MA138 Lecture Notes
MA138 Lecture Notes
Contents
1 Introduction 3
5 Complex Numbers 23
5.1 The field of complex numbers . . . . . . . . . . . . . . . . . . 23
5.2 Complex conjugation and the absolute value . . . . . . . . . . 24
5.3 The exponential function . . . . . . . . . . . . . . . . . . . . . 25
5.4 Trigonometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.5 The complex plane . . . . . . . . . . . . . . . . . . . . . . . . 28
5.6 The quadratic formula . . . . . . . . . . . . . . . . . . . . . . 28
5.7 Roots of unity . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.8 Polar form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
7 Polynomials 42
7.1 Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
7.2 Division with remainder for polynomials . . . . . . . . . . . . 44
7.3 Unique factorisation for polynomials . . . . . . . . . . . . . . 45
7.4 The Euclidean algorithm . . . . . . . . . . . . . . . . . . . . . 47
7.5 Roots of polynomials . . . . . . . . . . . . . . . . . . . . . . . 48
7.6 C is algebraically closed . . . . . . . . . . . . . . . . . . . . . 49
7.7 Algebraic numbers . . . . . . . . . . . . . . . . . . . . . . . . 50
1 Introduction
The modules MA132 Foundations and MA138 Sets and Numbers come with
the same set of lecture notes written by David Mond.
You’re now looking at my alternative lecture notes which cover more or
less the same topics. For your convenience I have used the same titles for
chapters. Titles for sections are different.
If you carefully study either set of lecture notes then you’re in good shape
for the exam. More important than looking at a second set of lecture notes,
is doing lots of exercises.
We assume that you are familiar with the basics of the integers
Z= . . . , −2, −1, 0, 1, 2, . . . .
You may study the integers from the ground up (and much deeper) in later
modules.
The natural numbers are 0, 1, 2, . . .. In words, they are the nonnegative
integers. We write N for the set of natural numbers.
Proof. Exercise.
Proof. Exercise.
Lemma 4. Let a, n be integers and 0 < a < n. Then n does not divide a.
Proof. We cannot prove this here because it goes too close to the axioms of
mathematics for our scope.
4 Chapter 2 December 24, 2018
Exercise (2.1) Use the WOP to prove that there exists a greatest integer n
such that n2 < 1000.
Proof. Suppose there are only finitely many prime numbers. Denote them by
p1 , . . . , pk and put n = p1 · · · pk + 1. Then pi ∤ n for all i. Also, n ≥ 2 so n
is divisible by a prime number q by lemma 7. So q is a prime number not in
{ p1 , . . . , pk }. This is a contradiction and finishes the proof.
n | (m − r2 ) − (m − r1 ) = r1 − r2 = s.
(1) In
this case, unique means that there is at most one such pair (q, r). This is the first
time we are seeing uniqueness. It is all over mathematics. Proofs often treat existence and
uniqueness separatedly, as is the case here.
MA132 Foundations / MA138 Sets and Numbers 5
2.2 Induction
Proposition 11: Induction, second form. For all natural numbers n let P(n)
denote a statement (2) . Assume that P(0) is true. Assume P(n) ⇒ P(n + 1)
for all n ≥ 0. Then P(n) holds for all n.
Proof. Let T be the set of natural numbers n such that P(n) is true. Then
conditions (a) and (b) of proposition 10 hold. So T = N. That is, P(n) is
true for all n.
Example 12. We are now ready for our first example of proof by induction.
Prove that for all n ≥ 0 we have
1 + 2 + · · · + n = n(n + 1)/2.
Solution. As this is our first proof of induction we shall use rather formal
language. For n ≥ 0 let P(n) denote the statement that 1 + 2 + · · · + n =
n(n + 1)/2.
The assertion P(0) states that 0 = 0 and is therefore true.
Assuming that P(n − 1) is true for a fixed n ≥ 1, we shall prove P(n):
1 + 2 + · · · + n = 1 + 2 + · · · + (n − 1) + n
= n(n − 1)/2 + n by P(n − 1)
= n(n + 1)/2
Solution. This is our second example of induction and is put in less formal
language.
Induction on n. For n = 0 the equation reads 0 = 0 which is true.
Assuming the result for n (n ≥ 0) we prove it for n + 1:
n n−1
Σ xk = Σ xk + xn
k= 0 k= 0
1 − xn
= + xn by the induction hypothesis
1−x
1 − xn + (1 − x) xn 1 − xn+1
= = .
1−x 1−x
Proposition 14: Induction, third form. Let T ⊂ N. Assume that for n ≥ 0
we have
{0, . . . , n − 1} ⊂ T ⇒ n ∈ T.
Then T = N. Note that for n = 0 the convention is {0, . . . , n − 1} = ∅.
Proof. Exercise.
Remark 16. Here is a more precise way to state the uniqueness part in the
theorem. Let p1 , . . . , pk and q1 . . . , qℓ be prime numbers and assume
p1 · · · pk = q1 · · · qℓ .
a = p1 · · · pk , b = q1 · · · qℓ
n = ab = p1 · · · pk q1 · · · qℓ
n = p1 · · · pk = q1 · · · qℓ
where pi and q j are prime numbers, and the two factorisations are distinct
even up to reordering.
Suppose p1 = q1 . Then n/ p1 would have two distinct factorisations,
contradicting minimality of our counterexample n. Likewise, pi 6= q j for all
i, j. Moreover, pi ∤ q j and q j ∤ pi for all i, j because the pi and the q j are
prime.
We may assume p1 < q1 . Put s = (q1 − p1 )q2 · · · qℓ . Then s < n so s
admits a unique factorisation by minimality of the counterexample n.
We have
s = q1 · · · qℓ − p1 q2 · · · qℓ
= p1 · · · pk − p1 q2 · · · qℓ = p1 ( p2 · · · pk − q2 · · · qℓ ).
Exercise (2.2) Prove existence in the above theorem using the WOP instead
of induction. Prove uniqueness in the above theorem using induction instead
of the WOP.
Corollary 17. Let p be a prime number. Let a, b ∈ Z be such that p | ab. Then
p | a or p | b.
Proposition/Definition 18.
(a) Let m, n ∈ Z not both be zero. A common divisor of m, n is an integer
dividing both m and n. Among the common divisors of m, n there is
greatest one. It is written hcf(m, n) or gcd(m, n) and called the highest
common factor or greatest common divisor of m, n.
(b) Let m, n ∈ Z r {0}. A common multiple of m, n is an integer divisible
by both m and n. Among the positive common multiples of m, n there is
least one. It is written lcm(m, n) and called the least common multiple
of m, n.
Proof. Exercise.
m= Πi pimi , n= Πi pini .
(a) We have m | n if and only if mi ≤ ni for all i.
(b) We have hcf(m, n) = ∏i min(mi , ni ).
(c) We have lcm(m, n) = ∏i max(mi , ni ).
8 Chapter 2 December 24, 2018
Proof. It is clear that (b)–(e) follow from (a). Also (f) follows from (b) and
(c).
Proof of (a). The proof of ⇐ is an easy exercise. Proof of ⇒. Suppose
m | n, say, n = md.
We may assume that { p1 , . . . , pk } contains all prime divisors of d (if not,
enlarge this set by more prime numbers and set m j = n j = 0 for the new
indices). Write
d
d = Π pi i
i
Let m, n ∈ Z>0 . One way to find hcf(m, n) is to factorise both and use
proposition 19(b). We won’t use that method because there is a much faster
method known as the Euclidean algorithm.
Example 22. We shall show how to compute hcf(226, 126) by the Euclidean
algorithm. We write n1 = 226, n2 = 126.
Next we put n3 = n1 − n2 = 226 − 126 = 100.
Next we put n4 = n2 − n3 = 126 − 100 = 26.
Next we put n5 = n3 − 4n4 = 100 − 4 × 26 = 100 − 104 = −4.
Next we put n6 = n4 + 6n5 = 26 − 6 × 4 = 26 − 24 = 2.
Finally, n7 = n5 + 2n6 = −4 + 2 × 2 = −4 + 4 = 0.
So hcf(n1 , n2 ) = |n6 | = 2.
Imagine though that we knew already that hcf(26, −4) = 2. Then com-
puting ni for i > 5 wouldn’t be necessary.
Remark 31. We could extend the above proposition by the following coun-
terpart to (c) which is completely trivial. We have C ⊂ A ∩ B if and only if
C ⊂ A and C ⊂ B. In words: A ∩ B is the greatest subgroup of Z contained
in A and B.
x − y = ( p + r) − (q + s) = ( p − q) + (r − s) ∈ A + B.
Finally our endeavours with subgroups pay off in the following theorem.
mZ + nZ = hZ.
(b) Retain the notation of (a). Then x is a common divisor of m, n if and only
if it divides h.
(c) Suppose m and n are both nonzero and put ℓ = lcm(m, n). Then
mZ ∩ nZ = ℓZ.
We’ve proved parts (b) and (d) of the above theorem before (using unique
factorisation of integers) but we prove them again as a byproduct of (a) and
(c). Parts (a) and (c) are new.
Proof. Proof of (a) and (b). By proposition 27 there exists unique d > 0 with
mZ + nZ = dZ. We must prove: x is a common divisor of m, n if and only if
12 Chapter 3 December 24, 2018
x | d. Well,
x|m and x | n
⇔ xZ ⊃ mZ and xZ ⊃ nZ by lemma 28
⇔ xZ ⊃ mZ + nZ by proposition 30(c)
⇔ xZ ⊃ dZ
⇔ x|d by lemma 28.
Proof of (c) and (d). By proposition 27 there exists unique d > 0 with
mZ ∩ nZ = dZ. We must prove: x is a common multiple of m, n if and only
if d | x. Well,
m | x and n | x
⇔ xZ ⊂ mZ and xZ ⊂ nZ by lemma 28
⇔ xZ ⊂ mZ ∩ nZ
⇔ xZ ⊂ dZ
⇔ d|x by lemma 28.
Corollary 33. Let m, n ∈ Z, not both zero. Then there are a, b ∈ Z such that
am + bn = hcf(m, n).
Remark 34. Let m, n ∈ Z. So far, hcf(m, n) is only defined if (m, n) 6= (0, 0),
and lcm(m, n) is only defined if m, n are nonzero.
It is common to define them in the remaining cases by the conditions that
parts (a) and (c) of theorem 32 hold for all integers m, n; and hcf(m, n) ≥ 0
and lcm(m, n) ≥ 0.
Equivalently: hcf(m, 0) = |m| and lcm( p, q) = 0 if pq = 0.
Let m, n ∈ Z be not both zero and put d = hcf(m, n). Recall the Euclidean
algorithm (example 22 and the material just before it) for computing d.
The extended Euclidean algorithm does the same, then goes on to find
a, b such that am + bn = d which we know to exist by corollary 33. It works
by going backwards through the sequence nm , nm−1 , . . . , n1 .
Let’s do this in the case of example 22. Recall: n1 = 226, n2 = 126 and
n3 = n1 − n2 n4 = n2 − n3 n5 = n3 − 4n4
n6 = n4 + 6n5 = 2 = d n7 = n5 + 2n6 = 0.
Definition 35. Let A be a set and write A2 or A × A for the set of pairs (a, b)
with a, b ∈ A. A binary operation on A is a map f : A2 → A.
A Z C C R {maps R → R}
f (a, b) a + b a + b ab max(a, b) a◦b
name sum sum product maximum composite.
Definition 38. A field is a commutative ring A with the property that for all
nonzero a ∈ A there exists b ∈ A such that ab = 1. Notation: b = a−1 . We
say that b is an inverse to a. We pronounce a−1 as: a inverse.
Remark 40. You don’t need to remember the definition of ring by heart for
this module. You should know the names and meanings of associative and
commutative.
Definitions as the above one, and their study, belong to what is sometimes
called abstract algebra. It is supposed to be outside our scope. On the other
14 Chapter 3 December 24, 2018
hand, maybe you find things easier to understand when the full definitions
are at least briefly presented.
All our examples of rings are commutative.
Exercise (3.4) This is outside our scope. Some basic consequences of the
axioms of rings and fields. Let A be a ring.
(a) Suppose that A is a field. Let a, b ∈ A be such that ab = 0. Prove that
a = 0 or b = 0. (This fails in some rings.)
(b) Prove (a + b) − b = a for all a, b ∈ A.
(c) Prove 0a = 0 = a0 for all a ∈ A.
(d) Let z ∈ A and suppose z + a = a for all a ∈ A. Prove: z = 0. In words:
There is only one zero.
(e) Let e ∈ A and suppose ea = a = ae for all a ∈ A. Prove: e = 1. In
words: There is only one identity.
(f) Prove that distributivity holds for − instead of +, that is, a(b − c) =
ab − ac and (a − b)c = ac − bc for all a, b, c ∈ A.
(g) Higher associativity. For a1 , . . . , ak ∈ A we define the product a1 · · · ak
recursively by a1 · · · ak = (a1 · · · ak−1 )ak . Prove:
a1 · · · ak (b1 · · · bℓ )c 1 · · · c m = a1 · · · ak b1 · · · bℓ c 1 · · · c m
whenever k, ℓ, m ≥ 0 and ai , bi , ci ∈ A. Hint: induction on k + ℓ + m.
(h) State and prove higher associativity for addition in a ring (it is almost
the same as for multiplication and possibly easier because addition is
commutative).
We shall not construct Z. In later modules you may learn how it is con-
structed, or study it more deeply.
In later modules you may learn how Q and R are constructed from Z. We
spend a few words on this later on in chapter 4. In chapter 5 we construct C
from R even though our tool set is limited.
But first we use Z to construct a new ring Zn for n ∈ Z>0 , in the next
section.
3.5 Zn
So n divides a1 := R( x) R( y) − xy.
But n | R( z) − z for all z ∈ Z. So n divides a2 := R( R( x) R( y)) −
R( x) R( y) and a3 := xy − R( xy).
So n divides a1 + a2 + a3 (by lemma 2 used twice), that is, n divides
R( R( x) R( y)) − R( xy). So R( R( x) R( y)) and R( xy) are two elements of Zn
whose difference is divisible by n. They must be equal and the proof is
finished.
x + n y : = Rn ( x + y ), x ×n y := Rn ( xy).
Example 45. Here are the tables for addition and multiplication in Z3 :
+ 0 1 2 × 0 1 2
0 0 1 2 0 0 0 0
1 1 2 0 1 0 1 2
2 2 0 1 2 0 2 1
Proof. We’ll do just one axiom. The others are similar. Let x, y, z ∈ Zn . We
prove
( x × n y) × n z = x × n ( y × n z)
as follows:
( x ×n y) ×n z
= R( R( xy) z) by definition of ×n
= R( R( xy) R( z)) by lemma 42 and since R2 = R
16 Chapter 3 December 24, 2018
Exercise (3.8) Outside our scope. Let n ≥ 1. In this exercise you will prove
that Zn is a field if and only if n is a prime number.
(a) Prove ⇒. Hint: if a, b belong to a field and ab = 0 then a = 0 or b = 0.
(b) Assume n is a prime number and let x ∈ Zn r {0}. Prove that x, n are
coprime. Use theorem 32(a) to deduce that there exists y ∈ Zn such
that x ×n y = 1.
(c) Assume n is a prime number and let x ∈ Zn r {0}. Define f : Zn → Zn
by f ( y) = x ×n y. Prove that f is injective (see definition 133). It fol-
lows that f is surjective by the pigeon hole principle (proposition 136).
Deduce that there exists y ∈ Zn such that x ×n y = 1.
Exercise (3.10) For all odd prime numbers p with p ≤ 17 find out if there
exists x ∈ Z p such that x2 + 1 = 0. (The values of x are less important).
Guess a pattern.
so x2 y2 ≡n x1 y1 .
The other parts are left as an exercise.
Example 52. Let’s compute R29 (219 ) without computing 219 . Write n = 29:
25 ≡n 32 ≡n 3,
215 ≡n (25 )3 ≡n 33 ≡n −2,
219 ≡n 215 · 24 ≡n (−2) · 24 ≡n −32 ≡n 26.
Informally, (a, b) stands for ab−1 but we cannot use that (yet). For (a, b) ∈ Q
and x ∈ Z r {0} we define
(a) Let u, v ∈ Z not both be zero. Prove that [u, v] is well-defined and
express it in u, v only.
(b) Prove:
We’ve seen the definition of fields in definition 38. And Q is a field. There are
other fields than Q but only a small number of additional axioms are needed
to single out Q.
Definition 53. Two fields K, L are isomorphic if there exists a bijective (4)
map f : K → L such that f ( x + y) = f ( x) + f ( y) and f ( xy) = f ( x) f ( y) for
all x, y ∈ K.
Theorem 55: Axioms for Q, first version. There exists an infinite field Q
without subfields other than itself. It is unique up to isomorphism.
Definition 56. An ordered field is a field K along with a total ordering (5)
Theorem 58: Axioms for Q, second version. There exists an ordered field
Q without subfields other than itself. It is unique up to isomorphism.
We shall construct R≥0 using Dedekind cuts, then construct R from R≥0 .
Definition 60. We define R≥0 to be the set of Dedekind cuts that are con-
tained in Q>0 . We put 0R := Q>0 and 1R := Q>1 . For A, B ∈ R≥0 we
define
Exercise (4.4) Prove that A + B, AB and A−1 defined above are again in
R≥0 , that is, are Dedekind cuts contained in Q>0 .
Exercise (4.6) There exists a natural bijection between R and the set of
Dedekind cuts. What would go wrong if we defined the sum and product of
any two Dedekind cuts by (61)?
A = x ∈ Q> 0 x 2 > 2 ,
B = Q> 2 .
Prove: A2 = B.
Exercise (4.8) Prove that there exists a bijection K × R≥0 → R≥0 × R≥0
defined by (( A, B), C) 7→ ( A + C, B + C). So [ X, Y ] is well-defined.
( A, B) + (C, D ) := [ A + C, B + D ],
( A, B)(C, D ) := [ AC + BD, AD + BC].
Also, ( A, 0)−1 = ( A−1 , 0) and (0, A)−1 = (0, A−1 ). Finally, 0K := (0R , 0R )
and 1K := (1R , 0R ).
20 Chapter 4 December 24, 2018
Definition 63. Two ordered fields K, L are isomorphic if there exists a bijec-
tive map f : K → L such that f ( x + y) = f ( x) + f ( y) and f ( xy) = f ( x) f ( y)
for all x, y ∈ K; and such that f takes positive elements in K to positive
elements in L.
Theorem 64: Axioms for R. There exists an ordered field R with supremums.
Moreover, any two such are isomorphic.
Proof. This is outside our scope. It is a nice topic for a 2nd year essay.
Theorem 65. There exists a bijective map from the set of the Dedekind cuts to
R taking a Dedekind cut to its infimum.
There are two ways for constructing the real numbers. One uses Dedekind
cuts as we’ve done in section 4.3. The other uses Cauchy sequences which
we won’t go into here.
Proof. Exercise.
MA132 Foundations / MA138 Sets and Numbers 21
S(k1 , k2 , . . . ) : = Σ kn · 10−n .
n≥1
Remark 69. You are certainly familiar with the decimal expansions of real
numbers x such that x ≥ 1; they involve digits to the left of the decimal
point. We won’t be interested in these. It is easy and left to you to generalise
our results to real numbers outside the range 0 ≤ x < 1.
Proposition 70. Every real number x with 0 ≤ x < 1 is given by at least one
decimal expansion.
Proof. Exercise.
The identity
follows from proposition 67 (geometric series) and shows that some real
numbers admit more than one decimal expansion.
Proposition 72. Every real number x with 0 ≤ x < is given by exactly one re-
duced decimal expansion. Conversely, the sum y associated to a reduced decimal
expansion satisfies 0 ≤ y < 1.
Proof. Exercise.
S(k1 , k2 , . . . )
p
= (1 − 10− p )−1 Σ kr · 10−r (this was proved in (75))
r=1
1 p p t
= p 10 Σ kr · 10−r = = S(ℓ1 , ℓ2 , . . .).
10 − 1 r=1 s
Suppose first that x admits only one decimal expansion. Then kn = ℓn and
the result follows.
Suppose finally that x admits more than one decimal expansion. In fact,
there are then two decimal expansions, one ending in zeroes only and one
ending in nines only. Both decimal expansions are periodic as required.
MA132 Foundations / MA138 Sets and Numbers 23
so 3/7 = .428571.
5 Complex Numbers
5.1 The field of complex numbers
Definition 77. A complex number is a pair of real numbers (ba). The set
of complex numbers is written C. Addition and multiplication of complex
numbers are defined as follows:
+c −bd
(ba) + (dc ) = (ba+ d ), (ba)(dc ) = (ac
ad+bc ) .
You are probably used to a different notation for complex numbers. We’ll
come to that. The above definition is better than saying something like a
complex number is an expression of the form a + bi because that would use +
and i before they’ve even been defined!
a c e a c+e a(c+e)−b(d+ f )
x( y + z) = (b )(( d ) + ( f )) = (b )( d+ f ) = ( )
a(d+ f )+b(c+e)
ac−bd ae−b f a c a e
= ( ad+bc ) + ( a f +be ) = (b )(d ) + (b )( f ) = xy + xz.
24 Chapter 5 December 24, 2018
Lemma 79. Let a, b ∈ R. Then (0a) + (0b) = (a+0 b), (0a)(0b) = (ab0 ).
Proof. This is easy.
Proof. We have
0 2
i 2 = (1) by definition of i
−1
=(0) by multiplication in C
= −1 ;
a b 0 b 0
a + bi = (0) + (0)(1) because b = (0) and i = ( 1)
a 0
= (0) + ( b) by multiplication in C
a
= (b) by addition in C.
a + bi = a − bi.
Proof. Exercise.
MA132 Foundations / MA138 Sets and Numbers 25
Lemma 86.
(a) Let n ≥ 0 and define f : C → C by f ( x) = xn . Then f ( x ) = f ( x) for all
x ∈ C.
(b) Let n ≥ 0 and a0 , . . . , an ∈ R. Define f : C → C by
n
f ( x) = Σ ak xk
k= 0
h( x) = an xn = an xn = an x n = h( x ).
f ( x) = g( x) + h( x) = g( x) + h( x) = g( x ) + h( x ) = f ( x ).
Definition 87.
(a) For x ∈ R we define | x| := max( x, − x).
(b) For x ∈ C we define | x| = ( x x )1/2 . Note that x x ∈ R≥0 ; we take the
positive square root.
We call | x| the absolute value of x. Note that for real numbers x the two
definitions of | x| agree.
Theorem 88.
(a) For a, b ∈ R we have | a + bi |2 = a2 + b2 .
(b) For x ∈ C we have | x| = 0 if and only if x = 0.
(c) For x, y ∈ C we have | xy| = | x| · | y|.
(d) For x, y ∈ C we have the triangle inequality | x + y| ≤ | x| + | y|.
Proof. Exercise.
In this section and the next, agreement 66 about infinite sums applies again.
26 Chapter 5 December 24, 2018
xn
Definition 89. For x ∈ C we define exp( x) = Σ .
n ≥ 0 n!
Proof. Exercise. Hint: First prove the recursion formula (nk) = (n− 1 n−1
k ) + ( k− 1 )
for 1 ≤ k ≤ n assuming (nk) = 0 if 0 ≤ n < k. Then prove the binomial
theorem by induction on n.
A = (n, k) ∈ Z2 0 ≤ k ≤ n
B = (k, ℓ) ∈ Z2 k ≥ 0, ℓ ≥ 0 .
( x + y)n
exp( x + y) = Σ
n≥0 n!
1 n n! xk yn−k
= Σ Σ xk yn−k = Σ
n ≥ 0 n! k= 0 k! (n − k)! (n,k)∈ A k! (n − k)!
xk yℓ xk yℓ
= Σ k! ℓ!
= Σ Σ = exp( x) exp( y).
(k,ℓ)∈ B k≥ 0 k! ℓ≥ 0 ℓ!
Second proof of theorem 90. This uses all sorts of analysis including com-
plex differentiation. Fix y ∈ C and put f ( x) = exp( x + y) for x ∈ C. Using
the transformation n − 1 = k we find
( x + y)n ′ ( x + y)n ′
f ′ ( x) = Σ = Σ
n≥0 n! n≥1 n!
n( x + y )n − 1 ( x + y)k
= Σ = Σ k! = f (x).
n≥1 n! k≥ 0
Remark 93. If you’re unwilling to use complex differentation but are willing
to use real differentiation then you may still benefit from the second proof.
It then proves the functional equation for real x, y.
It is possible to make sense of the functional equation and its two proofs
in a way that rules out all analysis, keyword: power series. This is outside
our scope.
5.4 Trigonometry
The best way to develop trigonometry uses complex numbers and the expo-
nential function.
Theorem/Definition 97. There exists unique least π > 0 such that e2π i = 1.
For x ∈ C we have e x = 1 if and only if x ∈ 2π iZ.
= Re ecos x (cos (sin x) + i sin (sin x)) = ecos x cos (sin x).
For the present purpose, let’s identify R2 with C. Choose a point on the
blackboard and call it the orign. A number or vector (a, b) = a + bi ∈ C = R2
defines the following point on the blackboard. Starting at the origin, go right
by a and go up by b.
The proof of this is left to you. It is the same proof as for real numbers. We
shall soon see that every complex number has a square root (see proposi-
tion 109). We call b2 − 4ac the discriminant. If the discriminant is nonzero
then our quadratic equation has precisely two complex solutions.
Remark 104. This is outside our scope and aimed at those of you who know
about groups. The above proposition follows from the fact that µ (n) is a
group under multiplication. It is cyclic of order n and generated by ζn .
Figure 1. Let n > 1. We shall not define convex hull or regular n-gon but just say that
the convex hull of µ (n) is an example of a regular n-gon. Here are pictures of them.
ζ3 b
i
b
b
−1 b b
1 1
b
b
ζ32 −i
ζ5 ζ62 ζ6
b
b b
ζ52 b
ζ63
b b b
1 1
b
ζ53 b
b b
Proposition 109. Let a be a nonzero complex number. Then there are n distinct
solutions x ∈ C to xn = a. The solution set is { x0 y | y ∈ µ (n)} whenever x0
is one solution. One nth root of aeix (with a, x ∈ R and a > 0) is given by
a1/n eix/n .
Definition 110. Let Ω be a set of two elements. Common notations for the
elements of Ω are:
0 = F = False, 1 = T = True.
Example 112. It may be surprising that there are only two statements. For
example, the statement P(n): n is divisible by 3 seems neither true nor false.
What’s going on is, this is a map P: Z → Ω!
Remark 113. If you study logic at a higher level then you learn that, in-
avoidably, there are statements which are neither true nor false (and not of
the sort in example 112). See section 8.5 for an example. For us it doesn’t
do any harm to ignore this.
Exercise (6.2) Write down the table for a or b but not both. This is different
from a ∨ b!
Example 119. A widely used method for proving theorems is proof by con-
tradiction. Behind it is the fact that (a ⇒ b) is ( b ⇒ a).
Proof by contradiction works as follows. We want to prove a ⇒ b. In
order to do this we assume a and b. Then the actual arguments follow. We
conclude a. This is a contradiction. This proves a ⇒ b.
a⇔b = ( a ⇒ b) ∧ (b ⇒ a).
Proof. Exercise.
which is absurd.
6.2 Sets
x ∈ A∪B ⇔ x ∈ A or x ∈ B
x ∈ A∩B ⇔ x ∈ A and x ∈ B
x ∈ ArB ⇔ x ∈ A and x 6∈ B.
( A ∪ B) ∪ C = A ∪ ( B ∪ C) ∪ is associative
( A ∩ B) ∩ C = A ∩ ( B ∩ C) ∪ is associative.
Proof. This is easy.
A Venn diagram means that every set is a subset of the blackboard. You
start with two or three discs called A, B, C. Then more sets are obtained
using the set operations such as union and intersection. In the following, to
make clear which set is meant we colour it gray.
A B A B A B
C B C B C B
A A A
A∪B A∪C ( A ∪ B) ∩ ( A ∪ C ).
The picture in the top left corner shows that all 8 cases are dealt with as
required.
Definition 128. From now on we assume that the union of all sets exists,
and is written U (the universe). For us, this assumption doesn’t do any
harm though it is known that it isn’t necessarily correct (see Russell’s paradox
below). The complement of a set A is Ac := U r A.
Proposition 130. Suppose A, B are sets and x an element. Consider the state-
ments a := ( x ∈ A), b := ( x ∈ B). Then
(a) x ∈ ( A ∪ B) ⇔ a ∨ b.
(b) x ∈ ( A ∩ B) ⇔ a ∧ b.
( a ∨ b) ∧ ( a ∨ c ) = a ∨ (b ∧ c )
( a ∧ b) ∨ ( a ∧ c ) = a ∧ (b ∨ c ).
Proof. A first proof is by inspection of 8 cases for each law. A second proof is
by using the analogy between logic and sets.
There exist sets A, B, C and an element x such that a = ( x ∈ A), b =
( x ∈ B), c = ( x ∈ C). Using proposition 130 and the distributivity laws for
sets (proposition 127)
( a ∨ b) ∧ ( a ∨ c )
⇔ [( x ∈ A) ∨ ( x ∈ B)] ∧ [( x ∈ A) ∨ ( x ∈ C)]
⇔ ( x ∈ A ∪ B) ∧ ( x ∈ A ∪ C) ⇔ x ∈ ( A ∪ B) ∩ ( A ∪ C)
⇔ x ∈ A ∪ ( B ∩ C) ⇔ ( x ∈ A) ∨ ( x ∈ B ∩ C)
⇔ ( x ∈ A) ∨ [( x ∈ B) ∧ ( x ∈ C)] ⇔ a ∨ (b ∧ c).
Likewise for the other law.
Use the results of exercise 6.1 to prove p ⇔ q (don’t use proposition 131).
Example 132. Here are some more examples of the analogy between logic
and sets. Let A, B be sets and x an element. Define a := ( x ∈ A), b := ( x ∈
B). Then
a = ( x ∈ Ac ), ( a ⇒ b) = ( x ∈ Ac ∪ B).
6.5 Mappings
The naive definition of map is good enough for us, and goes like this. Let
A, B be sets. A map f : A → B is something that provides us with an element
f ( x) of B for each x ∈ A.
We call A the source or domain of f and B the target or codomain of f .
Don’t confuse the target of f with the image { f ( x) | x ∈ A}. In some areas
of mathematics, including most of these notes, source and target are ‘part’
of the map f , that is, a map has exactly one source and exactly one target.
Other areas of mathematics are more relaxed about this.
6.6 Bijections
Proposition 136: Pigeon hole principle. Let A, B be finite sets such that
#A = n = #B. Let f : A → B be a map. Then f is injective if and only if it is
surjective.
Proof. This cannot be proved here because it goes too close to the axioms of
mathematics for our scope.
Example 137. There are maps N → N which are injective or surjective but
not both. An injective example is f : N → N defined by f ( x) = x + 1. A
surjective example is g: N → N defined by g( x) = max( x − 1, 0). We say
that the pigeon hole principle is false for infinite sets.
A f
B
g
g◦ f
f g h
Proposition 139. Let sets and maps A −−−→ B −−−→ C −−−→ D be given.
Then (h ◦ g) ◦ f = h ◦ ( g ◦ f ).
This shows that (h ◦ g) ◦ f and h ◦ ( g ◦ f ) give the same output on any input.
So they are equal.
38 Chapter 6 December 24, 2018
Remark 140.
(a) Consider maps f : A → B and g: C → D. Then g ◦ f is only defined if
B = C. In words, the target of f is the source of g. So far for the strict
view. A less strict view says g ◦ f is even defined if B ⊂ C.
(b) Composition is a bit like a binary operation on the ‘set’ of maps. It is not
quite because a composite g ◦ f may not be defined as we saw in (a).
(c) Proposition 139 is known as associativity of composition.
Proof. Exercise.
Example 145. A map may have two or more right inverses as the following
example shows. Let f be the unique map {1, 2} → {3}. Define g1 , g2 : {3} →
{1, 2} by gi (3) = i. Then f ◦ gi = id{3} for all i. So f has two right inverses.
Exercise (6.8) Show by an example that a map can have two or more left
inverses.
6.9 Inverses
Proposition 149.
(a) Let g be a map having a left inverse f and a right inverse h. Then f = h.
(b) Let g be a map having two inverses f , h. Then f = h.
6.10 Relations
Definition 155. Here are a few properties that a relation R on a set A may
or may not have.
◦ Reflexive: xRx for all x ∈ A. (156)
◦ Anti-reflexive: xRx for no x ∈ A. (157)
◦ Symmetric: For all x, y ∈ A we have xRy ⇒ yRx. (158)
◦ Anti-symmetric: For all x, y ∈ A we have (159)
( xRy and yRx) ⇒ x = y.
◦ Transitive: For all x, y, z ∈ A we have ( xRy and yRz) ⇒ xRz. (160)
Exercise (6.12) Which of the above five properties are satisfied by (R, ≤)?
Recall that two sets are said to be disjoint if their intersection is empty.
(1) xRy.
(2) C( x) = C( y).
(3) C( x) ∩ C( y) 6= ∅.
Proof of (1) ⇒ (2). We shall prove C( x) ⊃ C( y). Let z ∈ C( y), that is,
yRz. From xRy and yRz it follows that xRz by transitivity of R. But xRz
says z ∈ C( x). This proves C( x) ⊃ C( y). The reverse inclusion C( x) ⊂ C( y)
follows by ( x, y)-symmetry of the situation and since clearly yRx. This proves
(1) ⇒ (2).
Proof of (2) ⇒ (3). Assume C( x) = C( y). By reflexivity of R we have
xRx and so x ∈ C( x). So x ∈ C( x) = C( x) ∩ C( y). This proves (2) ⇒ (3).
Proof of (3) ⇒ (1). Say w ∈ C( x) ∩ C( y). That is, xRw and yRw. By
symmetry of R we have yRw. From xRw and wRy and transitivity of R it
follows that xRy.
Exercise (6.14) Let Q be a set of sets. Prove that the following are equiva-
lent.
(1) There exist a set A and an equivalence relation R on A such that Q =
A/ R. (In the previous exercise you proved that A, R are unique.)
(2) For all distinct X, Y ∈ Q we have X ∩ Y = ∅. Moreover, ∅ 6∈ Q.
Sometimes people say that Q is a partition of A if these hold.
Example 169.
42 Chapter 7 December 24, 2018
7 Polynomials
Throughout this chapter we fix a field K. If you don’t know what a field is,
simply choose one of the examples of fields that you surely know, like Q, R,
C, Z p with p prime. The reasoning is usually the same for all fields.
7.1 Polynomials
for all x ∈ K. In the above we often define ai := 0 for all i > m. Then we can
write
f ( x ) = Σ ai x i .
i≥0
This is a finite sum in the sense that only finitely many terms are nonzero.
Proof. Write
f ( x) = Σ ai x i ,
i≥0
g( x) = Σ bj xj
j≥0
so h is a polynomial mapping.
f ( x) = Σ ai x i , g( x) = Σ bj xj
i≥0 j≥0
Definition 174.
(a). A polynomial over K is a sequence
a = ( a0 , a1 , . . .)
such that ai ∈ K for all i, and only finitely many ai are nonzero.
(b). The polynomial mapping associated with a is the mapping K → K
taking x to ∑i ≥0 ai xi .
(c). Let a, b be polynomials over K and write
For k ≥ 0 define
k
ck = Σ ai b k − i .
i=0
We define ab := (c0 , c1 , . . .). This is again a polynomial, that is, only finitely
many of the ck are nonzero. Also, we define a + b := (a0 + b0 , a1 + b1 , . . .)
which also a polynomial.
(d). Another notation for the polynomial a = (a0 , a1 , . . .) is
a( x) = Σ ai x i
i≥0
(this is a finite sum). This is the notation we use from now on. Any other
variable than x is also good. The set of polynomials (in the variable x) is
written K [ x]. So multiplication and addition, defined in (c), are binary oper-
ations on K [ x].
44 Chapter 7 December 24, 2018
Exercise (7.2) Let K be a finite field. Prove that there exists a polynomial
f ∈ K [ x] such that f (α ) = 0 for all α ∈ K but f 6= 0.
Proof. Exercise.
Proof. Equivalent is: there is unique r ∈ K [ x] such that g | f − r and deg(r) <
deg( g). We shall prove it in this form.
Uniqueness. Say r1 and r2 both have the above properties. Put s = r1 − r2 .
Then
g | ( f − r2 ) − ( f − r1 ) = r1 − r2 = s.
Also deg(s) < deg( g). By lemma 176 it follows that s = 0. So r1 = r2 . This
proves uniqueness.
Existence. Let A be the set of r ∈ K [ x] such that g | f − r. Note first that
A 6= ∅, because f ∈ A. By the well-ordering principle there exists r ∈ A
with minimal degree. Choose r like that.
We claim that deg(r) < deg( g). Suppose not. Let axd be the leading term
of g and bxd+e of r (a, b ∈ K r {0}). Then r and ba−1 xe g have equal leading
terms. Put
r1 = r − ba−1 xe g.
MA132 Foundations / MA138 Sets and Numbers 45
Then
g | ( f − r) + ba−1 xe g = f − r1 .
So r1 ∈ A but deg(r1 ) < deg(r) contradicting the minimality of deg(r). This
proves existence.
x2 − 2x | x5 + 3x3 + 1 | x3 + 2x2 + 7x + 14
x5 − 2x4
2x4 + 3x3
2x4 − 4x3
7x3
7x3 − 14x2
14x2
14x2 − 28x
28x + 1
Definition 179.
(a) Let f , g ∈ K [ x]. We say f divides g and write f | g if there exists
h ∈ K [ x] such that f h = g.
(b) Let f ∈ K [ x] be nonconstant. We say that f is irreducible if it has no
divisors of degree different from 0 and deg( f ). It is reducible other-
wise.
(c) A polynomial is monic if it is nonzero and its leading coefficient is 1.
Example 180.
(a) Every polynomial of degree 1 is irreducible.
(b) Here are some factorisations of monic polynomials into monic irre-
ducible polynomials:
x3 − 1 = ( x − 1)( x2 + x + 1) if K = Q
3 2
x − 1 = ( x − 1)( x − α )( x − α ) if K = C, α = exp(2π i /3)
x6 − 1 = ( x − 1)( x + 1)·
· ( x2 + x + 1)( x2 − x + 1) if K = Q
x2 − 2 = ( x − 21/2 )( x + 21/2 ) if K = R
x2 − 2 is irreducible if K = Q
f = p1 · · · ps = q1 · · · qt .
Here pi , q j are irreducible monic. Suppose that these are different factorisa-
tions even up to reordering.
Suppose p1 = q1 . Then f / p1 has two distinct factorisations
f / p1 = p2 · · · ps = q2 · · · qt
p1 ( p2 · · · ps − rq2 · · · qt ) = p1 · · · ps − p1 rq2 · · · qt
= q1 · · · qt − p1 rq2 · · · qt = (q1 − p1 r)q2 · · · qt =: g.
We call the latter g. Then deg( g) < deg( f ). By minimality of deg( f ) the
polynomial g has only one factorisation. Also p1 divides g. So p1 must divide
one of the factors in (q1 − p1 r)q2 · · · qt . So p1 divides q1 − p1 r. So p1 | q1 . So
p1 = q1 . Contradiction. This proves uniqueness.
Exercise (7.3) Let K be a finite field of q elements. For n ≥ 1 let a(n) be the
number of monic irreducible elements of K [ x] of degree n. Prove:
Just like for integers there is a Euclidean algorithm to compute hcf( f , g). At
the same time that we look at it, we shall reprove the uniqueness of hcf( f , g)
(corollary 182).
The following has been proved before. Here is our second proof.
Proof. Exercise.
f 3 := f 1 − x f 2 = ( x4 − 1) − x( x3 + x + 2) = − x2 − 2x − 1.
f 4 := f 2 + x f 3 = ( x3 + x + 2) + x(− x2 − 2x − 1) = −2x2 + 2.
−4d = −4( x + 1) = f 5 = 2 f 3 − f 4 = 2 f 3 − ( f 2 + x f 3 )
= − f 2 + (2 − x) f 3 = − f 2 + (2 − x)( f 1 − x f 2 )
= (2 − x) f 1 + ( x2 − 2x − 1) f 2 .
Lemma 191: Remainder lemma. Let f ∈ K [ x]. Let α ∈ K. Then the remain-
der on division of f by x − α is f (α ).
h := ( x − α 1 )d1 · · · ( x − α k )dk .
Exercise (7.5) Prove that a nonzero polynomial cannot have infinitely many
zeroes.
MA132 Foundations / MA138 Sets and Numbers 49
Example 196. The field R of the real numbers is not algebraically closed
because x2 + 1 has no real roots.
Remark 197. Let K be a field. Then there exists an algebraically closed field
L containing K as a subfield. This belongs to Galois Theory (year 3).
Lemma 199. Let f ∈ C[ x] and put A = {| f ( x)| : x ∈ C}. Then A has a least
element, that is, an element p ∈ A such that p ≤ q for all q ∈ A.
Example 201.
√ √
(a) Let f = x2 − 2. Then 2 is a root of f . Therefore 2 is algebraic.
(b) Let n ∈ Z>0 . Let α ∈ C be an nth root of unity. Then α is algebraic: it
is a root of f ( x) = xn − 1.
so f ( x) = x2 + x − 1 works.
√ √
Example 204. We know that √ 2 and√ 3 are algebraic. Assuming theo-
rem 202, it follows that α := 2 + 3 is algebraic. Let’s find nonzero
f ∈ Q[ x] such that f (α√) = 0.
Firstly, α 2 = 2 + 2 6 + 3. So (α 2 − 5)2 = 24. So f = ( x2 − 5)2 − 24 =
x − 10x2 + 1 does it.
4
Proof. We have
n
n n
0 = q f (α ) = q f ( p/q) = Σ ak pk qn − k .
k= 0
Let tk denote the kth term in the right-hand sum: tk = ak pk qn−k . For k > 0
it is clear that p | tk . It follows that p divides the remaining term t0 , that is,
p | a0 qn . But hcf( p, q) = 1 so p | a0 . Likewise q | an .
√ √
Example 206. Let α = 2 + 3. We shall prove that α is irrational. Let
f = x4 − 10x2 + 1. Then f (α ) = 0 as we saw in example 204. Say α = p/q
with p, q ∈ Z coprime and q > 0. By proposition 205 it follows that p, q are
divisors of 1. So α ∈ {−1, 1}. This is a contradiction because α > 1. So α is
irrational.
Proof. For k ≥ 1 and α ∈ R let dk (α ) denote the kth digit after the decimal
dot of the decimal of α .
Let f : Z≥1 → R be any map. For k ≥ 1 let xk ∈ {0, . . . , 8} be any digit
different from dk ( f (k)) (we avoid 9). Let x be the (unique) real number such
that dk ( x) = xk for all k ≥ 1 and 0 ≤ x < 1.
Then x 6= f (k) for all k ≥ 1 because dk ( x) = xk 6= dk ( f (k)) by our choice
of xk . So x is not in the image of f .
Definition 209. Two sets A, B are said to be equipotent or have the same
cardinality if there exists a bijection A → B. We write #A = #B or A ∼ B if
and only if such a bijection exists.
So every finite set is countable. Note also that a set A is infinite if and
only if there exists an injection N → A.
b b
b b b
b b b b
b b b b b
Exercise (8.2) Prove the following results which help construct countable
sets.
(a) Suppose there exists a surjection N → A. Then A is countable.
(b) Suppose A1 , A2 , · · · are countable. Then their union is countable. Hint:
Use that N2 ∼ N by example 212.
(c) Suppose A, B are countable. Then A × B is countable.
We don’t use countable as a synonym for countably infinite (it is not). It’s
just that most sets in our examples are clearly infinite.
This argument by Cantor caused a shitstorm from people saying that this
is not what mathematics is supposed to be like.
h( x) = f ( x) if x ∈ A A
h( x) = g−1 ( x) if x ∈ A B
h( x) = f ( x) if x ∈ A∞ .
54 Chapter 8 December 24, 2018
Then h is bijective.
Definition 222. Let A be a set. Its power set P( A) is the set of subsets of A.
For n ≥ 1 we also define, recursively, Pn+1 ( A) = P( Pn ( A)).