You are on page 1of 260

Foundations

Undergraduate Mathematics

Sets and Logic, Relations and Functions,


Number Systems, Infinite Sets

Amber Habib

Ver 1.0 2019


Amber Habib
Department of Mathematics
Shiv Nadar University
Gautam Buddha Nagar
201314
India

Email: amber.habib@snu.edu.in

Undergraduate Mathematics: Foundations


c 2019 Amber Habib
Version 1.0 January 2019.

Book website: https://sites.google.com/a/snu.edu.in/amber-habib/my-books/

Cover image by Abha Dev Habib

This work is distributed free under the Creative Commons Attribution-NonCommercial-


ShareAlike 4.0 International License. To view a copy of this license, visit

http://creativecommons.org/licenses/by-nc-sa/4.0/

or send a letter to Creative Commons, PO Box 1866, Mountain View, CA 94042, USA.
About this Book
Modern mathematics is an immensely successful enterprise. Over the last
century the discipline has expanded exponentially in its depth and scope.
At the start of this period its serious applications were almost entirely in
physics and engineering. Now they extend to computer science, chemistry,
biology, medicine, economics, nance, management, and even the ne arts.
This growing popularity appears, somewhat paradoxically, to be due to an
increased abstraction. By removing material interpretations from the core of
mathematics, this process of abstraction has made the subject universal in
its scope. This book presents a course on the reformulation of mathematics
that has made this possible. It is suitable for being taught during the rst
year of an undergraduate mathematics program.

An early goal is to get students to realize that very often what we say or
write does not convey what we feel and think, and that clarity of expression
is benecial to both parties in a conversation. The basic topics from logic and
set theory are ideal for illuminating these aspects of doing mathematics. Of
course they also lay the foundation for further studies in mathematics, but
the immediate impact that the student can see relates to clarity of thought
and to self-criticism. This is the task of the rst chapter.

In the second chapter, we take up the notions of relations and functions.


These are the foundational concepts in modern mathematics, in the sense
that every mathematical activity is sought to be expressed in their terms.
The third chapter applies this language of sets, relations and functions to the
creation and study of numbers. The main goal here is the description of real
numbers as a complete ordered eld, required for a rigorous development of
Calculus. Thus the rst three chapters can be seen as laying the foundation
for any mathematical activity whatsoever.

A parallel goal is to give students a glimpse of mathematics as a liv-


ing subject, created by humans and determined by our choices. Set the-

i
ii

ory may now be taught to undergraduates in a manner that suggests it is


a perfect model for all of mathematics, but a hundred years ago the best
mathematicians of the time held very diverse views about it. In 1924, when
Alfred Tarski submitted for publication an equivalence involving the Axiom
of Choice, one editor returned it because he objected to the Axiom, and an-
1
other returned it because he thought it was pointless to derive a true axiom!
Such disagreements were not conned to the rules by which Set theory was
to be governed. They extended to the question of its place in mathematics.
If David Hilbert saw in Set theory the potential to provide a rm foundation
for all of mathematics, Henri Poincaré viewed it as less fundamental than
arithmetic.

The desire to touch upon these aspects has inuenced the choice of topics
in the fourth and nal chapter. To appreciate Set theory as a theory, and not
just as a convenient tool for expressing mathematical statements, we must
consider what it has to say about innite sets. The simplest of questions,
such as Is there a smallest innite set?, then force us to face the Axiom of
Choice and its consequences, the useful as well as the bizarre.

Readers will prot more from the formal approach of this book if they
have prior exposure to the basic notions regarding sets and and operations
with them, at the level of the Class XI NCERT textbook [19], or the books
[1, 2] in our References.

The rst three chapters already oer enough material for a typical 45
lecture course. If another 15 lectures are available, the fourth chapter could
also be covered. Alternately, one could skip some of the material of Chapter
3 in order to include the basics of innite sets from Chapter 4 in a 45 lecture
course. One such possibility is to cover all of the rst two chapters, the rst
two sections of the third chapter, and the rst four sections of the fourth
chapter.

Any comments, suggestions or corrections would be gratefully received.


Please send them to amber.habib@snu.edu.in.

1 Gregory H. Moore, Zermelo's Axiom of Choice, Dover, 2013.

Undergraduate Mathematics: Foundations


iii

Acknowledgments

• Prof. S. Kumaresan and all participants  instructors and students 


2
of the Mathematical Training & Talent Search programme. My rst
opportunity to teach many of the topics covered in this book was at an
MTTS summer camp, and I beneted greatly from the notes of Prof.
Kumaresan's classes. Those have now been expanded into a full book,
co-authored by Ajit Kumar, S. Kumaresan and Bhaba K. Sarma [6].

• Sushil Singla, who educated me on the richness of the mathematics


related to the Axiom of Choice while I was supervising his MSc project.
His comments on the original draft also helped clarify my ideas on the
material and led to signicant changes in its presentation. In fact ŸA.1
Truth and Provability is essentially due to him.

• The students of the Monsoon 2017 class of MAT100 Foundations at


Shiv Nadar University. Prior to my time with them, I did not think
the Axiom of Choice could be a suitable topic for undergraduates in
their rst semester.

• AT X typesetting system. Among the many L


The creators of the L E
AT X
E
packages used in composing this book, special mention has to be made
of Tik Z, which was used for most of the gures. Ingo H. de Boer's
3
Winshell editor provided a convenient interface and project organizer.

Signage

We have used the following icons to mark certain portions of text:

« An indication of an issue that needs special attention.

7 A motivational remark, usually preceding a proof or a part of a proof.

’ An assertion that needs to be worked out by the reader.

 End of an example.

 End of a proof.

2 The main MTTS activity consists of month-long summer camps for undergraduate
students. MTTS implements a discussion-based pedagogy that provides to most of its
students their rst experience of genuine mathematical activity. Visit the MTTS website
http://mtts.org.in to view their latest activities.
3 http://www.winshell.de.
Contents
1 Sets and Logic 1
1.1 Statements and Operations . . . . . . . . . . . . . . . . . . . 3
1.2 Implication . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Statement Forms and Equivalence . . . . . . . . . . . . . . . 15
1.4 Open Sentences and Quantiers . . . . . . . . . . . . . . . . . 21
1.5 Axioms for Sets . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.6 Sets, Quantiers and Numbers . . . . . . . . . . . . . . . . . 40

2 Relations and Functions 49


2.1 Cartesian Products . . . . . . . . . . . . . . . . . . . . . . . . 51
2.2 Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.3 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
2.4 Indexed Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
2.5 Bijections, Inverse Functions and Cardinality . . . . . . . . . 88
2.6 Binary Operations . . . . . . . . . . . . . . . . . . . . . . . . 97

3 Number Systems 105


3.1 Natural Numbers: Denitions . . . . . . . . . . . . . . . . . . 106
3.2 Natural Numbers: Properties . . . . . . . . . . . . . . . . . . 116
3.3 Integers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
3.4 Rational Numbers . . . . . . . . . . . . . . . . . . . . . . . . 131
3.5 Integral Domains and Fields . . . . . . . . . . . . . . . . . . . 138
3.6 Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
3.7 Complex Numbers . . . . . . . . . . . . . . . . . . . . . . . . 154

4 Innity 165
4.1 Constructing the Natural Numbers . . . . . . . . . . . . . . . 166
4.2 Finite and Innite Sets . . . . . . . . . . . . . . . . . . . . . . 170
4.3 The Axiom of Choice . . . . . . . . . . . . . . . . . . . . . . . 176
4.4 Countable and Uncountable Sets . . . . . . . . . . . . . . . . 181
4.5 Zorn's Lemma and Well-Ordering Theorem . . . . . . . . . . 186

A Supplements 193

v
vi CONTENTS

A.1 Truth and Provability . . . . . . . . . . . . . . . . . . . . . . 193


A.2 Sets and Classes . . . . . . . . . . . . . . . . . . . . . . . . . 195
A.3 Constructing the Real Numbers . . . . . . . . . . . . . . . . . 198

B Solutions 203
References 245
Index 249

Undergraduate Mathematics: Foundations


1 | Sets and Logic
That which renders Logic possible, is the existence in our minds
of general notions, our ability to conceive of a class, and to des-
ignate its individual members by a common name.
1
 George Boole

The rst philosopher stepped forward to start the disputation.


`What', he asked the Mulla, `is the centre of the Earth?'
The Mulla pointed with his pen. `The exact centre of the Earth
is the centre of the spot upon which my donkey, yonder, has his
foot.'
`How can you prove it?'
`On the contrary, you disprove it. Get a measuring tape!'
2
 Idries Shah

In this chapter we shall explore the fundamentals of mathematical rea-


soning, and the language in which we carry out this reasoning. This is the
language of set theory and logic. To begin with, let us consider some typi-
cal expressions or sentences one may encounter while discussing mathematics
in school:

1. 2 + 2 = 4.
2. 0/1 = 0.
3. There is no real number whose square is −1.
4. If the diagonals of a parallelogram are equal then it is a rectangle.

1 George Boole, The Mathematical Analysis of Logic, Macmillan, London, 1847. (http:
//www.gutenberg.org/ebooks/36884)
2 Idries Shah, The Pleasantries of the Incredible Mulla Nasrudin, The Octagon Press,
1995.

1
2 CHAPTER 1. SETS AND LOGIC

These are all assertions of facts, and one goal of a discussion involving them
would be to gure out whether they are correct or not. The rst two are
assertions about mathematical processes, about the result of carrying out an
operation with known numbers. The third is an assertion about the possible
existence of a number with a special property. The fourth claims a rela-
tionship between the property of a parallelogram being rectangular and the
lengths of its diagonals. It is in fact a combination of two simpler assertions
about parallelograms (being rectangular and having equal diagonals). At
the simplest level, we will learn how to correctly understand and express
mathematical assertions such as these. An important aim will be to remove
the ambiguities that are present in our common speech.

As to determining the correctness of a given mathematical statement,


this is where logic makes an entrance. Logic provides techniques by which
the known correctness of some statements can be used to establish the cor-
rectness of others. We need a starting point for this process. The starting
point consists of certain statements which are assumed to be correct (these
are called axioms). Then the techniques of logic are used to establish the
correctness of other statements (which are called theorems). A sequence
of logically justied statements that leads from the axioms to a theorem is
called a proof of that theorem.

Now let us look at a few more sentences that might crop up in a school
mathematics class.

1. There are innitely many prime numbers.



2. 2 is an irrational number.

3. The interior angles of a triangle always sum to 180◦ .


Each statement involves a collection and its members. Let's take them one
by one.

1. We consider the collection of all prime numbers and we say something


about how many members this collection has.

2. We have two collections: the real numbers and the rational members,
and we assert that there is a real number whose square is 2 and this
real number is not a rational number.

3. We have the collection of all triangles, as well as the collection of real


numbers. To each triangle we associate a real number  the sum of its
interior angles  and we make a claim about this associated number.

Undergraduate Mathematics: Foundations


1.1. STATEMENTS AND OPERATIONS 3

So our next task will be to bring clarity to the concept of collection. It


may not seem like there is much to do here. Mathematicians used to think
so too, until just over a hundred years ago, when Bertrand Russell pointed
out how an unbridled acceptance of all conceivable collections would lead to
contradictions.

In the rst four sections of this chapter, we have an introduction to the


basic concepts of symbolic logic. Here, we learn how to make convincing
arguments in mathematics. This will require setting down the rules for ex-
pressing combinations of statements and the implications between them. In
the fth and sixth sections, we will develop the concept of sets, which are
essentially those collections whose unrestricted use is permissible in mathe-
matics. On the one hand we will apply logic to prove facts about sets. On
the other, we will use sets to bring greater clarity to logic and its connections
with mathematics.

The steps we will take to bring more clarity to mathematics are modeled
on Euclid's formulation of geometry over two thousand years ago. Consider
the problem of dening points and lines in a plane. We may consider at-
tempts like A point occupies no space or A line has no width but we soon
realize these are all insucient. In fact, points and lines are fundamental
concepts of geometry and can't be dened in terms of simpler ones. What
matters is not the attempted denition of points and lines but the relation-
ships between them, such as There is a unique line joining any two points.
Euclid called these fundamental relationships postulates (the modern term
for such fundamental relationships is `axioms'). With them as a starting
point, he proceeded to derive all the theorems of geometry. In the same way
we will seek to set down some fundamental notions and relationships from
3
which all of mathematics is to be obtained.

1.1 Statements and Operations


In logic, we are concerned with rules for deciding which arguments are valid.
Our notion of valid will be extremely strict. Modes of argument which we
accept in everyday conversation, may well be considered invalid in logic.
Fortunately, this strictness is precisely what we need in mathematics.

As an example, consider the expression n2 + n + 41. Substituting succes-


sive whole numbers for n, we nd the following values:

3 Will we succeed? See ŸA.1 for some unexpected twists in this story.
4 CHAPTER 1. SETS AND LOGIC

n 0 1 2
n2 + n + 41 41 43 47

Now I say to you, Look, this expression always produces a prime number
when we substitute a whole number for n! Are you convinced? Of course
not. You feel that three pieces of data are too little for rm conclusions and
you demand more evidence. I provide the following:

n 3 4 5 6 7 8 9 10 11
2
n + n + 41 53 61 71 83 97 113 131 151 173

These are all prime numbers too. Perhaps now you are wavering  surely
something more than coincidence is involved here? Suppose I keep showing
you more and more instances of prime numbers being produced by this ex-
pression. Would there be a point when you would say, Please don't show
me any more instances! Now I believe you?

In our everyday lives, even in our professional lives, we do argue and


accept claims in this manner. We rely on some intuition of reasonable
evidence. However, in logic and in mathematics, this mode of arguing is
not accepted. Here, it is all or nothing. To make a claim about all whole
numbers we must somehow cover every whole number. If we miss even one,
we have not succeeded.

Task 1.1.1 Check that 402 + 40 + 41 is not a prime number.

Simple Statements

We wish to set down precise rules for working out the truth or falsity of
mathematical assertions. First, we have the notion of a simple statement,
which has the form Object A has property P. For example, the following are
simple statements:

The number 7 is a prime number.


Four equals two plus two.

We may attach a truth value to each simple statement. The truth value is
either true (T) or false (F). If the object actually has the stated property
the statement is termed true, else it is termed false. We cannot allot both
F and T to the same simple statement. It may also happen that we do not

Undergraduate Mathematics: Foundations


1.1. STATEMENTS AND OPERATIONS 5

know which one to allot, for we simply may not know enough to make the
4
determination.

More generally, a statement is a sentence that makes an assertion that


can be evaluated for whether it is true or false. That is, it should be clear
what it means for the statement to be true and it should also be clear what
it means for the statement to be false.

The purpose of logic is not so much to gure out whether given statements
are individually true or false, but rather to delineate the consequences of their
being true or false.

Compound Statements

Compound statements are created by combining or altering other state-


ments:

Two plus two equals four and ve.


Two plus two equals four or ve.
Not all rectangles are squares.

The rst two statements are combinations of Two plus two equals four with
Two plus two equals ve. The third is an alteration of All rectangles are
squares.

We have to be careful while using the words and, or, not. In common
English usage there are ambiguities which we must remove. For instance,
the word or can be used in dierent ways as illustrated below:

Every student in this class has taken algebra or analysis.


You will get 60% in this exam or you will fail the course.

In the rst, the usage of or is inclusive: it allows for the possibility that
a student may have taken both analysis and algebra. In the second, it is
exclusive: the possibility of getting 60% and failing is excluded. The dierent
usage is sensed from one's experience with the language and from the context
 it is not detected from the structure of the sentence. This is obviously a
possible cause of confusion.

Similarly, consider the sentence:

All apples are not black.

4 See ŸA.1.
6 CHAPTER 1. SETS AND LOGIC

Did the writer mean to convey that no apple is black, or that some apples
are not black? It is dicult to be certain.

(a) No apple is black. (b) Some apples are not black.

To avoid such problems, we shall specify the exact usage of each connective
within the subject of logic.

Conjunction, Disjunction and Negation

It is convenient to introduce some notation:

1. Individual statements will be denoted by letters such as P , Q, R. These


statement variables.
letters will be called

2. P ∧ Q denotes `P and Q' and is called the conjunction of P , Q.

3. P ∨ Q denotes `P or Q' and is called the disjunction of P , Q.

4. ¬P denotes `not P ' and is called the negation of P .

We use truth tables to dene when a compound statement such as P ∧Q


is true, depending on the truth values of the component statements. This
will x the exact meaning of each connective.

P Q P ∧Q P Q P ∨Q
T T T T T T
T F F T F T P ¬P
F T F F T T T F
F F F F F F F T

(a) Conjunction (b) Disjunction (c) Negation

Table 1.1.1: Truth tables for the three connectives.

Undergraduate Mathematics: Foundations


1.1. STATEMENTS AND OPERATIONS 7

Let us take up the truth table for conjunction. In each row, the rst
entry is a possible truth value of P while the second entry is a possible truth
value of Q. The third entry gives the truth value of P ∧Q corresponding
to the given truth values of P , Q. We need four rows to account for all the
possible combinations of the truth values of P and Q. We see that P ∧Q is
dened to be true exactly when both P, Q are true. If one or both of them
are false then P ∧Q is false.

The truth table for P ∨Q is constructed similarly. We see that for ∨ we


have chosen the inclusive sense: P ∨ Q is true if one or both of P , Q are true.
The truth table for the negation ¬P only has to account for the truth
values of P and so needs just two rows. We see that ¬P is dened to be true
when P is false, and false when P is true.

Negation is applied to a single statement, conjunction and disjunction to


pairs of statements. So the use of brackets is very important when dealing
« with combinations of multiple statements. For example, we should not
write P ∨ Q ∧ R: It has to be one of (P ∨ Q) ∧ R and P ∨ (Q ∧ R), and
these two carry dierent meanings.

In order to keep brackets from proliferating, we adopt the convention that


a negation symbol placed just before a statement variable is applied only to
that variable. For example, ¬A ∨ ¬B is read as (¬A) ∨ (¬B).

Exercises for Ÿ1.1

1 Express the following symbolically, using standard mathematical notation


(like 2 + 2 = 4) and the operators ∧, ∨, ¬.
(a) Two plus two equals four and ve.

(b) Two plus two equals four or ve.

(c) It is false that two plus two equals ve.

(d) It is false that two plus two does not equal ve.

2 Which of the following are true? (Answer by combining your existing


knowledge of common mathematical facts and fallacies with the rules you
have learnt for disjunction, conjunction and negation.)

(a) Two plus two equals four and ve.

(b) Two plus two equals four or ve.

(c) It is false that two plus two equals ve.


8 CHAPTER 1. SETS AND LOGIC

(d) It is false that two plus two does not equal ve.

3 Let P, Q be two statements. Express Exactly one of P, Q is true using


the symbols P , Q, ∧, ∨, ¬.

4 A logician wishes to use the exclusive or operator for combining state-


ments, and decides to denote it by t. (That is, P tQ means Either P is
true or Q is true but not both.) Fill the truth table for P t Q.

5 Can combinations of the operators ¬, ∧, ∨ generate every possible truth


table with two statement variables?

6 Let P and Q be two statements. Are the following possible?

(a) P ∧Q is true but P ∨Q is false. (b) P ∨Q is true but P ∧Q is false.

7 Let P, Q and R be statements. Is it possible to allocate truth values to


them so that the following happen?

(a) (P ∨ Q) ∧ R is false but P ∨ (Q ∧ R) is true.

(b) (P ∨ Q) ∧ R is true but P ∨ (Q ∧ R) is false.

8 I am given some statements P , Q, R and am asked to prove that at least


one of them is true. I manage to show that if P and Q are false then R is
true. Is my proof complete?

9 Let P and Q be the following statements:

P : The function f (x) is continuous.

Q : The function f (x) is dierentiable.

Express the following statements symbolically, using the symbols P , Q, and


the operators ∧, ∨, ¬.
(a) f (x) is both continuous and dierentiable.

(b) f (x) is neither continuous nor dierentiable.

(c) f (x) is continuous but not dierentiable.

Undergraduate Mathematics: Foundations


1.2. IMPLICATION 9

10 Let P , Q and R be the following statements about a certain plane gure:

P : The gure is a polygon.

Q : The gure is a triangle.

R : The gure is a square.

Express the following statements symbolically, using the symbols P , Q, R,


and the operators ∧, ∨, ¬.
(a) The gure is a polygon and is either a triangle or a square.

(b) The gure is a polygon but neither a triangle nor a square.

(c) The gure is a triangle or a polygon which is not a square.

11 Is the following a well-phrased statement: The given gure is a triangle


or a polygon but not a square?

1.2 Implication
Another kind of compound statement is the  P implies Q form. The deni-
tion of this form requires more careful thought than we needed for conjuction
and its companions. The reason is that words like `if ' and `implies' have
subtle connotations which we must untangle before we can settle on their
appropriate use in logic. Consider the statement

The number 6 is even implies that the number 36 is divisible by 4.

You can surely see how the divisibility of 36 by 4 follows from the divisibility
of 6 by 2. Therefore, you would accept this as a correct statement. Let's
consider a similar statement:

The number 5 is odd implies that the number 36 is divisible by 4.

This may throw you for a second but not longer. From 5 being odd you
deduce that 6 is even and so you again obtain the divisibility of 36 by 4. It
seems that you, and anyone else, would be comfortable saying  P implies Q
if the truth of Q could be seen as being caused by the truth of P, perhaps
through a chain of deductions.

The number 5 is odd implies that the number 36 is divisible by 5.

Now here is an implication we would reject. We know 5 is indeed odd and


that 36 is not divisible by 5. A true statement cannot be allowed to imply a
false statement. Are all statements involving implication so easy to settle?
10 CHAPTER 1. SETS AND LOGIC


The real number 2 is irrational implies that the number 36 is divisible
by 4.

This statement involves two true statements, but does it seem reasonable to

claim that the irrationality of 2 is responsible for the divisibility of 36 by
4? It does not. Yet, can we show that there is no chain of deductions that

moves from the irrationality of 2 to the divisibility of 36 by 4? This task
also seems daunting.

We resolve this dilemma by deciding that in logic we shall not worry


about causation. So we interpret  P implies Q in a much narrower way: Is
it possible for P to be true and Q to be false? If yes, then the implication
does not hold. If no, then it does hold.

To remind ourselves that implication in logic does not include causation,


we refer to it as logical implication or material implication. We further
strengthen the distinction by preferring to write If P then Q rather than
 P implies Q. And, nally, we adopt the notation P =⇒ Q for If P then
Q and call it a conditional statement.
The discussion so far gives us two rows of the truth table for logical
implication:

P Q P =⇒ Q
T T T
T F F
F T
F F

Here are two statements, corresponding to the last two rows of this truth
table:

If the real number
√2 is rational then the number 36 is divisible by 4.
If the real number 2 is rational then the number 36 is divisible by 5.

In each case, you may again feel uneasy about accepting the implication.
Can a false statement be allowed to imply anything? We argue our way out
of this as follows. The statement P =⇒ Q is concerned with what happens
when P is true. It has nothing to say about what happens when P is false.
A statement that has nothing to say has to be accepted as true rather than
false! (Think about 1 = 1) Hence we will accept both these statements as
true.

Here is an amusing example of how a false statement can be used to prove

Undergraduate Mathematics: Foundations


1.2. IMPLICATION 11

any other, such as starting from 2 + 2 = 5 and showing that a certain person
5
M is the Pope:

For if 2 and 2 make 5, since they also make 4, we would conclude


that 5 is equal to 4. Consequently, subtracting 3 from both sides,
we conclude that 2 would be equal to 1. But if this were true,
since M and the Pope are two, they would be one, and obviously
then M would be the Pope.

Thus the full truth table for logical implication is

P Q P =⇒ Q
T T T
T F F
F T T
F F T

We see that we have dened the logical implication P =⇒ Q to be false


only when P (the hypothesis) is true but Q (the conclusion) is false. It is
true under any other conditions.

This truth table gives the nal and complete denition of =⇒ . We have
tried to motivate why we have made these particular choices. From now
« on you must use this denition alone and not be confused by expectations
arising from other usages of if  or implies in ordinary language.

Task 1.2.1 Write the following statements in symbolic form using ∧, ∨, ¬


and =⇒ :
(a) If 5 and 7 are odd then 35 is odd.

(b) If a>0 then, if ab > ac, we have b > c.

In mathematics texts you will nd P =⇒ Q represented by various


phrases, such as the following:

5 Recounted by P. E. B. Jourdain in The Philosophy of Mr B*rtr*nd R*ss*ll, George


Allen & Unwin, 1918. https://archive.org/details/philosophyrussel00jouruoft
12 CHAPTER 1. SETS AND LOGIC

If P then Q, Q whenever P,
Q if P, Q follows from P.

Further, P =⇒ Q is also represented by the following expressions:

1. P sucient condition for Q.


is a

2. Q is a necessary condition for P .


From your earlier study of mathematics, you know we often have to set
up a chain of implications. The following result underlies such work.

Theorem 1.2.2 If P =⇒ Q and Q =⇒ R are true then P =⇒ R is


also true.

Proof: Let us set up a truth-table for these three statements. Since three
statement variables are involved, the table would have eight rows.

P Q R P =⇒ Q Q =⇒ R P =⇒ R
T T T T T T
T T F T F F
T F T F T T
T F F F T F
F T T T T T
F T F T F T
F F T T T T
F F F T T T

We have shaded the rows in which P =⇒ Q and Q =⇒ R are both true.


In each of these rows, P =⇒ R is also true. This proves the theorem. 
From the last theorem, we can say that (P =⇒ Q) ∧ (Q =⇒ R) is a
sucient condition for P =⇒ R. However, it is not a necessary condition
as there are cases where (P =⇒ Q) ∧ (Q =⇒ R) is false but P =⇒ R is
still true.

Task 1.2.3 Give a specic example of statements P , Q, R for which (P =⇒


Q) ∧ (Q =⇒ R) is false but P =⇒ R is true.

Truth-table proofs are easy to carry out in the sense that they are mechan-
ical and do not strain our creative ability. But, as the last proof illustrates,

Undergraduate Mathematics: Foundations


1.2. IMPLICATION 13

they can be tedious and that creates its own possibility of error. They also
do not give the sense of gaining insight into why something is true. A little
later, we will see other ways of proving this result.

This is a suitable moment to discuss a common convention. In mathe-


matical texts one often sees a chain of implications, such as:

x = x2 =⇒ x2 − x = 0 =⇒ x(x − 1) = 0
=⇒ x = 0 or x − 1 = 0 =⇒ x = 0 or x = 1.

This is a compressed way of writing the following:

(x = x2 =⇒ x2 − x = 0)
2
∧ (x − x = 0 =⇒ x(x − 1) = 0)
∧ (x(x − 1) = 0 =⇒ x = 0 or x − 1 = 0)
∧ (x = 0 or x − 1 = 0 =⇒ x = 0 or x = 1).

By repeatedly applying Theorem 1.2.2 we can conclude from these implica-


tions that
x = x2 =⇒ x = 0 or x = 1.

Biconditional Statements

Consider the conditional statements P =⇒ Q and Q =⇒ P . Let us


inspect their combined truth table:

P Q P =⇒ Q Q =⇒ P
T T T T
T F F T
F T T T
F F T F

We see that P =⇒ Q and Q =⇒ P are not equivalent. We dene the


biconditional statement P ⇐⇒ Q to be (P =⇒ Q) ∧ (Q =⇒ P ). We
read this statement as  P if and only if Q. Its truth table is

P Q P =⇒ Q Q =⇒ P P ⇐⇒ Q
T T T T T
T F F T F
F T T F F
F F T T T
14 CHAPTER 1. SETS AND LOGIC

Thus, the biconditional P ⇐⇒ Q is true when both P and Q have the


same truth value, and is false when they have dierent truth values.

Exercises for Ÿ1.2

1 Let P and Q be two statements.

(a) Suppose P =⇒ Q and P are true. Can we conclude that Q is true?

(b) Suppose P =⇒ Q and Q are true. Can we conclude that P is true?

(c) Suppose P and Q are true. Can we conclude that P =⇒ Q is true?

2 Fill the truth tables for the following:

(a) P =⇒ (P =⇒ P ), (b) (P =⇒ P ) =⇒ P ,

(c) (P ∧ Q) =⇒ (P ∨ Q), (d) (P ∨ Q) =⇒ (P ∧ Q).

3 Express the following symbolically using ∧, ∨, ¬, =⇒ :


(a) If not P or not Q, then it is not the case that P or Q.
(b) If P implies that Q implies R, then P and Q together imply R.

4 Let P and Q be the following statements:

P : The function f (x) is continuous.

Q : The function f (x) is dierentiable.

Express the following statements symbolically, using the symbols P , Q, and


the operators ∧, ∨, ¬, =⇒ .
(a) If f (x) is dierentiable then it is continuous.

(b) f (x) is dierentiable only if it is continuous.

(c) If f (x) is not continuous then it is not dierentiable.

5 Let P, Q and R be the following statements about a certain plane gure:

P : The gure is a rectangle.

Q : The gure is a rhombus.

R : The gure is a square.

Express the following statements symbolically, using the symbols P , Q, R,


and the operators ∧, ∨, ¬, =⇒ .

Undergraduate Mathematics: Foundations


1.3. STATEMENT FORMS AND EQUIVALENCE 15

(a) The gure is a rectangle and is either a rhombus or a square.

(b) If the gure is a rectangle then it is either a rhombus or a square.

(c) If the gure is a rectangle and a rhombus then it is a square.

(d) The gure is a square only if it is a rectangle.

(e) If the gure is a rectangle but not a rhombus then it is not a square.

6 Rewrite the following in the If . . . then . . .  format:

(a) For a triangle to be equilateral, it is sucient that the angles of the


triangle be congruent.

(b) The divisibility of x by 3 is necessary for the divisibility of x by 6.

7 Show that P =⇒ Q and (¬P ) ∨ Q have the same truth table.

8 Show that ¬(P =⇒ Q) and P ∧ ¬Q have the same truth table.

9 Show that P =⇒ Q and ¬Q =⇒ ¬P have the same truth table.

10 Show that P ⇐⇒ Q and (P ∧ Q) ∨ (¬P ∧ ¬Q) have the same truth


table.

1.3 Statement Forms and Equivalence


A statement form is an expression such as P =⇒ (Q∧(¬ R)), obtained by
combining some statement variables with connectives. Depending on what
actual statements are substituted for the variables, a form may take on value
T or F. We have already seen how this works for simple forms such as ¬P ,
P =⇒ Q, etc.

Some statement forms always take on the value T, no matter what actual
statements are substituted for the statement variables. These are called
tautologies. Others, called contradictions, only take on the value F.
Example 1.3.1 The statement form P ∨(¬P ) is a tautology, while P ∧(¬P )
is a contradiction, as seen from their truth tables:

P ¬P P ∨ (¬P ) P ¬P P ∧ (¬P )
T F T T F F
F T T F T F 
16 CHAPTER 1. SETS AND LOGIC

Checking whether a statement form is a tautology by truth tables gets


cumbersome when it has many variables and connectives. It is more ecient
to develop ways of simplifying the statement form till it becomes obvious
whether it is a tautology. First, we consider two statement forms A and B
to be equivalent (and write A ≡ B ) if A is true exactly when B is true.
For example, all tautologies are equivalent to each other. And all contra-
dictions are equivalent to each other.

Task 1.3.2 Let T be a tautology and C a contradiction. Show that

T ∨ P ≡ T, T ∧ P ≡ P, C ∨ P ≡ P, C ∧ P ≡ C.

Example 1.3.3 The statement forms P =⇒ Q and ¬P ∨ Q are equivalent:

P Q P =⇒ Q P Q ¬P ¬P ∨ Q
T T T T T F T
T F F T F F F
F T T F T T T
F F T F F T T

Both the forms take value F when P Q is false, and


is true and the value
T in all other cases. Hence we can write P =⇒ Q ≡ ¬P ∨ Q. 

Task 1.3.4 Show that P =⇒ (Q =⇒ R) is not equivalent to (P =⇒


Q) =⇒ R.

You may have noticed that the equivalence of statement forms appears
similar to biconditionality. Let A and B be two statement forms. They
are equivalent if A and B always have the same truth values. That is, if
A ⇐⇒ B is a tautology.

How is the biconditional dierent from equivalence? The biconditional is


an operation, it combines any two statements to create another statement,
« or any two statement forms to create another statement form. Equiva-
lence is a relation between certain statement forms. It holds when the
requirements described above are met.

Let us put down some fundamental equivalences.

Theorem 1.3.5 (The Laws of Logic)

Undergraduate Mathematics: Foundations


1.3. STATEMENT FORMS AND EQUIVALENCE 17

1. (Idempotence) A ∨ A ≡ A ≡ A ∧ A.
2. (Commutativity) A ∨ B ≡ B ∨ A, A ∧ B ≡ B ∧ A.
3. (Associativity) A ∨ (B ∨ C) ≡ (A ∨ B) ∨ C , A ∧ (B ∧ C) ≡ (A ∧ B) ∧ C .
4. (Distributivity) A ∧ (B ∨ C) ≡ (A ∧ B) ∨ (A ∧ C), A ∨ (B ∧ C) ≡
(A ∨ B) ∧ (A ∨ C).
5. (Absorption) A ∧ (A ∨ B) ≡ A ∨ (A ∧ B) ≡ A.

Proof: Each of these can be proved by drawing up the truth-tables and


checking that the left side is true exactly when the right side is true. They
can also be proved by assuming that one side is true or false, and showing
that the other side has the same truth-value. We illustrate this for the
distributive law A ∧ (B ∨ C) ≡ (A ∧ B) ∨ (A ∧ C).
First, assume that A∧(B ∨C) is true. Then A and B ∨C are individually
true. The truth of B ∨ C gives us two cases:
1. B is true: Then A∧B is true, hence (A ∧ B) ∨ (A ∧ C) is true.

2. C is true: Then A∧C is true, hence (A ∧ B) ∨ (A ∧ C) is true.

Next, suppose that A ∧ (B ∨ C) is false. Again we have two cases.

1. A is false: Then A∧B and A∧C are false. Hence (A ∧ B) ∨ (A ∧ C)


is false.

2. B∨C is false: Then B and C are both false. Since B is false, A ∧ B is


false. Since C A ∧ C is false. Hence (A ∧ B) ∨ (A ∧ C) is false.
is false,

Task 1.3.6 Prove all the other parts of Theorem 1.3.5.

Negations

The correct interpretation of negations is very important for mathematical


proofs. So we shall now explore how to negate compound statements.

Task 1.3.7 Use a truth table to conrm that for any statement form A we
have

¬(¬A) ≡ A.
18 CHAPTER 1. SETS AND LOGIC

Theorem 1.3.8 (de Morgan's Laws) For any two statement forms A and
B we have:

¬(A ∧ B) ≡ ¬A ∨ ¬B, ¬(A ∨ B) ≡ ¬A ∧ ¬B.

Proof: We work out the truth tables for ¬(A ∧ B) ≡ ¬A ∨ ¬B . The other
law can be derived in a similar fashion.

A B A∧B ¬(A ∧ B) A B ¬A ¬B ¬A ∨ ¬B
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

The nal columns of each table have the same truth-values. This estab-
lishes the desired equivalence. 

Task 1.3.9 Are any of the following statement forms logically equivalent to
each other?

(a) P =⇒ (Q =⇒ R),
(b) Q =⇒ (P =⇒ R),
(c) (P ∧ Q) =⇒ R.

Converse and Contrapositive

P =⇒ Q. Its converse is the statement


Consider a statement of the form
Q =⇒ P . We have already seen that the statement forms P =⇒ Q and
Q =⇒ P are not logically equivalent.

Task 1.3.10 Check that under any assignation of truth values to P and
Q, at least one of P =⇒ Q and its converse is true. Hence, (P =⇒
Q) ∨ (Q =⇒ P ) is a tautology.

Consider a statement of the form P =⇒ Q. Its contrapositive is


¬Q =⇒ ¬P .

Task 1.3.11 Check that P =⇒ Q is equivalent to its contrapositive.

Undergraduate Mathematics: Foundations


1.3. STATEMENT FORMS AND EQUIVALENCE 19

The equivalence of a statement with its contrapositive forms the basis of a


convenient and popular strategy called proof by contrapositive. Suppose
we have to prove that P =⇒ Q has truth value T for some choice of
P and Q. A direct proof would start by assuming the truth of P and
then deriving various consequences until the truth of Q was also obtained.
However, a proof by contrapositive starts by assuming that Q is false and
obtains as a consequence that P is false. Either way, we show that it is
impossible for both P to be true and Q to be false.

A related technique is proof by contradiction. Here, to prove P =⇒


Q, we assume that P and ¬Q are true. From these two assumptions we
proceed to derive that some statement R is both true and false. Since R
cannot be simultaneously true and false, it follows that P and ¬Q cannot
both be true. This establishes the truth of P =⇒ Q.
These basic proof techniques will be encountered many times during your
reading of this book. Let us illustrate how we can use them to approach
Theorem 1.2.2 and bypass the need for truth-tables. First, we restate that
theorem.

Theorem 1.3.12 If P =⇒ Q and Q =⇒ R are true then P =⇒ R is


also true.

Proof: (Direct) We assume that (P =⇒ Q) ∧ (Q =⇒ R) is true. Now,

(P =⇒ Q) ∧ (Q =⇒ R) ≡ (¬P ∨ Q) ∧ (¬Q ∨ R)
≡ (¬P ∧ ¬Q) ∨ (¬P ∧ R) ∨ (Q ∧ ¬Q) ∨ (Q ∧ R)
≡ (¬P ∧ ¬Q) ∨ (¬P ∧ R) ∨ (Q ∧ R).

At least one of the three bracket terms in the last expression is true. Let's
do the three cases:

1. ¬P ∧ ¬Q is true. Then P is false, hence P =⇒ R is true.

2. ¬P ∧ R is true. Then P is false, hence P =⇒ R is true.

3. Q∧R is true. Then R is true, hence P =⇒ R is true. 


Proof: (By Contrapositive) We assume that P =⇒ R is false. Then P is
true and R is false. We have two cases:

1. Q is true. Since R is false, in this case Q =⇒ R is false.

2. Q is false. Then P is true, in this case P =⇒ Q is false.


20 CHAPTER 1. SETS AND LOGIC

In both cases, (P =⇒ Q) ∧ (Q =⇒ R) is false. 


Proof: (By Contradiction) Assume P =⇒ Q and Q =⇒ R are true and
P =⇒ R is false. Then P is true andR is false.
Since P is true and P =⇒ Q is true, Q is true.

Since R is false and Q =⇒ R is true, Q is false.

Hence Q is both true and false. 


Did you like any of these proofs more than the others?

Exercises for Ÿ1.3

1 Recall the exclusive or operator t dened in Exercise 4 on page 8. Find


a statement form that is equivalent to P tQ and uses only the operators ¬,
∧, ∨.

2 Formulate a statement form P ∗Q using only ¬, ∧, ∨ whose truth table


will be

P Q P ∗Q P Q P ∗Q
T T T T T T
(a) T F F (b) T F T
F T T F T T
F F F F F F

3 Is (P ∧ Q) =⇒ (P ∨ Q) a tautology?

4 Prove the logical equivalence [(P ∧ ¬Q) ∨ Q] ⇐⇒ (P ∨ Q) using:

(a) Truth tables,

(b) The laws of logic (including de Morgan's laws), stating each law or
previously established consequence where it is used.

5 Are the following pairs of statements logically equivalent?

(a) P =⇒ (Q ∨ R) and (P =⇒ Q) ∨ (P =⇒ R),


(b) (P ∨ Q) =⇒ R and (P =⇒ R) ∨ (Q =⇒ R).

6 Are the following logical equivalences correct?


   
(a) P =⇒ (Q ∧ R) ≡ (P =⇒ Q) ∧ (P =⇒ R) ,

Undergraduate Mathematics: Foundations


1.4. OPEN SENTENCES AND QUANTIFIERS 21

  
(b) (P ∧ Q) =⇒ R] ≡ (P =⇒ R) ∧ (Q =⇒ R) .

7 Formulate the following symbolically (using ∧, ∨, ¬, =⇒ , P , Q and R),


and then write their converse and contrapositive:

(a) If not P or not Q, then it is not the case that either P or Q.


(b) If P implies that Q implies R, then P and Q together imply R.

8 For each of the six statement forms you have written in the previous
exercise, nd out whether it is a tautology or a contradiction (or neither).

9 Show that if P =⇒ Q is true then P =⇒ (P ∧ Q) is true, by

(a) A direct proof,

(b) A proof by contrapositive,

(c) A proof by contradiction.

10 Prove that P ⇐⇒ ¬Q is a necessary and sucient condition so that


P ∨Q is true and P ∧ Q is false.

11 Show that any statement form is logically equivalent to one that only
uses the implication and negation operators.

1.4 Open Sentences and Quantiers


Consider the sentence

The integer n is an odd number.

We cannot assign this a truth value because we don't know the value of n.
On the other hand, if we knew that n had a certain value then this sentence
would become a statement. For example, if we knew that n = 99 then the
sentence would become a true statement:

The integer 99 is an odd number.

An open sentence is a sentence involving variables (such as x, y, . . . )


that becomes a statement whenever all the variables are replaced by actual
values. In general, the truth value of an open sentence changes with the
values of the concerned variables. We use notation like P (x, y, . . . ) to denote
an open sentence that involves variables x, y , ...
22 CHAPTER 1. SETS AND LOGIC

Example 1.4.1 Consider the open sentence

P (n) : The integer n is an odd number.

Then P (99) is the statement

The integer 99 is an odd number.

Or, consider the open sentence Q(x, y): x + y = 0. Then Q(2, 3) is the (false)
statement 2 + 3 = 0. 

Given some open sentences, we can combine them using ∧, ∨, ¬ and =⇒


to create compound open sentences.

Example 1.4.2 We shall write the following open sentence in symbolic


form:

If a gure x is a rhombus, then it is a square.

Let R(x) denote the open sentence The gure x is a rhombus and S(x)
denote The gure x is a square. Then the given sentence is expressed by
R(x) =⇒ S(x). 

Task 1.4.3 Let P (n) be the open sentence The integer n is an odd number.
Express the following symbolically: If 5 and 7 are odd then 35 is odd.

Consider an open sentence P (x). We can convert it into a statement by


substituting a specic value of x. Two other typical statements that can be
obtained from this open sentence are:

1. For every value of x, P (x) is a true statement.

2. For some value of x, P (x) is a true statement.

To cover these examples we introduce the symbols ∀ ( for all) and ∃ (there
exists): 6

1. ∀x P (x) means For every value of x, P (x) is a true statement.

2. ∃x P (x) means For some value of x, P (x) is a true statement.

The symbols ∀ and ∃ are called quantiers. The symbol ∀ is called the
universal quantier while ∃ is called the existential quantier.
6 Did you notice that these symbols are just A (`All') and E (`Exists') rotated by 180◦ ?

Undergraduate Mathematics: Foundations


1.4. OPEN SENTENCES AND QUANTIFIERS 23

Example 1.4.4 The statement All squares are rectangles can be written
as
∀x (S(x) =⇒ R(x)),
where S(x) is the statement The object x is a square and R(x) is The
object x is a rectangle. The symbolic expression ∀x (S(x) =⇒ R(x)) is
read as Every object x which is a square is also a rectangle.

Similarly, the statement Some rectangles are squares would be expressed


as
∃x (R(x) ∧ S(x)).
This is read as There is at least one object x which is a rectangle and also
a square. 

A point of convention here is that some is interpreted as at least one and


doesn't imply plurality.

Task 1.4.5 In the notation of the above example, would ∃x (R(x) =⇒


S(x)) be a correct rendering of Some rectangles are squares?

Task 1.4.6 Is the following a true statement?

There is an object which, if it is a square, is also a circle.

(Don't hurry over this one! It may help to express it symbolically.)

The previous tasks illustrate that ∃x (P (x) =⇒ Q(x)) does not have
the meaning we may expect it to have. While it is allowed to write such
« a statement, it is unlikely to come up in any meaningful discussion. For
it becomes true as soon as there is a value of x which makes P (x) false.
Implication goes naturally with ∀.

Example 1.4.7 Let us express the following statement symbolically:

There is a natural number whose factors are all even.

First, we identify the open sentences that are used in this statement. Let
N (x) stand for  x is a natural number and E(x) stand for  x is an even
number. Also, let F (x, y) stand for  y is a factor of x.
The statement asserts the existence of a natural number with certain
properties. So it will have the form

∃x [N (x) ∧ · · · ],
24 CHAPTER 1. SETS AND LOGIC

where the dots ··· indicate the property that we have to include. The prop-
erty is about every natural number which is a factor of x. So our statement
becomes
∃x [N (x) ∧ ∀y [(N (y) ∧ F (x, y)) · · · ]].
We have brought in all the factors y of x. Now we just have to assert that if
something is such a factor then it is even. So the nal form of the statement
is
∃x [N (x) ∧ ∀y [(N (y) ∧ F (x, y)) =⇒ E(y)]].


When a statement has two or more quantiers, the order in which they
appear is critical to the meaning of the statement. The consequences of
« writing ∀x ∃y (· · · ) can be quite dierent from those of ∃y ∀x (· · · ). For
an example, see Exercise 6 below.

Task 1.4.8 Express the following statements symbolically using ∧, ∨, ¬,


=⇒, ∀ and ∃, as well as appropriate mathematical notation:

1. There is an even prime number.

2. Not all integers are positive.

3. Every whole number is an integer.

4. There is an integer that is both negative and positive.

5. There is a natural number with at least one odd factor.

6. The product of any two odd integers is an odd integer.

Next, we consider the negations of statements with quantiers. It is


obvious that a statement of the form ∀x A(x) is false if and only if there is
some x such that A(x) is false. In other words, the negation of ∀x A(x) is
∃x ¬A(x). Similarly, the negation of ∃x A(x) is ∀x ¬A(x).

Example 1.4.9 Let us negate the statement All squares are rectangles.
Recall that it can be written as

∀x (S(x) =⇒ R(x)),

where S(x) is the sentence  x is a square and R(x) is  x is a rectangle. The


negation is
∃x ¬(S(x) =⇒ R(x)).

Undergraduate Mathematics: Foundations


1.4. OPEN SENTENCES AND QUANTIFIERS 25

Now ¬(S(x) =⇒ R(x)) ≡ ¬(¬S(x)∨R(x)) ≡ S(x)∧¬R(x). So the negation


can also be expressed as

∃x [S(x) ∧ ¬R(x)],
which is read as There is a square that is not a rectangle.

Similarly, the statement Some rectangles are squares is expressed as


∃x (R(x) ∧ S(x)) and has negation ∀x ¬(R(x) ∧ S(x)). We have

¬(R(x) ∧ S(x)) ≡ ¬R(x) ∨ ¬S(x) ≡ R(x) =⇒ ¬S(x).


So the negation is ∀x [R(x) =⇒ ¬S(x)], which is read as Everything that
is a rectangle is not a square or No rectangle is a square. 

Task 1.4.10 Are the following pairs of statements equivalent?

(a) Every rectangle is a square and Every square is a rectangle.

(b) Some rectangle is a square and Some square is a rectangle.

(c) No rectangle is a square and No square is a rectangle.

Task 1.4.11 For each statement in Task 1.4.8 express the negation symbol-
ically and then in English.

Task 1.4.12 Are the following statements true or false?

(a) All unicorns are white.

(b) No unicorns are white.

(c) Some unicorns are white.

Task 1.4.13 Can you imagine circumstances under which No roses are red
would be the negation of All roses are red? Or is that impossible?

Often we want to make a stronger assertion than the existence of an object


with a property: We also want to assert the uniqueness of that object. For
example, zero not only is an integer that doesn't change other integers on
addition, it is the only such integer. For this we create the symbol ∃!. The
statement ∃!x P (x) is read as There exists exactly one x such that P (x) is
true.

For example, the existence and uniqueness of zero among the integers
can be expressed as

∃!x [I(x) ∧ ∀n (I(n) =⇒ x + n = n)],


26 CHAPTER 1. SETS AND LOGIC

where I(t) is the open sentence The object t is an integer.

The notions of converse and contrapositive are adapted to universally


quantied implications as follows:

• The converse of ∀x (P (x) =⇒ Q(x)) is ∀x (Q(x) =⇒ P (x)).


• The contrapositive of ∀x (P (x) =⇒ Q(x)) is ∀x (¬Q(x) =⇒
¬P (x)).

Task 1.4.14 Show that ∀x (P (x) =⇒ Q(x)) is logically equivalent to its


contrapositive.

Task 1.4.15 Is it true that either ∀x (P (x) =⇒ Q(x)) or its converse must
be true?

Free and Bound Variables

Consider the familiar open sentence:

The integer x is even.

To express this formally, we need to bring in the meaning of `even'. This


requires the introduction of a new variable, say n, which takes the value of
an integer such that x is twice of n:
 
∃n I(n) ∧ (x = 2n) .

Here, I(n) is  n is an integer. The two variables x and n have very dierent
roles. The value of x is imposed from outside the sentence and determines
whether it will become true or false. If we write

 
∃n I(n) ∧ (y = 2n) ,

we have a sentence about y and its truth value is to be determined indepen-


dently of that for the sentence with x. We call x a free variable. On the
other hand, n is internal to the sentence and its values are not externally
imposed. If we used some other variable instead, for example by writing

 
∃m I(m) ∧ (x = 2m) ,

the truth value of the sentence would not be aected. We call n a bound
variable.

Undergraduate Mathematics: Foundations


1.4. OPEN SENTENCES AND QUANTIFIERS 27

This terminology covers the universal quantier as well. The statement


Every prime number is odd can be expressed as

∀x (P (x) =⇒ ∃n (I(n) ∧ x = 2n + 1))

where P (x) is  x is a prime number. Here, x and n are bound variables


and replacing them by other symbols has no impact on the content of the
statement. This statement has no free variables.

Exercises for Ÿ1.4

1 Let N (x) be the open sentence  x is a natural number. Using only this
open sentence, with a suitable choice of variables and quantiers, express the
following symbolically:

(a) M is an even natural number.

(b) M is an odd natural number.

(c) M is a prime number.

2 Let P (n) be an open sentence in which n varies over the natural numbers.
Will the truth of the following imply the truth of P (99)?

(a) ∀n P (n). (b) ∃n P (n).

3 Does the truth of Every dragon is red imply the truth of There is a
red dragon? (Here, we take a break from our decision to only talk about
collections of mathematical objects.)

4 Do the quantiers distribute over conjunction and disjunction? In other


words, are the following statement forms equivalent?

(a) ∃x (P (x) ∧ Q(x)) and (∃x P (x)) ∧ (∃x Q(x)),


(b) ∃x (P (x) ∨ Q(x)) and (∃x P (x)) ∨ (∃x Q(x)),
(c) ∀x (P (x) ∧ Q(x)) and (∀x P (x)) ∧ (∀x Q(x)),
(d) ∀x (P (x) ∨ Q(x)) and (∀x P (x)) ∨ (∀x Q(x)).

5 Formulate the following in logical symbolism using quantiers, and then


write their negations:

(a) For any integer x there is at least one integer y such that x + y = 1.
(b) For any integer x there is exactly one integer y such that x + y = 1.
28 CHAPTER 1. SETS AND LOGIC

6 Express the following expressions in words as simply as you can. It is


given that the variables x and y take values from the natural numbers.

(a) ∃x ∀y (x ≥ y), (b) ∀y ∃x (x ≥ y).

7 Are the following statement forms equivalent?

(a) ∃x ∃y P (x, y) and ∃y ∃x P (x, y),


(b) ∀x ∀y P (x, y) and ∀y ∀x P (x, y),
(c) ∃x ∀y P (x, y) and ∀y ∃x P (x, y).

8 P (x) be the open sentence The object x is a point in the Euclidean


Let
plane,L(m) be The object m is a line in the Euclidean plane, and O(x, m)
be The point x lies on the line m. Express the following axioms of Euclidean
geometry symbolically:

(a) Given any two distinct points in the plane, there is a unique line
through them.

(b) Every line in the plane has at least two points.

(c) The plane has at least three non-collinear points.

(d) Given a line m and a point x not on m, there is a unique line through
x m.
that is parallel to

9 Are the following always true?

  
(a) ∀x P (x) =⇒ (Q(x) ∨ R(x)) ⇐⇒ ∀x (P (x) =⇒ Q(x))∨

∀x (P (x) =⇒ R(x))
 
(b) ∀x (P (x) ∨ Q(x)) =⇒ R(x)] ⇐⇒ ∀x (P (x) =⇒ R(x))∨

∀x (Q(x) =⇒ R(x))
  
(c) ∀x P (x) =⇒ (Q(x) ∧ R(x)) ⇐⇒ ∀x (P (x) =⇒ Q(x))∧

∀x (P (x) =⇒ R(x))

10 Let x take values over real numbers and let f vary over the real functions
(functions whose domain and codomain consist of real numbers). Consider
the following open statements:

C(f, x) : f is continuous at x
D(f, x) : f is dierentiable at x

Undergraduate Mathematics: Foundations


1.5. AXIOMS FOR SETS 29

Express the following statements symbolically using the logical operators,


quantiers, the variables x and f, and the open statements C(f, x) and
D(f, x):
(a) If any real function f is dierentiable at a real number x then it is
continuous at x.
(b) There is a real function f which is continuous at every real number
but is dierentiable at no real number.

11 Identify the free and bound variables in the following:

(a) ∃x (M = 2x),
(b) ∀x ∀y [(M = xy) =⇒ (x = 1 ∨ x = M )],
(c) (∀x (x = x)) ∧ (∃y (y 6= x)).

1.5 Axioms for Sets


Our initiation of set theory in this section will help us to consistently describe
all the mathematical objects we'll study in this book. Indeed, over the course
of the book, we will glimpse a remarkable fact  that it is possible to represent
every object studied in mathematics as a set. The importance of the notion of
a set was pointed out by George Boole in the nineteenth century. Set theory
was established a little later by Georg Cantor, who used it to bring clarity
to the notion of innity and whose results astonished and even oended his
contemporaries.

You have probably seen a set being described in one or more of the
following ways:

• A collection of well-dened objects.

• A well-dened collection of objects.

• A collection of objects with some common properties.

None of these is very satisfactory. What does well-dened mean, and does
it matter whether it qualies the collection or the objects? The usual inter-
pretation is that well-dened refers to clarity about whether any particular
object belongs to the candidate set or not. Given a set and an object we
should be able to say, without ambiguity or doubt, whether the object is a
member of the set.
30 CHAPTER 1. SETS AND LOGIC

How do we actually decide whether an object is in a set? This is where


the third description comes in. It says that we should have a list of properties
such that an object belongs to the set if and only if it has all those proper-
ties. What is a property ? This is just another name for an open sentence.
For example, the open sentence I(x) :  x is an integer corresponds to the
property of being an integer.

Remarkably, this common-sense idea of creating sets through properties


can create problematic situations.

Example 1.5.1 (Russell's Paradox) The members of a set can be sets


themselves. Therefore we can ask whether a set can be a member of itself.
It is dicult to imagine such a set, but a set being a member of itself would
be a property and so we should be able to use it to dene a set.

Let C be the collection of all sets that are not members of themselves.

The critical question is whether C is a member of itself. The answer ought


to be either yes or no. But something else happens. For suppose C is a
member of itself. Then it violates the condition for being a member of C and
we have a contradiction. On the other hand, if it is not a member of itself,
it does satisfy the condition for being a member of C and we again have a
contradiction. So the answer is neither yes nor no!

The only way to resolve this paradox is to not call C a set. 

Russell's Paradox tells us that we need a careful description of sets, one


that keeps out collections that could create paradoxes, yet includes all the
collections that we need in mathematics. It is not enough to just keep out
the C of the above example, as there could be other paradoxes.

The Axioms of Set Theory

When we rst encounter sets in a mathematics classroom, the typical illus-


trative examples are collections of certain everyday objects. For example,
we might be asked to consider the following:

• The set of all students in the school in which you study.

• The set of all citizens of India.

• The set of all member countries of the United Nations.

• The set of all cars registered in Delhi.

Undergraduate Mathematics: Foundations


1.5. AXIOMS FOR SETS 31

Although such collections are useful for organizing our thoughts they are
not, strictly speaking, part of mathematics. For the objects in these sets
are not mathematical objects. Cars, dogs, tables, countries and citizens
cannot be dened purely by mathematical means. On the other hand, the
set of all even integers can be conceived purely within mathematics. Set
theory is concerned only with such sets. You may well ask here, How am
I to recognise a mathematical object? The answer we adopt, the standard
answer for about a century now, is that sets themselves are the mathematical
objects. All the objects and operations of mathematics can be viewed as sets!
This may seem unreasonable at the moment but as you read this book the
details will become clear.

The concepts of sets and the relationship of membership between sets,


are the fundamental concepts of set theory and are not dened further. All
we can do is x the axioms they follow. The rst axiom system for set
theory was created by Ernst Zermelo, and was rened over subsequent years
by Adolf Fraenkel, Thoralf Skolem, John von Neumann, Paul Bernays and
Kurt Gödel. We will present a version of the axioms of set theory that
follows the approach of Zermelo, Fraenkel and Skolem and which is known
as the Zermelo-Fraenkel system or ZF system. We won't present all
the standard ZF axioms right away. We defer the Axioms of Regularity and
Innity till we really need them (to create the set of natural numbers, in
Chapter 4). Further, the Axiom of Replacement is not required in this book
and we state it only in the Appendix (page 197) and then in the context of
a dierent axiomatization.

These axioms assume a universal domain called B consisting of sets. In


general, we will denote sets by upper case letters such as A
B . Further,
and
there is a relationship called membership between sets. If A and B are sets
then A may be a member of B , or B may be a member of A, or neither may
be a member of the other. If A is a member of B we write A ∈ B and treat
it as a statement with truth value T. If A is not a member of B we say that
A ∈ B has truth value F. In the last case, we may also write A ∈
/ B and call
it a true statement. Thus A ∈ / B is the negation of A ∈ B .

We shall often use lower case letters such as x and y for sets that occur
as members of other sets in our discussion. This is merely a convention
that provides visual cues which are useful for following an argument. The
notation x∈A is also read as  x is an element of A or  x belongs to A or
 x is in A.

We consider two sets to be equal if they can substitute for each other
32 CHAPTER 1. SETS AND LOGIC

in any statement without altering its truth value. Our rst axiom describes
just when two sets can be considered to be equal in this manner.

ZF1. Axiom of Extension: Two sets are equal if and only if they have the
same members.

Given the work we have put in to learn how to express statements symboli-
cally, we should try our hand at expressing the ZF axioms symbolically. For
that, we need to include the quantiers ∀ and ∃, with the understanding
that they apply to our universe B . That is, ∀A stands for for every set A
and ∃A stands for there is a set A. The Axiom of Extension can then be
expressed as:

∀A ∀B [A = B ⇐⇒ ∀x (x ∈ A ⇐⇒ x ∈ B)].

The next axiom gives us a set which we can use as a basic building block,
and the following two axioms give us ways to combine known sets to get new
sets:

ZF2. Axiom of the Empty Set: There is a set that has no members. It is
called the empty set.
This axiom can be expressed symbolically as follows: ∃A ∀x (x ∈
/ A).
ZF3. Axiom of Pairs: Given sets x, y there is a set whose members are x
and y alone.

Task 1.5.2 Express the Axiom of Pairs symbolically.

ZF4. Axiom of Union: If F is a set then there is a set called the union of
F whose members are exactly the members of the members of
S S F. The
union of F is denoted by F or A∈F A.

Task 1.5.3 Express the Axiom of Union symbolically.

Let us note some consequences of these axioms:

• Due to ZF1, the empty set of ZF2 is unique. We shall denote it by ∅.


• Due to ZF1, the set postulated by ZF3 is unique. We denote it by
{x, y}.
• If we take x=y in ZF3, we get the set {x} with x as the only member.
• Due to ZF1, the union described by ZF4 is unique.

Undergraduate Mathematics: Foundations


1.5. AXIOMS FOR SETS 33

S
• Consider sets A and B . Using ZF3 we obtain F = {A, B}. Then F
is also denoted by A ∪ B and is called the union of A and B .

Task 1.5.4 Show that for any sets x, A and B, we have x ∈ A ∪ B ⇐⇒


(x ∈ A) ∨ (x ∈ B).

Task 1.5.5 Show that ∅ 6= {∅}.

Our method of using {x} to represent the set whose only member is x,
and {x, y} to represent the set whose only members are x and y , is called the
roster notation for describing a set. One can use this for sets with more
elements as well. For example if the set has only a, b, c, d, e as members
then we can represent the set by {a, b, c, d, e}. In the other direction, the
empty set is represented by {}.

Example 1.5.6 It would be inconvenient to describe a large set by actually


listing all its members. If the members of a set follow some pattern, that
pattern can be used to describe the set:

Description of the set Roster notation


All the natural numbers from 1 to 100 {1, 2, 3, . . . , 100}
The rst fty even numbers {2, 4, 6, . . . , 100}
The rst ten perfect squares {12 , 22 , 32 , . . . , 102 }
All the natural numbers {1, 2, 3, . . . }
All the integers {. . . , −2, −1, 0, 1, 2, . . . }

The group of three dots is called ellipsis and stands for and so on or and
so on, until. 

Starting with the empty set ∅ we can use ZF3 and ZF4 to create many
more sets:

E=∅ (ZF2)
0
E = {∅} (ZF3)
00 0 0
E = E ∪ {E } = {∅, {∅}} (ZF3, ZF4)

E 000 = E 00 ∪ {E 00 } = {∅, {∅}, {∅, {∅}}} (ZF3, ZF4)

.
.
.
34 CHAPTER 1. SETS AND LOGIC

These axioms allow us to build larger sets from smaller ones. Another
possibility is to extract smaller sets from a given set by selecting some of
its elements. To some extent this can be achieved by the axioms we already
have. For example, suppose we are given that A = {α, β, γ} is a set. We can
use Axiom ZF3 to create the sets {α}, {β}, {γ}, {α, β}, {α, γ}, {β, γ}.
A ⊆ B to mean  A is a subset of B , i.e. every member of
We shall use
A B . Note that this allows a set to be a subset of itself.
is also a member of
When A is a subset of B we also say that B is a superset of A and denote
this by B ⊇ A. The usage is akin to that of ≤ and ≥ for numbers.

You must distinguish between ∈ (is a member of ) and ⊆ (is a subset of ).


For example, consider A = {1, 2} and B = {1, 2, 3}. We look at each
member of A in turn, rst 1 and then 2, and nd they are also members
« of B. A ⊆ B is true. But A ∈ B is false as it would require {1, 2}
Hence
to be present as a single object in B , and it is not. On the other hand, if
we take C = {1, {1, 2}} then we see that A ∈ C is true and A ⊆ C is false.

A related warning is that the truth of x∈A and A∈B does not imply
«
x ∈ B.

The negations of these relationships are represented by striking through


their symbols:

A*B:A is not a subset of B,


A+B:A is not a superset of B.
When A is a subset of B but is not all of B , we call it a proper subset.
7
The notation A B or B ! A is used to denote A being a proper subset.
These denitions related to subsets can be expressed in symbols as follows:

A ⊆ B : ∀x (x ∈ A =⇒ x ∈ B),
A B : ∀x (x ∈ A =⇒ x ∈ B) ∧ ∃x (x ∈ B ∧ x ∈
/ A),
A * B : ∃x (x ∈ A ∧ x ∈
/ B).

Theorem 1.5.7 For any given set A, show that ∅ ⊆ A.

Proof: The statement ∅⊆A is equivalent to

∀x (x ∈ ∅ =⇒ x ∈ A).
7 It is common to use the simpler notation A ⊂ B to denote subsets. Unfortunately,
some authors use ⊂ to denote only proper subsets, while others allow it to include equality.
For this reason, we shall not use this notation in this book.

Undergraduate Mathematics: Foundations


1.5. AXIOMS FOR SETS 35

Since x∈∅ is always false, the implication is always true. 

Task 1.5.8 Show that A⊆B and B⊆C implies A ⊆ C.

A general approach for obtaining subsets is provided by the fth axiom:

ZF5. Axiom of Separation: Given a set A, the collection of those members


of A that satisfy a given property, is a set. The property should not
involve the subset we are trying to dene.

Let P (x) be the open sentence corresponding to a property that a set x may
8
have. Then the Axiom of Separation can be stated as follows:

∀A ∃B ∀x [x ∈ B ⇐⇒ (x ∈ A ∧ P (x))].
It is worth noting what ZF5 does not allow. It does not allow applying a
property to all objects of the universal domain B and then collecting the
ones that satisfy that property into a set. A property can only be used to
separate out subsets from something that is already known to be a set, and
we shall see a little later (Theorem 1.5.15) that B cannot be a set.

If A is a set then the subset consisting of members x for which P (x) is


true is denoted by { x ∈ A | P (x) } { x | (x ∈ A) ∧ P (x) }. This is called
or
the set-builder notation for describing a set. For example, let A be the
set of natural numbers from 1 to 100. Then the set of even natural numbers
between 1 and 100 can be expressed in set-builder notation by

{ x ∈ A | ∃k (k ∈ A ∧ x = 2k) }.

Example 1.5.9 Here are some sets expressed in both roster and set-builder
notations, with A being the set of natural numbers from 1 to 100.

Roster Set-builder
{} {x ∈ A | x2 = 2 }
{1} {x ∈ A | x2 = 1 }
{1, 2} {x ∈ A | x2 − 3x + 2 = 0 }
{1, 2, 3, . . . , 10} {x ∈ A | 1 ≤ x ≤ 10 } 

The requirement that the property should not involve the subset we are
trying to dene, is needed to avoid circular reasoning. For example, it would
be silly to dene a subset B of A by the expression B = {x ∈ A | x ∈ B }
and so our axioms must explicitly ban it.

8 We have a separate statement for each property. So this axiom is actually a collection
of axioms, one for each property, and is called an axiom schema.
36 CHAPTER 1. SETS AND LOGIC

Task 1.5.10 Show that dening B = {x ∈ A | x ∈


/ B} creates a contradic-
tion if A 6= ∅.

One application of ZF5 is to dene various set operations:

• intersection of two sets is A ∩ B = { x ∈ A | x ∈ B }.


The

• The dierence of A and B is A \ B = { x ∈ A | x ∈ / B }.


• If A ⊆ X , we dene the complement of A in X by Ac = X \ A.
If A ∩ B = ∅ we say that A and B are disjoint.

Task 1.5.11 Show that x ∈ A ∩ B ⇐⇒ (x ∈ A) ∧ (x ∈ B).

Task 1.5.12 Show that x ∈ A \ B ⇐⇒ (x ∈ A) ∧ (x ∈


/ B).

Theorem 1.5.13 If F is a set then there is a set whose members are exactly
those sets that are members of every member of
T F. (This set is called the
intersection
T
of F and is denoted by F or A∈F A.)

Proof: We use the Axioms of Union and Separation to dene the desired set
as follows:
\ [
F = {x ∈ F | ∀A (A ∈ F =⇒ x ∈ A) }.


The sixth axiom allows us to collect the subsets of a set into a set.

ZF6. Axiom of the Power Set: If A is a set then there is a set whose members
are precisely the subsets of A. This set is called the power set of A
and is denoted by P(A).
As an example, let A = {α, β, γ} be a set. Then

P(A) = {∅, {α}, {β}, {γ}, {α, β}, {α, γ}, {β, γ}, A}.

Task 1.5.14 Express the Axiom of the Power Set symbolically.

All the axioms so far are oriented towards allowing us to construct various
sets. Zermelo claimed that if we use only these axioms for constructing

Undergraduate Mathematics: Foundations


1.5. AXIOMS FOR SETS 37

9
sets, then we would not run into paradoxes. For example, he showed that
Russell's Paradox would be avoided by rst proving the following result:

Theorem 1.5.15 Every set M has a subset M0 that is not a member of M.

Proof: Consider the set M0 = { x ∈ M | x ∈


/ x }, dened using the Axiom of
Separation. Suppose M0 ∈ M .

This is analogous to the use of C in Russell's Paradox. The dierence is


7 that the members of M0 are being chosen from a set, so M0 is a valid set.
We then continue with the same basic question as in Russell's Paradox.

Does M0 belong to itself ? If yes, then by its denition, it must not belong
to itself, a contradiction. If no, then by its denition, it must belong to itself,
also a contradiction. So M0 ∈
/ M. 
This shows that the universal domain B cannot be taken to be a set in
this system of axioms. If it were a set, then every set would be a member
of the set B, contradicting the last theorem. One consequence is that we
cannot take the C of Russell's Paradox to be a set, since the property A∈
/A
is not being applied to select from a set.

The following tasks are meant to hone your skill in comprehending and
manipulating sets. Let us start with an example.

Example 1.5.16 We shall prove the following:

A = B ⇐⇒ (A ⊆ B and A ⊇ B).

First, suppose that A⊆B and A⊇B are both true. To prove A=B we
need to check that these two sets have the same members (by the Axiom of
Extension). Consider any member x of A. Since A ⊆ B , x is also a member
ofB . So every member of A is also a member of B . Now consider a member
y of B . Since A ⊇ B , y is a member of A. So every member of B is also
a member of A. Hence A and B have exactly the same members and we
obtain A = B .

Conversely, suppose that A = B is true. Then the two sets have the same
members. Since members of A are members of B , we have A ⊆ B . And since
members of B are members of A, we have A ⊇ B . 
9 And so far, he has not been proved wrong!
38 CHAPTER 1. SETS AND LOGIC

You may consider this a long-winded way to establish something obvious,


and you would be right. However, careful conrmation of the obvious is
an important part of learning new mathematics.

Task 1.5.17 Let A be a set. Use the ZF axioms to show the following:

(a) A ∪ ∅ = A, (b) A ∩ ∅ = ∅.

Task 1.5.18 Let A, B be sets. Show that each of the following implies the
others:

(a) A ⊆ B, (b) A ∪ B = B, (c) A ∩ B = A.

Task 1.5.19 Let A, B ⊆ X . Then A⊆B if and only if B c ⊆ Ac .

Theorem 1.5.20 Let A, B , C be subsets of a set U. Let all complements


be with respect to U. Then:

1. (Idempotence) A ∪ A = A ∩ A = A.
2. (Commutativity) A ∪ B = B ∪ A, A ∩ B = B ∩ A.
3. (Associativity) A ∪ (B ∪ C) = (A ∪ B) ∪ C , A ∩ (B ∩ C) = (A ∩ B) ∩ C .
A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C),
4. (Distributivity) A ∪ (B ∩ C) =
(A ∪ B) ∩ (A ∪ C).
5. (Absorption) A ∩ (A ∪ B) = A ∪ (A ∩ B) = A.
6. (Double complement) (Ac )c = A.
7. (de Morgan's Laws) (A ∪ B)c = Ac ∩ B c , (A ∩ B)c = Ac ∪ B c .

Proof: Each of these can be proved by patient applications of the Axiom


of Extension. But we can also prove them by utilizing the connections be-
tween union and disjunction, intersection and conjunction, complement and
negation. For example, the rst distributive law can be derived as follows:

x ∈ A ∩ (B ∪ C) ⇐⇒ (x ∈ A) ∧ (x ∈ B ∪ C)
⇐⇒ (x ∈ A) ∧ [(x ∈ B) ∨ (x ∈ C)]
⇐⇒ [(x ∈ A) ∧ (x ∈ B)] ∨ [(x ∈ A) ∧ (x ∈ C)]
⇐⇒ [x ∈ A ∩ B] ∨ [x ∈ A ∩ C]
⇐⇒ x ∈ (A ∩ B) ∪ (A ∩ C).

Undergraduate Mathematics: Foundations


1.5. AXIOMS FOR SETS 39

The rst of de Morgan's laws is obtained similarly:

x ∈ (A ∪ B)c ⇐⇒ (x ∈ U ) ∧ ¬(x ∈ A ∪ B)
⇐⇒ (x ∈ U ) ∧ ¬[(x ∈ A) ∨ (x ∈ B)]
⇐⇒ (x ∈ U ) ∧ [(x ∈
/ A) ∧ (x ∈
/ B)]
⇐⇒ [(x ∈ U ) ∧ (x ∈
/ A)] ∧ [(x ∈ U ) ∧ (x ∈
/ B)]
⇐⇒ (x ∈ Ac ) ∧ (x ∈ B c )
⇐⇒ x ∈ (Ac ∩ B c ).

The others are left for you to work out. 


What kind of proof was this: direct, by contrapositive, or by contradic-
tion?

Exercises for Ÿ1.5

1 Express the following sets using the roster notation:

(a) The integer solutions of (x2 − 1)(x2 − 4) = 0,


(b) The subsets of {a, b},
(c) The prime factors of 60,

(d) The set of odd natural numbers.

2 Express the following sets using the set-builder notation:

(a) The lines parallel to a given line m in the plane (Use the notation of
Example 1.6.1),

(b) The collection of non-empty and proper subsets of a given set A,

3 Let A, B and C be sets. Explain, in detail, how you would use the set
theory axioms to dene:

(a) A ∪ B ∪ C, (b) A ∩ B ∩ C.

4 Let A, B and C be given sets. Write the converse of each of the following:

(a) If A⊆C and B⊆C then A ∪ B ⊆ C.


(b) If A⊆B and A⊆C then A ⊆ B ∩ C.
(c) If A⊆B and B⊆C then A ⊆ C.
40 CHAPTER 1. SETS AND LOGIC

What can you say about the truth values of the above statements and their
converses?

5 Suppose A ⊆ B. Show that the following hold for any set C:

(a) A ∪ C ⊆ B ∪ C, (b) A ∩ C ⊆ B ∩ C.

6 Let A = { x ∈ R | x2 > x + 6 } and B = { x ∈ R | x > 3 }, where R is the


set of real numbers. Are the following true?

(a) A ⊆ B, (b) B ⊆ A.

7 Show that the following are true for any sets A, B , C :


(a) (A ∪ B) ∩ C ⊆ A ∪ (B ∩ C),
(b) (A ∪ B) ∩ (A ∪ C) ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C) ∪ (B ∩ C).

8 Show that the following equalities fail for some sets A, B . Further, modify
the right hand side so that equality becomes correct for all sets.

(a) A \ (A \ B) = B , (b) A \ (B \ A) = A \ B .

9 Prove the following:

(a) B ⊆ A =⇒ A \ (A \ B) = B ,
(b) A ∩ B = ∅ =⇒ A \ (B \ A) = A \ B .

10 The symmetric dierence of two sets is dened by A4B = (A \ B) ∪


(B \ A). Prove the following:

(a) A4B = (A ∪ B) \ (A ∩ B),


(b) A4∅ = A and A4A = ∅,
(c) A4(B4C) = (A4B)4C ,
(d) A ∩ (B4C) = (A ∩ B)4(A ∩ C).

1.6 Sets, Quantiers and Numbers


Typically, we need to x the set from which the values of a variable x may
be taken. For instance, we may want to say that

x < 0 =⇒ x(x − 1) > 0,

Undergraduate Mathematics: Foundations


1.6. SETS, QUANTIFIERS AND NUMBERS 41

provided that x is a real number. We introduce the notation ∀x ∈ A for


x in the set A, and we write (∀x ∈ A)P (x) as an alternative to
for all
(∀x)(x ∈ A =⇒ P (x)).
If we let R denote the set of real numbers, the above statement can be
written as:
(∀x ∈ R) [x < 0 =⇒ x(x − 1) > 0].
Earlier, we would have used an open sentence for this purpose, by setting
R(x) to be  x is a real number:
 
∀x (R(x) ∧ (x < 0)) =⇒ x(x − 1) > 0 .

The new notation is easier to read and write.

Example 1.6.1 Consider Exercise 8 of Ÿ1.4, on expressing some axioms of


Euclidean geometry. Let P be the set of points in the plane and L the set
of lines in the plane. Note that each line is a subset of P.
The axiom Given any two distinct points in the plane, there is a unique
line through them can now be expressed symbolically as:
 
(∀x ∈ P )(∀y ∈ P ) x 6= y =⇒ (∃!` ∈ L)(x ∈ ` ∧ y ∈ `) .

There is no need to introduce multiple new open sentences. The improvement


is even more marked for the parallel line axiom Given a line m and a point
x not on m, there is a unique line through x that is parallel to m:
 
(∀m ∈ L)(∀x ∈ P ) x ∈
/ m =⇒ (∃!` ∈ L)(x ∈ ` ∧ ` ∩ m = ∅) .

Task 1.6.2 Using the notation and approach of Example 1.6.1, express the
following axioms symbolically:

(a) Every line in the plane has at least two points.

(b) The plane has at least three non-collinear points.

How will negation work in this notation? Consider the statement that every
member of A has property P . It will be false if and only if there is a member
ofA that does not have property P . Hence, the negation of (∀x ∈ A) P (x) is
(∃x ∈ A) ¬P (x). Similarly, the negation of (∃x ∈ A) Q(x) is (∀x ∈ A) ¬Q(x).
We dene the converse and contrapositive of (∀x ∈ A) [P (x) =⇒ Q(x)]
as follows:
42 CHAPTER 1. SETS AND LOGIC

• The converse is (∀x ∈ A) [Q(x) =⇒ P (x)].


• The contrapositive is (∀x ∈ A) [¬Q(x) =⇒ ¬P (x)].

Task 1.6.3 Consider the following statements:

(a) (∀x ∈ R) [x < 0 =⇒ x(x − 1) > 0],


(b) (∀x ∈ R) [x > 0 =⇒ x(x − 1) > 0].
For each of these, write the converse and contrapositive. State whether any
of the original statement, the converse, and the contrapositive, are true.

Task 1.6.4 Prove that (∀x ∈ A) [P (x) =⇒ Q(x)] and its contrapositive
are equivalent.

Sets of Numbers

Let us undertake a brief review of some basic sets that we use all the time
in mathematics, and which we have already been referring to in this book
without much comment. These are the various sets of numbers, and their
subsets. For the time being we shall take it for granted that these are sets
and that they can be constructed on the basis of the axioms of set theory.
We'll return to this issue in Chapter 3. Our desire right now is to just collect
the basic notation and facts so that we can use them as examples to illustrate
and explore the various concepts of set theory.

Natural Numbers : N = {1, 2, 3, . . . }


Whole Numbers : W = N ∪ {0} = {0, 1, 2, 3, . . . }

: Z = {. . . , −2, −1, 0, 1, 2, . . . }
Integers
 
p
Rational Numbers : Q = p ∈ Z, q ∈ Z, q 6= 0
q

Do you see the diculty in accepting {1, 2, 3, . . . } as a denition of N? The


axioms mentioned so far only allow us to construct nite sets. Similarly
« the description of Q does not follow our approved format for the set-builder
notation, where the members have to be selected from a pre-existing set.
So these are informal descriptions and not formal denitions.

One can see that the natural numbers are the fundamental objects. Every
other system of numbers is eventually dependent on them. In school, the
natural numbers are introduced as a means of counting, and the integers

Undergraduate Mathematics: Foundations


1.6. SETS, QUANTIFIERS AND NUMBERS 43

and rational numbers help us to frame and answer more involved questions
related to counting. For example, the question If two loaves of bread are
completely and equally shared among ve girls, how much will each girl
get? has no sensible answer if we only know natural or whole numbers, but
is easily answered if we understand rational numbers.

Apart from counting, we need numbers for measuring magnitudes such as


lengths, areas, weights, and so on. And here we nd that the rational num-
bers, though capable of very ne precision, are unable to exactly represent
all magnitudes. This is known from the following famous example.

Consider a right triangle ABC whose perpendicular sides AB and BC


have unit length. Then, by the theorem of Pythagoras, the length ` of the
hypotenuse AC must satisfy `2 = 12 + 12 = 2. But there is no rational
number whose square is 2!

The set of real numbers, denoted by R, is introduced to resolve this lack


in rational numbers. The initial approach, the one you would have seen in
school, is geometric. We rst mark o a unit length on a straight line and let
it represent the number 1. We then declare that each length marked on this
`real line' represents a `real number'. We create positive and negative real
numbers by allowing each length to have a choice of direction. This geometric
approach suces for the needs of school mathematics, but it seems to set
the real numbers apart from the earlier number systems (N, W, Z, Q) which
do not need geometry.

The essentials of this geometric approach go back to the ancient Greeks.


However, they did not consider these geometric entities as genuine num-
bers. Credit is usually assigned as follows: Eudoxus gave the basic princi-
ples governing magnitudes, Euclid wrote a comprehensive treatment, and
10
Archimedes applied them to resolve many problems of area and volume.

It was only in the nineteenth century that the German mathematician


Richard Dedekind, dissatised with what passed for proofs in typical calculus
courses, identied a way to base the real numbers also on the natural numbers
without invoking geometry. In Chapter 3 we will catch a small glimpse of
his work, insofar as we will describe the axiomatic approach to real numbers
11
initiated by him.

The set of complex numbers is denoted by C. It consists of expressions

10 Eudoxus of Cnidus (c. 390 to c. 340 BC), Euclid of Alexandria (born c. 300 BC),
Archimedes of Syracuse (c. 287-212 BC). The dates are approximate.
11 A construction of real numbers from rational numbers is described in ŸA.3.
44 CHAPTER 1. SETS AND LOGIC

of the form a + ib where a and b are real numbers and the symbol i is dened
to have the property i2 = −1.
Now we quickly summarize some denitions and notation involving num-
bers and sets of numbers. We will use these in our examples and exercises.
However, we shall not use these sets in any theorems until we have described
them formally in Chapter 3.

• a and b are members of Z, we say a divides b, denoted a | b, if there


If
k ∈ Z such that b = ka. If a | b, we call a a factor of b, and b a
is
multiple of a.
• A number p ∈ N is called a prime number if p 6= 1 and the only
factors of p in N are 1 and p.
• If a ∈ N, we dene aN to be the set of multiples of a in N. We similarly
dene aZ.
• Z+ = Z>0 = N is the set of positive integers.Z− = Z<0 is the set of

negative integers, also denoted by −N. Z = { n ∈ Z | n 6= 0 } is the
set of non-zero integers.

• Z≥0 = W is the set of non-negative integers. Z≤0 = {0} ∪ Z<0 is the


set of non-positive integers.

• Similarly, for real numbers:

R>0 = R+ = { x ∈ R | x > 0 }, R<0 = R− = { x ∈ R | x < 0 },


R≥0 = { x ∈ R | x ≥ 0 }, R≤0 = { x ∈ R | x ≤ 0 },
R∗ = { x ∈ R | x 6= 0 }.

• The notation introduced above for R can also be adapted for Q. For
example, Q∗ = { x ∈ Q | x 6= 0 } and Q+ = { x ∈ Q | x > 0 }.
• Given two real numbers a and b, the greater one is called their max-
imum and is denoted by max{a, b}. The smaller one is called their
minimum and is denoted by min{a, b}. If a and b are equal their
common value serves as both maximum and minimum.

• A bounded interval in R is a subset of R consisting of all real num-


bers between two given real numbers. Depending on whether the two
given numbers are included or not, we get dierent kinds of intervals.
Suppose the given real numbers are a and b. Then they are called the
end-points of the following intervals:

Undergraduate Mathematics: Foundations


1.6. SETS, QUANTIFIERS AND NUMBERS 45

◦ (a, b) = { x ∈ R | a < x < b } open interval.


is called an

◦ [a, b] = { x ∈ R | a ≤ x ≤ b } is called a closed interval.


◦ [a, b) = { x ∈ R | a ≤ x < b } and (a, b] = { x ∈ R | a < x ≤ b }
half-open intervals.
are called

• An unbounded interval in R is either the full set R or consists of


all the numbers lying to one side of a given end-point a:
◦ (a, ∞) = { x ∈ R | a < x } and [a, ∞) = { x ∈ R | a ≤ x },
◦ (−∞, a) = { x ∈ R | x < a } and (−∞, a] = { x ∈ R | x ≤ a },
◦ (−∞, ∞) = R.

Task 1.6.5 Express the sets R>0 and R≥0 in interval notation.

In this context, ∞ and −∞ are just symbols that help to identify particular
« unbounded intervals. They do not stand for any actual objects called
innity or minus innity.

Exercises for Ÿ1.6

1 Express the following sets using the set-builder notation:

(a) {1, 8, 27, . . . , 1000},


(b) The set of odd natural numbers,

(c) The set of prime numbers.

2 The following sets are described in set-builder notation. Express them in


roster notation:

(a) { x ∈ N | x | 100 },
(b) { x ∈ R | x2 + x + 1 = 0 },
(c) { x ∈ R | x2 − 1 > 0 } .

3 Express the following statements symbolically using quantiers and set


notation, and also state whether they are true or false:

(a) The square of every non-zero integer is positive.

(b) Let U be a set and A, B , C be any subsets of U . If A and B are


disjoint, and B and C are disjoint, then A and C are disjoint.
(c) Every positive rational number has a rational square root.
46 CHAPTER 1. SETS AND LOGIC

4 What is the negation of (∀x ∈ A) [P (x) =⇒ Q(x)]?

5 Express the following in symbols, then write out the negation in symbols,
and nally write the negation as a sentence in English:

(a) For every real number x there is a natural number n that is greater
than x.
(b) If m, a and b are natural numbers such that m divides ab then m
divides either a or b.
(c) If two lines meet at two distinct points then they are the same line.

6 Express the following in symbols, then write out the converse in symbols,
and nally write the converse as a sentence in English:

(a) The square of every non-zero integer is positive.

(b) The product of an even natural number with any natural number is
even.

(c) Larger natural numbers have larger squares.

7 Is it true that either (∀x ∈ A) [P (x) =⇒ Q(x)] or its converse must be


true?

8 Write out the contrapositive in symbols and as an English sentence:

(a) The square of every non-zero integer is positive.

(b) The product of an even natural number with any natural number is
even.

(c) Larger natural numbers have larger squares.

9 Let X be the set of intervals (0, x) with x ∈ (0, ∞). Justify the following
statements:

S T
(a) X = (0, ∞), (b) X = ∅.

10 Let Y be the set of intervals (−x, x) with x ∈ (0, ∞). Justify the following
statements:

S T
(a) Y = R, (b) Y = {0}.

11 Express the following sets as unions of intervals:

Undergraduate Mathematics: Foundations


1.6. SETS, QUANTIFIERS AND NUMBERS 47

(a) { x ∈ R | x(x − 1) > 0 },


(b) { x ∈ R | x(x − 1)(x − 2) > 0 }.

12 Let a, b, c, d be real numbers. Prove that:

(a) [a, b] ∩ [c, d] = [max{a, c}, min{b, d}],


(b) (a, b) ∩ (c, d) = (max{a, c}, min{b, d}).

13 Let p and q be rational numbers with p < q. Dene the interval (p, q)Q
by
(p, q)Q = { x ∈ Q | p < x < q }.
(a) Let F be the set whose members are the intervals
T (0, 1/n)Q with
n ∈ N. Show that F = ∅.
(b) Let G be the set whose
S members are the intervals (0, 1 − 1/n)Q with
n ∈ N. Show that G = (0, 1)Q .
2 | Relations and Functions
Good mathematicians see analogies between theorems or theories;
the very best ones see analogies between analogies.
1
 Stefan Banach

For example, even at age four, many children can share sweets
fairly between two recipients by using correspondences: they share
giving one-for-you, one-for-me, until there are no sweets left.
They do sometimes make mistakes but they know that, when the
sharing is done fairly, the two people will have the same amount
of sweets at the end.
2
 Terezinha Nunes and Peter Bryant

We began our rst chapter by looking at some typical sentences we en-


counter when we study mathematics. Our rst observation was that these
sentences made assertions and could be true or false. We decided to call them
statements and made a detailed study of how they could be combined to
make new statements or be arranged into proofs. Our second observation
was that the typical mathematical statement is an assertion about the prop-
erties of the members of some collection. We may assert, for instance, that
they all have a certain property, or that at least one of them has it, or that
exactly one of them has it. Russell's Paradox alerted us to the dangers of us-
ing the words set and collection too freely and so we undertook a careful
axiomatic approach to sets and what we can do with them.

1 Quoted in Analogies Between Analogies: The Mathematical Reports of S.M. Ulam


and his Los Alamos Collaborators. Berkeley: University of California Press, 1990. (http:
//ark.cdlib.org/ark:/13030/ft9g50091s/)
2 T. Nunes and P. Bryant, Key understandings in mathematics learn-
ing, Nueld Foundation, 2009. (http://www.nuffieldfoundation.org/
key-understandings-mathematics-learning)

49
50 CHAPTER 2. RELATIONS AND FUNCTIONS

For our next step forward, we again look at some mathematical state-
ments:

Fundamental Theorem of Arithmetic Every natural number except 1


is a product of some prime numbers.

Fundamental Theorem of Algebra Every non-constant complex poly-


nomial has a complex number as a root.

Fundamental Theorem of Calculus Every continuous function dened


on a closed interval [a, b] has an anti-derivative.

You may not understand all the terms used in these theorems. Do not worry
about that. We are interested in the structure of these theorems, and we
can appreciate the main features even without a detailed knowledge of the
terms. And the essentials will be reviewed below.

First of all, we note that these are existence theorems. The Fundamental
Theorem of Algebra asserts the existence of a complex number that is a root
of the given complex polynomial. The Fundamental Theorem of Calculus
asserts the existence of a function whose derivative is the given continuous
function. The Fundamental Theorem of Arithmetic asserts the existence of
some primes that can be multiplied with each other to get the given natural
number.

We know from earlier experience that mathematical statements tend to


be about sets and their members. Looking at each theorem in turn, we nd
that each involves two sets A and B as follows:

A B
FT of Arithmetic Natural Numbers Prime Numbers
FT of Algebra Complex Polynomials Complex Numbers
FT of Calculus Continuous functions Dierentiable functions

There is a common pattern involving these two sets. Each theorem considers
the elements of A and asserts the existence of corresponding elements of B
which are in a certain relationship to the elements of A.
In this chapter we will learn how to work systematically in such frame-
works. First, we need a language for talking about pairs of objects, where
each object is chosen from a particular set. This language, of cartesian prod-
ucts and ordered pairs, is developed in the rst section of the chapter. In the
remaining sections we carry out a description of relations and functions, and
put them to work. When you have completed this chapter you will possess

Undergraduate Mathematics: Foundations


2.1. CARTESIAN PRODUCTS 51

a working knowledge of the basic vocabulary of modern mathematics, and


will be equipped to take up the study of any aspect of it.

2.1 Cartesian Products


You have already encountered the notion of an ordered pair while studying
coordinate geometry. An ordered pair of two real numbers x and y is written
(x, y). It diers from the set {x, y} in that not only the members matter
but also the order in which they are written. We have {1, 2} = {2, 1} but
(1, 2) 6= (2, 1). The Cartesian plane consists of all possible ordered pairs of
real numbers.

We could hope to do this for any two sets A and B . Their cartesian
product would consist of all ordered pairs (a, b) with a ∈ A and b ∈ B .
Such pairs would be useful as a means of discussing connections between the
members of A and B . The rst task is to show that the collection of ordered
pairs constitutes a set in the sense of Zermelo's axioms.

We begin by dening an individual ordered pair as a set:

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


This denition involves repeated use of the Axiom of Pairs (Z3) to create
rst {a}, then {a, b}, and nally {{a}, {a, b}}.

Theorem 2.1.1 Two ordered pairs are equal if and only if the corresponding
elements are equal:

(a, b) = (c, d) ⇐⇒ a = c and b = d.

Proof: Suppose (a, b) = (c, d). Then {{a}, {a, b}} = {{c}, {c, d}}. Equality
of these two sets means their elements must match.

One is tempted to argue that since the elements match, we have {a} = {c}
7 and {a, b} = {c, d}, due to the size of the sets. But then we are making
the error of assuming, for instance, that a 6= b.

Since the elements must match we have {a} = {c} or {a} = {c, d}. In both
cases, we conclude that a = c, and replacing c by a gives us
{{a}, {a, b}} = {{a}, {a, d}}.
• Ifb 6= a we have {a, b} =
6 {a} and hence {a, b} = {a, d}. It follows that
b = d.
52 CHAPTER 2. RELATIONS AND FUNCTIONS

• If b = a we have {a, b} = {a} and hence {{a}} = {{a}, {a, d}}. It


follows that {a} = {a, d} and so b = a = d. 
Let X = A ∪ B , a ∈ A and b ∈ B . Then {a} and {a, b} are members of
the power set P(X) of X . Hence
(a, b) = {{a}, {a, b}} ∈ P(P(X)).

Thus all the ordered pairs are present in the set P(P(X)). This opens
the prospect of using the Axiom of Separation to show that the collection of
ordered pairs constitutes a set:

A × B = { S ∈ P(P(X)) | (∃a ∈ A)(∃b ∈ B)[S = {{a}, {a, b}}] }.


Hence the cartesian product A×B is a set and consists exactly of the ordered
pairs.

Suppose A⊆X and B ⊆ Y. Then A×B can be viewed as a subset of


X ×Y as follows:

A × B = { (x, y) ∈ X × Y | x ∈ A ∧ y ∈ B }.

For some practice in working with cartesian products, we take up the


already familiar example of the cartesian plane R2 = R × R. Let us start by
2
recalling that we visualize R as a plane with two perpendicular number lines
superimposed on it. These number lines are called the coordinate axes and
meet at their respective origins. This meeting point is then called the origin
O of the cartesian plane. Typically, we draw one coordinate axis horizontally
and call it the x-axis. The other coordinate axis is vertical and is called the
y -axis. To each point in the plane we associate a unique ordered pair as
shown below:

y -axis

b (a, b)

x-axis
O a

Example 2.1.2 Let us sketch the following subsets of R2 on the cartesian


plane:

Undergraduate Mathematics: Foundations


2.1. CARTESIAN PRODUCTS 53

(a) Z × {1}, (b) Z × R.

First, we work out a detailed description of Z × {1}:

Z × {1} = { (x, y) ∈ R2 | x ∈ Z ∧ y ∈ {1} } = { (x, y) ∈ R2 | x ∈ Z ∧ y = 1 }.

Hence Z × {1} consists of ordered pairs (x, 1) with x ∈ Z.

y -axis

x-axis
−1 O 1 2 3

For the other set, we have

Z × R = { (x, y) ∈ R2 | x ∈ Z ∧ y ∈ R } = { (x, y) ∈ R2 | x ∈ Z }.

For a xed x∈Z our set includes all the points (x, y) with y varying over
R, and these points constitute a vertical line.

y -axis

x-axis
−1 0 1 2 3

Task 2.1.3 Sketch the following subsets of R×R on the xy -plane:


54 CHAPTER 2. RELATIONS AND FUNCTIONS

(a) Z × {0, 1}, (b) Z × Z,

(c) (Z × R) ∪ (R × Z), (d) [0, 1] × [1, 2].

We have been working with subsets of R2 that have the form A×B . Does
2
every subset of R have this form?

Example 2.1.4 2 2 2
Consider the subset S = { (x, y) ∈ R | x + y = 1 } of
R2 . Can it be of the form A × B ? If yes, then we would have S =
{ (x, y) ∈ R2 | x ∈ A ∧ y ∈ B }. This enables the following implications:

02 + 12 = 1 =⇒ (0, 1) ∈ S =⇒ 0 ∈ A,
12 + 02 = 1 =⇒ (1, 0) ∈ S =⇒ 0 ∈ B,
0 ∈ A ∧ 0 ∈ B =⇒ (0, 0) ∈ A × B = S.

But (0, 0) ∈ S is false! Hence S is not of the form A × B. 

Task 2.1.5 For each of the following subsets of R2 , check if it is of the form
A × B:
(a) { (x, y) ∈ R2 | 0 ≤ x ≤ 1, 0 ≤ y ≤ 1 },
(b) { (x, y) ∈ R2 | y ≥ 0 },
(c) { (x, y) ∈ R2 | x2 + y 2 ≤ 1 },
(d) { (x, y) ∈ R2 | |x| + |y| ≤ 1 }.

Now we turn to a general cartesian product A × B. We can visualize this


in a way similar to the cartesian plane R × R. We lay out the points of A
along a horizontal line, the points of B along a vertical line, and the points
of A × B in a corresponding grid pattern. For example, let A = {a, b, c, d}
and B = {1, 2, 3}:

B

× × × × A
a b c d

Undergraduate Mathematics: Foundations


2.1. CARTESIAN PRODUCTS 55

In such a diagram, we can indicate a subset of A × B by lling in the corre-


sponding circles. Suppose S = {(a, 1), (b, 2), (c, 3), (d, 2)}. It is represented
by:

B

× × × × A
a b c d

Let us consider the rules relating cartesian product with the set operations
of union, intersection and complement.

Theorem 2.1.6 For any sets A, B , C , we have:

1. A × (B ∪ C) = (A × B) ∪ (A × C),

2. A × (B ∩ C) = (A × B) ∩ (A × C).

Proof: We shall prove the rst of these `distributive laws' and leave the
second to the reader:

(x, y) ∈ A × (B ∪ C) ⇐⇒ x ∈ A ∧ y ∈ B ∪ C
⇐⇒ x ∈ A ∧ (y ∈ B ∨ y ∈ C)
⇐⇒ (x ∈ A ∧ y ∈ B) ∨ (x ∈ A ∧ y ∈ C)
   
⇐⇒ (x, y) ∈ A × B ∨ (x, y) ∈ A × C
 
⇐⇒ (x, y) ∈ (A × B) ∪ (A × C) .

Exercises for Ÿ2.1

1 Show that the following can also be used as a denition of ordered pair:

(x, y) = {{x, ∅}, {{∅}, y}}.


56 CHAPTER 2. RELATIONS AND FUNCTIONS

2 Consider the denition P = {{x}, {x, y}} of the ordered pair (x, y). Show
that the coordinates x and y are the unique objects satisfying the following
properties:

(a) For x: x ∈ ∩P ,
(b) For y : (y ∈ ∪P ) ∧ [(∃S1 ∈ P )(∃S2 ∈ P )(S1 6= S2 ) =⇒ (y ∈
/ ∩P )].

3 Prove the following for any sets A, B, C :


(a) A × ∅ = ∅ × A = ∅.
(b) A × B = B × A ⇐⇒ A = B or A=∅ or B = ∅.
(c) Suppose A 6= ∅. Then A × B = A × C ⇐⇒ B = C .

4 Let A = {−2, −1, 0, 1, 2}. Express the given subsets of A×A using the
set-builder notation:

2 × 2 × 2 ×
1 × 1 × 1 ×
0 × 0 × 0 ×
−1 × −1 × −1 ×
−2 × −2 × −2 ×
× × × × × × × × × × × × × × ×
−2 −1 0 1 2 −2 −1 0 1 2 −2 −1 0 1 2
(a) (b) (c)

5 For each of the following subsets of R2 , check if it is of the form A × B:


2
(a) X = { (x, y) ∈ R | xy = 0 },
(b) Y = { (x, y) ∈ R2 | |x + y| + |x − y| ≤ 2 }.

6 Are the following identities valid?

(a) (A × B) ∩ (C × D) = (A ∩ C) × (B ∩ D),
(b) (A × B) ∪ (C × D) = (A ∪ C) × (B ∪ D).

7 Let X and Y be sets. Let A ⊆ X, so that A×Y ⊆ X ×Y. Is it true that


(A × Y )c = Ac × Y ?

8 Recall that symmetric dierence is dened by A4B = (A \ B) ∪ (B \ A).


Does the cartesian product distribute over symmetric dierence?

Undergraduate Mathematics: Foundations


2.2. RELATIONS 57

9 Let A ⊆ X, B ⊆ Y , so that A × B ⊆ X × Y . Are any of the following


always equal to (A × B)c ?

(a) Ac × B c , (b) (Ac × B) ∪ (A × B c ),

(c) (Ac × Y ) ∪ (X × B c ), (d) (Ac ×B)∪(A×B c )∪(Ac ×B c ).

10 (a) Extend the denition (a, b) = {{a}, {a, b}} of an ordered pair to an
ordered triple (a, b, c). Your denition can be counted as successful if
you succeed in proving that (a, b, c) = (a0 , b0 , c0 ) if and only if a = a0
0 0
and b=b and c=c.
(b) Use part (a) to dene the Cartesian product A×B ×C of three sets.

2.2 Relations
Ordered pairs and cartesian products give us a way to talk about pairs of
elements where each member is from a particular set. Now consider the
Fundamental Theorem of Algebra (page 50). Let C be the set of complex
numbers and let C[z] be the set of complex polynomials in the variable z.
The Fundamental Theorem of Algebra can be seen as a statement about
certain ordered pairs in C[z] × C. It says that if a polynomial p(z) ∈ C[z]
is non-constant then there is an ordered pair (p(z), w) ∈ C[z] × C such that
p(w) = 0.
The pairs (p(z), w) such that p(w) = 0, give a complete description of
the relationship of a complex number being a root of a complex polynomial.

A relation from a set X to a set Y is a subset R of X × Y . We denote


(x, y) ∈ R by xRy and say that x is related to y . We denote (x, y) ∈/ R
by xRy/ . If X = Y we say that R is a relation on X . Consider a relation
R ⊆ X × Y . We have two associated sets:
• The domain of R is dom(R) = { x ∈ X | (∃y ∈ Y ) xRy }.
• The range of R is ran(R) = { y ∈ Y | (∃x ∈ X) xRy }.

Example 2.2.1 Consider the relation S shown by the following diagram:


58 CHAPTER 2. RELATIONS AND FUNCTIONS

B

× × × × A
a b c d

We see that:

aS1 =⇒ a ∈ dom(S) ∧ 1 ∈ ran(S),


bS2 =⇒ b ∈ dom(S) ∧ 2 ∈ ran(S),
dS2 =⇒ d ∈ dom(S) ∧ 2 ∈ ran(S).

Hence dom(S) = {a, b, d} and ran(S) = {1, 2}. 

Task 2.2.2 Consider the relation R from C[z] to C dened by p(z)Rw if


p(w) = 0. What is the domain of R?

Equivalence Relations

In mathematics we often ask whether two objects are of the same type in
some sense. For example, in plane geometry, we may consider two triangles
to be of the same type if they are congruent. Or, we may consider them to
be of the same type if they are similar. Among integers, we may consider
two numbers to be of the same type if they have the same remainder when
divided by 2. This splits the integers into the even and odd ones.

We wish to formalize this common practice. First we identify some ob-


jects who are to be classied into types. This is done by xing a set from
which they will be taken. Next we need a rule that can be applied to pairs of
objects and decides whether they are of the same type or not. The operation
of this rule should follow certain patterns: (a ) Every object should be of the
same type as itself; (b ) If A is of the same type as B then B is of the same
type as A; and (c ) If A is of the same type as B and B is of the same type
as C then A is of the same type as C.
It should be evident by now that we are leading up to dening a relation
with certain properties. Here is the formal denition:

Undergraduate Mathematics: Foundations


2.2. RELATIONS 59

A relation R on a set X is called an equivalence relation if it has the


following three properties:

Reexivity: ∀x ∈ X , xRx.

Symmetry: xRy =⇒ yRx.

Transitivity: xRy ∧ yRz =⇒ xRz .


When R is an equivalence relation, we read xRy as  x is equivalent to y . It
is typical to use the name ∼ for an equivalence relation and to write x∼y
for  x is equivalent to y . As usual, we use x 6∼ y for  x is not equivalent to
y .
Let ∼ denote an equivalence relation on X and let x ∈ X. Then we
dene the equivalence class of x to be the set
[x] = { y ∈ X | x ∼ y }.

Any member of an equivalence class is called a representative of that class.

Example 2.2.3 Let us dene a relation on R2 by (x1 , y1 ) ∼ (x2 , y2 ) if x1 =


x2 . Let us rst check that this is an equivalence relation:

Reexivity: (x1 , y1 ) ∼ (x1 , y1 ) since x1 = x1 .

Symmetry: (x1 , y1 ) ∼ (x2 , y2 ) =⇒ x1 = x2 =⇒ x2 = x1 =⇒ (x2 , y2 ) ∼


(x1 , y1 ).

Transitivity:

(x1 , y1 ) ∼ (x2 , y2 ) and (x2 , y2 ) ∼ (x3 , y3 ) =⇒ x1 = x2 and x2 = x3


=⇒ x1 = x3
=⇒ (x1 , y1 ) ∼ (x3 , y3 )

The equivalence class of (x1 , y1 ) is

[(x1 , y1 )] = { (x, y) ∈ R2 | x = x1 }.

This is just the vertical line through the point (x1 , y1 ).


60 CHAPTER 2. RELATIONS AND FUNCTIONS

(x1 , y1 )

x = x1

Note that two distinct equivalence classes are parallel lines and have no
common points. 

Task 2.2.4 2 2 2
Let (xj , yj ) ∈ R , j = 1, 2. We say (x1 , y1 ) ∼ (x2 , y2 ) if x1 +y1 =
2 2 2
x2 + y2 . Show that this denes an equivalence relation on R . Identify the
equivalence classes.

Theorem 2.2.5 Let ∼ be an equivalence relation on X , and x, y ∈ X . Then

1. [x] = [y] ⇐⇒ x ∼ y .
2. [x] 6= [y] =⇒ [x] ∩ [y] = ∅.

Proof:

1. First, suppose [x] = [y]. Then y ∈ [y] =⇒ y ∈ [x] =⇒ x ∼ y . For the


converse, assume x ∼ y . If z ∈ [x] then x ∼ z . By transitivity, x ∼ y
and x ∼ z implies y ∼ z and z ∈ [y]. Therefore [x] ⊆ [y]. Similarly,
[y] ⊆ [x] and hence [x] = [y].
2. We shall prove the contrapositive. Suppose [x] ∩ [y] 6= ∅. Then there
is a z ∈ [x] ∩ [y]. Now z ∈ [x] =⇒ x ∼ z and z ∈ [y] =⇒ y ∼ z . By
transitivity, we get x ∼ y and hence [x] = [y]. 
Consider a set X and a collection C of some non-empty subsets of X. We
call C a partition of X if

1. For each x∈X there is a set A∈C such that x ∈ A. ( C covers X )


2. If A, B ∈ C and A 6= B then A ∩ B = ∅. (Members of C are pairwise
disjoint.)

For example, if ∼ is an equivalence relation on X , then its equivalence classes


are a partition of X.

Undergraduate Mathematics: Foundations


2.2. RELATIONS 61

Task 2.2.6 Let C be a partition of X. Show that there is a unique equiv-


alence relation on X whose equivalence classes are exactly the members of
C.

Example 2.2.7 For m, n ∈ Z let us say that m∼n i m−n is a multiple


of 3. You should check that this denes an equivalence relation on Z. Now
let us consider the equivalence classes. We use n for [n]:

0 = {. . . , −6, −3, 0, 3, 6, . . . },
1 = {. . . , −3, 1, 1, 4, 7, . . . },
2 = {. . . , −4, −1, 2, 5, 8, . . . },
3 = {. . . , −3, 0, 3, 6, 9, . . . } = 0.

We see that the equivalence classes 0, 1, and 2, are distinct but 3 = 0. In fact
we can see that 0 consists of integers of the form 3k , 1 consists of integers of
the form 3k + 1, and 2 consists of integers of the form 3k + 2. Since every
integer is one of these forms, these three are all the equivalence classes. 

Ordered Sets

Often we want to rank objects in some way. If two integers are not equal,
we wish to know which is greater. If two sets are dierent, we wonder if
one is contained in the other. We may not always have a clear answer. For
example, given two sets it may happen that neither is a subset of the other.

A relation ≤ on a set X is called a partial order if it has the following


properties:

Reexivity: ∀x ∈ X , x ≤ x.
Anti-Symmetry: x ≤ y ∧ y ≤ x =⇒ x = y .
Transitivity: x ≤ y ∧ y ≤ z =⇒ x ≤ z.
We read x≤y as  x is less than or equal to y  or  y is greater than or equal
to x. We also use y≥x as equivalent to x ≤ y.

Example 2.2.8 The set Z is partially ordered by m≤n i n − m ∈ W. 

Example 2.2.9 Any power set P(X) is partially ordered by ⊆. We call


this ordering by inclusion. As an example, consider X = {1, 2, 3}. Then
the partial ordering of P(X) by ⊆ is represented by the following diagram,
where a sequence of arrows points from A to B if A ⊆ B:
62 CHAPTER 2. RELATIONS AND FUNCTIONS

{1, 2, 3}

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

{1} {2} {3}

This ordering really is partial, in the sense that there are pairs of elements
A, B that are unrelated: Neither A⊆B nor B ⊆ A. 

Example 2.2.10 The set N is partially ordered by m ≤ n i m | n. In


this ordering, we have sequences like 1 ≤ 2 ≤ 6 ≤ 30. This ordering is also
genuinely partial. For example, 2 and 3 are unrelated as neither divides the
other. Since N is innite we can't draw the full diagram for this ordering,
but here is a small part of it:

8 12

4 6 9 10

2 3 5 7 11

1 

If ≤ is a partial order on X we can dene the associated strict order < on


X by x<y if x≤y and x 6= y . If x < y we say that  y is strictly greater
than x.

Task 2.2.11 Let ≤ be a partial order on X . Show that the associated strict
order < has the following properties:

(a) (Irreexivity) ∀x ∈ X , x ≮ x,
(b) (Transitivity) x < y ∧ y < z =⇒ x < z .

Undergraduate Mathematics: Foundations


2.2. RELATIONS 63

Task 2.2.12 Let < be a strict order on X, i.e., < is a relation on X that
is irreexive and transitive. Dene a relation ≤ on X by x≤y if x < y or
x = y. Show that ≤ is a partial order.

Once we start ranking objects, we become interested in the extreme cases.


This leads to the following denitions. Let ≤ be a partial order on X, and
let S ⊆ X. Then:

• We call m∈S the greatest element of S if s ≤ m for every s ∈ S .


The greatest element is also called the maximum of S and is denoted
by max(S).
• We call m0 ∈ S the least element of S if m0 ≤ s for every s ∈ S .
The least element is also called the minimum of S and is denoted by
min(S).

Task 2.2.13 Show that the greatest and least elements are unique, if they
exist.

Example 2.2.14 Let us order the power set P(X) of X = {1, 2, 3} by


inclusion. Consider the following subsets of P(X):
n o
S = {2}, {1, 2}, {2, 3}, {1, 2, 3} ,
n o
S 0 = {2}, {1, 2}, {1, 3}, {2, 3}, {1, 2, 3} ,
n o
S 0 = {2}, {1, 2}, {1, 3}, {2, 3} .

These subsets are illustrated below:

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

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

{2} {2} {2}

We observe the following:

• S has greatest element {1, 2, 3} and least element {2}.


64 CHAPTER 2. RELATIONS AND FUNCTIONS

• S0 has greatest element {1, 2, 3} and no least element.

00
• S has neither a greatest element nor a least element.

Can you see why {2} is not the least element of S0 and S 00 ? 

This example illustrates that the greatest element may not exist. Some-
times all we need is a weaker condition  the existence of an element that
has nothing above it, even though it may not itself be greater than all other
elements. This leads to the following denition:

• We call m∈X amaximal element of S if m ≮ s for every s ∈ S .


• We call m0 ∈ X a minimal element of S if s ≮ m for every s ∈ S .

Task 2.2.15 Show that if a set has a greatest element then that element is
also maximal. And a least element is also minimal.

Example 2.2.16 Consider the subsets in Example 2.2.14.

• S has greatest element {1, 2, 3} so this element is also maximal. And


{2} is least, hence minimal.

• S0 has greatest element {1, 2, 3} so this element is also maximal. The


minimal elements are {2} and {1, 3}.
• The maximal elements of S 00 are {1, 2}, {1, 3} and {2, 3}. The minimal
elements are {2} and {1, 3}.
Thus there can be multiple maximal and minimal elements. 

Always be careful to distinguish between maximum and maximal, mini-


«
mum and minimal!

Even if a set has no greatest or maximal element, it may still be conned


in the sense that there may be an external element that sits above all its
members. This leads to the following denition:

• We call u∈X upper bound of S if s ≤ u for every s ∈ S . If S


an
has an upper bound, we say S is bounded above.
• We call ` ∈ X a lower bound of S if ` ≤ s for every s ∈ S . If S has
a lower bound, we say S is bounded below.

• If S is both bounded above and bounded below, we say it is bounded.


Otherwise we call it unbounded.

Undergraduate Mathematics: Foundations


2.2. RELATIONS 65

Note that the formal denition allows the upper and lower bounds to be
outside of S.

Task 2.2.17 Show that if an upper bound of S belongs to S then it is the


greatest element of S. Make and prove the corresponding statement for lower
bounds.

A mathematician studies a certain subset of N and concludes that 49 is an


upper bound. Another mathematician studies the same set and shows that 43
is an upper bound. Which one has done better? The second mathematician's
work conveys more information. For example, it implies that 44 is not a
member of the subset, a possibility that the rst mathematician was unable
to rule out.

In general, given several upper bounds for a set A, the lower ones give
more information, and so we wonder whether there is a least upper bound.
This would be the minimum element of the collection of all upper bounds of
A, if it exists. The least upper bound of A is also called the supremum of
A and is denoted by sup(A).

Example 2.2.18 Consider a partially ordered set X, whose ordering is de-


picted below.

1 2

a b

Let A = {a, b}. The upper bounds of A form the set {1, 2}. This set has no
minimum so sup(A) does not exist. 

Example 2.2.19 Consider a partially ordered set X, whose ordering is de-


picted below.

α β γ

a b c
66 CHAPTER 2. RELATIONS AND FUNCTIONS

Let A = {a, b, c}. The upper bounds of A form the set {0, α, β, γ}. This set
has 0 as its minimum, so sup(A) = 0. 

Similarly, when it comes to the lower bounds of a set A, we are inter-


ested in the highest one. The greatest lower bound is dened to be the
maximum element of the collection of all lower bounds of A. The greatest
lower bound of A is also called the inmum of A and is denoted by inf(A).
A relation ≤ on a set X is called a total order if it is a partial order
and has the following property:

Totality: x, y ∈ X =⇒ x ≤ y or y ≤ x.
Of the orders we have considered so far the only total order is the one dened
on Z by n≤m if m − n ∈ W. We shall call this the standard order on Z.

Example 2.2.20 (Dictionary Order) Consider the relation on Z2 dened


0 0
by (x, y) ≤ (x , y ) if one of the following happens (where we use the standard
order on Z):
1. x = x0 and y = y 0 ,
2. x < x0 ,
3. x = x0 and y < y 0 .
Let us check that this is a partial order:

Reexive: If (x, y) = (x0 , y 0 ), we get x = x0 and y = y0 , hence (x, y) ≤


(x0 , y 0 ) by the rst condition.

Anti-Symmetric: Suppose (x, y) ≤ (x0 , y 0 ) and (x0 , y 0 ) ≤ (x, y). Now,

(x, y) ≤ (x0 , y 0 ) [(x, y) = (x0 , y 0 ) ∨ x < x0 ∨ (x = x0 ∧ y < y 0 )]


∧ ⇐⇒ ∧
(x0 , y 0 ) ≤ (x, y) [(x, y) = (x0 , y 0 ) ∨ x0 < x ∨ (x = x0 ∧ y 0 < y)]

⇐⇒ (x, y) = (x0 , y 0 ).
The last equivalence is obtained by applying the distributive law and
dropping every term that is a contradiction.

Transitive: The proof is similar to that of anti-symmetry and is left to you.

Translation: The author worked it out in his head, but realized typing it
out would be drudgery. He asks you to spare him and work it out yourself.
7 To give him his due, he honestly believes this will be good exercise and
will strengthen your mathematical muscles.

Undergraduate Mathematics: Foundations


2.2. RELATIONS 67

The dictionary order is also a total order. If (x, y) 6= (a, b) the two ordered
pairs dier in at least one coordinate and that will ensure that one is greater
than the other. In the diagram drawn below, the white circles mark the
ordered pairs that are greater than a certain (a, b), and the grey circles mark
the ones that are lesser.

y=b

x=a 

Theorem 2.2.21 Let ≤ be a total order on X, and let < be the associated
strict order. Then < has the following properties:

Irreexivity: ∀x ∈ X , x ≮ x.
Trichotomy: ∀x, y ∈ X , exactly one of the following holds: x < y , y < x,
x = y.
Transitivity: x < y ∧ y < z =⇒ x < z .

Proof: The property of being irreexive is part of the denition of <. (Put
y=x in the denition.)

For trichotomy, we rst check that no two of x < y , y < x, x = y , can


hold simultaneously. x = y cannot hold
From irreexivity it follows that
together with either x < y or y < x. Further, if both x < y and y < x hold,
then both x ≤ y and y ≤ x hold, and hence x = y , in contradiction to what
we just established.

Transitivity of < follows from the transitivity of ≤ combined with tri-


chotomy. ’


Exercises for Ÿ2.2

1 Let R be a relation on X. What can you say about its domain and range
if
68 CHAPTER 2. RELATIONS AND FUNCTIONS

(a) R is reexive, (b) R is symmetric.

2 Dene a relation on N×N by setting (m, n) ∼ (r, s) if m + s = n + r.


Show that this is an equivalence relation.

3 Dene a relation on Z×Z by (m, n) ∼ (a, b) if mb = na. Show that this


does not dene an equivalence relation.

4 Dene a relation on (Z × Z)∗ = (Z × Z) \ {(0, 0)} by (m, n) ∼ (a, b) if


mb = na. Show that this denes an equivalence relation.

5 Fix k ∈ N. For m, n ∈ Z let us say that m ∼ n i m − n is a multiple of k .


Show that this denes an equivalence relation on Z, and list all the distinct
equivalence classes.

6 Dene a relation on R2 by (x1 , y1 ) ∼ (x2 , y2 ) if there exists α ∈ R+ such


that x1 = αx2 and y1 = αy2 . Show that this denes an equivalence relation
on R2 . Identify the equivalence classes.

7 Let R be a relation on X. Suppose the sets x̄ = { y ∈ X | xRy } satisfy


the following:

(a) ∀x ∈ X , x ∈ x̄, and

(b) x̄ 6= ȳ =⇒ x̄ ∩ ȳ = ∅.
Prove that R is an equivalence relation.

8 In the previous exercise, weaken assumption (a) to ∀x ∈ X , x̄ 6= ∅. Is it


still true that R must be an equivalence relation?

9 Would it be correct to argue as follows that symmetry and transitivity


together imply reexivity?

aRb =⇒ bRa =⇒ (aRb ∧ bRa) =⇒ aRa.

10 Let X be a partially ordered set. Show that if X has a greatest element


then that element is the unique maximal element. Is the converse true?

11 Consider N ordered by divisibility. Show that if A and B are subsets of


N which are bounded above then A∪B is bounded above.

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 69

12 Let A and B be subsets of a partially ordered set X. Are the following


necessarily true?

(a) If A and B are bounded above then A∪B will be bounded above.

(b) If A and B have greatest elements then A∪B will have a greatest
element.

(c) If A and B have maximal elements then A∪B will have a maximal
element.

13 Will your answers to the previous exercise change if X is totally ordered?

14 Let X be a non-empty partially ordered set.

(a) Will ∅ be bounded above in X?


(b) If S ⊆X has upper bound u and lower bound ` can we claim that
` ≤ u?

2.3 Functions
Consider a device that takes in some kind of input and converts it into an
output. What kind of abstract framework would be useful for describing such
devices? First, we need to know what are the allowed inputs into the device.
In the language of mathematics, we need to know the set from which the
inputs can be taken. Similarly, we need to know the set of possible outputs.
And we also need to know what happens for each possible input  what is
the corresponding output?

Input Output
BP

BP

BP

X Y

(Just for fun: What do you think this `device' is doing to the people passing
through it?)

Let X be the set of allowed inputs and Y be the set of possible outputs.
Clearly, we are again looking for a relation from X to Y. What special
70 CHAPTER 2. RELATIONS AND FUNCTIONS

properties should the relation have? We must know the set of allowed inputs
exactly. We don't want to feed in something that the device cannot handle!
This means the domain of the relation must be all of X. On the other hand,
no harm may be done if we have extra elements in what we think is the
set of outputs. Those elements will just never show up as outputs, and so
what? The implication is that the range of the relation need not be all of
Y. Finally, we usually expect a device to produce one output per input. So
each element of X should be related to only one element of Y.
You can imagine devices which don't t this narrow framework. Maybe
they have a random element. Maybe their allowed inputs change with time.
Still, this simple type of device of ours is common enough and it has been
very fruitful to set up a corresponding formal denition, as done below:

A relation F from a set X to a set Y is called a function if


• For each x ∈ X, there is a y∈Y such that xF y . (That is, dom(F ) =
X)
• If xF y and xF y 0 then y = y0 .
The rules for F ensure that for each x∈X there is exactly one y∈Y such
that xF y . We give this unique y the name F (x) and call it the image of x
. Conversely, we call x a pre-image of y.
Thus, the statement xF y can also be stated as y = F (x) and this is our
preferred notation for functions. We think of F as a rule for associating a
unique member F (x) of Y to each member x of X.
The set X is called the domain of F , while Y is called the codomain.
We write F : X → Y to indicate a function together with its domain and
codomain.

Example 2.3.1 Consider the relations depicted by the following pictures.


In each pair of ovals, the one on the left represents the domain, while the one
on the right represents the codomain. The arrows describe the associations
given by the relation. Which diagrams represent functions?

(A) (B)

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 71

(C) (D)

(A) does not represent a function because there is a point in the domain that
has no image. (B) also does not represent a function, since there is a point in
the domain that has two images. (C) and (D) do represent functions, since
it is permitted for points in the codomain of a function to have no pre-image
as well as to have multiple pre-images. 

Example 2.3.2 Consider the following relations.

(a) F = { (m, n) ∈ Z × N | m = n },
(b) G = { (m, n) ∈ N × Z | m = n }.
The relation F has dom(F ) = N 6= Z, since (m, n) ∈ F ⇐⇒ m = n =⇒
m∈N and m ∈ N =⇒ (m, m) ∈ F . Hence F is not a function.
For G, we see that the same reasoning gives dom(G) = N. Also (m, n) ∈
G =⇒ m = n and hence each point of the domain is associated to only one
point of the codomain. Hence G is a function. 

Task 2.3.3 Which of the following relations are functions?

(a) { (m, n) ∈ Z × Z | m2 = n2 }, (b) { (m, n) ∈ N × Z | m2 = n2 },

(c) { (m, n) ∈ Z × N | m2 = n2 }, (d) { (m, n) ∈ N × N | m2 = n2 }.

In practice, functions are usually described as explicit rules for associating


members of the codomain to members of the domain. For example one might
say that f : Z → N is given by f (x) = x2 + 1. This means that to any integer
x we associate the natural number x2 + 1. For example f (0) = 02 + 1 = 1,
f (1) = 12 + 1 = 2, f (−1) = (−1)2 + 1 = 2, and so on. This is a convenient
2
manner of expressing the fact that f is the set { (x, y) ∈ Z × N | y = x + 1 }.

Example 2.3.4 Let X be any set and dene 1X : X → X by 1X (x) = x.


Then 1X is called the identity function of X . 
72 CHAPTER 2. RELATIONS AND FUNCTIONS

Example 2.3.5 Let X and Y be any two sets. We x a member C ∈Y


and dene f: X →Y by f (x) = C . Then f is called a constant function.
We usually denote a constant function by its xed value, and talk of the
constant function C . 

Example 2.3.6 An empty function has domain ∅ and some codomain


Y. Since ∅ × Y = ∅, as a relation this function is just the empty set. It
may appear useless, but it is a handy thing to have around during proofs! 

Let f: X → Y be a function. The subset of Y consisting of the values


actually taken by the function is called the range of f. This is the same
denition that we had for the range of any relation.

Example 2.3.7 Consider f : N → N dened by f (n) = 2n. The domain and


codomain are both N, but the range is the set 2N of even natural numbers.


We have been using the name f for a function. This is simply the most
commonly used notation for a function (as x is for a variable), but we are
free to use any other letter, symbol, or word. Other popular choices are g,
h, u, v , F , G, H , η , θ and so on. Functions that are particularly important
have their own names such as sin, cos, exp and log.
A function may also be called a map or mapping. If f is a function and
f (x) = y we say that f maps x to y.
Further, suppose f: X → Y is given by a clear rule, such as f (x) = x2 .
2
Then we represent the rule by a special arrow: x 7→ x .

For example, suppose the function f: R × R → R just adds the two


coordinates of an ordered pair. Then we represent the function by (x, y) 7→
x + y.
Finally, two functions f, g : X → Y are considered equal if they are equal
as subsets of X ×Y. That is,

f = g ⇐⇒ (∀x ∈ X, f (x) = g(x)).

1-1 Functions

A function f: X →Y is called one-one or 1-1 or injective if distinct points


in X have distinct images in Y: If a, b ∈ X and a 6= b then f (a) 6= f (b).

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 73

Task 2.3.8 Show that f: X → Y is 1-1 if and only if f (a) = f (b) implies
a = b.

Example 2.3.9 Consider the functions depicted below.

a
b
(A) (B)

The function in (A) is 1-1 because distinct points in the domain are mapped
to distinct points in the codomain. The function in (B) is not 1-1 because
the points a and b are mapped to the same value. 

Task 2.3.10 Let f: X → Y be a map. Express the condition for f to be


1-1 as a universally quantied statement. Also express its negation as an
existentially quantied statement.

Task 2.3.11 Can a constant function be 1-1?

To prove that a function is 1-1 we can choose between the usual ap-
proaches:

Direct proof: Assume a 6= b and prove f (a) 6= f (b).


Proof by contrapositive: Assume f (a) = f (b) and prove a = b.
Proof by contradiction: Assume f (a) = f (b) and a 6= b, and derive a
contradiction.

Since it is easier to manipulate equalities, the proofs by contrapositive or


contradiction are usually easier to implement.

Example 2.3.12 Consider the function f : R → R dened by f (x) = x3 . To


see if it is 1-1 we take up the contrapositive approach and assume f (a) = f (b).
We hope to deduce a = b.
Our assumption is a 3 = b3 . We derive some consequences as follows:

a3 = b3 =⇒ a3 − b3 = 0
=⇒ (a − b)(a2 + ab + b2 ) = 0
=⇒ a − b = 0 or a2 + ab + b2 = 0.
74 CHAPTER 2. RELATIONS AND FUNCTIONS

Ifa2 + ab + b2 = 0 then ab ≤ 0, and hence a3 b3 ≤ 0. Now a3 b3 < 0 violates


a = b3 so we must have a3 b3 = 0 and so a = 0 = b.
3 2 2
And a + ab + b 6= 0
anyway implies a − b = 0 and hence a = b.

Thus, a3 = b3 implies a=b and so f is 1-1. 

On the other hand, to prove that a function f is not 1-1 we need to nd
some specic a and b such that a 6= b and f (a) = f (b). This usually calls
for more than manipulating the rule for the function. We need to recognize
some key characteristic that may lead to repeated values.

Example 2.3.13 Consider the function f : R → R, x 7→ x2 . We realize


that squaring converts negative to positive, and so −a and a would have
the same image. Just stating this general observation would not suce as a
proof that the function is not 1-1. We have to provide two specic points in
the domain. For example, we can complete the proof by noting that 1 6= −1
and f (−1) = f (1) = 1. 

Task 2.3.14 Give an example of a cubic polynomial function on R that is


not 1-1.

Task 2.3.15 Let X be any nonempty set and let P(X) be its power set.
Give a 1-1 map f : X → P(X).

Onto Functions

A function f : X → Y is called onto or surjective if Y = ran(f ), that is,


for each b ∈ Y there exists a ∈ X such that f (a) = b.

Example 2.3.16 Consider the functions depicted below.

(A) (B)

The function in (A) is not onto because the point z in the codomain has no
pre-image. The function in (B) is onto. 

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 75

Task 2.3.17 Consider a function f: X → Y. Express the condition for f


to be onto as a universally quantied statement. Also express its negation
as an existentially quantied statement.

To prove that a function f : X → Y is onto, we typically choose between


the direct proof (take any b ∈ Y and nd a corresponding a ∈ X such that
f (a) = b) or the proof by contradiction (assume there is a b ∈ Y such that
f (a) 6= b for every a ∈ X , and reach a contradiction). In our initial examples,
we will take the direct approach. However, as we take up more complicated
functions later on, we will nd that proofs by contradiction are often easier
because nding a specic pre-image may be hard to pull o.

Example 2.3.18 p(x) = x2 +2x+1 as a relation from R to [0, ∞).


Consider
2 2
First we note that p(x) = x + 2x + 1 = (x + 1) ≥ 0, hence [0, ∞) is a valid
choice of codomain and p is a function from R to [0, ∞). Next, we will show
that it is an onto function.

We are going to use the fundamental fact that every non-negative real

number y has a non-negative square root y. You would not have seen
7 the proof of this fact in school! We also just assume it for now. Proofs are
indicated in Exercises 5 and 6 of Ÿ3.6.

Let y ∈ [0, ∞). Consider the equation p(x) = y , or (x + 1)2 = y . We can


√ √
solve it for x: x = ± y − 1. Hence every y ∈ [0, ∞) has pre-image y − 1.


Task 2.3.19 Find out whether the following functions are 1-1 or onto.

(a) f : [0, 1] → [a, b], f (x) = bx + (1 − x)a.


(b) f : [0, 1] → [0, 1], f (x) = (1 − x)/(1 + x).
(c) f : R → R, f (x) = x2 + x + 1.
(d) f : Z → N, f (n) = 2n if n≥0 and f (n) = −2n − 1 if n < 0.

Do you feel surprised by the results of any of these calculations?

Task 2.3.20 Let 2N be the set of all even natural numbers. Give examples
of functions from 2N to N that are
76 CHAPTER 2. RELATIONS AND FUNCTIONS

(a) 1-1 but not onto, (b) Onto but not 1-1,

(c) Neither 1-1 nor onto, (d) Both 1-1 and onto.

Images and Inverse Images of Sets

Consider a function f: X →Y and a subset A ⊆ X. The image of A under


f is the set
f (A) = { y ∈ Y | (∃x ∈ X)(y = f (x)) }.

A f (A)

X Y

Task 2.3.21 Consider the function f: N → Z dened by f (n) = (−1)n .


Describe f (O) and f (E), where O is the set of odd natural numbers and E
is the set of even natural numbers.

Task 2.3.22 Show that for any function f : X → Y , f (∅) = ∅.

Theorem 2.3.23 Consider a function f: X → Y and subsets A, B of X.


Then

1. A ⊆ B =⇒ f (A) ⊆ f (B),
2. f (A ∪ B) = f (A) ∪ f (B),
3. f (A ∩ B) ⊆ f (A) ∩ f (B) and equality may fail.

Proof:

1. Take any element y ∈ f (A). We have to show that y ∈ f (B). First,


y ∈ f (A) implies there is an x ∈ A such that y = f (x). And A ⊆
B =⇒ x ∈ B . Hence y = f (x) ∈ f (B).
2. A, B ⊆ A ∪ B =⇒ f (A), f (B) ⊆ f (A ∪ B) =⇒ f (A) ∪ f (B) ⊆
f (A ∪ B).

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 77

For the other inclusion, let y ∈ f (A ∪ B). Then there is x ∈ A ∪ B such


thaty = f (x). If x ∈ A we have y ∈ f (A). If x ∈ B we have y ∈ f (B).
Hence y ∈ f (A) ∪ f (B).

3. A ∩ B ⊆ A, B =⇒ f (A ∩ B) ⊆ f (A), f (B) =⇒ f (A ∩ B) ⊆
f (A) ∩ f (B).
Here is an example of failure of equality. Consider the constant function
f : Z → Z, n 7→ 0, A = Z+ , B = Z− . Then A∩B = ∅ =⇒ f (A∩B) =
∅. But, f (A) ∩ f (B) = {0} ∩ {0} = {0}. 

Task 2.3.24 Consider a function f: X →Y and a subset A of X . Will any


of the following denitely happen?

(a) f (Ac ) = f (A)c , (b) f (Ac ) ⊆ f (A)c , (c) f (Ac ) ⊇ f (A)c .

Task 2.3.25 Let f : X → Y be a 1-1 function and A ⊆ X. Show that


f (X \ A) = f (X) \ f (A).

Consider a function f: X →Y and a subset C ⊆Y. The inverse image


of C under f is the set

f −1 (C) = { x ∈ X | f (x) ∈ C }.

f −1 (C)
C

X Y

Example 2.3.26 Consider the function f: N → N dened by f (n) = 2n.


Then,

f −1 ({1, 2, 3, 4}) = {1, 2}, f −1 (E) = N, f −1 (O) = ∅.

Task 2.3.27 Consider the function f: N → Z dened by f (n) = (−1)n .


Describe the inverse images of the following:
78 CHAPTER 2. RELATIONS AND FUNCTIONS

(a) {1}, (b) Z+ , (c) Z− , (d) 2Z.

Task 2.3.28 Let f : (0, ∞) → (0, ∞) be dened by f (x) = 1/x. Describe


f −1 ((0, 1)) and f −1 ((1, ∞)).

Task 2.3.29 Show that for any function f : X → Y , f −1 (∅) = ∅ and


−1
f (Y ) = X .

Task 2.3.30 Let f : X → Y , f (x) = c, be a constant map. What is f −1 (D)


for D ⊆Y?

Let us consider the relationship between inverse images and the set op-
erations. As we see below, the situation is much neater than it was for
images!

Theorem 2.3.31 Consider a function f: X → Y and subsets C, D of Y.


Then

1. C ⊆ D =⇒ f −1 (C) ⊆ f −1 (D),
2. f −1 (C ∪ D) = f −1 (C) ∪ f −1 (D),
3. f −1 (C ∩ D) = f −1 (C) ∩ f −1 (D),
4. f −1 (C c ) = [f −1 (C)]c .

Proof:

1. x ∈ f −1 (C) =⇒ f (x) ∈ C =⇒ f (x) ∈ D =⇒ x ∈ f −1 (D).


2. x ∈ f −1 (C ∪ D) ⇐⇒ f (x) ∈ C ∪ D
⇐⇒ f (x) ∈ C ∨ f (x) ∈ D
⇐⇒ x ∈ f −1 (C) ∨ x ∈ f −1 (D)
⇐⇒ x ∈ f −1 (C) ∪ f −1 (D).
3. Similar to the proof for union. Just ip each ∪ to ∩ and each ∨ to ∧!
4. x ∈ f −1 (C c ) ⇐⇒ f (x) ∈ C c ⇐⇒ f (x) ∈ / f −1 (C) ⇐⇒
/ C ⇐⇒ x ∈
−1 c
x ∈ [f (C)] .


Theorem 2.3.32 Let f : X → Y , A ⊆ X, and C ⊆Y. Then

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 79

1. A ⊆ f −1 (f (A)),

2. f (f −1 (C)) ⊆ C .

Proof:

1. Let x ∈ A. Then f (x) ∈ f (A) =⇒ x ∈ f −1 (f (A)).

2. Let x ∈ f −1 (C). We need to show that f (x) ∈ C . This is true by the


−1
denition of f (C).

Equality can fail in these inclusions. For example, consider f: R → R
dened by f (x) = x2 . Then

1. f −1 (f (R≥0 )) = f −1 (R≥0 ) = R 6= R≥0 ,

2. f (f −1 (R)) = f (R) = R≥0 6= R.


Here is a diagrammatic representation of what can go wrong:

C
f −1 (f (A)) f −1 (C)
f (f −1 (C))
A f (A)

X Y X Y

In the rst diagram, the fact that f is not 1-1 allows an extra point to creep
into f −1 (f (A)). In the second diagram, the fact that f is not onto makes
−1
f (f (C)) lose a point.

Theorem 2.3.33 Let f: X →Y. The following hold:

1. If f is 1-1 then f −1 (f (A)) = A for all A ⊆ X.

2. If f is onto then f (f −1 (C)) = C for all C ⊆Y.

Proof: We already know that A ⊆ f −1 (f (A)) and f (f −1 (C)) ⊆ C . Let us


do the other inclusions.
80 CHAPTER 2. RELATIONS AND FUNCTIONS

1. x ∈ f −1 (f (A)) =⇒ f (x) ∈ f (A)


=⇒ (∃a ∈ A) f (x) = f (a)
=⇒ (∃a ∈ A) x = a (∵ f is 1 − 1)
=⇒ x ∈ A.

2. y ∈ C =⇒ (∃x ∈ X) f (x) = y (∵ f is onto)


−1
=⇒ (∃x ∈ f (C)) f (x) = y (∵ y ∈ C =⇒ x ∈ f −1 (C))
=⇒ y ∈ f (f −1 (C)).


Task 2.3.34 Let f: X →Y. Prove the converses of the results in the above
theorem:

1. If f −1 (f (A)) = A for all A⊆X then f is 1-1.

2. If f (f −1 (C)) = C for all C⊆Y then f is onto.

(Hint: Select appropriate A and C .)

Composition of Functions

Consider two functions f : X → Y and g : Y → Z . Their composition


g◦f: X →Z is dened by g ◦ f (x) = g(f (x)).

x f f (x) g g(f (x))

X Y Z

g◦f

Example 2.3.35 Consider f : N → Z, n 7→ (−1)n , and g : Z → N, n 7→


2
n + 1. Then g◦f: N→N is dened by

g ◦ f (n) = g(f (n)) = g((−1)n ) = ((−1)n )2 + 1 = (−1)2n + 1 = 1 + 1 = 2.

In this case g◦f is the constant function 2. 

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 81

Theorem 2.3.36 Consider functions f: X →Y and g : Y → Z, with com-


position g ◦ f : X → Z. Then,

1. If both f and g are 1-1, then g◦f is 1-1.

2. If both f and g are onto, then g◦f is onto.

Proof:

1. g ◦ f (a) = g ◦ f (b) =⇒ g(f (a)) = g(f (b))


=⇒ f (a) = f (b) (∵ g is 1-1)

=⇒ a = b. (∵ f is 1-1)

2. c ∈ Z =⇒ (∃b ∈ Y ) g(b) = c (∵ g is onto)

=⇒ (∃a ∈ X) f (a) = b (∵ f is onto)

=⇒ g ◦ f (a) = g(f (a)) = g(b) = c.



This theorem establishes sucient conditions for the composition to be
1-1 or onto. Naturally we ask if these conditions are also necessary. If you
can produce the examples asked for below, you will see that they are not
necessary.

Task 2.3.37 Dene functions f, g : Z → Z such that

(a) g is not 1-1 but g◦f is 1-1, (b) f is not onto but g◦f is onto.

Here are conditions that are necessary:

Theorem 2.3.38 Let f: X →Y and g : Y → Z. Then

1. If g◦f is 1-1, then f is 1-1.

2. If g◦f is onto, then g is onto.

Proof:

1. f (a) = f (b) =⇒ g(f (a)) = g(f (b)) =⇒ g◦f (a) = g◦f (b) =⇒ a = b.
2. c ∈ Z =⇒ ∃a ∈ X, g ◦ f (a) = c =⇒ g(f (a)) = c =⇒ g(b) =
c, where b = f (a). 
82 CHAPTER 2. RELATIONS AND FUNCTIONS

Finally, let us see the interaction of composition with images and inverse
images of sets.

Theorem 2.3.39 Consider functions f: X → Y and g : Y → Z. Let A⊆


X and C ⊆ Z. Then

1. g ◦ f (A) = g(f (A)),


2. (g ◦ f )−1 (C) = f −1 (g −1 (C)).

Proof: Left for you! 

Exercises for Ÿ2.3

1 Consider the relations given below as subsets of X ×Y. Which of them


represent functions from X to Y?

2 × 2 × 2 ×
1 × 1 × 1 ×
0 × 0 × 0 ×
−1 × −1 × −1 ×
−2 × −2 × −2 ×
× × × × × × × × × × × × × × ×
−2 −1 0 1 2 −2 −1 0 1 2 −2 −1 0 1 2
(a) (b) (c)

2 Let f: X → Y be a map. We say that x1 ∼ x2 i f (x1 ) = f (x2 ). Show


that this denes an equivalence relation on X.

3 Consider f, g : R × R → R dened by f (x, y) = x + y and g(x, y) =


x2 + y 2 . Find the equivalence classes induced by these functions. (See the
last exercise)

4 Let R be a relation on X . Dene its `inverse relation' R−1 by aR−1 b ⇐⇒


bRa. Which of the following are true?

(a) If R is an equivalence relation, so is R−1 .


(b) If R is a partial order, so is R−1 .
(c) If R is a function, so is R−1 .

5 Let I = [−1, 1]. Which of the following relations are functions?

Undergraduate Mathematics: Foundations


2.3. FUNCTIONS 83

(a) { (x, y) ∈ I × I | x2 + y 2 = 1 },
(b) { (x, y) ∈ I × I | x2 + y 2 = 1 ∧ x ≥ 0 },
(c) { (x, y) ∈ I × I | x2 + y 2 = 1 ∧ y ≥ 0 }.

6 Determine whether the given functions f, g : R → R are 1-1 or onto:

x2
 
1/x if x 6= 0 if x≥0
(a) f (x) = , (b) g(x) = .
0 if x=0 −x2 if x≤0

7 Let f, g, h : R → R be dened by

x x2 x3
f (x) = , g(x) = , h(x) = .
1 + x2 1 + x2 1 + x2
(a) Determine which of them are injective.

(b) Show that f and g are not surjective.

8 Show that f : R → R, x 7→ x4 + x3 + x2 + x + 1 is not surjective by


showing that 0 is not in the image of f . (Hint: What happens if you multiply
by x − 1?)

9 Prove that f: X →Y is surjective if and only if f (X \ A) ⊇ Y \ f (A) for


all A ⊆ X.

10 Show that f: X →Y is injective if and only if f (A ∩ B) = f (A) ∩ f (B)


for all A, B ⊆ X .

11 f: N → A
Suppose and g: N → B are onto. Show that there is an onto
map h N → A ∪ B.
:

12 Let f: R→R be dened by f (x) = x2 . Describe the inverse images of:

(a) {1}, (b) [0, 1], (c) [−1, 1], (d) Z.

13 2 2
Let f : R → R be given by f (x, y) = x + y .
2
What are f −1 ({t}) and
f −1 ([a, b])? Draw pictures of these inverse images.

14 Consider functions f, g : X → X . Will the compositions f ◦g and g◦f


necessarily be equal?
84 CHAPTER 2. RELATIONS AND FUNCTIONS

15 Consider functions f : W → X, g : X → Y and h : Y → Z. Will the


`associative law' h ◦ (g ◦ f ) = (h ◦ g) ◦ f hold?

16 Consider a function f : X → X. Show that

(a) f ◦f is 1-1 if and only if f is 1-1.

(b) f ◦f is onto if and only if f is onto.

17 Consider a function f: X →Y. Prove that:

(a) f is onto if ∃g : Y → X such that f ◦ g = 1Y . 3


(b) f is 1-1 if and only if ∃h : Y → X such that h ◦ f = 1X .
(c) If f ◦ g = 1Y and h ◦ f = 1X then g = h.

2.4 Indexed Sets


We say a set indexed by a set I if there is an onto function f : I → X .
X is
For any element i ∈ I , we denote its image f (i) ∈ X by xi . We may represent
this indexing by writing X = { xi | i ∈ I }, and we call I the index set
while X is called the indexed set. The members of the set I act as labels
for the members of X . This is especially useful when we want to arrange the
members of X in some way.

For example, if X has three elements and we want to think of them as


a rst element, a second element and a third element, we may say, Let
the members of X be x1 , x2 and x3 . This statement uses the set {1, 2, 3}
to index X . As described above, we may write X = { xi | i ∈ {1, 2, 3} } or
X = { xi | i = 1, 2, 3 }.
E of even integers. Each even inte-
For another example, consider the set
ger has the form 2n where n is an integer, and this creates an indexing of E
by Z. We dene xn = 2n and write E = { xn | n ∈ Z }. We can also combine
both denitions in the previous sentence by writing E = { 2n | n ∈ Z }.

Task 2.4.1 Carry out an indexing of the set O of odd integers by the index
set Z.

The requirement of an onto indexing function ensures that every member


of the indexed set gets a label. Note that we did not require the indexing
function to be one-one. This allows an element to have multiple labels.

3 The `if ' direction is also true, provided we assume the Axiom of Choice. See Ÿ4.3.

Undergraduate Mathematics: Foundations


2.4. INDEXED SETS 85

S = { n2 | n ∈ Z }. Then S is the set of all


Consider the description
2 2
square integers and is indexed by Z. Since n = (−n) , an element labeled
by n is also labeled by −n. For example, 4 is labeled by both 2 and −2.
To describe the set of square integers, we would earlier have written S =
{ m ∈ Z | (∃n ∈ Z)(n2 = m) }. This illustrates how the use of index sets
4
leads to descriptions that are more ecient and easier to read.

Union and Intersection

Consider an indexed set X = { Xi | i ∈ I } whose members are sets Xi . We


represent the union and intersection of all the members of X by

[ [ \ \
X = Xi and X = Xi .
i∈I i∈I

respectively. If we use X to index itself, this notation becomes

[ [ \ \
X = X and X = X.
X∈X X∈X

Task 2.4.2
[
Let X = { Xn | n ∈ N } with Xn = {1, 2, . . . , n}. Find Xn
n∈N
\
and Xn .
n∈N

The laws for union and intersection, which we stated earlier for combinations
of two or three sets at a time, can be generalized to arbitrary collections of
sets.

Task 2.4.3 (Distributive Laws) Let X = { Xi | i ∈ I } and U be any set.


Then

[ [ \ \
(a) U ∩( Xi ) = (U ∩ Xi ), (b) U ∪( Xi ) = (U ∪ Xi ).
i∈I i∈I i∈I i∈I

Task 2.4.4 (de Morgan's Laws) Let X = { Xi | i ∈ I } such that each


Xi is a subset of a given set U. Then the complements relative to U obey:

4 Galileo noted that the set {1, 2, 3 . . . } can be used to label, without repetition, the
visibly smaller collection {1, 4, 9, . . . }. He appears to have considered this paradoxical and
counted it as a reason to not consider innite collections as mathematical entities.
86 CHAPTER 2. RELATIONS AND FUNCTIONS

[ \ \ [
(a) ( Xi )c = Xic , (b) ( Xi )c = Xic .
i∈I i∈I i∈I i∈I

Cartesian Product

Given an indexed set


Q { Xi | i ∈ I }S cartesian product
we can dene its

i∈I Xi . f: I →Q i∈I Xi with the prop-


It is the collection of all functions
erty that f (i) ∈ Xi for every i ∈ I . Given f ∈ i∈I Xi , it is convenient
to represent it by (xi )i∈I with xi = f (i). This resembles the ordered pair
notation and we can use it in a similar way. Each member of the carte-
sian product represents a simultaneous selection of one member from each
5
constituent set.

Example 2.4.5 X1 = {a}, X2 = {a, b} and X3Q= {a, b, c}. These


Suppose
sets are indexed by I = {1, 2, 3}. The cartesian product i∈I Xi consists of
functions f : I → X1 ∪X2 ∪X3 such that f (i) ∈ Xi for each i = 1, 2, 3. Such a
function is determined by its three values f (1), f (2), f (3). We can therefore
represent f by the notation (f (1), f (2), f (3)), which we call an ordered triple.

For example, (a, b, c) represents the function dened by f (1) = a, f (2) =


b and f (3) = c. The full list of members of the cartesian product is:

(a, a, a), (a, a, b), (a, a, c), (a, b, a), (a, b, b), (a, b, c).

This cartesian product is also denoted by X1 × X2 × X3 . 

Consider an indexed set { Xi | i ∈ I }. Here is some commonly used


notation for special cases of the cartesian product:

Q
i∈I Xi
n
Y
I = {1, . . . , n} Xi or X1 × · · · × Xn
i=1
Y∞
I=N Xi or X1 × X2 × · · ·
i=1
I = {1, . . . , n} and X = X1 = · · · = Xn Xn
I = N and X = X1 = X2 = · · · X∞
5 It has been a matter of intense debate whether it is reasonable to allow arbitrary
amounts of simultaneous selections in a proof in mathematics. For more on this, see Ÿ4.3
on the Axiom of Choice.

Undergraduate Mathematics: Foundations


2.4. INDEXED SETS 87

Exercises for Ÿ2.4

1 Suppose X is indexed by I and Y is indexed by J.


(a) Can we always index X ∪Y by I ∪ J?
(b) Can we always index X ∩Y by I ∩ J?

2 Suppose X is indexed by I, Y is indexed by J, and I ∩ J = ∅.


(a) Can we always index X ∪Y by I ∪ J?
(b) Can we always index X ∩Y by I ∩ J?

3
[ \
Find Xn and Xn :
n∈N n∈N

(a) Xn = nN = { nk | k ∈ N }, (b) Xn = nZ = { nk | k ∈ Z }.

4
[ \
Find Xr and Xr :
r∈[0,1] r∈[0,1]

(a) Xr = [r, 1], (b) Xr = (r − 1, r + 1),

(c) Xr = (r − 1/2, r + 1/2), (d) Xr = [r − 1/2, r + 1/2].

5 Let{ Xi | i ∈ I } be an indexed set, and let A be any set. Show that:

[ [ \ \
(a) A ∪ ( Xi ) = (A ∪ Xi ), (b) A∩( Xi ) = (A ∩ Xi ).
i∈I i∈I i∈I i∈I

6 Let { Xi | i ∈ I } be an indexed set, and let A be any set. Show that:

(a) If
[Xi ⊆ A for each i, then (b) If A\⊆ Xi for each i, then

Xi ⊆ A. A⊆ Xi .
i∈I i∈I

7 Let { Xi | i ∈ I } be an indexed set, and let J ⊆ I . Show that:


[ [ \ \
(a) Xi ⊆ Xi , (b) Xi ⊆ Xi .
i∈J i∈I i∈I i∈J

8 Let f: X →Y be a function and let { Xi | i ∈ I } be an indexed set whose


members are subsets of X. Prove the following:
88 CHAPTER 2. RELATIONS AND FUNCTIONS

[ [
(a) f( Xi ) = f (Xi ),
i∈I i∈I
\ \
(b) If f is 1-1 then f( Xi ) = f (Xi ).
i∈I i∈I

9 Let f: X →Y be a function and let { Yi | i ∈ I } be an indexed set whose


members are subsets of Y. Prove the following:

[ [ \ \
(a) f −1 ( Yi ) = f −1 (Yi ), (b) f −1 ( Yi ) = f −1 (Yi ).
i∈I i∈I i∈I i∈I

10 Let U be any set, and let { Xi | i ∈ I } be an indexed set. Then we have


the following distributive laws for the cartesian product:

[ [ \ \
(a) U ×( Xi ) = (U × Xi ), (b) U ×( Xi ) = (U × Xi ).
i∈I i∈I i∈I i∈I

11 Let { Xi | i ∈ I } and { Yi | i ∈ I } be indexed sets with the same index


set I. Then Y Y Y
Xi ∩ Yi = (Xi ∩ Yi ).
i∈I i∈I i∈I

12 Let X = { Xi | i ∈ I } be an indexed set such


S that eachTXi is a subset of
a given set U . If I = ∅, what denitions of i∈I Xi and i∈I Xi would be
consistent with the laws for union and intersection?

2.5 Bijections, Inverse Functions and Cardinal-


ity
As the quote at the start of this chapter illustrates, it is possible to have a
clear idea of equal or greater amounts without having an idea of number.
To see if a group of men has the same number of members as a group of
women, we pair them two-by-two, and see if either group has members left
over at the end. Then we can say whether one group had more than the
other, without actually counting how many members were in each group.

As you may have guessed, this leads to a discussion of a certain kind of


relation. (Everything does!) Each group is a set. In pairing members we
create a relation. When we complete the process, all elements of at least
one set (the smaller one) have been paired o. We can view this set as the
domain. A member from the domain is related to exactly one member from

Undergraduate Mathematics: Foundations


2.5. BIJECTIONS, INVERSE FUNCTIONS AND CARDINALITY 89

the other set, the one it was paired with. Thus our relation is actually a 1-1
function.

If the two groups have the same amount of people, there will be no
leftover members, and the function will be onto. Thus, a perfect pairing o
is represented by a function that is 1-1 and onto.

BP

BP

A function f: X → Y that is both 1-1 and onto is called a 1-1 corre-


spondence or a bijection.
Task 2.5.1 Can there be a bijection from:

(a) {a, b, c} to {1, 2, 3}?


(b) {a, b, c} to {1, 2}?
(c) {a, b} to {1, 2, 3}?

These tasks illustrate how the notion of bijection helps capture the in-
tuition of same amount. However, this simple idea has consequences that
seem unnatural at rst glance. For instance, we have seen earlier that there
is a bijection from N to Z, and so we should think of them as having the
same amount of points, even though Z has many points that are not in N!
We have also seen that all closed intervals [a, b] with a < b are in bijection
with each other, even though their widths vary indenitely. This is perhaps
to be expected as the numbers associated with an interval are due to our
choice of unit and hence are quite arbitrary. The next example is much more
surprising.
90 CHAPTER 2. RELATIONS AND FUNCTIONS

Example 2.5.2 We shall construct a bijection f : (−1, 1) → R, using the


following diagram:

y -axis

( ) x-axis
−1 x 1 f (x)

The point x on the horizontal axis is rst lifted vertically to the point P on
the semicircle and then sent out radially until it hits the horizontal axis at
f (x).
The semicircle has equationx2√+ (y − 1)2 = 1. Hence the point P corre-
sponds to the ordered pair (x, 1 − 1 − x2 ). By considering similar triangles
we see that f (x) is given by

f (x) f (x) − x x
= √ or f (x) = √ .
1 1 − 1 − x2 1 − x2
If you visualize sliding the point x along the interval (−1, 1) you should be
able to see that f (x) will range over the whole real line. We can check this
by calculations.

First we check that the function f is 1-1:

a b a2 b2
f (a) = f (b) =⇒ √ =√ =⇒ 2
=
1 − a2 1 − b2 1−a 1 − b2
=⇒ a2 (1 − b2 ) = b2 (1 − a2 ) =⇒ a2 = b2 =⇒ b = ±a.
Note that f (−a) = −f (a). If a 6= 0 we have f (a) 6= 0 and hence f (−a) =
−f (a) 6= f (a). So b 6= −a, and hence b = a. If a = 0 we have b = ±0 = 0 = a.
Next we check that f is onto. Take any y∈R and try to solve y = f (x)
for x. We have:

x y2
y=√ =⇒ y 2 (1 − x2 ) = x2 =⇒ y 2 = (1 + y 2 )x2 =⇒ x2 = .
1 − x2 1 + y2
y
By substituting, you can check that x = p works. 
1 + y2

Undergraduate Mathematics: Foundations


2.5. BIJECTIONS, INVERSE FUNCTIONS AND CARDINALITY 91

Task 2.5.3 Investigate the one-one, onto, or bijective nature of the following
functions. Also, consider what shrinking of the domain or the range or both.
is needed to make the function one-one or onto or bijective.

(a) f: Q→Q given by f (x) = x2 ,


(b) f: R→R given by f (x) = x2 ,
(c) f: C→C given by f (z) = z 2 .

Inverse Functions

Consider a bijection f : X → Y . Since it represents a perfect pairing of the


elements of X and Y , we can switch the roles of X and Y, and obtain a
function g : Y → X dened by:

g(y) = x ⇐⇒ f (x) = y.

The function g is called the inverse function of f , and is denoted by f −1 .


The inverse function reverses the action of the original function: If f maps
x to y then g = f −1 maps y to x.

f f −1

X Y X Y

Theorem 2.5.4 Let f : X → Y be a bijection and g = f −1 . Then g is a


bijection and f = g −1 .

Proof: g is 1-1: Let g(y) = g(y 0 ) = x. Then y = y 0 = f (x).


g is onto: Let x ∈ X. Consider y = f (x). Then g(y) = x.
Finally, the condition for f = g −1 is the same as the condition for g=
f −1 . 

Theorem 2.5.5 Let f : X → Y and g : Y → X be functions such that


g(y) = x ⇐⇒ f (x) = y . Then f is a bijection and g = f −1 .
92 CHAPTER 2. RELATIONS AND FUNCTIONS

Proof: The proof is very similar to the previous theorem, hence is left as an
exercise. 
The notion of inverse function can also be expressed neatly in terms of
composition:

Theorem 2.5.6 Consider functionsf : X → Y and g : Y → X . Then g is


the inverse function of f ⇐⇒ g ◦ f = 1X and f ◦ g = 1Y .

Proof: First, suppose g = f −1 . Then

f (x) = y =⇒ g(y) = x =⇒ g(f (x)) = x =⇒ g ◦ f = 1X ,


g(y) = x =⇒ f (x) = y =⇒ f (g(y)) = y =⇒ f ◦ g = 1Y .
For the converse, rst we note that g ◦ f = 1X implies that f is 1-1, while
f ◦ g = 1Y implies that f is onto. Hence f is a bijection and f −1 exists.
Therefore,

f ◦ g = 1Y =⇒ f −1 ◦ (f ◦ g) = f −1 ◦ 1Y
=⇒ (f −1 ◦ f ) ◦ g = f −1
=⇒ 1X ◦ g = f −1
=⇒ g = f −1 .


Task 2.5.7 Show that f: X →Y has an inverse function if and only if f is


a bijection.

We have two dierent interpretations of f −1 (A). The rst is that for


−1
any function f, f (A) is the set of pre-images of points of A. The second
interpretation is possible when f is a bijection, and it consists of treating
f −1 (A) as the image of A under the inverse function f −1 . Fortunately, these
two interpretations do not clash:

• If f: X →Y is not a bijection, only the rst interpretation is possible.

• If f: X →Y is a bijection, both interpretations lead to the same result:

x ∈ Image of A under f −1 ⇐⇒ ∃y ∈ A, x = f −1 (y)


⇐⇒ ∃y ∈ A, f (x) = y
⇐⇒ f (x) ∈ A
⇐⇒ x ∈ Inverse image of A under f.

Undergraduate Mathematics: Foundations


2.5. BIJECTIONS, INVERSE FUNCTIONS AND CARDINALITY 93

−1
A common error is to forget the rst, and main, interpretation of f (A).
−1
Often, we come across f (A) and conclude that this indicates the avail-
« −1
ability of the inverse function f . It does not, unless we have separate
information that f is a bijection.

The other common error is to confuse the notation for an inverse function
with that of the reciprocal of a real number. This confusion leads to
calculations like the following:
«
f (x) = x =⇒ f −1 (x) = x−1 = 1/x.

But the inverse function of f (x) = x is itself !

Task 2.5.8 Suppose f: X →Y and g: Y → Z are bijections. Show that:

(a) g◦f is a bijection, (b) (g ◦ f )−1 = f −1 ◦ g −1 .

Cardinality

We say that two sets A and B have the same cardinality if there is a
bijection f : A → B, and we denote this by |A| = |B|.

Task 2.5.9 Consider sets A, B , C . Show that:

(a) |A| = |A|,


(b) |A| = |B| =⇒ |B| = |A|,
(c) |A| = |B| and |B| = |C| =⇒ |A| = |C|.

We have already seen many examples of sets with the same cardinality,
for instance:

• |N| = |Z| = |O| = |E|,


• |[0, 1]| = |[a, b]| if a < b,
• |(−1, 1)| = |R|.
Note that we have dened same cardinality (|A| = |B|) but not cardinality
(|A|) by itself. This is in accordance with the realization that same amount
can be dened without dening amount or number.

Continuing in this vein, we say |A| ≤ |B| if there is a 1-1 function f : A →


B. And |A| < |B| if |A| ≤ |B| but |A| =
6 |B|.
94 CHAPTER 2. RELATIONS AND FUNCTIONS

BP

BP

X Y

An example of |X| ≤ |Y |.

In the diagram drawn just above, there is a function that is 1-1 but not

« onto. The fact that this function is not onto is not sucient for claiming
that |X| < |Y |, as there may be another function that is both 1-1 and
onto.

Example 2.5.10 We illustrate this terminology by means of a very simple


example. Let A = {1} and B = {2, 3}. Then f (1) = 2 denes a 1-1 function
from A to B . Hence |A| ≤ |B|. On the other hand, there is no onto function
from A to B , since such a function would have to map 1 to both 2 and 3.
Hence |A| =
6 |B|. Therefore |A| < |B|. 

The following theorem of Cantor is remarkable, as it shows there is no


`largest size' for sets.

Theorem 2.5.11 (Cantor's Theorem) For any set X , |X| < |P(X)|.

Proof: It is easy to create a 1-1 function f : X → P(X) by setting f (x) =


{x}. Therefore |X| ≤ |P(X)|. To prove that the inequality is strict, we will
show that there is no onto map from X to P(X).
The proof is by contradiction. Suppose g : X → P(X) is an onto map.
Dene a subset of X by

C = {x ∈ X | x ∈
/ g(x) }.

Undergraduate Mathematics: Foundations


2.5. BIJECTIONS, INVERSE FUNCTIONS AND CARDINALITY 95

Since C ∈ P(X), there is a point a∈X such that g(a) = C . Does a belong
to C? We have the following:

a ∈ C ⇐⇒ a ∈ g(a) ⇐⇒ a ∈
/ C.

This contradiction implies that g must not be onto. 


Until the time of Cantor, an innite collection was simply one that could
never be completely counted. This theorem shows that even if X is innite,
P(X) must represent an innity of a higher order. By repeatedly taking
power sets (P(P(X)) etc.) we see there must be an innity of innities.
Cantor's theorem is the rst really striking truth in this book. Prior to
this, we were just learning how to think and communicate with clarity and
accuracy. Finally our work begins to bear fruit.

Exercises for Ÿ2.5

1 Give an example of functions f, g : N → N such that g ◦ f = 1N but


f ◦ g 6= 1N .

2 Suppose f : X → Y and g : Y → Z such that g◦f is a bijection. Does


this imply that f and g are bijections?

3 Suppose f: X →Y and g : Y → X such that g ◦ f and f ◦g are bijections.


Does this imply that f and g are bijections?

4 Prove that f: X →Y is a bijection if and only if f (X \ A) = Y \ f (A) for


all subsets A ⊂ X.

5 Let a, b ∈ R with a < b. Find a bijection f : [a, b] → [0, 1].

6 Suppose that a function f : R → R, f (x) = ax + b, is its own inverse.


What are the possible values of a and b?

7 The following diagram illustrates a variation of Example 2.5.2. Use it to


derive bijections from (−1, 1) to R and from R to (−1, 1).
96 CHAPTER 2. RELATIONS AND FUNCTIONS

y -axis

( ) x-axis
−1 x 1 f (x)

8 Give bijections between

(a) R+ and (0, 1), (b) R+ and R.

9 Give a bijection between the set P of planes passing through the origin of
R3 and the set L of lines passing through the origin of R3 .

10 Suppose |A| = |C| and |B| = |D|. Which of the following are always
true?

(a) |A ∪ B| = |C ∪ D|, (b) |A ∩ B| = |C ∩ D|,

(c) |A × B| = |C × D|.

11 Suppose A and B are subsets of X, with |A| = |B|. Can we assert the
following?

(a) |A| = |B| =⇒ |X \ A| = |X \ B|.


(b) |X \ A| = |X \ B| =⇒ |A| = |B|.

12 Suppose f: X →Y, and dene an equivalence relation on X by x∼y if



f (x) = f (y). Let X be the set of equivalence classes for this relation.

(a) Show that there is a unique function f ∗ : X∗ → Y such that f ∗ ([x]) =


f (x) for each x ∈ X.

(b) Show that f is 1-1.

(c) If f is onto, show that f∗ is a bijection.

13 Give a bijection between the plane R2 and the open unit ball D =
2 2 2
{ (x, y) ∈ R | x + y < 1 }.

14 Does  A ∼B if |A| = |B| dene an equivalence relation?

Undergraduate Mathematics: Foundations


2.6. BINARY OPERATIONS 97

2.6 Binary Operations


A very common procedure in mathematics is to take two objects and combine
them in some way to get a third object. For example, we add two numbers,
or take the conjunction of two statements, or carry out the union of two
sets. This can also be expressed in the language of functions. Considering
two objects to be combined amounts to choosing an ordered pair from a
cartesian product. The process of combining them can be represented by a
function.

A binary operation on a set X is a function from X × X to X . We can


use the familiar notation f (a, b) to represent the output corresponding to an
input (a, b). However, the usual way is to adopt a symbol as the function
name and then place the symbol between the two inputs. For example, if we
use ∗ as the name for the binary operation then a ∗ b is used to represent the
function value corresponding to (a, b).

X a
a∗b
b X

You are already familiar with many binary operations, even if you did
not know that term earlier! For example:

1. Addition of natural numbers, + : N × N → N, (m, n) 7→ m + n.


2. Multiplication of natural numbers: × : N × N → N, (m, n) 7→ m × n.
3. Addition and multiplication of other numbers: integers, rational num-
bers, real numbers, complex numbers.

4. Subtraction of integers, − : Z × Z → Z, (m, n) 7→ m − n.


5. Subtraction of other numbers: rational numbers, real numbers, com-
plex numbers.

6. Union of subsets of a set X:

∪ : P(X) × P(X) → P(X), (A, B) 7→ A ∪ B.


98 CHAPTER 2. RELATIONS AND FUNCTIONS

7. Intersection of subsets of a set X:


∩ : P(X) × P(X) → P(X), (A, B) 7→ A ∩ B.

8. If X is a totally ordered set, then (a, b) 7→ max{a, b} is a binary oper-


ation on X.
We do not have subtraction as a binary operation for natural numbers,
because some subtractions do not lead to natural numbers. Similarly, we do
not have division as a binary operation on natural numbers since one natural
number may not be divisible by another.

Example 2.6.1 Let F(X) be the set of functions from X to X. Then


composition of functions gives a binary operation ◦ on F(X). 

Example 2.6.2 If you have studied vectors in high school, you may recall
the cross product for three-dimensional vectors. It is a binary operation on
R3 . 

Task 2.6.3 Can division be viewed as a binary operation on any of the


standard types of numbers?

A binary operation ∗ on a set X has various terms associated with it:

• It is commutative if a ∗ b = b ∗ a for all a, b ∈ X .


• It is associative if a ∗ (b ∗ c) = (a ∗ b) ∗ c for all a, b, c ∈ X .

• An element e ∈ X is an identity for ∗ if e ∗ x = x ∗ e = x for every


x ∈ X.
• If X has an identity element e, then b is the inverse of a if a ∗ b =
b ∗ a = e.
• The binary operation ∗ is said to distribute over another binary oper-
ation ◦ on X , if a ∗ (b ◦ c) = (a ∗ b) ◦ (a ∗ c) and (b ◦ c) ∗ a = (b ∗ a) ◦ (c ∗ a)
for every a, b, c ∈ X .

Task 2.6.4 Let ∗ be a binary operation on X . Show that if ∗ has an identity


then the identity is unique.

Theorem 2.6.5 (Cancellation Law) Let ∗ be an associative binary oper-


ation on X, with identity e, and let a have an inverse. Then

x ∗ a = y ∗ a =⇒ x = y and a ∗ x = a ∗ y =⇒ x = y.

Undergraduate Mathematics: Foundations


2.6. BINARY OPERATIONS 99

Proof: Let b be the inverse of a. Then

x ∗ a = y ∗ a =⇒ (x ∗ a) ∗ b = (y ∗ a) ∗ b =⇒ x ∗ (a ∗ b) = y ∗ (a ∗ b)
=⇒ x ∗ e = y ∗ e =⇒ x = y.

The other implication is proved similarly. ’




Example 2.6.6 Consider a power set P(X) with the binary operations ∪
and ∩. Then:

1. Both ∪ and ∩ are commutative and associative.

2. The empty set ∅ is the identity for ∪, while X is the identity for ∩.
3. Only ∅ has an inverse relative to ∪, and only X has an inverse relative
to ∩.
4. The two operations distribute over each other:

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

Example 2.6.7 Consider a power set P(X) of a non-empty set X, with


the binary operation of set dierence: (A, B) 7→ A \ B . Then:
1. The binary operation \ is neither commutative nor associative. (Give
examples)

2. Set dierence has no identity. (Prove this) 

Example 2.6.8 Let F(X) be the set of functions from X to X, with com-
position of functions as the binary operation.

Composition is not commutative if X has at least two distinct pointsa


and b. Let f: X →X be the constant function with value a and g : X → X
be the constant function with value b. Then f ◦ g = f and g ◦ f = g .
Composition is associative, since both (f ◦ (g ◦ h))(x) and ((f ◦ g) ◦ h)(x)
equal f (g(h(x))).
The identity function 1X serves as the identity for composition.

The inverse of a function relative to composition is just its inverse function


as dened earlier. So not all functions have inverses. 
100 CHAPTER 2. RELATIONS AND FUNCTIONS

Example 2.6.9 Let S3 be the set of all bijections of F3 = {1, 2, 3} with


itself. Members of S3 are of three types:

1. Bijections that switch two numbers and leave the third one unchanged,
for example 1 7→ 2, 2 7→ 1, 3 7→ 3.

2. Bijections that move all three numbers, for example: 1 7→ 2, 2 7→ 3,


3 7→ 1.

3. The identity function.

Members of S3 can be visualized as the symmetries of an equilateral triangle


whose vertices are labeled by 1, 2 and 3. First, let us use the name Rk for
the bijection that maps k to k and switches the other two numbers. Each
of these can be viewed as a reection of the triangle about an altitude, as
shown below.

1 1 1

2 3 2 3 2 3

R1 R2 R3

The other bijections can be viewed as counterclockwise rotations by 0, 120


and 240 degrees around the centre of the triangle. We name them T0 , T120
and T240 respectively. Of course T0 is just the identity map. And T240 is the
same as a clockwise rotation by 120 degrees.

1 1 1

2 3 2 3 2 3

T0 T120 T240

Undergraduate Mathematics: Foundations


2.6. BINARY OPERATIONS 101

We leave it to you to check that these six bijections exhaust S3 .


As usual, we take composition as the binary operation. This works be-
cause the composition of two bijections is a bijection. What are its proper-
ties?

1. Composition on S3 is not commutative. For example, R1 ◦ R2 = T120


while R2 ◦ R1 = T240 .
2. Composition is associative.

3. T0 is the identity element.

4. Every element has an inverse. Each reection is its own inverse, while
T120 and T240 are the inverses of each other. 

Let∗ be a binary operation on X . A subset A ⊆ X is called closed with


respect to ∗ if x, y ∈ A =⇒ x ∗ y ∈ A. For example, in S3 the sets A =
{T0 , R1 } and B = {T0 , T120 , T240 } are closed with respect to composition.
On the other hand, C = {R1 } is not closed since R1 ◦ R1 = T0 ∈ / C.
When X is a nite set, a binary operation ∗
X can be represented by
on
a multiplication table. X as {x1 , . . . , xn }. Then
We list the members of
the entry in the ith row and j th column of the multiplication table is xi ∗ xj .

Example 2.6.10 Let X = {0, 1, 2}. Dene ∗ by 0 ∗ k = k ∗ 0 = 0 for each


k, 1 ∗ k = k ∗ 1 = k for eachk , and 2 ∗ 2 = 1. The multiplication table for ∗
is

0 1 2
0 0 0 0
1 0 1 2
2 0 2 1

Suppose someone were provided this multiplication table. Certain things


would be immediately apparent. For example:

(a) 1 is the identity as the row and column for 1 just replicate the list of
members of X in the same order.

(b) 0 has no inverse since the row (and column) for 0 has no entry of 1.

(c) 1 and 2 are their own inverse.

(d) The subset {1, 2} is closed since the entries for combinations of 1 and
2 only consist of 1 and 2.
102 CHAPTER 2. RELATIONS AND FUNCTIONS

(e) The operation is commutative since the table is symmetric about the
diagonal running from the top left to the bottom right. 

Exercises for Ÿ2.6

1 Let B(X) be the set of all bijections from X to X, with composition as


the binary operation. Show that:

(a) If |X| ≤ 2 composition is commutative.

(b) If |X| ≥ 3 composition is not commutative.

(c) Composition is associative.

(d) 1X is the identity element.

(e) Every member of B(X) has an inverse.

2 Recall that the symmetric dierence of two sets is dened by A4B =


(A \ B) ∪ (B \ A). X = {a, b}.
Let Draw the multiplication table for 4 as a
binary operation on P(X).

3 Consider the symmetric set dierence 4 as a binary operation on a power


set P(X). Prove the following:

(a) Symmetric dierence is commutative and associative.

(b) The empty set is the identity element for 4.


(c) Every member of P(X) has an inverse.

(d) Intersection ∩ distributes over 4.

4 Let ∗ be a binary operation on X . An element el ∈ X is a left identity


ifel ∗ x = x for every x ∈ X . Also, er ∈ X is a right identity if x ∗ er = x
for every x ∈ X . Show that if ∗ has both a left identity and a right identity
then they are equal.

5 Let ∗ be an associative binary operation on X with identity element e. If


a ∗ b = e we call a a left inverse of b, and we call b a right inverse of a.
(a) Give an example of ∗ where some elements fail to have a left inverse,
and some elements fail to have a right inverse.

(b) Give an example of ∗ where some elements have multiple left inverses.
(c) Show that if an element has both a left inverse and a right inverse
then they are equal, and hence are the inverse of that element. (In
particular, an element can have at most one inverse)

Undergraduate Mathematics: Foundations


2.6. BINARY OPERATIONS 103

6 Let ∗ be an associative binary operation on X with identity element e.


Suppose a has inverse a0 and b has inverse b0 . Show that a∗b has inverse
b0 ∗ a0 .

7 Let X = {T, F}.


(a) How many binary operations are there on X?
(b) Taking T as `True' and F as `False', consider the binary operation
given by ∧. (For example, T ∧F=F and T ∧ T = T) Is this binary
operation associative? Commutative? Does it have an identity?

(c) Similarly, consider the binary operation given by ∨. Is it associative?


Commutative? Does it have an identity?

8 Consider S3 , the symmetries of an equilateral triangle, with composition


as the binary operation. Prove the following:

(a) Every rotation is the composition of two reections.

(b) If R is a reection and T is a rotation then R◦T ◦R is a rotation, in


fact it is the inverse of T.

9 Draw the multiplication table for S3 with composition as the binary op-
eration.

10 Which non-empty subsets of S3 are closed under composition?

11 Consider the equivalence relation on k ∼ l if n | (k − l).


Z dened by
Let Zn 0, 1,. . . , n − 1. Show
be the set of corresponding equivalence classes
that we can dene a binary relation + on Zn by k + l = k + l. What are its
properties? (Zn with this + is known as the integers modulo n)

12 Consider Z6 with binary operation + as dened in the previous exercise.


Which subsets of Z6 are closed under +?

13 Let D4 denote the set of symmetries of a square with its vertices labeled
1, 2, 3, 4. Which bijections of {1, 2, 3, 4} correspond to members of D4 ?
Which are reections and which are rotations?

14 Let ∗ be a binary operation on a nite set G such that it is associative,


has an identity, and every element of G has an inverse. Show that each
row and column of the multiplication table of ∗ contains each element of
G exactly once. (Can you express this in terms of certain functions being
bijections?)
104 CHAPTER 2. RELATIONS AND FUNCTIONS

15 Let G = {0, 1, 2}. Let ∗ be a binary operation on G that is associative,


has 0 as identity, and such that every element of G has an inverse. Show
that the multiplication table for G must be as follows, and conclude that ∗
is commutative:

0 1 2
0 0 1 2
1 1 2 0
2 2 0 1

16 Let G = {0, a, b, c}. Let ∗ be a binary operation on G that is associative,


has 0 as identity, and such that every element of G has an inverse.
(a) Prove that the number of elements of G that are their own inverse is
either two or four.

(b) Suppose G has exactly two elements that are their own inverse. Then
there is a pair of elements which are the inverses of each other. Name
one of them 1. Also dene 2 = 1 + 1, and name the last member 3.
With these renamings, show that the multiplication table becomes

0 1 2 3
0 0 1 2 3
1 1 2 3 0
2 2 3 0 1
3 3 0 1 2

(c) Now consider the other case, when every member of G is its own
inverse. Show that the multiplication table for G is the following:

0 1 2 3
0 0 1 2 3
1 1 0 3 2
2 2 3 0 1
3 3 2 1 0

(d) Check that ∗ is necessarily commutative.

Undergraduate Mathematics: Foundations


3 | Number Systems
The less is to be taken from the greater, positive from positive,
negative from negative. When the greater, however, is subtracted
from the less, the dierence is reversed. Negative, taken from
cipher, becomes positive; and armative, becomes negative. Neg-
ative, less cipher, is negative; positive is positive; cipher, nought.
When armative is to be subtracted from negative, and negative
from positive, they must be thrown together.
1
 Brahmagupta

It came to appear that, between two truths of the real domain, the
easiest and shortest path quite often passes through the complex
domain.
2
 Paul Painlevé

In earlier chapters, we have set up a framework within which the concepts


and arguments of mathematics can be expressed. Our goal has been to
express all concepts in the language of set theory and all arguments using the
tools of symbolic logic. In order to motivate our procedures and decisions
we have used many examples from the world of numbers. We have used
the sets and subsets of natural numbers, whole numbers, integers, rational
numbers and real numbers. We have also carried out calculations with the
members of these sets. In doing so, we took the basic knowledge of these
objects and their properties for granted, on the grounds that they would
have been familiar to you from school.

It is time that all these numbers were also brought within our house
of set theory and logic. In the rst section of this chapter we shall give a
1 From his text Brahmasphutasiddhanta composed in 628 AD. Translation by Henry
Thomas Colebrooke, 1817. https://archive.org/details/algebrawitharith00brahuoft
2 See http://homepage.math.uiowa.edu/~jorgen/hadamardquotesource.html.

105
106 CHAPTER 3. NUMBER SYSTEMS

formal description of the set of natural numbers via the Peano axioms. In
later sections we shall use the natural numbers and set theory methods such
as cartesian product and equivalence relation to build the other kinds of
numbers.

3.1 Natural Numbers: Denitions


Peano Axioms

The most essential set of numbers is that of natural numbers. The formal ap-
proach to natural numbers originates with Charles Pierce, Richard Dedekind
and Giuseppe Peano in the late 19th century. The following axioms for the
set N of natural numbers are called the Peano Axioms:
P1. (Successor Function) There is a 1-1 function S : N → N.

P2. (Unity) There is a natural number denoted by 1, which is not in the


range of S.

P3. (Principle of Mathematical Induction) If A ⊆ N satises 1 ∈ A and


S(A) ⊆ A then A = N.

The rst axiom represents the intuition that each natural number has a
`next' number. (For example, S(1) = 2, S(2) = 3 and so on) The second,
that there is an initial natural number which does not succeed any other.
The third, that all natural numbers can be obtained by taking consecutive
successors starting with 1. The following theorem expands on the last point:

Theorem 3.1.1 The Peano axioms for N imply that if y ∈ N \ {1} then
∃!x ∈ N such that S(x) = y .

Proof: Let A = {1} ∪ S(N). Then 1 ∈ A. Suppose y ∈ A, and consider S(y).


Obviously S(y) ∈ A. So, by mathematical induction, A = N. 

Peano showed that all the basic properties of arithmetic with natural
numbers could be deduced from these axioms, and that is our topic of study
in this section. An obvious question at this stage concerns the actual exis-
tence of a set with a successor function that satises the Peano axioms, and
hence can serve as N. We leave that for the next chapter.

Undergraduate Mathematics: Foundations


3.1. NATURAL NUMBERS: DEFINITIONS 107

Recursion

To construct special functions and relations (addition, multiplication, total


order) involving N, our basic idea would be to use the successor function
to cover elements of N one-by-one. This may sound like induction but the
format is dierent. We are not proving that a given set is all of N. Instead,
we wish to expand a construction till it covers all of N. The next result shows
us how to do this.

Theorem 3.1.2 (Recursion Theorem) Consider a set X with a function


G : X → X and an element c ∈ X. Then there is a unique function f: N→
X such that
1. f (1) = c,
2. ∀n ∈ N, f (S(n)) = G(f (n)).

We are trying to justify a process of constructing a function by starting


7 with a given value f (1) = c, and then using a rule G for getting f (S(n))
from f (n) to extend the denition of f to all of N.

Proof: To construct f we rst consider all relations that have the same
essential properties. Thus, we let M be the set of all R⊆N×X such that

(a) (1, c) ∈ R,
(b) (n, x) ∈ R =⇒ (S(n), G(x)) ∈ R.
For example, N × X ∈ M. Now take the intersection of all these relations:
\
f= M.

Clearly f ∈ M. We need to show f is a function. At this stage, our only


available tool is mathematical induction. So we let

A = { n ∈ N | (n, x) ∈ f ∧ (n, x0 ) ∈ f =⇒ x = x0 }.

Let R ⊆ N × X be the relation obtained by excluding points of the form


(1, x) with x 6= c:

R = { (n, x) ∈ N × X | n 6= 1 ∨ x = c }.

Then R ∈ M. And (1, x) ∈ f =⇒ (1, x) ∈ R =⇒ x = c. Therefore 1 ∈ A.


We need a little preparation for the induction step. We'll show that if
(S(n), x) ∈ f then there is a t∈X such that x = G(t) and (n, t) ∈ f . For
108 CHAPTER 3. NUMBER SYSTEMS

if not, then consider the relation R = f \ {(S(n), x)}. Our assumption gives
R ∈ M, which contradicts the minimality of f.

n ∈ A. Let (S(n), x) ∈ f and (S(n), x0 ) ∈ f . There exist


Now suppose
t, t ∈ X such that G(t) = x, G(t0 ) = x0 and (n, t), (n, t0 ) ∈ f . Since n ∈ A
0
0 0
we have t = t , and hence x = x . This completes the induction step.

Hence, by mathematical induction, A = N. To complete the proof that


f is a function we have to conrm that its domain is all of N. This is easily
done by induction, as is the proof of uniqueness. ’


Now we shall recover the arithmetic and order properties of N from the
Peano Axioms. So we begin by forgetting everything about N except that it
is a set with a successor function S: N → N which obeys the Peano Axioms.
We have no addition, no order, and we can't make statements like take the
rst n elements.

We do have the Recursion Theorem. This will be our main tool.

Addition

In order to apply recursion to dening addition and multiplication, we need


to adapt it to binary operations. Let's understand the situation visually
rst. Consider the product N × N:

(1,1) (1,2) (1,3) (1,4) ···

(2,1) (2,2) (2,3) (2,4) ···

(3,1) (3,2) (3,3) (3,4) ···


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

We can do a recursion along each row, using its rst point as the base point.
This means that we begin by setting the value cn of the binary operation for
each (n, 1) pair.

Undergraduate Mathematics: Foundations


3.1. NATURAL NUMBERS: DEFINITIONS 109

(1,1) (1,2) (1,3) (1,4) ···

(2,1) (2,2) (2,3) (2,4) ···

(3,1) (3,2) (3,3) (3,4) ···


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

Next, we need a rule that nds the value for any (n, S(m)) by using the value
for (n, m). This requires functions Gn : N → N for each n.
To begin the recursion for dening addition, we rst dene n + 1 = S(n).
Then to obtain n + (m + 1) from n + m, we set n + (m + 1) = (n + m) + 1.

Theorem 3.1.3 There is a unique binary operation (called addition)

+ : N × N → N, (n, m) 7→ n + m,
such that n + 1 = S(n) and n + S(m) = S(n + m) for all natural numbers x
and y.

Proof: Fix n ∈ N. By the Recursion Theorem, there is a unique function


fn : N → N such that

(a) fn (1) = S(n),


(b) fn (S(m)) = S(fn (m)).
Dene the binary operation of addition by n + m = fn (m). 
If m + k = n we say k = n − m. In particular, if y = S(x) then x = y − 1.

Multiplication

To begin the recursion for dening multiplication, we rst dene n • 1 = n.


Then to obtain n • (m + 1) from n • m, we set n • (m + 1) = (n • m) + n.

Theorem 3.1.4 There is a unique binary operation (called multiplication),

• : N × N → N, (n, m) 7→ n • m,
such that n•1 = n and n • S(m) = (n • m) + n for all natural numbers n and
m.
110 CHAPTER 3. NUMBER SYSTEMS

Proof: Fix n ∈ N. By the Recursion Theorem, there is a unique function


fn : N → N such that

(a) fn (1) = n,
(b) fn (S(m)) = fn (m) + n.
Dene the binary operation of multiplication by n • m = fn (m). 

Theorem 3.1.5 Addition and multiplication satisfy the following laws:

1. (Commutative Law) ∀x, y ∈ N, x + y = y + x and x • y = y • x.

2. (Associative Law) ∀x, y, z ∈ N, (x + y) + z = x + (y + z) and (x • y) • z =


x • (y • z).

3. (Cancellation Law) ∀x, y, z ∈ N, x + y = x + z =⇒ y = z .

4. (Distributive Law) ∀x, y, z ∈ N, x • (y + z) = x • y + x • z .

Proof: Now that the operations have been established, their properties can
be veried by the usual induction approach. From our brief experience of
binary operations, we know that associativity is very helpful in proving other
facts. So we begin with the associative law for addition. Let

A = { n ∈ N | (` + m) + n = ` + (m + n) ∀`, m ∈ N }.

Now, the recursive property of + gives (` + m) + 1 = ` + (m + 1), which


implies that 1 ∈ A. Let n ∈ A. Then

(` + m) + (n + 1) = ((` + m) + n) + 1 (Theorem 3.1.3)

= ((` + (m + n)) + 1 (n ∈ A)
= ` + ((m + n) + 1) (Theorem 3.1.3)

= ` + (m + (n + 1)) (Theorem 3.1.3)

Hence n + 1 ∈ A, and so A = N, establishing the associativity of addition.

For the commutative law of addition, we rst use induction to establish


that n + 1 = 1 + n for every n. Let B = { n ∈ N | n + 1 = 1 + n }. Clearly
1 ∈ B . Let n ∈ B . Then 1 + S(n) = S(1 + n) = S(n + 1) = S(S(n)) =
S(n) + 1, and hence S(n) ∈ B . Therefore B = N.
Now let C = { n ∈ N | m + n = n + m ∀m ∈ N }. We have just seen that

Undergraduate Mathematics: Foundations


3.1. NATURAL NUMBERS: DEFINITIONS 111

1 ∈ C. Let n ∈ C. Then we show n + 1 ∈ C:

m + (n + 1) = (m + n) + 1 (Theorem 3.1.3)

= 1 + (m + n) (1 ∈ C)
= 1 + (n + m) (n ∈ C)
= (1 + n) + m (Associativity)

= (n + 1) + m (1 ∈ C)

We leave the other laws to be worked out by you. Exercise 2 of this section
provides one sequence for doing so. 

Task 3.1.6 Show that m, n ∈ N =⇒ m 6= m + n.

Zero

Many calculations and formulas involving natural numbers are simplied if


zero is also available to us. We can incorporate it as follows. First, we
decide to denote zero by the symbol 0. Then we combine it with the set of
natural numbers to create the set W of whole numbers:
W = {0} ∪ N

We extend the binary operations of addition and multiplication from N to


W by setting

0+n=n+0=n and 0•n=n•0=0 for all n∈W

Task 3.1.7 Check that addition and multiplication of whole numbers have
the following properties:

(a) Addition and multiplication are commutative and associative.

(b) 0 is the identity for addition and 1 is the identity for multiplication.

(c) Multiplication distributes over addition.

The successor function can be extended to W by dening S(0) = 1. Induction


and recursion also extend to W as follows:

Theorem 3.1.8 (Principle of Mathematical Induction for W) If A⊆


W such that 0∈A and n ∈ A =⇒ S(n) ∈ A, then A = W.
112 CHAPTER 3. NUMBER SYSTEMS

Proof: ’


Theorem 3.1.9 (Recursion Theorem for W) Consider a set X, a func-


tion G : X → X , and an element c ∈ X. Then there is a unique function
f : W → X such that
1. f (0) = c,
2. ∀n ∈ W, f (S(n)) = G(f (n)).

Proof: ’


Exponentiation

We will now develop the notion of `repeated multiplication' of a number with


itself. We shall develop it not just for N or W but in a general setting that
can then be applied to each system of numbers in turn.

Theorem 3.1.10 Let ∗ be a binary operation on X with an identity element


1. Let a ∈ X . Then there is a unique function Ea : W → X such that
Ea (0) = 1 and Ea (n + 1) = Ea (n) ∗ a.

Proof: Apply the Recursion Theorem for W to X with G : X → X dened


by G(b) = b ∗ a, and the initial condition Ea (0) = 1. 
We shall apply this theorem to the binary operation of multiplication.
With that in mind, we introduce the notation an = Ea (n). Note that

a0 = Ea (0) = 1,
a1 = Ea (0) ∗ a = 1 ∗ a = a,
a2 = Ea (1) ∗ a = a ∗ a,
.
.
.

an+1 = E(n) ∗ a = an ∗ a.

Theorem 3.1.11 Let ∗ be an associative binary operation on X with an


identity element 1. Let a ∈ X and m, n ∈ W. We have the following:

Undergraduate Mathematics: Foundations


3.1. NATURAL NUMBERS: DEFINITIONS 113

(a) am+n = am ∗ an , (b) (am )n = am • n


.

Proof: To establish (a), we let A = { n ∈ W | (∀m ∈ W)(am+n = am ∗ an ) }


and apply mathematical induction for W.

First, we have 0∈A since am+0 = am = am ∗ 1 = am ∗ a0 .


Now suppose n ∈ A. Then

m+(n+1) (m+n)+1
a =a = am+n ∗ a = (am ∗ an ) ∗ a = am ∗ (an ∗ a) = am ∗ an+1 .
Hence n+1∈A and the proof is complete.

We can prove (b) in a similar fashion. ’




Theorem 3.1.12 Let ∗ be an associative and commutative binary operation


on X with an identity element 1. Let a, b ∈ X and n ∈ W. Then

(a ∗ b)n = an ∗ bn .

Proof: We let A = { m ∈ W | (a ∗ b)m = am ∗ bm } and apply mathematical


induction for W. First, (a ∗ b)0 = 1 = 1 ∗ 1 = a0 ∗ b0 =⇒ 0 ∈ A. Now
assume n ∈ A. Then,

(a ∗ b)n+1 = (a ∗ b)n ∗ (a ∗ b)
= (an ∗ bn ) ∗ (a ∗ b)
= ((an ∗ bn ) ∗ a) ∗ b
= (a ∗ (an ∗ bn )) ∗ b
= ((a ∗ an ) ∗ bn ) ∗ b
= an+1 ∗ (bn ∗ b)
= an+1 ∗ bn+1 .
Therefore n + 1 ∈ A. Hence, A = W. 

When we apply this denition to the binary operation of multiplication on


W, one special value that we get is 00 = 1. If you have studied Calculus
0
you may have been taught that 0 is an `indeterminate form' and that it
« 0
has no xed value. In fact, calling 0 an indeterminate form means only
y
that the function x does not have a limit at (0, 0). It does not preclude
0
giving that function a value at (0, 0). Indeed, the value 0 = 1 is used in
0
other places in Calculus itself, for example while using the function x .
114 CHAPTER 3. NUMBER SYSTEMS

Order

Now we equip N with an order.

Theorem 3.1.13 The relation on N dened by m≤n if m=n or ∃k ∈ N


such that m + k = n, is a total order.

Proof: We check the required properties:

(a) Reexivity is directly built into the denition.

(b) For anti-symmetry, suppose m ≤ n and n ≤ m. If m 6= n, we have


m + k = n and n + ` = m for some k, ` ∈ N. But then n + (` + k) = n,
which is impossible.

(c) Transitivity is obvious.

(d) For totality, x m∈N and consider the set

A = { n ∈ N | (m ≤ n) ∨ (n ≤ m) }.

We need to prove that A = N.


If m=1 then of course 1 ∈ A. If m 6= 1 then m = m0 + 1 for some
0
m and hence 1 ≤ m. In this case too, 1 ∈ A.

Let n∈A and consider n + 1.


Case 1: Let m ≤ n. Then m ≤ n ≤ n + 1.
Case 2: Let n ≤ m and n 6= m. Then there is k ∈ N such that
n + k = m. If k = 1 we are done. Else, we have k = k 0 + 1 and so
(n + 1) + k 0 = m, implying n + 1 ≤ m. 
Recall that any order ≤ has an associated strict order < dened by x < y
ifx ≤ y and x 6= y . In our case, this gives us m < n if ∃k ∈ N such
that m + k = n. The strict order automatically possesses the properties of
irreexivity, trichotomy and transitivity.

Theorem 3.1.14 Let n ∈ N. Then there is no m∈N such that n<m<


n + 1.

Proof: If n < m then ∃k ∈ N such that n + k = m. If k = 1 we get


m = n + 1 which contradicts m < n + 1. If k 6= 1 we have k = k 0 + 1, hence
(n + 1) + k 0 = m, and so n + 1 < m which again contradicts m < n + 1. 

Theorem 3.1.15 For any `, m, n ∈ N,

Undergraduate Mathematics: Foundations


3.1. NATURAL NUMBERS: DEFINITIONS 115

(a) n 6= 1 =⇒ 1 < n,
(b) n < n + `,
(c) m < n ⇐⇒ m + ` < n + `,
(d) m < n ⇐⇒ m • ` < n • `,

Proof:

(a) We know that if n 6= 1 then it is a successor: there is m∈N such


that m + 1 = n. This gives 1 < n.
(b) Obvious.

(c) The =⇒ direction is obvious. For the converse, apply the cancella-
tion law for addition.

(d) For the =⇒ direction, Let m + k = n. Then

m+k = n =⇒ (m+k) • ` = n • ` =⇒ m • `+k • ` = n • ` =⇒ m • ` < n • `

For the converse, we apply proof by contrapositive. Suppose n≤m


(the negation of m < n). Both n = m and n < m lead to n • ` ≤ m • `,
the negation of m • ` < n • `. 

Task 3.1.16 Prove the Cancellation Law for multiplication: If `, m, n ∈ N


then m • ` = n • ` =⇒ m = n.

From here on, we shall revert to our familiar notation of writing the
product of a and b as simply ab, rather than a • b.

Exercises for Ÿ3.1

1 Show that n 6= S(n) for each n ∈ N.

2 The following tasks provide one path for completing the proof of Theorem
3.1.5:

(a) Prove the cancellation law for addition.

(b) Prove the distributive laws for addition and multiplication.

(c) Prove the associativity of multiplication.

(d) Prove the commutativity of multiplication.

3 By denition, 2 = S(1) = 1 + 1, 3 = S(2) = 2 + 1 and 4 = S(3) = 3 + 1.


Prove that 4 = 2 + 2.
116 CHAPTER 3. NUMBER SYSTEMS

4 Let `, m, n ∈ N. Show that:

(a) m ≤ n ⇐⇒ m + ` ≤ n + `, (b) m ≤ n ⇐⇒ m` ≤ n`.

5 Let m, n ∈ W. Prove the following:

(a) m + n = 0 =⇒ m = n = 0, (b) mn = 0 =⇒ (m = 0∨n = 0).

6 Use induction to prove that 0n = 0 for each n ∈ N.

7 Prove the Principle of Mathematical Induction for W.

8 Prove the Recursion Theorem for W.

9 Let f : X → X . Show that for each n ∈ W there is a unique function


f n : X → X such that f 0 = 1X and f n+1 = f n ◦ f . (This exercise sets up
`repeated composition' of a function)

10 (Fibonacci Sequence) F : N → N
Show there is a unique function
such that F (1) = F (2) = 1 and for each n ∈ N, F (n + 2) = F (n) + F (n + 1).
(Hint: Consider the function G(m, n) = (n, m + n) on X = N × N)

11 (Factorial) F : W → N such that


Show there is a unique function
F (0) = 1 n ∈ W, F (n + 1) = (n + 1)F (n). (Hint: Consider the
and for each
function G(m, n) = (m + 1, (m + 1)n) on X = N × N)

3.2 Natural Numbers: Properties


In this section we develop some of the basic results and proof techniques
related to the set of natural numbers. We begin by looking at applications
of the Principle of Mathematical Induction, one of Peano's Axioms. We
prove a variation that is known as the Principle of Strong Induction and
which provides greater exibility. We use strong induction to prove the
Well-Ordering Principle, a useful result for identifying numbers with special
properties.

The Principle of Mathematical Induction can be expressed in terms of


possible properties of natural numbers by setting P (n) to be the property
that n ∈ A. Then its statement becomes:

Suppose P (1) is true, and (∀n ∈ N)(P (n) =⇒ P (n + 1)) is true. Then
P (n) is true for every n ∈ N.

Undergraduate Mathematics: Foundations


3.2. NATURAL NUMBERS: PROPERTIES 117

Example 3.2.1 We'll prove by induction that the sum of the rst n natural
numbers is n(n + 1)/2. We set

n
X n(n + 1)
P (n) : k= .
2
k=1

P (1) is the statement 1 = 1(1 + 1)/2, which is true. Now we assume the
truth of P (n) and use it to show the truth of P (n + 1):

n+1 n
X X n(n + 1) (n + 1)(n + 2)
k= k + (n + 1) = +n+1= .
2 2
k=1 k=1

This conrms P (n + 1). Hence the formula is conrmed for every n. 

Task 3.2.2 Use induction to prove that the sum of the squares of the rst
n natural numbers is n(n + 1)(2n + 1)/6.

Example 3.2.3 (Binomial formula) We'll use induction to prove that for
any real numbers a and b, and any natural number n,

n    
X n n n!
(a + b)n = ak bn−k , where = .
k k k!(n − k)!
k=0

We denote this formula by P (n). 3 While our formal description of the real
numbers happens later, in Ÿ3.6, in this example we only use their familiar
properties like commmutativity and associativity. We should also note that
the binomial formula assumes 00 = 1.

P (1) is the statement (a + b)1 = a + b, which is true. Now we assume the

3 Recall that n! is called the factorial of n and is dened recursively by 0! = 1 and


n! = (n − 1)! × n. See Exercise 11 of Ÿ3.1.
118 CHAPTER 3. NUMBER SYSTEMS

truth of P (n) and use it to show the truth of P (n + 1):


(a + b)n+1 = (a + b)n (a + b)
 
n  
X n
= ak bn−k  (a + b)
k
k=0
n   n  
X n k+1 (n+1)−(k+1)
X n
= a b + ak b(n+1)−k
k k
k=0 k=0
n+1
X  n  
n X n k (n+1)−k
= ak b(n+1)−k + a b
k−1 k
k=1 k=0
n
"   #
n+1
X n n
=a + + ak b(n+1)−k + bn+1
k−1 k
k=1
n  
X n + 1 k (n+1)−k
= an+1 + a b + bn+1
k
k=1
n+1
X n + 1

= ak b(n+1)−k .
k
k=0

This conrms P (n + 1). Hence the formula is conrmed for every n. 

Task 3.2.4 Suppose that P (k) is true for some k ∈ N, and (P (n) =⇒
P (n + 1)) is true for every n ≥ k . Show that P (n) is true for every n ≥ k.

Theorem 3.2.5 (Principle of Strong Induction) Suppose P (1) is true,


and for every n∈N the statement ((∀k ≤ n) P (k)) =⇒ P (n + 1)) is true.
Then P (n) is true for every n ∈ N.

Proof: Let P (n) be a property that may be associated ton. Suppose P (1)
is true, and for every n ∈ N ((∀k ≤ n) P (k)) =⇒ P (n + 1)) is true.
Dene a new property Q(n) = (∀k ≤ n) P (k). We apply the Principle of
Mathematical Induction to Q(n).
First, we see that Q(1) = P (1) is true.

Next, we assume that Q(n) is true. Then we have (∀k ≤ n) P (k). By the
assumptions about P, we nd that P (n + 1) is true. Therefore Q(n + 1) =
(∀k ≤ n + 1) P (k) is true.

So, by Principle of Mathematical Induction, Q(n) is true for all n ∈ N.

Undergraduate Mathematics: Foundations


3.2. NATURAL NUMBERS: PROPERTIES 119

Since Q(n) implies P (n), we see that P (n) is also true for all n ∈ N. 
The advantage of the Principle of Strong Induction is greater exibility
in carrying out proofs. It allows us to prove P (n + 1) from any convenient
prior case, or by their combinations, and not from P (n) alone. The very
next theorem illustrates this gain.

Theorem 3.2.6 (Well-Ordering Principle) Every non-empty subset of


N has a least element.

Proof: Every non-empty subset of N contains some natural number n. Hence


the Well-Ordering Principle will be proved if we can prove that each n∈N
has the following property:

P (n) : (n ∈ A) ∧ (A ⊆ N) =⇒ A has a least element.

P(1) is true because if 1∈A then 1 is the least element of A.


Suppose P (k) is true for each k ≤ n. A ⊆ N and n + 1 ∈ A. If A has
Let
no element which is less than or equal to n n + 1 is the least element
then
of A. On the other hand, if A has an element k which is less than or equal
to n then the truth of P (k) implies that A has a least element.

Hence P (n + 1) is true. So, by strong induction, P (n) is true for every


n. 
The Well-Ordering Principle gives easy justications of many of our basic
facts about natural numbers. We begin with division.

Theorem 3.2.7 Let a, b ∈ N. Then there are unique q, r ∈ W such that


b = aq+r and r < a. (We say that q is the quotient and r is the remainder
when we divide b by a.)

Proof: First, we prove existence. Let A = { n ∈ N | an > b }. A is non-


empty because b + 1 ∈ A. Hence A has a least member q 0 . Dene q = q 0 − 1.
Then aq ≤ b < a(q + 1). Dene r = b − aq . Then b = aq + r, and

aq ≤ b < a(q + 1) =⇒ 0 ≤ b − aq < a =⇒ 0 ≤ r < a.

Now we take up uniqueness. Suppose we also have b = aq 0 + r0 and r0 < a


0 0 0
with q > q . Subtracting the two expressions for b gives 0 = a(q−q )+(r−r ).
0 0 0 0
Hence r = r + a(q − q ) ≥ a, a contradiction. Therefore q = q and r = r .


120 CHAPTER 3. NUMBER SYSTEMS

If the remainder is zero when we divide b by a we say that `a divides b'


and we represent this by a | b, we say that a is a factor or
a | b. Moreover, if
divisor of b, while b is called a multiple of a. Recall that a prime number
is a natural number that is greater than 1 and whose only factors are 1 and
itself.

If a does not divide b we denote the situation by a - b.

Theorem 3.2.8 Suppose a∈N and a ≥ 2. Then there is a prime number


p such that p | a.

Proof: Let A = { n ∈ N | n | a } \ {1}. A 6= ∅ since a ∈ A. Let p be the


least element of A. You can complete the proof by checking that p is a prime
number. 

Theorem 3.2.9 Let a, b ∈ N. Then there is a number ` ∈ N that is a


multiple of both a and b and which divides every other common multiple of
a and b.

Proof: Let A = { n ∈ N | a | n and b | n } be the set of common multiples of


a and b. A 6= ∅ because ab ∈ A. Let ` be the least element of A. Now take
any n ∈ A. Then n = `q + r with 0 ≤ r < `.
Suppose r 6= 0. We have r = n − `q . Since a divides n and `, it divides
r. Similarly b | r. Then r ∈ A, contradicting the choice of ` as the least
element of A. So we have r = 0. 
The number ` in Theorem 3.2.9 is called the least common multiple
or lcm of a and b. It is denoted by lcm(a, b).

Theorem 3.2.10 Let a, b ∈ N. Then there is a number h ∈ N that is a


factor of both a and b and also is a multiple of every other common factor
of a and b.

Proof: Let ` = lcm(a, b). Then ` | ab implies there is h ∈ N such that


h` = ab.
The rst claim is that h is a common factor of a and b. Let am = `.
Then amh = ab, hence mh = b, hence h | b. Similarly h | a.
The second claim is that h is a multiple of every common factor of a and
b. Let k be a common factor of a and b. Let km = a and kn = b. Then

Undergraduate Mathematics: Foundations


3.2. NATURAL NUMBERS: PROPERTIES 121

kmb = ab = akn implies mb = na. Therefore `0 = mb is a common multiple


0 0 0
of a and b, and so ` | ` . We have k` = h` and ` | ` . Therefore k | h. 

The number h in Theorem 3.2.10 is called the highest common factor


or hcf of a and b. It is also called their greatest common denominator
or gcd. It is denoted by gcd(a, b) or just (a, b). If gcd(a, b) = 1 we say that
a and b are coprime numbers.
p is a prime number then the only possible values of gcd(a, p) are 1
If
and p. Now gcd(a, p) = p ⇐⇒ p | a. Hence we can say that if p - a then
gcd(a, p) = 1.
Note that the proof of Theorem 3.2.10 shows that lcm(a, b) • gcd(a, b) =
a • b.

Decimal Representation

Let us see how our abstract description of N matches with our everyday
use of these numbers, when we work with concrete numbers. How do we
represent an arbitrary member of N in the decimal notation? The key is the
Well-Ordering Principle.

We begin by setting names for the rst few natural numbers: 2 = 1 + 1,


3 = 2 + 1, 4 = 3 + 1, 5 = 4 + 1, 6 = 5 + 1, 7 = 6 + 1, 8 = 7 + 1, 9 = 8 + 1
and 10 = 9 + 1.

Fix a natural number M ∈ N. Our goal is to show thatM can be


PNk
expressed as a linear combination of powers of 10: k=0 mk 10 with
M =
mk ∈ {0, 1, . . . , 9} and mN > 0. The numbers mk will be the digits of the
decimal representation of M.
Let A = { n ∈ N | M < 10n }. First, we show M ∈A and so A is non-
empty:
M  
X M M −k k
10M = (9 + 1)M = 1 9 > 9M > M.
k
k=0
By the Well-Ordering Principle, A has a least element a. Let N = a − 1.
Then 10N ≤ M < 10N +1 . On dividing M by 10N we get

M = mN 10N + M1 , with 0 ≤ M1 < 10N and 1 ≤ mN ≤ 9.


Repeat this process, starting with M1 .

Task 3.2.11 Express the above discussion as a formal proof by mathemat-


ical induction.
122 CHAPTER 3. NUMBER SYSTEMS

Exercises for Ÿ3.2

1 Let A⊆N be bounded above. Show that A has a greatest element.

2 Let k, a, m ∈ N. Show that m is a multiple of ka if and only if m equals


k times a multiple of a.

3 Let k, a, b ∈ N. Show that lcm(ka, kb) = k • lcm(a, b) and gcd(ka, kb) =


k • gcd(a, b).

4 Let k, a, b ∈ N and gcd(k, a) = 1. Show that k | ab implies k | b.

5 Let a, b ∈ N and p be a prime number. Show that p | ab implies p|a or


p | b.

6 Let p be a prime number and a1 , . . . , an ∈ N. Show that p | (a1 • · · · • an )


implies p | ai for some i = 1, . . . , n.

7 Let a, b, k ∈ N and gcd(a, b) = 1. Show that if a|k and b|k then ab | k .

8 Let p be a prime number and a ∈ N such that p | a. Then there is a


unique natural number α such that pα | a but pα+1 - a.

9 Prove the following two statements, which constitute the Fundamental


Theorem of Arithmetic :

(a) Let p1 , . . . , pn be the distinct prime factors of a natural number a


(a ≥ 2), with p1 < p2 < · · · < pn . Then a is the product of powers of
α1 α2 α
these prime factors: a = p1 p2 · · · pn n , with αi ∈ N.

(b) The factorization given above is unique.

10 Let a, b ∈ N with a < b. Let b = aq + r with 0 ≤ r < a. Show that

(a) If r=0 then gcd(a, b) = a


(b) If r 6= 0 then gcd(a, b) = gcd(a, r)

11 Use the last exercise to create a process for calculating the gcd of a and
b without nding their factors.

12 Consider gcd as a binary operation on N. Show that it is commutative


and associative. Does it have an identity?

Undergraduate Mathematics: Foundations


3.3. INTEGERS 123

13 Consider a prime power pk , where p is a prime number and k ∈ N. Prove


k k
that the number of natural numbers n such that n ≤ p and gcd(n, p ) = 1,
is pk − pk−1 .

14 Show that the sum of two consecutive prime numbers (like 7 and 11)
cannot be twice of a prime number.

15 Let T : N → N such that T (m + n) = T (m) + T (n) for every m, n ∈ N.


Show that ∃k ∈ N such that T (n) = k · n for every n ∈ N.

16 Let T : N → N such that T (m + n) = T (m) · T (n) for every m, n ∈ N.


n
Show that ∃k ∈ N such that T (n) = k for every n ∈ N.

3.3 Integers
While discussing natural numbers in the previous section, we had a few oc-
casions when we incorporated zero into our discussion. This added exibility
4
and allowed some arguments to run in a smoother fashion with fewer cases.
Now we will take the next step, of incorporating the negative integers as well.
How were you taught this in school? You already knew the natural numbers
and zero, and were told to just put a `minus sign' in front of these numbers
to create the new `negative numbers'. The positive and negative integers
were visualized as being laid out along a line, extending in both directions
from 0.

−3 −2 −1 0 1 2 3
• • • • • • •

This approach can be formalized as follows:

Z = ({+, −} × N) ∪ {0}.

The diculty with dening integers as being made of two (in fact three)
intrinsically dierent types of objects is that all denitions and proofs take a
case-by-case form. And this is where it gets confusing. For a reminder, just
read the description of subtraction of integers given over a thousand years
ago by the Indian mathematician and astronomer Brahmagupta, and quoted
at the start of this chapter.

4 For this reason, many mathematicians prefer to include zero among the natural num-
bers.
124 CHAPTER 3. NUMBER SYSTEMS

We'll take a dierent tack. One motivation for creating negative numbers
is the desire to be able to carry out subtraction of arbitrary natural numbers
(If I have Rs 8 and I owe Rs 10, how much do I really have? Well, I need to
subtract 10 from 8, and so . . . ) To begin with, we can simply represent any
dierence by an ordered pair: for example (8, 10) can represent 8 − 10. Next
we observe that certain dierences should have the same result: for example
10 − 8 and 9−7 should be viewed as having the same value. Overall, this
gives an approach of viewing integers as ordered pairs with an equivalence
relation.

1. Dene a relation ∼ on N×N by (m, n) ∼ (i, j) if m + j = n + i.

2. Check that ∼ is an equivalence relation. ’


3. Dene Z to be the set of equivalence classes of ∼. Each equivalence
class is called an integer.

We'll use the notation (m, n) to represent the equivalence class of (m, n).
For example,

(1, 1) = { (i, j) ∈ N × N | i = j } = {(1, 1), (2, 2), (3, 3), . . . },


(5, 2) = { (i, j) ∈ N × N | i = 3 + j } = {(4, 1), (5, 2), (6, 3), . . . },
(2, 5) = { (i, j) ∈ N × N | i + 3 = j } = {(1, 4), (2, 5), (3, 6), . . . }.

The equivalence classes form diagonal rays in the N×N grid:

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

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

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

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

Addition and Multiplication

We dene a binary operation, called addition and denoted by ⊕, on Z by

(m, n) ⊕ (i, j) = (m + i, n + j).

Undergraduate Mathematics: Foundations


3.3. INTEGERS 125

We say we have dened a binary operation, but have we succeeded? The


problem here is that we are calculating something associated to two sets
« (the equivalence classes) by using a member from each. If our answer
depends on which members we use, we will have failed to give a valid
denition.

Task 3.3.1 Let (m, n) ∼ (m0 , n0 ) and (i, j) ∼ (i0 , j 0 ). Show that

(m + i, n + j) ∼ (m0 + i0 , n0 + j 0 ).

Theorem 3.3.2 Addition of integers has the following properties:

1. Addition is commutative and associative.

2. The element 0Z = (1, 1) serves as identity. It is called `zero'.

3. Every integer x has an `additive inverse' called y such that x + y = 0Z .


We denote y by −x.

Proof: We leave the proofs of commutativity, associativity, and identity, to


you. We just note that if x = (m, n) then y = (n, m) serves as its additive
inverse. 
Given x, y ∈ Z we dene their dierence by x y = x ⊕ (−y). The binary
operation is called subtraction.
We dene a binary operation, called multiplication and denoted by ⊗, on
Z by
(m, n) ⊗ (i, j) = (mi + nj, mj + ni).

Task 3.3.3 Let (m, n) ∼ (m0 , n0 ) and (i, j) ∼ (i0 , j 0 ). Show that

(mi + nj, mj + ni) ∼ (m0 i0 + n0 j 0 , m0 j 0 + n0 i0 ).

Theorem 3.3.4 Multiplication of integers has the following properties:

1. Multiplication is commutative and associative.

2. The element 1Z = (2, 1) serves as identity. It is called `one' or `unity'.

3. Let x, y ∈ Z. Then x ⊗ y = 0Z if and only if x = 0Z or y = 0Z .


4. Multiplication distributes over addition: If x, y, z ∈ Z then x⊗(y ⊕z) =
(x ⊗ y) ⊕ (x ⊗ z).
126 CHAPTER 3. NUMBER SYSTEMS

Proof: These are straightforward calculations which you can carry out. We
have done the third one as an illustration.

Letx = (m, n) and y = (i, j). Suppose x 6= 0Z , i.e., m 6= n. If m>n


then m = n + k and

x ⊗ y = 0Z ⇐⇒ (mi + nj, mj + ni) ∼ (1, 1) ⇐⇒ mi + nj = mj + ni


⇐⇒ n(i + j) + ki = n(i + j) + kj ⇐⇒ ki = kj
⇐⇒ i = j ⇐⇒ y = 0Z .

There is a similar calculation for the m<n case. 

Order

The next step is to supply Z with an order. First we identify integers that
we will call positive:
Z+ = { x ∈ Z | x = (m, n) such that n < m }.

The additive inverses of positive integers are called negative and their col-
lection is denoted by Z− . So,

Z− = { x ∈ Z | x = (m, n) such that m < n }.

Task 3.3.5 Prove the following:

(a) Z = Z+ ∪ {0Z } ∪ Z− is a disjoint union.


+
(b) Z is closed under ⊕ and ⊗.

Given x, y ∈ Z we say y < x (or x > y) if x y ∈ Z+ . And we say y≤x (or


x ≥ y ) if y < x or y = x.

Task 3.3.6 Prove the following:

(a) x > 0Z ⇐⇒ x ∈ Z+ and x < 0Z ⇐⇒ x ∈ Z− .


(b) ≤ is a total order on Z.
(c) ∀x, y, z ∈ Z, x < y ⇐⇒ x ⊕ z < y ⊕ z .
(d) ∀x, y, z ∈ Z, x ≤ y ⇐⇒ x ⊕ z ≤ y ⊕ z .
(e) ∀x, y ∈ Z and z ∈ Z+ , x < y ⇐⇒ x ⊗ z < y ⊗ z .

Undergraduate Mathematics: Foundations


3.3. INTEGERS 127

(f) ∀x, y ∈ Z and z ∈ Z+ , x ≤ y ⇐⇒ x ⊗ z ≤ y ⊗ z .


(g) ∀x, y ∈ Z and z ∈ Z− , x < y ⇐⇒ x ⊗ z > y ⊗ z .
(h) ∀x, y ∈ Z and z ∈ Z− , x ≤ y ⇐⇒ x ⊗ z ≥ y ⊗ z .

Z Enlarges N
Let us again look at the equivalence classes that constitute Z. In the gure
below, the ones in Z+ are coloured dark gray while the ones in Z− are
coloured light gray. The zero class 0Z is coloured black.

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

(2,1) (2,2) (2,3) (2,4) −3Z

(3,1) (3,2) (3,3) (3,4) −2Z

(4,1) (4,2) (4,3) (4,4) −1Z

3Z 2Z 1Z 0Z

The classes have been labeled as follows. We already know 0Z = (1, 1) and
1Z = (2, 1). We further dene

2Z = 1Z ⊕ 1Z = (2, 1) ⊕ (2, 1) = (2 + 2, 1 + 1) = (4, 2) = (3, 1)


3Z = 2Z ⊕ 1Z = (3, 1) ⊕ (2, 1) = (3 + 2, 1 + 1) = (5, 2) = (4, 1)

These observations motivate the following correspondence between N and


Z+ .

Theorem 3.3.7 The function ϕ : N → Z+ dened by ϕ(n) = (n + 1, 1) has


the following properties:

1. It is a bijection,

2. ϕ(1) = 1Z ,
3. ϕ(m + n) = ϕ(m) ⊕ ϕ(n),
4. ϕ(m · n) = ϕ(m) ⊗ ϕ(n).
128 CHAPTER 3. NUMBER SYSTEMS

Proof: Since n + 1 > 1, we have ϕ(n) ∈ Z+ .


ϕ is 1-1: ϕ(n) = ϕ(m) =⇒ (n + 1, 1) ∼ (m + 1, 1) =⇒ n + 2 =
m + 2 =⇒ n = m.
ϕ is onto: x ∈ Z+ =⇒ x = (m, n) with m > n =⇒ x = (n + k, n) =⇒
x = ϕ(k).
Thus ϕ is a bijection. The subsequent verications are again left for you.


+
Thus Z can be thought of as a copy of N inside Z and we shall do so. In
particular, the Principle of Mathematical Induction applies to Z+ as follows:
+
Let A ⊆ Z such that 1Z ∈ A and if x ∈ A then x ⊕ 1 Z ∈ A. Then
A = Z+ .
The correspondence between N and Z+ allows us to change our notation
to the familiar one. n ∈ N, we allow ourselves to also say n ∈ Z by
If
identifying n with (n + 1, 1). Similarly, we create the symbol −n and identify
it with (1, n + 1). 0 is identied with 0Z = (1, 1). And we replace ⊕, and
⊗ with +, − and · respectively.
With these identications, our calculations generate the familiar results.
For example,

3 − 7 = 3 + (−7) = (4, 1) ⊕ (1, 8) = (4 + 1, 1 + 8) = (5, 9) = (1, 5) = −4,


(−1) · 7 = (1, 2) ⊗ (8, 1) = (1 · 8 + 2 · 1, 1 · 1 + 2 · 8) = (10, 17) = (1, 8) = −7.

Division and GCD

We will now extend the concept of division and related results to integers.

Theorem 3.3.8 Let b ∈ Z and a ∈ N. Then there are unique integers q and
r such that b = qa + r and 0 ≤ r < a.

Proof: We have already proved this result for b > 0. If b = 0 we just take
q = r = 0. If b < 0, we rst apply the established case to −b: −b = q 0 a + r0
0
with 0 ≤ r < a.

If r0 = 0 we have b = −q 0 a with q = q0 and r = 0.


0 < r0 < a, we have b = −q 0 a−r0 = (−1−q 0 )a+(a−r0 ). Set q = −1−q 0
If
0 0 0
and r = a − r . Then b = qa + r and 0 < r < a =⇒ −a < −r < 0 =⇒
0
0 < a − r < a. 

Undergraduate Mathematics: Foundations


3.3. INTEGERS 129

We have the same notation as for N: If b = qa we say that a divides b,


or a is a factor of b, or b is a multiple of a, and we denote the fact by a | b.
The negation of a | b is a - b.

Task 3.3.9 Every integer is a factor of 0. If 0 is a factor of an integer n


then n = 0.

Let a, b ∈ Z. An integer g is called their greatest common divisor or


gcd and denoted by gcd(a, b), if
1. g|a and g | b,
2. If h|a and h|b then h | g,
3. g ≥ 0.
For example, consider 4 and −6. Their common factors are 1, −1, 2, −2. Of
these, only 2 satises all three requirements, and hence 2 = gcd(4, −6).

Task 3.3.10 Show the following:


5

(a) gcd(0, 0) = 0, (b) If a 6= 0 then gcd(a, 0) = a.

Theorem 3.3.11 Let a, b ∈ Z. Then they have a unique gcd and it can be
expressed in the form ma + nb with m, n ∈ Z.

Proof: If a = b = 0 we have gcd(a, b) = 0 = 1 · a + 1 · b. So assume that at


least one of a, b is non-zero.
A = { x ∈ Z+ | x = ma + nb, m ∈ Z, n ∈ Z }. A 6= ∅ since a2 +b2 ∈
Let
A. Let g be the least element of A. On dividing a by g we get a = qg + r,
0 ≤ r < g . If r > 0 then a − qg ∈ A and contradicts the minimality of g ,
Hence r = 0 and g | a. Similarly, g | b.

If h divides a and b then it divides any ma + nb and so divides g. 

Theorem 3.3.12 Leta, b ∈ Z. Then gcd(a, b) = 1 if and only if there exist


m, n ∈ Z such that ma + nb = 1.
5 An alternative denition of gcd(a, b) is that it is the greatest amongst the common di-
visors of a and b. The existence and non-negativity of the gcd are immediate consequences.
The divisibility property can be be obtained via the division algorithm. However, gcd(0, 0)
becomes undened. This is the only dierence, and it is inconsequential.
130 CHAPTER 3. NUMBER SYSTEMS

Proof: If gcd(a, b) = 1 then the previous theorem gives the existence of m


and n.
Conversely, suppose ma + nb = 1 for some m, n ∈ Z. Let d = gcd(a, b).
Then, since d|a and d | b, we have d | ma + nb = 1 and hence d = 1. 

Absolute Value


n n≥0
The absolute value of an integer n is dened by |n| = −n
if
if n<0
.

Task 3.3.13 Let m, n ∈ Z. Show that:

(a) |m| ≥ 0, (b) |m| = 0 ⇐⇒ m = 0,

(c) −|m| ≤ m ≤ |m|, (d) |mn| = |m| |n|,

(e) |m + n| ≤ |m| + |n|, (f) |m − n| ≥ |m| − |n|.

Exercises for Ÿ3.3

In Exercises 1 to 3, use the description of Z as equivalence classes of N×N


with the operations ⊕ and ⊗.

1 Let x ∈ Z. Show that

(a) 0Z ⊗ x = 0Z , (b) (−1Z ) ⊗ x = −x.

2 Let x, y ∈ Z. Use the notation x2 = x ⊗ x. Show that

2 2
(a) x ≥ 0, and x = 0 ⇐⇒ x = 0.
2 2
(b) x ⊕ y ≥ 0, and x2 ⊕ y 2 = 0 ⇐⇒ x = y = 0.
(c) x ∈ Z+ and y ∈ Z− implies x ⊗ y ∈ Z− .
(d) x, y ∈ Z− implies x ⊗ y ∈ Z+ .

3 Which elements of Z have multiplicative inverses? (That is, for which


x∈Z is there an element y∈Z such that x ⊗ y = 1Z ?)

In Exercise 4 and after, use the description of Z with Z+ identied with N


and with the operations + and ·.

Undergraduate Mathematics: Foundations


3.4. RATIONAL NUMBERS 131

4 Let n ∈ N. Use mathematical induction to prove that (−1)n = 1 if n is


n
even and (−1) = −1 if n is odd.

5 Let T : Z → Z such that T (m + n) = T (m) + T (n) for every m, n ∈ Z.


Show that ∃k ∈ Z such that T (n) = kn for every n ∈ Z.

6 Leta, b ∈ Z. Use mathematical induction to prove that an − bn is divisible


by a − b for every n ∈ N.

7 Prove the following version of the division algorithm: Let a, b ∈ Z with


a 6= 0. Then there are unique integers q and r such that b = qa + r and
0 ≤ r < |a|.

8 Let m, a, b ∈ Z. Prove that gcd(ma, mb) = |m| gcd(a, b).

9 Let m, a, b ∈ Z and gcd(m, a) = 1. Prove that if m | ab then m | b.

10 Let I⊆Z have the following properties:

(a) a, b ∈ I =⇒ a + b ∈ I , (b) a ∈ I =⇒ −a ∈ I .

Show that I = nZ = { na | a ∈ Z } for some n ∈ Z.

11 Consider Zn , the integers modulo n. (See page 103) For any number
k ∈ {1, 2, . . . , n − 1} dene hki = { mk | m ∈ N }. Show that hki = Zn if and
only if gcd(k, n) = 1.

3.4 Rational Numbers


As we enlarged the system of natural numbers to the integers, so we shall
now enlarge the integers to the rational numbers. In the earlier episode
the desire was to allow unrestricted subtraction and so we took pairs of
natural numbers and thought of each pair as representing the dierence of
its members. In the same manner we shall now take pairs of integers and
think of each pair as representing the ratio of its members.

What do we have to be careful about? First, the member that repre-


sents the denominator must be non-zero. Second, two pairs may represent
the same ratio, and this creates an equivalence relation. Thus we have the
following steps:
132 CHAPTER 3. NUMBER SYSTEMS

1. Let X = Z × Z∗ (recall that Z∗ = Z \ {0}).


2. Dene a relation on X by (m, n) ∼ (i, j) if mj = ni.
3. Check that ∼ is an equivalence relation. ’
4. Let Q be the set of equivalence classes of ∼. The members of Q are
called rational numbers or just rationals.
m
We'll denote the equivalence class of (m, n) by m/n or . Here are some
n
examples of equivalent members of Z × Z∗ :

(0, 1) ∼ (0, 2), (2, 1) ∼ (6, 3), (3, −2) ∼ (−3, 2), (−3, −2) ∼ (3, 2).

They lead to the following equalities of rational numbers:

0 0 2 6 3 −3 −3 3
= , = , = , = .
1 2 1 3 −2 2 −2 2

Task 3.4.1 Show that every rational number can be expressed in the form
m/n where m∈Z and n ∈ N.

Task 3.4.2 Show that every rational number can be expressed in the form
m/n where gcd(m, n) = 1. (This is called a reduced form of the rational
number. How many reduced forms can a rational have?)

Addition and Multiplication

Given two rational numbers m/n and i/j we dene their sum by
m i mj + ni
+ = .
n j nj

As we are working with equivalence classes we have to make sure our deni-
tion is independent of the choice of representatives.

Task 3.4.3 (m, n) ∼ (m0 , n0 )


Suppose that and (i, j) ∼ (i0 , j 0 ). Show that
(mj + ni, nj) ∼ (m j + n i , n0 j 0 ).
0 0 0 0

We have created a binary operation + : Q → Q, called addition. Its prop-


erties are derived below:

Theorem 3.4.4 Addition of rational numbers has the following properties:

Undergraduate Mathematics: Foundations


3.4. RATIONAL NUMBERS 133

1. Addition is commutative and associative.

2. The element 0Q = 0/1 serves as identity. It is called zero.

3. Every rational x has an additive inverse called y such that x+y = 0Q .


We denote y by−x.

m −m
Proof: We leave the proofs to you. We just note that if x= then y=
n n
m −m m
serves as its additive inverse. That is, − = = . 
n n −n

We dene the dierence of x, y ∈ Q by x − y = x + (−y). The binary


operation − subtraction.
is called

We dene a binary operation, called multiplication and denoted by , •

on Q by
m i mi
• = .
n j nj
If x, y ∈ Q then x•y is called the product of x and y.
Task 3.4.5 Let (m, n) ∼ (m0 , n0 ) and (i, j) ∼ (i0 , j 0 ) in Z × Z∗ . Show that

(mi, nj) ∼ (m0 i0 , n0 j 0 ).

We use the notation Q∗ = Q \ {0Q }.

Theorem 3.4.6 Multiplication of rationals has the following properties:

1. Multiplication is commutative and associative.

2. The element 1Q = 1/1 serves as identity. It is called one or unity.


3. Let x∈Q . Then ∃y ∈ Q∗ such that x • y = 1Q . The rational number
y is called the multiplicative inverse of x and is denoted by x−1 or
1/x.
4. Multiplication distributes over addition: If x, y, z ∈ Q then x • (y + z) =
(x • y) + (x • z).

Proof: You must have predicted that this proof would be left to you! We

just mention, for the record, a fact already known to you: If x = m/n ∈ Q
then x−1 = n/m. 
134 CHAPTER 3. NUMBER SYSTEMS

Task 3.4.7 Let x, y ∈ Q. Show that:

(a) 0Q · x = 0,
(b) x • y = 0Q if and only if x = 0Q or y = 0Q ,
(c) (−1Q ) • x = −x.

Suppose x ∈ Q∗ and y ∈ Q. Then there is a unique z ∈ Q such that


−1
z x = y . It
• is z = y • x . We call z the quotient of y by x and denote it by
y
y/x or .
x

m i y in
Task 3.4.8 Let x= ∈ Q∗ and y= ∈ Q. Show that = .
n j x jm

The notation x•y for multiplication will usually be abbreviated to xy .

Q Enlarges Z

Theorem 3.4.9 Dene ψ: Z → Q by ψ(n) = n/1. Then,

1. ψ is 1-1,

2. ψ(0) = 0Q and ψ(1) = 1Q ,

3. For every m, n ∈ Z, ψ(m + n) = ψ(m) + ψ(n),

4. For every m, n ∈ Z, ψ(mn) = ψ(m)ψ(n).

Proof: ’

Therefore we can think of ψ(Z) as a copy of Z inside Q, and thus Q
becomes an enlargement of Z. We start directly combining integers and
rationals through the following conventions, which are based on ψ:

k n k k n k
For n, k ∈ Z, ` ∈ Z∗ , n+ = + and n• = • .
` 1 ` ` 1 `

We also start writing 0Q and 1Q as simply 0 and 1 respectively.

Undergraduate Mathematics: Foundations


3.4. RATIONAL NUMBERS 135

Exponentiation

On page 112 we saw how to dene an , for n ∈ W, with respect to any binary
operation with identity. In particular, this leads to a denition of an for
multiplication of rational numbers, with a ∈ Q and n ∈ W. Since every
non-zero rational number has a multiplicative inverse, we can extend this
denition to negative powers of non-zero rational numbers:

If a ∈ Q∗ and n ∈ Z− then we dene an = (a−1 )−n .


 −3  −1 !3  3
2 2 3 33
For example, = = = 3.
3 3 2 2

Task 3.4.10 Let a ∈ Q∗ and n ∈ Z. Show that a−n = (a−1 )n = (an )−1 .

Theorem 3.4.11 Let a, b ∈ Q∗ and m, n ∈ Z. We have the following:

(a) am+n = am an , (b) (am )n = amn , (c) (ab)n = an bn .

Proof:

(a) We note that we already have the result when m, n ∈ W. The other
cases are:

• m, n ∈ Z− :

am+n = (a−1 )−m−n = (a−1 )−m (a−1 )−n = am an .

• m ∈ W, n ∈ Z− , m + n ∈ W:

am+n a−n = am =⇒ am+n = am an .

• m ∈ W, n ∈ Z− , m + n ∈ Z− :

am a−(m+n) = a−n =⇒ am = am+n a−n =⇒ am an = am+n .

(b) ’
(c) ’

136 CHAPTER 3. NUMBER SYSTEMS

Order

To introduce an order on Q we follow the same process as for Z. The rst


step is to decide which rationals will be called positive.

A rational number is called positive if it can be expressed as m/n with


m, n ∈ N. The set of positive rationals is denoted by Q+ . The additive in-
verse of a positive rational is called negative and the set of negative rationals
is denoted by Q− .

Theorem 3.4.12 The positive and negative rationals have the following
properties:

1. Q = Q+ ∪ {0} ∪ Q− is a disjoint union.

+
2. Q is closed under addition and multiplication.

Proof: ’

Given x, y ∈ Q we say y < x (or x > y) if x − y ∈ Q+ . And we say y≤x
(or x ≥ y ) if y < x or y = x.

Task 3.4.13 Let x, y ∈ Q. Express them as x = m/n and y = i/j with


m, i ∈ Z and n, j ∈ N. Show that

y < x ⇐⇒ ni < mj.

Task 3.4.14 Prove the following:

(a) x > 0 ⇐⇒ x ∈ Q+ ,
(b) x < 0 ⇐⇒ x ∈ Q− ,
(c) ≤ is a total order on Q.

Exercises for Ÿ3.4

1 Show that there is no rational number whose square is 2.

2 Show that there is no rational number whose square is 6.

3 Show that there is no rational number whose cube is 2.

4 Prove the following for every x, y, z ∈ Q:

Undergraduate Mathematics: Foundations


3.4. RATIONAL NUMBERS 137

(a) x < y ⇐⇒ x + z < y + z , (b) x ≤ y ⇐⇒ x + z ≤ y + z .

5 Let x ∈ Q. Prove that:

(a) x > 0 ⇐⇒ x−1 > 0, (b) x < 0 ⇐⇒ x−1 < 0.

6 Let x, y ∈ Q. Prove that:

(a) x>0 and y < 0 =⇒ xy < 0,


(b) x<0 and y < 0 =⇒ xy > 0.

7 Prove the following for every x, y, z ∈ Q:


+
(a) If z∈Q then x < y ⇐⇒ xz < yz ,
+
(b) If z∈Q then x ≤ y ⇐⇒ xz ≤ yz ,

(c) If z∈Q then x < y ⇐⇒ xz > yz ,

(d) If z∈Q then x ≤ y ⇐⇒ xz ≥ yz .

8 Is the following statement correct?

m i
If x= and y= with m, i ∈ Z and n, j ∈ Z∗ then y < x ⇐⇒ ni < mj .
n j

9 Prove the following (which are collectively known as the Archimedean


property of Q)
(a) N has no upper bound in Q.
(b) Let x, y ∈ Q with x > 0. Then ∃n ∈ N such that nx > y .
x
(c) Let x, y ∈ Q+ . Then ∃n ∈ N such that < y.
n

10 Let p and q be rational numbers with p < q. Dene the interval (p, q)Q
by
(p, q)Q = { x ∈ Q | p < x < q }.
(a) Let F be the set whose members are the intervals
T (0, 1/n)Q with
n ∈ N. Show that F = ∅.
(b) Let G be the set whose
S members are the intervals (0, 1 − 1/n)Q with
n ∈ N. Show that G = (0, 1)Q .

11 Let T : Q → Q such that T (x + y) = T (x) + T (y) for every x, y ∈ Q.


Show that ∃k ∈ Q such that T (x) = kx for every x ∈ Q.
138 CHAPTER 3. NUMBER SYSTEMS

12 Let E : Z → Q such that E(m + n) = E(m)E(n) for every m, n ∈ Z and


E(0) = 1. Show that ∃k ∈ Z such that E(n) = k n for every n ∈ Z.

13 Can there be a function E: Q → Q such that E(x + y) = E(x)E(y) for


every x, y ∈ Q and E(1) 6= 0?

3.5 Integral Domains and Fields


In this section we shall abstract formal denitions out of our various exam-
ples. First we note that all our versions of number have two operations
with certain properties. We rst take up properties that are common to Z
and Q.
An integral domain is a set R with two binary operations + (addition)
and • (multiplication) such that

1. Addition and multiplication are commutative: If a, b ∈ R then a+b =


b+a and a • b = b • a.
2. Both operations are associative: If a, b, c ∈ R then (a+b)+c = a+(b+c)
and a • (b • c) = (a • b) • c.
3. There is an additive identity 0: a+0=a for every a ∈ R.
4. Every a∈R has an additive inverse b such that a + b = 0. We denote
b by −a.
5. There is a multiplicative identity 1 (1 6= 0): 1 • a = a for every a ∈ R.
6. Multiplication distributes over addition: If a, b, c ∈ R then a • (b + c) =
a • b + a • c.
7. There are no `zero divisors': If a, b ∈ R and a•b = 0 then a=0 or
b = 0.
We usually shorten a•b to ab.

Example 3.5.1 N is not an integral domain as it has no additive identity. Z


and Q are integral domains, for we have checked all the required properties.

Example 3.5.2 Let us see if the integers modulo n, denoted by Zn , are an


integral domain. We already know that addition and multiplication are com-
mutative and associative, and that multiplication distributes over addition.

Undergraduate Mathematics: Foundations


3.5. INTEGRAL DOMAINS AND FIELDS 139

The equivalence classes of 0 and 1 serve as the additive and multiplicative


identities, respectively. Take any member k of Zn . Then n−k is its additive
inverse. So the only issue is the presence or absence of zero divisors, that is,
of non-zero elements whose product is zero.

Let us rst take the example of Z2 = {0, 1}. The only non-zero member
is the multiplicative identity 1. Now 1 • 1 = 1 shows that Z2 is an integral
domain.

Next we consider Z3 = {0, 1, 2}. We have 1 • 1 = 1, 1 • 2 = 2 and 2 • 2 = 1.


So there are no zero divisors and Z3 is an integral domain.

The news is not always good. In Z4 = {0, 1, 2, 3} we do have zero divisors,


since 2 • 2 = 0.
A natural question to ask is which Zn are integral domains and which are
not. We leave this for the Exercises. For the present, it is enough to have
seen one example of the presence of zero divisors, and hence of the absence
of the cancellation law for multiplication. 

A eld is an integral domain F such that every non-zero a ∈ F has a


−1 1
multiplicative inverse b: ab = 1. We denote b by a or 1/a or . We also
a
x
denote xy −1 by x/y or .
y
The set of non-zero elements of a eld F is denoted by F∗ .

Example 3.5.3 N is not an integral domain, hence not a eld. Z is an


integral domain but it is not a eld as the only non-zero members that have
multiplicative inverse are 1 and −1. Q is a eld, as we have checked all the
required properties earlier. 

Task 3.5.4 Verify that Z2 and Z3 are elds.

From the original denition of a eld, we can deduce many other prop-
erties that it automatically possesses:

Theorem 3.5.5 Let F be a eld.

1. F has only one additive identity.

2. F has only one multiplicative identity.

3. (Cancellation Law for Addition) a+b=a+c implies b = c.


140 CHAPTER 3. NUMBER SYSTEMS

4. (Cancellation Law for Multiplication) ab = ac and a 6= 0 implies b = c.


5. 0a = 0 for every a ∈ F.
6. If ab = 0 then a=0 or b = 0.
7. −(−a) = a.
8. (−a)b = −(ab). (In particular, (−1)b = −b.)
9. (−a)(−b) = ab. (In particular, (−a)2 = a2 .)
10. 0 has no multiplicative inverse.

Proof:

1. If 00 is also an additive identity, then 0 + 00 = 0 and also 0 + 00 = 00 ,


0
hence 0=0.
2. Similar.

3. a + b = a + c =⇒ (−a) + a + b = (−a) + a + c =⇒ 0 + b = 0 + c =⇒
b = c.
4. Similar.

5. 0a = (0 + 0)a = 0a + 0a.
6. Suppose a 6= 0. Then a−1 ab = a−1 0 = 0 =⇒ 1b = 0 =⇒ b = 0.
7. a + (−a) = 0 implies a is additive inverse of −a, implies a = −(−a).
8. ab + (−a)b = (a + (−a))b = 0b = 0.
9. Apply previous result.

10. If 0b = 1 then 0 = 1. 
Since Q is a eld, it has all the properties listed above. But there is more
to Q. Not only do we have rules for combining rational numbers, we also
have a notion of comparing the size of two rational numbers. We know that
with respect to this notion of order, Q splits into three types of numbers:
0, positive and negative. We know that the sum and product of positive
numbers are positive, and that x is positive if and only if −x is negative.

If A is a subset of a eld F then −A is the set { x ∈ F | −x ∈ A } consisting


of the additive inverses of the members of A.
A eld F is called an ordered eld if it has a subset F+ such that

Undergraduate Mathematics: Foundations


3.5. INTEGRAL DOMAINS AND FIELDS 141

1. F = F+ ∪ {−F+ } ∪ {0} is a disjoint union.

2. F+ is closed under addition and multiplication.

Members of F+ are called positive and members of −F+ are called nega-
tive. The
+ −
set −F is also denoted by F .

If F is an ordered eld and x, y ∈ F we say x<y if y − x ∈ F+ .

Task 3.5.6 c ∈ F+ i 0 < c.

We have already seen that Q is an ordered eld.

Task 3.5.7 Verify that there is no way to make Z2 into an ordered eld.

The order relation < has the following properties:

Theorem 3.5.8 Let F be an ordered eld with order relation <. Then we
have

1. (Trichotomy) Given any x, y ∈ F exactly one of the following holds:


x < y, y < x and x = y.
2. (Transitivity) If x<y and y<z then x < z.
3. If a<b then a + c < b + c.
4. If a<b and 0<c then ac < bc.

Proof:

1. The condition x < y is equivalent to y − x ∈ F+ . The condition y<x


+
is equivalent to y − x ∈ −F . The condition x = y is equivalent to
y − x ∈ {0}. Therefore, by the rst requirement in the denition of an
ordered eld, exactly one of these three conditions will be satised.

2. First, x<y implies y − x ∈ F+ . Second, y <z implies z − y ∈ F+ .


Therefore

z − x = (z − y) + (y − x) ∈ F+ .

3. If a<b then b − a ∈ F+ . Hence (b + c) − (a + c) = b − a ∈ F+ .


4. We have b − a ∈ F+ and c ∈ F+ . Hence bc − ac = (b − a)c ∈ F+ . 
142 CHAPTER 3. NUMBER SYSTEMS

We also use the following notation:

x>y if y<x
x≤y if x<y or x=y
x≥y if x>y or x=y

The four basic properties of < can be used to prove other important prop-
erties:

Theorem 3.5.9 Let F be an ordered eld with order relation <. Then we
have

1. If x≤y and x≥y then x = y.


2. If x 6= 0 then x2 > 0.
3. 1 > 0.
4. If x<y then −y < −x.
5. If x<y and c<0 then cx > cy .
6. If x<0 and y<0 then xy > 0.
7. If xy > 0 then x, y are both positive or both negative.

8. If x<y and a<b then x + a < y + b.

Proof:

1. Proof by contradiction. Suppose x 6= y . Then

(x ≤ y ∧ y ≤ x) =⇒ (x < y ∧ y < x).

This contradicts Trichotomy, which says that x<y and y<x cannot
be true simultaneously. Hence we must have x = y.
2. If x ∈ F+ then x2 ∈ F+ by denition of ordered eld. If x ∈ −F+ then
−x ∈ F+ and x2 = (−x)2 ∈ F+ .
3. 1 = 1 2 > 0, by previous result since 1 6= 0..
4. If y−x>0 then (−y) − (−x) = x − y < 0.
5. Apply previous result.

6. −x > 0 and −y > 0 and xy = (−x)(−y).

Undergraduate Mathematics: Foundations


3.5. INTEGRAL DOMAINS AND FIELDS 143

7. Suppose x>0 and y < 0. then −y > 0. Hence −xy = x(−y) > 0, and
so xy < 0.
8. Sum of positives is positive. 

Task 3.5.10 Let F be an ordered eld, x, y ∈ F and 0 < x < y. Show that:

(a) 0 < y −1 < x−1 , (b) 0 < x2 < y 2 .

Theorem 3.5.11 Every ordered eld F contains a copy of N.

Proof: We dene a function ϕ: N → F by recursion as follows:

• ϕ(1) = 1, • ϕ(n + 1) = ϕ(n) + 1.

We can prove the following by induction on n:


1. ϕ(n) ∈ F+ , ’
2. ϕ(n + k) = ϕ(n) + ϕ(k), ’
3. ϕ(nk) = ϕ(n)ϕ(k). ’
It follows from the rst two assertions above that m < n implies ϕ(m) <
ϕ(n). Therefore ϕ is 1-1. Thus ϕ is a bijection of N with ϕ(N) which
preserves unity, addition, multiplication, and order. So ϕ(N) is a copy of N
inside F. 
One consequence is that every ordered eld is innite.

Exercises for Ÿ3.5

1 Show that the cancellation laws hold in an integral domain:

a + b = a + c =⇒ b = c, ab = ac and a 6= 0 =⇒ b = c.

2 Prove that Zn is an integral domain if and only if n is prime.

3 Prove that every nite integral domain is a eld. (Hint: Can the powers an
all be distinct for a xed non-zero a and with n varying over N?)

4 Prove that Zn is a eld if and only if n is prime.


144 CHAPTER 3. NUMBER SYSTEMS

5 Let F be an ordered eld. Dene 2 = 1 + 1. Show that 2 ∈ F∗ and if x<y


x+y
then x< < y.
2

6 Let F be a eld with a strict total order < such that

(a) If a<b then a + c < b + c,


(b) If 0<a and 0<b then 0 < ab.
Prove that choosing F+ = { a ∈ F | a > 0 } makes F an ordered eld. (This
is the converse of Theorem 3.5.8)

7 Let F be an ordered eld, a, x ∈ F.


a2
(a) Show that x > 0 =⇒ x + ≥ 2a.
x
1 1
(b) Show that 1 ≤ x ≤ a =⇒ x + ≤ a + .
x a

Before trying out Exercises 8 to 12, you may want to review the material on bounds
in pages 6366.

8 Let F be an ordered eld and A = { x ∈ F | x < 1 }. Show that A has no


maximum, but sup(A) = 1.

9 Let F be an ordered eld and A ⊆ F. Let −A = { y ∈ F | −y ∈ A }. Show


that

(a) A has an upper bound if and only if −A has a lower bound.

(b) A has a maximum if and only if −A has a minimum, and then


− max(A) = min(−A).
(c) A has a supremum if and only if −A has an inmum, and then
− sup(A) = inf(−A).

10 Let F be an ordered eld, A ⊆ F, and c ∈ F. Dene

cA = { y ∈ F | y = cx for some x ∈ A }.

Suppose sup(A) exists. Show that

(a) c > 0 =⇒ sup(cA) = c sup(A),


(b) c < 0 =⇒ inf(cA) = c sup(A).

Undergraduate Mathematics: Foundations


3.6. REAL NUMBERS 145

11 Let F be an ordered eld and A, B ⊆ F. Dene

A + B = {y ∈ F | y = a + b for some a∈A and b ∈ B }.

Suppose sup(A) and sup(B) exist. Show that sup(A+B) = sup(A)+sup(B).


Also formulate and prove the analogous result for inmum.

12 Let F be an ordered eld and A, B ⊆ F. Dene

AB = { y ∈ F | y = ab for some a∈A and b ∈ B }.

Suppose sup(A) and sup(B) exist. Will sup(AB) = sup(A) sup(B) always
hold? Will it hold if A, B ⊆ F+ ?

13 Show that every ordered eld F contains a copy of Z.

14 Show that every ordered eld F contains a copy of Q.

3.6 Real Numbers


The natural numbers capture our intuition about counting. The integers and
rational numbers make incremental technical improvements, and allow us to
subtract and divide without restrictions. Most of the things we do with
integers and rational numbers in daily life, we could also do with natural
numbers alone but the calculations would be lengthier and more awkward.
For example, in the world of integers we go from x + 7 = 2x + 1 to x = 2x − 6
without a second thought, but in the world of natural numbers alone we
would have to worry about whether it is possible that 2x ≤ 6.
At rst glance, the rational numbers also seem to suce for measuring
continuously varying quantities like length. On the one hand, they are un-
bounded and so can reach out to very large values. On the other hand, by
using large denominators we can give as ne a resolution as desired to our
measurements and calculations. For example, when Aryabhata wanted to
describe π to a high level of accuracy, he expressed his result as follows:

Add 4 to 100, multiply by 8, and add 62,000. The result is ap-


proximately the circumference of a circle of which the diameter
6
is 20,000.
6 In the Aryabhatiya, composed around 500 AD. This translation by Walter Eugene
Clark, 1930. (https://archive.org/details/The_Aryabhatiya_of_Aryabhata_Clark_
1930/)
146 CHAPTER 3. NUMBER SYSTEMS

This amounts to the value 62, 832/20, 000 = 3.1416. The precision to 4
decimal places is enabled by the large diameter of 20,000.

Example 3.6.1 We describe an ancient technique for approximating the


square root of 2 by rational numbers. It is named Hero's method after the
Greek mathematician who described it in the 1st century AD. But values
that could have been obtained from it are also found much earlier from
Mesopotamia and India.

The original approach was based on geometry, on manipulating areas of


squares and rectangles. A briefer algebraic account is as follows. Take any
rational number a1 > 0, with a21 6= 2. Dene b1 = 2/a1 , so that 2 lies
2 2
between a1 and b1 . Then the midpoint a2 = (a1 + b1 )/2 lies between a1 and
b1 , and a22 should be a better approximation to 2 than the squares of a1 and
b1 . Repeat the process with a2 to get a3 and so on.

For example, take a1 = 1. Then:

1+2 3
a1 = 1 =⇒ b1 = 2 =⇒ a2 = = ,
2 2
3 4 3/2 + 4/3 17
a2 = =⇒ b2 = =⇒ a3 = = ,
2 3 2 12
17 24 17/12 + 24/17 577
a3 = =⇒ b3 = =⇒ a4 = = .
12 17 2 408
The value 577/408 ≈ 1.414216 is accurate to ve decimal places! It is found
in a clay tablet from Mesopotamia from about 1800 BC. It is also found in
the Indian Sulvasutra texts from 800BC or later, where it is expressed in a
particularly convenient form:

1 1 1
1+ + + .
3 3 × 4 3 × 4 × 34
(And the rst three terms give the previous approximation of 17/12) 

However, we run into trouble when we want to capture certain quantities


exactly and not just accurately. The well-known example of this is the length
of the diagonal of a unit square. By the Pythagoras theorem, the square of
this length is 2, but then there is no rational number whose square is 2. Thus
there are lengths that cannot be represented exactly by rational numbers.
The rational numbers, too, need to be enlarged.

From the modern viewpoint, we are seeking an ordered eld (which au-
tomatically contains Q) that also suces for geometry. If a quantity can

Undergraduate Mathematics: Foundations


3.6. REAL NUMBERS 147

be represented by rationals to arbitrary precision then we should be able to


represent it exactly by a member of our expanded eld.

The next example provides insight into what it means for some quantities
to be beyond Q and the additional requirement that would bring them into
the fold. This step was rst taken by the German mathematician Richard
Dedekind.

Example 3.6.2 F be an ordered eld and dene A = {x ∈ F : x2 ≤ 2}.


Let
A is non-empty as 1 ∈ A. It is also bounded above by 2. For if x > 2 then
x2 > 4 > 2. Since A is bounded above, we can ask if it has a least upper
2
bound or supremum. We'll show that if A has a supremum c then c = 2.
Thus the issue of whether 2 has a square root in F is linked to the issue of
whether a certain set has a least upper bound.

So let c = sup(A). We have already seen that 1 ≤ c ≤ 2. In particular,


c ∈ F+ .
First we rule out c2 > 2:
Suppose c2 > 2 . We apply Hero's method to c. First, b = 2/c satises
b+c
0 < b < c. Consider d= . Then b < d < c. Now, using c2 > 2 and
2
Exercise 7 of Ÿ3.5:

1 2 1 c2
d2 = (b + c2 + 4) = 2 + + 1 > 2.
4 c 4
Hence d is an upper bound of A. But then c is not the least upper bound, a
contradiction.

Next we rule out c2 < 2:


Suppose c2 < 2. Then b = 2/c satises 0 < c < b. The d we used in the
rst part of the proof won't work here since it again satises d2 > 2. Instead,
we consider a point which is closer to c:
1 2 1
e = c + (b − c) = c+ .
3 3 c
Then, using 1 ≤ c2 < 2 and Exercise 7 of Ÿ3.5:

4 1  4 1 
e2 = c2 + + 2 ≤ 2 + + 2 = 2.
9 c2 9 2
Hence e∈A and c is not an upper bound of A. Since c2 > 2 and c2 < 2
2
have been ruled out, the only possibility is c = 2. 
148 CHAPTER 3. NUMBER SYSTEMS

We say that an ordered eld F is order complete or has the lub


property if every non-empty subset of F that is bounded above has a least
upper bound.

So far, the only ordered eld we have seen is Q and it is not order com-
plete, since the set {x ∈ Q : x2 < 2} is bounded above by 2 but has no
supremum.

Theorem 3.6.3 There exists an order complete eld. It is unique up to


renaming of its elements.

The unique order complete eld described above is called the eld of
real numbers and is denoted by R. We will take this result for granted .
Assuming it, we will try to prove all properties of real numbers without any
7
further assumptions.

Task 3.6.4 Show that R contains a unique positive number c such that
c2 = 2.

If x, y ∈ R and x2 = y , we say that x is a square root of y . If x is a


square root of y, then−x is also a square root of y . The argument for 2 can
be adjusted to show that every positive real number has a unique positive
square root in R. We have left this generalization to the exercises. If y ∈ R+


then its unique positive square root is denoted by y. We also write 0=0
since 02 = 0.

Task 3.6.5 Let A ⊆ R be non-empty and bounded below. Show that inf(A)
exists.

Archimedean Property

Theorem 3.6.6 (Archimedean Property of R, ver.1) N is unbounded


in R.

Proof: Suppose N is bounded. Let α = sup(N). Then α − 1 is not an upper


bound of N. So there is an N ∈ N such that N > α − 1. But then N + 1 ∈ N
and N + 1 > α, contradicting α being an upper bound of N. 

Task 3.6.7 Z has neither an upper nor a lower bound in R.


7 We describe the main points of Dedekind's construction in ŸA.3.

Undergraduate Mathematics: Foundations


3.6. REAL NUMBERS 149

Theorem 3.6.8 (Archimedean Property of R, ver.2) Let x, y ∈ R+ .


Then there exists N ∈N such that N x > y.

Proof: Consider y/x. By the Archimedean Property there is N ∈ N such


that N > y/x, and hence N x > y. 

Theorem 3.6.9 (Archimedean Property of R, ver.3) Let x, y ∈ R+ .


y
Then there exists N ∈N such that 0< < x.
N
y
Proof: There is N ∈N such that N x > y > 0. Hence 0< < x. 
N
Theorem 3.6.10 (Greatest Integer Function) Let x ∈ R. Then there
is a unique [x] ∈ Z such that [x] ≤ x < [x] + 1.

Proof: Let S = {n ∈ Z : n ≤ x}. S is bounded above by x, so it has a


supremum α. Now α − 1 is not an upper bound of S , hence there is m ∈ S
such that α − 1 < m. Then α < m + 1 and so m + 1 ∈ / S.
Now m ∈ S implies m ≤ x. And m+1 ∈
/S implies x < m + 1. So we
can take [x] = m.
For uniqueness, supposem0 is any integer that satises m0 ≤ x < m0 + 1.
0 0 0
If m − m > 0 then m − m ≥ 1 implies m + 1 ≤ m ≤ x, a contradiction.
0 0
Similarly, m−m > 0 implies m +1 ≤ m ≤ x, also a contradiction. Therefore
m = m0 by Trichotomy. 

Theorem 3.6.11 (Denseness of Q in R) Let x, y ∈ R and y > x. Then


there is p∈Q such that x < p < y.

1
Proof: Since y − x > 0, there exists N ∈ N< y − x.
such that 0 <
N
Then 1 < N y − N x, hence N x + 1 < N y . Dene M = [N x] + 1. Then,
N x < [N x] + 1 ≤ N x + 1 < N y , i.e. N x < M < N y . So we can take
p = M/N . 
Members of Qc = R \ Q are called irrational numbers.
Theorem 3.6.12 (Denseness of Qc in R) Let x, y ∈ R and y > x. Then
there is t ∈ Qc such that x < t < y.
√ √
Proof: By the denseness of
√ Q, we have p∈Q such 2x < p < 2 y. Take
t = p/ 2. 
150 CHAPTER 3. NUMBER SYSTEMS

Absolute Values

The absolute value |x| of a real number x is dened by



x if x≥0
|x| = .
−x if x<0

Some obvious properties of absolute value are

1. |x| ≥ 0,
2. |x| = 0 ⇐⇒ x = 0,
3. |xy| = |x||y|.
We view |x| as giving the distance between x and 0. If we have two real
numbers x, y then the distance between them is |x − y|.

Task 3.6.13 Let a ≥ 0. Show that |x| ≤ a if and only if −a ≤ x ≤ a. (Hint:


consider the cases x≥0 and x < 0)

Theorem 3.6.14 (Triangle Inequality) Let x, y ∈ R. Then |x + y| ≤


|x| + |y|.

Proof: We have −|x| ≤ x ≤ |x| and −|y| ≤ y ≤ |y|.


Therefore −(|x| + |y|) = −|x| − |y| ≤ x + y ≤ |x| + |y|.
Therefore |x + y| ≤ |x| + |y|. (Using the previous Task) 

Task 3.6.15 Let x, y ∈ R. Show that |x − y| ≥ |x| − |y|.

Task 3.6.16 Let x, y ∈ R. Show that |x − y| ≥ ||x| − |y||.

Decimal Representations

In this portion, we take up the task of representing individual real numbers


in a concrete manner that can be used for computations.

x ∈ R+ . Using the greatest integer function, we can express x =


Let
[x] + r1 where [x] ∈ N ∪ {0} and 0 ≤ r1 < 1. We know [x] has a decimal
representation, an expression in terms of powers of 10: units, tens, hundreds,

Undergraduate Mathematics: Foundations


3.6. REAL NUMBERS 151

thousands, and so on. To describe r we use negative powers of 10: tenths,


hundredths, thousandths, and so on. This is how we do it. Let

a1 = [10r1 ], r2 = 10r1 − a1 ,
a2 = [10r2 ], r3 = 10r2 − a2 ,
. .
. .
. .

an = [10rn ], rn+1 = 10rn − an ,


. .
. .
. .

Note that for each n, 0 ≤ rn < 1 and so an ∈ {0, 1, . . . , 9}. For each n,
consider
n
X
xn = [x] + ai 10−i .
i=1

1. One can use mathematical induction to prove that for each n, 0 ≤


x − xn < 10−n . ’
2. It follows that x = sup{ xn | n ∈ N }. ’
Our usual decimal notation is just a convenient way of representing this
fact. When we say x = 3.333 . . . , we mean that x is the supremum of the set
whose members are 3.3, 3.33, 3.333, etc. The general form of this notation
is d.a1 a2 a3 . . . where d stands for the decimal representation of the integer
[x].
The decimal representation of 0 is of course 0.000. . .

Conversely, given any expression of the form d.a1 a2 a3 . . . , we have a


corresponding non-negative real number x. It is given by

n
X
x = sup{ d + ai 10−i | n ∈ N }.
i=1

If x < 0, its decimal representation is just the decimal representation of −x


with a minus sign in front.

What we have just done is to describe one process for generating a decimal
representation of a given number. Could there be others, and could they lead
to a dierent decimal representation? Is the decimal representation of a real
number unique? It surprises most people to learn that the answer to the last
question is not always. For an example, consider the decimal representation
0.999 . . . of some real number x. Let x1 = 0.9, x2 = 0.99, and in general
152 CHAPTER 3. NUMBER SYSTEMS

Pn −i
xn = i=1 9 · 10 . Dene A = { xn | n ∈ N }. Then x = sup(A). But we
can easily compute the supremum in another way. First we note that 1 is
−n
an upper bound of A. Second, 1 − xn = 10 . Hence 1 = sup(A) = x.
Therefore 0.999 . . . can also be expressed as 1.000 . . . !
On completing the next Task you will have established that this is the
only kind of double representation of a real number in decimal notation, and
so we almost have a bijection between the real numbers and their decimal
representations.

Task 3.6.17 Suppose a real number has two dierent decimal representa-
tions. Then one representation eventually consists of only 9's and the other
eventually consists of only 0's, and both become eventually constant at the
same stage.

8
The fact that we mostly count in powers of 10 is a historical accident.
Mathematically, we may equally well count in powers of 2 or any higher
natural number. A representation of a real number as a sum of powers of 2
is called a binary representation. For example,

1. We can express ten as 0 · 20 + 1 · 21 + 0 · 22 + 1 · 23 and hence we say


that the binary representation of ten is 1010.

0 1 0 1
2. We can express one-third as 0+ + 2 + 3 + 4 + ··· and hence
21 2 2 2
we say that its binary representation is 0.0101 . . .

Here too, one can have two distinct representations of the same number,
such as 0.111 . . . = 1.000 . . . .

Task 3.6.18 Suppose a real number has two dierent binary representa-
tions. Then one representation eventually consists of only 1's and the other
eventually consists of only 0's, and both become eventually constant at the
same stage.

Exercises for Ÿ3.6

1 Let A, B be non-empty subsets of R such that a ∈ A, b ∈ B implies a < b.


Show there is a real number r such that a ≤ r ≤ b for every a ∈ A and
b ∈ B. (Also show that the claim is false if we replace R by Q.)

8 The rst fully developed place value system came about in Mesopotamia around
1800BC, and it used powers of sixty!

Undergraduate Mathematics: Foundations


3.6. REAL NUMBERS 153

2 Let A, B be non-empty subsets of R such that A∪B = R and a ∈ A, b ∈ B


implies a < b. Show there is a unique real number r such that a ≤ r ≤ b for
every a ∈ A and b ∈ B .

3 Suppose F is an ordered eld with the following property: If A, B ⊆ F


are non-empty, A ∪ B = F and a ∈ A, b ∈ B implies a < b, then there is a
unique r ∈ F such that a ≤ r ≤ b for every a ∈ A and b ∈ B . Show that F is
an order complete eld.

4 Let x ∈ R+ . Prove that

(a) x < 1 =⇒ x2 < x < 1, and x > 1 =⇒ x2 > x > 1,


√ p
(b) y = x =⇒ 1/y = 1/x.

5 (Existence of Square Roots) Prove the following sequence of claims,


which will establish that every positive real number has a unique positive
square root. Fix a ∈ R+ and dene A = { x ∈ R+ | x2 ≤ a }.
(a) If a has a positive square root, it is unique.

(b) For existence, it is enough to do the a > 1 case. (So we assume a>1
in the subsequent steps)

(c) 1 ∈ A (∴ A 6= ∅) and a is an upper bound for A.


(d) Let c = sup(A). Then 1 ≤ c ≤ a.
1 a
(e) Assume c2 > a. Let d = c+ . Then a < d < c, a contradiction.
2 c
a  1
(f) Assume c2 < a. Let e = c+ . Then c < e but e2 ≤ a, a
a+1 c
contradiction.

6 Here is an alternate approach to proving the existence of square roots.


Again, x a > 1, dene A = { x ∈ R | x2 ≤ a }, and let c = sup(A).
(a) Assume c2 > a. Then c2 = a + δ with δ > 0. Let h > 0 and compute
(c − h) = c2 − 2ch + h2 > a + δ − 2ch. Find a value of h that ensures
2

(c − h)2 > a. This rules out c2 > a.


(b) Assume c2 < a. Then c2 = a − δ with δ > 0. Let h > 0 and compute
(c + h) = c2 + 2ch + h2 = a − δ + h(2c + h). Find a value of h that
2
2 2
ensures (c + h) < a. This rules out c < a.

7 Show that every real number has a unique cube root.


154 CHAPTER 3. NUMBER SYSTEMS

8 Let x, y ∈ R. Prove that



(a) x2 = |x|,
√ √ √
(b) If x, y ≥ 0 then xy = x y.

c
9 Let a ∈ R, c ∈ R+ , and 0≤a≤ for every n ∈ N. Show that a = 0.
n
1
10 Let x, y ∈ R such that |x − y| ≤ for every n ∈ N. Prove that x = y.
n

n n
11
X X
Let x 1 , . . . , x n ∈ R. Show that
x i

|xi |.
i=1 i=1

12 (Cantor's Intersection Theorem) In = [an , bn ] be a family of


Let
closed and bounded intervals indexed by N, such that In+1 ⊆ In for each
n ∈ N. Let A = { an | n ∈ N } and B = { bn | n ∈ N }. Prove that sup(A) ≤
inf(B) and
\
In = [sup(A), inf(B)].
n∈N

In particular, the intersection is non-empty.

3.7 Complex Numbers


As you may remember from school, the rst step in creating the `complex
numbers' is to declare the existence of a new object named i that has the
property that i2 = −1. We then combine this new object with real numbers
to get the complex numbers. The combinations have the form x + iy where
x, y ∈ R. We call x the real part of x + iy, while y is called the imaginary
part. These new numbers have an addition and a multiplication dened by

(x+iy)+(u+iv) = (x+u)+i(y+v), (x+iy)(u+iv) = (xu−yv)+i(xv+yu).

Complex numbers were forced on mathematicians when they tried to solve


cubic equations. A cubic equation has the form x3 + ax2 + bx + c = 0. If the
coecients a, b, c are real numbers, such an equation can have up to three
real solutions. Yet when mathematicians searched for formulas that would
capture all solutions they kept getting stuck. Ultimately, they realized that
the only resolution was to carry out the standard algebraic manipulations
while treating the square root of −1 as an actually existing quantity which

Undergraduate Mathematics: Foundations


3.7. COMPLEX NUMBERS 155

is subject to the same rules as any real number. We tell this story later in
this section.

Two questions arise: Can we t complex numbers within our formal


approach of sets and binary operations, and are they more than a formal
game with symbols? Let us start with the rst question.

Consider the cartesian plane R2 = R × R, and dene binary operations


on it as follows:

Addition: (x, y) + (u, v) = (x + u, y + v),


Multiplication: (x, y)(u, v) = (xu − yv, xv + yu).

We use C to denote the cartesian plane with these operations and call it the
complex plane.
Task 3.7.1 Consider the map ϕ : R → C, x 7→ (x, 0). Verify that it is
1-1, that its image is the  x-axis, and that for every x, y ∈ R we have
ϕ(x + y) = ϕ(x) + ϕ(y) and ϕ(xy) = ϕ(x)ϕ(y).

Thus C contains a copy of the eld R. Moreover, C is itself a eld:

1. Addition and multiplication are commutative and associative. ’


2. (0, 0) is the additive identity, and (−x, −y) is the additive inverse of
(x, y). ’
3. (1, 0) is the multiplicative identity, and a nonzero (x, y) has multiplica-
x −y 
tive inverse , 2 . ’
x + y x + y2
2 2

4. Multiplication distributes over addition. ’


The member (x, y) of C = R2 is the formal version of the complex number
x + iy . With these denitions and observations we have created the complex
numbers using already known objects, much as we created integers from
natural numbers.

We denote (0, 0) by 0, and (1, 0) by 1. This is our usual notation for


additive and multiplicative identities, and it is also consistent with our iden-
tication of R with the x-axis of C.

Task 3.7.2 Check that (0, 1)2 = −(1, 0). Thus (0, 1) is the formal version
of i.
156 CHAPTER 3. NUMBER SYSTEMS

Now that we have seen how complex numbers t in our scheme of set
theoretic constructions, let us return to using the x + iy notation.

If z = x + iy with x, y ∈ R, we dene its absolute value to be |z| =


x2 + y 2 and its complex conjugate to be z = x − iy = x + i(−y).
p

Task 3.7.3 Let z be a complex number. Show that (a ) z = 0 ⇐⇒ |z| = 0,


(b ) zz = |z|2 , and (c ) z 6= 0 =⇒ z −1 = z/|z|2 .

Visualizing Complex Numbers

The algebraic manipulations of complex numbers are easy enough to carry


out. The remaining question pertains to their nature. Integers and rational
numbers describe counting, real numbers describe continuous measurement.
What notion or activity do complex numbers capture?

We have already seen that a formal description of complex numbers can


be obtained by identifying them with ordered pairs of real numbers. This
9
also shows us how to visualise them. Each complex number becomes a
point in the cartesian plane, which is called the complex plane in this con-
text. The operations with complex numbers can then be given geometric
interpretations.

Imaginary Axis

z = x + iy
y
2
y
2 +
p x
| = Real Axis
−x |z x

−y
−z = −x − iy z = x − iy

This diagram shows a complex number z = x + iy which is identied with


the point (x, y) of the cartesian plane. The real part x is plotted along
the horizontal axis, which is called the real axis. The imaginary part y is

9 This visualisation is due to Jean-Robert Argand (1768-1822), who published it in


1806, two and a half centuries after Rafael Bombelli worked out the algebra of complex
numbers and applied them to cubic equations.

Undergraduate Mathematics: Foundations


3.7. COMPLEX NUMBERS 157

plotted along the vertical axis, which is called the imaginary axis. The
distance
p of z from the origin is called the magnitude of z and is given by
|z| = x2 + y 2 . The number −z corresponds to reection with respect to
the origin. The complex conjugate z is obtained by reection in the real axis.

Imaginary Axis Imaginary Axis


z+w
y+v
w w
v v

y z y z
Real Axis Real Axis
u x u x x+u

This pair of diagrams depicts the addition of complex numbers. The left
diagram shows two complex numbers z = x + iy
w = u + iv . The second
and
diagram shows their sum z + w . We see that adding w to z shifts z by w .
Geometrically, if we treat z as origin and then plot w , we get z +w . Addition
in real numbers has the same eect, if we add 2 to a real number it shifts by
2 units.

Next, we consider multiplication by the imaginary unit i. It sends x + iy


to −y + ix, as shown below.

Imaginary Axis

−y + ix x

y x + iy
Real Axis
−y x

One can use the Pythagoras theorem to check that the points x + iy , −y + ix
and 0 form a right angled triangle. ’
Thus, multiplication by i causes each complex number to undergo a ro-
tation of 90 degrees around the origin. Multiplying by i twice gives a total
rotation of 180 degrees and hence sends z to −z . This is the geometric de-
scription of what i does and how it can be interpreted as a square root of
−1.
158 CHAPTER 3. NUMBER SYSTEMS

Task 3.7.4 Find all the complex square roots of −1.

Task 3.7.5 Find all the complex square roots of i.

Next, we wish to look at multiplication of arbitrary complex numbers.


As rotation is involved, let us describe complex numbers using quantities
related to rotation: length and angle.

Imaginary Axis

y = |z| sin θ z = x + iy

|z |

θ
Real Axis
x = |z| cos θ

We join the number z = x + iy to the origin by a straight line segment. This


line segment has length |z|. The angle θ is measured counterclockwise from
the positive direction of the real axis to this line segment and is called the
argument of z . The length |z| and the angle θ can be used to locate z , and
are called the polar coordinates of z . We have the following relationships
between the polar and cartesian coordinates:

p
x = |z| cos θ, |z| = x2 + y 2 ,
y = |z| sin θ, tan θ = y/x.

In this discussion of visualizing complex numbers, we are taking geometric


notions like length and angle for granted. Similarly, we are taking the
trigonometric functions and their properties for granted. A purist may
« demand that we not introduce new undened notions and that we should
phrase these too in the formal language of sets. That is a fair demand, but
the sensible place to meet it is in a book on calculus or real analysis.

We shall now use the polar coordinates to understand multiplication. Let


z have argument θ and let w have argument φ. Then z = |z| cos θ + i|z| sin θ

Undergraduate Mathematics: Foundations


3.7. COMPLEX NUMBERS 159

and w = |w| cos φ + i|w| sin φ. Let us multiply z and w:

zw = (|z| cos θ + i|z| sin θ)(|w| cos φ + i|w| sin φ)


= |z||w| (cos θ + i sin θ)(cos φ + i sin φ)
= |z||w| [(cos θ cos φ − sin θ sin φ) + i(cos θ sin φ + sin θ cos φ)]
= |z||w|(cos(θ + φ) + i sin(θ + φ)).

Imaginary Axis Imaginary Axis

zw

z z

|
|w
φ

|z |
| |
|z |z
θ w θ w
|w | |w |
φ φ
Real Axis Real Axis

We obtain the following:

1. The magnitude of zw is the product of the magnitudes of z and w.


2. The argument of zw is the sum of the arguments of z and w.
We can visualize multiplication by w as having two eects: a scaling by |w|
and a rotation by φ.

Solving Cubic Equations

We had remarked earlier that complex numbers became essential to mathe-


matics with the rst successful attempts at solving general cubic equations.
10
We shall now study the formula given by Cardano and observe the remark-
able fact that a real solution of a cubic equation requires the use of complex
11
numbers for its calculation.
10 Girolamo Cardano (1501-1576) was the rst to publish this formula, in 1545. He gave
credit for its discovery to Scipione del Ferro (1465-1526) and Niccoló Tartaglia (1500-
1557). The simple calculations we present here appeared much more complicated in the
16th century due to an insistence on avoiding negative numbers.
11 We follow the exposition in Lectures on Elementary Mathematics by Joseph Louis
Lagrange, translated by Thomas J. McCormack, The Open Court Publishing Company,
1901. http://www.gutenberg.org/ebooks/36640
160 CHAPTER 3. NUMBER SYSTEMS

Task 3.7.6 Find α ∈ R such that substituting x + α for x converts the


general cubic equation x3 +ax2 +bx+c = 0 to the special form x3 +dx+e = 0.

This Task establishes that it is enough to be able to solve cubic equations


that don't have a second degree term. Therefore we shall work with cubics
of the following form:
x3 = 3px + 2q.
We shall see later that the use of 3p and 2q simplies the expression for the
solutions. Now substitute x = y + z:
y 3 + z 3 + 3y 2 z + 3yz 2 = 3p(y + z) + 2q
=⇒ y 3 + z 3 + 3(yz − p)(y + z) = 2q.
We resolve this into two simultaneous equations:

yz = p and y 3 + z 3 = 2q.
Now substitute the rst into the second:

y 3 + (p/y)3 − 2q = 0
=⇒ y 6 − 2qy 3 + p3 = 0.

This is a quadratic in y3 . Its solution is y 3 = q+(q 2 −p3 )1/2 , where we choose


2 3 3 3 2 3 1/2
either of the square roots of q −p . We now get z = 2q−y = q−(q +p ) .
This nally gives Cardano's formula:

h i1/3 h i1/3
x = y + z = q + (q 2 − p3 )1/2 + q − (q 2 − p3 )1/2 .

We see that the choice of square root of q 2 − p3 does not matter, as both
roots nally appear in the formula for x.
Let us see what Cardano's formula produces in some simple situations.
Currently we are looking only for real solutions.

Example 3.7.7 Consider the cubic equation x3 = q . Then Cardano's for-


mula provides the solution

 !1/2 1/3  !1/2 1/3


2 2
q q q q
x= +  + − 
2 4 2 4
 1/3  1/3
q q q q
= + + − = q 1/3 .
2 2 2 2

Undergraduate Mathematics: Foundations


3.7. COMPLEX NUMBERS 161

Nothing surprising here. There was only one solution, and the formula found
it. 

Example 3.7.8 Consider the equation x3 = 6x + 6. Let us see what Car-


dano's formula can nd:

h i1/3 h i1/3
x = 3 + (32 − 23 )1/2 + 3 − (32 − 23 )1/2
= [3 + 1]1/3 + [3 − 1]1/3 = 41/3 + 21/3 .

This would probably be hard to guess! 

One drawback of Cardano's formula is that it only produces one real root,
whereas there may be up to three. This is not a serious issue because once
we know one real root a, we can divide the cubic by x−a and reduce it to
a quadratic.

Task 3.7.9 Let x=a be a real solution of x3 = 3px + 2q . Show that the
other two solutions are obtained from the quadratic equation x2 + ax + a2 =
3p.

The real trouble arises when q 2 − p3 is negative. Then, if we are only


using real numbers, the formula fails to provide us any solution. Here is an
instance:

Example 3.7.10 Let us consider the equation x3 = 9x. We rearrange it to


2
x(x − 9) = 0. The solutions are x = 0, ±3. Now let us see what Cardano's
formula can nd:

h i1/3 h i1/3 √ 1/3 √


x = (−33 )1/2 + −(−33 )1/2 = −27 + (− −27)1/3 .

We can't proceed further if we use only real numbers because we can't make
√ √
sense of −27. But, if −27 did make sense, we would have a cancellation
and we would get the x = 0 solution. This is an example of complex numbers
being forced upon us in a problem whose terms and solutions are all real. 

Now we shall explore the benets accruing from using complex numbers
in these calculations.
12
The rst is that the q 2 − p3 < 0 case is no longer a

12 The application of complex numbers to this context was rst carried out by Rafael
Bombelli (1526-1572).
162 CHAPTER 3. NUMBER SYSTEMS

dead end. For instance, in Example 3.7.10, we can continue the calculation
as follows:
√ √
1/3 √ √
x= + (− −27)1/3 = ( 27i)1/3 + (− 27i)1/3
−27
√ √
= 3 i1/3 + 3(−i)1/3 .
| {z } | {z }
y z

Here we have a dilemma. A non-zero complex number has three distinct


cube roots. Which combinations of the cube roots of i and −i shall we use
in our calculations? Let us work out the available choices. We begin by
identifying the complex cube roots of 1. These are solutions of z 3 − 1 = 0,
2
which can be expressed as (z − 1)(z + z + 1) = 0. So apart from 1 itself, the
other cube roots are solutions of z 2 + z + 1 = 0. Hence they are given by
√ √
−1 ± −3 −1 ± 3 i
z= = .
2 2

√ !2 √
−1 + 3 i −1 − 3 i
Task 3.7.11 Verify that = . Hence, if we set
2 2

−1 + 3 i 2
w = , then the cube roots of 1 are 1, w and w . Further, show
2
2
that if a is a cube root of z , then the list of cube roots of z is a, wa and w a.

Since i3 = −i and (−i)3 = i, we have the following:


√ √
± 3+i ± 3−i
i1/3 = −i, and (−i)
1/3
= i, .
2 2
Recall that the terms of our solution x=y+z were chosen to satisfy yz =
p = 3. Hence the valid combinations are as follows:

y z x
√ √
− √ 3i 3√i 0
3 + 3i 3 − 3i
3
2√ 2√
−3 + 3i −3 − 3i
−3
2 2

Thus, in this example, the use of complex numbers has paid o twice. First,
we found a way past the diculty of having to compute the square root of a

Undergraduate Mathematics: Foundations


3.7. COMPLEX NUMBERS 163

negative number. Second, the multiplicity of cube roots of a complex number


enabled Cardano's formula to generate all three solutions of our equation.

Now that we have had a glimpse of the benets of using complex numbers,
let us reconsider the cubic equation x3 = 3px + 2q . Setting x = y + z, we
once more reach the conditions

yz = p,
y + z 3 = 2q.
3

p p
2 3
Let q 2 − p3 q − p
be a square root of
p , y be a cube root of q + q 2 − p3 ,
and z be a cube root of q − 2 3
q − p . Then the choice of square root does
3 3
not matter. Also, each combination of y and z satises y + z = 2q . Let us
consider the product yz :

p p
(yz)3 = (q + q 2 − p3 )(q − q 2 − p3 ) = p3 =⇒ yz = p or wp or w2 p.

Hence, for a given y there is exactly one choice of z such that yz = p is also
satised and x = y + z solves the cubic. ’

Task 3.7.12
p p
y 3 = q + q 2 − p3 and z 3 = q − q 2 − p3 such that
Let
x = y + z is a solution of x3 = 3px + 2q . Show that the other solutions are
x = wy + w2 z and x = w2 y + wz .

Exercises for Ÿ3.7

1 Let z, w ∈ C. Show that

(a) z + w = z + w, (b) −z = −z ,

(c) zw = z w, (d) z/w = z/w if w 6= 0.

2 Let z, w ∈ C. Show that

(a) |zw| = |z||w|, (b) |z + w| ≤ |z| + |w|.

3 Show that C cannot be an ordered eld.

4 Sketch the subsets of the complex plane whose elements z satisfy the given
condition:
164 CHAPTER 3. NUMBER SYSTEMS

(a) z = z, (b) z = −z ,

(c) |z| < 1, (d) 1 ≤ |z| < 2.

5 Consider the function f : C → C dened by f (z) = z 2 . Find the images


of the following subsets of C under this function:

(a) Real Axis, (b) Imaginary Axis,

(c) Lines passing through origin, (d) Circles centered at origin.

6 Consider a polynomial P (x) = a0 + a1 x + a2 x2 + · · · + an xn where the


coecients ai are complex numbers.

(a) Show that P (α) = 0 ⇐⇒ P (x) = (x − α)Q(x) where Q(x) is also a


polynomial with complex coecients.

(b) Show that the equation P (x) = 0 has at most n distinct complex
solutions.

7 Consider a polynomial P (x) = a0 + a1 x + a2 x2 + · · · + an xn where the


coecients ai are real numbers. Let α ∈ C. Show that P (α) = 0 ⇐⇒
P (α) = 0.

8 Express the square roots of x + iy in cartesian and polar coordinates.

9 Use polar coordinates to nd the three complex roots of 1.

10 Show that:

(a) The unity 1 has exactly n distinct nth roots and they have the form
1, w, w2 , ..., wn−1 .
(b) The nth roots of unity satisfy 1 + w + w2 + · · · + wn−1 = 0.

11 Assume that every positive real number a has a unique positive nth root.
Show that every non-zero complex number has exactly n distinct complex
nth roots.

Undergraduate Mathematics: Foundations


4 | Innity
The second philosopher asked: `How many stars are there in the
sky?'.

Nasrudin immediately replied: `Exactly the same number as there


are hairs in the coat of my donkey. Anyone who disbelieves this
is at liberty to count both.'
1
 Idries Shah

The Axiom of Choice (together with the Continuum Hypothe-


sis) is probably the most interesting and most discussed axiom
in mathematics after Euclid's Axiom of Parallels.
2
 P. Bernays and A.A. Fraenkel

When it comes to innite sets, we have been a bit evasive. We have


referred to certain innite sets such as N, Z, Q, R and C, but we haven't
dened them from the set theory axioms. Rather, we have used the set
theory axioms to describe Z, Q, R and C in terms of N, but N itself has
not been constructed from the set theory axioms. Instead, we took it to be
independently given via the Peano axioms.

At the same time we have been careful in not allowing the words nite
and innite into our theorems. It is time we come to grips with these matters.
First, we need axioms that allow us to speak of innite sets, by providing
for the existence of at least one such set.

1 Idries Shah, The Pleasantries of the Incredible Mulla Nasrudin, The Octagon Press,
1995.
2 Quoted in H. Herrlich, Axiom of Choice. Springer, 2006.

165
166 CHAPTER 4. INFINITY

4.1 Constructing the Natural Numbers


To create innite sets we need a process that can create new elements without
end. The simplest way to create a new element is to make {x} from x, using
the Axiom of Pairs . But is this denitely new? Could we have x = {x}?
None of our earlier axioms rule this out. We introduce a new axiom that
helps avoid this strange situation.

ZF7. Axiom of Regularity: If A is a non-empty set then one of its members


is disjoint from A.

If A = {A}, then we rst note that A 6= ∅. Hence A ∩ {A} = A ∩ A =


A 6= ∅, violating the Axiom of Regularity.

Task 4.1.1 Prove that the Axiom of Regularity rules out the possibility of
A ∈ A. (Hint: Consider the set {A})

ZF8. Axiom of Innity: There is a set I with the following properties:

(a) ∅ ∈ I,
(b) If x∈I then x ∪ {x} ∈ I .

Due to the Axiom of Regularity we have x∈


/ x, and hence x x ∪ {x}.
Any set with the properties listed in the Axiom of Innity will be called an
inductive set.
We observe that I has the following distinct members, obtained by suc-
cessive applications of property (c) to ∅:

∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}, . . .

It may also, conceivably, have members that are not obtained in this fashion.
We have no control over these members, so we would like to eliminate them.
We do this as follows:

1. For each x∈I dene its successor S(x) = x ∪ {x}. Since S(x) ∈ I ,
we obtain a function S: I → I called the successor function.

2. A subset A of I will be an inductive set if ∅∈A and S(A) ⊆ A.


3. Use the separation and power set axioms to dene the set of all induc-
tive subsets of I:

J = { A ∈ P(I) | ∅ ∈ A and S(A) ⊆ A }

Undergraduate Mathematics: Foundations


4.1. CONSTRUCTING THE NATURAL NUMBERS 167

T
4. Intersect all the inductive subsets: N = J. (Then x ∈ N ⇐⇒
(∀A ∈ J ) (x ∈ A))
5. Verify that N is inductive. ’
6. Thus N is the smallest inductive subset of I and hence the best candi-
date for not having any extraneous elements.

3
We name the members of N as follows:

1=∅={}
2 = S(1) = 1 ∪ {1} = ∅ ∪ {∅} = {∅} = {1}
3 = S(2) = 2 ∪ {2} = {1} ∪ {2} = {1, 2}
4 = S(3) = 3 ∪ {3} = {1, 2} ∪ {3} = {1, 2, 3}
and so on

Theorem 4.1.2 N satises the Peano axioms.

Proof: We have already dened the successor function S : N → N.


First, we prove the axiom of mathematical induction. Let A ⊆ N such
that1 = ∅ ∈ A and S(A) ⊆ A. Then A is an inductive subset of I . Hence
N ⊆ A, and so A = N.
Second, we prove that 1 is not in the range of S. We just observe that
for any x ∈ N, x ∈ S(x), hence S(x) 6= ∅ = 1.
Finally, we need to show that S is 1-1. Now,

S(x) = S(y) =⇒ x ∪ {x} = y ∪ {y}


=⇒ (x ∈ y ∨ x = y) ∧ (y ∈ x ∨ y = x)
=⇒ x = y ∨ (x ∈ y ∧ y ∈ x).

To complete the proof we will show by mathematical induction that every


z∈N has the following property:

x∈y and y∈z implies x ∈ z. (4.1)

It will follow from this that (x ∈ y ∧ y ∈ x) =⇒ x ∈ x, hence x∈y∧y ∈x


is impossible, hence x = y.
So we let A be the subset of N consisting of elements that have the
property given by 4.1. First we note that ∅ ∈ A. Now, suppose z ∈ A. Let

3 Some prefer to start with 0 = ∅. It depends on whether we wish to include 0 in N.


168 CHAPTER 4. INFINITY

x ∈ y and y ∈ S(z) = z ∪ {z}. Then y ∈ z or y = z . In the rst case we have


x ∈ z because z has property 4.1. In the second case we have x ∈ z because
x ∈ y = z . Hence, by mathematical induction, A = N. 
We have just seen one way of creating a set with a successor function
that satises the Peano Axioms. There could be other ways of achieving
this. Then we would have competing versions of the natural numbers with
potentially dierent properties. Clearly, this is undesirable.

Let us dene a Peano system to be a triple (P, S, e) of a set P, a


function S : P → P, and an element e ∈ P, that satisfy the Peano axioms:

1. S is a 1-1 function.

2. The element e is not in the range of S.

3. If A⊆P satises e∈A and S(A) ⊆ A then A = P.


Given two Peano systems (P, S, e) and (P , S , e ) we call them isomorphic
0 0 0

if there is a bijection f : P → P 0 such that f (e) = e0 and S 0 (f (x)) = f (S(x))


for every x ∈ P. The function f is called an isomorphism Isomorphic
Peano systems can be viewed as being essentially the same, the only change
from one to the other is a renaming of the elements.

Our earlier results about N, since they are based only on the Peano
axioms, hold for all Peano systems. In particular, Theorems 3.1.1 and 3.1.2
do so. For convenience, we restate them in our new terminology:

Theorem 4.1.3 If (P, S, e) is a Peano system then for each y ∈ P \ {e}


there is a unique x ∈ P such that S(x) = y .

Theorem 4.1.4 (Recursion Theorem) Consider a set X with a function


G: X → X and an element c ∈ X . Let (P, S, e) be a Peano system. Then
there is a unique function f : P → X such that
1. f (e) = c,

2. ∀n ∈ P , f (S(n)) = G(f (n)).

With the help of these results we can show that there is essentially only
one Peano system.

Theorem 4.1.5 All Peano systems are isomorphic.

Undergraduate Mathematics: Foundations


4.1. CONSTRUCTING THE NATURAL NUMBERS 169

Proof: Let (P, S, e) and (P 0 , S 0 , e0 ) be two Peano systems. In Theorem 4.1.4,


0 0 0 0
set X = P , G = S and c = e . Then we obtain a unique function f : P → P
0 0
such that f (e) = e and ∀n ∈ P , f (S(n)) = S (f (n)). We need to show f is
a bijection.

To show that f is onto, apply the principle of mathematical induction to


the subset f (P ) of P 0. ’
To show that f is one-one, apply the principle of mathematical induction
to the subset A = { x ∈ P | f (x) = f (y) =⇒ x = y } of P . (Theorem 4.1.3
may help) ’


Exercises for Ÿ4.1

1 Express the Axioms of Regularity and Innity symbolically using logical


operations and quantiers.

2 Show that the Axiom of Regularity implies that there cannot be sets A
and B such that A∈B and B ∈ A.

3 Show that the Axiom of Regularity implies that there cannot be sets
A0 , A1 , . . . , An with A0 ∈ A1 ∈ · · · ∈ An ∈ A0 .

4 Which of the following choices of a set P and a function S:P →P can


be Peano systems (P, S, e), with an appropriate choice of e?
(a) P = N, S(n) = n + 2.
(b) P = W, S(n) = n + 1.
(c) P = Z, S(n) = n + 1.
(d) P = 2N = { 2n | n ∈ N }, S(n) = n + 2.
(e) P = { n ∈ N | n ≥ 5 }, S(n) = n + 1.

5 Given an inductive set I , dene NI to be the intersection of all the induc-


tive subsets of I. Prove that if I and J are inductive sets then NI = NJ .

6 A set A is called transitive if x ∈ A implies x ⊆ A. Show that A is


transitive if and only if x ∈ y ∧ y ∈ A implies x ∈ A.

7 Let X be an inductive set. Show that the following are also inductive sets:
170 CHAPTER 4. INFINITY

(a) { x ∈ X | x ⊆ X }, (b) {x ∈ X | x is transitive }.

8 Let N be dened as the smallest inductive subset of a given inductive set


I. Show that:

(a) Every member of N is transitive,

(b) N is transitive.

9 Show that X is transitive if and only if its power set P(X) is transitive.

10 Consider the following variant of the Axiom of Innity: There is a set A


such that ∅∈A and X ∈ A =⇒ {X} ∈ A. Show how a Peano system can
be obtained from this axiom.

4.2 Finite and Innite Sets


Now we will take up the formal denitions of nite and innite sets and how
the two are distinct from each other. We begin with nite sets. Our intuition
for a nite set is that its members can be counted one-by-one and the process
will ultimately stop. Counting means that we rst associate 1 to an element,
then 2 to another, and so on. If we have associated the natural number m
to an element, then in the next step we will associate S(m) = m + 1. When
the process stops with some n ∈ N, every natural number smaller than n
will already have been used.

Therefore, for every n ∈ N, we dene Fn = { m ∈ N | m ≤ n }. And we


dene F0 = ∅ . We say that a set A is nite if it is in bijection with Fn for
some n ∈ W.

Theorem 4.2.1 If there is a 1-1 map f : Fm → Fn then m ≤ n.

Proof: We shall proceed by mathematical induction.

We have little choice in this. Mathematical induction is the only tool we


possess for proving any non-trivial statement about all natural numbers.
7 The only issue is whether we shall apply mathematical induction to m or
n. Below, we oer a proof based on induction on n. You may want to see
if you can develop a proof based on m.

Dene

A = { n ∈ W | (∃f )(f : Fm → Fn is 1-1) =⇒ m ≤ n }.

Undergraduate Mathematics: Foundations


4.2. FINITE AND INFINITE SETS 171

First, we check that 0 ∈ A. The only function with codomain F0 = ∅ is the


`empty function' with domain ∅. In this case we have m = n = 0 and so
0∈A is true.

Next, suppose n ∈ A. We need to show that if f : Fm → Fn+1 is 1-1 then


m ≤ n + 1.
Case 1: Suppose n+1 ∈ / f (Fm ). Then f (Fm ) ⊆ Fn and f denes a 1-1
function from Fm intoFn . Since n ∈ A we get m ≤ n ≤ n + 1.
Case 2: Suppose n+1 ∈ f (Fm ). Then ∃m0 ∈ Fm such that f (m0 ) = n+1.
The other members of Fm map into Fn . Dene g : Fm−1 → Fm by

k if k < m0
g(k) = .
k + 1 if k ≥ m0

Then g is 1-1. Now g(k) 6= m0 =⇒ f (g(k)) 6= n+1, since f is 1-1. Therefore


f ◦g maps Fm−1 into Fn . Since f ◦ g is also 1-1, and n ∈ A, we get m − 1 ≤ n
and hence m ≤ n + 1. 

Theorem 4.2.2 If there is a bijection f : Fm → Fn then m = n.

Proof: Apply the previous theorem to f and f −1 . 

Theorem 4.2.3 If f : Fn → Fn is 1-1 then it is onto. Hence Fn is not in


bijection with any proper subset of itself.

Proof: Suppose f : Fn → Fn is 1-1 but not onto. The existence of such f


implies n > 0.
Case 1: Suppose n∈
/ f (Fn ). Then f is a 1-1 map of Fn into Fn−1 , and
so n ≤ n − 1, which is a contradiction.

Case 2: Suppose n ∈ f (Fn ). Then ∃m0 ≤ n such that f (m0 ) = n. Also,


since f is not onto, ∃n0 < n such that n0 ∈ / f (Fn ). Dene g : Fn → Fn by

f (k) if k 6= m0
g(k) = .
n0 if k = m0

Then g is 1-1 with n∈


/ g(Fn ), and so we are in the Case 1 situation, again
leading to a contradiction. 

Theorem 4.2.4 A nite set is not in bijection with any proper subset of
itself.
172 CHAPTER 4. INFINITY

Proof: Suppose A is a nite set. Then there is a bijection g : A → Fn .


Further, let f : A → A be 1-1. Then h = g ◦ f ◦ g −1 : Fn → Fn is 1-1, hence
−1
onto, hence a bijection. Therefore f = g ◦ h ◦ g is a bijection. 
We call a set innite if it is not a nite set.
Theorem 4.2.5 The set of prime numbers is innite.

7 This is a famous proof by contradiction, of the same league as the proof


that

2 is irrational.

Proof: Suppose there are nitely many primes. Since their set is in bi-
jection with some Fn we can label them as p1 , p2 , . . . , pn . Now consider
q = p1 p2 · · · pn + 1. Since q ≥ 2 it has a prime factor p. But none of the pi
divide q (they all leave remainder 1). So this p is not present in our list of
all primes, a contradiction! 

Theorem 4.2.6 A set is innite if it is in bijection with some proper subset


of itself.

Proof: This is just the contrapositive of Theorem 4.2.4. 

Task 4.2.7 Show that N, Z, Q and R are innite.

Finally, if A is a nite set and has a bijection with Fn we say that A


has cardinality n and we write |A| = n. (In particular, |∅| = 0) Our
results show that the cardinality of A is well-dened, and that the notation
|A| ≤ |B| is consistent with the ordering of natural numbers.

Now, suppose we have A ⊆ B ⊆ C and |A| = |C|. We suspect that


sinceA and C are the same size and B sits between them, we should have
|A| = |B| = |C|. If this turns out to be correct, we'll have a powerful tool for
proving two sets have the same cardinality without nding an actual bijection
between them. For example, (0, 1) ⊆ [0, 1] ⊆ (−1, 2) and f : (0, 1) → (−1, 2),
x 7→ 3x − 1, is a bijection. This will imply that (0, 1) and [0, 1] have the
same cardinality.

If Cantor's Theorem was our rst striking theorem, the theorem which
conrms our suspicion has our rst really clever proof !

Undergraduate Mathematics: Foundations


4.2. FINITE AND INFINITE SETS 173

Theorem 4.2.8 Let A⊆B⊆C and |A| = |C|. Then |A| = |B| = |C|.

Proof: Let f: C →A be a bijection. We wish to use f to create a bijection


from C to B.
Applying f to A ⊆ B ⊆ C, we get f (A) ⊆ f (B) ⊆ f (C). Substituting
f (C) = A gives
f (f (C)) ⊆ f (B) ⊆ f (C) ⊆ B ⊆ C.
This can be extended repeatedly in a pattern with alternating B and C:

· · · ⊆ f (f (f (C))) ⊆ f (f (B)) ⊆ f (f (C)) ⊆ f (B) ⊆ f (C) ⊆ B ⊆ C.

This sequence of shrinking sets is depicted below. It creates bands which


we have coloured gray and white, alternately. The outermost gray band
represents C \ B . The next gray band represents f (C) \ f (B). After that
comes f (f (C)) \ f (f (B)), and so on.

f (f (B)) f (f (C)) f (B) f (C) B C

Let X1 = C \ B . Since f is 1-1, we get f (X1 ) = f (C) \ f (B), f (f (X1 )) =

S∞\ f (f (B)), and so on. Dene Xn+1 = f (Xn ) for each n ∈ N and set
f (f (C))
X = n=1 Xn . Observe that
S∞
1. f (X) = n=2 Xn ⊂ B .

2. C \ X ⊆ B, since x∈ / X =⇒ x ∈/ X1 =⇒ x ∈ B .

f (x) if x ∈ X
Dene g: C → B by g(x) = .
x if x ∈
/X
From the two observations it follows that g indeed maps C into B. Fur-
ther, it maps X into X and Xc into X c. Let us check that g is a bijection:

1. Onto: Let x ∈ B . If x ∈ / X weShave g(x) = x. If x ∈ X , then rst



x ∈ B =⇒ x ∈ / X1 . Hence x ∈ n=2 Xn = f (X). Hence ∃t ∈ X such
that g(t) = f (t) = x.
174 CHAPTER 4. INFINITY

2. 1-1: g(x) = g(x0 ) = y . If y ∈ X then x, x0 ∈ X . Therefore


Let
f (x) = g(x) = g(x0 ) = f (x0 ). Since f is 1-1 we get x = x0 . On the
c 0 c 0 0
other hand, if y ∈ X then x, x ∈ X . Hence x = g(x) = g(x ) = x .

Task 4.2.9 An interval in R is called non-degenerate if it has more than


one point. Show that all non-degenerate intervals have the same cardinality.

Theorem 4.2.10 (Schröder-Bernstein Theorem) Suppose that |A| ≤


|B| and |B| ≤ |A|. Then |A| = |B|.

Proof: We have 1-1 maps f: A → B and g : B → A. We can use these to


create the following subsets of A:

g(f (A)) ⊆ g(B) ⊆ A.

Now g ◦ f : A → g(f (A)) |g(f (A))| = |A|. From the


is a bijection. Hence
previous theorem, we can conclude that |g(B)| = |A|. Since g : B → g(B) is
a bijection, we also have |g(B)| = |B|. Hence |A| = |B|. 

Exercises for Ÿ4.2

1 Let f : A → Fn be a 1-1 map. Show that A is nite.

2 Prove that:

(a) A subset of a nite set is nite.

(b) A set with an innite subset is innite.

3 Let A and B be nite sets, and assume the denition of cardinality as a


whole number. Show that

(a) If x∈
/A then |A ∪ {x}| = |A| + 1.
(b) |A ∪ B| ≤ |A| + |B|.
(c) If A∩B =∅ then |A ∪ B| = |A| + |B|.
(d) |A \ B| = |A| − |A ∩ B|.
(e) |A ∪ B| = |A| + |B| − |A ∩ B|.

4 Show that |Fm × Fn | = |Fmn |.

Undergraduate Mathematics: Foundations


4.2. FINITE AND INFINITE SETS 175

5 Suppose |A| = n.
Qn
(a) Give a bijection between P(Fn ) and k=1 F2 .
n
(b) Show that |P(A)| = 2 .

6 Show that if f: A→B and A is nite then f (A) is nite.

7 Show that if f: N→A is a 1-1 function then A is innite.

8 Prove Theorem 4.2.1 via induction on m (instead of n).

9 Let A be a set.

(a) Prove that if there is a 1-1 map gn : Fn → A for every n∈N then A
is innite.

(b) Use (a) to prove that there are innitely many rational numbers be-
tween any two rational numbers.

10 Give a bijection between N and N × Fn .

11 Give a bijection between N and N × N. The diagram below gives a hint


for one such bijection:

(1,1) (1,2) (1,3) (1,4) ···


3 6
1

(2,1) (2,2) (2,3) (2,4) ···


5
2

(3,1) (3,2) (3,3) (3,4) ···


4

(4,1) (4,2) (4,3) (4,4) ···


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

12 Show that the following sets have the same cardinality:

(a) R and R∗ , (b) R and R \ Z.

13 Show that
176 CHAPTER 4. INFINITY

(a) N is in bijection with Q+ , (b) N is in bijection with Q.

4.3 The Axiom of Choice


It is time to take up a thorny issue that has troubled generations of mathe-
maticians and for which the evidence suggests that we have to learn to live
with it rather than hope for a denite resolution. This may surprise you, be-
cause mathematics is usually presented as a science that embodies certainty.
The current status of this issue is akin to questions like Does God exist?
We do not have an answer that convinces everybody, and probably there is
no way to get such an answer. Yet we are unable to ignore the question.
Everyone has their own response and this response aects how we live our
lives and do our work, and also how we view the way others live and work.

Our dilemma begins with a simple act, the act of selecting a member of
a set. One can hardly write any signicant amount of mathematics without
including a phrase of the form let x be a member of A. In formal work this
is accomplished by the existentially quantied statement ∃x ∈ A. Diculties
arise when we wish to make innitely many such selections during a proof.
If we express them one by one, we will have a proof that never ends. Can
we express them all at once? Let us see what that entails.

First, suppose we have a set A whose members are non-empty sets. Let
choice function on A is a
S
A be the union of all the members of
S A. A
function c: A → A such that c(S) ∈ S for every S ∈ A. A choice function
therefore selects a member of each set S from the collection A.
If we want to use choice functions to make selections, we should have
some guarantee that they exist. One way to accomplish this is through an
axiom:

Z9. Axiom of Choice: Let A be a set whose members are non-empty sets.
Then there is a a choice function on A.

Before we go on to consider some implications of the Axiom of Choice,


let us also see that not every selection requires this axiom. First, we do not
require it to make nitely many selections. Second, if we have suciently
explicit knowledge of the sets, we may be able to give a direct formula and not
have to resort to the Axiom of Choice. For example, suppose the members
of A are the real intervals In = (0, 2/n). Then we just dene c(In ) = 1/n,
and do not have to invoke the Axiom of Choice.

Undergraduate Mathematics: Foundations


4.3. THE AXIOM OF CHOICE 177

Here is a simple illustration of the use of the Axiom of Choice:

Theorem 4.3.1 Let f : X → Y be an onto function. Then there is a 1-1


function g: Y → X such that f ◦ g = 1Y .

Let's consider the situation before formulating the proof. For f to reverse
g, g needs to map each point of Y
to one of its pre-images under f . In
7 other words, we need to select a member of f
−1
({y}) for each y ∈ Y . This
requires a choice function.

Proof: Let A = P(X) \ {∅}. Dene h : Y → A by h(y) = f −1 ({y}). Since


f is onto, each h(y) 6= ∅. By Axiom of Choice, there is a choice function c
on A. Then c maps A to ∪A = X . Dene g : Y → X by g = c ◦ h.

Suppose y y 0 are distinct elements of Y . Then h(y) and h(y 0 ) are


and
0 0
disjoint subsets of X . Since c(h(y)) ∈ h(y) and c(h(y )) ∈ h(y ), we see that
0 0
g(y) = c(h(y)) 6= c(h(y )) = g(y ). Therefore g is 1-1.
Finally, c(h(y)) ∈ h(y) = f −1 ({y}) gives f (g(y)) = f (c(h(y))) = y .
Hence f ◦ g = 1Y . 

Recall that we dened |A| ≤ |B| to mean the existence of a 1-1 function
f : A → B.

Task 4.3.2 Show that |A| ≤ |B| if and only if there is an onto function
g : B → A.

Task 4.3.3 Show that if B = N then the proof of Task 4.3.2 does not require
the Axiom of Choice. (Hint: Use the Well-Ordering Principle instead of
Axiom of Choice)

Now we shall see that the Axiom of Choice helps us complete our earlier
descriptions of nite and innite sets.

Theorem 4.3.4 If A is an innite set then there is a 1-1 map from N to


A.(Hence |N| ≤ |A|)

Pick an element a ∈ A. {a} is nite, so is proper in A. Therefore


The set
we can pick a dierent element b ∈ A \ {a}. {a, b} is nite, so is proper in
A, and we can pick c ∈ A \ {a, b}. Because A is innite this process cannot
7 terminate, and we get an innite sequence of points in A. We can map 1
to a, 2 to b, and so on. Again, to make this idea work formally, we use the
Axiom of Choice to select the points.
178 CHAPTER 4. INFINITY

Proof: Let A = P(A) \ {∅} and let c be a choice function on A. (Thus c


selects a member from every non-empty subset of A.)
Dene f (1) = c(A). For each n ∈ N \ {1} dene f (n) = c(A \ f (Fn−1 )).
Let us understand why this denition works.

(a) It may appear cyclical to use f in its own denition. However, what
happens here is that we use the values of f (k) for k < n to dene
f (n). We use f (1) to x f (2). Then we use f (1) and f (2) to x f (3),
and we proceed in this manner. You should try to justify it formally
using the Recursion Theorem (Theorem 3.1.2).

(b) f (Fn−1 ) is nite, hence A \ f (Fn−1 ) 6= ∅ and f (n) is dened.

(c) Suppose f (n) = f (m) with n < m. Then f (n) ∈ f (Fm−1 ) and
f (m) ∈ A \ f (Fm−1 ). Hence f (n) 6= f (m). Therefore f is 1-1. 

Theorem 4.3.5 Every innite subset of N is in bijection with N.

Proof: Let A be an innite subset of N. Then the inclusion map ι : A → N,


x 7→ x, is 1-1. Further, by Theorem 4.3.4 there is a 1-1 map f : N → A.
Now apply the Schröder-Bernstein Theorem. 

Theorem 4.3.6 A set is innite if and only if it is in bijection with a proper


subset of itself.

Proof: We already know that such a bijection implies the set is innite. For
the converse, we need the Axiom of Choice, specically the implication that
if a set is innite then there is an injection from N into that set.

Let A be an innite set. Then there is a 1-1 function f : N → A. Dene


g : A → A \ {f (1)} by

x x ∈ A \ f (N)
g(x) = .
f (n + 1) x = f (n)

We leave it to you to check that g is a bijection. 

Theorem 4.3.7 A set is nite if and only if it has no bijection with a proper
subset of itself.

Proof: This is the contrapositive of the previous theorem. Note that the
implication in one direction was already established and did not need the
Axiom of Choice. 

Undergraduate Mathematics: Foundations


4.3. THE AXIOM OF CHOICE 179

These are the only implications of the Axiom of Choice that we will
develop for now. You may wonder at this stage why we made a fuss before
presenting the Axiom and these few consequences. It may seem quite like
others for, on the face of it, all it does is formalize the act of selection.

The Axiom of Choice has been shown to be a powerful tool for proving
results about all sets, or all sets satisfying some general conditions. For
example, in the study of cardinality, it further implies that any two sets A
and B will be comparable: either |A| ≤ |B| or |B| ≤ |A|. Applications like
4
this make it popular with most mathematicians.

The objections to the Axiom also relate to its power: that it is too pow-
erful, that it allows the use of actions (unrestricted selections) that we can
never come close to performing in practice. This power has led to results that
are so startling that some deem them unacceptable. Consider the Banach-
Tarski Paradox. This is a proof that, using the Axiom of Choice, one can
divide a solid sphere in R3 into a few parts and then reassemble them, us-
ing only transformations that do not change volume, into two spheres of
the same radii as the original! In other words, we can double volume, even
though we are using transformations that do not change volume. This para-
dox is possible because the Axiom of Choice allows us to construct strange
sets whose volume is not dened. Since their volume is not dened, the fact
that we use transformations that preserve volume becomes irrelevant. Thus
the Banach-Tarski Paradox does not cause any logical issues within mathe-
matics. But it does cause mathematicians to wonder if the Axiom of Choice
is too much of a good thing.

A nal twist is that in some parts of mathematics the theory is more


satisfactory if we do not use the Axiom of Choice, and instead use other
axioms which may even contradict it. We have, as it were, parallel universes
with varying rules of set theory. We can't be in two of these universes at the
same time, but we can visit and appreciate them in turn.

Exercises for Ÿ4.3

1 Let A = {{1}, {1, 2}, {1, 2, 3}}.


(a) Give an example of a choice function for A.
(b) How many choice functions are there for A?
4 In practice, many of the applications rely on a statement called Zorn's Lemma, which
is equivalent to the Axiom of Choice. We have included a description of Zorn's Lemma
and some consequences in Ÿ4.5.
180 CHAPTER 4. INFINITY

2 Let A be a nite collection of non-empty sets. Proving, without using the


Axiom of Choice, that A has a choice function.

3 Let A consist of the nite non-empty subsets of R. Give a formula for a


choice function, thus bypassing the Axiom of Choice.

4 Various collections of non-empty sets are given below, and we wish to have
a choice function for each. For which collections is it necessary to invoke the
Axiom of Choice and for which can we avoid its use?

(a) All non-empty subsets of N.


(b) All non-empty subsets of Z.
(c) All open intervals (a, b) of R with a < b.
(d) An innite collection of pairs of shoes. (Each pair has a left shoe and
a right shoe)

(e) An innite collection of pairs of socks. (Each pair has two identical
socks)

5 Prove that the Axiom of Choice is equivalent to the following:

(a) For each surjection f: X →Y there is a map g: Y → X with f ◦g =


1Y .
(b) For each function f: X →Y and surjection g: Z → Y there is a map
k: X → Z with f = g ◦ k.

6 Prove that the Axiom of Choice is equivalent to the following:

(a) If A is a set of non-empty and pairwise disjoint sets then there is a


set C that contains exactly one element from each member of A.
(b) If A is a non-empty set and { Xa | a ∈ A } is a collection
Q of non-empty
sets indexed by A, then the cartesian product a∈A Xa is non-empty

7 Show that the following are equivalent:

(a) Axiom of Choice.

(b) For every relation R there is a function F ⊆R such that dom(F ) =


dom(R).
(c) For every set A there is a function f : P(A) \ {∅} → A such that for
every non-empty B⊆A we have f (B) ∈ B .

Undergraduate Mathematics: Foundations


4.4. COUNTABLE AND UNCOUNTABLE SETS 181

8 Consider the statement (c) of the previous exercise. Was it necessary to


exclude the empty set from the domain of f?

9 Assuming the Axiom of Choice, prove the Principle of Dependent Choice:


If R is a relation on a non-empty set X with dom(R) = X then there is a
sequence (xn )n∈N in X such that xn Rxn+1 for every n ∈ N.

10 Read the graphic novel Logicomix: an epic search for truth by Apostolos
Doxiadis and Christos H. Papadimitriou [3].

4.4 Countable and Uncountable Sets


In the previous section we saw that, if we assume the Axiom of Choice, then
the set N of natural numbers is the smallest innite set. (In the sense that
if A is innite then there is a 1-1 map f : N → A)
We call a set countable if it is in bijection with a subset of N, and
countably innite if it is in bijection with N. A set that is not countable
is called uncountable.

It follows from Theorem 4.3.5 that every subset of N is either nite or


countably innite.

Task 4.4.1 Show that a set is countable if and only if it is nite or countably
innite.

Task 4.4.2 Show that a set A is countable if and only if there is an onto
function f : N → A.

We have seen that Z is in bijection with N and hence Z is countably


innite.

Let us take up a few more examples. If you have completed Exercise 11


of Ÿ4.2 you know that N × N is in bijection with N and hence is countably in-
nite. Similarly, Exercise 13 of Ÿ4.2 establishes that Q+ and Q are countably
innite.

We have already mentioned the notation |A| = n for a nite set A which is
in bijection with Fn . IfA is countably innite we use the notation |A| = ℵ0 . 5
So far, we know that |N| = |Z| = |Q| = ℵ0 .
5ℵ is a Hebrew letter called `aleph' and so ℵ0 is read as `aleph-nought' or `aleph-null'.
182 CHAPTER 4. INFINITY

Theorem 4.4.3 Any countable union of countable sets is countable.

Proof: Let A be a countable set whose members are also countable sets. We
can assume that each member is non-empty.

Since A is countable, there is an onto map φ : N → A. For each m ∈ N,


φ(m) is countableSso there is an onto map ψm : N → φ(m). Now we dene a
map g : N × N → A by

g(m, n) = ψm (n).

We claim that g is onto, leaving the justication for you to work out. Com-
posing
S g with a bijection from
S N to N × N, we get an onto function from N
to A. Hence A is countable. 

Task 4.4.4 The proof of Theorem 4.4.3 uses the Axiom of Choice in one
place. Locate it.

Can we create some uncountable sets? Cantor's theorem on power sets


implies that |N| < |P(N)| and hence P(N) is uncountable. In keeping with
the pattern for nite sets, where |A| = n =⇒ |P(A)| = 2n , we denote the
cardinality of |P(N)| by 2ℵ0 .
Now we give another method for showing the uncountability of |P(N)|,
known as Cantor's diagonal method.
First we note that P(N) is in bijection with the set X consisting of
functions from N to {0, 1}. The bijection takes A⊆N to χA : N → {0, 1}
dened by

1 if n∈A
χA (n) = .
0 if n∈
/A
Now we concentrate on X. Each member f of X can be viewed as a sequence
f (1), f (2), f (3), . . . of 0s and 1s.

We easily create a 1-1 function ψ : N → X , n 7→ ψn , by dening



1 if k = n
ψn (k) = χ{n} (k) = .
0 if k 6= n

Therefore X is innite.

Now, suppose we have a function ϕ : N → X , n 7→ ϕn . We represent ϕ as


an innite array of 0s and 1s, where the rst row represents ϕ1 , the second
row represents ϕ2 , and so on:

Undergraduate Mathematics: Foundations


4.4. COUNTABLE AND UNCOUNTABLE SETS 183

ϕ1 (1) ϕ1 (2) ϕ1 (3) ϕ1 (4) · · ·

ϕ2 (1) ϕ2 (2) ϕ2 (3) ϕ2 (4) · · ·

ϕ3 (1) ϕ3 (2) ϕ3 (3) ϕ3 (4) · · ·

ϕ4 (1) ϕ4 (2) ϕ4 (3) ϕ4 (4) · · ·


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

We dene a new member g∈X by using the diagonal values (circled in the
diagram) and changing them. A value 1 is changed to 0 and a value 0 is
changed to 1. This is done by setting

g(n) = 1 − ϕn (n).

Now the function g does not equal any of the functions ϕn , since g(n) 6=
ϕn (n). Therefore the function ϕ: N → X is not onto. Hence there is no onto
function from N to X. So there is no bijection from N to X. Hence X is not
countably innite. Therefore it is uncountable.

Theorem 4.4.5 If A is innite and a∈A then |A| = |A \ {a}|.

Proof: There is a 1-1 map f : N → A. We can adjust its values so that


f (1) = a. Dene g : A → A \ {a} by

x if x∈/ f (N)
g(x) = .
f (n + 1) if x = f (n)

Then g is a bijection. 

Task 4.4.6 Show that if A is innite, and B is a nite subset of A, then


|A| = |A \ B|.

Theorem 4.4.7 If A is uncountable and B is a countable subset of A then


|A| = |A \ B|.

Proof: The case of nite B is already done (if you have completed Task
4.4.6). So we assume that B is countably innite.
184 CHAPTER 4. INFINITY

If A \ B were countable then A = B ∪ (A \ B) would be countable.


So A \ B is uncountable, in particular it is innite. So it has a countably
innite subset C . Now B ∪ C is countably innite, so there is a bijection
f : B ∪ C → C . Dene g : A → A \ B by

x if x ∈/ B∪C
g(x) = .
f (x) if x ∈ B ∪ C

Then g is a bijection. 

Cardinality of R
Theorem 4.4.8 R has the same cardinality as P(N).

Proof: We have seen earlier that R is in bijection with the interval [0, 1).
Also, we can identify P(N) with the set X of functions from N to{0, 1}.
Therefore it is enough to show that [0, 1) has the same cardinality as X .

Take any real number x ∈ [0, 1). It has a unique binary representation
x = 0.x1 x2 . . . xi is 0 or 1 and the xi 's are not eventually 1. (That
where each
is, there is no stage after which all the xi equal 1) We dene a corresponding
function fx ∈ X by fx (i) = xi .

Thus we have a 1-1 map from [0, 1) into X. Is this map onto? No, for
the constant function f (i) = 1 is not in its range. The functions that are
missing from the range are exactly those which are eventually 1, i.e. there
is some N such that for n≥N the function takes only the value 1.

Hence, we dene:

B = { f ∈ X | (∃N ∈ N)(n ≥ N =⇒ f (n) = 1) },


BN = { f ∈ X | n ≥ N =⇒ f (n) = 1 }.
S
Each BN is nite. B = N ∈N BN is a countable union of nite sets, hence
it is countable. And the mapping x 7→ fx maps [0, 1) onto X \ B . Therefore,
by Theorem 4.4.7, |X| = |X \ B| = | [0, 1) |. 

It is standard to use the notation |R| = c (c stands for `continuum'). We


have just proved that c = 2ℵ0 .
An interesting consequence is that since R is uncountable and Q is count-
able, R has the same cardinality as R \ Q. The irrationals are uncountable
and thus far more numerous than the rationals!

Undergraduate Mathematics: Foundations


4.4. COUNTABLE AND UNCOUNTABLE SETS 185

We have sets with cardinality ℵ0 and others with cardinality2ℵ0 . Is there


a set A whose cardinality is strictly in between, i.e. ℵ0 < |A| < 2ℵ0 ? Cantor
felt that there would not be such sets and this expectation became known
as the Continuum Hypothesis. The Continuum Hypothesis became one
of the main puzzles of set theory and the eorts to resolve it have led to a
much deeper understanding of axiomatic systems.

Exercises for Ÿ4.4

1 Suppose A is countable and f: A → B is a surjection. Show that B is


countable.

2 Show that a nite union of nite sets is nite.

3 The dyadic rationals have the form m/2n , m ∈ Z and n ∈ N. Show that
the set of dyadic rationals is countably innite.

4 Show that if A and B are countably innite then A∪B is countably


innite.

5 Show that if A and B are countably innite then A×B is countably


innite.

6 Prove that the set of all nite subsets of a countable set is countable.

7 Consider the set of sequences of 0's and 1's that are constant after some
entry. Show this set is countably innite.

8 Show that the following sets are countably innite:

(a) The set of polynomials with integer coecients and degree n or less.

(b) The set of polynomials with integer coecients.

(c) The set of real numbers which are roots of polynomials with integer
coecients. (Such numbers are called algebraic numbers )

9 A real number is called transcendental if it is not the root of any poly-


nomial with integer coecients. Show that the collection of transcendental
numbers is uncountable.

10 Show that |P(N)| = |P(N) × P(N)|.


186 CHAPTER 4. INFINITY

11 Show that |R| = |R × R| = |N × R|.

12 Consider the equivalence relation on R dened by x∼y if x − y ∈ Q.


Let X be the set of equivalence classes for this relation. Is X countable?

13 Let A be a countable subset of R. Let S be the set of all nite sums of


distinct elements of A. Give two examples of A such that S is bounded in
one case and unbounded in the other.

14 Let A be an uncountable subset of R. Let S be the set of all nite sums


of distinct elements of A. Show that S is unbounded.

15 Show that if A is an innite set then |A| = |A ∪ N|.

4.5 Zorn's Lemma and Well-Ordering Theorem


Zorn's Lemma

A subset of a partially ordered set is called a chain if it is totally ordered. For


example, in the order depicted below, the subset {∅, {2}, {1, 2}, {1, 2, 3}}
is a chain.

{1, 2, 3}

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

{1} {2} {3}

Task 4.5.1 Show that

(a) For any partially ordered set, the empty set is a chain.

(b) Every subset of a chain is a chain.

(c) If C is a chain and m is an upper bound of C then C ∪ {m} is a chain.

Undergraduate Mathematics: Foundations


4.5. ZORN'S LEMMA AND WELL-ORDERING THEOREM 187

(d) If C and D are chains in X and every element in C is less than every
element in D then C ∪ D is a chain.

We now state and prove a result known as Zorn's Lemma. The proof
requires the Axiom of Choice. Zorn's Lemma is often the means through
which the Axiom of Choice is applied to a proof. It is suitable for showing
the existence of objects whose construction can be visualized in a piece-by-
piece manner.

Theorem 4.5.2 (Zorn's Lemma) Let X be a non-empty partially ordered


set. If every chain in X has an upper bound then X has a maximal element.

6
Proof: Let us assume that X has no maximal element and then strive to
reach a contradiction.

Let C X . It has an upper bound u. Since u is not


be any chain in
maximal, ∃m ∈ X u < m. Then m is an upper bound of C but
such that
does not belong to C . Let Cb be the set of upper bounds of C which do not
belong to it. We have just shown that C b is non-empty. We use the Axiom
of Choice to select an element fC ∈ Cb for each C .

We now dene an f -chain to be a non-empty chain K such that if C ⊆ K


and b ∩K 6= ∅ then fC is minimal in C
C b ∩K . For example, {f∅ } is an f -chain.
In fact f∅ is the least element of every f -chain.

If K is an f -chain, then K ∪ {fK } is a larger f -chain. If we can show


that there is a largest f -chain K0 then we will have reached a contradiction.

The rst step is to show that, given two f -chains, their union is a chain.
Thus, let K and K 0 be f -chains with K K 0 . We have K∪K 0 = K∪(K 0 \K).
Let x ∈ K 0 \ K . Dene
0

C = {x ∈ K ∩ K 0 : x < x0 }.

Then C ⊆ K 0 and x0 ∈ C b ∩ K 0 . Hence fC is minimal in C


b ∩ K 0 . Now
Cb ∩ K = ∅, for otherwise fC is in K , hence in C , which contradicts its
choice. Therefore any x ∈ K is not an external upper bound for C : ∃c ∈ C
0
such that x ≤ c. But then x ≤ c < x . So every element of K is less than
0 0
every element of K \ K . Therefore K ∪ K is a chain.

6 This proof is due to J. D. Weston, A short proof of Zorn's Lemma, Archiv der Math-
ematik, October 1957.
188 CHAPTER 4. INFINITY

K0 = { K ∈ P(X) | K is an f -chain } is a
S
It follows immediately that
0 0
chain. Further, if x, x ∈ K0 with x < x, then for any f -chain K , x ∈ K
0 0 0 0 0 0
implies x ∈ K . (Let x belong to an f -chain K . If x ∈ K \ K then x < x ,
a contradiction)

We will now show that K0 is an f -chain. Consider a chain C ⊆ K0


such that b ∩ K0 =
C 6 ∅. Then C b ∩ K 6= ∅ for some f -chain K . By our
observation about K0 , since K contains an upper bound of C , it contains all
of C : C ⊆ K . Therefore fC is minimal in C b ∩ K . Hence fC is minimal in
b ∩ K0 . (Since x < fC implies x ∈ K )
C
Thus K0 is the largest f -chain and we have the desired contradiction. 
We have just used the Axiom of Choice to prove Zorn's Lemma. In fact,
they are equivalent:

Theorem 4.5.3 Zorn's Lemma implies the Axiom of Choice.

This is a prototypical example of how Zorn's Lemma is used in existence


7 proofs. The main ideas of the proof are provided below. Several details
are left for you to ll in.

Proof: Let A be a set whose members are non-empty sets.

Let X be the set consisting of pairs (B, fB ) where B ⊆ A, and fB : B →


∪B is a choice function. Dene a relation on X by (B, fB ) ≤ (C, fC ) if
B ⊆ C and fC (b) = fB (b) for every b ∈ B . At this stage, we need to check
the following:

(a) That X is a set (use the Axiom of Separation). ’


(b) That X is non-empty. ’
(c) That ≤ denes a partial order on X. ’
Now let C ⊆ A be a chain. We need to S show the existence of an upper bound
(B, fB ) of C . First, we dene B = { C ∈ P(X) | (∃fC )((C, fC ) ∈ C) }.
Then, if (C, fC ) ∈ C and x ∈ C , we dene fB (x) = fC (x). We have to check:

(a) That fB is well-dened. The issue is that if x ∈ B then there may


be more than one C such that x∈C and (C, fC ) ∈ C. Will they all
give the same value of fC (b)? ’
(b) That fB is a choice function on B. ’
(c) That (B, fB ) is an upper bound of C. ’

Undergraduate Mathematics: Foundations


4.5. ZORN'S LEMMA AND WELL-ORDERING THEOREM 189

We have found an upper bound for every chain. Hence, by Zorn's Lemma,
X has a maximal element (M, fM ). We need to show that M = A. If not,
S a ∈ A \ M.
there is a set Pick an element ta of a. Dene N = M ∪ {a} and
fN : N → N by 
fM (x) x ∈ M
fN (x) = .
ta x=a
Then (M, fM ) < (N, fN ), which is a contradiction. Hence M =A and fM
is a choice function on A. 

Here is an application of Zorn's Lemma. Given two sets A and B we


wonder if they can be compared in terms of cardinality. Will one of |A| ≤ |B|
and |B| ≤ |A| hold? Zorn's Lemma gives a positive answer.

Theorem 4.5.4 (Comparability of Cardinality) Let A and B be two


sets. Then one of |A| ≤ |B| and |B| ≤ |A| is true.

Proof: We'll show that if |A| ≤ |B| is false then |B| ≤ |A| is true. Suppose
|A| ≤ |B| is false. Then there is no 1-1 map from A to B . Consequently
there is no onto map from B to A (This requires Axiom of Choice).

Consider the set X of pairs (C, φC ) such that C ⊆ B and φC : C → A is


1-1. The set X is non-empty, as the empty function belongs to it. Dene a
relation on X by (C, φC ) ≤ (D, φD ) if C ⊆ D and φC is the restriction of
φD | to C.
Check that ≤ denes a partial order on X. ’
Let C be a chain in X. Let U be the union of all the subsets of B that
are present in C. Dene φU : U → A by φM (x) = φC (x) if x ∈ C and
(C, φC ) ∈ C . Now we check:

(a) That φU is well-dened and 1-1, ’


(b) That (U, φU ) is an upper bound of C. ’
We have shown that every chain has an upper bound. Hence, by Zorn's
Lemma, X (M, φM ). If φM were onto A we could
has a maximal element
extend it to an onto function from B to A and reach a contradiction. So φM
is not onto A.

Now suppose M B . Choose b ∈ B \ M and a ∈ A \ φM (M ). Dene


N = M ∪ {b} and φN : N → A by φN |M = φM and φN (b) = a. Then
(N, φN ) ∈ X and (M, φM ) < (N, φN ). This contradicts the maximality of
(M, φM ). So we must have M = B and so φM is a 1-1 map from B to A. 
190 CHAPTER 4. INFINITY

Task 4.5.5 Show that the rule for comparing cardinality follows trichotomy.
That is, for any two sets A and B, exactly one of the following holds: |A| =
|B|, |A| < |B|, |B| < |A|.

Well-Ordering Theorem

A totally ordered set is called well-ordered if every non-empty subset has


a least element. Any nite totally ordered set is well-ordered. The usual
order on N makes it a well-ordered set as well. (This is the Well-Ordering
Principle, Theorem 3.2.6) On the other hand, the sets Z, Q and R are not
well-ordered by the usual orders as none of these sets has a least element.

Now, although Z is not well-ordered by the usual order, we can put a


dierent order on it so that it does become well-ordered. Take a bijection
ϕ: Z → N and dene an order ≤ϕ on Z by m ≤ϕ n if ϕ(m) ≤ ϕ(n).

Task 4.5.6 Show that Z is well-ordered by ≤ϕ .

In this way, any countable set can be well-ordered by a suitable choice of


order. But what about R? It has no bijection with N and so this technique
cannot be applied. In fact, nobody has yet come up with a way to make R a
well-ordered set! This makes the following theorem all the more remarkable
(or unreasonable?):

Theorem 4.5.7 (Well-Ordering Theorem) Every set has a total order


that makes it a well-ordered set.

Finite subsets can be well-ordered. We use Zorn's Lemma to patch these


up into a single well-ordering of the whole set. The scheme is similar to the
earlier applications and we just give the main steps. You should provide
7 proofs of the following claims: (a ) That ≤ is a partial order on X , (b ) The
existence of (B, ≤B ) in Step 3, (c ) That (U, ≤U ) ∈ X and is an upper
bound of C , and (d ) That M A leads to a contradiction.

Proof: Let A be a set. We shall show that it has a total order that makes it
well-ordered.

1. Let X be the set of pairs (B, ≤B ) such that B ⊆ A and ≤B is a


well-ordering of B. The set X is non-empty because nite sets can be
well-ordered.

2. Dene a partial order on X by (B, ≤B ) ≤ (C, ≤C ) if

Undergraduate Mathematics: Foundations


4.5. ZORN'S LEMMA AND WELL-ORDERING THEOREM 191

(a) B ⊆ C,
(b) For any x, y ∈ B we have x ≤B y ⇐⇒ x ≤C y , and
(c) If x ∈ B , y ∈ C , y ≤C x, then y ∈ B .

{ B ∈ P(A) | (∃ ≤B )((B, ≤B ) ∈ C) }.
S
3. If C is a chain in X , dene U =
Note that if x, y ∈ U then there is a (B, ≤B ) ∈ C such that x, y ∈ B .
Dene x ≤U y if x ≤B y .

4. Then (U, ≤U ) ∈ X and is an upper bound of C.


5. By Zorn's Lemma, X has a maximal element (M, ≤M ).
6. Assume M A and reach a contradiction. Hence M =A and ≤M is
a well-ordering of A. 

Like Zorn's Lemma, the Well-Ordering Theorem too implies the Axiom
of Choice. Thus these three form a single package once the ZF axioms are
assumed.

Theorem 4.5.8 The Well-Ordering Theorem implies the Axiom of Choice.

Proof: Let A be a set whose members are non-empty sets.


S Let ≤ be a
well-ordering of A.

We have selected one partial order out of possibly many that qualify as
7 well-orderings of
S
A. However, as it is just a single selection, the Axiom
of Choice is not required.

S
Dene c: A → A by letting c(A) be the least element of A with respect
to ≤. 

Exercises for Ÿ4.5

1 Give an example of a partially ordered set that has a maximal element


but for which it is not true that every chain has an upper bound.

2 Give an example of a partially ordered set in which every chain has an


upper bound but the set does not have a greatest element.

3 Prove that Zorn's Lemma is equivalent to Hausdor 's Maximal Principle :


Every partially ordered set has a maximal chain.
192 CHAPTER 4. INFINITY

4 Prove that Hausdor 's Maximal Principle is equivalent to the following


statement: Every chain in a partially ordered set is contained in a maximal
chain.

5 Which of the following totally ordered sets are well-ordered?

(a) W with the usual order.

(b) {0, 1} × N with the dictionary order.

(c) N×N with the dictionary order.

(d) (0, 1) × (0, 1) with the dictionary order.

6 Provide a well-ordering of Q.

7 X be a well-ordered set
Prove the Principle of Transnite Induction : Let
and let S ⊆ X have the property that whenever { x ∈ X | x < s } is a subset
of S then s ∈ S . Then S = X .

8 Suppose that X is a totally ordered set but is not well-ordered. Prove


that there is a decreasing sequence of elements in X : x1 > x2 > x3 > · · ·

9 Suppose that X is a partially ordered set in which every non-empty count-


able subset has a least element. Show that X is well-ordered.

10 Let X and Y be well-ordered sets. Show that X ×Y is also well-ordered


if it is given the dictionary order.

11 Let A be an innite set. Show that |A × N| = |A|.

Undergraduate Mathematics: Foundations


A | Supplements

A.1 Truth and Provability


When we described statements or propositions in Ÿ1.1 we said they are as-
sertions to which we can assign a truth value, either T (true) or F (false).
We said only one of these values can be assigned, but also that sometimes
we may not know which one to assign. In this section, we explore the deeper
reasons behind this qualication.

We must begin by considering the signicance of the words `true' and


`false'. Their import in mathematics has to be dierent from their usage
in daily life, for the claims of mathematics are purely in the world of ideas.
Consider the question of whether the statement There is a set which has
no members is true or false. Should we set out, armed with our senses and
assorted devices, to look for evidence which would settle this question? This
will not help, for such a set is purely a mental construct. The issue is whether
it is a mental construct that can or should be included in mathematics. In
this sense, the truth or falsity of the statement at hand is chosen by us.
However, we do not have complete freedom while choosing. We have to be
consistent with our other choices of what is allowed in mathematics.

For example, in our presentation of the Zermelo-Fraenkel or ZF axioms


for set theory, the existence of the empty set is stipulated by the second
axiom. What would happen if the existence of the empty set is not directly
given? Well, as long as at least one set exists, let us call it A, and if we
also have the Axiom of Separation, we can create a set with no members as
follows:
{ x ∈ A | x 6= x }
In the second approach the existence of the empty set is a theorem. That is,
it follows from the axioms. This is what truth means to a mathematician. A

193
194 APPENDIX A. SUPPLEMENTS

mathematical statement is considered true if it is a theorem. It is considered


false if its negation is a theorem.

Thus our framework for a mathematical theory has three components.


The rst is a collection of laws of reasoning, for example the ones we de-
scribed in the rst chapter. The second is a collection of axioms. The third
is the process of applying the rules of reasoning to the axioms to derive
theorems. To carry out this process we must have a clear way of identify-
ing whether a given statement is an axiom. If there are only nitely many
axioms, this is easy. If there are innitely many axioms (as in ZF) then
we need an algorithm, a mechanical procedure, which can establish in nite
steps whether a given statement is an axiom. If a theory has a set of axioms
with such an algorithm, we say that it has an axiomatization. Further,
we say a statement φ is derivable from a collection S of statements if there
is a nite sequence φ1 , . . . , φn such that each φi is either a member of S
or is obtained from the earlier members of the list via one of the rules of
1
reasoning.

This raises the question of whether a statement can fall through the
cracks. Can a statement be such that neither it nor its negation is derivable?

As our rst example, let us take the dening properties of an integral


domain (page 138) as our axioms. The statement 1+1 = 0 cannot be
derivable from these axioms since it is false in the integral domain Z. Nor is
its negation derivable, since 1+1=0 is true in the integral domain Z2 .
Next, consider the statement There is a set which is a member of itself .
This can be neither proved nor disproved from the rst six ZF axioms. The
Axiom of Regularity had to be introduced to prevent such a set existing.
This also illustrates a general trend. If we come across such a statement,
which can be neither proved nor disproved, we seek to add axioms which
would settle the issue. This can lead to controversy, for there may not be
unanimity on whether we desire the statement to be true or false!

The Axiom of Choice is a famous instance. We have already remarked


that the remarkable consequences of this statement make it popular with
most mathematicians yet repugnant to some. But is it true or false? In
1938, Kurt Gödel showed that it is consistent with the ZF axioms. So it
became possible to add it to these axioms and create a new system, called
ZFC. But then, in 1963, Paul Cohen showed that it is independent of the
ZF axioms (that is, it is not provable from them). Thus it is entirely our

1 For a detailed description see Ÿ5.2 of Goldrei [4].

Undergraduate Mathematics: Foundations


A.2. SETS AND CLASSES 195

choice whether to consider it true or not, in the sense that logic places no
restrictions on it.

Another instance is the Continuum Hypothesis, which asserts that there


is no set whose cardinality is strictly between those of the natural numbers
and the real numbers. Again, Gödel showed that the Continuum Hypothesis
is consistent with the ZF axioms while Cohen showed it is independent. It
is still a matter of argument whether ZF should be expanded so that the
Continuum Hypothesis becomes true, or so that it becomes false. Unlike
the Axiom of Choice, the Continuum Hypothesis has no impact on everyday
mathematics and so this debate has not raised the same passions.

As we have seen over the course of this book, the natural numbers N are
fundamental to mathematics. So we would very much want an axiomatiza-
tion of N that has the following properties:

• It is consistent: There is no statement P made in the language of N


such that both P and ¬P are derivable.

• It is complete: For every statement P made in the language of N,


either P or ¬P is derivable.

• The standard properties of N are derivable.

A very remarkable achievement was the proof given by Gödel in 1930 that
this is impossible! His First Incompleteness Theorem implies that any con-
sistent axiomatization that can prove the standard properties of N would be
incomplete.

And this is why in mathematics a statement is not something which is


either true or false but rather something for which it is meaningful to ask
whether it is true or false.

These very brief remarks on the subtle topic of provability should not be
taken as being complete, in any sense. If the issues mentioned here pique
your interest you can nd more details in books on symbolic or mathematical
logic, such as Goldrei [4], Leary and Kristensen [18] or Wilder [24].

A.2 Sets and Classes


The Zermelo-Fraenkel or ZF axioms for set theory, which we presented in
Ÿ1.5 and 4.1, seek to eliminate contradictions such as Russell's Paradox by
restricting the kinds of collections that are allowed to be used in mathematics.
This is primarily achieved by the Axiom of Regularity which, for example,
196 APPENDIX A. SUPPLEMENTS

rules out the mention of such things as `the set of all sets'. Many other
conceivable collections are also ruled out as sets. For example, we cannot
have `the set of all sets with exactly two elements'. For if this was a set
then the union over its members would be a set, and that union is the entire
universal domain B. Since any set with two members can be viewed as a
eld (set one member as `0' and the other as `1') we also cannot have `the
set of all elds'.

While this approach does eliminate the known paradoxes, it also causes
inconvenience since we often wish to make statements about arbitrary sets
of a common type. For example, consider the statement that the additive
identity of any eld is unique. If we were allowed to use the collection F of
all elds then we could easily express this symbolically using quantiers as

(∀F ∈ F)(∀a ∈ F)(∀b ∈ F)[(∀x ∈ F)(a + x = b + x = x) =⇒ a = b].

The scientist John von Neumann proposed a larger system that encom-
passes both sets as well as certain useful collections that do not qualify as
2
sets and whose use is correspondingly restricted. His system was recast and
simplied by Paul Bernays and Kurt Gödel, and the resulting collection of
axioms is known as the von Neumann-Bernays-Gödel or NBG system. In
the NBG system, our starting point is a universe of things called classes
together with a relationship called membership. A class is called a set if
it is a member of some class. A class which is not a set is called a proper
class.
We'll not give a systematic development of the NBG system. If you wish
to see one, we recommend Pinter [9]. For our part, we merely present some
of the more striking facts. Some of these are axioms, others are theorems. If
they whet your appetite or arouse your suspicions, do consult Pinter for the
full story.

First, the sets of NBG are the same as the sets of ZF, in that they follow
the same axioms. If a theorem about sets is true in ZF then it is also true
in NBG and vice-versa. Thus NBG can be seen as an enlargement of ZF.

Just as the ZF axioms allow the construction of sets in various ways, the
NBG axioms allow various constructions of classes in general. The rst NBG
axiom for classes describes equality:

2 John von Neumann (190357) made fundamental contributions to Mathematics,


Physics and Computer Science! As happened with Kabir, two communities (Mathemati-
cians and Physicists, in this case) squabble about which one he truly belonged to.

Undergraduate Mathematics: Foundations


A.2. SETS AND CLASSES 197

• Two classes are equal if they have the same members.


Another axiom allows us to take complements with respect to the entire
universe:

• Every class C has a complement whose members are the sets that do
not belong to C.

Together with the existence of the empty set, this implies that we have a
class whose members are all the sets. So, while we do not have a `set of all
sets' we do have the `class of all sets'. Since the class of all sets is not a set,
it is a proper class and cannot be a member of a class. Hence there is no
`class of all classes'.

More generally, any property can be used to dene a class:

• Let P (x) be a property of sets. Then there is a class whose members are
the sets that satisfy this property. We denote this class by { x | P (x) }.

This fact is reminiscent of the Axiom of Separation in ZF. In diers in not


requiring the elements to be chosen from a set, but then it does not neces-
sarily produce a set. With its assistance we can set up classes such as `the
class of all sets with two elements' or `the class of all elds'. It can also
be used to extend the standard machinery of sets to classes: union, inter-
section, cartesian product, relations and functions can all be dened with
classes substituting for sets. The following is needed so that functions have
good behaviour:

• (Axiom of Replacement) A function maps sets to sets.

In the NBG context one may choose to deploy a stronger version of the
Axiom of Choice:

• (Axiom of Global Choice) The class of non-empty sets has a choice


function.

The extension of sets to classes has been crucial to the development of


an area of mathematics called Category theory. While Set theory gives ways
of constructing mathematical objects, Category theory concentrates on the
relationships between them. It has provided a language and a way of thinking
that has created fresh connections between dierent realms of mathematics.
198 APPENDIX A. SUPPLEMENTS

A.3 Constructing the Real Numbers


In Ÿ3.6 we dened the real numbers as a complete ordered eld, leaving
open two questions. The rst is whether any complete ordered eld exists
at all. The second is whether there might be multiple such elds, leading
to confusion about just what is a real number. We faced exactly the same
situation with the natural numbers in Chapters 3 and 4. We rst gave
the Peano axioms and used them to develop the properties of the natural
numbers. Only later did we show how to create a set that denitely satises
the Peano axioms, followed by a proof that all Peano systems are isomorphic.

We present a minor modication of Dedekind's construction of the real


3
numbers. In our approach, we shall rst use certain sets of rational numbers
to represent positive real numbers. Then we shall expand to the full set of
real numbers using the same technique that created the integers from the
natural numbers.

First, we dene a cut to be a subset A of Q+ with the following proper-


4, 5
ties:

1. A is non-empty and bounded above.

2. A has no maximum element.

3. If x ∈ Q+ and y∈A and x<y then x ∈ A.


For example, each q ∈ Q denes a corresponding cut hqi = { x ∈ Q+ | x < q }.

Task A.3.1 Show that a cut has the form hqi if and only if it has a supre-
mum (in Q).

Task A.3.2 Show that A = { x ∈ Q+ | x2 < 2 } is a cut which is not of the


form hqi.
3 See R. Dedekind, Essays on the Theory of Numbers, translated by W. W. Beman,
The Open Court Publishing Company, 1901 http://www.gutenberg.org/2/1/0/1/21016/.
Dedekind has given a detailed description of how he was led to his approach.
4 A good case can be made out for speaking not of the Dedekind cut but of the Eudoxian
cut. The crux of Dedekind's procedure was to use cuts as a working denition of real
numbers; and this is what Eudoxus had done, over two thousand years before. Edwin
Moise, Elementary Geometry from an Advanced Standpoint, 3rd ed., Addison-Wesley,
1990. Moise describes how the Greeks essentially developed the real numbers as cuts.
5 Our exposition is based on the one by E. Landau [7]. Following Landau, we rst
develop the positive real numbers as cuts and then add on zero and the negative real
numbers, although our extension beyond the positive reals uses a dierent technique.

Undergraduate Mathematics: Foundations


A.3. CONSTRUCTING THE REAL NUMBERS 199

Task A.3.3 A be a
Let cut and ε ∈ Q+ . Show that ∃a ∈ A such that
x > a + ε implies x ∈
/ A.

Let R be the set of all cuts. We equip it with binary operations + (addition)
and • (multiplication):

A + B = {a + b | a ∈ A and b∈B}
A B = { ab | a ∈ A
• and b∈B}

Let us check that A + B is indeed a cut. First, since A and B are non-empty,
we have an element a ∈ A and an element b ∈ B . Then a + b ∈ A + B and so
A + B is non-empty. Second, if a is an upper bound of A and b is an upper
bound of B then a+b is an upper bound of A+B . Third, if x = max(A+B),
let x = a + b with a ∈ A and b ∈ B . Then

a0 ∈ A =⇒ a0 + b ∈ A + B =⇒ x = a + b ≥ a0 + b
=⇒ a ≥ a0 =⇒ a = max(A)

yields a contradiction. Finally, let 0<x<y and y ∈ A + B. Then y = a+b


with a∈A and b ∈ B. Hence,

ax bx ax bx
<a and < b =⇒ ∈ A and ∈B
a+b a+b a+b a+b
ax bx
=⇒ x = + ∈ A + B.
a+b a+b
We can similarly check that A•B is a cut.

Task A.3.4 Show that if p, q ∈ Q then hpi+hqi = hp+qi and hpi • hqi = hpqi.

The binary operations on R have the following properties:

DC1. Addition is commutative and associative. ’


DC2. Multiplication is commutative and associative. ’
DC3. Multiplication distributes over addition. ’
DC4. The cut h1i is the multiplicative identity. ’
DC5. Each A∈R has a multiplicative inverse given by ’
A−1 = { x ∈ Q+ | x−1 ∈
/ A ∪ {sup A} }.
200 APPENDIX A. SUPPLEMENTS

Next, we dene a relation on R by setting A≤B if A ⊆ B. We have the


following:

DC6. The relation ≤ is a total order on R. ’


DC7. If B<C then A + B < A + C. ’
DC8. If B<C then A • B < A • C. ’
DC9. If B<C then ∃A ∈ R such that A + B = C. ’
From DC68 we get the cancellation laws:

DC10. A + B = A + C =⇒ B = C . ’
DC11. A • B = A • C =⇒ B • C . ’
Finally, we have the existence of supremum and inmum:

DC12. If A⊆R is non-empty and bounded above, it has a supremum.

DC13. If A⊆R is non-empty and bounded below, it has an inmum.

Let us prove DC12 and DC13. Suppose


S A ⊆ R is non-empty and bounded
above. Let M= A. M is a cut. It is clearly non-empty.
First we show that
Let A be an upper bound of A in R , and let a be an upper bound of A in
Q+ . Then a is an upper bound of M . If x ∈ M then x ∈ A for some A ∈ A.
Then ∃y ∈ A with x < y . But then y ∈ M . So x is not the maximum of
M . Finally, suppose x ∈ Q+ and x < y for some y ∈ M . Then y ∈ A for
some A ∈ A. Since A is a cut, we get x ∈ A and hence x ∈ M . This shows
M ∈ R . It is easy to see that M is the least upper bound of A. This proves
DC12.

For DC13, suppose


T T A ⊆ R is non-empty and bounded below. Then
L= A \ {max( A)} is the greatest lower bound of A.
Dene a relation ∼ on R×R by (A, B) ∼ (C, D) if A + D = B + C . It
is easy to check that this is an equivalence relation. Let R be the set of its
equivalence classes. The members of R will be called real numbers.

We denote the equivalence class of (A, B) ∈ R × R by (A, B). Then we


equip R with addition and multiplication:

(A, B) + (C, D) = (A + C, B + D)
(A, B) • (C, D) = (A • C + B • D, A • D + B • C)

This is the same process by which we created Z out of N. The arguments


we gave then also apply now, and lead to the following for R:

Undergraduate Mathematics: Foundations


A.3. CONSTRUCTING THE REAL NUMBERS 201

R1. Addition and multiplication are well-dened.

R2. Addition is commutative and associative.

R3. The element (h1i, h1i) is the additive identity. (We call it zero and
denote it by 0.)
R4. The additive inverse of (A, B) is (B, A). We denote the additive inverse
by −(A, B).
R5. Multiplication is commutative and associative.

R6. Multiplication distributes over addition.

R7. The element (h2i, h1i) is the multiplicative identity. (We call it unity
or one and denote it by 1.)
R8. If (A, B) 6= 0 then it has a multiplicative inverse.

In comparison with the facts for Z, R8 is new and so we should prove it.
First we note that if (A, B) 6= 0 then A 6= B . If A > B then A = B + C
for some C ∈ R . Let C be the multiplicative inverse of C in R . Then
−1

(C −1 + h1i, h1i) is the multiplicative inverse of (A, B) in R. If B > A then


B = A + D and (h1i, D−1 + h1i) is the multiplicative inverse.
At this point, we have established that R is a eld. Now we dene the
positive real numbers by

R+ = { (A, B) | B ≤ A }.

The additive inverses of the positive real numbers are called negative and
their set is denoted by R− .
R9. R = R+ ∪ {0} ∪ R− is a disjoint union.

R10. R+ is closed under addition and multiplication.

Thus R becomes an ordered eld.

Theorem A.3.5 R is a complete ordered eld.

Proof: We already know that R is an ordered eld. The subset R+ of positive


elements is in bijection with the set R of cuts via the map φ : R → R+
dened by A 7→ (A + h1i, h1i). This map also respects the binary operations
of addition and multiplication, as well as the respective orders on R and R+ :

φ(A + B) = φ(A) + φ(B), φ(A • B) = φ(A) • φ(B), A < B =⇒ φ(A) < φ(B).
202 APPENDIX A. SUPPLEMENTS

Therefore R+ has the same order properties as R. In particular, if a non-


empty subset of R+ is bounded above (below) in R+ then it has a supremum
(inmum).

A ⊆ R which is bounded above. If A∩R+ 6=


Now consider any non-empty
∅ then A and A ∩ R have the same upper bounds. Since A ∩ R+ has a
+
+
least upper bound, so does A. If A ∩ R = ∅ then we can assume that
0 is not the supremum of A. In that case, A has a negative upper bound
b. Then −A = { −x | x ∈ A } is a subset of R+ and is bounded below by
−b ∈ R+ . Hence −A has an inmum. Hence A has a supremum, in fact
sup(A) = − inf(−A). 
We say two ordered elds F and K are isomorphic if there is a bijection
φ:F→K such that for every x, y ∈ F:

φ(x+y) = φ(x)+φ(y), φ(x • y) = φ(x) • φ(y), and x < y =⇒ φ(x) < φ(y).

The bijection φ is called an isomorphism of the ordered elds.


Theorem A.3.6 Any two complete ordered elds are isomorphic.

Proof: We give a very brief sketch of the proof, as readers who have persisted
this far can likely ll in the details themselves! If help is needed, you can
consult Mendelson [8].

To begin, recall that every ordered eld contains an isomorphic copy of


Q. (Theorem 3.5.11 and Exercises 13 and 14 of Ÿ3.5.) Let F and K be two
complete ordered elds. Let QF and QK be the isomorphic copies of Q in
F and K respectively. Then there is an isomorphism θ : QF → QK of the
ordered elds QF and QK . We use it to dene a function ψ : F → K as
follows:
ψ(x) = sup{ θ(q) | q ∈ QK and q < x }.
The function ψ is an isomorphism of the ordered elds F and K. 

Undergraduate Mathematics: Foundations


B | Solutions
Ÿ1.1

1 We are required to express the given statements symbolically, using logical


operations and standard mathematical notation.

(a) Two plus two equals four and ve.

This compound statement is made of two statements `Two plus two


equals four' and `Two plus two equals ve' connected by a conjunc-
tion. So it can be expressed symbolically as (2 + 2 = 4) ∧ (2 + 2 = 5).

One should not just translate into symbols as one reads. That may result
« in 2 + 2 = 4 ∧ 5, which would not make sense as `5' is not a statement.

(b) Two plus two equals four or ve.

Here, the two statements are connected by a disjunction. The sym-


bolic expression becomes (2 + 2 = 4) ∨ (2 + 2 = 5).
(c) It is false that two plus two equals ve.

This is the negation of `Two plus two equals ve'. So its symbolic
expression is ¬(2 + 2 = 5).
(d) It is false that two plus two does not equal ve.

This is the negation of `Two plus two does not equal ve', which in
turn is the negation of `Two plus two equals ve'. Hence it can be
expressed as ¬(2 + 2 6= 5) or as ¬(¬(2 + 2 = 5)).
3 If the given statement had been simply `One of P , Q is true', the interpre-
tation would have been that at least one of them is true, and the symbolic
version would be P ∨ Q. The use of `Exactly one' means we have to exclude
the case when both are true. This can be done by incorporating the negation

203
204 APPENDIX B. SOLUTIONS

of P ∧Q as follows:
(P ∨ Q) ∧ ¬(P ∧ Q)
Another way to think about the statement is that it allows for two possibil-
ities: P is true and Q is false, or else P is false and Q is true. Thus we can
also express it as:
(P ∧ ¬Q) ∨ (¬P ∧ Q)

5 Suppose the two statement variables are P and Q. A truth table for some
combination P ∗Q will have the following format:

P Q P ∗Q
T T
T F
F T
F F

Each position in the third column has two possible entries (T and F), hence
the number of possible tables is 24 = 16. We can study how to generate
them by considering various cases.

1. P ∧ ¬P can never be true, so it gives the table where all the P ∗Q


entries are false.

2. P ∨ ¬P is always true, so it gives the table where all the P ∗Q entries


are true.

3. P ∧ Q gives a table where exactly one P ∗ Q entry is true. By replacing


P and/or Q by their negations we can generate all four such tables.
4. P ∨ Q gives a table where exactly one P ∗ Q entry is false. By replacing
P and/or Q by their negations we can generate all four such tables.
This leaves six cases. In these cases the P ∗Q column has two T entries and
two F entries. Four of these cases are covered by the statements P , Q, ¬P
and ¬Q. The remaining two cases are still left as an exercise for you!

7 We are asked whether P , Q and R can be allocated truth values so that


the given expressions have certain values. A mechanical approach would be
to create the truth tables for the two expressions and then see if the desired
combinations ever happen together. But that means lling 8 rows! Let us
try analysis instead.

Before we begin, let us also note that in this context `but' is to be taken
as `and'.

Undergraduate Mathematics: Foundations


205

(a) (P ∨ Q) ∧ R is false but P ∨ (Q ∧ R) is true.

If P P ∨ (Q ∧ R) is true regardless of what truth values


is true then
are given to Q R. Whereas, (P ∨ Q) ∧ R can be made false by
and
setting R to be false. So the aim can be achieved by setting P to T,
R to F, and giving Q either value.
(b) (P ∨ Q) ∧ R is true but P ∨ (Q ∧ R) is false.

For (P ∨ Q) ∧ R to be true we must set theR value to T. Once R is


set to T, we must set P and Q to F in order to make P ∨ (Q ∧ R)
false. But then (P ∨ Q) ∧ R becomes false! So this combination is
impossible.

9 (a) f (x) is both continuous and dierentiable: P ∧ Q.


(b) f (x) is neither continuous nor dierentiable: ¬P ∧ ¬Q.
(c) f (x) is continuous but not dierentiable. P ∧ ¬Q.
11 The issue of well-phrased is whether the statement is framed unambigu-
ously. Here, we can group the words in dierent ways:

1. The given gure is a (triangle or a polygon) but not a square.

2. The given gure is a triangle or (a polygon but not a square).

We can see if the two versions always have the same meaning by considering
various possibilities for the given gure:

(a) If the gure is not a polygon, both are false.

(b) If the gure is a polygon but not a square, both are true.

(c) If the gure is a square, both are false.

So the statement is well-phrased.

Ÿ1.2

1 This can be answered on the basis of the truth table of P =⇒ Q:

P Q P =⇒ Q
T T T
T F F
F T T
F F T
206 APPENDIX B. SOLUTIONS

(a) Suppose P =⇒ Q and P are true. Can we conclude that Q is true?

Yes. Only in the rst row are both P =⇒ Q and P true. And in
that row Q is true.

(b) Suppose P =⇒ Q and Q are true. Can we conclude that P is true?

No. Look at the third row.

(c) Suppose P and Q are true. Can we conclude that P =⇒ Q is true?

Yes. ’
3 We have to express the following symbolically using ∧, ∨, ¬, =⇒ :
(a) If not P or not Q, then it is not the case that P or Q.
As a rst step let us put some brackets to help understand the overall
structure:

If (not P or not Q) then (it is not the case that (P or Q)).


Hence the symbolic expression is (¬P ∨ ¬Q) =⇒ (¬(P ∨ Q)).
(b) If P implies that Q implies R, then P and Q together imply R.
First insert brackets: If (P implies that (Q implies R)) then ((P and
Q together) imply R).
Hence the symbolic expression is (P =⇒ (Q =⇒ R)) =⇒ ((P ∧
Q) =⇒ R).
5 (a) The gure is a rectangle and is either a rhombus or a square : P ∧
(Q ∨ R).
(b) If the gure is a rectangle then it is either a rhombus or a square :
P =⇒ (Q ∨ R).
(c) If the gure is a rectangle and a rhombus then it is a square : (P ∧
Q) =⇒ R.
(d) The gure is a square only if it is a rectangle : R =⇒ P .
(e) If the gure is a rectangle but not a rhombus then it is not a square :
(P ∧ ¬Q) =⇒ ¬R.
7 The truth-table for (¬P ) ∨ Q is obtained below. The shaded columns
match the truth-table for P =⇒ Q.

P Q ¬P (¬P ) ∨ Q
T T F T
T F F F
F T T T
F F T T

Undergraduate Mathematics: Foundations


207

9 The truth-table for ¬Q =⇒ ¬P is obtained below:

P Q ¬Q ¬P ¬Q =⇒ ¬P
T T F F T
T F T F F
F T F T T
F F T T T

Ÿ1.3

1 See the solution to Exercise 3 of Ÿ1.1.


3 To answer whether (P ∧ Q) =⇒ (P ∨ Q) is a tautology, one approach
is to create its truth-table and see whether the nal value is always T. This
proceeds as follows:

P Q P ∧Q P ∨Q (P ∧ Q) =⇒ (P ∨ Q)
T T T T T
T F F T T
F T F T T
F F F F T

Alternately, we use analysis. The statement (P ∧ Q) =⇒ (P ∨ Q) can only


be false if it is possible for P ∧Q to be true and P ∨ Q to be false under the
same values of P and Q. Now, if P ∧ Q is true, then P is true, and hence
P ∨Q is true. So (P ∧ Q) =⇒ (P ∨ Q) can never be false.

Does either approach appeal to you more?

5 We have to check if the given statements are equivalent. As seen in other


exercises, we can use either truth-tables or analysis. For practice, let us do
the rst one by truth-table and the second one by analysis.

(a) P =⇒ (Q ∨ R) and (P =⇒ Q) ∨ (P =⇒ R).


208 APPENDIX B. SOLUTIONS

P Q R Q∨R P P P (P =⇒ Q)
=⇒ =⇒ =⇒ ∨
(Q ∨ R) Q R (P =⇒ R)
T T T T T T T T
T F T T T T T T
F T T T T T T T
F F T T T T T T
T T F T T T F T
T F F F F F F F
F T F T T T T T
F F F F T T T T

The two gray columns match exactly and so they are equivalent.

(b) (P ∨ Q) =⇒ R and (P =⇒ R) ∨ (Q =⇒ R).


IfR is true then all the implications are true, and both sides are true.
If both P and Q are false, then again all the implications are true.
So let us consider the case when P is true, Q is false, and R is false.
Then P ∨ Q is true and (P ∨ Q) =⇒ R is false. On the other hand,
P =⇒ R is false and Q =⇒ R is true, so (P =⇒ R) ∨ (Q =⇒ R)
is true. Hence, they are not equivalent.

7 (a) If not P or not Q, then it is not the case that either P or Q.


The symbolic expression for this is: (¬P ∨ ¬Q) =⇒ ¬(P ∨ Q).
Converse: ¬(P ∨ Q) =⇒ (¬P ∨ ¬Q).
Contrapositive: (P ∨ Q) =⇒ ¬(¬P ∨ ¬Q). Using de Morgan's Laws,
this can be simplied to (P ∨ Q) =⇒ (P ∧ Q).
(b) If P implies that Q implies R, then P and Q together imply R.
The symbolic expression is: [P =⇒ (Q =⇒ R)] =⇒ [(P ∧ Q) =⇒
R].
Converse: [(P ∧ Q) =⇒ R] =⇒ [P =⇒ (Q =⇒ R)].
Contrapositive: ¬[(P ∧ Q) =⇒ R] =⇒ ¬[P =⇒ (Q =⇒ R)].
(A =⇒
Incidentally, using de Morgan's Laws, and the equivalence
B) ≡ (¬A ∨ B), both sides of the implication can be simplied to
P ∧ Q ∧ ¬R. Hence this is a tautology.

9 We have to provide dierent proofs of If P =⇒ Q is true then P =⇒


(P ∧ Q) is true .

Undergraduate Mathematics: Foundations


209

(a) Direct proof: Assume P =⇒ Q is true. We have to show that


whenever P is true, then P ∧ Q is true. By our assumption, if P is
true then Q is true, and so P ∧ Q is true.

(b) Proof by contrapositive: Suppose P =⇒ (P ∧ Q) is false. Then P is


true and P ∧Q is false. Hence Q is false. Hence P =⇒ Q is false.

(c) Proof by contradiction: Suppose P =⇒ Q is true and P =⇒ (P ∧Q)


is false. The falsity of P =⇒ (P ∧ Q) establishes that P is true and
Q is false. But once P is true, the truth of P =⇒ Q means Q is
true. The two values of Q yield a contradiction.

11 It is enough to see that conjunction and disjunction can be replaced by


combinations of negation and implication. This happens as follows:

P ∨ Q ≡ ¬P =⇒ Q and P ∧ Q ≡ ¬(P =⇒ ¬Q).

Ÿ1.4

1 (a) M is an even natural number : ∃x (N (x) ∧ M = 2x).


(b) M is an odd natural number : ∃x (N (x) ∧ M = 2x − 1).
(c) M is a prime number :

N (M ) ∧ M 6= 1 ∧ [∀x ∀y [(N (x) ∧ N (y) ∧ M = xy) =⇒ (x = 1 ∨ x = M )]].

3 If there were any dragons at all, then this implication would hold. But if
there are actually no dragons, then Every dragon is red is true but There
is a red dragon is false.

5 Let I(x) be the open statement that The object x is an integer.


(a) For any integer x there is at least one integer y such that x + y = 1.
Statement: ∀x [I(x) =⇒ ∃y (I(y) ∧ x + y = 1)].
Negation: ∃x [I(x) ∧ ∀y (I(y) =⇒ x + y 6= 1)].
(b) For any integer x there is exactly one integer y such that x + y = 1.
Statement: ∀x [I(x) =⇒ ∃!y (I(y) ∧ x + y = 1)].
Negation:

∃x [I(x) ∧ ∀y (I(y) =⇒ x + y 6= 1)] ∨


∃x ∃y ∃z [I(x) ∧ I(y) ∧ I(z) ∧ y 6= z ∧ x + y = 1 ∧ x + z = 1].
Here, we have to take care of two cases of the failure of `unique
existence': one case is lack of existence and the other case is multiple
existence.
210 APPENDIX B. SOLUTIONS

7 If any pair among the following consists of equivalent statements they will
have the same value no matter what variable values x, y or open statement
P we consider. We also have to remember that the quantiers are applied
from left to right. This is important for (c).

(a) ∃x ∃y P (x, y) and ∃y ∃x P (x, y).


Equivalent. ’
(b) ∀x ∀y P (x, y) and ∀y ∀x P (x, y).
Equivalent. ’
(c) ∃x ∀y P (x, y) and ∀y ∃x P (x, y).
Not equivalent. Let P (x, y) be the statement N (x) ∧ N (y) ∧ (x > y),
where N (x) is the open statement The object x is a natural num-
ber. Then ∃x ∀y P (x, y) asserts the existence of a natural number x
that is greater than every other natural number, and is false. While
∀y ∃x P (x, y) asserts that for every natural number y there is a cor-
responding natural number x which is greater than it, and this is true
(just take x = y + 1). The order of the quantiers is crucial here. In
∃x ∀y there is a single x that works for all y . In ∀y ∃x the x can vary
with y .

9
  
(a) ∀x P (x) =⇒ (Q(x)  ∨ R(x)) ⇐⇒ ∀x (P (x) =⇒ Q(x)) ∨
∀x (P (x) =⇒ R(x)) .
Suppose x varies over the integers. Let P (x) be the statement x 6= 0,
Q(x) be x > 0, and R(x) be x < 0. Then ∀x P (x) =⇒ (Q(x) ∨
R(x)) is true but ∀x (P (x) =⇒ Q(x)) ∨ ∀x (P (x) =⇒ R(x)) is
false.
 
(b) ∀x (P (x) ∨ Q(x)) =⇒ R(x)] ⇐⇒ ∀x (P (x) =⇒ R(x)) ∨
∀x (Q(x) =⇒ R(x)) .
Suppose x varies over the integers. Let P (x) be the statement x 6= 0,
 be x > 0. Then (P (−1) ∨ Q(−1)) =⇒ R(−1) is false,
Q(x) and R(x)
and so ∀x (P (x) ∨ Q(x)) =⇒ R(x)] is false. But ∀x (Q(x) =⇒
R(x)) is true and so ∀x (P (x) =⇒ R(x)) ∨ ∀x (Q(x) =⇒ R(x)) is
true.
  
(c) ∀x P (x) =⇒ (Q(x)  ∧ R(x)) ⇐⇒ ∀x (P (x) =⇒ Q(x)) ∧
∀x (P (x) =⇒ R(x)) .
This is always true. ’
11 (a) M is free and x is bound (by ∃x).

Undergraduate Mathematics: Foundations


211

(b) M is free and x, y are bound.

(c) This exercise is intended to illustrate that it is the occurrence of a


variable that is free or bound. In ∀x (x = x), x is bound but in
∃y (y 6= x) it is free. In its bound occurrence it could be replaced
by another variable name without aecting the content of the full
expression. For example, (∀z (z = z)) ∧ (∃y (y 6= x)) would be
equivalent to the given expression.

Ÿ1.5

1 (a) Integer solutions of (x2 − 1)(x2 − 4) = 0: {1, −1, 2, −2} or {±1, ±2}.
(b) Subsets of {a, b}: {∅, {a}, {b}, {a, b}}.
(c) Prime factors of 60: {2, 3, 5}.
(d) Odd natural numbers: {1, 3, 5, 7, . . . }.
3 (a) First, use the Axiom of Pairs to create {A, B} and {C}. Use the
Axiom of Pairs again to createX = {{A, B}, {C}}. Apply the Axiom
of Union to X to obtain Y = {A, B, C}. Finally, apply the Axiom of
Union to Y to get A ∪ B ∪ C .

(b) Use the Axiom of Separation: A∩B ∩C = { x ∈ A | x ∈ B ∧ x ∈ C }.


5 We are given that A ⊆ B .
(a) A ∪ C ⊆ B ∪ C : Consider any x ∈ A ∪ C . Then x ∈ A or x ∈ C . If
x ∈ A then x ∈ B (due to A ⊆ B ). Hence, x ∈ B or x ∈ C . Therefore
x ∈ B ∪ C.
(b) A ∩ C ⊆ B ∩ C : Consider any x ∈ A ∩ C . Then x ∈ A and x ∈ C.
Since x ∈ A and A ⊆ B , we have x ∈ B . Hence, x ∈ B and x ∈ C.
Therefore x ∈ B ∩ C .

7 (a) (A ∪ B) ∩ C ⊆ A ∪ (B ∩ C):
Consider any x ∈ (A ∪ B) ∩ C . Then x ∈ A ∪ B and x ∈ C . Consider
both cases ofx ∈ A ∪ B . If x ∈ A then x ∈ A ∪ (B ∩ C). If x ∈ B
then x ∈ B ∩ C , and so again x ∈ A ∪ (B ∩ C). Hence in both cases
we have x ∈ A ∪ (B ∩ C).

(b) (A ∪ B) ∩ (A ∪ C) ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C) ∪ (B ∩ C):
212 APPENDIX B. SOLUTIONS

We shall apply the distributive law:

(A ∪ B) ∩ (A ∪ C) ∩ (B ∪ C) = [(A ∪ B) ∩ (A ∪ C)] ∩ (B ∪ C)
= [A ∪ (B ∩ C)] ∩ (B ∪ C)
= [A ∩ (B ∪ C)] ∪ [(B ∩ C) ∩ (B ∪ C)]
= [(A ∩ B) ∪ (A ∩ C)] ∪ [B ∩ C]
= (A ∩ B) ∪ (A ∩ C) ∪ (B ∩ C)

9 (a) B ⊆ A =⇒ A \ (A \ B) = B :
Since B ⊆ A, A\B is just the complement of B in A. Hence A\(A\B)
is the complement of the complement of B , i.e., B itself.
(b) A ∩ B = ∅ =⇒ A \ (B \ A) = A \ B :
Let us consider A and B as subsets of some universal set U (for
example, we could take U = A ∪ B ). Then, taking complements with
respect to U, we calculate as follows:

A \ (B \ A) = A ∩ (B ∩ Ac )c = A ∩ (B c ∪ A).

Now A∩B =∅ implies A ⊆ Bc. Hence,

A \ (B \ A) = A ∩ (B c ∪ A) = A ∩ B c = A \ B.

Ÿ1.6

1 (a) { x ∈ N | (∃n ∈ {1, 2, . . . , 10}) x = n3 }.


(b) { x ∈ N | (∃m ∈ N) x = 2m − 1 }.
(c) { x ∈ N | (x > 1) ∧ (∀n ∈ N)(n | x =⇒ (n = 1 ∨ n = x) }.
3 (a) The square of every non-zero integer is positive.

(∀n ∈ Z)(n2 > 0). True.

(b) Let U be a set and A, B , C be any subsets of U . If A and B are


disjoint, and B and C are disjoint, then A and C are disjoint.
(∀A ∈ P(U ))(∀B ∈ P(U ))(∀C ∈ P(U ))[((A ∩ B = ∅) ∧ (B ∩ C =
∅)) =⇒ (A ∩ C = ∅)]. This is false: for a counterexample, consider
A = C = {0} and B = ∅.
(c) Every positive rational number has a rational square root.

(∀x ∈ Q)[x > 0 =⇒ (∃y ∈ Q)(y 2 = x)]. False, as 2 does not have a
rational square root.

Undergraduate Mathematics: Foundations


213

5 (a) For every real number x there is a natural number n that is greater
than x.
(∀x ∈ R)(∃n ∈ N)(n > x).
Negation: (∃x ∈ R)(∀n ∈ N)(n ≤ x).
Negation in words: There is a real number that is greater than or
equal to every natural number.

(b) If m, a and b are natural numbers such that m divides ab then m


divides either a or b.
(∀m ∈ N)(∀a ∈ N)(∀b ∈ N)[(m | ab) =⇒ (m | a ∨ m | b)].
Negation: (∃m ∈ N)(∃a ∈ N)(∃b ∈ N)[m | ab ∧ m - a ∧ m - b].
Negation in words: There are natural numbers m, a, b such that m
divides ab but does not divide a or b.
Another version of the negation: A natural number may divide a
product without dividing either factor.

(c) If two lines meet at two distinct points then they are the same line.

Let P be the set of points and L be the set of lines. Then the given
statement is expressed symbolically as:

(∀` ∈ L)(∀m ∈ L)[(∃x ∈ P )(∃y ∈ P )(x 6= y ∧ {x, y} ⊆ ` ∩ m) =⇒


` = m].
Negation: (∃` ∈ L)(∃m ∈ L)[` 6= m ∧ (∃x ∈ P )(∃y ∈ P )(x 6= y ∧
{x, y} ⊆ ` ∩ m)].
Negation in words: There are two distinct lines with two distinct
points in common.

7 No. Consider (∀x ∈ Z)(x > 0 =⇒ x < 0).


9 X is the set of intervals (0, x) with x ∈ (0, ∞).
S S
(a) X = (0, ∞): First, let a ∈ X . Then aS∈ (0, x) for some x > 0.
Hence a > 0 and so a ∈ (0, ∞). This gives X ⊆ (0, ∞). Second, let
a ∈ (0,
S ∞). Then 0 < a < a + 1 implies a ∈ (0, a + 1) ∈ X and hence
S
a ∈ X . Therefore, X = (0, ∞).
T T
(b) X = ∅: Suppose a ∈ X . Then a ∈ (0, 1) implies 0 < a. Hence
0 < a/2 < a. But then a ∈/ (0, 1/2) ∈ X , a contradiction.
11 (a) { x ∈ R | x(x − 1) > 0 } = (−∞, 0) ∪ (1, ∞).
(b) { x ∈ R | x(x − 1)(x − 2) > 0 } = (0, 1) ∪ (2, ∞).
214 APPENDIX B. SOLUTIONS

13 (a)
T
F =T{ (0, 1/n)Q | n ∈ N }. We have to show that F = ∅. Let
x ∈ F . Then x ∈ (0, 1)Q and so x is a positive rational. Hence
x = p/q with p, q ∈ N. But then x ∈ / (0, 1/q), a contradiction.
S
(b) G = { (0, 1 − 1/n)Q | n ∈ N }. We have to show that S G = (0, 1)Q .
First, since each (0, 1 − 1/n)Q ⊆ (0, 1)Q , we have G ⊆ (0, 1)Q . In
the other direction, consider any x ∈ (0, 1)Q . It has the form x = p/q
with p, q ∈ N and p < q . Hence x ∈ (0, 1 − 1/(2q)).

Ÿ2.1

1 We have to show that (x, y) = {{x, ∅}, {{∅}, y}} can also dene an ordered
pair. That is, we have to show that (x, y) = (a, b) if and only if x = a and
y = b. The `if ' direction is trivial. So let us assume that (x, y) = (a, b). Then
{{x, ∅}, {{∅}, y}} = {{a, ∅}, {{∅}, b}}
Case I: {x, ∅} = {a, ∅} and {{∅}, y} = {{∅}, b}. It is immediate that
x=a and y = b.
Case II:{x, ∅} = {{∅}, b} and {{∅}, y} = {a, ∅}. Here, x = {∅} = a
and b = ∅ = y.
3 A, B, C are any sets.

(a) A × ∅ = ∅ × A = ∅:
If (x, y) ∈ A × ∅ then y ∈ ∅, a contradiction.

(b) A × B = B × A ⇐⇒ A = B or A=∅ or B = ∅:
The reverse implication is trivial. For the forward implication, sup-
pose that A × B = B × A, A 6= ∅ and B 6= ∅. We have to show
A = B. Leta ∈ A. Since B 6= ∅, we can choose some b ∈ B . Then

(a, b) ∈ A × B = B × A =⇒ a ∈ B.

Hence A ⊆ B. Similarly, B ⊆ A.
(c) Suppose A 6= ∅. Then A × B = A × C ⇐⇒ B = C :
Similar to (b). ’
5 Are the given sets of the form A × B ?
(a) X = { (x, y) ∈ R2 | xy = 0 }:
X = A × B . Then (1, 0) ∈ X =⇒ 1 ∈ A and (0, 1) ∈
Suppose
X =⇒ 1 ∈ B . Hence (1, 1) ∈ A × B = X , a contradiction.

Undergraduate Mathematics: Foundations


215

(b) Y = { (x, y) ∈ R2 | |x + y| + |x − y| ≤ 2 }:
By trying out various cases of the signs of x and y we can establish
that |x + y| + |x − y| always equals the greater among 2|x| and 2|y|.
For example, suppose x ≥ y ≥ 0. Then x±y ≥ 0 and |x+y|+|x−y| =
(x + y) + (x − y) = 2x = 2|x| = max(2|x|, 2|y|).
Hence, Y = { (x, y) ∈ R2 | |x| ≤ 1 ∧ |y| ≤ 1 } = [−1, 1] × [−1, 1].
7 We are given A ⊆ X and are asked whether this implies that (A × Y )c =
c
A ×Y. Let us consider a point (x, y) ∈ X × Y . Then,

(x, y) ∈ (A × Y )c ⇐⇒ (x, y) ∈
/ A × Y ⇐⇒ x ∈
/A
⇐⇒ x ∈ Ac ⇐⇒ (x, y) ∈ Ac × Y.

So the answer is `Yes'.

9 A ⊆ X, B ⊆ Y . We are asked whether any of the following


Let always
equal(A × B)c . We shall see that (c) and (d) qualify. For the others,
A = B = [0, 1] and X = Y = [0, 2] serve as counterexamples.
(a) Ac × B c : (0, 2) ∈ (A × B)c but / Ac × B c .
(0, 2) ∈
(b) (Ac ×B)∪(A×B c ): (2, 2) ∈ (A×B)c but / (Ac ×B)∪(A×B c ).
(2, 2) ∈
(c) Let (x, y) ∈ X × Y . Then

(x, y) ∈ (A × B)c ⇐⇒ x ∈
/ A∨y ∈
/B
⇐⇒ (x, y) ∈ Ac × Y ∨ (x, y) ∈ X × B c
⇐⇒ (x, y) ∈ (Ac × Y ) ∪ (X × B c ).

(d)

(Ac × B)∪(A × B c ) ∪ (Ac × B c )


= [(Ac × B) ∪ (Ac × B c )] ∪ [(A × B c ) ∪ (Ac × B c )]
= [Ac × (B ∪ B c )] ∪ [(A × Ac ) × B c ]
= (Ac × Y ) ∪ (X × B c ) = (A × B)c

Ÿ2.2

1 R is a relation on X .
(a) R is reexive: Then x ∈ X =⇒ xRx =⇒ x ∈ dom(R)∧x ∈ ran(R).
Hence dom(R) = ran(R) = X .
216 APPENDIX B. SOLUTIONS

(b) R is symmetric: Then x ∈ dom(X) ⇐⇒ ∃y xRy ⇐⇒ ∃y yRx ⇐⇒


x ∈ ran(R). Hence dom(R) = ran(R).
3 We have dened a relation on Z×Z by (m, n) ∼ (a, b) if mb = na. It
is easily checked to be reexive and symmetric. We'll show that it fails to
be transitive. First, observe that (0, 0) ∼ (a, b) (a, b) ∈ Z × Z. If
for every
the relation were transitive, we would have have (m, n) ∼ (0, 0) ∼ (a, b) and
hence (m, n) ∼ (a, b) for any two members (m, n) and (a, b) of Z × Z. But
(2, 1) 6∼ (1, 2).
5 We have xed k ∈ N and dened a relation on Z by m ∼ n if and only if
n−m is a multiple of k, that is, k | (m − n).
Reexive: k | 0 =⇒ k | (n − n) =⇒ n ∼ n.
Symmetric: m ∼ n =⇒ k | (n − m) =⇒ k | (m − n) =⇒ n ∼ m.
Transitive: Let m ∼ n and n ∼ p. Then k | (n − m) and k | (p − n). Hence
k | (n − m) + (p − n). Therefore k | (p − m), and so m ∼ p.
For any m ∈ Z, let [m] denote its equivalence class. Then

[m] = { n ∈ Z | (∃` ∈ Z)(n − m = `k) } = { m + `k | ` ∈ Z }

Since integers that dier by multiples of k are in the same equivalence class,
the distinct equivalence classes are [0], [1], . . . , [k − 1].
7 We are given a relation R on X such that the sets x̄ = { y ∈ X | xRy }
are pairwise disjoint and x ∈ x̄ for each x ∈ X. We shall prove that R is an
equivalence relation:

Reexive: x ∈ x̄ =⇒ xRx.
Symmetric: xRy =⇒ y ∈ x̄ =⇒ x̄ = ȳ =⇒ x ∈ ȳ =⇒ yRx.
Transitive: xRy ∧ yRz =⇒ xRy ∧ zRy =⇒ y ∈ x̄ ∩ z̄ =⇒ x̄ = z̄ =⇒ z ∈
x̄ =⇒ xRz .
9 We are asked whether the following is a correct argument that symmetry
and transitivity together imply reexivity: aRb =⇒ bRa =⇒ (aRb ∧
bRa) =⇒ aRa. In order to conclude aRa, the argument requires the
existence of some b such that aRb holds. There may not be such a b! As an
extreme example, consider the empty relation.

11 We have ordered N by divisibility. Suppose A is bounded above by a,


and B is bounded above by b. Then both A andB are bounded above by
ab. Hence A∪B is bounded above by ab.
13 We are given a totally ordered set X and subsets A, B ⊆ X .

Undergraduate Mathematics: Foundations


217

(a) If A and B are bounded above then A∪B will be bounded above.

True. If A and B are bounded above by a and b respectively, then


A∪B is bounded above by max{a, b}.
(b) If A and B have greatest elements then A∪B will have a greatest
element.

True. If A and B have greatest element a and b respectively, then


A∪B has greatest element max{a, b}.
(c) If A and B have maximal elements then A∪B will have a maximal
element.

True. In a totally ordered set, maximal elements are also greatest


elements.

Ÿ2.3

1 2 × 2 × 2 ×
1 × 1 × 1 ×
0 × 0 × 0 ×
−1 × −1 × −1 ×
−2 × −2 × −2 ×
× × × × × × × × × × × × × × ×
−2 −1 0 1 2 −2 −1 0 1 2 −2 −1 0 1 2
(a) (b) (c)

(a) Not a function. The points (0, 0) and (0, 1) violate the requirement
(xRy ∧ xRz) =⇒ y = z .
(b) This is a function. ’
(c) Not a function. There is no y such that 0Ry and so the domain is
not all of X.
3 Equivalence classes for f (x, y) = x + y : Parallel straight lines with slope
−1.
Equivalence classes for g(x, y) = x2 + y 2 : Concentric circles with centre
at origin.

5 We have I = [−1, 1]. We are asked if the following relations represent


functions.

(a) { (x, y) ∈ I × I | x2 + y 2 = 1 }: No. The points (0, ±1) satisfy the


relation but violate the requirement of unique image.
218 APPENDIX B. SOLUTIONS

(b) { (x, y) ∈ I × I | x2 + y 2 = 1 ∧ x ≥ 0 }: No. Again consider (0, ±1).


2 2
(c) { (x, y) ∈ I × I | x + y √= 1 ∧ y ≥ 0 }: Yes. For each x∈I we have
the unique image y = − 1 − x2 .
7 Functions f, g, h : R → R are dened by
x x2 x3
f (x) = , g(x) = , h(x) = .
1 + x2 1 + x2 1 + x2
(a) We have to determine which of them are injective.

x y
f (x) = f (y) =⇒ =
1 + x2 1 + y2
=⇒ x + xy 2 = y + x2 y =⇒ x − y = xy(x − y).

Hence, repeated values can occur for f when xy = 1. For example,


f (2) = f (1/2) = 2/5 and so f is not injective.

It is easy to see that g is not injective either, since its values are the
same for ±x. For example, g(1) = g(−1) = 1/2.
For h, we rst note that h(x) has the same sign as x. Hence h(x) =
h(y) implies xy ≥ 0. So,

x3 y3
h(x) = h(y) =⇒ 2
= =⇒ x3 + x3 y 2 = y 3 + x2 y 3
1+x 1 + y2
=⇒ (x − y)(x2 + xy + y 2 + x2 y 2 ) = 0 =⇒ x = y.

So the function h is injective.

(b) We have to show that f and g are not surjective.

Since f (x) = −f (−x) we can concentrate on x ≥ 0. Then,

x 1
0 ≤ (x − 1)2 = 1 + x2 − 2x =⇒ 2x ≤ 1 + x2 =⇒ ≤ .
1 + x2 2
Hence f (x) has no values greater than 1/2 and is not surjective.

The function g satises 0 ≤ g(x) < 1 and hence is not surjective.

9 We have to prove that f : X → Y is onto ⇐⇒ f (X \ A) ⊇ Y \ f (A) for


all A ⊆ X.
( =⇒ ): Suppose f is onto and A ⊆ X . Let y ∈ Y \ f (A). Then ∃x ∈ X
such thatf (x) = y . If x ∈ A then y = f (x) ∈ f (A), a contradiction. Hence,
x ∈ X \ A and y = f (x) ∈ f (X \ A).

Undergraduate Mathematics: Foundations


219

( ⇐= ): Suppose that f (X \ A) ⊇ Y \ f (A) for all A ⊆ X. If we consider


A = ∅, we get f (X) ⊇ Y , and so f (X) = Y .
11 Let f : N → A and g : N → B be onto functions. Dene h: N → A ∪ B by

f (x/2) x is even
h(x) = .
g((x + 1)/2) x is odd

Then h is onto. ’
13 Let f : R2 → R be given by f (x, y) = x2 + y2 . Pictures of f −1 ({t}) and
−1
f ([a, b]) are shown below:

f −1 ({t}) f −1 ([a, b])

15 Composition is associative:
(h ◦ (g ◦ f ))(x) = h((g ◦ f )(x)) = h(g(f (x))) = (h ◦ g)(f (x)) = ((h ◦ g) ◦ f )(x)

17 Consider a function f : X → Y .
(a) ∃g : Y → X such that f ◦ g = 1Y only if f is onto.

Suppose there is a function g : Y → X such that f ◦ g = 1Y . Now let


b ∈ Y. Dene a = g(b). Then f (a) = f (g(b)) = 1Y (b) = b, so f is
onto.

(b) ∃h : Y → X such that h ◦ f = 1X if and only if f is 1-1.

Suppose there is a function h: Y → X such that h ◦ f = 1X . Then

f (a) = f (b) =⇒ h(f (a)) = h(f (b))


=⇒ (h ◦ f )(a) = (h ◦ f )(b) =⇒ a = b,
and so f is 1-1.

Conversely, suppose f is 1-1. We dene h : Y → X as follows. If


y ∈ f (X) h(y) to be the unique pre-image of y under f .
we dene
Then we x some x0 ∈ X and for every y ∈ Y \ f (X) we dene
h(y) = x0 . We claim that h ◦ f = 1X . ’
220 APPENDIX B. SOLUTIONS

(c) If f ◦ g = 1Y and h ◦ f = 1X then g = h.


h ◦ (f ◦ g) = h ◦ 1Y = h
(h ◦ f ) ◦ g = 1X ◦ g = g
By associativity of composition, h = g.

Ÿ2.4

1 Suppose X is indexed by I and Y is indexed by J.


(a) Can we always index X ∪Y by I ∪ J?
No. If and J overlap then I ∪ J may not have enough elements
I
to index X ∪ Y . For example, suppose I = J = {1}, X = {x1 },
Y = {y1 }, and x1 6= y1 . Then I ∪ J = {1} cannot serve as an index
set for X ∪ Y = {x1 , y1 }.

(b) Can we always index X ∩Y by I ∩ J?


No. I ∩ J may be too small. For example, consider I = {1}, J = {2},
X = {x1 }, Y = {y2 }, with x1 = y2 . Then I ∩ J = ∅ cannot index
X ∩ Y = {x1 }.
3 We
[ \
have to nd Xn and Xn .
n∈N n∈N
[ \
(a) Xn = nN = { nk | k ∈ N }: Xn = N and Xn = ∅.
n∈N n∈N
[ \
(b) Xn = nZ = { nk | k ∈ Z }: Xn = Z and Xn = {0}.
n∈N n∈N

5 { Xi | i ∈ I } is an indexed set and A is any set. We have to show that:


[ [
(a) A∪( Xi ) = (A ∪ Xi ):
i∈I i∈I

[
x∈A∪( Xi ) ⇐⇒ x ∈ A ∨ (∃i ∈ I)(x ∈ Xi )
i∈I
⇐⇒ (∃i ∈ I)(x ∈ A ∨ x ∈ Xi )
[
⇐⇒ (∃i ∈ I)(x ∈ A ∪ Xi ) ⇐⇒ x ∈ (A ∪ Xi )
i∈I

\ \
(b) A∩( Xi ) = (A ∩ Xi ). ’
i∈I i∈I

Undergraduate Mathematics: Foundations


221

7 Let { Xi | i ∈ I } be an indexed set and let J ⊆ I . Show that:


[ [ [
(a) Xi ⊆ Xi : Suppose x∈ Xi . Then ∃i0 ∈ J such that x ∈ Xi0 .
i∈J i∈I i∈J
[
But J ⊆I means i0 ∈ I . Hence x∈ Xi .
i∈I
\ \
(b) Xi ⊆ Xi . ’
i∈I i∈J

9 We have f : X →Y and { Yi | i ∈ I } a family of subsets of Y. We have


to prove:
[ [
(a) f −1 ( Yi ) = f −1 (Yi ):
i∈I i∈I
[ [
x ∈ f −1 ( Yi ) ⇐⇒ f (x) ∈ Yi ⇐⇒ (∃i ∈ I)(f (x) ∈ Yi )
i∈I i∈I
[
⇐⇒ (∃i ∈ I)(x ∈ f −1 (Yi )) ⇐⇒ x ∈ f −1 (Yi )
i∈I

\ \
(b) f −1 ( Yi ) = f −1 (Yi ). ’
i∈I i∈I

11 { Xi | i ∈ I } and { Yi | i ∈ I } are indexed families of sets. We have to


show that Y Y Y
Xi ∩ Yi = (Xi ∩ Yi ).
i∈I i∈I i∈I

Letx = (xi )i∈I . Then


Y Y
x∈ Xi ∩ Yi ⇐⇒ (∀i ∈ I)(xi ∈ Xi ) ∧ (∀i ∈ I)(xi ∈ Yi )
i∈I i∈I
Y
⇐⇒ (∀i ∈ I)(xi ∈ Xi ∩ Yi ) ⇐⇒ x ∈ (Xi ∩ Yi )
i∈I

Ÿ2.5

1 Consider the following functions f, g : N → N:



1 n=1
f (n) = n + 1, g(n) = .
n−1 n≥2

Then g ◦ f = 1N and f ◦ g 6= 1N . ’
222 APPENDIX B. SOLUTIONS

3 Suppose f : X → Y and g : Y → X such that g ◦ f and f ◦ g are bijections.


Then g◦f is 1-1 implies that f is 1-1, and f ◦ g is onto implies that f is
onto. Hence f is a bijection. Similarly, g is a bijection.

5 Let a, b ∈ R with a < b. Consider the equation of the line joining (a, 0) to
(b, 1):
y−a
f (x) =
b−a
This denes a bijection f : [a, b] → [0, 1]. ’
7 We are asked to use the following diagram to derive bijections f : (−1, 1) →
R and g : R → (−1, 1).
y -axis

1
P

−1 ( x
)
1 f (x) x-axis

The coordinates of P are (x, |x|). By comparing slopes we get

|x| − 0 1−0 x
= =⇒ f (x) = .
x − f (x) 0 − f (x) 1 − |x|
y
The inverse function is g(y) = . ’
1 + |y|
9 Each plane passing through the origin is mapped to its normal line at the
origin. Specically, the plane given by Ax + By + Cz = 0 is mapped to the
x y z
line given by = = .
A B C
11 We have A, B ⊆ X , with |A| = |B|. Can we assert the following?

(a) |A| = |B| =⇒ |X \ A| = |X \ B|.


No. Consider X=A=N and B = 2N. ’
(b) |X \ A| = |X \ B| =⇒ |A| = |B|.
No. Consider X = N, A = ∅, B = {1}. ’
Note that the counter-examples have innite sets. Can there be a
counter-example with X nite?

Undergraduate Mathematics: Foundations


223


13 For any x = (a, b) ∈ R2 dene its length by kxk = a2 + b2 . Consider
2
the map f: D →R dened by

1
f (x) = x.
1 − kxk

Then f (x) is a bijection between D and R2 . ’

Ÿ2.6

1 B(X) is the set of all bijections from X to X, with composition as the


binary operation.

(a) If |X| ≤ 2 composition is commutative.

If X = {a} we have B(X) = {1X }, and 1X commutes with itself.

If X = {a, b} we have B(X) = {1X , σ}, where σ switches a and b.


Note 1X commutes with all functions, and σ of course commutes with
itself.

(b) If |X| ≥ 3 composition is not commutative.

Let X = {a, b, c}. Consider ra , rb ∈ B(X) dened by ra (a) = a,


ra (b) = c, ra (c) = b, rb (a) = c, rb (b) = b, rb (c) = a. Then ra ◦ rb (b) =
c and rb ◦ ra (b) = a, hence ra and rb do not commute.
(c) Composition is associative.

See solution of Exercise 15 in Ÿ2.3.

(d) 1X is the identity element. ’


(e) Every member of B(X) has an inverse.

If f ∈ B(X) then f is a bijection, hence has an inverse function f −1 .


This is also its inverse with respect to composition.

3 For ease of calculation, view all the sets involved as lying in a common
`universal set' U.
(a) Symmetric dierence is commutative and associative.

A4B = (A \ B) ∪ (B \ A) = B4A.

A4(B4C) =A4[(B ∩ C c ) ∪ (C ∩ B c )]
=[A ∩ (B c ∪ C) ∩ (B ∪ C c )] ∪ [(B ∩ C c ) ∪ (C ∩ B c )] ∩ Ac
=(A ∩ B c ∩ C c ) ∪ (A ∩ B ∩ C) ∪ (Ac ∩ B ∩ C c )
∪ (Ac ∩ B c ∩ C)
224 APPENDIX B. SOLUTIONS

Since the last expression is symmetric in A, B , C , it follows That we


can regroup it into symmetric dierences of A, B , C in any order we
like, in particular as (A4B)4C .

(b) The empty set is the identity element for 4. ’


(c) Every member of P(X) has an inverse.

The inverse of a subset is itself.

(d) Intersection ∩ distributes over 4. ’


5 (a) A ∗ where some elements lack a left inverse, and some elements lack
a right inverse :

Let F(N) be all the functions from N to N, with composition as the


binary operation. Then the identity function 1N serves as the identity
for the operation. Functions which are 1-1 but not onto have a left
inverse but not a right inverse. Functions which are onto but not 1-1
have a right inverse but not a left inverse.

(b) A ∗ where some elements have multiple left inverses :

Again, consider F(N) under composition. Then the member f (n) =


n+1 has multiple left inverses. ’
(c) If an element has both a left inverse and a right inverse then they are
equal :

Let a ∗ x = e and x ∗ b = e. Now (a ∗ x) ∗ b = e ∗ b = b and


a ∗ (x ∗ b) = a ∗ e = a. Then, associativity implies a = b.

7 (a) A binary operation is a function X × X → X. Here, X ×X has 4


elements. For each element we have a choice of 2 possible images.
Hence the number of binary operations is 24 = 16.

(b) We already know, from the `laws of logic' in the rst chapter, that ∧
is associative and commutative. Since T ∧F =F and T ∧ T = T, we
see that T is the identity for ∧.

(c) ’
9 The multiplication table for S3 :

Undergraduate Mathematics: Foundations


225

T0 T120 T240 R1 R2 R3
T0 T0 T120 T240 R1 R2 R3
T120 T120 T240 T0 R3 R1 R2
T240 T240 T0 T120 R2 R3 R1
R1 R1 R2 R3 T0 T120 T240
R2 R2 R3 R1 T240 T0 T120
R3 R3 R1 R2 T120 T240 T0

11 First, we have to check that we have a well-dened binary operation.


Since the denition involves a choice of representative of each equivalence
class, we need to establish that the result is independent of the choice. If
k=a and l=b then we need k + l = a + b. Now,

k=a and l = b =⇒ n | (k − a) and n | (l − b)


=⇒ n | (k − a + l − b)
=⇒ n | ((k + l) − (a + b))
=⇒ k + l = a + b

This binary operation is commutative, associative, has 0 as identity, and the


inverse of k is n − k. ’
13 Suppose the square is labeled as follows:
2 1

3 4

A symmetry σ of the square has to preserve the original relationships. For


example, σ(2) and σ(4) must be adjacent to σ(1), while σ(3) should be
diagonally opposite to it.

Let us count how many members D4 will have. We have 4 choices for
σ(1). Once σ(1) has been xed, we have two choices for σ(2). And then σ(3)
has to be opposite σ(1) while σ(4) has to be opposite σ(2). So the number
of possibilities is 4 × 2 = 8. (While the total number of bijections is 4! = 24)

The possible symmetries are listed below. Each is given as a grid of two
rows. Every number in the lower row is the image of the number directly
226 APPENDIX B. SOLUTIONS

above it.
   
1 2 3 4 1 2 3 4
1 7→ 1 :
1 2 3 4 1 4 3 2
Identity Reection

   
1 2 3 4 1 2 3 4
1 7→ 2 :
2 3 4 1 2 1 4 3
Rotation Reection

   
1 2 3 4 1 2 3 4
1 7→ 3 :
3 4 1 2 3 2 1 4
Rotation Reection

   
1 2 3 4 1 2 3 4
1 7→ 4 :
4 1 2 3 4 3 2 1
Rotation Reection

15 G = {0, 1, 2}, with a binary operation ∗ that is associative, has 0 as


identity, and such that every element of G has an inverse. We have to
develop its multiplication table. First, since 0 is identity, we have:

0 1 2
0 0 1 2
1 1
2 2

Note that under the given conditions, the cancellation law holds.

If 1 ∗ 2 = 1, we have 1 ∗ 2 = 1 ∗ 0 and the cancellation law gives 2 = 0, a


contradiction. Similarly 1 ∗ 2 = 2 gives 1 = 0, a contradiction. So we must
have 1 ∗ 2 = 0. We similarly conclude that 2 ∗ 1 = 0. We now have

0 1 2
0 0 1 2
1 1 0
2 2 0

Undergraduate Mathematics: Foundations


227

If 1 ∗ 1 = 0, we get 1 ∗ 1 = 1 ∗ 2, and so 1 = 2. If 1 ∗ 1 = 1, we get 1 ∗ 1 = 1 ∗ 0,


and so 1 = 0. Therefore we must have 1 ∗ 1 = 2. Similarly, 2 ∗ 2 = 1. This
completes the table:

0 1 2
0 0 1 2
1 1 2 0
2 2 0 1

The symmetry of the table about the diagonal shows the commutativity of
∗.

Ÿ3.1

1 We shall apply mathematical induction. Let A = { n ∈ N | n 6= S(n) }.


We know that 1 is not in the image of S . Hence 1 6= S(1) and 1 ∈ A. Now
suppose n ∈ A. If S(n) = S(S(n)) then n = S(n) since S is 1-1, and this is
false. Therefore S(n) 6= S(S(n)) and so S(n) ∈ A.

3 4=3+1 (denition of 4)

= (2 + 1) + 1 (denition of 3)

= 2 + (1 + 1) (associativity of +)

=2+2 (denition of 2)

5 We have m, n ∈ W. Recall that 0 is the only member of W that is not of


the form k + 1.
(a) Suppose m + n = 0. If m 6= 0 then m = m0 + 1 for some m0 ∈ W.
0 0
Then m + n = (m + 1) + n = (m + n) + 1 6= 0. Hence m = 0.
Similarly, n = 0.

(b) We shall prove the contrapositive. Suppose m 6= 0 and n 6= 0. Then


m = m0 + 1 and n = n0 + 1 for some m0 , n0 ∈ W. Hence mn =
(m0 + 1)(n0 + 1) = (m0 n0 + m0 + n0 ) + 1 6= 0.
7 Suppose A ⊆ W such that 0 ∈ A and n ∈ A =⇒ n + 1 ∈ A. First,
0 ∈ A =⇒ 1 ∈ A. Let B = A \ {0}. Then 1 ∈ B . Further, n ∈ B =⇒ n ∈
A =⇒ n + 1 ∈ A. Since n + 1 6= 0 we have n + 1 ∈ B . Therefore B = N
and A = B ∪ {0} = W.

9 Let f : X → X . Consider the set Y of all functions from X to X. Dene


G: Y → Y by G(g) = g ◦ f . By the Recursion Theorem there is a function
228 APPENDIX B. SOLUTIONS

Φ: W → Y such that Φ(0) = 1X and Φ(n + 1) = G(Φ(n)) = Φ(n) ◦ f . Let


f n = Φ(n).
11 Dene a function G : W × N → W × N by G(m, n) = (m + 1, (m + 1)n).
Apply the Recursion Theorem to get a function Φ : W → W × N such that
Φ(0) = (0, 1) and Φ(n+1) = G(Φ(n)). Further, let π : W×N → N be dened
by π(m, n) = n. Now dene F : W → N by F = π ◦ Φ.

For example, this achieves the following:

Φ(0) = (0, 1) F (0) = 1


Φ(1) = (1, 1 • 1) F (1) = 1
Φ(2) = (2, 2 • 1) F (2) = 2 • 1
Φ(3) = (3, 3 2 1)• • F (3) = 3 • 2 • 1

One can use mathematical induction to prove that the rst coordinate of
Φ(n) is n. Hence Φ(n) = (n, F (n)) for each n ∈ W. Then Φ(n + 1) =
G(n, F (n)) = (n + 1, (n + 1)F (n)) implies that F (n + 1) = (n + 1)F (n).
The function F is called the factorial function, and we commonly write
n! for F (n).

Ÿ3.2

1 Let U = { n ∈ N | (∀a ∈ A)(n ≥ a) } be the set of upper bounds of A.


Since A is bounded above, U is non-empty. Hence U has a least element M .
This M is also the greatest element of A. ’
3 Let k, a, b ∈ N. M = lcm(ka, kb). Then ka | M =⇒ k | M =⇒ M =
Let
km for some m ∈ N.
Next, ka | km =⇒ a | m. Similarly, b | m. Hence m
is a common multiple of a and b. If ` is also a common multiple of a and
b, then k` is a common multiple of ka and kb, which implies km = M | k`.
Therefore m | ` and so m = lcm(a, b).

We can similarly prove that gcd(ka, kb) = k • gcd(a, b), or derive it as a


consequence of the lcm property. ’
5 This is an immediate consequence of the previous exercise. So we provide
a solution of that exercise. First, gcd(k, a) = 1 =⇒ gcd(kb, ab) = b. Then,
since k | kb and k | ab, we have k | b.
7 Since k is a common multiple of a and b we have lcm(a, b) | k . Hence
ab = lcm(a, b) gcd(a, b) = lcm(a, b) divides k .

Undergraduate Mathematics: Foundations


229

9 (a) Let a∈N and a ≥ 2. We have to show that a is a product of powers


of primes.

First we show that a has a prime factor. Consider the set

A = {n ∈ N | 1 < n ≤ a and n | a }.

It is non-empty. Let p be its least element. Then p is a prime factor


of a.
Now let p1 , . . . , pn be the distinct prime factors of a, with p1 < p2 <
· · · < pn . By Exercise 8 there are natural numbers α1 , . . . , αn such
αi αi +1
that pi | a and pi - a.
α
Note that i 6= j =⇒ gcd(pα j
i , pj ) = 1.
i
’
α1 α
Hence, by repeated application of Exercise 7, p1 · · · pn n | a. We
α1 αn α αn
claim that a = p1 · · · pn . If not, then a = mp1 · · · pn with m ≥ 2.
1

Let p be a prime factor of m.

Then p is a prime factor of a that is dierent from each pi . ’


This contradiction leads to the desired equality.

(b) Suppose that a = q1β1 · · · qm


βm
is another factorization into prime pow-
ers, with q1 < · · · < qm .
Since the pi consist of all the prime factors of a, we have {q1 , . . . , qm } ⊆
{p1 , . . . , pn }. Consider any qj . It equals some pi0 . Then we must have
βj ≤ αi0 . By repeated application of the cancellation law for multi-
plication, we see that m = n, each qi equals pi , and each βi equals
αi .
11 Suppose a, b ∈ N with b ≤ a. We wish to calculate gcd(a, b). The steps
are as follows:

(a) Using division algorithm, obtain a = q1 b + r1 with 0 ≤ r1 < b. If


r1 = 0 then b|a and gcd(a, b) = b.
(b) If r1 6= 0 then, using the previous exercise, we have gcd(a, b) =
gcd(b, r1 ). Now the division algorithm gives b = q2 r1 + r2 with
0 ≤ r2 < r1 . If r2 = 0 then gcd(a, b) = gcd(b, r1 ) = r1 .
(c) If r2 6= 0 then apply the division algorithm to get r1 = q3 r2 + r3 , and
proceed as before.

In these iterative steps, since the remainder ri decreases at each step, we


must eventually get rn = 0. Then the gcd is rn−1 and the process stops.
230 APPENDIX B. SOLUTIONS

13 There are pk natural numbers that are less than or equal to pk . The ones
which are not coprime to pk are the ones that have p as a factor. These have
the form pm with 1 ≤ m ≤ pk−1 , hence are pk−1 in number.
15 Proof by induction:
Dene k = T (1). Then T (1) = k · 1 is true. Assume for some n ∈ N that
T (n) = k · n is true. Then, T (n + 1) = T (n) + T (1) = k · n + k = k(n + 1).

Ÿ3.3

1 (a) 0Z ⊗ x = (1, 1) ⊗ (m, n) = (m + n, m + n) = 0Z .


(b) (−1Z ) ⊗ x = (1, 2) ⊗ (m, n) = (m + 2n, n + 2m) = (n, m) = −x
3 Let x = (m, n) and y = (k, l). Then, assuming m>n and k > l,

x ⊗ y = 1Z ⇐⇒ (mk + nl, ml + nk) = (2, 1)


⇐⇒ mk + nl + 1 = ml + nk + 2
⇐⇒ m(k − l) = n(k − l) + 1 ⇐⇒ (m − n)(k − l) = 1
⇐⇒ m = n + 1 and k = l + 1 ⇐⇒ x = y = 1Z .

Similarly, the m<n and k<l case gives x = y = −1Z . ’


The other cases don't lead to any solutions. ’
5 In the solution to Exercise 15 of Ÿ3.2 we showed that T (n) = kn for every
n ∈ N, with k = T (1). Further,

T (0 + 0) = T (0) + T (0) =⇒ T (0) = T (0) + T (0)


=⇒ T (0) = 0 = k · 0,
T (n + (−n)) = T (n) + T (−n) =⇒ 0 = T (n) + T (−n)
=⇒ T (−n) = −T (n) = −k · n = k · (−n).

7 Apply the usual division algorithm to b and |a| to get b = p|a| + r with
0 ≤ r < |a|. Then b = qa + r, where q = p|a|/a.
9 Since gcd(m, a) = 1 we have xm + ya = 1 for some integers x and y. Then
xmb + yab = b. If m | ab then m | xmb + yab = b.
11 hki = Zn ⇐⇒ 1 ∈ hki ⇐⇒ (∃m ∈ N) 1 = mk
⇐⇒ (∃m ∈ N) n | (mk − 1)
⇐⇒ (∃m, m0 ∈ N) mk − m0 n = 1 ⇐⇒ gcd(k, n) = 1.

Undergraduate Mathematics: Foundations


231

Ÿ3.4

1 You must have seen the standard proof in school. It goes like this. Suppose
there is a rational number x x2 = 2. We can assume it is positive.
such that
Then it has the reduced form p/q where p, q ∈ N and gcd(p, q) = 1. Now we
2 2 2 2 2
have p /q = 2, hence p = 2q . Therefore p is even. Therefore p is even.
2 2 2 2 2
Hence p = 2k for some k ∈ N. Then 4k = 2q implies 2k = q , hence q is
even, hence q is even.

Now 2 is a common factor of p and q , contradicting our initial assumption


1
of their gcd being 1.

3 Similar to the proof of 2 not having a rational square root.


5 (a) x > 0 ⇐⇒ x = m/n with m, n ∈ N ⇐⇒ x−1 = n/m with m, n ∈
N ⇐⇒ x−1 > 0.
(b) x < 0 ⇐⇒ −x > 0 ⇐⇒ (−x)−1 > 0 ⇐⇒ −x−1 > 0 ⇐⇒ x−1 <
0.
7 First, let z ∈ Q+ . Note that this implies z −1 ∈ Q+ .
(a) x < y ⇐⇒ y − x ∈ Q+ ⇐⇒ (y − x)z ∈ Q+ ⇐⇒ yz − xz ∈
Q+ ⇐⇒ xz < yz .
(b) x ≤ y ⇐⇒ x = y or x < y ⇐⇒ xz = yz or xz < yz ⇐⇒ xz ≤ yz .
Next, let z ∈ Q− . Note that this implies −z ∈ Q+ .
(c) x < y ⇐⇒ x(−z) < y(−z) ⇐⇒ xz − yz > 0 ⇐⇒ xz > yz .
(d) x ≤ y ⇐⇒ x(−z) ≤ y(−z) ⇐⇒ xz − yz ≥ 0 ⇐⇒ xz ≥ yz .
9 (a) N has no upper bound in Q.
We have to show that no member of Q can serve as an upper bound
for N. It is enough to prove that no positive rational can be an upper
bound. If x ∈ Q+ then x = m/n with m, n ∈ N. Then m + 1 > x.
(b) Let x, y ∈ Q with x > 0. Then ∃n ∈ N such that nx > y .
Since y/x ∈ Q cannot be an upper bound for N we have ∃n ∈ N such
that n > y/x. Then nx > y .
x
(c) Let x, y ∈ Q+ . Then ∃n ∈ N such that < y.
n
x
Using (b), we have an n∈N such that ny > x. Then < y.
n
1 See http://www.cut-the-knot.org/proofs/sq_root.shtml for a variety of other
proofs of this fact, some of them based on geometry rather than arithmetic.
232 APPENDIX B. SOLUTIONS

11 From Exercise 5 of Ÿ3.3, we know that T (n) = kn for every n ∈ Z, with


k = T (1). Consider any x = 1/n with n ∈ N. Then

1 1 1 1 1
k = T (1) = T ( + · · · + ) = T ( ) + · · · + T ( ) = nT ( )
n n n n n
1 k
=⇒ T (x) = T ( ) = = kx.
n n
Now consider any x = m/n with m, n ∈ N. Then

m 1 1 1 1 1
T (x) = T () = T ( + · · · + ) = T ( ) + · · · + T ( ) = mT ( )
n n n n n n
k
=⇒ T (x) = m • = kx.
n
Finally, for any x ∈ Q+ :

0 = T (0) = T (x − x) = T (x) + T (−x) =⇒ T (−x) = −T (x) = k(−x).

13 Suppose E : Q → Q satises E(1) 6= 0 as well as E(x + y) = E(x)E(y)


for every x, y ∈ Q. Note that E(1) = E((1/2) + (1/2)) = E(1/2)2 > 0. So
αk
E(1) = m/n with m, n ∈ N and gcd(m, n) = 1. Let m = pα 1 · · · pk and
1

β1 βl
n = q1 · · · ql be the factorizations into powers of distinct primes. Then the
pi and the qj are all distinct from each other.
From E(1/n)n = E(1) it follows that each αi and βj is divisible by every
n. This is impossible.

Ÿ3.5

1 Let R be an integral domain, and a, b, c ∈ R. First,

a + b = a + c =⇒ (−a) + (a + b) = (−a) + (a + c)
=⇒ ((−a) + a) + b = ((−a) + a) + c
=⇒ 0 + b = 0 + c =⇒ b = c.

For multiplication, we have ab = ac =⇒ ab − ac = 0 =⇒ a(b − c) = 0. If


a 6= 0 we must have b − c = 0, since R has no zero divisors. Then b = c.
3 Let R be a nite integral domain. We need to show every non-zero member
has a multiplicative inverse. a ∈ R and a 6= 0. Consider the subset
Let
{ an | n ∈ W } of R. If the elements an were all distinct this would be an
m
innite set. So some of them must equal each other. Let a = an with

Undergraduate Mathematics: Foundations


233

m > n. Then am−n an = an =⇒ am−n = 1 due to the cancellation law.


m−n−1
Hence a is the inverse of a. (Note that m − n − 1 ≥ 0)

5 Since F is an ordered eld, we have 1 > 0. Hence 2 = 1 + 1 > 0. Also


2−1 > 0.
x+y
Next, x < y =⇒ 2x < x + y =⇒ x < 2−1 (x + y) = .
2
x+y
Similarly, x < y =⇒ < y. ’
2
7 (a)
2 2 2 2
First, (x − a) ≥ 0 =⇒ x + a − 2ax ≥ 0 =⇒ x + a ≥ 2ax.
2
Since
x > 0 we divide both sides by x to get the desired inequality.
(b) We have 1 ≤ x ≤ a. Hence ax ≥ 1. Further,

a−x 1 1
ax ≥ 1 and a ≥ x =⇒ a − x ≥ =⇒ a − x ≥ −
ax x a
1 1
=⇒ x + ≤ a + .
x a
9 Let F be an ordered eld and A ⊆ F.
(a) A has an upper bound if and only if −A has a lower bound :

M is an upper bound of A if and only if −M is a lower bound for


−A. ’
(b) A has a maximum if and only if −A has a minimum, and then
− max(A) = min(−A):
Combine the solution to (a) with: M ∈A if and only if −M ∈ −A.
(c) A has a supremum if and only if −A has an inmum, and then
− sup(A) = inf(−A):
M = sup(A) ⇐⇒ M is an upper bound of A
and M ≤u for every upper bound u of A
⇐⇒ − M is a lower bound of −A
and −M ≥ −u for every upper bound u of A
⇐⇒ − M is a lower bound of −A
−M ≥ ` for every
and lower bound ` of −A
⇐⇒ − M = inf(−A)

11 Let M = sup(A) and N = sup(B). Then,

y ∈ A + B =⇒ y = a + b with a∈A and b ∈ B =⇒ y ≤ M + N


234 APPENDIX B. SOLUTIONS

Hence M +N is an upper bound of A+B . We need to show that if u < M +N


then u is not an upper bound of A + B . Let 2ε = M + N − u. (Every ordered
eld has 2 = 1 + 1 > 0) Note that ε > 0.

Since M − ε is not an upper bound of A, there is a ∈ A such that


a > M − ε. Similarly, there is b ∈ B such that b > N − ε. Then a + b >
M + N − 2ε = u.
The analogous result for inmum is: If A and B have inma, then A+B
has an inmum and inf(A + B) = inf(A) + inf(B). ’
13 In Theorem 3.5.11 we showed that F contains a copy of N by providing
a 1-1 function ϕ : N → F that preserved addition and multiplication and
mapped 1 to 1. Dene ψ : Z → F by

 ϕ(n) n>0
ψ(n) = 0 n=0 .
−ϕ(−n) n < 0

By its denition, ϕ(n) ∈ F+ for each n ∈ N. This implies that ψ is 1-1.

Let us check that ψ preserves addition, i.e., ψ(m + n) = ψ(m) + ψ(n) for
all m, n ∈ Z. This is trivial if either number is zero, or if both have the same
sign. The remaining case is when m>0 and n < 0. Then:

m + n > 0 =⇒ ψ(m + n) − ψ(n) = ϕ(m + n) + ϕ(−n) = ϕ(m + n + (−n))


= ϕ(m) = ψ(m)
m + n < 0 =⇒ ψ(m + n) − ψ(m) = −ϕ(−(m + n)) − ϕ(m)
= −[ϕ(−(m + n)) + ϕ(m)]
= −ϕ(−(m + n) + m)
= −ϕ(−n) = ψ(n).

One can similarly prove that ψ preserves multiplication, i.e., ψ(mn) =


ψ(m)ψ(n) for all m, n ∈ Z. ’

Ÿ3.6

1 Each b ∈ B is an upper bound for A. Therefore sup(A) exists. Call it r.


Since it is an upper bound of A we have a ≤ r for every a ∈ A. Since it is
the least upper bound of A, we have r ≤ b for every b ∈ B .
To see that the claim is false in Q, consider A = { x ∈ Q+ | x 2 < 2 } and
+ 2
B = { x ∈ Q | x > 2 }.

Undergraduate Mathematics: Foundations


235

3 Let X ⊆ F be non-empty and bounded above. Let B be the set of all strict
upper bounds of X : B = { b ∈ F | (∀x ∈ X)(x < b) }. Let A be the set of all
strict lower bounds of B : A = { a ∈ F | (∀b ∈ B)(a < b) }. Then:

1. X ⊆ A =⇒ A 6= ∅.
2. If b is an upper bound of A then b+1 is a strict upper bound. Hence
6 ∅.
B=
3. a ∈ A, b ∈ B =⇒ a < b.
4. Suppose a∈
/ B. If there is b ∈ B such that b ≤ a, then a is a strict
upper bound of A and hence a ∈ B . So for every b ∈ B we have a < b.
Therefore a ∈ A. Hence A ∪ B = F.
Applying the given property of F we obtain a unique r ∈ F such that a≤
r≤b for every a∈A and b ∈ B . We claim that r = sup(X). ’
5 Fix a ∈ R+ and dene A = { x ∈ R+ | x2 ≤ a }.
(a) From Task 3.5.10 we know that 0 < x < y =⇒ 0 < x2 < y 2 . Hence
a can have at most one positive square root.

(b) Suppose every a > 1 has a positive square root. Now consider b < 1.
Then 1/b > 1 has a positive square root y , and 1/y is a positive
square root of b.

Hence, it is enough to prove the existence of positive square roots for


a > 1. We assume a>1 in the remainder of the proof.

(c) Since 1 ∈ A, we have A 6= ∅. Also, from the previous exercise we


know that a < a2 . If x > a then x2 > a2 > a. Hence a is an upper
bound of A.

(d) Since A is non-empty and bounded above, it has a supremum. Let


c = sup(A). Then 1 ≤ c ≤ a.
2 1 a
(e) Suppose c > a. Let d = c+ . Then
2 c
1  2 a2  1  2 a2  1 a 2
d2 − a = c + 2 + 2a − a = c + 2 − 2a = c− > 0,
4 c 4 c 4 c
hence d2 > a and d is an upper bound of A. But,

1 c2 
d< c+ = c,
2 c
which contradicts c being the least upper bound of A.
236 APPENDIX B. SOLUTIONS

a  1
(f) Suppose c2 < a. Let e= c+ . Then, using Exercise 7 of
a+1 c
Ÿ3.5,

 a 2  1   a 2  1 
e2 = c2 + 2 + 2 < a + + 2 = a,
a+1 c a+1 a
hence e ∈ A. Also,
a 1 a
e= c+ ,
a+1 a+1c
implies that e lies strictly between c and a/c. Now c2 < a implies
c < a/c and hence c < e. Since e ∈ A, this contradicts c being an
upper bound of A.
Combining (e) and (f ) with trichotomy gives c2 = a.
7 Imitate the steps outlined for the previous exercise. ’
c
9 We are given a ∈ R and c ∈ R+ such that 0≤a≤ for every n ∈ N. If
n
+
a 6= 0 then a ∈ R and, by the Archimedean property, there is n∈N such
that na > c. This contradicts the given information. Hence a = 0.
11 We can prove this by mathematical induction. The n=1 case is trivial.
The n=2 case has been proved earlier. Now consider the induction step:

n+1 n X n n n+1
X X  X X
xi = xi +xn+1 ≤ xi +|xn+1 | ≤ |xi |+|xn+1 | = |xi |.


i=1 i=1 i=1 i=1 i=1

Ÿ3.7

1 Let z = x + iy and w = i + iv with x, y, u, v ∈ R.


(a) z + w = (x + iy) + (u + iv) = (x + u) + i(y + v) = (x+u)−i(y+v) =
(x − iy) + (u − iv) = z + w.
(b) −z = −x − iy = −x + iy = −(x − iy) = −z .
(c) zw = z w. ’
 z   zw   1  1 zw z
(d) = = 2
zw = 2
zw = = .
w ww |w| |w| ww w
3 Suppose that C has an order < that makes it an ordered eld. We know
that in any ordered eld 0<1 and hence −1 < 0. Further, for any non-zero
x in an ordered eld, we have x2 > 0. But in C, we have i2 = −1 < 0.

Undergraduate Mathematics: Foundations


237

5 We have to nd the images of the given subsets of C under the function
f (z) = z 2 . Note that if z = x + iy then z 2 = (x2 − y 2 ) + i(2xy).
(a) Real Axis: y = 0 =⇒ z 2 = x2 and so the image is the positive real
axis.

(b) Imaginary Axis: x = 0 =⇒ z 2 = −y 2 and so the image is the


negative real axis.

(c) Lines passing through origin: Squaring a number doubles its argu-
ment, so a line passing through origin and making an angle θ with
the positive real axis will be sent to the ray starting at origin and
making an angle 2θ with the positive real axis.

(d) Circles centered at origin: Squaring a number squares its modulus,


so the circle with radius R will be sent to the circle with radius R2 .
7 We have ai xi = ai xi since ai is real. Hence,

P (α) = 0 ⇐⇒ P (α) = 0
⇐⇒ a0 + a1 α + a2 α2 + · · · + an αn = 0
⇐⇒ P (α) = 0.

9 Let z = |z|(cos θ + i sin θ). z 3 = 1 =⇒ |z|3 (cos 3θ + i sin 3θ) = 1.


Then
Therefore |z| = 1. Also, cos 3θ = 1 and sin 3θ = 0. Hence θ = 0, 2π/3, 4π/3.
Hence the three cube roots of 1 are

cos 0 + i sin 0 = 1,

2π 2π −1 + i 3
cos + i sin = ,
3 3 2 √
4π 4π −1 − i 3
cos + i sin = .
3 3 2

11 Let α ∈ C and consider the polynomial equation z n − α = 0. Each nth


root of α is a solution of this equation. Applying Exercise 6 we see that α
has at most n distinct complex nth roots.

We can show the existence of the n roots with the help of the `roots of
unity' 1, w, . . . , wn−1 of Exercise 10. Let α = r(cos θ + i sin θ) with r = |α| >
0 and 0 ≤ θ < 2π . Then β = r1/n (cos θ/n + i sin θ/n) satises β n = α and
n−1
is an nth root of α. Further, all the numbers β, wβ, . . . , w β are distinct
and are nth roots of α.
238 APPENDIX B. SOLUTIONS

Ÿ4.1

1 In the following symbolic expressions the quantiers range over the uni-
versal domain B of all sets:

Axiom of Regularity: (∀A)[A 6= ∅ =⇒ (∃x ∈ A)(x ∩ A = ∅)].


Axiom of Innity: (∃I)[∅ ∈ I ∧ (∀x)(x ∈ I =⇒ x ∪ {x} ∈ I)].
3 If we have sets A0 , A1 , . . . , An such that A0 ∈ A1 ∈ · · · ∈ An ∈ A0 , then
the set M = {A0 , A1 . . . , An } will violate the Axiom of Regularity.

5 Let I and J be inductive sets. Dene NI to be the intersection of all the


inductive subsets of I, and NJ to be the intersection of all the inductive
subsets of J. We observe that

1. ∅ ∈ NI ∩ NJ .
2. x ∈ NI ∩ NJ =⇒ S(x) ∈ NI ∩ NJ .
ThereforeNI ∩ NJ is an inductive set. Hence NI ⊆ NI ∩ NJ , and so
NI = NI ∩ NJ . Similarly, NJ = NI ∩ NJ .
7 X is an inductive set. We have to show that the following are also inductive:

(a) A = { x ∈ X | x ⊆ X }: Since X is inductive, ∅ ∈ X . Hence ∅ ∈ A.


Now let x ∈ A. We have to show S(x) ∈ A. We have S(x) ∈ X since
X is inductive. And x ⊆ X =⇒ S(x) = x ∪ {x} ⊆ X .
(b) B = {x ∈ X | x is transitive } ’
9 Suppose X is transitive. Let A ∈ P(X). Then A ⊆ X. And,

x ∈ A =⇒ x ∈ X =⇒ x ⊆ X =⇒ x ∈ P(X).
Therefore A ⊆ P(X). This shows P(X) is transitive.

Now suppose P(X) is transitive. Then,

x ∈ X =⇒ {x} ∈ P(X) =⇒ {x} ⊆ P(X) =⇒ x ∈ P(X) =⇒ x ⊆ X.

Ÿ4.2

1 Let F be the set of whole numbers n such that the existence of a 1-


1 function f : A → Fn implies the niteness of A. Apply mathematical
induction to prove that F = W. ’
3 Since A and B are nite sets, we have bijections ϕA : A → Fn and ϕB : B →
Fm .

Undergraduate Mathematics: Foundations


239

(a) If x ∈/ A then |A ∪ {x}| = |A| + 1: Dene ϕ : A ∪ {x} → Fn+1 by


ϕ(t) = ϕA (t) if t ∈ A, and ϕ(x) = n + 1.
(b) |A ∪ B| ≤ |A| + |B|: Dene f : A ∪ B → Fn+m by

ϕA (x) if x∈A
f (x) =
ϕB (x) + n if x∈B\A

(c) If A∩B = ∅ then |A ∪ B| = |A| + |B|: In this case the function f


dened in (b) is a bijection.

(d) |A \ B| = |A| − |A ∩ B|: The sets A \ B and A ∩ B are disjoint and


their union isA. Therefore, |A| = |A \ B| + |A ∩ B|.
(e) |A ∪ B| = |A| + |B| − |A ∩ B|: The sets A and B \ A are disjoint and
A∪B . Hence, |A∪B| = |A|+|B\A| = |A|+|B|−|A∩B|.
their union is

5 Let |A| = n.
Qn
(a) Dene g : P(Fn ) → k=1 F2 by

1 if k∈A
g(A)k = .
0 if k∈
/A

(b) Apply Exercise 4.

7 Suppose f : N → A is 1-1. If A is nite there is a bijection g : A → Fn .


Apply Exercise 1 to g ◦ f : N → Fn .
9 (a) If there is a 1-1 map gn : Fn → A for every n ∈ N then A is innite :
If A is nite there is a bijection g : A → Fm for some m. Then
g ◦ gm+1 : Fm+1 → Fm is 1-1, contradicting Theorem 4.2.1.
(b) There are innitely many rational numbers between any two rational
numbers : Let p, q be any two rational numbers, with p < q . Let A =
(p, q)Q = { x ∈ Q | p < x < q }. Apply (a) to gn : Fn → A dened by
k n+1−k
gn (k) = p+ q, k ∈ Fn .
n+1 n+1

11 The diagram itself should be intuitively convincing. Our task is to convert


it into a formula for a bijection f : N × N → N. Here is an analysis

1. Observe that the k th diagonal has k numbers, each of whose coordi-


nates adds up to k + 1. Hence the ordered pair (m, 1) is preceded by
1 + 2 + · · · + (m − 1) = m(m − 1)/2 values, and we should map (m, 1)
to 1 + m(m − 1)/2.
240 APPENDIX B. SOLUTIONS

2. Hence we should map (m − k, 1 + k) to k + 1 + m(m − 1)/2 with


k = 0, . . . , m − 1.
3. Substitute u=m−k and v =1+k to get a function of u and v:

(u + v − 1)(u + v − 2)
f (u, v) = v + .
2

4. This f is a bijection. ’
13 (a) + +
N is in bijection with Q : Each x ∈ Q has a unique reduced form
mx /nx with mx , nx ∈ N and gcd(mx , nx ) = 1. Mapping x to (mx , nx )
+
creates a 1-1 map from Q to N × N. Composing this with the
+
bijection from N × N to N creates a 1-1 map from Q to N. There is
+
also an obvious 1-1 map from N to Q . Apply the Schröder-Bernstein
Theorem.

(b) N is in bijection with Q: Let f : Q+ → N be a bijection (its existence


was established in (a)). Dene a bijection g: Q → N as follows:

 2f (x) if x>0
g(x) = 1 if x>0 .
2f (−x) + 1 if x<0

Ÿ4.3

1 A = {{1}, {1, 2}, {1, 2, 3}}.


(a) A choice function for A: c({1}) = 1, c({1, 2}) = 2, c({1, 2, 3}) = 3.
(b) Number of choice functions for A: 1 • 2 • 3 = 6.
3 A consists of the nite non-empty subsets of R. Each nite non-empty
subset of real numbers has a unique minimum element. Dene c: A → R by
c(A) = min(A).
5 We have to show that the Axiom of Choice is equivalent to each of the
following.

(a) For each surjection f: X →Y there is a map g: Y → X with f ◦g =


1Y .
Theorem 4.3.1 has already established that this statement follows
from the Axiom of Choice. We need to prove the converse. Let A be
a set of non-empty sets and let X be the union of all the members
of A. Dene f: X → A by A = f (x) if x ∈ A. Since each A ∈ A is

Undergraduate Mathematics: Foundations


241

non-empty this function is a surjection. So we have g: A → X such


thatf (g(A)) = A for each A ∈ A. This means g(A) ∈ A for each
A ∈ A.
(b) For each function f: X →Y and surjection g: Z → Y there is a map
k: X → Z with f = g ◦ k.
First, we assume the Axiom of Choice. Let c : P(Z) \ {∅} → Z be a
choice function. Dene k: X → Z by k(x) = c(g −1 {f (x)}).
For the converse, note that (b) implies (a). ’
7
(a) =⇒ (b) Let R be a subset of X × Y . It is also a subset of dom(R) ×
Y . For each x ∈ dom(R) dene Rx = { y ∈ Y | xRy }. Then A =
{ Rx | x ∈ X } is a collection of non-empty sets. Let c : A → Y be a
choice function. Dene F : dom(R) → Y by F (x) = c(Rx ).

(b) =⇒ (a) Let A be a set of non-empty sets. Let X = A. Dene a


S
relation R ⊆ A × X by ARx ⇐⇒ x ∈ A. It has domain A. Then
there is a function F : A → X such that F ⊆ R, and x = F (A) =⇒
ARx =⇒ x ∈ A.

(a) =⇒ (c) ’
(c) =⇒ (a) ’
9 R is a relation X with dom(R) = X . For each x ∈ X let Rx =
on
{ y ∈ X | xRy }. A = { Rx | x ∈ X } is a collection of non-empty sets.
Then
By Axiom of Choice, there is a choice function c : A → X . Dene G : X → X
by G(x) = c(Rx ). Apply the Recursion Theorem to G.

Ÿ4.4

1 We are given that A is countable and f : A → B is a surjection. Since A


is countable, there is a surjection g : N → A, and so h = f ◦g is a surjection
of N onto B. Therefore B is countable.

3 If D is the set of dyadic rationals, we have Z ⊆ D ⊆ Q. Now Z ⊆ D


implies that D is innite and D⊆Q implies that D is countable.
5 Since A and B are countably innite, we have bijections fA : N → A and
fB : N → B . Then f : N × N → A × B, dened by f (m, n) = (fA (m), fB (n)),
is a bijection.
242 APPENDIX B. SOLUTIONS

7 Let X be the set of sequences of 0's and 1's that are constant after some
entry. Then X = A∪B is a disjoint union, where A consists of the sequences
that terminate in 0's and B consists of the sequences that terminate in 1's.
It suces to show that A and B are countably innite. In fact, by symmetry,
it suces to show A is countably innite.

First, we show A is innite by giving an injection from N: The natural


number n is mapped to the sequence in which the nth term is 1 and all other
terms are 0.

Next we show A is countable by exhibiting it as a countable union of


nite sets An (n ∈ N). Dene An to consist of the sequences whose terms
after the nth one are all 0. Then |An | = 2n .
9 The transcendental numbers are the complement of the algebraic numbers.
By the previous exercise, the set of algebraic numbers is countable. Hence,
by Theorems 4.4.7 and 4.4.8 the transcendental numbers are uncountable.

11 Since R and P(N) are in bijection, we apply the previous exercise to


obtain |R| = |R × R|.
To complete the proof, observe that |R| ≤ |R × N| ≤ |R × R|. ’
13 First, consider A = { 1/2n | n ∈ N }. If F is a nite subset of A, it has a
least element 1/2k . Then the sum σ of the members of F is bounded above
by 1:

1 1 1 1 1 − 1/2k 1 1
σ≤ + 2 + ··· + k = < = 1.
2 2 2 2 1 − 1/2 2 1 − 1/2

For the second example, consider A = { 1/ n | n ∈ N }. Further, let Fn =
√ √
{1, 1/ 2, . . . , 1/ n}. Then,

1 1 1 1 1 √
1 + √ + · · · + √ > √ + √ + · · · + √ = n.
2 n n n n

15 If A is countably innite then |A ∪ N| = |N| = |A| since we have a nite


union of countable sets. If A is uncountable, apply Theorem 4.4.7.

Ÿ4.5

1 Consider the disjoint union X = N ∪ {α} where N has the usual order
while α is not related to any other element of X. Then α is maximal while
N is a chain that is not bounded above.

Undergraduate Mathematics: Foundations


243

3 ( =⇒ ): Assume Zorn's Lemma. Let X be a partially ordered set. Let C


be the set of all chains in
S X, ordered by inclusion. If A is a chain in C then
A is a chain which is an upper bound of A. By Zorn's Lemma, C has a
maximal element.

( ⇐= ): Assume Hausdor 's Maximal Principle. Let X be a partially ordered


set in which every chain has an upper bound. By the Principle, this X has
a maximal chain C. Let u be an upper bound of C. Then u is a maximal
element of X.
5 Set Order Well-Ordered

W Usual Ë
{0, 1} × N Dictionary Ë
N×N Dictionary Ë
(0, 1) × (0, 1) Dictionary é

7 Let X be a well-ordered set. Let S ⊆ X have the property that whenever


{ x ∈ X | x < s } is a subset of S then s ∈ S . If X \ S is non-empty it has a
least element u. Then { x ∈ X | x < u } is a subset of S , and so u ∈ S ! This
contradiction implies S = X .

9 Let ≤ be the partial order on X . We rst show that ≤ is a total order. If


x, y ∈ X then {x, y} is a non-empty countable subset of X , hence it has a
least element. Depending on which of the elements is least, we have x ≤ y
or y ≤ x.

Suppose X is not well-ordered. Then it has a subset C that has no


least element. Consider the > relation on S. Since S has no least element,
we have dom(>) = S . Therefore, by the Principle of Dependent Choice
there is a sequence (xn )n∈N in S such that xn > xn+1 for each n. Then
A = { xn | n ∈ N } is a countable set that has no least element.
11 We'll use Zorn's Lemma. Let

X = { (D, f ) | D ⊆ A and f :D×N→D is a bijection }.

X is non-empty because A has a countably innite subset D, and there will


be a bijection from D × N to N.
0 0 0 0
Give X S (D, f ) ≤ (D , f ) if D ⊆ D and f |D =
a partial order by dening
f . If C is a chain in X , dene E = (D,f )∈C D. Also dene fE E × N → E
:
by fE (x, n) = f (x, n) when x ∈ D and (D, f ) ∈ C . Check that fE is well-
dened and a bijection. Then (E, fE ) is an upper bound for C . So, by Zorn's
Lemma, X has a maximal element (M, g).
244 APPENDIX B. SOLUTIONS

A \ M is innite we have an injection h : N → A \ M . Then M 0 =


If
M ∪ h(N) is a disjoint union. Let ϕ : h(N) × N → N be a bijection. Dene
g 0 : M 0 × N → M 0 by

g(x, n) if x ∈ M
g 0 (x, n) =
ϕ(x, n) if x ∈ h(N)

Check that g 0 is a bijection. Since g 0 |M = g this violates the maximality of


(M, g). Therefore A \ M is nite. But then |A| = |M | and so |A × N| =
|M × N| = |M | = |A|.

Undergraduate Mathematics: Foundations


References
We have assumed that the reader has already had a rst encounter with
sets and set operations. If required, these topics can be reviewed with the
help of [1]. In the chapter on Logic and Sets, our examples are almost entirely
drawn from the world of mathematics. You may nd this a bit dry and may
long for real life examples, in which case you should consult [1, 2]. More
elaborate treatments of how to discover proofs and write mathematics can
be found in [10, 13, 14, 16, 17, 20]. The website [22] has a rich collection of
examples and exercises in propositional calculus. A compact introduction to
symbolic logic can be found in [15]. The book [24] not only introduces logic
but discusses its relationship with mathematics, as well as the relationship
of mathematics to the general human culture.

Reference [6] is excellent for self-study. It is probably better appreciated


after one or two semesters of studying mathematics at a university, as many
of its examples and exercises refer to material usually taught in that period.

More advanced treatments of Set Theory can be found in [5, 9]. The
books [8, 11] have a similar aim to this book of moving from basic logic and
sets to number systems, while the classic [7] takes the natural numbers as
given and then constructs all the number systems up to the real and complex
numbers. There is little `chat' in [7] but then no awkward details are left for
the reader to sort out.

The text [12] has a chapter on Sets, Logic and Computation which tells
the story of the attempted axiomatization of mathematics in the 19th and
20th centuries, with connections to current developments. Finally, [3] has an
engrossing depiction of the human side of logic and mathematics.

245
246 REFERENCES

Books
[1] Shobha Bagai, Amber Habib, and Geetha Venkataraman. A Bridge to Math-
ematics. New Delhi: Sage Publications India, 2017. isbn: 978-9386446121.
[2] Keith Devlin. Introduction to Mathematical Thinking. Lightning Source Inc.,
2011. isbn: 978-0615653631.
[3] A. Doxiadis and C. H. Papadimitriou. Logicomix: an epic search for truth.
Bloomsbury, 2009. isbn: 978-0747597209.
[4] Derek Goldrei. Propositional and Predicate Calculus: A Model of Argument.
London: Springer, 2005.

[5] Paul R. Halmos. Naive Set Theory. Undergraduate Texts in Mathematics.


New York: Springer, 1974. isbn: 978-8184892062.
[6] Ajit Kumar, S. Kumaresan, and B. K. Sarma. A Foundation Course in Math-
ematics. New Delhi: Narosa, 2017. isbn: 978-8184876109.
[7] Edmund Landau. Foundations of Analysis. Third edition. Vol. 79. AMS
Chelsea Publishing Series. Providence, Rhode Island: American Mathemat-
ical Society, 2001. isbn: 978-0821826935.
[8] Eliott Mendelson. Number Systems and the Foundations of Analysis. Mine-
ola, N.Y.: Dover Publications, 2008.isbn: 978-0486457925.
[9] Charles C. Pinter. A Book of Set Theory. Dover, 2014. isbn: 978-0486497082.

[10] George Polya. How To Solve It. Reprint edition. First published in 1945.
Princeton University Press, 2014. isbn: 978-0691164076.
[11] Inder K. Rana. From Numbers to Analysis. Singapore: World Scientic, 1998.
isbn: 978-9810233044.
[12] John Stillwell. Mathematics and Its History. Third edition. Undergraduate
Texts in Mathematics. New York: Springer, 2010. isbn: 978-1441960528.
[13] Daniel J. Velleman. How to Prove it. Second edition. New York: Cambridge
University Press, 2006. isbn: 978-0521675994.
[14] Franco Vivaldi. Mathematical Writing. Springer Undergraduate Mathematics
Series. London: Springer, 2014. isbn: 978-1447165262.

Websites and E-books


Websites are liable to migrate or die and references to them are inherently
unstable. The ones given below are correct as of December 2018.

[15] A. H. Basson and D. J. O'Connor.Introduction to Symbolic Logic. url:


https://archive.org/details/introductiontosy00bass.

Undergraduate Mathematics: Foundations


WEBSITES AND E-BOOKS 247

[16] Richard Hammack. Book of Proof. url: http : / / www . people . vcu . edu /
~rhammack/BookOfProof/.
[17] Patrick Keef and David Guichard. An Introduction to Higher Mathematics.
url: https://www.whitman.edu/mathematics/higher_math_online/.
[18] A Friendly Introduction to Math-
Christopher C. Leary and Lars Kristensen.
ematical Logic. url: https://minerva.geneseo.edu/a-friendly-introdu
ction-to-mathematical-logic.
[19] Mathematics Text-
National Council of Educational Research and Training.
book for Class XI. url: http://ncert.nic.in/textbook/textbook.htm?
kemh1=0-16.
[20] Ted Sundstrom. Mathematical Reasoning: Writing and Proof. url: https:
//scholarworks.gvsu.edu/books/9/.
[21] Proof Designer. url: https://app.cs.amherst.edu/
Daniel J. Velleman.
~djvelleman/pd/pd.html.
[22] Finite Mathematics: introduction to
Stefan Waner and Steven R. Costenoble.
logic. url: https://www.zweigmedia.com/RealWorld/logic/logicintro.
html.
[23] Douglas West.The Grammar According to West. url: https://faculty.
math.illinois.edu/~west/grammar.html.
[24] Raymond L. Wilder.Introduction to the Foundations of Mathematics. url:
https://archive.org/details/IntroductionToTheFoundationsOfMathem
atics.
Index
ℵ0 , 181 consistent, 195
1-1 correspondence, see bijection
Banach-Tarski Paradox, 179
absolute value, 130, 150, 156 bijection, 89, 91
absorption, 17, 38 binary operation, 97
addition binomial formula, 117
of complex numbers, 155 bound
of integers, 124 lower, 64
of natural numbers, 109 upper, 64
of rational numbers, 132 bound variable, 26
of real numbers, 200 bounded, 64
and, 5 above, 64
anti-symmetry, 61 below, 64
Archimedean Property, 137, 148149
associative, 17, 38, 98, 110 cancellation, 98, 110, 115, 226
axiom, 2, 3, 193 Cantor's diagonal method, 182
Axiom of Cantor's Theorem, 94, 182
Choice, 84, 86, 176, 179, 182, Cardano's formula, 160
187, 188, 191, 194, 197 cardinality, 93, 172
Empty Set, 32 cartesian product, 51, 52, 56, 86
Extension, 32 chain, 186
Global Choice, 197 choice function, 176
Innity, 166, 169 class, 196
Pairs, 32, 166, 211 proper, 196
Power Set, 36, 166 closure, 101
Regularity, 166, 169, 194 codomain, 70
Replacement, 197 commutative, 17, 38, 98, 110
Separation, 35, 52, 166, 188, Comparability of Cardinality, 189
193, 211 complement, 36, 197
Union, 32, 211 complex numbers, 43, 154
axiom schema, 35 composition, 80, 98, 99
axiomatization, 194 condition
complete, 195 necessary, 12

249
250 INDEX

sucient, 12 ordered, 140


conjunction, 6 nite set, 170, 171, 178, 190
Continuum Hypothesis, 185, 195 free variable, 26
contradiction, 15, 16 function, 6970
contrapositive, 18, 26, 42 1-1, 72, 177
converse, 18, 26, 42 bijective, see bijection
coprime numbers, 121 constant, 72
countable set, 181, 183 empty, 72
countably innite set, 181 identity, 71
cut, 198 injective, see function, 1-1
inverse, 91, 92
D4 , 103 one-one, see function, 1-1
de Morgan's Laws, 18, 38, 85 onto, 74, 177
decimal representation surjective, see function, onto
of natural numbers, 121 Fundamental Theorem of
of real numbers, 150 Algebra, 50, 57
Dedekind cut, 198 Arithmetic, 50, 122
dierence Calculus, 50
of rational numbers, 133
dierence (of sets), 36, 99 gcd, 121, 122, 129
disjoint, 36 glb, see greatest lower bound
disjunction, 6 greatest common denominator, see
distributive, 17, 38, 55, 85, 98, 110 gcd
division greatest element, 63
of integers, 128 greatest integer function, 149
of natural numbers, 119 greatest lower bound, see inmum,
divisor, see factor 66
domain
of a function, 70 Hausdor 's Maximal Principle, 191

of a relation, 57 hcf, see gcd


Hero's method, 146
empty set, 32, 193 highest common factor, see gcd
equality of functions, 72
equivalence class, 59 idempotence, 17, 38

representative of, 59 image (of a point under a function),

equivalence of statement forms, 16, 70

26 image(of a set under a function), 76

equivalence relation, 59 implication, 9, 11

exponentiation, 112, 135 index set, 84


indexed set, 84
factor, 44, 120, 129 inductive set, 166
factorial, 116, 117 inmum, 66
Fibonacci sequence, 116 innite set, 172, 177, 183
eld, 139 integers, 42
order complete, 148, 201, 202 integers modulo n, 103, 131

Undergraduate Mathematics: Foundations


INDEX 251

integral domain, 138 negative elements (of an ordered


intersection, 36, 85 eld), 141
interval negative integers, 126
bounded, 44 negative rationals, 136
closed, 45 not, 5
half-open, 45
open, 45 open sentence, 21, 30
unbounded, 45 or, 5
interval (of real numbers), 4445 exclusive, 5, 8, 20
inverse image (of a set under a func- inclusive, 5
tion), 77, 92 order
irreexivity, 62, 67 dictionary, 66, 67, 192, 243
isomorphism partial, 61
of ordered elds, 202 strict, 62, 67
of Peano systems, 168 total, 66, 67
ordered pair, 51, 55
Laws of Logic, 16
lcm, 120 partition, 60
least common multiple, see lcm Peano Axioms, 106, 108, 116, 167
least element, 63 Peano system, 168
least upper bound, see supremum, 65 polynomial
left identity, 102 cubic, 74
left inverse, 102 positive elements (of an ordered
logical implication, see implication eld), 141
lub, see least upper bound positive integers, 126
positive rationals, 136
map, 72 postulates, 3
material implication, see implication power set, 36, 94, 99
maximal element, 64 pre-image (of a point under a func-
maximum, 63, 144 tion), 70
membership, 31, 196 prime number, 44, 120, 122, 172
minimal element, 64 Principle of Dependent Choice, 181,
minimum, 63 243
multiple, 44, 120, 129 Principle of Mathematical Induction,
multiplication 106, 107, 116, 170
of complex numbers, 155 Principle of Strong Induction, 118
of integers, 125 Principle of Transnite Induction,
of natural numbers, 109 192
of rational numbers, 133 proof, 2
of real numbers, 200 by contradiction, 19, 20, 73
multiplication table, 101 by contrapositive, 19, 73
direct, 19, 73
natural numbers, 42 proper subset, 34
negation, 6, 17, 24 property, 30
252 INDEX

quantier, 22, 40 statement form, 15


existential, 22 statement variable, 6
universal, 22 subset, 34
quotient, 119 successor, 166
successor function, 106, 166
range superset, 34
of a function, 72 supremum, 65
of a relation, 57 symmetric dierence, 40, 56, 102
rational numbers, 42, 132 symmetry, 59
real numbers, 43, 200
recursion, 107, 112, 116, 168
tautology, 15, 16
Recursion Theorem, 107, 112, 116,
theorem, 2
168, 241
totality, 66
reduced form (of a rational number),
transitive set, 169
132
transitivity, 59, 61, 62, 67, 141
reexivity, 59, 61
triangle inequality, 150
relation, 57
trichotomy, 67, 141, 190
remainder, 119
truth table, 6, 8, 11, 18
right identity, 102
truth value, 4
right inverse, 102
roots of unity, 164
unbounded, 64
roster notation, 33
uncountable set, 181, 183
Russell's Paradox, 30, 37
union, 32, 85

S3 , 100, 103
Schröder-Bernstein Theorem, 174, well-ordered set, 190

178 Well-Ordering Principle, 119, 121,

set, 31, 196 177

set-builder notation, 35, 56 Well-Ordering Theorem, 190

square root, 75, 148, 153 whole numbers, 42, 111

statement, 5
biconditional, 13 zero, 111, 125, 133
compound, 5 to the power of zero, 113
conditional, 10 ZF system of axioms, 31, 38
derivable, 194 Zn , 103, 138
simple, 4 Zorn's Lemma, 187, 191

Undergraduate Mathematics: Foundations

You might also like