Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Algebraic Systems, Spring 2014,

January, 2014 Edition


Todd Cochrane
Gabriel Kerr

Contents
Notation

Chapter 0.

Peano Axioms for Natural Numbers An Introduction to Proofs


0.1. Sets and Logic
Exercises
0.2. Peano Axioms
Exercises
0.3. Relations
Exercises

Chapter 1. Basic Arithmetic


1.1. The set Z
Exercises
1.2. The ring Z

7
7
9
10
15
15
18
19
19
21
21

Notation

N = {1, 2, 3, 4, 5, . . . } = Natural numbers


Z = {0, 1, , 2, 3, . . . } = Integers
E = {0, 2, 4, 6, . . . } = Even integers
O = {1, 3, 5, . . . } = Odd integers
Q = {a/b : a, b Z, b 6= 0} = Rational numbers
R = Real numbers
C = Complex numbers
Zm = Ring of integers mod m
[a]m = {a + mx : x Z} = Residue class of a mod m
Um = Multiplicative group of units mod m
1

(mod m) = multiplicative inverse of a (mod m)


(m) = Euler phi-function

(a, b) = gcd(a, b) = greatest common divisor of a and b


[a, b] = lcm[a, b] = least common multiple of a and b
a|b = a divides b
M2,2 (R) = Ring of 2 2 matrices over a given ring R
R[x] = Ring of polynomials over R
|S| = order or cardinality of a set S
Sn = n-th symmetric group

intersection

union

empty set

there exists

for all

equivalent to

subset

there exists a unique


implies

iff if and only if

element of

congruent to

CHAPTER 0

Peano Axioms for Natural Numbers An Introduction to Proofs


We begin this course with the construction of the natural number system. The
Peano Axioms form the heart of much of mathematics and lay the foundation for
algebra and analysis. Every attempt will be made to stay away from unnecessary
abstraction. This should be a guiding principle when working out the exercises as
well!
0.1. Sets and Logic
This section should serve as a very short introduction to 20th century mathematics. That is, it is an introduction to proofs involving sets. Set theory is in fact
a subject within itself that was initiated and studied by many in the early 20th
century. This was in response to several paradoxes that had come up. For us, a set
is a collection of things, called elements. For example, Atoms could be the set of
all atoms in the universe while Adams could be the set of all people named Adam
in your family. You say there is no Adam in your family? Then Adams is known
as the empty set which is the set that contains no elements at all. The notation
for this important set is
Adams = .
Unfortunately, this is just the beginning of new notation. Now that we know
what sets are, we need a slew of symbols to describe how they work and interact
with each other. First, we often write a set with only a few elements by enclosing
the elements inside curly brackets. For example, if we wanted to write the set S of
letters in the alphabet that occur before g we would write
S = {a, b, c, d, e, f }.
If you have a set with a lot of elements that are ordered, you can use the dot-dotdot notation... For example the set of integers T between 3 and 600 and the set of
integers T 0 greater than 5 can be written
T 0 = {6, 7, . . .}.

T = {3, 4, ..., 600}

In general, we will be interested in stating whether something is or is not an


element of a given set. If a is in A (which is the same thing as saying a is an element
of A), we write a A while if it is not, we write a 6 A. For the above examples we
could say a S but a 6 T and 5 T but 5 6 T 0 .
Sometimes we want to define a set that has elements in another set which
satisfy a property. For example, if we want to consider atoms H that have only one
proton we can write
H = {a Atoms : a has one proton }.
7

0. PEANO AXIOMS FOR NATURAL NUMBERS - AN INTRODUCTION TO PROOFS

It is clear that any element of the set H is also an element of Atoms, after all, that
was how H was defined. This is precisely what it means for H to be a subset of
Atoms. To express this relationship we write
H Atoms.
If we have two sets A and B, we can define a whole lot of new sets. We will
sum many of these constructions (and one property) up in the following definition.
Definition 0.1.1. Suppose A and B are sets.
(1) The set A B is the intersection of A and B. It consists of elements c
such that c A and c B.
(2) Two sets A and B are called disjoint if A B = .
(3) The set A B is the union of A and B. It consists of elements c such
that c A or c B.
(4) The set A B is the Cartesian product of A and B. It consists of
elements c = (a, b) where a A and b B.
Enough of the definitions. Lets try proving something algebraic.
Proposition 0.1.1 (Commutativity of intersection). If A and B are sets then
AB =BA
You may think about this a second and say Whats the big deal? Of course this
is true!, but us mathematicians really need something better than of course... we
need a proof! The way to prove a statement like this is to go back to the definition
and methodically show that the definitions force the statement to be correct. This
way of proving something is called a direct proof.
Proof. Suppose c A B. Then, by definition, c A and c B which
implies that c B and c A. But this means, again by definition, that c B A.
Thus A B B A.
Conversely, suppose c B A. Then, by definition, c B and c A which
implies that c A and c B. Again by definition, we get that c A B. Thus
B A A B.
So every element in A B is an element of B A and vice-versa. This means
that these two sets consist of the same elements and are therefore equal.

What should not be lost in this discussion is that a new term was introduced
in the title of the proposition, namely commutativity. It simply means that a
combined with b equals b combined with a for some way of combining things. It is
one of the key ideas in algebra that can sometimes fail, and will come up repeatedly
in the course.
Leaving this fascinating stuff for later, let us return to sets. If you are generous
and have a set that you would like to share with others, then you may be tempted
to break it up into subsets and pass those subsets around. In fact, the idea of
breaking up a set into subsets is a precise and important notion in mathematics
whose definition is given below.
Definition 0.1.2. A partition P of a set A is a collection of non-empty
subsets, P = {Ai }iI indexed by I such that
(1) For any element a A, there is an i I such that a Ai .
(2) For any two distinct elements i, i0 I, the Ai and Ai0 are disjoint.

EXERCISES

In this definition, the indexing set is arbitrary could be called J or S2 or


anything you want. The important thing is that you have a set P of subsets of
Asatisfying (1) and (2). We will encounter many partitions as we progress through
this course, but for now, lets look at a simple example.
Example 0.1.1. For our set Atoms, we can form the partition
P = {An }n{1,...,103}
where An = {atoms with n protons}.
Now lets return back to relationships between sets. One way of relating two sets
is by defining a function or a map from one to another. Here is the mathematical
definition.
Definition 0.1.3. Suppose A and B are sets. A function f from A to B is
a subset f A B such that for every a A there exists exactly one element
c = (a, b) f . A function can be denoted
f :AB
There is a lot of notation that comes along with a function. For example, we
write f (a) as the unique element b for which (a, b) f . We also call A the domain
of f and B the codomain. This latter term should be prevalent in secondary
school but is frequently confused with the different notion of range. The range of
f is defined as the subset
range(f ) = {b B : there is an a A such that (a, b) f }.
Connection 0.1.1. Most middle and high school texts (and too many college
texts) content themselves with saying a function is an assignment. This is a good
description of what a function does, but perhaps not what a function is. Nevertheless, a teacher can usually pull off this type of definition and use it successfully at
the high school and early college level. A worse situation occurs when high school
students are taught that functions always send real numbers to real numbers. This
is a sad injustice that no student of Math 511 will perpetuate! However, a question
from this type of thinking arises. Why does the vertical line test for graphs mean
that a graph is defined by a function?
Functions are most useful when they are combined and compared. The most
common way of combining two functions f : A B and g : B C is by composition.
Definition 0.1.4. If f : A B and g : B C are functions then gf : A C
is defined as the set
g f = {(a, g(f (a))) : a A}.
Exercises
(1) Using the notation developed, write the set of vowels V and the set of
integers I between 100 and 100.
(2) Using your previous notation, write the set EI of even integers between
100 and 100.
(3) Prove that if A B then A B = A.

10

0. PEANO AXIOMS FOR NATURAL NUMBERS - AN INTRODUCTION TO PROOFS

(4) Prove that intersection is an associative operation. I.e. prove that if A, B


and C are sets, then
A (B C) = (A B) C
(5) Suppose P = {Ai }iI is a partition of A. Prove that the set
= {(a, Ai ) : a Ai } A P
defines a function : A P.
(6) Give an example of a set A and two functions f and g with domain and
codomain A such that f g 6= g f .
(7) A function f : A B is one to one if the equality f (a1 ) = f (a2 ) implies
a1 = a2 .
(a) Give an example of a function that is one to one.
(b) Give an example of a function that is not one to one.
(8) A function f : A B is onto if for every b B there exists an a A
such that f (a) = b.
(a) Give an example of a function that is onto.
(b) Give an example of a function that is not onto.
(c) Give an example of a function that is both onto and one to one.
0.2. Peano Axioms
Let us now write down the basic ingredients that come together to produce the
set of the natural numbers which is denoted N.
Axiom 1. The number 0 is a natural number.
Another way of writing this axiom is 0 N. Even though we have just begun,
the first axiom is not without some controversy. To some mathematicians, the
natural numbers start with 1 instead of 0. We will adopt the more mainstream
attitude though and keep Axiom 1 as it is written.
Axiom 2. If n is a natural number, then n++ is also a natural number.
In the context of Peano axioms, the notation n++ is stolen from Terrance Taos
book on real analysis. He, in turn, stole it from the world of computer programming
where it means to add 1 to the number n. The mathematical way of writing Axiom
2 is to say that there is a function ++ : N N. Combining Axioms 1 and 2 gives
us a way of writing some potentially new natural numbers!
Definition 0.2.1. The number 3 is ((0++)++)++.
Let us try now to prove our first proposition about N.
Proposition 0.2.1. The number 3 is a natural number.
Proof. We work our way step by step to show that the proposition is true.
0 is a natural number (Axiom 1).
0++ is a natural number (Axiom 2). Lets call this number 1 from now
on!
1++ is a natural number (Axiom 2). Lets call this number 2 from now
on!
2++ is a natural number.

0.2. PEANO AXIOMS

11

We can appeal to Definition 0.2.1 for the meaning of 3 as ((0++)++)++ = 2++.


So we conclude that it is indeed a natural number.

Before moving on, lets pause and look over that proof again. Each step appealed to either a definition or an axiom. We never made anything up (except the
notation 1 and 2), and directly concluded that the statement in the proposition
was true. Recall that this type of argument, i.e. an unraveling of definitions and
axioms, was called a direct proof.
Lets return to the axioms.
Axiom 3. The number 0 is not n++ for any natural number n.
This axiom says that 0 is not in the range of the function ++.
Axiom 4. If n++ is the same natural number as m++ then n and m are the
same.
To use the language of the last section, Axiom 4 states that the function ++
is one to one. We can now prove the following proposition.
Proposition 0.2.2. 1 does not equal 2.
We will do this by using what is called a proof by contradiction. Heres how it
works.
Step 1) Assume the proposition is false.
Step 2) Arrive at a contradiction.
Step 3) Conclude that the assumption in Step 1) was false and therefor the proposition is true.
Proof. Following Step 1) we will assume that the proposition is false and that
1 does equal 2. Then,
1 = 0++ and 2 = 1++ (Definitions of 1 and 2)
0 = 1 (Axiom 4)
0 = 0++ (Definition of 1 again)
0 is n++ for a natural number n .
This conclusion contradicts Axiom 3. We have reached a contradiction, so our
initial assumption that 1 = 2 must be false. This means that the proposition is
true.

Now that we are getting somewhere, lets throw a real winner into the mix.
Axiom 5 (Induction). Given statements Pn for every natural number n. If
(A) P0 is true,
(B) Pn implies Pn++ for every natural number n,
then Pn is true for all natural numbers n.
This axiom has a bit of vocabulary associated with it. Part (A) is usually called
the base case and one can think of it as the first step of a ladder. Part (B) as a
whole is called the induction step and can be thought of as saying: if you can
get to the n-th step on the ladder, then you can climb to the (n + 1)-st step. The
assumption in Part (B) that Pn is true is called the induction hypothesis.
One immediate consequence to the induction axiom is the following proposition.

12

0. PEANO AXIOMS FOR NATURAL NUMBERS - AN INTRODUCTION TO PROOFS

Proposition 0.2.3. If n N then either n = 0 or n is obtained by applying


++ to 0 a finite number of times, but not both.
Proof. Lets try using induction here. The statement in the proposition can
be written
(Pn ) Either n = 0 or n = ( (0++) )++ but not both.
(A) Base case P0 . The P0 case is true since 0 = 0 and 0 6= m++ by Axiom 3.
(B) Now assume the statement Pn is true (or, with our new vocabulary, assume
the induction hypothesis). Is Pn++ true? If n = 0 then n++ = 0++ and the
statement is true. Otherwise n is obtained by successively applying ++ to 0. But
then n++ is obtained by applying ++ to zero exactly one more time, which is still
a finite number of times. Also, by Axiom 3 we again have that n 6= 0 since it is
the result of applying ++. Thus Pn++ is true and the induction step is proven.
Since we proved the base case and the induction step, we have proved that Pn is
true for all n N by Axiom 5. But Pn being true for all n is the proposition, so
the proposition is proved.

The upshot of this is that we can almost write down the natural numbers as
the set
N = {0, 1, 2, 3, 4, 5, . . .}.
Connection 0.2.1. Induction is used throughout high school education. Examples range from Gauss trick for adding the first 100 (or 103,211, or ...) numbers
together to the binomial theorem. A high school calculus course can use induction
to prove several formulas such as the power rule
d n
x = nxn1 .
dx
One should think of it as an essential instrument in the mathematical toolkit!
We now use induction to prove that addition and multiplication are well defined operations. What does this mean? It means that they make sense. Before
we can be sure they make sense though, we have to define them.
Definition 0.2.2. Addition and multiplication, denoted + and respectively,
are binary operations (i.e. functions from N N to N). Addition is defined as the
operation that satisfies the following two properties for any m N:
(i) m + 0 = m.
(ii) m + (n++) = (m + n)++.
Multiplication is defined as the operation that satisfies the following two properties
for any m N
(i) m 0 = 0,
(ii) m (n + +) = m n + m.
This is what is known as an inductive definition. Lets prove it makes sense.
Proposition 0.2.4. For any natural numbers m and n, there exists a unique
number m + n and a unique number m n.
Proof. We will prove the statement involving addition and leave the multiplication case as an exercise. We prove this by induction (which means we use the
axiom of induction to prove the statement). Take m to be any natural number and
let Pn be the statement

0.2. PEANO AXIOMS

13

(Pn ) There is a unique number m + n.


(A) To prove P0 we just use property (i) to see m + 0 = m.
(B) Now assume m + n is defined and unique. Then by property (ii), m +
(n++) = (m + n)++ so that it is defined. Since m + n is unique, (m + n)++ is
also uniquely defined by Axiom 3.
Having proved both conditions (A) and (B), we have that Pn is true for all natural
numbers n and the proposition is proved.

Try to prove this next proposition out for fun.
Proposition 0.2.5. If n N then n++ = n + 1.
Rather than being coy about these operations and forestalling the inevitable,
lets write down straightaway the most important properties. We do this with the
next two theorems which should be taken as foundational and important. In fact,
without these theorems, practical arithmetic would be nearly impossible.
Theorem 0.2.1 (Algebraic properties of (N, +)). The following properties hold.
Additive identity: For any n N, 0 + n = n = n + 0.
Associativity of addition: For any three natural numbers l, m, n N,
(0.1)

(l + m) + n = l + (m + n)
Commutativity of addition: For any two natural numbers m, n N,

(0.2)

m+n=n+m
Cancellation law for addition: For any natural number n N, if n +
m1 = n + m2 , then m1 = m2 .

Proof. We prove these in order.


Additive identity: The right hand equality n = n + 0 follows from Definition 0.2.2, part (i). To see that 0 + n = n, we use induction. The base
case is 0 + 0 = 0 which again follows from Definition 0.2.2, part (i). Now
assume 0 + n = n. We need to prove that 0 + (n++) = n++. For this,
we see
0 + (n++) = (0 + n)++
= n++

Definition 0.2.2, part (ii)


Induction hypothesis

Associativity of addition: Here we use induction on n with l and m fixed.


For the base case, we need to prove (l + m) + 0 = l + (m + 0). But by the
additive identity result we just proved, we see that (l + m) + 0 = l + m and
l + (m + 0) = l + (m) = l + m. So the base case is proven. Now assume
equation (0.1) holds for n. We must prove
(l + m) + (n++) = l + (m + (n++)).
Observe,
(l + m) + (n++) = ((l + m) + n)++

Definition 0.2.2, part (ii)

= (l + (m + n))++

Induction hypothesis

= l + (m + n)++

Definition 0.2.2, part (ii)

= l + (m + (n++))

Definition 0.2.2, part (ii)

14

0. PEANO AXIOMS FOR NATURAL NUMBERS - AN INTRODUCTION TO PROOFS

Commutativity of addition: Induction again! First lets show that m +


1 = m + + = 1 + m by induction on m. The base case of m = 0 is true
by the additive identity and the definition of 1. Now the induction step
can be shown by observing that
(m++) + 1 = (m + 1) + 1

Proposition 0.2.5

= (1 + m) + 1

Induction hypothesis

= 1 + (m + 1)

Associativity of addition

= 1 + (m++)

Definition of m++

So we have shown that m + 1 = 1 + m for any m N. Now we want to


do this for any n. Again we use induction, this time on n.
The base case is simply the fact that 0 is an additive identity. Now
for the induction step,
m + (n++) = m + (n + 1)

Proposition 0.2.5

= (m + n) + 1

Associativity of addition

= (n + m) + 1

Induction hypothesis

= n + (m + 1)

Associativity of addition

= n + (1 + m)

Commutativity of 1 and m

= (n + 1) + m

Associativity of addition

= (n++) + m

Proposition 0.2.5

Cancellation law for addition: Guess what we use... you got it, induction on n! Base case is the statement that if 0 + m1 = 0 + m2 , then
m1 = m2 . This follows immediately from 0 being the additive identity.
Now for the induction step. Let us assume the statement is true for n and
suppose that (n++) + m1 = (n++) + m2 . Then
(n + m1 )++ = (n + m1 ) + 1

Proposition 0.2.5

= (n + 1) + m1

Associativity and commutativity

= (n++) + m1

Proposition 0.2.5

= (n++) + m2

Assumption

= (n + 1) + m2

Proposition 0.2.5

= (n + m2 ) + 1

Associativity and commutativity

= (n + m2 )++

Proposition 0.2.5

But by Axiom 4, this equality implies that n + m1 = n + m2 . By the


induction hypothesis, this means that m1 = m2 .

And now its multiplications turn.
Theorem 0.2.2 (Algebraic properties of (N, )). following properties hold.
Multiplicative identity: For any n N, 1 n = n = n 1.
Associativity of multiplication: For any three natural numbers l, m, n
N,
(0.3)

(l m) n = l (m n)

0.3. RELATIONS

15

Commutativity of multiplication: For any two natural numbers m, n


N,
mn=nm

(0.4)

Cancellation rule for multiplication: For any natural number n, if n


m1 = n m2 6= 0 then m1 = m2 .
Exercises
(1)
(2)
(3)
(4)

Prove
Prove
Prove
Using

the power rule by induction.


Proposition 0.2.4 for the operation of multiplication.
Proposition 0.2.5.
the proof of Theorem 0.2.1 as inspiration, prove Theorem 0.2.2.
0.3. Relations

We saw in the first section that a function is defined as a subset of a Cartesian


product of sets satisfying a particular property. The idea of a relation is similar to
this, but without the additional property.
Definition 0.3.1. A binary relation on a set A is a subset R A A.
There are some key properties that a relation might satisfy. These terms permeate common language because of their relationship to basic logic.
Definition 0.3.2. Let R A A be a binary relation.
(1) R is called reflexive if (a, a) R for every a A.
(2) R is called symmetric if
(a, b) R implies (b, a) R
for every a, b A.
(3) R is called antisymmetric if
(a, b), (b, a) R implies a = a
(4) R is called transitive if
(a, b), (b, c) R implies (a, c) R
for every a, b, c A.
Lets take some time to humanize these properties.
Example 0.3.1. Suppose A is the set of all of the people in Kansas. The
relation R(blank) is defined as
(a, b) R(blank) if and only if a -blank- b
Which of the properties in Definition 0.3.2 are satisfied by
R(has had lunch with),
R(has the same color hair as),
R(is a step sibling of) and
R(loves)?
There are a couple of relations that have great utility in mathematics. First,
lets go back to preschool and make sure that we understand the size of a number.

16

0. PEANO AXIOMS FOR NATURAL NUMBERS - AN INTRODUCTION TO PROOFS

Definition 0.3.3. A (non-strict) partial order on a set A is a binary relation


R that is reflexive, antisymmetric and transitive.
A set with a partial order is sometimes called a poset. Now, as promised, we
return to our early youth with the following definition.
Definition 0.3.4. If a, b N we say that a is greater than or equal to b, written
a b if there exists c N such that a = b + c.
This definition is the same as giving the relation R N N defined by
a b if and only if (a, b) R .
Lets establish that is indeed a partial order on N.
Theorem 0.3.1. The relation a b is a partial order on N.
Proof. We need to show that the relation is reflexive, antisymmetric and
transitive.
Reflexive: Exercise.
Antisymmetric: We need to prove
If a b and b a then a = b.
By definition, there are natural numbers c1 and c2 such that
a = b + c1
b = a + c2
So that a + 0 = a = b + c1 = (a + c2 ) + c1 = a + (c2 + c1 ). But by the
cancellation law for addition we have that 0 = c2 + c1 . If c1 6= 0, then
c1 = c + + for some c N by Proposition 0.2.3. But then 0 = (c2 + c)++
which contradicts Axiom 3. Thus c1 = 0 and a = b + c1 = b + 0 = b which
proves antisymmetry.
Transitive: Exercise.

The following theorem is a very useful and important property of subsets of N.
Theorem 0.3.2 (Well ordering principle). Given any non-empty subset A N,
there exists a unique smallest element a A.
By the smallest element, we mean an element a A such that if b A then
b a.
Proof. We use induction for this theorem. First, assume that A N does
not contain a smallest element. Now let Pn be the statement
If m N such that n m then m 6 A
Let us show the base case is true. Clearly, if 0 m then m = 0, but m is less
than or equal to all natural numbers and so m 6 A (for otherwise it would be a
smallest element). Now if Pn is true, but Pn++ is false, then n A and there is no
m strictly less than n such that m A. But then for every m A we must have
m n which means n is a smallest element. This is a contradiction, so Pn++ is
true and we have proven Pn for all n. This means that there is no natural number
n A and thus, since A N, A must be the empty set.
In conclusion, if A is non-empty, then it must have a smallest element (for
otherwise the above argument would show it to be empty).


0.3. RELATIONS

17

A partial order is not the only type of binary relation that is important to us.
As we will see very shortly, the following notion may be even more important in
algebra.
Definition 0.3.5. A binary relation R on a set A is an equivalence relation
if it is reflexive, symmetric and transitive.
As we saw above with , often it is convenient to denote a relation with a
symbol separating two elements of the set. We can do this generally for a binary
relation R A A by writing
a R b if and only if (a, b) R.
This is just notation to indicate the relation as a relationship between elements.
For example, in this notation, R is a transitive relation if and only if
a R b and b R c implies a R c.
So what can we do with a relation R on A? Well, we can take each element a A
and make it into a subset [a]R by defining
(0.5)

[a]R := {b A : a R b}.

If we have a fixed relation that we know about, we will just write [a] instead of
[a]R . By the way, the notation := means that we define the left hand side by the
right hand side. A cool fact comes up when R is an equivalence relation.
Proposition 0.3.1. Assume R is an equivalence relation on A. For any a, b
A either [a] = [b] or [a] is disjoint from [b].
Proof. We can do this directly. Suppose [a] and [b] are not disjoint, then
there is a c [a] [b].
Now suppose d [a]. Since c is in [a] we have a R c and since R is symmetric
c R a. Since d is in [a] we have a R d. Thus c R a and a R d which implies
c R d by the transitivity of R. On the other hand, since c [b] we have b R c.
Since b R c and c R d we get b R d which implies d [b]. Thus [a] [b].
But switching the a and b in the above argument shows [b] [a]. Thus [a] = [b]
as was to be shown.

For an equivalence relation R on A we can define
(0.6)

A
:= {S A : there is an a A such that S = [a]}.
R

If Cartesian products are the set theory analog of products of numbers (which they
are), then AR is the set theory analog of a quotient.
Theorem 0.3.3. If R is an equivalence relation then

A
R

is a partition of A.

Proof. By definition, we have that if S and S 0 are in AR then there is an a


and b such that S = [a] and S 0 = [b]. By Proposition 0.3.1, we know that either
S = S 0 which means they are the same element in AR or they are disjoint. Thus
property (2) in Definition 0.1.2 is satisfied. To see property (1), namely that every
element a A is an element of some S AR , simply observe that [a] AR by
definition. But since R is reflexive, a R a and a [a].


18

0. PEANO AXIOMS FOR NATURAL NUMBERS - AN INTRODUCTION TO PROOFS

It is hard to overstate the importance of this last theorem. It manifests itself in


a huge number of constructions in algebra, geometry and analysis. The idea that
an equivalence relation makes partitions means that a R b can be thought of as
saying a is equal to b in some R sense. So if we want to think about a in the R
sense of equality, we only need to think of the element [a] AR . For example, say
we think of the set of students S in the class. We can define an equivalence relation
R as a R b if and only if student a and student b get the same letter grade. Then
the quotient SR is the list of grades that the students will receive. Mike and Mary
are equal from the perspective of R if they get the same grade, otherwise they are
not equal. The map : S SR from the exercises in Section 0.1 will be equal on
R-equivalent elements and will be distinct on R-inequivalent elements.
Exercises
(1) Give an example of a binary relation on a set A that satisfies exactly two
of the conditions in Definition 0.3.2.
(2) Prove the reflexive and transitive properties in Theorem 0.3.1.
(3) Give an example of a partially ordered set that does not satisfy the well
ordering principle.
(4) Show that there is a converse to Theorem 0.3.3 in the following sense. If
P is a partition of A, define a relation
R = {(a, b) A A : there exists S P such that a, b S}.
Prove that
(a) R is an equivalence relation.
(b) P = AR .

CHAPTER 1

Basic Arithmetic
While the title of this chapter may strike the college upper classmen as slightly
offensive, it is my hope that the impression will be overcome by a study of its
contents. For many mathematicians, a modern viewpoint on arithmetic is more
subtle and complicated than several other advanced sounding subjects. So what
do I mean by arithmetic? I mean working with integers and rational numbers and
their basic operations.

1.1. The set Z


In this section we utilize the construction of the natural numbers N to construct
the integers Z. Let us first define the binary relation on N N via
(1.1)

(a, b) (c, d)

if and only if

a + d = b + c.

There are some fundamental facts about this relation that we now establish.
Proposition 1.1.1. The following statements hold with respect to relation 1.1.
(1) The relation is an equivalence relation.
(2) For every (a, b) there exists a unique natural number c N for which
either (a, b) (0, c) or (a, b) (c, 0) with both occurring if and only if
c = 0.
(3) If (a1 , b1 ) (a2 , b2 ) and (c, d) N N then
(1.2)

(a1 + c, b1 + d) (a2 + c, b2 + d).


(4) If (a1 , b1 ) (a2 , b2 ) and (c, d) N N then

(1.3)

(a1 c + b1 d, a1 d + b1 c) (a2 c + b2 d, a2 d + b2 c)

Proof. We will prove the first two properties and leave the last two as exercises.
(1) We need to prove that is reflexive, symmetric and transitive.
reflexive: Let (a, b) N N then a + b = a + b since addition is well
defined. By definition of , this implies (a, b) (a, b) and so is
reflexive.
symmetric: Suppose (a, b) (c, d). Then c + b = b + c = a + d = d + a
by commutativity of addition. Thus (c, d) (a, b) which shows that
is symmetric.
19

20

1. BASIC ARITHMETIC

transitive: Now assume (a, b) (c, d) and (c, d) (e, f ). Then


d + (a + f ) = (d + a) + f,

associativity

= (a + d) + f,

commutativity

= (b + c) + f,

definition of

= b + (c + f ),

associativity

= b + (d + e),

definition of

= (d + e) + b,

commutativity

= d + (e + b),

associativity

= d + (b + e).

commutativity

By the cancellation property in Theorem 0.2.1, this implies that a +


f = b + e which in turn yields (a, b) (e, f ). Thus is transitive.
(2) We first show the existence of such a c. Let us consider the equivalence
class of the pair (a, b) which is defined as
[(a, b)] := {(c, d) N N : (a, b) (c, d)}.
This is the set of elements that are -equivalent to (a, b). Now we define
another set,
S(a,b) = {c N : there exists d N such that (c, d) [(a, b)] }.
Note that S(a,b) is non-empty since a S(a,b) . Thus, by the Well Ordering
Principal of N, there is a smallest element e S(a,b) . If e = 0 then there is
an element (0, c) [(a, b)] and we have shown existence. If e > 0, then
we claim (e, 0) [(a, b)] . If not, then (e, f ) [(a, b)] with e > 0 and
f > 0. Thus f = f + 1 and e = e + 1 for natural numbers e, f. Note e > e
and
e + f = (
e + 1) + f,
definition of e
= e + (1 + f),
= e + (f + 1),

associativity

= e + f,

commutativity
definition of f

= f + e.

commutativity

Thus (e, f ) (
e, f) implying (
e, f) [(a, b)] and e S(a,b) . But since
e > e and e was assumed to be the smallest element of S(a,b) , we have
achieved a contradiction. So we must have that (e, 0) [(a, b)] showing
the existence of c.
Now we come to uniqueness. There are three options to consider.
Suppose (c, 0) (a, b) (c0 , 0). Then, since is transitive (c, 0)
(c0 , 0), we have c = c + 0 = 0 + c0 = c0 .
Suppose (c, 0) (a, b) (0, c0 ). Then, since is transitive (c, 0)
(0, c0 ), we have c + c0 = 0 + 0 = 0. But this implies that c0 = 0 = c
(otherwise 0 = n++ for some natural number n, violating Axiom 3).
Suppose (0, c) (a, b) (0, c0 ). Then, since is transitive (0, c)
(0, c0 ), we have c0 = 0 + c0 = c + 0 = c.


1.2. THE RING Z

21

Do not worry, it is OK if you are feeling lost. This theorem may have looked
looked arbitrary and unneccessary, but now lets see the motivation by thinking
about the next definition.
Definition 1.1.1. The set of integers, denoted Z is the quotient
NN
(1.4)
Z=

If (a, b) (c, 0), we denote [(a, b)] by c. If (a, b) (0, c) for c > 0, we denote
[(a, b)] by c.
Thus, we have introduced negative numbers by partitioning relative to the
equivalence relation . Before moving on, we should assess what we have and what
we do not have! What we have is the set of integers Z. However, we do not yet have
arithmetic of the integers. In fact, we do not even know how to add or multiply
two integers, much less whether these operations satisfy the properties in Theorem
0.2.1.
Exercises
(1) Prove part 3) of Proposition 1.1.1.
(2) Prove part 4) of Proposition 1.1.1.
(3)
1.2. The ring Z

You might also like