 2m 2m( p 0  m )

and, similarly,

cosh   1 ( p 0  m)
sinh(  / 2)  
2 2m

( p 0 )2  m 2 p
 
2m( p 0  m ) 2m( p 0  m )

It is worth remembering that boosts do not form a subgroup because the


product of two Hermitian matrices is not Hermitian unless they commute with
each other.

91
Block 5 Group Theory

SAQ 6
A particle of rest mass zero moves with four-momentum
p (0)   ( , 0, 0, ),   0. Find the SL(2, C) matrix which represents the
boost along the 3-direction and which takes (, 0, 0, ) to ( p 0 , 0, 0, p 0 ) with
p 0  0.

19.5.4 Rotations and SU(2)


Unitary SL(2,C) matrices form a subgroup of SL(2,C) and they correspond to
rotations. Let us demonstrate this in Example 19.3.

Example 19.3

a) Show that unitary matrices of SL(2, C) form a subgroup.

b) Show that for A  SU(2), the Lorentz matrix L(R) is of the form:

1 0 0 0
 
0 
 
L( A)   
0 Rij 
 
 
0 

where R ij is a 3  3 rotation matrix RT R  1, det R  1.

Solution : a) The product of two unitary matrices is unitary:

U 2U1 † U 2U1   U1† U 2† U 2U1  U1† U1  1

As the inverse U 1 is equal to hermitian adjoint U † and

U † † U †  UU †  1

the inverse matrix is unitary. The identity is trivially a unitary matrix.

This subgroup of SL(2, C) is denoted by SU(2) (for special unitary matrices


of dimension 2).

b) If A is unitary A †  A 1 , then

1 1
L00 ( A)  Tr ( 0 A 0 A † )  Tr ( AA 1 )  1
2 2

Therefore, (see SAQ 4b in Sec. 19.4.1).

L10  L20  L30  L01  L02  L03  0


92
Unit 19 Continuous Groups
The Lorentz matrix has the form:

1 0 0 0 
 
0 R11 R12 R13 

 
0 R 21 R 22 R 23 
 
 
0 R 31 R 23 R 33 

and the fact that column vectors of a Lorentz matrix form an orthonormal
basis, shows that R is an orthogonal matrix.

Any SU(2) matrix can be written as

  
A   0 cos  in .  sin
2 2

where

n .   n11  n 22  n 3 3

and n  (n1, n 2 , n 3 ) is a unit vector. This follows from the fact that for such a
matrix A †  A 1 , so that if

a b
A  , ad  bc  1
c d 

then

d  b  a * c *
A 1     A†
 c a  b * d * 

Comparing the two sides, we get two conditions

d  a*, c  b *

Therefore, a general SU(2) matrix has the form

 a b
A , a2  b2 1
 b * a *

If a  a1  ia2 and b  b1  ib2 , (the real and imaginary parts), then

 a1  ia2 b1  ib2 
A   a1 0  ia2  3  ib2 1  ib1 2
 b  ib a1  ia2 
 1 2

with the condition that

a 2  b 2  a12  a 22  b12  b22  1


93
Block 5 Group Theory
We can see that the coefficients of 1,  2 ,  3 are i times b2 , b1, a2 . We
choose a1  cos  / 2. Then

b22  b12  a 22  1  cos 2  / 2  sin 2  / 2

We define the unit vector n by

(b2 , b1, a2 )  n sin( / 2)

so that A can be written as:



A  cos( / 2)0  i sin( / 2)n. 

The nature of rotation R can be found by looking at simple examples. If

e i / 2 0   
A    0 cos  i 3 sin
 0 e i / 2  2 2

then n = (0, 0, 1) and

1 0 0 0
 
 
0 cos  sin  0
L(A)   
0  sin  cos  0
 
 
0 1 
 0 0

which is a rotation about the axis in the direction of n.

19.5.5 Form of a General SL(2,C) Matrix

Any SL(2, C) matrix can be written as a product of a boost and an SU(2)


matrix. Specifically,

A  V H, H  ( A† A)1/ 2 , V  A( A† A) 1/ 2

The square-root (and square root inverse) of ( A † A) is well defined because it


is a positive definite hermitian matrix, that is, a matrix with strictly positive
eigenvalues. The proof is especially easy because A † A it is a 2  2 hermitian
matrix. It has determinant (equal to product of the two eigenvalues) equal to 1
and the trace (equal to the sum of the eigenvalues) is
 a 2  c 2  b 2  d 2   0 where a, b, c, d are elements of A.
 

To calculate the square root of a positive definite hermitian matrix, we first


diagonalize it by a similarity transformation. Then take the square-root of the
positive eigenvalues along the diagonal. Then go back to the original basis by
the inverse of the same similarity transformation.
94
Unit 19 Continuous Groups

19.5.6 The Matrix  and Other Useful Formulas

A very useful and interesting matrix is the SU(2) element

0 1
  .
 
 1 0

which corresponds to a rotation by  about the y-axis. It is antisymmetric real


matrix whose negative is its inverse

(  )   1

It converts any SL(2, C) matrix to its inverse-transpose as follows:

 A  1  ( A 1 )T ,

as can be immediately verified.

Parity inverted Pauli matrices  P


 are defined as:

1 0 0  1
 0P  0   , 1P  1   ,
   
0 1 1 0

0 i 1 0
P      , P      .
2 2   3 3  
 i 0 0 1

These matrices are related to the standard Pauli matrices by

 
    1   P
 *

which can be verified directly.


The basic formula connecting SL(2, C) to L is:

A  A †     

Multiplying by   and summing over , we get:

  A  A †      ( T ) 

   ( T )      

  
1 P

Taking the complex conjugate of the above equation and multiplying by  and
1 on two sides, we get

( A 1 )   A 1  
1    ( 1T )
B B  95
Block 5 Group Theory
Therefore, the SL(2, C) matrix which generates 1T is the inverse-dagger of
the matrix which generates .

Let us now summarise what we have studied in this unit.

19.6 SUMMARY
In this unit we have discussed the following concepts:

 continuous groups: an introduction;

 active and passive transformations;

 simplest continuous groups;

 orthogonal group O(3);

 the subgroup SO(3) or rotation group;

 Lie groups or continuous groups;

 one parameter subgroups of SO(3);

 generators of one-parameter subgroups;

 commutators of generators;

 Lie algebra of generators;

 Minkowski space and the Lorentz group;

 Minkowski space;

 Lorentz group;

 SL(2,C) and the Lorentz group L ;


 homomorphism SL(2,C)  L is 2:1;

 six parameters of SL(2,C);

 boosts and hermitian SL(2,C) matrices;

 rotations and SU(2);

 form of a general SL(2,C) matrix; and

 the matrix  and other useful formulas.

19.7 TERMINAL QUESTIONS


1. Find the rotation matrix (of active transformation) which corresponds to a
rotation about an axis in the direction of a unit vector n by an angle .

2. Calculate the structure constants for the Lie algebra of SO(3).


3. Prove that if L  L then for a positive time-like vector x  (x 0 , x ), x 0  x ,
96 the vector x   Lx , is also positive time like: x  0  x  .
Unit 19 Continuous Groups
4. Show that the four column vectors of a Loretnz matrix form an orthonormal
basis in Minkowski space. (This result is similar to that for the orthogonal
matrices.)

5. Calculate L(A) for A = cosh (/2) + n .  sinh(  / 2).

6. Show that

  
A   0 cos  in .  sin
2 2

corresponds to a passive rotation by an angle  about the axis in the


direction n. Compare the result with the SO(3) matrix of TQ 1 after changing
the sign of .

19.8 SOLUTIONS AND ANSWERS


Self-Assessment Questions
1. Yes, because if R  SO(3) and T  O(3) then

TRT 1  T R T 1  R  1

There are two cosets. Apart from SO(3), the other is obtained by
multiplying elements of SO(3) by the matrix I s  1. The factor group is
isomorphic to the multiplicative group of two elements (1, 1).

2. An active rotation about the z-axis in the right-handed screw sense


changes only the x-and y-coordinates (see Fig. 19.3).

x   R cos(    )  x cos   y sin ,



y   R sin(    )  y cos   x cos 


  x = R cos, y = R sin 

Active x

Fig. 19.3: Diagram for solution of SAQ 2.

x   x cos   y sin 

y   y cos   x sin 

z  z 97
Block 5 Group Theory
Therefore, the SO(2) matrices are:

cos   sin 
 
 sin  cos  

and they generate a one-parameter additive group:

cos   sin  cos   sin   cos(   )  sin(   )


    
 sin  cos    sin  cos    sin(   ) cos(   ) 
 

3. The structure constants are zero for i = j. Because of antisymmetry,


c ij k  c ji k , so there are n(n – 1)/2 independent pairs for each k which
can take n values.

4. a) (LT )L  (LT L)    1.

The Lorentz group matrices L can be divided into two mutually


exclusive subsets:

i) those for which det L = +1, and

ii) those for which det L = 1.

They are called, respectively, the proper and improper Lorentz


groups.

The former, written as L with det = +1 form a subgroup, but those


with det = 1 (called L ) do not form a subgroup because the product
of two matrices with determinant –1 is +1, not –1.

These two subsets are further divided into subsets, as can be seen
below:

The equation LT L  , when written for 00 element, is

00  1   (LT )0 () L0


, 

 (L00 )2  (L10 )2  (L20 )2  (L30 )2

In other words

(L00 )2  1  (L10 )2  (L20 )2  (L30 )2 .

so that (L00 )2  1 and therefore, either L00  1 or L00  1.

Thus, denoting matrices with L00  1 by an ‘up arrow’ and those with
L00  1 by a ‘down arrow’,

L  L  L , L  L  L


98
Unit 19 Continuous Groups

Members of L , that is, Lorentz matrices with determinant +1 and


L00  1 are called proper orthochronus transformations. They form a
subgroup within L .

b) We look at the inverse matrix L1. Then L001  L00 and Li 01  L0i .
Therefore,

L00  L001   (L10


1 )  (L1 ) 2  (L1 ) 2   L2  L2  L2
20 30 01 02 03

5. As A †  A, the expression for L(A) is

1 1
L ( A)  Tr (  A  A)  Tr (  A  A)  L ( A)
2 2

using the property of the trace Tr(AB) = Tr(BA).

6. The boost in 3-direction is of the forms

e / 2 0 
B3   ,
   / 2 
 0 e 

leading to Lorentz transformation which must change  to p 0 .

 p 0   cosh  0 0 sinh    
     
     
 0   0 1 0 0  0
    
   0 0 1 0  0
 0     
     
 p 0   sinh  cosh    
   0 0  

Therefore,

p 0  (cosh   sinh )  e  , that is e  / 2  p 0 / 

and

 p0 /  0 
 
B3   
 0  / p 0 

Terminal Questions

1. Let the fixed point of rotation be at the origin, and P = r = (x, y, z) be the
point which gets rotated about axis n by  to a new point
P = r = (x, y, z). We have to find P = r. (See Fig. (19.4a).

From Fig. 19.4b, we can arrive at the vector OP as the following sum:
   
OP   OA  AB  BP 

OA is in the direction of the axis n. 99


Block 5 Group Theory
AB is perpendicular to n and in the plane OAP.

BP is perpendicular to the plane OAP.

The length OA is equal to n . r = r cos where  is the angle AOP


between n and r and r  r .

The length AP = AP is equal to n  r  r sin 

Length AB = AP cos = AP cos = n  r cos 

Length BP = AP sin = n  r sin 

The unit vector in the direction of BP is m = (n  r)/ n  r

The unit vector in the direction of AB is 1 = m  n = (n  r)  n/ n  r


   
Thus, r   OP   OA  AB  BP 

 n . r n  n  r cos  1  n  r m

n  r   n n  r 
 n . r  n  n  r cos   n  r sin 
nr nr

 n . r n  cos  n  r   n  sin  n  r

 (1  cos ) n . r  n  (cos ) r  (sin ) n  r

P
P
A

P


A B P

(a) (b)

Fig. 19.4: a) Diagram for solution of TQ 1; b) Enlarged part of a) showing the


100 region ABPP.
Unit 19 Continuous Groups
2. If we take the basis J1, J2, J3 for the Lie algebra of SO(3), then in
[J i , J j ]   c ij k J k , the nine structure constants are:
k

c12 k  0, 0,1 for k  1, 2, 3, respectively

c 23 k  1, 0, 0 for k  1, 2, 3, respectively

c 31k  0,1, 0 for k  1, 2, 3, respectively

3. We already know that x 0   x 2  x  0   x  2  0. We only need to


2 2

prove that x 0  0 .

x  0  L00 x 0  L01x 1  L02 x 2  L03 x 3

On the right-hand side the terms L01x 1  L02 x 2  L03 x 3 are like the dot
product of two three-dimensional vectors a  (L01, L02 , L03 ) and
x  ( x 1, x 2 , x 3 ). Therefore,

L01x 1  L02 x 2  L03 x 3  x 0   a . x  a x


2

The lowest value that L01x 1  L02 x 2  L03 x 3 can have is:

 a x   L201  L202  L203 x

But

L00  L201  L202  L203 and x0  x

Therefore,

( x )0  L00 x 0  L01x1  L02 x 2  L03 x 3

 L00 x 0  L201  L202  L203 x

0

On the other hand, members of L flip the time sense. That is so because
every member of L is equal to a matrix of L multiplied by the time
inversion matrix I t   . The former does not flip the time sense but the
latter does. The product of two L matrices is, of course, in L .

4. The defining equation for the Lorentz matrix is:

LT L   101
Block 5 Group Theory
We call the column vectors of the matrix as:

L00  L01  L02  L03 


       
L  L  L  L 
 10   11   12   13 
E0   , E1   , E2   , E3   ,
L20  L21  L22  L23 
       
       
L30  L31  L32  L33 

Then the defining equations can be written as:

E , E  E E   L  L  (LT L)  

5. Let n = (n1, n2 , n3 ) be the unit vector with n12  n 22  n 32  1. We write


n  n1  in2 , c  cosh(  / 2) and s  sinh(  / 2) . Then

 c  n3 s ns 
A   
 n s c  n 3 s 

Note that a and d are real and

n   n1, n   n 2 , n 2  n12  n 22 , n 2  2n1n 2 .

To calculate the various products in the expressions for the matrix


elements L(A), it is helpful to first calculate quantities like ab*, etc.

a* b* c* d*

a (c  n 3 s ) 2 (c  n3 s )n  s (c  n3 s )n  s (c 2  n 32 s 2 )

b (c  n 3 s ) n  s (n12  n 22 ) s 2 n 2 s 2 (c  n 3 s ) n s

c (c  n3 s )n  s n 2 s 2 (n12  n 22 )s 2 (c  n 3 s )n  s

d (c 2  n 32 s 2 ) (e  n3 s )n  s (c  n3 s )n  s (c  n 3 s ) 2

You can show that

L00  cosh 

L11  1  n12 (cosh   1)

L22  1  n 22 (cosh   1)

L33  1  n 32 (cosh   1)

L01  n1 sinh   L10


102
Unit 19 Continuous Groups
L02  n2 sinh   L20

L03  n3 sinh   L30

L12  n1n2 (cosh   1)  L21

L13  n1n3 (cosh   1)  L31

L23  n2n3 (cosh   1)  L32

This kind of boost is called direct boost like the B3 above. It is used to
connect the state of a particle (of proper mass m) at rest when it is boosted
in the direction of p to a state when its three momentum becomes p. The
Lorentz matrix in the direction of momentum given by the unit vector
n  p / p , so that it acquires a speed

p
v  tanh  
p 2  m2

6. With notation

 
n  (n1, n 2 , n 3 ), n12  n 22  n 32  1, c  cos , s  sin ,
2 2

our matrix is,

a b   c  in3 s n 2  in1  s 
A   d  a*, c  b*
   
c d    n 2  in1  s c  in3 s 

We use the formulas connecting A to L(A).

We have already proved that a unitary matrix like this will be a pure
rotation for which L00  1, L0i  Li 0  0. For the remaining:

L11  (ad * b * c )  (a 2  c 2 )  cos   (1  cos )n12

L12  (ad * b * c )  (a 2  c 2 )  u 3 sin   (1  cos )n1n2

L13  2(ac*)  n2 sin   (1  cos )n1n3

L21  (a * 2 c 2 )  n3 sin   (1  cos )n1n2

L22  (a * 2 c * 2 )  cos   (1  cos )n 22

L23  2(ac*)  n1 sin   (1  cos )n2n3

L31  2(ac )  n2 sin   (1  cos )n1n3

L32  2(ac )  n1 sin   (1  cos )n2n3

L33  a 2  c 2  cos   (1  cos )n 23


103
Block 5 Group Theory
Comparison with the result of TQ 1 with     is

r   (1  cos ) (n.r ) n  (cos ) r  (sin ) n  r

or, written in detail, it is:

x 1  (1  cos ) n1x 1  n2 x 2  n3 x 3 n1

 x 1 cos   n2 x 3  n3 x 2 sin 

 
 cos   (1  cos ) n12 x 1  n 3 sin   (1  cos ) n1n 2  x 2

  n2 sin   (1  cos )n1n3 

 L11x 1  L12 x 2  L13 x 3 .

Similarly, you can determine for x  2 and x  3 .

104

You might also like