Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

FoundationsofAbstractMathematics PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 145

Foundations of Abstract Mathematics

Part1: Groups, Rings, Fields and Modules

Raoof Mirzaei

(Last revised: September 23, 2015)

Copyright
c 2015 by Raoof Mirzaei
All rights reserved.

i
ii

Dedicated to Neda
Contents

Preface v

0 What is Abstract Algebra 1

I Sets and First-Order Logic 5

1 Set Theory 7
1.1 Origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Basic Definitions and Axioms . . . . . . . . . . . . . . . . . . . . 12
1.4 Operations on Sets . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.5 Families of Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.6 Representation of sets and subsets . . . . . . . . . . . . . . . . . 25
1.7 Paradoxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.8 ZFC Axioms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2 Enough Number Theory 39


2.1 Divisibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.2 Congruence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.3 Modular Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . 44

3 Relations and Functions 49


3.1 Ordered pairs and Cartesian products . . . . . . . . . . . . . . . . 49
3.2 Relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Partitions of a Set . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.4 Equivalence relations, equivalence classes . . . . . . . . . . . . . 55
3.5 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.6 Set of all functions . . . . . . . . . . . . . . . . . . . . . . . . . 68

iii
Contents iv

II Group Theory 73
4 Algebraic Structure * 77

5 Introduction to groups 81
5.1 Binery operation . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.2 Cayley Table . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.3 Definition and Examples of Groups . . . . . . . . . . . . . . . . . 94
5.4 Groups of modular arithmetic . . . . . . . . . . . . . . . . . . . 95
5.5 Groups of permutation . . . . . . . . . . . . . . . . . . . . . . . 97
5.6 Basic Properties Of Groups . . . . . . . . . . . . . . . . . . . . . 97

6 Subgroup And Cyclic group 103


6.1 Subgroup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.2 Cyclic Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.3 Isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

7 Matrix Group And The Group Of Circle 113

8 Symmetric Group 117


8.1 Cycle Decomposition . . . . . . . . . . . . . . . . . . . . . . . . 118
8.2 Composition of two permutations . . . . . . . . . . . . . . . . . . 119
8.3 The Alternating group . . . . . . . . . . . . . . . . . . . . . . . . 126
8.4 Dihedral Groups and Symmetries Of Objects . . . . . . . . . . . . 129
8.5 Cayley Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . 134

9 Cosets and Lagrange’s Theorem 135


9.1 Coset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Preface

This book is intended for mathematicians who are not satisfied with ambiguity
of definitions and consideration in Abstract Mathematics, in particular Abstract
Algebra. As it is know Logic and set theory is the backbone of pure mathematics
and every branch of study is build upon it, So The Part I, incuding chapter 1 to 4 is
the foundation of set theory from the begining, without almost any harm!
As we proceed in following chapters I try to take back every modern consept
and definition into our base language we defined in set theory.
In Part II We strat to giving an structure to a set which we will call Groups, with
the consept of an n-ary operation which is a generalisation of a function that we
defined before and some other axioms.
In Part III We will continue with structures, with two n-ary operation, and
introduce Fields, Ring and their Modules.
In Part IV We will discuss the famous "told story" of Galois theory in complete
detail.
In Part V We will talk enough about what we have established Using Category
theory.
In Part VI We will eventually get to the main goal of this book: thinking over
new stories about Galois theory.

v
Chapter -1. Preface vi
Chapter 0

What is Abstract Algebra

It was a shocking discovery of course that Newton’s laws are


“ wrong ... we now have a much more humble point of view of our
physical laws : EVERYTHING CAN BE WRONG.

Richard Feynman,

We all are familiar with the notion of algebra from high school. The word
abstract might seems weird at the first look. Iranian Mathematiton Mohammad
Khawarizmi first used the word al-jabr meaning balencing or reduction. Although
the modern algebra is completly different from the ancient algebra,but it has
something to do with it. If you ask someone randomly about algebra you will
probably get : An unknown stupid game with the unknown variable x. Then,
what if you ask about abstract algebra ? Probably if you ask this question they
dont notice the change. If someone asks me what is abstract algebra ? I simply
answer : It is an abstract approach to algebra. What does this mean? Is it just a
nice fancy statement ? I claim no. Abstract Algebra is the formal structure of our
mathematical intuition. At the beggining we simply used mathematics to satisfy
our intentions. At first we would like to have measurement , well that is not abstract
at all, on the contrary it is very intutive. That is why we first start to introduce
numbers, to count our sheep for example. I call this event The birth of Mathematics.
That is what we call today a natural number. The set of natural numbers is called
N. Then we generalise this idea to get an integer number system, the one including

1
Chapter 0. What is Abstract Algebra 2

the natural numbers, and their negatives, together with zero. We do not stop, we
keep on going to generalise this to the rational number system, the set consisting
of all fractions except when the denominator is zero. Then we construct other
number systems for our perposes, the real numbers, consisting of all points in a
continuous geometric line. We did it because the rational numbers were not enough.
I used the word enough. Enough for what? Enough for who? In Mathematics
everything must be defined precisely. If someone uses the number of his sheep,
the integers are more than enough. the number of sheeps can not be 2.576 . If he
owes someone we can say he has −4 sheep for example. In Mathematics we would
like to solve equations and real numbers are not enough to contain the solution of
every algebraic equation. So again we extend real numbers to get a another space
for which every algebraic equation has a solution in that number system which is
called the complex numbers. The first number system which is in 2 Dimensions.
Before this we worked with 1 Dimensional world. Now, is this enough ? Enough
for what? Enough for who? If we want to solve an algebraic polynomial equation
it turns out that the fundamental theorem of algebra guarantees every such equation
has all of its solutions in complex numbers.

But we can go on and generalise again to get number systems in higher dimen-
sions. One such generalisation is Cayley Dickson Construction which from we get
hypercomplex systems in in 4 (Quaternions); 8 (Octonions); 16 (sedenions); and in
general 2n dimensions ( an algebra is merely a vector space, A, equipped with a
bilinear map which prescribes the multiplication of vectors. The Cayley-Dickson
Process affords a means of building a larger algebra, B, which contains A as a
subalgebra)
N⊂Z⊂Q⊂R⊂C⊂H⊂O⊂S

N = {1, 2, 3, . . . }
Z = {. . . , −3, −2, −1, 0, 1, 2, 3, . . . }
n
Q = { : n, m ∈ Z and m 6= 0}
m

Every highschool student obviusly knows on this number system we have


a + b = b + a ∀a, b ∈ R
3

(a × b) × c = a × (b × c)
They think that is always so , but it is not.
For quaternions, a number system in 4 dimensions, the multiplication commuta-
tivity fails i.e. a × b 6= b × a
But associativity still holds i.e.
(a × b) × c = a × (b × c)
But interestingly for octionions associativity fails too. (a × b) × c 6= a × (b × c)
Abstract algebra is about generalizing these number systems and studying the
properties of these number systems and classifiying them up to their structure. We
do so by considering sets toghter with operation(s) defined on that set and cheking
axioms they satisfy. For example
(Z, +) ⇒ Group
(Z, +, ×) ⇒ Ring
(Q, +, ×) ⇒ Field
(Rn , +) ⇒ Vector spaces over R
Chapter 0. What is Abstract Algebra 4
Part I

Sets and First-Order Logic

5
Chapter 1

Set Theory

Zeno was concerned with three problems . . . These are the


“ problem of the infinitesimal, the infinite, and continuity . . .

Bertrand Russell,

Set theory is the foundation of mathematics. Almost all branches of study are
built up from the concepts somehow derived from it. So as a backbone of all
mathematics it needs to be well established. The beginning of mathematics is
concerned with counting and measurement, and these are described in terms of
numbers and arithmetic. Even arithmetic can be derived from sets, Namely those
finite sets whose elements are also finite sets, the elements of which are also finite,
and so on, which is called pure sets is formally equivalent to arithmetic. Since set
theory plays the role of forerunner to arithmetic, it seems to be a good candidate
for the foundation of almost all mathematics.
In fact, set theory can be studied as:
• Set Theory as a foundation for mathematics
• Set Theory as a subject
Set Theory as Foundation
At the end of the 19th century, mathematicians became concerned that the did
not fully understand the nature of the basic mathematical notions. For example:
What is a number? We understand the set N of natural numbers intuitively,
but what is the real foundation for that? Even if we accept naturals and

7
Chapter 1. Set Theory 8

rationals, we have so much difficulties with real numbers ? What is the


meaning of a function f : R2 → R2 ?
The reason that mathematicians become so concern was because of rising some
paradoxes in set theory.
So the most logical way seems to be : seeking for a strong foundations. What
better than set theory, since it starts with a limited number of axioms from logic
that is based on humans pure mind atleast apparently.
Once we stablished a good foundations, it is turn for the other subjects, such as:
Number theory, Analysis, Probability , etc to build on it.
Or—we can do something simpler: show that all of mathematics can be reduced
to (encoded in) set theory. We must show that the notions of real number, function,
and so on can be defined inside set theory in such a way that these notions have the
correct properties. For the reals, this includes:
• Laws of arithmetic in ;
• Completeness of the ordering (least upper bound principle)
Calculus can be developed rigorously in terms of the theory of limits, if the
reals have these fundamental properties.
Set Theory as a subject
This involves understanding the sizes of infinite sets in two ways:
• Comparison of size: cardinality
• Counting: ordinals
It leads to the study of well ordering and the axiom of choice and eventually to
a picture of the universe of all sets called the Cumulative Hierarchy.

1.1 Origin

Set theory had some revolutions from the word go. Set theory, began with the
work of Georg Cantor1 and Richard Dedekind in the 1870’s.it came to born by a
single paper in 1874 by Georg Cantor: "On a Characteristic Property of All Real
Algebraic Numbers".[1]
1
Georg Cantor is a great German mathematician who opend the door of thinking about different
kinds of infinity..
9 1.2. Classification

Cantor himself gave in a publication in 1895 a description of the term set.

A set is a gathering together into a whole of definite, distinct objects of our


perception or of our thought. These objects are called elements of the set. The
question about what actually the word set means is a fundamental question and
fundamental problem in the subject. One of the difficulties is that the word set is
common word in any informal language, and some of the definitions of set are as
following:

1. A number of things of the same kind that belong or are used together (
Webster)

2. A group of things that belong together ( Cambridge)

3. A group or collection of things that belong together or resemble one another


or are usually found together ( Oxford)

In these definition the notion of a set is very general, because there is no


restriction on the nature of the things which may be elements of a set. If we
want to take this word and apply it to mathematics, we need to be much more
precise to define what it means in a mathematical sense. In the naive set theory we
consider every well-defined collection of objects to be a set. The problem with this
definition is that the terms ‘collection’ , ‘object’ and ‘well-defined’ are themselves
undefined. there is a problem of circularity for defining the primitive concepts like
set,point,line,etc. for example you can start by: a set is a collection, then it comes
to ask what is a collection, then you say a collection is a family, a family is a group
, a group is a class and eventually we get back to where we started. So we consider
set as a primitive concept and defining everything else in terms of sets. I don’t
know whether it is possible or not, but most of the community of mathematicians
agree with that.

1.2 Classification

Set theory is classified in several different ways, but in general there are two
approaches to set theory:

1. Naive Set Theory

2. Axiomatic Set Theory


Chapter 1. Set Theory 10

Naive Set Theory

Set theory was introduced to clarify paradoxes of infinite size.


For example, The paradox of the Even Numbers (Bolzano Paradox) : There are
as many even natural numbers as natural numbers?
• The answer is yes, because for every natural number n there is a even natural
number 2n, and Vice versa.
• The answer is no, because the evens are a proper subset of the naturals: every
even number is a natural number but there are natural numbers, which are
even.
To solve this paradox and in general whenever it comes to infinity we need to
be precise about what is meant by size.
Cantor’s notion of sets makes this clear. The first development of set theory is
called naive set theory . It was created at the end of the 19th century by Georg
Cantor as part of his study of infinite sets.
He established the notion of a one-to-one correspondence and proved a theorem
now called the Cantor-Schroeder-Bernstein Theorem which is used for proving
equivalence of two sets. the theorem states that if there exists a one-to-one function
from A 7−→ B (not necessarily onto) and a one-to-one function from B 7−→ A
(not necessarily onto), then there exists a bijection from one set to the other; i.e. the
two sets are equivalent. Using this theorem paradoxes like the former are solved.
Cantor showed there was no one-to-one correspondence from N to R , and hence
the natural numbers have a smaller cardinality than the real numbers.
Cantor thought if we can add numbers with each other then we can be able
to add infinity and infinity. He realized that it was actually possible to add and
subtract infinities, and that beyond what was normally thought of as infinity existed
another, larger infinity, and then other infinities beyond that.
Cantor showed that there is more than one kind of infinity. Before him it
was assumed that all kinds of infinity having the same number of elements as it
seems even today with our intuition. Cantor proved using his well known diagonal
argument, that the collection of real numbers has more elements than integers, but
surprisingly the integers and rationals have the same number of elements. In other
words,Unlike integers and rational numbers, the real numbers are not countable.
So, however infinite the infinity is, there are infinite kinds of infinity, so he divided
numbers into two classes, countable and uncountable. A set that has a larger
cardinality than the natural numbers is called uncountable.
11 1.2. Classification

In set theory all existing objects are sets. If an object exists it is a set otherwise
it does not exist. To remind us of the fact that sets include elements we sometimes
refer to sets as a collection of sets, or as a families of sets. This is just a “human
factors” trick since the theory makes no distinction between sets, families, collec-
tions or elements. In axiomatic set theory elements, collections, and families are
just sets.
So Cantor’s set theory solved these kinds of paradoxes,but nothing is over in
the world of mathematics. Yes, it solved some paradoxes, but created others So
after Cantor another mathematician came along and tried to rebuild set theory in a
way that got free of these new paradoxes.
In naive set theory (NST) we define concepts informally based on natural
language.
The words and, or, if ... then, not, for some, for every are not here subject to
rigorous definition, it is an approach to set theory which assumes the existence of a
universal set, despite the fact that such an assumption leads to paradoxes. Frege
constructed a formal theory which is the axiomatization of naive set theory.

Axiomatic Set Theory

Axiomatic set theory was developed with the purpose of eliminating paradoxes
which occured, with the goal of determining precisely what operations were allowed
and when.
Structure of set theory and other fields constructed upon it has the three follow-
ing following elements:
1. Variables (e.g., a, b, ..., x, y, z) which stand for sets.
2. The predicate ∈ , which stands for element inclusion
3. Logical operators and symbols include
(a) ¬P , where ¬ is the logical “negation” operator.(not)
(b) ∧P , where ∧ is the logical “and” operator.(and)
(c) ∨P , where ∨ is the logical “or” operator.(or)
(d) P =⇒ P , where =⇒ is the logical “implication” operator.(implies)
(e) P ⇐⇒ P , where ⇐⇒ is the logical “bijection” operator.(if and only if)
(f) ∀ x P is the logical “for-all” quantifier.(for all)
Chapter 1. Set Theory 12

(g) ∃ x P is the logical “exists” quantifier.(there exists)


(h) ∃! x P is the logical “exists” quantifier.(there exists exactly one,unique )
In set theory all propositions are constructed from the non-logical predicate ∈
and connected to each other using these logical operators. propositions.
Tautology table of propositions is here:
P Q P ∨ Q P ∧ Q ∼ (P ∧ Q) (P ∨ Q)∧ ∼ (P ∧ Q)
T T T T F F
T F T F T T
F T T F T T
F F F F T F

P Q P =⇒ Q Q =⇒ P P ⇐⇒ Q ∼ (P ⇐⇒ Q)
T T T T T F
T F F T F T
F T T F F T
F F T T T F

In Zermelo-Fraenkel set theory this notion of equality is taken as an axiom. In


fact, all of the definition and axiom in naive set theory with the pioneer work of
Cantor is axiomatized as 3 axiom in ZFC (Zermelo-Fraenkel with the axiom of
choice)
has two axioms: The Axiom of Extensionality.
First, I use the approach of naive set theory, That is defining concepts intuitively
but for avoiding duplication I try to make it similar to axiomatic approach as
much as possible. I call this axiomatic approach to naive set theory or Frege
axomatization of NFS.
Then we see some paradoxes will occur, and naive set theory falls. Then we
proceed in a way to rescue set theory from these paradoxes. After that I briefly talk
about ZFC.

1.3 Basic Definitions and Axioms

A set is an unordered collection of distinct elements or objects .2 These elements


can be anything like numbers, letters, logical proposition, points in arbitrary
2
In this definition we assume these elements or objects are distinct.
13 1.3. Basic Definitions and Axioms

space ,etc. Sets even may include sets as their elements. You are familiar with
sets of numbers, that is a set consisting of numbers. We have several kind of
number system. We can generalize this notion of a set to consists things other than
numbers. These kind of generalization is called abstraction. Number themselves
are abstraction of things. We could simply count our sheep without having a formal
theory of natural numbers.
If A is a set, then the objects in the collection A are called either the members
of A or the elements of A.
When we talk about a set as a collection of objects it is natural to ask "what
does it consist of ?".
This motivates the notion of membership. Membership is a binary relation 3 ∈
between an object o and a set A.
We write o ∈ A to indicate that the object o is an element, or a member, of the
set A . We also say that o belongs to A , that o is contained in A , or simply that o
is in A .The negation of o ∈ A is ¬( o ∈ A) which is denoted by o ∈ / A.
We use capital letters to display a set and small letters to denote its element
and put the elements of a set between brackets , For example S = {a, b, c} and
T = {{a, b, c}, {a, c}, b, c} .
A set can consists of only one element, so given an object x we can form the set
that has x as its only element which is denoted by {x} .
Likewise, for any two objects x and y, we can form the pair set {x, y} consisting
of just the elements x and y.
More generally, for any objects x, y, z, . . . we can form the set having x, y, z, . . .
as its elements, which we denote by {x, y, z, . . . }

A set containing only one element is called a singleton. This set {x} is not the
same as x, because, x is an element while the singleton {x} is like a basket that
has x as its element. That is, x ∈ {x}, but not that x ∈ x .
Even worse, a set that has no element at all is called an empty set or null set
denoted by { } or ∅ . The existence of empty set is stated as an axiom in Axiomatic
set theory called axiom of empty set4 and the uniqness of empty set is obtained
from the axiom of extensionality which we will discuss below.
3
Consider it as just a relation for now, we will define binary relation in terms of sets later.
4
This axiom is stronger condition from the axiom of specification.
Chapter 1. Set Theory 14

Definition 1.3.1 (Cardinality) The number of elements in a set A is called its


cardinality or order, and denoted by | A | or n(A).
For example, if T = {{a, b, c}, {a, c}, b, c} , then | T |= 4
A set which has only a finite number of elements is called a finite set .i.e. A set
is said to be a finite set if all of its elements can be listed, this include the condition
that set has no elements.
otherwise A is an infinite set.
Definition 1.3.2 (Equivalent Sets)
Two sets A and B are said to be equivalent if their cardinal number is same, i.e.,
n(A) = n(B). The symbol for denoting an equivalent set is ‘’.
For example:
A = {a, b, c} ; n(A) = 3
B = {1, 2, 3} ; n(B) = 3
Therefore, A ←→ B
Since sets are objects themselves, the membership relation can be defined on
two set. A binary relation between two sets is the subset relation ⊆ .
Definition 1.3.3 (Subset) Suppose A and B are sets. the expression A ⊆ B , i.e.
B contains all the elements of A is defined in terms of membership by

∀x(x ∈ A ⇒ x ∈ B) (1.1)

Then A is said to be a subset of B .


Furthermore, A ⊂ B if

∀x(x ∈ A ⇒ x ∈ B) ∧ ∃x ∈ Bs.t. x ∈
/A
(1.2)

, i.e. there is an element of B which is not in A . In this case we call A a


proper subset of B (or that A is strictly contained in B) . In other word, a subset of
A is proper if it is neither A itself nor ∅.
The set B is also called a superset of A .
15 1.3. Basic Definitions and Axioms

Remark According to this definition, for every set A , A ⊆ A , ∅ ⊆ A


Clearly, two sets A and B are equal if they contain the same elements , i.e.

A = B if and only if ∀x(x ∈ A ⇒ x ∈ B) ∧ ∀x(x ∈ B ⇒ x ∈ A)

(1.3)

where ∧ denotes logical AND.


Equivalently

A = B if and only if (A ⊆ B ∧ B ⊆ A)
(1.4)

This is called the axiom of extentionality.


Axiom of Extensionality The axiom of extensionality says that a set is
defned by its members, so sets formed by the same elements are equal,

∀A∀B[∀x(x ∈ A ⇐⇒ x ∈ B) ⇒ A = B]. (1.5)

This implies, for example, that empty set (if it exists) is unique. Since ∅ has no
elements at all and it is the only set with no elements, so by extensionality any two
such sets must be equal.
This axiom and definition has the direct consequence that within sets, there is
no order. For example if A = {a, b} and B = {b, a} then A = B
What about repetition? Does it matter?
Are A = {a, b} , B = {a, a, b, b, b, c} and C = {a, b, c, b, b} equal?
According to equality definition

∀x ∈ A ⇒ x ∈ B, C ∧ ∀x ∈ B, C ⇒ x ∈ A.

But according to cardinality definition of a set the number of elements in A, B


and C are 3 , 6 and 5 respectivly.
Counting the number of elements in a set is like you have a basket and you close
your eyes and take element one by one so you wont notice if you pick an element
Chapter 1. Set Theory 16

more than once. Against all of the elementry set theory books, unlike ordering
repetiton in a set matters. So non of the A, B and C are equal.
Note that, we dont consider collections that have multiple copies of the same
element in the set. In other word a set is a collection that has distinct elements. So
by this contract B and C are not sets, they are called multisets.
Multisets are extension of the idea of a set that unlike a set can have multiples
of the elements.
The number of repetition of an element in amultiset is called multiplicity. The
elements of multisets usually represented in square brackets [ ]. for example
C = [a, b, c, b, b] is not a set but a multiset.
The idea of multisets arise from the work of Dedekind, and is a quite rich
subject. We dont want to get into this but here I give a formal definition of multiset
and its relation with our familiar sets.
Multiset is a pair (A, m) where A is a set and m : A → N is a function from A
to the set N = {1, 2, 3, . . . }
For each a ∈ A the multiplicity of a is the number m(a).
If
∀a m(a) = 1
then we have a simple set. In other word a set is a mutiset (A, 1) , i.e. each
element in the multiset occurs just once.

In A the multiplicity of elements are : a = 1 , b = 1 and c = 1


In B the multiplicity of elements are : a = 2 , b = 3 and c = 1
In C the multiplicity of elements are : a = 1 , b = 3 and c = 1

1.4 Operations on Sets

There are various natural operations we can perform on sets. (They are analogue
of addition, multiplication, and subtraction for numbers.) There are a number of
simple operations that can be performed on sets, forming new sets from given
sets.(They are analogue of addition, multiplication, and subtraction for numbers.)
Definition 1.4.1 (Union and Intersection) Given any two sets A and B;
17 1.4. Operations on Sets

(i) The Union A ∪ B of A and B is the set which consists of all the elements
of A and all the elements of B such that no element is repeated. The union of two
given sets is the smallest set which contains all the elements of both the sets.

A ∪ B = {x|(x ∈ A) ∨ (x ∈ B)}

(ii) The Intersection A ∩ B of A and B is the set which consists of all elements
that belong to both A and B. That is, the largest set which contains all the elements
that are common to both the sets And has the formal definition

A ∩ B = {x|(x ∈ A) ∧ (x ∈ B)}

Furthermore, the operations are commutative. That is,

A ∪ B = B ∪ A, A ∩ B = B ∩ A

And associative. That is,

A ∪ (B ∪ C) = (A ∪ B) ∪ C, A ∩ (B ∩ C) = (A ∩ B) ∩ C

This is an analogy between sets and numbers5 , in numbers addition + and


multiplication × is commutative.
We have commutative laws x + y = y + x and x × y = y × x, and also
associative laws x + (y + z) = (x + y) + z and x × (y × z) = (x × y) × z.
The set theoretic distributive laws are

A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C), A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C)

Which are called Demorgan law and corresponds to number distributive law. x ×
(y + z) = x × y + x × z.
But this analogy is not always true. For example, A ∩ A = A is an identity for
sets but in algebra a × a = a is not an identity.
Example 1.4.2 (i) Let A = {−3, 7, 0, 8, 10} and B = {1, 8, 3, 7, 0, 6} .
Then A ∪ B = {−3, 7, 7, 0, 8, 10, 1, 3, 6} and A ∩ B = {0, 8}.
5
Numbers that are subsets of complex number, because for example Quaternion are not commu-
tative. Octonion are not even associative.
Chapter 1. Set Theory 18

(ii) Let X = {a, b, c} and Y = {∅}.


Then A ∪ B = {a, b, c, {∅}} and A ∩ B = ∅.
Two set A and B are called disjoint or not connected if and only if A ∩ B =
∅, i.e. they have no elements in common. For example, the set of even numbers
and the set of odd numbers are disjoint.
Definition 1.4.3 (Universal Set)
A Universal Set is a fixed set under consideration and consisting of all objects
considered in the theory.
It is not necessary the set of everything in the world. for example, when we talk
about ordinals, natural numbers can be considered as a universal set. It is usualy
denoted by U .
The complement of the universal set is the empty set
Definition 1.4.4 (Complement)
Suppose U is the universal set and A is a subset, then the complement of A
denoted by A or Ac is the set consisting of all the elements of U which are not the
elements of A.
A = {x ∈ U : x ∈
/ A}
Definition 1.4.5 (Relative Complement)
If A and B are sets, then the set theoretic difference of A and B, also known as
relative complement of B relative to A is the set of all those elements of A which
do not belong to B, denoted by A \ B or A − B

A \ B = A − B = {x|x ∈ A ∧ x ∈
/ B}

This is analoge of subtraction in numbers but not exactly, so we call it set


theoretic difference.
Definition 1.4.6 (Symmetric differences)
The symmetric difference of two sets A and B is the set of all objects that are
in one and only one of the sets. The symmetric difference is written A∆B .

A∆B = {(A \ B) ∪ (B \ A)

1. Commutative Laws
(a)A ∪ B = B ∪ A
19 1.4. Operations on Sets

(b)A ∪ B = B ∪ A

2. Associative Laws

(a)A ∪ (B ∪ C) = (A ∪ B) ∪ C

(b)A ∩ (B ∩ C) = (A ∩ B) ∩ C

3. Distributive Laws

(a)A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C)

(b)A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C)

4. Identity Laws
(a)A ∪ ∅ = A
(b)A ∩ U = A

5. Inverse Laws
(a)A ∪ A = U
(b)A ∩ A = ∅

6. Double Negation Law


A=A

7. Idempotent Laws
(a)A ∪ A = A
(b)A ∩ A = A

8. DeMorgan’s Laws
(a)A ∪ B = A ∩ B
(b)A ∩ B = A ∪ B

9. Domination Laws
(a)A ∪ U = U
(b)A ∩ ∅ = ∅

10. Absorption Laws


(a)A ∪ (A ∩ B) = A
(b)A ∩ (A ∪ B) = A
Chapter 1. Set Theory 20

11. Complements of U and ∅

(a)U = ∅
(b)∅ = U

12. Consistency Principle

(a)A ⊆ B ⇐⇒ A ∪ B = B

(b)A ⊆ B ⇐⇒ A ∩ B = A
Proof We shall proof (2.a), (3.a), (5.a), (5.b), 6 and (8.a). The others left to readers.

2.(a) A ∪ (B ∪ C) = (A ∪ B) ∪ C

x ∈ A ∪ (B ∪ C) ⇐⇒ x ∈ A ∨ (x ∈ B ∨ x ∈ C)
⇐⇒ (x ∈ A ∨ x ∈ B) ∨ x ∈ C
⇐⇒ x ∈ (A ∪ B) ∪ C

Therefore, x ∈ A ∪ (B ∪ C) if and only if x ∈ (A ∪ B) ∪ C.

3.(a) We shall prove two cases :

I. A ∩ (B ∪ C) ⊂ (A ∩ B) ∪ (A ∩ C)

II. (A ∩ B) ∪ (A ∩ C) ⊂ A ∩ (B ∪ C)

I. A ∩ (B ∪ C) ⊂ (A ∩ B) ∪ (A ∩ C)
Let x ∈ A ∩ (B ∪ C) =⇒ x ∈ A ∧ x ∈ (B ∪ C)
=⇒ x ∈ A ∧ {x ∈ B ∨ x ∈ C}
=⇒ {x ∈ A ∧ x ∈ B} ∨ {x ∈ A ∧ x ∈ C}
=⇒ x ∈ (A ∩ B) ∨ x ∈ (A ∩ C)
=⇒ x ∈ (A ∩ B) ∪ (A ∩ C)
∴ x ∈ A ∩ (B ∪ C) =⇒ x ∈ (A ∩ B) ∪ (A ∩ C)
∴ A ∩ (B ∪ C) ⊂ (A ∩ B) ∪ (A ∩ C)
II. (A ∩ B) ∪ (A ∩ C) ⊂ A ∩ (B ∪ C)
21 1.4. Operations on Sets

Let x ∈ (A ∩ B) ∪ (A ∩ C) =⇒ x ∈ (A ∩ B) ∨ x ∈ (A ∩ C)
=⇒ {x ∈ A ∧ x ∈ B} ∨ {x ∈ A ∧ x ∈ C}
=⇒ x ∈ A ∧ {x ∈ B ∨ x ∈ C}
=⇒ x ∈ A ∧ {B ∪ C}
=⇒ x ∈ A ∩ (B ∪ C)
∴ x ∈ (A ∩ B) ∪ (A ∩ C) =⇒ x ∈ A ∩ (B ∪ C)
∴ (A ∩ B) ∪ (A ∩ C) ⊂ A ∩ (B ∪ C)
∴ A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C)

5.(a)

x ∈ A ∪ A ⇐⇒ x∈A∨x∈A
⇐⇒ x∈A∨x∈U \A
⇐⇒ x ∈ A ∨ (x ∈ U ∧ x ∈
/ A)
⇐⇒ x∈U
We have shown, A ∪A = U

5.(b) A ∩ A = ∅

x ∈ A ∩ A ⇐⇒ x ∈ A ∧ x ∈ A
⇐⇒ x ∈ A ∧ x ∈ U \ A
⇐⇒ x ∈ A ∧ (x ∈ U ∧ x ∈
/ A)

that is a contradiction. So, there is no x , such that x ∈ A ∩ A , hence A ∩ A = ∅

6. A = A

x ∈ A ⇐⇒ x ∈ U \ A
⇐⇒ x ∈ U ∧ x ∈
/A
⇐⇒ x ∈ U ∧ x ∈ A
⇐⇒ x ∈ A
Chapter 1. Set Theory 22

8.(a) A ∪ B = A ∩ B

x ∈ A ∪ B ⇐⇒ x ∈ U \ A ∪ B
⇐⇒ x ∈ U ∧ x ∈
/ A∪B
⇐⇒ x ∈ U ∧ x ∈
/ A∧x∈/B
⇐⇒ x ∈ U ∧ x ∈ A ∧ x ∈ B
⇐⇒ x ∈ U ∧ x ∈ A ∩ B
⇐⇒ x ∈ A ∩ B

1.5 Families of Sets

Sets can be elements of other sets. Instead of calling them set of sets we refer to
them as collection or families of sets .
A family of sets is not a set necessarily, because we allow repeated elements, so
a family is a multiset.
Definition 1.5.1 (Power set) Let A be a set. The set consisting of all subsets of A
is called the power set of A and denoted by P(A).
P (A) = {X : X ⊆ A}

Remark 1. ∅ ⊆ A for any set A, So P (A) 6= ∅


In particular, P (∅) = ∅
2. According to definition, A ⊆ A, therefore A ∈ P (A) but an element a ∈ A
could not be an element of P (A) ,i.e. a ∈
/ P (A)
Example 1.5.2 If A = {a, b, 1}, then

P (A) = {∅, {a}, {b}, {1}, {a, b} {a, 1}, {b, 1}, {a, b, 1}}.

Note that:
23 1.5. Families of Sets

a ∈ A; {a} ⊆ A; {a} ∈ P (A) ; a∈


/ P (A)

∅ ⊆ A; ∅∈
/ A; ∅ ∈ P (A) ; ∅ ⊆ P (A)

Arbitrary Unions and Intersections


The union operation ∪ allows us to form the union A ∪ B of two sets. As a
result of associativity laws we can form the union and intersection of finitely many
arbitrarily collections of sets.

A1 ∪ A2 ∪ A3 ∪ · · · ∪ An−1 ∪ An = A1 ∪ (A2 ∪ A3 ) ∪ . . . (An−2 ∪ An−1 ) ∪ An

But what about infinite collections of sets ? The above approach does’t work
for the infinite case.
Definition 1.5.3 Suppose I is a set, called the index set, and with each i ∈ I we
associate a set Ai . We call {Ai : i ∈ I} an indexed family of sets. Sometimes this
is denoted by {Ai }i∈I .
The size of the collection of sets being "unioned over" is whatever the size
of I is. This notion makes perfect sense even if I is an infinite set (countable or
uncountable).
For example, if I = N, the set of natural numbers, we could write {Ai }i∈N ,
meaning that we have a countable number of sets which are being considered.
(Note, in general, it is not necessary that be even countable. The set of all real
numbers denoted by R is an example of an uncountable set as compared to N,
which is a countable set.)
Example 1.5.4 I = {A1 , A2 , A3 } with A1 = {a, b, 2}, A2 = {a, b}, A3 = {a, d}
and is a family of sets.
Given an indexed family {Ai }i∈I we can define the intersection and union of
the sets Ai
Definition 1.5.5 (Extended Union and Intersection) Let I be a family of sets. Then
we define:
Chapter 1. Set Theory 24

The union over I by:


[
Ai = {x | ∃i ∈ I : x ∈ Ai }
i∈I

and the intersection over I by:


\
Ai = {x | ∀i ∈ I : x ∈ Ai }
i∈I

S
That is, i∈I Ai is the set of
T all those elements which belongs to one or more of
the sets Ai in the family, and i∈I Ai is the set of those elements which belongs to
every Ai .
Note that if I has two elements, like I = {1, 2}, then we have only two sets A1
and A2 .
So the union of the collection {Ai : i ∈ I} = {A1 , A2 } is just
[
Ai = A1 ∪ A2
i∈I

And the intersection \


Ai = A1 ∩ A2
i∈I

So the definition of arbitrary union and intersection over indexed families


reduses to the former notions of union and intersection of pairs of sets.
If I = N, is the set of natural numbers we can form the union of the collection
over an infinite countable sequence of sets A1 , A2 , A3 , . . . then {Ai : i ∈ I} is
A1 ∪ A2 ∪ A3 ∪ . . . .
We can go further and take the union of the collection over an infinite uncount-
able sequence. Let I = R the set of real numbers. 6
Example 1.5.6 let I = {1, 2, 3, ..., n}.

[
Ai = N
i∈I


\
Ai = {1}
i=1
6
(Although the sets I used in indexing collections of sets are often sets of numbers, they don’t
have to be; this notion makes perfect sense for any set I whatsoever.)
25 1.6. Representation of sets and subsets

8
\
Ai = {1, 2, 3, 4, 5}
i=5

Example 1.5.7 Let I = {A1 , A2 , A3 } where A1 = {a, b, 2}, A2 = {a, b}, A3 =


{a, d} :
[ \
Ai = {a, b, 2, d} Ai = {a}
i∈I i∈I

Note that
S 
I⊆P i∈I Ai = P ({a, b, 2, d})

Example 1.5.8 Let R denote the set of real numbers, N denote the set of natural
numbers and S denote the set of all finite subsets of R.
For each n ∈ N consider the subset An = {n} of R.
For all n ∈ N, An ∈ T .
But [ [
{An : n ∈ N } = {n : n ∈ N } = N ∈
/T
Because it is neither finite, nor equal to the entire set of real numbers.
A family with index set N is called a sequence.

1.6 Representation of sets and subsets

There are some way of describing a set. The elements of a finite sets can be
described by listing them all, but for the sets with large number of elements or
infinite sets the best way to describing them is giving the property which defines
the set.
Intuitively, a set is a collection of all elements that satisfy a certain given
property, so If P (x) is some property we write {x : P (x)}, or x ∈ A to mean the
set of all those x that satisfy P (x), so that for all x,

x ∈ A ⇐⇒ P (x)

Formally,
{x ∈ A |P (x)} := {x|x ∈ A ∧ P (x)},
Chapter 1. Set Theory 26

For example,
A = {x|x is a prime number}
, the closed form of {2, 3, 5,7, 11, . . . }, is the set of all prime number which is an
infinite set.
The set of natural numbers can be described :

N = {x ∈ Z |x > 0}

i.e. the set of all x belonging to integer numbers with this restriction that x is
positive integer.
The empty set
∅ = {x ∈ R |x2 < 0}

Suppose A and B are sets and A ⊆ B . Another way of Specifying a subset of


a set B is of the form:
{F (x) : x ∈ A}
where F (x) is some formula depends on x. It consists of all objects obtained by
putting members of the set A into the formula F. For example,

{x2 : x ∈ N} = {1, 4, 9, 16, 25, . . . } = B

and if A = {−3, 0, 1, 2, 5}, then

C = {F (x) : x ∈ A} = {x2 : x ∈ A} = {9, 0, 1, 4, 25} ⊆ B

In formal,
{F (x) : x ∈ A}
can be defined as
{y ∈ B : ∃x ∈ A s.t. y = F (x)}

The empty set can be specified in many ways, e.g.,

∅ = {x ∈ R |x2 = −1}
∅ = {x |x 6= x}

This is an axiom called The Axiom of Unrestricted Comprehension.


27 1.6. Representation of sets and subsets

1. Axiom Schema of Unrestricted Comprehension This says that, for any


property, there exists a set of all and only those things that have that property,
Y = {x : P (x)}. More precisely, we restrict to properties that can be
defined by formulae in the language of set theory with parameters:

∃A ∀x (x ∈ A ⇐⇒ P (x)) (1.6)

This axiom guarantee for any proposition there exists a set with some elements
corresponding to that property.
Now consider following sets:

A = {x |x is a number}

A = {0, 9, 1, −15, . . . } = the set of all real numbers R

B = {x |x is not a number}

B = {{0}, {9}, Grothendick, Stars, Sky, logical proposition, lines, . . . B }


= The set of every thing except numbers
Notice that A ∈
/ A and B ∈ B.
A∈
/ A , because A is not a number.
B ∈ B , because the set B is not a number so it contains even itself!
So there are sets that Do belong to themselves and there are sets that Do not
belong to themselves
Now consider the collection of all sets U = {all sets}.
Since U is a set it contains itself as a set, i.e. U ∈ U .
Lemma 1.6.1 For every A, there is a unique set B such that x ∈ B if and only if
x ∈ A and P (x).
Proof Suppose B 0 is another set such that x ∈ B 0 if and only if x ∈ A and P (x).
If x ∈ B implies x ∈ A and P (x), then x ∈ B 0 .
If x ∈ B 0 implies x ∈ A and P (x), then x ∈ B.
Thus we have x ∈ B and only if x ∈ B 0 .
Therefore B = B 0 .
Chapter 1. Set Theory 28

Although this axiom seems quite obvious but it leads to several paradoxes.
Before I get to these paradoxes I introduce some basic operation on sets.
29 1.7. Paradoxes

1.7 Paradoxes

The Axiom Schema of Unrestricted Comprehension allows us to generate collection


of sets.
We can form, "the collection of all sets x that satisfy a condition P (x). In set
builder notation, this is written Y = {x : P (x)}. The question is can we assume
this new collection of sets is a set itself? It turns out NO.
The set of all sets which have the property of:
• Not being members of themselves =⇒ Russell’s Paradox
• Being a set =⇒ Cantor’s Paradox
• Being an ordinal =⇒ Burali-Forti’s Paradox
Therefore, we don’t consider the set of all sets to be a set, but a proper class.

Russell’s Paradox
Bertrand Russell considered the following set:

R = {x : x ∈
/ x}

That is, the collection of all sets x that are not members of themselves.
So, the set x is a member of R if and only if x is not a member of x ,i.e.

∀x(x ∈ R ⇐⇒ x 6∈ x)

And asked a question: Is R a member of itself or not ?


• If R ∈ R, then according to the definition of R we have R ∈
/ R.
• If R ∈
/ R, then according to the definition of R we have R ∈
/ R.
If we accept either of them we get contradiction.

R ∈ R ⇐⇒ R 6∈ R

It is always both true and false or it is neither true nor false! must satisfy the
defining condition, so R ∈ / R, and that is also a contradiction.
Thus we must confess that {x : x ∈
/ x} is not a set.
Chapter 1. Set Theory 30

Cantor’s Paradox
Cantor prooved for every set A the power set of A; that is, the set of all subsets
of A denoted by P (A), has a larger cardinality than A itself.7
P (A) has |2A | elements. So |A| < |2A |
Now let A be the set of all sets,

|A| < |2A |

but 2A is a subset of A, because every set in 2A is in A. Therefore

|2A | ≤ |A|

A contradiction. Therefore the set of all sets doesn’t exist.


So we must give set theory an axiomatic foundation which does not lead to
contradictions.
Exercise (Curry’s Paradox) P be the statement “The earth is flat.” Let S be the set

{x|x ∈ x =⇒ P }

(a) Prove that S ∈ S =⇒ P .


(b) Prove that S ∈ S
(c) Prove P !

Naive Set Theory failed because it allowed us to form sets that comprehend
arbitrary properties. One way to solve paradoxes of this type is to abandon the
Axiom Schema of Unrestricted Comprehension and restrict it to, the Axiom Schema
of Restricted Comprehension (also called the Schema of Separation). In ZFC, we
are allowed to form subsets that comprehend any property definable by a formulae
with parameters.
Now this axiom says, for each such property of sets P (x) and given any set A
the set Y = {x ∈ A : P (x)} exists.
Let’s consider the Russele’s Paradox again:

R = {x : x ∈ A ∧ x ∈
/ x}
7
In the next section we will proof it.
31 1.7. Paradoxes

• If R ∈ R, then R ∈ A & R ∈
/ R (Contradiction).
• Therefore, R ∈
/ R, and either R ∈
/ A or R ∈ R.
We conclude that R ∈
/ R&R∈
/ A.
So R ∈ R is false, and R ∈
/ R is true.
In other word in :
R ∈ R ⇐⇒ R 6∈ R

we can only prove R ∈ R =⇒ R 6∈ R not R ∈


/ R =⇒ R ∈ R.
As we see Russele’s Paradox is not a Paradox anymore. Since the existence of
the set {x : x ∈
/ x} is not valid in ZFC becuase of Axiom Schema of Restricted
Comprehension.
This axiom only allows us to create such sets as {x : x ∈ A ∧ P (x)}. By
another axiom called the axiom of foundation this set is just the empty set.
As a matter of fact, it is the concept of the set of all sets that is paradoxical, not
the idea of comprehension itself.
Axiom Schema of Restricted Comprehension 8
If P is a property (with parameter p), then for any X and p there exists a set
Y = {x ∈ X : P (x, p)} that contains all those x ∈ X that have property P .
In other word, for a given set X and proposition P (x), there exists a subset A
of X such that:

∀X ∃A ∀x(x ∈ A ⇐⇒ x ∈ X ∧ P (x))

Restricted Comprehension allows us to define some other set-theoretical con-


structions.
• Intersection
If we apply the axiom to the set A and the property P (x) : x ∈ B we get the
definition for intersection of two sets.

Y = {x |x ∈ A ∧ x ∈ B}

Y =A∩B
8
Also called axiom schema of specification or separation
Chapter 1. Set Theory 32

• Relative complement
If we make the restriction that P (x) : x ∈
/ B , we get definition of relative
complement.

Y = {x |x ∈ A ∧ x ∈
/ B}

Y =A\B

• Empty set
Taking P (x) : x 6= x we get :

Y = {x |x ∈ A ∧ x 6= x}

Y =∅

Empty set can be derived from some other axioms and sometimes the existence
of empty set is taken as an axiom in ZFC.
Not to mention that the Cantor paradox fails if we do not assume the power set
axiom, namely it might be that not all sets have a power set. So if the universal set
does not have a power set, there is no problem in terms of cardinality.
This seems quite coherent so naive set theory takes it as an axiom, precisly:
As we mentioned before, for any two objects x and y, we can form the pair set
{x, y} consisting of just the elements x and y.
In ZFC we take the existence of this new set as axiomatic, called Pairing axiom.
Axiom of pairing The axiom of pairing says that if x and y exist (i.e., if they
are sets) there also exists a set whose only elements are x and y.

∀A ∀B ∃C [∀x(x ∈ C ⇔ x = A or x = B]. (1.7)

Again, by Extensionality, there is a unique such set. We will represent such


set as {x, y}. The set consists of the sets a is {a} and is called the singleton
whose only element is a . So again having {a} it follows that the set {{a}}
exists applying pairing axiom again the set {{{a}}} exists.
This axiom will allow us to set up the definition of union between two sets.
33 1.7. Paradoxes

Axiom of union For every set A, there exists a set, denoted by , whose
elements are all the elements of the elements of A.

∀A ∃B [∀x(x ∈ B ⇐⇒ ∃y(y ∈ A ∧ x ∈ y)] (1.8)


S
For example, if x = {{1, 2,Sa, b}, {2, 3, b, d}} then x = {1, 2, 3, a, b, d}
Using the axiom of equality x is unique. This axiom together with the pair
set axiom, allows us to take the union of finite sequence x1 , . . . , xn . From
the pairing axiom, we have:
{A, B}
From the union axiom, we have the union of {A, B} :
[
{A, B}

There are only two members in {A, B} . Therefore:


[
x ∈ {A, B} ⇐⇒ x ∈ A ∨ x ∈ B ⇐⇒ x ∈ (A ∪ B)

So
S as a special Case we define the union of the two sets as follows :
{A, B} = A ∪ B
Just as we defined the union of only one set we can also define the intersection
of one set. But we don’t need an axiom for that becuase it can be derived from the
axiom of separation and Axiom of pairing.
T
The axiom of separation guarantee if A exists then A also exists.
\
A = {x : ∀y((y ∈ A) =⇒ (x ∈ y))}

As a special Case we define the intersection of the two sets as follows :


T
{A, B} = A ∩ B Beacuse :

\
x∈ {A, B} ⇐⇒ x ∈ A ∧ x ∈ B ⇐⇒ x ∈ (A ∩ B)
T
For example, if x = {{1, 2, a, b}, {2, 3, b, d}} then x = {2}
Remark
S
∅=∅
T
∅=U
Beacuse \
∅ = {x : x ∈ ∅} = {x : x ∈ U}
Chapter 1. Set Theory 34

is a paradox since the set of all set dosen’t exist.


But this paradox will not occur anymore since ZFC axioms doesn’t allow us
to make that set. In ZFC that is :
n \ o
x∈X:x∈ ∅ =X

Which dosen’t come up with a paradox.


35 1.8. ZFC Axioms

1.8 ZFC Axioms


1. Axiom of extensionality:

∀x∀y[∀z(z ∈ x ⇔ z ∈ y) ⇒ ∀w(x ∈ w ⇔ y ∈ w)]

2. Axiom Schema of Restricted Comprehension (specification):

∀X ∃A ∀x(x ∈ A ⇐⇒ x ∈ X ∧ P (x))

3. Axiom of pairing:

∀A ∀B ∃C [∀x(x ∈ C ⇔ x = A or x = B]

4. Axiom of union:

∀A ∃B [∀x(x ∈ B ⇐⇒ ∃y(y ∈ A ∧ x ∈ y)]

5. Axiom of foundation:

∀x[∃a(a ∈ x) ⇒ ∃y(y ∈ x ∧ ¬∃z(z ∈ y ∧ z ∈ x))]

6. Axiom schema of replacement:

∀x ∈ X ∃!y P (x, y) ⇒ [ ∃Y ∀y (y ∈ Y ⇔ ∃x ∈ X (P (x, y))) ]

7. Axiom of power set :

∀X ∃Y ∀Z [ Z ∈ Y ⇔ ∀z(z ∈ Z ⇒ z ∈ X) ]

8. Axiom of infinity:

∃X [ ∅ ∈ X and ∀x(x ∈ X ⇒ x ∪ {x} ∈ X) ]

9. Axiom of choice:

∀X [ ∅ ∈
/ X and ∀Y, Z ∈ X(Y 6= Z ⇒ Y ∩Z = ∅) ⇒ ∃Y ∀Z ∈ X ∃!z ∈ Z (z ∈ Y ) ]
Chapter 1. Set Theory 36

Exercise
1. uppose that A ⊆ B with |B| = n and |A| = m. Compute the number of
subsets of B that contain A.
2. Proof that following conditions are equivalent.
(a) A ⊆ B
(b) A ∪ B = B
(c) A \ B = ∅
(d) A ∩ B = A
3. Show that
(a) A ∩ (B∆C) = (A ∩ B)∆(A ∩ C)
(b) A∆(B∆C) = (A∆B)∆C
4. Proof for two sets A, B for which A ∩ B is non-empty
\ \ \
A ∩ B 6= (A ∩ B)

5. Let F be a family of sets. Prove that


F is the smallest set B such that A ⊆ B for all A ∈ F .
S
6. Show that for any set A, P (A) = A
S
7. Show that for any set A, A ⊆ P A
8. Show that for any set A, {∅, {∅}} ∈ P(P(P (A)))
9. Proof P (A) = P (B) ⇐⇒ A = B
37 1.8. ZFC Axioms

Georg Ferdinand Ludwig Philipp Cantor (1845 – 1918)

He was born in Saint Petersburg. Cantor, father was German and his mother
was Russian and Roman catholic. When CANTOR was eleven years old
,his family moved to Germany, although cantor was never at ease in this
country. Canton studied at the Gymnasium here and graduated with an
outstanding report In 1860, He entered the Polytechnique of Zurich in 1862,
where he studied mathematics with his parents approval, he studied there
for a couple of years. In 1863, after the death of his father, Cantor moved
to the university of Berlin. He studied at the University of Gottingen over
summer and completed His first dissertation on the number theory named
‘De aequationibus secondi gradus indeterminatis’. He received his doctorate
in mathematics in 1867. He continued working an separate dissertations
on the number theory and analysis. Cantor solved the problem proving the
uniqueness of the representation. Cantor proved that rational numbers were
countable and could be placed in correspondence to the natural numbers
.Cantor had proved that real algebric numbers, were also countable. He
loved to play The violin. He was awarded with the Sylvester medal for his
work in mathematics in 1913. Some of his Major Works in mathematics
are: Infinite sets, Uncountable sets, Cantor set, Cardinals and Ordinals,
The Continuum hypothesis. Georg Cantor died on 1918 in Halle, after a
prolonged mental illness. There were many publications on Cantor such
as ‘Men of Mathematics’ and the ‘history of mathematics’. He laid the
foundation for Modern Mathematics and most of his works have survived to
date.
Chapter 1. Set Theory 38
Chapter 2

Enough Number Theory

No two persons ever read the same book.


“ ”
Edmund Wilson,

This is not a book in number theory, but we need some basic definiton and
theorm that will be used in following chapters. So I try to talk briefly as much as
possible.

Well Ordering Principle


A nonempty set S of nonnegative integers always contains a smallest element
m. That is, m satisfies the following two conditions: (i)m ∈ S and (ii)m < n for
every number n in S.

Assume a 6= 0 and b are integers. if the remainder of division of b by a is zero


then b = ka. This motivates following definition.

2.1 Divisibility
In number theory divisibility is the key idea and every ideas based on its notion.
Definition 2.1.1 For a, b ∈ Z such that a 6= 0, we say "a divides b" if there exists
an integer q such that b = qa.
a | b ⇐⇒ ∃q : b = qa

39
Chapter 2. Enough Number Theory 40

If a divides b, i.e. a | b we say b is divisible by a or a is a factor of b. If a doesn’t


divide b, then we write a - b. For example 4 | 12 and 3 | 21, while 2 - 3.
As an example take a, b ∈ Q such that a 6= 0, then a always divides b, i.e. a | b
because ∃q : q = b/a ∈ Q ,i.e. there always exists such q in the rationals. Even
b | a since ∃q : q = a/b ∈ Q, That is I call this trivial divisibility because every
non-zero element always divides any other one.
Definition 2.1.2 Any even integer has the form 2k for some integer k, while any
odd integer has the form 2k + 1 for some integer k.
Example 2.1.3 2|n if n is even, while 2 - n if n is odd.

Here we state some Basic Properties of Divisibility.


Theorem 2.1.4 For all a, b, c ∈ Z, we have
(1) a | a, 1 | a, and a | 0

(2) 0 | a ⇐⇒ a = 0

(3) a | b ⇐⇒ −a | b ⇐⇒ a | −b

(4) a | b & a | c =⇒ a | (b ± c)

(5) a | b & b | c =⇒ a | c

(6) a | b & b | a ⇐⇒ a = ±b

Proof These can be directy concluded from the definition, and the proof is left for
readers.
The product of any two non-zero integers is non-zero. This implies the usual
cancellation law: if a, b, and c are integers such that a 6= 0 and ab = ac, then we
must haveb = c; indeed, ab = ac implies a(b − c) = 0, and so a 6= 0 implies
b − c = 0, and hence b = c.
More generally, we have the following fundamental theorem, called The Divi-
sion Algorithm.
Theorem 2.1.5 [The Division Algorithm]
41 2.1. Divisibility

Given any a, b ∈ Z, b 6= 0, there exist unique q, r ∈ Z such that a = qb + r and


0 ≤ r < |b|.

∀a, b ∈ Z, b 6= 0 : ∃!q, r ∈ Z : a = qb + r, 0 ≤ r < b

Here q is called quotient of the integer division of a by b, and r is called remainder.


The number r will be denoted by a mod b.
Existence
Consider the set S = {a − bq|q ∈ Z and a − bq ≥ 0}. If 0 ∈ S, then b divides
a and we get q = a/b and r = 0.
Now assume 0 ∈ / S. According to Well Ordering Principle Since S is a
nonempty set, we know that S has a smallest member r = a − bq Then a = bq + r
and r ≥ 0.
Moreover, if r ≥ b, then r − b ≥ 0, so a − bq − b ≥ 0, or a − b(q + 1) ≥ 0 So
a − b(q + 1) ∈ T , but r = a − bq > a − b(q + 1). This contradicts the assumption
that r was the smallest element of S
Uniqueness
Suppose we have another pair q 0 and r0 such that a = bq 0 + r0 , 0 ≤ r0 < b.
Then bq + r = bq 0 + r0 .
So r − r0 = b(q 0 − q), so b | (r − r0 ).
Since 0 ≤ r < b and 0 ≤ r0 < b, we have r − r0 = 0, and therefore r = r0 and
q = q0
Example 2.1.6 For a = 23 and b = 3, then 23 = 3 · 6 + 5. Here q = 6 and r = 5.
For a = −23 and b = 3, then −23 = 3 · (−8) + 1. Here q = −8 and r = 1.
Remark The Division Algorithm is a consequence of the Well-Ordering Axiom
for the positive integers.
Definition 2.1.7 Let a, b ∈ Z, and write a = qn + r. We denote the remainder r
by ā, or [a]n , and call it the remainder of a mod n.
Example 2.1.8 Let a = 31 and n = 7. Then a = 31 = 4.7 + 3 so the remainder
is [a]n = [37]7 = 3
We consider cases that two integers have the same remainder divided by n .
This leads us to define the notion of congruence mod n.
Chapter 2. Enough Number Theory 42

2.2 Congruence

Definition 2.2.1 Given integers a and n, with n > 0, a mod n is defined to be the
remainder when a is divided by n.
The most important application of the division algorithm that we use in everyday
life is modular arithmetic. For example, if you measure time with a 12-hour clock,
then you are calculating the hour modulo 12. Now that I’m writing this paper, the
time is 11 o’clock. If someone asks me to call him 4 hours later,it means 3 o’clock.
What we do without thinking is in fact this : 11 + 4 = 15 ≡ 3 (mod 12) .
Today is Saturday, they say the university start then 20 days from today. 20 ≡ 6
(mod 7) will be Friday.
Today is Saturday August 29 2015. To say what day of the week it will be in
2017, we can calculate 730 ≡ 2 (mod 7) 2 · 365 = 730 ≡ 2 (mod 7)
Definition 2.2.2 (Gauss) Let a, b ∈ Z and n ∈ N. Then a is congruent to b modulo
(or mod) n, if n | (b − a). That is, a and b have the same remainder when divided
by n, i.e. [a]n = [b]n And we write

a ≡ b (mod n).

or a ≡n b
For example, 26 ≡ 4 (mod 11). since 26 - 4 = 22 is divisible by 11.
We say 26 and 4 are congruent mod 11.

Example 2.2.3
23 mod 4 ≡ 3 since 23 = (5)4 + 3 and 06263
−23 mod 4 ≡ 1 since −23 = (−4)4 + 1 and 06164

Lemma 2.2.4 Let a, b, n ∈ Z.Then the following statements are equivalent .


(a) a ≡ b (mod n).

(b) a − b = kn for some integer k.

(c) a = kn + b for some integer k.


43 2.2. Congruence

(d) a (mod n) = b (mod n).

Proof The proof is obtained from the definition, and is left to readers.
Remark
If n = 0 then a ≡ b (mod n) ⇐⇒ a = b. So equality is a special kind of
congruency.
If n = 1 then a ≡ b (mod n) ⇐⇒ a − b = k is always true.
So congruence mod n = 0, 1 is not interesting.
We proof a theorem which states congruence modulo n is an equivalence
relation.
According to division algorithm, if a, b ∈ Z we can talk about quotient and
remainder of division a by b.Take b as a fixed integer, then the remainder of division
a by b is a cyclic set.
Since the remainder is between zero and b, i.e. 0 ≤ r < b the set
{a (mod n) : n ∈ Z} is exactly the same as {0, 1, . . . , n − 1} these are all
possible remainders when n is divided by m.
Taking n = 2, every integer is congruent mod 2 to exactly one of 0 and 1.
Saying a ≡ 0 (mod 2) means n = 2k for some integer k, so n is even.
Andsaying a ≡ 1 (mod 2)means n = 2k + 1 for some integer k, so n is odd.
We have a ≡ b (mod 2) precisely when a and b have the same parity, i.e.
both are even or both are odd.
Take b = 2. the remainder of division 0, 1, 2, 3, 4, . . . by 2 is 0, 1, 0, 1, 0, . . .

Take b = 3. the remainder of division a ∈ Z by 3 is one of {0, 1, 2}.

Take b = 4. the remainder of division a ∈ Z by 3 is one of {0, 1, 2, 3}.

Take b = n. the remainder of division a ∈ Z by n is one of {0, 1, . . . , n − 1}.

Theorem 2.2.5 Let n be a positive integer. Every integer is congruent modulo n


to exactly one integer between 0 and n − 1.
Chapter 2. Enough Number Theory 44

Proof Let n be a positive integer. If a is an integer, then the Division Algorithm


gives us integers q and r with a = qn + r and 0 6 r < n. Thus, a − r = qn is
divisible by n, so a ≡ r (mod n).
If s is between 0 and n−1 and a ≡ s (mod n) , according to Division Algorithm
a and s have the same remainders after division by n. But, since s = 0.n + s and
a = qn + r, the lemma tells us that s = r. Thus, a is congruent modulo n to
exactly one integer between 0 and n − 1.
All the other numbers can be found congruent to one of the n numbers.

Remark The mod operation is derived from the Division Algorithm: If we divide
the integer a by the positive integer b, we get a unique quotient q and remainder r
satisfying a = bq + r and 0 6 r < b. The remainder r is defined to be the value of
a mod m.
Definition 2.2.6 Let n ∈ N and a ∈ Z. The unique integer between 0 and n − 1
to which a is congruent modulo n is called the least residue of a modulo n.
Definition 2.2.7 The set of remainders of integer division by n is denoted by Zn .
For any integer a , its remainder after division by n,i.e. [a]n ∈ Z
For example, let n = 4 . Then [7]4 = [11]4 = [3]4 = [−1]4 = 3
Corollary 2.2.8
The set of all elements congruent to x modulo n is called the congruence class
containing x and is denoted x̄ or [x]n . The set of all congruences class modulo n is
denoted Zn = Z/nZ = {0̄, 1̄, . . . , n − 1}.
Since size of the set Zn is equal to n and every integer must be in one of the n
congruences classes.
The additive identity, additive inverse, and multiplicative identity always exist. In
particular, if we want to solve x + a ≡ b (mod n), then we are guaranteed that
the additive inverse of a, called −a (or n − a), exists and we are allowed to write
x ≡ b − a (mod n).

2.3 Modular Arithmetic


We can add, subtract or multiply congruences. That is, if a ≡ b (mod n) and c ≡ d
(mod n), then
a + c ≡ b + d (mod n) and ac ≡ bd (mod n).
45 2.3. Modular Arithmetic

Definition 2.3.1 Let a, b ∈ Zn . Take x, y ∈ Z such that a = [x], b = [y]. The


addition of two elements is defined as

a + b = [x] + [y] := [x + y]

or sometimes
a +n b = [x] +n [y] := [x + y]n
Also The multiplication of two elements is defined as

a.b = [x].[y] := [x.y]

or
a.n b = [x].n [y] := [x.y]n
Example 2.3.2 The set Z4 consists of [0], [1], [2], [3]. Addition and multiplication
rule can be given by a table.

+4 0 1 2 3 ×4 0 1 2 3
0 0 1 2 3 0 0 0 0 0
1 1 2 3 4 1 0 1 2 3
2 2 3 0 1 2 0 2 0 2
3 3 4 1 2 3 0 3 2 1
Table 2.1: Addition and multiplication mod 4

Theorem 2.3.3 Addition and multiplication defined above is well-defined.


Theorem 2.3.4 Let a, b, c and d denote integers. Let m be a positive integers.
Then:
1. If a ≡ b(mod m), then b ≡ a(mod m).
2. If a ≡ b(mod m) and b ≡ c(mod m), then a ≡ c(mod m).
3. If a ≡ b(mod m), then a + c ≡ b + c(mod m).
4. If a ≡ b(mod m), then a − c ≡ b − c(mod m).
5. If a ≡ b(mod m), then ac ≡ bc(mod m).
6. If a ≡ b(mod m), then ac ≡ bc(mod mc), for c > 0.
7. If a ≡ b(mod m) and c ≡ d(mod m) then a + c ≡ (b + d)(mod m).
8. If a ≡ b(mod m) and c ≡ d(mod m) then a − c ≡ (b − d)(mod m).
9. If a ≡ b(mod m) and c ≡ d(mod m) then ac ≡ bd(mod m).
Chapter 2. Enough Number Theory 46

1. If a ≡ b(mod m), then m | (a − b). Thus there exists integer k such that
Proof
a − b = mk, this implies b − a = m(−k) and thus m | (b − a). Consequently
b ≡ a(mod m).
2. Since a ≡ b(mod m), then m | (a − b). Also, b ≡ c(mod m), then m | (b − c).
As a result, there exit two integers k and l such that a = b + mk and b = c + ml,
which imply that a = c + m(k + l) giving that a = c(mod m).
3. Since a ≡ b(mod m), then m | (a − b). So if we add and subtract c we get

m | ((a + c) − (b + c))

and as a result
a + c ≡ b + c(mod m).

4. Since a ≡ b(mod m), then m | (a − b) so we can subtract and add c and we get

m | ((a − c) − (b − c))

and as a result
a − c ≡ b − c(mod m).

5. If a ≡ b(mod m), then m | (a−b). Thus there exists integer k such that a−b = mk
and as a result ac − bc = m(kc). Thus

m | (ac − bc)

and hence
ac ≡ bc(mod m).

6. If a ≡ b(mod m), then m | (a−b). Thus there exists integer k such that a−b = mk
and as a result
ac − bc = mc(k).
Thus
mc | (ac − bc)
and hence
ac ≡ bc(mod mc).

7. Since a ≡ b(mod m), then m | (a − b). Also, c ≡ d(mod m), then m | (c − d).
As a result, there exits two integers k and l such that a − b = mk and c − d = ml.
Note that
(a − b) + (c − d) = (a + c) − (b + d) = m(k + l).
47 2.3. Modular Arithmetic

As a result,
m | ((a + c) − (b + d)),
hence
a + c ≡ b + d(mod m).

8. If a = b + mk and c = d + ml where k and l are integers, then

(a − b) − (c − d) = (a − c) − (b − d) = m(k − l).

As a result,
m | ((a − c) − (b − d)),
hence
a − c ≡ b − d(mod m).

9. There exit two integers k and l such that a − b = mk and c − d = ml and thus
ca − cb = m(ck) and bc − bd = m(bl). Note that

(ca − cb) + (bc − bd) = ac − bd = m(kc − lb).

As a result,
m | (ac − bd),
hence
ac ≡ bd(mod m).

Properties of arithmetic modulo n

Because of the simple formula for addition and multiplication in Zn , there are
some we analogues of the common properties of integer arithmetic. In particular,
the following properties hold for all elements a, b, c ∈ Zn .

Commutativity of Addition: [a + b] = [b + a]
Associativity of Addition: ([a] + [b]) + [c] = [a] + ([b] + [c])
Existence of an Additive Identity:[a] + [0] = [a]
Existence of Additive Inverses:[a] + ([−a]) = 0
Commutativity of Multiplication:[a] · b = b · a
Chapter 2. Enough Number Theory 48

Associativity of Multiplication:([a] · [b]) · [c] = [a](·[b] · [c])


Existence of a Multiplicative Identity: [a] · 1 = [a]
Distributivity:[a] · ([b] + [c]) = ([a] · [b]) + [c]
Chapter 3

Relations and Functions

“Whoever loves to meet God, God loves to meet him1 .”


“ ”
Prophet Muhammad,

In this chapter we continue our program to build up useful notion and concepts
from set theory, with some basic set-theoretic definitions of ordered pairs, relations,
and functions.

3.1 Ordered pairs and Cartesian products


In euclidean geometry a point is described by two co-ordinates x, y , that is, the
point is described by a pair (x, y) and (x, y, z) in plane and space respectively. the
pair (x, y) is called an ordered pair because In a pair, the order of the terms are
important . (x, y) and (y, x) represent different points in the plane. But as we
mentined in the first chapter because of axiom of extensionality the sets {x, y} and
{y, x} are the same, since they have the same elements. Therefore, the ordered
pair (x, y) is not the same as the set {x, y} which is defined to be unordered. So
we come up with following definition :
Definition 3.1.1 (Equality of ordered sets) The ordered pair construct (a, b) with
first component a and second component b have the property that

(a, b) = (c, d) ⇐⇒ a = c ∧ b = d
1
This is a symmetric relation

49
Chapter 3. Relations and Functions 50

We would like to define ordered pairs in terms of sets but how ? Because sets
don’t respect order.
Consider a set {a, b, c, d}. Suppose we want to propose an order in a set-
theoretic manner for the elements of the set as follows : d first, a second, b third,
c forth Take the first element and build a set out of it, that is {d}.Then take a
set and put the next element and everything that is before it, that is , {d, a}. Do-
ing so again we have {d, a, b} and the last one is {d, a, b, c}. Now collect them
all in a new set O = {{d}, {d, a}, {d, a, b}, {d, a, b, c}} In this way we have a
notion of order because d, a, b, c are repeated 4, 3, 2, 1 respectively. Changing
the order of members of O is irrelevant. For example if you write this, O =
{{d}, {a, d}, {a, b, d}, {a, b, c, d}} or even O = {{d}, {a, d}, {d, a, b}, {d, c, b, a}}
nothing is wrong and we can find out what the order of elements are.
This motivates Kuratowski definition of Ordered Pair.
Definition 3.1.2 The ordered pair (a, b) is defined as follows:
(a, b) = {{a}, {a, b}}

Notice that its existence follows from the Axiom of Pair Set alone.
We have to prove that Kuratowski definition obeys the definition of Equality of
ordered sets. That is,
{{a} , {a, b}} = {{c} , {c, d}} ⇐⇒ a = c ∧ b = d
Proof ⇐=
This follows from our definition.
n o n o
If a = c ∧ b = d then (a,b) = {a}, {a, b} = {c}, {c, d} = (c, d).
⇐==⇒
(i) If a = b, then {a, b} = {a, a} = {a}, so the set (a, b) = {{a}, {a}} = a is
a singleton, so the set (c, d) is also a singleton, so that c = d and (c, d) = {{c}},
and since this last singleton is equal to {{a}}, we have a = c and, hence, also
a = b = c = d.
(ii) If a 6= b ...
Take it as an exersice.
 
Remark (i) {a}, {a, b} = {a, b}, {a} because two sets are equal if and only
if they have the same elements and the order dosen’t matter.
(ii)
51 3.1. Ordered pairs and Cartesian products

We can further define ordered triples

(a, b, c) = ((a, b), c) = {{{a}, {a, b}}, {{{a}, {a, b}}, c}}

or simply
(a, b, c) = {{a}, {a, b}, {a, b, c}
Generally, the ordered n-tuples

(a1 , . . . , an ) = {{a1 }, {a1 , a2 }, . . . , {a1 , . . . an−1 }, {a1 , . . . an }}

Cartesian product
Suppose we have two sets A and B and we form ordered pairs by taking an
element of A as the first member of the pair and an element of B as the second
member.
Definition 3.1.3 The Cartesian product of two sets A and B, denoted A × B is
the set of all ordered pairs (a, b) with a ∈ A, b ∈ B. and defined as follows

A × B = {(a, b) : a ∈ A ∧ b ∈ B}

Theorem 3.1.4 The Cartesian product of two sets A and B exists


Proof We can use the axioms to show that the set A × B exists. By Axiom of Pair,
A ∪ B exists as a unique set. Thus P(A ∪ B) exists. Using Axiom of Power Set
again guarantee P(P(A ∪ B)) exists.
If a ∈ A and b ∈ B then {a} ⊂ A and {b} ⊂ B, therefore {a, b} ⊂ A ∪ B
so we can conclude {a, b} ∈ P(A ∪ B) since also {a} ⊂ A ∪ B we can say
{a} ∈ P(A ∪ B) . We can conclude again, {{a}, {a, b}} ⊂ P(A ∪ B) and hence
{{a}, {a, b}} ∈ P(P(A ∪ B)).
So we proved {{a}, {a, b}} = (a, b) is an element of P(P(A ∪ B)). Now apply
the Axiom of Seperation with the properties P (a) : a ∈ A and P (b) : b ∈ B to
construct A × B .
In general, we can extend the definition to n arbitrary sets A1 , A2 , . . . , An .

A1 × A2 × · · · × An = {(a1 , a2 , . . . , an ) : a1 ∈ A1 ∧ a2 ∈ A2 . . . ∧ an ∈ An }

If A1 = A2 = · · · = An we write An istead of A1 × A2 × · · · × An .
For example {(x, y) : x ∈ R ∧ y ∈ R} is the set of coordinates of points in
the plane . We often write it as (x, y) ∈ R × R or (x, y) ∈ R2 .
Chapter 3. Relations and Functions 52

Example 3.1.5 Let A = {a, 1, 2} and B = {b, c} . Then


A × B = {(a, b), (a, c), (1, b), (1, c), (2, b), (2, c)}
But
B × A = {(b, a), (c, a), (b, 1), (c, 1), (b, 2), (c, 2)}
The Cartesian product A × B is not commutative,
A × B 6= B × A, Unless A = B or one of them is the empty set.
It is not even associative.
(A × B) × C 6= A × (B × C) 6= A × B × C
If A, B are any finite sets, then |A × B| = |A||B|, because there are |A| ways of
choosing to first component of an element (x, y) of |A × B|, and for each choice
|B| ways of choosing the second component.
That is why we use × to denote Cartesian product of sets.

3.2 Relation
As it comes from its name a relation is a relation; that is it considers two objects.
Even when we have only an element x and a relation R we are comparing x with
itself and generalizing the idea of relation between two objects to one object with
itself. So it is pointless to define a relation on a singleton. It is convinent to define
a relation as a subset of cartesian product since it deals with the set of two things
related toghter, If we want to define a relation on a set X we may define it in the
product of a set with itself, i.e. X × X.
Because a relation might be consists of some pairs or include entire set it is a
subset of X × X.
If a point (a, b) ∈ X × X is in R we write aRb. for example a relation > is a
subset of R2 if (a, b) ∈ > then we write a > b.
Examples of relations on the set of real numbers include <, =, and ≥ . And of
relations on P(X), the power set of X, include = and ⊆ .
In fact, given sets A and B, a relation between A and B makes a special link
between elements of A with elements of B.
Definition 3.2.1 Let A and B be two sets. A (binery) relation from A to B,
denoted by A → B is a triple, (A, R, B) where R ⊆ A × B is any subset of A × B
53 3.2. Relation

. If A = B then we call such a relation a relation on A. If (a, b) ∈ R then we say a


and b are related and we write aRb .
Note that by this definition a relation is just any subset of A × B. If R is a
relation on A × B, we call the set A the domain of R and B the codomain of R.
The relation of equality in a nonempty set A. R = {(a, a) : a ∈ A} thus
(a, b) ∈ R ⊆ A × B ⇐⇒ a = b
Like ordered pairs, two relations, (A, R, B) and (A0 , R0 , B 0 ), are equal ⇐⇒
A = k 0 , B = B0 and R = R0
Definition 3.2.2 (Domain and range of relations.) Let A and B be two sets and R
be a relation from A → B, i.e. R ⊆ A × B .
(i) Domain of the relation R: The domain of R is defined by
Dom(R) := {a ∈ A : ∃b ∈ B , (a, b) ∈ R}

(ii) Range of the relation R: The range of R is defined by


Rng(R) := {b ∈ B : ∃a ∈ A , (a, b) ∈ R}

Here for some reasons we would like to compare objects from the same set, i.e.
we consider relation on one set A × A.
In the case where R is a relation for which the domain and codomain are the
same set A, that set A is called the underlying set for R.
Definition 3.2.3 (Inverse relation.) Let A and B be two sets and R be a relation
from A → B, i.e. R ⊆ A × B . The inverse of R (denoted by R−1 ) is defined by
R−1 := {(b, a) ∈ B × A (a, b) ∈ R}

What we have defined is abinary relations, i.e., sets of ordered pairs. We can
also define ternary, quaternary or just n-place relations consisting respectively of
ordered triples, quadruples or n-tuples.

Example 3.2.4 Let A = {1, 2, 3, 4} then each of the following is a relation on A.


R1 = {(1, 3), (2, 4)}
R2 = {(1, 1), (2, 2), (3, 3), (4, 4)} = {(a, b) ∈ A × A : a = b}
In relation R1 : 1 is related to 3 and 2 is related to 4 ,i.e. (1, 3) ∈ R1 and
(2, 4) ∈ R1 respectivly. But 4 is not related to 1 (4, 1) ∈
/ R1 . We say relation R1 is
not symmetric. (a, b) ∈ R1 but (b, a) ∈ / R1 or a R1 b but b 6R1 a .
Chapter 3. Relations and Functions 54

In relation R2 : 1 is related to 1 , 2 is related to 2,etc. That is, every element is


related to itself. (a, a) ∈ R2 or a R2 a .
There are some differences between relations. Consider the “sisterhood” rela-
tion. Neda could be a sister of Ali, but Ali is not a sister of Neda. We say that
sisterhood relation is not symmetric.
However, the “neighborhood” relation is symmetric: if A is the neighbor of B
then B is the neighbor of A.
In this case, the relation has a further property. If A is a neighbor of B and B is
a neighbor of C then A is a neighbor of C. We call this property transitivity. An
important, and very useful, class of relations are the relations that are reflexive,
symmetric and transitive. We call this kind of relation, equivalence relations. For
equivalence relations instead of a R b we use the special notation a ∼ b.

3.3 Partitions of a Set


Consider the set of integers Z.
We can divide it into two parts in different ways:
• The set of negative and non-negative integers.
Z = {. . . , −5, −3, −1} ∪ {0, 2, 4, . . . }

• The set of even integers and the set of odd integers.


Z = {. . . , −5, −3, −1, 1, 3, 5, . . . } ∪ {. . . , −4, −2, 0, 2, 4, 5, . . . }

In both cases every integer is a member of one of these subsets, and no integer
is a member of both, so this gives a partition of Z:
Definition 3.3.1 A partition of a set A is a collection P of non-empty subsets
A1 , A2 , . . . , An satisfying the following properties.
(a) A is the union of all the A0i : A =A1 ∪ A2 ∪ . . . An

(b) the Ai are disjoint: Ai ∩ Aj = ∅ ; for all i 6= j, 1 ≤ i, j ≤ n.


The elements of P are called the blocks of the partition.
The sequence of sets A1 , A2 , . . . could be finite or infinite, so we had
S better
defined partition of A to be collection of subsets {Ai }i∈I of A such that i∈I Ai .
55 3.4. Equivalence relations, equivalence classes

Remark
(a) For any set A, P = {A} is a partition of A, called the trivial partition.
(b) For any A ⊂ U , the set A with its complement A form a partition of U .
Example 3.3.2
(a) The empty set {} and singleton {x} have only one partition, {{}} and {{x}}
respectivly
(b) A set with 2 elements, say , {1, 2} has 2 different partition:

{1, 2}

{{1}, {2}}

(c) A set with 3 elements, say , {1, 2, 3} has 5 different partition:

{1, 2, 3}

{{1}, {2}, {3}}


{{1, 2}, {3}}
{{1, 3}, {2}}
{{1}, {2, 3}}

The number of different partition of a set with n element is called the Bell
number Bn .

n  
X n
Bn+1 = Bk
k=0
k
The number of partition is exponentialy increasing as the cardinal goes higher, for
example the set | A |= 7 has 877 partition.

3.4 Equivalence relations, equivalence classes

The equality relation = between two or more objects is associate with the concept
of being the same or being identical but sometimes we classify non-identical object
into a collecton. This abstraction of equality is called Equivalence relation.
Chapter 3. Relations and Functions 56

Equivalence relation is generalization of the notion of equality. Indeed, the


usual notion of equality among the set of integers is an example of an equivalence
relation. This is the " equivalence The more interesting concepts arose from trying
to abstract some of the most useful properties of equality.
Definition 3.4.1 Let ∼ be a relation on a set A. We call ∼ an equivalence relation
on A if it satisfies the following properties:

• (i) Reflexivity: a ∼ a (∀a ∈ A)

• (ii) Symmetry: a ∼ b =⇒ b ∼ a (∀a, b ∈ A)

• (iii) Transitivity: a ∼ b ∧ b ∼ c =⇒ a ∼ c (∀a, b, c ∈ A)


Remark Since an equivalence relation is reflexive, it implise that the set A is not
empty. That is why we didn’t note being nonempty in the definition.
In the language of sets, we can rewrite these properties as follows:
• (i) Reflexivity: (a, a) ∈∼ (∀a ∈ A)

• (ii) Symmetry: (a, b) ∈∼ =⇒ (b, a) ∈∼ (∀a, b ∈ A)

• (iii) Transitivity: (a, b) ∈∼ ∧ (b, c) ∈∼ =⇒ (a, c) ∈∼ (∀a, b, c ∈ A)


Example 3.4.2
(a) Let A be a set and consider equality = relation. Then = is obviously an
equivalence relation on A since
• (i) Reflexivity: a = a (∀a ∈ A)

• (ii) Symmetry: if a = b =⇒ b = a (∀a, b ∈ A)

• (iii) Transitivity: if a = b ∧ b = c =⇒ a = c (∀a, b, c ∈ A)

(b) The relation < on Z is transitive, but not reflexive or symmetric. The relation
≤ is transitive and reflexive, but not symmetric.
(c) The relation ∼ on R defined by a ∼ b ⇐⇒ a − b ∈ Z is an equivalence
relation, but the same relation defined by a ∼ b ⇐⇒ a − b ∈ N is not, since it
57 3.4. Equivalence relations, equivalence classes

dosen’t satisfy symmetry property.


As we mentioned equality is a special case of equivalence relation. Any kind of
equality, such as congruence of triangles, is generally an equivalence relation.
If ∼ is an equivalence relation on a set A , then x ∼ y is more than just being
related, they are equivalent.

Recall, the next definition which is another example of an equivalence relation.


Definition 3.4.3 (Gauss) Let a, b ∈ Z and n ∈ N. Then a is congruent to b modulo
(or mod) n, if n | (b − a). That is, a and b have the same remainder when divided
by n, i.e. [a]n = [b]n And we write

a ≡ b (mod n).

we define the modular relation ≡n , by a ≡n b if n | a − b, i.e.

≡n : = {(a, b) : n | a − b}.

Theorem 3.4.4 Congruence modulo n is an equivalence relation on Z. That is:


∀a, b, c ∈ Z and n ∈ N

• (i) Reflexivity: a ≡ a (mod n)


(ii) Symmetry: a ≡ b (mod n) =⇒ b ≡ a (mod n)
(iii) Transitivity: a ≡ b (mod n) ∧ b ≡ c (mod n) =⇒ a ≡ c (mod n).
Proof
(i) a − a = 0 and n | 0 =⇒ a ≡ a (mod n)
(ii) a ≡ b (mod n) means a − b = nk for some k ∈ Z. Therefore, b − a = −nk =
n(−k) so b ≡ a (mod n)
(iii) a ≡ b (mod n) means a − b = nk and b ≡ c (mod n) means b − c = nk 0
Therefore, a − c = n(k − k 0 ) , so a ≡ c (mod n)

Equivalence relation partitions a set into disjoint subsets, each of them consists
of equivalent elements. These subsets are called equivalence classes.
Chapter 3. Relations and Functions 58

Definition 3.4.5 Let ∼ be an equivalence relation on a set A. For each a ∈ A


the equivalence class of a is defined to be the set of all elements of A that are
equivalent to a. The equivalence class of a will be denoted by [a] or [a]∼ or a.

[a] = {x ∈ A | x ∼ a}
[a] is not just an element, it is a set of elements. An element of an equivalence
class X ⊆ A is called a representative of X. Here a is a representative of the
equivalence class [a].
Beacuse of reflexivity we have a ∈ [a]. So any a ∈ A is a representative of its
own equivalence class.
Every equivalence relation makes equivalence classes.
The equivalence classes [a] are subsets of A. The set of all equivalence classes
is called the quotient space, denoted by A/∼ .
Equivalence classes for congruence mod n are also called congruence classes.
Let a be an integer. By the definition of an equivalence class we have [a] = {x ∈
Z | x ≡n a} = {x ∈ Z | x = a + kn ;k ∈ Z }.
Consider the relation ≡2 on Z.Then

[0] = {n ∈ Z | 0 ≡2 n} = {n ∈ Z | 2 | n} = {n ∈ Z |n = 2k, k ∈ Z} = {2k|k ∈ Z}

[1] = {n ∈ Z | 1 ≡2 n} = {n ∈ Z | 2 | (n−1)} = {n ∈ Z |n−1 = 2k, k ∈ Z} = {2k+1|k ∈ Z}


[2] = {n ∈ Z | 2 ≡2 n} = {n ∈ Z | 2 | (n − 2)} = {n ∈ Z |n − 2 = 2k 0 , k 0 ∈ Z}

= {2k|k ∈ Z} = [0]

[3] = {a ∈ Z | 3 ≡2 a} = {a ∈ Z | 2 | (n − 3)} = {a ∈ Z |n − 3 = 2k 0 , k 0 ∈ Z}

= {2k + 1|k ∈ Z} = [1]


We have :
Let R = {(x, y)| x and y have the same parity} be a relation on Z

[0] = [2] = ... = [2k] = {0, ±2, ±4, ±6, . . . , ±2k, . . . }

[−1] = [1] = ... = [2k + 1] = {±1, ±3, ±5, . . . , ±(2k + 1), . . . }


59 3.4. Equivalence relations, equivalence classes

This shows when n = 2, the equivalence classes under congruence mod 2


partition Z into two class, the even integers and the odd integers.
Therefore, the quotient space (the set of all equivalence classes) has only 2
elements.
Z/≡2 = {[0], [1]} = {[2k], [2k + 1]}

Now consider the relation ≡3 on Z. The equivalence relation ∼ on Z has three


equivalence (congruence) classes:

[0] = {n ∈ Z | 0 ≡3 n} = {n ∈ Z | 3 | n} = {n ∈ Z |n = 3k, k ∈ Z} = {3k|k ∈ Z}

[1] = {n ∈ Z | 1 ≡3 n} = {n ∈ Z | 3 | (n−1)} = {n ∈ Z |n−1 = 3k, k ∈ Z} = {3k+1|k ∈ Z}


[2] = {n ∈ Z | 1 ≡3 n} = {n ∈ Z | 3 | (n−2)} = {n ∈ Z |n−2 = 3k, k ∈ Z} = {3k+2|k ∈ Z}

Z/≡3 = {[0], [1], [2]} = {[3k], [3k + 1], [3k + 2]}

In general, for any a ∈ Z, the equivalence class of a is

[a] = {a + kn |k ∈ Z}

Therefore, the quotient space (the set of all equivalence classes also called congru-
ence class representatives modulo n) has only n elements.

Z/≡n = {[0], [1], [2], . . . , [n − 1]} = {[nk], [nk + 1], [nk + 2], . . . , [nk + (n − 1)]}

The quotient space Z/≡n is also denoted by Z/nZ or Zn . This is one of the
most important and useful example of equivalence classes so I make a definition
out of it.
Definition 3.4.6
The set of all elements congruent to a modulo n is called the congruence class
containing a and is denoted [a] or a. The set of all congruences class modulo n is
denoted Z/≡n = Z/nZ = Zn = {[0], [1], [2], . . . , [n − 1]}.
Operations on equivalence classes
An operation like addition or multiplication defined on the original set A can
also be defined on the equivalence classes. For example, in Z3 we have three
equivalence classes, namely

[0] = {3k|k ∈ Z} = {. . . , −3, 0, 3, . . . }


Chapter 3. Relations and Functions 60

[1] = {3k + 1|k ∈ Z} = {. . . , −2, 1, 4, . . . }


[2] = {3k + 2|k ∈ Z} = {. . . , −1, 2, 5, . . . }

Every integer is in one and only one of these classes.


we can say for example, −3 + 4 = 1 but since −3 ∈ [0] and 4 ∈ [1] and every
element of [0] and [1] is of the form 3k and 3k + 1 respectivly, 3n + 3m + 1 =
3(n + m) + 1 = 3k 0 + 1 = [1] so the addition we defined in perivious chapter
makes sence here. −3 + 4 = [0] + [1] = [0 + 1] = [1]
Example 3.4.7 Let A be the set of fractions: A = { pq : p, q ∈ Z, q 6= 0} Define a
relation ∼ on A by:
a c
∼ ⇐⇒ ad = bc
b d
This relation is an equivalence relation. Beacuse
a a
1) For any fraction b
∼ b
since ab = ba. (Reflexitivity)

a
2) If b
∼ dc , then ad = bc, so cb = da and c
d
∼ ab . (Symmetry)

3) If ab ∼ dc , and dc ∼ fe , then ad = bc and cf = de. Multiply the first equation by


f and divide by d to get af = be, so ab ∼ fe . (Transitivity)
The equivalence classes of this equivalence relation, for example:
1 1 2 3 n
[ ] = { , , ,..., ,...}
1 1 2 3 n
1 1 2 3 n
[ ] = { , , ,..., ,...}
2 2 4 6 2n
3 3 6 9 3n
[ ] = { , , ,..., ,...}
7 7 14 21 7n
The set of all the equivalence classes are called rational numbers denoted by Q.

A/∼ = Q

Since ∼ is reflexive x ∼ x, therefore x ∈ [x].


Theorem 3.4.8 Let ∼ be an equivalence relation on a set A. Then the following
are equivalent:
61 3.4. Equivalence relations, equivalence classes

(i)a ∼ b

(ii)[a] = [b]

(iii)[a] ∩ [b] 6= ∅

Proof (i) ⇒ (ii) . Suppose a, b ∈ A and a ∼ b. We must show that [a] = [b].
Suppose x ∈ [a]. Then, by definition of [a], x ∼ a. Since ∼ is symmetric and
a ∼ b,then x ∼ b. Therefore, x ∈ [b].
Suppose x ∈ [b]. Then b ∼ x. Since a ∼ b and ∼ is transitive, a ∼ b. Thus,
x ∈ [a].
We have shown that x ∈ [a] ⇐⇒ x ∈ [b]. Thus, [a] = [b].
(ii) ⇒ (iii). Suppose a, b ∈ A and [a] = [b]. Then [a] ∩ [b] = [a]. Since ∼ is
reflexive, a ∼ a; that is a ∈ [a]. Thus [a] ∩ [b] = [a]
(iii) ⇒ (i). Suppose [a] ∩ [b] 6= ∅. Then there is an x ∈ [a] ∩ [b]. By definition,
a ∼ x and b ∼ x. Since ∼ is symmetric, x ∼ b. By transitive a ∼ x and x ∼ b,
hence a ∼ b.

Equivalence relations and partitions


Equivalence relations and artitions are related closely together. The following
theorem connects the concepts of an equivalence relation on a set A (which, is a
subset of A × A with certain properties) and a partition on A (which is a collection
of subsets of A with certain properties).
Theorem 3.4.9 There is a one to one correspondence between equivalence rela-
tions defined on a set A and the set of partitiona of A.
• Equivalence relations ⇒ partitions:
If ∼ is any equivalence relation on the set A then the collection of all distinct
equivalence classes (i.e. A/∼ ) is a partition of A.
According to the previous theorem, we have:
(a) For any x ∈ A there exists an equivalence class that contains x (namely, the
class [x]).
(b) For any a, b ∈ A, the associated equivalence classes [x] and [y] are either equal
[a] = [b]
Chapter 3. Relations and Functions 62

or disjoint [a] ∩ [b] = ∅

• Partitions ⇒ equivalence relations:


Given any partition P of the set A into sets A1 , A2 , . . . , An , the relation defined by

a ∼ b ⇐⇒ a, b ∈ Ai for some block Ai ∈ P

is an equivalence relation on A. in which case the equivalence classes of ∼ are the


blocks of the partition
According to the definition of a partition, each element of A is in exactly one block.
(i) Since x is in the same block as itself, x ∼ x , so ∼ is reflexive.
(ii) If a ∼ b, then a and b are in the same block, which is the same thing as b ∼ a,
so . So ∼ is symmetric.
(iii) If a ∼ b and b ∼ c then a and b are in the same block also as b and c. again
with the definition of a partition a and c must be in the same block, so a ∼ c Hence
∼ is transitive.

3.5 Functions

function is generally represented in set-theoretic terms as a special kind of relation.


A function is a special type of relation. It is a relation with a special restriction
on the first coordinate. For a relation to be a function, it must be the case that every
element in its domain is associated to one and only one element in its range.
Definition 3.5.1 A function ( mapping or transformation) f from A to B, denoted
f : A → B, is a relationf ⊆ A × B such that every element a ∈ A is related to
exactly one element b ∈ B.

In formal notation :
1. ∀a ∈ A ∃! b ∈ B : (a, b) ∈ f

2. (a, b1 ) ∈ f ∧ (a, b2 ) ∈ f ⇒ b1 = b2
By condition 1 we have for any a ∈ A there exists a unique b ∈ B such that
(a, b) ∈ f . We usually write f (a) = b or a 7−→ f (a) instead of (a, b) ∈ f .
63 3.5. Functions

If f : A → B, we refer to the set A as the domain of f and the set B as the


codomain. and if b = f (a), then b is called the image of a and a is a pre-image of
b under f .
Definition 3.5.2 Let f : A → B be a function. Suppose X is a subset of A,
X ⊆ A. The image set f [X] is defined by

Imgf = f [X] = {b ∈ B|(∃a ∈ A)(b = f (a))} = {f (x) ∈ B | x ∈ X}

If Y ⊆ B then the pre-image f −1 [f ] is defined by

f −1 [Y ] = {x ∈ A | f (x) ∈ Y }

A Function Must be Defined


Let f : A → B be a function. Note that by definition, f assgns to each element
a ∈ A a unique element b ∈ B. Thus, for f to be a function f must be defined on
its entire domain.
Example 3.5.3 Let A = {a, b, c} and B = {1, 2}. The relation ϕ : A → B
defined by ϕ(a) = 2, ϕ(b) = 1 is not a function unless we also define ϕ(3) =?

This motivates the following definition which is a generalization of the concept


of a function by defining f to just some elements of the domain not all of it.
Definition 3.5.4 A partial functionf from A to B, denoted f : A * B, is a
relationf ⊆ A × B such that some element a ∈ A is related to exactly one element
b ∈ B.

In formal notation :
1. ∃!b ∈ B ∀a ∈ A : (a, b) ∈ f

2. (a, b1 ) ∈ f ∧ (a, b2 ) ∈ f ⇒ b1 = b2
Be careful about the difference between definition of a function and a partial
function. The only difference is that in condition 1, I swap ∀ and ∃!
A partial function from A to B denoted f : A * B is a function f : X → B,
for some subset X ⊂ A.
Chapter 3. Relations and Functions 64

A Function Must be Well-Defined


Note that by definition, f assgns to each element a ∈ A a unique element
b ∈ B. That is, f : A → B is well-defined ⇐⇒ ∀a1 , a2 ∈ A, if a1 = a2 then
f (a1 ) = f (a2 ).
By condition 2, a function must be well-defined.
Definition 3.5.5 (Equality of functions ) Let f : A → B and g : C → D be
functions.
f = g ⇐⇒ A = C, B = D and f (a) = g(a) ∀a ∈ A
Definition 3.5.6 (Injective, surjective, bijective functions.) Let f : A → B be a
function.
(1) Injective function
f : A  B is injective if and only if for each b ∈ B there is at most one a ∈ A
with f (a) = b
(∀a, a0 ∈ A) [f (a) = f (a0 ) ⇒ a = a0 ]
Equivalently
(∀a, a0 ∈ A) [a 6= a0 ⇒ f (a) 6= f (a0 )]

We also say that f is an injection, or a one-to-one correspondence.


(2) Surjective function
f : A  B is surjective if and only if every element of B is in the image of f :

(∀b ∈ B) (∃a ∈ A) [b = f (a)]

That is Imgf = B .
We also say that f is a surjection, or f is a map onto B
(3) Bijective function
f is bijective if and only if it is both injective and surjective.

The notation  is used to denote that f is an injective map from A to B.


Similarly  means that f is a surjective map.

Remark 1.Any function, f : ∅ → B is trivially injective.


2. f : ∅ → ∅ is trivially surjective.
3. Since Imgf = {b ∈ B|(∃a ∈ A)(b = f (a))}, a function f : A  B is
always surjective onto its image.
65 3.5. Functions

Example 3.5.7 Let A = {a, b, c} and B = {1, 2, 3}. The Cartesian product is

A × B = {(a, 1), (a, 2), (a, 3), (b, 1), (b, 2), (b, 3), (c, 1), (c, 2), (c, 3)}
A relation is just any subset of A × B. It could be the entire set. Let fi ⊆ A × B
be defined by

1. f1 = {(a, 1), (b, 3), (c, 2)} is a function

Imgf1 = {1, 3, 2}

f −1 [B] = {a, b, c}

Also f1 is injective and surjective.


2. f2 = {(a, 1), (b, 3), (c, 1)} is a function

Imgf2 = {1, 3}

f −1 [B] = {a, b}

Also f3 is not injective and not surjective.


4. f4 = {(a, 1), (b, 3)} is not a function but a partial function.

Imgf4 = {1, 3}

f −1 [B] = {a, b}

Also f4 is not injective and not surjective.


5. f5 = {(a, 1), (b, 3), (c, 1)} is not a function nor a partial function but a
relation.

Imgf5 = {1, 3}

f −1 [B] = {a, b, c}
Chapter 3. Relations and Functions 66

Also f5 is injective but and not surjective.


Note that it is not possible in this example to define a function that is injective
but not surjective or converse, because of the condition 2 in the definition of a
function.
Example 3.5.8 Let A = {a, b, c} and B = {1, 2}. It is not possible to define a
function f : A → B that is injective, since |A| > |B| .
Think how can you prove it in terms of set theory, although it turns out not be
possible to prove, so we make it as an axiom or principle.

Pigeonhole Principle
Let f : A → B be a function, where A and B are finite. If |A| > |B|, then f
cannot be an injective function.
Use Pigeonhole Principle to following theorem
Theorem 3.5.9 Let A and B be finite sets, and let f : A → B.
1. If f is one-to-one, then |A| ≤ |B|.
2. If f is onto, then |A| ≥ |B|.
3. If f is a bijection, then |A| = |B|.
Definition 3.5.10 Let A be a set. We define the function idA : A → A by the rule

idA (x) = x ∀x ∈ A.

ιA is called the identity function.


Sometimes 1A or ιA are also used instead of idA to indicate the identity function
on A.
Definition 3.5.11 If f : A → B and g : C → A then the rule defined by

f ◦ g(a) = f (g(a)) ∀a ∈ A

defines a function f ◦ g: C → B. This function is called the composition of f


and g.
Theorem 3.5.12 (Associativity) Let f : A → B, g : B → C and h : C → D be
functions.
Then
h ◦ (g ◦ f ) = (h ◦ g) ◦ f
67 3.5. Functions

Proof Let a ∈ A. Then


(h ◦ (g ◦ f ))(a) = h((g ◦ f )(a)) = h(g(f (a))) = (h ◦ g)(f (a)) = ((h ◦ g) ◦ f )(a)

The associativity rule can be extended to any finite arbitrary function. We can
calculate
f1 ◦ (f2 ◦ (. . . (fn−1 ◦ fn )))
in any oreder for example,
f3 ◦ f1 ◦ f2 ◦ . . . fn ◦ fn−1
Definition 3.5.13 A left inverse g (if it exists) to a function f : A → B is a
function f : B → A such that g ◦ f = idA .
A right inverse g (if it exists) to a function f : A → B is a function g : B → A
such that f ◦ g = idB ,
Theorem 3.5.14 A function f : A → B has

(i) a left inverse g : B → A ⇐⇒ it is injective .


(ii) a right inverse g : B → A ⇐⇒ it is surjective.
(iii) an inverse ⇐⇒ it is bijective.
Proof (i) ⇒ First, assume that there is such a g. To proof f is injective, suppose
f (x) = f (y). Then
g(f (x)) = g(f (y)) ⇒ x = y Therefore, f is injective.
⇐ Assume that f is injective. ∀y ∈ Im f ∃!x s.t. f (x) = y
Define g : B → A as follows:
(
f −1 (y) y ∈ Im (f )
g (y) =
a y∈/ Im (f )
g ◦ f = x = idA .
Proof (ii) & (iii) as an exsersice.
Hint: Use the seconed form of the Axiom of choice, That is
For any relation R there is a function H ⊆ R with Dom H = Dom R
Corollary 3.5.15 From previous definition and theorem it turns out that left and
right inverses dosen’t have to be unique, but if f : A → B has an inverse, it is
unique. In this case f is called an invertible function and it’s inverse denoted by
f −1 .
Chapter 3. Relations and Functions 68

Theorem 3.5.16 If f : A → B is a bijection then the rule

∀b ∈ B f −1 (b) = a ⇐⇒ f (a) = b
defines a function f −1 : B → A. The function f −1 is a bijection itself and satisfies

f ◦ f −1 = idB and f −1 ◦ f =idA


The function f −1 defined in the above theorem is called the inverse of f .
Proof It is straightforward.
Theorem 3.5.17 Let f : A → B and g : B → C.
(i) If f and g are injective then g ◦ f : A → C is injective.

(ii) If f and g are serjective then g ◦ f : A → C is serjective.

(iii) If f and g are bijective then g ◦ f : A → C is also bijective. bijective.

Proof 1. Let f : A → B and g : B → C be injections. Let a1 , a2 ∈ A such that


g ◦ f (a1 ) = g ◦ f (a2 ) .
g is injective f is injective
Then g(f (a1 )) = g(f (a2 )) ======⇒ f (a1 ) = f (a2 ) ======⇒ a1 = a2
2. Let f : A → B and g : B → C be surjections. Let c ∈ C.
Since g is serjective, there exists b ∈ B such that g(b) = c.
Since f is serjective, there exists a ∈ A such that f (a) = b. By definition of a
composition of function

g ◦ f (a) = g(f (a)) = g(b) = c

3. It is a consequence of 1 & 2.

3.6 Set of all functions

Let A = {a, b, c} and B = {1, 2} .

A × B = {(a, 1), (a, 2), (b, 1), (b, 2), (c, 1), (c, 2)}
69 3.6. Set of all functions

List the set of all functions from A to B.


f1 = {(a, 1), (b, 1), (c, 1)} is a function.
f2 = {(a, 2), (b, 1), (c, 1)} is a function.
f3 = {(a, 1), (b, 1), (c, 2)} is a function.
f4 = {(a, 1), (b, 2), (c, 2)} is a function.
f5 = {(a, 2), (b, 1), (c, 2)} is a function.
f6 = {(a, 1), (b, 2), (c, 1)} is a function.
f7 = {(a, 2), (b, 2), (c, 1)} is a function.
f8 = {(a, 2), (b, 2), (c, 2)} is a function.

The set of all functions f : A → B has 23 = 8 elements.


Now list the set of all functions from B to A. The set of all functions f : B → A
has 32 = 9 elements.
An ordered pair x = (x1 , x2 ) ∈ R2 , is a function x : {1, 2} → R.
An n-tuple x = (x1 , x2 , . . . , xn ) ∈ Rn is a function x : {1, 2, . . . , n} → R.
Theorem 3.6.1 Let A and B be two set. The set of all functions f : A → B is
denoted B A or Hom(A, B) or F un(A, B) has |B A | elements.
Proof Suppose |A| = m and |B| = n . Each element of A has n choice to be
mapped to. since each element has n choice the number of all functions from A to
B is n × n × · · · × n = nm = |B||A| .
Chapter 3. Relations and Functions 70

Exercise

1. Let n ≥ 2 be an integer. Define a relation R on by a R b if and only if a2 ≡ b2


(mod n).
(a) Prove that R is an equivalence relation when n = 4 and determine the
distinct equivalence classes.
(b) In general, under what conditions on n is R an equivalence relation?
(c) When R is an equivalence relation, what are the equivalence classes?
2. Let R be a relation from a set A to a set B.
(a) Prove that a relation R on a set A is symmetric if and only if R = R−1 .
(b) Prove that a relation R on a set A is antisymmetric if and only if R ∩ R−1 is
a subset of the diagonal relation ∆ = {(a, a) | a ∈ A}.
3. Let A, B, C, D be sets. Then:
(a) (A × B) ∩ (C × D) = (A ∩ C) × (B ∩ D)
(b) (A × B) ∪ (C × D) ⊆ (A ∪ C) × (B ∪ D)
Consider the case (b) when the sets are empty.
4. Let A = {1, 2, 3}. Define a relation on A that is:
(a) reflexive, not symmetric, not transitive

(b) not reflexive, symmetric, not transitive

(c) not reflexive, not symmetric, transitive

(d) not reflexive, symmetric, transitive

(e) reflexive, not symmetric, transitive

(f) reflexive, symmetric, not transitive

(g) not reflexive, not symmetric, not transitive


71 3.6. Set of all functions

(h) reflexive, symmetric, transitive.


5. Define R on N by mRn ⇐⇒ 3|m + n. Prove that R is not an equivalence relation
on N.
6. Let F be the collection of all finite sets. Define ' on F by: X ' Y if and only if
there is a bijection f : X → Y
(a) Prove that ' is an equivalence relation on F .

(b) What are the equivalence classes of the sets , {}, {1, 2, 3}?

(c) What is the equivalence class of the set {a1 , . . . , an } for any distinct objects
a1 , , a n ?
7. Let f, g be functions.
(a) Show that f ∩ g is a function.
(b) Show that f ∪ g is a function if and only if

f (x) = g(x) ∀x ∈ (Domf ) ∩ (Domg)


Chapter 3. Relations and Functions 72
Part II

Group Theory

73
75

Evariste Galois (1811–1832)

Galois was a tragic and romantic figure who died in a duel at age twenty. He
was also one of the foremost mathematicians of all time. In his very brief life he
created one of the great edifices of mathematics—Galois Theory—of fundamental
importance to this day. He was born October 25, 1811, in Bourg-la-Reine, a village
near Paris. His father was a progressive thinker who headed the liberal party in the
town. He was elected mayor in 1815, the year in which Napoleon returned from
exile on the island of Elba and took control for the period known as the Hundred
Days. (Later in the year Napolean was exiled by the British to St. Helena.) This
was the year the monarchy was restored, and, unlike in the eighteenth century,
came to accept a Charter confirming most of the gains of the French Revolution.
Galois’ mother came from a family of jurists and received an education in the
classics. She was his sole teacher for the first 12 years of his life, stressing the study
of Greek and Latin. There is no evidence she taught him any mathematics beyond
rudimentary arithmetic. But these were happy years for Galois, with no hint of the
troubled times to come. His formal education began in 1823, when he enrolled
in the Collège Royal de Louis-le-Grand, a Paris preparatory school (the alma
mater of Robespierre and Hugo). It was at this time that he acquired his political
consciousness. During the first term the students rebelled and refused to take part
in the required religious observances. Scores were expelled for their disobedience.
Galois was not among them, but the severity and apparent arbitrariness of the
action made a deep impression on him. Galois’ first two years at Louis-le-Grand
were academically successful. He won several prizes in Greek and Latin and a
number of honorable mentions. During his third year his work in rhetoric was poor
and he had to repeat the year. Following this reversal he enrolled, at age fifteen, in
his first mathematics course. This awakened his mathematical talent. The standard
76

mathematics texts were no challenge. He soon came across Legendre’s Elements


of Geometry and Lagrange’s “The resolution of algebraic equations” and Theory
of Analytic Functions. These fired the young Galois’ imagination. Undoubtedly he
was influenced in his subsequent work on Galois Theory by Lagrange’s important
paper “The resolution of algebraic equations”. He later also read Abel’s work on
the subject. After these encounters with masterful mathematical works he seems
to have lost all interest in his normal classes at the school. There soon followed a
series of events which proved to be traumatic for the young Galois and soured him
on authority. In 1828, at age sixteen, he applied, a year earlier than normal, to the
very prestigious Ecole Polytechnique. But he failed the competitive entry exams.
He blamed the failure on the ignorance of the examiners, but the most likely reason
was his lack of preparation and communication skills. That year his father, whom
he loved dearly, committed suicide as a result of persecution by authorities for his
liberal views.
Galois began to make fundamental breakthroughs in the study of solvability of
equations, and in early 1829 submitted a paper on the topic to the French Academy
of Sciences. The referee was Cauchy, who did not present the paper to the Academy.
Many sources have claimed that he lost the paper, but recent research has found
otherwise. Here is Cauchy: I was supposed to present to the Academy . . . a report
on the work of the young Galois . . .. Am indisposed at home. I regret not to be
able to attend today’s session and I would like you to schedule me for the following
session.
Chapter 4

Algebraic Structure *

I think there is some change. If you went back to the 19th cen-
“ tury or earlier, mathematicians and physicists tended to be the
same people. But in the 20th century, mathematics became much
broader and in many ways much more abstract. What has hap-
pened in the last 20 years or so is that some areas of mathematics
that seemed to be so abstract that they were no longer connected
with physics instead turn out to be related to the new quantum
physics, the quantum gauge theories, and especially the supersym-
metric theories and string theories that physicists are developing
now.

Edward Witten, Frontline, 2001



A non-empty set together with one or more binary operation +, ×, . . . that sat-
isfies certain property is called an algebraic structure. The set is called underlying
set and the properties are called axioms.

Simple structures: no binary operation


Set A : A single set without having any operation.

Group-like structures: one binary operation


• Magma or groupoid (M, ·):
M and a single binary operation over M .

77
Chapter 4. Algebraic Structure * 78

• Semigroup:

1. (M, ·) is a magma.

2. (g1 · g2 ) · g3 = g1 · (g2 · g3 ) ∀g1 , g2 , g3 ∈ M

• Monoid (M, ·):

1. g1 · g2 ∈ M ∀g1 , g2 ∈ M

2. (g1 · g2 ) · g3 = g1 · (g2 · g3 ) ∀g1 , g2 , g3 ∈ M

3. e · g = g · e = g ∀g, e ∈ M ,

• Group (G, ·):

1. (G, ·) is a monoid

2. g · g −1 = g −1 · g = e ∀g, g −1 ∈ G

• Abelian (commutative) group (G, +):


g1 + g2 = g2 + g1

Ring-like structures or Ringoids: two binary operations

• Ring (R, +, ·) :

1. (R, +) is an Abelian group

2. (R, ·) is a monoid

3. r1 · (r2 + r3 ) = r1 · r2 + r1 · r3 , (r1 + r2 ) · r3 = r1 · r3 + r2 · r3

• Division ring:

1. 1 6= 0

2. (R \ {0}, ·) is a group

• Field:
R is a division ring with commutative multiplication.
79

Module-like structures: Systems involving two sets and with at least two binary
operations.
• Module(M, R, +) or M over the ring R (R-module):
1. (M, +) is an Abelian group
2. ∀n, m ∈ R & ∀a, b ∈ M, na ∈ M
(i) n(a + b) = na + nb
(ii) (n + m)a = na + ma
(iii) n(ma) = (nm)a
R itself is a special (one-dimensional) module over R.

• Vector space over R (R-vector space):


R is a field. R itself is a special (one-dimensional) vector space over R.
• Algebra (A, R, +, ·) or A (R-algebra):
1. R is a commutative ring (with unity),
2. (A, R, +) is a module
3. (A, ·) is a monoid, and ∀a, b, c ∈ A and ∀n ∈ R
(i) a · (b + c) = a · b + a · c (a + b) · c = ac + b · c
(ii) n(a · b) = (na) · b = a · (nb)
R itself is a special (one-dimensional) algebra over R.
Chapter 4. Algebraic Structure * 80
Chapter 5

Introduction to groups

Mathematicians do not study objects, but relations between objects.


“ Thus, they are free to replace some objects by others so long as the
relations remain unchanged. Content to them is irrelevant: they
are interested in form only.

Henri Poincar,

In this chapter, along the way of our journey in mathematical structures we
come up with one of the first and maybe for most important algebraic structure
called Group.
It is hard to say when groups first appeared in mathematics since its application
were used long before we even give an abstract definition of a group.
Euler (1761) and Gauss (1801) studied modular arithmetic, and Lagrange (1770)
and Cauchy (1815) studied groups of permutations. Important moves towards a
more formal, abstract theory were taken by Cauchy (1845), von Dyck (1882) and
Burnside (1897). But it came to work magicaly as we shall see.
The story all begins with the pioneer work of young French mathematician
Evariste Galois, who was working on insolvability of quintic equations. He died in
May 31, 1832 at the young age of 20 in a duel over a whore girl. Before he dies he
invented a language called Group Theory which is seems to be a good candidate to
describe symmetry in mathematical structures. The notion of symmetry is universal
,all mankind have a universal understanding of the notion of symmetry. Moreover,
every living creature knows what symmetry is. Even subatomic particles treatment
correspond with symmetries of mathematical structures.

81
Chapter 5. Introduction to groups 82

Groups are seems to be a good candidate to measuring the symmetry.


Symmetry appears everywhere in the world, from atomic scale ( the quantum
mechanics world) to galaxy (the relativistic world), So its mathematical co-exist
ought to be very important .
Abstract Algebra is abstraction of our thought , it translate informal notions
into its own magical language ( which is build up from basic logic and set theory)
and gives us new results which does not come up to our mind without it. As I
mentioned language of Abstract Algebra is completely based on logic and set
theory, so everything must be converted to its language.
The nature of symmetry is different from numbers or even object itself, yes
we can associate a scale to a object for its number of symmetries, but in nature
symmetry is not a number. Symmetry is more like a kind of transformation which
can be applied to an object. If after the transformation the object looks the same as
before that transformation is considered as a symmetry of the object. For example,
one symmetry of a square is rotation by π2 . After rotating by π2 the square looks the
same as before. In set theoretic language this is a one-to-one and onto function (a
bijection) from the set of vertices of square to itself. Another symmetry is rotation
by π, it is a bijective function again. The set of all bijections from a set to itself, i.e.

Hom(A) = {f | f : A → A}

has some properties :


1. The composition of functions is closed. ◦ : A × A → A.
2. The composition of functions is associative.
3. The composition of functions contains an identity element which does noting.
4. The composition of functions contains an inverse for each transformation which
gets back to the identity.
This Motivates the formal definition of a Group, but first we need some prelimi-
naries.

5.1 Binery operation

Binery operation is a rule which takes two elements in the set A, combine them
and give another element from the same set A.
83 5.1. Binery operation

Definition 5.1.1 A binary operation ∗ on a set A is a mapping or function ∗ :


A × A → A. That is, it sends elements of the Cartesian product A × A to A to A
If a, b, c ∈ A according to Axiom of Pairing (a, b) ∈ A × A exists. Instead of using
the functional notation ∗(a, b) = c, we write a ∗ b = c
Binary operations are usually denoted by special symbols such as

+, −, ., ×, ◦, ∪, ∩, ∨, ∧

By the definition of function a binary operation is a triple (A × A, A, ∗) but as


is usual in mathematics we write the binary operation ∗ instead of the binary
operation (A × A, A, ∗)
Remark In the definition of binary operation it is not necessary to point out the set
must be non-empty. Beacuse a binary operation on A is a function ∗ : A × A → A.
If A = ∅ we have ∅ × ∅ = ∅. so a binary operation on ∅ is a function f : ∅ → ∅ .
There is exactly one such function, the empty function.
The notion of operation is a generalization of function. In general, a function
f A : A × A × · · · × A → A or f A : An → A is called an n-ary operation and n is
| {z }
n times
called the order of the nullary or its its arity. It is called nullary if n = 0, unary if
n = 1 (it is a function), binary if n = 2, ternary if n = 3.
In the case of nullary n = 0, since A0 is a singleton like {∅}.

A0 = {∅ : ∅ → A}

Since a binary operation is a function, it has to obey it’s definition. The following
lemma explains
Lemma 5.1.2 Let ∗ be a binary operation on a set A. This must satisfy the
following conditions:
(a) a ∈ A and b ∈ A =⇒ a ∗ b ∈ A. [A is closed under ∗.]
(b) For all a, b, c, d in A
a = c and b = d =⇒ a ∗ b = c ∗ d. [Substitution is permissible.]
(c) For all a, b, c, d in A
a = b =⇒ a ∗ c = b ∗ c. [Multiplication both side on the right.]
(d) For all a, b, c, d in A
c = d =⇒ a ∗ c = a ∗ d. [Multiplication both side on the left.]
Proof multiply both sides of an equation on the right by the the same element
Chapter 5. Introduction to groups 84

As we mentined if a, b ∈ A according to Axiom of Pairing (a, b) ∈ A × A


exists. Since ∗ is a function ∗ : A × A → A it assgns to (a, b) an element a ∗ b ∈ A

∗(a, b) = a ∗ b ∈ A

This proofs (a).


For ∗ to be a function, it must satisfy well-defined condition

x = y =⇒ ∗(x) = ∗(y)
and a function acts on ordered pairs, (a, b) where a, b ∈ A which are a subset of
the Cartesian product A × A. Equality of ordered pairs is defined by the rule

a = c ∧ b = d ⇐⇒ (a, b) = (c, d)

To satisfy well-defined condition we must have

(a, b) = (c, d) =⇒ a ∗ b = c ∗ d

According to equality of ordered pairs we have

a = c and b = d =⇒ a ∗ b = c ∗ d.

This Proofs (b).


To prove (c) Take a = b and by reflexivity of equality we have c = c. Substitu-
tion in part (b) we get a ∗ c = b ∗ c.
To prove (d) Take c = d and by reflexivity of equality we have a = a. Substitu-
tion in part (b) we get a ∗ c = a ∗ d.
Binary operations are usually denoted by symbols such as

+, +n , −, ., .n , ×, ◦, ·, ∪, ∩, ∨, ∧, ]

When we say A is closed under operation ∗ we mean ∗(a, b) = a ∗ b is on the


same set A.
Let A be a set and f be a function
Definition 5.1.3 Let ∗ be a binary operation on a set A.
85 5.1. Binery operation

1. ∗ is Commutative if
a ∗ b = b ∗ a ∀a, b ∈ A

2. ∗ is Anti-commutative if

a ∗ b = −b ∗ a ∀a, b ∈ A

3. ∗ is Associative if

a ∗ (b ∗ c) = (a ∗ b) ∗ c ∀a, b, c ∈ A.

4. ∗ satisfies Jacobi identity

a ∗ (b ∗ c) + c ∗ (a ∗ b) + b ∗ (c ∗ a) = 0 ∀a, b, c ∈ A.

As we mentioned a binery operation is a function that is defined on a set, so


it does not make sense to call a single operation ∗ is a binary operation without
specifying underling set.
The most familiar example of binary operations are ordinary addition and
multiplication.
Example 5.1.4
1. Addition is a binary operation on N, Z, Zn , Q, R and C.
Let + be the addition operation on Z.
+ : Z × Z 7−→ Z defined by +(a, b) = a + b
2. Multiplication is a binary operation on N, Z, Zn , Q, R and C.
× : Q × Q 7−→ Q defined by ×(a, b) = a + b
3. Subtraction is a binary operation on Z, Q, R and C. But subtraction is not a binary
operation on N since, for example, 7 − 5 ∈ N but 5 − 7 ∈ / N.
4. Addition and multiplication are both associative and commutative operations on N,
Z, Zn , Q, R and C but Subtraction is niether associative nor commutative .
5. Division is not a binary operation on R because 1, 0 ∈ R but 10 ∈ / R. Thus R is not
closed under division. If ∗ is not a function, but is a partial function instead, it is
called a partial binary operation. Here division is a partial binary operation. In
other word division is a binary operation on R∗ = R \ {0}.
But addition is no longer a binary operation on R∗ ; because a, (−a) ∈ R∗ but
a + (−a) = 0 ∈ / R∗ . So it does not worth to count division as a binary operation.
Chapter 5. Introduction to groups 86

6. Let Mn (K) be the set of all n × n matrices with entries from K. And let K denote
each one of the following: Z, Zn , Q, R and C.
We denote (i, j) entry of an n × n matrix A by aij ,
Matrix addition and multiplication are defined by the following rules

     
a11 a12 · · · a1m b11 b12 · · · b1m a11 + b11 a12 + b12 · · · a1m + b1m
 a21 a22 · · · a2m   b21 b22 · · · b2m   a21 + b21 a22 + b22 · · · a2m + b2m 
..  +  .. ..  =
     
 .. .. ... .. ... .. .. ... .. 
 . . .   . . .   . . . 
an1 an2 · · · anm bn1 bn2 · · · bnm an1 + bn1 an2 + bn2 · · · anm + bnm

     
a11 a12 · · · a1m b11 b12 · · · b1m a11 b11 a12 b12 · · · a1m b1m
 a21 a22 · · · a2m   b21 b22 · · · b2m   a21 b21 a22 b22 · · · a2m b2m 
.. ×  .. ..  =  ..
     
 .. .. ... .. .. .. .. .. 
 . . .   . . . .   . . . . 
an1 an2 · · · anm bn1 bn2 · · · bnm an1 bn1 an2 bn2 · · · anm bnm

where ai , bj ∈ K. Matrix addition and multiplication is a binary operation on


Mn (K).

7. Recall the power set of the set A; that is, the set of all subsets of A denoted by
P (A). If A1 , A2 are subsets of A, then A1 A2 , A1 ∩ A2 and A1 \ A2 are also
subsets of A. Therefore, union, intersection and set theoretic difference are binary
operations on the set P (A). Moreover, ∪, ∩ are both commutative and associative.
V W
8. Let P be the set of all logical propositions. Then and are binary operation on
P.
9. Recall the set of all functions f : A → A denoted by AA or Hom(A, A) or
F un(A, A) or simply F un(A) .
F un(A) = {f | f : A → A}
Define a operations on F un(A) by composition of functions ◦. This is a binary
operation because given functions f : A → A and g : A → A, i.e. f, g ∈ F un(A)
their composition f ◦ g is also a function f ◦ g : A → A.

f /
A A
g
g◦f  
A
87 5.1. Binery operation

Moreover, composition of functions ◦ is associative but not commutative .


10. The usual addition of vectors in Euclidean n-space as Rn , n ∈ N.

Rn = {(x1 , x2 , . . . , xn ) | xi ∈ R ∀i}.

Thus R2 is the set of vectors in the plane, and R3 is the set of vectors in the space.
Componentwise addition is defined by the rule

(x1 , x2 , . . . , xn ) + (y1 , y2 , . . . , yn ) = (x1 + y1 , x2 + y2 , . . . , xn + yn ).

is a binary operation since (x1 + y1 , x2 + y2 , . . . , xn + yn ) ∈ Rn .


But scalar multiplication is not a binary operation, because if λ ∈ R and
x = (x1 , x2 , . . . , xn ) ∈ Rn scalar multiplication is defined as

λx = (λx1 , λx2 , . . . , λxn )

Despite of the fact that λx ∈ Rn but here we do not combine two elements of Rn
so it is not a binary operation.
11. The dot product X · Y of vectors X and Y in Rn is not a binary operation. If

X = (x1 , x2 , . . . , xn )
Y = (y1 , y2 , . . . , yn )

their dot product is defined to be

X · Y = x 1 y 1 + x2 y 2 + · · · + xn y n

Notice that the result is not in Rn , so the dot product is not a binary operation.
12. The cross product A × B of vectors A and B in R3 is a binary operation since it
takes two vectors in R3 and produce another vector in R3 . Recall that if

A = (a1 , a2 , a3 )
B = (b1 , b2 , b3 )

then A × B is defined by the formula


 
a2 a3 a1 a3 a1 a2
A × B = ,−
b1 b3 , b 1 b2

b2 b3

However, the cross product of two vectors is not commutative a × b 6= b × a, but


anti-commutative A × B = −B × A or [A, B] = [B, A]
Chapter 5. Introduction to groups 88

it is not even associative A × (B × C) 6= (A × B) × C, but it satisfies Jacobi


identity

A × (B × C) + B × (C × A) + C × (A × B) = 0
or in closed form

[A, [B, C]] + [B, [C, A]] + [C, [A, B]] = 0

The set R3 = {(ai , bi , ci ) a, b, c ∈ R} equipped with the cross product(Lie


bracket) defined is a Lie algebra
Lie algebras are algebras satisfying anticommutativity and the Jacobi identity and
two additional property, called Bilinearity and Alternativity.
Definition 5.1.5 If A is a non-empty set with binary operation ∗, then an element
e ∈ A is called a
(a) right identity if a ∗ e = a ∀a ∈ A
(b) left identity if e ∗ a = a ∀a ∈ A
If e ∈ A is the right and left identity, e is called an identity element with respect to
∗. That is,
a ∗ e = a = e ∗ a ∀a ∈ A
Example 5.1.6
1. a  b = max(a, b).
e = 1 is an identity element: 1  a = a  1 = a ∀a ∈N
2. The identity element for addition and multiplication of Mn (K) is e = On (the
zero matrix) and e = In (the identity matrix) respectivly.
3. ◦ on F un(A) = id (the identity map defined by id(x) = x∀x ∈ A). Since, for
any function f : A → A we have f ◦ id = id ◦ f = f .
Definition 5.1.7 If A is a non-empty set with binary operation ∗ and identity
element e and a ∈ A. Then an element a−1 ∈ A is called a
1. right inverse if a ∗ a−1 = e ∀a ∈ A
2. left inverse if a−1 ∗ a = e ∀a ∈ A
If a−1 ∈ A is the right and left inverse, a−1 is called an (two-side)inverse with
respect to ∗. That is,
a ∗ a−1 = e = a−1 ∗ a ∀a ∈ A
89 5.1. Binery operation

a−1 is called an invertible element of a


In additive notation the inverse of an element is shown by −a and in multiplica-
tive notation by a−1
Example 5.1.8
1. The identity element for addition of Z is e = 0, since
0 + a = a = a + 0 and for multiplication in Z is e = 1, since
1.a = a = a.1 But the identity element for subtracttion on Z does not exist.
Beacuse we should have e − a = a = a − e but e − a = a ⇒ e = 2a and
a − e = a ⇒ e = 0 Therefore, the inverse is not unique, so it does not exists.
2. 0 is a right identity for subtraction, but subtraction has no left identity.
Theorem 5.1.9 The identity element on a set A with respect to binary operation ∗
(if it exists) is unique.
Proof Suppose e1 , e2 be identity elements with respect to ∗.
Definition 5.1.10 Let G be a set and ∗ a binary operation on G. A non-empty
subset H of G is said to be closed respect to ∗ if, for every pair (x, y) of elements
of H, also x ∗ y belongs to H. That is,

∀a, b ∈ H ⇒ a ∗ b ∈ H

Example 5.1.11 Recall Zn for the set of all congruences classes modulo n.

Zn = Z/nZ = {0̄, 1̄, . . . , n − 1} = {[0], [1], . . . , [n − 1]}

In previous chapters we introduced operations of addition and multiplication


for congruence classes.
Definition 5.1.12 Let a, b ∈ Zn . Take x, y ∈ Z such that a = [x], b = [y]. The
addition of two elements is defined as

a + b = [x] + [y] := [x + y]

or sometimes
a +n b = [x] +n [y] := [x + y]n
Also The multiplication of two elements is defined as

a.b = [x].[y] := [x.y]

or
a.n b = [x].n [y] := [x.y]n
Chapter 5. Introduction to groups 90

Every congruence class can be represented by any one of its elements, we have
to show the operations defined does not depend on representative of the class (It is
well-defined) .

Theorem 5.1.13 The addition and multiplication defined for congruence classes
is a binary operation .

+n : Zn × Zn → Zn where +n ([a], [b]) 7−→ [a +n b] and

.n : Zn × Zn → Zn where .n ([a], [b]) 7−→ [a.n b]

Proof Let x0 ∈ [x] and Let y 0 ∈ [y]. We show [x+y] = [x0 +y 0 ] and [x.y] = [x0 .y 0 ].
x0 = x + mk and y 0 = y + mt forsome k, t ∈ Z. Therefore, x0 + y 0 = (x + y) +
m(k + t) and x0 .y 0 = (x.y) + m(xt + yk + mkt).

So x0 + y 0 = x + y (mod m) and x0 .y 0 = x.y (mod m).

Thus, [x0 + y 0 ] = [x + y] and [x0 .y 0 ] = [x.y].

The operaton +n and .n are commutative and associative. 0 is additive identity


and 1 is multiplicative identity.

Every element of Zn is invertible with respect to +n

(
n − a if a 6= 0
−a =
0 if a = 0

5.2 Cayley Table

Recall the definition of a binary operation on a set A, it is a rule which ∀a, b ∈ A


produces a third element back in A again. So if the set A is finite, this is possible
to determine by means of a composition table, for any pair of elements of A what
the third element is.

Let A = {a1 , a2 , ..., an } be a finite set, and let ∗ be a binary operation on A.


The multiplication table of ∗ is a square table of size |A| × |A|.
91 5.2. Cayley Table

◦ a1 a2 a3 ... an
a1 a1 ◦ a1 a1 ◦ a2 a1 ◦ a3 ... a1 ◦ an
a2 a2 ◦ a1 a2 ◦ a2 a2 ◦ a3 ... a2 ◦ an
a3 a3 ◦ a1 a3 ◦ a2 a3 ◦ a3 ... a3 ◦ an
.. .. .. .. ..
. . . . .
an an ◦ a1 a3 ◦ a2 a3 ◦ a3 ... a3 ◦ an

This multiplication table usually called a Cayley table1 .


Remark A binary operation is commutative if and only if its multiplication table
is symmetric with respect to the main diagonal.
Example 5.2.1 The set Z8 consists of [0], [1], [2], [3], [4], [5], [6], [7]. Addition and
multiplication rule can be given by a table.

+8 0 1 2 3 4 5 6 7 .8 0 1 2 3 4 5 6 7
0 0 1 2 3 4 5 6 7 0 0 0 0 0 0 0 0 0
1 1 2 3 4 5 6 7 0 1 0 1 2 3 4 5 6 7
2 2 3 4 5 6 7 0 1 2 0 2 4 6 0 2 4 6
3 3 4 5 6 7 0 1 2 3 0 3 6 1 4 7 2 5
4 4 5 6 7 0 1 2 3 4 0 4 0 4 0 4 0 4
5 5 6 7 0 1 2 3 4 5 0 5 2 7 4 1 6 3
6 6 7 0 1 2 3 4 5 6 0 6 4 2 0 6 4 2
7 7 0 1 2 3 4 5 6 7 0 7 6 5 4 3 2 1

Number of binary operation on a set


Since a binary operation is a function f : A × A → A we can apply the rules of
counting to determine how many binary operations can be defined on a given set
with n element.
The number of binary operation on a set |A| = n f : A × A → A is |A||A×A| =
n2
n . Furthermore, we can also determine the number of commutative binary
operation on a given set.
Let A = {a1 , a2 , . . . , an } then recall the operation ∗ is said to be commutative
if
ai ∗ aj = aj ∗ ai ∀ai , aj ∈ A
1
In the honors of British mathematician Arthur Cayley 1821 - 1895.
Chapter 5. Introduction to groups 92
   
n n
In A × A there are n ordered pairs (ai , aj ) and (ai , aj ) and (aj , ai ), so
2 2
the number of all elements (ordered pairs) are
   
n n
n+ + = n2
2 2

Because we want commutative binary operations, that is ∗(ai , aj ) and ∗(a


 j, ai )
n
must map to the same value in the set. So we should throw away one of of
  2
n
ordered pairs. Now the restricted set A × A has caedinality n +
2
Therefore, commutative binary operations from restricted set A × A → A has
cardinality  
n
n+ 2
n 2 = n n 2+n

Corollary 5.2.2 More than half binary operations on a finite set are commutative.
Lemma 5.2.3 For a given set |A| = n
2 +1
(a) The number of operations containing an identity is n(n−1)
Proof Exersice
Unfortunetly, it appears that there is no closed formula for counting the number
of associative binary operations on a set.
Example 5.2.4 Let A = {a, b} be a with two elements. The number of different
2
binary operations on this set is 22 = 16. Those operations are:
∗ ∗ ∗ ∗ ∗
1) a a 2) a a 3) a a 4) a b 5) b a
a a a b b a a a a a
∗ ∗ ∗ ∗ ∗
6) a a 7) a b 8) b a 9) a b 10) b a
b b a b a b b a b a
∗ ∗ ∗ ∗ ∗
11) b b 12) a b 13) b a 14) b b 15) b b
a a b b b b a b b a

16) b b
b b
93 5.2. Cayley Table

Eight operations are commutative. See tables 1, 2, 7, 8, 9, 10, 15 and 16. Eight
operations are associative. See tables 1, 2, 4, 6, 7, 8, 10 and 16. Six operations
are both commutative and associative. See tables 1, 2, 7, 8, 10 and 16. For four
operations there exists an identity in S. See tables 2, 7, 8 and 10. For four operations
there exists a zero in S. See tables 1, 2, 8 and 16. For two operations there exist
both an identity and a zero in S. See tables 2 and 8.
Chapter 5. Introduction to groups 94

5.3 Definition and Examples of Groups

Definition 5.3.1 [Group]


A group is a pair (G, ∗) consisting of a (non-empty) set G together with a
defined binary operation ∗ that satisfies following axioms:
1. Associativity: ∀a, b, c ∈ G, a ∗ (b ∗ c) = (a ∗ b) ∗ c
2. Identity: ∃e ∈ G, ∀a ∈ G, such that e ∗ a = a = a ∗ e
3. Inverse: ∀a ∈ G, ∃a−1 ∈ G, such that a−1 ∗ a = e = a ∗ a−1
Notation Remark
1. If binary operation in the group is multiplication (G, .), for a, b ∈ G we write
a.b. In this case we denote the identity element by 1 and the inverse of a ∈ G with
a−1 .
2. If binary operation in the group is addition (G, +), for a, b ∈ G we write
a + b. In this case we denote the identity element by 0 and the inverse of a ∈ G
with −a.
3. For the sake of simplicity, when it is clear what the binary operation is, then
the group (G, ∗) may be referred to by its underlying set G alone and for a, b ∈ G
we write a.b or even simpler ab instead of a ∗ b.

Remark 1. Some books mention closure as first property, but since we defined binary
operation to be a function G × G → G, closure is direct consequence of a binary
operation.
2. The empty set does not admit a group structure because it does not contain the
identity element.
3. An identity element e ∈ G is a condition for all elements of the group since it
depends on the binary operation ∗ not on the elements, whereas an inverse element
a−1 ∈ G is defined relative to a single element of G.
4. The order of axioms 3 and 4 in the definition above is important, since it is
impossible to talk about an inverse of an element until existence of an identity is
unknown.
5. In the definition of a group G we do not require commutativity as the composition
of functions is not commutative.
95 5.4. Groups of modular arithmetic

Definition 5.3.2 A group G is called abelian2 (or commutative), if the group oper-
ation is commutative . That is,

a ∗ b = b ∗ a ∀a, b ∈ G

If the group we are considering is abelian we frequently use additive notation,


and for non-abelian group we use multiplicative notation. The reason for this dis-
tinction is that we are used to multiplication being non-commutative (multiplication
of matrices and of permutations are non-commutative).
Definition 5.3.3 The order of a group G, denoted by |G|, is the number of elements
in G.
If |G| is a finite number, then the group is a finite group. If |G| is infinite and if
its elements are labeled by n continuous real parameters, then G is a Continuous
group or Topological group with dimension n.
1. (Z, +), (Q, +), (R, +), (C, +), (Rn , +), (nZ, +) are examples of group
structures on familiar number systems. The inverse of x ∈ is −x. Lets check
(Z, +) for example,
(a) + is a binary operation on Z, i.e. + : Z × Z → Z.
(b) Numbers addition is associative a + (b + c) = (a + b) + c ∀a, b, c ∈ Z
(c) ∃ 0 ∈ Z satisfying 0 + n = n + 0 = n ∀n ∈ Z
(d) ∀n ∈ Z ∃ − n ∈ Z such that (−n) + n = n + (−n) = 0
2. (N, +) is not a group, since it does not contain inverse element, −n ∈
/ N.
3. Let R∗ = R \ {0} the set of all non-zero real numbers. (R∗ , .) is a group. Also as
(Q∗ , .),(C∗ , .) with the identity element being 1 and the inverse of x being x1
4.

5.4 Groups of modular arithmetic


Recall congruent class mod n

Zn = {0̄, 1̄, . . . , n − 1} = {[0], [1], . . . , [n − 1]}

Convention
2
In the honors of norwegian mathematician Niels Henrik Abel (1802-1829).
Chapter 5. Introduction to groups 96

1. For the sake of simplicity, I drop the overline for the element Zn =
{0, 1, . . . , n − 1} but we mean as before the set of equivalence classes not just
single elements.
2. Since it is clear, we may use just + and . instead of +n and .n respectively.

(Zn , +n ) forms a group, Because


1. + is a binary operation on Zn , i.e. + : Zn × Zn → Zn .
2. a + (b + c) = (a + b) + c ∀a, b, c ∈ Zn
3. ∃ 0 ∈ Z satisfying 0 + n = n + 0 = n ∀n ∈ Zn
4. ∀a ∈ Zn ∃(−a) = (n − a) ∈ Z such that (−a) + a = a + (−a) = 0
Zn is called the cyclic group of order n.

But (Zn , .n ) is not a group, Because for example 0 does not have a multiplicative
inverse.
The Cayley table of addition and multiplication for the set
Z8 = {0, 1, 2, 3, 4, 5, 6, 7} is as follows.

+8 0 1 2 3 4 5 6 7 .8 0 1 2 3 4 5 6 7
0 0 1 2 3 4 5 6 7 0 0 0 0 0 0 0 0 0
1 1 2 3 4 5 6 7 0 1 0 1 2 3 4 5 6 7
2 2 3 4 5 6 7 0 1 2 0 2 4 6 0 2 4 6
3 3 4 5 6 7 0 1 2 3 0 3 6 1 4 7 2 5
4 4 5 6 7 0 1 2 3 4 0 4 0 4 0 4 0 4
5 5 6 7 0 1 2 3 4 5 0 5 2 7 4 1 6 3
6 6 7 0 1 2 3 4 5 6 0 6 4 2 0 6 4 2
7 7 0 1 2 3 4 5 6 7 0 7 6 5 4 3 2 1

Only (Zn , +n ) is a group.


Here we use some terminology to make a multiplicative group out of Zn .
Definition 5.4.1 Let a ∈ Zn . a is called a unit if there is an element b ∈ Zn such
that
ab = 1 (mod n)

We denote the set of all units in Zn by Un . We call Un the group of units of


Zn .
97 5.5. Groups of permutation

Theorem 5.4.2 a ∈ Zn has a unit ⇐⇒ (a, n) = 1.


Proof ⇐= Suppose a ∈ Zn has a unit, so ak = 1 for some k ∈ Zn or ak = 1
(mod n). Therefore, ak − 1 = mn for some m ∈ Z. So, ak − mn = 1
This is a linear combination of a and n which gives 1. Therefore (a, n) = 1.
=⇒ Conversely, suppose (a, n) = 1. There exists k and m such that
ak + mn = 1. That is, ak = 1 (mod n) or

ak = 1 for some k ∈ Zn
Corollary 5.4.3 For n ≥ 2, Un = φ(n) = {a ∈ Zn : (a, n) = 1}.
In number theory the order of Un is obtained from, is called the Euler totient
function φ(n).
φ(n) is the of numbers which are relatively prime to n.
In particular, Up , where p is a prime is a group with p − 1 elements. namely

Up = {1, 2, . . . , p − 1}

Here, the group of units of Z8


.8 1 3 5 7
1 1 3 5 7
3 3 1 7 5
5 5 7 1 3
7 7 5 3 1

It is easy to check that each element of this group is self-inverse.

5.5 Groups of permutation

Next chapters

5.6 Basic Properties Of Groups

Theorem 5.6.1 Let G be a group and a, b, c, ∈ G. Then


Chapter 5. Introduction to groups 98

(i) ∃!e ∈ G .
(ii) ∀a ∈ G ∃!a−1 ∈ G .
(iii) If ab = ac then b = c. (left cancellation law)
(iv) If ba = ca then b = c. (right cancellation law)
Proof (i) Suppose e1 and e2 are identities of G.
e1 .a = a.e1 = a & e2 .a = a.e2 = a
Put e2 instead of a, we have

e1 .e2 = e1 = e2 .e1 = e2

(ii) Suppose a ∈ G has two inverses b, c ∈ G

b = be = b(ac) = (ba)c = ec = c

(iii) Suppose ab = ac.

a−1 (ab) = a−1 (ac)


By associative law,

(a−1 a)b = (a−1 a)c


eb = ec
b=c

(iv) The proof is similar to (iii)


Corollary 5.6.2
Theorem 5.6.3 [The Generalized Associative Law] Let ∗ be an associative binary
operation on a set A. If a1 , a2 , . . . , an is a sequence of n ≥ 3 elements of A, then
the product
a1 ∗ a2 ∗ · · · ∗ an
is unambiguous; that is, the same element will be obtained regardless of how
parentheses are inserted in the product (in a legal manner).
Proof The general case can be proved by induction on n. Do it as an excersice.
99 5.6. Basic Properties Of Groups

The Generalized Associative Law allows us to define integral powers of an


element a ∈ G inductively.

an = a × a × · · · × a
| {z }
n times

Definition 5.6.4 For any a ∈ G we define


 n 
 a for n > 0 
an = e for n = 0
 −1 n
(a ) for n < 0

In additive group

na = a {z· · · + a}
|+a+
n times

For any a ∈ G we define


 
 na for n > 0 
na = 0 for n = 0
−na for n < 0
 

Theorem 5.6.5 Let G be a group and a, b, c, ∈ G. Then


(i) ∀a ∈ G (a−1 )−1 = a
(ii) ∀a, b ∈ G (ab)−1 = b−1 a−1
Proof (i) If a ∈ G then since

aa−1 = a−1 a = e

a is an inverse of a−1 . Since inverses are unique then (a−1 )−1 = a.


(ii) Let x be the inverse of ab. Then (ab)x = e By associativity, we have
a(bx) = aa−1 . By left cancellation law, we have bx = a−1 . But

bx = ea−1 = b(b−1 a−1 )

using left cancellation again x = (b−1 a−1 ). Therefore, (ab)−1 = (b−1 a−1 ).
Proposition 5.6.6 If G is a group and a, b ∈ G each of the equations ax = b and
xa = b has a unique solution in G.
Chapter 5. Introduction to groups 100

Proof
x = ex = (a−1 a)x = a−1 (ax) = a−1 b
x = xe = x(aa−1 ) = (xa)(a−1 ) = ba−1

These are unique since inverses are unique.

a−1 b = ba−1

Corollary 5.6.7 If G is a group and a, b ∈ G. if either ab = e or ba = e then


b = a−1 .
Theorem 5.6.8 [Laws of Exponents for Groups] Let G be a group with identity e.
Then ∀a, b ∈ G and n, m ∈ Z we have
(i) an am = an+m
(ii) (an )−1 = a−n
(iii) (an )m = anm
(iv) (an )m = anm
(v) (ab)n = an bn if ab = ba
Proof Exersice
According to these facts, we have an improtant result about Cayley table (multi-
plication table) of a finite group.
Lemma 5.6.9 Let G = {a1 , a2 , . . . , an } be a finite group. Then in any row and
column of the multiplication table of G, each element of G appears exactly once.
Assume that the elements a1 , a2 , . . . , an of G are distinct. The i − th row of the
multiplication table is
ai a1 , ai a2 , . . . , ai an
All these elements are distinct as listed. Indeed, if ai aj = ai ak for j 6= k, then
the Cancelation Lemma implies that aj = ak contradicting the assumption that
aj 6= ak for j 6= k. Hence the n elements are all distinct, so each element of must
appear, and it can only appear once.
Proof Exersice
101 5.6. Basic Properties Of Groups

Niels Henrik Abel (1802-1829)

Was a Norwegian mathematician who was born on Augusts’5 1802 and Died 6
April 1829. He entered the cathedral school at the age of 13. Anew mathematics
teacher, Bront Michael Holmloe saw his talent in mathematics and encouraged
him to study the subject to advanced level. Bernt Holmboe supported Abel with
a scholarship to remain at school and raised many from his friends to enable him
to study at the Royal Fredrick University. Abel started working on the quintic
equation in radicals At the age of 19, he solved a problem that had vexed leading
mathematicians for hundreds of year. He proved that, unlike the situation for
equations of degree 4 or less, there is no finite formula for the solution of the
general fifth degree equation. This question had been unresolved for 250 years.
Most of his work was done in six or seven years of his working life. Abel thought
that he had solved the problem and submitted his work for publication. Unable to
find an error and understand his arguments he was asked by the editor to illustrate
his method. In 1824, during the process of illustration he discovered an error. This
discovery led Abel to a proof that no such solution exists. He also worked on
elliptic functions and essence revolutionized the theory of elliptic functions. He
traveled to Paris and Berlin in order to find a teaching position .Then poverty took
its toll, and Abel died from tuberculosis on April, 1829. Two days later a letter
from Crelle reached his address, conveying the news of his appointment to the
professor ship of mathematics at the University of Berlin. Abel is honored by such
terms as Abelian group and Abelian function.
Chapter 5. Introduction to groups 102
Chapter 6

Subgroup And Cyclic group

By a small sample we may judge of the whole piece.


“ ”
Miguel De Cervantes, New Scientist

6.1 Subgroup

Despite of the basic definiton of a group, it has a very reach structure. Recognition
of these structures help us to recognize the entire group. We are interested in
breaking up a group into smaller peaces which have the same structure as a group,
and instead of studying the whole group consider those smaller peaces. For some
reason as I mentioned we restrict those peaces to be groups themselves.
Definition 6.1.1 Let (G, ∗) be a group. H is a Subgroup of a group (G, ∗) if
H ⊆ G, and (H, ∗) is a group.
We usually write to denote that H is a subgroup of G.
If H ≤ G and {e} ⊂ H ⊂ G we say that H is a proper subgroup of G and we
write H < G.
Remark Note that form the definition it arise that the subset H is a subgroup of
G with respect to the same operation that makes G a group. For example Zn ⊂ Z
but Zn  Z .
The trivial subgroup e and G are always subgroups of G. Because according

103
Chapter 6. Subgroup And Cyclic group 104

to axiom of extensionality every set is subset of itself so G ≤ G and since G is a


group, it is a subgroup of itself.
A subgroup is therefore a group which sits inside the original group.
Since (H, ∗) is a subgroup it must satisfy group properties. That is,
1. Closure: ∀a, b ∈ H ab ∈ H.
2. Associativity: a(bc) = (ab)c
3. Identity : eH = eG
4. Inverse: ∀a ∈ H, ∃a−1 ∈ H, such that, a ∗ a−1 = a−1 ∗ a = eH
But we do not need to check all of them. Only two condition 1 and 4 are
sufficient. Because :
1.If a set is associativity every subset of it is associativity again.
2.If a ∈ H and a−1 ∈ H then aa−1 = a−1 a = e ∈ H
Lets make it as a theorem.
Theorem 6.1.2 Let G be a group. H ⊆ G is a group if and only if
(i) ∀a, b ∈ H ab ∈ H.
(ii) ∀a ∈ H a−1 ∈ H.
Proof

=⇒ If H is a subgroup then obviously (i), (ii) are hold.


⇐= Assume (i), (ii) are hold. From (i), (ii) and substituting b = a−1 we can conclude
e ∈ H. So H is a subgroup.
This theorem holds in general group whether finite or infinite. But in finite case
it is just enough to check Closure.
Theorem 6.1.3 Let G be a finite group. H ⊆ G is a group if and only if
(i) ∀a, b ∈ H ab ∈ H.
Proof

=⇒ If H is a subgroup then by definition it is closed.


⇐= We are assuming that H is closed, so it needs to contain at least an element.
If H only contains e, trivially H is a subgroup.
If H contains an element a 6= e. Since H is closed, the elements
105 6.1. Subgroup

a2 = a.a, a3 = a.a.a. . . , an = a.a . . . a are in H.


Since G is finite, the powers of a must coincide somewhere, so for some
1 ≤ m < n we have

am = an
an−m = e

Therefore, e ∈ H. We also have an−m−1 = a−1 ∈ H.


Hence, H is a subgroup of G.
Corollary 6.1.4 (The Subgroup Criterion) Let G be a group and let H be a
nonempty subset of G. Then

H ≤ G ⇐⇒ ab−1 ∈ H ∀a, b ∈ H

Proof =⇒ Suppose H ≤ G, and a, b ∈ H. Then b−1 ∈ H, so ab−1 ∈ H since


H is closed.
⇐= Suppose ab−1 ∈ H ∀a, b ∈ H. Since H 6= ∅ take a = b.
Then e = aa−1 ∈ H. Since e ∈ H, for any a ∈ H, a−1 = ea−1 ∈ H.
Now if a, b ∈ H, then b−1 ∈ H, therefore ab = a(b−1 )−1 ∈ H. Hence H ≤ G
Example 6.1.5
1. The set 2Z = (Even integers, +) is a subgroup of (Z, +). Because
(a) Closure : ∀a, b ∈ 2Z a = 2n, b = 2m for some n, m ∈ Z.
(b) Inverse : ∀a ∈ 2Z a = 2n for some n ∈ Z, and − a = −2n = 2(−n) ∈ 2Z.
Therefore, (2Z, +) < (Z, +).
2. In general nZ = {na : a ∈ Z} is a subgroup of Z. Both are subgroups of (Q, +).

nZ ≤ Z ≤ Q

3. The subset H = {0, 2} is the only proper subgroup of Z4 , because it is closed

0 +n 0 = 0, 0 +n 2 = 2 +n 0 = 2, 2 +n 2 = 0

and contains inverses


0 +n 0 = 2 +n 2 = 0
Chapter 6. Subgroup And Cyclic group 106

If H is any other subgroup that contains another element, say 1, it must contains
1 +n 1 = 2, 1 +n 1 +n 1 = 3, 1 +n 1 +n 1 + n 1 = 0
so H is includes entire group, That is H = Z4 .

Z4

{0}

Definition 6.1.6 If G is a group and S is the set of all subgroups of it, then (G, ⊆)
is called a subgroup lattice. The largest subgroup is at the top, the smallest is at the
bottom, and the relationship between two subgroups K ⊆ H given by a vertical
line.
Lattice of a subgroup and Cayley Diagrams are the best way of visualizing a
group.
Example: H = ({0, 2, 4}, +6 ) and K = ({0, 3}, +6 ) are subgroups of (Z6 , +6 ).
Z6

H K

{0}

Example 6.1.7 Draw the subgroup lattice of (Z16 , +16 ). The subgroups of Z16 are
h0i, h2i, h4i, h8i, h1i = Z16
The only chain is h0i < h8i < h4i < h2i < h1i.
The lattice of subgroups is :

h1i

h2i

h4i

h8i

h0i
107 6.2. Cyclic Group

The union of subgroups is not necessarily a subgroup, but the intersection of


subgroups is always a subgroup.
Example 6.1.8
1. 2Z ≤ Z and 3Z ≤ Z but 2Z ∪ 3Z
Z since, 2 + 3 = 5
2Z ∪ 3Z .
2. Consider (R2 , +) which forms a group. The following sets are subgroups of R2 .

A = {(a, 0)|a ∈ R} and B = {(0, b)|b ∈ R}

But A ∪ B is not a subgroup because it’s not closed under addition. That is,
(a, 0) ∈ A and (0, b) ∈ B, but

(a, 0) + (0, b) = (a, b) ∈


/ A∪B

Theorem 6.1.9 The intersection of any arbitrary collection of subgroups of a


group is again a subgroup.
ProofTLet{Hi }i∈I be an be an arbitrary collection of subgroups of a group G. We
show i∈I Hi is also a subgroup of G.
T T T
Since ∀i ∈ I, eT∈ Hi , we have e ∈ i∈I Hi . So i∈I Hi 6= ∅ and i∈I Hi ⊆ G.
Now let a, b ∈ i∈I Hi ,so we have a, b ∈ Hi for all i ∈ I. Then ab−1 ∈ Hi for all
i ∈ I, since each Ha is a subgroup.
This shows that ab−1 ∈ Hi for all i ∈ I. From the subgroup criterion
\
Hi ≤ G
i∈I

6.2 Cyclic Group

There is a certain kind of group which arise when we collect the set of all integral
exponents of an element a ∈ G which forms a subgroup of G.
Theorem 6.2.1 Let G be a group and a ∈ G. Then the set

hai = {an : n ∈ Z}

is a subgroup of G. Also hai contains a and is the smallest subgroup of G that


contains a.
Chapter 6. Subgroup And Cyclic group 108

Proof The set hai is nonempty since a0 = e ∈ hai.


∀x, y ∈ hai x = am and y = an for some integers m and n. Then

xy −1 = am a−n = am−n ∈ hai.

Since a = a1 it is clear that a ∈ hai. If H is any subgroup of G that contains a,


since H is closed under taking products and taking inverses, an ∈ hai for every
n ∈ Z. So hai ⊆ H. That is, every subgroup of G that contains a also contains
hai. This implies that hai is the smallest subgroup of G that contains a.
Definition 6.2.2 Let a be an element of the group G. Define

hai = {an : n ∈ Z} = {. . . , a−3 , a−2 , a−1 , a0 , a1 , a2 , a3 , . . .}

And in additive group

hai = {na : n ∈ Z} = {. . . , −3a, −2a, −a, 0a, a, 2a, 3a, . . .}

We call hai the cyclic subgroup of G generated by a.


Remark Note that

hai = {. . . , a−3 , a−2 , a−1 , a0 , a1 , a2 , a3 , . . .}.

In particular, a = a1 and e = a0 are in hai.


Definition 6.2.3 A group G is Cyclic if ∃a ∈ G, such that,

G = hai = {an : n ∈ Z}

Definition 6.2.4 Let a ∈ G. the order of a is the smallest positive integer n ∈ N


such that an = e. That is,

o(a) = min{n ∈ N | an = e}

If there exists such n we say that a has finite order.


And we define If an 6= e for all n ∈ N, we say that a has infinite order and we
define
o(a) = ∞.
In either case we call o(a) the order of a is denoted by o(a) or |a|.
1. (Z, +) is a cyclic group of infinite order.
Example 6.2.5

h1i = h−1i = hZi


109 6.2. Cyclic Group

2. (Zn , +) is a cyclic group of finite order.

h{a : (a, n) = 1}i = hZn i

3. The cyclic subgroup of C∗ generated by i is

hii = {. . . , i−4 , i−3 , i−2 , i−1 , i0 , i1 , i2 , i3 , i4 , . . . } = {i, −1, −i, 1}

Example 6.2.6 Let G = Z12 = {0, 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11}.

h1i = {n : n ∈ Z} = {. . . , −3, −2, −1, 0, 1, 2, 3, . . .} = Z12


h2i = {2n : n ∈ Z} = {. . . , −6, −4, −2, 0, 2, 4, 6, . . .} = {0, 2, 4, 6, 8, 10}
h3i = {3n : n ∈ Z} = {. . . , −9, −6, −3, 0, 3, 6, 9, . . .} = {0, 3, 6, 9}
h4i = {4n : n ∈ Z} = {. . . , −12, −8, −4, 0, 4, 8, 12, . . .} = {0, 4, 8}
h5i = {5n : n ∈ Z} = {. . . , −15, −10, −5, 0, 5, 10, 15, . . .} = Z12
h6i = {6n : n ∈ Z} = {. . . , −18, −12, −6, 0, 6, 12, 18, . . .} = {0, 6}
h7i = {7n : n ∈ Z} = {. . . , −21, −14, −7, 0, 7, 14, 21, . . .} = Z12

o(1) = 12 , o(2) = 6, o(3) = 4, o(4) = 3, o(5) = 12 and o(6) = 2, etc.


Exersice In Z12 , show that if (a, n) = d, then

hai = hdi

For example, h8i = h4i since (8, 12) = 4 and h5i = h1i since (5, 12) = 1.
This is not an accident that order of each element divides the order of the group.
According to Lagrange theorem not just in cyclic group, but in every finite group
the order of each element and the order of each subgroup divides the order of
group.
We see the subgroup generated by h2i is a subgroup of the entire set Z12 , also
as h3i,h4i,h6i,h8i,h9i,h10i.
But the subgroup generated by h1i,h5i,h7i,h11i generates the entire set Z12 .
Lemma 6.2.7 Let a be an element of finite order n in a group G. Then
(i) the elements e, a, a2 , . . . , an−1 are distinct and o(a) = |hai|;
(ii) every the elements a ∈ G has finite order;
(iii) If i, j ∈ Z, then ai = aj ⇐⇒ i ≡ j (mod n);
Chapter 6. Subgroup And Cyclic group 110

(iv) am = e ⇐⇒ n|m.
Proof
(i) Suppose ai = aj for some i, j ∈ N. Without loss of generality assume
0 ≤ i < j ≤ n − 1. Then

aj−i = aj a−i = ai a−i = e

But j − i < n, so aj−i = e contradicts the fact that o(a) = n It shows that hai
contains exactly n elements, that is, o(a) = |hai|.
(ii) By (i) we have, aj−i = e. So o(a) ≤ j − i, therefore a has finite order.
(iii) Exersice
(iv) ⇐= If n|m then m = nk. Then

am = ank = (an )k = ek = e

=⇒ Suppose am = e. If m ≥ n, we can use Division algorithm, m = nq +


r with 0 ≤ r < n. Then

am = anq+r = anq ar = (an )q ar = ear = ar


So, ar = e.
contradicting the assumption that n is the smallest natural number such that an =
e. Hence, r = 0 and n|m.
Theorem 6.2.8 Every cyclic group is abelian.
Proof let a, b ∈ G. Since G is cyclic, there exists an element g ∈ G that generates
each element in G. So, a = g i , b = g j .

ab = g i g j = g i+j = g j+i = g j g i = ba

However, the converse of this theorem does not hold. That is, not every abelian
group is cyclic. Here we give an important example of an abelian group which is
not cyclic.
Klein 4-group1
1
Also simply called the 4-group, and denoted V or V4 for vierergruppe, "four-group" in German.
It is named for the mathematician Felix Christian Klein.
111 6.3. Isomorphism

Consider the set V4 = {e, a, b, c}. Define ◦ to be the composition of functions


by means of following multiplication table.
◦ e a b c
e e a b c
a a e c b
b b c e a
c c b a e
From the table we have, a2 = b2 = c2 and ab = c, bc = a, ca = b.
It is easy to show that all the elements of V4 can be generated by just two
elements, a, b. That is

V4 = h a, b | a2 = b2 = (ab)2 i

(V4 , ◦) is an abelian group. But it is not a cyclic group because, non of its
elements is of order 4, i.e. not a single element can generates the entire group.

hei = {e}, hai = {e, a}, hbi = {e, b}, hci = {e, c}

We will show later this group has a geometric interpretation, it represents the
group of symmetries of a rectangle.
The subgroup lattice of V4 is as follows.

6.3 Isomorphism
Chapter 6. Subgroup And Cyclic group 112
Chapter 7

Matrix Group And The Group Of


Circle

The universe is an enormous direct product of representations of


“ symmetry groups.

Steven Weinberg ,

Let Mn (K) or M (n, k) be the set of all n × n matrices with entries from K. And
let K denote each one of the following: Z, Q, R, C, Zp where p is a prime number.

 
a11 a12 · · · a1m
 a21 a22 · · · a2m 
 
 .. .. .. .. 
 . . . . 
an1 an2 · · · anm

We denote (i, j) entry of an n × n matrix A by aij .


Recall the identity In of usual addition and multiplication of matrices .
Recall the determinant function det: Mn (K) → k.
Theorem 7.0.1 det: Mn (K) → k has the following properties.
(i) ∀A, B ∈ Mn (K), det(AB) = detA.detB.
(ii) detIn = 1.

113
Chapter 7. Matrix Group And The Group Of Circle 114

(iii) A ∈ Mn (K) is invertible ⇐⇒ detA 6= 0.


Proof Exersice
(Mn (K), +) is a group. Because:
1. Closure: ∀A, B, C ∈ Mn (K), + : A + B 7−→ C
2. Associativity: ∀A, B, C ∈ Mn (K), A + (B + C) = (A + B) + C
3. Identity: ∃In ∈ Mn (K), ∀A ∈ Mn (K), s.t. In + A = a = A + In
4. Inverse: ∀A ∈ Mn (K), ∃A−1 ∈ Mn (K), s.t. A−1 + A = In = A + A−1
But (Mn (K), ×) is not a group. Because:
1. Closure: ∀A, B, C ∈ Mn (K), × : A × B 7−→ C
2. Associativity: ∀A, B, C ∈ Mn (K), A × (B × C) = (A × B) × C
3. Identity: ∃In ∈ Mn (K), ∀A ∈ Mn (K), s.t. In × A = a = A × In
But A ∈ Mn (K) is invertible ⇐⇒ detA 6= 0. So it does not hold for all
A ∈ Mn (K).
Therefore, (Mn (K), ×) does not form a group.
But as always in mathematics we would like to forget about the bad point and
make a group out of that. Since the only problem is that it might be det = 0 we
consider the matrices with non-zero determinant.
Let GLn (K) denote the set of all invertible matrices with entries in K.

GLn (K) = {A ∈ Mn (K) : detA 6= 0}

GLn (K) is a group under matrix multiplication, which is called the General
Linear Group of degree n over K.
Since if A, B ∈ GLn (K), then det(AB) = detA.detB 6= 0 so AB ∈ GLn (K).
In special case when n = 2 and K = R
 
a b
GL2 (R) = { : ad − bc 6= 0}
c d
 
a b
The inverse of ∈ GL2 (R) is
c d
 −1  d −b

a b ad−bc ad−bc
= −c a
c d ad−bc ad−bc
115

This is the
 group of linear transformation in 2-Dimensional Euclidean plane
0
  
x x
which maps to , where
y y0
 0     
x a b x ax + by
= =
y0 c d y cx + dy

The special linear group, SLn (K) is the group of all matrices with det = 1.

SLn (K) = {A ∈ GLn (K) : detA = 1}


This got the name special since if we consider GLn (K) as invertible linear trans-
formation from K to itself, then the elements of SLn (K) are those transformations
which preserve volime and orientation.
The Orthogonal subset of GLn (K) is defined by

On (K) = {A ∈ GLn (K) : AAT = 1}

from AAT = 1 we have det(A) = ±1 .


The subgroup of On (K) which has determinant 1 forms a group itself.

SOn (K) = {A ∈ On (K) : detA = 1}

SOn (K) is called the Special orthogonal group. It is easy to show that each
A ∈ SOn (R) is of the form
 
cos θ − sin θ
A= θ ∈ [0, 2π]
sin θ cos θ

SOn (R) is the group of rotation in 2-Dimensional Euclidean plane.


 0   
x cos θ − sin θ x
=
y0 sin θ cos θ y

We note that Matrix group are examples of continous or Lie group. We do


not want to get into the detail since it is not a book in Lie Algebra or Topological
group, but consider the fact that The determinant det: Mn (K) → k is a continuous
function.
We note that Matrix group are examples of continous or Lie group. We do
not want to get into the detail since it is not a book in Lie Algebra or Topological
Chapter 7. Matrix Group And The Group Of Circle 116

group, but consider the fact that The determinant det: Mn (K) → k is a continuous
function.
We also define another matrices which have a group structure like, Euclidean
group E(n), Unitary group U (n), Special unitary group SU (n), Symplectic Sp(n)
and etc.

Un (C) = {U ∈ GLn (C) : U U T = 1}


The elements of U1 (C) includes just numbers.
U1 (C) = {eiθ : θ ∈ [0, 2π]}
The subset of Un (C) which has determinant 1, forms a group called SUn (C). It is
easy to show that every U ∈ SU2 (C) has the following form
  
a b 2 2
U= , |a | + |b | = 1
−b∗ a∗

The matrix subgroup chain is as follows


SOn (R) ≤ On (R) ≤ GLn (R) ≤ GLn (C)
SUn (C) ≤ Un (C) ≤ GLn (C)
We can consider GLn (K) as a subgroup of GLn+1 (K) by identifying the n × n
matrix A = [aij ] with
a11 a12 · · · a1m 0
 
.
    a21 a22 · · · a2m .. 
A 0
≤  ... .. .. ..
 
0 1  . . . 

an1 an2 · · · anm 0
0 0 ··· 0 1
GLn (K) is closed in GLn+1 (K), hence GLn (K) is a matrix subgroup of GLn+1 (K).
We can consider GLn (K) as a subgroup of GLn+1 (K) by identifying the n × n
matrix A = [aij ] with

a11 a12 · · ·
 
a1m 0
.
   a21 a22 · · · a2m .. 
A 0
≤  ... .. .. ..
 
0 1  . . . 

an1 an2 · · · anm 0
0 0 ··· 0 1
GLn (K) is closed in GLn+1 (K), hence GLn (K) is a matrix subgroup of GLn+1 (K).
Chapter 8

Symmetric Group

The Universe is built on a plan the profound symmetry of which is


“ somehow present in the inner structure of our intellect

Paul Valery ,

One of the first group that came into study is the group of permutations. Per-
mutations groups were used in mathematics before the abstract concept of a group
had been formulated. Indeed in the 19th century “group” was synonymous with
“group of permutations”. In fact the key role of permutations of the roots of an
equation for solving them by Lagrange was the reason that we defined the notion
of a group. And later as it turns out by Cayley theorem that every group can be
regarded as a subgroup of the full permutations.
We will now study one-to-one correspondences in more detail, particularly for
finite sets.
Definition 8.0.2 Let A be a non-empty set. A function σ : A → A is called a
permutation of A if σ is both one-to-one and onto (bijection).
The set of all permutations of A will be denoted by Sym(A).
The set of all permutations of the finite set {1, 2, . . . , n} will be denoted by Sn .
If σ and τ are elements of Sn we define their product στ to be the composition
of σ and τ , that is,
στ (i) = σ(τ (i)) for all i ∈ [n].

Convention

117
Chapter 8. Symmetric Group 118

1. We use lower case Greek letters such as σ, τ , α, β, etc., to indicate elements


of Sn .
2. For each integer n ≥ 1 we let [n] = {1, 2, . . . , n}.

Let [n] = {1, 2, . . . , n}, and let Sn denote the set of all permutations of [n] to
itself. Then Sn is a group under composition of functions. Because
1. ∀σ, τ ∈ Sn =⇒ στ ∈ Sn .

2. ∀σ, τ, π ∈ Sn =⇒ σ(τ π) = (στ )π.

3. id ∈ Sn .

4. ∀σ ∈ Sn =⇒ σ −1 ∈ Sn .

Sn is called symmetric group of degree n .

8.1 Cycle Decomposition


We think of each permutation as a map σ : [n] → [n] that rearrange the order of the
elements. For example, σ : {1, 2, 3, 4} → {2, 3, 1, 4} is a permutation. In terms of
functions, it is nothing more than a function that

σ(1) = 2 σ(2) = 3 σ(3) = 1 σ(4) = 4

So σ is a function which it is determined by σ(1), σ(2), σ(3), σ(4).


A Simpler notation for the permutation σ which sends i → σ(i) is the two line
or two row representation.
 
1 2 ··· n
σ= .
σ(1) σ(2) . . . σ(n)

Where each σ(i) is the image of i under σ , i ∈ [n]


Since σ is injective, σ(1), σ(2), . . . , σ(n) are all different and therefore include
all n elements of the set A, so each element of A must appear once and only once
in the second row.
119 8.2. Composition of two permutations

The identity of Sn , which is called identity permutation, sends every element to


itself, is  
1 2 ··· n
id =
1 2 ··· n
And the inverse of an arbitrary permutation is
 
−1 σ(1) σ(2) . . . σ(n)
σ =
1 2 ··· n

To count the number of elements of Sn , if we define an element σ ∈ Sn , there


are n possible choices for the value of first element σ(1). Since σ is bijective, that
leaves (n − 1) remaining choices for σ(2), then (n − 2) choices for σ(3), and so
on. In particular,
|Sn | = n(n − 1)(n − 2) . . . 1 = n!

8.2 Composition of two permutations


For now we just get to some calculation in Sn in order to understand some facts
about the Group.
Example 8.2.1 Let and τ be defined as follows:
   
1 2 3 4 1 2 3 4
σ= τ=
3 1 4 2 1 3 2 4

It follows that
στ (1) = σ(τ (1)) = σ(1) =3
στ (2) = σ(τ (2)) = σ(3) =4
στ (3) = σ(τ (3)) = σ(2) =1
στ (4) = σ(τ (4)) = σ(4) =2
Thus we have  
1 2 3 4
στ =
3 4 1 2

But

τ σ(1) = τ (σ(1)) = τ (3) =2


τ σ(2) = τ (σ(2)) = τ (1) =1
τ σ(3) = τ (σ(3)) = τ (4) =4
τ σ(4) = τ (σ(4)) = τ (2) =3
Chapter 8. Symmetric Group 120

Thus we have  
1 2 3 4
στ =
3 4 1 2

Note that στ 6= τ σ , i.e., multiplication of permutations is not commutative.


There is a more compact and convinent way to use insead of two line represen-
tation. Note that in every permutation the first line is the same and only the second
line is different. So we could use a simpler form called Cycle Notation.
First we use some terminology.
Definition 8.2.2 The fixed points of a permutation σ are the elements i ∈ {1, 2, . . . , n}
such that σ(i) = i .
Definition 8.2.3 Let i1 , i2 , . . . , ik be k distinct elements from {i1 , i2 , . . . , in }. The
cycle of length k or a k-cycle (i1 , i2 , . . . , ik ) is a permutation σ ∈ Sn such that

σ(i1 ) = i2
σ(i2 ) = i3
σ(i3 ) = i4
.. .. ..
. . .
σ(ik−1 ) = ik
σ(ik ) = i1

and leaving all the other elements of {1, . . . , n}. That is,

σ(i) = i i∈
/ {i1 , i2 , . . . , ik }

If k = 1 then the cycle = (i1 ) is just the identity element in Sn .


k is called the length of the cycle.
Definition 8.2.4 Let σ ∈ Sn . the order of a permutation σ is the smallest positive
integer n ∈ N such that σ n = (1). That is,

o(σ) = min{n ∈ N | σ n = (1)}

Remark A k-cycle can be written in k different ways, since

(a1 a2 . . . ak ) = (a2 a3 . . . ak a1 ) = · · · = (ak a1 . . . ak−1 )

Example 8.2.5
121 8.2. Composition of two permutations
 
1 2 3 4
For example, the permutation σ = means σ(1) = 3, σ(2) =
3 1 4 2
1, σ(3) = 4, σ(4) = 2.
In cycle notation we start with something, say for example the smallest integer
and write what it is maps to as a cycle. Here 1 maps to 3, so we have for now (1 3).
And 3 maps to 4 so we have (1 3 4) and at last 4 maps to 2, (1 3 4 2). Since nothing
is left the cycle is over, and we encoded every useful information in the two line
representation.
In fact,
σ(1) = 3
σ 2 (1) = σ(σ(1)) = 4
σ 3 (1) = σ 2 (σ(1)) = 2
σ 4 (1) = σ 3 (σ(1)) = 1
That is , powers of σ(1) generates the entire permutation. But it is not always
as so as we shall see.
So we can write it as

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

σ 4 = (1) so the order of the group is m = 4.


Theorem 8.2.6 Define a relation ∼ on the set {1, 2, . . . , n} by

i ∼ j ⇐⇒ σ m (i) = j

Then ∼ is an equivalence relation. The equivalence classes would have the follow-
ing form.
[i] = {i, σ(i), σ 2 (i), σ 3 (i), . . . , σ m (i) = i}
Proof Exersice
One of the advanteges of a equivalence relation as we proved is that it splits
up the set into disjoint classes which we called equivalence classes. Here the
equivalence classes are exactly what we have defined a cycles.
Example 8.2.7 Let’s consider the permutation τ ∈ S11 given by
 
1 2 3 4 5 6 7 8 9 10 11
τ=
3 7 6 4 1 2 5 9 8 11 10
Chapter 8. Symmetric Group 122

 
σ(1) = 3

 σ 2 (1) = σ(3) = 6 
  

 σ 3 (1) = σ(6) = 2 
 σ(8) = 9 σ(10) = 11

 σ 4 (1) = σ(2) = 7  σ 2 (8) = σ(9) = 8
 σ 2 (10) = σ(11) = 10
 σ 5 (1) = σ(7) = 5 
σ 6 (1) = σ(5) = 1

The equivalence classes or cycles are:

[1] = {1, 3, 6, 2, 7, 5}
[8] = {8, 9}
[10] = {10, 11}

In equivalence class language

τ /∼ = {[1], [8], [10]}

So we can write it as
(1 3 6 2 7 5)(8 9)(10 11)

It is not a cycle, since is the product of three cycles. We can ignore the fixed
points of a permutation, here 4 is maps to itself, (4).
Lemma 8.2.8 Let σ ∈ Sn have order m . Then for all integers i, j we have

σ i = σ j ⇐⇒ i ≡ j (mod m)

Proof σ have order m, so m is the smallest positive integer such that σ m = (1).
=⇒ If σ i = σ j then σ i−j = (1). Using the division algorithm we can have
i − j = qm + r for integers q, r with 0 ≤ r < m . Then

σ i−j = σ qm+r = (σ m )q σ r = σ r

we have r = 0 because m is the smallest positive integer such that σ m = (1). Thus
m|(i − j) and so i ≡ j (mod m).
⇐= If i ≡ j (mod m), then i = j + km for some k .

σ i = σ j+km = σ j σ km = σ j (σ m )k = σ j
123 8.2. Composition of two permutations

We know that equivalence classes of a set are either identical or disjoint. As


cycles are just equivalence classes, so we can define the consept of being disjoint
for cycles.
Definition 8.2.9 Let σ = (a1 a2 . . . ak ) and τ = (b1 b2 . . . bk ) be cycles in Sn .
Then σ and τ are said to be disjoint if ai 6= bj ∀i, j .
That is, if the elements moved by one are left fixed by the other
For example, in the last example (1 3 6 2 7 5)(8 9)(10 11), these three cycles
are disjoint. But (7 3 6) and (2 6 1) are not since they both have the number 6 in
common.
Here we are going to point out two important facts.
First of all, every equivalence relation partition a set into equivalence classes. It
is the same as saying, if we have a permutation it can be written as disjoint cycles.
Theorem 8.2.10 Every permutation on a finite set can be written uniquely, except
for order of cycles or the different ways a cycle is written, as a product of disjoint
cycles.
Proof Exercise.

(2 3) = (1 2)(2 3)(1 3).

(2 3) = (1 2)(1 2)(2 3).


As we pointed out permutations are not abelian. As we proved every permutation
can be written as a product of disjoint cycles. But if two cycles are disjoint, they
commute.
Theorem 8.2.11 If σ and τ are disjoint cycles in Sn then στ = τ σ.
Proof Let A = {a1 , a2 , . . . , ak , b1 , b2 , . . . , bk , c1 , c2 , . . . , ck } and let σ = (a1 a2 . . . ak )
and τ = (b1 b2 . . . bk ) be cycles in Sn .
where {a1 , a2 , . . . , ak } ∩ {b1 , b2 , . . . bk } = ∅
For any element x ∈ A, we note
• If x ∈
/ {ai }, then xσ = x. Then (xσ)τ = xτ = (xτ )σ.
• If x ∈
/ {bi }, then xτ = x. Then (xσ)τ = xσ = (xτ )σ.
Hence, στ = τ σ.
Example 8.2.12 Let σ, τ ∈ S9 be given by σ = (2 5 3 6), τ = (1 4 7 9).
Chapter 8. Symmetric Group 124

To show that they commute we have to show that

στ (i) = τ σ(i) ∀i ∈ {1, 2, 3, . . . , 9}

στ (1) = σ(4) = 4
τ σ(1) = τ (1) = 4
So
στ (1) = τ σ(1)
Since σ fixes 1 we have σ(1) = 1.
Also for 2 :
στ (2) = σ(2) = 5
τ σ(2) = τ (5) = 5
So
στ (2) = τ σ(2)

Similary for all i ∈ {1, 2, 3, . . . , 9} we have στ (i) = τ σ(i).


Every permutation is a product of transpositions.
Definition 8.2.13 A cycle of length 2 is called a transposition. A transposition
(i j) is a permutation that interchanges i and j and leaves other elements fixed.

(i j) = (j i)

Lemma 8.2.14 Every permutation σ ∈ Sn is a product of (not necessarily disjoint)


transpositions.
Proof It is enough to show that every cycle is a product of transpositions, since
every permutation is a product of cycles.

(a1 a2 . . . ak ) = (a1 ak )(a1 ak−1 ) . . . (a1 a3 )(a1 a2 )

We can write it in several ways for example,

(a1 a2 . . . ak ) = (a1 a2 )(a2 a3 ) . . . (ak−2 ak−1 )(ak−1 ak )

1. If n = 1, then |S1 | = 1! = 1. Thus, S1 only contains the identity


Example 8.2.15
permutation.   
1 2
S1 = =e
1 2
125 8.2. Composition of two permutations

2. If n = 2, then |S2 | = 2! = 2. Thus, S2 contains the identity permutation and a


transposition.

    
1 2 1 2
S2 = =e , = (1 2)
1 2 2 1

3. If n = 3, then |S3 | = 3! = 6. Thus, S3 contains :

     
1 2 3 1 2 3 1 2 3
S3 = =e , = (2 3) , = (1 3) ,
1 2 3 1 3 2 3 2 1
     
1 2 3 1 2 3 1 2 3
= (1 2) , = (1 2 3) , = (1 3 2)
2 1 3 2 3 1 3 1 2

The multiplication for S3 can be written down as follows

e (1 2 3) (1 3 2) (2 3) (1 3) (1 2)
e e (1 2 3) (1 3 2) (2 3) (1 3) (1 2)
(1 2 3) (1 2 3) (1 3 2) e (1 2) (2 3) (1 3)
(1 3 2) (1 3 2) e (1 2 3) (1 3) (1 2) (2 3)
(2 3) (2 3) (1 3) (1 2) e (1 2 3) (1 3 2)
(1 3) (1 3) (1 2) (2 3) (1 3 2) e (1 2 3)
(1 2) (1 2) (2 3) (1 3) (1 2 3) (1 3 2) e

We usually show the odd cycles and the even cycles with τ and σ respectivly.

e σ1 σ2 τ1 τ2 τ3
e e σ1 σ2 τ1 τ2 τ3
σ1 σ1 σ2 e τ3 τ1 τ2
σ2 σ2 e σ1 τ2 τ3 τ1
τ1 τ1 τ2 τ3 e σ1 σ2
τ2 τ2 τ3 τ1 σ2 e σ1
τ3 τ3 τ1 τ2 σ1 σ2 e

It is easy to show that all the elements of S3 can be generated by just two elements,
any τ, σ. That is
hσ1 , τ1 i = S3
Chapter 8. Symmetric Group 126

There is also an interesting relation between elements of S3

τ12 = τ22 = τ32 = e

σ13 = σ23 = e
σ12 = σ2 , σ22 = σ1
σ 1 τ1 = τ2
σ12 τ1 = τ3

Now we can construct a table by using only the genetators of S3 .

e σ1 σ12 τ1 σ 1 τ1 σ12 τ1
e e σ1 σ12 τ1 σ 1 τ1 σ12 τ1
σ1 σ1 σ12 e σ12 τ1 τ1 σ 1 τ1
σ12 σ12 e σ1 σ 1 τ1 σ12 τ1 τ1
τ1 τ1 σ1 τ1 σ12 τ1 e σ1 σ12
σ1 τ1 σ1 τ1 σ12 τ1 τ1 σ12 e σ1
σ12 τ1 τ3 τ1 σ1 τ1 σ1 σ12 e

More generally, we can write S3 as a set of generators with the a relation between
them.
S3 = hσ1 , τ1 ; σ13 = τ12 = e, σ1 τ1 = τ1 σ12 i
Remark Note that the generators are not unique, for example one could write
hσ1 , τ2 i = S3 or hσ12 , τ1 i = S3 .

8.3 The Alternating group


We saw that the decomposition of permutations into transpositions are not unique
up to order, but it is unique up to parity. For example,

(1 2 3 4) = (1 4)(1 3)(1 2).

(1 2 3 4) = (1 4)(7 2)(7 2)(1 3)(1 2)(1 2)(1 2).


The number of transpositions in both of decomposition are odd.

(1 2 3) = (1 3)(1 2).
(1 2 3) = (2 4)(2 4)(1 3)(5 3)(5 3)(1 2)(1 3)(1 3).
127 8.3. The Alternating group

The number of transpositions in both of decomposition are even.


This gives us an important result.
Definition 8.3.1 A permutation σ ∈ Sn is called even if it can be written as a
product of an even number of transpositions, and odd if it can be written as a
product of an odd number of transpositions.
We define the function sgn : Sn → {1, −1} by

1 if σ is even
sgn(σ) =
−1 if σ is odd

Example 8.3.2 The permutation


 
1 2 3 4 5 6 7 8
π=
2 1 4 7 5 6 8 3

is even. Because

π = (1 2)(3 4 7)(5)(6) = (1 2)(3 8)(3 7)(3 4)

is the product of 4 (even number) transposition.


Remark. Given a square matrix A = [aij ] ∈ Mn×n . The determinant of A may be
defined by the sum over all permutations of n elements (i.e., over the symmetric
group) X
det(A) = sgn(σ)a1σ(1) a2σ(2) · · · anσ(n)
σ∈Sn

 
a11 a12
For n = 2, we have only two permutations S2 = {e, (1 2)}.
a21 a22

det(A) = sgn(e)a11 a22 + sgn(1 2)a12 a21

Since sgn(e) = 1 and sgn(1 2) = −1 we obtain

det(A) = a11 a22 − a12 a21

det(A) = sgn(e)a11 a22 + sgn(1 2)a12 a21

For n = 3,  
a11 a12 a13
a21 a22 a23 
a31 a32 a33
Chapter 8. Symmetric Group 128

we have 6 permutations .

S3 = {e, (2 3), (1 3), (1 2), (1 2 3), (1 3 2)}

In this case the determinant is :


det(A) = sgn(e)a11 a22 a33 + sgn(1 2)a12 a21 a33 + sgn(1 3)a13 a22 a31 +
sgn(2 3)a11 a23 a32 + sgn(1 2 3)a12 a23 a31 + sgn(1 3 2)a13 a21 a32 = a11 a22 a33 −
a12 a21 a33 + a13 a22 a31 − a11 a23 a32 + a12 a23 a31 − a13 a21 a32
Theorem 8.3.3 A permutation σ ∈ Sn cannot be both even and odd. It is either
even or odd.
Proof Suppose that σ is both even and odd that is σ = τ1 τ2 . . . τk = π1 π2 . . . πl ,
where k is even and l odd. Since every transposition is its own inverse, this would
imply that
e = τ1 τ2 . . . τk πl πl−1 . . . π1

Since k + l is odd, this contradicts the fact that the identity permutation is even.
Definition 8.3.4 The set of all even permutations in Sn is denoted by An called
the alternating group of degree n. That is,

An = {σ ∈ Sn : sgn(σ) = 1}

Theorem 8.3.5 The set An is a subgroup of Sn , An ≤ Sn .


Proof id ∈ An . If σ ∈ An then σ −1 ∈ An .
Therefore ∀σ, τ ∈ An =⇒ στ −1 ∈ An .
Theorem 8.3.6 If An denotes the set of even and Bn denotes the set of odd per-
mutations in Sn , Show
|Sn | |n!|
|An | = |Bn | = =
2 2
Proof Let σ ∈ Sn . Define a function φ : An → Bn , by φ(τ ) = στ . There is
a one-to-one correspondence between the number of elements of these two sets.
Becasue
• Injective : φ(τ1 ) = φ(τ2 ) =⇒ στ1 = στ2 =⇒ τ1 = τ2 .
• Surjective : Let π ∈ Sn . That is π is an odd permutation so σπ is an even
permutation. π = σ 2 = φ(σπ).
Hence, |An | = |Bn |.
129 8.4. Dihedral Groups and Symmetries Of Objects

We have An ∪ Bn = Sn and An ∩ Bn = ∅.
|Sn | = |An | = |Bn | = 2|An |. Therefore,

|Sn | |n!|
|An | = = 
2 2
Example 8.3.7
S1 = {e}
A1 = {e}

S2 = {e, (1 2)}
A2 = {e}

S3 = {e, (1 2 3), (1 3 2), (2 3), (1 3), (1 2)}


A3 = {e, (1 2 3), (1 3 2)}

8.4 Dihedral Groups and Symmetries Of Objects


As we mentioned groups can be used to determine or measure symmetry. By
symmetry of an object we mean transformations(bijections) that do not change the
object. To write it more formally, we need some terminology. Here we consider
objects in Euclidean geometry of the plane R2 .
Definition 8.4.1 An isometry or rigid motion of R2 is a permutation σ of R2 that
preserves distance and maps the object into itself. That is, if d is a metric

σ(F ) = F ⇐⇒ σ is a symmetry of F

d(x, y) = d(σ(x), σ(y)) ∀x, y ∈ F

Using composition of bijections as the binary operation, it is easy to show that


the set of rigid motions of the plane forms a group. There are only three kinds of
rigid motions: parallel translations, rotations about a point, reflections across a
line. Now given a figure in the plane F ⊆ R2 , the symmetry group of F is the set
of rigid motions of the plane mapping F to F . This means, the symmetry group of
F must map the origin of F to itself. So the set of symmetries does not contain
translations any more, it just consists of rotations and reflections.
Chapter 8. Symmetric Group 130

There is another approach to this intention. Consider the object as a graph. A


graph is an ordered pair G = (V, E) consisting of a set V of vertices together with
a set E of edges.
In this case E ⊆ V × V . E is the subset of cartisean product of vertices.
Becasue an edge is related with 2 vertices. If 2 vertices connected there is an edge
between them. The elements of E are of the form

E = {(Vi , Vj ) , Vi 6= Vj and Vi , Vj ∈ V }

The pair (V1 , V2 ) is called an edge. When (V1 , V2 ) ∈ E, we say V1 , V2 are con-
nected. Here we consider edges as unordered pais, that is (V1 , V2 ) = (V2 , V1 ).
If we represent the elements of V as points of the plane, and draw a line between
(V1 , V2 ) ∈ E we have an geometric object. It turns out we can consider a graph as
a group (G, ∗) which G = V and ∗ = E acts a binary operation. So we can define
the group of permuations for a graph (V, E) to be the set of all permutations of
V → V that preserves connectedness. That is, the set of all σ ∈ SV such that

(x, y) ∈ E ⇐⇒ (σ(x), σ(y)) ∈ E

This set is a subgroup of SV , and it is equal to SV if every two vertic is connected.


Proposition 8.4.2 An isometry of the plane R2 forms a group, called the Euclidean
group.
Lemma 8.4.3 If F is a geometric object in R2 then the set of symmetries of F
forms a subgroup of Sn called the symmetry group of F , which is denoted by
Sym(F ) or D2n or Dn .
Examples
If B is a butterfly

then Sym(B) = {e, τ } where τ is the reflection in the axis of symmetry.


While cyclic groups describe objects that have only rotational symmetry, dihe-
dral groups describe objects that have both rotational and bilateral symmetry.
131 8.4. Dihedral Groups and Symmetries Of Objects

A simple example occurs by taking F to be a cyclic group.


Now let F be a regular n-gon, That is a convex polygon with n sides of equal
length and n angles of the same size. For example, a regular 3-gon is an equilateral
triangle, and a regular 4-gon is a square.
It is convenient to label the vertices of the n-gon by 1, 2, . . . , n, so that each
symmetry may be represented by a permutation of {1, 2, . . . , n}, i.e., by an element
of Sn .
Let S be a square in the plane, that is a regular polygon with n = 4. Then there
are exactly eight symmetries of S. These are
1. e : the identity.
π
2. r : a rotation of 2
around the center.
3. r2 : a rotation of π around the center.

4. r3 : a rotation of 2
around the center.
5. t : a reflection over one of the sides.
6. tr : the composition of t and r.
7. tr2 : the composition of t and r2 .
8. tr3 : the composition of t and r3 .

D8 = {e, r, r2 , r3 , t, rt, r2 t, r3 t}

r4 = f 2 = e, f rf = r−1
Chapter 8. Symmetric Group 132

The square is symmetric about its axises, we call this kind of symmetry a
reflection. Let t denote reflection about horizontal axis and v about vertical axis.
But we do not need to write down the symmetries that come from v, because every
such symmetry can be obtained by composition of two rotation and reflection.
Corollary 8.4.4 We do not need 2 reflection as a basis that span the group of
symmetries, the composition of a rotation and a reflection generate the entire
group.
This is the group table for the symmetry group of a square D8 .

e r r2 r3 t rt r2 t r3 t
e e r r2 r3 t rt r2 t r3 t
2
r r r r3 e rt r2 t r3 t t
r2 r2 r3 e r r2 t r3 t t rt
r3 r3 e r r2 r3 t t rt r2 t
t t r3 t r2 t rt e a3 a2 a
rt rt t r3 t r2 t r e r3 r2
r2 t 2
r t rt t a3 t r2 r e r3
r3 t r3 t r2 t rt t r3 r2 r e

It turns out that D8 is a non-abelian group since for example, rt 6= tr. But
rt = tr3 .
In general,a regular n-gon has 2n different symmetries: n rotational and n
reflectional symmetries.
• If n is odd each axis of symmetry connects the midpoint of one side to the
opposite vertix.
• If n is even there are n2 axies of symmetry connecting the midpoint of
opposite side and n2 axies of symmetry connecting opposite vertices.

we can explicitly, write a formula depending on the 2 generators.


133 8.4. Dihedral Groups and Symmetries Of Objects

D2n = {e, r, r2 , . . . , rn−1 , t, rt, r2 t, . . . , rn−1 t}


rn = f 2 = e, f rf = r−1

Now consider the simplest polygon : An equilateral triangle

The set of symmetries of an equilateral triangle forms a group with 6 elements.


|D3 | = 6 and this is the same as the permutation group on 3 elements, |S3 | = 6.
The first question that comes up is:
Is the symmetry of a polygon is permutation of vertices like what we have for
an equilateral triangle ?
NO. Not every permutation is a symmetry because in a square we cannot swap
two opposite vertix but not swap the others. There is no action on Euclidean plane
to do that but in triangle it is all of permutation. If you increase sides, you do not
get Sn group for its symmetry, but dihedral group.
The second question: Is it just happen for a triangle that has permutations of
vetices for its symmetry group? The permutation group has n! elements and the
dihedral group has 2n elements. They happen to be equal n! = 2n if if n = 3.
Does it have any other solution? not in integers but in R.

Z ∞
n! = Γ(n + 1) = nΓ(n) = tn e−t dt = 2n
0

Γ(n) = 2
The solutions of this integral equation are:

n1 = 3, n2 = −3.9784, n3 = −3.0763, n4 = 0.4428

What do they mean? can we define a negative number to be a side?


Chapter 8. Symmetric Group 134

8.5 Cayley Diagrams


A Cayley diagram is a graph that visualize a group. There is one vertice in the graph
for each element in the group, and the edges (arrows) show how the generators
act on the elements of the group. That is, if the group has two generators, a and
b, then there will be one type of arrow for generator a and another type of arrow
for generator b. Here is a Cayley diagram of a group; you can see that it fits this
description.
Chapter 9

Cosets and Lagrange’s Theorem

It might be difficult, at this point, for students to see the extreme


“ importance of this result [Lagrange’s Theorem]. As we penetrate
the subject more deeply they will become more and more aware of
its basic character.

I. N. HERSTEIN, Topics in Algebra



9.1 Coset
Recall the notion of congruence in arithmetic. Two integers a, b are congruent if
they have the same remainder divided by n. That is,
a ∼ b ⇐⇒ a ≡ b (mod n)
We showed that congruence modulo n is an equivalence relation, and every such
relation partitions a set into equivalence classes. The consept of equivalence
relation is very powerful and has many applications.
Equivalence relation is not restricted to sets only but it can be generalized to any
group, since groups are just special kinds of sets. Note that in congruence relation,
the integers Z is a group and nZ is a subgroup. In group theory language a, b ∈ Z
are equivalent if their difference is in the subgroup nZ. That is,
a ∼ b ⇐⇒ a − b ∈ nZ
Using this similarity we now define equivalence relation in a group.

135
Chapter 9. Cosets and Lagrange’s Theorem 136

Definition 9.1.1 Let G be a group and H ≤ G. Define a relation ∼R on G by


a ∼R b ⇐⇒ ab−1 ∈ H

This relation a ∼R b is an equivalence relation, because


• (i) Reflexivity: Since e ∈ H and e = aa−1 , we have a ∼R a.

• (ii) Symmetry: if a ∼R b then ab−1 ∈ H then (ab−1 )−1 = ba−1 ∈ H so


b ∼R a.

• (iii) Transitivity: if a ∼R b and if b ∼R c then ab−1 ∈ H and bc−1 ∈ H. But


(ab−1 )(bc−1 ) = ac−1 ∈ H. Hence a ∼R c.

Since ∼R is an equivalence relation on the group G, the equivalence classes


partition G.
[a] = {b ∈ G | a ∼R b} = {b ∈ G | ab−1 ∈ H}
= {b ∈ G | ab−1 = h ∈ H}
= {hb | h ∈ H}

Note that since the group might be non-abelian we could define another equiva-
lence relation on G, which is
a ∼L b ⇐⇒ a−1 b ∈ H

[a] = {b ∈ G | b ∼L a} = {b ∈ G | b−1 a ∈ H}
= {b ∈ G | b−1 a = h ∈ H}
= {bh | h ∈ H}
Remark Note that both of ∼L and ∼R are equivalence relations but one cannot be
implied by another, so these two relations partition a group G in two different way.
Definition 9.1.2 Let G be a group and let H be a subgroup of G. For each element
a ∈ G . The set
aH = {ah | h ∈ H}.
is called the left coset of H in G determined by a.
Similarly, the right coset H in G determined by a is the set
Ha = {ha | h ∈ H}.
137 9.1. Coset

Remark In the case of additive notation the left and right cosets of H in G
determined by a is written respectivly

a + H = {a + h | h ∈ H}

H + a = {h + a | h ∈ H}

Here we give some important properties of cosets. The properties are given for
left cosets. Similar properties hold for right cosets.
Theorem 9.1.3 Let H be a subgroup of G, and a, b ∈ G. Then, the following
statements hold.
(i) a ∈ aH.
(ii) aH = H ⇐⇒ a ∈ H.
(iii) aH = bH ⇐⇒ a ∈ bH.
(iv) aH = bH or aH ∩ bH.
(v) aH = bH ⇐⇒ a−1 b ∈ H.
(vi) |aH| = |bH|.
(vii) aH = Ha ⇐⇒ a ∈ H = aHa−1 .
(viii) aH 6 G ⇐⇒ a ∈ H.
Proof
(i) Since H 6 G, e ∈ H. a = ae ∈ aH.

(ii) =⇒ Suppose aH = H, prove a ∈ H. From (i), a ∈ aH but aH =


H hence a ∈ H.
⇐= Assume that a ∈ H and show that aH = H. By clousre of H we have aH ⊆
H.To show that H ⊆ aH, let h ∈ H. Then, since a ∈ H and h ∈ H, we know that a−1 h ∈
H. Thus, h = eh = (aa−1 )h = a(a−1 h) ∈ aH.

(iii) If aH = bH, then a = ae ∈ aH = bH. Conversely, if a ∈ bH we have a =


bh where h ∈ H, and therefore aH = (bh)H = b(hH) = bH.

(iv) from (iii) if c ∈ aH ∩ bH, then cH = aH and cH = bH.


Chapter 9. Cosets and Lagrange’s Theorem 138

(v)aH = bH ⇐⇒ a ∈ bH ⇐⇒ a = bh ⇐⇒ ab−1 = h ⇐⇒ ab−1 ∈ H.

(vi) There is a bijection between |Ha| and |Hb|. The correspondence


φ : Ha → Hb which maps ha 7→ hb is clearly onto. It is one-to-one because :

ha = h0 ⇐⇒ h = h0 ⇐⇒ hb = h0 b ⇐⇒ φ(ha) = φ(h0 a)
(vii) aH = Ha ⇐⇒ (aH)a−1 = (Ha)a−1 = H(aa−1 ) = H
(viii) aH 6 G ⇐⇒ e ∈ aH ⇐⇒ e = ah ⇐⇒ a = h−1 ∈ H.
Remark Suppose G = {a1 , a2 , . . . , an } is a group with n elements and H ≤ G.
Then if we form the list of all cosets of H in G we have

a1 H, a2 H, . . . , an H.

That is, some of these cosets may have repeated several times. But as we proved in
part (iii)
ai H = aj H ⇐⇒ ai ∈ aj H
But as noted in the above examples some of the cosets in this list are repeated
several times. If we remove all repetitions from the list we are left with what we
shall call the distinct cosets of H in G. If there are s distinct cosets we may denote
them by a1 H, a2 H, . . . , as H.
Definition 9.1.4 The number of distinct left cosets of H in G is called the index
of H in G, and is denoted by [G : H].
Bibliography

[1] Lang, S., Algebra (3rd edition). Reading, Mass. : Addison-Wesley Publishing
Co. , Inc., 1993.
[2] T. W. Hungerford, Algebra, Springer Verlag, 1980.
[3] I. Stewart Why Beauty Is Truth: History of Symmetry,2007.
[4] E.Lim, Basic Algebra I and II, (SSymmetry in physics), King’s college
London, 2013.
[5] J. A. Gallian, Contemporary Abstract Algebra, (Third Edition), D.C. Heath,
1994.
[6] J. B. Fraleigh, A First Course in Abstract Algebra, (Fifth Edition), Addison-
Wesley, 1994.
[7] G. Birkhoff and S. MacLane, A Survey of Modern Algebra, A. K. Peters Ltd.,
1997.
[8] I. N. Herstein, Topics in Algebra, (Second Edition), Blaisdell, 1975.
[9] G. D. Birkhoff and T. C. Bartee, Modern Applied Algebra, McGraw-Hill
Book Company, 1970.
[10] L. Dornhoff and F. Hohn, Applied Modern Algebra, Macmillan, 1978.
[11] B. L. Van der Waerden, Modern Algebra, (Seventh Edition, 2 vols), Fredrick
Ungar Publishing Co., 1970.
[12] I. Stewart Why Beauty Is Truth: History of Symmetry,2007.
[13] E.Lim, Basic Algebra I and II, (SSymmetry in physics), King’s college
London, 2013.
[14] N. Jacobson, Basic Algebra I and II, (Second Edition, 2 vols), W. H. Freeman
and Company, 1989.

139

You might also like