Mth112 Notes 2
Mth112 Notes 2
Mth112 Notes 2
Everybody is a genius. But if you judge a fish by its ability to climb a tree, it will live its whole life
believing that it is stupid.
—Albert Einstein
This course highlights various types of mathematical thinking including direct, indirect and induc-
tive proofs, with careful treatment of quantified statements. Topics include sets, number theory
and relations. Development of the ability to write a precise mathematical proof is the primary
goal. The primary objective of this course is to develop a certain amount of fluency with the lan-
guage of mathematics and basic techniques of mathematical proof. Students should understand
the significance of a quantified statement, what it means if such a statement is true, and what it
takes to show that such a statement is false. The emphasis should be on the writing of proofs. We
want the student to learn the rules of logic, develop a good writing style, explore some interesting
mathematics and learn to make conjectures, decide if the conjecture is true or false, and write up
a justification (proof or counterexample). It is intended that a student completing this course will
have had success with proving and disproving mathematical statements. As a consequence, the
emphasis should be on the collecting and recollecting of attempts at proof, until the student has
some success.
Part of the reason you are likely to find this material difficult is that it will seem unmotivated.
It is bound to: the sole motivation is to provide you with the foundation on which to build the
mathematics that comes later – mathematics that you do not yet know about! So take it steadily,
and try to understand the new concepts as you meet them. There is little in the way of new facts
to learn, but a great deal to comprehend! (The actual facts contained in these notes could be listed
on three or four pages of notes.) And try the exercises as many of them as possible. They are
included for a purpose: to aid your understanding.
Discuss any difficulties that arise with your colleagues and with your lecturer. Do not give up.
Students all around the world managed it last year. Likewise the previous year, and the year before
that. So did I. So will you!
Chapter 1
What is Mathematics?
—Bertrand Russell
Fear is the main source of superstition, and one of the main sources of cruelty. To conquer fear is the
beginning of wisdom.
—Bertrand Russell
In fact, the answer to the question “What is mathematics?” has changed several times during the
course of history.
Up to 500BC or thereabouts, mathematics was indeed the study of number. This was the period of
Egyptian and Babylonian mathematics. In those civilizations, mathematics consisted almost solely
of arithmetic. It was largely utilitarian, and very much of a “cookbook” variety. (“Do such and
such to a number and you will get the answer.”)
The period from around 500BC to 300AD was the era of Greek mathematics. The mathematicians
of ancient Greece were primarily concerned with geometry. Indeed, they regarded numbers in a
geometric fashion, as measurements of length, and when they discovered that there were lengths to
which their numbers did not correspond (the discovery of irrational lengths), their study of number
largely came to a halt. For the Greeks, with their emphasis on geometry, mathematics was the
study of number and shape.
In fact, it was only with the Greeks that mathematics came into being as an area of study, and ceased
being merely a collection of techniques for measuring, counting, and accounting. Greek interest in
mathematics was not just utilitarian; they regarded mathematics as an intellectual pursuit having
both aesthetic and religious elements. Around 500BC, Thales of Miletus (now part of Turkey)
1
introduced the idea that the precisely stated assertions of mathematics could be logically proved
by a formal argument. This innovation marked the birth of the theorem, now the bedrock of
mathematics. For the Greeks, this approach culminated in the publication of Euclid’s Elements,
reputedly the most widely circulated book of all time after the Bible.
There was no major change in the overall nature of mathematics, and hardly any significant advances
within the subject, until the middle of the 17th century, when Isaac Newton (in England) and
Gottfried Leibniz (in Germany) independently invented calculus. In essence, calculus is the study
of continuous motion and change. Previous mathematics had been largely restricted to the static
issues of counting, measuring, and describing shape. With the introduction of techniques to handle
motion and change, mathematicians were able to study the motion of the planets and of falling
bodies on earth, the workings of machinery, the flow of liquids, the expansion of gases, physical
forces such as magnetism and electricity, flight, the growth of plants and animals, the spread of
epidemics, the fluctuation of profits, and so on.
After Newton and Leibniz, mathematics became the study of number, shape, motion, change, and
space. Most of the initial work involving calculus was directed toward the study of physics; indeed,
many of the great mathematicians of the period are also regarded as physicists. But from about
the middle of the 18th century there was an increasing interest in mathematics itself, not just
its applications, as mathematicians sought to understand what lay behind the enormous power
that calculus gave to humankind. Here the old Greek tradition of formal proof came back into
ascendancy, as a large part of present-day pure mathematics was developed. By the end of the 19th
century, mathematics had become the study of number, shape, motion, change, and space, and of
the mathematical tools that are used in this study.
The explosion of mathematical activity that has taken place over the past 100 years or so has been
dramatic. The growth has not just been a further development of previous mathematics; many quite
new branches have sprung up. At the start of the 20th century, mathematics could reasonably be
regarded as consisting of about 12 distinct subjects: arithmetic, geometry, calculus, and so on.
Today, between 60 and 70 distinct categories would be a reasonable figure. Some subjects, such as
algebra or topology, have split into various subfields; others, such as complexity theory or dynamical
systems theory, are completely new areas of study.
Given this tremendous growth in mathematical activity, for a while it seemed as though the only
simple answer to the question “What is mathematics?” was to say, somewhat fatuously, “It is
what mathematicians do for a living.” A particular study was classified as mathematics not so
much because of what was studied but because of how it was studied –that is, the methodology used.
It was only in the 1980s that a definition of mathematics emerged on which most mathematicians
now agree: mathematics is the science of patterns. What the mathematician does is examine
abstract patterns – numerical patterns, patterns of shape, patterns of motion, patterns of behavior,
voting patterns in a population, patterns of repeating chance events, and so on. Those patterns can
be either real or imagined, visual or mental, static or dynamic, qualitative or quantitative, purely
utilitarian or of little more than recreational interest. They can arise from the world around us,
from the depths of space and time, or from the inner workings of the human mind. Different kinds
of patterns give rise to different branches of mathematics. For example:
2
• Arithmetic and number theory study the patterns of numbers and counting.
• Geometry studies the patterns of shape.
• Calculus allows us to handle patterns of motion.
• Logic studies patterns of reasoning.
• Probability theory deals with patterns of chance.
• Topology studies patterns of closeness and position.
and so forth.
One aspect of modern mathematics that is obvious to even the casual observer is the use of ab-
stract notations: algebraic expressions, complicated-looking formulas, and geometric diagrams. The
mathematicians reliance on abstract notation is a reflection of the abstract nature of the patterns
she studies.
Different aspects of reality require different forms of description. For example, the most appropriate
way to study the lay of the land or to describe to someone how to find their way around a strange
town is to draw a map. Text is far less appropriate. Analogously, line drawings in the form of
blueprints are the appropriate way to specify the construction of a building. And musical notation
is the most appropriate medium to convey music, apart from, perhaps, actually playing the piece.
In the case of various kinds of abstract, formal patterns and abstract structures, the most appro-
priate means of description and analysis is mathematics, using mathematical notations, concepts,
and procedures. For instance, the symbolic notation of algebra is the most appropriate means of
describing and analyzing general behavioral properties of addition and multiplication.
For example, the commutative law for addition could be written in English as:
Such is the complexity and the degree of abstraction of the majority of mathematical patterns,
that to use anything other than symbolic notation would be prohibitively cumbersome. And so the
development of mathematics has involved a steady increase in the use of abstract notations.
3
The first systematic use of a recognizably algebraic notation in mathematics seems to have been
made by Diophantus, who lived in Alexandria some time around 250AD. His treatise Arithmetica,
of which only six of the original thirteen volumes have been preserved, is generally regarded as the
first algebra textbook. In particular, Diophantus used special symbols to denote the unknown in
an equation and to denote powers of the unknown, and he employed symbols for subtraction and
for equality.
These days, mathematics books tend to be awash with symbols, but mathematical notation no more
is mathematics than musical notation is music. A page of sheet music represents a piece of music;
the music itself is what you get when the notes on the page are sung or performed on a musical
instrument. It is in its performance that the music comes alive and becomes part of our experience;
the music exists not on the printed page but in our minds. The same is true for mathematics;
the symbols on a page are just a representation of the mathematics. When read by a competent
performer (in this case, someone trained in mathematics), the symbols on the printed page come
alive – the mathematics lives and breathes in the mind of the reader like some abstract symphony.
Given the strong similarity between mathematics and music, both of which have their own highly
abstract notations and are governed by their own structural rules, it is hardly surprising that many
(perhaps most) mathematicians also have some musical talent. Although some commentators make
more of this connection in terms of mental ability than I believe is warranted, it is true that for
most of the 2,500 years of Western civilization, starting with the ancient Greeks, mathematics and
music were regarded as two sides of the same coin. It was only with the rise of the scientific method
in the 17th century that the two started to go their separate ways.
For all the historical connections, however, there was, until recently, one very obvious difference
between mathematics and music. Although only someone well trained in music can read a musical
score and hear the music in her head, if that same piece of music is performed by a competent
musician, anyone with a sense of hearing can appreciate the result. It requires no musical training
to experience and enjoy music when it is performed.
For most of its history, however, the only way to appreciate mathematics was to learn how to “sight-
read” the symbols. Although the structures and patterns of mathematics reflect the structure of,
and resonate in, the human mind every bit as much as do the structures and patterns of music,
human beings have developed no mathematical equivalent to a pair of ears. Mathematics can only
be “seen” with the “eyes of the mind”. It is as if we had no sense of hearing, so that only someone
able to sight-read musical notation would be able to appreciate the patterns and harmonies of music.
Without its algebraic symbols, large parts of mathematics simply would not exist. Indeed, the issue
is a deep one having to do with human cognitive abilities. The recognition of abstract concepts and
the development of an appropriate language are really two sides of the same coin.
The use of a symbol such as a letter, a word, or a picture to denote an abstract entity goes hand
in hand with the recognition of that entity as an entity. The use of the numeral “7” to denote the
number 7 requires that the number 7 be recognized as an entity; the use of the letter m to denote
an arbitrary whole number requires that the concept of a whole number be recognized. Having the
symbol makes it possible to think about and manipulate the concept.
4
This linguistic aspect of mathematics is often overlooked. Indeed, one often hears the complaint that
mathematics would be much easier if it weren’t for all that abstract notation, which is rather like
saying that Shakespeare would be much easier to understand if it were written in simpler language.
Sadly, the level of abstraction in mathematics, and the consequent need for notations that can cope
with that abstraction, means that many, perhaps most, parts of mathematics will remain forever
hidden from the non-mathematician; and even the more accessible parts may be at best dimly
perceived, with much of their inner beauty locked away from view.
Inner beauty? What do I mean by that? In what sense is mathematics beautiful? Mathemat-
ical beauty is hard to convey to an outsider. Apart from computer graphical representations of a
few mathematical objects, where there is a visual beauty plain for all to see, mathematical beauty
is highly abstract — a beauty of abstract form, structure, and logic.
The mathematicians patterns, like the painters or the poets, must be beautiful, the ideas, like the colors or
the words, must fit together in a harmonious way. Beauty is the first test; there is no permanent place in
the world for ugly mathematics. . . . It may be very hard to define mathematical beauty, but that is just as
true of beauty of any kind we may not know quite what we mean by a beautiful poem, but that does not
prevent us from recognizing one when we read it.
The beauty to which Hardy was referring is a highly abstract, inner beauty, a beauty of abstract
form and logical structure, a beauty that can be observed, and appreciated. It is a beauty “cold
and austere”, according to Bertrand Russell, the famous English mathematician and philosopher,
who wrote in his 1918 book, Mysticism and Logic:
Mathematics, rightly viewed, possesses not only truth, but supreme beauty— a beauty cold and austere, like
that of sculpture, without appeal to any part of our weaker nature, without the gorgeous trappings of
painting or music, yet sublimely pure, and capable of a stern perfection such as only the greatest art can
show.
Mathematics, the science of patterns, is a way of looking at the world, both the physical, biological,
and sociological world we inhabit and the inner world of our minds and thoughts. Mathematics
greatest success has undoubtedly been in the physical domain, where the subject is rightly referred
to as both the queen and the servant of the (natural) sciences. Yet, as an entirely human creation,
the study of mathematics is ultimately a study of humanity itself. For none of the entities that
form the substrate of mathematics exists in the physical world; the numbers, the points, the lines
and planes, the surfaces, the geometric figures, the functions, and so forth are pure abstractions
that exist only in humanitys collective mind. The absolute certainty of a mathematical proof and
the indefinitely enduring nature of mathematical truth are reflections of the deep and fundamental
status of the mathematicians patterns in both the human mind and the physical world.
5
In an age when the study of the heavens dominated scientific thought, Galileo said,
The great book of nature can be read only by those who know the language in which it was written. And
this language is mathematics.
Striking a similar note in a much later era, when the study of the inner workings of the atom had
occupied the minds of many scientists for a generation, the Cambridge physicist John Polkinhorne
wrote, in 1986,
Mathematics is the abstract key which turns the lock of the physical universe.
We have answered the question “What is mathematics?” with the catch phrase “Mathematics
is the science of patterns.” The second fundamental question about mathematics — “What
does it do for us?”—can also be answered with a catchy phrase: “Mathematics makes the
invisible visible.”
Let me give you some examples of what I mean by this answer. Without mathematics, there is no
way you can understand what keeps a jumbo jet in the air. As we all know, large metal objects
don’t stay above the ground without something to support them. But when you look at a jet
aircraft flying overhead, you can’t see anything holding it up. It takes mathematics to “see” what
keeps an airplane aloft. In this case, one way to “see” the invisible is an equation discovered by
the mathematician Daniel Bernoulli early in the 18th century.
While I’m on the subject of flying, what is it that causes objects other than aircraft to fall to the
ground when we release them? “Gravity,” you answer. But that’s just giving it a name. It doesn’t
help us to understand it. It’s still invisible. We might as well call it magic. To understand it,
you have to “see” it. Thats exactly what Newton did with his equations of motion and mechanics
in the 17th century. Newton’s mathematics enabled us to “see” the invisible forces that keep the
earth rotating around the sun and cause an apple to fall from the tree onto the ground.
Both Bernoulli’s equation and Newton’s equations use calculus. Calculus works by making visible
the infinitesimally small. That’s another example of making the invisible visible.
6
Here are some human patterns:
• Aristotle used mathematics to try to “see” the invisible patterns of sound that we recognize
as music.
• He also used mathematics to try to describe the invisible structure of a dramatic performance.
• In the 1950s, the linguist Noam Chomsky used mathematics to “see” and describe the in-
visible, abstract patterns of words that we recognize as a grammatical sentence. He thereby
turned linguistics from a fairly obscure branch of anthropology into a thriving mathematical
science.
• Probability theory and mathematical statistics let us predict the outcomes of elections, often
with remarkable accuracy.
• Market analysts use various mathematical theories to try to predict the future behavior of the
stock market.
• Insurance companies uses statistics and probability theory to predict the likelihood of an
accident during the coming year, and set their premiums accordingly.
That then, is the big picture of mathematics. Hopefully now you have a good overall sense of what
the subject is and why it is important, even if you did not previously. But what does it mean to
do mathematics? What is it that you, the beginning student of post-calculus math, must learn to
do? Here too, the answer we give today is not the one that would have been given in the past.
Up to about 150 years ago, although mathematicians had long ago expanded the realm of objects
they studied beyond numbers and algebraic symbols for numbers, they still regarded mathematics
as primarily about calculation. That is, proficiency in mathematics depended above all at being
able to carry out calculations or manipulate symbolic expressions to solve problems. By and large,
high school mathematics is still very much based on that earlier tradition.
In the middle of the 19th century, however, a revolution took place. One of its epicenters was
the small university town of Gottingen in Germany, where the local revolutionary leaders were
the mathematicians Lejeune Dirichlet, Richard Dedekind, and Bernhard Riemann. In their new
conception of the subject, the primary focus was not performing a calculation or computing an
answer, but formulating and understanding abstract concepts and relationships. This was a shift in
7
emphasis from doing to understanding. Within a generation, this revolution would completely
change the way pure mathematicians thought of their subject. Nevertheless, it was an extremely
quiet revolution that was recognized as such only when it was all over. It is not even clear that the
leaders knew they were spearheading a major change.
For the Gottingen revolutionaries, mathematics was about “thinking in concepts”(Denken in Be-
griffen). Mathematical objects were no longer thought of as given primarily by formulas, but
rather as carriers of conceptual properties. Proving was no longer a matter of transforming terms
in accordance with rules, but a process of logical deduction from concepts.
This new approach to mathematics now permeates all university mathematics instruction beyond
calculus. And by and large, it is the post-Gottingen approach to mathematics that this is all about.
Among the new concepts that the revolution embraced are many with which today’s university
mathematics student must come to grips.
In modern, pure mathematics we are primarily concerned with statements about mathematical
objects. Mathematical objects are things such as integers, real numbers, sets, functions, etc.
Examples of mathematical statements are:
Not only are we interested in statements of the above kind, we are, above all, interested in knowing
which statements are true and which are false. (The truth or falsity in each case is demonstrated
not by measurement or experiment, as in most other sciences, but by a proof, of which more in due
course).
Clearly, before we can prove whether certain statements are true or false, we must be able to
understand precisely what a statement says. Above all, mathematics is a very precise subject,
where exactness of expression is required. This already creates a difficulty, for in real life our use
of language is rarely precise. Now, systematically to make the entire English language precise
(by defining exactly what each word is to mean) would be an impossible task. It would also be
unnecessary. It turns out that by deciding exactly what we mean by a few simple words, we can
obtain enough precision to enable us to make all (or at least most) of the mathematical statements
we need. The point is that in mathematics, we do not use all of the English language. Indeed, when
we restrict our attention to mathematical statements themselves (as opposed to our attempts to
explain them), we need only a very small part of the English language.
8
Chapter 2
Mathematical Discourse
—Srinivasa Ramanujan
Is the written and spoken language used by mathematicians and students of mathematics for com-
municating about mathematics and for communicating mathematical reasoning. The discourse is
also used to describe their own behaviour when doing mathematics and to describe their attributes
towards various aspects of mathematics. This is “communication” in a broad sense, including not
only communication of definitions and proofs but also communication about approaches to problem
solving, typical errors, and attitudes and behaviors connected with doing mathematics.
When communicating mathematical reasoning and facts, mathematicians speak and write in a spe-
cial register of the language suitable for communicating mathematical arguments. Uses special
technical words, as well as ordinary words, phrases and grammatical constructions with special
meanings that may differ from their meaning in ordinary English. It is typically mixed with ex-
pressions from the symbolic language (below).
9
2.1.2 Symbolic Language
This is a special purpose language. This consists of the symbolic expressions and statements used
d
in calculations and presentation of results. For example, the statement, dx sin x = cos x is part of
the symbolic language, whereas,“ The derivative of the sine function is the cosine function”, is not
part of it.
This consists of expressions such as conceptual proofs and intuitive. These communicate something
about the process of doing mathematics but do not themselves communicate mathematics.
There is a standard interpretation of the mathematical register, including the symbolic language,
in the sense that at least most of the time most mathematicians would agree on the meaning of
most statements made in the register. Students have various other interpretations of particular
constructions used in the mathematical register. One of their tasks as students is to learn how to
extract the standard interpretation from what is said and written. One of the tasks of lecturers is
to teach them how to do that.
Words and phrases in the mathematical register that name mathematical objects, relations or
properties. It includes the words, for example, divide, equivalence relation, function, positive.
Words, phrases and more elaborate syntactic constructions of the mathematical register that com-
municate the logical structure of a mathematical argument, for example, if, let, thus.
10
2.3.2 Types of Prose
Theorem : A theorem is a mathematical statement for which the truth can be established using
logical reasoning on the basis of certain assumptions that are explicitly given or implied in the
statement i.e., by constructing a proof. The word theorem shares its Greek root with the word
theater. Both words are derived from the root thea, which means “the act of seeing.” Indeed, the
proof of a theorem usually allows us to see further into the subject we are studying.
Lemma : A lemma is an auxiliary theorem proved beforehand so it can be used in the proof of
another theorem. This word comes from the Greek word that means “to grasp.” Indeed, in a lemma
one “grasps” some truth to be used in the proof of a larger result. The proofs of some theorems
are long and difficult to follow. In these cases, it is common for one or more of the intermediate
steps to be isolated as lemmas and to be proved ahead. Then, in the proof of the theorem we can
refer to the lemmas already established and use them to move to the next step. Often the results
stated in lemmas are not very interesting by themselves, but they play key roles in the proof of
more important results. On the other hand, some lemmas are used in so many different cases and
are so important that they are named after famous mathematicians.
Corollary : A corollary is a theorem that follows logically and easily from a theorem already proved.
Corollaries can be important theorems. The name, which derives from the Latin word for “little
garland,” underlines the fact that the result stated in a corollary follows naturally from another
theorem. The James and James Mathematics Dictionary defines a corollary as a “by-product of
another theorem.”
11
Chapter 3
—Georg Cantor
The purpose of this chapter is twofold: to provide an introduction to, or review of, the terminology,
notation, and basic properties of sets, and, perhaps more important, to serve as a starting point for
our primary goal — the development of the ability to discover and prove mathematical theorems.
The emphasis in this chapter is on discovery, with particular attention paid to the kinds of evidence
(e.g., specific examples, pictures) that mathematicians use to formulate conjectures about general
properties.
A set is an unordered collection of objects, and as such a set is determined by the objects it contains.
Before the 19th century it was uncommon to think of sets as completed objects in their own right.
Mathematicians were familiar with properties such as being a natural number, or being irrational,
but it was rare to think of say the collection of rational numbers as itself an object. (There were
exceptions. From Euclid mathematicians were used to thinking of geometric objects such as lines
and planes and spheres which we might today identify with their sets of points.) In the mid 19th
century there was a renaissance in Logic. For thousands of years, since the time of Aristotle and
before, learned individuals had been familiar with syllogisms as patterns of legitimate reasoning,
for example:
12
But syllogisms involved descriptions of properties. The idea of pioneers such as Boole was to assign
a meaning as a set to these descriptions. For example, the two descriptions “is a man” and “is
a male homo sapiens” both describe the same set, the set of all men. It was this objectification
of meaning, understanding properties as sets, that led to a rebirth of Logic and Mathematics in
the 19th century. Cantor took the idea of set to a revolutionary level, unveiling its true power.
By inventing a notion of size of set he was able compare different forms of infinity and, almost
incidentally, to shortcut several traditional mathematical arguments.
But the power of sets came at a price; it came with dangerous paradoxes. The work of Boole and
others suggested a programme exposited by Frege, and Russell and Whitehead, to build a foundation
for all of Mathematics on Logic. Though to be more accurate, they were really reinventing Logic
in the process, and regarding it as intimately bound up with a theory of sets. The paradoxes of
set theory were a real threat to the security of the foundations. But with a lot of worry and care
the paradoxes were sidestepped, first by Russell and Whiteheads theory of stratified types and then
more elegantly, in for example the influential work of Zermelo and Fraenkel. The notion of set is
now a cornerstone of Mathematics.
The formal development of set theory began in 1874 with the work of Georg Cantor (1845- 1918).
Since then, motivated particularly by the discovery of certain paradoxes (e.g., Russell’s paradox),
logicians have made formal set theory and the foundations of mathematics a vital area of mathe-
matical research, and mathematicians at large have incorporated the language and methods of set
theory into their work, so that it permeates all of modern mathematics.
The notion of set is a primitive, or undefined, term in mathematics, analogous to point and line
in plane geometry. Therefore, our starting point, rather than a formal definition, is an informal
description of how the term ”set” is generally viewed in applications to undergraduate mathematics.
Similar (but informal) words : collection, group, aggregate, bundle, ensemble, family, class.
Description : A set is a collection of objects which are called the members or elements of that
set. For example
2. The English alphabet may be viewed as the set of letters of the English language.
Sets can consist of elements of various nature : people, physical objects, numbers, signs,
other sets, e.t.c.
13
A set is an ABSTRACT object, its members do not have to be physically collected together for
them to constitute a set. The membership criteria for a set must in principle be well-defined, and
not vague. Sets can be finite or infinite.
Let’s look at different types of numbers that we can have in our sets.
1. Natural Numbers
The set of natural numbers is {1, 2, 3, 4, . . . } and is denoted by N.
2. Integers
The set of integers is {. . . , −4, −3, −2, −1, 0, 1, 2, 3, 4, . . . } and is denoted by Z. The Z symbol
comes from the German word, Zahlen, which means number. Define the non-negative
integers {0, 1, 2, 3, 4, . . . } often denoted by Z+ . All natural numbers are integers.
3. Rational Numbers
The set of rational numbers is denoted by Q and consists of all fractional numbers i.e., x ∈ Q
if x can be written in the form pq , where p, q ∈ Z with q 6= 0.
4. Real Numbers
The real numbers are denoted by R.
5. Complex Numbers
The complex numbers are denoted by C.
3.4 Notation
1. A, B, C, . . . for sets.
2. a, b, c, . . . or x, y, z, . . . for members.
3. b ∈ A, if b belongs to A.
4. c ∈
/ A, if c does not belong to A.
5. ∅ is used for the empty set. There is exactly one set, the empty set or null set, which has
no members at all.
6. A set with only one member is called a singleton or singleton set. for example, {x}.
14
3.5 Specification of Sets
One advantage of having an informal definition of the term set is that, through it, we can introduce
some other terminology related to sets. The term element is one example, and the notion well-
defined is another. The latter term relates to the primary requirement for any such description:
Given an object, we must be able to determine whether or not the object lies in the described set.
3. By defining a set of rules which generates (defines) its members (Recursive Rules).
List Notation
This is suitable for finite sets. It lists names of elements of a set, separated by commas and enclose
them in braces. For example
Two important facts are: (i) the order in which elements are listed is irrelevant and (ii) an object
should be listed only once in the list, since listing it more than once does not change the set. As an
example, the set {1, 1, 2} is the same as the set {1, 2} (so that the representation {1, 1, 2} is never
used) which, in turn, is the same as the set {2, 1}.
Predicate Notation
We describe a set in terms of one or more properties to be satisfied by objects in the set, and by
those objects only. Such a description is formulated in so-called set-builder notation, that is, in the
form A = {x|x satisfies some property or properties}, which we read “A is the set of all objects
x such that x satisfies . . . .”.
15
For example, {x|x is a natural number and x < 8}. Reading : the set of all x such that x is
a natural number and is less than 8.
For example, (i) {x|x is a positive number} (ii) {x|x is a letter of the Russian alphabet}.
In all these examples the vertical line is read “such that” and the set is understood to consist of all
objects satisfying the preceding description, and only those objects.
Recursive Rules
The collection, out of which all sets under consideration may be formed, is called the universe
of discourse or universal set, denoted by U. For our purposes a universal set is the set of all
objects under discussion in a particular setting. A universal set will often be specified at the start
of a problem involving sets, whereas in other situations a universal set is more or less clearly, but
implicitly, understood as background to a problem. The role then of a universal set is to put some
bounds on the nature of the objects that can be considered for membership in the sets involved in
a given situation.
It is in connection with the description method that “well definedness” comes into play. The rule
or rules used in describing a set must be (i) meaningful, that is, use words and/or symbols with an
understood meaning and (ii) specific and definitive, as opposed to vague and indefinite.
So like G = {x|x is a goople} or E = {x|x% & 3} or Z = {x|x is a large state in the U nited States}
do not define sets. The descriptions of G and E involve nonsense symbols or words, while the
description of Z gives a purely subjective criterion for membership.
We have that the fundamental property of a set is that we can assert of each object whether or not
it is a member of the set.
Consider a set constructed by asserting of each object that it is not a member of the set. This set
has no members and is therefore called the empty set.
16
Definition 3.6.1. The null or empty set is the set that does not contain any elements, denoted
by the Scandinavian letter ∅ = {} = {x|x 6= x}.
Two sets are identical if and only if (iff) they have the same elements or both are empty. So A = B
iff, for every x, x ∈ A ⇔ x ∈ B.
Example 3.7.1. {0, 2, 4} = {x|x is an even positive integer less than 5}.
As the above example shows, equality of sets does not mean they have identical definitions; there
are often many different ways of describing the same set. The definition of equality reflects rather
the fact that a set is just a collection of objects.
If we have to prove that the sets A and B are equal, it is often quite difficult to prove in one go
that they have the same elements. What is usually done is to split the proof into two parts:
The number of elements in a set A is called the cardinality of A, denoted by |A|. The cardinality
of a finite set is a natural number. Infinite sets also have cardinalities but they are not natural
numbers. The set A is said to be countable or enumerable if there is a way to list the elements
of A.
A paradox (antimony) is an apparently true statement that seems to lead to a logical self-
contradiction.
Its important to note that any given property, P (x) does not necessarily determine a set, i.e., we
cannot say that given an arbitrarily property P , there corresponds a set whose elements satisfy the
property P .
17
Consider the following, There was once a barber. Wherever he lived, all of the men in this town
either shaved themselves or were shaved by the barber. And this barber only shaved the men who
did not shave themselves. Did the barber shave himself?
Let’s say that he did shave himself. But from this he shaved only the men in town, who did not
shave themselves, therefore, he did not shave himself.
But we see that every men in town either shaved himself or was shaved by the barber. So he did
shave himself. We have a contradiction.
Russell observed that if S is a set, then either S ∈ S or S ∈/ S, since a given object is either a
member of a given set or is not a member of that set. Consider the set of all sets that are not
members of themselves, R = {x|x is a set and x ∈ / x}. R is an object, either R ∈ R or R ∈/ R.
In both cases we have inferred the paradox that R ∈ R iff R ∈ / R. In other words, the assumption
that R is a set has led to a contradiction and therefore there is no such thing, then, as the set of
all sets. To avoid unnecessary paradoxes, we assume the existence of the universal set, U. All this
leads to the following problems
After this paradox was described, set theory had to be reformulated axiomatically as axiomatic
set theory.
3.9 Inclusion
Definition 3.9.1. Having fixed our universal set, U, then for all x ∈ U. If A and B are sets (with
all members in U), we write A ⊆ B or B ⊇ A iff x ∈ A =⇒ x ∈ B. (⊆ , set inclusion symbol)
18
Theorem 3.9.1. If A ⊆ B and B ⊆ C then A ⊆ C.
Proof. Let x ∈ A, then since A ⊆ B, we have x ∈ B and given that B ⊆ C, we conclude that
x ∈ C, thus A ⊆ C.
Example 3.9.1. (i) {a, b} ⊆ {d, a, b, e} (ii) {a, b} ⊆ {a, b} (iii) {a, b} ⊂ {d, a, b, e}
(iv) {a, b} 6⊂ {a, b}.
Note that the empty set is a subset of every set, ∅ ⊆ A, for every set A and that for any set A, we
have A ⊆ A.
The set of all subsets of A is called the power set of A and is denoted by P(A) and |P(A)| = 2|A|
where |A| is finite.
Example 3.10.1. If A = {a, b}, then P(A) = {∅, {a}, {b}, {a, b}}.
Just as there is an “algebra of numbers” based on operations such as addition and multiplication,
there is also an algebra of sets based on several fundamental operations of set theory. We develop
properties of set algebra later in this chapter; for now our goal is to introduce the operations by
which we are able to combine sets to get another set, just as in arithmetic we add or multiply
numbers to get a number.
19
3.11.1 Union and Intersection
Let A and B be arbitrary sets. The union of A and B, written A ∪ B, is the set whose elements
are just the elements of A or B or both.
A ∪ B := {x|x ∈ A or x ∈ B}.
Example 3.11.1. Let K = {a, b}, L = {c, d}, M = {b, d}, then K ∪ L = {a, b, c, d},
K ∪ M = {a, b, d}, L ∪ M = {b, c, d}, (K ∪ L) ∪ M = K ∪ (L ∪ M ) = {a, b, c, d}, K ∪ K = K,
K ∪ ∅ = ∅ ∪ K = K = {a, b}.
The intersection of A and B, written A ∩ B, is the set whose elements are just the elements of
both A and B.
A ∩ B := {x|x ∈ A and x ∈ B}.
Example 3.11.2. K ∩ L = ∅, K ∩ M = {b}, L ∩ M = {d}, (K ∩ L) ∩ M = K ∩ (L ∩ M ) = ∅,
K ∩ K = K, K ∩ ∅ = ∅ ∩ K = ∅.
Observe also that the sets that result from the operation of union tend to be relatively large, whereas
those obtained through intersection are relatively small.
20
3.13 Difference and Complement
Definition 3.13.1. A minus B written A \ B or A − B, which subtracts from A all elements which
are in B. (also called relative complement, or the complement of B relative to A)
A − B := {x|x ∈ A and x ∈
/ B}.
The operation, complement, is unary rather than binary; we obtain a resultant set from a single
given set rather than from two such sets. The role of the universal set is so important in calculating
complements that we mention it explicitly in the following definition. The complement of a set
A, is the set of elements which do not belong to A, i.e., the difference of the universal set U and A.
A0 = {x|x ∈ U and x ∈
/ A} or A0 = U − A.
The complement of a set consists of all objects in the universe at hand that are not in the given set.
Clearly the complement of A is very much dependent on the universal set, as well as on A itself. If
A = {1}, then A0 is one thing if U = N, something quite different if U = R.
Example 3.13.2. Let E = {2, 4, 6, . . . }, the set of all even numbers. Then E c = {1, 3, 5, . . . }, the
set of odd numbers.
A simple and instructive way of illustrating the relationship between sets in the use of the so called
Venn-Euler diagrams or simply Venn diagrams.
21
A∩B
A B
A∩B
A B
A∪B
A B
A−B
A B
B−A
A B
22
3. Associative Laws (i) (X ∪ Y ) ∪ Z = X ∪ (Y ∪ Z) (ii) (X ∩ Y ) ∩ Z = X ∩ (Y ∩ Z).
4. Distributive Laws (i) X ∪(Y ∩Z) = (X ∪Y )∩(X ∪Z) (ii) X ∩(Y ∪Z) = (X ∩Y )∪(X ∩Z).
Example 3.16.1. Let A = {a, b, c, d, e} and B = {d, e, f, g, h, i}, so that A∪B = {a, b, c, d, e, f, g, h, i}
and A ∩ B = {d, e}. Since |A| = 5, |B| = 6, |A ∪ B| = 9, |A ∩ B| = 2, we have
|A ∪ B| = |A| + |B| − |A ∩ B| = 5 + 6 − 2 = 9.
23
3.17 The Algebra of Sets
We have considered the problem of showing that two sets are the same, however this technique
becomes tedious should the expressions involved be at all complicated. We shall develop an algebra
of sets, to assist us in simplifying a given expression. The following basic laws are easily established.
Law 1 : (Ac )c = A Law 2 : A ∪ B = B ∪ A Law 3 : A ∩ B = B ∩ A
Law 4 : A ∪ (B ∩ C) = (A ∪ B) ∪ C Law 5 : A ∩ (B ∩ C) = (A ∩ B) ∩ C
Law 6 : A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C) Law 7 : A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C)
Law 8 : (A ∪ B)c = Ac ∩ B c Law 9 : (A ∩ B)c = Ac ∪ B c Law 10 : U c = ∅
Law 11 : ∅c = U Law 12 : A ∪ ∅ = A Law 13 : A ∪ U = U Law 14 : A ∩ U = A
Law 15 : A ∩ ∅ = ∅ Law 16 : A ∪ Ac = U Law 17 : A ∩ Ac = ∅.
Proof.
A ∪ (B ∩ Ac ) = (A ∪ B) ∩ (A ∪ Ac ) by Law 6
= (A ∪ B) ∩ U by Law 16
= A ∪ B by Law 14.
Definition 3.18.1. Let n be any natural number and let a1 , a2 , . . . , an be any objects. Then
(a1 , a2 , . . . , an ) denotes the ordered n-tuple with first term a1 , second term a2 , . . . and nth term
an .
Example 3.18.1. (5, 7) denotes the ordered pair whose first term is 5 and second term 7. Note
that (5, 7, 2) is called an ordered triple, (5, 7, 2, 4) is called an ordered 4-tuple.
The idea of a product of sets can be extended to any finite number of sets. For any sets A1 , A2 , . . . , An ,
the set of all ordered n-tuples (a1 , a2 , . . . , an ) where a1 ∈ A1 , a2 ∈ A2 , . . . , an ∈ An is called the
product of sets A1 , A2 , . . . , An and is denoted
n
Y
A1 × A2 × · · · × An or Ai .
i=1
The fundamental statement we can make about an ordered n-tuple is that a given object is the ith
term of an ordered n-tuple.
24
Definition 3.18.2. Let A and B be any non-empty sets, then
If A and B are both finite sets, then |A × B| = |A| · |B|. If A = B, we sometimes write A2 for
A × A.
Example 3.18.2. 1. If A = {1, 2} and B = {2, 3, 4}, then A×B = {(1, 2), (1, 3), (1, 4), (2, 2), (2, 3), (2, 4)}
and B × A = {(2, 1), (2, 2), (3, 1), (3, 2), (4, 1), (4, 2)}.
Notice that A × B 6= B × A, in general.
2. The Cartesian product R × R = R2 is the set of all ordered pairs of real numbers and this
represents the 2-dimensional Cartesian plane.
1. A × (B ∪ C) = (A × B) ∪ (A × C).
2. A × (B ∩ C) = (A × B) ∩ (A × C).
3. (A × B) ∩ (C × D) = (A ∩ C) × (B ∩ D).
4. (A × B) ∪ (C × D) ⊆ (A ∪ C) × (B ∪ D).
5. (A − B) × C = (A × C) − (B × C).
25
3.20 Tutorial Questions
1. Let A = {0, {1, 2}, 3}. Make a list of all elements of A and a (separate) list of all the subsets
of A.
5. Denote by A the set of square numbers, B the set of natural numbers divisible by 3, C the
set of natural numbers less than 50 and D the set {3, 4, 7, 8, 15, 20, 30, 81, 100}.
A ∪ (B ∩ C) 6= (A ∪ B) ∩ C.
9. Let A, B, C be sets and U be the universal set. Prove the following, (i) A ∪ B = B ∪ A (ii)
A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C) (iii) (A ∪ B)c = Ac ∩ B c (iv) A ∩ U = A.
10. In each of the following cases determine whether A ⊂ B, A = B or A * B. Give reasons for
your answers.
(i) A is the set of all integers, B is the set of all real numbers x such that p(x) = 0 for some
quadratic polynomial p.
(ii) A is the set of all integer multipliers of 2, B is the set of all integer multiples of −2.
(iii) A = {∅, 1, 2}, B = {∅, {2}, {∅}}.
26
11. Prove that for all sets A, B and C we have the identity
(A \ B) ∩ C = (A ∩ C) \ (B ∩ C).
(E ∩ F ) ∪ (F ∩ G) ∪ (G ∩ E) = (E ∪ F ) ∩ (F ∪ G) ∩ (G ∪ E).
14. For the sets A = {1, 3, 5, 7, 9, 11} and B = {2, 3, 5, 7, 11}. Determine the following numbers
(a) | A | (b) | B | (c) | A ∪ B | (d) | A | + | B | − | A ∩ B |.
16. Let X, Y be subsets of a universal set U. Prove, using De Morgan’s Laws, that
X 4 Y = (X ∪ Y ) ∩ (X ∪ Y )c .
17. The empty set is a subset of every set, ∅ ⊆ A for every set A. Why?
18. Let A, B and C are subsets of some given set. Prove that if C ∩ B = ∅, then
(A ∪ C) ∩ (A ∪ B) = A.
X \ (A ∪ B) = (X \ A) ∩ (X \ B).
Find A0 .
26. Show that (i) P(∅) = {∅} (ii) P(P(∅)) = {∅, {∅}}.
27
27. If | X |= n is finite, prove that | P(X) |= 2n .
28. Let A, B be sets and define P(X) to be the power set of the set X.
29. Show that if A is any set and {0, 1} a two-element set, then
31. Given that A, B and C are subsets of U, use the algebra of sets to prove the following state-
ments,
(a) (A ∪ B)c = Ac ∩ B (b) [A ∪ (B ∩ C)]c = Ac ∩ (B c ∪ C c )
(c) [(A ∪ B)c ∪ C]c = (A ∪ B) ∩ C c (d) ((A ∪ B)c ∪ C)c = (A ∪ B) ∩ C c
(e) ((A ∪ B)c ∪ A)c = B ∩ Ac (f) (A ∩ B ∩ C) ∪ (Ac ∪ B c ∪ C c ) = U
(g) A ∪ B ∪ (Ac ∩ B c ) = U (h) A ∪ A = A (i) A ∩ A = A (j) A ∪ (B ∩ A) = A (k)
A ∩ (B ∪ A) = A.
32. Let A, B, C be subsets of a universal set U. Prove the following statement using the laws of
the algebra of sets,
A × B = B × A iff A = B.
28
39. If X and Y are sets, prove that X × Y ⊆ P(P(X ∪ Y )).
29
Chapter 4
Relations
—Richard Dedekind
In natural language relations are a kind of links existing between objects. For example, mother of,
neighbour of, part of, is older than, is an ancestor of, e.t.c. In mathematics there are endless ways
that two entities can be related to each other. Consider the following mathematical statements.
30
5 < 10 5≤5 6= 5|80
5
x 6= y 6∈Z X⊆Y π ≈ 3.14.
In each case two entities appear on either side of a symbol, and we interpret the symbol as expressing
some relationship between the two entities. Symbols such as <, ≤, =, |, ≥, >, ∈ e.t.c are called
relations because they convey relationship among things.
Although the concept of a relation is certainly one you have met many times, both in mathematics
and in your everyday activities, what you might well find novel is to regard relations as mathematical
objects of study in their own right, rather than as concepts that relate the objects under study.
Given a set A, a relation on A is some property that is either true or false, for any ordered pair
(x, y) ∈ A × A.
Example 4.1.1. Let A = {eggs, milk, corn} and B = {cows, goats, hens}. We can define a
relation R from A to B by (a, b) ∈ R if a is produced by b. In other words
With respect to this relation eggs R hens, milk R cows and so on.
30
Example 4.1.2. “greater than” is a relation on Z, denoted by >. It is true that for the pair (3, 2)
but false for the pairs (2, 2) and (2, 3).
Definition 4.1.1. Given sets A and B, a relation R between A and B is a subset of A × B i.e.,
R ⊆ A × B.
A binary relation is a set of ordered pairs. Any subset of A × A is called a relation on A. Since a
relation R on A is a subset of A × A, it is an element of the power set of A × A i.e., R ⊆ P(A × A).
All the following expressions mean the same thing
1. x bears relationship R to y.
2. x and y are in the R relationship.
We use what is known as infix notation, whereby the proposition that the elements x, y of the
domain and codomain satisfy the relation R is written in the form
xRy.
Example 4.1.3. 1. Let A be the set of people and B the set of dogs. Define a relation R on
A × B by aRb. In this case a is related to an object b if and only if a owns b.
2. Let X = Y . The equality is a relation , we say xRy if x = y.
3. Let X = Y = R. Then ≤, <, ≥, > are all relations between R and R.
4. Let X = Y = Z. Then divisibility is a relation between Z and Z, we say xRy if x|y.
If R is a relation on A × B, we call the set A the domain of R and B the range of R i.e.,
domR = {a ∈ A|there exists some b ∈ B such that (a, b) ∈ R},
and
ranR = {b ∈ B|there exists some a ∈ A such that (a, b) ∈ R}.
fldR = domR ∪ ranR is called the field of R. Observe that domR, ranR and f ldR are all subsets
of 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.
Example 4.1.4. Let A = {1, 2, 3, 4, 5, 6} and define R by xRy if and only if x < y and x divides
y. So R = {(1, 2), (1, 3), . . . , (1, 6), (2, 4), (2, 6), (3, 6)}. So domR = {1, 2, 3}, ranR = {2, 3, 4, 5, 6}
and f ldR = A.
31
4.1.2 Inverse Relations
Every relation R from A to B has inverse relation R−1 from B to A, which is defined by
R−1 = {(b, a)|(a, b) ∈ R}.
bR−1 a if and only if aRb.
Example 4.1.5. Let A = {1, 2, 3} and B = {a, b}. Then R = {(1, a), (1, b), (3, a)} is a relation
from A to B. The inverse relation is R−1 = {(a, 1), (b, 1), (a, 3)}.
Venn diagrams and arrows can be used for representing relations between given sets.
Example 4.2.1. If A = {a, b, c, d} and B = {1, 2, 3, 4} and R = {(a, 1), (b, 1), (c, 2), (c, 3)} is a
relation from A to B.
A B
- 1
a -
b : 2
c - 3
d 4
In this diagram an arrow from x to y means that x is related to y. This kind of graph is called
directed graph or digraph. Another example is given which represents the divisibility relation
on the set {1, 2, 3, 4, 5, 6, 7, 8, 9}.
Its rows are labelled with elements of A and its column are labelled with the elements of B. If a ∈ A
and b ∈ B we write 1 ia a row a and column b if aRb, otherwise we write 0. From the example
32
above, R = {(a, 1), (b, 1), (c, 2), (c, 3)} has the following matrix
1 0 0 0
1 0 0 0
.
0 1 1 0
1 0 0 0
Definition 4.3.1. A relation R on a set A is called reflexive, if for all, a ∈ A, aRa. More
concisely, for all a ∈ A, (a, a) ∈ R.
All the values are related to themselves. For example, the relation of equality =, is reflexive, for all
numbers a ∈ R, a = a. So = is reflexive. ≤ is also reflexive (a ≤ a for any a ∈ R).
Example 4.3.1. Consider the following five relations on the set A = {1, 2, 3, 4} :
R1 = {(1, 1), (1, 2), (2, 3), (1, 3), (4, 4)}
R2 = {(1, 1), (1, 2), (2, 1), (2, 2), (3, 3), (4, 4)}
R3 = {(1, 3), (2, 1)}
R4 = ∅, empty relation
R5 = A × A, universal relation
33
Question: Determine which of the relations are reflexive.
Solution: The relation (3) is not reflexive since no line is perpendicular to itself. Also (4) is not
reflexive since no line is parallel to itself. The other relations are reflexive, that is, x ≤ x for every
x ∈ Z, A ⊆ A for any set A ∈ C and n | n for every positive integer n in N.
Example 4.3.3. Let V = {1, 2, 3} and R = {(1, 1), (2, 4), (4, 4)}. Then R is not a reflexive relation,
since (2, 2) does not belong R.
One should note that all ordered pairs (a, a) must belong to R in order for R to be reflexive.
Definition 4.3.2. A relation R on a set A is called symmetric, if for all, a, b ∈ A, aRb implies
bRa.
Neither ≤ nor < are symmetric (2 ≤ 3 and 2 < 3 but not 3 ≤ 2 nor 3 < 2 is true).
Example 4.3.4. (a) Determine which of the relations in Example 4.3.1 are symmetric.
Solution: R1 is not symmetric since (1, 2) ∈ R1 but (2, 1) ∈ / R1 . R3 is not symmetric since
(1, 3) ∈ R3 but (3, 1) ∈
/ R3 . The other relations are symmetric.
Solution: The relation ⊥ is symmetric since if line a is perpendicular to line b then b is perpen-
dicular to a. Also k is symmetric since if line a is parallel to line b then b is parallel to a. The
others are not symmetric. For example, 3 ≤ 4 but 4
3, {1, 2} ⊆ {1, 2, 3} but {1, 2, 3} * {1, 2}
and 2 | 6 but 6 - 2.
Example 4.3.5. Let P = {1, 2, 3, 4} and R = {(1, 3), (4, 2), (2, 4), (2, 3), (3, 1)}. Then R is not a
symmetric relation, since (2, 3) ∈ R but (3, 2) 6∈ R.
Definition 4.3.3. A relation R on a set A is called anti-symmetric, if for all, a, b ∈ A, aRb and
bRa implies a = b.
34
Example 4.3.6. Determine which of the relations in Example 4.3.2 are antisymmetric.
The relation ⊥ is not anti-symmetric since we cannot have distinct lines a and b such that a ⊥ b
and b ⊥ a. Similarly k is not anti-symmetric.
Definition 4.3.4. A relation R on a set A is called transitive, if for all a, b, c ∈ A, aRb, bRc
implies aRc.
Example 4.3.7. Determine which of the relations in Example 4.3.2 are transitive.
Solution: The relations ≤, ⊆ and | are transitive, that is, (i) If a ≤ b and b ≤ c then a ≤ c (ii)
If A ⊆ B and B ⊆ C then A ⊆ C (iii) If a | b and b | c then a | c. On the other hand the relation
⊥ is not transitive. If a ⊥ b and b ⊥ c, then it is not true that a ⊥ c.
For example, = on R is an equivalence relation, the classification of animals by species, that is,
the relation “ is of the same species as” is an equivalence relation on the set of animals and the
35
relation ⊆ of set inclusion is not an equivalence relation. It is reflexive and transitive, but it is not
symmetric since A ⊆ B does not imply B ⊆ A. Not all relations are equivalence relations.
Example 4.4.1. Let U = Z and define R = {(x, y)|x and y have the same parity}, i.e., x and y
are either both even or both odd. The parity is an equivalence relation.
3. If (x, y) ∈ R and (y, z) ∈ R, then x and z have the same parity as y, so they have the same
parity as each other (if y is odd, both x and z are odd, if y is even both x and z are even),
thus (x, z) ∈ R.
Example 4.4.2. For any set S, the identity relation on S, IS = {(x, x)|x ∈ S} is an equivalence
relation.
1. Obvious.
Example 4.4.3. Let U = R and define the square relation R = {(x, y)|x2 = y 2 }. Square relation
is an equivalence relation.
xDy ⇐⇒ 3 | (x2 − y 2 )
is an equivalence relation.
(i) Reflexive : For any x ∈ Z we have x2 − x2 = 0 and since 3 | 0 it follows that xDx for all x ∈ Z.
(ii) Symmetric : Suppose xDy. Then 3 | (x2 − y 2 ) so x2 − y 2 = 3n for some n ∈ Z. It follows that
y 2 − x2 = 3(−n) and hence 3 | (y 2 − x2 ). Consequently yDx, so D is symmetric.
(iii) Transitive : Suppose xDy and yDz. There there exists n, m ∈ Z such that x2 − y 2 = 3n and
y 2 − z 2 = 3m. It follows that x2 − (3m + z 2 ) = 3n or x2 − z 2 = 3m + 3n = 3(m + n) and so xDz,
that is, D is transitive.
36
Example 4.4.5. Modular Arithmetic
We say an integer a is congruent to another integer b modulo a positive integer n, denoted as,
a ≡ b mod n, if b − a is an integer multiple of n.
Let n = 3 and let A be the set of integers from 0 to 11. Then x ≡ y mod 3 if x and y belongs to
A0 = {0, 3, 6, 9} or both belong to A1 = {1, 4, 7, 10} or both belong to A3 = {2, 5, 8, 11}. Congruence
modulo 3 is in fact an equivalence relation on A.
Reflexive
Since x − x = 0 · 3 we know that x ≡ x mod 3.
Symmetric
If x ≡ y mod 3, then y − x = 3k for some integer k. Hence x − y = −3k and since −k is an
integer we have y ≡ x mod 3.
Transitive
Let x ≡ y mod 3 and y ≡ z mod 3. Then there are integers k and l such that y − x = 3k and
z − y = 3l. It follows that z − x = 3k + 3l = 3(k + l) and since k + l is an integer we have x ≡ z
mod 3.
Definition 4.5.1. Let A be a set and R an equivalence relation on A. Let a ∈ A, then the set of
all elements b ∈ A such that aRb is called an equivalence class.
Denoted by [x], meaning, equivalence class of x where the element x is said to be a representative
of the class.
0 ≡ 2 ≡ 4 ≡ ··· mod 2
1 ≡ 3 ≡ 5 ≡ ··· mod 2.
37
All even numbers are equivalent to each other under R. In set notation [0] = {0, 2, 4, · · · } and
[1] = {1, 3, 5, · · · }, the sets are called equivalence classes under modulo n.
We call the act of doing this grouping with respect to some equivalence relation partitioning
(partitioning a set A into equivalence classes under a relation R).
For example, we have partitioned Z into equivalence classes [0] and [1] under the relation of con-
gruence modulo 2. The collection of equivalence classes, denoted A/R = {[x]|x ∈ A} is called the
quotient set of A by R.
4.5.1 Partitions
Partitioning a set, in English means to break it into non-overlapping pieces. A partition of a set
A is the collection S of non overlapping non-empty subsets of A whose union is the whole A.
For instance a partition of A = {1, 2, 3, 4, 5, 6, 7, 8, 9, 10} could be S = {{1, 2, 4, 8}, {3, 6}, {5, 7, 9, 10}}.
Given a partition S of a set A, every element of A belongs to exactly one member of S.
Example 4.5.2. The division on the integers Z into even and odd numbers is a partition :
S = {E, O} where E = {2n|n ∈ Z} and O = {2n + 1|n ∈ Z}.
Example 4.5.3. The division of the integers Z into negative integers, positive integers and zero is
a partition : S = {Z+ , Z− , {0}}.
Definition 4.5.2. The partition of a set A is a collection of non-empty subsets A1 , A2 , · · · , of A
which are pairwise disjoint and whose union is A.
1. Ai is non-empty.
2. Ai ∩ Aj = ∅ for i 6= j.
3. ∪n An = A.
Proposition 4.5.1. Let A be a set. If R is an equivalence relation on A, then the equivalence
classes of R form a partition of A.
Let R be an equivalence relation on A and let {[x] : x ∈ A} be the family of equivalence classes of
R. Then [x] 6= ∅, for all x, because x ∈ [x].
For pairwise disjointness, we prove that [x] ∩ [y] 6= ∅ implies [x] = [y]. Let [x], [y] be so that there
is a z ∈ [x] ∩ [y]. Then xRz and zRy implies xRy. Now let u ∈ [x]. Then uRx and xRy so
uRy and u ∈ [y]. Hence [x] ⊆ [y] and we prove [y] ⊆ [x] similarly. Thus the equivalence classes
of R are pairwise disjoint. Finally, for ∪{[x] : x ∈ A} = A, first note that ∪{[x] : x ∈ A} ⊆ A
is clear. For the reverse containment, let x ∈ A. Then x ∈ [x] and [x] ⊆ ∪{[x] : x ∈ A} so
A ⊆ ∪{[x] : x ∈ A} = A.
38
Example 4.5.4. Let A = {1, 2, 3, . . . , 14, 15}. Let R be the equivalence relation on A defined by
congruence modulo 4.
Solution (a) Accordingly [1] = {1, 5, 9, 13}, [2] = {2, 6, 10, 14}, [3] = {3, 7, 11, 15}, [4] = {4, 8, 12}.
Then [1], [2], [3], [4] are all the equivalence classes since they all include all the elements of A.
(b) Choose an element in each equivalence class which is a multiple of 3. Thus B = {9, 6, 3, 12} or
B = {9, 6, 15, 12}.
Example 4.5.5. Let A be a set of non-zero integers and let ≈ be the relation on A × A defined as
follows :
(a, b) ≈ (c, d) whenever ad = bc.
Prove that ≈ is an equivalence relation.
(ii) Symmetry : Suppose (a, b) ≈ (c, d). Then ad = bc. Accordingly, cb = da and hence
(c, d) ≈ (a, b). Thus ≈ is symmetric.
(iii) Transitivity : Suppose (a, b) ≈ (c, d) and (c, d) ≈ (e, f ). Then ad = bc and cf = de. Multi-
plying corresponding terms of the equations gives (ad)(cf ) = (bc)(de). Cancelling c 6= 0 and d 6= 0
from both sides of the equation yields af = be, and hence (a, b) ≈ (e, f ). Thus ≈ is transitive.
Accordingly, ≈ is an equivalence relation.
Let A = {1, 2, 3, . . . , 14, 15}. Let ≈ be the equivalence relation on A × A defined by (a, b) ≈ (c, d) if
ad = bc. Find the equivalence class of (3, 2).
We seek all (m, n) such that (3, 2) ≈ (m, n), that is, such that 3n = 2m or 23 = m n
.[In other words,
if (3, 2) is written as the fraction 2 , then we seek all fractions n which are equal to 32 .] Thus
3 m
[(3, 2)] = {(3, 2), (6, 4), (9, 6), (12, 8), (15, 10)}.
One of the main properties of an equivalence relation on a set A is that the quotient set, i.e., the
collection of equivalence classes is a partition of A. An equivalence class of elements is sometimes
called a block.
Theorem 4.5.2. Let R be an equivalence relation on a set A. Then the blocks partition A.
39
4.6 Order Relations
Ordering to numbers.
Definition 4.6.1. A relation that is reflexive, anti-symmetric and transitive is called a par-
tial order.
For example, the relations, ≤, ≥ and | on the set Z as well as the relation ⊆ on the power set of
any set A are partial orders.
A set A with a partial order on A is called a partially ordered set or commonly a poset, and
write (A, ≤).
Used to visualize the structure of a partial ordering and it draws its transitive reduction. Here,
each element of A as a vertex, in the plane draws a line segment arc curve that goes upward from
x to y whenever y covers x (x < y and there is no such z such that x < z < y).
Example 4.6.1. 1. Divide ordering on the set A = {1, 2, 3, 4, 6, 8, 12, 24} where A is the set of
all elements of N which divide 24.
2. If A = P({1, 2, 3}), the power set of {1, 2, 3} and ≤ is the subset relation ⊆.
3. The set a = {1, 2, 3, 4, 5, 6, 10, 12, 15, 20, 30, 60} of all divisors of 60, partially ordered by di-
visibility
2. Consider the set N of natural numbers. Determine whether or not the following relations on
N are transitive
40
(i) reflexive, not transitive, not anti-symmetric.
(ii) not reflexive, not symmetric, not transitive.
(iii) symmetric and transitive.
(iv) both symmetric and anti-symmetric.
4. Which of these relations on {0, 1, 2, 3} are equivalence relations? Determine the properties of
an equivalence relation that the other lack
11. Suppose that A is a non empty set, and if f is a function that has A as its domain. Let R be
the relation on A consisting of all ordered pairs (x, y) where f (x) = f (y). Show that R is an
equivalence relation on A.
12. Consider the relation R = {(a, b) | a = b or a = −b}. How many equivalence classes are there?
13. Consider the relation R = {(a, b) | a ≡ b mod 2}. Write down the equivalence classes and the
partition of R.
14. Partition {x | 1 ≤ x ≤ 9} into equivalence classes under the equivalence relation xRy ⇔ x ≡
y mod 6.
41
15. Which of the following are partitions of the set of integers?
(i) The set of even integers and the set of odd integers.
(ii) The set of positive integers and the set of negative integers.
(iii) The set of integers divisible by 3, the set of integers leaving a remainder of 1 when divided
by 3 and the set of integers leaving a remainder of 2 when divided by 3.
(iv) The set of integers less −100, the set of integers with absolute value not exceeding 100
and the set of integers greater than 100.
(v) The set of integers not divisible by 3, the set of even integers and the set of integers that
leave a remainder of 3 when divided by 6.
16. A relation R is defined on Z by xRy ⇔ ∃n ∈ Z such that x = 2n y, (x, y ∈ N). Prove that R
is an equivalence relation.
19. Let A = {1, 2, . . . , 10} and define the relation R on A by xRy iff x is a multiple of y. Show
that R is a partial order on A and draw its diagram.
21. Let A be a family of sets. Prove that the relation on A defined by X is a subset of Y is a
partial order of A.
22. Draw the Hasse diagram if A = P({1, 2, 3}), the power set of {1, 2, 3} and ≤ is the subset
relation ⊆.
23. Draw the Hasse diagram for the set A = {1, 2, 3, 4, 5, 6, 10, 12, 15, 20, 30, 60} of all the divisors
of 60, partially ordered by divisibility.
42
Chapter 5
Functions
It is not worth an intelligent man’s time to be in the majority. By definition, there are already enough
people to do that.
—G.H Hardy
A complete history of the modern concept of a function would take some time and would have to
include reference to a great many mathematicians whose work contributed in one way or another.
But arguably the two most significant figures in the story are Jean Baptise Joseph Fourier, who lived
in France from 1768 to 1830, and Johann Peter Gustav Lejeune Dirichlet, a Belgian mathematician
who lived from 1805 to 1859.
Among a number of mathematicians inspired by Fouriers work to reflect on the nature of functions,
Dirichlet was the one who made the final decisive move that gave us the modern function concept.
It was in 1837 that Dirichlet proposed that mathematicians should stop thinking of functions as
computational processes, but instead concentrate simply on their overall behavior. He wrote:
If a variable y is so related to a variable x that whenever a numerical value is assigned to x, there is a rule
according to which a unique value of y is determined, then y is said to be a function of the independent
variable x.
For today’s beginning student of modern mathematics, the main lesson to be learned from all of
this is that the high degree of abstraction that you have to master, of which the general definition of
a function is just one example, did not come about easily, nor without cause. When you think that
Fourier’s work – which forced mathematicians to reconceptualize the notion of a function— lies at
the heart of modern music synthesizers, Internet transfers of music files, MP3 music players, and
43
a host of other applications, including the original questions of heat flow that motivated Fourier
himself, you realize that the abstraction was the result of some decidedly practical applications.
Let X and Y be sets. A function f : X → Y is a special kind of relation between X and Y . Its
a relation R ⊂ X × Y satisfying the following condition : for all x ∈ X, there exists exactly one
y ∈ Y such that (x, y) ∈ R.
Functions are also called mappings, maps, or transformations (although each of these words tends
to be reserved for special kinds of functions).
Definition 5.2.1. Let X and Y be sets. A function f from X to Y is a relation from X to Y such
that
(i) the set X is called the domain (source) of f and the set Y is called the co-domain (target)
of f .
(ii) range of f is the set
Range (f )= {y ∈ Y |there is an x ∈ X such that y = f (x)} = {f (x)|x ∈ X} = f (X). The
set f (X) is the image of X under f .
One of the requirements that any function must fulfill (in order to be a function) is that any argu-
ment produces at most one value (in fact, exactly one value). Although advanced-level mathematics
textbooks sometimes speak of “multiple-valued functions,” such terminology is restricted to highly
specialized purposes, and may be ignored here. For our purposes (and for practically all of mathe-
matics), no function can produce two different values for the same argument. If it does, it simply
is not a function.
Definition 5.2.3. Let f : X → Y be a function. Then
(i) f is said to be surjective (onto) if f (X) = Y i.e., f is surjective if and only if, for any
y ∈ Y there is an x ∈ X such that f (x) = y.
(ii) f is said to be injective (one-to-one, 1 − 1) if f maps different points of X to different
points of Y i.e., f is injective if and only if, for any x, x0 ∈ X, f (x) = f (x0 ) =⇒ x = x0 i.e.,
x 6= x0 then f (x) 6= f (x0 ).
(iii) A function f : X → X is called a bijection (or bijective) if f is both injective and
surjective. Two sets X and Y are said to be in 1 − 1 correspondence (i.e., to every element
of X there corresponds an element of Y and vice versa) if there is a bijection between them.
44
5.3 Composition of Functions
Example 5.3.1. If f (x) = x2 and g(x) = x+1, then g(f (x)) = x2 +1 whereas f (g(x)) = (x+1)2 =
x2 + 2x + 1.
Proof. (a) We must show that for all x1 , x2 ∈ X if g(f (x1 )) = g(f (x2 )), then x1 = x2 . But put
y1 = f (x1 ) and y2 = f (x2 ). Then g(y1 ) = g(y2 ). Since g is assumed to be injective, this implies
that f (x1 ) = y1 = y2 = f (x2 ). Since f is also assumed to be injective, this implies that x1 = x2 .
(b) We must show that for all z ∈ Z, there exists at least one x in X such that g(f (x)) = z. Since
g : Y → Z is surjective, there exists y ∈ Y such that g(y) = z. Since f : X → Y is surjective, there
exists x ∈ X such that f (x) = y. Then g(f (x)) = g(y) = z.
Definition 5.3.2. Let X be any set. Let a function f : X → X be defined by f (x) = x i.e., let f
mapping to each element in X, to itself.
45
5.3.2 Inverse of a Function
Let f : X → Y be a function and let y ∈ Y , then the inverse of y, denoted f −1 (y) consists of those
elements of X which are mapped onto y i.e., elements in X which have y as their image.
f −1 (y) = {x|x ∈ X and f (x) = y} = {x ∈ X|y = f (x)}.
f −1 (y) is a subset of X.
If f : X → Y is both one-to-one function and onto function, then f −1 : Y → X and call f −1 the
inverse function of f . We say that a function g : Y → X is the inverse function to f : X → Y
if both of the following hold :
(i) f is bijective.
(ii) the inverse relation f −1 ; Y → X is a function.
(iii) f has an inverse function g.
(b) When the equivalent conditions of part (a) hold then the inverse function g is uniquely deter-
mined and it is the function f −1 .
Assume (ii) i.e., the inverse relation f −1 is a function. We claim that it is then the inverse function
to f in the sense that f −1 ◦ f = IX and f ◦ f −1 = IY for some x ∈ X, f −1 (f (x)) is the unique
element of X which get mapped under f to f (x). Since x is such an element and the uniqueness is
assumed, we must have f −1 (f (x)) = x. Similarly, for y ∈ Y, f −1 (y) is the unique element x of X
such that f (x) = y, so f (f −1 (y)) = f (x) = y.
(iii) =⇒ (i)
(b) Suppose that we have any function g : Y → X such that g ◦ f = IX and f ◦ g = IY . We know
that f is bijective and thus the inverse relation f −1 is a function such that f −1 ◦f = IX , f ◦f −1 = IY .
Thus
g = g ◦ IY = g ◦ (f ◦ f −1 ) = (g ◦ f ) ◦ f −1 = IX ◦ f −1 = f −1 .
46
5.4 Tutorial Questions
1. Let the functions f : R → R and g : R → R be defined by f (x) = 2x + 1 and g(x) = x2 − 2.
Find the product functions g ◦ f and f ◦ g.
6. Two sets A and B are said to be equipollent if and only if there is a bijection f : A → B.
Show that for any two non-empty real intervals (a, b) and (c, d) are equipollent, if we take the
bijection f : (a, b) → (c, d), defined by
d−c bc − ad
f (x) = x+ .
b−a b−a
47
Chapter 6
Symbolic Logic
A persons first duty, a young persons at any rate, is to be ambitious, and the noblest ambition is that of
leaving behind something of permanent value.
—G.H Hardy
Uses words and phrases that have a bearing on the truth or falsity of the sentence in which they
occur. Such words or phrases are aptly called logical connectives. For example, not, or, and,
if, then, if and only if, . . . .
For example, consider the sentence : It is cold and the sun is shining. Sentence is obtained by
joining the two sentences : It is cold and The sun is shining. The resulting sentence is called a
compound sentence and is true provided that each of the two component sentences is true.
Definition 6.0.1. A proposition/statement is a declarative sentence which is true or false (but
not both).
Notation is useful in the study of compound statements. If we let p denote the statement “All cows
eat grass” and let q denote the statement “Columbus discovered America”, the we can write the
compound statement p and q.
6.1 Abbreviations
∧ denotes and , ⇐⇒ denotes if and only if , ∨ denotes or, =⇒ denotes if, ¬p denotes
not p.
For example, If p denotes the proposition “It is raining”, then ¬p denotes “It is not raining”. The
truthfulness or falsity of the statement is called its truth value. Denoting “true” by “T ” and
48
“false” by “F ”, the logical connectives are conveniently defined by means of a truth-table which
spells out the truth value of a compound statement in each possible truth-value cases.
6.1.1 Conjunction, p ∧ q
Two statements can be combined by the word “and” to form a composite statement which is called
the conjugation of the original statements.
Denoted by p ∧ q.
Example 6.1.1. 1. Let p be it is raining and q be it is overcast. Then p ∧ q denotes it is raining
and it is overcast.
2. The symbol ∧ can be used to define the intersection of two sets,
C ∩ D = {x|x ∈ C ∧ x ∈ D}.
Truth value of a composite statement satisfies the following property : If p is true and q is true,
the p ∧ q is true, otherwise, p ∧ q is false. Conjugation of two statements is true if and only if each
component is true.
Truth-table
p q p∧q
T T T
T F F
F T F
F F F
6.2 Disjunction, p ∨ q
Truth value of the composite statement p ∨ q satisfies the property : If p is true or q is true or both
p and q are true, then p ∨ q is true, otherwise, p ∨ q is false.
49
Truth-table
p q p∨q
T T T
T F T
F T T
F F F
6.3 Negation, ¬p
Given any statement p, another statement “not p”, called the negation of p and is denoted by ¬p.
Negation (i) It is false that Chinhoyi is in Zimbabwe or (ii) Chinhoyi is not in Zimbabwe.
Truth-table
p ¬p
T F
T F
F T
F T
If p then q, also read as (a) p implies q (b) p only if q (c) p is sufficient for q (d) q is
necessary for p.
The truth value of the conditional statement p =⇒ q satisfies the following property : p =⇒ q is
true unless p is true and q is false, a true statement cannot imply a false statement.
Truth-table
p q p =⇒ q
T T T
T F F
F T T
F F T
50
6.4.1 The Bi-conditional,p ⇐⇒ q
p if and only if q.
Truth value satisfied if : p and q have the same truth value, then p ⇐⇒ q is true, otherwise, it is
false.
Truth-table
p q p ⇐⇒ q
T T T
T F F
F T F
F F T
p q ¬p ¬p ∨ q (¬p ∨ q) =⇒ p
T T F T T
Example 6.4.1. 1. Find the truth values for (¬p∨q) =⇒ p. T F F F T
F T T T F
F F T T F
p q p =⇒ q ¬p ¬p ∨ q (p =⇒ p) ⇐⇒ (¬p ∨ q)
T T T F T T
2. Construct a truth table for (p =⇒ q) ⇐⇒ (¬p∨q). T F F F F T
F T T T T T
F F T T T T
Some sentences are not statements because they contain unspecified variables, for example, we
cannot assign a truth value to the sentence, He was a president of the United States, until a proper
name is substituted for the pronoun he. We call a sentence that contains unspecified variables a
predicate.
For example, S is green and X discovered America are predicates. S and X are unspecified variables
that may be replaced by various nouns. The predicate is neither true nor false, its truth value
depends upon the name that replaces X. In these examples, he, S and X are called free variables
in their respective predicates. A statement that is always false is called an absurdity. A statement
that may be true or false, depending upon the values of its constituent statements, is called a
contigency.
51
6.5 Logical Equivalence
Definition 6.5.1. The propositions p and q are said to be logically equivalent if their truth tables
are identical. Denoted by p ≡ q.
Example 6.5.1. ¬(p ∧ q) ≡ ¬p ∨ ¬q.
p q p∧q ¬(p ∧ q) p q ¬p ¬q ¬p ∨ ¬q
T T T F T T F F F
T F F T T F F T T
F T F T F T T F T
F F F T F F T T T
Consider the statement, “It is not the case that roses are red and violets are blue”. This
statement can be written in the form ¬(p ∧ q) where p is “roses are red” and q is “violets are blue”.
However, as noted above, ¬(p ∧ q) ≡ ¬p ∨ ¬q. Thus the statement, Roses are not red, or violets are
not blue, has the same meaning as the given statement.
Idempotent Laws
(1a) p ∨ p ≡ p (1b) p ∧ p ≡ p.
Associative Laws
(2a) (p ∨ q) ∨ r ≡ p ∨ (q ∨ r) (2b) (p ∧ q) ∧ r ≡ p ∧ (q ∧ r).
Commutative Laws
(3a) p ∨ q ≡ q ∨ p (3b) p ∧ q ≡ q ∧ p.
Distributive Laws
(4a) p ∨ (q ∧ r) ≡ (p ∨ q) ∧ (p ∨ r) (4b) p ∧ (q ∨ r) ≡ (p ∧ q) ∨ (p ∧ r).
Identity Laws
(5a) p ∧ T ≡ p (5b) p ∨ F ≡ p.
(6a) p ∨ T ≡ T (6b) p ∧ F ≡ F .
Complement Laws
(7a) p ∨ ¬p ≡ T (8a) ¬T ≡ F .
(7b) p ∧ ¬p ≡ F (8b) ¬F ≡ T .
Involution Law
(9) ¬¬p ≡ p.
De Morgan’s Laws
(10a) ¬(p ∨ q) ≡ ¬p ∧ ¬q (10b) ¬(p ∧ q) ≡ (¬p ∨ ¬q).
52
6.6 The Converse
p q p =⇒ q ¬p ¬q ¬q =⇒ ¬p
T T T F F T
T F F F T F
F T T T F T
F F T T T T
6. (p ∧ q) =⇒ p Decomposing a conjunction.
7. (p ∧ q) =⇒ q Decomposing a conjunction.
8. p =⇒ (p ∨ q) Constructing a disjunction.
9. q =⇒ (p ∨ q) Constructing a disjunction.
53
12. (p ∨ q) ⇐⇒ (q ∨ p) Commutative law for ∨.
13. (p =⇒ q) ⇐⇒ (¬p ∨ q) Conditional disjunction.
14. [(p ∨ q) ∧ ¬p] =⇒ q Disjunctive syllogism.
15. (p ∨ p) ⇐⇒ p Simplification.
Set A is called the domain of p(x) and the set Tp of all elements of A for which p(a) is true is
called the truth set of p(x), i.e.,
Tp = {x|x ∈ A, p(x) is true} or {x|p(x)}.
When A is some set of numbers, the condition p(x) has the form of an equation or inequality
involving the variable x.
Example 6.9.1. Find the truth set Tp of each propositional function p(x) defined on the set
P = {1, 2, 3 . . . }.
(a) Let p(x) be x + 2 > 7. Then Tp = {x|x ∈ P, x + 2 > 7} = {6, 7, 8, . . . }. Consisting of all integers
greater than 5.
(b) Let p(x) be x + 5 < 3. Then Tp = {x|x ∈ P, x + 5 < 3} = ∅, the empty set.
(c) Let p(x) be x + 5 > 1. Then Tp = {x|x ∈ P, x + 5 > 1} = P.
From the above example, shows that if p(x) is a propositional function defined on a set A, then
p(x) could be true for all x ∈ A, for some x ∈ A or for no x ∈ A.
Let p(x) be a propositional function defined on a set A. Consider the expression, (∀x ∈ A)p(x) or
∀x, p(x) which reads “For every x in A, p(x) is a true statement”, or simply “For all x, p(x)”.
The symbol ∀, (for all, for every) is called the universal quantifier. (∀x ∈ A)p(x) is equivalent
to the statement Tp = {x|x ∈ A, p(x)} = A, i.e., the truth set of p(x) is the entire set of A. If
{x|x ∈ A, p(x)} = A, then ∀x, p(x) is true, otherwise, ∀x, p(x) is false.
Example 6.10.1. 1. The proposition (∀n ∈ P)(n + 4 > 3) is true since {n|n + 4 > 3} =
{1, 2, 3, . . . } = P.
54
2. The proposition (∀n ∈ P)(n + 2 > 8) is false since {n|n + 2 > 8} = {7, 8, . . . } =
6 P.
Let p(x) be the propositional function defined on a set A. Consider the expression (∃x ∈ A)p(x) or
∃x, p(x), which reads “There exists an x in A such that p(x)” is a true statement or simply, “For
some x, p(x)”.
The symbol ∃ (there exists, for some, for at least one) is called the existential quantifier.
(∃x ∈ A)p(x) is equivalent to the statement Tp = {x|x ∈ A, p(x)} = 6 ∅, i.e., the truth set of p(x) is
not empty. If {x|p(x)} =
6 ∅ then ∃x, p(x) is true, otherwise, ∃x, p(x) is false.
Example 6.11.1. 1. The proposition (∃n ∈ P)(n + 4 < 7) is true since {n|n + 4 < 7} = {1, 2} =
6
∅.
2. The proposition (∃n ∈ P)(n + 6 < 4) is false since {n|n + 6 < 4} = ∅.
6.11.1 Notation
Let A = {2, 3, 5} and let p(x) be the sentence “x is a prime number” or simply x is prime. Then the
proposition “Two is prime and three is prime and five is prime”, can be denoted by p(2)∧p(3)∧p(5) or
∧(a ∈ A, p(a)), which is equivalent to the statement, “Every number in A is prime or ∀a ∈ A, p(a)”.
Similarly, the proposition, “Two is prime or three is prime or five is prime”, can be denoted by
p(2) ∨ p(3) ∨ p(5) or ∨(a ∈ A, p(a)), which is equivalent to the statement “At least one number in
A is prime or ∃a ∈ A, p(a)”.
Consider the statement “All Mathematics majors are male”. Its negation is either the following
equivalent statements
Symbolically, using M to denote the set of Mathematics major, the above can be written as
¬(∀x ∈ M )(x is male) ≡ (∃x ∈ M )(x is not male),
55
or when p(x) denotes “x is a male”, we have ¬(∀x ∈ M )p(x) ≡ (∃x ∈ M )¬p(x) or
¬∀x, p(x) ≡ ∃x, ¬p(x).
Theorem 6.12.1 (De Morgan).
i.e., (a) It is not true that, for all x ∈ A, p(x) is true and (b) There exists an x ∈ A such that p(x)
is false.
Theorem 6.12.2 (De Morgan).
i.e., (a) It is not true for some x ∈ A, p(x) is true and (b) For all x ∈ A, p(x) is false.
The following statements are also negatives of each other : There exists a college student who is 60
years old, Every college student is not 60 years old.
The opposite of “For all x, p(x) is true”, is “There exists x for which p(x) is not true”.
The opposite of “There exists x for which p(x) is true ”, is “For all x, p(x) is not true”.
For example, All rational numbers equal one, the opposite (negation) is, There exists a rational
number that does not equal one.
All eleven-legged crocodiles are orange with blue spots is true, if it was false, then there would exist
an eleven-legged crocodile that is not orange with blue spots.
6.13 Proofs
In Italy its said that it requires two men to make a good salad dressing; a generous man to add
the oil and a mean man the vinegar. Constructing proofs in mathematics is similar. Often a
tolerant openness and awareness is important in discovering or understanding a proof, while a
strictness and discipline is needed in writing it down. There are many different styles of thinking,
even amongst professional mathematicians, yet they can communicate well through the common
medium of written proof. Its important not to confuse the rigour of a well-written-down proof with
the human and very individual activity of going about discovering it or understanding it. Too much
of a straightjacket on your thinking is likely to stymie anything but the simplest proofs. On the
other hand too little discipline, and writing down too little on the way to a proof, can leave you
uncertain and lost.
When you cannot see a proof immediately (this may happen most of the time initially), it can
help to write down the assumptions and the goal. Often starting to write down a proof helps you
discover it. You may have already experienced this in carrying out proofs by induction. It can
happen that the induction hypothesis one starts out with isnt strong enough to get the induction
56
step. But starting to do the proof even with the wrong induction hypothesis can help you spot
how to strengthen it. Of course, theres no better way to learn the art of proof than to do proofs,
no better way to read and understand a proof than to pause occasionally and try to continue the
proof yourself. For this reason you are very strongly encouraged to do the exercises, most of them
are placed strategically in the appropriate place in the text.
Mathematicians solve problems and proofs is the guarantee that our solutions are correct. A proof
is an explanation of why a statement is true. A conjecture is a statement which we believe to be
true for which we have no proof. An axiom is a basic assumption about a mathematical situation.
Proof. If m is odd, then m = 2r + 1 for some integer r. Then m2 = (2r + 1)2 = 4r2 + 4r + 1 =
2(2r2 + 2r) + 1, i.e., m2 is odd.
a
Proof. By assumption p = for some integers a and b, where the fraction is in its lowest form.
b
a 2 a2
Thus p2 = = 2 . Since p2 ∈ Z and the fraction is in its lowest form so we have that b2 = 1.
b b
a
Thus b = ±1 ⇒ p = = ±a ∈ Z.
±1
Example 6.14.1. Prove that the square of every odd number is of the form 8a + 1 for some a ∈ N.
Proof. Any odd number n is of the form n = 2l + 1 for some l ∈ Z. Therefore n2 = (2l + 1)2 =
4l2 + 4l + 1 = 4(l2 + l) + 1. Thus it is enough to show that l2 + 1 is even. If l is even, then l = 2m for
some m ∈ Z, so l2 + l = 4m2 + 2m = 2(2m2 + m) which is divisible by 2. If l is odd, then l = 2m + 1
for some m ∈ Z, so l2 + l = 4m2 + 4m + 1 + 2m + 1 = 2(2m2 + 3m + 1) which is also divisible by 2.
Thus l2 + l is always even and so n2 is of the form 8a + 1 for some a ∈ Z. But n2 ≥ 1 so a ∈ N.
The biggest mistake is assuming what has to be proved and incorrect use of equivalence.
57
6.15.1 Don’t Assume What Has to be Proved
Suppose that we had to prove the statement P . If we assume it is true, then it is not surprising
that we can deduce it is true, P ⇒ P , would seem to be obviously true. P is assumed to be true
and this is used to deduce something that is true and so it is concluded that P is true.
Example 6.15.1. Consider the following statement ; If a and b are real numbers, then a2 +b2 ≥ 2ab.
For example, x = y can be proved by that x ≤ y and y ≤ x. We have broken the problem into two
cases.
Proof. Divide into two cases (i) n is even and (ii) n is odd.
As you see, this method of cases involves exhausting all the possibilities and so this method is
also known as exhaustion.
6.17 Contradiction
The law of the excluded middle asserts that a statement is true or false, it cannot be anything
in between. The name comes from the fact that assuming that the statement is false is later
contradicted by some other fact. Also called reductio ad absurdum (reduction to the absurd).
58
Proof. Assume to the contrary, i.e., we suppose that n is an odd integer but that the conclusion is
false, i.e., n2 is an even integer. As n is odd, n = 2k + 1 for some k ∈ Z. Thus n2 = (2k + 1)2 =
4k 2 + 2k + 1 which contradicts n2 is even. Thus our assumption that n2 is even must be wrong, i.e.,
n2 must be odd.
√
Example 6.17.2. Prove that 3 is irrational.
√ √ a
Proof. Suppose a contradiction that 3 is rational. Then we can write 3= for some integers a
b
a2
and b. Assume that a and b have no common divisors. Now squaring both sides, we get 3 = 2 and
b
so 3b2 = a2 . This implies that a2 is divisible by 3 and so a is also divisible by 3. Thus we can write
a = 3c for some integer c. Replacing this in the above equation we get 3b2 = 9c2 and so b2 = 3c2 .
Hence b2 is divisible√by 3 and so is b. But this contradicts the fact that a and b have no common
divisors. Therefore 3 has to be irrational.
6.18 Induction
Is applied when we have an infinite number of statements indexed by the natural numbers, for
example, n5 − n is even for all n ∈ N.
Let A(n) be an infinite collection of statements with n ∈ N. Suppose (i) A(1) is true and
(ii) A(k) ⇒ A(k + 1) for all k ∈ N. Then A(n) is true for all n ∈ N.
Checking condition (i) is called the initial step and checking condition (ii) is called the inductive
step. assuming that A(k) is true for some k in (ii) is called the inductive hypothesis.
Example 6.18.1. 6n − 1 is divisible by 5 for all n ∈ N.
Inductive Step: Assume statement is true for some k ∈ N, that means that 6k −1 = 5m for some
m ∈ N. Then 6k+1 −1 = 6(6k )−1 = 6(5m+1)−1 (by inductive hypothesis)= 30m+6−1 = 5(6m+1).
This is divisible by 5 and so the statement is true for k+1. Hence statement is true for all n ∈ N.
Example 6.18.2. Show that 2n−1 ≤ n! for all n ∈ N.
Then for n = k + 1, 2(k+1) − 1 = 2k = 2(2k−1 ) ≤ 2(k!) (by inductive hypothesis) ≤ (k + 1)(k!) (as
2 ≤ k + 1) = (k + 1)!.
59
6.19 The Contrapositive Method
Proof. If x ∈
/ D, then x ∈ A (the contrapositive). Let us suppose that x ∈
/ D. Since x ∈ C is
assumed, then x ∈ C \ D. Because C \ D ⊂ A ∩ B ⇒ x ∈ A ∩ B, i.e., x ∈ A.
Example 6.19.2. Let a be any integer. prove that if a2 is divisible by 3, then a is divisible by 3.
6.20 Counterexamples
For example, Is all multiples of 3 are multiples of 6, true or false? Prove your answer. It is false
because 9 is a multiple of 3, but is not a multiple of 6. (9 is called a counterexample to the “all”
statement, All multiples of 3 are multiples of 6).
6.21 Divisors
An integer a divides the integer b if there exists an integer k such that b = ka. In this case we say
b is divisible by a and write a | b. We also say that a is a divisor of b.
Proof. By assumption, there exists integers k1 and k2 such that b = k1 a and c = k2 a. For any
integers m and n, we have mb + nc = m(k1 a) + n(k2 a) (by assumption) = (mk1 + nk2 )a. Thus
mb + nc is divisible by a.
60
Theorem 6.21.2. Let a, b, c ∈ Z. Then
Exercise 6.21.3. For each positive integer, show that x3 − x is divisible by 3 and x5 − x is divisible
by 5. Can you generalise this? Is xn − x divisible by n?
2. Let p represent the statement “Simon is married” and let q represent the statement “Michael
is married.” Assume that “married” and “single” are opposites, that is to be not married is
to be single. Translate each of the following statements into symbols.
(a) Both Michael and Simon are married (b) It is not the case that both Simon and Michael
are married (c) Simon is married or Michael is single
(d) Simon is married and Michael is single (e) Either Simon is married or Michael is married
but not both.
3. Let p represent the statement “John is a mathematics teacher” and let q represent the state-
ment “John enjoys reading mathematics book.” Translate each of the following symbolic
statements into words.
(a) p ∧ q (b) ¬p ∧ q (c) ¬(p ∧ q) (d) p ∨ (¬q) (e) ¬p ∨ (¬q).
61
5. Write the converse and the contrapositive of each of the following
(a) If you break the law, then you go to jail (b) If p, then q (c) The sun shines only if
you are happy (d) If wishes were cars, then beggars would drive
(e) Our grandchildren will not survive if we do not control HIV and AIDS in Zimbabwe.
6. These conditional and bi-conditionals are tautologies. Construct truth tables to verify each
of the laws.
(a) p ⇔ p (b) p ⇔ ¬(¬p) (c) p ∨ (¬p) (d) ((p ⇒ q) ∧ p) ⇒ q
(e) (p ∨ q) ∧ (¬p) ⇒ q (Disjunctive Syllogism).
7. Show that (a) ¬(p ∨ ¬q) ∨ (¬p ∧ ¬q) ≡ ¬p (b) p ∨ (p ∧ q) ≡ p (c) p ∨ q ≡ ¬(¬p ∧ ¬q)
(d) ¬(p ∨ (¬p ∧ q)) ≡ ¬p ∧ ¬q (e) p ⇒ q ≡ ¬p ∨ ¬q.
8. Let p(x) be the predicate ‘x > 0’ with x ∈ Z, n(x) be the predicate ‘x < 0’ with x ∈ Z and
q(x, y) be the predicate ‘x < y’ with x, y ∈ Z. Determine whether each of the following is true
or false, giving reasons for your answers.
(a) (∀x)(p(x) ∨ n(x)).
(b) ((∀x)(p(x)) ⇒ (∃x)n(x)).
(c) (∃x)(∃y)(p(x) ∧ n(y)).
(d) (∃x)(∀y)(q(x, y)).
9. For any positive integers x and y, let p(x) be the predicate ‘x is prime’, q(x) be the predicate
’x is odd’ and r(x, y) be the predicate ‘x > y’. Decide whether the following statements are
true or false. Justify your answers.
(a) ¬((∀x)p(x) ⇒ q(x)).
(b) (∃x)(p(x) ∧ (¬q(x)) ∧ r(x, 2)).
(c) (∃x)(∀y)r(x, y).
(d) (∀y)(∃x)(p(x) ∧ r(x, y)).
10. Consider the statement (∀k ∈ N)[6k − 1 is prime or 6k + 1 is prime]. Write down its
negation and decide whether the statement or its negation is true.
11. What is the negation of (∀x ∈ R)(∀y ∈ R)(∃z ∈ Q)[x + y ≥ z]?
12. Give the negation of the following statement
(∃x ∈ R)(∃n ∈ N)(∀q ∈ Q)[x + q < xn ].
13. Show that the statement (∀x ∈ Z)(∀y ∈ Z)(∃z ∈ Z)[x2 +y 2 = z 2 ] is false by means of a specific
example. Write down, in symbolic form not involving ¬, the negation of the statement.
14. Write down the following statements using the quantifiers ∀, ∃ and state which are true and
which are false.
(a) For every real number x there is a real number y such that y 3 = x. (b) There is a real
number y such that for every real number x, the sum x + y is positive
(c) For each irrational number x, there is an integer n satisfying x < n < x + 1
(d) For each pair of real numbers, x, y, if x3 = y 3 , then x = y (e) For each pair of real
numbers, x, y, if x2 = y 2 , then x = y or x = −y.
62
15. Write in words the statements (a) (∀x ∈ R)[x2 − 1 = (x − 1)(x + 1)]
(b) (∃x ∈ R)[x2 − 5x + 6 = 0].
16. Prove that, there are no positive integers x and y such that x2 − y 2 = 1.
17. Prove that the sum of a rational number and an irrational number is an irrational number.
18. Prove that the square of any integer is of the form 3k or 3k + 1 for some k ∈ Z.
√
19. Show that 5 is irrational.
20. Is x4 − x divisible by 4 and is x7 − x divisible by 7 for all positive integers x? Give proofs or
counter examples to justify the answers given.
√
21. Prove that (a) 3 irrational (b) 4n − 1 is divisible by 3, for all n = 1, 2, 3, . . . .
22. Use mathematical induction to prove that for all n ∈ N, the number 42n+1 + 3n+2 is a multiple
of 13.
23. Use induction to prove that 2n > n3 for every integer n ≥ 10.
24. For a natural number n prove that n < 2n , for all n by induction.
31. Prove that for every positive integer n, that 33n−2 + 23n+1 is divisible by 19.
63
Chapter 7
A linguist would be shocked to learn that if a set is not closed this does not mean that it is open, or again
that “E is dense in E” does not mean the same thing as “E is dense in itself”.
We have the following sets, C-set of complex numbers, Z-set of integer numbers, N-set of natural
numbers, Q-set of rational numbers and R-set of real numbers.
7.1 Operations
Operations (such as addition) that involve two input values, for example, 2 + 3 are called binary
operations.
Those
√ that involve only one input value, such as finding the square root of a number (for example,
8) are called unary operations.
Others that involve three input values are called ternary operations.
7.2 Idea
64
Binary operations are usually represented by symbols like ∗, ·, +, ◦. For example, the binary oper-
ations of addition (+) and multiplication (·) in the set Z of integers.
Example 7.2.1. (i) Addition (+) is an operation on Z since + : Z × Z −→ Z i.e., for example,
1 + 2 = 3.
(ii) Let P(S) be the power set of S, i.e., P(S) = {T |T ⊆ S}. ∪ and ∩ are both operations on
P(S).
7.2.1 Properties
For
example,
operation
of multiplication
applied
to the 2 × 2 matrices.
set of
1 2 3 7 5 17 3 7 1 2 3 34
= and = , so a ∗ b 6= b ∗ a.
0 4 1 5 4 20 1 5 0 4 1 22
If on the
other
hand, ∗ as
we take (+),
then
a∗ b =b ∗ a,for all pairs of matrices a and b, i.e., for
1 2 3 7 3 7 1 2 4 9
example + = + = .
0 4 1 5 1 5 0 4 1 9
Definition 7.2.2. ∗ is associative on A, if for all a, b, c ∈ A we have (a ∗ b) ∗ c = a ∗ (b ∗ c).
For example, (i) Addition is both commutative and associative on Z. (ii) ∪ and ∩ are commutative
and associative on P(S).
7.2.2 Closure
Suppose we have a binary operation ∗ and apply to the elements a, b ∈ A to produce a ∗ b. Its
important to know whether or not a ∗ b belongs to the original set A. If a ∗ b ∈ A for every pair
a, b ∈ A, then we say A is closed under the binary operation ∗.
An identity element is an element that when involved in an operation with another element does
not change the value of that element.
Example 7.2.2. Find a the identity element for the operation defined as a ◦ b = a + b + 2.
Let a + b + 2 = a, then b = −2, identity element = −2.
65
Consider the product a ∗ b = c of a, b ∈ A which is closed under the binary operation ∗. a ∗ b is
called the right hand product of b with a and b ∗ a is called the left hand product of b with a. For
example, in the set of R under the operation +, 3 + 0 = 3. We have formed the right hand sum of
0 and 3, addition of 0 does not affect the number 3.
If in any set A, there is an element i for which, for all elements a ∈ A, a ∗ i = i ∗ a = a, then we call
i an identity element or neutral element for set A under the operation ∗.
For example, consider P(S), the power set of S. The ∅ acts as an identity for the operation of
union on P(S).
An inverse element is an element that, when involved in an operation with another results in the
identity element for that operation.
Consider the set of R i.e., given a real number say 2 21 we can obtain another real number −2 12 with
the property that 2 21 + (−2 21 ) = 0.
For all x ∈ R, x + (−x) = 0 and (−x) + x = 0. An element (−x) is called an inverse element. We
only seek inverse elements in sets which have an identity element.
Take a set A with binary operation ∗, (A, ∗). Suppose that there is an identity element i ∈ A and
a ∈ A. Then there is an element b in A such that a ∗ b = b ∗ a = i, then b is called the inverse of a
and write b = a−1 .
Suppose that A is a small finite set and ∗ is an operation on A. We can write out a complete
description of ∗ as follows ;
66
The situation is easily improved by recording the information in the form of a table.
∗ a b c
a a b c
b b c a
c c a b
It is easy to determine the existence of special elements from the table and also whether or not the
operation is commutative.
Commutativity of the operation corresponds to symmetry in the table with respect to the diagonal
i.e., if we fold the table along the diagonal, then the operation is commutative only when the
corresponding entries are the same.
The existence of a left identity can be determined simply by locating a row that is a repetition of
the column headings. Element a is a left identity for ∗. A right identity is found (if one exists) by
locating a column that is a repeat of the row headings. Again a is a right identity for ∗.
Putting these two together yields a method for determining the existence (non-existence) of an
identity relative to a given operation.
Example 7.3.1. The Cayley table for the set S = {1, i, −1, i} with operation multiplication is
× 1 i −1 −i
1 1 i −1 −i
i i −1 −i 1
−1 −1 −i 1 i
−i −i 1 i −1
Test four properties, closed set, associative, identity and multiplicative inverse.
All the results are members of the original set {1, i, −1, −i}. Therefore a closed set.
There is a 1 in every row of the table, so each element has a unique inverse. Note the table is
symmetric about the leading diagonal, hence it is also commutative.
Two operations in the real number system are addition and multiplication. These are called binary
operations because they serve to combine two elements (numbers) in prescribed ways.
67
Let R, the set of real numbers, we have
A1: Every pair of numbers a, b ∈ R have a unique sum a + b ∈ R. (Closure Law of addition)
A2: For a, b ∈ R, a + b = b + a. (Commutative Law for addition)
A3: For a, b, c ∈ R, a + (b + c) = (a + b) + c. (Associative Law for addition)
A4: There is a number in R, called zero, 0, such that for every a ∈ R, a + 0 = a. (Existence of an
additive identity or zero element in R)
A5: For every a ∈ R, there exists a number in R, −a, such that a + (−a) = 0. (Existence of an
additive inverse or negative for each real number a)
The difference between a and b is defined as a + (−b) and the indicated operation is called sub-
traction.
M1: Every pair of numbers a, b ∈ R have a unique product ab ∈ R. (Closure law for multiplication)
M2: For a, b ∈ R, ab = ba. (Commutative Law for multiplication)
M3: For a, b, c ∈ R, a(bc) = (ab)c. (Associative Law for multiplication)
M4: There exists a number in R, 1, such that for each a ∈ R, a·1 = a. (Existence of a multiplicative
identity or unit element in R)
M5: For every a 6= 0 in R, there exists a number, a−1 such that a · a−1 = 1. (Existence of a
multiplicative inverse for each real number a 6= 0)
The quotient of a and b(b 6= 0), is defined as a · b−1 or b−1 · a and the indicated operation is called
division.
These axioms are called field axioms and any algebraic structure which satisfies the field axioms
is called a field.
Theorem 7.4.1 (Cancellation Law for Addition). b + a = c + a =⇒ b = c.
Proof. a + a = a by hypothesis
a + 0 = a by A4
a + a = a + 0 by equality of 1 and 2
a = 0 by A2 and above Theorem
Theorem 7.4.3. The additive inverse of a number is unique.
Restatement : If a number b has the property a + b = 0 then b = −a.
Proof. a + b = 0 by hypothesis
a + (−a) = 0 by A5
68
a + b = a + (−a) by equality of 1 and 2
b + a = (−a) + a by A2
b = −a by Cancellation Theorem
Theorem 7.4.4. a · 0 = 0, where a ∈ R.
Proof. 0 + 0 = 0 by A4
a · (0 + 0) = a · 0 by M1
a · (0 + 0) = a · 0 = a · 0 by D
a · 0 = a · 0 = a · 0 by equality of 2 and 3
a · 0 = 0 by Theorem 6.4.2
Corollary 7.4.1. Zero has no reciprocal.
Proof. If there exists a number 0−1 such that 0−1 · 0 = 1, then Theorem 7.4.4 would be false.
Corollary 7.4.2. If a 6= 0 then a−1 6= 0.
Proof. If a−1 = 0 then a · a−1 = 0 by Corollary 6.4.1 but this contradicts M5. Hence a−1 cannot
be 0.
Theorem 7.4.5. If ab = 0 then either a = 0 or b = 0.
Proof. ab = 0 by Hypothesis
a−1 (ab) = 0 by M1
(a−1 a)b = 0 by M3
1 · b = 0 by M2, M5
b = 0 by M2, M4.
An algebraic structure consists of a non-empty set, one or more operations, one or more relations
and some particular basic properties that these operation and/ or relations possess.
7.6 Groups
Definition 7.6.1. A group (G, ∗) is an algebraic structure, where G is a set and ∗ is a composition
on that set, satisfying the following laws
1. Closure : ∀a, b ∈ G, a ∗ b ∈ G.
69
2. Associativity : ∀a, b, c ∈ G, a ∗ (b ∗ c) = (a ∗ b) ∗ c.
A group can be infinite, the number |G| denotes the order of G, i.e., the number of elements in the
group G.
Definition 7.6.2. An abelian group or commutative group, is a group for which the commu-
tative axiom holds. i.e., a ∗ b = b ∗ a for every a, b ∈ G.
Theorem 5. Suppose that a∗c = b∗c, then (a∗c)∗c−1 = (b∗c)∗c−1 . Hence a∗(c∗c−1 ) = b∗(c∗c−1 )
(by the associativity law). Hence a ∗ i = b ∗ i (by the inverse property), and so a = b (by the identity
property).
Definition 7.7.1. A ring (R, +, ·) is an algebraic system consisting of a non-empty set R and two
binary operations usually denoted by + and ·, so that
2. · is associative on R and
70
Our definition of a ring requires (R, +) to be an abelian group. This guarantees in (R, +, ·) the
uniqueness of the additive identity, 0 and of the additive inverse −a, for each a ∈ R.
For example, (Z, +, ·), (Q, +, ·), (0, +, ·), (P(S), +, ·), where P(S) is the powerset of S and addition
and multiplication are defined as follows, for A, B ∈ P(S), A + B = (A ∪ B) \ (A ∩ B) and
A · B = A ∩ B.
Definition 7.7.2. 1. (R, +, ·) is a commutative ring if ab = ba for all a, b ∈ R.
1. a(−b) = −(ab).
2. (−a)b = −(ab).
3. (−a)(−b) = (ab).
Proof. 1. −(ab) is the unique additive inverse of ab. Now ab + a(−b) = a[b + (−b)] = a(0) = 0.
Thus a(−b) is also the unique additive inverse of ab and we must have a(−b) = −(ab).
3. (−a)(−b) = −[(−a)b] by 1 = −(−(ab)) by 2 = ab.
Theorem 7.7.2. Let R be a ring and a, b, c ∈ R. Then
1. −(a + b) = −a − b.
2. −(a − b) = b − a.
3. (a − b) − c = (a − c) − b.
4. a(b − c) = ab − ac.
5. (a − b)c = ac − bc.
Proof. Addition
Identity : a ⊕ 1 = 1 ⊕ a = 1 + a − 1 = a, thus one is the additive identity.
Inverse : a ⊕ (2 − a) = (2 − a) ⊕ a = 2 − a + a − 1 = 1, thus (2 − a) is the additive inverse of a.
Commutative : a ⊕ b = a + b − 1 = b + a − 1 = b ⊕ a.
71
Associative : (a ⊕ b) ⊕ c = (a + b − 1) ⊕ c = a + b − 1 + c − 1
= a + (b + c − 1) − 1 = a ⊕ (b + c − 1) = a ⊕ (b ⊕ c).
Multiplication
Identity : a 0 = 0 a = a + 0 − a · 0 = a, thus 0 is the multiplicative identity.
Commutative : a b = a + b − ab = b + a − ba = b a.
Associative : (a b) c = (a + b − ab) c = (a + b − ab) + c − (a + b − ab)c
= a + (b + c − bc) − a(b + c − bc) = a (b + c − bc) = a (b c).
7.8 Fields
Definition 7.8.1. A non-zero element a in a commutative ring R is called a zero divisor if there
exists a non-zero element b ∈ R such that ab = 0.
Theorem 7.8.1 (The Cancellation law for Multiplication). Let a, b, c be elements of a commutative
ring R. If a 6= 0 is not a zero divisor and ab = ac, then b = c.
Proof. If ab = ac, then ab − ac = 0 and a(b − c) = 0. Now a 6= 0 and a is not a zero divisor,
therefore b − c = 0 and b = c.
Definition 7.8.2. An integral domain is a commutative ring with identity that has no zero
divisors.
The rational and real number systems have a structure only slightly more sophisticated than an
integral domain. The difference is that every non-zero element has a multiplicative inverse.
Definition 7.8.3. A field is a commutative ring with identity in which the non-zero elements form
a multiplicative group.
72
7.9 Integers
3. a0 6= 1 for any a ∈ N.
(a) 1 ∈ S.
(b) If a ∈ S, then a0 ∈ S. Then S = N.
The fifth axiom is called the axiom of induction or the principle of induction. 10 = 2, 20 =
3, 30 = 4 and so on.
7.10 Addition
Addition on N is defined by
73
7.11 Multiplication
Defined by
(i) n · 1 = n.
A proposition P (m) is true for all m ∈ N provided P (1) is true and for each k ∈ N, P (k) is true
implies P (k 0 ) is true.
Proof. We want to prove that n + m is defined (is a natural number) by (i) and (ii) for all m, n ∈ N.
Suppose n to be a fixed natural number and consider the proposition P (m) : n + m ∈ N for every
m ∈ N. Now P (1) : n + 1 ∈ N is true since n + 1 = n0 ∈ N by (i) and n0 ∈ N by 2. Suppose next
that for some k ∈ N P (k) : n + k ∈ N is true. Then P (k 0 ) : n + k 0 ∈ N is true since n + k 0 = (n + k)0
by (ii) and (n + k)0 ∈ N whenever n + k ∈ N by 3. By induction P (m) is true for all m ∈ N.
Theorem 7.12.1. There exists one and only one binary operation “+” in N satisfying (i) a+1 = a0
(ii) a + b0 = (a + b)0 for all a, b ∈ N.
74
Proof. 2 + 2 = 2 + 10 by definition = (2 + 1)0 = (20 )0 = 30 = 4.
Theorem 7.12.2. There exists one and only one binary operation “ · ” in N satisfying (i) a · 1 = a
(ii) a · b0 = a · b + a.
Proof. 2 · 2 = 2 · 10 = 2 · 1 + 2 = 2 + 2 = 4.
7. Verify that the set of odd integers does not form a group under addition.
8. Let S = {A, B, C, D} where A = ∅, B = {a, b}, C = {a, c}, D = {a, b, c}. Construct tables to
show that ∪ is a binary operation on S but ∩ is not.
9. Let S = {A, B, C, D} where A = ∅, B = {a}, C = {a, b}, D = {a, b, c}. Construct tables to
show that ∪ and ∩ are binary operation on S.
(i) Is ∗ commutative on Z?
(ii) Is ∗ associative on Z?
(iii) Does there exists an identity relative to ∗?
(iv) Does a ∈ Z have an inverse relative to ∗?
11. Suppose p and q are fixed integers. Define an operation ∗ on Z by m ∗ n = pm + qn for all
m, n ∈ Z. What should p, q be to ensure that Z has an identity element with respect to ∗?
ab
12. The binary operation ∗ : N × N → Z is defined, for all a, b ∈ N, by a ∗ b = . Prove that
a+b
the binary operation is associative but does not have an identity element.
75
13. Prove the following
(a) The additive inverse of a − b is b − a i.e., −(a − b) = b − a. (b) The additive identity
is unique (c) The additive inverse of the additive inverse of a number b is b itself i.e.,
−(−b) = b (d) The negative of zero is zero itself i.e., −0 = 0 (e) The cancellation law
for multiplication holds i.e., ba = ca and a 6= 0 ⇒ b = c (f) The multiplicative inverse of
a non-zero number is unique (g) The multiplicative identity is unique (h) The additive
inverse of a + b is −a − b i.e., −(a + b) = −a − b (i) If a 6= 0 then (a−1 )−1 = a (j)
(−1) · a = −a (k) (−a)(−b) = ab (l) (−a) · b = −(ab).
15. Let G = {[1]13 , [3]13 , [4]13 , [9]13 , [10]13 , [12]13 } be a set of equivalence classes modulo 13. Let
(G, ∗) be the set with binary operation of multiplication modulo 13.
76