You are on page 1of 293

Analysis I:

Calculus of One Real Variable


Peter Philip∗

Lecture Notes
Originally Created for the Class of Winter Semester 2015/2016 at LMU Munich
Includes Subsequent Corrections and Revisions†

September 17, 2021

Contents
1 Foundations: Mathematical Logic and Set Theory 6
1.1 Introductory Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 Propositional Calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Statements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Logical Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.3 Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3 Set Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4 Predicate Calculus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2 Functions and Relations 26


2.1 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2 Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

3 Natural Numbers, Induction, and the Size of Sets 41


3.1 Induction and Recursion . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

E-Mail: philip@math.lmu.de

Resources used in the preparation of this text include [Kön04, Kun80, Wal04].

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
CONTENTS 2

3.2 Cardinality: The Size of Sets . . . . . . . . . . . . . . . . . . . . . . . . . 48

4 Real Numbers 51
4.1 The Real Numbers as a Complete Totally Ordered Field . . . . . . . . . 51
4.2 Important Subsets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

5 Complex Numbers 56
5.1 Definition and Basic Arithmetic . . . . . . . . . . . . . . . . . . . . . . . 56
5.2 Sign and Absolute Value (Modulus) . . . . . . . . . . . . . . . . . . . . . 58
5.3 Sums and Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.4 Binomial Coefficients and Binomial Theorem . . . . . . . . . . . . . . . . 63

6 Polynomials 66
6.1 Arithmetic of K-Valued Functions . . . . . . . . . . . . . . . . . . . . . . 66
6.2 Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

7 Limits and Convergence of Real and Complex Numbers 71


7.1 Sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
7.2 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
7.2.1 Definitions and First Examples . . . . . . . . . . . . . . . . . . . 81
7.2.2 Continuity, Sequences, and Function Arithmetic . . . . . . . . . . 83
7.2.3 Bounded, Closed, and Compact Sets . . . . . . . . . . . . . . . . 86
7.2.4 Intermediate Value Theorem . . . . . . . . . . . . . . . . . . . . . 90
7.2.5 Inverse Functions, Existence of Roots, Exponential Function, Log-
arithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
7.3 Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.3.1 Definition and Convergence . . . . . . . . . . . . . . . . . . . . . 102
7.3.2 Convergence Criteria . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.3.3 Absolute Convergence and Rearrangements . . . . . . . . . . . . . 109
7.3.4 b-Adic Representations of Real Numbers . . . . . . . . . . . . . . 115

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
CONTENTS 3

8 Convergence of K-Valued Functions 117


8.1 Pointwise and Uniform Convergence . . . . . . . . . . . . . . . . . . . . . 117
8.2 Power Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
8.3 Exponential Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.4 Trigonometric Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
8.5 Polar Form of Complex Numbers, Fundamental Theorem of Algebra . . . 136

9 Differential Calculus 140


9.1 Definition of Differentiability and Rules . . . . . . . . . . . . . . . . . . . 140
9.2 Higher Order Derivatives and the Sets C k . . . . . . . . . . . . . . . . . 147
9.3 Mean Value Theorem, Monotonicity, and Extrema . . . . . . . . . . . . . 148
9.4 L’Hôpital’s Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
9.5 Convex Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

10 The Riemann Integral on Intervals in R 158


10.1 Definition and Simple Properties . . . . . . . . . . . . . . . . . . . . . . 158
10.2 Important Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
10.2.1 Fundamental Theorem of Calculus . . . . . . . . . . . . . . . . . 172
10.2.2 Integration by Parts Formula . . . . . . . . . . . . . . . . . . . . 174
10.2.3 Change of Variables . . . . . . . . . . . . . . . . . . . . . . . . . . 175
10.3 Application: Taylor’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . 176
10.4 Improper Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

A Axiomatic Set Theory 186


A.1 Motivation, Russell’s Antinomy . . . . . . . . . . . . . . . . . . . . . . . 186
A.2 Set-Theoretic Formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
A.3 The Axioms of Zermelo-Fraenkel Set Theory . . . . . . . . . . . . . . . . 189
A.3.1 Existence, Extensionality, Comprehension . . . . . . . . . . . . . 190
A.3.2 Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
A.3.3 Pairing, Union, Replacement . . . . . . . . . . . . . . . . . . . . . 194

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
CONTENTS 4

A.3.4 Infinity, Ordinals, Natural Numbers . . . . . . . . . . . . . . . . . 198


A.3.5 Power Set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
A.3.6 Foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
A.4 The Axiom of Choice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
A.5 Cardinality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
A.5.1 Relations to Injective, Surjective, and Bijective Maps; Schröder-
Bernstein Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 214
A.5.2 Finite Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
A.5.3 Power Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

B Associativity and Commutativity 227


B.1 Associativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
B.2 Commutativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230

C Algebraic Structures 233


C.1 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
C.2 Rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
C.3 Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239

D Construction of the Real Numbers 240


D.1 Natural Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
D.2 Interlude: Orders on Groups . . . . . . . . . . . . . . . . . . . . . . . . . 246
D.3 Integers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
D.4 Rational Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
D.5 Real Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
D.6 Uniqueness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262

E Series: Additional Material 268


E.1 Riemann Rearrangement Theorem . . . . . . . . . . . . . . . . . . . . . . 268
E.2 b-Adic Representations of Real Numbers . . . . . . . . . . . . . . . . . . 270

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
CONTENTS 5

F Cardinality of R and Some Related Sets 275

G Partial Fraction Decomposition 279

H Irrationality of e and π 285


H.1 Irrationality of e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
H.2 Irrationality of π . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286

I Trigonometric Functions 288


I.1 Additional Trigonometric Formulas . . . . . . . . . . . . . . . . . . . . . 288

J Differential Calculus 289


J.1 Continuous, But Nowhere Differentiable Functions . . . . . . . . . . . . . 289

References 292

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 6

1 Foundations: Mathematical Logic and Set Theory

1.1 Introductory Remarks


The task of mathematics is to establish the truth or falsehood of (formalizable) state-
ments using rigorous logic, and to provide methods for the solution of classes of (e.g.
applied) problems, ideally including rigorous logical proofs verifying the validity of the
methods (proofs that the method under consideration will, indeed, provide a correct
solution).
The topic of this class is calculus, which is short for infinitesimal calculus, usually un-
derstood (as it is here) to mean differential and integral calculus of real and complex
numbers (more generally, calculus may refer to any method or system of calculation
guided by the symbolic manipulation of expressions, we will briefly touch on another
example in Sec. 1.2 below). In that sense, calculus is the beginning part of the broader
field of (mathematical) analysis, the section of mathematics concerned with the notion
of a limit (for us, the most important examples will be limits of sequences (Def. 7.1
below) and limits of functions (Def. 8.17 below)).
Before we can properly define our first limit, however, it still needs some preparatory
work. In modern mathematics, the objects under investigation are almost always so-
called sets. So one aims at deriving (i.e. proving) true (and interesting and useful)
statements about sets from other statements about sets known or assumed to be true.
Such a derivation or proof means applying logical rules that guarantee the truth of the
derived (i.e. proved) statement.
However, unfortunately, a proper definition of the notion of set is not easy, and neither
is an appropriate treatment of logic and proof theory. Here, we will only be able to
briefly touch on the bare necessities from logic and set theory needed to proceed to the
core matter of this class. We begin with logic in Sec. 1.2, followed by set theory in
Sec. 1.3, combining both in Sec. 1.4. The interested student can find an introductory
presentation of axiomatic set theory in Appendix A and he/she should consider taking
a separate class on set theory, logic, and proof theory at a later time.

1.2 Propositional Calculus


1.2.1 Statements

Mathematical logic is a large field in its own right and, as indicated above, a thorough in-
troduction is beyond the scope of this class – the interested reader may refer to [EFT07],
[Kun12], and references therein. Here, we will just introduce some basic concepts using

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 7

common English (rather than formal symbolic languages – a concept touched on in Sec.
A.2 of the Appendix and more thoroughly explained in books like [EFT07]).
As mentioned before, mathematics establishes the truth or falsehood of statements. By
a statement or proposition we mean any sentence (any sequence of symbols) that can
reasonably be assigned a truth value, i.e. a value of either true, abbreviated T, or false,
abbreviated F. The following example illustrates the difference between statements and
sentences that are not statements:
Example 1.1. (a) Sentences that are statements:

Every dog is an animal. (T)


Every animal is a dog. (F)
The number 4 is odd. (F)
2 + 3 = 5. (T)

2 < 0. (F)
x + 1 > 0 holds for each natural number x. (T)

(b) Sentences that are not statements:

Let’s study calculus!


Who are you?
3 · 5 + 7.
x + 1 > 0.
All natural numbers are green.

The fourth sentence in Ex. 1.1(b) is not a statement, as it can not be said to be either
true or false without any further knowledge on x. The fifth sentence in Ex. 1.1(b) is
not a statement as it lacks any meaning and can, hence, not be either true or false. It
would become a statement if given a definition of what it means for a natural number
to be green.

1.2.2 Logical Operators

The next step now is to combine statements into new statements using logical operators,
where the truth value of the combined statements depends on the truth values of the
original statements and on the type of logical operator facilitating the combination.
The simplest logical operator is negation, denoted ¬. It is actually a so-called unary
operator, i.e. it does not combine statements, but is merely applied to one statement.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 8

For example, if A stands for the statement “Every dog is an animal.”, then ¬A stands
for the statement “Not every dog is an animal.”; and if B stands for the statement “The
number 4 is odd.”, then ¬B stands for the statement “The number 4 is not odd.”, which
can also be expressed as “The number 4 is even.”
To completely understand the action of a logical operator, one usually writes what is
known as a truth table. For negation, the truth table is
A ¬A
T F (1.1)
F T
that means if the input statement A is true, then the output statement ¬A is false; if
the input statement A is false, then the output statement ¬A is true.
We now proceed to discuss binary logical operators, i.e. logical operators combining
precisely two statements. The following four operators are essential for mathematical
reasoning:
Conjunction: A and B, usually denoted A ∧ B.
Disjunction: A or B, usually denoted A ∨ B.
Implication: A implies B, usually denoted A ⇒ B.
Equivalence: A is equivalent to B, usually denoted A ⇔ B.
Here is the corresponding truth table:
A B A∧B A∨B A⇒B A⇔B
T T T T T T
T F F T F F (1.2)
F T F T T F
F F F F T T

When first seen, some of the assignments of truth values in (1.2) might not be completely
intuitive, due to the fact that logical operators are often used somewhat differently in
common English. Let us consider each of the four logical operators of (1.2) in sequence:
For the use in subsequent examples, let A1 , . . . , A6 denote the six statements from Ex.
1.1(a).
Conjunction: Most likely the easiest of the four, basically identical to common language
use: A ∧ B is true if, and only if, both A and B are true. For example, using Ex. 1.1(a),
A1 ∧ A4 is the statement “Every dog is an animal and 2 + 3 = 5.”, which is true since
both A1 and A4 are true. On the other hand, A1 ∧ A3 is the statement “Every dog is
an animal and the number 4 is odd.”, which is false, since A3 is false.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 9

Disjunction: The disjunction A ∨ B is true if, and only if, at least one of the statements
A, B is true. Here one already has to be a bit careful – A ∨ B defines the inclusive or,
whereas “or” in common English is often understood to mean the exclusive or (which is
false if both input statements are true). For example, using Ex. 1.1(a), A1 ∨ A4 is the
statement “Every dog is an animal or 2 + 3 = 5.”, which is true since both A1 and A4
are true. The statement A1 ∨ A3 , i.e. “Every dog is an animal or the number 4 is odd.”
is also true,
√ since A1 is true. However, the statement A2 ∨ A5 , i.e. “Every animal is a
dog or 2 < 0.” is false, as both A2 and A5 are false.
As you will have noted in the above examples, logical operators can be applied to
combine statements that have no obvious contents relation. While this might seem
strange, introducing contents-related restrictions is unnecessary as well as undesirable,
since it is often not clear which seemingly unrelated statements might suddenly appear
in a common context in the future. The same occurs when considering implications and
equivalences, where it might seem even more obscure at first.
Implication: Instead of A implies B, one also says if A then B, B is a consequence
of A, B is concluded or inferred from A, A is sufficient for B, or B is necessary for
A. The implication A ⇒ B is always true, except if A is true and B is false. At first
glance, it might be surprising that A ⇒ B is defined to be true for A false and B true,
however, there are many examples of incorrect statements implying correct statements.
For instance, squaring the (false) equality of integers −1 = 1, implies the (true) equality
of integers 1 = 1. However, as with conjunction and disjunction, it is perfectly valid
to combine statements without any obvious context relation: For example, using Ex.
1.1(a), the statement A1 ⇒ A6 , i.e. “Every dog is an animal implies x + 1 > 0 holds for
each natural number x.” is true, since A6 is true, whereas the statement A4 ⇒ A2 , i.e.
“2 + 3 = 5 implies every animal is a dog.” is false, as A4 is true and A2 is false.
Of course, the implication A ⇒ B is not really useful in situations, where the truth
values of both A and B are already known. Rather, in a typical application, one tries
to establish the truth of A to prove the truth of B (a strategy that will fail if A happens
to be false).

Example 1.2. Suppose we know Sasha to be a member of a group of students, taking


a class in Analysis. Then the statement A “Sasha has taken a class in Analysis before.”
implies the statement B “There is at least one student in the group, who has taken the
class before”. A priori, we might not know if Sasha has taken the Analysis class before,
but if we can establish that Sasha has, indeed, taken the class before, then we also know
B to be true. If we find Sasha to be taking the class for the first time, then we do not
know, whether B is true or false.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 10

Equivalence: A ⇔ B means A is true if, and only if, B is true. Once again, using
input statements from Ex. 1.1(a), we see that A1 ⇔ A4 , i.e. “Every dog is an animal
is equivalent to 2 + 3 = 5.”, is true as well as A2 ⇔ A3 , i.e. “Every animal is a dog is
equivalent to √the number 4 is odd.”. On the other hand, A4 ⇔ A5 , i.e. “2 + 3 = 5 is
equivalent to 2 < 0, is false.
Analogous to the situation of implications, A ⇔ B is not really useful if the truth values
of both A and B are known a priori, but can be a powerful tool to prove B to be true
or false by establishing the truth value of A. It is obviously more powerful than the
implication as illustrated by the following example (compare with Ex. 1.2):
Example 1.3. Suppose we know Sasha has obtained the highest score among the stu-
dents registered for the Analysis class. Then the statement A “Sasha has taken the
Analysis class before.” is equivalent to the statement B “The student with the highest
score has taken the class before.” As in Ex. 1.2, if we can establish Sasha to have taken
the class before, then we also know B to be true. However, in contrast to Ex. 1.2, if we
find Sasha to have taken the class for the first time, then we know B to be false.
Remark 1.4. In computer science, the truth value T is often coded as 1 and the truth
value F is often coded as 0.

1.2.3 Rules

Note that the expressions in the first row of the truth table (1.2) (e.g. A ∧ B) are not
statements in the sense of Sec. 1.2.1, as they contain the statement variables (also known
as propositional variables) A or B. However, the expressions become statements if all
statement variables are substituted with actual statements. We will call expressions of
this form propositional formulas. Moreover, if a truth value is assigned to each statement
variable of a propositional formula, then this uniquely determines the truth value of the
formula. In other words, the truth value of the propositional formula can be calculated
from the respective truth values of its statement variables – a first justification for the
name propositional calculus.
Example 1.5. (a) Consider the propositional formula (A ∧ B) ∨ (¬B). Suppose A is
true and B is false. The truth value of the formula is obtained according to the
following truth table:

A B A ∧ B ¬B (A ∧ B) ∨ (¬B)
(1.3)
T F F T T

(b) The propositional formula A ∨ (¬A), also known as the law of the excluded middle,
has the remarkable property that its truth value is T for every possible choice of

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 11

truth values for A:


A ¬A A ∨ (¬A)
T F T (1.4)
F T T
Formulas with this property are of particular importance.

Definition 1.6. A propositional formula is called a tautology or universally true if,


and only if, its truth value is T for all possible assignments of truth values to all the
statement variables it contains.

Notation 1.7. We write φ(A1 , . . . , An ) if, and only if, the propositional formula φ
contains precisely the n statement variables A1 , . . . , An .

Definition 1.8. The propositional formulas φ(A1 , . . . , An ) and ψ(A1 , . . . , An ) are called
equivalent if, and only if, φ(A1 , . . . , An ) ⇔ ψ(A1 , . . . , An ) is a tautology.

Lemma 1.9. The propositional formulas φ(A1 , . . . , An ) and ψ(A1 , . . . , An ) are equiva-
lent if, and only if, they have the same truth value for all possible assignments of truth
values to A1 , . . . , An .

Proof. If φ(A1 , . . . , An ) and ψ(A1 , . . . , An ) are equivalent and Ai is assigned the truth
value ti , i = 1, . . . , n, then φ(A1 , . . . , An ) ⇔ ψ(A1 , . . . , An ) being a tautology implies it
has truth value T. From (1.2) we see that either φ(A1 , . . . , An ) and ψ(A1 , . . . , An ) both
have truth value T or they both have truth value F.
If, on the other hand, we know φ(A1 , . . . , An ) and ψ(A1 , . . . , An ) have the same truth
value for all possible assignments of truth values to A1 , . . . , An , then, given such an
assignment, either φ(A1 , . . . , An ) and ψ(A1 , . . . , An ) both have truth value T or both
have truth value F, i.e. φ(A1 , . . . , An ) ⇔ ψ(A1 , . . . , An ) has truth value T in each case,
showing it is a tautology. 

For all logical purposes, two equivalent formulas are exactly the same – it does not
matter if one uses one or the other. The following theorem provides some important
equivalences of propositional formulas. As too many parentheses tend to make formulas
less readable, we first introduce some precedence conventions for logical operators:

Convention 1.10. ¬ takes precedence over ∧, ∨, which take precedence over ⇒, ⇔.


So, for example,
(A ∨ ¬B ⇒ ¬B ∧ ¬A) ⇔ ¬C ∧ (A ∨ ¬D)
is the same as
    
A ∨ (¬B) ⇒ (¬B) ∧ (¬A) ⇔ (¬C) ∧ A ∨ (¬D) .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 12

Theorem 1.11. (a) (A ⇒ B) ⇔ ¬A ∨ B. This means one can actually define impli-
cation via negation and disjunction.

(b) (A ⇔ B) ⇔ (A ⇒ B) ∧ (B ⇒ A) , i.e. A and B are equivalent if, and only if, A
is both necessary and sufficient for B. One also calls the implication B ⇒ A the
converse of the implication A ⇒ B. Thus, A and B are equivalent if, and only if,
both A ⇒ B and its converse hold true.

(c) Commutativity of Conjunction: A ∧ B ⇔ B ∧ A.

(d) Commutativity of Disjunction: A ∨ B ⇔ B ∨ A.

(e) Associativity of Conjunction: (A ∧ B) ∧ C ⇔ A ∧ (B ∧ C).

(f ) Associativity of Disjunction: (A ∨ B) ∨ C ⇔ A ∨ (B ∨ C).

(g) Distributivity I: A ∧ (B ∨ C) ⇔ (A ∧ B) ∨ (A ∧ C).

(h) Distributivity II: A ∨ (B ∧ C) ⇔ (A ∨ B) ∧ (A ∨ C).

(i) De Morgan’s Law I: ¬(A ∧ B) ⇔ ¬A ∨ ¬B.

(j) De Morgan’s Law II: ¬(A ∨ B) ⇔ ¬A ∧ ¬B.

(k) Double Negative: ¬¬A ⇔ A.

(l) Contraposition: (A ⇒ B) ⇔ (¬B ⇒ ¬A).

Proof. Each equivalence is proved by providing a truth table and using Lem. 1.9.
(a):
A B ¬A A ⇒ B ¬A ∨ B
T T F T T
T F F F F
F T T T T
F F T T T

(b) – (h): Exercise.


(i):
A B ¬A ¬B A ∧ B ¬(A ∧ B) ¬A ∨ ¬B
T T F F T F F
T F F T F T T
F T T F F T T
F F T T F T T

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 13

(j): Exercise.
(k):
A ¬A ¬¬A
T F T
F T F

(l):
A B ¬A ¬B A ⇒ B ¬B ⇒ ¬A
T T F F T T
T F F T F F
F T T F T T
F F T T T T
Having checked all the rules completes the proof of the theorem. 

The importance of the rules provided by Th. 1.11 lies in their providing proof techniques,
i.e. methods for establishing the truth of statements from statements known or assumed
to be true. The rules of Th. 1.11 will be used frequently in proofs throughout this class.

Remark 1.12. Another important proof technique is the so-called proof by contradic-
tion, also called indirect proof. It is based on the observation, called the principle of
contradiction, that A ∧ ¬A is always false:

A ¬A A ∧ ¬A
T F F (1.5)
F T F

Thus, one possibility of proving a statement B to be true is to show ¬B ⇒ A ∧ ¬A for


some arbitrary statement A. Since the right-hand side of the implication is false, the
left-hand side must also be false, proving B is true.

Two more rules we will use regularly in subsequent proofs are the so-called transitivity
of implication and the transitivity of equivalence (we will encounter equivalence again
in the context of relations in Sec. 1.3 below). In preparation for the transitivity rules,
we generalize implication to propositional formulas:

Definition 1.13. In generalization of the implication operator defined in (1.2), we say


the propositional formula φ(A1 , . . . , An ) implies the propositional formula ψ(A1 , . . . , An )
(denoted φ(A1 , . . . , An ) ⇒ ψ(A1 , . . . , An )) if, and only if, each assignment of truth values
to the A1 , . . . , An that makes φ(A1 , . . . , An ) true, makes ψ(A1 , . . . , An ) true as well.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 14

Theorem 1.14. (a) Transitivity of Implication: (A ⇒ B) ∧ (B ⇒ C) ⇒ (A ⇒ C).


(b) Transitivity of Equivalence: (A ⇔ B) ∧ (B ⇔ C) ⇒ (A ⇔ C).

Proof. According to Def. 1.13, the rules can be verified by providing truth tables that
show that, for all possible assignments of truth values to the propositional formulas on
the left-hand side of the implications, either the left-hand side is false or both sides are
true. (a):
A B C A ⇒ B B ⇒ C (A ⇒ B) ∧ (B ⇒ C) A ⇒ C
T T T T T T T
T F T F T F T
F T T T T T T
F F T T T T T
T T F T F F F
T F F F T F F
F T F T F F T
F F F T T T T

(b):
A B C A ⇔ B B ⇔ C (A ⇔ B) ∧ (B ⇔ C) A ⇔ C
T T T T T T T
T F T F F F T
F T T F T F F
F F T T F F F
T T F T F F F
T F F F T F F
F T F F F F T
F F F T T T T
Having checked both rules, the proof is complete. 
Definition and Remark 1.15. A proof of the statement B is a finite sequence of
statements A1 , A2 , . . . , An such that A1 is true; for 1 ≤ i < n, Ai implies Ai+1 , and An
implies B. If there exists a proof for B, then Th. 1.14(a) guarantees that B is true.
Remark 1.16. Principle of Duality: In Th. 1.11, there are several pairs of rules that
have an analogous form: (c) and (d), (e) and (f), (g) and (h), (i) and (j). These
analogies are due to the general law called the principle of duality: If φ(A1 , . . . , An ) ⇒
ψ(A1 , . . . , An ) and only the operators ∧, ∨, ¬ occur in φ and ψ, then the reverse im-
plication Φ(A1 , . . . , An ) ⇐ Ψ(A1 , . . . , An ) holds, where one obtains Φ from φ and Ψ
from ψ by replacing each ∧ with ∨ and each ∨ with ∧. In particular, if, instead of an
implication, we start with an equivalence (as in the examples from Th. 1.11), then we
obtain another equivalence.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 15

1.3 Set Theory


In the previous section, we have had a first glance at statements and corresponding truth
values. In the present section, we will move our focus to the objects such statements
are about. Reviewing Example 1.1(a), and recalling that this is a mathematics class
rather than one in zoology, the first two statements of Example 1.1(a) are less relevant
for us than statements 3–6. As in these examples, we will nearly always be interested in
statements involving numbers or collections of numbers or collections of such collections
etc.
In modern mathematics, the term one usually uses instead of “collection” is “set”. In
1895, Georg Cantor defined a set as “any collection into a whole M of definite and
separate objects m of our intuition or our thought”. The objects m are called the ele-
ments of the set M . As explained in Appendix A, without restrictions and refinements,
Cantor’s set theory is not free of contradictions and, thus, not viable to be used in the
foundation of mathematics. Axiomatic set theory provides these necessary restrictions
and refinements and an introductory treatment can also be found in Appendix A. How-
ever, it is possible to follow and understand the rest of this class, without having studied
Appendix A.

Notation 1.17. We write m ∈ M for the statement “m is an element of the set M ”.

Definition 1.18. The sets M and N are equal, denoted M = N , if, and only if, M and
N have precisely the same elements.

Definition 1.18 means we know everything about a set M if, and only if, we know all its
elements.

Definition 1.19. The set with no elements is called the empty set; it is denoted by the
symbol ∅.

Example 1.20. For finite sets, we can simply write down all its elements,
√ for example,
A := {0}, B := {0, 17.5}, C := {5, 1, 5, 3}, D := {3, 5, 1}, E := {2, 2, −2}, where the
symbolism “:=” is to be read as “is defined to be equal to”.
Note C = D, since both sets contain precisely the same elements. In particular, the
order in which the elements are written down plays no role and a set does not change if
an element is written down more than once.
If a set has many elements, instead of writing down all its elements, one might use
abbreviations such as F := {−4, −2, . . . , 20, 22, 24}, where one has to make sure the
meaning of the dots is clear from the context.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 16

Definition 1.21. The set A is called a subset of the set B (denoted A ⊆ B and also
referred to as the inclusion of A in B) if, and only if, every element of A is also an
element of B (one sometimes also calls B a superset of A and writes B ⊇ A). Please
note that A = B is allowed in the above definition of a subset. If A ⊆ B and A 6= B,
then A is called a strict subset of B, denoted A ( B.
If B is a set and P (x) is a statement about an element x of B (i.e., for each x ∈ B,
P (x) is either true or false), then we can define a subset A of B by writing

A := {x ∈ B : P (x)}. (1.6)

This notation is supposed to mean that the set A consists precisely of those elements of
B such that P (x) is true (has the truth value T in the language of Sec. 1.2).

Example 1.22. (a) For each set A, one has A ⊆ A and ∅ ⊆ A.

(b) If A ⊆ B, then A = {x ∈ B : x ∈ A}.

(c) We have {3} ⊆ {6.7, 3, 0}. Letting A := {−10, −8, . . . , 8, 10}, we have {−2, 0, 2} =
{x ∈ A : x3 ∈ A}, ∅ = {x ∈ A : x + 21 ∈ A}.

Remark 1.23. As a consequence of Def. 1.18, the sets A and B are equal if, and only
if, one has both inclusions, namely A ⊆ B and B ⊆ A. Thus, when proving the equality
of sets, one often divides the proof into two parts, first proving one inclusion, then the
other.

Definition 1.24. (a) The intersection of the sets A and B, denoted A ∩ B, consists of
all elements that are in A and in B. The sets A, B are said to be disjoint if, and
only if, A ∩ B = ∅.

(b) The union of the sets A and B, denoted A ∪ B, consists of all elements that are in
A or in B (as in the logical disjunction in (1.2), the or is meant nonexclusively). If
A and B are disjoint, one sometimes writes A ∪˙ B and speaks of the disjoint union
of A and B.

(c) The difference of the sets A and B, denoted A\B (read “A minus B” or “A without
B”), consists of all elements of A that are not elements of B, i.e. A \ B := {x ∈
A: x∈ / B}. If B is a subset of a given set A (sometimes called the universe in
this context), then A \ B is also called the complement of B with respect to A.
In that case, one also writes B c := A \ B (note that this notation suppresses the
dependence on A).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 17

Example 1.25. (a) Examples of Intersections:

{1, 2, 3} ∩ {3, 4, 5} = {3}, (1.7a)



{ 2} ∩ {1, 2, . . . , 10} = ∅, (1.7b)
{−1, 2, −3, 4, 5} ∩ {−10, −9, . . . , −1} ∩ {−1, 7, −3} = {−1, −3}. (1.7c)

(b) Examples of Unions:

{1, 2, 3} ∪ {3, 4, 5} = {1, 2, 3, 4, 5}, (1.8a)


˙
{1, 2, 3}∪{4, 5} = {1, 2, 3, 4, 5}, (1.8b)
{−1, 2, −3, 4, 5} ∪ {−99, −98, . . . , −1} ∪ {−1, 7, −3}
= {−99, −98, . . . , −2, −1, 2, 4, 5, 7}. (1.8c)

(c) Examples of Differences:

{1, 2, 3} \ {3, 4, 5} = {1, 2}, (1.9a)



{1, 2, 3} \ {3, 2, 1, 5} = ∅, (1.9b)
{−10, −9, . . . , 9, 10} \ {0} = {−10, −9, . . . , −1} ∪ {1, 2, . . . , 9, 10}. (1.9c)

With respect to the universe {1, 2, 3, 4, 5}, it is

{1, 2, 3}c = {4, 5}; (1.9d)

with respect to the universe {0, 1, . . . , 20}, it is

{1, 2, 3}c = {0} ∪ {4, 5, . . . , 20}. (1.9e)

As mentioned
 earlier, it will
often be unavoidable
to consider sets of sets. Here are first
examples: ∅, {0}, {0, 1} , {0, 1}, {1, 2} .
Definition 1.26. Given a set A, the set of all subsets of A is called the power set of A,
denoted P(A) (for reasons explained later (cf. Prop. 2.18), the power set is sometimes
also denoted as 2A ).
Example 1.27. Examples of Power Sets:

P(∅) = {∅}, (1.10a)



P({0}) = ∅, {0} , (1.10b)
   
P P({0}) = P ∅, {0} = ∅, {∅}, {{0}}, P({0}) . (1.10c)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 18

So far, we have restricted our set-theoretic examples to finite sets. However, not sur-
prisingly, many sets of interest to us will be infinite (we will have to postpone a mathe-
matically precise definition of finite and infinite to Sec. 3.2). We will now introduce the
most simple infinite set.
Definition 1.28. The set N := {1, 2, 3, . . . } is called the set of natural numbers (for a
more rigorous construction of N, based on the axioms of axiomatic set theory, see Sec.
A.3.4 of the Appendix, where Th. A.46 shows N to be, indeed, infinite). Moreover, we
define N0 := {0} ∪ N.

The following theorem compiles important set-theoretic rules:


Theorem 1.29. Let A, B, C, U be sets.

(a) Commutativity of Intersections: A ∩ B = B ∩ A.


(b) Commutativity of Unions: A ∪ B = B ∪ A.
(c) Associativity of Intersections: (A ∩ B) ∩ C = A ∩ (B ∩ C).
(d) Associativity of Unions: (A ∪ B) ∪ C = A ∪ (B ∪ C).
(e) Distributivity I: A ∩ (B ∪ C) = (A ∩ B) ∪ (A ∩ C).
(f ) Distributivity II: A ∪ (B ∩ C) = (A ∪ B) ∩ (A ∪ C).
(g) De Morgan’s Law I: U \ (A ∩ B) = (U \ A) ∪ (U \ B).
(h) De Morgan’s Law II: U \ (A ∪ B) = (U \ A) ∩ (U \ B).
(i) Double Complement: If A ⊆ U , then U \ (U \ A) = A.

Proof. In each case, the proof results from the corresponding rule of Th. 1.11:
(a):
Th. 1.11(c)
x∈A∩B ⇔x∈A∧x∈B ⇔ x ∈ B ∧ x ∈ A ⇔ x ∈ B ∩ A.

(g): Under the general assumption of x ∈ U , we have the following equivalences:


 Th. 1.11(i)
x ∈ U \ (A ∩ B) ⇔ ¬(x ∈ A ∩ B) ⇔ ¬ x ∈ A ∧ x ∈ B ⇔ ¬(x ∈ A) ∨ ¬(x ∈ B)
⇔ x ∈ U \ A ∨ x ∈ U \ B ⇔ x ∈ (U \ A) ∪ (U \ B).

The proofs of the remaining rules are left as an exercise. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 19

Remark 1.30. The correspondence between Th. 1.11 and Th. 1.29 is no coincidence.
One can actually prove that, starting with an equivalence of propositional formulas
φ(A1 , . . . , An ) ⇔ ψ(A1 , . . . , An ), where both formulas contain only the operators ∧, ∨, ¬,
one obtains a set-theoretic rule (stating an equality of sets) by reinterpreting all state-
ment variables A1 , . . . , An as variables for sets, all subsets of a universe U , and replacing
∧ by ∩, ∨ by ∪, and ¬ by U \ (if there are no multiple negations, then we do not need
the hypothesis that A1 , . . . , An are subsets of U ). The procedure also works in the op-
posite direction – one can start with a set-theoretic formula for an equality of sets and
translate it into two equivalent propositional formulas.

1.4 Predicate Calculus


Now that we have introduced sets in the previous section, we have to return to the
subject of mathematical logic once more. As it turns out, propositional calculus, which
we discussed in Sec. 1.2, does not quite suffice to develop the theory of calculus (nor
most other mathematical theories). The reason is that we need to consider statements
such as

x + 1 > 0 holds for each natural number x. (T) (1.11a)


All real numbers are positive. (F) (1.11b)
There exists a natural number bigger than 10. (T) (1.11c)
There exists a real number x such that x2 = −1. (F) (1.11d)
For all natural numbers n, there exists a natural number bigger than n. (T) (1.11e)

That means we are interested in statements involving universal quantification via the
quantifier “for all” (one also often uses “for each” or “for every” instead), existential
quantification via the quantifier “there exists”, or both. The quantifier of universal
quantification is denoted by ∀ and the quantifier of existential quantification is denoted
by ∃. Using these symbols as well as N and R to denote the sets of natural and real
numbers, respectively, we can restate (1.11) as

∀ x + 1 > 0. (T) (1.12a)


x∈N
∀ x > 0. (F) (1.12b)
x∈R
∃ n > 10. (T) (1.12c)
n∈N
∃ x2 = −1. (F) (1.12d)
x∈R
∀ ∃ m > n. (T) (1.12e)
n∈N m∈N

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 20

Definition 1.31. A universal statement has the form

∀ P (x), (1.13a)
x∈A

whereas an existential statement has the form

∃ P (x). (1.13b)
x∈A

In (1.13), A denotes a set and P (x) is a sentence involving the variable x, a so-called
predicate of x, that becomes a statement (i.e. becomes either true or false) if x is substi-
tuted with any concrete element of the set A (in particular, P (x) is allowed to contain
further quantifiers, but it must not contain any other quantifier involving x – one says
x must be a free variable in P (x), not bound by any quantifier in P (x)).
The universal statement (1.13a) has the truth value T if, and only if, P (x) has the truth
value T for all elements x ∈ A; the existential statement (1.13b) has the truth value T
if, and only if, P (x) has the truth value T for at least one element x ∈ A.
V W
Remark 1.32. Some people prefer to write instead of ∀ and instead of ∃ .
x∈A x∈A x∈A x∈A
Even though this notation has the advantage of emphasizing that the universal statement
can be interpreted as a big logical conjunction and the existential statement can be
interpreted as a big logical disjunction, it is significantly less common. So we will stick
to ∀ and ∃ in this class.

Remark 1.33. According to Def. 1.31, the existential statement (1.13b) is true if, and
only if, P (x) is true for at least one x ∈ A. So if there is precisely one such x, then
(1.13b) is true; and if there are several different x ∈ A such that P (x) is true, then
(1.13b) is still true. Uniqueness statements are often of particular importance, and one
sometimes writes
∃! P (x) (1.14)
x∈A

for the statement “there exists a unique x ∈ A such that P (x) is true”. This notation
can be defined as an abbreviation for
 

∃ P (x) ∧ ∀ P (y) ⇒ x = y . (1.15)
x∈A y∈A

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 21

Example 1.34. Here are some examples of uniqueness statements:

∃! n > 10. (F) (1.16a)


n∈N
∃! 12 > n > 10. (T) (1.16b)
n∈N
∃! 11 > n > 10. (F) (1.16c)
n∈N
∃! x2 = −1. (F) (1.16d)
x∈R
∃! x2 = 1. (F) (1.16e)
x∈R
∃! x2 = 0. (T) (1.16f)
x∈R

Remark 1.35. As for propositional calculus, we also have some important rules for
predicate calculus:

(a) Consider the negation of a universal statement, ¬ ∀ P (x), which is true if, and
x∈A
only if, P (x) does not hold for each x ∈ A, i.e. if, and only if, there exists at least
one x ∈ A such that P (x) is false (such that ¬P (x) is true). We have just proved
the rule
¬ ∀ P (x) ⇔ ∃ ¬P (x). (1.17a)
x∈A x∈A

Similarly, consider the negation of an existential statement. We claim the corre-


sponding rule is
¬ ∃ P (x) ⇔ ∀ ¬P (x). (1.17b)
x∈A x∈A

Indeed, we can prove (1.17b) from (1.17a):


Th. 1.11(k) (1.17a) Th. 1.11(k)
¬ ∃ P (x) ⇔ ¬ ∃ ¬¬P (x) ⇔ ¬¬ ∀ ¬P (x) ⇔ ∀ ¬P (x).
x∈A x∈A x∈A x∈A
(1.18)
One can interpret (1.17) as a generalization of the De Morgan’s laws Th. 1.11(i),(j).
One can actually generalize (1.17) even a bit more: If a statement starts with several
quantifiers, then one negates the statement by replacing each ∀ with ∃ and vice versa
plus negating the predicate after the quantifiers (see the example in (1.21e) below).

(b) If A, B are sets and P (x, y) denotes a predicate of both x and y, then ∀ ∀ P (x, y)
x∈A y∈B
and ∀ ∀ P (x, y) both hold true if, and only if, P (x, y) holds true for each x ∈ A
y∈B x∈A
and each y ∈ B, i.e. the order of two consecutive universal quantifiers does not
matter:
∀ ∀ P (x, y) ⇔ ∀ ∀ P (x, y) (1.19a)
x∈A y∈B y∈B x∈A

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 22

In the same way, we obtain the following rule:


∃ ∃ P (x, y) ⇔ ∃ ∃ P (x, y). (1.19b)
x∈A y∈B y∈B x∈A

If A = B, one also uses abbreviations of the form


∀ P (x, y) for ∀ ∀ P (x, y), (1.20a)
x,y∈A x∈A y∈A

∃ P (x, y) for ∃ ∃ P (x, y). (1.20b)


x,y∈A x∈A y∈A

Generalizing rules (1.19), we can always commute identical quantifiers. Caveat:


Quantifiers that are not identical must not be commuted (see Ex. 1.36(d) below).
Example 1.36. (a) Negation of universal and existential statements:
¬(x+1>0)
z }| {
Negation of (1.12a) : ∃ x + 1 ≤ 0 . (F) (1.21a)
x∈N
¬(x>0)
z }| {
Negation of (1.12b) : ∃ x ≤ 0 . (T) (1.21b)
x∈R
¬(n>10)
z }| {
Negation of (1.12c) : ∀ n ≤ 10 . (F) (1.21c)
n∈N
¬(x2 =−1)
z }| {
Negation of (1.12d) : ∀ x2 6= −1 . (T) (1.21d)
x∈R
¬(m>n)
z }| {
Negation of (1.12e) : ∃ ∀ m ≤ n . (F) (1.21e)
n∈N m∈N

(b) As a more complicated example, consider the negation of the uniqueness statement
(1.14), i.e. of (1.15):
 

¬ ∃! P (x) ⇔ ¬ ∃ P (x) ∧ ∀ P (y) ⇒ x = y
x∈A x∈A y∈A
 
(1.17b), Th. 1.11(a) 
⇔ ∀ ¬ P (x) ∧ ∀ ¬P (y) ∨ x = y
x∈A y∈A
 
Th. 1.11(i) 
⇔ ∀ ¬P (x) ∨ ¬ ∀ ¬P (y) ∨ x = y
x∈A y∈A
 
(1.17a) 
⇔ ∀ ¬P (x) ∨ ∃ ¬ ¬P (y) ∨ x = y
x∈A y∈A
 
Th. 1.11(j),(k) 
⇔ ∀ ¬P (x) ∨ ∃ P (y) ∧ x 6= y
x∈A y∈A
 
Th. 1.11(a) 
⇔ ∀ P (x) ⇒ ∃ P (y) ∧ x 6= y . (1.22)
x∈A y∈A

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 23

So how to decode the expression, we have obtained at the end? It states that if
P (x) holds for some x ∈ A, then there must be at least a second, different, element
y ∈ A such that P (y) is true. This is, indeed, precisely the negation of ∃! P (x).
x∈A

(c) Identical quantifiers commute:

∀ ∀ x2n ≥ 0 ⇔ ∀ ∀ x2n ≥ 0, (1.23a)


x∈R n∈N n∈N x∈R
∀ ∃ ∃ ny > x2 ⇔ ∀ ∃ ∃ ny > x2 . (1.23b)
x∈R y∈R n∈N x∈R n∈N y∈R

(d) The following example shows that different quantifiers do, in general, not commute
(i.e. do not yield equivalent statements when commuted): While the statement

∀ ∃ y>x (1.24a)
x∈R y∈R

is true (for each real number x, there is a bigger real number y, e.g. y := x + 1 will
do the job), the statement
∃ ∀ y>x (1.24b)
y∈R x∈R

is false (for example, since y > y is false). In particular, (1.24a) and (1.24b) are not
equivalent.
(e) Even though (1.14) provides useful notation, it is better not to think of ∃! as a
quantifier. It is really just an abbreviation for (1.15), and it behaves very differently
from ∃ and ∀: The following examples show that, in general, ∃! commutes neither
with ∃, nor with itself:

∃ ∃! m < n 6⇔ ∃! ∃ m<n
n∈N m∈N m∈N n∈N

(the statement on the left is true, as one can choose n = 2, but the statement on
the right is false, as ∃ m < n holds for every m ∈ N). Similarly,
n∈N

∃! ∃! m < n 6⇔ ∃! ∃! m < n
n∈N m∈N m∈N n∈N

(the statement on the left is still true and the statement on the right is still false
(there is no m ∈ N such that ∃! m < n)).
n∈N

Remark 1.37. One can make the following observations regarding the strategy for
proving universal and existential statements:

(a) To prove that ∀ P (x) is true, one must check the truth of P (x) for every element
x∈A
x ∈ A – examples are not enough!

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 24

(b) To prove that ∀ P (x) is false, it suffices to find one x ∈ A such that P (x) is
x∈A
false – such an x is then called a counterexample and one counterexample is always
enough to prove ∀ P (x) is false!
x∈A

(c) To prove that ∃ P (x) is true, it suffices to find one x ∈ A such that P (x) is true
x∈A
– such an x is then called an example and one example is always enough to prove
∃ P (x) is true!
x∈A

The subfield of mathematical logic dealing with quantified statements is called predicate
calculus. In general, one does not restrict the quantified variables to range only over
elements of sets (as we have done above). Again, we refer to [EFT07] for a deeper
treatment of the subject.
As an application of quantified statements, let us generalize the notion of union and
intersection:

Definition 1.38. Let I 6= ∅ be a nonempty set, usually called an index set in the present
context. For each i ∈ I, let Ai denote a set (some or all of the Ai can be identical).

(a) The intersection  


\
Ai := x : ∀ x ∈ Ai (1.25a)
i∈I
i∈I

consists of all elements x that belong to every Ai .

(b) The union  


[
Ai := x : ∃ x ∈ Ai (1.25b)
i∈I
i∈I

consists of all elements x that belong to at least one Ai . The union is called disjoint
if, and only if, for each i, j ∈ I, i 6= j implies Ai ∩ Aj = ∅.

Proposition 1.39. Let I 6= ∅ be an index set, let M denote a set, and, for each i ∈ I,
let Ai denote a set. The following set-theoretic rules hold:
 
T T
(a) Ai ∩M = (Ai ∩ M ).
i∈I i∈I
 
S S
(b) Ai ∪M = (Ai ∪ M ).
i∈I i∈I

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
1 FOUNDATIONS: MATHEMATICAL LOGIC AND SET THEORY 25

 
T T
(c) Ai ∪M = (Ai ∪ M ).
i∈I i∈I
 
S S
(d) Ai ∩M = (Ai ∩ M ).
i∈I i∈I
T S
(e) M \ Ai = (M \ Ai ).
i∈I i∈I
S T
(f ) M \ Ai = (M \ Ai ).
i∈I i∈I

Proof. We prove (c) and (e) and leave the remaining proofs as an exercise.
(c):
!
\ (∗) 
x∈ Ai ∪ M ⇔ x ∈ M ∨ ∀ x ∈ Ai ⇔ ∀ x ∈ Ai ∨ x ∈ M
i∈I i∈I
i∈I
\
⇔ x∈ (Ai ∪ M ).
i∈I

To justify the equivalence at (∗), we make use of Th. 1.11(b) and verify ⇒ and ⇐. For
⇒ note that the truth of x ∈ M implies x ∈ Ai ∨ x ∈ M is true for each i ∈ I. If x ∈ Ai
is true for each i ∈ I, then x ∈ Ai ∨ x ∈ M is still true for each i ∈ I. To verify ⇐, note
that the existence of i ∈ I such that x ∈ M implies the truth of x ∈ M ∨ ∀ x ∈ Ai .
i∈I
If x ∈ M is false for each i ∈ I, then x ∈ Ai must be true for each i ∈ I, showing
x ∈ M ∨ ∀ x ∈ Ai is true also in this case.
i∈I

(e):
\
x∈M\ Ai ⇔ x ∈ M ∧ ¬ ∀ x ∈ Ai ⇔ x ∈ M ∧ ∃ x ∈
/ Ai
i∈I i∈I
i∈I
[
⇔ ∃ x ∈ M \ Ai ⇔ x ∈ (M \ Ai ),
i∈I
i∈I

completing the proof. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 26

Example 1.40. We have the following identities of sets:


\
N = N, (1.26a)
x∈R
\
{1, 2, . . . , n} = {1}, (1.26b)
n∈N
[
N = N, (1.26c)
x∈R
[
{1, 2, . . . , n} = N, (1.26d)
n∈N
[ \ 
N\ {2n} = {1, 3, 5, . . . } = N \ {2n} . (1.26e)
n∈N n∈N

2 Functions and Relations

2.1 Functions
Definition 2.1. Let A, B be sets. Given x ∈ A, y ∈ B, the set
n o
(x, y) := {x}, {x, y} (2.1)

is called the ordered pair (often shortened to just pair) consisting of x and y. The set of
all such pairs is called the Cartesian product A × B, i.e.

A × B := {(x, y) : x ∈ A ∧ y ∈ B}. (2.2)

Example 2.2. Let A be a set.

A × ∅ = ∅ × A = ∅, (2.3a)
{1, 2} × {1, 2, 3} = {(1, 1), (1, 2), (1, 3), (2, 1), (2, 2), (2, 3)} (2.3b)
6 {1, 2, 3} × {1, 2} = {(1, 1), (1, 2), (2, 1), (2, 2), (3, 1), (3, 2)}.
= (2.3c)

Also note that, for x 6= y,


 
(x, y) = {x}, {x, y} 6= {y}, {x, y} = (y, x). (2.4)

Definition 2.3. Given sets A, B, a function or map f is an assignment rule that assigns
to each x ∈ A a unique y ∈ B. One then also writes f (x) for the element y. The set A

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 27

is called the domain of f , denoted dom(f ), and B is called the codomain of f , denoted
codom(f ). The information about a map f can be concisely summarized by the notation

f : A −→ B, x 7→ f (x), (2.5)

where x 7→ f (x) is called the assignment rule for f , f (x) is called the image of x, and
x is called a preimage of f (x) (the image must be unique, but there might be several
preimages). The set

graph(f ) := (x, y) ∈ A × B : y = f (x) (2.6)

is called the graph of f (not to be confused with pictures visualizing the function f ,
which are also called graph of f ). If one wants to be completely precise, then one
identifies the function f with the ordered triple (A, B, graph(f )).
The set of all functions with domain A and codomain B is denoted by F(A, B) or B A ,
i.e.

F(A, B) := B A := (f : A −→ B) : A = dom(f ) ∧ B = codom(f ) . (2.7)

Caveat: Some authors reserve the word map for continuous functions, but we use func-
tion and map synonymously.
Definition 2.4. Let A, B be sets and f : A −→ B a function.

(a) If T is a subset of A, then

f (T ) := {f (x) ∈ B : x ∈ T } (2.8)

is called the image of T under f .

(b) If U is a subset of B, then

f −1 (U ) := {x ∈ A : f (x) ∈ U } (2.9)

is called the preimage or inverse image of U under f .

(c) f is called injective or one-to-one if, and only if, every y ∈ B has at most one
preimage, i.e. if, and only if, the preimage of {y} has at most one element:
 
−1
f injective ⇔ ∀ f {y} = ∅ ∨ ∃! f (x) = y
y∈B x∈A

⇔ ∀ x1 6= x2 ⇒ f (x1 ) 6= f (x2 ) . (2.10)
x1 ,x2 ∈A

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 28

(d) f is called surjective or onto if, and only if, every element of the codomain of f has
a preimage:

f surjective ⇔ ∀ ∃ y = f (x) ⇔ ∀ f −1 {y} 6= ∅. (2.11)


y∈B x∈A y∈B

(e) f is called bijective if, and only if, f is injective and surjective.

Example 2.5. Examples of Functions:

f : {1, 2, 3, 4, 5} −→ {1, 2, 3, 4, 5}, f (x) := −x + 6, (2.12a)


g : N −→ N, g(n) := 2n, (2.12b)
h : N −→ {2, 4, 6, . . . }, h(n) := 2n, (2.12c)
(
n for n even,
h̃ : N −→ {2, 4, 6, . . . }, h̃(n) := (2.12d)
n + 1 for n odd,
G : N −→ R, G(n) := n/(n + 1), (2.12e)

F : P(N) −→ P P(N) , F (A) := P(A). (2.12f)

Instead of f (x) := −x + 6 in (2.12a), one can also write x 7→ −x + 6 and analogously


in the other cases. Also note that, in the strict sense, functions g and h are different,
since their codomains are different (however, using the following Def. 2.4(a), they have
the same image in the sense that g(N) = h(N)). Furthermore,

f ({1, 2}) = {5, 4} = f −1 ({1, 2}), h̃−1 ({2, 4, 6}) = {1, 2, 3, 4, 5, 6}, (2.13)

f is bijective; g is injective, but not surjective; h is bijective; h̃ is surjective, but not


injective. Can you figure out if G and F are injective and/or surjective?

Example 2.6. (a) For each nonempty set A, the map Id : A −→ A, Id(x) := x, is
called the identity on A. If one needs to emphasize that Id operates on A, then one
also writes IdA instead of Id. The identity is clearly bijective.

(b) Let A, B be nonempty sets. A map f : A −→ B is called constant if, and only if,
there exists c ∈ B such that f (x) = c for each x ∈ A. In that case, one also writes
f ≡ c, which can be read as “f is identically equal to c”. If f ≡ c, ∅ 6= T ⊆ A, and
U ⊆ B, then (
A for c ∈ U ,
f (T ) = {c}, f −1 (U ) = (2.14)
∅ for c ∈/ U.
f is injective if, and only if, A = {x}; f is surjective if, and only if, B = {c}.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 29

(c) Given A ⊆ X, the map


ι : A −→ X, ι(x) := x, (2.15)
is called inclusion (also embedding or imbedding). An inclusion is always injective;
it is surjective if, and only if A = X, i.e. if, and only if, it is the identity on A.

(d) Given A ⊆ X and a map f : X −→ B, the map g : A −→ B, g(x) = f (x), is called


the restriction of f to A; f is called the extension of g to X. In this situation, one
also uses the notation f ↾A for g (some authors prefer the notation f |A or f |A).

Theorem 2.7. Let f : A → B be a map, let ∅ 6= I be an index set, and assume S, T, Si ,


i ∈ I, are subsets of A, whereas U, V, Ui , i ∈ I, are subsets of B. Then we have the
following rules concerning functions and set-theoretic operations:

f (S ∩ T ) ⊆ f (S) ∩ f (T ), (2.16a)
!
\ \
f Si ⊆ f (Si ), (2.16b)
i∈I i∈I
f (S ∪ T ) = f (S) ∪ f (T ), (2.16c)
!
[ [
f Si = f (Si ), (2.16d)
i∈I i∈I
f −1 (U ∩ V ) = f (U ) ∩ f −1 (V ),
−1
(2.16e)
!
\ \
f −1 Ui = f −1 (Ui ), (2.16f)
i∈I i∈I
−1
f (U ∪ V ) = f (U ) ∪ f −1 (V ),
−1
(2.16g)
!
[ [
f −1 Ui = f −1 (Ui ), (2.16h)
i∈I i∈I
f (f (U )) ⊆ U, f −1 (f (S)) ⊇ S,
−1
(2.16i)
f −1 (U \ V ) = f −1 (U ) \ f −1 (V ). (2.16j)

Proof. We prove (2.16b) (which includes (2.16a) as a special case) and the second part
of (2.16i), and leave the remaining cases as exercises.
For (2.16b), one argues
!
\  \
y∈f Si ⇔ ∃ ∀ x ∈ Si ∧ y = f (x) ⇒ ∀ y ∈ f (Si ) ⇔ y ∈ f (Si ).
x∈A i∈I i∈I
i∈I i∈I

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 30

The observation

x ∈ S ⇒ f (x) ∈ f (S) ⇔ x ∈ f −1 (f (S)).

establishes the second part of (2.16i). 

It is an exercise to find counterexamples that show one can not, in general, replace the
four subset symbols in (2.16) by equalities (it is possible to find examples with sets that
have at most 2 elements).

Definition 2.8. The composition of maps f and g with f : A −→ B, g : C −→ D, and


f (A) ⊆ C is defined to be the map

g ◦ f : A −→ D, (g ◦ f )(x) := g f (x) . (2.17)

The expression g ◦ f is read as “g after f ” or “g composed with f ”.

Example 2.9. Consider the maps

f : N −→ R, n 7→ n2 , (2.18a)
g : N −→ R, n 7→ 2n. (2.18b)

We obtain f (N) = {1, 4, 9, . . . } ⊆ dom(g), g(N) = {2, 4, 6, . . . } ⊆ dom(f ), and the


compositions

(g ◦ f ) : N −→ R, (g ◦ f )(n) = g(n2 ) = 2n2 , (2.19a)


(f ◦ g) : N −→ R, (f ◦ g)(n) = f (2n) = 4n2 , (2.19b)

showing that composing functions is, in general, not commutative, even if the involved
functions have the same domain and the same codomain.

Proposition 2.10. Consider maps f : A −→ B, g : C −→ D, h : E −→ F , satisfying


f (A) ⊆ C and g(C) ⊆ E.

(a) Associativity of Compositions:

h ◦ (g ◦ f ) = (h ◦ g) ◦ f. (2.20)

(b) One has the following law for forming preimages:

∀ (g ◦ f )−1 (W ) = f −1 (g −1 (W )). (2.21)


W ∈P(D)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 31

Proof. (a): Both h ◦ (g ◦ f ) and


 (h ◦ g) ◦ f map A into F . So it just remains to prove
h ◦ (g ◦ f ) (x) = (h ◦ g) ◦ f (x) for each x ∈ A. One computes, for each x ∈ A,
  
h ◦ (g ◦ f ) (x) = h (g ◦ f )(x) = h g(f (x)) = (h ◦ g)(f (x))

= (h ◦ g) ◦ f (x), (2.22)

establishing the case.


(b): Exercise. 

Definition 2.11. A function g : B −→ A is called a right inverse (resp. left inverse)


of a function f : A −→ B if, and only if, f ◦ g = IdB (resp. g ◦ f = IdA ). Moreover,
g is called an inverse of f if, and only if, it is both a right and a left inverse. If g is
an inverse of f , then one also writes f −1 instead of g. The map f is called (right, left)
invertible if, and only if, there exists a (right, left) inverse for f .

Example 2.12. (a) Consider the map

f : N −→ N, f (n) := 2n. (2.23a)

The maps
(
n/2 if n even,
g1 : N −→ N, g1 (n) := (2.23b)
1 if n odd,
(
n/2 if n even,
g2 : N −→ N, g2 (n) := (2.23c)
2 if n odd,

both constitute left inverses of f . It follows from Th. 2.13(c) below that f does not
have a right inverse.

(b) Consider the map


(
n/2 for n even,
f : N −→ N, f (n) := (2.24a)
(n + 1)/2 for n odd.

The maps

g1 : N −→ N, g1 (n) := 2n, (2.24b)


g2 : N −→ N, g2 (n) := 2n − 1, (2.24c)

both constitute right inverses of f . It follows from Th. 2.13(c) below that f does
not have a left inverse.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 32

(c) The map (


n − 1 for n even,
f : N −→ N, f (n) := (2.25a)
n + 1 for n odd,
is its own inverse, i.e. f −1 = f . For the map


 2 for n = 1,

3 for n = 2,
g : N −→ N, g(n) := (2.25b)


 1 for n = 3,

n for n∈/ {1, 2, 3},

the inverse is


 3 for n = 1,

1 for n = 2,
g −1 : N −→ N, g −1 (n) := (2.25c)


 2 for n = 3,

n for n∈/ {1, 2, 3}.

While Examples 2.12(a),(b) show that left and right inverses are usually not unique,
they are unique provided f is bijective (see Th. 2.13(c)).

Theorem 2.13. Let A, B be nonempty sets.

(a) f : A −→ B is right invertible if, and only if, f is surjective (where the implication
“⇐” makes use of the axiom of choice (AC), see Appendix A.4).

(b) f : A −→ B is left invertible if, and only if, f is injective.

(c) f : A −→ B is invertible if, and only if, f is bijective. In this case, the right inverse
and the left inverse are unique and both identical to the inverse.

Proof. (a): If f is surjective, then, for each y ∈ B, there exists xy ∈ f −1 {y} such that
f (xy ) = y. By AC, we can define the choice function

g : B −→ A, g(y) := xy . (2.26)

Then, for each y ∈ B, f (g(y)) = y, showing g is a right inverse of f . Conversely, if


g : B −→ A is a right inverse of f , then, for each y ∈ B, it is y = f (g(y)), showing that
g(y) ∈ A is a preimage of y, i.e. f is surjective.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 33

(b): Fix a ∈ A. If f is injective, then, for each y ∈ B with f −1 {y} 6= ∅, let xy denote
the unique element in A satisfying f (xy ) = y. Define
(
xy for f −1 {y} 6= ∅,
g : B −→ A, g(y) := (2.27)
a otherwise.

Then, for each x ∈ A, g(f (x)) = x, showing g is a left inverse of f . Conversely, if


g : B −→ A is a left inverse of f and x1 , x2 ∈ A with f (x1 ) = f (x2 ) = y, then
x1 = (g ◦ f )(x1 ) = g(f (x1 )) = g(f (x2 )) = (g ◦ f )(x2 ) = x2 , showing y has precisely one
preimage and f is injective.
(c): Assume g to be a left inverse of f and h to be a right inverse of f . Then, for each
y ∈ B,  
g(y) = g ◦ (f ◦ h) (y) = (g ◦ f ) ◦ h (y) = h(y), (2.28)
showing g = h. In particular, if f has an inverse f −1 , then g = h = f −1 . If f is
invertible, then f is bijective by (a) and (b). If f is bijective, then, by (a) and (b), f has
a left inverse g and a right inverse h (here, this follows without using AC, since, if f is
both injective and surjective, then, for each y ∈ B, the element xy ∈ f −1 {y} is unique,
and (2.26) can be defined without AC). By (2.28), g = h, i.e. f is invertible. 
Theorem 2.14. Consider maps f : A −→ B, g : B −→ C. If f and g are both injective
(resp. both surjective, both bijective), then so is g ◦ f . Moreover, in the bijective case,
one has
(g ◦ f )−1 = f −1 ◦ g −1 . (2.29)

Proof. Exercise. 
Definition 2.15. (a) Given an index set I and a set A, a map f : I −→ A is sometimes
called a family (of elements in A), and is denoted in the form f = (ai )i∈I with
ai := f (i). When using this representation, one often does not even specify f and
A, especially if the ai are themselves sets.

(b) A sequence in a set A is a family of elements in A, where the index set is the set of
natural numbers N. In this case, one writes (an )n∈N or (a1 , a2 , . . . ). More generally,
a family is called a sequence, given a bijective map between the index set I and a
subset of N.

(c) Given a family of sets (Ai )i∈I , we define the Cartesian product of the Ai to be the
set of functions
( ! )
Y [
Ai := f : I −→ Aj : ∀ f (i) ∈ Ai . (2.30)
i∈I
i∈I j∈I

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 34

QI has precisely n elements with n ∈ N, then the elements of the Cartesian product
If
i∈I Ai are called (ordered) n-tuples, (ordered) triples for n = 3.

Example
T 2.16. (a) Using
S the notion of family, we can now say that the intersection
A
i∈I i and union i∈I Ai as defined in Def. 1.38 are the intersection and union of
the family of sets (Ai )i∈I , respectively. As a concrete example, let us revisit (1.26b),
where we have
\
(An )n∈N , An := {1, 2, . . . , n}, An = {1}. (2.31)
n∈N

(b) Examples of Sequences:


Sequence in {0, 1} : (1, 0, 1, 0, 1, 0, . . . ), (2.32a)
Sequence in N : (n2 )n∈N = (1, 4, 9, 16, 25, . . . ), (2.32b)
n
√   √ √ 
Sequence in R : (−1) n n∈N = −1, 2, − 3, . . . , (2.32c)
 
1 1
Sequence in R : (1/n)n∈N = 1, , , . . . , (2.32d)
2 3

Finite Sequence in P(N) : {3, 2, 1}, {2, 1}, {1}, ∅ . (2.32e)
Q
(c) The Cartesian product i∈I A, where all sets Ai =QA, is the same as AI , the set
of all functions from I into A. So, for example, n∈N R = RN is the set of all
sequences in R. If I = {1, 2, . . . , n} with n ∈ N, then
Y n
Y
{1,2...,n}
A=A =: A =: An (2.33)
i∈I i=1

is the set of all n-tuples with entries from A.


In the following, we explain the common notation 2A for the power set P(A) of a set
A. It is related to a natural identification between subsets and their corresponding
characteristic function.
Definition 2.17. Let A be a set and let B ⊆ A be a subset of A. Then
(
1 if x ∈ B,
χB : A −→ {0, 1}, χB (x) := (2.34)
0 if x ∈/ B,

is called the characteristic function of the set B (with respect to the universe A). One
also finds the notations 1B and 1B instead of χB (note that all the notations suppress
the dependence of the characteristic function on the universe A).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 35

Proposition 2.18. Let A be a set. Then the map

χ : P(A) −→ {0, 1}A , χ(B) := χB , (2.35)

is bijective (recall that P(A) denotes the power set of A and {0, 1}A denotes the set of
all functions from A into {0, 1}).

Proof. χ is injective: Let B, C ∈ P(A) with B 6= C. By possibly switching the names


of B and C, we may assume there exists x ∈ B such that x ∈ / C. Then χB (x) = 1,
whereas χC (x) = 0, showing χ(B) 6= χ(C), proving χ is injective.
χ is surjective: Let f : A −→ {0, 1} be an arbitrary function and define B := {x ∈ A :
f (x) = 1}. Then χ(B) = χB = f , proving χ is surjective. 

Proposition 2.18 allows one to identify the sets P(A) and {0, 1}A via the bijective map
χ. This fact together with the common practice of set theory to identify the number
2 with the set {0, 1} (cf. the first paragraph of Sec. D.1 in the Appendix) explains the
notation 2A for P(A).

2.2 Relations
Definition 2.19. Given sets A and B, a relation is a subset R of A × B (if one wants
to be completely precise, a relation is an ordered triple (A, B, R), where R ⊆ A × B).
If A = B, then we call R a relation on A. One says that a ∈ A and b ∈ B are related
according to the relation R if, and only if, (a, b) ∈ R. In this context, one usually writes
a R b instead of (a, b) ∈ R.

Example 2.20. (a) The relations we are probably most familiar with are = and ≤.
The relation R of equality, usually denoted =, makes sense on every nonempty set
A:
R := ∆(A) := {(x, x) ∈ A × A : x ∈ A}. (2.36)
The set ∆(A) is called the diagonal of the Cartesian product, i.e., as a subset of
A × A, the relation of equality is identical to the diagonal:

x = y ⇔ x R y ⇔ (x, y) ∈ R = ∆(A). (2.37)

Similarly, the relation ≤ on R is identical to the set

R≤ := {(x, y) ∈ R2 : x ≤ y}. (2.38)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 36

(b) Every function f : A −→ B is a relation, namely the relation


Rf = {(x, y) ∈ A × B : y = f (x)} = graph(f ). (2.39)
Conversely, if B 6= ∅, then every relation R ⊆ A × B uniquely corresponds to the
function
fR : A −→ P(B), fR (x) = {y ∈ B : x R y}. (2.40)
Definition 2.21. Let R be a relation on the set A.

(a) R is called reflexive if, and only if,


∀ x R x, (2.41)
x∈A

i.e. if, and only if, every element is related to itself.


(b) R is called symmetric if, and only if,

∀ xRy ⇒ yRx , (2.42)
x,y∈A

i.e. if, and only if, each x is related to y if, and only if, y is related to x.
(c) R is called antisymmetric if, and only if,

∀ (x R y ∧ y R x) ⇒ x = y , (2.43)
x,y∈A

i.e. if, and only if, the only possibility for x to be related to y at the same time that
y is related to x is in the case x = y.
(d) R is called transitive if, and only if,

∀ (x R y ∧ y R z) ⇒ x R z , (2.44)
x,y,z∈A

i.e. if, and only if, the relatedness of x and y together with the relatedness of y and
z implies the relatedness of x and z.
Example 2.22. The relations = and ≤ on R (or N) are reflexive, antisymmetric, and
transitive; = is also symmetric, whereas ≤ is not; < is antisymmetric (since x < y∧y < x
is always false) and transitive, but neither reflexive nor symmetric. The relation

R := (x, y) ∈ N2 : (x, y are both even) ∨ (x, y are both odd) (2.45)
on N is not antisymmetric, but reflexive, symmetric, and transitive. The relation
S := {(x, y) ∈ N2 : y = x2 } (2.46)
is not transitive (for example, 2 S 4 and 4 S 16, but not 2 S 16), not reflexive, not sym-
metric; it is only antisymmetric.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 37

Definition 2.23. A relation R on a set A is called an equivalence relation if, and only
if, R is reflexive, symmetric, and transitive. If R is an equivalence relations, then one
often writes x ∼ y instead of x R y.

Example 2.24. (a) The equality relation = is an equivalence relation on each A 6= ∅.

(b) The relation R defined in (2.45) is an equivalence relation on N.



(c) Given a disjoint union A = i∈I Ai with every Ai 6= ∅ (which is sometimes called a
decomposition of A), an equivalence relation on A is defined by

x ∼ y ⇔ ∃ x ∈ Ai ∧ y ∈ Ai . (2.47)
i∈I

Conversely, given an equivalence


Ṡ relation ∼ on a nonempty set A, we can construct
a decomposition A = i∈I Ai such that (2.47) holds: For each x ∈ A, define

[x] := {y ∈ A : x ∼ y}, (2.48)

called the equivalence class of x; each y ∈ [x] is called a representative of [x]. One
verifies that the properties of ∼ guarantee
 
[x] = [y] ⇔ x ∼ y ∧ [x] ∩ [y] = ∅ ⇔ ¬(x ∼ y) . (2.49)

The set of all equivalence classes I := A/ ∼:= {[x] : x ∈ A} is called the quotient set

of A by ∼, and A = i∈I Ai with Ai := i for each i ∈ I is the desired decomposition
of A.

Definition 2.25. A relation R on a set A is called a partial order if, and only if, R is
reflexive, antisymmetric, and transitive. If R is a partial order, then one usually writes
x ≤ y instead of x R y. A partial order ≤ is called a total or linear order if, and only if,
for each x, y ∈ A, one has x ≤ y or y ≤ x.

Notation 2.26. Given a (partial or total) order ≤ on A 6= ∅, we write x < y if, and
only if, x ≤ y and x 6= y, calling < the strict order corresponding to ≤ (note that the
strict order is never a partial order).

Definition 2.27. Let ≤ be a partial order on A 6= ∅, ∅ 6= B ⊆ A.

(a) x ∈ A is called lower (resp. upper) bound for B if, and only if, x ≤ b (resp. b ≤ x)
for each b ∈ B. Moreover, B is called bounded from below (resp. from above) if, and
only if, there exists a lower (resp. upper) bound for B; B is called bounded if, and
only if, it is bounded from above and from below.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 38

(b) x ∈ B is called minimum or just min (resp. maximum or max) of B if, and only if,
x is a lower (resp. upper) bound for B. One writes x = min B if x is minimum and
x = max B if x is maximum.

(c) A maximum of the set of lower bounds of B (i.e. a largest lower bound) is called
infimum of B, denoted inf B; a minimum of the set of upper bounds of B (i.e. a
smallest upper bound) is called supremum of B, denoted sup B.

Example 2.28. (a) For each A ⊆ R, the usual relation ≤ defines a total order on A.
For A = R, we see that N has 0 and 1 as lower bound with 1 = min N = inf N. On
the other hand, N is unbounded from above. The set M := {1, 2, 3} is bounded
with min M = 1, max M = 3. The positive real numbers R+ := {x ∈ R : x > 0}
have inf R+ = 0, but they do not have a minimum (if x > 0, then 0 < x/2 < x).

(b) Consider A := N × N. Then

(m1 , m2 ) ≤ (n1 , n2 ) ⇔ m1 ≤ n1 ∧ m2 ≤ n2 , (2.50)

defines a partial order on A that is not a total order (for example, neither (1, 2) ≤
(2, 1) nor (2, 1) ≤ (1, 2)). For the set

B := (1, 1), (2, 1), (1, 2) , (2.51)

we have inf B = min B = (1, 1), B does not have a max, but sup B = (2, 2) (if
(m, n) ∈ A is an upper bound for B, then (2, 1) ≤ (m, n) implies 2 ≤ m and
(1, 2) ≤ (m, n) implies 2 ≤ n, i.e. (2, 2) ≤ (m, n); since (2, 2) is clearly an upper
bound for B, we have proved sup B = (2, 2)).
A different order on A is the so-called lexicographic order defined by

(m1 , m2 ) ≤ (n1 , n2 ) ⇔ m1 < n1 ∨ (m1 = n1 ∧ m2 ≤ n2 ). (2.52)

In contrast to the order from (2.50), the lexicographic order does define a total
order on A.

Lemma 2.29. Let ≤ be a partial order on A 6= ∅, ∅ 6= B ⊆ A. Then the relation ≥,


defined by
x ≥ y ⇔ y ≤ x, (2.53)
is also a partial order on A. Moreover, using obvious notation, we have, for each x ∈ A,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 39

x ≤-lower bound for B ⇔ x ≥-upper bound for B, (2.54a)


x ≤-upper bound for B ⇔ x ≥-lower bound for B, (2.54b)
x = min≤ B ⇔ x = max≥ B, (2.54c)
x = max≤ B ⇔ x = min≥ B, (2.54d)
x = inf ≤ B ⇔ x = sup≥ B, (2.54e)
x = sup≤ B ⇔ x = inf ≥ B. (2.54f)

Proof. Reflexivity, antisymmetry, and transitivity of ≤ clearly imply the same properties
for ≥, respectively. Moreover

x ≤-lower bound for B ⇔ ∀ x ≤ b ⇔ ∀ b ≥ x ⇔ x ≥-upper bound for B,


b∈B b∈B

proving (2.54a). Analogously, we obtain (2.54b). Next, (2.54c) and (2.54d) are implied
by (2.54a) and (2.54b), respectively. Finally, (2.54e) is proved by

x = inf ≤ B ⇔ x = max≤ {y ∈ A : y ≤-lower bound for B}


⇔ x = min≥ {y ∈ A : y ≥-upper bound for B} ⇔ x = sup≥ B,

and (2.54f) follows analogously. 

Proposition 2.30. Let ≤ be a partial order on A 6= ∅, ∅ 6= B ⊆ A. The elements


max B, min B, sup B, inf B are all unique, provided they exist.

Proof. Exercise. 

Definition 2.31. Let A, B be nonempty sets with partial orders, both denoted by ≤
(even though they might be different). A function f : A −→ B, is called (strictly)
isotone, order-preserving, or increasing if, and only if,

∀ x < y ⇒ f (x) ≤ f (y) (resp. f (x) < f (y)) ; (2.55a)
x,y∈A

f is called (strictly) antitone, order-reversing, or decreasing if, and only if,



∀ x < y ⇒ f (x) ≥ f (y) (resp. f (x) > f (y)) . (2.55b)
x,y∈A

Functions that are (strictly) isotone or antitone are called (strictly) monotone.

Proposition 2.32. Let A, B be nonempty sets with partial orders, both denoted by ≤.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
2 FUNCTIONS AND RELATIONS 40

(a) A (strictly) isotone function f : A −→ B becomes a (strictly) antitone function


and vice versa if precisely one of the relations ≤ is replaced by ≥.
(b) If the order ≤ on A is total and f : A −→ B is strictly isotone or strictly antitone,
then f is one-to-one.
(c) If the order ≤ on A is total and f : A −→ B is invertible and strictly isotone (resp.
antitone), then f −1 is also strictly isotone (resp. antitone).

Proof. (a) is immediate from (2.55).


(b): Due to (a), it suffices to consider the case that f is strictly isotone. If f is strictly
isotone and x 6= y, then x < y or y < x since the order on A is total. Thus, f (x) < f (y)
or f (y) < f (x), i.e. f (x) 6= f (y) in every case, showing f is one-to-one.
(c): Again, due to (a), it suffices to consider the isotone case. If u, v ∈ B such that u < v,
then u = f (f −1 (u)), v = f (f −1 (v)), and the isotonicity of f imply f −1 (u) < f −1 (v) (we
are using that the order on A is total – otherwise, f −1 (u) and f −1 (v) need not be
comparable). 
Example 2.33. (a) f : N −→ N, f (n) := 2n, is strictly increasing, every constant map
on N is both increasing and decreasing, but not strictly increasing or decreasing.
All maps occurring in (2.25) are neither increasing nor decreasing.
(b) The map f : R −→ R, f (x) := −2x, is invertible and strictly decreasing, and so is
f −1 : R −→ R, f −1 (x) := −x/2.
(c) The following counterexamples show that the assertions of Prop. 2.32(b),(c) are no
longer correct if one does not assume the order on A is total. Let A be the set from
(2.51) (where it had been called B) with the (nontotal) order from (2.50). The map

f (1, 1) := 1,

f : A −→ N, f (1, 2) := 2, (2.56)


f (2, 1) := 2,
is strictly isotone, but not one-to-one. The map

f (1, 1) := 1,

f : A −→ {1, 2, 3}, f (1, 2) := 2, (2.57)


f (2, 1) := 3,

is strictly isotone and invertible, however f −1 is not isotone (since 2 < 3, but
f −1 (2) = (1, 2) and f −1 (3) = (2, 1) are not comparable, i.e. f −1 (2) ≤ f −1 (3) is not
true).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 41

3 Natural Numbers, Induction, and the Size of Sets

3.1 Induction and Recursion


One of the most useful proof techniques is the method of induction – it is used in
situations, where one needs to verify the truth of statements φ(n) for each n ∈ N, i.e.
the truth of the statement
∀ φ(n). (3.1)
n∈N

Induction is based on the fact that N satisfies the so-called Peano axioms:

P1: N contains a special element called one, denoted 1.

P2: There exists an injective map S : N −→ N \ {1}, called the successor function (for
each n ∈ N, S(n) is called the successor of n).

P3: If a subset A of N has the property that 1 ∈ A and S(n) ∈ A for each n ∈ A, then
A is equal to N. Written as a formula, the third axiom is:

∀ 1 ∈ A ∧ S(A) ⊆ A ⇒ A = N .
A∈P(N)

Remark 3.1. In Def. 1.28, we had introduced the natural numbers N := {1, 2, 3, . . . }.
The successor function is S(n) = n + 1. In axiomatic set theory, one starts with the
Peano axioms and shows that the axioms of set theory allow the construction of a
set N which satisfies the Peano axioms. One then defines 2 := S(1), 3 := S(2), . . . ,
n + 1 := S(n). The interested reader can find more details in Appendix D.1.

Theorem 3.2 (Principle of Induction). Suppose, for each n ∈ N, φ(n) is a statement


(i.e. a predicate of n in the language of Def. 1.31). If (a) and (b) both hold, where

(a) φ(1) is true,



(b) ∀ φ(n) ⇒ φ(n + 1) ,
n∈N

then (3.1) is true, i.e. φ(n) is true for every n ∈ N.

Proof. Let A := {n ∈ N : φ(n)}. We have to show A = N. Since 1 ∈ A by (a), and


(b)
n ∈ A ⇒ φ(n) ⇒ φ(n + 1) ⇒ S(n) = n + 1 ∈ A, (3.2)

i.e. S(A) ⊆ A, the Peano axiom P3 implies A = N. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 42

Remark 3.3. To prove some φ(n) for each n ∈ N by induction according to Th. 3.2
consists of the following two steps:

(a) Prove φ(1), the so-called base case.

(b) Perform the inductive step, i.e. prove that φ(n) (the induction hypothesis) implies
φ(n + 1).

Example 3.4. We use induction to prove the statement


 
n(n + 1)
∀ 1 + 2 + ··· + n = : (3.3)
n∈N 2
| {z }
φ(n)

1·2
Base Case (n = 1): 1 = 2
, i.e. φ(1) is true.
n(n+1)
Induction Hypothesis: Assume φ(n), i.e. 1 + 2 + · · · + n = 2
holds.
Induction Step: One computes

φ(n) n(n + 1) n(n + 1) + 2n + 2
1 + 2 + · · · + n + (n + 1) = +n+1=
2 2
2
n + 3n + 2 (n + 1)(n + 2)
= = , (3.4)
2 2
i.e. φ(n + 1) holds and the induction is complete.

Corollary 3.5. Theorem 3.2 remains true if (b) is replaced by


  
∀ ∀ φ(m) ⇒ φ(n + 1) . (3.5)
n∈N 1≤m≤n

Proof. If, for each n ∈ N, we use ψ(n) to denote ∀ φ(m), then (3.5) is equivalent to
1≤m≤n

∀ ψ(n) ⇒ ψ(n + 1) , i.e. to Th. 3.2(b) with φ replaced by ψ. Thus, Th. 3.2 implies
n∈N
ψ(n) holds true for each n ∈ N, i.e. φ(n) holds true for each n ∈ N. 

Corollary 3.6. Let I be an index set. Suppose, for each i ∈ I, φ(i) is a statement. If
there is a bijective map f : N −→ I and (a) and (b) both hold, where

(a) φ f (1) is true,
  
(b) ∀ φ f (n) ⇒ φ f (n + 1) ,
n∈N

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 43

then φ(i) is true for every i ∈ I.


Finite Induction: The above assertion remains true if f : {1, . . . , m} −→ I is bijective
for some m ∈ N and N in (b) is replaced by {1, . . . , m − 1}.

Proof. If, for each n ∈ N, we use ψ(n) to denote φ f (n) , then Th. 3.2 shows ψ(n) is
true for every n ∈ N. Given i ∈ I, we have n := f −1 (i) ∈ N with f (n) = i, showing that
φ(i) = φ f (n) = ψ(n) is true.

For the finite induction, let ψ(n) denote n ≤ m ∧ φ f (n) ∨ n > m. Then, for 1 ≤
n < m, we have ψ(n) ⇒ ψ(n + 1) due to (b). For n ≥ m, we also have ψ(n) ⇒ ψ(n + 1)
due to n ≥ m ⇒ n + 1 > m. Thus, Th. 3.2 shows ψ(n) is true for every n ∈ N. Given 
i ∈ I, it is n := f −1 (i) ∈ {1, . . . , m} with f (n) = i. Since n ≤ m ∧ ψ(n) ⇒ φ f (n) , we
obtain that φ(i) is true. 

Apart from providing a widely employable proof technique, the most important ap-
plication of Th. 3.2 is the possibility to define sequences inductively, using so-called
recursion:

Theorem 3.7 (Recursion Theorem). Let A be a nonempty set and x ∈ A. Given a


sequence of functions (fn )n∈N , where fn : An −→ A, there exists a unique sequence
(xn )n∈N in A satisfying the following two conditions:

(i) x1 = x.

(ii) ∀ xn+1 = fn (x1 , . . . , xn ).


n∈N

The same holds if N is replaced by an index set I as in Cor. 3.6.

Proof. To prove uniqueness, let (xn )n∈N and (yn )n∈N be sequences in A, both satisfying
(i) and (ii), i.e.

x1 = y1 = x and (3.6a)

∀ xn+1 = fn (x1 , . . . , xn ) ∧ yn+1 = fn (y1 , . . . , yn ) . (3.6b)
n∈N

We prove by induction (in the form of Cor. 3.5) that (xn )n∈N = (yn )n∈N , i.e.

∀ xn = yn : (3.7)
n∈N| {z }
φ(n)

Base Case (n = 1): φ(1) is true according to (3.6a).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 44

Induction Hypothesis: Assume φ(m) for each m ∈ {1, . . . , n}, i.e. xm = ym holds for
each m ∈ {1, . . . , n}.
Induction Step: One computes

(3.6b) φ(1),...,φ(n) (3.6b)
xn+1 = fn (x1 , . . . , xn ) = fn (y1 , . . . , yn ) = yn+1 , (3.8)

i.e. φ(n + 1) holds and the induction is complete.


To prove existence, we have to show that there is a function F : N −→ A such that the
following two conditions hold:

F (1) = x, (3.9a)

∀ F (n + 1) = fn F (1), . . . , F (n) . (3.9b)
n∈N

To this end, let


 
  
F := B ⊆ N × A : (1, x) ∈ B ∧ ∀ n + 1, fn (a1 , . . . , an ) ∈ B (3.10)
 n∈N, 
(1,a1 ),...,(n,an )∈B

and \
G := B. (3.11)
B∈F

Note that G is well-defined, as N × A ∈ F. Also, clearly, G ∈ F. We would like to


define F such that G = graph(F ). For this to be possible, we will show, by induction,

∀ ∃! (n, xn ) ∈ G . (3.12)
n∈N xn ∈A
| {z }
φ(n)

Base Case (n = 1): From the definition of G, we know (1, x) ∈ G. If (1, a) ∈ G with
a 6= x, then H := G \ {(1, a)} ∈ F, implying G ⊆ H in contradiction to (1, a) ∈
/ H.
This shows a = x and proves φ(1).
Induction Hypothesis: Assume φ(m) for each m ∈ {1, . . . , n}.
Induction Step: From the induction hypothesis, we know

∃! (1, x1 ), . . . , (n, xn ) ∈ G.
(x1 ,...,xn )∈An

Thus, if we let xn+1 := fn (x1 , . . . , xn ), then (n + 1, xn+1 ) ∈ G by the definition of G. If


(n + 1, a) ∈ G with a 6= xn+1 , then H := G \ {(n + 1, a)} ∈ F (using the uniqueness of

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 45

the (1, x1 ), . . . , (n, xn ) ∈ G), implying G ⊆ H in contradiction to (n + 1, a) ∈


/ H. This
shows a = xn+1 , proves φ(n + 1), and completes the induction.
Due to (3.12), we can now define F : N −→ A, F (n) := xn , and the definition of G then
guarantees the validity of (3.9). 

Example 3.8. In many applications of Th. 3.7, one has functions gn : A −→ A and
uses 
∀ fn : An −→ A, fn (a1 , . . . , an ) := gn (an ) . (3.13)
n∈N

Here are some important concrete examples:

(a) The factorial function F : N0 −→ N, n 7→ n!, is defined recursively by

0! := 1, 1! := 1, ∀ (n + 1)! := (n + 1) · n!, (3.14a)


n∈N

i.e. we have A = N and gn (x) := (n + 1) · x. So we obtain

(n!)n∈N0 = (1, 1, 2, 6, 24, 120, . . . ). (3.14b)

(b) For each a ∈ R and each d ∈ R, we define the following arithmetic progression (also
called arithmetic sequence) recursively by

a1 := a, ∀ an+1 := an + d, (3.15a)
n∈N

i.e. we have A = R and gn = g with g(x) := x + d. For example, for a = 2 and


d = −0.5, we obtain

(an )n∈N = (2, 1.5, 1, 0.5, 0, −0.5, −1, −1.5, . . . ). (3.15b)

(c) For each a ∈ R and each q ∈ R \ {0}, we define the following geometric progression
(also called geometric sequence) recursively by

x1 := a, ∀ xn+1 := xn · q, (3.16a)
n∈N

i.e. we have A = R and gn = g with g(x) := x · q. For example, for a = 3 and


q = −2, we obtain
(xn )n∈N = (3, −6, 12, −24, 48, . . . ). (3.16b)

For the time being, we will continue to always specify A and the gn or fn in subsequent
recursive definitions, but in the literature, most of the time, the gn or fn are not provided
explicitly.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 46

Example 3.9. (a) The Fibonacci sequence consists of the Fibonacci numbers, defined
recursively by
F0 := 0, F1 := 1, ∀ Fn+1 := Fn + Fn−1 , (3.17a)
n∈N
i.e. we have A = N0 and
(
1 for n = 1,
fn : An −→ A, fn (a1 , . . . , an ) := (3.17b)
an + an−1 for n ≥ 2.
So we obtain
(Fn )n∈N0 = (0, 1, 1, 2, 3, 5, 8, 13, 21, 34, 55, 89, . . . ). (3.17c)

(b) For A := N, x := 1, and


fn : An −→ A, fn (a1 , . . . , an ) := a1 + · · · + an , (3.18a)
one obtains
x1 = 1, x2 = f1 (1) = 1, x3 = f2 (1, 1) = 2, x4 = f3 (1, 1, 2) = 4,
(3.18b)
x5 = f4 (1, 1, 2, 4) = 8, x6 = f5 (1, 1, 2, 4, 8) = 16, . . .
Definition 3.10. (a) Summation Symbol: On A = R (or, more generally, on every set
A, where an addition + : A × A −→ A is defined), define recursively, for each given
(possibly finite) sequence (a1 , a2 , . . . ) in A:
1
X n+1
X n
X
ai := a1 , ai := an+1 + ai for n ≥ 1, (3.19a)
i=1 i=1 i=1

i.e.
fn : An −→ A, fn (x1 , . . . , xn ) := xn + an+1 . (3.19b)
In (3.19a), one can also use other symbols for i, except a and n; for a finite sequence,
n needs to be less than the maximal index of the finite sequence.
More generally, if I is an index set and φ : {1, . . . , n} −→ I a bijective map, then
define n
X X
ai := aφ(i) . (3.19c)
i∈I i=1

The commutativity of addition implies that the definition in (3.19c) is actually


independent of the chosen bijective map φ (cf. Th. B.5). Also define
X
ai := 0 (3.19d)
i∈∅

(for a general A, 0 is meant to be an element such that a + 0 = 0 + a = a for each


a ∈ A and we can even define this if 0 ∈/ A).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 47

(b) Product Symbol: On A = R (or, more generally, on every set A, where a multiplica-
tion · : A × A −→ A is defined), define recursively, for each given (possibly finite)
sequence (a1 , a2 , . . . ) in A:
1
Y n+1
Y n
Y
ai := a1 , ai := an+1 · ai for n ≥ 1, (3.20a)
i=1 i=1 i=1

i.e.
fn : An −→ A, fn (x1 , . . . , xn ) := an+1 · xn . (3.20b)
In (3.20a), one can also use other symbols for i, except a and n; for a finite sequence,
n needs to be less than the maximal index of the finite sequence.
More generally, if I is an index set and φ : {1, . . . , n} −→ I a bijective map, then
define n
Y Y
ai := aφ(i) . (3.20c)
i∈I i=1

The commutativity of multiplication implies that the definition in (3.20c) is actually


independent of the chosen bijective map φ (cf. Th. B.5). Also define
Y
ai := 1 (3.20d)
i∈∅

(for a general A, 1 is meant to be an element such that a · 1 = 1 · a = a for each


a ∈ A and we can even define this if 1 ∈
/ A).
Example 3.11. (a) Given a, d ∈ R, let (an )n∈N be the arithmetic sequence as defined
in (3.15a). It is an exercise to prove by induction that

∀ an = a + (n − 1)d, (3.21a)
n∈N
n
X n n 
∀ Sn := ai = (a1 + an ) = 2 a + (n − 1) d , (3.21b)
n∈N
i=1
2 2

where the Sn are called arithmetic sums.

(b) Given a ∈ R and q ∈ R \ {0}, let (xn )n∈N be the geometric sequence as defined in
(3.16a). We will prove by induction that

xn = a q n−1 ,
∀ (3.22a)
n∈N
n n n−1
(
X X X na for q = 1,
∀ Sn := xi = (a q i−1 ) = a q i = a (1−qn ) (3.22b)
n∈N
i=1 i=1 i=0 1−q
for q 6= 1,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 48

where the Sn are called geometric sums.


For the induction proof of (3.22a), φ(n) is xn = a q n−1 . The base case, φ(1), is the
statement x1 = a q 0 = a, which is true. For the induction step, we assume φ(n)
and compute 
φ(n)
xn+1 = xn · q = a q n−1 · q = a q n , (3.23)
showing φ(n) ⇒ φ(n + 1) and completing the proof.
For q = 1, the sum Sn is actually arithmetic with d = 0, i.e. Sn = na can be
obtained from (3.21b). For the induction proof of (3.22b) with q 6= 1, φ(n) is
n)
Sn = a(1−q
1−q
. The base case, φ(1), is the statement S1 = a(1−q)1−q
= a, which is true.
For the induction step, we assume φ(n) and compute

φ(n) a(1 − q n ) a(1 − q n ) + aq n (1 − q) a(1 − q n+1 )
Sn+1 = Sn + xn+1 = + aq n = = ,
1−q 1−q 1−q
(3.24)
showing φ(n) ⇒ φ(n + 1) and completing the proof.

3.2 Cardinality: The Size of Sets


Cardinality measures the size of sets. For a finite set A, it is precisely the number of
elements in A. For an infinite set, it classifies the set’s degree or level of infinity (it turns
out that not all infinite sets have the same size).

Definition 3.12. (a) The sets A, B are defined to have the same cardinality or the
same size if, and only if, there exists a bijective map ϕ : A −→ B. One can show
that this defines an equivalence relation on every set of sets (see Th. A.54 of the
Appendix).

(b) The cardinality of a set A is n ∈ N (denoted #A = n) if, and only if, there exists
a bijective map ϕ : A −→ {1, . . . , n}. The cardinality of ∅ is defined as 0, i.e.
#∅ := 0. A set A is called finite if, and only if, there exists n ∈ N0 such that
#A = n; A is called infinite if, and only if, A is not finite, denoted #A = ∞ (in the
strict sense, this is an abuse of notation, since ∞ is not a cardinality – for example
#N = ∞ and #P(N) = ∞, but N and P(N) do not have the same cardinality,
since the power set P(A) is always strictly bigger than A (see Th. A.70 of the
Appendix) – #A = ∞ is merely an abbreviation for the statement “A is infinite”).
The interested student finds additional material regarding the uniqueness of finite
cardinality in Th. A.62 and Cor. A.63, and regarding characterizations of infinite
sets in Th. A.55 of the Appendix.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 49

(c) The set A is called countable if, and only if, A is finite or A has the same cardinality
as N. Otherwise, A is called uncountable.

In the rest of the section, we present a number of important results regarding the natural
numbers and countability.

Theorem 3.13. (a) Every nonempty finite subset of a totally ordered set has a mini-
mum and a maximum.

(b) Every nonempty subset of N has a minimum.

Proof. (a): Let A be a set and let ≤ denote a total order on A. Moreover, let ∅ 6= B ⊆ A.
We show by induction 
∀ #B = n ⇒ B has a min .
n∈N | {z }
φ(n)

Base Case (n = 1): For n = 1, B contains a unique element b, i.e. b = min B, proving
φ(1).
Induction Step: Suppose φ(n) holds and consider B with #B = n + 1. Let b be one
element from B. Then C := B \ {b} has cardinality n and, according to the induction
hypothesis, there exists c ∈ C satisfying c = min C. If c ≤ b, then c ≤ x for each x ∈ B,
proving c = min B. If b ≤ c, then b ≤ x for each x ∈ B, proving b = min B. In each
case, B has a min, proving φ(n + 1) and completing the induction.
(b): Let ∅ 6= A ⊆ N. We have to show A has a min. If A is finite, then A has a min by (a).
If A is infinite, let n be an element from A. Then the finite set B := {k ∈ A : k ≤ n}
must have a min m by (a). Since m ≤ x for each x ∈ B and m ≤ n < x for each
x ∈ A \ B, we have m = min A. 

Proposition 3.14. Every subset A of N is countable.

Proof. Since ∅ is countable, we may assume A 6= ∅. From Th. 3.13(b), we know that
every nonempty subset of N has a min. We recursively define a sequence in A by
(
min An if An := A \ {ai : 1 ≤ i ≤ n} 6= ∅,
a1 := min A, an+1 :=
an if An = ∅.

This sequence is the same as the function f : N −→ A, f (n) = an . An easy induction


shows that, for each n ∈ N, an 6= an+1 implies the restriction f ↾{1,...,n+1} is injective.
Thus, if there exists n ∈ N such that an = an+1 , then f ↾{1,...,k} : {1, . . . , k} −→ A is

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
3 NATURAL NUMBERS, INDUCTION, AND THE SIZE OF SETS 50

bijective, where k := min{n ∈ N : an = an+1 }, showing A is finite, i.e. countable. If


there does not exist n ∈ N with an = an+1 , then f is injective. Another easy induction
shows that, for each n ∈ N, f ({1, . . . , n}) ⊇ {k ∈ A : k ≤ n}, showing f is also
surjective, proving A is countable. 
Proposition 3.15. For each set A 6= ∅, the following three statements are equivalent:

(i) A is countable.

(ii) There exists an injective map f : A −→ N.

(iii) There exists a surjective map g : N −→ A.

Proof. Directly from the definition of countable in Def. 3.12(c), one obtains (i)⇒(ii) and
(i)⇒(iii). To prove (ii)⇒(i), let f : A −→ N be injective. Then f : A −→ f (A) is
bijective, and, since f (A) ⊆ N, f (A) is countable by Prop. 3.14, proving A is countable
as well. To prove (iii)⇒(i), let g : N −→ A be surjective. Then g has a right inverse
f : A −→ N. One can obtain this from Th. 2.13(a), but, here, we can actually construct
f without the axiom of choice: For a ∈ A, let f (a) := min g −1 ({a}) (recall Th. 3.13(b)).
Then, clearly, g ◦ f = IdA . But this means g is a left inverse for f , showing f is injective
according to Th. 2.13(b). Then A is countable by an application of (ii). 
Q
Theorem 3.16. If (A1 , . . . , An ), n ∈ N, is a finite family of countable sets, then ni=1 Ai
is countable.

Proof. We first consider the special case n = 2 with A1 = A2 = N and show the map

ϕ : N × N −→ N, ϕ(m, n) := 2m · 3n ,

is injective: If ϕ(m, n) = ϕ(p, q), then 2m · 3n = 2p · 3q . Moreover m ≤ p or p ≤ m.


If m ≤ p, then 3n = 2p−m · 3q . Since 3n is odd, 2p−m · 3q must also be odd, implying
p − m = 0, i.e. m = p. Moreover, we now have 3n = 3q , implying n = q, showing
(m, n) = (p, q), i.e. ϕ is injective.
We now come back to the general case stated in the theorem. If at least one of the Ai is
empty, then A is empty. So it remains to consider the case, where all Ai are nonempty.
The proof is conducted by induction by showing
n
Y
∀ Ai is countable .
n∈N
i=1
| {z }
φ(n)

Base Case (n = 1): φ(1) is merely the hypothesis that A1 is countable.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
4 REAL NUMBERS 51

Q
Induction Step: Assuming φ(n), Prop. 3.15(ii) provides injective maps f1 : Q ni=1 Ai −→
N and f2 : An+1 −→ N. To prove φ(n+1), we provide an injective map h : n+1 i=1 Ai −→
N: Define
n+1
Y 
h: Ai −→ N, h(a1 , . . . , an , an+1 ) := ϕ f1 (a1 , . . . , an ), f2 (an+1 ) .
i=1

The injectivity of f1 , f2 , and ϕ clearly implies the injectivity of h, thereby proving


φ(n + 1) and completing the induction. 

S sets (i.e. ∅ 6= I is countable


Theorem 3.17. If (Ai )i∈I is a countable family of countable
and each Ai , i ∈ I, is countable), then the union A := i∈I Ai is also countable (this
result makes use of AC, cf. Rem. 3.18 below).

Proof. It suffices to consider the case that all Ai are nonempty. Moreover, according to
Prop. 3.15(iii), it suffices to construct a surjective map ϕ : N −→ A. Also according
to Prop. 3.15(iii), the countability of I and the Ai provides us with surjective maps
f : N −→ I and gi : N −→ Ai (here AC is used to select each gi from the set of all
surjective maps from N onto Ai ). Define

F : N × N −→ A, F (m, n) := gf (m) (n).

Then F is surjective: Given x ∈ A, there exists i ∈ I such that x ∈ Ai . Since f is


surjective, there is m ∈ N satisfying f (m) = i. Moreover, since gi is surjective, there
exists n ∈ N with gi (n) = x. Then F (m, n) = gi (n) = x, verifying that F is surjective.
As N × N is countable by Th. 3.16, there exists a surjective map h : N −→ N × N. Thus,
F ◦ h is the desired surjective map from N onto A. 

Remark 3.18. The axiom of choice is, indeed, essential for the proof of Th. 3.17. It
is shown in [Jec73, Th. 10.6] that it is consistent with the axioms of ZF (i.e. with the
axioms of Sec. A.3 of the Appendix) that, e.g., the uncountable sets P(N) and R (cf.
Th. F.2 of the Appendix) are countable unions of countable sets.

4 Real Numbers

4.1 The Real Numbers as a Complete Totally Ordered Field


The set of real numbers, denoted R, is a set with special properties, namely a so-called
complete totally ordered field, which, after some preliminaries, will be defined in Def. 4.3
below.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
4 REAL NUMBERS 52

Definition 4.1. A total order ≤ on a nonempty set A is called complete if, and only if,
every nonempty subset B of A that is bounded from above has a supremum, i.e.
  
∀ ∃ ∀ b≤x ⇒ ∃ s = sup B . (4.1)
B∈P(A)\{∅} x∈A b∈B s∈A

Lemma 4.2. A total order ≤ on a nonempty set A is complete if, and only if, every
nonempty subset B of A that is bounded from below has an infimum.

Proof. According to Lem. 2.29, it suffices to prove one implication. We show that (4.1)
implies that every nonempty B bounded from below has an infimum: Define

C := {x ∈ A : x is lower bound for B}. (4.2)

Then every b ∈ B is an upper bound for C and (4.1) implies there exists s = sup C ∈ A.
To verify s = inf B, it remains to show s ∈ C, i.e. that s is a lower bound for B.
However, every b ∈ B is an upper bound for C and s = sup C is the min of all upper
bounds for C, i.e. s ≤ b for each b ∈ B, showing s ∈ C. 

Definition 4.3. Let (A, +, ·) be a field and let ≤ be a total order on A. Then A (or,
more precisely, (A, +, ·, ≤)) is called a totally ordered field if, and only if, the order is
compatible with addition and multiplication, i.e. if, and only if,

∀ x≤y ⇒ x+z ≤y+z , (4.3a)
x,y,z∈A

∀ 0 ≤ x ∧ 0 ≤ y ⇒ 0 ≤ xy . (4.3b)
x,y∈A

Finally, A is called a complete totally ordered field if, and only if, A is a totally ordered
field that is complete in the sense of Def. 4.1.

Theorem 4.4. There exists a complete totally ordered field R (it is called the set of
real numbers). Moreover, R is unique up to isomorphism, i.e. if A is a complete totally
ordered field, then there exists an isomorphism φ : A −→ R, i.e. a bijective map φ :
A −→ R, satisfying

∀ φ(x + y) = φ(x) + φ(y), (4.4a)


x,y∈A

∀ φ(xy) = φ(x)φ(y), (4.4b)


x,y∈A

∀ x < y ⇒ φ(x) < φ(y) . (4.4c)
x,y∈A

It also turns out that the isomorphism is unique.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
4 REAL NUMBERS 53

Proof. To really prove the existence of the real numbers by providing a construction
is tedious and not easy. One possible construction is provided in Appendix D (the
existence proof is completed in Th. D.41, the results regarding the isomorphism can be
found in Th. D.45). 
Theorem 4.5. The following statements and rules are valid in the set of real numbers
R (and, more generally, in every totally ordered field):

(a) x ≤ y ⇒ −x ≥ −y.

(b) x ≤ y ∧ z ≥ 0 ⇒ xz ≤ yz holds as well as x ≤ y ∧ z ≤ 0 ⇒ xz ≥ yz.

(c) x 6= 0 ⇒ x2 := x · x > 0. In particular 1 > 0.

(d) x > 0 ⇒ 1/x > 0, whereas x < 0 ⇒ 1/x < 0.

(e) If 0 < x < y, then x/y < 1, y/x > 1, and 1/x > 1/y.

(f ) x < y ∧ u < v ⇒ x + u < y + v.

(g) 0 < x < y ∧ 0 < u < v ⇒ xu < yv.


x+y
(h) x < y ∧ 0 < λ < 1 ⇒ x < λx + (1 − λ)y < y. In particular x < 2
< y.

Proof. (a): Using (4.3a): x ≤ y ⇒ 0 ≤ y − x ⇒ −y ≤ −x.


(b): One argues, for z ≥ 0,
(4.3b)
x ≤ y ⇒ 0 ≤ y − x ⇒ 0 ≤ (y − x)z = yz − xz ⇒ xz ≤ yz,

and, for z ≤ 0,
(4.3b)
x ≤ y ⇒ 0 ≤ y − x ⇒ 0 ≤ (y − x)(−z) = xz − yz ⇒ xz ≥ yz.

(c): From (b), one obtains x2 ≥ 0. From Th. C.10(i), one then gets x2 > 0.
(d): If x > 0, then x−1 < 0 implies the false statement 1 = xx−1 < 0, i.e. x−1 > 0. The
case x < 0 is treated analogously.
(e): Using (d), we obtain from 0 < x < y that x/y = xy −1 < yy −1 = 1 and 1 = xx−1 <
yx−1 = y/x.
(f): x < y ⇒ x + u < y + u and u < v ⇒ y + u < y + v; both combined yield
x + u < y + v.
(g): 0 < x < y ∧ 0 < u < v ⇒ xu < yu ∧ yu < yv ⇒ xu < yv.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
4 REAL NUMBERS 54

(h): Since 0 < λ and 1 − λ > 0, x < y implies


λx < λy ∧ (1 − λ)x < (1 − λ)y.
Using (4.3a), we obtain
x = λx + (1 − λ)x < λx + (1 − λ)y < λy + (1 − λ)y = y,
completing the proof of the theorem. 
Theorem 4.6. Let ∅ 6= A, B ⊆ R, λ ∈ R, and define
A + B := {a + b : a ∈ A ∧ b ∈ B}, (4.5a)
λA := {λa : a ∈ A}. (4.5b)
If A and B are bounded, then
sup(A + B) = sup A + sup B, (4.6a)
inf(A + B) = inf A + inf B, (4.6b)
(
λ · sup A for λ ≥ 0,
sup(λA) = (4.6c)
λ · inf A for λ < 0,
(
λ · inf A for λ ≥ 0,
inf(λA) = (4.6d)
λ · sup A for λ < 0.

Proof. Exercise. 

4.2 Important Subsets


Remark 4.7. We would like to recover the natural numbers N as a subset of R. Indeed,
if we start with 1 as the neutral element of multiplication and define 2 := 1+1, 3 := 2+1,
. . . , then N := {1, 2, . . . } is a subset of R, satisfying the Peano axioms P1, P2, P3 of Sec.
3.1. However, if one does actually construct R according to the axioms of axiomatic
set theory, then one starts by constructing N first, constructing R from N in several
steps (cf. Appendix D). Depending on the construction used, the original set of natural
numbers will typically not be the same set as the natural numbers as a subset of R.
However, both sets will satisfy the Peano axioms and you will have a canonical bijection
between the two sets. Which one you consider the “genuine” set of natural numbers
depends on your personal taste and philosophy and is completely irrelevant. Any two
models of N will always produce equivalent results, since they must both satisfy the
three Peano axioms.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
4 REAL NUMBERS 55

We now introduce a zoo of important subsets of R together with corresponding notation:

N := {1, 2, 3, . . . } (natural numbers), (4.7a)


N0 := N ∪ {0}, (4.7b)
Z− := {−n : n ∈ N} (negative integers), (4.7c)
Z := Z− ∪ N0 (integers), (4.7d)
Q+ := {m/n : m, n ∈ N} (positive rational numbers), (4.7e)
Q+ +
0 := Q ∪ {0} (nonnegative rational numbers), (4.7f)
Q− := {−q : q ∈ Q+ } (negative rational numbers), (4.7g)
Q− −
0 := Q ∪ {0} (nonpositive rational numbers), (4.7h)

Q := Q+0 ∪Q (rational numbers), (4.7i)
R+ := {x ∈ R : x > 0} (positive real numbers), (4.7j)
R+
0 := {x ∈ R : x ≥ 0} (nonnegative real numbers), (4.7k)
R− := {x ∈ R : x < 0} (negative real numbers), (4.7l)
R−
0 := {x ∈ R : x ≤ 0} (nonpositive real numbers). (4.7m)

For a, b ∈ R with a ≤ b, one also defines the following intervals:

[a, b] := {x ∈ R : a ≤ x ≤ b} (bounded closed interval), (4.8a)


]a, b[ := {x ∈ R : a < x < b} (bounded open interval), (4.8b)
]a, b] := {x ∈ R : a < x ≤ b} (bounded half-open interval), (4.8c)
[a, b[ := {x ∈ R : a ≤ x < b} (bounded half-open interval), (4.8d)
] − ∞, b] := {x ∈ R : x ≤ b} (unbounded closed interval), (4.8e)
] − ∞, b[ := {x ∈ R : x < b} (unbounded open interval), (4.8f)
[a, ∞[ := {x ∈ R : a ≤ x} (unbounded closed interval), (4.8g)
]a, ∞[ := {x ∈ R : a < x} (unbounded open interval). (4.8h)

For a = b, one says that the intervals defined by (4.8a) – (4.8d) are degenerate or
trivial, where [a, a] = {a}, ]a, a[=]a, a] = [a, a[= ∅ – it is sometimes convenient to have
included the degenerate cases in the definition. It is sometimes also useful to abandon
the restriction a ≤ b, to let c := min{a, b}, d := max{a, b}, and to define

[a, b] := [c, d], ]a, b[:=]c, d[, ]a, b] := [c, d] \ {a}, [a, b[:= [c, d] \ {b}. (4.8i)

Theorem 4.8 (Archimedean Property). Let ǫ, x be real numbers. If ǫ > 0 and x > 0,
then there exists n ∈ N such that n ǫ > x.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 56

Proof. We conduct the proof by contradiction: Suppose x is an upper bound for the set
A := {n ǫ : n ∈ N}. Since the order ≤ on R is complete, according to (4.1), there exists
s ∈ R such that s = sup A. In particular, s − ǫ is not an upper bound for A, i.e. there
exists n ∈ N satisfying n ǫ > s − ǫ. But then (n + 1) ǫ > s in contradiction to s = sup A.
This shows x is not an upper bound for A, thereby establishing the case. 

5 Complex Numbers

5.1 Definition and Basic Arithmetic


According to Th. 4.5(c), x2 ≥ 0 holds for every real number x ∈ R, i.e. the equation
x2 + 1 = 0 has no solution in R. This deficiency of the real numbers motivates the
effort to try to extend the field of real numbers to a larger field C, the so-called complex
numbers. The two requirements that C is to be a field containing R and that there is to
be some complex number i ∈ C satisfying i2 = −1 already dictates the following laws
of addition and multiplication for complex numbers z = x + iy and w = u + iv with
x, y, u, v ∈ R:

z + w = x + iy + u + iv = x + u + i(y + v), (5.1a)


zw = (x + iy)(u + iv) = xu − yv + i(xv + yu). (5.1b)

Moreover, if x + iy = u + iv, then (x − u)2 = −(v − y)2 , i.e. x − u = 0 = v − y,


implying x = u and y = v. This suggests to try defining complex numbers as pairs of
real numbers. Indeed, this works:

Definition 5.1. We define the set of complex numbers C := R × R, where, keeping in


mind (5.1), addition on C is defined by

+ : C × C −→ C, (x, y), (u, v) 7→ (x, y) + (u, v) := (x + u, y + v), (5.2)

and multiplication on C is defined by



· : C × C −→ C, (x, y), (u, v) →7 (x, y) · (u, v) := (xu − yv, xv + yu). (5.3)

Theorem 5.2. (a) The set of complex numbers C with addition and multiplication as
defined in Def. 5.1 forms a field, where (0, 0) and (1, 0) are the neutral elements
with respect to addition and multiplication, respectively,

−z := (−x, −y) (5.4a)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 57

is the additive inverse to z = (x, y), whereas


 
−1 1 x −y
z := := , (5.4b)
z x2 + y 2 x2 + y 2
is the multiplicative inverse to z = (x, y) 6= (0, 0).
(b) Defining subtraction and division in the usual way, for each z, w ∈ C, by w − z :=
w + (−z), and w/z := wz −1 for z 6= (0, 0), respectively, all the rules stated in Th.
C.10 are valid in C.
(c) The map
ι : R −→ C, ι(x) := (x, 0), (5.5)
is a monomorphism, i.e. it is injective and satisfies
∀ ι(x + y) = ι(x) + ι(y), (5.6a)
x,y∈R

∀ ι(xy) = ι(x) · ι(y). (5.6b)


x,y∈R

It is customary to identify R with ι(R), as it usually does not cause any confusion.
One then just writes x instead of (x, 0).

Proof. All computations required for (a) and (c) are straightforward and are left as an
exercise; (b) is a consequence of (a), since Th. C.10 is valid for every field. 
Notation 5.3. The number i := (0, 1) is called the imaginary unit (note that, indeed,
i2 = i · i = (0, 1) · (0, 1) = (0 · 0 − 1 · 1, 0 · 1 + 1 · 0) = (−1, 0) = −1). Using i, one obtains
the commonly used representation of a complex number z = (x, y) ∈ C:
z = (x, y) = x · (1, 0) + y · (0, 1) = x + iy, (5.7)
where one calls Re z := x the real part of z and Im z := y the imaginary part of z.
Moreover, z is called purely imaginary if, and only if, Re z = 0 (as a consequence of this
convention, one has the (harmless) pathology that 0 is both real and purely imaginary).
Remark 5.4. There does not exist a total order ≤ on C that makes C into a totally
ordered field (i.e. no total order on C can be compatible with addition and multiplication
in the sense of (4.3)): Indeed, if there were such a total order ≤ on C, then all the rules
of Th. 4.5 had to be valid with respect to that total order ≤. In particular, 0 < 12 = 1
and 0 < i2 = −1 had to be valid by Th. 4.5(c), and, then, 0 < 1 + (−1) = 0 had to
be valid by Th. 4.5(f). However, 0 < 0 is false, showing that there is no total order on
C that satisfies (4.3). Caveat: Of course, there do exist total orders on C, just none
compatible with addition and multiplication – for example, the lexicographic order on
R × R (defined as it was in (2.52) for N × N) constitutes a total order on C.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 58

Definition and Remark 5.5. Conjugation: For each complex number z = x + iy, we
define its complex conjugate or just conjugate to be the complex number z̄ := x − iy.
We then have the following rules that hold for each z = x + iy, w = u + iv ∈ C:

(a) z + w = x+u−iy−iv = z̄+ w̄ and zw = xu−yv−(xv+yu)i = (x−iy)(u−iv) = z̄ w̄.

(b) z + z̄ = 2x = 2 Re z and z − z̄ = 2yi = 2i Im z.

(c) z = z̄ ⇔ x + iy = x − iy ⇔ y = 0 ⇔ z ∈ R.

(d) z z̄ = (x + iy)(x − iy) = x2 + y 2 ∈ R+


0.

5.2 Sign and Absolute Value (Modulus)


We face a certain conundrum regarding the handling of square roots. The problem
is that we will need√the notion of a continuous function to prove the existence of a
unique square root x for every nonnegative real number x and, in consequence, we
will have to wait until Section 7.2.5 below to carry out this proof. On the other hand, it
is extremely desirable to present the theory of convergence simultaneously for real and
for complex numbers, which requires the notion of the absolute value or modulus of a
complex number, to be defined in Def. 5.7(b) below as the square root of a nonnegative
real number.
Faced with this difficulty, we will introduce the notion of square root now, assuming the
existence, until we can add the proof in Section 7.2.5. Some students might be worried
that this might lead to a circular argument, where our later proof of the existence of
square roots would somehow make use of our previous assumption of that existence. Of
course, we will be careful not to make such a circular (and, thereby, logically invalid)
argument. The point is that for real numbers the notion of absolute value does in no
way depend on the notion of a square root (see Lem. 5.8 below).
Definition and Remark 5.6. We define a nonnegative real number y ∈ R+ 0 to be the
+ 2
square root of the nonnegative real number x √ ∈ R0 if, and only if, y = x. If y is the
square root of x, then one uses the notation x := y. We will see in Rem. and Def.
7.61 that√every x ∈ R+ +
0 has a unique square root and that the function f : R0 −→ R0 ,
+

f (x) := x, is strictly increasing (in particular, injective).


Definition 5.7. (a) The sign function is defined by

1
 for x > 0,
sgn : R −→ R, sgn(x) := 0 for x = 0, (5.8)


−1 for x < 0.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 59

It is emphasized that the sign function is only defined for real numbers (cf. Rem.
5.4)!

(b) The absolute value or modulus function is defined by


√ p
abs : C −→ R+ 0, z = x + iy 7→ |z| := z z̄ = x2 + y 2 , (5.9)

where the term absolute value is often preferred for real numbers z ∈ R and the
term modulus is often preferred if one also considers complex numbers z ∈
/ R.

Lemma 5.8. For each x ∈ R, one has


(
x for x ≥ 0,
|x| = x · sgn(x) = (5.10)
−x for x < 0.

Proof. One has (


√ x for x ≥ 0,
|x| = x2 = (5.11)
−x for x < 0,
as claimed. 

Theorem 5.9. The following rules hold for each z, w ∈ C:

(a) z 6= 0 ⇒ |z| > 0.

(b) ||z|| = |z|.

(c) |z| = |z̄|.

(d) max{| Re z|, | Im z|} ≤ |z| ≤ | Re z| + | Im z|.

(e) |zw| = |z||w|.


|z|
(f ) For w 6= 0, one has | wz | = |w|
.

(g) Triangle Inequality:


|z + w| ≤ |z| + |w|. (5.12)

(h) Inverse Triangle Inequality:



|z| − |w| ≤ |z − w|. (5.13)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 60

Proof. We carry out the proofs for z, w ∈ C. However, for z, w ∈ R, everything can
easily be shown directly from (5.10), without making use of square roots.
Let z = x + iy with x, y ∈ R.
(a): If z 6= 0, then x 6= 0 or y 6= p 0, i.e. x2 > 0 or y 2 > 0 by Th. 4.5(c), implying
2 2
x + y > 0 by Th. 4.5(f), i.e. |z| = x2 + y 2 > 0.

(b): Since a := |z| ∈ R+0 , we have |a| = a2 = a = |z|.
p p
(c): Since z̄ = x − iy, we have |z̄| = x2 + (−y)2 = x2 + y 2 = |z|.
(d): It is x = Re z, y = Im z. Let a := max{|x|, |y|}. As remarked in Def. and Rem.
5.6, the square root function is increasing and, thus, taking square roots in the chain of
inequalities a2 ≤ x2 + y 2 ≤ (|x| + |y|)2 implies a ≤ |z| ≤ |x| + |y| as claimed.
(e): As remarked in Def. and Rem. 5.6, the square root function is injective, and, thus,
(e) follows from
Def. and Rem. 5.5(a)
|zw|2 = zw zw = zwz̄ w̄ = z z̄ ww̄ = |z|2 |w|2 .

(f): Let w = u + iv with u, v ∈ R. We first consider the special case z = 1. Applying


the formula (5.4b) for the inverse to w, one obtains

u2 v2 1 2
|w−1 |2 = 2 2 2
+ 2 2 2
= 2 2
= |w|−1 ,
(u + v ) (u + v ) u +v
|z|
i.e. |w−1 | = |w|−1 . Now (f) follows from (e): | wz | = |zw−1 | = |z||w−1 | = |z||w|−1 = |w|
.
(g) follows from

|z + w|2 = (z + w)(z̄ + w̄) = z z̄ + wz̄ + z w̄ + ww̄


Def. and Rem. 5.5(b)
= |z|2 + 2 Re(z w̄) + |w|2
(d) 2
≤ |z|2 + 2|z w̄| + |w|2 = |z| + |w| ,

once again using that the square root function is increasing.


(h): Using (g), we obtain

|z| = |z − w + w| ≤ |z − w| + |w| ⇒ |z| − |w| ≤ |z − w|,


|w| = |w − z + z| ≤ |z − w| + |z| ⇒ −(|z| − |w|) ≤ |z − w|,

implying |z| − |w| ≤ |z − w| by (5.10) (notice |z| − |w| ∈ R). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 61

Remark 5.10. Each complex number (x, y) = x + iy can be visualized as a point in


the so-called complex plane, where the horizontal x-axis represents real numbers and
the vertical y-axis represents purely imaginary numbers. Then the addition of complex
numbers is precisely the vector addition of 2-dimensional vectors in the complex plane,
and conjugation is represented by reflection through the x-axis. Moreover, the modulus
|z| of a complex number is precisely its distance from the origin (0, 0), and |z − w|
is the distance between the points z = (x, y) and w = (u, v) in the plane. Complex
multiplication can also be interpreted geometrically in the plane: If φ denotes the angle
that the vector representing z = (x, y) forms with the x-axis, and, likewise, ψ denotes
the angle that the vector representing w = (u, v) forms with the x-axis, then zw is
the vector of length |zw| that forms the angle φ + ψ with the x-axis (we will better
understand this geometrical interpretation of complex multiplication later (see Def. and
Rem. 8.29), when writing complex numbers in the polar form z = x + iy = |z| exp(iφ),
making use of the exponential function exp).

5.3 Sums and Products


Here we compile some important rules involving sums and products. We are mostly
interested in applying them to real and complex numbers. However, most of the rules,
without extra difficulty, can be proved to hold in more general structures. We will pro-
vide the more general statements, but the reader will not lose much by merely thinking
of C rather than a general ring in Th. 5.11, and of R rather than a general totally
ordered field in Th. 5.12. Some rules involve exponentiation as defined in (C.10a) of the
Appendix (in particular, it is used that z 0 = 1 for each z ∈ R, where R is a ring with
unity), and some proofs make use of the corresponding exponentiation rules of Th. C.6.

Theorem 5.11. Let (R, +, ·) be a ring (cf. Def. C.7 of the Appendix).

(a) For each n ∈ N and each λ, µ, zj , wj ∈ R, j ∈ {1, . . . , n}:


n
X n
X n
X
(λ zj + µ wj ) = λ zj + µ wj .
j=1 j=1 j=1

(b) If R is a ring with unity, then, for each n ∈ N0 and each z ∈ R:


n
X
2 n
(1 − z)(1 + z + z + · · · + z ) = (1 − z) z j = 1 − z n+1 .
j=0

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 62

(c) If R is a commutative ring with unity, then, for each n ∈ N0 and each z, w ∈ R:
n
X
wn+1 − z n+1 = (w − z) z j wn−j = (w − z)(wn + zwn−1 + · · · + z n−1 w + z n ).
j=0

Proof. In each case, the proof can be conducted by an easy induction. We carry out
(c) and leave the other cases as exercises. For (c), the base case (n = 0) is provided by
the true statement w0+1 − z 0+1 = w − z = (w − z)z 0 w0−0 . For the induction step, one
computes
n+1 n
!
X X
(w − z) z j wn+1−j = (w − z) z n+1 w0 + z j wn+1−j
j=0 j=0
n
X
n+1
= (w − z)z + (w − z) w z j wn−j
j=0

ind. hyp.
= (w − z) z n+1 + w(wn+1 − z n+1 ) = wn+2 − z n+2 ,
completing the induction. 
Theorem 5.12. Let (F, +, ·) be a totally ordered field (cf. Def. 4.3).

(a) For each n ∈ N and each xj , yj ∈ F , j ∈ {1, . . . , n}:


  X n n
X
∀ xj ≤ yj ⇒ xj ≤ yj ,
j∈{1,...,n}
j=1 j=1

where equality can only hold if xj = yj for each j ∈ {1, . . . , n}.


(b) For each n ∈ N and each xj , yj ∈ F , j ∈ {1, . . . , n}:
  Yn n
Y
∀ 0 < xj ≤ yj ⇒ xj ≤ yj ,
j∈{1,...,n}
j=1 j=1

where equality can only hold if xj = yj for each j ∈ {1, . . . , n}.

Proof. Both cases are proved by simple inductions, where Th. 4.5(f) is used in (a) and
Th. 4.5(g) is used in (b). 
Theorem 5.13. Triangle Inequality: For each n ∈ N and each zj ∈ C, j ∈ {1, . . . , n}:

Xn X n

z j ≤ |zj |.

j=1 j=1

Proof. Another very simple induction that is left to the reader. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 63

5.4 Binomial Coefficients and Binomial Theorem


The goal in this section is to expand (z +w)n into a sum. This sum involves the so-called
binomial coefficients nk , which are also useful in other contexts. To obtain an idea for
what to expect, let us compute the cases n = 0, 1, 2, 3: (z + w)0 = 1, (z + w)1 = z + w,
(z + w)2 = z 2 + 2zw + w2 , (z + w)3 = z 3 + 3z 2 w + 3zw2 + w3 . One finds that the
coefficients form what is known as Pascal’s triangle, which we write for n = 0, . . . , 5:
n=0: 1
n=1: 1 1
n=2: 1 2 1
(5.14)
n=3: 1 3 3 1
n=4: 1 4 6 4 1
n=5: 1 5 10 10 5 1
 
The entries of the nth row of Pascal’s triangle are denoted by n0 , . . . , nn . One also
observes that one obtains each entry of the (n + 1)st row, except the first and last entry,
by adding the corresponding entries in row n to the left and to the right of the considered
entry in row n + 1. The first and last entry of each row are always set to 1. This can
be summarized as
          
n n n+1 n n
∀ = = 1, = + for k ∈ {1, . . . , n} . (5.15)
n∈N0 0 n k k−1 k
The following Def. 5.14 provides a different and more general definition of binomial
coefficients. We will then prove in Prop. 5.15 that the binomial coefficients as defined
in Def. 5.14 do, indeed, satisfy (5.15).
Definition 5.14. For each α ∈ C and each k ∈ N0 , we define the binomial coefficient
    k
α α Y α+1−j α(α − 1) · · · (α − k + 1)
:= 1, := = for k ∈ N. (5.16)
0 k j=1
j 1 · 2···k

Proposition 5.15. (a) For each α ∈ C and each k ∈ N:


       
α α+1 α α
= 1, = + . (5.17)
0 k k−1 k

(b) For each n ∈ N0 :  


n
= 1. (5.18)
n

The above statements include (5.15) as a special case.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 64

Proof. (a): The first identity is part of the definition in (5.16). For the second identity,
we first observe, for each k ∈ N,
  Y k k−1  
α α+1−j α+1−k Yα+1−j α α+1−k
= = = , (5.19)
k j=1
j k j=1
j k−1 k

which implies
        
α α α α+1−k α α+1
+ = 1+ =
k−1 k k−1 k k−1 k
k−1 k  
α+1Yα+1−j Yα+2−j α+1
= = = . (5.20)
k j=1 j j=1
j k


(b): 00 = 1 according to (5.16). For n ∈ N, (5.18) is proved by induction. The base
 1+1−1
1
case (n = 1) is provided by the true statement 1 = 1 = 1. For the induction step,
one computes
  n+1 n  
n+1 Y n+1+1−j n+1Yn+1−j n ind. hyp.
= = = = 1, (5.21)
n+1 j=1
j n + 1 j=1 j n

which completes the induction. 

Theorem 5.16 (Binomial Theorem). Let R be a commutative ring with unity (cf. Def.
C.7 of the Appendix – for us, R = C is the most important example). For each z, w ∈ R
and each n ∈ N0 , the following formula holds:
n      
n
X n n−k k n n−1
n n
(z + w) = z w =z + z w + ··· + zwn−1 + wn . (5.22)
k=0
k 1 n−1

Proof. The proof is conducted via induction


 on n. The base case (n = 0) is provided by
0 0 0−0 0
the correct statement (z + w) = 1 = 0 z w . For the induction step, we first observe

(z + w)n+1 = (z + w) (z + w)n = z (z + w)n + w (z + w)n . (5.23)

Using the induction hypothesis, we now further manipulate the two terms on the right-

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
5 COMPLEX NUMBERS 65

hand side of (5.23):


n   n  
n ind. hyp.
X n n−k k
X n
z (z + w) = z z w = z n+1−k wk
k=0
k k=0
k
n+1
X 
(n+1
n
)=0 n n+1−k k
= z w , (5.24)
k=0
k
n   n  
ind. hyp.
X n n−k k X n n−k k+1
w (z + w)n = w z w = z w
k=0
k k=0
k
n+1  
X n
= z n+1−k wk . (5.25)
k=1
k−1

Plugging (5.24) and (5.25) into (5.23) yields


  n+1    
n+1 n n+1 0 X n n
(z + w) = z w + + z n+1−k wk
0 k=1
k k − 1
  n+1  
Prop. 5.15 n + 1 n+1 0 X n + 1 n+1−k k
= z w + z w
0 k=1
k
n+1  
X n + 1 n+1−k k
= z w , (5.26)
k=0
k

completing the induction. 

The binomial theorem can now be used to infer a few more rules that hold for the
binomial coefficients:

Corollary 5.17. One has the following identities:


n        
X n n n n
∀ = + + ··· + = 2n , (5.27a)
n∈N0
k=0
k 0 1 n
n          
X n k n n n n n
∀ (−1) = − + − + · · · + (−1) = 0. (5.27b)
n∈N
k=0
k 0 1 2 n

Proof. (5.27a) is just (5.22) with z = w = 1; (5.27b) is just (5.22) with z = 1 and
w = −1. 

The formulas provided by the following proposition are also sometimes useful.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
6 POLYNOMIALS 66

Proposition 5.18. (a) For each α ∈ C and each k ∈ N0 :


k          
X α+j α α+1 α+k α+k+1
= + + ··· + = . (5.28)
j=0
j 0 1 k k

(b) For each n, k ∈ N0 with k ≤ n:


 
n n!
= . (5.29)
k k!(n − k)!

Moreover, for n ≥ 1, one has nk = #Pk ({1, . . . , n}), where

Pk (A) := B ∈ P(A) : #B = k (5.30)

denotes the set of all subsets of a set A that have precisely k elements.

(c) For each n, k ∈ N0 :


k          
X n+j n n+1 n+k n+k+1
= + + ··· + = . (5.31)
j=0
n n n n n+1

Proof. The induction proofs of (a) and (b) are left as exercises. For (c), one computes
k   k k  
X n+j (5.29) X (n + j)! (5.29) X n + j
= =
j=0
n j=0
n!(n + j − n)! j=0
j
   
(5.28) n + k + 1 (5.29) (n + k + 1)! n+k+1
= = = ,
k k!(n + 1)! n+1

thereby establishing the case. 

6 Polynomials

6.1 Arithmetic of K-Valued Functions


Notation 6.1. We will write K in situations, where we allow K to be R or C.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
6 POLYNOMIALS 67

Notation 6.2. If A is any nonempty set, then one can add and multiply arbitrary
functions f, g : A −→ K, and one can define several further operations to create new
functions from f and g:

(f + g) : A −→ K, (f + g)(x) := f (x) + g(x), (6.1a)


(λf ) : A −→ K, (λf )(x) := λf (x) for each λ ∈ K, (6.1b)
(f g) : A −→ K, (f g)(x) := f (x)g(x), (6.1c)
(f /g) : A −→ K, (f /g)(x) := f (x)/g(x) (assuming g(x) 6= 0), (6.1d)
Re f : A −→ R, (Re f )(x) := Re(f (x)), (6.1e)
Im f : A −→ R, (Im f )(x) := Im(f (x)). (6.1f)

For K = R, we further define



max(f, g) : A −→ R, max(f, g)(x) := max f (x), g(x) , (6.1g)

min(f, g) : A −→ R, min(f, g)(x) := min f (x), g(x) , (6.1h)
f + : A −→ R, f + := max(f, 0), (6.1i)
f − : A −→ R, f − := max(−f, 0). (6.1j)

Finally, once again also allowing K = C,

|f | : A −→ R, |f |(x) := |f (x)|. (6.1k)

One calls f + and f − the positive part and the negative part of f , respectively. For
R-valued functions f , we have
|f | = f + + f − . (6.1l)

6.2 Polynomials
Definition 6.3. Let n ∈ N0 . Each function from K into K, x 7→ xn , is called a
monomial. A function P from K into K is called a polynomial if, and only if, it is a
linear combination of monomials, i.e. if, and only if P has the form
n
X
P : K −→ K, P (x) = aj x j = a0 + a1 x + · · · + an x n , aj ∈ K. (6.2)
j=0

The aj are called the coefficients of P . The largest number d ≤ n such that ad 6= 0 is
called the degree of P , denoted deg(P ). If all coefficients are 0, then P is called the zero
polynomial; the degree of the zero polynomial is defined as −1 (in Th. 6.6(b) below, we

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
6 POLYNOMIALS 68

will see that each polynomial of degree n ∈ N0 is uniquely determined by its coefficients
a0 , . . . , an and vice versa).
Polynomials of degree ≤ 0 are constant. Polynomials of degree ≤ 1 have the form
P (x) = a + bx and are called affine functions (often they are also called linear functions,
even though this is not really correct for a 6= 0, since every function P that is linear (in
the sense of linear algebra) must satisfy P (0) = 0). Polynomials of degree ≤ 2 have the
form P (x) = a + bx + cx2 and are called quadratic functions.
Each ξ ∈ K such that P (ξ) = 0 is called a zero or a root of P .
A rational function is a quotient P/Q of two polynomials P and Q.
Remark 6.4. Let λ ∈ K and let P, Q be polynomials. Then λP , P +Q, and P Q defined
according to Not. 6.2 are polynomials as well. More precisely, if λ = 0 or P ≡ 0, then
λP = 0; if P ≡ 0, then P + Q = Q; if Q ≡ 0, then P + Q = P ; if P ≡ 0 or Q ≡ 0, then
P Q = 0. If λ 6= 0 and
n
X m
X
P (x) = aj x j , Q(x) = bj xj ,
j=0 j=0 (6.3)
with deg(P ) = n ≥ 0, deg(Q) = m ≥ 0, n ≥ m ≥ 0,

then, defining bj := 0 for each j ∈ {m + 1, . . . , n} in case n > m,


n
X
(λP )(x) = (λ aj ) xj , deg(λP ) = n, (6.4a)
j=0
Xn
(P + Q)(x) = (aj + bj ) xj , deg(P + Q) ≤ n = max{m, n}, (6.4b)
j=0
m+n
X
(P Q)(x) = c j xj , deg(P Q) = m + n, (6.4c)
j=0

where, setting ak := 0 for each k ∈ {n + 1, . . . , m + n} and bk := 0 for each k ∈


{m + 1, . . . , m + n},
j
X
∀ cj = ak bj−k . (6.4d)
j∈{0,...,n+m}
k=0

Formula (6.4c) can be proved by induction on m = deg(Q) ∈ N0 as follows: For m = 0,


we compute
n
X n+0
X
j
(P Q)(x) = b0 aj x = b 0 aj x j ,
j=0 j=0

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
6 POLYNOMIALS 69

P
i.e. cj = b0 aj = jk=0 ak bj−k , which establishes the base case, remembering bj−k = 0 for
j > k. For the induction step, we compute, for deg(Q) = m + 1,
n m+1 n m
!
X X X X
(P Q)(x) = aj x j bα xα = aj xj bm+1 xm+1 + bα xα
j=0 α=0 j=0 α=0
n m+n j
!
ind. hyp.
X X X
= aj bm+1 xm+1+j + ak bj−k xj
j=0 j=0 k=0
m+n+1 m+n j
!
X X X
= aj−m−1 bm+1 xj + ak bj−k xj
j=m+1 j=0 k=0
m+n+1 j
!
X X
= ak bj−k xj ,
j=0 k=0

which completes the induction step. There is a notational issue in the second and third
line in of the above computation, since, in both lines, the bm+1 in the first sum is the
actual bm+1 from Q, but bm+1 = 0 in the second sum in both lines, which is due to the
induction hypothesis being applied for m < m+1. This is actually used when combining
j ≤ m + n: aj−m−1 bm+1 xj + aj−m−1 ·
both sums in the last step, computing, for m + 1 ≤ P
0 · x = aj−m−1 bm+1 x . For j = m + n + 1, one has m+n+1
j j
k=0 ak bm+n+1−k = an bm+1 , since
bm+n+1−k = 0 for n > k and ak = 0 for k > n.
Finally, deg(P Q) = m + n follows from cm+n = am bn 6= 0.

Theorem 6.5. (a) For each polynomial P given in the form of (6.3) and each ξ ∈ K,
we have the identity
Xn
P (x) = bj (x − ξ)j , (6.5)
j=0

where
n  
X k k−j
∀ bj = ak ξ , in particular b0 = P (ξ), b n = an . (6.6)
j∈{0,...,n}
k=j
j

(b) If P is a polynomial with n := deg(P ) ≥ 1, then, for each ξ ∈ K, there exists a


polynomial Q with deg(Q) = n − 1 such that

P (x) = P (ξ) + (x − ξ) Q(x). (6.7)

In particular, if ξ is a zero of P , then P (x) = (x − ξ) Q(x).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
6 POLYNOMIALS 70

Proof. (a): For ξ = 0, there is nothing to prove. For ξ 6= 0, defining the auxiliary
variable η := x − ξ, we obtain x = ξ + η and
n   k
n X n n  
X k k−j j X X
k (5.22)
X k k−j j
P (x) = ak (ξ + η) = ak ξ η = ak ξ η
k=0 k=0 j=0
j k=0 j=0
j
n X n   n n  
X k k−j j X X k k−j j
= ak ξ η = ak ξ η , (6.8)
j=0 k=0
j j=0 k=j
j

which is (6.5).
(b): According to (a), we have
n
X n−1
X
j−1
P (x) = P (ξ) + (x − ξ) Q(x), with Q(x) = bj (x − ξ) = bj+1 (x − ξ)j , (6.9)
j=1 j=0

proving (b). 

Theorem 6.6. (a) If P is a polynomial with n := deg(P ) ≥ 0, then P has at most n


zeros.

(b) Let P, Q be polynomials as in (6.3) with n = m, deg(P ) ≤ n, and deg(Q) ≤ n. If


P (xj ) = Q(xj ) at n + 1 distinct points x0 , x1 , . . . , xn ∈ K, then aj = bj for each
j ∈ {0, . . . , n}.
Consequence 1: If P, Q with degree ≤ n agree at n + 1 distinct points, then P = Q.
Consequence 2: If we know P = Q, then they agree everywhere, in particular at
max{deg(P ), deg(Q)} + 1 distinct points, which implies they have the same coeffi-
cients.

Proof. (a): For n = 0, P is constant, but not the zero polynomial, i.e. P ≡ a0 6= 0 with
no zeros as claimed. For n ∈ N, the proof is conducted by induction. The base case
(n = 1) is provided by the observation that deg(P ) = 1 implies P is the affine function
with P (x) = a0 + a1 x, a1 6= 0, i.e. P has precisely one zero at ξ = −a0 /a1 . For the
induction step, assume deg(P ) = n + 1. If P has no zeros, then the assertion of (a)
holds true. Otherwise, P has at least one zero ξ ∈ K, and, according to Th. 6.5(b),
there exists a polynomial Q such that deg(Q) = n and

P (x) = (x − ξ) Q(x). (6.10)

From the induction hypothesis, we gather that Q has at most n zeros, i.e. (6.10) implies
P has at most n + 1 zeros, which completes the induction.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 71

(b): If P (xj ) = Q(xj ) at n + 1 distinct points xj , then each of these points is a zero of
P − Q. Thus P − Q is a polynomial of degree ≤ n with at least n + 1 zeros. Then (a)
implies deg(P − Q) = −1, i.e. P − Q is the zero polynomial, i.e. aj − bj = 0 for each
j ∈ {0, . . . , n}. 
Remark 6.7. Let P be a polynomial with n := deg(P ) ≥ 0. According to Th. 6.6(a), P
has at most n zeros. Using Th. 6.5(b) for an induction shows there exists k ∈ {0, . . . , n}
and a polynomial Q of degree n − k such that
k
Y
P (x) = Q(x) (x − ξj ) = (x − ξ1 )(x − ξ2 ) · · · (x − ξk )Q(x), (6.11a)
j=1

where Q does not have any zeros in K and {ξ1 , . . . , ξk } = {ξ ∈ K : P (ξ) = 0} is the set
of zeros of P . It can of course happen that P does not have any zeros and P = Q (no
ξj exist). It can also occur that some of the ξj in (6.11a) are identical. Thus, we can
rewrite (6.11a) as
l
Y
P (x) = Q(x) (x − λj )mj = (x − λ1 )m1 (x − λ2 )m2 · · · (x − λl )ml Q(x), (6.11b)
j=1
Pl
where λ1 , . . . , λl , l ∈ {0, . . . , k}, are the distinct zeros of P , and mj ∈ N with j=1 mj =
k. Then mj is called the multiplicity of the zero λj of P .

7 Limits and Convergence in the Real and Complex


Numbers

7.1 Sequences
Recall from Def. 2.15(b) that a sequence in K is a function f : N −→ K, in this context
usually denoted as f = (zn )n∈N or (z1 , z2 , . . . ) with zn := f (n). Sometimes the sequence
also has the form (zn )n∈I , where I 6= ∅ is a countable index set (e.g. I = N0 ) different
from N (in the context of convergence (see the following Def. 7.1), I must be N or it
must have the same cardinality as N, i.e. finite I are not permissible).
Definition 7.1. The sequence (zn )n∈N in K is said to be convergent with limit z ∈ K if,
and only if, for each ǫ > 0, there exists an index N ∈ N such that |zn − z| < ǫ for every
index n > N . The notation for (zn )n∈N converging to z is limn→∞ zn = z or zn → z for
n → ∞. Thus, by definition,
lim zn = z ⇔ ∀ ∃ ∀ |zn − z| < ǫ. (7.1)
n→∞ ǫ∈R+ N ∈N n>N

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 72

The sequence (zn )n∈N in K is called divergent if, and only if, it is not convergent.
Example 7.2. (a) For every constant sequence (zn )n∈N = (a)n∈N with a ∈ K, one has
limn→∞ zn = limn→∞ a = a: Since, for each n ∈ N, |zn − a| = |a − a| = 0, one can
choose N = 1 for each ǫ > 0.
1
(b) limn→∞ n+a = 0 for each a ∈ C: Here zn := 1/(n + a) (if n = −a, then set zn := w
with w ∈ C arbitrary). Given ǫ > 0, choose an arbitrary N ∈ N with N ≥ ǫ−1 + |a|.
Then, for each n ≥ N , we compute |n + a| = |n − (−a)| ≥ |n − |a|| = n − |a| >
N − |a| ≥ ǫ−1 , and, thus, |zn | = |n + a|−1 < ǫ as desired.
(c) ((−1)n )n∈N is not convergent: We have zn = 1 for each even n and zn = −1 for each
odd n. Thus, for each z 6= 1 and each even n, |zn − z| = |1 − z| > |1 − z|/2 =: ǫ > 0,
i.e. z is not a limit of (zn )n∈N . However, z = 1 is also not a limit of the sequence,
since, for each odd n, |zn − 1| = | − 1 − 1| = 2 > 1 =: ǫ > 0, proving that the
sequence has no limit.
Theorem 7.3. (a) Let (zn )n∈N be a sequence in C. Then (zn )n∈N is convergent in C
if, and only if, both (Re zn )n∈N and (Im zn )n∈N are convergent in R. Moreover, in
that case,
lim zn = z ⇔ lim Re zn = Re z ∧ lim Im zn = Im z. (7.2)
n→∞ n→∞ n→∞

(b) Let (xn )n∈N be a sequence in R and z ∈ C. Then


lim xn = z ⇒ z ∈ R. (7.3)
n→∞

Proof. (a): Suppose (zn )n∈N converges to z ∈ C. Then, given ǫ > 0, there exists N ∈ N
such that, for each n > N , |zn − z| < ǫ. In consequence, for each n > N ,
Th. 5.9(d)
| Re zn − Re z| = | Re(zn − z)| ≤ |zn − z| < ǫ, (7.4)
proving limn→∞ Re zn = Re z. The proof of limn→∞ Im zn = Im z is completely anal-
ogous. Conversely, suppose there are x, y ∈ R such that limn→∞ Re zn = x and
limn→∞ Im zn = y. Here we encounter, for the first time, what is sometimes called an ǫ/2
argument: Given ǫ > 0, there exists N ∈ N such that, for each n > N , | Re zn − x| < ǫ/2
and | Im zn − y| < ǫ/2, implying, for each n > N ,
|zn − (x + iy)| = | Re zn + i Im zn − (x + iy)|
≤ | Re zn − x| + |i|| Im zn − y| < ǫ/2 + ǫ/2 = ǫ, (7.5)
proving limn→∞ zn = x + iy.
(b) is a direct consequence of (a). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 73

Example 7.4. (a) According to Th. 7.3(a), we have


 
√ i Ex. 7.2(a),(b) √ √
lim 2+ = 2 + 0i = 2.
n→∞ n − 17

(b) According to Th. 7.3(a) and Ex. 7.2(c), the sequence ( n1 + (−1)n i)n∈N is divergent.

Another important example relies on the following inequality:


Proposition 7.5 (Bernoulli’s Inequality). For each n ∈ N0 and each x ∈ [−1, ∞[, we
have
(1 + x)n ≥ 1 + nx, (7.6)
with strict inequality whenever n > 1 and x 6= 0.

Proof. For n = 0, (7.6) reads 1 ≥ 1, for n = 1, (7.6) reads 1 + x ≥ 1 + x, for n = 2,


(7.6) reads (1 + x)2 = 1 + 2x + x2 ≥ 1 + 2x, all three statements being trivially true, in
the case n = 2 with strict inequality for x 6= 0. We now proceed by induction for n ≥ 2.
For the induction step, one estimates
ind. hyp., x ≥ −1
(1 + x)n+1 = (1 + x)n (1 + x) ≥ (1 + nx) (1 + x) = 1 + (n + 1)x + nx2
≥ 1 + (n + 1)x, (7.7)
with strict inequality for x 6= 0. 
Example 7.6. We have, for each q ∈ C,
|q| < 1 ⇒ lim q n = 0 : (7.8)
n→∞

For q = 0, there is nothing to prove. For 0 < |q| < 1, it is |q|−1 > 1, i.e. h := |q|−1 −1 > 0.
Thus, for each ǫ > 0 and N ≥ 1/(ǫh), we obtain
(7.6)
n>N ⇒ |q|−n = (1 + h)n ≥ 1 + nh > nh > 1/ǫ ⇒ |q n | = |q|n < ǫ. (7.9)
Definition 7.7. (a) Given z ∈ K and ǫ ∈ R+ , we call the set Bǫ (z) := {w ∈ K :
|w − z| < ǫ} the ǫ-neighborhood of z or, in anticipation of Analysis II, the (open) ǫ-
ball with center z (in fact, for K = C, Bǫ (z) represents an open disk in the complex
plane with center z and radius ǫ, whereas, for K = R, Bǫ (z) =]z − ǫ, z + ǫ[ is the
open interval with center z and length 2ǫ). More generally, a set U ⊆ K is called
a neighborhood of z if, and only if, there exists ǫ > 0 with Bǫ (z) ⊆ U (so, for
example, for ǫ > 0, Bǫ (z) is always a neighborhood of z, whereas R and [z − ǫ, ∞[
are neighborhoods of z for K = R, but not for K = C ([z − ǫ, ∞[ not even being
defined for z ∈
/ R); the sets {z}, {w ∈ K : Re w ≥ Re z}, {w ∈ K : Re w ≥ Re z +ǫ}
are never neighborhoods of z).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 74

(b) If φ(n) is a statement for each n ∈ N, then φ(n) is said to be true for almost all
n ∈ N if, and only if, there exists a finite subset A ⊆ N such that φ(n) is true for
each n ∈ N \ A, i.e. if, and only if, φ(n) is always true, with the possible exception
of finitely many cases.
Remark 7.8. In the language of Def. 7.7, the sequence (zn )n∈N converges to z if, and
only if, every neighborhood of z contains almost all zn .
Definition 7.9. The sequence (zn )n∈N in K is called bounded if, and only if, the set
{|zn | : n ∈ N} is bounded in the sense of Def. 2.27(a).
Proposition 7.10. Let (zn )n∈N be a sequence in K.

(a) Limits are unique, that means if z, w ∈ K such that limn→∞ zn = z and limn→∞ zn =
w, then z = w.

(b) If (zn )n∈N is convergent, then it is bounded.

Proof. (a): Exercise.


(b): If limn→∞ zn = z, then A := {|zn | : |zn − z| ≥ 1} ∪ {|z1 |} is nonempty and finite.
According to Th. 3.13(a), A has an upper bound M . Then max{M, |z| + 1} is an upper
bound for {|zn | : n ∈ N}, and 0 is always a lower bound, showing that the sequence is
bounded. 
Proposition 7.11. Let (zn )n∈N be a sequence in C with limn→∞ zn = 0.

(a) If (bn )n∈N is a sequences in C such that there exists C ∈ R+ with |bn | ≤ C|zn | for
almost all n, then limn→∞ bn = 0.

(b) If (cn )n∈N is a bounded sequence in C, then limn→∞ (cn zn ) = 0.

Proof. (a): Given ǫ > 0, there exists N ∈ N such that |zn | < ǫ/C and |bn | ≤ C|zn | for
each n > N . Then, for each n > N , |bn | ≤ C|zn | < ǫ, proving limn→∞ bn = 0.
(b): If (cn )n∈N is bounded, then there exists C ∈ R+ such that |cn | ≤ C for each n ∈ N.
Thus, |cn zn | ≤ C|zn | for each n ∈ N, implying limn→∞ (cn zn ) = 0 via (a). 
Example 7.12. The sequences ((−1)n )n∈N and (b)n∈N with b ∈ C are bounded. Since,
1
for each a ∈ C, limn→∞ n+a = 0 by Example 7.2(b), we obtain

(−1)n b
lim = lim =0 (7.10)
n→∞ n + a n→∞ n + a

from Prop. 7.11(b).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 75

Theorem 7.13. (a) Let (zn )n∈N and (wn )n∈N be sequences in C. Moreover, let z, w ∈ C
with limn→∞ zn = z and limn→∞ wn = w. We have the following identities:
lim (λzn ) = λz for each λ ∈ C, (7.11a)
n→∞
lim (zn + wn ) = z + w, (7.11b)
n→∞
lim (zn wn ) = zw, (7.11c)
n→∞
lim zn /wn = z/w given all wn 6= 0 and w 6= 0, (7.11d)
n→∞
lim |zn | = |z|, (7.11e)
n→∞
lim z̄n = z̄, (7.11f)
n→∞
lim znp = z p for each p ∈ N. (7.11g)
n→∞

(b) Let (xn )n∈N and (yn )n∈N be sequences in R. Moreover, let x, y ∈ R with limn→∞ xn =
x and limn→∞ yn = y. Then
lim max{xn , yn } = max{x, y}, (7.12a)
n→∞
lim min{xn , yn } = min{x, y}. (7.12b)
n→∞

(c) If, in the situation of (b) (i.e. for real sequences), xn ≤ yn holds for almost all
n ∈ N, then x ≤ y. In particular, if almost all xn ≥ 0, then x ≥ 0.

Proof. We start with the identities of (a).


(7.11a): For λ = 0, there is nothing to prove. For λ 6= 0 and ǫ > 0, there exists N ∈ N
such that, for each n > N , |zn − z| < ǫ/|λ|, implying
∀ |λ zn − λ z| = |λ| |zn − z| < ǫ. (7.13a)
n>N

(7.11b): Given ǫ > 0, there exists N ∈ N such that, for each n > N , |zn − z| < ǫ/2 and
|wn − w| < ǫ/2, implying
∀ |zn + wn − (z + w)| ≤ |zn − z| + |wn − w| < ǫ/2 + ǫ/2 = ǫ. (7.13b)
n>N

(7.11c): Let M1 := max{|z|, 1}. According to Prop. 7.10(b), there exists M2 ∈ R+ such
that M2 is an upper bound for {|wn | : n ∈ N}. Moreover, given ǫ > 0, there exists
N ∈ N such that, for each n > N , |zn − z| < ǫ/(2M2 ) and |wn − w| < ǫ/(2M1 ), implying
 
|zn wn − zw| = (zn − z)wn + z(wn − w)
 
∀  M2 ǫ M1 ǫ  (7.13c)
n>N
≤ |wn | · |zn − z| + |z| · |wn − w| < + = ǫ.
2M2 2M1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 76

(7.11d): We first consider the case, where all zn = 1. Given ǫ > 0, there exists N ∈ N
such that, for each n > N , |wn − w| < ǫ |w|2 /2 and |wn − w| < |w|/2 (since w 6= 0 for
this case), implying |w| ≤ |w − wn | + |wn | < |w|/2 + |wn | and |wn | > |w|/2. Thus,

1 1 wn − w 2 |wn − w| 2 ǫ |w|2
∀ − = ≤ < = ǫ. (7.13d)
n>N wn w wn w |w|2 |w|2 2
The general case now follows from (7.11c).
(7.11e): This is a consequence of the inverse triangle inequality (5.13): Given ǫ > 0,
there exists N ∈ N such that, for each n > N , |zn − z| < ǫ, implying

∀ |zn | − |z| ≤ |zn − z| < ǫ. (7.13e)
n>N

(7.11f): Write zn = xn + iyn and z = x + iy with xn , yn , x, y ∈ R, n ∈ N. Then we know


limn→∞ xn = x and limn→∞ yn = y from (7.2), and
(7.11a),(7.11b)
lim z̄n = lim (xn − iyn ) = x − iy = z̄, (7.13f)
n→∞ n→∞

which establishes the case.


(7.11g) follows by induction from (7.11c) (cf. (7.16b) below).
The proofs for the two identities of (b) are left as exercises.
(c): Proceeding by contraposition, assume x > y and set s := (x+y)/2. Then y < s < x
and yn < s < xn holds for almost all n, i.e. xn ≤ yn does not hold for almost all n. 
Example 7.14. (a) limn→∞ n+a
n+b
= 1 for each a, b ∈ C: Here zn := (n + a)/(n + b) (if
n = −b, then set zn := w with w ∈ C arbitrary). Using (7.11b) and (7.11d), one
obtains
n+a 1 + a/n lim 1 + lim na 1+0
n→∞ n→∞
lim = lim = b
= = 1. (7.14)
n→∞ n + b n→∞ 1 + b/n lim 1 + lim n 1+0
n→∞ n→∞

(b) Using (7.11b), (7.11d), and (7.11g), one obtains


2n5 − 3in3 + 2i 2 − 3i/n2 + 2i/n5 2+0+0 2
lim 5
= lim 4
= = . (7.15)
n→∞ 3n + 17n n→∞ 3 + 17/n 3+0 3
(1) (k)
Corollary 7.15. For k ∈ N, let (zn )n∈N , . . . , (zn )n∈N be sequences in C. Moreover,
(j)
let z (1) , . . . , z (k) ∈ C with limn→∞ zn = z (j) for each j ∈ {1, . . . , k}. Then
k
X k
X
lim zn(j) = z (j) , (7.16a)
n→∞
j=1 j=1
k
Y k
Y
lim zn(j) = z (j) . (7.16b)
n→∞
j=1 j=1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 77

Proof. (7.16) follows by simple inductions from (7.11b) and (7.11c), respectively. 

Theorem 7.16 (Sandwich Theorem). Let (xn )n∈N , (yn )n∈N , and (an )n∈N be sequences
in R. If xn ≤ an ≤ yn holds for almost all n ∈ N, then

lim xn = lim yn = x ∈ R ⇒ lim an = x. (7.17)


n→∞ n→∞ n→∞

Proof. Given ǫ > 0, there exists N ∈ N such that, for each n > N , xn ≤ an ≤ yn ,
|xn − x| < ǫ, and |yn − x| < ǫ, implying

∀ x − ǫ < xn ≤ an ≤ yn < x + ǫ, (7.18)


n>N

which establishes the case. 


1 1
Example 7.17. Since, 0 < n!
≤ n
holds for each n ∈ N, the Sandwich Th. 7.16 implies

1
lim = 0. (7.19)
n→∞ n!

Definition 7.18. Let (xn )n∈N be a sequence in R. The sequence is said to diverge to
∞ (resp. to −∞), denoted limn→∞ xn = ∞ (resp. limn→∞ xn = −∞) if, and only if, for
each K ∈ R, almost all xn are bigger (resp. smaller) than K. Thus,

lim xn = ∞ ⇔ ∀ ∃ ∀ xn > K, (7.20a)


n→∞ K∈R N ∈N n>N
lim xn = −∞ ⇔ ∀ ∃ ∀ xn < K. (7.20b)
n→∞ K∈R N ∈N n>N

Theorem 7.19. Suppose S := (xn )n∈N is a monotone sequence in R (increasing or


decreasing). Defining A := {xn : n ∈ N}, the following holds:


 sup A if S is increasing and bounded,

∞ if S is increasing and not bounded,
lim xn = (7.21)
n→∞ 

 inf A if S is decreasing and bounded,

−∞ if S is decreasing and not bounded.

Proof. We treat the increasing case; the decreasing case is proved completely analo-
gously. If A is bounded and ǫ > 0, let K := sup A − ǫ; if A is unbounded, then let
K ∈ R be arbitrary. In both cases, since K can not be an upper bound, there exists
N ∈ N such that xN > K. Since the sequence is increasing, for each n > N , xN ≤ xn ,
showing | sup A − xn | < ǫ in the bounded case, and xn > K in the unbounded case. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 78

Example 7.20. Theorem 7.19 implies


 
∀ lim nk = ∞, lim (−nk ) = −∞ . (7.22)
k∈N n→∞ n→∞

It is sometimes necessary to consider so-called subsequences and reorderings of a given


sequence. Here, we are interested in sequences in R or C, but for subsequences and
reorderings it is irrelevant in which set A the sequence takes its values. As it presents
virtually no extra difficulty to introduce the notions for general sequences, and since we
will need to consider sequences with values in sets other than R or C in Analysis II, we
admit general sequences in the following definition.
Definition 7.21. Let A be an arbitrary nonempty set. Consider a sequence σ : N −→
A. Given a function φ : N −→ N (that means (φ(n))n∈N constitutes a sequence of
indices), the new sequence (σ ◦ φ) : N −→ A is called a subsequence of σ if, and
only if, φ is strictly increasing (i.e. 1 ≤ φ(1) < φ(2) < . . . ). Moreover, σ ◦ φ is
called a reordering of σ if, and only if, φ is bijective. One can write σ in the form
(zn )n∈N by setting zn := σ(n), and one can write σ ◦ φ in the form (wn )n∈N by setting
wn := (σ ◦ φ)(n) = zφ(n) . Especially for a subsequence of (zn )n∈N , it is also common
to write (znk )k∈N . This notation corresponds to the one above if one lets nk := φ(k).
Analogous definitions work if the index set N of σ is replaced by a general countable
nonempty index set I.
Example 7.22. Consider the sequence (1, 2, 3, . . . ). Then (2, 4, 6, . . . ) constitutes a
subsequence and (2, 1, 4, 3, 6, 5, . . . ) constitutes a reordering. Using the notation of Def.
7.21, the original sequence is given by σ : N −→ N, σ(n) := n; the subsequence
is selected via φ1 : N ( −→ N, φ1 (n) := 2n; and the reordering is accomplished via
n + 1 if n is odd,
φ2 : N −→ N, φ2 (n) :=
n − 1 if n is even.
Proposition 7.23. Let (zn )n∈N be a sequence in C. If limn→∞ zn = z, then every
subsequence and every reordering of (zn )n∈N is also convergent with limit z.

Proof. Let (wn )n∈N be a subsequence of of (zn )n∈N , i.e. there is a strictly increasing
function φ : N −→ N such that wn = zφ(n) . If limn→∞ zn = z, then, given ǫ > 0, there
is N ∈ N such that zn ∈ Bǫ (z) for each n > N . For Ñ choose any number from N that
is ≥ N and in φ(N). Take M := φ−1 (Ñ ) (where φ−1 : φ(N) −→ N). Then, for each
n > M , one has φ(n) > Ñ ≥ N , and, thus, wn = zφ(n) ∈ Bǫ (z), showing limn→∞ wn = z.
Let (wn )n∈N be a reordering of (zn )n∈N , i.e. there is a bijective function φ : N −→ N
such that wn = zφ(n) . Let ǫ and N be as before. Define
M := max{φ−1 (n) : n ≤ N }. (7.23)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 79

As φ is bijective, it is φ(n) > N for each n > M . Then, for each n > M , one has
wn = zφ(n) ∈ Bǫ (z), showing limn→∞ wn = z. 

Definition 7.24. Let (zn )n∈N be a sequence in K. A point z ∈ K is called a cluster


point or an accumulation point of the sequence if, and only if, for each ǫ > 0, Bǫ (z)
contains infinitely many members of the sequence (i.e. #{n ∈ N : zn ∈ Bǫ (z)} = ∞).

Example 7.25. The sequence ((−1)n )n∈N has cluster points 1 and −1.

Proposition 7.26. A point z ∈ K is a cluster point of the sequence (zn )n∈N in K if,
and only if, the sequence has a subsequence converging to z.

Proof. If (wn )n∈N is a subsequence of (zn )n∈N , limn→∞ wn = z, then every Bǫ (z), ǫ > 0,
contains infinitely many wn , i.e. infinitely many zn , i.e. z is a cluster point of (zn )n∈N .
Conversely, if z is a cluster point of (zn )n∈N , then, inductively, define φ : N −→ N as
follows: For φ(1), choose the index k of any point zk in B1 (z) (such a point exists, since
z is a cluster point of the sequence). Now assume that n > 1 and that φ(m) have already
been defined for each m < n. Let M := max{φ(m) : m < n}. Since B 1 (z) contains
n
infinitely many zk , there must be some zk ∈ B 1 (z) such that k > M . Choose this k as
n
φ(n). Thus, by construction, φ is strictly increasing, i.e. (wn )n∈N with wn := zφ(n) is a
subsequence of (zn )n∈N . Moreover, for each ǫ > 0, there is N ∈ N such that 1/N < ǫ.
Then, for each n > N , wn ∈ B 1 (z) ⊆ B 1 (z) ⊆ Bǫ (z), showing limn→∞ wn = z. 
n N

Theorem 7.27 (Bolzano-Weierstrass). Every bounded sequence S := (xn )n∈N in K


has at least one cluster point in K. Moreover, for K = R, the set A := {x ∈ R :
x is cluster point of S} has a max x∗ ∈ R and a min x∗ ∈ R, i.e. every bounded sequence
in R has a largest and a smallest cluster point. In addition, for each ǫ > 0, the inequality
x∗ − ǫ < xn < x∗ + ǫ holds for almost all n.

Proof. We first consider the case K = R. Define

A∗ := {x ∈ R : xn ≤ x for almost all n}, (7.24a)


A∗ := {x ∈ R : xn ≥ x for almost all n}. (7.24b)

We claim A∗ 6= ∅ is bounded from below and x∗ = max A = inf A∗ ; A∗ 6= ∅ is bounded


from above and x∗ = min A = sup A∗ . We prove the claim for A∗ – the proof for A∗ is
conducted completely analogous. Let m, M ∈ R be a lower and an upper bound for S,
respectively. Then M ∈ A∗ , showing A∗ 6= ∅; and m is a lower bound for A∗ . Since A∗
is bounded from below, a := inf A∗ ∈ R by the completeness of R. Moreover, for each
/ A∗ , as a is a lower bound for A∗ , i.e. xn > a − ǫ holds for infinitely many
ǫ > 0, a − ǫ ∈
n ∈ N. On the other hand, a + ǫ/2 ∈ A∗ follows from a being the largest lower bound

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 80

of A∗ , i.e. xn > a + ǫ/2 holds for only finitely many n (if any). In particular, we have
shown xn < a + ǫ holds for almost all n, and a − ǫ < xn < a + ǫ must hold for infinitely
many n, showing a is a cluster point of S. To see that a is the largest cluster point of
S (i.e. a = max A), we have to show that x > a implies x is not a cluster point of S.
However, letting ǫ := x − a > 0, we had seen above that xn > a + ǫ/2 holds for only
finitely many n, i.e. Bǫ/2 (x) contains only finitely many xn , showing x is not a cluster
point of S.
It now remains to consider the complex case, i.e. a bounded sequence S := (zn )n∈N in C.
For each n ∈ N, let zn = xn +iyn with xn , yn ∈ R. Due to Th. 5.9(d), we have |xn | ≤ |zn |
and |yn | ≤ |zn |, i.e. the boundedness of S implies the boundedness of both (xn )n∈N and
(yn )n∈N . Then we know that (xn )n∈N has a cluster point x and, by Prop. 7.26, S
has a subsequence (znj )j∈N such that x = limj→∞ xnj . As the subsequence (ynj )j∈N is
still bounded, it must have a cluster point y and a subsequence (ynjk )k∈N such that
y = limk→∞ ynjk . Since x = limk→∞ xnjk as well, we now have limk→∞ znjk = x + iy =: z,
i.e. S has a subsequence converging to z. According to Prop. 7.26, z is a cluster point
of S. 
Definition 7.28. A sequence (zn )n∈N in C is defined to be a Cauchy sequence if, and
only if, for each ǫ ∈ R+ , there exists N ∈ N such that |zn − zm | < ǫ for each n, m > N ,
i.e.
(zn )n∈N Cauchy ⇔ ∀+ ∃ ∀ |zn − zm | < ǫ. (7.25)
ǫ∈R N ∈N n,m>N

Theorem 7.29. The sequence (zn )n∈N in C is convergent if, and only if, it is a Cauchy
sequence.

Proof. Suppose the sequence is convergent with limn→∞ zn = z. Then, given ǫ > 0,
there is N ∈ N such that zn ∈ B 2ǫ (z) for each n > N . If n, m > N , then |zn − zm | ≤
|zn − z| + |z − zm | < 2ǫ + 2ǫ = ǫ, establishing that (zn )n∈N is a Cauchy sequence.
Conversely, suppose the sequence is a Cauchy sequence. Using similar reasoning as in
the proof of Prop. 7.10(b), we first show the sequence is bounded. If the sequence is
Cauchy, then there exists N ∈ N such that |zn − zm | < 1 for all n, m > N . Thus, the
set A := {|zn | : |zn − zN +1 | ≥ 1} ∪ {|z1 |} ⊆ R+
0 is nonempty and finite. According to
Th. 3.13(a), A has an upper bound M . Then max{M, |zN +1 | + 1} is an upper bound for
{|zn | : n ∈ N}, showing that the sequence is bounded. From Th. 7.27, we obtain that
the sequence has a cluster point z. It remains to show limn→∞ zn = z. Given ǫ > 0,
choose N ∈ N such that |zn − zm | < ǫ/2 for all n, m > N . Since z is a cluster point,
there exists k > N such that |zk − z| < ǫ/2. Thus,
ǫ ǫ
∀ |zn − z| ≤ |zn − zk | + |zk − z| < + = ǫ, (7.26)
n>N 2 2
proving limn→∞ zn = z. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 81

Example 7.30. Consider the sequence S := (sn )n∈N defined by


n
X 1 1 1
sn := =1+ + ··· + . (7.27)
k=1
k 2 n

We claim S is not a Cauchy sequence and, thus, not convergent by Th. 7.29: For each
N ∈ N, we find n, m > N such that sn −sm > 1/2, namely m = N +1 and n = 2(N +1):
2(N +1)
X 1 1 1 1
s2(N +1) − sN +1 = = + + ··· +
k=N +2
k N +2 N +3 2(N + 1)
1 1
> (N + 1) · = . (7.28)
2(N + 1) 2
While we have just seen that S is not convergent, it is clearly increasing, i.e. Th. 7.19
implies S is unbounded and limn→∞ sn = ∞. Sequences defined by longer and longer
sums are known as series and will be studied further in Sec. 7.3 below. The series of the
present example is known as the harmonic series. It has become famous as the simplest
example of a series that does not converge even though its summands converge to 0. In
terms of the notation introduced in Sec. 7.3 below, we have shown

X 1 1 1
=1+ + + · · · = ∞. (7.29)
k=1
k 2 3

7.2 Continuity
7.2.1 Definitions and First Examples

Roughly, a function is continuous if a small change in its input results in a small change
of its output. For functions defined on an interval, the notion of continuity makes
precise the idea of a function having no jump – no discontinuity – at some point x in
its domain. For example, we would say the sign function of (5.8) has precisely one
jump – one discontinuity – at x = 0, whereas quadratic functions (or, more generally,
polynomials) do not have any jumps – they are continuous.
Definition 7.31. Let M ⊆ C. If ζ ∈ M , then a function f : M −→ K is said to be
continuous in ζ if, and only if, for each ǫ > 0, there is δ > 0 such that the distance
between the values f (z) and f (ζ) is less than ǫ, provided the distance between z and ζ
is less than δ, i.e. if, and only if,

∀ + ∃ + ∀ |z − ζ| < δ ⇒ |f (z) − f (ζ)| < ǫ . (7.30)
ǫ∈R δ∈R z∈M

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 82

Moreover, f is called continuous if, and only if, f is continuous in every ζ ∈ M . The set of
all continuous functions from f : M −→ K is denoted by C(M, K), C(M ) := C(M, R).
Example 7.32. (a) Every constant map f : M −→ K, ∅ 6= M ⊆ C, is continuous: In
this case, given ǫ, we can choose any δ > 0 we want, say δ := 42: If ζ, z ∈ M , then
|f (ζ) − f (z)| = 0 < ǫ, which holds independently of δ, in particular, if |ζ − z| < δ.
(b) Every affine function f : K −→ K, f (z) := az + b is continuous: For a = 0, this
follows from (a). For a 6= 0, given ǫ > 0, choose δ := ǫ/|a|. Then,
 ǫ 
|z − ζ| < δ = ⇒ f (z) − f (ζ) = az + b − aζ − b
 |a| 
∀  ǫ . (7.31)
ζ,z∈K
= |a| |z − ζ| < |a| =ǫ
|a|

(c) The sign function of (5.8) is not continuous: It is continuous in each ξ ∈ R\{0}, but
not continuous in 0: If ξ 6= 0, then, given ǫ > 0, choose δ := |ξ|. If |x − ξ| < δ, then
sgn(x) = sgn(ξ), i.e. | sgn(x) − sgn(ξ)| = 0 < ǫ, proving continuity in ξ. However,
at 0, for ǫ := 1/2, we have
1
∀ sgn(0) − sgn(δ/2) = |0 − 1| = 1 > = ǫ, (7.32)
δ>0 2
showing sgn is not continuous in 0.

Some subtleties arise from the possibility that f can be defined on subsets of C with
very different properties. The notions introduced in Def. 7.33 help to deal with these
subtleties.
Definition 7.33. Let M ⊆ C.

(a) The point z ∈ C is called a cluster point or accumulation point of M if, and only if,
each ǫ-neighborhood of z, ǫ ∈ R+ , contains infinitely many points of M , i.e. if, and
only if,
∀ + #(M ∩ Bǫ (z)) = ∞. (7.33)
ǫ∈R

Note: A cluster point of M is not necessarily in M .


(b) The point z is called an isolated point of M if, and only if, there is ǫ ∈ R+ such
that Bǫ (z) ∩ M = {z}. Note: An isolated point of M is always in M .
Proposition 7.34. If M ⊆ C, then each point of M is either a cluster point or an
isolated point of M , i.e.
˙ ∈ M : z isolated point of M }.
M = {z ∈ M : z cluster point of M } ∪{z (7.34)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 83

Proof. Consider z ∈ M that is not a cluster point of M . We have to show that z is an


isolated point of M . Since z is not a cluster point of M , there exists ǫ̃ > 0 such that
A := (M ∩ Bǫ̃ (z)) \ {z} is finite. Define
(
min{|a − z| : a ∈ A} if A 6= ∅,
ǫ := (7.35)
ǫ̃ if A = ∅.

Then Bǫ (z) ∩ M = {z}, showing z is an isolated point of M . Finally, the union in (7.34)
is clearly disjoint. 
Lemma 7.35. Let M ⊆ C, f : M −→ K. If ζ is an isolated point of M , then f is
always continuous in ζ.

Proof. Independently of the concrete definition of f , we know there is δ > 0 such that
Bδ (ζ) ∩ M = {ζ}. In other words, if z ∈ M with |z − ζ| < δ, then z = ζ, implying
|f (z) − f (ζ)| = 0 < ǫ for each ǫ > 0, showing f to be continuous in ζ. 
Example 7.36. (a) The sign function restricted to the set M :=]−∞, −1]∪{0}∪[1, ∞[,
i.e. 
1
 for x ∈ [1, ∞[,
sgn(x) = 0 for x = 0,


−1 for x ∈] − ∞, −1]
is continuous: As in Ex. 7.32(c), one sees that sgn is continuous in each ξ ∈ M \{0}.
However, now it is also continuous in 0, since 0 is an isolated point of M .

(b) Every function f : N −→ K is continuous, since every n ∈ N is an isolated point of


N (due to {n} = N ∩ B 1 (n)).
2

7.2.2 Continuity, Sequences, and Function Arithmetic

To make available the power of the results on convergent sequences from Sec. 7.1 to
investigations regarding the continuity of functions, we need to understand the relation-
ship between both notions. The core of this relationship is the contents of the following
Th. 7.37, which provides a criterion allowing one to test continuity in terms of convergent
sequences:
Theorem 7.37. Let M ⊆ C, f : M −→ K. If ζ ∈ M , then f is continuous in ζ if, and
only if, for each sequence (zn )n∈N in M with limn→∞ zn = ζ, the sequence (f (zn ))n∈N
converges to f (ζ), i.e.

lim zn = ζ ⇒ lim f (zn ) = f (ζ). (7.36)


n→∞ n→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 84

Proof. If ζ ∈ M is an isolated point of M , then there is δ > 0 such that M ∩Bδ (ζ) = {ζ}.
Then every f : M −→ K is continuous in ζ according to Lem. 7.35. On the other hand,
every sequence in M converging to ζ must be finally constant and equal to ζ, i.e. (7.36)
is trivially valid at ζ. Thus, the assertion of the theorem holds if ζ ∈ M is an isolated
point of M .
If ζ ∈ M is not an isolated point of M , then ζ is a cluster point of M according to Prop.
7.34. So, for the remainder of the proof, let ζ ∈ M be a cluster point of M . Assume
that f is continuous in ζ and (zn )n∈N is a sequence in M with limn→∞ zn = ζ. For each
ǫ > 0, there is δ > 0 such that z ∈ M and |z − ζ| < δ implies |f (z) − f (ζ)| < ǫ. Since
limn→∞ zn = ζ, there is also N ∈ N such that, for each n > N , |zn − ζ| < δ. Thus,
for each n > N , |f (zn ) − f (ζ)| < ǫ, proving limn→∞ f (zn ) = f (ζ). Conversely, assume
that f is not continuous in ζ. We have to construct a sequence (zn )n∈N in M with
limn→∞ zn = ζ, but (f (zn ))n∈N does not converge to f (ζ). Since f is not continuous
in ζ, there must be some ǫ0 > 0 such that, for each 1/n, n ∈ N, there is at least one
zn ∈ M satisfying |zn − ζ| < 1/n and |f (zn ) − f (ζ)| ≥ ǫ0 . Then (zn )n∈N is a sequence in
M with limn→∞ zn = ζ and (f (zn ))n∈N does not converge to f (ζ). 

We can now apply the rules of Th. 7.13 to see that all the arithmetic operations defined
in Not. 6.2 preserve continuity:

Theorem 7.38. Let M ⊆ C, f, g : M −→ K, λ ∈ K, ζ ∈ M . If f, g are both continuous


in ζ, then λf , f + g, f g, f /g for g 6= 0, |f |, Re f , and Im f are all continuous in ζ. If
K = R, then max(f, g), min(f, g), f + and f − , are also all continuous in ζ.

Proof. Let (zn )n∈N be a sequence in M such that limn→∞ zn = ζ. Then the continuity
of f and g in ζ yields limn→∞ f (zn ) = f (ζ) and limn→∞ g(zn ) = g(ζ). Then

(7.11a) ⇒ lim (λf )(zn ) = (λf )(ζ),


n→∞
(7.11b) ⇒ lim (f + g)(zn ) = (f + g)(ζ),
n→∞
(7.11c) ⇒ lim (f g)(zn ) = (f g)(ζ),
n→∞
(7.11d) ⇒ lim (f /g)(zn ) = (f /g)(ζ),
n→∞
(7.11e) ⇒ lim |f |(zn ) = |f |(ζ),
n→∞
(7.2) ⇒ lim (Re f )(zn ) = (Re f )(ζ),
n→∞
(7.2) ⇒ lim (Im f )(zn ) = (Im f )(ζ).
n→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 85

If f, g are both R-valued, then we also have

(7.12a) ⇒ lim max(f, g)(zn ) = max(f, g)(ζ),


n→∞
(7.12b) ⇒ lim min(f, g)(zn ) = min(f, g)(ζ),
n→∞

and, finally, the continuity of f + and f − follows from the continuity of max(f, g). 

Corollary 7.39. A function f : M −→ C, M ⊆ C, is continuous in ζ ∈ M if, and


only if, both Re f and Im f are continuous in ζ.

Proof. If f is continuous in ζ, then Re f and Im f are both continuous in ζ by Th. 7.38.


If Re f and Im f are both continuous in ζ, then, as

f = Re f + i Im f, (7.37)

f is continuous in ζ, once again, by Th. 7.38. 

Example 7.40. (a) The continuity of the absolute value function z 7→ |z| on K can be
concluded directly from (7.11e) and, alternatively, from combining the continuity
of f : K −→ K, f (z) = z, according to Ex. 7.32(b), with the continuity of |f |
according to Th. 7.38.
P
(b) Every polynomial P : K −→ K, P (x) = nj=0 aj xj , aj ∈ K, is continuous: First
note that every monomial x 7→ xj is continuous on K by (7.11g). Then Th. 7.38
implies the continuity of x 7→ aj xj on K. Now the continuity of P follows from
(7.16a) or, alternatively, by an induction from the f + g part of Th. 7.38.

(c) Let P, Q : K −→ K, be polynomials and let A := Q−1 {0} the set of all zeros of
Q (if any). Then the rational function (P/Q) : K \ A −→ K is continuous as a
consequence of (b) plus the f /g part of Th. 7.38.

Theorem 7.41. Let Df , Dg ⊆ C, f : Df −→ C, g : Dg −→ K, f (Df ) ⊆ Dg . If f


is continuous in ζ ∈ Df and g is continuous in f (ζ) ∈ Dg , then g ◦ f : Df −→ K is
continuous in ζ. In consequence, if f and g are both continuous, then the composition
g ◦ f is also continuous.

Proof. Let ζ ∈ Df and assume f is continuous in ζ and g is continuous in f (ζ). If (zn )n∈N
is a sequence in Df such that limn→∞ zn = ζ, then the continuity of f in ζ implies that
limn→∞ f (zn ) = f (ζ). Then the continuity of g in f (ζ) implies limn→∞ g(f (zn )) =
g(f (ζ)), thereby establishing the continuity of g ◦ f in ζ. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 86

7.2.3 Bounded, Closed, and Compact Sets

Subsets A of C (and even subsets of R) can be extremely complicated. If the set A has
one or more of the benign properties defined in the following, then this can often be
exploited in some useful way (we will see an important example in Th. 7.54 below).
Definition 7.42. Consider A ⊆ C.

(a) A is called bounded if, and only if, A = ∅ or the set {|z| : z ∈ A} is bounded in R
in the sense of Def. 2.27(a), i.e. if, and only if,
∃ A ⊆ BM (0).
M ∈R+

(b) A is called closed if, and only if, every sequence in A that converges in C has its
limit in A (note that ∅ is, thus, closed).
(c) A is called compact if, and only if, A is both closed and bounded.
Example 7.43. (a) Clearly, ∅ and sets containing single points {z}, z ∈ C are com-
pact. The sets C and R are simple examples of closed sets that are not bounded.
(b) Let a, b ∈ R, a < b. Each bounded interval ]a, b[, ]a, b], [a, b[, [a, b] is, indeed,
bounded (e.g. by M := ǫ + max{|a|, |b|} for each ǫ ∈ R+ ). If (xn )n∈N is a sequence
in [a, b], converging to x ∈ R, then Th. 7.13(c) shows a ≤ x ≤ b, i.e. x ∈ [a, b] and
[a, b] is, indeed, closed. Analogously, one sees that the unbounded intervals [a, ∞[
and ] − ∞, a] are also closed. On the other hand, open and half-open intervals are
not closed: For sufficiently large n, the convergent sequence (b − n1 )n∈N is in [a, b[,
but limn→∞ (b − n1 ) = b ∈ / [a, b[, and the other cases are treated analogously. In
particular, only intervals of the form [a, b] (and trivial intervals) are compact.
(c) For each ǫ > 0 and each z ∈ C, the set Bǫ (z) is bounded (since Bǫ (z) ⊆ Bǫ+|z| (0)
by the triangle inequality), but not closed (since, for sufficiently large n ∈ N,
(z + ǫ − n1 )n∈N is a sequence in Bǫ (z), converging to z + ǫ ∈
/ Bǫ (z)). In particular,
Bǫ (z) is not compact.
Proposition 7.44. (a) Finite unions of bounded (resp. closed, resp. compact) sets are
Sn if A1 , . . . , An ⊆ C, n ∈ N, are bounded
bounded (resp. closed, resp. compact), i.e.
(resp. closed, resp. compact), then A := j=1 Aj is also bounded (resp. closed, resp.
compact).
(b) Arbitrary (i.e. finite or infinite) intersections of bounded (resp. closed, resp. com-
pact) sets are bounded (resp. closed, resp. compact), i.e. if I 6= ∅ is an arbitrary
set and, for each j ∈ I, Aj ⊆ C is bounded (resp. closed, resp. compact), then
index T
A := j∈I Aj is also bounded (resp. closed, resp. compact).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 87

Proof. (a): Exercise.


(b): Fix j0 ∈ I. If all Aj , j ∈ I,Tare bounded, then, in particular, there is M ∈ R+ such
that Aj0 ⊆ BM (0). Thus, A = j∈I Aj ⊆ Aj0 ⊆ BM (0) shows A is also bounded. If all
Aj , j ∈ I, are closed and (an )n∈N is a sequence in A that converges to some z ∈ C, then
(an )n∈N is a sequenceT in each Aj , j ∈ I, and, since each Aj is closed, z ∈ Aj for each
j ∈ I, i.e. z ∈ A = j∈I Aj . If all Aj , j ∈ I, are compact, then they are all closed and
bounded and, thus, A is closed and bounded, i.e. A is compact. 
Example 7.45. (a) According to Prop. 7.44(a), all finite subsets of C are compact.
S
(b) N = n∈N S {n}1 shows1 that infinite unions of compact sets can be unbounded, and
]0, 1[= n∈N [ n , 1 − n ] shows that infinite unions of compact sets are not always
closed.

Many more examples of closed sets can be obtained as preimages of closed sets under
continuous maps according to the following remark:
Remark 7.46. In Analysis II, it will be shown in the more general context of maps
f between topological spaces that a map f is continuous if, and only if, all preimages
f −1 (A) under f of closed sets A are closed. Here, we will only prove the following special
case:
f : C −→ K continuous and A ⊆ K closed ⇒ f −1 (A) ⊆ C closed. (7.38)
Indeed, suppose f is continuous and A ⊆ K is closed. If (zn )n∈N is a sequence in f −1 (A)
with limn→∞ zn = z ∈ C, then (f (zn ))n∈N is a sequence in A. The continuity of f then
implies limn→∞ f (zn ) = f (z) and, then, f (z) ∈ A, since A is closed. Thus, z ∈ f −1 (A),
showing f −1 (A) is closed.
Example 7.47. (a) For each z ∈ C and each r > 0, the closed disk B r (z) := {w ∈ C :
|z − w| ≤ r} with radius r and center z is, indeed, closed by (7.38), since
B r (z) = f −1 [0, r], (7.39)
where f is the continuous map f : C −→ R, f (w) := |z − w|. Since B r (z) is clearly
bounded, it is also compact.
(b) For each z ∈ C and each r > 0, the circle (also called a 1-sphere) Sr (z) := {w ∈ C :
|z − w| = r} with radius r and center z is closed by (7.38), since Sr (z) = f −1 {r},
where f is the same map as in (7.39). Moreover, Sr (z) is also clearly bounded, and,
thus, compact.
(c) According to (7.38), for each x ∈ R, the closed half-spaces {z ∈ C : Re z ≥ x} =
Re−1 [x, ∞[ and {z ∈ C : Im z ≥ x} = Im−1 [x, ∞[ are, indeed, closed.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 88

Theorem 7.48. A subset K of C is compact if, and only if, every sequence in K has a
subsequence that converges to some limit z ∈ K.

Proof. If K is closed and bounded, and (zn )n∈N is a sequence in K, then the boundedness,
the Bolzano-Weierstrass Th. 7.27, and Prop. 7.26 yield a subsequence that converges to
some z ∈ C. However, since K is closed, z ∈ K.
Conversely, assume every sequence in K has a subsequence that converges to some
limit z ∈ K. Let (zn )n∈N be a sequence in K that converges to some w ∈ C. Then this
sequence must have a subsequence that converges to some z ∈ K. However, according to
Prop. 7.23, it must be w = z ∈ K, showing K is closed. If K is not bounded, then there
exists a sequence (zn )n∈N in K such that limn→∞ |zn | = ∞. Every subsequence (znk )k∈N
then still has the property that limk→∞ |znk | = ∞, in particular, each subsequence is
unbounded and can not converge to some z ∈ C (let alone in K). 

Caveat 7.49. In Analysis II, we will generalize the notion of compactness to subsets
of so-called metric spaces. In metric spaces, it is still true that a set K is compact if,
and only if, every sequence in K has a subsequence that converges to some limit in
K. However, while it remains true that every compact set is closed and bounded, the
converse does not(!) hold in general metric spaces (in general, even in closed sets, there
exist bounded sequences that do not have convergent subsequences).

One reason that compact sets are useful is that real-valued continuous functions on
compact sets assume a maximum and a minimum, which is the contents of Th. 7.54
below. In preparation, we now define maxima and minima for real-valued functions.

Definition 7.50. Let M ⊆ C, f : M −→ R.

(a) Given z ∈ M , f has a (strict) global min at z if, and only if, f (z) ≤ f (w) (f (z) <
f (w)) for each w ∈ M \ {z}. Analogously, f has a (strict) global max at z if, and
only if, f (z) ≥ f (w) (f (z) > f (w)) for each w ∈ M \{z}. Moreover, f has a (strict)
global extreme value at z if, and only if, f has a (strict) global min or a (strict)
global max at z.

(b) Given z ∈ M , f has a (strict) local min at z if, and only if, there exists ǫ > 0
such that f (z) ≤ f (w) (f (z) < f (w)) for each w ∈ {w ∈ M : |z − w| < ǫ} \ {z}.
Analogously, f has a (strict) local max at z if, and only if, there exists ǫ > 0 such
that f (z) ≥ f (w) (f (z) > f (w)) for each w ∈ {w ∈ M : |z − w| < ǫ} \ {z}.
Moreover, f has a (strict) local extreme value at z if, and only if, f has a (strict)
local min or a (strict) local max at z.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 89

Remark 7.51. In the context of Def. 7.50, it is immediate from the respective definitions
that f has a (strict) global min at z ∈ M if, and only if, −f has a (strict) global max
at z. Moreover, the same holds if “global” is replaced by “local”. It is equally obvious
that every (strict) global min/max is a (strict) local min/max.

Theorem 7.52. If K ⊆ C is compact, and f : K −→ C is continuous, then f (K) is


compact.

Proof. If (wn )n∈N is a sequence in f (K), then, for each n ∈ N, there is some zn ∈ K
such that f (zn ) = wn . As K is compact, there is a subsequence (an )n∈N of (zn )n∈N
with limn→∞ an = a for some a ∈ K. Then (f (an ))n∈N is a subsequence of (wn )n∈N and
the continuity of f yields limn→∞ f (an ) = f (a) ∈ f (K), showing that (wn )n∈N has a
convergent subsequence with limit in f (K). By Th. 7.48, we have therefore established
that f (K) is compact. 

Lemma 7.53. If K is a nonempty compact subset of R, then K contains a smallest


and a largest element, i.e. there exist m, M ∈ K such that m ≤ x ≤ M for each x ∈ K.

Proof. Since the compact set K is bounded, we know that

−∞ < m := inf K ≤ sup K =: M < ∞.

According to the definition of the inf and sup as largest lower bound and smallest upper
bound, respectively, for each n ∈ N, there must be elements xn , yn ∈ K such that
m ≤ xn ≤ m + n1 and M − n1 ≤ yn ≤ M . Since the compact set K is also closed, we get
m = limn→∞ xn ∈ K and M = limn→∞ yn ∈ K. 

Theorem 7.54. If K ⊆ C is compact, and f : K −→ R is continuous, then f assumes


its max and its min, i.e. there are zm ∈ K and zM ∈ K such that f has a global min at
zm and a global max at zM . In particular, the continuous function f assumes its max
and min on each compact interval K = [a, b] ⊆ R, a, b ∈ R.

Proof. Since K is compact and f is continuous, f (K) ⊆ R is compact according to


Th. 7.52. Then, by Lem. 7.53, f (K) contains a smallest element m and a largest
element M . This, in turn, implies that there are zm , zM ∈ K such that f (zm ) = m and
f (zM ) = M . 

Example 7.55. On an unbounded set, a continuous function does not necessarily have
a global max or a global min, as one can already see from x 7→ x. An example for a
continuous function on a bounded, but not closed, interval, that does not have a global
max is f : ]0, 1] −→ R, f (x) := 1/x, which is continuous by Th. 7.38.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 90

7.2.4 Intermediate Value Theorem

Theorem 7.56 (Bolzano’s Theorem). Let a, b ∈ R with a < b. If f : [a, b] −→ R is


continuous with f (a) > 0 and f (b) < 0, then f has at least one zero in ]a, b[. More
precisely, the set A := f −1 {0} has a min ξ1 and a max ξ2 , a < ξ1 ≤ ξ2 < b, where f > 0
on [a, ξ1 [ and f < 0 on ]ξ2 , b].

Proof. Let ξ1 := inf f −1 (R−


0 ).

(a): f (ξ1 ) ≤ 0: This is clear if ξ1 = b. If ξ1 < b, then, for each n ∈ N sufficiently large,
there exists xn ∈ [ξ1 , ξ1 + 1/n[⊆ [a, b] such that f (xn ) ≤ 0). Then limn→∞ xn = ξ1 and
the continuity of f implies limn→∞ f (xn ) = f (ξ1 ). Now f (ξ1 ) ≤ 0 is a consequence of
Th. 7.13(c). In particular, (a) yields a < ξ1 and f > 0 on [a, ξ1 [.
(b): f (ξ1 ) ≥ 0: The continuity of f implies limn→∞ f (ξ1 − 1/n) = f (ξ1 ) and, since we
have already seen f (ξ1 − 1/n) > 0 for each n ∈ N sufficiently large, f (ξ1 ) ≥ 0 is again a
consequence of Th. 7.13(c). In particular, we have ξ1 < b.
Combining (a) and (b), we have f (ξ1 ) = 0 and a < ξ1 < b.
Defining ξ2 := sup f −1 (R+0 ), f (ξ2 ) = 0 and a < ξ2 < b is shown completely analogous.
Then f < 0 on ]ξ2 , b] is also clear as well as ξ1 ≤ ξ2 . 
Theorem 7.57 (Intermediate Value Theorem). Let a, b ∈ R with a < b. If f : [a, b] −→
R is continuous, then f assumes every value between f (a) and f (b), i.e.
h i 
min{f (a), f (b)}, max{f (a), f (b)} ⊆ f [a, b] . (7.40)

Proof. If f (a) = f (b), then there is nothing to prove. If f (a) < f (b) and η ∈]f (a), f (b)[,
then consider the auxiliary function g : [a, b] −→ R, g(x) := η − f (x). Then g is
continuous with g(a) = η − f (a) > 0 and g(b) = η − f (b) < 0. According to Bolzano’s
Th. 7.56, there exists ξ ∈]a, b[ such that g(ξ) = η − f (ξ) = 0, i.e. f (ξ) = η as claimed.
If f (b) < f (a) and η ∈]f (b), f (a)[, then consider the auxiliary function g : [a, b] −→ R,
g(x) := f (x)−η. Then g is continuous with g(a) = f (a)−η > 0 and g(b) = f (b)−η < 0.
Once again, according to Bolzano’s Th. 7.56, there exists ξ ∈]a, b[ such that g(ξ) =
f (ξ) − η = 0, i.e. f (ξ) = η. 
Theorem 7.58. If I ⊆ R is an interval (of one of the 8 types listed in (4.8)) and
f : I −→ R is continuous, then f (I) is also an interval (it can degenerate to a single
point if f is constant). More precisely, if ∅ 6= I = [a, b] is a compact interval, then
∅ 6= f (I) = [min f (I), max f (I)]; if I is not a compact interval, then one of the following
9 cases occurs:

f (I) = R, (7.41a)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 91

f (I) =] − ∞, sup f (I)], (7.41b)


f (I) =] − ∞, sup f (I)[, (7.41c)
f (I) = [inf f (I), ∞[ (7.41d)
f (I) = [inf f (I), sup f (I)], (7.41e)
f (I) = [inf f (I), sup f (I)[, (7.41f)
f (I) =] inf f (I), ∞[, (7.41g)
f (I) =] inf f (I), sup f (I)], (7.41h)
f (I) =] inf f (I), sup f (I)[. (7.41i)

Proof. If I is a compact interval, then we merely combine Th. 7.54 with Th. 7.57.
Otherwise, let η ∈ f (I). If f (I) has an upper bound, then Th. 7.57 implies [η, sup f (I)[⊆
f (I) and f (I) ∩ [η, ∞[⊆ [η, sup f (I)]. If f (I) does not have an upper bound, then Th.
7.57 implies f (I) ∩ [η, ∞[= [η, ∞[. Analogously, one obtains f (I)∩] − ∞, η] =] − ∞, η]
or f (I)∩] − ∞, η] = [inf f (I), η] or f (I)∩] − ∞, η] =] inf f (I), η],showing that there
 are
precisely the 9 possibilities of (7.41) for f (I) = f (I)∩] − ∞, η] ∪ f (I) ∩ [η, ∞[ . 

The above results will have striking consequences in the following Sec. 7.2.5.

Example 7.59. The piecewise affine function


 
(−1)n · n − 2n+1
1 1 x − n1 for x ∈ [ n1 , n−1
1
], n even,

f : ]0, 1] −→ R, f (x) := n−1 n

(−1)n · n + 2n+1
1 x − n1 for x ∈ [ n1 , n−1
1
], n ≥ 3 odd,
−1 n−1 n

satisfies f (1/n) = (−1)n · n for each n ∈ N and is an example of a continuous function


on the bounded half-open interval I :=]0, 1] with f (I) = R.

7.2.5 Inverse Functions, Existence of Roots, Exponential Function, Loga-


rithm

Theorem 7.60. Let I ⊆ R be an interval (of one of the 8 types listed in (4.8)). If
f : I −→ R is strictly increasing (resp. decreasing), then f has an inverse function f −1
defined on J := f (I), i.e. f −1 : J −→ I, and f −1 is continuous and strictly increasing
(resp. decreasing). If f is also continuous, then J must be an interval.

Proof. From Prop. 2.32(b), we know f : I −→ R is one-to-one. Then f : I −→ f (I)


is invertible and Prop. 2.32(c) shows f −1 is strictly monotone in the same sense as f .
We need to prove the continuity of f −1 . We assume f to be strictly increasing (the case
where f is strictly decreasing then follows by considering −f ). Let η ∈ J, ǫ > 0, and

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 92

ξ ∈ I with f (ξ) = η. If I = {ξ}, then J = {η}, and there is nothing to prove. It remains
to consider three cases:
(a) ξ = min I, i.e. ξ is the left endpoint of I (and ξ 6= max I),
(b) ξ = max I, i.e. ξ is the right endpoint of I (and ξ 6= min I),
(c) ξ is neither the min nor the max of I.
We carry out the proof for (c) and leave the (very similar) cases (a) and (b) as exercises.
In case (c), there are points ξ1 , ξ2 ∈ Bǫ (ξ) ∩ I such that

ξ − ǫ < ξ1 < ξ < ξ2 < ξ + ǫ. (7.42)

As f is strictly increasing, this implies

f (ξ1 ) < η < f (ξ2 ).

Choose δ > 0 such that

f (ξ1 ) < η − δ < η < η + δ < f (ξ2 ).

Then
 
f −1 str. inc. (7.42)
∀ f (ξ1 ) < y < f (ξ2 ) ⇒ ξ1 < f −1 (y) < ξ2 ⇒ f −1 (y) ∈ Bǫ (ξ) ,
y∈J∩Bδ (η)

proving the continuity of f −1 in η. That J must be an interval if f is continuous was


already shown in Th. 7.58. 

Remark and Definition 7.61 (Roots). We are now in a position to fulfill the promise
made in Def. and Rem. 5.6, i.e. to prove the existence of unique roots for nonnegative
real numbers: For each n ∈ N, the function f : R+ n
0 −→ R, f (x) := x , is continuous
and strictly increasing with J := f (R+ +
0 ) = R0 . Then Th. 7.60 implies the existence
of a continuous and strictly increasing inverse function f −1 : R+ +
0 −→ R0 . For each
√ 1
x ∈ R+ 0 , we call f
−1
(x) the nth root of x and write n x := x n := f −1 (x). Then
√ 1
( n x)n = (x n )n = x is immediate from the definition. Caveat: By definition, roots are
always nonnegative and they are only defined for nonnegative numbers (when studying
complex numbers and C-valued functions more deeply in the field of Complex Analysis,
one typically extends the notion of root, but we will not
√ pursue this√ route in this√class).
As anticipated in Def. and Rem. 5.6, one also writes x instead of x and calls x the
2

square root of x.

Remark and Definition √ 7.62. It turns out that 2 (and many other roots)
√ are not
rational numbers, i.e. 2 ∈ √ by contradiction: If 2 ∈ Q, then
/ Q. This is easily proved
there exist natural numbers m, n ∈ N such that 2 = m/n. Moreover, by canceling
possible factors of 2, we may assume at least one of the numbers m, n is odd. Now

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 93


2 = m/n implies m2 = 2n2 , i.e. m2 and, thus, m must be even. In consequence, there
exists p ∈ N such that m = 2p, implying 2n2 = m2 = 4p2 and n2 = 2p2 . Thus n2 and n
must also be even, in contradiction to m, n not both being even.
The elements of R \ Q are called irrational numbers. It turns out that most real num-
bers are irrational numbers – one can show that Q is countable, whereas R \ Q is not
countable (actually, every interval contains countably many rational and uncountably
many irrational numbers, see Appendix F, in particular, Th. F.1(c) and Cor. F.4).

Theorem 7.63 (Inequality Between the Arithmetic Mean and the Geometric Mean).
If n ∈ N and x1 , . . . , xn ∈ R+
0 , then

√ x1 + · · · + xn
n
x1 · · · xn ≤ , (7.43)
n
where the left-hand side is called the geometric mean and the right-hand side is called
the arithmetic mean of the numbers x1 , . . . , xn . Equality occurs if, and only if, x1 =
· · · = xn .

Proof. If at least one of the xj is 0, then (7.43) becomes the true statement 0 ≤ x1 +···+x
n
n

with strict equality if at least one xj > 0. If x1 = · · · = xn = x, then (7.43) also holds
since both sides are equal to x. Thus, for the remainder of the proof, we assume all
xj > 0 and not all xj are equal. First, we consider the special case, where x1 +···+x
n
n
= 1.
Since not all xj are equal, there exists k with xk 6= 1. We prove (7.43) by induction for
n ∈ {2, 3, . . . } in the form
n
! n
X Y
xj = n ∧ ∃ xk 6= 1 ⇒ xj < 1.
k∈{1,...,n}
j=1 j=1

Base Case (n = 2): Since x1 + x2 = 2, 0 < x1 , x2 and not both x1 and x2 are equal to
1, there is ǫ > 0 such that x1 = 1 + ǫ and x2 = 1 − ǫ, i.e. x1 x2 = 1 − ǫ2 < 1, which
Pn+1 the base case. Induction Step: We now have n ≥ 2 and 0 < x1 , . . . , xn+1
establishes
with j=1 xj = n + 1 plus the existence of k, l ∈ {1, . . . , n + 1} such that xk = 1 + α,
xl = 1 − β with α, β > 0. Then define y := xk + xl − 1 = 1 + α − β. One observes y > 0
(since β < 1) and
n+1
X n+1
X n+1
Y
ind. hyp.
y+ xj = −1 + xj = n ⇒ y xj ≤ 1
j=1, j=1 j=1,
j6=k,l j6=k,l

(we can not exclude equality as y and all the remaining xj might be equalQto 1). Since
xk xl = (1 + α)(1 − β) = 1 + α − β − αβ = y − αβ < y, we now infer n+1 j=1 xj < 1,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 94

concluding the induction proof. It remains to consider the case x1 +···+x n


n
= λ > 0, not
all xj equal. One estimates
r
√ x1 xn special case x1 + · · · + xn x1 + · · · + xn
n
x1 · · · xn = λ n ··· < λ = ,
λ λ λn n
completing the proof of the theorem. 
Corollary 7.64. For each a ∈ R+
0 \ {1}, n ∈ {2, 3, . . . }, p ∈ {1, . . . , n − 1}:


n p √ a−1
ap < 1 + (a − 1); p = 1 yields n
a<1+ . (7.44)
n n
Proof. The simple application
v
u n−p

n
u Y Th. 7.63 p a + n − p p
p
a = a ·
n
t p 1 < = 1 + (a − 1) (7.45)
j=1
n n

of Th. 7.63 establishes the case. 


Example 7.65. We use (7.43) to show

n
lim n=1: (7.46)
n→∞
√ √
First note 0 < x < 1 ⇒ 0 < xn < 1, i.e.
√ √ Qn−2
n
n > 1 for each ( n
n)n = n > 1. Now write n
as the product of n factors n = n n · k=1 1. Then, for n > 1,
v
Y Th. 7.63 2√n + n − 2
u
√ u√ √ n−2 2
n
n= tn
n n· 1 < <1+ √ . (7.47)
k=1
n n

It is an exercise to show
1
lim √ = 0. (7.48)
n→∞ n

Now this together with 1 ≤ n
n≤1+ √2 and the Sandwich Th. 7.16 proves (7.46).
n

Example 7.66 (Euler’s Number). We use Th. 7.63 to prove the limit
 n
1
e := lim 1 + (7.49)
n→∞ n

exists. It is known as Euler’s number. One can show it is an irrational number (see
Appendix H.1) and its first digits are e = 2.71828 . . . It is of exceptional importance for

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 95

analysis and mathematics in general, as it pops up in all kinds of different mathematical


contexts. From Th. 7.63, we obtain
  n+1
x n  x n x
∀ ∀ 1+ =1· 1+ < 1+ , (7.50)
n∈N x∈[−n,∞[, n n n+1
x6=0

where we have used that, on both sides of the inequality in (7.50), there are n + 1 factors
having the same sum, namely n + 1 + x; and the inequality in (7.43) is strict, unless
all factors are equal. We now apply (7.50) to the sequences (an )n∈N , (bn )n∈N , (cn )n∈N ,
where   n  n 
1 1
an := 1 + , bn := 1 − ,

 n n 

∀ 
  −1 !n+1  :
n+1  (7.51)
n∈N
c := b−1 = 1 1 
n n+1 1− = 1+
n+1 n
Applying (7.50) with x = 1 and x = −1, respectively, yields (an )n∈N and (bn )n∈N are
strictly increasing, and (cn )n∈N is strictly decreasing. On the other hand, an < cn holds
for each n ∈ N, showing (an )n∈N is bounded from above by c1 , and (cn )n∈N is bounded
from below by a1 . In particular, Th. 7.19 implies the convergence of both (an )n∈N and
(cn )n∈N . Moreover, limn→∞ cn = limn→∞ an (1 + 1/n) = e · 1 = e, which, together with
an < e < cn for each n ∈ N, can be used to compute e to an arbitrary precision.

Definition 7.67. Let A ⊆ R be a subset of the real numbers. Then A is called dense
in R if, and only if, every ǫ-neighborhood of every real number contains a point from A,
i.e. if, and only if,
∀ ∀ + A ∩ Bǫ (x) 6= ∅.
x∈R ǫ∈R

Theorem 7.68. (a) Q is dense in R.

(b) R \ Q is dense in R.

(c) For each x ∈ R, there exist sequences (rn )n∈N and (sn )n∈N in the rational numbers
Q such that x = limn→∞ rn = limn→∞ sn , (rn )n∈N is strictly increasing and (sn )n∈N
is strictly decreasing.

Proof. (a): Since each Bǫ (x) is an interval, it suffices to prove that every interval ]a, b[,
a < b, contains a rational number. If 0 ∈]a, b[, then there is nothing to prove. Suppose
0 < a < b and set δ := b − a > 0. Choose n ∈ N such that 1/n < δ and let
 
k k
q := max : k∈N ∧ <b .
n n

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 96

Then q ∈ Q and a < q < b. If a < b < 0, choose δ and n as above, but let
 
k k
q := min − : k ∈ N ∧ − > a .
n n
Then, once again, q ∈ Q and a < q < b.
(b): Analogous to (a), we show that every interval ]a, b[, a < b, contains an irrational
√ According to (a), we choose q ∈
number: √ Q∩]a, b[, δ := b − q > 0 and n ∈ N such
that
√ 2/n < δ. Then a < λ := q + 2/n < b and also λ ∈ R \ Q (otherwise,
2 = n(λ − q) ∈ Q).
(c): Using (a), for each n ∈ N, we choose rational numbers rn and sn such that
   
1 1 1 1
rn ∈ x − , x − , sn ∈ x + ,x + .
n n+1 n+1 n
Then, clearly, (rn )n∈N is strictly increasing, (sn )n∈N is strictly decreasing, and the Sand-
wich Th. 7.16 implies x = limn→∞ rn = limn→∞ sn . 
Definition and Remark 7.69 (Exponentiation). We have already used that Not. C.5
of the Appendix defines ax for (a, x) ∈ C×N0 and for (a, x) ∈ (C\{0})×Z (using G = C
and G = C \ {0}, respectively). We will now extend the definition to (a, x) ∈ R+ × R
(later, we will further extend the definition to (a, z) ∈ R+ × C). The present extension
to (a, x) ∈ R+ × R is accomplished in two steps – first, in (a), for rational x, then, in
(b), for irrational x.

(a) For rational x = k/n with k ∈ Z and n ∈ N, define


k √
n
ax := a n := ak . (7.52)

For this definition to make sense, we have to check it does not depend on the special
representation of x, i.e., we have to verify x = nk = nmkm
with k ∈ Z and m, n ∈ N
k km
implies a n = a nm . To this end, observe, using Rem. and Def. 7.61,
k √
n km √
nm
(a n )nm = ( ak )nm = akm and (a nm )nm = ( akm )nm = akm , (7.53)
k km
proving a n = a nm (here, as in Rem. and Def. 7.61, we used that λ 7→ λN is one-
to-one on R+0 for each N ∈ N). The exponentiation rules of Th. C.6 now extend to
rational exponents in a natural way, i.e., for each a, b > 0 and each x, y ∈ Q:

ax+y = ax ay , (7.54a)
(ax )y = ax y , (7.54b)
ax bx = (ab)x . (7.54c)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 97

For the proof, by possibly multiplying numerator and denominator by some natural
number, we can assume x = k/n and y = l/n with k, l ∈ Z and n ∈ N. Then
k+l Th. C.6(a) k l Th. C.6(c)
(ax+y )n = (a n )n = ak+l = ak al = (a n )n (a n )n = (ax ay )n ,

proving (7.54a);
Th. C.6(b)
 l
n k Th. C.6(b) k Th. C.6(b)
x y n2 x n
((a ) ) = ((a ) ) n = ((a n )l )n = ((a n )n )l = akl
kl 2 2
= (a n2 )n = (ax y )n ,

proving (7.54b);
Th. C.6(c) k k Th. C.6(c) k Th. C.6(b)
(ax bx )n = (a n )n (b n )n ak bk = (ab)k = (ab) n ·n = ((ab)x )n ,

proving (7.54c).
Moreover, we obtain the following monotonicity rules for each a, b ∈ R+ and each
x, y ∈ Q:
 
∀ a < b ⇒ ax < b x , (7.55a)
x>0
 
∀ a < b ⇒ ax > b x , (7.55b)
x<0
 
∀ x < y ⇒ ax < ay , (7.55c)
a>1
 
∀ x < y ⇒ ax > ay . (7.55d)
0<a<1

If x = k/n with k, n ∈ N and a < b, then a1/n < b1/n according to Rem. and
Def. 7.61, which, in turn, implies ax = (a1/n )k < (b1/n )k = bx , proving (7.55a); and
a−1 > b−1 implies a−x = (a−1 )x > (b−1 )x = b−x , proving (7.55b). If x < y, set q :=
y − x > 0. Then 1 < a and (7.55a) imply 1 = 1q < aq , i.e. ax < ax aq = ay , proving
(7.55c). Similarly, 0 < a < 1 and (7.55a) imply aq < 1q = 1, i.e. ay = ax aq < ax ,
proving (7.55d).
The following estimates will also come in handy: For a ∈ R+ and x, y ∈ Q:

a>1 ∧ x>0 ⇒ ax − 1 < x · ax+1 , (7.56)



∀ x, y ∈ [−m, m] ⇒ |ax − ay | ≤ L |x − y|,
m∈N
 (7.57)
m+1 m+1
where L := max{a , (1/a) } .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 98

For x ≥ 1, (7.56) is proved by ax < ax+1 < x · ax+1 + 1; for x < 1, write x = p/n
with p, n ∈ N and p < n, and apply (7.44) to obtain ax < 1 + x(a − 1) < 1 + xa <
1 + x · ax+1 . For the proof of (7.57), first consider a > 1. Moreover, by possibly
renaming x and y, we may assume x < y, i.e. z := y − x > 0. Thus, (7.56) holds
with x replaced by z. Multiplying the resulting inequality by ax yields

ax az − ax = ay − ax < z · ax az+1 = (y − x) ay+1 ≤ (y − x) am+1 ,

proving (7.57) for a > 1. For a = 1, it is clearly true, and for a < 1, it is a−1 > 1,
i.e.
|ax − ay | = |(a−1 )−x − (a−1 )−y | ≤ |y − x| (a−1 )m+1 ,
finishing the proof of (7.57).

(b) We now define ax for irrational x by letting

ax := lim aqn , where (qn )n∈N is a sequence in Q with lim qn = x. (7.58)


n→∞ n→∞

For this definition to make sense, we have to know such sequences (qn )n∈N exist,
which we do know from Th. 7.68(c). We also know from Th. 7.68(c) that there
exists an increasing sequence (qn )n∈N in Q converging to x, in particular, bounded
by x. Then, by (7.55c) and (7.55d), respectively, (aqn )n∈N is increasing for a > 1
and decreasing for 0 < a < 1. Moreover, the sequence is bounded from above by
aN with N ∈ N, N > x, for a > 1; and bounded from below by 0 for 0 < a < 1.
In both cases, Th. 7.19 implies convergence of the sequence to some limit that we
may call ax . However, we still need to verify that, for each sequence (rn )n∈N in Q
with limn→∞ rn = x, the sequence (arn )n∈N converges to the same limit ax in R. If
limn→∞ rn = x, then limn→∞ |qn − rn | = 0. Since (rn )n∈N and (qn )n∈N are bounded,
(7.57) implies
∃ + ∀ |aqn − arn | ≤ L |qn − rn |, (7.59)
L∈R n∈N

such that Prop. 7.11(a) implies limn→∞ |aqn − arn | = 0 and

lim arn = lim (arn − aqn + aqn ) = 0 + ax = ax , (7.60)


n→∞ n→∞

showing (7.58) does not depend on the chosen sequence.

Proposition 7.70. The exponentiation rules (7.54), the monotonicity rules (7.55), and
the estimates (7.56) and (7.57) remain valid if x, y ∈ Q is replaced by x, y ∈ R. More-
over, for each a > 0 and each sequence (xn )n∈N in R:

lim xn = x ∈ R ⇒ lim axn = ax . (7.61)


n→∞ n→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 99

Proof. Given x, y ∈ R, let (pn )n∈N and (qn )n∈N be sequences in Q such that limn→∞ pn =
x and limn→∞ qn = y.
We start by verifying (7.57). As we can assume (pn )n∈N and (qn )n∈N to be monotone,
we may also assume pn , qn ∈ [−m, m] for each n ∈ N. Then the rational case of (7.57)
implies
∀ |apn − aqn | ≤ L |pn − qn |,
n∈N

and Th. 7.13(c) establishes the case. Then (7.61) also follows, since

0 ≤ |axn − ax | ≤ L |xn − x| → 0.

We deal with (7.54) next. For each a, b > 0:


(7.54a)
ax+y = lim apn +qn = lim (apn aqn ) = ax ay ,
n→∞ n→∞
(7.54c)
ax bx = lim apn lim bpn = lim (apn bpn ) = lim (ab)pn = (ab)x ,
n→∞ n→∞ n→∞ n→∞
(7.54b)
∀ (ax )qk = lim (apn )qk = lim apn qk = ax qk ,
k∈N n→∞ n→∞
(7.57)
⇒ (ax )y = lim (ax )qn = lim ax qn = ax y ,
n→∞ n→∞

thereby proving (7.54).


Proceeding to (7.55c), let a > 1 and h > 0. If (qn )n∈N is a strictly increasing sequence
in Q+ with limn→∞ qn = h, then ah = limn→∞ aqn > aq1 > 1. Thus, if x < y, let
h := y − x > 0 to obtain ay = ax ah > ax , i.e. (7.55c). If 0 < a < 1 and x < y, then
(1/a)x < (1/a)y , yielding (7.55d). For (7.55a), consider x > 0 and 0 < a < b. Then
 x
b bx b
>1 ⇒ x
= > 1 ⇒ b x > ax ,
a a a

proving (7.55a). If x < 0 and 0 < a < b, then ax = (1/a)−x > (1/b)−x = bx , proving
(7.55b).
Finally, it remains to verify (7.56). For x ≥ 1, the proof for rational x still works for
irrational x. For 0 < x < 1, one uses the usual sequence (qn )n∈N in Q with limn→∞ qn = x
and obtains (recalling a > 1)
(7.44) 
ax = lim aqn ≤ lim 1 + qn (a − 1) = 1 + x(a − 1) < 1 + x · ax+1 ,
n→∞ n→∞

proving (7.56). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 100

Definition 7.71 (Exponential and Power Functions). (a) Each function of the form

f : R+ −→ R, f (x) := xα , α ∈ R, (7.62)

is called a power function. For α > 0, the power function is extended to x = 0 by


setting 0α := 0; for α ∈ Z, it is defined on R \ {0}; for α ∈ N0 even on R.

(b) Each function of the form

f : R −→ R+ , f (x) := ax , a > 0, (7.63)

is called a (general) exponential function. The case where a = e with e being Euler’s
number from (7.49) is of particular interest and importance. Most of the time, when
referring to an exponential function, one actually means x 7→ ex . It is also common
to write exp(x) instead of ex .

Theorem 7.72. (a) Every power function as defined in Def. 7.71(a) is continuous on
its respective domain. Moreover, for each α > 0, it is strictly increasing on [0, ∞[;
for each α < 0, it is strictly decreasing on ]0, ∞[.

(b) Every exponential function as defined in Def. 7.71(b) is continuous. Moreover, for
each a > 1, it is strictly increasing; for each 0 < a < 1, it is strictly decreasing.

Proof. (a): The monotonicity claims are provided by (7.55a) and (7.55b), respectively.
For each α ∈ N0 , the power function is a polynomial, for each α ∈ Z, a rational function,
i.e. continuity is provided by Ex. 7.40(b) and Ex. 7.40(c), respectively. For a general
α ∈ R, the continuity proof on R+ will be postponed to Ex. 7.76(a) below, where it can
be accomplished more easily. So it remains to show the continuity in x = 0 for α > 0.
However, if (xn )n∈N is a sequence in R+ with limn→∞ xn = 0 and k ∈ N with 1/k ≤ α,
1/k
then, at least for n sufficiently large such that xn ≤ 1, 0 < xαn ≤ xn by (7.55d). Then
1/k
the continuity of x 7→ x1/k implies limn→∞ xn = 0 and the Sandwich Th. 7.16 implies
limn→∞ xαn = 0, proving continuity in x = 0.
(b): Everything has already been proved – continuity is provided by (7.61), monotonicity
is provided by (7.55c) and (7.55d). 

Remark and Definition 7.73 (Logarithm). According to Th. 7.72(b), for each a ∈
R+ \ {1}, the exponential function f : R −→ R+ , f (x) := ax , is continuous and strictly
monotone with f (R) = R+ (verify that the image is all of R+ as an exercise). Then
Th. 7.60 implies the existence of a continuous and strictly monotone inverse function
f −1 : R+ −→ R. For each x ∈ R+ , we call f −1 (x) the logarithm of x to base a and write
loga x := f −1 (x). The most important special case is where the base is Euler’s number,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 101

a = e. This is called the natural logarithm. Bases a = 2 and a = 10 also carry special
names, binary and common logarithm, respectively. The notation is
ln x := loge x, lb x := log2 x, lg x := log10 x, (7.64)
however, the notation in the literature varies – one finds log used instead of ln, lb, and
lg; one also finds lg instead of lb. So you always need to verify what precisely is meant
by either notation.
Corollary 7.74. For each a ∈ R+ \ {1}, the logarithm function f : R+ −→ R, f (x) =
loga x is continuous. For a > 1, it is strictly increasing; for 0 < a < 1, it is strictly
decreasing. 
Theorem 7.75. One obtains the following logarithm rules:
∀ loga 1 = 0, (7.65a)
a∈R+ \{1}

∀ loga a = 1, (7.65b)
a∈R+ \{1}

∀ ∀ aloga x = x, (7.65c)
a∈R+ \{1} x∈R+

∀ ∀ loga ax = x, (7.65d)
a∈R+ \{1} x∈R

∀ ∀ loga (xy) = loga x + loga y, (7.65e)


a∈R+ \{1} x,y∈R+

∀ ∀ ∀ loga (xy ) = y loga x, (7.65f)


a∈R+ \{1} x∈R+ y∈R

∀ ∀ loga (x/y) = loga x − loga y, (7.65g)


a∈R+ \{1} x,y∈R+

√ 1
∀ ∀+ ∀ loga n
x= loga x, (7.65h)
+
a∈R \{1} x∈R n∈N n
∀ ∀ logb x = (logb a) loga x. (7.65i)
a,b∈R+ \{1} x∈R+

Proof. All the rules are easy consequences of the logarithm being defined as the inverse
function to f : R −→ R+ , f (x) := ax .
(7.65a): It is loga 1 = f −1 (1) = 0, as f (0) = a0 = 1.
(7.65b): It is loga a = f −1 (a) = 1, as f (1) = a1 = a.
(7.65c): It is aloga x = f (f −1 (x)) = x.
(7.65d): It is loga ax = f −1 (f (x)) = x.

(7.65e): It is loga (xy) = f −1 (xy) = f −1 f (loga x + loga y) = loga x + loga y, since
(7.65c)
f (loga x + loga y) = aloga x+loga y = aloga x aloga y = xy.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 102


(7.65f): It is loga (xy ) = f −1 (xy ) = f −1 f (y loga x) = y loga x, since

(7.65c)
f (y loga x) = ay loga x = (aloga x )y = xy .

(7.65g) is just a combination of (7.65e) and (7.65f): loga (x/y) = loga (xy −1 ) = loga x −
loga y.

(7.65h) is just a special case of (7.65f): loga n x = loga x1/n = n1 loga x.
(7.65i): One computes
(7.65f) (7.65c)
(logb a) loga x = logb aloga x = logb x.

Thus, we have verified all the rules and concluded the proof. 

Example 7.76. (a) For each α ∈ R, the power function

f : R+ −→ R, f (x) := xα = eα ln x , (7.66)

is continuous, which follows from Th. 7.41, since f = exp ◦(α ln), ln is continuous
by Cor. 7.74, and exp is continuous by Th. 7.72(b).

(b) As a consequence of Th. 7.41, each of the following functions f1 , f2 , f3 , where



f1 : R −→ R, f1 (x) := exp(λ + x2 ) ,
1
f2 : R −→ R, f2 (x) := ,
eαx

x5
f3 : R −→ R, f3 (x) := ,
(λ + |x|)α

is continuous for each α ∈ R and each λ ∈ R+ .

7.3 Series
7.3.1 Definition and Convergence

Series are a special type of sequences, namely sequences whose members arise from
summing up the members of another sequence. We have, on occasion, already encoun-
tered series, for example the harmonic series (sn )n∈N , whose members sn were defined
in (7.27). In the present section, we will study series more systematically.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 103

Definition 7.77. Given a sequence (an )n∈N in K (or, more generally, in any set A,
where an addition is defined), the sequence (sn )n∈N , where
n
X
∀ sn := aj , (7.67)
n∈N
j=1

is called an (infinite) series and is denoted by



X X
aj := aj := (sn )n∈N . (7.68)
j=1 j∈N

The anPare called the summands of the series, the sn its partial sums. Moreover, each
series ∞j=k aj with k ∈ N is called a remainder (series) of the series (sn )n∈N .

The example of the remainder series already shows that it is useful to allow countable
index sets other than N. Thus, if (aj )j∈I , where I is a countable index set and φ : N −→
I a bijective map, then define
X X∞
aj := aφ(j) (7.69)
j∈I j=1

(compare the definition in (3.19c) for finite


P sums). Note that the definition depends on
φ, which is suppressed in the notation j∈I aj .

For sequences in K, the notion of convergence is available, and, thus, it is also available
for series arising from real or complex sequences (as such series are, again, sequences in
K).

Definition 7.78. If (sn )n∈N is a series with the sn defined as in (7.67) and with sum-
mands aj ∈ K, then the series is called convergent with limit s ∈ K if, and only if,
limn→∞ sn = s in the sense of (7.1). In that case, one writes

X
aj = s (7.70)
j=1

and calls s the sum of thePseries. The series P is called divergent if, and only if, it is

not convergent. We write j=1 aj = ∞ (resp. ∞ j=1 aj = −∞) if, and only if, (sn )n∈N
diverges to ∞ (resp. −∞) in the sense of Def. 7.18.
P
Caveat 7.79. One has to use care as the symbol ∞ j=1 aj is used with two completely
different meanings. If it is used according to (7.68), then it means a sequence; if it is

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 104

used according to (7.70), then it means a real or complex number (or, possibly, ∞ or
−∞). It should always be clear from the Pcontext, if it means a sequence or a number.
For example, in the statement “the series ∞j=1 2
−j
is convergent”, it means a sequence;
P∞ −j
whereas in the statement “ j=1 2 = 1”, it means a number.
P
Example 7.80. (a) For each q ∈ C with |q| < 1, ∞ j
j=0 q is called a geometric series.
From (3.22b) (the reader is asked to go back and check that (3.22b) andPits proof,
indeed, remain valid for each q ∈ C), we obtain the partial sums sn = nj=0 q j =
1−q n+1
1−q
Since |q| < 1, we know limn→∞ q n+1 = 0 from Ex. 7.6. Thus, the series is
.
convergent with

X 1 − q n+1 1
∀ q j = lim sn = lim = . (7.71)
|q|<1 n→∞ n→∞ 1 − q 1−q
j=0

(b) In Ex. 7.30, we obtained the divergence of the harmonic series:



X 1
= ∞. (7.72)
k=1
k
P∞ P∞
Corollary 7.81. Let j=1 aj and j=1 bj be convergent series in C.

(a) Linearity:

X ∞
X ∞
X
∀ (λ aj + µ bj ) = λ aj + µ bj . (7.73)
λ,µ∈C
j=1 j=1 j=1

(b) Complex Conjugation:



X ∞
X
aj = aj . (7.74)
j=1 j=1

(c) Monotonicity:
  ∞
X ∞
X
∀ aj , b j ∈ R ∧ aj ≤ b j ⇒ aj ≤ bj . (7.75)
j∈N
j=1 j=1

P∞ P∞
(d) Each P
remainder seriesP j=n+1 a j , n ∈ N, converges, and, letting S := j=1 aj ,
sn := nj=1 aj , rn := ∞ a
j=n+1 j , one has
 
∀ S = s n + rn , lim an = lim rn = 0. (7.76)
n∈N n→∞ n→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 105

Proof. (a) follows from the first two identities of Th. 7.13(a), (b) is due to

X n
X n
X n
X ∞
X
Def. and Rem. 5.5(a) (7.11f)
aj = lim aj = lim aj = lim aj = aj ,
n→∞ n→∞ n→∞
j=1 j=1 j=1 j=1 j=1

(c) follows from Th. 7.13(c), and, for (d), one computes
lim an = lim (sn − sn−1 ) = S − S = 0,
n→∞ n→∞
Xk
∀ rn = lim aj = lim (sk − sn ) = S − sn ,
n∈N k→∞ k→∞
j=n+1

lim rn = lim (S − sn ) = S − S = 0,
n→∞ n→∞

completing the proof. 

7.3.2 Convergence Criteria


P∞ +
Pn
Corollary 7.82.PLet j=1 aj be series such that all aj ∈ R0 . If sn := j=1 aj are the

partial sums of j=1 aj , then
(
sup{sn : n ∈ N} if (sn )n∈N is bounded,
lim sn = (7.77)
n→∞ ∞ if (sn )n∈N is not bounded.

Proof. Since (sn )n∈N is increasing, (7.77) is a consequence of (7.21). 


P∞ P∞
Theorem 7.83. Let j=1 aj and j=1 bj be series in C such that |aj | ≤ |bj | holds for
each j ≥ k for some fixed k ∈ N.
P∞ P∞
(a) If j=1 |bj | is convergent, then aj is convergent as well, and, moreover,
j=1
∞ ∞
X X

aj ≤ |bj |. (7.78)

j=k j=k

P∞ P∞
(b) If j=1 aj is divergent, then j=1 |bj | is divergent as well.

Proof. Since
Pn(b) is merely the P contraposition of (a), it suffices toP
prove (a). ToP
this end,
let sn := j=1 aj and tn := j=1 |bj | be the partial sums of j=1 aj and ∞
n ∞
j=1 |bj |,
respectively. Since (tn )n∈N converges, it must be a Cauchy sequence by Th. 7.29. Thus,
∀ ∃ ∀ |tn − tm | = |bm+1 | + · · · + |bn | < ǫ
ǫ∈R+ N ∈N, n>m>N
N ≥k

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 106

and the triangle inequality for finite sums implies

|sn − sm | = |am+1 + · · · + an | ≤ |am+1 | + · · · + |an |


∀+ ∃ ∀
ǫ∈R N ∈N, n>m>N ≤ |bm+1 | + · · · + |bn | < ǫ,
N ≥k

showing (sn )n∈N is a Cauchy sequence as well, i.e. P convergent


Pby Th. 7.29. Since the
triangle inequality for finite sums also implies nj=k aj ≤ nj=k |bj | for each n ≥ k,
(7.78) is now a consequence of Th. 7.13(c). 
P
Definition 7.84. A series ∞ j=1 aj in R is called alternating if, and only if, its summands
alternate between positive and negative signs, i.e. if sgn(aj+1 ) = − sgn(aj ) 6= 0 for each
j ∈ N.
P
Theorem 7.85 (Leibniz Criterion). Let ∞ j=1 aj be an alternating series. If the sequence
(|an |)n∈N of absolute values is strictly decreasing and limn→∞ an = 0, then the series is
convergent and
X∞
∀ ∃ rn := aj = θn an+1 , (7.79)
n∈N 0<θn <1
j=n+1

that means the error made when approximating the limit by the partial sum sn has the
same sign as the first neglected summand an+1 , and its absolute value is less than |an+1 |.

Proof. We first consider the case where a1 > 0, i.e. where there exists a strictly de-
creasing sequence of positive numbers (bn )n∈N such that an = (−1)n+1 bn . As the bn are
strictly decreasing, we obtain bn − bn+1 > 0 for each n ∈ N, such that the sequences
(un )n∈N and (vn )n∈N , defined by
n
X
∀ un := s2n = (b2j−1 − b2j ) = (b1 − b2 ) + (b3 − b4 ) + · · · + (b2n−1 − b2n ),
n∈N
j=1
n
X
∀ vn := s2n+1 = b1 − (b2j − b2j+1 )
n∈N
j=1

= b1 − (b2 − b3 ) − (b4 − b5 ) − · · · − (b2n − b2n+1 ),

are strictly monotone, namely (un )n∈N strictly increasing and (vn )n∈N strictly decreasing.
Since, 0 < un < un + b2n+1 = vn < b1 for each n ∈ N, both sequences (un )n∈N and
(vn )n∈N are also bounded, and, thus, convergent by Th. 7.19, i.e. U := limn→∞ un ∈ R
and V := limn→∞ vn ∈ R. Since

V − U = lim (vn − un ) = lim (s2n+1 − s2n ) = lim a2n+1 = 0,


n→∞ n→∞ n→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 107

we obtain U = V P and limn→∞ sn = U and 0 < U < b1 = a1 . In particular, there is


θ ∈]0, 1[ satisfying ∞ j=1 aj = θ a1 .
P∞ P∞
In
P∞ the case a 1 < 0, the above proof yields convergence of − j=1 a j = j=1 (−aj ) with

Pj=1 (−a j ) = θ (−a 1 ) for a suitable θ ∈]0, 1[. However, this then yields, as before,

j=1 aj = θ a1 .
P
Applying the above result to each remainder series ∞ j=n+1 aj , n ∈ N, completes the
proof of (7.79) and the theorem. 
Example 7.86. (a) Each of the following alternating series clearly converges, as the
Leibniz criterion of Th. 7.85 clearly applies in each case:

X (−1)j+1 1 1
=1− + − +..., (7.80a)
j=1
j 2 3

X (−1)j+1 1 1
=1− + − +..., (7.80b)
j=1
2j − 1 3 5

X (−1)j+1 1 1 1
= − + − +... (7.80c)
j=1
ln(j + 1) ln 2 ln 3 ln 4

(b) To see that Th. 7.85P is false without its monotonicity requirement, take any diver-
gent series with ∞ j=1 aj = ∞, 0 <P aj , limj→∞ aj = 0 (for example the harmonic
series), any convergent series with ∞ +
j=1 cj = s ∈ R and 0 < cj (for example any
geometric series with 0 < q < 1), and define
(
a(n+1)/2 for n odd,
dn :=
−cn/2 for n even.
P
an exercise to show that ∞
It is P j=1 dj is an alternating series with limn→∞ dn = 0

and j=1 dj = ∞.
P∞
Definition
P∞ 7.87. The series j=1 aj in C is said to be absolutely convergent if, and
only if, j=1 |aj | is convergent.
P∞
Corollary 7.88. Every absolutely convergent series j=1 aj is also convergent and
satisfies the triangle inequality for infinite series:
∞ ∞
X X

aj ≤ |aj |. (7.81)

j=1 j=1

Proof. The corollary is given by the special case aj = bj for each j ∈ N of Th. 7.83(a). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 108

P∞
Theorem 7.89. We consider the series j=1 aj in C.
P
(a) If ∞ +
P∞cj is a convergent series such that cj ∈ R0 and |aj | ≤ cj for each j ∈ N,
j=1
then j=1 aj is absolutely convergent.

(b) Root Test:


 
p
n
∃ ( |an | ≤ q < 1 for almost all n ∈ N)
0<q<1

X
⇒ aj is absolutely convergent, (7.82a)
j=1
n o ∞
X
p
n
# n∈N: |an | ≥ 1 = ∞ ⇒ aj is divergent. (7.82b)
j=1

(c) Ratio Test: If all an 6= 0, then


  
an+1
∃ an ≤ q < 1 for almost all n ∈ N

0<q<1

X ∞
⇒ aj is absolutely convergent, (7.83a)
j=1

an+1 X
an ≥ 1 for almost all n ∈ N ⇒ aj is divergent. (7.83b)

j=1

Proof. (a) is just another special case of Th. 7.83(a).


p
(b): If there is q ∈]0, 1[ and N ∈ N such that n |an | ≤ q for each n > N , i.e. |an | ≤ q n
P PN
for each n > N , then, by (7.71), ∞ 1
j=1 |aj | is bounded by 1−q + j=1 |aj | and, thus,
p
convergent. If |an | ≥ 1 for infinitely many n ∈ N, then |an | ≥ 1 for infinitely
n
P∞many
n ∈ N, showing that (an )n∈N does not converge to 0, proving the divergence of j=1 aj .

(c): If there is q ∈]0, 1[ and N ∈ N such that an+1

an
≤ q for each n > N , then,
letting C := |aN +1 |, an induction
P PNshows |aN +1+k | ≤ Cq k for each k ∈ N, i.e., by (7.71),
∞ C +1
j=1 |aj | is bounded by 1−q + j=1 |aj | and, thus, convergent. If there is N ∈ N such
an+1
that an ≥ 1 for each n > N , then |an | ≥ |aN +1 | > 0 for each n > N , showing (an )n∈N
P
does not converge to 0 and proving the divergence of ∞ j=1 aj . 
p
Caveat 7.90. In (7.82a), it does not suffice to have n |an | < 1 to conclude convergence,
and, likewise, | an+1an
| < 1 does not suffice in (7.83a): As a counterexample, consider

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 109

p
the harmonic series, which does not converge, but n
1/n < 1 for each n ≥ 2 and
1/(n+1) n
1/n
= n+1 < 1 for each n ∈ N.
P
Example 7.91. (a) For each z ∈ C with |z| < 1√and each p ∈ N0 , the series ∞ p n
n=1 n z
is absolutely convergent: We have limn→∞ n p
n = 1 as a consequence of Ex. 7.65.
p p
This implies limn→∞ |an | = limn→∞ n |z|n = |z| < 1. Thus, the root test of
n n p

(7.82a) applies and proves convergence of the series.


P
(b) Let z ∈ C. The series ∞ z n n!
n=1 nn is absolutely convergent for |z| < e and divergent
for |z| ≥ e, where e is Euler’s number from (7.49). We have, for each n ∈ N,

an+1 |z| (n + 1) nn |z| |z|
= =  n → for n → ∞. (7.84)
an (n + 1)n+1 1 + n1 e

Thus, the ratio test of (7.83a) applies and proves absolute convergence of the series
for |z| < e. For |z|
n > e, (7.83b) applies and proves divergence. Since, according to
Ex. 7.66, 1 + n1 < e for each n ∈ N, (7.83b) applies to prove divergence also for
|z| = e.

7.3.3 Absolute Convergence and Rearrangements

In general, one has to use care when dealing with infinite series, as convergence properties
and even the limit in case of convergence can depend on the order of the summands (in
obvious contrast to the situation of finite sums). For real series that are convergent, but
not absolutely convergent, one has the striking Riemann rearrangement Th. 7.93, that
states one can choose an arbitrary number S ∈ R∪{−∞, ∞} and reorder the summands
such that the new series converges to S (actually, Th. 7.93 says even more, namely that
one can prescribe an entire interval of cluster points for the rearranged series). However,
the situation is better for absolutely convergent series. In Th. 7.95, we will see that the
sum of absolutely convergent series does not depend on the order of the summands.
P
Proposition 7.92. Let ∞ j=1 aj be a series in R. Defining
 
∀ a+
j := max{aj , 0}, a−
j := max{−a j , 0} , (7.85)
j∈N

the following assertions (a) and (b) hold:


P∞ P∞ P∞
(a) aj is absolutely convergent if, and only if, both series
j=1 j=1 a+
j and j=1 a−
j
are convergent.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 110

P∞
(b) If j=1 aj is convergent, but not absolutely convergent, then

X ∞
X
a+
j = a−
j = ∞. (7.86)
j=1 j=1

Proof. The key observation is that (7.85) implies, for each j ∈ N,



a+
j + aj = |aj |, (7.87a)

a+
j − aj = aj , (7.87b)

0≤ a+
j , aj ≤ |aj |. (7.87c)
P∞ P∞
(a): If j=1 a+
j and j=1 a−
j are convergent, then


X ∞
X ∞
X
(7.87a),(7.73)
|aj | = a+
j + a−
j , (7.88)
j=1 j=1 j=1
P P
and, in particular, ∞j=1 aj is absolutely convergent. Conversely, if ∞ j=1 aj is absolutely
P∞ + P∞ −
convergent, then j=1 aj and j=1 aj are convergent by (7.87c) and Th. 7.83(a).
P P∞ + P∞ −
(b): If ∞ j=1 aj and j=1
P∞ a j are convergent, then (7.87b) implies that j=1
P∞ aj is also
convergent and, thus, j=1 aj absolutely convergent by (a). Likewise, if j=1 aj and
P∞ − P∞ +
j=1 aP
j are convergent, then (7.87b) implies that j=1 aj is
Palso convergent and, once
∞ ∞
again, j=1 aj absolutely convergent by (a). Therefore, if j=1 aj is convergent, but
not absolutely convergent, then (7.86) must hold by (7.77). 
Theorem
P 7.93 (Riemann Rearrangement
P∞ Theorem a.k.a. Riemann Series Theorem).
Let ∞ a
j=1 j be a series in R. If a
j=1 j is convergent, but not absolutely
P∞ convergent,
then, given x, y ∈ R ∪ {−∞, ∞} with x ≤ y, there exists P a rearrangement j=1 bj of the
series (i.e. a reordering (bj )j∈N of (aj )j∈N ) such that ∞ j=1 bj has precisely all elements of
[x, y] as cluster points (where we call −∞ (resp. ∞) a cluster point of the real sequence
(tn )n∈N if, and only if, #{n ∈ N : tn < −N } = ∞ (resp. #{n ∈ N : tn > N } = ∞) for
each N ∈ N). In particular,P choosing S := x = y ∈ R ∪ {−∞, ∞}, one can prescribe an
arbitrary limit S such that ∞ j=1 bj = S.

Sketch of Proof. Here we just give a sketch of the proof to convey its fairly simple idea;
a detailed proof is provided in Appendix E.1. According to Prop. 7.92(b), (7.86) must

hold, where the a+j and aj are as defined in (7.85). Thus, we can define
 

 −k for x = −∞, −k for y = −∞,

∀ xk := x for x ∈ R, yk := y for y ∈ R,
k∈N 
 

k for x = ∞, k for y = ∞,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 111

and, noting xk ≤ yk for almost all k ∈ N, alternate between adding summands a+ j until

the partial sum exceeds yk and subtracting summands aj until the partial sum falls
below xk . If k is sufficiently large such that xk ≤ yk , then, at each switching point (from
adding to subtracting or vice versa), the absolute value of the difference between the
last partial sum and xk or yk , respectively, is less than the value of the last contributing
nonzero summand. Since

lim a+j = lim aj = 0,
j→∞ j→∞

the partial sums corresponding to the switching points converge to the respective end-
points x or y, respectively, and precisely all points between x and y are cluster points. 

We will now study the more benign situation of absolutely convergent series.
P P∞
Theorem 7.94. Let ∞ j=1 aj and j=1 bj be
Pseries in C such that (bn )n∈N is a reordering

of (an )n∈N inPthe sense P
P of Def. 7.21. If j=1 aj is absolutely convergent, then so is
∞ ∞ ∞
b
j=1 j and a
j=1 j = b
j=1 j .
Pn Pn Pn
Proof. Let sn := j=1 aj , s̃n := j=1 |aj |, and tn := j=1 bj denote the respective
partial sums. We will show that limn→∞ (sn − tn ) = 0. Given ǫ > 0, since (s̃n )n∈N is a
Cauchy sequence by Th. 7.29, there exists N ∈ N, such that
∀ |s̃n − s̃m | = |am+1 | + · · · + |an | < ǫ.
n>m>N

Since (bn )n∈N is a reordering of (an )n∈N , there exists a bijective map φ : N −→ N such
that bn = aφ(n) for each n ∈ N. Since φ is bijective, there exists M ∈ N such that
{1, 2, . . . , N + 1} ⊆ φ{1, 2, . . . , M }. Then n > M implies φ(n) > N + 1, and
∀ ∃ |sn − tn | ≤ |aN +2 | + · · · + |aN +k | < ǫ,
n>M k∈N

since all aj with j ≤ N +1 occur in both sn and tn and cancel in sn −tn (i.e. all aj that do
not cancel must have an index j > N + 1). So we have shown that limn→∞ (sn − tn ) = 0,
which, in turn, implies

X ∞
X ∞
X
bj = lim tn = lim (tn − sn + sn ) = 0 + aj = aj .
n→∞ n→∞
j=1 j=1 j=1
Pn P∞ P∞
ApplyingPthis to s̃n := j=1 |aj | yields j=1 |bj | = j=1 |aj |, proving absolute conver-
gence of ∞j=1 bj . 
Theorem 7.95. Let I be an arbitrary infinite countable index set and let

I= In (7.89)
n∈N

be a disjoint decomposition of I into (empty, finite, or infinite) countable index sets In .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 112

P
(a) If the series j∈I aj (cf. (7.69)) is absolutely convergent, then

X ∞ X
X
aj = aα . (7.90)
j∈I n=1 α∈In

(b) The following statements are equivalent:


P
(i) j∈I aj is absolutely convergent.
P
(ii) There exists a constant C ∈ R+
0 such that j∈J |aj | ≤ C for each finite subset
J of I.
P∞ P
(iii) n=1 α∈In |aα | < ∞.

P
Proof. (a):PFirst notePthat Th. 7.94 implies that, for absolutely convergent j∈I aj ,

the limit j∈I aj = j=1 aφ(j) does not depend on the bijective map φ : N −→ I:

P ∞ P∞ map ψ : N −→ I, (aψ(j) )j∈N is a reordering of (aφ(j) )j∈N and, thus,


For each bijective
a
j=1 ψ(j) = j=1 aφ(j) .
P
Analogously, the sums α∈In aα do not depend on the order of the indices in In .
P
Claim 1. If M ⊆ I, then S(I) = S(M ) + S(I \ M ), where S(J) := j∈J aj for each
J ⊆ I.

Proof. If M = ∅, then there is nothing to prove. If #M = n ∈ N, then let φ1 :


{1, . . . , n} −→ M and φ2 : {n + 1, n + 2, . . . } −→ I \ M be bijective maps. Then
(
φ1 (j) for j ≤ n,
φ : N −→ I, φ(j) :=
φ2 (j) for j > n,

is a bijective map. Moreover,



X n
X ∞
X
(7.76)
S(I) = aφ(j) = aφ(j) + aφ(j) = S(M ) + S(I \ M ),
j=1 j=1 j=n+1

establishing the case.


If #M = #(I \ M ) = #N, then let φ1 : {1, 3, 5, . . . } −→ M and φ2 : {2, 4, 6, . . . } −→
I \ M be bijective maps. Then
(
φ1 (j) for j odd,
φ : N −→ I, φ(j) :=
φ2 (j) for j even,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 113

is a bijective map. Define,


( (
aφ(j) for j odd, aφ(j) for j even,
bφ(j) := cφ(j) :=
0 for j even, 0 for j odd.

One then obtains



X ∞
X ∞
X
(7.73)
S(I) = aφ(j) = bφ(j) + cφ(j) = S(M ) + S(I \ M ),
j=1 j=1 j=1

establishing the case. N


Ṡk
Claim 2. If I = n=1 Mn with k ∈ N is a decomposition of I, then, using the notation
P
introduced in Cl. 1, S(I) = kn=1 S(Mn ).

Proof. Follows by an induction from Cl. 1. N

Coming back to (7.89), Cl. 2 implies


k
!
[
∀ S(I) = S(I1 ) + S(I2 ) + · · · + S(Ik ) + S(Mk ), where Mk := I \ Ij .
k∈N
j=1

(7.90), fix a bijective φ : N −→ I, and let ǫ > 0. Due to Cor.


To prove the equality in P
7.81(d), the sums rn := ∞ j=n+1 |aφ(j) | of the remainder series converge to 0, i.e. there
exists N ∈ N such that rn < ǫ for each n > N . More generally, for each (empty, finite,
or infinite) subset J ⊆ {N + 2, N + 3, . . . },

X ∞
X
|aφ(j) | ≤ |aφ(j) | = rN +1 < ǫ.
j∈J j=N +2

Next, we choose M ∈ N sufficiently large such that {φ(1), . . . , φ(N + 1)} ⊆ I1 ∪ · · · ∪ IM .


Then, for each k > M ,

X (7.81) X X

|S(Mk )| = aj ≤ |aj | ≤ |aφ(j) | = rN +1 < ǫ,

j∈Mk j∈Mk j=N +2

proving
X k
X ∞ X
X
aj = S(I) = lim S(In ) = aα ,
k→∞
j∈I n=1 n=1 α∈In

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 114

which is (7.90).
P
(b): (i) implies (ii) with C := j∈I |aj | using Cl. 1 (with aj replaced by |aj |). (i) implies
P (with aj replaced by |aj |). (ii) implies (i) via (7.77), as C is an upper
(iii) using (7.90)
bound for (P nj=1 |a Pφ(j) |)n∈N for each bijection φ : N −→ I. Finally, (iii) implies (ii)

with C := n=1 α∈In |aα |, since, given a finite J ⊆ I, there exists k ∈ N such that
J ⊆ I1 ∪ · · · ∪ Ik , i.e.

X k
X X ∞ X
X
|aj | ≤ |aα | ≤ |aα | = C,
j∈J n=1 α∈In n=1 α∈In

thereby completing the proof. 

Example 7.96. We apply Th. 7.95 to so-called double series, i.e. to series with index
set I := N × N. The following notation is common:

X X
amn := a(m,n) , (7.91)
m,n=1 (m,n)∈N×N

where one writes amn (also am,n ) instead of a(m,n) . Recall from Th. 3.16 that N × N is
countable. In general, the convergence properties of the double series and, if it exists,
the value of the sum, will depend on the chosen bijection φ : N −→ N × N.
However, we will now assume our double series to be absolutely convergent. Then Th.
7.94 guarantees the sum does not depend on the chosen bijection and we can apply Th.
7.95. Applying Th. 7.95 to the decompositions
˙
[
N×N= {(m, n) : n ∈ N}, (7.92a)
m∈N
˙
[
N×N= {(m, n) : m ∈ N}, (7.92b)
n∈N
˙
[
N×N= {(m, n) ∈ N × N : m + n = k}, (7.92c)
k∈N

yields
X ∞ X
X ∞ ∞ X
X ∞
(7.92a) (7.92b)
a(m,n) = amn = amn
(m,n)∈N×N m=1 n=1 n=1 m=1

X X ∞ X
X k−1
(7.92c)
= amn := am,k−m . (7.93)
k=2 m+n=k k=2 m=1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 115

Theorem 7.97. ItP is possible to compute


P∞ the product of two absolutely convergent (real

or complex) series m=1 am and m=1 bm as a double series:

! ∞ ! ∞ ∞ Xk−1 ∞
X X X X X
am bm = am b n = am bk−m = ck ,
m=1 m=1 m,n=1 k=2 m=1 k=2
(7.94)
k−1
X
where ck := am bk−m = a1 bk−1 + a2 bk−2 + · · · + ak−1 b1 .
m=1

This form of computing the product is known as a Cauchy product.


P∞
Proof.
P∞ We first showPthat m,n=1 am bn is absolutely convergent: By letting A :=

m=1 |am | and B := m=1 |bm |, we obtain

∞ X
X ∞ ∞
X 
|am bn | = |am | B = AB < ∞,
m=1 n=1 m=1
P
i.e. ∞m,n=1 am bn is absolutely convergent according to Th. 7.95(b)(iii). Now the second
equality in (7.94) is just the third equality in (7.93), and the first equality in (7.94) also
follows from (7.93):
∞ ∞ X ∞ ∞ ∞ ∞
! ∞ !
X X X X X X
am b n = am b n = am bn = am bm ,
m,n=1 m=1 n=1 m=1 n=1 m=1 m=1

completing the proof. 

Theorem 7.97 will be useful in Sec. 8.2 below.

7.3.4 b-Adic Representations of Real Numbers

We are mostly used to representing real numbers in the decimal system. For example,
we write ∞
395 X
x= 2 1 0
= 131.6 = 1 · 10 + 3 · 10 + 1 · 10 + 6 · 10−n , (7.95a)
3 n=1

where ∞  
X
−n (7.71) 1 1 2
6 · 10 = 6· 1 −1 =6· = .
n=1
1 − 10 9 3

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
7 LIMITS AND CONVERGENCE OF REAL AND COMPLEX NUMBERS 116

The decimal system represents real numbers as, in general, infinite series of decimal
fractions. Digital computers represent numbers in the dual system, using base 2 instead
of 10. For example, the number from (7.95a) has the dual representation

X
x = 10000011.10 = 2 + 2 + 2 + 7 1 0
2−(2n+1) , (7.95b)
n=0

where it is an exercise to verify



X 2
2−(2n+1) = .
n=0
3

Representations with base 16 (hexadecimal) and 8 (octal) are also of importance when
working with digital computers. More generally, each natural number b ≥ 2 can be used
as a base.
Definition 7.98. Let b ≥ 2 be a natural number.

(a) Given an integer N ∈ Z and a sequence (dN , dN −1 , dN −2 , . . . ) in {0, . . . , b − 1}, the


series ∞
X
dN −ν bN −ν (7.96)
ν=0

is called a b-adic series. The number b is called the base or the radix, and the
numbers dν are called digits.

(b) If x ∈ R+0 is the sum of the b-adic series given by (7.96), than one calls the b-adic
series a b-adic representation or a b-adic expansion of x.
Theorem 7.99. Given a natural number b ≥ 2 and a nonnegative real number x ∈
R+0 , there exists a b-adic series representing x, i.e. there is N ∈ Z and a sequence
(dN , dN −1 , dN −2 , . . . ) in {0, . . . , b − 1} such that

X
x= dN −ν bN −ν . (7.97)
ν=0

If one introduces the additional requirement that 0 6= dN , then each x > 0 has either a
unique b-adic representation or precisely two b-adic representations. More precisely, for
0 6= dN and x > 0, the following statements are equivalent:

(i) The b-adic representation of x is not unique.

(ii) There are precisely two b-adic representations of x.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 117

(iii) There exists a b-adic representation of x such that dn = 0 for each n ≤ n0 for
some n0 < N .

(iv) There exists a b-adic representation of x such that dn = b − 1 for each n ≤ n0 for
some n0 ≤ N .

Proof. The proof is a bit lengthy and is provided in Appendix E.2. 

Example 7.100. Every natural number has precisely two decimal (i.e. 10-adic) repre-
sentations. For instance,
∞  
X
−n (7.71) 1 1
2 = 2.0 = 1.9 = 1 + 9 · 10 = 1+9· 1 −1 =1+9· , (7.98)
n=1
1 − 10 9

and analogously for all other natural numbers.

8 Convergence of K-Valued Functions

8.1 Pointwise and Uniform Convergence


So far we have studied the convergence of sequences in K. We will now also need to
study the convergence of sequences (fn )n∈N , where each member fn of the sequence is
a function fn : M −→ K, M ⊆ C. Here, for the first time, we encounter the situation
that there exist different useful notions of convergence for such sequences.

Definition 8.1. Let (fn )n∈N be a sequence of functions, fn : M −→ K, ∅ 6= M ⊆ C.

(a) We say (fn )n∈N converges pointwise to f : M −→ K if, and only if, limn→∞ fn (z) =
f (z) for each z ∈ M , i.e. if, and only if,

∀ ∀ ∃ ∀ |fn (z) − f (z)| < ǫ. (8.1)


z∈M ǫ∈R+ N ∈N n>N

So, in general, N in (8.1) depends on both z and ǫ.

(b) We say (fn )n∈N converges uniformly to f : M −→ K if, and only if,

∀ ∃ ∀ ∀ |fn (z) − f (z)| < ǫ. (8.2)


ǫ∈R+ N ∈N n>N z∈M

In (8.2), N is still allowed to depend on ǫ, but, in contrast to the situation of (8.1),


not on z – in that sense, the convergence is uniform in z.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 118

Remark 8.2. It is immediate from Def. 8.1(a),(b) that uniform convergence implies
pointwise convergence, but Ex. 8.3(b) below will show the converse is not true.

Example 8.3. (a) Let ∅ =


6 M ⊆ C (for example M = [0, 1] or M = B1 (0)), and
fn : M −→ K, fn (z) = 1/n for each n ∈ N. Then, clearly, (fn )n∈N converges
uniformly to f ≡ 0.

(b) The sequence (fn )n∈N , where fn : [0, 1] −→ R, fn (x) := xn , converges pointwise,
but not uniformly, to
(
0 for 0 ≤ x < 1,
f : [0, 1] −→ R, f (x) := (8.3)
1 for x = 1 :

For x = 1, limn→∞ xn = limn→∞ 1 = 1, and, for 0 ≤ x < 1, limn→∞ xn = 0 by Ex.


7.6. To see that the convergence is not uniform, consider ǫ := 21 . Then, for every
n ∈ N, according to the intermediate value Th. 7.57, there exists ξn ∈]0, 1[ such
that fn (ξn ) = ξnn = 12 , i.e.

1
∀ |fn (ξn ) − f (ξn )| = ξnn = = ǫ, (8.4)
n∈N 2
proving the convergence is not uniform.

Theorem 8.4. Let (fn )n∈N be a sequence of functions, fn : M −→ K, ∅ 6= M ⊆ C. If


(fn )n∈N converges uniformly to f : M −→ K and all fn are continuous at ζ ∈ M , then
f is also continuous at ζ. In particular, if each fn is continuous, then so is f (uniform
limits of continuous functions are continuous).

Proof. Let ǫ > 0. Due to the uniform convergence of (fn )n∈N ,


ǫ
∃ ∀ |fm (z) − f (z)| < . (8.5)
m∈N z∈M 3
Due to the continuity of fm in ζ,
ǫ
∃ ∀ |fm (z) − fm (ζ)| < . (8.6)
δ>0 z∈M ∩Bδ (ζ) 3

Thus,
ǫ
∀ |f (z)−f (ζ)| ≤ |f (z)−fm (z)|+|fm (z)−fm (ζ)|+|fm (ζ)−f (ζ)| < 3· = ǫ, (8.7)
z∈M ∩Bδ (ζ) 3

proving continuity of f in ζ. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 119

8.2 Power Series


Definition 8.5. (a) In Def. 7.77, it was mentioned that series can be formed from each
sequence in a set A, where an addition is defined. Letting ∅ 6= M ⊆ C, we now
consider A := F(M, K), i.e. the set of functions from M into K. Then the addition
on A is defined according to (6.1a) and, given a sequence of functions (fn )n∈N in A,
the series ∞
X
fj := (sn )n∈N (8.8)
j=1
Pn
is defined as the sequence of partial sums sn := j=1 fj .

(b) Given a sequence of functions (fn )n∈N0 , where fn : K −→ K, fn (z) = an z n with


an ∈ K, the series
X∞ ∞
X
aj z j := fj (8.9)
j=0 j=0

is called a power series


P∞ and the aj are called the coefficients of the power series.
j
Note: The notation j=0 aj z introduced in (8.9) is very common, but not entirely
correct, since one writes aj z j = fj (z) for the summands, even though one actually
means fj . Moreover, one uses the same notation if one actually
P P∞ does mean the series
∞ j
j=0 fj (z) in K, so one has to see from the context if j=0 aj z means a series of
K-valued functions or a series of numbers.
P
Definition 8.6. P Consider a series of K-valued functions ∞ j=1 fj as in Def. 8.5(a), in
particular, sn := nj=1 fj for each n ∈ N.

(a) The series converges pointwise to f : M −→ K if, and only if, it (i.e. (sn )n∈N )
converges pointwise in the sense of Def. 8.1(a). In that case, we use the notation

X
f= fj . (8.10)
j=1

If (8.10) holds, then the series is sometimes called a series expansion of f , in par-
ticular, a power series expansion if the series happens to be a power series.
P
Analogous to the situation of series in K, the notation ∞ j=1 fj is also used with
two different meanings – it can mean the sequence of partial sums as in (8.8) or, in
the case of convergent series, the limit function as in (8.10) (cf. Caveat 7.79).

(b) The series converges uniformly to f : M −→ K if, and only if, it converges uniformly
in the sense of Def. 8.1(b).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 120

P∞
Corollary 8.7. Consider a function series j=1 fj with fj : M −→ K, ∅ 6= M ⊆ C.

(a) The series converges uniformly to some f : PM −→ K if, and only if, for each
n ∈ N and each z ∈ M , the remainder series ∞j=n+1 fj (z) in K converges to some
rn (z) ∈ K such that
∀ + ∃ ∀ ∀ |rn (z)| < ǫ. (8.11)
ǫ∈R N ∈N n>N z∈M
P∞
(b) If j=1 aj is a convergent series in R+
0 , then the condition

∀ ∀ |fj (z)| ≤ aj (8.12)


z∈M j∈N
P∞
implies uniform convergence of j=1 fj .
(c) If each fj is continuous in ζ ∈ M and the series converges uniformly to f : M −→
K, then f is continuous in ζ. In particular, if each fj is continuous, then f is
continuous.
P P∞
Proof. (a): If ∞ j=1 fj converges uniformly to f , then f (z) = j=1 fj (z) holds for each
z ∈ M, P rn (z) = f (z) − sn (z) for each n ∈ N, z ∈ M according to (7.76), where
sn (z) := nj=1 fj (z). Then (8.11) is just (8.2), where the sn now play the role of the fn
series converge for each z ∈ M , then we can define
in (8.2). Conversely, if the remainder P
f : M −→ K, f (z) := f1 (z)+r1 (z) = ∞ j=1 fj (z). Then, once again, rn (z) = f (z)−sn (z)
for each n ∈ N, z ∈ M , and (8.11) is just (8.2), yielding the uniform convergence of the
series.
P
(b): First, (8.12) implies each remainder series ∞ j=n+1 fj (z) converges absolutely. Thus,
with rn (z) as in (a),
∞ ∞
(7.81) X X
∀ |rn (z)| ≤ |fj (z)| ≤ aj → 0 for n → ∞,
z∈M
j=n+1 j=n+1

such that (a) yields uniform convergence.


(c) is immediate from Th. 8.4. 
P
RemarkP8.8. Given a function series ∞ j=1 fj with fj : M −→ K, ∅ 6= M ⊆ C; for each

z ∈PM , j=1 fj (z) constitutes a series in K. Typically, one will only have convergence
of ∞ j=1 fj (z) in K on a subset C ⊆ M . The series then converges pointwise in the
sense
P of Def. 8.6(a) if all fj are restricted to C. It can be very difficult to determine
if ∞ j=1 fj (z) converges or diverges for some z ∈ M , and such investigations are often
of particular interest in the context of function series. Even for power series, studying
convergence can still be difficult, but the availability of the following Th. 8.9 does help
to (at least partially) settle the question in many cases.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 121

P
Theorem 8.9. For each power series ∞ j
j=0 aj z , aj ∈ K, there exists a number r ∈
[0, ∞] := R+
0 ∪ {∞}, called the radius of convergence of the power series, such that

  ∞
X
z ∈ K ∧ |z| < r ⇒ aj z j converges absolutely in K, (8.13a)
j=0
  X∞
z ∈ K ∧ |z| > r ⇒ aj z j diverges in K (8.13b)
j=0

P∞ r = j ∞, (8.13a) claims absolute convergence for each z ∈ K). In particular,


(for
j=0 aj z converges pointwise in the sense of Def. 8.6(a) for each z ∈ Br (0) (cf. Def.
7.7(a)). Moreover,

!
X
j converges uniformly on B r0 (0) (cf. Ex. 7.47(a))
∀ aj z . (8.14)
0<r0 <r
j=0
in the sense of Def. 8.6(b)

For the radius of convergence, one has the formula


1 p
n
r= , where L := lim sup |an |. (8.15)
L n→∞

In (8.15), lim sup denotes the pso-called limit superior, which is defined as the largest
cluster point of the sequence ( n |an |)n∈N if the sequence is bounded (cf. Th. 7.27) and
∞ if the sequence is unbounded. As the limit superior can be 0 or ∞, we also define
1/0 := ∞ and 1/∞ := 0 in (8.15).
One has the simpler formula
an
r = lim
, (8.16)
n→∞ an+1

provided all an are nonzero and provided the limit in (8.16) either exists in R+
0 or is ∞.

Proof. For the proof of (8.15), we apply theproot test from Th. 7.89(b). Here, for the
root test, we have to consider the sequence ( n |an ||z|n )n∈N . As a consequence of (7.11a)
and Prop. 7.26, lim supn→∞ (λxn ) = λ lim supn→∞ xn for each λ > 0 and each sequence
(xn )n∈N in R (with λ · ∞ := ∞, this also holds if the limit superior is infinite). Thus,
p p
lim sup n |an ||z|n = |z| lim sup n |an | = |z| L.
n→∞ n→∞

If |z| > 1/L, then |z| L > 1 and (7.82b) applies, i.e. (8.13b) holds for r = 1/L. If
|z| < 1/L, then |z| L < 1, and, recalling the Bolzano-Weierstrass Th. 7.27, one sees that
(7.82a) applies, i.e. (8.13a) holds for r = 1/L.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 122

P
Next, if 0 < r0 < r, then ∞ j
j=0 |aj r0 | converges according to (8.13a). Since, for each
z ∈ Br0 (0) and each j ∈ N, we have |aj z j | ≤ |aj r0j |, (8.14) is a consequence of Cor.
8.7(b).
The validity of (8.16) follows from the ratio test of Th. 7.89(c): If all an 6= 0 and z 6= 0,
then
an+1 z n+1 an+1 |z|
lim
n
= |z| lim
=
an .

n→∞ an z n→∞ an
limn→∞ an+1

an
If |z| < l := limn→∞ an+1 , then |z|/l < 1, i.e. (7.83a) applies, proving (8.13a) for r = l.
If |z| > l, then |z|/l > 1, i.e. (7.83b) applies, proving (8.13b) for r = l. 
P∞
Corollary 8.10. If j=0 aj z j , aj ∈ K, is a power series with radius of convergence
r ∈]0, ∞], then the function

X
f : Br (0) −→ K, f (z) := aj z j , (8.17)
j=0

is continuous. In particular, if r = ∞, then f is continuous on K.


P
Proof. Each partial sum z 7→ nj=0 aj z j is a polynomial, i.e. continuous on K. Moreover,
if ζ ∈ Br (0), then the power series converges uniformly on M := B|ζ| (0) by (8.14), i.e.
it is continuous at ζ ∈ M by Th. 8.4. 
P∞ α n
Example 8.11. (a) For each α ∈ R, the radius of convergence of n=1 n z is r = 1,
since p √
lim sup n |an | = lim n nα = 1, (8.18)
n→∞ n→∞

which, for each α ∈ Z, follows from (7.46) and Th. 7.13(a), and, then, for all α ∈ R
from the Sandwich Th. 7.16.
Let us investigate what can happen for |z| = r = 1 for some cases: The series
P ∞ n
n=1 z (α = 0) is divergent for each z ∈ C with z = 1 by the observation that
(z )n∈NPdoes not converge to 0 for n → ∞ (as |z n | = 1 for each n ∈ N); the
n

series ∞ n=1 n
−1 n
z (α = −1) is the harmonic series, i.e. divergent, for z = 1, but
convergent for z = −1 according to Ex. 7.86(a).
P P∞ z n
(b) The radius of convergence of both ∞ zn
n=0 n! and n=0 nn is r = ∞ by (8.16) and
(8.15), respectively, since

an
lim = lim (n + 1)! = lim (n + 1) = ∞, (8.19a)
n→∞ an+1 n→∞ n! n→∞
r
p n 1 1
lim sup n |an | = lim n
= lim = 0. (8.19b)
n→∞ n→∞ n n→∞ n

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 123

P
(c) The radius of convergence of ∞ n
n=0 n! z is r = 0 by (8.16), since

an n! 1
lim = lim = lim = 0. (8.20)
n→∞ an+1 n→∞ (n + 1)! n→∞ n + 1
P
Caveat 8.12. Theorem 8.9 does not claim the uniform convergence of ∞ aj z j on
P∞ j=0
Br (0), which is usually not true (e.g., it is an exercise to show that j=0 z j does not
converge uniformlyP∞on B1 (0)). Theorem 8.9 also claims nothing about the convergence
j
or divergence of j=0 aj z for |z| = r, which has to be determined case by case (cf. Ex.
8.11(a)).
P∞ j
Definition and Remark 8.13. Given two power series p := j=0 aj z and q :=
P∞ j
j=0 bj z in K, we define their Cauchy product

∞ j
X X
j
p ∗ q := cj z , where cj := ak bj−k = a0 bj + a1 bj−1 + · · · + aj b0 . (8.21)
j=0 k=0

Note that we have not assumed any convergence of the series so far, i.e. p, q, and p ∗ q
are not K-valued functions, but sequences of K-valued functions according to Def. 8.5
(sequences of polynomials, actually). Sometimes one also calls the Cauchy product p ∗ q
the convolution of p and q.
Now, if we do assume p and q to have some nonzero radii of convergence, say rp , rq ∈
]0, ∞], respectively, then, by (8.13a), both series are absolutely convergent for each
z ∈ Br (0), where r := min{rp , rq }. Thus, the functions

X ∞
X
f : Br (0) −→ K, f (z) := aj z j , g : Br (0) −→ K, g(z) := bj z j , (8.22)
j=0 j=0

are well-defined, and (7.94) implies



X
∀ f (z)g(z) = cj z j with cj as in (8.21). (8.23)
z∈Br (0)
j=0

8.3 Exponential Functions


The notion of power series allows us to extend the definition of exponential functions to
complex arguments:
Definition and Remark 8.14. We define the exponential function

X zn z2 z3
exp : C −→ C, exp(z) := =1+z+ + + ... (8.24)
n=0
n! 2! 3!

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 124

From Ex. 8.11(b), we already know the radius of convergence of the power series in
(8.24) is ∞, such that the function in (8.24) is well-defined.
For the time being, we also redefine Euler’s number as e := exp(1) > 1 > 0 and, for each
x ∈ R+ , ln x := logexp(1) (x). This, as well as calling the function of (8.24) exponential
function, will be justified as soon as we will have proved
 n X ∞
1 1 1 1 1
lim 1 + = = 1 + + + + ... (8.25)
n→∞ n n=0
n! 1! 2! 3!

and ∞
x
X xn x2 x3
∀ e = =1+x+ + + ... (8.26)
x∈R
n=0
n! 2! 3!
in (8.36) of Th. 8.18 and in Th. 8.16(c) below, respectively.

Proposition 8.15. If a continuous function E : R −→ R satisfies

a := E(1) > 0 and (8.27a)


∀ E(x + y) = E(x)E(y), (8.27b)
x,y∈R

then f is an exponential function – more precisely

∀ E(x) = ax . (8.28)
x∈R

Proof. First, a = E(1) = E(0 + 1) = E(0)E(1) = E(0) a and a > 0 shows E(0) = 1.
−1
Then, for each x ∈ R, 1 = E(0) = E(x − x) = E(x)E(−x), i.e. E(−x) = E(x) ,
showing E(x) 6= 0 for each x ∈ R. Thus, E(1) > 0, the continuity of E, and the
intermediate value Th. 7.57 imply E(x) > 0 for each x ∈ R. Next, an induction shows
n
∀ ∀ E(n · x) = E(x) : (8.29)
x∈R n∈N

The base case is trivially true and the induction step is


ind. hyp.
E((n + 1)x) = E(nx)E(x) = (E(x))n E(x) = (E(x))n+1 .

Applying (8.29) with x = 1 shows E(n) = an for each n ∈ N. Applying (8.29) with
x = 1/n, n ∈ N, shows a = E(1) = (E(1/n))n , i.e. E(1/n) = a1/n since E(1/n) > 0.
Next,
(8.29) k k
∀ E(k/n) = E(1/n) = (a1/n )k = a n ,
n,k∈N

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 125

showing (8.28) holds for each x ∈ Q+ . Then (8.28) also holds for each x ∈ R+ , since, if
(qn )n∈N is a sequence in Q+ with limn→∞ qn = x, then the continuity of E implies

ax = lim aqn = lim E(qn ) = E(x).


n→∞ n→∞

Finally, if x ∈ R− , then
−1
ax = (a−x )−1 = E(−x) = E(x),

completing the proof that (8.28) holds for each x ∈ R. 

Theorem 8.16. We consider the exponential function exp as defined in (8.24). The
following holds:

(a) exp is continuous on C.

(b) exp(z + w) = exp(z) exp(w) is valid for all z, w ∈ C.

(c) With e := exp(1) (cf. Def. and Rem. 8.14), it is



x
X xn
∀ e = exp(x) = .
x∈R
n=0
n!

Proof. (a) holds by Cor. 8.10; for (b), we compute (using (7.94)),
 ∞ 
X
exp(z) exp(w) = cn , 
 n=0 
∀  n n   ;
z,w∈C 
 X z j wn−j 1 X n j n−j (5.22) (z + w)n 

where cn = = z w =
j=0
j! (n − j)! n! j=0 j n!
(8.30)
and then (c) is an immediate consequence of (a), (b), and Prop. 8.15. 

Definition 8.17. Let M ⊆ C. If ζ ∈ C is a cluster point of M , then a function


f : M −→ K is said to tend to η ∈ K (or to have the limit η ∈ K) for z → ζ
(denoted by limz→ζ f (z) = η) if, and only if, for each sequence (zk )k∈N in M \ {ζ} with
limk→∞ zk = ζ, the sequence (f (zk ))k∈N converges to η ∈ K, i.e.
 
lim f (z) = η ⇔ ∀ lim zk = ζ ⇒ lim f (zk ) = η . (8.31)
z→ζ (zk )k∈N in M \{ζ} k→∞ k→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 126

Theorem 8.18. We consider the exponential function exp as defined in (8.24). With
ez := exp(z) for each z ∈ C and ln x := logexp(1) (x) for each x ∈ R+ (cf. Th. 8.16(c)
and Def. and Rem. 8.14), we have the following limits:
ez − 1 
lim =1 z ∈ M := C \ {0} , (8.32)
z→0 z
ln(1 + x) 
lim =1 x ∈ M :=] − 1, ∞[\{0} , (8.33)
x→0 x
1 
∀ lim ln(1 + ξ x) x = ξ x ∈ M := {x ∈ R : 1 + ξ x > 0} \ {0} , (8.34)
ξ∈R x→0
1 
∀ lim (1 + ξ x) x = eξ x ∈ M := {x ∈ R : 1 + ξ x > 0} \ {0} , (8.35)
ξ∈R x→0

 x n x
X xn
∀ lim 1 + =e = . (8.36)
x∈R n→∞ n n=0
n!

Proof. (8.32): From (8.24) and ez = exp(z), we obtain



ez − 1 X z n z z2
∀ = =1+ + + ...,
z6=0 z n=0
(n + 1)! 2! 3!
P∞ zn
which, since z 7→ n=0 (n+1)! is continuous on C by Cor. 8.10, implies (8.32).
(8.33): Consider the auxiliary function f : ] − 1, ∞[: −→ R, f (x) := ln(x + 1), with
f −1 (x) = ex − 1. Now, given a sequence (xk )k∈N in ] − 1, ∞[\{0} with limk→∞ xk = 0,
one obtains
 
ln(1 + xk ) ln 1 + f −1 (f (xk )) ln 1 + ef (xk ) − 1
lim = lim = lim
k→∞ xk k→∞ f −1 (f (xk )) k→∞ ef (xk ) − 1
f (xk ) (8.32)
= lim f (x ) = 1,
k→∞ e k − 1

where, in the last step, it was used that limk→∞ xk = 0 and the continuity of f implies
limk→∞ f (xk ) = ln 1 = 0.
Similarly, but simpler, one obtains (8.34) and (8.35) (exercise). Finally, for the sequence
(xn )n∈N with xn := 1/n, (8.35) implies (8.36). 

Definition 8.19 (Exponentiation with Complex Exponents). For each (a, z) ∈ R+ × C,


we define
az := exp(z ln a), (8.37)
where exp is the function defined in (8.24). For a = e, (8.37) yields ez = exp(z), i.e.
(8.37) is consistent with (8.26).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 127

Theorem 8.20. (a) The first two exponentiation rules of (7.54) still hold for each
a, b > 0 and each z, w ∈ C:

az+w = az aw , (8.38a)
az bz = (ab)z . (8.38b)

(b) For each a ∈ R+ , the exponential function

f : C −→ C, f (z) := az , (8.39a)

is continuous, and, for each ζ ∈ C, the power function

g : R+ −→ C, g(x) := xζ , (8.39b)

is continuous.

(c) The limit in (8.36) extends to complex numbers:



 z n z
X zn
∀ lim 1 + =e = . (8.40)
z∈C n→∞ n n=0
n!

Proof. (a): We compute


(8.37)
az+w = exp((z + w) ln a) = exp(z ln a + w ln a)
Th. 8.16(b) (8.37)
= exp(z ln a) exp(w ln a) = az aw ,

proving (8.38a), and


(8.37) Th. 8.16(b)
az b z = exp(z ln a) exp(z ln b) = exp(z ln a + z ln b)
(7.65e)  (8.37)
= exp z ln(ab) = (ab)z ,

proving (8.38b).
(b): The continuity of both functions follows from the continuity of exp (according to
Th. 8.16(a)) and from the fact that continuity is preserved by compositions (according
to Th. 7.41): The exponential function f , given by f (z) = ez ln a , is the composition of
the continuous functions z 7→ z ln a and w 7→ ew , whereas (analogous to Ex. 7.76(a)),
the power function g, given by g(x) = eζ ln x , is the composition g = exp ◦(ζ ln), where
ln is continuous by Cor. 7.74.
(c): Exercise. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 128

8.4 Trigonometric Functions


The first “definition” of the trigonometric functions sine and cosine is the one based on
geometric visualization usually given in high school: cos x and sin x are the coordinates
of the point p = (p1 , p2 ) ∈ R2 on the unit circle, such that x is the angle measured in
radian between the line segment between (0, 0) and (1, 0) and the line segment between
(0, 0) and p.
While this “definition” allows to obtain many important properties of sine and cosine
using geometric arguments, it is not mathematically rigorous, and, for example, provides
no clue how to compute values like sin 1. The problem is related to the fact that the
angle measured in radian between the line segment between (0, 0) and (1, 0) and the line
segment between (0, 0) and p is supposed to be the length of the segment of the unit
circle between (1, 0) and p (taken in the counter-clockwise direction).
In the following Def. and Rem. 8.21, we will provide a mathematically rigorous definition
of sine and cosine using power series, and we will then verify that the functions have the
familiar properties one learns in high school. However, as the computation of lengths of
curved paths is actually beyond the scope of this lecture, we will not be able to see that
our sine and cosine functions are precisely the same we visualized in high school (the
interested reader is referred to Ex. 1 in Sec. 5.14 of [Wal02] and to [Phi17, Ex. 3.13(b)]).

Definition and Remark 8.21. We define the sine function, denoted sin, and the
cosine function, denoted cos by

X (−1)n z 2n+1 z3 z5
sin : C −→ C, sin z := =z− + − +..., (8.41a)
n=0
(2n + 1)! 3! 5!

X (−1)n z 2n z2 z4
cos : C −→ C, cos z := =1− + − +.... (8.41b)
n=0
(2n)! 2! 4!

(a) sin and cos are well-defined and continuous: For both series and each z ∈ C, we can
estimate the absolute value of the nth summand by the nth summand of the series
for the exponential function e|z| (cf. (8.36)), which we know to be convergent from
Ex. 8.11(b). Thus, by Th. 8.9, both series in (8.41) have radius of convergence ∞
and are continuous by Cor. 8.10.

(b) cos : R −→ R (i.e. cos↾R ) has a smallest positive zero α ∈ R+ . We define π := 2α.
One can show π is an irrational number (see Appendix H.2) and its first digits are
π = 3.14159 . . .
To see cos has a smallest positive zero and to obtain a first (very coarse) estimate,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 129

note
 
xk xk+1 x
∀ ∀ > ⇔ 1> ⇔ k+1>x ,
x∈R+ k∈N k! (k + 1)! k+1
k k+1
showing xk! > (k+1)!
x
holds for each k ≥ 2 and each x ∈]0, 3[. In particular, the
summands of the series in (8.41) converge monotonically to 0 (for k ≥ 2) and, since
the series are alternating for x 6= 0, Th. 7.85 applies and (7.79) yields
 
x2 x2 x4
f (x) := 1 − 2 < cos x < 1 − 2 + 24 =: g(x),
∀   (8.42)
0<x<3  x3 x3 x5 
x− < sin x < x − + .
6 6 120
√ √ √
The zeros of x 7→ f p(x) are − 2,p2, i.e. 2p is its smallest positive zero;pthe zeros
√ √ √ p √ √
of x 7→ g(x) are − 6 − 2 3, − 6 + 2 3, 6 − 2 3, 6 + 2 3, i.e. 6 − 2 3
is its smallest positive zero. Thus, as f (0) = g(0) = 1, the intermediate value Th.
7.57 implies cos has a smallest positive zero α and
q
√ π √
1.4 < 2 < := α < 6 − 2 3 < 1.6 (8.43)
2
Theorem 8.22. We have the following identities:
sin 0 = 0, cos 0 = 1, (8.44a)
∀ sin z = − sin(−z), cos z = cos(−z), (8.44b)
z∈C
∀ sin(z + w) = sin z cos w + cos z sin w, (8.44c)
z,w∈C

∀ cos(z + w) = cos z cos w − sin z sin w, (8.44d)


z,w∈C

∀ (sin z)2 + (cos z)2 = 1, (8.44e)


z∈C
π π
cos = 0, sin = 1, ∀ cos x > 0, (8.44f)
2 2 x∈[0, π2 [
 π  π
∀ sin z + = cos z, cos z + = − sin z, (8.44g)
z∈C 2 2
∀ sin(z + π) = − sin z, cos(z + π) = − cos z, (8.44h)
z∈C
∀ sin(z + 2π) = sin z, cos(z + 2π) = cos z, (8.44i)
z∈C
sin z cos z − 1 1
lim = 1, lim 2
=− . (8.44j)
z→0 z z→0 z 2
Identities (8.44i) can be restated as sine and cosine being periodic functions with period
2π.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 130

Proof. (8.44a) is immediate from (8.41) since, for z = 0, all summands of the sine
series are 0 and all summands of the cosine series are 0, except the first one, which is
(−1)0 00
0!
= 1.
(8.44b) is also immediate from (8.41), since (−z)2n+1 = (−1)2n+1 z 2n+1 = −z 2n+1 and
(−z)2n = (−1)2n z 2n = z 2n .
(8.44c) and (8.44d) can be verified using the Cauchy product: According to (7.94),
 ∞ ∞

X X
 sin z cos w = cn , cos z sin w = dn , 
 n=0 n=0

 
 n j 2j+1 n−j 2(n−j) 
 X (−1) z (−1) w 
∀ where cn = , ,
z,w∈C 
 j=0
(2j + 1)! (2(n − j))!  
 n

 X (−1)j z 2j (−1)n−j w2(n−j)+1 
 dn = ,
j=0
(2j)! (2(n − j) + 1)!

that means, for each z, w ∈ C,


n n
X (−1)n z 2j+1 w2(n−j) X (−1)n z 2j w2(n−j)+1
cn + d n = +
j=0
(2j + 1)! (2(n − j))! j=0 (2j)! (2(n − j) + 1)!
n n
X (−1)n z 2j+1 w2n+1−(2j+1) X (−1)n z 2j w2n+1−2j
= +
j=0
(2j + 1)! (2n + 1 − (2j + 1))! j=0 (2j)! (2n + 1 − 2j)!
2n+1 2n+1  
n
X z j w2n+1−j (−1)n X 2n + 1 j 2n+1−j
= (−1) = z w
j=0
j! (2n + 1 − j)! (2n + 1)! j=0 j

(−1)n (z + w)2n+1
= ,
(2n + 1)!

proving (8.44c). Similarly, according to (7.94),


 ∞ ∞

X X
cos z cos w = cn , sin z sin w = dn , 
 n=0 n=0

 
 n j 2j n−j 2(n−j) 
 X (−1) z (−1) w 
∀ where cn = , ,
z,w∈C 
 j=0
(2j)! (2(n − j))! 

 n

 X (−1)j z 2j+1 (−1)n−j w2(n−j)+1 
 dn = ,
j=0
(2j + 1)! (2(n − j) + 1)!

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 131

that means, for each z, w ∈ C,

c0 = 1 and
n n−1
X (−1)n z 2j w2(n−j) X (−1)n−1 z 2j+1 w2(n−1−j)+1
∀ cn − dn−1 = −
n∈N
j=0
(2j)! (2(n − j))! j=0 (2j + 1)! (2(n − 1 − j) + 1)!
n n−1
X (−1)n z 2j w2n−2j X (−1)n z 2j+1 w2n−(2j+1)
= +
j=0
(2j)! (2n − 2j)! j=0 (2j + 1)! (2n − (2j + 1))!
2n 2n  
X z j w2n−j
n (−1)n X 2n j 2n−j
= (−1) = z w
j=0
j! (2n − j)! (2n)! j=0
j

(−1)n (z + w)2n
= ,
(2n)!

proving (8.44d).
(8.44e): One computes for each z ∈ C:
(8.44d)
(sin z)2 + (cos z)2 = cos z cos(−z) − sin z sin(−z) = cos(z − z) = cos 0 = 1.

(8.44f): cos π2 = 0 and cos x > 0 for 0 ≤ x < π


2
hold according to the definition of π in
Def. and Rem. 8.21(b). Then
 π 2 (8.44e)

π 2 π π (π/2)3 (8.43) (1.6)3
sin = 1− cos = 1 and sin > − > 1.4− > 0.7 > 0.
2 2 2 2 6 6

(8.44g) is immediate from (8.44c), (8.44d), and (8.44f).


(8.44h): One obtains
π π  (8.44c)
sin π = sin + = 1 · 0 + 0 · 1 = 0,
2 2
 π π  (8.44d)
cos π = cos + = 0 · 0 − 1 · 1 = −1,
2 2
(8.44c)
∀ sin(z + π) = − sin z + 0 = − sin z,
z∈C
(8.44d)
∀ cos(z + π) = − cos z + 0 = − cos z.
z∈C

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 132

(8.44i): One obtains


(8.44c)
sin(2π) = sin(π + π) = 0 + 0 = 0,
(8.44d)
cos(2π) = cos(π + π) = (−1)(−1) − 0 = 1,
(8.44c)
∀ sin(z + 2π) = sin z + 0 = sin z,
z∈C
(8.44d)
∀ cos(z + 2π) = cos z − 0 = cos z.
z∈C

(8.44j): One obtains


 ∞ 
sin z X (−1)n z 2n z2 z4
 = =1− + − +..., 
 z n=0
(2n + 1)! 3! 5! 
∀ 

.
 n+1 2n 2 4 
z∈C\{0}  cos z − 1 X (−1) z 1 z z 
= = − + − + − . . .
z2 n=0
(2(n + 1))! 2! 4! 6!

For both series on the right-hand side and each z ∈ C, we can estimate the absolute
value of each summand by the corresponding summand of the exponential series for e|z|
(cf. (8.36)), showing they have radius of convergence ∞ and are continuous by Cor. 8.10.
In particular, their continuity in z = 0 proves (8.44j). 

Theorem 8.23. One has sin(R) = cos(R) = [−1, 1], i.e. the image of both sine and
cosine is [−1, 1]. Moreover, for each k ∈ Z:
h π π i
sin is strictly increasing on − + 2kπ, + 2kπ , (8.45a)
 2 2 
π 3π
sin is strictly decreasing on + 2kπ, + 2kπ , (8.45b)
2 2
cos is strictly increasing on [(2k − 1)π, 2kπ], (8.45c)
cos is strictly decreasing on [2kπ, (2k + 1)π], (8.45d)

which, due to (8.44e), can be summarized (and visualized) by saying that, if x runs from
2kπ to 2(k + 1)π, then (cos x, sin x) runs once counterclockwise through the unit circle,
starting at (1, 0).

Proof. From (8.44e), we know sin(R) ⊆ [−1, 1] and cos(R) ⊆ [−1, 1]. As
π (8.44f)  π  (8.44b) (8.44a) (8.44h)
sin = 1, sin − = −1, cos 0 = 1, cos π = − cos 0 = −1,
2 2

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 133

the continuity of sine and cosine together with the intermediate value Th. 7.57 implies
sin(R) = cos(R) = [−1, 1].
3 2 4
From (8.42), we know 0 < x − x6 < sin x and cos x < 1 − x2 + x24 < 1 for each x ∈]0, π2 ],
implying
∀ cos(x + y) = cos x cos y − sin x sin y ≤ cos x cos y < cos x,
0≤x<x+y≤ π2

showing cos is strictly decreasing on [0, π2 ]. Then cos is strictly increasing on [− π2 , 0] by


(8.44b), sin is strictly increasing on [0, π2 ] and strictly decreasing on [ π2 , π] by (8.44g),
implying sin is strictly increasing on [− π2 , 0] and strictly decreasing on [−π, − π2 ] by
(8.44b), i.e. sin is strictly increasing on [ 3π2
, 2π] and strictly decreasing on [π, 3π 2
] by
π
(8.44i), implying cos is strictly decreasing on [ 2 , π] and strictly increasing on [−π, − π2 ]
by (8.44g). Since this fixes the monotonicity properties of both sine and cosine over
more than one period, the general statements in (8.45) are provided by (8.44i). 

We now come to important complex number relations between sine, cosine, and the
exponential function.
Theorem 8.24. One has the following formulas, relating the (complex) sine, cosine,
and exponential function:
∀ eiz = cos z + i sin z (Euler formula), (8.46a)
z∈C

eiz + e−iz
∀ cos z = , (8.46b)
z∈C 2
eiz − e−iz
∀ sin z = . (8.46c)
z∈C 2i

Proof. Let z ∈ C. For (8.46a), one computes


∞ ∞  
iz (8.37),(8.24)
X (iz)n X (−1)n z 2n i (−1)n z 2n+1
e = = +
n=0
n! n=0
(2n)! (2n + 1)!
∞ ∞
Th. 7.95(a) X (−1)n z 2n X (−1)n z 2n+1 (8.41)
= +i = cos z + i sin z.
n=0
(2n)! n=0
(2n + 1)!
Then
(8.46a)
eiz + e−iz = cos z + i sin z + cos(−z) + i sin(−z) = 2 cos z
proves (8.46b), and
(8.46a)
eiz − e−iz = cos z + i sin z − cos(−z) − i sin(−z) = 2i sin z
proves (8.46c). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 134

As a first application of (8.46), we can now determine all solutions to the equation
ez = 1 and all zeros (if any) of exp, sin, and cos:
Theorem 8.25. The set of (complex) solutions to the equation ez = 1 consists precisely
of all integer multiples of 2πi, the exponential function has no zeros (neither in R nor
in C), and the set of all (real or complex) zeros of sine and cosine consists of a discrete
set of real numbers. More precisely:

exp−1 {1} = {2kπi : k ∈ Z}, (8.47a)


exp−1 {0} = ∅, (8.47b)
sin−1 {0} = {kπ : k ∈ Z}, (8.47c)

cos−1 {0} = (2k + 1) π2 : k ∈ Z . (8.47d)

Proof. Exercise. 
Definition and Remark 8.26. We define tangent and cotangent by
sin z
tan : C \ cos−1 {0} −→ C, tan z := , (8.48a)
| {z } cos z
C \ {(2k + 1) π2 : k ∈ Z} by (8.47d)
cos z
cot : C \ sin−1 {0} −→ C, cot z := , (8.48b)
| {z } sin z
C \ {kπ : k ∈ Z} by (8.47c)

respectively. Since sine and cosine are both continuous, tangent and cotangent are also
both continuous on their respective domains. Both functions have period π, since, for
each z in the respective domains,
sin(z + π) (8.44h) − sin z (8.44h) − cos z
tan(z + π) = = = tan z, cot(z + π) = = cot z.
cos(z + π) − cos z − sin z
(8.49)
Since
   
π 1 π π 1 π
lim sin − = sin = 1 ∧ lim cos − = cos = 0
n→∞ 2 n 2 n→∞ 2 n 2
   
π 1 π 1
∧ cos − >0 ⇒ lim tan − = ∞,
2 n n→∞ 2 n

   π    π
π 1 π 1
lim sin − + = sin − = −1 ∧ lim cos − + = cos − =0
n→∞ 2 n 2 n→∞ 2 n 2
   
π 1 π 1
∧ cos − + >0 ⇒ lim tan − + = −∞,
2 n n→∞ 2 n

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 135

1 1 1
lim sin = sin 0 = 0 ∧ lim cos = cos 0 = 1 ∧ sin > 0
n→∞ n n→∞ n n
1
⇒ lim cot = ∞,
n→∞ n
   
1 1
lim sin π − = sin π = 0 ∧ lim cos π − = cos π = −1
n→∞ n n→∞ n
   
1 1
∧ sin π − >0 ⇒ lim cot π − = −∞,
n n→∞ n
we obtain tan(R \ cos−1 {0}) = cot(R \ sin−1 {0}) = R.
For each k ∈ Z,
i π π h
tan is strictly increasing on − + kπ, + kπ , (8.50a)
i2 2 h
cot is strictly decreasing on kπ, (k + 1)π : (8.50b)
On ]0, π2 [, sin is strictly increasing and cos is strictly decreasing, i.e. tan is strictly
increasing and cot is strictly decreasing. Since tan(−x) = sin(−x)/ cos(−x) = − tan(x),
on ] − π2 , 0[, tan is strictly increasing and cot is strictly decreasing. Taking into account
the signs of tan and cot on the respective intervals and their π-periodicity according to
(8.49) proves (8.50).
Definition and Remark 8.27. Since we have seen sin to be strictly increasing on
[− π2 , π2 ] with image [−1, 1], cos to be strictly decreasing on [0, π] with image [−1, 1], tan
to be strictly increasing on ] − π2 , π2 [ with image R, and cot to be strictly decreasing on
]0, π[ with image R; and since all four functions are continuous, Th. 7.60 implies the
existence of inverse functions, denoted by
arcsin : [−1, 1] −→ [−π/2, π/2], (8.51a)
arccos : [−1, 1] −→ [0, π], (8.51b)
arctan : R −→] − π/2, π/2[, (8.51c)
arccot : R −→]0, π[, (8.51d)
respectively, where all four inverse functions are continuous, arcsin is strictly increasing,
arccos is strictly decreasing, arctan is strictly increasing, and arccot is strictly decreasing.
Of course, using (8.45) and (8.50), respectively, one can also obtain the inverse functions
on different intervals, and, in the literature, such inverse functions are, indeed, considered
as well. Somewhat confusingly, it is common to denote all these different functions by
the same symbols, namely the ones introduced in (8.51). Here, we will not need to pursue
this any further, i.e. we will only consider the inverse functions precisely as defined in
(8.51), which are also known as the principle inverse functions of sin, cos, tan, and cot,
respectively.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 136

8.5 Polar Form of Complex Numbers, Fundamental Theorem


of Algebra
Theorem 8.28. For each complex number z ∈ C, there exist real numbers r ≥ 0 and
ϕ ∈ R such that
z = r eiϕ . (8.52)
Moreover, if (8.52) holds with r ≥ 0 and ϕ ∈ R, then r is the modulus of z and, for
z 6= 0, ϕ is uniquely determined up to addition of an integer multiple of 2π, i.e.
 
∀ z = r eiϕ1 = r eiϕ2 ∧ r ≥ 0 ⇒ r = |z| ∧ ∃ ϕ1 − ϕ2 = 2πk . (8.53)
z∈C\{0} k∈Z

Proof. For z = 0, there is nothing to prove, so we assume z 6= 0 and set r := |z|. We


write z = x + iy with x, y ∈ R, first assuming y ≥ 0. Then
z x y
= ξ + iη, where ξ = , η= ≥ 0, ξ 2 + η 2 = 1. (8.54)
r r r
In particular, −1 ≤ ξ ≤ 1. Thus, letting

ϕ := arccos ξ,

we obtain ϕ ∈ [0, π], ξ = cos ϕ, and sin ϕ ≥ 0, yielding


p p (8.54)
sin ϕ = 1 − (cos ϕ)2 = 1 − ξ 2 = η.

In consequence,
z (8.46a)
= ξ + iη = cos ϕ + i sin ϕ = eiϕ ,
r
as desired. If y ≤ 0, then the above shows the existence of ψ ∈ R such that z̄ = x − iy =
reiψ = r cos ψ + ir sin ψ. Letting ϕ := −ψ, we, once again, have z = r cos ψ − ir sin ψ =
re−iψ = reiϕ , as desired, completing the existence proof for the representation (8.52).
Now assume (8.52) holds with r ≥ 0. Then
p
|z| = r|eiϕ | = r (sin ϕ)2 + (cos ϕ)2 = r.

Finally, if r eiϕ1 = r eiϕ2 with r > 0, then ei(ϕ1 −ϕ2 ) = 1, i.e. i(ϕ1 − ϕ2 ) ∈ {2kπi : k ∈ Z}
by (8.47a). 
Definition and Remark 8.29. The representation of z ∈ C given by (8.52) is called its
polar form, where (r, ϕ) are also called polar coordinates of z, ϕ is called an argument of
z. For z 6= 0, one can fix the argument uniquely by the additional requirement ϕ ∈ [0, 2π[
(but one also finds other choices, for example ϕ ∈] − π, π], in the literature). The above

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 137

terminology is consistent with the common use of calling (r, ϕ) polar coordinates of the
vector z = (x, y) ∈ R2 (= C) (in contrast to the Cartesian coordinates (x, y)), where r
constitutes the distance of the point z = (x, y) from the origin (0, 0) and ϕ is the angle
between the vector z = (x, y) and the x-axis (cf. the three introductory paragraphs
of the previous Sec. 8.4). As promised, we can now better understand the geometric
interpretation of complex multiplication already described in Rem. 5.10: If z1 = r1 eiϕ1
and z2 = r2 eiϕ2 , then z1 z2 = r1 r2 ei(ϕ1 +ϕ2 ) , i.e. complex multiplication, indeed, means
multiplying absolute values and adding arguments.

Corollary 8.30. If z ∈ C, then |z| = 1 holds if, and only if, there exists ϕ ∈ R such
that z = eiϕ – in other words, the map

f : R −→ {z ∈ C : |z| = 1}, f (ϕ) := eiϕ , (8.55)

is surjective. Moreover f (ϕ1 ) = f (ϕ2 ) holds if, and only if, ϕ1 − ϕ2 = 2πk for some
k ∈ Z.

Proof. Everything is immediate from Th. 8.28. 

Corollary 8.31 (Roots of Unity). For each n ∈ N, the equation z n = 1 has precisely n
distinct solutions ζ1 , . . . , ζn ∈ C, where

(8.46a) k2π k2π


∀ ζk := ek2πi/n = cos + i sin = ζ1k . (8.56)
k=1,...,n n n

The numbers ζ1 , . . . , ζn defined in (8.56) are called the nth roots of unity.

Proof. It is ζkn = ek2πi = 1 for each k ∈ {1, . . . , n} and the ζ1 , . . . , ζn are all distinct
by Cor. 8.30, since, for k, l ∈ {1, . . . , n} with k 6= l, (k − l)/n ∈ / Z. As ζ1 , . . . , ζn are
n distinct zeros of the polynomial P : C −→ C, P (z) := z n − 1, and P has at most n
zeros by Th. 6.6(a), ζ1 , . . . , ζn constitute all solutions to z n = 1. 

We are now in a position to prove one of the central results of analysis and algebra,
namely the fundamental theorem of algebra. The following proof does not need any tools
beyond the ones provided by this class – it is actually mainly founded on continuous
functions attaining a min and a max on compact sets according to Th. 7.54 and the
existence of nth roots of unity according to Cor. 8.31.

Theorem P8.32 (Fundamental Theorem of Algebra). Every polynomial P : C −→ C,


P (z) := nj=0 aj z j , of degree n ≥ 1 (i.e. a0 , . . . , an ∈ C with an 6= 0) has at least one
zero z0 ∈ C.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 138

Proof. Dividing the equation P (z) = 0 by an 6= 0, it suffices to consider the case an = 1.


We therefore assume

∀ P (z) = z n + an−1 z n−1 + · · · + a1 z + a0 .


z∈C

Claim 1. The function |P | attains its global min on C, i.e. there exists z0 ∈ C such that
|P | is minimal in z0 .

Proof. We first note


 an−1 a0
∀ P (z) = z n 1 + r(z) , where r(z) := + ··· + n.
z6=0 z z

Set M := |a0 | + · · · + |an−1 | and R := max{1, 2M }.


Then
|z|≥1 M |z|≥2M 1
∀ |r(z)| ≤ ≤
|z|≥R |z| 2
and, thus,
|z|n
∀ |P (z)| = |z|n 1 + r(z) ≥ ≥ M.
|z|≥R 2
This estimate together with |P (0)| = |a0 | ≤ M shows that the min of |P | on the compact
disk B R (0) (see Ex. 7.47(a)) (such a min z0 ∈ B R (0) exists due to Th. 7.54) must be
the global min of |P | on C. N
Claim 2. If |P | has a min in z0 ∈ C, then P (z0 ) = 0.

Proof. Proceeding by contraposition, we assume P (z0 ) 6= 0 and show that |P | does not
have a min in z0 . We need to construct z1 ∈ C such that |P (z1 )| < |P (z0 )|. To this end,
define
P (z0 + z)
p : C −→ C, p(z) := .
P (z0 )
Then p is still a polynomial of degree n. Since p(0) = 1,
n
X
∃ ∃ ∀ p(z) = 1 + bj z j , bk 6= 0.
k∈{1,...,n} bk ,...,bn ∈C z∈C
j=k

Write −b−1 −1
k in polar form, i.e. −bk = re

with r ∈ R+ and ϕ ∈ R. Define
√ 
β := k r eiϕ/k i.e. β k = reiϕ = −b−1
k

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
8 CONVERGENCE OF K-VALUED FUNCTIONS 139

and
n
X
k k
q : C −→ C, q(z) := p(βz) = 1 + bk β z + bj β j z j = 1 − z k + z k+1 S(z),
j=k+1

where S is the polynomial


n−k−1
X
S : C −→ C, S(z) := bk+1+j β k+1+j z j (S ≡ 0 in case k = n).
j=0

Then, according to Th. 7.54,

∃ ∀ |S(z)| ≤ C.
C∈R+ z∈B 1 (0)

Letting
c := min{1, C −1 },
one obtains k+1
∀ z S(z) ≤ C |z|k+1 < |z|k
0<|z|<c

and, thus,
∀ |q(x)| ≤ 1 − xk + xk+1 S(x) < 1 − xk + xk = 1.
x∈]0,c[

Thus, finally,
|P (z0 + βx)|
∀ = |p(βx)| = |q(x)| < 1,
x∈]0,c[ |P (z0 )|
showing |P | does not have a min in z0 . N

Combining Claims 1 and 2 completes the proof of the theorem. 

Corollary 8.33. (a) For every polynomial P : C −→ C of degree n ≥ 1, there exist


numbers c, ζ1 , . . . , ζn ∈ C such that
n
Y
P (z) = c (z − ζj ) = c(z − ζ1 )(z − ζ2 ) · · · (z − ζn ) (8.57)
j=1

(the ζ1 , . . . , ζn are precisely all the zeros of P , some or all of which might be iden-
tical).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 140

(b) For every polynomial P : R −→ R of degree n ≥ 1, there exist numbers n1 , n2 ∈ N0


and c, ξ1 , . . . , ξn1 , α1 , . . . , αn2 , β1 , . . . , βn2 ∈ R such that

n = n1 + 2n2 , (8.58a)

and
n1
Y n2
Y
P (x) = c (x − ξj ) (x2 + αj x + βj ). (8.58b)
j=1 j=1

Proof. For (a), one just combines Th. 8.32 with Rem. 6.7.
(b): If P has only real coefficients, then we can take complex conjugates to obtain

P (ζ) = 0 ⇒ P (ζ) = P (ζ) = 0, (8.59)

showing that the nonreal zeros of P (if any) must occur in conjugate pairs. Moreover,

(x − ζ)(x − ζ) = x2 − (ζ + ζ)x + ζζ = x2 − 2x Re ζ + |ζ|2 , (8.60)

showing that (8.57) implies (8.58). 

9 Differential Calculus

9.1 Definition of Differentiability and Rules


The basic idea of differential calculus is to locally approximate nonlinear functions f by
linear functions. In our case, f will be defined on a subset M of R and, given ξ ∈ M
and R-valued f , we will investigate the question if we can define a number f ′ (ξ) ∈ R
that represents the slope of the graph of f at ξ such that the line through ξ with slope
f ′ (ξ) (called the tangent of f in ξ) can be considered as a local approximation of the
graph of f .
If such a local approximation of f in ξ is at all reasonable, then, for x 6= ξ,

f (x) − f (ξ)
x−ξ

should provide “good” approximations of f ′ (ξ) if x tends to ξ. This leads to the following
Def. 9.1, where we also allow C-valued functions (while the above-described geometric
interpretation only works for R-valued functions, it can be applied to both the real and

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 141

the imaginary parts of a C-valued function, cf. Rem. 9.2 below); but note that we do
not consider differentiability of functions f : C −→ C, which would lead to the notion
of complex differentiability or holomorphicity, which is studied in the field of Complex
Analysis and is beyond the scope of this class.
Definition 9.1. Let a < b, f : ]a, b[−→ K (a = −∞, b = ∞ is admissible), and ξ ∈]a, b[.
Then f is said to be differentiable at ξ if, and only if, the following limit in (9.1) exists
in the sense of Def. 8.17 (where x 7→ f (x)−f
x−ξ
(ξ)
plays the role of x 7→ f (x) in Def. 8.17).
The limit is then called the derivative of f in ξ. Many symbols are used in the literature
to denote derivatives, the following provides a selection:
df (ξ) f (x) − f (ξ) f (ξ + h) − f (ξ)
f ′ (ξ) := ∂x f (ξ) := := lim = lim . (9.1)
dx x→ξ x−ξ h→0 h
Note both limits occurring in (9.1) are, indeed, identical, since the sequence (xk )k∈N in
]a, b[ converges to ξ if, and only if, the sequence (hk )k→∞ with hk := xk − ξ converges
to 0. The number in (9.1) (if it exists) is also called a differential quotient, whereas
f (x)−f (ξ)
x−ξ
is known as a difference quotient.
f is called differentiable if, and only if, it is differentiable at each ξ ∈]a, b[. In that case,
one calls the function
f ′ : ]a, b[−→ K, x 7→ f ′ (x), (9.2)
the derivative of f .
Remark 9.2. In the situation of Def. 9.1, the complex-valued function f : ]a, b[−→ C
is differentiable at ξ ∈]a, b[ if, and only if, both functions Re f, Im f : ]a, b[−→ R are
differentiable, and, in that case
f ′ (ξ) = (Re f )′ (ξ) + i (Im f )′ (ξ). (9.3)
Indeed, we merely have to note
f (x) − f (ξ) Re f (x) − Re f (ξ) Im f (x) − Im f (ξ)
∀ = +i (9.4)
x,ξ∈]a,b[, x−ξ x−ξ x−ξ
x6=ξ

and that, by (7.2) a sequence (zn )n∈N in C converges to ζ ∈ C if, and only if, both
limn→∞ Re zn = Re ζ and limn→∞ Im zn = Im ζ hold.
Definition 9.3. If f : ]a, b[−→ R as in Def. 9.1 is differentiable at ξ ∈]a, b[, then the
graph of the affine function
L : R −→ R, L(x) := f (ξ) + f ′ (ξ)(x − ξ), (9.5)
i.e. the line through (ξ, f (ξ)) with slope f ′ (ξ) is called the tangent to the graph of f at
ξ.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 142

Theorem 9.4. If f : ]a, b[−→ K as in Def. 9.1 is differentiable at ξ ∈]a, b[, then it is
continuous at ξ. In particular, if f is everywhere differentiable, then it is everywhere
continuous.

Proof. Let (xk )k∈N be a sequence in ]a, b[\{ξ} such that limk→∞ xk = ξ. Then

 (xk − ξ) f (xk ) − f (ξ)
lim f (xk ) − f (ξ) = lim = 0 · f ′ (ξ) = 0, (9.6)
k→∞ k→∞ xk − ξ
proving the continuity of f in ξ. 
Example 9.5. (a) For each a, b ∈ K, the affine function f : R −→ K, f (x) := ax + b,
is differentiable with f ′ (x) = a for each x ∈ R: If x ∈ R and (hk )k∈N is a sequence
with hk 6= 0 such that limk→∞ hk = 0, then
f (x + hk ) − f (x) a(x + hk ) + b − ax − b a hk
lim = lim = lim = a. (9.7)
k→∞ hk k→∞ hk k→∞ hk

In particular, each constant function f ≡ b has derivative f ′ ≡ 0.

(b) For each c ∈ K, the function f : R −→ K, f (x) := ecx , is differentiable with


f ′ (x) = c ecx for each x ∈ R (in particular, c = 1 yields f ′ (x) = ex for f (x) = ex ,
and c = ln a yields f ′ (x) = (ln a) ax for f (x) = ax = ex ln a , a ∈ R+ ): The case c = 0
was treated in (a). Thus, let c 6= 0. If x ∈ R and (hk )k∈N is a sequence with hk 6= 0
such that limk→∞ hk = 0, then
f (x + hk ) − f (x) ecx+chk − ecx echk − 1 (8.32) cx
lim = lim = c ecx lim = c e . (9.8)
k→∞ hk k→∞ hk k→∞ chk

(c) The sine and the cosine function f, g : R −→ R, f (x) := sin x, g(x) := cos x, are
differentiable with f ′ (x) = cos x and g ′ (x) = − sin x for each x ∈ R: If x ∈ R and
(hk )k∈N is a sequence with hk 6= 0 such that limk→∞ hk = 0, then
f (x + hk ) − f (x) sin(x + hk ) − sin x
lim = lim
k→∞ hk k→∞ hk
(8.44c) sin x cos hk + cos x sin hk − sin x
= lim
k→∞ hk
hk (cos hk − 1) sin hk
= sin x lim 2
+ cos x lim
k→∞ h k→∞ hk
 k
(8.44j) 1
= (sin x) · 0 · − + (cos x) · 1 = cos x. (9.9)
2
The proof of g ′ (x) = − sin x is left as an exercise.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 143

(d) The absolute value function f : R −→ R, f (x) := |x|, is not differentiable at ξ = 0:

f (0 + n1 ) − f (0)
lim 1 = lim 1 = 1, (9.10a)
n→∞ n→∞
n
1 1
f (0 − n
)− f (0) n
lim = lim = −1, (9.10b)
n→∞ − n1 n→∞ − 1
n

f (0+h)−f (0)
showing that h
does not have a limit for h → 0.
Remark 9.6. As just seen in Ex. 9.5(d), the absolute value function shows continuous
functions do not need to be differentiable. Somewhat surprisingly, with a bit more effort,
one can even construct continuous functions f : R −→ R that are not differentiable at
any x ∈ R (see Appendix J.1).
Theorem 9.7. Let a < b, f, g : ]a, b[−→ K (a = −∞, b = ∞ is admissible), and
ξ ∈]a, b[. Assume f and g are differentiable at ξ.

(a) For each λ ∈ K, λf is differentiable at ξ and (λf )′ (ξ) = λf ′ (ξ).


(b) f + g is differentiable at ξ and (f + g)′ (ξ) = f ′ (ξ) + g ′ (ξ).
(c) Product Rule: f g is differentiable at ξ and (f g)′ (ξ) = f ′ (ξ)g(ξ) + f (ξ)g ′ (ξ).
(d) Quotient Rule: If g(ξ) 6= 0, then f /g is differentiable at ξ and
f ′ (ξ)g(ξ) − f (ξ)g ′ (ξ) g ′ (ξ)
(f /g)′ (ξ) = , in particular (1/g)′ (ξ) = − .
(g(ξ))2 (g(ξ))2

Proof. Let (hk )k∈N be a sequence with hk 6= 0 such that limk→∞ hk = 0.


For (a), one computes
(λf )(ξ + hk ) − (λf )(ξ) λf (ξ + hk ) − λf (ξ)
lim = lim
k→∞ hk k→∞ hk
f (ξ + hk ) − f (ξ)
= λ lim = λ f ′ (ξ).
k→∞ hk
For (b), one computes
(f + g)(ξ + hk ) − (f + g)(ξ) f (ξ + hk ) − f (ξ) + g(ξ + hk ) − g(ξ)
lim = lim
k→∞ hk k→∞ hk
f (ξ + hk ) − f (ξ) g(ξ + hk ) − g(ξ)
= lim + lim = f ′ (ξ) + g ′ (ξ).
k→∞ hk k→∞ hk

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 144

For (c), one computes

(f g)(ξ + hk ) − (f g)(ξ)
lim
k→∞ hk
f (ξ + hk )g(ξ + hk ) − f (ξ)g(ξ + hk ) + f (ξ)g(ξ + hk ) − f (ξ)g(ξ)
= lim
k→∞ hk
f (ξ + hk ) − f (ξ) g(ξ + hk ) − g(ξ)
= lim g(ξ + hk ) lim + f (ξ) lim
k→∞ k→∞ hk k→∞ hk
= f ′ (ξ)g(ξ) + f (ξ)g ′ (ξ),

where, in the last equality, we used the continuity of g in ξ according to Th. 9.4.
For (d), one first proves the special case f ≡ 1 by

(1/g)(ξ + hk ) − (1/g)(ξ) g(ξ) − g(ξ + hk ) g ′ (ξ)


lim = lim =− ,
k→∞ hk k→∞ g(ξ + hk )g(ξ)hk (g(ξ))2

which implies the general case using (c):


 ′
′ 1 f ′ (ξ) f (ξ)g ′ (ξ) f ′ (ξ)g(ξ) − f (ξ)g ′ (ξ)
(f /g) (ξ) = f · (ξ) = − = ,
g g(ξ) (g(ξ))2 (g(ξ))2

completing the proof. 

Example 9.8. (a) Each polynomial is differentiable and the derivative is, again, a
polynomial. More precisely,
n
X
P : R −→ K, P (x) = aj x j , aj ∈ K
j=0
n (9.11)
X
⇒ P ′ : R −→ K, P ′ (x) = j aj xj−1 :
j=1

The cases n = 0, 1 are provided by Ex. 9.5(a). To complete the induction P proof of
(9.11), we carry out the induction step for each n ∈ N: Writing P (x) = nj=0 aj xj +
an+1 x xn and applying the induction hypothesis as well as the rules of Th. 9.7 yields
n
X n+1
X
′ j−1 n n−1
P (x) = j aj x + an+1 (1 · x + x · n · x )= j aj xj−1 ,
j=1 j=1

which establishes the case.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 145

(b) Clearly, the derivatives of rational functions P/Q with polynomials P and Q can
be computed from (9.11) and the quotient rule of Th. 9.7(d).
(c) The functions tan and cot as defined in (8.48) and restricted to R \ cos−1 {0} and
R \ sin−1 {0}, respectively, are differentiable and one obtains
1
tan′ : R \ cos−1 {0} −→ R, tan′ x = = 1 + (tan x)2 , (9.12a)
| {z } (cos x)2
R\{(2k+1) π2 : k∈Z}

1
cot′ : R \ sin−1 {0} −→ R, cot′ x = − 2
= −(1 + (cot x)2 ) : (9.12b)
| {z } (sin x)
R\{kπ: k∈Z}

One merely needs the derivatives of sin and cos from Ex. 9.5(c) and the quotient
rule of Th. 9.7(d):
cos x cos x − sin x(− sin x) (8.44e) 1 (8.44e)
tan′ x = = = 1 + (tan x)2 ,
(cos x)2 (cos x)2
− sin x sin x − cos x cos x (8.44e) 1 (8.44e)
cot′ x = 2
= − = −(1 + (cot x)2 ).
(sin x) (sin x)2
Theorem 9.9 (Derivative of Inverse Functions). Let a < b, I :=]a, b[ (a = −∞, b = ∞
is admissible). If f : I −→ R is differentiable and strictly increasing (resp. decreasing),
then f has a continuous, strictly increasing (resp. decreasing) inverse function f −1 de-
fined on the interval J := f (I), i.e. f −1 : J −→ I, and, for each ξ ∈ I with f ′ (ξ) 6= 0,
f −1 is differentiable at η := f (ξ) with
1 1
(f −1 )′ (η) = = . (9.13)
f ′ (ξ) f′ f −1 (η)

Proof. As a differentiable function, f is continuous by Th. 9.4, i.e. Th. 7.60 provides
all the present assertions, except differentiability at η and (9.13). Let (yk )k∈N be a
sequence in J \ {η} such that limk→∞ yk = η. Then, as f −1 is bijective and continuous,
(f −1 (yk ))k∈N is a sequence in I \ {ξ} such that limk→∞ f −1 (yk ) = ξ, and one obtains
f −1 (yk ) − f −1 (η) f −1 (yk ) − f −1 (η) 1
lim = lim  = , (9.14)
k→∞ yk − η k→∞ f f −1 −1
(yk ) − f f (η) ′ −1
f f (η)
establishing the case. 
Example 9.10. (a) The function ln : R+ −→ R is differentiable and, for each x ∈ R+ ,
ln′ x = 1/x: If f (x) = ex , then f ′ (x) = ex 6= 0 for each x ∈ R, ln x = f −1 (x), and
(9.13) yields
1 1 1
ln′ x = ′ = ln x = .
f (ln x) e x

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 146

(b) The function arcsin : ] √ − 1, 1[−→] − π/2, π/2[ is differentiable and, for each x ∈

] − 1, 1[, arcsin x = 1/ 1 − x2 : If f (x) = sin x, then f ′ (x) = cos x 6= 0 for each
x ∈] − π/2, π/2[, arcsin x = f −1 (x), and (9.13) yields
1 1 (∗) 1 1
arcsin′ x = ′ = = p =√ ,
f (arcsin x) cos arcsin x 1 − (sin arcsin x) 2 1 − x2
where, at (∗), it was used that cos2 = 1−sin2 and cos t > 0 for each t ∈]−π/2, π/2[.
(c) The function arccos
√ : ] − 1, 1[−→]0, π[ is differentiable and, for each x ∈] − 1, 1[,
arccos′ x = −1/ 1 − x2 : If f (x) = cos x, then f ′ (x) = − sin x 6= 0 for each x ∈]0, π[,
arccos x = f −1 (x), and (9.13) yields
1 1 (∗) 1 1
arccos′ x = ′ = = −p = −√ ,
f (arccos x) − sin arccos x 1 − (cos arccos x)2 1 − x2
where, at (∗), it was used that sin2 = 1 − cos2 and sin t > 0 for each t ∈]0, π[.
(d) The function arctan : R −→] − π/2, π/2[ is differentiable and, for each x ∈ R,
arctan′ x = 1/(1 + x2 ): Apply Th. 9.9 with f (x) = tan x as an exercise.
(e) The function arccot : R −→]0, π[ is differentiable and, for each x ∈ R, arccot′ x =
−1/(1 + x2 ): Apply Th. 9.9 with f (x) = cot x as an exercise.
Theorem 9.11 (Chain Rule). Let a < b, c < d, f : ]a, b[−→ R, g : ]c, d[−→ K,
f (]a, b[) ⊆]c, d[ (a, c = −∞; b, d = ∞ is admissible). If f is differentiable in ξ ∈]a, b[
and g is differentiable in f (ξ) ∈]c, d[, then g ◦ f : ]a, b[−→ K is differentiable in ξ and
(g ◦ f )′ (ξ) = f ′ (ξ)g ′ (f (ξ)). (9.15)

Proof. Let η := f (ξ) and define the auxiliary function


(
g(x)−g(η)
x−η
for x 6= η,
g̃ : ]c, d[−→ K, g̃(x) := ′
(9.16)
g (x) for x = η.
Then
∀ g(x) − g(η) = g̃(x)(x − η). (9.17)
x∈]c,d[

Let (xk )k∈N be a sequence in ]a, b[\{ξ} such that limk→∞ xk = ξ. One obtains
   
g f (xk ) − g f (ξ) (9.17) g̃ f (xk ) f (xk ) − f (ξ)
lim = lim
k→∞ xk − ξ k→∞ xk − ξ
 f (xk ) − f (ξ)
= lim g̃ f (xk ) lim
k→∞ k→∞ xk − ξ
= f ′ (ξ)g ′ (f (ξ)), (9.18)
establishing the case. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 147

Example 9.12. (a) According to the chain rule of Th. 9.11, the function h : R −→ R,
h(x) := sin(−x3 ) is differentiable and, for each x ∈ R, h′ (x) = −3x2 cos(−x3 ).

(b) According to the chain rule of Th. 9.11, each power function h : R+ −→ K, h(x) :=
xα = eα ln x , α ∈ K, is differentiable and, for each x ∈ R+ , h′ (x) = αx eα ln x = α xα−1 .
Indeed, h = g ◦f , where f : R+ −→ R, f (x) := ln x with f ′ : R+ −→ R, f ′ (x) := x1 ,
according to Ex. 9.5(b), and g : R −→ K, g(x) := eαx , with g ′ : R −→ K,
g ′ (x) := α eαx according to Ex. 9.10(a).

9.2 Higher Order Derivatives and the Sets C k


Definition 9.13. Let a < b, I :=]a, b[, f : I −→ K (a = −∞, b = ∞ is admissible). If f
is differentiable, then f ′ might or might not itself be differentiable. If f ′ is differentiable,
then its derivative is denoted by f ′′ and is called the second derivative of f . Clearly, this
process can be iterated, leading to the following general recursive definition of higher-
order derivatives:
Let f (0) := f . For k ∈ N0 assume the kth derivative of f , denoted by f (k) exists on
I. Then f is said to have a derivative of order k + 1 at ξ ∈ I if, and only if, f (k) is
differentiable at ξ. In that case, define

f (k+1) (ξ) := (f (k) )′ (ξ). (9.19)

If f (k+1) (ξ) exists for all ξ ∈ I, then f is said to be (k + 1)-times differentiable and the
function f (k+1) : I −→ K, x 7→ f (k+1) (ξ), is called the (k + 1)st derivative of f . It is
common to write f ′ := f (1) , f ′′ := f (2) , f ′′′ := f (3) , but f (k) if k ≥ 4.
If f (k) exists, it might or might not be continuous (cf. Ex. 9.14(c) below). One defines
n o
k (k)
∀ C (I, K) := f ∈ F(I, K) : f exists and is continuous on I , (9.20)
k∈N0
\
C ∞ (I, K) := C k (I, K) (9.21)
k∈N0

(note C 0 (I, K) = C(I, K) and C(I, K) ⊇ C 1 (I, K) ⊇ C 2 (I, K) ⊇ . . . ). Finally, we define


the notation C k (I) := C k (I, R) for k ∈ N0 ∪ {∞}.

Example 9.14. (a) One has sin ∈ C ∞ (R) with sin′ = cos, sin′′ = − sin, sin′′′ = − cos,
sin(4) = sin, . . .
P
(b) A simple induction shows, for each polynomial P : R −→ K, P (x) = nj=0 aj xj ,
aj ∈ K, n ∈ N0 , that P (n) (x) = n! an . In particular, P ∈ C ∞ (R, K).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 148

(c) It is an exercise to show the following function f is differentiable, but f ′ is not


continuous, i.e. f ∈/ C 1 (R):
( 
x2 cos x1 for x 6= 0,
f : R −→ R, f (x) :=
0 for x = 0.

9.3 Mean Value Theorem, Monotonicity, and Extrema


Theorem 9.15. Let a < b. If f : ]a, b[−→ R is differentiable in ξ ∈]a, b[ and f has a
local min or max in ξ, then f ′ (ξ) = 0.

Proof. Suppose f has a local max at ξ. Then there exists ǫ > 0 such that |h| < ǫ implies
f (ξ + h) − f (ξ) ≤ 0. Now let (hk )k∈N be a sequence in ]0, ǫ[ with limk→∞ hk = 0. Then
f (ξ ± hk ) − f (ξ) ≤ 0 for all k ∈ N implies
f (ξ + hk ) − f (ξ) f (ξ − hk ) − f (ξ)
f ′ (ξ) = lim ≤ 0, f ′ (ξ) = lim ≥ 0, (9.22)
k→∞ hk k→∞ −hk
showing f ′ (ξ) = 0. Now, if f has a local min at ξ, then −f has a local max at ξ, and
f ′ (ξ) = −(−f )′ (ξ) = 0 establishes the case. 
Remark 9.16. For f : R −→ R, f (x) := x3 , it is f ′ (0) = 0, but f does not have a local
min or max at 0, showing that, while being necessary for an differentiable function f to
have a local extremum at ξ, f ′ (ξ) = 0 is not a sufficient condition for such an extremum
at ξ. Points ξ with f ′ (ξ) = 0 are sometimes called stationary or critical points of f .

Now, we first prove an important special case of the mean value theorem:
Theorem 9.17 (Rolle’s Theorem). Let a < b. If f : [a, b] −→ R is continuous on the
compact interval [a, b], differentiable on the open interval ]a, b[, and f (a) = f (b), then
there exists ξ ∈]a, b[ such that f ′ (ξ) = 0.

Proof. If f is constant, then f ′ (ξ) = 0 holds for each ξ ∈]a, b[. If f is nonconstant,
then there exists x ∈]a, b[ with f (x) 6= f (a). If f (x) > f (a), then Th. 7.54 implies the
existence of ξ ∈]a, b[ such that f attains its (global and, thus, local) max in ξ. Then
Th. 9.15 yields f ′ (ξ) = 0. The case f (x) < f (a) is treated analogously. 
Theorem 9.18 (Mean Value Theorem). Let a < b. If f, g : [a, b] −→ R are continuous
on the compact interval [a, b], differentiable on the open interval ]a, b[, and g ′ (x) 6= 0 for
each x ∈]a, b[, then there exists ξ ∈]a, b[ such that
f (b) − f (a) f ′ (ξ)
= ′ . (9.23a)
g(b) − g(a) g (ξ)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 149

In the special case that g : [a, b] −→ R, g(x) = x, one obtains the standard form

f (b) − f (a)
= f ′ (ξ). (9.23b)
b−a

Proof. First note that Rolle’s Th. 9.17 and g ′ 6= 0 imply g(b) − g(a) 6= 0. Next, one
applies Rolle’s Th. 9.17 to the auxiliary function
 f (b) − f (a)
h : [a, b] −→ R, h(x) := f (x) − g(x) − g(a) . (9.24)
g(b) − g(a)

Since f and g are continuous on [a, b] and differentiable on ]a, b[, so is h. Moreover,
h(a) = f (a) = h(b), i.e. Rolle’s Th. 9.17 applies and yields ξ ∈]a, b[ satisfying h′ (ξ) = 0.
However, (9.24) implies h′ (ξ) = 0 is equivalent to (9.23a). 

Corollary 9.19. Let c < d and f : ]c, d[−→ R be differentiable (c = −∞, d = ∞ is


admissible).

(a) If f ′ ≥ 0 (resp. f ′ ≤ 0), then f is increasing (resp. decreasing). Moreover, if the


inequalities are strict, then the monotonicity of f is strict as well.

(b) If f ′ ≡ 0, then f is constant.

Proof. If c < a < b < d and f ′ ≥ 0 (resp. f ′ ≤ 0, resp. f ′ ≡ 0), then (9.23b) implies
f (b) ≥ f (a) (resp. f (b) ≤ f (a), resp. f (b) = f (a)). Moreover, strict inequalities for f ′
yield strict inequality between f (b) and f (a). 

Lemma 9.20. Let a < b, f : ]a, b[−→ R, ξ ∈]a, b[, and assume f is differentiable at ξ.
If f ′ (ξ) > 0 (resp. f ′ (ξ) < 0), then there exists ǫ > 0 such that ]ξ − ǫ, ξ + ǫ[⊆]a, b[ and

∀ ∀ f (a1 ) < f (ξ) < f (b1 ) resp. f (a1 ) > f (ξ) > f (b1 ) .
a1 ∈]ξ−ǫ,ξ[ b1 ∈]ξ,ξ+ǫ[

Proof. If there does not exist ǫ > 0 such that f (a1 ) < f (ξ) < f (b1 ) for each a1 ∈]ξ − ǫ, ξ[
and each b1 ∈]ξ, ξ + ǫ[, then then there exists a sequence (xk )k∈N in ]a, b[\{ξ} such that
limk→∞ xk = ξ and
f (xk ) − f (ξ)
∀ ≤ 0,
k∈N xk − ξ
showing f ′ (ξ) ≤ 0. Analogously, one obtains that f ′ (ξ) ≥ 0 provided there does not exist
ǫ > 0 such that f (a1 ) > f (ξ) > f (b1 ) for each a1 ∈]ξ − ǫ, ξ[ and each b1 ∈]ξ, ξ + ǫ[. 

Caveat 9.21. The hypotheses of Lem. 9.20 are not sufficient for f to be increasing or
decreasing in any neighborhood of ξ: It is an exercise to find a counterexample.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 150

Theorem 9.22 (Sufficient Conditions for Extrema). Let c < d, let f : ]c, d[−→ R be
differentiable, and assume f ′ (ξ) = 0 for some ξ ∈]c, d[.

(a) If f ′ (x) > 0 for each x ∈]c, ξ[ and f ′ (x) < 0 for each x ∈]ξ, d[, then f has a strict
max at ξ. Likewise, if f ′′ (ξ) exists and is negative, then f has a strict max at ξ.
(b) If f ′ (x) < 0 for each x ∈]c, ξ[ and f ′ (x) > 0 for each x ∈]ξ, d[, then f has a strict
min at ξ. Likewise, if f ′′ (ξ) exists and is positive, then f has a strict min at ξ.

Proof. We just present the proof for (a); (b) is proved analogously. If f ′ (x) > 0 for each
x ∈]c, ξ[, then (9.23b) shows f (ξ)−f (a) > 0 for each c < a < ξ; analogously, if f ′ (x) < 0
for each x ∈]ξ, d[, then (9.23b) shows f (ξ) − f (b) > 0 for each ξ < b < d. Altogether, we
have shown f to have a strict max at ξ. If f ′′ (ξ) exists and is negative, then Lem. 9.20
yields the existence of ǫ > 0 such that f ′ is positive on ]ξ − ǫ, ξ[ and negative on ]ξ, ξ + ǫ[.
Applying what we have already proved with c := ξ − ǫ and d := ξ + ǫ establishes the
case. 
Example 9.23. One obtains
f : R −→ R, f (x) := x ex , (9.25a)
f ′ : R −→ R, f ′ (x) = ex + x ex = (1 + x) ex , (9.25b)
f ′′ : R −→ R, f ′′ (x) = 2ex + x ex = (2 + x) ex . (9.25c)
From Th. 9.15, we know that f can have at most one extremum, namely at ξ = −1,
where f ′ (ξ) = 0. Since f ′′ (ξ) = e−x > 0, Th. 9.22(b) implies that f has a strict min at
−1.

The following Th. 9.24, the intermediate value theorem for derivatives, is another ap-
plication of Lem. 9.20. Even though Ex. 9.14(c) shows that derivatives do not have
to be continuous, not every function can occur as a derivative: The following Th. 9.24
shows that derivatives always satisfy an intermediate value property, even if they are
not continuous (that also means that discontinuities in derivatives are always due to
oscillations rather than jumps).
Theorem 9.24 (Intermediate Value Theorem for Derivatives). Let a, b, c, d ∈ R with
a < c ≤ d < b. If f : ]a, b[−→ R is differentiable, then f ′ assumes every value between
f ′ (c) and f ′ (d), i.e.
h i 
min{f ′ (c), f ′ (d)}, max{f ′ (c), f ′ (d)} ⊆ f ′ [c, d] . (9.26)

Proof. Exercise (hint: use a suitable auxiliary function and apply Lem. 9.20 together
with Th. 7.54 and Th. 9.15). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 151

9.4 L’Hôpital’s Rule


L’Hôpital’s rule is a result that can help to determine (function) limits (cf. Def. 8.17).
Theorem 9.25 (L’Hôpital’s Rule). Let ξ ∈ R and either I =]a, ξ[ with a < ξ or
I :=]ξ, b[ with ξ < b. Moreover, assume f, g : I −→ R are differentiable, g ′ (x) 6= 0 for
each x ∈ I, and one of the following two conditions (a), (b) is satisfied:

(a) limx→ξ f (x) = limx→ξ g(x) = 0.

(b) limx→ξ g(x) = ∞ or limx→ξ g(x) = −∞, where Def. 8.17 is extended to the case
η ∈ {−∞, ∞} in the obvious way.

Then
f ′ (x) f (x)
lim =η ⇒ lim = η. (9.27)
x→ξ g ′ (x) x→ξ g(x)

The above statement also holds for ξ ∈ {−∞, ∞} and/or η ∈ {−∞, ∞} if, as in (b),
one extends Def. 8.17 to these cases in the obvious way.

Proof. First, we assume (a). Consider the case ξ ∈ R. Since f and g are continuous, (a)
implies f and g remain continuous, if we extend them to ξ by letting f (ξ) := g(ξ) = 0.
This extension will now allow us to apply Th. 9.18 to f and g. To prove (9.27), let
(xk )k∈N be a sequence in I with limk→∞ xk = ξ. Then (9.23a) yields, for each k ∈ N,
some ξk ∈]xk , ξ[ if xk < ξ and some ξk ∈]ξ, xk [ if ξ < xk , satisfying

f (xk ) f (xk ) − f (ξ) f ′ (ξk )


= = ′ . (9.28)
g(xk ) g(xk ) − g(ξ) g (ξk )
f ′ (x)
From the Sandwich Th. 7.16, we obtain limk→∞ ξk = ξ, i.e. (9.28) and limx→ξ g ′ (x)

imply limx→ξ fg(x)
(x)
= η (also for η ∈ {−∞, ∞}). Now consider the case ξ ∈ {−∞, ∞}
and let (xk )k∈N be as before. If ξ = ∞, then choose 1 ≤ c ∈ I and set I˜ :=]0, c−1 [; if
ξ = −∞, then choose −1 ≥ c ∈ I and set I˜ :=]c−1 , 0[. We apply what we have already
proved above to the auxiliary functions

f˜ : I˜ −→ R, f˜(x) := f (1/x), g̃ : I˜ −→ R, g̃(x) := g(1/x)


′ ′
at ξ˜ := 0. From the chain rule (9.15), we know f˜′ (x) = − f (1/x)
x2
and g̃ ′ (x) = − g (1/x)
x2
for
′ (x)
each x ∈ I.˜ Thus, limx→ξ ′ = η implies,
f
g (x)

f ′ (xk ) −x2k f ′ (xk ) f˜′ (1/xk ) f˜(1/xk ) f (xk )


η = lim ′
= lim 2 ′
= lim ′
= lim = lim ,
k→∞ g (xk ) k→∞ −x g (xk ) k→∞ g̃ (1/xk ) k→∞ g̃(1/xk ) k→∞ g(xk )
k

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 152

f (x)
proving limx→ξ g(x)
= η.
We now assume (b), still letting (xk )k∈N be as before. Note that g ′ 6= 0 implies g
is injective by Rolle’s Th. 9.17. Then the intermediate value theorem implies g is
either strictly increasing or strictly decreasing. We proceed with the proof for the case
I =]a, ξ[, the proof for I =]ξ, b[ can be done completely analogous. We first consider
the case where g is strictly increasing, i.e. limx→ξ g(x) = ∞. Assume η ∈ R and ǫ > 0.
′ (x)
Then limx→ξ fg′ (x) = η and limx→ξ g(x) = ∞ imply
 
ǫ f ′ (x) ǫ
∃ ∀ g(x) > 0 ∧ η− < ′ <η+ .
c∈]a,ξ[ x∈]c,ξ[ 2 g (x) 2

Since limk→∞ xk = ξ, there exists N0 ∈ N such that, for each k > N0 , c < xk < ξ. Next,
according to Th. 9.18,

ǫ f (xk ) − f (c) f ′ (ξk ) ǫ


∀ ∃ η− < = ′ <η+ .
k>N0 ξk ∈]c,xk [ 2 g(xk ) − g(c) g (ξk ) 2

In consequence, using g(xk ) > g(c), as g is strictly increasing,


 ǫ  ǫ
∀ η− (g(xk ) − g(c)) < f (xk ) − f (c) < η + (g(xk ) − g(c))
k>N0 2 2
and
 
 ǫ  f (c) − η − 2ǫ g(c) f (xk )  ǫ  f (c) − η + 2ǫ g(c)
∀ η− + < < η+ + .
k>N0 2 g(xk ) g(xk ) 2 g(xk )

Since limk→∞ g(xk ) = ∞,


!
f (c) − η − ǫ  g(c) ǫ f (c) − η + ǫ  g(c) ǫ
2 2
∃ ∀ < ∧ < ,
N ≥N0 k>N g(xk ) 2 g(xk ) 2

that means
f (xk )
∀ η−ǫ< < η + ǫ,
k>N g(xk )
f (x)
proving limx→ξ g(x)
= η. For η = ∞ and given n ∈ N, the argument is similar:
′ (x)
limx→ξ fg′ (x) = η and limx→ξ g(x) = ∞ imply
 
f ′ (x)
∃ ∀ g(x) > 0 ∧ n< ′ .
c∈]a,ξ[ x∈]c,ξ[ g (x)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 153

As before, since limk→∞ xk = ξ, there exists N0 ∈ N such that, for each k > N0 ,
c < xk < ξ. Again, according to Th. 9.18,

f (xk ) − f (c) f ′ (ξk )


∀ ∃ n< = ′ .
k>N0 ξk ∈]c,xk [ g(xk ) − g(c) g (ξk )

In consequence, using g(xk ) > g(c), as g is strictly increasing,

∀ n (g(xk ) − g(c)) < f (xk ) − f (c)


k>N0

and
f (c) − n g(c) f (xk )
∀ n+ < .
k>N0 g(xk ) g(xk )
Since limk→∞ g(xk ) = ∞,

f (c) − n g(c)
∃ ∀ < 1,
N ≥N0 k>N g(xk )

that means
f (xk )
∀ n−1< ,
k>N g(xk )
f (x)
proving limx→ξ g(x)
= η. If η = −∞, then, using what we have already shown,

f ′ (x) −f ′ (x) −f (x) f (x)


lim =η ⇒ lim = ∞ = lim ⇒ lim = η.
x→ξ g ′ (x) ′
x→ξ g (x) x→ξ g(x) x→ξ g(x)

Finally, if g strictly decreasing, then −g is strictly increasing and we obtain

f ′ (x) f ′ (x) f (x) f (x)


lim =η ⇒ lim = −η = lim ⇒ lim = η,
x→ξ g ′ (x) ′
x→ξ −g (x) x→ξ −g(x) x→ξ g(x)

concluding the proof. 

Example 9.26. (a) Applying L’Hôpital’s rule to f : ] − π/2, π/2[−→ R, f (x) := tan x,
g : ] − π/2, π/2[−→ R, g(x) := ex − 1, with ξ = 0 yields

tan x 1 + tan2 x 1
lim x
= lim x
= =1 (9.29)
x→0 e − 1 x→0 e 1
(note g ′ (x) = ex 6= 0 for each x ∈] − π/2, π/2[).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 154

(b) It can happen that a single application of L’Hôpital’s rule does not, yet, yield a
useful result, but that a repeated application does. An example is provided by
considering α > 0, n ∈ N, and f : R+ −→ R, f (x) := eα x , g : R+ −→ R,
g(x) := xn , ξ := ∞. Applying L’Hôpital’s rule n times yields
eα x α n eα x
∀+ ∀ lim = lim =∞ (9.30)
α∈R n∈N x→∞ xn x→∞ n!
(note g (k) (x) = n(n − 1) · · · (n − k + 1)xn−k 6= 0 for each k ∈ {1, . . . , n} and each
x ∈ R+ ).
(c) It can also happen that even repeated applications of L’Hôpital’s rule do not
help at all, even though limx→ξ fg(x)(x)
does exist and the hypotheses of Th. 9.25
are all satisfied. A simple example is given by f : R −→ R, f (x) := ex , g :
R −→ R, g(x) := 2ex , and ξ = −∞. Even though limx→−∞ fg(x) (x)
= 21 , one has
limx→−∞ f (n) (x) = limx→−∞ g (n) (x) = 0 for every n ∈ N.

9.5 Convex Functions


In the present section, we provide an introduction to (one-dimensional) convex functions.
They have many important applications, some of which we well need in Analysis II, when
studying so-called norms on Kn .
The idea is to call a function f : I −→ R (where I ⊆ R is an interval) convex if, and
only if, each line segment connecting two points on the graph of f lies above this graph,
and to call f concave if, and only if, each such line segment lies below the graph of f .
Noting that, for x1 < x2 , the line through the two points (x1 , f (x1 )) and (x2 , f (x2 )) is
represented by the equation
x2 − x x − x1
L(x) = f (x1 ) + f (x2 ), (9.31)
x2 − x1 x2 − x1
this leads to the following definition:
Definition 9.27. Let I ⊆ R be an interval (I can be open, closed, or half-open, it can
be for finite or of infinite length) and f : I −→ R. Then f is called convex if, and only
if, for each x1 , x, x2 ∈ I such that x1 < x < x2 , one has
x2 − x x − x1
f (x) ≤ f (x1 ) + f (x2 ); (9.32a)
x2 − x1 x2 − x1
f is called concave if, and only if, for each x1 , x, x2 ∈ I such that x1 < x < x2 , one has
x2 − x x − x1
f (x) ≥ f (x1 ) + f (x2 ). (9.32b)
x2 − x1 x2 − x1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 155

Moreover, f is called strictly convex (resp. strictly concave) if, and only if, (9.32a) (resp.
(9.32b)) always holds with strict inequality.

Lemma 9.28. Let I ⊆ R be an interval. Then f : I −→ R is (strictly) concave if, and


only if, −f is (strictly) convex.

Proof. Merely multiply (9.32b) by (−1) and compare with (9.32a). 

The following Prop. 9.29 provides equivalences for convexity. One can easily obtain the
corresponding equivalences for concavity by combining Prop. 9.29 with Lem. 9.28.

Proposition 9.29. Let I ⊆ R be an interval and f : I −→ R. Then the following


statements are equivalent:

(i) f is (strictly) convex.

(ii) For each a, b ∈ I such that a 6= b and each λ ∈]0, 1[, the following estimate holds
(with strict inequality):

f λa + (1 − λ)b ≤ λf (a) + (1 − λ)f (b). (9.33)

(iii) For each x1 , x, x2 ∈ I such that x1 < x < x2 , one has (with strict inequality)

f (x) − f (x1 ) f (x2 ) − f (x)


≤ . (9.34)
x − x1 x2 − x

(iv) For each x1 , x, x2 ∈ I such that x1 < x < x2 , one has (with strict inequality)

f (x) − f (x1 ) f (x2 ) − f (x1 ) f (x2 ) − f (x)


≤ ≤ . (9.35)
x − x1 x2 − x1 x2 − x

Proof. We leave the proof as an exercise. Suggestion: Establish the following implica-
tions: (i) ⇔ (ii), (i) ⇔ (iii), (iv) ⇒ (iii), (i) ⇒ (iv). 

Example 9.30. Since |λx + (1 − λ)y| ≤ λ|x| + (1 − λ)|y| for each 0 < λ < 1 and each
x, y ∈ R, the absolute value function is convex. This example also shows that a convex
function does not need to be differentiable.

For differentiable functions, one can formulate convexity criteria in terms of the deriva-
tive:

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 156

Proposition 9.31. Let a < b, and suppose that f : [a, b] −→ R is continuous on [a, b]
and differentiable on ]a, b[. Then f is (strictly) convex (resp. (strictly) concave) on [a, b]
if, and only if, the derivative f ′ is (strictly) increasing (resp. (strictly) decreasing) on
]a, b[.

Proof. Since (−f )′ = −f ′ and −f ′ is (strictly) increasing if, and only if, f ′ is (strictly)
decreasing, it suffices to consider the (strictly) convex case. So assume that f is (strictly)
convex. Then for each x1 , x, x0 , y, x2 ∈]a, b[ such that x1 < x < x0 < y < x2 , applying
Prop. 9.29(iv), one has (with strict inequalities),

f (x) − f (x1 ) f (x0 ) − f (x1 ) f (x2 ) − f (x1 ) f (x2 ) − f (y)


≤ ≤ ≤ . (9.36)
x − x1 x0 − x1 x2 − x1 x2 − y
Thus,

f (x) − f (x1 ) f (x0 ) − f (x1 ) (∗) f (x2 ) − f (x1 )


f ′ (x1 ) = lim ≤ ≤
x↓x1 x − x1 x0 − x1 x2 − x1
f (x2 ) − f (y)
≤ lim = f ′ (x2 ) (9.37)
y↑x2 x2 − y
(where the inequality at (∗) is strict if it is strict in (9.36)), showing that f ′ is (strictly)
increasing on ]a, b[. On the other hand, if f ′ is (strictly) increasing on ]a, b[, then for
each x1 , x, x2 ∈ [a, b] such that x1 < x < x2 , Th. 9.18 yields ξ1 ∈]x1 , x[ and ξ2 ∈]x, x2 [
such that
f (x) − f (x1 ) f (x2 ) − f (x)
= f ′ (ξ1 ) and = f ′ (ξ2 ). (9.38)
x − x1 x2 − x
As ξ1 < ξ2 and f ′ is (strictly) increasing, (9.38) implies (9.34) (with strict inequality)
and, thus, the (strict) convexity of f . 
Proposition 9.32. Let a < b, and suppose that f : [a, b] −→ R is continuous on [a, b]
and twice differentiable on ]a, b[.

(a) f is convex (resp. concave) on [a, b] if, and only if, f ′′ ≥ 0 (resp. f ′′ ≤ 0) on ]a, b[.

(b) If f ′′ > 0 (resp. f ′′ < 0) on ]a, b[, then f is strictly convex (resp. strictly concave)
(as a caveat we remark that, here, the converse does not hold – for example x 7→ x4
is strictly convex, but its second derivative x 7→ 12x2 is 0 at x = 0).

Proof. Since −f ′′ ≥ 0 if, and only if f ′′ ≤ 0; and −f ′′ > 0 if, and only if f ′′ < 0, if
suffices to consider the convex cases. Moreover, for (a), one merely has to combine Prop.
9.31 with the fact that f ′ is increasing on ]a, b[ if, and only if, f ′′ ≥ 0 on ]a, b[. The
proof of (b) is left as an exercise. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
9 DIFFERENTIAL CALCULUS 157

Example 9.33. (a) Since for f : R −→ R, f (x) = ex , it is f ′′ (x) = ex > 0, the


exponential function is strictly convex on R.

(b) Since for f : R+ −→ R, f (x) = ln x, it is f ′′ (x) = −1/x2 < 0, the natural logarithm
is strictly concave on R+ .

Theorem 9.34 (Jensen’s inequality). Let I ⊆ R be an interval and let f : I −→ R be


convex. If n ∈ N and λ1 , . . . , λn > 0 such that λ1 + · · · + λn = 1, then

∀ f (λ1 x1 + · · · + λn xn ) ≤ λ1 f (x1 ) + · · · + λn f (xn ). (9.39a)


x1 ,...,xn ∈I

If f is concave, then

∀ f (λ1 x1 + · · · + λn xn ) ≥ λ1 f (x1 ) + · · · + λn f (xn ). (9.39b)


x1 ,...,xn ∈I

If f is strictly convex or strictly concave, then equality in the above inequalities can only
hold if x1 = · · · = xn .

Proof. If one lets a := min{x1 , . . . , xn }, b := max{x1 , . . . , xn }, and x̄ := λ1 x1 +· · ·+λn xn ,


then n n
X X
a= λj a ≤ x̄ ≤ λj b = b ⇒ x̄ ∈ I. (9.40)
j=1 j=1

Since f is (strictly) concave if, and only if, −f is (strictly) convex, it suffices to consider
the cases where f is convex and where f is strictly convex. Thus, we assume that f is
convex and prove (9.39a) by induction. For n = 1, one has λ1 = 1 and there is nothing
to prove. For n = 2, (9.39a) reduces to (9.33), which holds due to the convexity of f .
Finally, let n > 2 and assume that (9.39a) already holds for each 1 ≤ l ≤ n − 1. Set

λ1 λn−1
λ := λ1 + · · · + λn−1 , x := x1 + · · · + xn−1 . (9.41)
λ λ
Then x ∈ I follows as in (9.40). One computes
n−1
!
X
f (λ1 x1 + · · · + λn xn ) = f λj xj + λn xn = f (λx + λn xn )
j=1
n−1
l=2 l=n−1 X λj
≤ λf (x) + λn f (xn ) ≤ λ f (xj ) + λn f (xn )
j=1
λ
= λ1 f (x1 ) + · · · + λn f (xn ), (9.42)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 158

thereby completing the induction, and, thus, the proof of (9.39a). If f is strictly convex
and (9.39a) holds with equality, then one can also proceed by induction to prove the
equality of the xj . Again, if n = 1, then there is nothing to prove. If n = 2, and x1 6= x2 ,
then strict convexity requires (9.39a) to hold with strict inequality. Thus x1 = x2 . Now
let n > 2. It is noted that (9.42) still holds. By hypothesis, the first and last term
in (9.42) are now equal, implying that all terms in (9.42) must be equal. Using the
induction hypothesis for l = 2 and the corresponding equality in (9.42), we conclude
that x = xn . Using the induction hypothesis for l = n−1 and the corresponding equality
in (9.42), we conclude that x1 = · · · = xn−1 . Finally, x = xn and x1 = · · · = xn−1 are
combined using (9.41) to get x1 = xn , finishing the proof of the theorem. 

Theorem 9.35 (Inequality Between the Weighted Arithmetic Mean and the Weighted
Geometric Mean). If n ∈ N, x1 , . . . , xn ≥ 0 and λ1 , . . . , λn > 0 such that λ1 +· · ·+λn = 1,
then
xλ1 1 · · · xλnn ≤ λ1 x1 + · · · + λn xn , (9.43)
where equality occurs if, and only if, x1 = · · · = xn . In particular, for λ1 = · · · = λn =
1
n
, one recovers the inequality between the arithmetic and the geometric mean without
weights, known from Th. 7.63.

Proof. If at least one of the xj is 0, then (9.43) becomes the true statement 0 ≤
P n
j=1 λj xj with strict inequality if, and only if, at least one xj > 0. Thus, it remains
to consider the case x1 , . . . , xn > 0. As we noted in Ex. 9.33(b), the natural logarithm
ln : R+ −→ R is concave and even strictly concave. Employing Jensen’s inequality
(9.39b) yields

ln(λ1 x1 + · · · + λn xn ) ≥ λ1 ln x1 + · · · + λn ln xn = ln(xλ1 1 · · · xλnn ). (9.44)

Applying the exponential function to both sides of (9.44), one obtains (9.43). Since
(9.44) is equivalent to (9.43), the strict concavity of ln yields that equality in (9.44)
implies x1 = · · · = xn . 

10 The Riemann Integral on Intervals in R

10.1 Definition and Simple Properties


R
Given a nonnegative function f : M −→ R+ 0 , M ⊆ R, we aim to compute the area M f
of the set “under the graph” of f , i.e. of the set

(x, y) ∈ R2 : x ∈ M and 0 ≤ y ≤ f (x) . (10.1)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 159

R
This area M f (if it exists) will be called the integral of f over M . Moreover, for
functions f : M −→ R that are not necessarily nonnegative, we would like to count
areas of sets of the form (10.1)
(which are below the graph of f and above the set

M = (x, 0) ∈ R : x ∈ M ⊆ R2 ) with a positive sign, and whereas we would like to
2

count areas of sets above the graph of f and below the set M with a negative sign. In
other words, making use of the positive and negative parts f + and f − of f = f + − f −
as defined in (6.1i) and (6.1j), respectively, we would like our integral to satisfy
Z Z Z
f= +
f − f −. (10.2)
M M M

Difficulties arise from the fact that both the function f and the set M can be extremely
complicated. To avoid dealing with complicated sets M , we restrict ourselves to the
situation of integrals over compact intervals, i.e. to integrals over sets of the form M =
[a, b]. Moreover, we will also restrict ourselves to bounded functions f , which we now
define:

Definition 10.1. Let ∅ 6= M ⊆ R and f : M −→ R. Then f is called bounded if, and


only if, the set {|f (x)| : x ∈ M } ⊆ R+
0 is bounded, i.e. if, and only if,

kf ksup := sup{|f (x)| : x ∈ M } ∈ R+


0. (10.3)
R
The basic idea for the definition of the Riemann integral M f Ris rather simple: De-
compose the set M into small pieces I1 , . . . , IN and approximate M f by the finite sum
PN R
j=1 f (xj )|Ij |, where xj ∈ Ij and |Ij | denotes the size of the set Ij . Define M f as the
limit of such sums as the size of the Ij tends to zero (if the limit exists). However, to
carry out this idea precisely and rigorously does require some work.
As stated before, we will assume that M is a closed bounded interval, and we will choose
the Ij to be closed bounded intervals as well. To emphasize we are dealing with intervals,
in the following, we will prefer to use the symbol I instead of M .

Definition 10.2. If a, b ∈ R, a ≤ b, and I := [a, b], then we call

|I| := b − a = |a − b|, (10.4)

the length or the (1-dimensional) size, volume, or measure of I.

Definition 10.3. Given a real interval I := [a, b] ⊆ R, a, b ∈ R, a < b, the (N + 1)-tuple


∆ := (x0 , . . . , xN ) ∈ RN +1 , N ∈ N, is called a partition of I if, and only if, a = x0 <
x1 < · · · < xN = b. We call x0 , . . . , xN the nodes of ∆, and let ν(∆) := {x0 , . . . , xN }
be the set of all nodes. A tagged partition of I is a partition together with an N -tuple

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 160

(t1 , . . . , tN ) ∈ RN such that tj ∈ [xj−1 , xj ] for each j ∈ {1, . . . , N }. Given a partition ∆


(with or without tags) of I as above and letting Ij := [xj−1 , xj ], the number

|∆| := max |Ij | : j ∈ {1, . . . , N } , (10.5)
is called the mesh size of ∆. It is sometimes convenient, if we extend our definitions to
trivial intervals, consisting of just one point: For a = b, we have I = [a, a] = {a}. We
then define ∆ = x0 = a to be a partition of I, ν(∆) = {x0 }, and a is then the only tag
that makes ∆ into a tagged partition. We also set I0 := I = {a}, and the mesh size in
this case is |∆| := 0.
Definition 10.4. Let ∆ be a partition of I = [a, b] ⊆ R, a ≤ b, as in Def. 10.3. Given
a function f : I −→ R that is bounded according to Def. 10.1, define
mj := mj (f ) := inf{f (x) : x ∈ Ij }, Mj := Mj (f ) := sup{f (x) : x ∈ Ij }, (10.6)
and
N
X N
X
r(∆, f ) := mj |Ij | = mj (xj − xj−1 ), (10.7a)
j=1 j=1
N
X N
X
R(∆, f ) := Mj |Ij | = Mj (xj − xj−1 ), (10.7b)
j=1 j=1

where r(∆, f ) is called the lower Riemann sum and R(∆, f ) is called the upper Riemann
sum associated with ∆ and f . If ∆ is tagged by τ := (t1 , . . . , tN ), then we also define
the intermediate Riemann sum
N
X N
X
ρ(∆, f ) := f (tj ) |Ij | = f (tj )(xj − xj−1 ). (10.7c)
j=1 j=1

Note that, for a = b, all the above sums are empty and we have r(∆, f ) = R(∆, f ) =
ρ(∆, f ) = 0.
Definition 10.5. Let I = [a, b] ⊆ R be an interval, a ≤ b, and suppose f : I −→ R is
bounded.

(a) Define

J∗ (f, I) := sup r(∆, f ) : ∆ is a partition of I , (10.8a)

J ∗ (f, I) := inf R(∆, f ) : ∆ is a partition of I . (10.8b)
We call J∗ (f, I) the lower Riemann integral of f over I and J ∗ (f, I) the upper
Riemann integral of f over I.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 161

(b) The function f is called Riemann integrable over I if, and only if, J∗ (f, I) = J ∗ (f, I).
If f is Riemann integrable over I, then
Z b Z Z b Z
f (x) dx := f (x) dx := f := f := J∗ (f, I) = J ∗ (f, I) (10.9)
a I a I

is called the Riemann integral of f over I. The set of all functions f : I −→ R that
are Riemann integrable over I is denoted by R(I, R) or just by R(I).

(c) The function g : I −→ C is called Riemann integrable over I if, and only if, both
Re g and Im g are Riemann integrable. The set of all Riemann integrable functions
g : I −→ C is denoted by R(I, C). If g ∈ R(I, C), then
Z Z Z  Z Z
g := Re g, Im g = Re g + i Im g ∈ C (10.10)
I I I I I

is called the Riemann integral of g over I.


Remark 10.6. If I = [a, b] ⊆ R, ∆ is a partition of I, and f : I −→ R is bounded,
then (10.6) implies
(4.6c) (4.6d)
mj (f ) = −Mj (−f ) and mj (−f ) = −Mj (f ), (10.11a)

(10.7) implies

r(∆, f ) = −R(∆, −f ) and r(∆, −f ) = −R(∆, f ), (10.11b)

and (10.8) implies

J∗ (f, I) = −J ∗ (−f, I) and J∗ (−f, I) = −J ∗ (f, I). (10.11c)

Example 10.7. (a) If I = [a, b] ⊆ R as before and f : I −→ R is constant, i.e. f ≡ c


with c ∈ R, then f ∈ R(I) and
Z b
f = c (b − a) = c |I| : (10.12)
a

We have, for each partition ∆ of I,


N
X N
X N
X
r(∆, f ) = mj |Ij | = c |Ij | = c |I| = c (b − a) = Mj |Ij | = R(∆, f ),
j=1 j=1 j=1
(10.13)
proving J∗ (f, I) = c (b − a) = J ∗ (f, I).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 162

(b) An example of a function that is not Riemann integrable for a < b is given by the
Dirichlet function
(
0 for x irrational,
f : [a, b] −→ R, f (x) := a < b. (10.14)
1 for x rational,
P
Since r(∆, f ) = 0 and R(∆, f ) = N j=1 |Ij | = b − a for every partition ∆ of I, one

obtains J∗ (f, I) = 0 6= (b − a) = J (f, I), showing that f ∈ / R(I).

Definition 10.8. (a) If ∆ is a partition of [a, b] ⊆ R as in Def. 10.3, then another


partition ∆′ of [a, b] is called a refinement of ∆ if, and only if, ν(∆) ⊆ ν(∆′ ), i.e. if,
and only if, the nodes of ∆′ include all the nodes of ∆.

(b) If ∆ and ∆′ are partitions of [a, b] ⊆ R, then the superposition of ∆ and ∆′ , denoted
∆ + ∆′ , is the unique partition of [a, b] having ν(∆) ∪ ν(∆′ ) as its set of nodes. Note
that the superposition of ∆ and ∆′ is always a common refinement of ∆ and ∆′ .

Lemma 10.9. Let a, b ∈ R, a < b, I := [a, b], and suppose f : I −→ R is bounded with
M := kf ksup ∈ R+ ′
0 . Let ∆ be a partition of I and assume
 

α := # ν(∆ ) \ {a, b} ≥ 1 (10.15)

is the number of interior nodes that occur in ∆′ . Then, for each partition ∆ of I, the
following holds:

r(∆, f ) ≤ r(∆ + ∆′ , f ) ≤ r(∆, f ) + 2 α M |∆|, (10.16a)


R(∆, f ) ≥ R(∆ + ∆′ , f ) ≥ R(∆, f ) − 2 α M |∆|. (10.16b)

Proof. We carry out the proof of (10.16a) – the proof of (10.16b) can be conducted
completely analogous. Consider the case α = 1 and let ξ be the single element of
ν(∆′ ) \ {a, b}. If ξ ∈ ν(∆), then ∆ + ∆′ = ∆, and (10.16a) is trivially true. If ξ ∈
/ ν(∆),
then xk−1 < ξ < xk for a suitable k ∈ {1, . . . , N }. Define

I ′ := [xk−1 , ξ], I ′′ := [ξ, xk ] (10.17)

and
m′ := inf{f (x) : x ∈ I ′ }, m′′ := inf{f (x) : x ∈ I ′′ }. (10.18)
Then we obtain

r(∆ + ∆′ , f ) − r(∆, f ) = m′ |I ′ | + m′′ |I ′′ | − mk |Ik | = (m′ − mk ) |I ′ | + (m′′ − mk ) |I ′′ |.


(10.19)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 163

Together with the observation

0 ≤ m′ − mk ≤ 2M, 0 ≤ m′′ − mk ≤ 2M, (10.20)

(10.19) implies

0 ≤ r(∆ + ∆′ , f ) − r(∆, f ) ≤ 2M |I ′ | + |I ′′ | ≤ 2M |∆|. (10.21)

The general form of (10.16a) follows by an induction on α. 


Theorem 10.10. Let a, b ∈ R, a ≤ b, I := [a, b], and let f : I −→ R be bounded.

(a) Suppose ∆ and ∆′ are partitions of I such that ∆′ is a refinement of ∆. Then

r(∆, f ) ≤ r(∆′ , f ), R(∆, f ) ≥ R(∆′ , f ). (10.22)

(b) For arbitrary partitions ∆ and ∆′ , the following holds:

r(∆, f ) ≤ R(∆′ , f ). (10.23)

(c) J∗ (f, I) ≤ J ∗ (f, I).

(d) For each sequence of partitions (∆n )n∈N of I such that limn→∞ |∆n | = 0, one has

lim r(∆n , f ) = J∗ (f, I), lim R(∆n , f ) = J ∗ (f, I). (10.24)


n→∞ n→∞

In particular, if f ∈ R(I), then


Z
lim r(∆n , f ) = lim R(∆n , f ) = f, (10.25a)
n→∞ n→∞ I

and if f ∈ R(I) and the ∆n are tagged, then also


Z
lim ρ(∆n , f ) = f. (10.25b)
n→∞ I

Proof. (a): If ∆′ is a refinement of ∆, then ∆′ = ∆ + ∆′ . Thus, (10.22) is immediate


from (10.16).
(b): This also follows from (10.16):
(10.16a) (10.7) (10.16b)
r(∆, f ) ≤ r(∆ + ∆′ , f ) ≤ R(∆ + ∆′ , f ) ≤ R(∆′ , f ). (10.26)

(c): One just combines (10.8) with (b).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 164

(d): For a = b, there is nothing to show. For a < b, let (∆n )n∈N be a sequence of
partitions of I such that limn→∞ |∆n | = 0, and let ∆′ be an arbitrary partition of I with
numbers α and M defined as in Lem. 10.9. Then, according to (10.16a):

r(∆n , f ) ≤ r(∆n + ∆′ , f ) ≤ r(∆n , f ) + 2 α M |∆n | for each n ∈ N. (10.27)



From (b), we conclude the sequence r(∆n , f ) n∈N is bounded. According to the Bolza-
no-Weierstrass Th. 7.27, if we can show that the sequence has J∗ (f, I) as its only cluster
point, then the first equality of (10.24) must hold. Thus, according to Prop. 7.26, it suf-
fices to show that every converging subsequence of (r(∆n , f ))n∈N converges to J∗ (f, I).
To this end, suppose (r(∆nk , f ))k∈N is a converging subsequence of (r(∆n , f ))n∈N with
β := limk→∞ r(∆nk , f ). First note β ≤ J∗ (f, I) due to the definition of J∗ (f, I). More-
over, (10.27) implies limk→∞ r(∆nk + ∆′ , f ) = β. Since r(∆′ , f ) ≤ r(∆nk + ∆′ , f ) and ∆′
is arbitrary, we obtain J∗ (f, I) ≤ β, i.e. J∗ (f, I) = β. Thus, we have shown that, indeed,
every subsequence of (r(∆n , f ))n∈N converges to β = J∗ (f, I). In the same manner, one
conducts the proof of J ∗ (f, I) = limn→∞ R(∆n , f ). Then (10.25a) is immediate from
the definition of Riemann integrability, and (10.25b) follows from (10.25a), since (10.7)
implies r(∆, f ) ≤ ρ(∆, f ) ≤ R(∆, f ) for each tagged partition ∆ of I. 

Theorem 10.11. Let a, b ∈ R, a ≤ b, I := [a, b].

(a) The integral is linear: More precisely, if f, g ∈ R(I, K) and λ, µ ∈ K, then λf +µg ∈
R(I, K) and Z Z Z
(λf + µg) = λ f + µ g. (10.28)
I I I

(b) Let ∆˜ = (y0 , . . . , yM ), M ∈ N, be a partition of I, Jk := [yk−1 , yk ]. Then f ∈


R(I, K) if, and only if, f ∈ R(Jk , K) for each k ∈ {1, . . . , M }. If f ∈ R(I, K),
then
Z b Z M Z
X XM Z yk
f= f= f= f. (10.29)
a I k=1 Jk k=1 yk−1

(c) Monotonicity of the Integral: If f, g : I −→ R are bounded and f ≤ g (i.e. f (x) ≤


g(x) for each x ∈ I), then J∗ (f, I) ≤ J∗ (g, I) and J ∗ (f, I) ≤ J ∗ (g, I). In particular,
if f, g ∈ R(I) and f ≤ g, then Z Z
f≤ g. (10.30)
I I

(d) Triangle Inequality: For each f ∈ R(I, C), one has


Z Z

f ≤ |f |. (10.31)

I I

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 165

Proof. (a): First, consider K = R, i.e. f, g : I −→ R and λ, µ ∈ R. For a = b,


there is nothing to prove, so let a < b. Let (∆n )n∈N be a sequence of partitions of I,
∆n = (xn,0 , . . . , xn,Nn ), In,j := [xn,j−1 , xn,j ], satisfying limn→∞ |∆n | = 0. Note that, for
each n ∈ N and each j ∈ {1, . . . , Nn },

mn,j (f + g) = inf{f (x) + g(x) : x ∈ In,j }


≥ inf{f (x) : x ∈ In,j } + inf{g(x) : x ∈ In,j }
= mn,j (f ) + mn,j (g), (10.32a)
Mn,j (f + g) = sup{f (x) + g(x) : x ∈ In,j }
≤ sup{f (x) : x ∈ In,j } + sup{g(x) : x ∈ In,j }
= Mn,j (f ) + Mn,j (g), (10.32b)
∀ mn,j (λf ) = inf{λf (x) : x ∈ In,j }
λ∈R
(
(4.6d) λ inf{f (x) : x ∈ In,j } = λ mn,j (f ) for λ ≥ 0,
= (10.32c)
λ sup{f (x) : x ∈ In,j } = λ Mn,j (f ) for λ < 0,
∀ Mn,j (λf ) = sup{λf (x) : x ∈ In,j }
λ∈R
(
(4.6c) λ sup{f (x) : x ∈ In,j } = λ Mn,j (f ) for λ ≥ 0,
= (10.32d)
λ inf{f (x) : x ∈ In,j } = λ mn,j (f ) for λ < 0.

Thus,
Nn
X
(10.24) (10.7a)
J∗ (f + g, I) = lim r(∆n , f + g) = lim mn,j (f + g) |In,j |
n→∞ n→∞
j=1
(10.32a) 
≥ lim r(∆n , f ) + r(∆n , g) = J∗ (f, I) + J∗ (g, I), (10.33a)
n→∞
Nn
X
∗ (10.24) (10.7b)
J (f + g, I) = lim R(∆n , f + g) = lim Mn,j (f + g) |In,j |
n→∞ n→∞
j=1
(10.32b) 
≤ lim R(∆n , f ) + R(∆n , g) = J ∗ (f, I) + J ∗ (g, I),(10.33b)
n→∞

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 166

Nn
X
(10.24) (10.7a)
∀ J∗ (λf, I) = lim r(∆n , λf ) = lim mn,j (λf ) |In,j |
λ∈R n→∞ n→∞
j=1
(
(10.32c) λ limn→∞ r(∆n , f ) = λ J∗ (f, I) for λ ≥ 0,
= (10.33c)
λ limn→∞ R(∆n , f ) = λ J ∗ (f, I) for λ < 0,
Nn
X
∗ (10.24) (10.7b)
∀ J (λf, I) = lim R(∆n , λf ) = lim Mn,j (λf ) |In,j |
λ∈R n→∞ n→∞
j=1
(
(10.32d) λ limn→∞ R(∆n , f ) = λ J ∗ (f, I) for λ ≥ 0,
= (10.33d)
λ limn→∞ r(∆n , f ) = λ J∗ (f, I) for λ < 0.

Thus, if f and g are both Riemann integrable over I, then we obtain J∗ (f + g, I) ≥


J∗ (f, I)+J∗ (g, I) = J ∗ (f, I)+J ∗ (g, I) ≥ J ∗ (f +g, I), i.e., by Th. 10.10(c), (f +g) ∈ R(I);
and J∗ (λf, I) = λ J∗ (f, I) = λ J ∗ (f, I) for λ ≥ 0, J∗ (λf, I) = λ J ∗ (f, I) = λ J∗ (f, I) for
λ < 0, i.e. (λf ) ∈ R(I) in each case. In particular, for each λ, µ ∈ R,
Z Z Z
(λf + µg) = J∗ (λf + µg, I) = λJ∗ (f, I) + µJ∗ (g, I) = λ f + µ g,
I I I

proving (10.28) for K = R. It remains to consider f, g ∈ R(I, C) and λ, µ ∈ C. One


computes, using the real-valued case,
Z Z Z 
(λf ) = (Re λ Re f − Im λ Im f ), (Re λ Im f + Im λ Re f )
I I I
 Z Z Z Z 
= Re λ Re f − Im λ Im f, Re λ Im f + Im λ Re f
I I I I
Z
=λ f
I

and
Z Z Z  Z Z Z Z 
(f + g) = Re(f + g), Im(f + g) = Re f + Re g, Im g + Im g
I I I I I I I
Z Z  Z Z  Z Z
= Re f, Im f + Re g, Im g = f + g.
I I I I I I

(b): Once again, consider K = R first. For a = b, there is nothing to prove, so let
a < b. For M = 1, there is still nothing to prove. For M = 2, we have a = y0 <
y1 < y2 = b. Consider a sequence (∆n )n∈N of partitions of I, ∆n = (xn,0 , . . . , xn,Nn ),
such that limn→∞ |∆n | = 0 and y1 ∈ ν(∆n ) for each n ∈ N. Define ∆′n := (xn,0 , . . . , y1 ),

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 167

∆′′n := (y1 , . . . , xn,Nn ). Then ∆′n and ∆′′n are partitions of J1 and J2 , respectively, and
limn→∞ |∆′n | = limn→∞ |∆′′n | = 0. Moreover,
 
′ ′′ ′ ′′
∀ r(∆n , f ) = r(∆n , f ) + r(∆n , f ), R(∆n , f ) = R(∆n , f ) + R(∆n , f ) ,
n∈N

∗ ∗ ∗
implying
R R J∗ (f,RI) = J∗ (f, J1 )+J∗ (f, J2 ) and J (f, I) = J (f, J1 )+J (f, J2 ). This proves
I
f = J1 f + J2 f provided f ∈ R(I) ∩ R(J1 ) ∩ R(J2 ). So it just remains to show the
claimed equivalence between f ∈ R(I) and f ∈ R(J1 ) ∩ R(J2 ). If f ∈ R(J1 ) ∩ R(J2 ),
then J∗ (f, I) = J∗ (f, J1 )+J∗ (f, J2 ) = J ∗ (f, J1 )+J ∗ (f, J2 ) = J ∗ (f, I), showing f ∈ R(I).
Conversely, J∗ (f, I) = J ∗ (f, I) implies J∗ (f, J1 ) = J ∗ (f, J1 ) + J ∗ (f, J2 ) − J∗ (f, J2 ) ≥
J ∗ (f, J1 ), showing J∗ (f, J1 ) = J ∗ (f, J1 ) and f ∈ R(J1 ); f ∈ R(J2 ) follows completely
analogous. The general case now follows by induction on M . If, f ∈ R(I, C), then one
computes, using the real-valued case,
Z Z Z  M Z M Z
! M Z
X X X
f= Re f, Im f = Re f, Im f = f.
I I I k=1 Jk k=1 Jk k=1 Jk

(c): If f, g : I −→ R are bounded and f ≤ g, then, for each partition ∆ of I, r(∆, f ) ≤


r(∆, g) and R(∆, f ) ≤ R(∆, g) are immediate from (10.7). As these inequalities are
preserved when taking the sup and the inf, respectively, all claims of (c) are established.
(d): We will see in Th. 10.18(c) below, that f ∈ R(I, K) implies |f | ∈ R(I). Let ∆ be
an arbitrary partition of I, tagged by (t1 , . . . , tN ). Then, using the same notation as in
Def. 10.3 and Def. 10.4,
!
N N
 X
X

ρ(∆, Re f ), ρ(∆, Im f ) := Re f (tj ) |Ij |, Im f (tj ) |Ij |

j=1 j=1

N
X 
≤ Re f (tj ), Im f (tj ) |Ij |
j=1
N
X
= |f (tj )| |Ij | =: ρ(∆, |f |). (10.34)
j=1

Since the intermediate Riemann sums in (10.34) converge to the respective integrals by
(10.25b), one obtains
Z Z
 (10.34)
f = lim ρ(∆, Re f ), ρ(∆, Im f ) ≤ lim ρ(∆, |f |) = |f |,
|∆|→0 |∆|→0
I I

proving (10.31). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 168

Theorem 10.12 (Mean Value Theorem). Let a, b ∈ R, a ≤ b, I := [a, b]. If f, p ∈ R(I)


and p ≥ 0, then, for each m, M ∈ R with m ≤ f ≤ M :
Z Z Z
m p ≤ f p ≤ M p. (10.35a)
I I I

In particular, if f is continuous, then


Z Z
∃ f p = f (ξ) p. (10.35b)
ξ∈I I I

Returning to a general f ∈ R(I), if p ≡ 1, then we obtain the theorem’s classical form:


Z b Z
m (b − a) = m |I| ≤ f= f ≤ M |I| = M (b − a). (10.35c)
a I
R
The theorem’s name comes from the fact that, for a < b, |I|−1 I
f is sometimes referred
to as the mean value of f on I.

Proof. Since mp ≤ f p ≤ M p, we compute


Z Th. 10.11(c) Z Th. 10.11(c)
Z
m p ≤ fp ≤ M p. (10.36)
I I I
R R
If I p = 0, then R(10.35a) implies I f p = 0 and (10.35b) is immediate. It remains
to consider that I p > 0. If f is continuous, then we can let m, M be such that
f (I) = [m, M ] by Th. 7.54 and Rthe intermediate value Th. 7.57. Then (10.35a) shows
fp
there is ξ ∈ I such that f (ξ) = RI p , i.e. (10.35b) holds. 
I

Theorem 10.13 (Riemann’s Integrability Criterion). Let I = [a, b] ⊆ R and suppose


f : I −→ R is bounded. Then f is Riemann integrable over I if, and only if, for each
ǫ > 0, there exists a partition ∆ of I such that

R(∆, f ) − r(∆, f ) < ǫ. (10.37)

Proof. Suppose, for each ǫ > 0, there exists a partition ∆ of I such that (10.37) is
satisfied. Then
J ∗ (f, I) − J∗ (f, I) ≤ R(∆, f ) − r(∆, f ) < ǫ, (10.38)
showing J ∗ (f, I) ≤ J∗ (f, I). As the opposite inequality always holds, we have J ∗ (f, I) =
J∗ (f, I), i.e. f ∈ R(I) as claimed. Conversely, if f ∈ R(I) and (∆n )n∈N is a sequence of
partitions of I with limn→∞ |∆n | = 0, then (10.25a) implies that, for each ǫ > 0, there
is N ∈ N such that R(∆n , f ) − r(∆n , f ) < ǫ for each n > N . 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 169

The previous theorem will allow us to prove that every continuous function on [a, b] is
Riemann integrable. However, we will also need to make use of the following result:
Proposition 10.14. Let I = [a, b] ⊆ R, a ≤ b, f : I −→ R. If f is continuous, then f
is even uniformly continuous, i.e.

∀ + ∃ + ∀ |x − y| < δ ⇒ |f (x) − f (y)| < ǫ . (10.39)
ǫ∈R δ∈R x,y∈I

Proof. Arguing by contraposition, we assume f not to be uniformly continuous on I.


Then the negation of (10.39) must hold, i.e.

∃ + ∀ + ∃ |x − y| < δ ∧ |f (x) − f (y)| ≥ ǫ0 . (10.40)
ǫ0 ∈R δ∈R x,y∈I

In particular, for each n ∈ N, there exist xn , yn ∈ I such that

|xn − yn | < δn := 1/n (10.41)

and |f (xn ) − f (yn )| ≥ ǫ0 . Then the sequence (xn )n∈N is bounded and the Bolzano-
Weierstrass Th. 7.27 provides a convergent subsequence (xφ(n) )n∈N , i.e. there is ξ ∈ R
with limn→∞ xφ(n) = ξ. Clearly, ξ ∈ [a, b] and (10.41) implies limn→∞ yφ(n) = ξ as
well. However, due to |f (xφ(n) ) − f (yφ(n) )| ≥ ǫ0 > 0, the sequences f (xφ(n) ) n∈N and

f (yφ(n) ) n∈N can not both converge to f (ξ), showing that f can not be continuous. 
Caveat 10.15. It is important in Prop. 10.14 that f is defined on a compact interval I.
The examples f : ]0, 1] −→ R, f (x) := 1/x, and f : R −→ R, f (x) := x2 are examples
of continuous functions that are not uniformly continuous.
Theorem 10.16. Let I = [a, b] ⊆ R, a ≤ b.

(a) If f : I −→ C is continuous, then f is Riemann integrable over I.

(b) If f : I −→ R is increasing or decreasing, then f is Riemann integrable over I.

Proof. (a): As f is continuous if, and only if, Re f and Im f are both continuous, it
suffices to consider a real-valued continuous f . For a = b, there is nothing to prove, so
let a < b. First note that, if f is continuous on I = [a, b], then f is bounded by Th.
7.54. Moreover, f is uniformly continuous due to Prop. 10.14. Thus, given ǫ > 0, there
is δ > 0 such that |x − y| < δ implies |f (x) − f (y)| < ǫ/|I| for each x, y ∈ I. Then, for
each partition ∆ of I satisfying |∆| < δ, we obtain
N N
X ǫ X
R(∆, f ) − r(∆, f ) = (Mj − mj )|Ij | ≤ |Ij | = ǫ, (10.42)
j=1
|I| j=1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 170

as |∆| < δ implies |x − y| < δ for each x, y ∈ Ij and each j ∈ {1, . . . , N }. Finally,
(10.42) implies f ∈ R(I) due to Riemann’s integrability criterion of Th. 10.13.
(b): Suppose f : [a, b] −→ R is increasing. Then f is bounded, as f (a) ≤ f (x) ≤ f (b)
for each x ∈ [a, b]. If f (a) = f (b), then f is constant. Thus, assume f (a) < f (b).
Moreover, if ∆ = (x0 , . . . , xN ) is a partition of I as in Def. 10.3, then
N
X N
X  
R(∆, f ) − r(∆, f ) = (Mj − mj )|Ij | = f (xj ) − f (xj−1 ) |Ij | ≤ |∆| f (b) − f (a) .
j=1 j=1
(10.43)
Thus, given ǫ > 0, we have R(∆, f ) − r(∆, f ) < ǫ for each partition ∆ of I satisfying
|∆| < ǫ/(f (b) − f (a)). In consequence, f ∈ R(I), once again due to Riemann’s integra-
bility criterion of Th. 10.13. If f is decreasing, then −f is increasing, and Th. 10.11(a)
establishes the case. 
Definition and Remark 10.17. Let M ⊆ C. A function f : M −→ C f is called
Lipschitz continuous in M with Lipschitz constant L if, and only if,
∃ ∀ |f (x) − f (y)| ≤ L |x − y|. (10.44)
L∈R+
0
x,y∈M

Every Lipschitz continuous function is, indeed, continuous, since, if ξ ∈ M and (yn )n∈N
is a sequence in M with limn→∞ yn = ξ, then (10.44) implies
∀ |f (ξ) − f (yn )| ≤ L |ξ − yn |, (10.45)
n∈N

proving limn→∞ f (yn ) = f (ξ). Moreover, it is not too much harder to prove Lipschitz
continuous functions are even uniformly continuous,√but we will not pursue this right
now. On the other hand, f : R+ 0 −→ R, f (x) := x, is an example of a continuous
function (actually, even uniformly continuous) that is not Lipschitz continuous.
Theorem 10.18. Let a, b ∈ R, a ≤ b, I := [a, b].

(a) If f ∈ R(I, R) and φ : f (I) −→ C is Lipschitz continuous, then φ ◦ f ∈ R(I, C).


(b) If f ∈ R(I, C) and φ : f (I) −→ R is Lipschitz continuous, then φ ◦ f ∈ R(I, R).
(c) If f ∈ R(I), then |f |, f 2 , f + , f − ∈ R(I). In particular, we, indeed, have (10.2)
from the introduction (with M replaced by I). If, in addition, there exists δ > 0
such that f (x) ≥ δ for each x ∈ I, then 1/f ∈ R(I). Moreover, |f | ∈ R(I) also
holds for f ∈ R(I, C).
(d) If f, g ∈ R(I), then max(f, g), min(f, g) ∈ R(I). If f, g ∈ R(I, K), then f¯, f g ∈
R(I, K). If, in addition, there exists δ > 0 such that |g(x)| ≥ δ for each x ∈ I, then
f /g ∈ R(I, K).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 171

Proof. (a),(b): We just carry out the proof for the case f ∈ R(I), φ : f (I) −→ R, and
leave the extension to the case that f or φ is C-valued as an exercise. Thus, let f ∈ R(I)
and let φ : f (I) −→ R be Lipschitz continuous. Then there exists L ≥ 0 such that

φ(x) − φ(y) ≤ L |x − y| for each x, y ∈ f (I). (10.46)
As f ∈ R(I), given ǫ > 0, Th. 10.13 provides a partition ∆ of I such that R(∆, f ) −
r(∆, f ) < ǫ/L, and we obtain
N
X 
R(∆, φ ◦ f ) − r(∆, φ ◦ f ) = Mj (φ ◦ f ) − mj (φ ◦ f ) |Ij |
j=1
N
X 
≤ L Mj (f ) − mj (f ) |Ij |
j=1

= L R(∆, f ) − r(∆, f ) < ǫ. (10.47)
Thus, φ ◦ f ∈ R(I) by another application of Th. 10.13.
(c): |f |, f 2 , f + , f − ∈ R(I) follows from (b) (for |f | also for f ∈ R(I, C)), since each of
the maps x 7→ |x|, x 7→ x2 , x 7→ max{x, 0}, x 7→ − min{x, 0} is Lipschitz continuous
on the bounded set f (I) (recall that f ∈ R(I) implies that f is bounded). Since
f = f + − f − , (10.2) is implied by (10.28). Finally, if f (x) ≥ δ > 0, then x 7→ x−1 is
Lipschitz continuous on the bounded set f (I), and f −1 ∈ R(I) follows from (b).
(d): For f, g ∈ R(I), we note that, due to
1
f g = (f + g)2 − (f − g)2 , (10.48a)
4
max(f, g) = f + (g − f )+ , (10.48b)
min(f, g) = g − (f − g)− , (10.48c)
everything is a consequence of (c). For f, g ∈ R(I, C), due to
f¯ = (Re f, − Im f ), (10.48d)
f g = (Re f Re g − Im f Im g, Re f Im g + Im f Re g), (10.48e)
1/g = (Re g/|g|2 , − Im g/|g|2 ), (10.48f)
everything follows from the real-valued case together with (c) and Th. 10.11(a), where
|g| ≥ δ > 0 guarantees |g|2 ≥ δ 2 > 0). 

10.2 Important Theorems


This section compiles a number of important theorems on Riemann integrals, which, in
particular, provide powerful tools to actually evaluate such integrals.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 172

10.2.1 Fundamental Theorem of Calculus

We provide two variants of the fundamental theorem with slightly different flavors: In
the first variant, Th. 10.20(a), we start with a function f , obtain another function F
by means of integrating f , and recover f by taking the derivative of F . In the second
variant, Th. 10.20(b), one first differentiates the given function F , obtaining f := F ′ ,
followed by integrating f , recovering F up to an additive constant.
Notation 10.19. If a, b ∈ R, a ≤ b, I := [a, b], f : I −→ C, then denote
Z b Z Z a Z b
f := f, f := − f, (10.49a)
a I b a

[f (t)]ba := [f ]ba := f (b) − f (a), [f (t)]ab := [f ]ab := f (a) − f (b), (10.49b)

where f ∈ R(I, C) for (10.49a).


Theorem 10.20. Let a, b ∈ R, a < b, I := [a, b].

(a) If f ∈ R(I, K) is continuous in ξ ∈ I, then, for each c ∈ I, the function


Z x
Fc : I −→ K, Fc (x) := f (t) dt , (10.50)
c

is differentiable in ξ with Fc′ (ξ) = f (ξ). In particular, if f ∈ C(I, K), then Fc ∈


C 1 (I, K) and Fc′ (x) = f (x) for each x ∈ I.
(b) If F ∈ C 1 (I, K) or, alternatively, F : I −→ K is differentiable with integrable
derivative F ′ ∈ R(I, K), then
Z b
b
F (b) − F (a) = [F (t)]a = F ′ (t) dt , (10.51a)
a

and
Z x
F (x) = F (c) + F ′ (t) dt for each c, x ∈ I. (10.51b)
c

Proof. It suffices to prove the case K = R, since the case K = C then follows by applying
the case K = R to Re Fc and Im Fc (for (a)) and to Re F and Im F (for (b)). Thus, for
the rest of the proof, we assume K = R.
(a): We need to show that
Fc (ξ + h) − Fc (ξ)
lim A(h) = 0, where A(h) := − f (ξ). (10.52)
h→0 h

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 173

One computes
Z ξ+h Z ξ+h Z ξ+h
1 1 1 
A(h) = f (t) dt − f (ξ) dt = f (t) − f (ξ) dt . (10.53)
h ξ h ξ h ξ

Now, given ǫ > 0, the continuity of f in ξ allows us to find δ > 0 such that |(f (t)−f (ξ)| <
ǫ/2 for each t ∈ I with |t − ξ| < δ. Thus, for each h with |h| < δ, we obtain
Z ξ+h
1 f (t) − f (ξ) dt ≤ ǫh < ǫ,

|A(h)| ≤ (10.54)
h ξ 2h

thereby proving limh→0 A(h) = 0, i.e. f (ξ) = Fc′ (ξ).


(b): First assume F ∈ C 1 (I). Then F ′ is continuous on I, and we can apply (a) to the
function Z x
G : I −→ R, G(x) := F ′ (t) dt , (10.55)
a

to obtain G = F . Thus, for H := F − G, we obtain H ′ ≡ 0, showing that H must be


′ ′

constant on I, i.e. H(x) = H(a) = F (a) − G(a) = F (a) for each x ∈ I. Evaluating at
x = b yields Z b
F (a) = H(b) = F (b) − F ′ (t) dt , (10.56)
a
thereby establishing the case.
Now we consider the remaining case of a differentiable F with integrable derivative
F ′ ∈ R(I). Consider a partition ∆ = (x0 , . . . , xN ) of I as in Def. 10.3. Then, for
each j ∈ {1, . . . , N }, the mean value theorem provides ξj ∈]xj−1 , xj [ such that F (xj ) −
F (xj−1 ) = (xj − xj−1 ) F ′ (ξj ). Thus,
N
X N
 X
F (b) − F (a) = F (xj ) − F (xj−1 ) = (xj − xj−1 ) F ′ (ξj ) = ρ(∆, F ′ ). (10.57)
j=1 j=1

If we choose a sequence of partitions ∆ of I such that |∆| → 0, then the integrability of


Rb
f implies that the right-hand side of (10.57) converges to a F ′ , once again establishing
the case. 

Definition 10.21. If I ⊆ R, f : I −→ K, and F : I −→ K is a differentiable function


with F ′ = f , then F is called a primitive or antiderivative of f .

Example 10.22. (a) Due to the fundamental theorem, if we know a function’s an-
tiderivative, we can easily compute its integral over a given interval. Here are three

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 174

simple examples:
Z 1  1
5 x6 3x2 1 3 4
(x − 3x) dx = − = − =− , (10.58a)
6 2 0 6 2 3
Z0 e
1
dx = [ln x]e1 = ln e − ln 1 = 1, (10.58b)
x
Z1 π
sin x dx = [− cos x]π0 = 2. (10.58c)
0

(b) As a more complicated example, consider a rational function x 7→ r(x) = p(x)/q(x)


with polynomials p and q. To find an antiderivative of r, one first writes r in the
form r = s + p̃/q̃ with polynomials s, p̃, q̃ such that deg(p̃) < deg(q̃) (this can be
done using so-called polynomial long division). One then applies the partial fraction
decomposition of Sec. G of the Appendix to p̃/q̃: As a concrete example, consider
3x8 − 9x6 − 6x5 + x3 + 5x2 − 1
r : R \ {−1, 2} −→ R, r(x) := .
x3 − 3x − 2
One then finds
5x2 + 3x + 1
r(x) = 3x5 + 1 + ,
(x + 1)2 (x − 2)
and the partial fraction decomposition according to (G.5) is
2 1 3
r(x) = 3x5 + 1 + − 2
+ .
x + 1 (x + 1) x−2
One can now easily provide an antiderivative for each summand, and putting ev-
erything together yields the antiderivative
1 1
R : R \ {−1, 2} −→ R, R(x) := x6 + x + 2 ln |x + 1| + + 3 ln |x − 2|,
2 x+1
for r. We note that, to apply partial fraction decomposition, one always needs the
zeros of the denominator (i.e. of q̃). For polynomials of high degree, it will usually
not be possible to determine theses zeros exactly, but merely approximately.

10.2.2 Integration by Parts Formula

Theorem 10.23. Let a, b ∈ R, a < b, I := [a, b]. If f, g ∈ C 1 (I, C), then the following
integration by parts formula holds:
Z b Z b
′ b
f g = [f g]a − f ′ g. (10.59)
a a

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 175

Proof. If f, g ∈ C 1 (I, C), then, according to the product rule, f g ∈ C 1 (I, C) with
(f g)′ = f ′ g + f g ′ . Applying (10.51a), we obtain
Z b Z b Z b
′ ′
b
[f g]a = (f g) = f g+ f g′, (10.60)
a a a

which is precisely (10.59). 


R 2π
Example 10.24. We compute the integral 0
sin2 t dt :
Z 2π Z 2π Z 2π
2 2π 2
sin t dt = [− sin t cos t]0 + cos t dt = cos2 t dt . (10.61)
0 0 0
R 2π
Adding 0
sin2 t dt on both sides of (10.61) and using sin2 + cos2 ≡ 1 yields
Z 2π Z 2π
2
2 sin t dt = 1 dt = 2π, (10.62)
0 0
R 2π
i.e. 0
sin2 t dt = π.

10.2.3 Change of Variables

Theorem 10.25. Let I, J ⊆ R be intervals, φ ∈ C 1 (I) and f ∈ C(J, C). If φ(I) ⊆ J,


then the following change of variables formula holds for each a, b ∈ I:
Z φ(b) Z φ(b) Z b Z b

f= f (x) dx = f (φ(t)) φ (t) dt = (f ◦ φ) φ′ . (10.63)
φ(a) φ(a) a a

Proof. Let
Z x
F : J −→ C, F (x) := f (t) dt . (10.64)
φ(a)

According to Th. 10.20(a) and the chain rule of Th. 9.11, we obtain

(F ◦ φ)′ : I −→ C, (F ◦ φ)′ (x) = φ′ (x)f (φ(x)). (10.65)

Thus, we can apply (10.51a), which yields


Z φ(b) Z b
f = F (φ(b)) − F (φ(a)) = (f ◦ φ) φ′ , (10.66)
φ(a) a

proving (10.63). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 176

R1 √
Example 10.26. We compute the integral 0 t2 1 − t dt using the change of variables
x := φ(t) := 1 − t, φ′ (t) = −1:
Z 1 Z 0 Z 1
2
√ 2
√ √ √ √
t 1 − t dt = − (1 − x) x dx = ( x − 2x x + x2 x) dx
0 1 0
" 3 5 7
#1
2x 2 4x 2 2x 2 16
= − + = . (10.67)
3 5 7 105
0

10.3 Application: Taylor’s Theorem


Theorem 10.27 (Taylor’s Theorem). Let I ⊆ R be an open interval and a, x ∈ I,
x 6= a. If m ∈ N0 and f ∈ C m+1 (I, K), then
f (x) = Tm (x, a) + Rm (x, a), (10.68)
where
m
X f (k) (a) k f ′′ (a) ′ 2 f (m) (a)
Tm (x, a) := (x−a) = f (a)+f (a)(x−a)+ (x−a) +· · ·+ (x−a)m
k=0
k! 2! m!
(10.69)
is the mth Taylor polynomial and
Z x
(x − t)m (m+1)
Rm (x, a) := f (t) dt (10.70)
a m!
is the integral form of the remainder term. For K = R, one can also write the remainder
term in Lagrange form:
f (m+1) (θ)
Rm (x, a) = (x − a)m+1 with some suitable θ ∈]x, a[. (10.71)
(m + 1)!

Proof. The integral form (10.70) of the remainder term we prove by using induction on
m: For m = 0, the assertion is
Z x
f (x) = f (a) + f ′ (t) dt , (10.72)
a

which holds according to the fundamental theorem of calculus in the form Th. 10.20(b).
For the induction step, we assume (10.68) holds for fixed m ∈ N0 with Rm (x, a) in
integral form (10.70) and consider f ∈ C m+2 (I, K). For fixed x ∈ I, we define the
function
(x − t)m+1 (m+1)
g : I −→ K, g(t) := f (t). (10.73)
(m + 1)!

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 177

Using the product rule, its derivative is


(x − t)m+1 (m+2) (x − t)m (m+1)
g ′ : I −→ K, g ′ (t) = f (t) − f (t). (10.74)
(m + 1)! m!
Applying the fundamental theorem to g then yields
Z x
(10.74)
−g(a) = g(x) − g(a) = g ′ (t) dt = Rm+1 (x, a) − Rm (x, a), (10.75)
a

with Rm (x, a) and Rm+1 (x, a) defined according to (10.70). Thus,


(10.75) f (m+1) (a)
Tm+1 (x, a) + Rm+1 (x, a) = Tm (x, a) + (x − a)m+1 + Rm (x, a) − g(a)
(m + 1)!
ind. hyp.
= Tm (x, a) + Rm (x, a) = f (x), (10.76)
thereby completing the induction and the proof of (10.70).
It remains to prove the Lagrange form (10.71) of the remainder term for K = R. Since
(x − t)m > 0 for x > t or m = 0, and (x − t)m < 0 for x < t and m > 0, (10.70) and
(10.35b) imply the existence of some θ ∈ [x, a] such that
Z x
(m+1) (x − t)m f (m+1) (θ)
Rm (x, a) = f (θ) dt = (x − a)m+1 , (10.77)
a m! (m + 1)!
proving the Lagrange form, where it remains an exercise to show one can always choose
θ ∈]x, a[. 
Remark 10.28. The importance of Taylor’s Th. 10.27 does not lie in the decomposition
f = Tm + Rm , which can be accomplished simply by defining Rm := f − Tm . The
importance lies rather in the specific formulas for the remainder term.
Example 10.29. Depending on f ∈ C ∞ (I) and x, a ∈ I, the Taylor series

 X f (k) (a)
Tm (x, a) m∈N
= (x − a)k
k=0
k!

can converge to f (x) (see (a) below), diverge (see (b) below), or converge to η 6= f (x)
(see (c) below):

(a) For f : R −→ R, f (x) = ex , x ∈ R, a = 0, we have f (k) (a) = e0 = 1 for each


k ∈ N0 , recovering the power series for the exponential function, which we already
know to converge from Def. and Rem. 8.14:
m
X 1 k
lim Tm (x, 0) = lim x = ex .
m→∞ m→∞
k=0
k!

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 178

(b) For f : R+ −→ R, f (x) = ln x, we have f (m) (x) = (−1)m−1 (m − 1)! x−m for
each m ∈ N0 and each x ∈ R+ . Thus, for x = 23 and a = 12 , using f (k) (a) =
(−1)k−1 (k − 1)! 2k , we have
m m
X f (k) (a) k
X (−1)k−1 2k
Tm (x, a) = (x − a) = ,
k=0
k! k=0
k
which diverges for m → ∞, since the summands do not converge to 0.
(c) For ( 2
e−1/x for x 6= 0,
f : R −→ R, f (x) :=
0 for x = 0,
it is an exercise to show f ∈ C ∞ (R) with f (k) (0) = 0 for each k ∈ N0 (hint: for
2
x 6= 0, one obtains f (k) (x) = Pk (x−1 )e−1/x , where Pk is a polynomial of degree 3k).
Thus, Tm (x, 0) = 0 for each x ∈ R and each m ∈ N0 , implying
lim Tm (x, 0) = 0 6= f (x) for each x ∈ R \ {0}.
m→∞

10.4 Improper Integrals


For our definition of the Riemann integral in Def. 10.5, it was important that we con-
sidered bounded functions on compact intervals (where the boundedness of the intervals
was more important than the closedness) – for unbounded functions and/or unbounded
intervals, even Def. 10.4 of lower and upper Riemann sums no longer makes sense.
Still, for sufficiently benign functions, it is possible to extend the notion of a definite
Riemann integral to both unbounded intervals and unbounded functions, and in such
situations we will speak of improper integrals (cf. Def. 10.33 below).
Definition 10.30. Let ∅ 6= I ⊆ R be an interval. We call f : I −→ R to be locally
Riemann integrable if, and only if, f ∈ R(J) for each compact interval J ⊆ I. Let
Rloc (I) denote the set of all locally Riemann integrable functions on I.
Remark 10.31. In particular, locally Riemann integrable functions are bounded on
every compact interval. Moreover, Rloc (I) = R(I) if, and only if, I is a compact interval.
For example, for each a, b ∈ R with a < b, the function given by the assignment rule
1
f (x) :=
(x − a)(x − b)
is clearly locally Riemann integrable, but not bounded on each of the intervals ] − ∞, a[,
]a, b[, and ]b, ∞[.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 179

Before we can define improper Riemann integrals, we define, in partial extension of Def.
8.17:

Definition 10.32. Let M ⊆ R. If M is unbounded from above (resp. below, then


f : M −→ K is said to tend to η ∈ K (or to have the limit η ∈ K) for x → ∞ (resp., for
x → −∞) (denoted by limx→±∞ f (x) = η) if, and only if, for each sequence (ξk )k∈N in M
with limk→∞ ξk = ∞ (resp. with limk→∞ ξk = −∞), the sequence (f (ξk ))k∈N converges
to η ∈ K, i.e.
 
lim f (x) = η ⇔ ∀ lim ξk = ±∞ ⇒ lim f (ξk ) = η . (10.78)
x→±∞ (ξk )k∈N in M k→∞ k→∞

Definition 10.33. Let a < c < b (a = −∞, b = ∞ is admissible).

(a) Let I := [c, b[, f ∈ Rloc (I), and assume b = ∞ or f is unbounded. Consider the
function Z x
F : I −→ R, F (x) := f.
c
If the limit Z x
lim F (x) = lim f (10.79)
x→b x→b c
exists in R, then we define
Z Z b Z b Z x
f := f (t) dt := f := lim f.
I c c x→b c

(b) Let I :=]a, c], f ∈ Rloc (I), and assume a = −∞ or f is unbounded. Consider the
function Z c
F : I −→ R, F (x) := f.
x
If the limit Z c
lim F (x) = lim f (10.80)
x→a x→a x
exists in R, then we define
Z Z c Z c Z c
f := f (t) dt := f := lim f.
I a a x→a x

(c) Let I =]a, b[ , f ∈ Rloc (I). If the conditions of both (a) and (b) hold, i.e. (i) – (iv),
where

(i) b = ∞ or f is unbounded on [c, b[,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 180

Rx
(ii) limx→b c
f exists in R,
(iii) a = −∞ or f is unbounded on ]a, c],
Rc
(iv) limx→a x f exists in R,
then we define Z Z Z Z Z
b b c b
f := f (t) dt := f := f+ f.
I a a a c

All the above limits of Riemann integrals (if they exist) are called improper Riemann
integrals. In each case, if the limit exists, we call f improperly Riemann integrable and
write f ∈ R(I).
Remark 10.34. (a) The definitions in Def. 10.33 are consistent with what occurs if
the limits are proper Riemann integrals: Let a, c, b ∈ R, a < c < b, and f ∈ R[a, b].
Then Z Z Z Z
x b c c
lim f= f and lim f= f. (10.81)
x→b c c x→a x a
Indeed, since f ∈ R[a, b], |f | is bounded by some M ∈ R+ ; and if (xk )k∈N is a
sequence in [a, b[ such that limk→∞ xk = b, then
Z b Z b


f ≤ |f | ≤ M (b − xk ) → 0 for k → ∞,
xk xk

implying
Z xk Z b Z b  Z b Z b
Th. 10.11(b)
lim f = lim f− f = f −0= f.
k→∞ c k→∞ c xk c c

An analogous argument shows the remaining equality in (10.81).


R∞
(b) In Def. 10.33(c), it can occur that −∞ f does not exist, even though the limit
Rx
limx→∞ −x f exists: For example, if f : R −→ R, f (x) = x, then f ∈ Rloc (R), and,
for each sequence (xk )k∈N in R such that limk→∞ xk = ∞ and each c ∈ R, one has
Z xk  2  xk
t x2 − x2k
lim t dt = lim = lim k = 0,
k→∞ −x k→∞ 2 k→∞ 2
k −xk
Z xk  2  xk
t x2 − c 2
lim t dt = lim = lim k = ∞,
k→∞ c k→∞ 2 k→∞ 2
c
Z c  2 c
t c2 − x2k
lim t dt = lim = lim = −∞,
k→∞ −x k→∞ 2 k→∞ 2
k −xk
Rx Rx Rc
i.e. limx→∞ −x t dt = 0, but neither limx→∞ c t dt nor limx→−∞ x t dt exists in
R.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 181

(c) Let a < c1 < c2 < b (a = −∞, b = ∞ is admissible). If I := [c1 , b[, f ∈ Rloc (I), and
Rb Rb
b = ∞ or f is unbounded, then c1 f exists if, and only if, c2 f exists. Moreover, if
the integrals exist, then Z Z Z b c2 b
f= f+ f. (10.82a)
c1 c1 c2
Rb
Indeed, if (xk )k→∞ is a sequence in [c1 , b[ such that limk→∞ xk = b and if c1
f exists,
then Z Z Z  Z Z
xk xk c2 b c2
Th. 10.11(b)
lim f = lim f− f = f− f,
k→∞ c2 k→∞ c1 c1 c1 c1
Rb Rb
proving c2
f exists and (10.82a) holds. Conversely, if c2
f exists, then
Z xk Z x k Z c2  Z b Z c2
Th. 10.11(b)
lim f = lim f+ f = f+ f,
k→∞ c1 k→∞ c2 c1 c2 c1
Rb
proving c1 f exists and (10.82a) holds. Analogously, one shows that if I :=]a, c2 ],
Rc Rc
f ∈ Rloc (I), and a = −∞ or f is unbounded, then a 1 f exists if, and only if, a 2 f
exists, where, if the integrals exist, then
Z c2 Z c2 Z c1
f= f+ f. (10.82b)
a c1 a

In particular, we see that neither the existence nor the value of the improper integral
in Def. 10.33(c) depends on the choice of c.
Example 10.35. (a) Let 0 < α < 1. We claim that
Z 1  Z 1 
1 1 1 1
α
dt = α = yields √ dt = 2 . (10.83)
0 t 1−α 2 0 t
Indeed, if (xk )k∈N is a sequence in ]0, 1] such that limk→∞ xk = 0, then
Z 1  1−α 1
1 t 1 − xk1−α 1
lim α
dt = lim = lim = .
k→∞ x t k→∞ 1 − α k→∞ 1 − α 1−α
k xk

(b) If (xk )k∈N is a sequence in ]0, 1] such that limk→∞ xk = 0, then


Z 1 h i1  
1
lim dt = lim ln t = lim 0 − ln xk = ∞,
k→∞ x t k→∞ xk k→∞
k

showing the limit does not exist in R, but diverges to ∞. Sometimes, this is stated
in the form Z 1
1
dt = ∞. (10.84)
0 t

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 182

(c) We claim that Z 0


et dt = 1. (10.85)
−∞

Indeed, if (xk )k∈N is a sequence in R− 0 such that limk→∞ xk = −∞, then


Z 0 h i0
lim et dt = lim et = lim (1 − exk ) = 1.
k→∞ xk k→∞ xk k→∞

(d) Consider the function


(
1
n for n ≤ t ≤ n + , n ∈ N,
f : R+
0 −→ R, f (t) := n2n
0 otherwise.

Then limt→∞ f (t) does not exist and f is not even bounded. However f ∈ R(R+
0)
and Z ∞ ∞ Z n+1/(n2n ) ∞
X X 1
f= n dt = 2−n = 1 − 1 = 1.
0 n=1 n n=1
1 − 2

Lemma 10.36. Let a < c < b (a = −∞, b = ∞ is admissible). Let I ⊆]a, b[ be one of
the three kinds of intervals occurring in Def. 10.33 (i.e. I = [c, b[, I =]a, c], or I =]a, b[),
and assume f, g : I −→ R to be improperly Riemann integrable over I.

(a) Linearity: For each λ, µ ∈ R, λf + µg is improperly Riemann integrable over I and


Z Z Z
(λf + µg) = λ f + µ g.
I I I

(b) Monotonicity: If f ≤ g, then Z Z


f≤ g.
I I

Proof. We conduct the proof for the case I = [c, b[ – the case I =]a, b] can be shown
analogously, and the case I =]a, b[ then also follows. Let (xk )k∈N be a sequence in I
such that limk→∞ xk = b.
(a): One computes
Z xk  Z xk Z xk  Z b Z b
Th. 10.11(a)
lim λf + µg) = lim λ f +µ g =λ f +µ g,
k→∞ c k→∞ c c c c

showing (λf + µg) ∈ R(I) and proving (a).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 183

(b): One estimates


Z b Z xk Th. 10.11(c)
Z xk Z b
f = lim f ≤ lim g= g,
c k→∞ c k→∞ c c

proving (b). 

Definition 10.37. Let a < c < b (a = −∞, b = ∞ is admissible). Let I ⊆]a, b[ be


one of the three kinds of intervals occurring in Def. 10.33 (i.e. I = [c, b[, IR =]a, c], or
I =]a, b[), and assume f ∈ Rloc (I). Then, by Th. 10.18(c), |fR | ∈ Rloc (I). If I |f | exists
as an improper integral, then we call the improper integral I f absolutely convergent.

Before we can proceed to Prop. 10.39 about convergence criteria for improper integrals,
we need to prove the analogon of Th. 7.19 for limits of functions.

Proposition 10.38. Let ∅ 6= M ⊆ R, a ∈ R ∪ {−∞}, b ∈ R ∪ {∞}, and assume


(
inf(M \ {a}) if M is bounded from below,
a= (10.86a)
−∞ if M is unbounded from below,
(
sup(M \ {a}) if M is bounded from above,
b= (10.86b)
∞ if M is unbounded from above.

Let f : M −→ R be monotone (increasing or decreasing). Defining A := f (M ) =


{f (x) : x ∈ M }, the following holds:


 sup A if f is increasing and A is bounded from above,

∞ if f is increasing and A is not bounded from above,
lim f (x) = (10.87a)
x→b 

 inf A if f is decreasing and A is bounded from below,

−∞ if f is decreasing and A is not bounded from below,


 sup A if f is decreasing and A is bounded from above,

∞ if f is decreasing and A is not bounded from above,
lim f (x) = (10.87b)
x→a 

 inf A if f is increasing and A is bounded from below,

−∞ if f is increasing and A is not bounded from below.

Proof. We prove (10.87a) for the case, where f is increasing – the remaining case of
(10.87a) as well as (10.87b) can be proved completely analogous. Let (xk )k∈N be a
sequence in M \ {b} such that limk→∞ xk = b. We have to show that limk→∞ f (xk ) = η,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 184

where η := sup A for A bounded from above and η := ∞ for A not bounded from above.
Seeking a contradiction, assume limk→∞ f (xk ) = η does not hold. Due to the choice of
b, there then must be ǫ > 0 and a subsequence (yk )k∈N of (xk )k∈N such that (yk )k∈N is
strictly increasing and
(
η − ǫ if η = sup A,
∀ f (yk ) ≤
k∈N ǫ if η = ∞.

Since limk→∞ yk = b and f is increasing, this means sup A ≤ η − ǫ or sup A = ǫ, which


means a contradiction in each case. Thus, limk→∞ f (xk ) = η must hold and the proof
is complete. 

Proposition 10.39. Let a < c < b (a = −∞, b = ∞ is admissible). Let I ⊆]a, b[ be


one of the three kinds of intervals occurring in Def. 10.33 (i.e. I = [c, b[, I =]a, c], or
I =]a, b[), and assume f ∈ Rloc (I).
R R
(a) If g ∈ Rloc (I), R0 ≤ f ≤ g, and I Rg exists, then I f exists as well. Conversely, if
0 ≤ g ≤ f and I g diverges, then I f diverges as well.
R
(b) If I f is an improper integral that is absolutely convergent, then it is also conver-
gent.

Proof. (a): We consider the case I = [c, b[ – the proof for the case I =]a, c] is completely
R
analogous, and the case I =]a, b[ then also follows. First, suppose 0 ≤ f ≤ g, and I g
exists. Since 0 ≤ f , the function
Z x
+
F : [c, b[−→ R0 , F (x) := f,
c

is increasing. Due to
Z x Z x Z b
∀ F (x) = f≤ g≤ g ∈ R+
0,
x∈[c,b[ c c c

F is also bounded from above (in the sense that {F (x) R: x ∈ [c, b[} is bounded from
x
above), i.e. Prop. 10.38 yields that limx→b F (x) = limx→b c f exists in R as claimed.
R
Now suppose 0 ≤ g ≤ f and I g diverges. As the function F above, the function
Z x
+
G : [c, b[−→ R0 , G(x) := g,
c

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
10 THE RIEMANN INTEGRAL ON INTERVALS IN R 185

is increasing. Since we assume that limx→b G(x) does not exist in R, Prop. 10.38 im-
plies limx→b G(x) = ∞. As a consequence, if (xk )k∈N is a sequence in [c, b[ such that
limk→∞ xk = b, then
Z xk Z xk
lim F (xk ) = lim f = lim g = ∞,
k→∞ k→∞ c k→∞ c
R
showing that f diverges as well.
I
R R
(b): We assume I f to converge absolutely, i.e. I |f |Rmust exist inRR. Since 0 ≤ f + ≤ |f |
and 0 ≤ f − ≤ |f |, R(a) then
R implies
R the existence of I f + and of I f − . Thus, according
to Lem. 10.36(a), I f = I f + − I f − must also exist. 

Example 10.40. (a) We will use Prop. 10.39(a) to show that the improper integral
Z ∞
2
e−t dt
0

exists. Indeed,
 2

∀ (t − 1)2 = t2 − 2t + 1 ≥ 0 ⇒ −t2 ≤ −2t + 1 ⇒ 0 ≤ e−t ≤ e−2t+1 ,
t∈R

and, since
Z ∞ Z x  x
−2t+1 −2t+1 e−2t+1 e − e−2x+1 e
e dt = lim e dt = lim − = lim = ,
0 x→∞ 0 x→∞ 2 0
x→∞ 2 2
R∞ 2
Prop. 10.39(a) implies that 0
e−t dt exists in R.

(b) We will use Prop. 10.39(a) to show that


Z ∞
2
et dt
0

diverges. Indeed,  
2 t2
∀ t ≥0 ⇒ e ≥1 ,
t∈R

and, since Z x
lim 1 dt = lim x = ∞,
x→∞ 0 x→∞
R∞ t2
Prop. 10.39(a) implies that 0
e dt = ∞.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 186

(c) We provide an example that shows an improper integral can converge without
converging absolutely: Consider the function
(
(−1)n+1 for n ≤ t ≤ n + n1 , n ∈ N,
f : [0, ∞[−→ R, f (t) := (10.88)
0 otherwise.

Then Z ∞ n Z k+ k1 n
X X 1 (7.72)
|f | = lim 1 dt = lim = ∞, (10.89)
0 k→∞
k=1 k k→∞
k=1
k
R∞
showing 0
f does not converge absolutely. However, we will show
Z ∞ ∞
X (−1)j+1
f= =: α > 0. (10.90)
0 j=1
j

We know α > 0 from Ex. 7.86(a) and Th. 7.85. Let (xk )k∈N be a sequence in R+ 0
such that limk→∞ xk = ∞. Given ǫ > 0, choose K ∈ N such that K1 < 2ǫ and N ∈ N
such that
∀ xk > K. (10.91)
k>N

Then, for each k > N , there exists K1 ∈ N such that K < K1 ≤ xk < K1 + 1. Thus
Z xk   KX1 −1
1 (−1)j+1
f (t) dt = min xk − K1 , + (10.92)
0 K1 j=1
j

and
X 
∞ (−1)j+1
Z xk 

α − 1 (7.79) 1 1
f (t) dt = − min xk − K1 , < +

0
j=K
j K1 K1 K1
1

2 ǫ
< < 2 · = ǫ, (10.93)
K 2
thereby proving (10.90).

A Axiomatic Set Theory

A.1 Motivation, Russell’s Antinomy


As it turns out, naive set theory, founded on the definition of a set according to Cantor
(as stated at the beginning of Sec. 1.3) is not suitable to be used in the foundation of

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 187

mathematics. The problem lies in the possibility of obtaining contradictions such as


Russell’s antinomy, after Bertrand Russell, who described it in 1901.
Russell’s antinomy is obtained when considering the set X of all sets that do not con-
tain themselves as an element: When asking the question if X ∈ X, one obtains the
contradiction that X ∈ X ⇔ X ∈ / X:
Suppose X ∈ X. Then X is a set that contains itself. But X was defined to contain
only sets that do not contain themselves, i.e. X ∈
/ X.
So suppose X ∈
/ X. Then X is a set that does not contain itself. Thus, by the definition
of X, X ∈ X.
Perhaps you think Russell’s construction is rather academic, but it is easily translated
into a practical situation. Consider a library. The catalog C of the library should contain
all the library’s books. Since the catalog itself is a book of the library, it should occur
as an entry in the catalog. So there can be catalogs such as C that have themselves as
an entry and there can be other catalogs that do not have themselves as an entry. Now
one might want to have a catalog X of all catalogs that do not have themselves as an
entry. As in Russell’s antinomy, one is led to the contradiction that the catalog X must
have itself as an entry if, and only if, it does not have itself as an entry.
One can construct arbitrarily many versions, which we will not do. Just one more:
Consider a small town with a barber, who, each day, shaves all inhabitants, who do not
shave themselves. The poor barber now faces a terrible dilemma: He will have to shave
himself if, and only if, he does not shave himself.
To avoid contradictions such as Russell’s antinomy, axiomatic set theory restricts the
construction of sets via so-called axioms, as we will see below.

A.2 Set-Theoretic Formulas


The contradiction of Russell’s antinomy is related to Cantor’s sets not being hierarchical.
Another source of contradictions in naive set theory is the imprecise nature of informal
languages such as English. In (1.6), we said that
A := {x ∈ B : P (x)}
defines a subset of B if P (x) is a statement about an element x of B. Now take
B := N := {1, 2, . . . } to be the set of the natural numbers and let
P (x) := “The number x can be defined by fifty English words or less”. (A.1)
Then A is a finite subset of N, since there are only finitely many English words (if you
think there might be infinitely many English words, just restrict yourself to the words

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 188

contained in some concrete dictionary). Then there is a smallest natural number n that
is not in A. But then n is the smallest natural number that can not be defined by
fifty English words or less, which, actually, defines n by less than fifty English words, in
contradiction to n ∈/ A.
To avoid contradictions of this type, we require P (x) to be a so-called set-theoretic
formula.

Definition A.1. (a) The language of set theory consists precisely of the following sym-
bols: ∧, ¬, ∃, (, ), ∈, =, vj , where j = 1, 2, . . . .

(b) A set-theoretic formula is a finite string of symbols from the above language of set
theory that can be built using the following recursive rules:

(i) vi ∈ vj is a set-theoretic formula for i, j = 1, 2, . . . .


(ii) vi = vj is a set-theoretic formula for i, j = 1, 2, . . . .
(iii) If φ and ψ are set-theoretic formulas, then (φ) ∧ (ψ) is a set-theoretic formula.
(iv) If φ is a set-theoretic formulas, then ¬(φ) is a set-theoretic formula.
(v) If φ is a set-theoretic formulas, then ∃vj (φ) is a set-theoretic formula for
j = 1, 2, . . . .

Example A.2. Examples of set-theoretic formulas are (v3 ∈ v5 ) ∧ (¬(v2 = v3 )),


∃v1 (¬(v1 = v1 )); examples of symbol strings that are not set-theoretic formulas are
v1 ∈ v2 ∈ v3 , ∃∃¬, and ∈ v3 ∃.

Remark A.3. It is noted that, for a given finite string of symbols, a computer can, in
principle, check in finitely many steps, if the string constitutes a set-theoretic formula
or not. The symbols that can occur in a set-theoretic formula are to be interpreted
as follows: The variables v1 , v2 , . . . are variables for sets. The symbols ∧ and ¬ are
to be interpreted as the logical operators of conjunction and negation as described in
Sec. 1.2.2. Similarly, ∃ stands for an existential quantifier as in Sec. 1.4: The statement
∃vj (φ) means “there exists a set vj that has the property φ”. Parentheses ( and ) are
used to make clear the scope of the logical symbols ∃, ∧, ¬. Where the symbol ∈ occurs,
it is interpreted to mean that the set to the left of ∈ is contained as an element in the
set to the right of ∈. Similarly, = is interpreted to mean that the sets occurring to the
left and to the right of = are equal.

Remark A.4. A disadvantage of set-theoretic formulas as defined in Def. A.1 is that


they quickly become lengthy and unreadable (at least to the human eye). To make
formulas more readable and concise, one introduces additional symbols and notation.
Formally, additional symbols and notation are always to be interpreted as abbreviations

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 189

or transcriptions of actual set-theoretic formulas. For example, we use the rules of Th.
1.11 to define the additional logical symbols ∨, ⇒, ⇔ as abbreviations:
(φ) ∨ (ψ) is short for ¬((¬(φ)) ∧ (¬(ψ))) (cf. Th. 1.11(j)), (A.2a)
(φ) ⇒ (ψ) is short for (¬(φ)) ∨ (ψ) (cf. Th. 1.11(a)), (A.2b)
(φ) ⇔ (ψ) is short for ((φ) ⇒ (ψ)) ∧ ((ψ) ⇒ (φ)) (cf. Th. 1.11(b)). (A.2c)
Similarly, we use (1.17a) to define the universal quantifier:
∀vj (φ) is short for ¬(∃vj (¬(φ))). (A.2d)
Further abbreviations and transcriptions are obtained from omitting parentheses if it is
clear from the context and/or from Convention 1.10 where to put them in, by writing
variables bound by quantifiers under the respective quantifiers (as in Sec. 1.4), and by
using other symbols than vj for set variables. For example,
∀ (φ ⇒ ψ) transcribes ¬(∃v1 (¬((¬(φ)) ∨ (ψ)))).
x

Moreover,
vi 6= vj is short for ¬(vi = vj ); vi ∈
/ vj is short for ¬(vi ∈ vj ). (A.2e)
Remark A.5. Even though axiomatic set theory requires the use of set-theoretic for-
mulas as described above, the systematic study of formal symbolic languages is the
subject of the field of mathematical logic and is beyond the scope of this class (see, e.g.,
[EFT07]). In Def. and Rem. 1.15, we defined a proof of statement B from statement A1
as a finite sequence of statements A1 , A2 , . . . , An such that, for 1 ≤ i < n, Ai implies
Ai+1 , and An implies B. In the field of proof theory, also beyond the scope of this class,
such proofs are formalized via a finite set of rules that can be applied to (set-theoretic)
formulas (see, e.g., [EFT07, Sec. IV], [Kun12, Sec. II]). Once proofs have been formal-
ized in this way, one can, in principle, mechanically check if a given sequence of symbols
does, indeed, constitute a valid proof (without even having to understand the actual
meaning of the statements). Indeed, several different computer programs have been
devised that can be used for automatic proof checking, for example Coq [Wik15a], HOL
Light [Wik15b], and Isabelle [Wik15c] to name just a few.

A.3 The Axioms of Zermelo-Fraenkel Set Theory


Axiomatic set theory seems to provide a solid and consistent foundation for conducting
mathematics, and most mathematicians have accepted it as the basis of their everyday
work. However, there do remain some deep, difficult, and subtle philosophical issues
regarding the foundation of logic and mathematics (see, e.g., [Kun12, Sec. 0, Sec. III]).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 190

Definition and Remark A.6. An axiom is a statement that is assumed to be true


without any formal logical justification. The most basic axioms (for example, the stan-
dard axioms of set theory) are taken to be justified by common sense or some underlying
philosophy. However, on a less fundamental (and less philosophical) level, it is a common
mathematical strategy to state a number of axioms (for example, the axioms defining
the mathematical structure called a group), and then to study the logical consequences
of these axioms (for example, group theory studies the statements that are true for all
groups as a consequence of the group axioms). For a given system of axioms, the ques-
tion if there exists an object satisfying all the axioms in the system (i.e. if the system
of axioms is consistent, i.e. free of contradictions) can be extremely difficult to answer.

We are now in a position to formulate and discuss the axioms of axiomatic set the-
ory. More precisely, we will present the axioms of Zermelo-Fraenkel set theory, usually
abbreviated as ZF, which are Axiom 0 – Axiom 8 below. While there exist various
set theories in the literature, each set theory defined by some collection of axioms, the
axioms of ZFC, consisting of the axioms of ZF plus the axiom of choice (Axiom 9, see
Sec. A.4 below), are used as the foundation of mathematics currently accepted by most
mathematicians.

A.3.1 Existence, Extensionality, Comprehension

Axiom 0 Existence:
∃ (X = X).
X

Recall that this is just meant to be a more readable transcription of the


set-theoretic formula ∃v1 (v1 = v1 ). The axiom of existence states that there
exists (at least one) set X.

In Def. 1.18 two sets are defined to be equal if, and only if, they contain precisely the
same elements. In axiomatic set theory, this is guaranteed by the axiom of extensionality:

Axiom 1 Extensionality:
 
∀ ∀ ∀ (z ∈ X ⇔ z ∈ Y ) ⇒ X = Y .
X Y z

Following [Kun12], we assume that the substitution property of equality is part of the
underlying logic, i.e. if X = Y , then X can be substituted for Y and vice versa without

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 191

changing the truth value of a (set-theoretic) formula. In particular, this yields the
converse to extensionality:
 
∀ ∀ X = Y ⇒ ∀ (z ∈ X ⇔ z ∈ Y ) .
X Y z

Before we discuss further consequences of extensionality, we would like to have the


existence of the empty set. However, Axioms 0 and 1 do not suffice to prove the exis-
tence of an empty set (see [Kun12, I.6.3]). This, rather, needs the additional axiom of
comprehension. More precisely, in the case of comprehension, we do not have a single
axiom, but a scheme of infinitely many axioms, one for each set-theoretic formula. Its
formulation makes use of the following definition:

Definition A.7. One obtains the universal closure of a set-theoretic formula φ, by


writing ∀ in front of φ for each variable vj that occurs as a free variable in φ (recall
vj
from Def. 1.31 that vj is free in φ if, and only if, it is not bound by a quantifier in φ).

Axiom 2 Comprehension Scheme: For each set-theoretic formula φ, not containing Y


as a free variable, the universal closure of
 
∃ ∀ x ∈ Y ⇔ (x ∈ X ∧ φ)
Y x

is an axiom. Thus, the comprehension scheme states that, given the set X,
there exists (at least one) set Y , containing precisely the elements of X that
have the property φ.

Remark A.8. Comprehension does not provide uniqueness. However, if both Y and
Y ′ are sets containing precisely the elements of X that have the property φ, then
 

∀ x ∈ Y ⇔ (x ∈ X ∧ φ) ⇔ x ∈ Y ,
x

and, then, extensionality implies Y = Y ′ . Thus, due to extensionality, the set Y given
by comprehension is unique, justifying the notation

{x : x ∈ X ∧ φ} := {x ∈ X : φ} := Y (A.3)

(this is the axiomatic justification for (1.6)).

Theorem A.9. There exists a unique empty set (which we denote by ∅ or by 0 – it is


common to identify the empty set with the number zero in axiomatic set theory).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 192

Proof. Axiom 0 provides the existence of a set X. Then comprehension allows us to


define the empty set by
0 := ∅ := {x ∈ X : x 6= x},
where, as explained in Rem. A.8, extensionality guarantees uniqueness. 

Remark A.10. In Rem. A.4 we said that every formula with additional symbols and
notation is to be regarded as an abbreviation or transcription of a set-theoretic formula
as defined in Def. A.1(b). Thus, formulas containing symbols for defined sets (e.g. 0
or ∅ for the empty set) are to be regarded as abbreviations for formulas without such
symbols. Some logical subtleties arise from the fact that there is some ambiguity in the
way such abbreviations can be resolved: For example, 0 ∈ X can abbreviate either
   
ψ : ∃ φ(y) ∧ y ∈ X or χ : ∀ φ(y) ⇒ y ∈ X , where φ(y) stands for ∀ (v ∈ / y).
y y v

Then ψ and χ are equivalent if ∃! φ(y) is true (e.g., if Axioms 0 – 2 hold), but they can
y
be nonequivalent, otherwise (see discussion between Lem. 2.9 and Lem. 2.10 in [Kun80]).

At first glance, the role played by the free variables in φ, which are allowed to occur
in Axiom 2, might seem a bit obscure. So let us consider examples to illustrate that
allowing free variables (i.e. set parameters) in comprehension is quite natural:

Example A.11. (a) Suppose φ in comprehension is the formula x ∈ Z (having Z as a


free variable), then the set given by the resulting axiom is merely the intersection
of X and Z:
X ∩ Z := {x ∈ X : φ} = {x ∈ X : x ∈ Z}.

(b) Note that it is even allowed for φ in comprehension to have X as a free variable, so
one can let φ be the formula ∃ (x ∈ u ∧ u ∈ X) to define the set
u
n o
X ∗ := x ∈ X : ∃ (x ∈ u ∧ u ∈ X) .
u

Then, if 0 := ∅, 1 := {0}, 2 := {0, 1}, we obtain

2∗ = {0} = 1.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 193

It is a consequence of extensionality that the mathematical universe consists of sets


and only of sets: Suppose there were other objects in the mathematical universe, for
example a cow C and a monkey M (or any other object without elements, other than
the empty set) – this would be equivalent to allowing a cow or a monkey (or any other
object without elements, other than the empty set) to be considered a set, which would
mean that our set-theoretic variables vj were allowed to be a cow or a monkey as well.
However, extensionality then implies the false statement C = M = ∅, thereby excluding
cows and monkeys from the mathematical universe.
Similarly, {C} and {M } (or any other object that contains a non-set), can not be inside
the mathematical universe. Indeed, otherwise we had
 
∀ x ∈ {C} ⇔ x ∈ {M }
x

(as C and M are non-sets) and, by extensionality, {C} = {M } were true, in contradic-
tion to a set with a cow inside not being the same as a set with a monkey inside. Thus,
we see that all objects of the mathematical universe must be so-called hereditary sets,
i.e. sets all of whose elements (thinking of the elements as being the children of the sets)
are also sets.

A.3.2 Classes

As we need to avoid contradictions such as Russell’s antinomy, we must not require the
existence of a set {x : φ} for each set-theoretic formula φ. However, it can still be
useful to think of a “collection” of all sets having the property φ. Such collections are
commonly called classes:

Definition A.12. (a) If φ is a set-theoretic formula, then we call {x : φ} a class,


namely the class of all sets that have the property φ (typically, φ will have x as a
free variable).

(b) If φ is a set-theoretic formula, then we say the class {x : φ} exists (as a set) if, and
only if   
∃ ∀ x∈X ⇔ φ (A.4)
X x

is true. Then X is actually unique by extensionality and we identify X with the


class {x : φ}. If (A.4) is false, then {x : φ} is called a proper class (and the usual
interpretation is that the class is in some sense “too large” to be a set).

Example A.13. (a) Due to Russell’s antinomy of Sec. A.1, we know that R := {x :
x∈
/ x} forms a proper class.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 194

(b) The universal class of all sets, V := {x : x = x}, is a proper class. Once again,
this is related to Russell’s antinomy: If V were a set, then

R = {x : x ∈
/ x} = {x : x = x ∧ x ∈
/ x} = {x : x ∈ V ∧ x ∈
/ x}

would also be a set by comprehension. However, this is in contradiction to R being


a proper class by (a).

Remark A.14. From the perspective of formal logic, statements involving proper
classes are to be regarded as abbreviations for statements without proper classes. For
example, it turns out that the class G of all sets forming a group is a proper class. But
we might write G ∈ G as an abbreviation for the statement “The set G is a group.”

A.3.3 Pairing, Union, Replacement

Axioms 0 – 2 are still consistent with the empty set being the only set in existence (see
[Kun12, I.6.13]). The next axiom provides the existence of nonempty sets:

Axiom 3 Pairing:
∀ ∀ ∃ (x ∈ Z ∧ y ∈ Z).
x y Z

Thus, the pairing axiom states that, for all sets x and y, there exists a set Z
that contains x and y as elements.

In consequence of the pairing axiom, the sets

0 := ∅, (A.5a)
1 := {0}, (A.5b)
2 := {0, 1} (A.5c)

all exist. More generally, we may define:

Definition A.15. If x, y are sets and Z is given by the pairing axiom, then we call

(a) {x, y} := {u ∈ Z : u = x ∨ u = y} the unordered pair given by x and y,

(b) {x} := {x, x} the singleton set given by x,

(c) (x, y) := {{x}, {x, y}} the ordered pair given by x and y (cf. Def. 2.1).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 195

We can now show that ordered pairs behave as expected:


Lemma A.16. The following holds true:
 
∀′ ′ (x, y) = (x′ , y ′ ) ⇔ (x = x′ ) ∧ (y = y ′ ) .
x,y,x ,y

Proof. “⇐” is merely


x=x′ , y=y ′
(x, y) = {{x}, {x, y}} = {{x′ }, {x′ , y ′ }} = (x′ , y ′ ).

“⇒” is done by distinguishing two cases: If x = y, then

{{x}} = (x, y) = (x′ , y ′ ) = {{x′ }, {x′ , y ′ }}.

Next, by extensionality, we first get {x} = {x′ } = {x′ , y ′ }, followed by x = x′ = y ′ ,


establishing the case. If x 6= y, then

{{x}, {x, y}} = (x, y) = (x′ , y ′ ) = {{x′ }, {x′ , y ′ }},

where, by extensionality {x} 6= {x, y} 6= {x′ }. Thus, using extensionality again, {x} =
{x′ } and x = x′ . Next, we conclude

{x, y} = {x′ , y ′ } = {x, y ′ }

and a last application of extensionality yields y = y ′ . 

While we now have the existence of the infinitely many different sets 0, {0}, {{0}}, . . . ,
we are not, yet, able to form sets containing more than two elements. This is remedied
by the following axiom:

Axiom 4 Union:  
∀ ∃ ∀ ∀ (x ∈ X ∧ X ∈ M) ⇒ x ∈ Y .
M Y x X

Thus, the union axiom states that, for each set of sets M, there exists a set
Y containing all elements of elements of M.
Definition A.17. (a) If M is a set and Y is given by the union axiom, then define
[ [  
M := X := x ∈ Y : ∃ x∈X .
X∈M
X∈M

(b) If X and Y are sets, then define


[
X ∪ Y := {X, Y }.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 196

(c) If x, y, z are sets, then define

{x, y, z} := {x, y} ∪ {z}.

Remark A.18. (a) The definition of set-theoretic unions as


[  
Ai := x : ∃ x ∈ Ai
i∈I
i∈I

in (1.25b) will be equivalent to the definition in Def. A.17(a) if we are allowed to


form the set
M := {Ai : i ∈ I}.
If I is a set and Ai is a set for each i ∈ I, then M as above will be a set by Axiom
5 below (the axiom of replacement).

(b) In contrast to unions, intersections can be obtained directly from comprehension


without the introduction of an additional axiom: For example

X ∩ Y := {x ∈ X : x ∈ Y },
\  
Ai := x ∈ Ai0 : ∀ x ∈ Ai ,
i∈I
i∈I

where i0 ∈ I 6= ∅ is an arbitrary fixed element of I.

(c) The union [ [ [


∅= X= Ai = ∅
X∈∅ i∈∅

is the empty set – in particular, a set. However,


\     \
∅ = x : ∀ x ∈ X = V = x : ∀ x ∈ Ai = Ai ,
X∈∅ i∈∅
i∈∅

i.e. the intersection over the empty set is the class of all sets – in particular, a proper
class and not a set.
Definition A.19. We define the successor function

x 7→ S(x) := x ∪ {x} (for each set x).

Thus, recalling (A.5), we have 1 = S(0), 2 = S(1); and we can define 3 := S(2), . . . In
general, we call the set S(x) the successor of the set x.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 197

In Def. 2.3 and Def. 2.19, respectively, we define functions and relations in the usual
manner, making use of the Cartesian product A × B of two sets A and B, which,
according to (2.2) consists of all ordered pairs (x, y), where x ∈ A and y ∈ B. However,
Axioms 0 – 4 are not sufficient to justify the existence of Cartesian products. To obtain
Cartesian products, we employ the axiom of replacement. Analogous to the axiom
of comprehension, the following axiom of replacement actually consists of a scheme of
infinitely many axioms, one for each set-theoretic formula:

Axiom 5 Replacement Scheme: For each set-theoretic formula, not containing Y as a


free variable, the universal closure of
   
∀ ∃! φ ⇒ ∃ ∀ ∃ φ
x∈X y Y x∈X y∈Y

is an axiom. Thus, the replacement scheme states that if, for each x ∈ X,
there exists a unique y having the property φ (where, in general, φ will depend
on x), then there exists a set Y that, for each x ∈ X, contains this y with
property φ. One can view this as obtaining Y by replacing each x ∈ X by
the corresponding y = y(x).

Theorem A.20. If A and B are sets, then the Cartesian product of A and B, i.e. the
class  
A × B := x : ∃ ∃ x = (a, b)
a∈A b∈B

exists as a set.

Proof. For each a ∈ A, we can use replacement with X := B and φ := φa being the
formula y = (a, x) to obtain the existence of the set

{a} × B := {(a, x) : x ∈ B} (A.6a)

(in the usual way, comprehension and extensionality were used as well). Analogously,
using replacement again with X := A and φ being the formula y = {x} × B, we obtain
the existence of the set
M := {{x} × B : x ∈ A}. (A.6b)
In a final step, the union axiom now shows
[ [
M= {a} × B = A × B (A.6c)
a∈A

to be a set as well. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 198

A.3.4 Infinity, Ordinals, Natural Numbers

The following axiom of infinity guarantees the existence of infinite sets (e.g., it will allow
us to define the set of natural numbers N, which is infinite by Th. A.46 below).

Axiom 6 Infinity:  
∃ 0∈X ∧ ∀ (x ∪ {x} ∈ X) .
X x∈X

Thus, the infinity axiom states the existence of a set X containing ∅ (iden-
tified with the number 0), and, for each of its elements x, its successor
S(x) = x ∪ {x}.

In preparation for our official definition of N in Def. A.41 below, we will study so-called
ordinals, which are special sets also of further interest to the field of set theory (the
natural numbers will turn out to be precisely the finite ordinals). We also need some
notions from the theory of relations, in particular, order relations (cf. Def. 2.19 and Def.
2.25).

Definition A.21. Let R be a relation on a set X.

(a) R is called asymmetric if, and only if,


 
∀ xRy ⇒ ¬(yRx) , (A.7)
x,y∈X

i.e. if x is related to y only if y is not related to x.

(b) R is called a strict partial order if, and only if, R is asymmetric and transitive. It is
noted that this is consistent with Not. 2.26, since, recalling the notation ∆(X) :=
{(x, x) : x ∈ X}, R is a partial order on X if, and only if, R \ ∆(X) is a strict
partial order on X. We extend the notions lower/upper bound, min, max, inf, sup
of Def. 2.27 to strict partial orders R by applying them to R ∪∆(X): We call x ∈ X
a lower bound of Y ⊆ X with respect to R if, and only if, x is a lower bound of Y
with respect to R ∪ ∆(X), and analogous for the other notions.

(c) A strict partial order R is called a strict total order or a strict linear order if, and
only if, for each x, y ∈ X, one has x = y or xRy or yRx.

(d) R is called a (strict) well-order if, and only if, R is a (strict) total order and every
nonempty subset of X has a min with respect to R (for example, the usual ≤
constitutes a well-order on N (see Th. D.5 below), but not on R (e.g., R+ does not
have a min)).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 199

(e) If Y ⊆ X, then the relation on Y defined by


xSy :⇔ xRy
is called the restriction of R to Y , denoted S = R↾Y (usually, one still writes R for
the restriction).
Lemma A.22. Let R be a relation on a set X and Y ⊆ X.

(a) If R is transitive, then R↾Y is transitive.


(b) If R is reflexive, then R↾Y is reflexive.
(c) If R is antisymmetric, then R↾Y is antisymmetric.
(d) If R is asymmetric, then R↾Y is asymmetric.
(e) If R is a (strict) partial order, then R↾Y is a (strict) partial order.
(f ) If R is a (strict) total order, then R↾Y is a (strict) total order.
(g) If R is a (strict) well-order, then R↾Y is a (strict) well-order.

Proof. (a): If a, b, c ∈ Y with aRb and bRc, then aRc, since a, b, c ∈ X and R is transitive
on X.
(b): If a ∈ Y , then a ∈ X and aRa, since R is reflexive on X.
(c): If a, b ∈ Y with aRb and bRa, then a = b, since a, b ∈ X and R is antisymmetric
on X.
(d): If a, b ∈ Y with aRb, then ¬bRa, since a, b ∈ X and R is asymmetric on X.
(e) follows by combining (a) – (d).
(f): If a, b ∈ Y with a = b and ¬aRb, then bRa, since a, b ∈ X and R is total on X.
Combining this with (e) yields (f).
(g): Due to (f), it merely remains to show that every nonempty subset Z ⊆ Y has a
min. However, since Z ⊆ X and R is a well-order on X, there is m ∈ Z such that m is
a min for R on X, implying m to be a min for R on Y as well. 
Remark A.23. Since the universal class V is not a set, ∈ is not a relation in the sense
of Def. 2.19. It can be considered as a “class relation”, i.e. a subclass of V × V, but
it is a proper class. However, ∈ does constitute a relation in the sense of Def. 2.19 on
each set X (recalling that each element of X must be a set as well). More precisely, if
X is a set, then so is
R∈ := {(x, y) ∈ X × X : x ∈ y}. (A.8a)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 200

Then
∀ (x, y) ∈ R∈ ⇔ x ∈ y. (A.8b)
x,y∈X

Definition A.24. A set X is called transitive if, and only if, every element of X is also
a subset of X:
∀ x ⊆ X. (A.9a)
x∈X

Clearly, (A.9a) is equivalent to


 
∀ x∈y ∧ y∈X ⇒ x∈X . (A.9b)
x,y

Lemma A.25. If X, Y are transitive sets, then X ∩ Y is a transitive set.

Proof. If x ∈ X ∩ Y and y ∈ x, then y ∈ X (since X is transitive) and y ∈ Y (since Y


is transitive). Thus y ∈ X ∩ Y , showing X ∩ Y is transitive. 
Definition A.26. (a) A set α is called an ordinal number or just an ordinal if, and only
if, α is transitive and ∈ constitutes a strict well-order on α. An ordinal α is called a
successor ordinal if, and only if, there exists an ordinal β such that α = S(β), where
S is the successor function of Def. A.19. An ordinal α 6= 0 is called a limit ordinal
if, and only if, it is not a successor ordinal. We denote the class of all ordinals by
ON (it is a proper class by Cor. A.33 below).

(b) We define

∀ (α < β :⇔ α ∈ β), (A.10a)


α,β∈ON

∀ (α ≤ β :⇔ α < β ∨ α = β). (A.10b)


α,β∈ON

Example A.27. Using (A.5), 0 = ∅ is an ordinal, and 1 = S(0), 2 = S(1) are both
successor ordinals (in Prop. A.43, we will identify N0 as the smallest limit ordinal). Even
though X := {1} and Y := {0, 2} are well-ordered by ∈, they are not ordinals, since
they are not transitive sets: 1 ∈ X, but 1 6⊆ X (since 0 ∈ 1, but 0 ∈ / X); similarly,
1 ∈ 2 ∈ Y , but 1 ∈
/ Y.
Lemma A.28. No ordinal contains itself, i.e.

∀ α∈
/ α.
α∈ON

Proof. If α is an ordinal, then ∈ is a strict order on α. Due to asymmetry of strict


orders, x ∈ x can not be true for any element of α, implying that α ∈ α can not be
true. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 201

Proposition A.29. Every element of an ordinal is an ordinal, i.e.


 
∀ X ∈ α ⇒ X ∈ ON
α∈ON

(in other words, ON is a transitive class).

Proof. Let α ∈ ON and X ∈ α. Since α is transitive, we have X ⊆ α. As ∈ is a


strict well-order on α, it must also be a strict well-order on X by Lem. A.22(g). In
consequence, it only remains to prove that X is transitive as well. To this end, let
x ∈ X. Then x ∈ α, as α is transitive. If y ∈ x, then, using transitivity of α again,
y ∈ α. Now y ∈ X, as ∈ is transitive on α, proving x ⊆ X, i.e. X is transitive. 
Proposition A.30. If α, β ∈ ON, then X := α ∩ β ∈ ON (we will see in Th. A.35(a)
below that, actually, α ∩ β = min{α, β}).

Proof. X is transitive by Lem. A.25, and, since X ⊆ α, ∈ is a strict well-order on X by


Lem. A.22(g). 
Proposition A.31. On the class ON, the relation ≤ (as defined in (A.10)) is the same
as the relation ⊆, i.e.
 
∀ α ≤ β ⇔ α ⊆ β ⇔ (α ∈ β ∨ α = β) . (A.11)
α,β∈ON

Proof. Let α, β ∈ ON.


Assume α ≤ β. If α = β, then α ⊆ β. If α ∈ β, then α ⊆ β, since β is transitive.
Conversely, assume α ⊆ β and α 6= β. We have to show α ∈ β. To this end, we set
X := β \ α. Then X 6= ∅ and, as ∈ well-orders β, we can let m := min X. We will show
m = α (note that this will complete the proof, due to α = m ∈ X ⊆ β). If µ ∈ m, then
µ ∈ β (since m ∈ β and β is transitive) and µ ∈/ X (since m = min X), implying µ ∈ α
(since X = β \ α) and, thus, m ⊆ α. Seeking a contradiction, assume m 6= α. Then
there must be some γ ∈ α \ m ⊆ α ⊆ β. In consequence γ, m ∈ β. As γ ∈ / m and ∈ is
a total order on β, we must have either m = γ or m ∈ γ. However, m 6= γ, since γ ∈ α
and m ∈ / α (as m ∈ X). So it must be m ∈ γ ∈ α, implying m ∈ α, as β is transitive.
This contradiction proves m = α and establishes the proposition. 
Theorem A.32. The class ON is well-ordered by ∈, i.e.

(i) ∈ is transitive on ON:


 
∀ α<β ∧ β<γ ⇒ α<γ .
α,β,γ∈ON

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 202

(ii) ∈ is asymmetric on ON:


 
∀ α < β ⇒ ¬(β < α) .
α,β∈ON

(iii) Ordinals are always comparable:


 
∀ α<β ∨ β<α ∨ α=β .
α,β∈ON

(iv) Every nonempty set of ordinals has a min.

Proof. (i) is clear, as γ is a transitive set.


(ii): If α, β ∈ ON, then α ∈ β ∈ α implies α ∈ α by (i), which is a contradiction to
Lem. A.28.
(iii): Let γ := α ∩ β. Then γ ∈ ON by Prop. A.30. Thus
Lem. A.31
γ⊆α ∧ γ⊆β ⇒ (γ ∈ α ∨ γ = α) ∧ (γ ∈ β ∨ γ = β). (A.12)
If γ ∈ α and γ ∈ β, then γ ∈ α ∩ β = γ, in contradiction to Lem. A.28. Thus, by (A.12),
γ = α or γ = β. If γ = α, then α ⊆ β. If γ = β, then β ⊆ α, completing the proof of
(iii).
(iv): Let X be a nonempty set of ordinals and consider α ∈ X. If α = min X, then
we are already done. Otherwise, Y := α ∩ X = {β ∈ X : β ∈ α} 6= ∅. Since α is
well-ordered by ∈, there is m := min Y . If β ∈ X, then either β < α or α ≤ β by (iii).
If β < α, then β ∈ Y and m ≤ β. If α ≤ β, then m < α ≤ β. Thus, m = min X,
proving (iv). 
Corollary A.33. ON is a proper class (i.e. there is no set containing all the ordinals).

Proof. If there is a set X containing all ordinals, then, by comprehension, β := ON =


{α ∈ X : α is an ordinal} must be a set as well. But then Prop. A.29 says that the set
β is transitive and Th. A.32 yields that the set β is well-ordered by ∈, implying β to be
an ordinal, i.e. β ∈ β in contradiction to Lem. A.28. 
Corollary A.34. For each set X of ordinals, we have:

(a) X is well-ordered by ∈.
(b) X is an ordinal if, and only if, X is transitive. Note: A transitive set of ordinals
X is sometimes called an initial segment of ON, since, here, transitivity can be
restated in the form
 
∀ ∀ α<β ⇒ α∈X . (A.13)
α∈ON β∈X

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 203

Proof. (a) is a simple consequence of Th. A.32(i)-(iv).


(b) is immediate from (a). 

Theorem A.35. Let X be a nonempty set of ordinals.


T
(a) Then γ := X is an ordinal, namely γ = min X. In particular, if α, β ∈ ON,
then min{α, β} = α ∩ β.
S
(b) Then δ := X is an ordinal, namely δ = sup X. In particular, if α, β ∈ ON, then
max{α, β} = α ∪ β.

Proof. (a): Let m := min X. Then γ ⊆ m, since m ∈ X. Conversely, if α ∈ X, then


m ≤ α implies m ⊆ α by Prop. A.31, i.e. m ⊆ γ. Thus, m = γ, proving (a).
(b): To show δ ∈ ON, we need to show δ is transitive (then δ is an ordinal by Cor.
A.34(b)). If α ∈ δ, then there is β ∈ X such that α ∈ β. Thus, if γ ∈ α, then γ ∈ β,
since β is transitive. As γ ∈ β implies γ ∈ δ, we see that δ is transitive, as needed. It
remains to show δ = sup X. If α ∈ X, then α ⊆ δ, i.e. α ≤ δ, showing δ to be an upper
bound for X. Now let u ∈ ON be an arbitrary upper bound for X, i.e.

∀ α ⊆ u.
α∈X

Thus, δ ⊆ u, i.e. δ ≤ u, proving δ = sup X. 

Next, we obtain some results regarding the successor function of Def. A.19 in the context
of ordinals.

Lemma A.36. We have


 
∀ x, y ∈ S(α) ∧ x ∈ y ⇒ x 6= α .
α∈ON

Proof. Seeking a contradiction, we reason as follows:


α∈α
/ y∈S(α) α transitive x∈y
x = α ⇒ y 6= α ⇒ y∈α ⇒ y ⊆ α ⇒ α ∈ α.

This contradiction to α ∈
/ α yields x 6= α, concluding the proof. 

Proposition A.37. For each α ∈ ON, the following holds:

(a) S(α) ∈ ON.

(b) α < S(α).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 204

(c) For each ordinal β, β < S(α) holds if, and only if, β ≤ α.

(d) For each ordinal β, if β < α, then S(β) < S(α).

(e) For each ordinal β, if S(β) < S(α), then β < α.

Proof. (a): Due to Prop. A.29, S(α) is a set of ordinals. Thus, by Cor. A.34(b), it
merely remains to prove that S(α) is transitive. Let x ∈ S(α). If x = α, then x =
α ⊆ α ∪ {α} = S(α). If x 6= α, then x ∈ α and, since α is transitive, this implies
x ⊆ α ⊆ S(α), showing S(α) to be transitive, thereby completing the proof of (a).
(b) holds, as α ∈ S(α) holds by the definition of S(α).
(c) is clear, since, for each ordinal β,

β < S(α) ⇔ β ∈ S(α) ⇔ β ∈ α ∨ β = α ⇔ β ≤ α.

(d): If β < α, then S(β) = β ∪ {β} ⊆ α, i.e. S(β) ≤ α < S(α).


(e) follows from (d) using contraposition: If ¬(β < α), then β = α or α < β, implying
S(β) = S(α) or S(α) < S(β), i.e. ¬(S(β) < S(α)). 

We now proceed to define the natural numbers:


Definition A.38. An ordinal n is called a natural number if, and only if,
 
n 6= 0 ∧ ∀ m ≤ n ⇒ m = 0 ∨ m is successor ordinal .
m∈ON

Proposition A.39. If n = 0 or n is a natural number, then S(n) is a natural number


and every element of n is a natural number or 0.

Proof. Suppose n is 0 or a natural number. If m ∈ n, then m is an ordinal by Prop.


A.29. Suppose m 6= 0 and k ∈ m. Then k ∈ n, since n is transitive. Since n is a natural
number, k = 0 or k is a successor ordinal. Thus, m is a natural number. It remains
to show that S(n) is a natural number. By definition, S(n) = n ∪ {n} 6= 0. Moreover,
S(n) ∈ ON by Prop. A.37(a), and, thus, S(n) is a successor ordinal. If m ∈ S(n), then
m ≤ n, implying m = 0 or m is a successor ordinal, completing the proof that S(n) is
a natural number. 
Theorem A.40 (Principle of Induction). If X is a set satisfying

0∈X ∧ ∀ S(x) ∈ X, (A.14)


x∈X

then X contains 0 and all natural numbers.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 205

Proof. Let X be a set satisfying (A.14). Then 0 ∈ X is immediate. Let n be a natural


number and, seeking a contradiction, assume n ∈
/ X. Consider N := S(n)\X. According
to Prop. A.39, S(n) is a natural number and all nonzero elements of S(n) are natural
numbers. Since N ⊆ S(n) and 0 ∈ X, 0 ∈ / N and all elements of N must be natural
numbers. As n ∈ N , N 6= 0. Since S(n) is well-ordered by ∈ and 0 6= N ⊆ S(n), N
must have a min m ∈ N , 0 6= m ≤ n. Since m is a natural number, there must be k
such that m = S(k). Then k < m, implying k ∈ / N . On the other hand

k < m ∧ m ≤ n ⇒ k ≤ n ⇒ k ∈ S(n).

Thus, k ∈ X, implying m = S(k) ∈ X, in contradiction to m ∈ N . This contradiction


proves n ∈ X, thereby establishing the case. 

Definition A.41. If the set X is given by the axiom of infinity, then we use compre-
hension to define the set

N0 := {n ∈ X : n = 0 ∨ n is a natural number}

and note N0 to be unique by extensionality. We also denote N := N0 \ {0}. In set theory,


it is also very common to use the symbol ω for the set N0 .

Corollary A.42. N0 is the set of all natural numbers and 0, i.e.


 
∀ n ∈ N0 ⇔ n = 0 ∨ n is a natural number .
n

Proof. “⇒” is clear from Def. A.41 and “⇐” is due to Th. A.40. 

Proposition A.43. ω = N0 is the smallest limit ordinal.

Proof. Since ω is a set of ordinals and ω is transitive by Prop. A.39, ω is an ordinal


by Cor. A.34(b). Moreover ω 6= 0, since 0 ∈ ω; and ω is not a successor ordinal (if
ω = S(n) = n ∪ {n}, then n ∈ ω and S(n) ∈ ω by Prop. A.39, in contradiction to
ω = S(n)), implying it is a limit ordinal. To see that ω is the smallest limit ordinal, let
α ∈ ON, α < ω. Then α ∈ ω, that means α = 0 or α is a natural number (in particular,
a successor ordinal). 

In the following Th. A.44, we will prove that N satisfies the Peano axioms P1 – P3 of
Sec. 3.1 (if one prefers, one can show the same for N0 , where 0 takes over the role of 1).

Theorem A.44. The set of natural numbers N satisfies the Peano axioms P1 – P3 of
Sec. 3.1.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 206

Proof. For P1 and P2, we have to show that, for each n ∈ N, one has S(n) ∈ N \ {1}
and that S(m) 6= S(n) for each m, n ∈ N, m 6= n. Let n ∈ N. Then S(n) ∈ N by Prop.
A.39. If S(n) = 1, then n < S(n) = 1 by Prop. A.37(b), i.e. n = 0, in contradiction to
n ∈ N. If m, n ∈ N with m 6= n, then S(m) 6= S(n) is due to Prop. A.37(d). To prove
P3, suppose A ⊆ N has the property that 1 ∈ A and S(n) ∈ A for each n ∈ A. We need
to show A = N (i.e. N ⊆ A, as A ⊆ N is assumed). Let X := A ∪ {0}. Then X satisfies
(A.14) and Th. A.40 yields N0 ⊆ X. Thus, if n ∈ N, then n ∈ X \ {0} = A, showing
N ⊆ A. 

Notation A.45. For each n ∈ N0 , we introduce the notation n + 1 := S(n) (more


generally, one also defines α + 1 := S(α) for each ordinal α).

Theorem A.46. Let n ∈ N0 . Then A := N0 \ n is infinite (see Def. 3.12(b)). In


particular, N0 and N = N0 \ {0} = N0 \ 1 are infinite.

Proof. Since n ∈ / n, we have n ∈ A 6= ∅. Thus, if A were finite, then there were a


bijection f : A −→ Am := {1, . . . , m} = {k ∈ N : k ≤ m} for some m ∈ N. However,
we will show by induction on m ∈ N that there is no injective map f : A −→ Am . Since
S(n) ∈ / n, we have S(n) ∈ A. Thus, if f : A −→ A1 = {1}, then f (n) = f (S(n)),
showing that f is not injective and proving the cases m = 1. For the induction step, we
proceed by contraposition and show that the existence of an injective map f : A −→
Am+1 , m ∈ N, (cf. Not. A.45) implies the existence of an injective map g : A −→ Am .
To this end, let m ∈ N and f : A −→ Am+1 be injective. If m + 1 ∈ / f (A), then f itself
is an injective map into Am . If m + 1 ∈ f (A), then there is a unique a ∈ A such that
f (a) = m + 1. Define
(
f (k) for k < a,
g : A −→ Am , g(k) := (A.15)
f (k + 1) for a ≤ k.

Then g is well-defined: If k ∈ A and a ≤ k, then k +1 ∈ A\{a}, and, since f is injective,


g does, indeed, map into Am . We verify g to be injective: If k, l ∈ A, k < l, then also
k < l + 1 and k + 1 6= l + 1 (by Peano axiom P2 – k + 1 < l + 1 then also follows, but we
do not make use of that here). In each case, g(k) 6= g(l), proving g to be injective. 

For more basic information regarding ordinals see, e.g., [Kun12, Sec. I.8].

A.3.5 Power Set

There is one more basic construction principle for sets that is not covered by Axioms 0
– 6, namely the formation of power sets. This needs another axiom:

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 207

Axiom 7 Power Set:  


∀ ∃ ∀ Y ⊆X ⇒ Y ∈M .
X M Y
Thus, the power set axiom states that, for each set X, there exists a set M
that contains all subsets Y of X as elements.
Definition A.47. If X is a set and M is given by the power set axiom, then we call
P(X) := {Y ∈ M : Y ⊆ X}
the power set of X. Another common notation for P(X) is 2X (cf. Prop. 2.18).

A.3.6 Foundation

Foundation is, perhaps, the least important of the axioms in ZF. It basically cleanses
the mathematical universe of unnecessary “clutter”, i.e. of certain pathological sets that
are of no importance to standard mathematics anyway.

Axiom 8 Foundation:
  
∀ ∃ (x ∈ X) ⇒ ∃ ¬∃ z ∈ x ∧ z ∈ X .
X x x∈X z

Thus, the foundation axiom states that every nonempty set X contains an
element x that is disjoint to X.
Theorem A.48. Due to the foundation axiom, the ∈ relation can have no cycles, i.e.
there do not exist sets x1 , x2 , . . . , xn , n ∈ N, such that
x1 ∈ x2 ∈ · · · ∈ xn ∈ x1 . (A.16a)
In particular, sets can not be members of themselves:
¬∃ x ∈ x. (A.16b)
x

Proof. If there were sets x1 , x2 , . . . , xn , n ∈ N, such that (A.16a) were true, then, by
using the pairing axiom and the union axiom, we could form the set
X := {x1 , . . . , xn }.
Then, in contradiction to the foundation axiom, X ∩ xi 6= ∅, for each i = 1, . . . , n:
Indeed, xn ∈ X ∩ x1 , and xi−1 ∈ X ∩ xi for each i = 2, . . . , n. 

For a detailed explanation, why “sets” forbidden by foundation do not occur in standard
mathematics, anyway, see, e.g., [Kun12, Sec. I.14].

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 208

A.4 The Axiom of Choice


In addition to the axioms of ZF discussed in the previous section, there is one more
axiom, namely the axiom of choice (AC) that, together with ZF, makes up ZFC, the
axiom system at the basis of current standard mathematics. Even though AC is used
and accepted by most mathematicians, it does have the reputation of being somewhat
less “natural”. Thus, many mathematicians try to avoid the use of AC, where possible,
and it is often pointed out explicitly, if a result depends on the use of AC (but this
practice is by no means consistent, neither in the literature nor in this class, and one
might sometimes be surprised, which seemingly harmless result does actually depend on
AC in some subtle nonobvious way). We will now state the axiom:

Axiom 9 Axiom of Choice (AC):


 
 
∀ ∅ ∈
/M ⇒ ∃ ∀ f (M ) ∈ M  .
M M ∈M
S
f : M−→ N
N ∈M

Thus, the axiom of choice postulates, for each nonempty set M, whose ele-
ments are all nonempty sets, the existence of a choice function, that means
a function that assigns, to each M ∈ M, an element m ∈ M .
Example A.49. For example, the axiom of choice postulates, for each nonempty set
A, the existence of a choice function on P(A) \ {∅} that assigns each subset of A one of
its elements.

The axiom of choice is remarkable since, at first glance, it seems so natural that one
can hardly believe it is not provable from the axioms in ZF. However, one can actually
show that it is neither provable nor disprovable from ZF (see, e.g., [Jec73, Th. 3.5, Th.
5.16] – such a result is called an independence proof, see [Kun80] for further material). If
you want to convince yourself that the existence of choice functions is, indeed, a tricky
matter, try to define a choice function on P(R) \ {∅} without AC (but do not spend too
much time on it – one can show this is actually impossible to accomplish).
Theorem A.52 below provides several important equivalences of AC. Its statement and
proof needs some preparation. We start by introducing some more relevant notions from
the theory of partial orders:
Definition A.50. Let X be a set and let ≤ be a partial order on X.

(a) An element m ∈ X is called maximal (with respect to ≤) if, and only if, there exists
no x ∈ X such that m < x (note that a maximal element does not have to be a
max and that a maximal element is not necessarily unique).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 209

(b) A nonempty subset C of X is called a chain if, and only if, C is totally ordered by
≤. Moreover, a chain C is called maximal if, and only if, no strict superset Y of C
(i.e. no Y ⊆ X such that C ( Y ) is a chain.

The following lemma is a bit technical and will be used to prove the implication AC
⇒ (ii) in Th. A.52 (other proofs in the literature often make use of so-called transfinite
recursion, but that would mean further developing the theory of ordinals, and we will
not pursue this route in this class).
Lemma A.51. Let X be a set and let ∅ 6= M ⊆ P(X) be a nonempty set of subsets of
X. We let M be partially ordered by inclusion, i.e. setting A ≤ B :⇔ A ⊆ B for each
A, B ∈ M. Moreover, define
[ [
∀ S := S (A.17)
S⊆M
S∈S

and assume  [ 
∀ C is a chain ⇒ C∈M . (A.18)
C⊆M

If the function g : M −→ M has the property that


 
∀ M ⊆ g(M ) ∧ #(g(M ) \ M ) ≤ 1 , (A.19)
M ∈M

then g has a fixed point, i.e.


∃ g(M ) = M. (A.20)
M ∈M

Proof. Fix some arbitrary M0 ∈ M. We call T ⊆ M an M0 -tower if, and only if, T
satisfies the following three properties

(i) M0 ∈ T .
S
(ii) If C ⊆ T is a chain, then C ∈T.

(iii) If M ∈ T , then g(M ) ∈ T .

Let T := {T ⊆ M : T is an M0 -tower}. If T1 := {M ∈ M : M0 ⊆ M }, then, clearly,


T1 is an M0 -tower and,
T in particular, T 6= ∅. Next, we note that the intersection of all
M0 -towers, i.e. T0 := T ∈T T , is also an M0 -tower. Clearly, no strict subset of T0 can
be an M0 -tower and
M ∈ T0 ⇒ M ∈ T1 ⇒ M0 ⊆ M. (A.21)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 210

The main work of the rest of the proof consists of showing that T0 is a chain. To show
T0 to be a chain, define
 
Γ := M ∈ T0 : ∀ (M ⊆ N ∨ N ⊆ M ) . (A.22)
N ∈T0

We intend to show that Γ = T0 by verifying that Γ is an M0 -tower. As an intermediate


step, we define

∀ Φ(M ) := {N ∈ T0 : N ⊆ M ∨ g(M ) ⊆ N }
M ∈Γ

and also show each Φ(M ) to be an M0 -tower. Actually, Γ and each Φ(M )Ssatisfy (i)
due to (A.21). To verify Γ satisfies (ii), let C ⊆ Γ be a chain and U := C. Then
U ∈ T0 , since T0 satisfies (ii). If N ∈ T0 , and C ⊆ N for each C ∈ C, then U ⊆ N . If
N ∈ T0 , and there is C ∈ C such that C 6⊆ N , then N ⊆ C (since C ∈ Γ), i.e. N ⊆ U ,
showing U ∈ Γ and Γ satisfyingS(ii). Now, let M ∈ Γ. To verify Φ(M ) satisfies (ii), let
C ⊆ Φ(M ) be a chain and U := C. Then U ∈ T0 , since T0 satisfies (ii). If U ⊆ M , then
U ∈ Φ(M ) as desired. If U 6⊆ M , then there is x ∈ U such that x ∈ / M . Thus, there
is C ∈ C such that x ∈ C and g(M ) ⊆ C (since C ∈ Φ(M )), i.e. g(M ) ⊆ U , showing
U ∈ Φ(M ) also in this case, and Φ(M ) satisfies (ii). We will verify that Φ(M ) satisfies
(iii) next. For this purpose, fix N ∈ Φ(M ). We need to show g(N ) ∈ Φ(M ). We already
know g(N ) ∈ T0 , as T0 satisfies (iii). As N ∈ Φ(M ), we can now distinguish three cases.
Case 1: N ( M . In this case, we cannot have M ( g(N ) (otherwise, #(g(N ) \ N ) ≥ 2
in contradiction to (A.19)). Thus, g(N ) ⊆ M (since M ∈ Γ), showing g(N ) ∈ Φ(M ).
Case 2: N = M . Then g(N ) = g(M ) ∈ Φ(M ) (since g(M ) ∈ T0 and g(M ) ⊆ g(M )).
Case 3: g(M ) ⊆ N . Then g(M ) ⊆ g(N ) by (A.19), again showing g(N ) ∈ Φ(M ). Thus,
we have verified that Φ(M ) satisfies (iii) and, therefore, is an M0 -tower. Then, by the
definition of T0 , we have T0 ⊆ Φ(M ). As we also have Φ(M ) ⊆ T0 (from the definition
of Φ(M )), we have shown
∀ Φ(M ) = T0 .
M ∈Γ

As a consequence, if N ∈ T0 and M ∈ Γ, then N ∈ Φ(M ) and this means N ⊆ M ⊆


g(M ) or g(M ) ⊆ N , i.e. each N ∈ T0 is comparable to g(M ), showing g(M ) ∈ Γ and Γ
satisfying (iii), completing the proof that Γ is an M0 -tower. As with the Φ(M ) above,
we conclude Γ = T0 , as desired. To conclude the proof of the lemma, we note Γ = T0
implies T0 is a chain. We claim that
[
M := T0

satisfies (A.20): Indeed, M ∈ T0 , since T0 satisfies (ii). Then g(M ) ∈ T0 , since T0


satisfies (iii). We then conclude g(M ) ⊆ M from the definition of M . As we always
have M ⊆ g(M ) by (A.19), we have established g(M ) = M and proved the lemma. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 211

Theorem A.52 (Equivalences to the Axiom of Choice). The following statements (i)
– (v) are equivalent to the axiom of choice (as stated as Axiom 9 above).
Q
(i) Every Cartesian product i∈I Ai of nonempty sets Ai , where I is a nonempty
index set, is nonempty (cf. Def. 2.15(c)).

(ii) Hausdorff’s Maximality Principle: Every nonempty partially ordered set X con-
tains a maximal chain (i.e. a maximal totally ordered subset).

(iii) Zorn’s Lemma: Let X be a nonempty partially ordered set. If every chain C ⊆ X
(i.e. every nonempty totally ordered subset of X) has an upper bound in X (such
chains with upper bounds are sometimes called inductive), then X contains a
maximal element (cf. Def. A.50(a)).

(iv) Zermelo’s Well-Ordering Theorem: Every set can be well-ordered (recall the defi-
nition of a well-order from Def. A.21(d)).

(v) Every vector space V over a field F has a basis B ⊆ V .

Proof. “(i) ⇔ AC”: Assume (i). Given a nonempty


Q set of nonempty sets M, let I := M
and, for each M ∈ M, let AM := M . If f ∈ M ∈I AM , then, according to Def. 2.15(c),
for each M ∈ I = M, one has f (M ) ∈ AM = M , proving AC holds. Conversely, assume
6= ∅ and eachSAi 6= ∅. Let M := {Ai : i ∈ I}.
AC. Consider a family (Ai )i∈I such that I S
Then, by AC, there is a map g : M −→ N ∈M N = j∈I Aj such that g(M ) ∈ M for
each M ∈ M. Then we can define
[
f : I −→ Aj , f (i) := g(Ai ) ∈ Ai ,
j∈I

to prove (i).
Next, we will show AC ⇒ (ii) ⇒ (iii) ⇒ (iv) ⇒ AC.
“AC ⇒ (ii)”: Assume AC and let X be a nonempty partially ordered set. Let M be the
set of all chains in X (i.e. the set of all nonempty totally ordered subsets of X). Then
∅∈/ M and M 6= ∅ (since X 6= ∅ and {x} ∈ M for each x ∈ X). Moreover, M satisfies
the hypothesis
S of Lem. A.51, since, if C ⊆ M is a chain of totally ordered subsets of X,
then C is a totally ordered subset of X, i.e. in M (here we have used the notation
of (A.17); also note that we are dealing with two different types of chains here, namely
those with respect to the order on X and those with respect to the order given by ⊆ on
M). Let f : P(X) \ {∅} −→ X be a choice function given by AC, i.e. such that

∀ f (Y ) ∈ Y.
Y ∈P(X)\{∅}

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 212

As an auxiliary notation, we set



∀ M ∗ := x ∈ X \ M : M ∪ {x} ∈ M .
M ∈M

With the intention of applying Lem. A.51, we define


(
M ∪ {f (M ∗ )} if M ∗ 6= ∅,
g : M −→ M, g(M ) :=
M if M ∗ = ∅.

Since g clearly satisfies (A.19), Lem. A.51 applies, providing an M ∈ M such that
g(M ) = M . Thus, M ∗ = ∅, i.e. M is a maximal chain, proving (ii).
“(ii) ⇒ (iii)”: Assume (ii). To prove Zorn’s lemma, let X be a nonempty set, partially
ordered by ≤, such that every chain C ⊆ X has an upper bound. Due to Hausdorff’s
maximality principle, we can assume C ⊆ X to be a maximal chain. Let m ∈ X be an
upper bound for the maximal chain C. We claim that m is a maximal element: Indeed,
if there were x ∈ X such that m < x, then x ∈ / C (since m is upper bound for C) and
C ∪ {x} would constitute a strict superset of C that is also a chain, contradicting the
maximality of C.
“(iii) ⇒ (iv)”: Assume (iii) and let X be a nonempty set. We need to construct a
well-order on X. Let W be the set of all well-orders on subsets of X, i.e.

W := (Y, W ) : Y ⊆ X ∧ W ⊆ Y × Y ⊆ X × X is a well-order on Y .

We define a partial order ≤ on W by setting



∀′ ′ (Y, W ) ≤ (Y ′ , W ′ ) :⇔ Y ⊆ Y ′ ∧ W = W ′ ↾Y
(Y,W ),(Y ,W )∈W

∧ (y ∈ Y, y ′ ∈ Y ′ , y ′ W ′ y ⇒ y ′ ∈ Y )

(recall the definition of the restriction of a relation from Def. A.21(e)). To apply Zorn’s
lemma to (W, ≤), we need to check that every chain C ⊆ W has an upper bound. To
this end, if C ⊆ W is a chain, let
[ [
UC := (YC , WC ), where YC := Y, WC := W.
(Y,W )∈C (Y,W )∈C

We need to verify UC ∈ W: If aWC b, then there is (Y, C) ∈ C such that aW b. In


particular, (a, b) ∈ Y × Y ⊆ YC × YC , showing WC to be a relation on YC . Clearly, WC
is a total order on YC (one just uses that, if a, b ∈ YC , then, as C is a chain, there is
(Y, W ) ∈ C such that a, b ∈ Y and W = WC ↾Y is a total order on Y ). To see that WC
is a well-order on YC , let ∅ 6= A ⊆ YC . If a ∈ A, then there is (Y, W ) ∈ C such that

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 213

a ∈ Y . Since W = WC ↾Y is a well-order on Y , we can let m := min Y ∩ A. We claim


that m = min A as well: Let b ∈ A. Then there is (B, U ) ∈ C such that b ∈ B. If
B ⊆ Y , then b ∈ Y ∩ A and mW b. If Y ⊆ B, then m, b ∈ B. If mU b, then we are
done. If bU m, then b ∈ Y (since (Y, W ) ≤ (B, U )), i.e., again, b ∈ Y ∩ A and mW b
(actually m = b in this case), proving m = min A. This completes the proof that WC is
a well-order on YC and, thus, shows UC ∈ W. Next, we check UC to be an upper bound
for C: If (Y, W ) ∈ C, then Y ⊆ YC and W = WC ↾Y are immediate. If y ∈ Y , y ′ ∈ YC ,
and y ′ WC y, then y ′ ∈ Y (otherwise, y ′ ∈ A with (A, U ) ∈ C, (Y, W ) ≤ (A, U ), y ′ U y, in
contradiction to y ′ ∈ / Y ). Thus, (Y, W ) ≤ UC , showing UC to be an upper bound for C.
By Zorn’s lemma, we conclude that W contains a maximal element (M, WM ). But then
M = X and WM is the desired well-order on X: Indeed, if there is x ∈ X \ M , then we
can let Y := M ∪ {x} and,
 
∀ aW b :⇔ (a, b ∈ M ∧ aWM b) ∨ b = x .
a,b∈Y

Then (Y, W ) ∈ W with (M, WM ) < (Y, W ) in contradiction to the maximality of


(M, WM ).
“(iv)
S ⇒ AC”: Assume (iv). Given a nonempty set of nonempty sets M, let X :=
M ∈M M . By (iv), there exists a well-order R on X. Then every nonempty Y ⊆ X
has a unique min. As every M ∈ M is a nonempty subset of X, we can define a choice
function
f : M −→ X, f (M ) := min M ∈ M,
proving AC.
“(v) ⇔ AC”: That every vector space has a basis is proved in [Phi19, Th. 5.23] by use
of Zorn’s lemma. That, conversely, (v) implies AC was first shown in [Bla84], but the
proof needs more algebraic tools than we have available in this class. There are also
some set-theoretic subtleties that distinguish this equivalence from the ones provided in
(i) – (iv), cf. Rem. A.53 below. 
Remark A.53. What is actually proved in [Bla84] is that Th. A.52(v) implies the
axiom of multiple choice (AMC), which states
 
 
  
∀ ∅ ∈
 /M ⇒ ∃ ∀ f (M ) ⊆ M ∧ 0 < #f (M ) < ∞  ,
M M ∈M
!
S
f : M−→P N
N ∈M

i.e., for each nonempty set M, whose elements are all nonempty sets, there exists a
function that assigns, to each M ∈ M, a finite subset of M . Neither the proof that (i)
– (iv) of Th. A.52 are each equivalent to AC nor the proof in [Bla84] that Th. A.52(v)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 214

implies AMC makes use of the axiom of foundation (Axiom 8 above). However, the
proof that AMC implies AC does require the axiom of foundation (cf., e.g., [Hal17, Ch.
6]).

A.5 Cardinality
A.5.1 Relations to Injective, Surjective, and Bijective Maps; Schröder-
Bernstein Theorem

Theorem A.54. Let M be a set of sets. Then the relation ∼ on M, defined by


A ∼ B :⇔ A and B have the same cardinality, (A.23)
constitutes an equivalence relation on M.

Proof. According to Def. 2.23, we have to prove that ∼ is reflexive, symmetric, and
transitive. According to Def. 3.12(a), A ∼ B holds for A, B ∈ M if, and only if, there
exists a bijective map f : A −→ B. Thus, since the identity Id : A −→ A is bijective,
A ∼ A, showing ∼ is reflexive. If A ∼ B, then there exists a bijective map f : A −→ B,
and f −1 is a bijective map f −1 : B −→ A, showing B ∼ A and that ∼ is symmetric.
If A ∼ B and B ∼ C, then there are bijective maps f : A −→ B and g : B −→ C.
Then, according to Th. 2.14, the composition (g ◦ f ) : A −→ C is also bijective, proving
A ∼ C and that ∼ is transitive. 

The next theorem provides two interesting, and sometimes useful, characterizations of
infinite sets:
Theorem A.55. Let A be a set. Using the axiom of choice (AC) of Sec. A.4, the
following statements (i) – (iii) are equivalent. More precisely, (ii) and (iii) are equivalent
even without AC (a set A is sometimes called Dedekind-infinite if, and only if, it satisfies
(iii)), (iii) implies (i) without AC, but AC is needed to show (i) implies (ii), (iii).

(i) A is infinite.
(ii) There exists M ⊆ A and a bijective map f : M −→ N.
(iii) There exists a strict subset B ( A and a bijective map g : A −→ B.

One sometimes expresses the equivalence between (i) and (ii) by saying that a set is
infinite if, and only if, it contains a copy of the natural numbers. The property stated in
(iii) might seem strange at first, but infinite sets are, indeed, precisely those identical in
size to some of their strict subsets (as an example think of the natural bijection n 7→ 2n
between all natural numbers and the even numbers).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 215

Proof. We first prove, without AC, the equivalence between (ii) and (iii).
“(ii) ⇒ (iii)”: Let E denote the even numbers. Then E ( N and h : N −→ E,
h(n) := 2n, is a bijection, showing that (iii) holds for the natural numbers. According to
(ii), there exists M ⊆ A and a bijective map f : M −→ N. Define B := (A\M ) ∪˙ f −1 (E)
and (
x for x ∈ A \ M ,
h : A −→ B, h(x) := −1
(A.24)
f ◦ h ◦ f (x) for x ∈ M .
Then B ( A since B does not contain the elements of M that are mapped to odd
numbers under f . Still, h is bijective, since h↾A\M = IdA\M and h↾M = f −1 ◦ h ◦ f is the
composition of the bijective maps f , h, and f −1 ↾E : E −→ f −1 (E).
“(iii) ⇒ (ii)”: As (iii) is assumed, there exist B ⊆ A, a ∈ A \ B, and a bijective map
g : A −→ B. Set
M := {an := g n (a) : n ∈ N}.
We show that an 6= am for each m, n ∈ N with m 6= n: Indeed, suppose m, n ∈ N with
n > m and an = am . Then, since g is bijective, we can apply g −1 m times to an = am
to obtain
a = (g −1 )m (am ) = (g −1 )m (an ) = g n−m (a).
Since l := n − m ≥ 1, we have a = g(g l−1 (a)), in contradiction to a ∈ A \ B. Thus, all
the an ∈ M are distinct and we can define f : M −→ N, f (an ) := n, which is clearly
bijective, proving (ii).
“(iii) ⇒ (i)”: The proof is conducted by contraposition, i.e. we assume A to be finite and
proof that (iii) does not hold. If A = ∅, then there is nothing to prove. If ∅ 6= A is finite,
then, by Def. 3.12(b), there exists n ∈ N and a bijective map f : A −→ {1, . . . , n}. If
B ( A, then, according to Th. A.64(a), there exists m ∈ N0 , m < n, and a bijective
map h : B −→ {1, . . . , m}. If there were a bijective map g : A −→ B, then h ◦ g ◦ f −1
were a bijective map from {1, . . . , n} onto {1, . . . , m} with m < n in contradiction to
Th. A.62.
“(i) ⇒ (ii)”: Inductively, we construct a strictly increasing sequence M1 ⊆ M2 ⊆ . . . of
subsets Mn of A n ∈ N, and a sequence of functions fn : Mn −→ {1, . . . , n} satisfying

∀ fn is bijective, (A.25a)
n∈N
 
∀ m ≤ n ⇒ f n ↾M m = f m : (A.25b)
m,n∈N

Since A 6= ∅, there exists m1 ∈ A. Set M1 := {m1 } and f1 : M1 −→ {1}, f1 (m1 ) := 1.


Then M1 ⊆ A and f1 bijective are trivially clear. Now let n ∈ N and suppose M1 , . . . , Mn
and f1 , . . . , fn satisfying (A.25) have already been constructed. Since A is infinite, there

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 216

must be mn+1 ∈ A\Mn (otherwise Mn = A and the bijectivity of fn : Mn −→ {1, . . . , n}


shows A is finite with #A = n; AC is used to select the mn+1 ∈ A \ Mn ). Set Mn+1 :=
Mn ∪ {mn+1 } and
(
fn (x) for x ∈ Mn ,
fn+1 : Mn+1 −→ {1, . . . , n + 1}, fn+1 (x) := (A.26)
n + 1 for x = mn+1 .

Then the bijectivity of fn implies the bijectivity of fn+1 , and, since fn+1 ↾Mn = fn holds
by definition of fn+1 , the implication

m≤n+1 ⇒ fn+1 ↾Mm = fm

holds true as well. An induction also shows Mn = {m1 , . . . , mn } and fn (mn ) = n for
each n ∈ N. We now define
[
M := Mn = {mn : n ∈ N}, f : M −→ N, f (mn ) := fn (mn ) = n. (A.27)
n∈N

Clearly, M ⊆ A, and f is bijective with f −1 : N −→ M , f −1 (n) = mn . 

Theorem A.56 (Schröder-Bernstein). Let A, B be sets. The following statements are


equivalent (even without assuming the axiom of choice):

(i) The sets A and B have the same cardinality (i.e. there exists a bijective map
φ : A −→ B).

(ii) There exist an injective map f : A −→ B and an injective map g : B −→ A.

We will give two proofs of the Schröder-Bernstein theorem. The first proof is rather
elegant, but also quite abstract. The second proof is longer, but less abstract. Even
though it is still nonconstructive in the general situation, in many concrete cases, it does
provide a method for actually constructing a bijective map from two injective maps. The
first proof is based on the following lemma:

Lemma A.57. Let A be a set. Consider P(A) to be endowed with the partial order
given by set inclusion, i.e., for each X, Y ∈ P(A), X ≤ Y if, and only if, X ⊆ Y . If
F : P(A) −→ P(A) is isotone with respect to that order, then F has a fixed point, i.e.
F (X0 ) = X0 for some X0 ∈ P(A).

Proof. Define \
A := {X ∈ P(A) : F (X) ⊆ X}, X0 := X (A.28)
X∈A

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 217

(X0 is well-defined, since F (A) ⊆ A). Suppose X ∈ A, i.e. F (X) ⊆ X and X0 ⊆ X.


Then F (X0 ) ⊆ F (X) ⊆ X due to the isotonicity of F . Thus, F (X0 ) ⊆ X for every
X ∈ A, i.e. F (X0 ) ⊆ X0 . Using the isotonicity of F again shows F (F (X0 )) ⊆ F (X0 ),
implying F (X0 ) ∈ A and X0 ⊆ F (X0 ), i.e. F (X0 ) = X0 as desired. 

First Proof of Th. A.56. (i) trivially implies (ii), as one can simply set f := φ and
g := φ−1 . It remains to show (ii) implies (i). Thus, let f : A −→ B and g : B −→ A
be injective. To apply Lem. A.57, define

F : P(A) −→ P(A), F (X) := A \ g B \ f (X) ,

and note

X ⊆ Y ⊆ A ⇒ f (X) ⊆ f (Y ) ⇒ B \ f (Y ) ⊆ B \ f (X)
 
⇒ g B \ f (Y ) ⊆ g B \ f (X) ⇒ F (X) ⊆ F (Y ).

Thus, by Lem. A.57, F has a fixed point X0 . We claim that a bijection is obtained via
setting (
f (x) for x ∈ X0 ,
φ : A −→ B, φ(x) := −1
g (x) for x ∈ / X0 .

First, φ is well-defined, since x ∈/ X0 = F (X0 ) implies x ∈ g B \ f (X0 ) . To verify
that φ is injective, let x, y ∈ A, x 6= y. If x, y ∈ X0 , then φ(x) 6= φ(y), as f is
injective. If x, y ∈ A \ X0 , then φ(x) 6= φ(y), as g −1 is well-defined. If x ∈ X0 and
y∈/ X0 , then φ(x) ∈ f (X0 ) and φ(y) ∈ B \ f (X0 ), once again, implying φ(x) 6= φ(y). If
remains to prove surjectivity. If b ∈ f (X0 ), then φ(f −1 (b)) = b. If b ∈ B \ f (X0 ), then
g(b) ∈
/ X0 = F (X0 ), i.e. φ(g(b)) = b, showing φ to be surjective. 

Second Proof of Th. A.56. As in the first proof, we only need to show (ii) implies (i).
We first assume that A and B are disjoint. To define φ, we first construct a suitable
partition of A ∪˙ B, where the subsets of the partition are given via sequences defined by
using f and g. The idea is to assign a unique sequence σ(a) to each a ∈ A and a unique
sequence σ(b) to each b ∈ B by alternately applying f and g to advance the sequence
to the right and by alternately applying f −1 and g −1 to advance the sequence to the
left, if possible (for a given a ∈ A, g −1 (a) might not be defined and, for a given b ∈ B,
f −1 (a) might not be defined). Thus, for a ∈ A, σ(a) has the form
 
. . . , f −1 g −1 (a) , g −1 (a), a, f (a), g f (a) , . . . (A.29)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 218

More precisely, for each a ∈ A, we define σ(a) = (σi (a))i∈Ia recursively by

σi (a) := a for i = 0, (A.30a)



σi (a) := f σi−1 (a) for i > 0 odd, (A.30b)

σi (a) := g σi−1 (a) for i > 0 even, (A.30c)
−1

σi (a) := g σi+1 (a) for i < 0 odd and σi+1 (a) ∈ g(B), (A.30d)
ma := i + 1, Ia := {k ∈ Z : ma ≤ k} for i < 0 odd and σi+1 (a) ∈
/ g(B), (A.30e)

σi (a) := f −1 σi+1 (a) for i < 0 even and σi+1 (a) ∈ f (A), (A.30f)
ma := i + 1, Ia := {k ∈ Z : ma ≤ k} for i < 0 even and σi+1 (a) ∈
/ f (A), (A.30g)

where the conditions in (A.30e) and (A.30g) are meant to implicitly require σi+1 (a) to
be defined for i+1. By induction, one shows σi−1 (a) ∈ A for each i > 0 odd, σi−1 (a) ∈ B
for each i > 0 even, σi+1 (a) ∈ A for each ma ≤ i < 0 odd, and σi+1 (a) ∈ B for each
ma ≤ i < 0 even, such that σi (a) is well-defined by (A.30) for each i ∈ Ia (with Ia = Z
if (A.30e) and (A.30g) are never satisfied). Analogously, for each b ∈ B, we define
σ(b) = (σi (b))i∈Ib recursively by

σi (b) := b for i = 0, (A.31a)



σi (b) := g σi−1 (b) for i > 0 odd, (A.31b)

σi (b) := f σi−1 (b) for i > 0 even, (A.31c)

σi (b) := f −1 σi+1 (b) for i < 0 odd and σi+1 (b) ∈ f (A), (A.31d)
mb := i + 1, Ib := {k ∈ Z : mb ≤ k} for i < 0 odd and σi+1 (b) ∈
/ f (A), (A.31e)

σi (b) := g −1 σi+1 (b) for i < 0 even and σi+1 (b) ∈ g(B), (A.31f)
mb := i + 1, Ib := {k ∈ Z : mb ≤ k} for i < 0 even and σi+1 (b) ∈
/ g(B), (A.31g)

where the conditions in (A.31e) and (A.31g) are meant to implicitly require σi+1 (b) to
be defined for i+1. By induction, one shows σi−1 (b) ∈ B for each i > 0 odd, σi−1 (b) ∈ A
for each i > 0 even, σi+1 (b) ∈ B for each mb ≤ i < 0 odd, and σi+1 (b) ∈ A for each
mb ≤ i < 0 even, such that σi (b) is well-defined by (A.31) for each i ∈ Ib (with Ib = Z if
(A.31e) and (A.31g) are never satisfied). The σ(a) and σ(b) now allow us to define the
sets
∀ Sx := {σi (x) : i ∈ Ix } ⊆ A ∪˙ B. (A.32)
˙B
x∈A ∪

Moreover, we call x ∈ A ∪˙ B an A-stopper if, and only if, σ(x) terminates to the left
with some element in A; a B-stopper, if, and only if, σ(x) terminates to the left with

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 219

some element in B; and a non-stopper, if σ(x) does never terminate to the left – thus,
 
x A-stopper ⇔ Ix 6= Z ∧ (x ∈ A ∧ mx even) ∨ (x ∈ B ∧ mx odd) ,
 
x B-stopper ⇔ Ix 6= Z ∧ (x ∈ A ∧ mx odd) ∨ (x ∈ B ∧ mx even) ,
x non-stopper ⇔ Ix = Z. (A.33)

Next, we prove that the Sx form a partition of A ∪˙ B. Since, for each x ∈ A ∪˙ B,


x = σ0 (x) ∈ Sx , it only remains to show
 
∀ Sx = Sy ∨ Sx ∩ Sy = ∅ . (A.34)
˙B
x,y∈A ∪

To prove (A.34), it clearly suffices to show


 
∀ z ∈ Sx ⇒ Sx = Sz . (A.35)
˙B
x,z∈A ∪

To verify (A.35), let z ∈ Sx . Then there exists i ∈ Ix such that z = σ0 (z) = σi (x) and
simple inductions show σk (z) = σk+i (x) for each k ∈ Iz and σk−i (z) = σk (x) for each
k ∈ Ix (in particular, i + Iz = Ix ), proving Sx = Sz .
We are now in a position to define the desired bijection φ : A −→ B:
(
f (a) if a is an A-stopper or a non-stopper,
φ : A −→ B, φ(a) := −1
(A.36)
g (a) if a is a B-stopper.

Indeed, φ is injective: If a1 , a2 ∈ {a ∈ A : a A-stopper or non-stopper} with a1 6= a2 ,


then φ(a1 ) 6= φ(a2 ) due to f being injective; if a1 , a2 ∈ {a ∈ A : a B-stopper} with
a1 6= a2 , then φ(a1 ) 6= φ(a2 ) due to g −1 being injective; and a1 , a2 ∈ A with a2 a B-
stopper and a1 not a B-stopper, Sa1 = Sf (a1 ) and Sa2 = Sg−1 (a2 ) , i.e. φ(a2 ) is also a
B-stopper, whereas φ(a1 ) is not a B-stopper, in particular, φ(a1 ) 6= φ(a2 ). Moreover,
φ is also surjective: If b ∈ B is a B-stopper, then, due to Sb = Sg(b) , so is g(b), and
b = g −1 (g(b)) = φ(g(b)); if b ∈ B is not a B-stopper, then f −1 (b) is defined and in Sb ,
i.e. f −1 (b) is not a B-stopper, either, and b = f (f −1 (b)) = φ(f −1 (b)).
To conclude the proof, we consider the case that A and B are not necessarily disjoint.
Since A × {0} and B × {1} are always disjoint with

f˜ : A × {0} −→ B × {1}, f˜(a, 0) := (f (a), 1), (A.37a)


g̃ : B × {1} −→ A × {0}, g̃(b, 0) := (g(b), 0), (A.37b)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 220

still being injective if f, g are, the first part of the proof yields a bijective function
φ̃ : A × {0} −→ B × {1}. Then, using the clearly bijective functions
α : A −→ A × {0}, α(a) := (a, 0), (A.38a)
β : B −→ B × {1}, β(b) := (b, 1), (A.38b)

φ := β −1 ◦ φ̃ ◦ α : A −→ B is also bijective. 
Remark A.58. In general, the second proof of the Schröder-Bernstein Th. A.56 is still
nonconstructive, since one has, in general, no algorithm to determine if a given element
is an A-stopper, a B-stopper, or a non-stopper. However, as the following Ex. A.59
shows, in particular situations, determining A-stoppers, B-stoppers, and non-stoppers
does not have to be difficult.
Example A.59. Let A := N0 , B := {n ∈ N0 : n even}. We consider A and B as being
made disjoint (for example, by using the trick employed in the last part of the proof
of Th. A.56 above), but, for the sake of readability, we will not reflect this in the used
notation. Define the maps
f : A −→ B, f (n) := 4n, (A.39a)
g : B −→ A, g(n) := n, (A.39b)
both being clearly injective, but not surjective. The goal is to, explicitly, find the
bijective map φ : A −→ B, given by (A.36). As an intermediate step, we determine
which elements of A are non-stoppers, A-stoppers, and B-stoppers, and likewise for the
elements of B. Clearly 0 ∈ A and 0 ∈ B are non-stoppers. We will see that all other
elements are either A-stoppers or B-stoppers. The precise claim is
A1 := {a ∈ A : a is A-stopper} = C1 := {a ∈ A : a = n 4k , n odd, k ∈ N0 }, (A.40a)
A2 := {a ∈ A : a is B-stopper} = C2 := A \ (C1 ∪ {0}), (A.40b)
B1 := {b ∈ B : b is A-stopper} = D1 := B \ (D2 ∪ {0}), (A.40c)
B2 := {b ∈ B : b is B-stopper} = D2 := {b ∈ B : b = n 2k ; n, k odd; n, k ≥ 1}.
(A.40d)

Indeed, if c = n 4k ∈ C1 , then (f −1 ◦ g −1 )k (c) = n is odd, i.e. n ∈ / g(B), showing c is


k
an A-stopper, proving C1 ⊆ A1 . If d = n 2 ∈ D2 , then k − 1 = 2m with m ∈ N0 , i.e.
d = n 2 · 4m and (g −1 ◦ f −1 )m (d) = 2n is not divisible by 4, i.e. 2n ∈
/ f (A), showing d is
a B-stopper, proving D2 ⊆ B2 . Clearly, each a ∈ N either has the form a = n 4k with n
odd and k ∈ N0 (i.e. a ∈ C1 ) or a = 2 · n4k with n odd and k ∈ N0 , i.e.
C2 = {a ∈ A : a = 2 · n(2 · 2)k ; n odd; k ∈ N0 }
= {a ∈ A : a = n 2k ; n, k odd; n, k ≥ 1} = g(D2 ). (A.41)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 221

Since D2 ⊆ B2 , all elements of D2 are B-stoppers, and, thus, so are all elements of C2 ,
proving C2 ⊆ A2 . Since A = C1 ∪˙ C2 ∪{0},
˙ we then also obtain A1 = C1 and A2 = C2 .
Clearly, each even b ∈ N either has the form b = n 2k with odd n, k ≥ 1 (i.e. b ∈ D2 ) or
b = n4k with n odd and k ∈ N, i.e.

D1 = {b ∈ B : b = n 4k , n odd, k ∈ N} = f (C1 ). (A.42)

Since C1 = A1 , all elements of C1 are A-stoppers, and, thus, so are all elements of D1 .
Since B = D1 ∪˙ D2 ∪{0},
˙ we then also obtain B1 = D1 and B2 = D2 .
Now that we have identified explicit formulas for A1 and A2 , we can write the assignment
rule for the bijective φ : A −→ B, given by (A.36), in the explicit form

0 if a = 0,

φ(a) := 4a if a = n 4k with n odd and k ∈ N0 , (A.43)


a if a = 2 · n4k with n odd and k ∈ N0 .

Thus, φ starts out with the assignments


0 1 2 3 4 5 6 7 8
φ: ↓ ↓ ↓ ↓ ↓ ↓ ↓ ↓ ↓ ... (A.44)
0 4 2 12 16 20 6 28 8

Theorem A.60. Let A, B be nonempty sets. Then the following statements are equiv-
alent (where the implication “(ii) ⇒ (i)” makes use of the axiom of choice (AC) of Sec.
A.4).

(i) There exists an injective map f : A −→ B.

(ii) There exists a surjective map g : B −→ A.

Proof. According to Th. 2.13(b), (i) is equivalent to f having a left inverse g : B −→ A


(i.e. g ◦ f = IdA ), which is equivalent to g having a right inverse, which, according to
Th. 2.13(a), is equivalent to (ii) (AC is used in the proof of Th. 2.13(a) to show each
surjective map has a right inverse). 
Corollary A.61. Let A, B be nonempty sets. Using AC, we can expand the two equiv-
alent statements of Th. A.56 to the following list of equivalent statements:

(i) The sets A and B have the same cardinality (i.e. there exists a bijective map
φ : A −→ B).

(ii) There exist an injective map f : A −→ B and an injective map g : B −→ A.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 222

(iii) There exist a surjective map f : A −→ B and a surjective map g : B −→ A.

(iv) There exist an injective map f1 : A −→ B and a surjective map f2 : A −→ B.

(v) There exist an injective map g1 : B −→ A and a surjective map g2 : B −→ A.

Proof. The equivalences are an immediate consequence of combining Th. A.56 with Th.
A.60. 

A.5.2 Finite Sets

It is intuitively clear that finite cardinalities are uniquely determined. Still one has to
provide a rigorous proof. The key is the following theorem:

Theorem A.62. If m, n ∈ N and the map f : {1, . . . , m} −→ {1, . . . , n} is bijective,


then m = n.

Proof. We conduct the proof via induction on m. If m = 1, then the surjectivity of f


implies n = 1. For the induction step, we now consider m > 1. From the bijective map
f , we define the map

n
 for x = m,
g : {1, . . . , m} −→ {1, . . . , n}, g(x) := f (m) for x = f −1 (n), (A.45)


f (x) otherwise.

Then g is bijective, since it is the composition g = h ◦ f of the bijective map f with the
bijective map

h : {f (m), n} −→ {f (m), n}, h(f (m)) := n, h(n) := f (m). (A.46)

Thus, the restriction g↾{1,...,m−1} : {1, . . . , m−1} −→ {1, . . . , n−1} must also be bijective,
such that the induction hypothesis yields m − 1 = n − 1, which, in turn, implies m = n
as desired. 

Corollary A.63. Let m, n ∈ N and let A be a set. If #A = m and #A = n, then


m = n.

Proof. If #A = m, then, according to Def. 3.12(b), there exists a bijective map f :


A −→ {1, . . . , m}. Analogously, if #A = n, then there exists a bijective map g :
A −→ {1, . . . , n}. In consequence, we have the bijective map (g ◦ f −1 ) : {1, . . . , m} −→
{1, . . . , n}, such that Th. A.62 yields m = n. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 223

Theorem A.64. Let A 6= ∅ be a finite set.

(a) If B ⊆ A with A 6= B, then B is finite with #B < #A.



(b) If a ∈ A, then # A \ {a} = #A − 1.

Proof. For #A = n ∈ N, we use induction to prove (a) and (b) simultaneously, i.e. we
show
 

∀ #A = n ⇒ ∀ ∀ #B ∈ {0, . . . , n − 1} ∧ # A \ {a} = n − 1 .
n∈N B∈P(A)\{A} a∈A
| {z }
φ(n)

Base Case (n = 1): In this case, A has precisely one element, i.e. B = A \ {a} = ∅, and
#∅ = 0 = n − 1 proves φ(1).
Induction Step: For the induction hypothesis, we assume φ(n) to be true, i.e. we assume
(a) and (b) hold for each A with #A = n. We have to prove φ(n + 1), i.e., we consider
A with #A = n + 1. From #A = n + 1, we conclude the existence of a bijective map ϕ :
A −→ {1, . . . , n + 1}. We have to construct a bijective map ψ : A \ {a} −→ {1, . . . , n}.
To this end, set k := ϕ(a) and define the auxiliary function

n + 1 for x = k,

f : {1, . . . , n + 1} −→ {1, . . . , n + 1}, f (x) := k for x = n + 1,


x for x ∈
/ {k, n + 1}.

Then f ◦ ϕ : A −→ {1, . . . , n + 1} is bijective by Th. 2.14, and

(f ◦ ϕ)(a) = f (ϕ(a)) = f (k) = n + 1.

Thus, the restriction ψ := (f ◦ ϕ) ↾A\{a} is the desired bijective map ψ : A \ {a} −→


{1, . . . , n}, proving # A \ {a} = n. It remains to consider the strict subset B of A.
Since B is a strict subset of A, there
 exists a ∈ A \ B. Thus, B ⊆ A \ {a} and, as
we have already shown # A \ {a}  = n, the induction hypothesis applies and yields B
is finite with #B ≤ # A \ {a} = n, i.e. #B ∈ {0, . . . , n}, proving φ(n + 1), thereby
completing the induction. 
Theorem A.65. For #A = #B = n ∈ N and f : A −→ B, the following statements
are equivalent:

(i) f is injective.

(ii) f is surjective.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 224

(iii) f is bijective.

Proof. It suffices to prove the equivalence of (i) and (ii).


If f is injective, then f : A −→ f (A) is bijective. Since #A = n, there exists a bijective
map ϕ : A −→ {1, . . . , n}. Then (ϕ ◦ f −1 ) : f (A) −→ {1, . . . , n} is also bijective,
showing #f (A) = n, i.e., according to Th. A.64(a), f (A) can not be a strict subset of
B, i.e. f (A) = B, proving f is surjective.
If f is surjective, then f has a right inverse g : B −→ A: One can obtain this from
Th. 2.13(a), but, here, we can actually construct g without the axiom of choice: We let
ϕ : A −→ {1, . . . , n} be the bijective map from above and, for b ∈ B, we let g(b) be the
unique a ∈ C := f −1 ({b}) such that ϕ(a) = min ϕ(C). Then, clearly, f ◦ g = IdB . But
this also means f is a left inverse for g, such that g must be injective by Th. 2.13(b).
According to what we have already proved above, g injective implies g surjective, i.e.
g must be bijective. From Th. 2.13(c), we then know the left inverse of g is unique,
implying f = g −1 . In particular, f is injective. 

Lemma A.66. For each finite set A (i.e. #A = n ∈ N0 ) and each B ⊆ A, one has
#(A \ B) = #A − #B.

Proof. For B = ∅, the assertion is true since #(A \ B) = #A = #A − 0 = #A − #B.


For B 6= ∅, the proof is conducted over the size of B, i.e. as a finite induction (cf. Cor.
3.6) over the set {1, . . . , n}, showing

∀ #B = m ⇒ #(A \ B) = #A − #B .
m∈{1,...,n} | {z }
φ(m)

Base Case (m = 1): φ(1) is precisely the statement provided by Th. A.64(b).
Induction Step: For the induction hypothesis, we assume φ(m) with 1 ≤ m < n. To
prove φ(m + 1), consider B ⊆ A with #B = m + 1. Fix an element b ∈ B and set
B1 := B \ {b}. Then #B1 = m by Th. A.64(b), A \ B = (A \ B1 ) \ {b}, and we compute

 Th. A.64(b) φ(m)
#(A \ B) = # (A \ B1 ) \ {b} = #(A \ B1 ) − 1 = #A − #B1 − 1
= #A − #B,

proving φ(m + 1) and completing the induction. 

Theorem A.67. If A, B are finite sets, then #(A ∪ B) = #A + #B − #(A ∩ B).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 225

Proof. The assertion is clearly true if A or B is empty. If A and B are nonempty, then
there exist m, n ∈ N such that #A = m and #B = n, i.e. there are bijective maps
f : A −→ {1, . . . , m} and g : B −→ {1, . . . , n}.
We first consider the case A∩B = ∅. We need to construct a bijective map h : A∪B −→
{1, . . . , m + n}. To this end, we define
(
f (x) for x ∈ A,
h : A ∪ B −→ {1, . . . , m + n}, h(x) :=
g(x) + m for x ∈ B.

The bijectivity of f and g clearly implies the bijectivity of h, proving #(A ∪ B) =


m + n = #A + #B.
˙
Finally, we consider the case of arbitrary A, B. Since A ∪ B = A ∪(B \ A) and B \ A =
B \ (A ∩ B), we can compute

#(A ∪ B) = # A ∪(B ˙ \ A) = #A + #(B \ A)
 Lem. A.66
= #A + # B \ (A ∩ B) = #A + #B − #(A ∩ B),
thereby establishing the case. 
Theorem A.68. If (A1 , . . . , An ), n ∈ N, is a finite sequence of finite sets, then
n
Y n
Y

# Ai = # A1 × · · · × An = #Ai . (A.47)
i=1 i=1

Proof. If at least one Ai is empty, then (A.47) is true, since both sides are 0.
The case where all Ai are nonempty is proved by induction over n, i.e. we know ki :=
#Ai ∈ N for each i ∈ {1, . . . , n} and show by induction
n
Y n
Y
∀ # Ai = ki .
n∈N
i=1 i=1
| {z }
φ(n)
Q1 Q1
Base Case (n = 1): i=1 Ai = #A1 = k1 = i=1 ki , i.e. φ(1) holds.
Induction Step: From the induction
Qn hypothesis φ(n),Qn we obtain a bijective map ϕ :
A −→ {1, . . . , N }, where A := i=1 Ai and N := i=1 ki . To prove φ(n + 1), we need
to construct a bijective map h : A × An+1 −→ {1, . . . , N · kn+1 }. Since #An+1 = kn+1 ,
there exists a bijective map f : An+1 −→ {1, . . . , kn+1 }. We define
h : A × An+1 −→ {1, . . . , N · kn+1 },

h(a1 , . . . , an , an+1 ) := f (an+1 ) − 1 · N + ϕ(a1 , . . . , an ).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
A AXIOMATIC SET THEORY 226

Since ϕ and f are bijective, and since every m ∈ {1, . . . , N · kn+1 } has a unique rep-
resentation in the form m = a · N + r with a ∈ {0, . . . , kn+1 − 1} and r ∈ {1, . . . , N }
(exercise), h is also bijective. This proves φ(n + 1) and completes the induction. 
Theorem A.69. For each finite set A (i.e. #A = n ∈ N0 ), one has #P(A) = 2n .

Proof. The proof is conducted by induction by showing



∀ #A = n ⇒ #P(A) = 2n .
n∈N0 | {z }
φ(n)

Base Case (n = 0): For n = 0, we have A = ∅, i.e. P(A) = {∅}. Thus, #P(A) = 1 = 20 ,
proving φ(0).
Induction Step: Assume φ(n) and consider A with #A = n + 1. Then A contains at
least one element a. ForB := A \ {a}, we then
know #B = n from Th. A.64(b).
Moreover, setting M := C ∪ {a} : C ∈ P(B) , we have the disjoint decomposition
P(A) = P(B) ∪˙ M. As the map ϕ : P(B) −→ M, ϕ(C) := C ∪ {a}, is clearly bijective,
P(B) and M have the same cardinality. Thus,

Th. A.67 φ(n)
#P(A) = #P(B) + #M = #P(B) + #P(B) = 2 · 2n = 2n+1 ,
thereby proving φ(n + 1) and completing the induction. 

A.5.3 Power Sets

Theorem A.70. Let A be a set. There can never exist a surjective map from A onto
P(A) (in this sense, the size of P(A) is always strictly bigger than the size of A; in
particular, A and P(A) can never have the same size).

Proof. If A = ∅, then there is nothing to prove. For nonempty A, the idea is to


conduct a proof by contradiction. To this end, assume there does exist a surjective map
f : A −→ P(A) and define
B := {x ∈ A : x ∈
/ f (x)}. (A.48)
Now B is a subset of A, i.e. B ∈ P(A) and the assumption that f is surjective implies
the existence of a ∈ A such that f (a) = B. If a ∈ B, then a ∈ / f (a) = B, i.e. a ∈ B
implies a ∈ B ∧ ¬(a ∈ B), so that the principle of contradiction tells us a ∈ / B must be
true. However, a ∈/ B implies a ∈ f (a) = B, i.e., this time, the principle of contradiction
tells us a ∈ B must be true. In conclusion, we have shown our original assumption that
there exists a surjective map f : A −→ P(A) implies a ∈ B ∧ ¬(a ∈ B), i.e., according
to the principle of contradiction, no surjective map from A into P(A) can exist. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
B ASSOCIATIVITY AND COMMUTATIVITY 227

B General Forms of the Laws of Associativity and


Commutativity

B.1 Associativity
In the literature, the general law of associativity is often stated in the form that
a1 a2 · · · an gives the same result “for every admissible way of inserting parentheses into
a1 a2 · · · an ”, but a completely precise formulation of what that actually means seems to
be rare. As a warm-up, we first prove a special case of the general law:
Proposition B.1. Let A be a set with a composition · : A × A −→ A (which we write
as a multiplication, but, clearly, this is not essential, and we could also write it as an
addition or with some other symbol). If the composition is associative, i.e. if

∀ (ab)c = a(bc), (B.1)


a,b,c∈A

then ! !
n
Y k−1
Y n
Y
∀ ∀ ∀ ai ai = ai , (B.2)
n∈N, a1 ,...,an ∈A k∈{2,...,n}
n≥2 i=k i=1 i=1

where the product symbol is defined according to (3.20a).

Proof. If k = n, then (B.2) is immediate from (3.20a). For 2 ≤ k < n, we prove (B.2)
by induction on n: For the base case, n = 2, there is nothing to prove. For n > 2, one
computes
n
! k−1 ! n−1
! k−1 ! n−1
! k−1 !!
Y Y (3.20a) Y Y (B.1) Y Y
ai ai = an · ai ai = an · ai ai
i=k i=1 i=k i=1 i=k i=1

n−1
Y n
Y
ind.hyp. (3.20a)
= an · ai = ai , (B.3)
i=1 i=1

completing the induction and the proof of the proposition. 

The difficulty in stating the general form of the law of associativity lies in giving a
precise definition of what one means by “an admissible way of inserting parentheses into
a1 a2 · · · an ”. So how does one actually proceed to calculate the value of a1 a2 · · · an , given
that parentheses have been inserted in an admissible way? The answer is that one does
it in n − 1 steps, where, in each step, one combines two juxtaposed elements, consistent
with the inserted parentheses. There can still be some ambiguity: For example, for

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
B ASSOCIATIVITY AND COMMUTATIVITY 228

(a1 a2 )(a3 (a4 a5 )), one has the freedom of first combining a1 , a2 , or of first combining
a4 , a5 . In consequence, our general law of associativity will show that, for each admissible
sequence of n − 1 directives for combining two juxtaposed elements, the final result is
the same (under the hypothesis that (B.1) holds). This still needs some preparatory
work.
In the following,
Qn one might see it as a slight notational inconvenience that we have
defined i=1 ai as an · · · a1 rather than a1 · · · an . For this reason, we will enumerate the
elements to be combined by composition from right to left rather then from left to right.
Definition and Remark B.2. Let A be a (nonempty) set with a composition · :
A × A −→ A, let n ∈ N, n ≥ 2, and let I be a totally ordered index set, #I = n,
I = {i1 , . . . , in } with i1 < · · · < in . Moreover, let F := (ain , . . . , ai1 ) be a family of n
elements of A.

(a) An admissible composition directive (for combining two juxtaposed elements of the
family) is an index ik ∈ I with 1 ≤ k ≤ n − 1. It transforms the family F into
the family G := (ain , . . . , aik+1 aik , . . . , ai1 ). In other words, G = (bj )j∈J , where
J := I \ {ik+1 }, bj = aj for each j ∈ J \ {ik }, and bik = aik+1 aik . We can write this
transformation as two maps
(1)
F 7→ δik (F ) := G = (ain , . . . , aik+1 aik , . . . , ai1 ) = (bj )j∈J , (B.4a)
(2)
I 7→ δik (I) := J = I \ {ik+1 }. (B.4b)
Thus, an application of an admissible composition directive reduces the length of
the family and the number of indices by one.
(b) Recursively, we define (finite) sequences of families, index sets, and indices as fol-
lows:
Fn := F, In := I, (B.5a)
(1) (2)
∀ Fα−1 := δjα (Fα ), Iα−1 := δjα (Iα ), where jα ∈ Iα \ {max Iα }.
α∈{2,...,n}
(B.5b)
The corresponding sequence of indices D := (jn , . . . , j2 ) in I is called an admissible
evaluation directive. Clearly,
∀ #Iα = α, i.e. Fα has length α. (B.6)
α∈{1,...,n}

In particular, I1 = {j2 } = {i1 } (where the second equality follows from (B.4b)),
F1 = (a), and we call
D(F ) := a (B.7)
the result of the admissible evaluation directive D applied to F .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
B ASSOCIATIVITY AND COMMUTATIVITY 229

Theorem B.3 (General Law of Associativity). Let A be a (nonempty) set with a com-
position · : A × A −→ A, let n ∈ N, n ≥ 2, and let I be a totally ordered index set,
#I = n, I = {i1 , . . . , in } with i1 < · · · < in . Moreover, let F := (ain , . . . , ai1 ) be a
family of n elements of A. If the composition is associative, i.e. if (B.1) holds, then,
for each admissible evaluation directive as defined in Def. and Rem. B.2(b), the result
is the same, namely
Yn
D(F ) = ai k . (B.8)
k=1

Proof. We conduct the proof via induction on n. For n = 3, there are only two possible
directives and (B.1) guarantees that they yield the same result. For the induction step,
let n > 3. As in Def. and Rem. B.2(b), we write D = (jn , . . . , j2 ) and obtain some
I2 = {i1 , im }, 1 < m ≤ n, as the corresponding penultimate index set. Depending on
im , we partition (jn , . . . , j3 ) as follows: Set
 
J1 := k ∈ {3, . . . , n} : jk < im , J2 := k ∈ {3, . . . , n} : jk ≥ im . (B.9)

Then, for k ∈ J1 , jk is a composition directive to combine two elements to the right of


aim and, for k ∈ J2 , jk is a composition directive to combine two elements to the left of
aim . Moreover, J1 and J2 might or might not be the empty set: If J1 = ∅, then jk 6= i1
for each k ∈ {3, . . . , n}, implying im = i2 ; if J2 = ∅, then, in each of the n − 2 steps to
obtain I2 , an ik with k < m was removed from I, implying im = in (in particular, as
n 6= 2, J1 and J2 cannot both be empty). If J1 6= ∅, then D1 := (jk )k∈J1 is an admissible
evaluation directive for (aim−1 , . . . , ai1 ) – this follows from
(2)
jk ∈ K ⊆ {i1 , . . . , im−1 } ⇒ δjk (K) ⊆ K ⊆ {i1 , . . . , im−1 }. (B.10)

Since m − 1 < n, the induction hypothesis applies and yields


m−1
Y
D1 (aim−1 , . . . , ai1 ) = ai k . (B.11)
k=1

Analogously, if J2 6= ∅, then D2 := (jk )k∈J2 is an admissible evaluation directive for


(ain , . . . , aim ) – this follows from
(2)
jk ∈ K ⊆ {im , . . . , in } ⇒ δjk (K) ⊆ K ⊆ {im , . . . , in }. (B.12)

Since m > 1, the induction hypothesis applies and yields


n
Y
D2 (ain , . . . , aim ) = aik . (B.13)
k=m

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
B ASSOCIATIVITY AND COMMUTATIVITY 230

Thus, if J1 6= ∅ and J2 6= ∅, then we obtain


n
! m−1
! n
j2 =i1
Y Y Prop. B.1
Y
D(F ) = D2 (ain , . . . , aim ) · D1 (aim−1 , . . . , ai1 ) = ai k ai k = aik
k=m k=1 k=1
(B.14)
as desired. If J1 = ∅, then, as explained above, im = i2 . Thus, in this case,
n
! n
j2 =i1
Y Prop. B.1
Y
D(F ) = D2 (ain , . . . , ai2 ) · ai1 = ai k · ai 1 = aik (B.15)
k=2 k=1

as needed. Finally, if J2 = ∅, then, as explained above, im = in . Thus, in this case,


n
Y
j2 =i1
D(F ) = ain · D1 (ain−1 , . . . , ai1 ) = ai k , (B.16)
k=1

again, as desired, and completing the induction. 

B.2 Commutativity
In the present section, we will generalize the law of commutativity ab = ba to a finite
number of factors, provided the composition is also associative. For this purpose, we
introduce the notion of permutation, also useful in many other mathematical contexts.
Definition and Remark B.4. Let n ∈ N. Each bijective map π : {1, . . . , n} −→
{1, . . . , n} is called a permutation of {1, . . . , n}. The set of permutations of {1, . . . , n}
forms a group with respect to the composition of maps, the so-called symmetric group
Sn : Indeed, the composition of maps is associative by Prop. 2.10(a); the neutral element
is the identity map e : {1, . . . , n} −→ {1, . . . , n}, e(i) = i; and, for each π ∈ Sn , its
inverse map π −1 is also its inverse element in the group Sn . Caveat: Simple examples
show that Sn is not commutative.
Theorem B.5 (General Law of Commutativity). Let A be a set with an associative
composition · : A × A −→ A (which we write as a multiplication, but, clearly, this is
not essential, and we could also write it as an addition or with some other symbol). If
the composition is commutative, i.e. if

∀ ab = ba, (B.17)
a,b∈A

then n n
Y Y
∀ ∀ ∀ ai = aπ(i) . (B.18)
n∈N π∈Sn a1 ,...,an ∈A
i=1 i=1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
B ASSOCIATIVITY AND COMMUTATIVITY 231

Before we can carry out the proof, we need to learn a bit more about permutations.

Definition B.6. Let k, n ∈ N, k ≤ n. A permutation π ∈ Sn is called a k-cycle if, and


only if, there exist k distinct numbers i1 , . . . , ik ∈ {1, . . . , n} such that

ij+1 if i = ij , j ∈ {1, . . . , k − 1},

π(i) = i1 if i = ik , (B.19)


i if i ∈/ {i1 , . . . , ik }.

If π is a cycle as in (B.19), then one writes

π = (i1 i2 . . . ik ). (B.20)

Each 2-cycle is also known as a transposition.

Theorem B.7. Let n ∈ N.

(a) Each permutation can be decomposed into finitely many disjoint cycles: For each
π ∈ Sn , there exists a decomposition of {1, . . . , n} into disjoint sets A1 , . . . , AN ,
N ∈ N, i.e.
N
[
{1, . . . , n} = Ai and Ai ∩ Aj = ∅ for i 6= j, (B.21)
i=1

such that Ai consists of the distinct elements ai1 , . . . , ai,Ni and

π = (aN 1 . . . aN,NN ) · · · (a11 . . . a1,N1 ). (B.22)

The decomposition (B.22) is unique up to the order of the cycles.

(b) If n ≥ 2, then every permutation π ∈ Sn is the composition of finitely many trans-


positions, where each transposition permutes two juxtaposed elements, i.e.

∀ ∃ ∃ π = τN ◦ · · · ◦ τ1 , (B.23)
π∈Sn N ∈N τ1 ,...,τN ∈T


where T := (i i + 1) : i ∈ {1, . . . , n − 1} .

Proof. (a): We prove the statement by induction on n. For n = 1, there is nothing to


prove. Let n > 1 and choose i ∈ {1, . . . , n}. We claim that
 
k l
∃ π (i) = i ∧ ∀ π (i) 6= i . (B.24)
k∈N l∈{1,...,k−1}

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
B ASSOCIATIVITY AND COMMUTATIVITY 232

Indeed, since {1, . . . , n} is finite, there must be a smallest k ∈ N such that π k (i) ∈ A1 :=
{i, π(i), . . . , π k−1 (i)}. Since π is bijective, it must be π k (i) = i and (i π(i) . . . , π k−1 (i))
is a k-cycle. We are already done in case k = n. If k < n, then consider B :=
{1, . . . , n} \ A1 . Then, again using the bijectivity of π, π↾B is a permutation onSB with
1 ≤ #B < n. By induction, there are disjoint sets A2 , . . . , AN such that B = N j=2 Aj ,
Aj consists of the distinct elements aj1 , . . . , aj,Nj and
π↾B = (aN 1 . . . aN,NN ) · · · (a21 . . . a2,N2 ).
Since π = (i π(i) . . . , π k−1 (i)) ◦ π ↾B , this finishes the proof of (B.22). If there were
another, different,
SM decomposition of π into cycles, say, given by disjoint sets B1 , . . . , BM ,
{1, . . . , n} = i=1 Bi , M ∈ N, then there were Ai 6= Bj and k ∈ Ai ∩ Bj . But then k
were in the cycle given by Ai and in the cycle given by Bj , implying Ai = {π l (k) : l ∈
N} = Bj , in contradiction to Ai 6= Bj .
(b): We first show that every π ∈ Sn is a composition of finitely many transpositions
(not necessarily transpositions from the set T ): According to (a), it suffices to show
that every cycle is a composition of finitely many transpositions. Since each 1-cycle is
the identity, it is (i) = Id = (1 2) (1 2) for each i ∈ {1, . . . , n}. If (i1 . . . ik ) is a k-cycle,
k ∈ {2, . . . , n}, then
(i1 . . . ik ) = (i1 i2 ) (i2 i3 ) · · · (ik−1 ik ) : (B.25)
Indeed,

i1
 for i = ik ,
∀ (i1 i2 ) (i2 i3 ) · · · (ik−1 ik )(i) = il+1 for i = il , l ∈ {1, . . . , k − 1}, (B.26)
i∈{1,...,n} 

i for i ∈
/ {i1 , . . . , ik },
proving (B.25). To finish the proof of (b), we observe that every transposition is a
composition of finitely many elements of T : If i, j ∈ {1, . . . , n}, i < j, then
(i j) = (i i + 1) · · · (j − 2 j − 1)(j − 1 j) · · · (i + 1 i + 2)(i i + 1) : (B.27)
Indeed,
∀ (i i + 1) · · · (j − 2 j − 1)(j − 1 j) · · · (i + 1 i + 2)(i i + 1)(k)
k∈{1,...,n}


 j for k = i,

i for k = j,
= (B.28)


 k for i < k < j,

k for k ∈ / {i, i + 1, . . . , j},
proving (B.27). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 233

Proof of Th. B.5. For n = 1, there is nothing to prove. So let n > 1. For l ∈
1, . . . , n − 1, let τl : {1, . . . , n} −→ {1, . . . , n} be the transposition that interchanges
l and l + 1 and leaves all other elements fixed (i.e. τl (l) = l + 1, τl (l + 1) = l, τ (α) = α
for each α ∈ {1, . . . , n} \ {l, l + 1}) and let T := {τ1 , . . . , τn−1 }. Then (B.17) and Th.
B.3 imply that the theorem holds for π = τ for each τ ∈ T . For a general permutation
π ∈ Sn , Th. B.7(b) provides a finite sequence (τ 1 , . . . , τ N ), N ∈ N, of elements of T
such that π = τ N ◦ · · · ◦ τ 1 . Thus, as we already know that the theorem holds for N = 1,
the case N > 1 follows by induction. 

The following example shows that, if the composition is not associative, then, in general,
(B.17) does not imply (B.18):
Example B.8. Let A := {a, b, c} with #A = 3 (i.e. the elements a, b, c are all distinct).
Let the composition · on A be defined according to the composition table

· a b c
a b b b
b b b a
c b a a

(the table entry at the intersection of the row labeled with factor x and the column
labeled with factor y is definition of x · y). Then, clearly, · is commutative. However, ·
is not associative, since, e.g.,

(cb)a = aa = b 6= a = cb = c(ba), (B.29)

and (B.18) does not hold, since, e.g.,

a(bc) = aa = b 6= a = cb = c(ba). (B.30)

C Algebraic Structures

C.1 Groups
Definition C.1. Let G be a nonempty set with a map

◦ : G × G −→ G, (x, y) 7→ x ◦ y (C.1)

(called a composition on G, the examples we have in mind are addition and multiplication
on R). Then (G, ◦) (or just G, if the composition ◦ is understood) is called a group if,
and only if, the following three conditions are satisfied:

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 234

(i) Associativity: x ◦ (y ◦ z) = (x ◦ y) ◦ z holds for all x, y, z ∈ G.

(ii) There exists a neutral element e ∈ G, i.e. an element e ∈ G such that

∀ x ◦ e = x. (C.2)
x∈G

(iii) For each x ∈ G, there exists an inverse element x ∈ G, i.e. an element x ∈ G such
that
x ◦ x = e.

G is called a commutative or abelian group if, and only if, it is a group and satisfies the
additional condition:

(iv) Commutativity: x ◦ y = y ◦ x holds for all x, y ∈ G.

Theorem C.2. The following statements and rules are valid in every group (G, ◦):

(a) If (C.2) holds for e ∈ G, then

∀ e ◦ x = x. (C.3)
x∈G

also holds.

(b) The neutral element is unique: If e, f ∈ G, then


   
∀ x◦e=x ∧ ∀ x◦f =x ⇒ e = f. (C.4)
x∈G x∈G

(c) If x, a ∈ G and x ◦ a = e (where e ∈ G is the neutral element), then a ◦ x = e as


well. Moreover, inverse elements are unique (for each x ∈ G, the unique inverse is
then denoted by x−1 ).

(d) (x−1 )−1 = x holds for each x ∈ G.

(e) y −1 ◦ x−1 = (x ◦ y)−1 holds for each x, y ∈ G.

(f ) x ◦ a = y ◦ a ⇒ x = y holds for each x, y, a ∈ G.

Proof. (a): Let x ∈ G. By Def. C.1(iii), there exists y ∈ G such that x ◦ y = e and, in
turn, z ∈ G such that y ◦ z = e. Thus,

e ◦ z = (x ◦ y) ◦ z = x ◦ (y ◦ z) = x ◦ e = x, (C.5)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 235

implying
x = e ◦ z = (e ◦ e) ◦ z = e ◦ (e ◦ z) = e ◦ x (C.6)
as desired.
(b): If e, f are both neutral elements, then, using (a), f = e ◦ f = e.
(c): Assume x ◦ a = e. Then there is b such that a ◦ b = e. One computes

e = a ◦ b = (a ◦ e) ◦ b = (a ◦ (x ◦ a)) ◦ b = a ◦ ((x ◦ a) ◦ b) = a ◦ (x ◦ (a ◦ b))


= a ◦ (x ◦ e) = a ◦ x, (C.7)

establishing the case. Now let a, b be inverses to x. Then a = a ◦ e = a ◦ x ◦ b = e ◦ b = b.


(d): x−1 ◦ x = e holds according to (c) and shows that x is the inverse to x−1 . Thus,
(x−1 )−1 = x as claimed.
(e) is due to y −1 ◦ x−1 ◦ x ◦ y = y −1 ◦ e ◦ y = e.
(f): If x ◦ a = y ◦ a, then x = x ◦ a ◦ a−1 = y ◦ a ◦ a−1 = y as claimed. 

Definition C.3. Let (G, ◦) and (H, ◦) be groups. A map φ : G −→ H is called a


(group) homomorphism if, and only if,

∀ φ(a ◦ b) = φ(a) ◦ φ(b). (C.8)


a,b∈G

Proposition C.4. Let (G, ◦) and (H, ◦) be groups and let φ : G −→ H be a group
homomorphism. Let e, e′ denote the neutral elements of G and H, respectively. Then
the following holds:

(a) φ(e) = e′ .

(b) φ(a−1 ) = (φ(a))−1 for each a ∈ G.

(c) If φ is bijective, then φ−1 is also a group homomorphism.

Proof. (a): We compute

φ(e) ◦ e′ = φ(e) = φ(e ◦ e) = φ(e) ◦ φ(e).

Applying (φ(e))−1 to both sides of the above equality then proves φ(e) = e′ .
(b): We compute
φ(a−1 ) ◦ φ(a) = φ(a−1 ◦ a) = φ(e) = e′ ,
proving (b).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 236

(c): Applying φ−1 to (C.8) yields

∀ a ◦ b = φ−1 (φ(a) ◦ φ(b)). (C.9)


a,b∈G

Thus, for each x, y ∈ H, we obtain


 (C.9) −1
φ−1 (x ◦ y) = φ−1 φ(φ−1 (x)) ◦ φ(φ−1 (y)) = φ (x) ◦ φ−1 (y),

establishing the case and completing the proof of the proposition. 

Notation C.5. Exponentiation with Integer Exponents: Let G be a nonempty set with
a composition · : G × G −→ G. Assume there exists a (unique) neutral element 1 ∈ G
(satisfying x · 1 = x for each x ∈ G). Define recursively for each x ∈ G and each n ∈ N0 :

x0 := 1, ∀ xn+1 := x · xn . (C.10a)
n∈N0

Moreover, if (G, ·) constitutes a group, then also define for each x ∈ G and each n ∈ N:

x−n := (x−1 )n . (C.10b)

Theorem C.6. Exponentiation Rules: Let G be a nonempty set with a composition


· : G × G −→ G. Assume that the composition satisfies the law of associativity and that
there exists a (unique) neutral element 1 ∈ G (satisfying x · 1 = x for each x ∈ G). Let
x, y ∈ G. Then the following rules hold for each m, n ∈ N0 . If (G, ·) is a group, then
the rules even hold for every m, n ∈ Z.

(a) xm+n = xm · xn .

(b) (xm )n = xm n .

(c) If the composition is commutative (i.e. xy = yx for each x, y ∈ G), then it holds
that xn y n = (xy)n .

Proof. (a): First, we fix n ∈ N0 and prove the statement for each m ∈ N0 by induction:
The base case (m = 0) is xn = xn , which is true. For the induction step, we compute
(C.10a) ind. hyp. (C.10a)
xm+1+n = x · xm+n = x · xm · xn = xm+1 xn ,

completing the induction step. Now assume G to be a group. Consider m ≥ 0 and


n < 0. If m + n ≥ 0, then, using what we have already shown,
(C.10b)
xm xn = xm (x−1 )−n = xm+n x−n (x−1 )−n = xm+n .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 237

Similarly, if m + n < 0, then


(C.10b) (C.10b)
xm xn = xm (x−1 )−n = xm (x−1 )m (x−1 )−n−m = xm+n .

The case m < 0, n ≥ 0 is treated completely analogously. It just remains to consider


m < 0 and n < 0. In this case,
(C.10b) (C.10b)
xm+n = x−(−m−n) = (x−1 )−m−n = (x−1 )−m · (x−1 )−n = xm · xn .

(b): First, we prove the statement for each n ∈ N0 by induction (for m < 0, we assume
G to be a group): The base case (n = 0) is (xm )0 = 1 = x0 , which is true. For the
induction step, we compute
(C.10a) ind. hyp. (a)
(xm )n+1 = xm · (xm )n = xm · xm n = xm n+m = xm (n+1) ,

completing the induction step. Now, let G be a group and n < 0. We already know
(xm )−1 = x−m . Thus, using what we have already shown,
(C.10b) −n
(xm )n = (xm )−1 = (x−m )−n = x(−m)(−n) = xm n .

(c): For n ∈ N0 , the statement is proved by induction: The base case (n = 0) is


x0 y 0 = 1 = (xy)0 , which is true. For the induction step, we compute
(C.10a) ind. hyp. (C.10a)
xn+1 y n+1 = x · xn · y · y n = xy · (xy)n = (xy)n+1 ,

completing the induction step. If G is a group and n < 0, then, using what we have
already shown,
(C.10b) Th. C.2(e) −n (C.10b)
xn y n = (x−1 )−n (y −1 )−n = (x−1 y −1 )−n = (xy)−1 = (xy)n ,

which completes the proof. 

C.2 Rings
Definition C.7. Let R be a nonempty set with two maps
+ : R × R −→ R, (x, y) 7→ x + y,
(C.11)
· : R × R −→ R, (x, y) 7→ x · y
(+ is called addition and · is called multiplication; often one writes xy instead of x · y).
Then (R, +, ·) (or just R, if + and · are understood) is called a ring if, and only if, the
following three conditions are satisfied:

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 238

(i) R is a commutative group with respect to +.

(ii) Multiplication is associative.

(iii) Distributivity:

∀ x · (y + z) = x · y + x · z, (C.12a)
x,y,z∈R

∀ (y + z) · x = y · x + z · x. (C.12b)
x,y,z∈R

A ring R is called commutative if, and only if, its multiplication is commutative. More-
over, a ring called a ring with unity if, and only if, R contains a neutral element with
respect to multiplication (i.e. there is 1 ∈ R such that 1 · x = x · 1 = x for each
x ∈ R) – some authors always require a ring to have a neutral element with respect to
multiplication.

Theorem C.8. The following statements and rules are valid in every ring (R, +, ·) (let
0 denote the additive neutral element and let x, y, z ∈ R):

(a) x · 0 = 0 = 0 · x.

(b) x(−y) = −(xy) = (−x)y.

(c) (−x)(−y) = xy.

(d) x(y − z) = xy − xz.

Proof. (a): One computes


(C.12a)
x·0+x·0 = x · (0 + 0) = x · 0 = 0 + x · 0,

i.e. x · 0 = 0 follows since (R, +) is a group. The second equality follows analogously
using (C.12b).
(b): xy +x(−y) = x(y −y) = x·0 = 0, where we used (C.12a) and (a). This shows x(−y)
is the additive inverse to xy. The second equality follows analogously using (C.12b).
(c): xy = −(−(xy)) = −(x(−y)) = (−x)(−y), where (b) was used twice.
(d): x(y − z) = x(y + (−z)) = xy + x(−z) = xy − xz. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
C ALGEBRAIC STRUCTURES 239

C.3 Fields
Definition C.9. Let (F, +, ·) be a ring with unity. Then (F, +, ·) (or just F , if + and
· are understood) is called a field if, and only if, F \ {0} is a commutative group with
respect to ·.

Theorem C.10. The following statements and rules are valid in every field (F, +, ·):

(a) Inverse elements are unique. For each x ∈ F , the unique inverse with respect to
addition is denoted by −x. Also define y − x := y + (−x). For each x ∈ F \ {0}, the
unique inverse with respect to multiplication is denoted by x−1 . For x 6= 0, define
the fractions xy := y/x := yx−1 with numerator y and denominator x.

(b) −(−x) = x and (x−1 )−1 = x for x 6= 0.

(c) (−x) + (−y) = −(x + y) and x−1 y −1 = (xy)−1 for x, y 6= 0.

(d) x + a = y + a ⇒ x = y and, for a 6= 0, xa = ya ⇒ x = y.

(e) x · 0 = 0.

(f ) x(−y) = −(xy).

(g) (−x)(−y) = xy.

(h) x(y − z) = xy − xz.

(i) xy = 0 ⇒ x = 0 ∨ y = 0.

(j) Rules for Fractions:

a b ad + bc a b ab a/c ad
+ = , · = , = ,
c d cd c d cd b/d bc

where all denominators are assumed 6= 0.



Proof. (a) follows by applying Th. C.2(c) to the groups (F, +) and F \ {0}, · .

(b) follows by applying Th. C.2(d) to the groups (F, +) and F \ {0}, · .

(c) follows by applying Th. C.2(e) to the groups (F, +) and F \ {0}, · , plus then using
commutativity of the groups.

(d) follows by applying Th. C.2(f) to the groups (F, +) and F \ {0}, · (in the latter
situation, the case x = y = 0 is also clear).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 240

(e) follows by applying Th. C.8(a) to the ring with unity (F, +, ·).
(f) follows by applying Th. C.8(b) to the ring with unity (F, +, ·).
(g) follows by applying Th. C.8(c) to the ring with unity (F, +, ·).
(h) follows by applying Th. C.8(d) to the ring with unity (F, +, ·).
(i): If xy = 0 and x 6= 0, then y = 1 · y = x−1 xy = x−1 · 0 = 0.
(j): One computes
a b ad + bc
+ = ac−1 + bd−1 = add−1 c−1 + bcc−1 d−1 = (ad + bc)(cd)−1 =
c d cd
and
a b ab
· = ac−1 bd−1 = ab(cd)−1 =
c d cd
and
a/c ad
= ac−1 (bd−1 )−1 = ac−1 b−1 d = ad(bc)−1 = ,
b/d bc
completing the proof. 

D Construction of the Real Numbers


In Th. 4.4, we have defined the set of real numbers R as a complete totally ordered field
and we claimed that such a complete totally ordered field does actually exist. In the
following, we will describe how R can be constructed. We will follow [EHH+ 95, Chs.
1,2], which contains several different approaches for the construction of R.

D.1 Natural Numbers


In the first step, one starts with the natural numbers N. The set of natural numbers
N was defined in Def. A.41 and it was shown in Th. A.44 that N satisfies the Peano
axioms P1 – P3 of Sec. 3.1. We denote natural numbers using the usual symbols 0 := ∅,
1 := S(0) = {0}, 2 := S(1) = {0, 1}, 3 := S(2) = {0, 1, 2}, . . . , n + 1 := S(n) =
n ∪ {n} = {0, 1, . . . , n} (which is consistent with previous definitions in (A.5) and Not.
A.45).
Theorem 3.7 allows to define addition and multiplication on N0 via recursion:
Definition D.1. (a) For each m, n ∈ N0 , m + n is defined recursively by
m + 0 := m, m + 1 := S(m), ∀ m + S(n) := S(m + n). (D.1)
n∈N

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 241

This fits into the framework of Th. 3.7, using A := N0 , x1 := S(m), and, for each
n ∈ N, fn : An −→ A, fn (x1 , . . . , xn ) := S(xn ) (due to the different initializations,
one obtains a different recursion for each m ∈ N0 ).

(b) For each m, n ∈ N0 , mn := m · n is defined recursively by

m · 0 := 0, m · 1 := m, ∀ m · (n + 1) := m · n + m. (D.2)
n∈N

This fits into the framework of Th. 3.7, using A := N0 , x1 := m, and, for each
m, n ∈ N, fm,n : An −→ A, fm,n (x1 , . . . , xn ) := xn + m.

Theorem D.2. The set N0 of the natural numbers (including 0) with the maps of
addition and multiplication

+ : N0 × N0 −→ N0 , (x, y) 7→ x + y,
· : N0 × N0 −→ N0 , (x, y) 7→ x · y,

as defined in Def. D.1(a) and Def. D.1(b), respectively, satisfies Def. C.1(i),(ii),(iv) for
both addition and multiplication, i.e. associativity, commutativity, and the existence of a
neutral element. This can be summarized as the statement that N0 forms a commutative
semigroup with respect to both addition and multiplication (however, no group, as the
existence of inverse elements is lacking). Moreover, distributivity, i.e. Def. C.7(iii) is
also satisfied.

Proof. Associativity of Addition: We have to show

∀ (k + m) + n = k + (m + n). (D.3a)
k,m,n∈N0

The proof of (D.3a) is carried out by induction on n. The base case (n = 0) follows
from the first definition in (D.1): (k + m) + 0 = k + m = k + (m + 0) for every k, m ∈ N0 .
For the induction step, one computes, for every k, m, n ∈ N0 ,
(D.1) (D.1) ind. hyp.
(k + m) + (n + 1) = (k + m) + S(n) = S((k + m) + n) = S(k + (m + n))
(D.1) (D.1)
= k + S(m + n) = k + (m + S(n))
(D.1)
= k + (m + (n + 1)), (D.3b)

completing the induction.


Neutral Element of Addition: That 0 is the neutral element of addition is immediate
from (D.1).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 242

Commutativity of Addition: We have to show

∀ m + n = n + m. (D.4a)
m,n∈N0

The proof of (D.4a) is also carried out by induction on n. More precisely, we prove
n = 0 separately, and then carry out the induction for n ∈ N. The case n = 0 is proved
by induction on m: The base case (m = 0) is the true statement 0 + 0 = 0 = 0 + 0. For
the induction step, one computes (m + 1) + 0 = m + 1 = S(m) = S(m + 0) = S(0 + m) =
0+S(m) = 0+(m+1). The base case for the induction on n, i.e. n = 1 is also proved by
induction on m: The base case (m = 0) is the true statement 0 + 1 = S(0) = 1 = 1 + 0.
For the induction step, one computes, for every m ∈ N0 ,
(D.1) ind. hyp. (D.1)
(m + 1) + 1 = S(m + 1) = S(1 + m) = (1 + m) + 1
(D.3a)
= 1 + (m + 1). (D.4b)

Now, for the induction step of the induction on n, one computes, for every (m, n) ∈
N0 × N,
(D.1) (D.1) ind. hyp. (D.1)
m + (n + 1) = m + S(n) = S(m + n) = S(n + m) = n + S(m)
(D.1) base case (D.3a)
= n + (m + 1) = n + (1 + m) = (n + 1) + m, (D.4c)

completing the induction.


Neutral Element of Multiplication: That 1 is the neutral element of addition is imme-
diate from (D.2).
Commutativity of Multiplication: We have to show

∀ m · n = n · m. (D.5a)
m,n∈N0

We start with some preparatory steps: We first show

∀ m = m · 1 = 1 · m. (D.5b)
m∈N0

We have m · 1 = m for each m ∈ N0 directly from (D.2). We prove 1 · m = m for each


m ∈ N0 via induction on m: 1 · 0 = 0 and 1 · 1 = 1 are immediate from (D.2). For the
induction step, one computes, for every m ∈ N,
(D.2) ind. hyp.
1 · (m + 1) = 1 · m + 1 = m + 1.

Next, we show
∀ n · m + m = (n + 1) · m (D.5c)
m,n∈N0

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 243

via induction on m. For the base case (m = 0), we note (n + 1) · 0 = 0 by (D.2), and
n · 0 + 0 = 0 by (D.1) and (D.2). For the induction step, we compute
(D.2) (D.4a)
n · (m + 1) + m + 1 = n·m+n+m+1 = n·m+m+n+1
ind. hyp. (D.2)
= (n + 1) · m + n + 1 = (n + 1) · (m + 1).

We are now in a position to carry out the proof of (D.5a) by induction on n. More
precisely, we prove n = 0 separately, and then carry out the induction for n ∈ N. Let
n = 0. We have m · 0 = 0 for each m ∈ N0 directly from (D.2). We prove 0 · m = 0 for
each m ∈ N0 via induction on m: 0 · 0 = 0 and 0 · 1 = 0 are immediate from (D.2). For
the induction step, one computes, for every m ∈ N,
(D.2) ind. hyp., (D.1)
0 · (m + 1) = 0 · m + 0 = 0.

The base case for the induction on n ∈ N is provided by (D.5b). For the induction step,
one computes, for every (m, n) ∈ N0 × N,
(D.2) ind. hyp. (D.5c)
m · (n + 1) = m · n + m = n · m + m = (n + 1) · m,

completing the proof of (D.5a).


Distributivity: As we have commutativity of multiplication, we only need to show

∀ (k + m) · n = k · n + m · n. (D.6a)
k,m,n∈N0

The proof of (D.6a) is carried out by induction on n. The base case (n = 0) follows
from (D.2): (k + m) · 0 = 0 = k · 0 + m · 0. For the induction step, one computes, for
every k, m, n ∈ N0 ,
(D.2)
(k + m) · (n + 1) = (k + m) · n + k + m
ind. hyp.
= k·n+m·n+k+m
(D.4a)
= k·n+k+m·n+m
(D.2)
= k · (n + 1) + m · (n + 1), (D.6b)

completing the induction.


Associativity of Multiplication: We have to show

∀ (k · m) · n = k · (m · n). (D.7a)
k,m,n∈N0

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 244

The proof of (D.7a) is carried out by induction on n. The base case (n = 0) follows
from (D.2): (k · m) · 0 = 0 = k · (m · 0) for every k, m ∈ N0 . For the induction step, one
computes, for every k, m, n ∈ N0 ,
(D.2) ind. hyp.
(k · m) · (n + 1) = (k · m) · n + k · m = k · (m · n) + k · m
(D.6a) (D.2)
= k · (m · n + m) = k · (m · (n + 1)), (D.7b)

completing the induction. 

Lemma D.3. We have


∀ ∀ m + n 6= m. (D.8)
m∈N0 n∈N

Proof. We fix n ∈ N and conduct an induction over m ∈ N0 . The base case (m = 0) is


clear, since 0 + n = n 6= 0. The induction step is also clear, since m + n 6= m implies
m + 1 + n = m + n + 1 = S(m + n) 6= S(m) = m + 1, where we used that S is injective
by Peano axiom P2. 

Next, one defines an order ≤ on N0 :

Definition D.4. For each n, m ∈ N0 , let

n≤m :⇔ ∃ n + k = m. (D.9)
k∈N0

Theorem D.5. The relation defined in (D.9) constitutes a well-order on N0 (in partic-
ular, a total order) that is compatible with addition and multiplication, i.e. it satisfies
(4.3).

Proof. ≤ is Reflexive: For each n ∈ N0 , we have n + 0 = n, showing n ≤ n.


≤ is Antisymmetric: If m ≤ n and n ≤ m, then m + k = n and n + l = m. Thus,
n = m + k = n + l + k. Thus, l + k = 0 by Lem. D.3, implying l = k = 0 (since
0∈/ S(N0 )) and m = n.
≤ is Transitive: If n ≤ m and m ≤ l, then there are kn , km ∈ N0 such that n + kn = m
and m + km = l. Then n + kn + km = m + km = l, showing m ≤ l.
≤ is Total Order: We have to show that, if m, n ∈ N0 , then m ≤ n or n ≤ m. To this
end, fix n ∈ N0 and conduct an induction over m. If m = 0, then m + n = n, showing
m ≤ n. For the induction step, let m ∈ N0 . If m ≤ n, then m + k = n. If k = 0, then
n ≤ n + 1 = m + 1. If k 6= 0, then k = S(l), i.e. n = m + l + 1 = m + 1 + l, showing
m + 1 ≤ n. If n ≤ m, then n ≤ m + 1 is immediate, completing the induction.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 245

≤ is Well-Order: We first show that, for each n ∈ N0 , the set An := {m ∈ N0 : m ≤ n}


is finite: Indeed, this follows by an induction over n if we can show An+1 = An ∪ {n + 1}.
Indeed, if m ≤ n, then m ≤ n ≤ n + 1, i.e. m ≤ n + 1 by transitivity, showing
An ∪ {n + 1} ≤ An+1 . If m ≤ n + 1 and m 6≤ n, then n < m, i.e. n + k = m with k 6= 0,
i.e. n + 1 + l = m, showing n + 1 ≤ m, i.e. m = n + 1 and An+1 ⊆ An ∪ {n + 1}. One now
finishes the proof that ≤ is a well-order as in the proof of Th. 3.13(b): Let ∅ 6= A ⊆ N0 .
We have to show A has a min. If A is finite, then A has a min by Th. 3.13(a). If A is
infinite, let n be an element from A. Then the finite set B := {k ∈ A : k ≤ n} = An ∩ B
must have a min m by Th. 3.13(a). Since m ≤ x for each x ∈ B and m ≤ n < x for
each x ∈ A \ B, we have m = min A.
Compatibility with Addition: We have to show

∀ k ≤m ⇒ k+n≤m+n . (D.10)
k,m,n∈N0

To this end, note that k ≤ m means that there is l ∈ N0 such that k + l = m. But then
k + n + l = k + l + n = m + n, showing k + n ≤ m + n.
Compatibility with Multiplication: As 0 ≤ n holds for every n ∈ N0 , there is nothing to
prove. 
Lemma D.6. Let a, n ∈ N. Then the following holds:

(a) a + n ∈ N.
(b) a · n ∈ N.

Proof. (a): Since a + n = 0 + (a + n), a + n 6= 0 is due to Lem. D.3.


(b): We conduct the proof via induction on n: For n = 1, a · 1 = a ∈ N by hypothesis.
For n ≥ 1, a · (n + 1) = a · n + a · 1 = a · n + a. Since a · n ∈ N by induction hypothesis,
we have a · n + a ∈ N by (a). 
Theorem D.7. (a) Monotonicity: Addition and multiplication on N0 are strictly in-
creasing in the sense that, for each a, b ∈ N0 and each n ∈ N,

a<b ⇒ a+n<b+n (D.11a)


a<b ⇒ a · n < b · n. (D.11b)

(b) Cancellation Laws: Given a, b ∈ N0 and n ∈ N, addition and multiplication on N0


satisfy

a+n=b+n ⇒ a=b (D.12a)


a·n=b·n ⇒ a = b. (D.12b)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 246

Proof. Let a, b ∈ N0 with a < b. Then there exists k ∈ N with a + k = b. Let n ∈ N.


(a): We have b + n = a + k + n. Since k + n > 0 by Lem. D.6(a), this shows a + n < b + n.
We have b · n = (a + k) · n = a · n + k · n. Since k · n > 0 by Lem. D.6(b), this shows
a · n < b · n.
(b): Arguing via contraposition, both laws are immediate from (a). 

D.2 Interlude: Orders on Groups


In the succeeding sections, we will construct the set of integers Z, the set of rational
numbers Q, and the set of real numbers R. In each case, we will use the same method
to define a total order on the constructed set, making use of the algebraic structure of
its additive group. It is therefore economical as well as mathematically interesting, to
study this construction once in its abstract form, which is the purpose of the present
section.
Recall the definition of a group from Def. C.1.
Theorem D.8. Assume (G, +) to be a group. Moreover, assume we have a disjoint
decomposition
˙
G = P ∪{0} ˙
∪(−P ), −P := {x ∈ G : −x ∈ P }, (D.13)
where −x denotes the inverse of x with respect to +. If P is closed under + (i.e. x, y ∈ P
implies x + y ∈ P ), then

y≤x :⇔ x − y ∈ P ∪ {0} (D.14)

defines a total order on G that is compatible with addition, i.e. it satisfies (4.3a). More-
over, if a multiplication is also defined on G and P ∪ {0} is closed under this multipli-
cation, then ≤ is also compatible with multiplication, i.e. it satisfies (4.3b). Of course,
one refers to the elements of P as positive and to the elements of −P as negative.

Proof. For each x ∈ G, one has x−x = 0 ∈ P ∪{0}, i.e. x ≤ x and the relation is reflexive.
If x, y ∈ G, x ≤ y and y ≤ x, then x − y ∈ P ∪ {0} and −(x − y) = y − x ∈ P ∪ {0},
and the disjointness of the union in (D.13) implies x − y = 0, i.e. x = y, showing the
relation is antisymmetric. If x, y, z ∈ G with x ≤ y and y ≤ z, then y − x ∈ P ∪ {0},
z −y ∈ P ∪{0}, and z −x = z −y +y −x ∈ P ∪{0} since P is closed under +, showing the
relation is transitive. So we have shown ≤ constitutes a partial order on G. It remains to
show the order is total. However, given the decomposition in (D.13), for each x, y ∈ G,
precisely one of the statements x − y ∈ P (i.e. y < x), x − y = 0 (i.e. x = y), x − y ∈ −P
(i.e. x < y) must be true, proving that the order is total. To see ≤ satisfies (4.3a), let
x, y, z ∈ G. If x ≤ y, then y − x ∈ P ∪ {0}, i.e. y + z − (x + z) = y + z − z − x ∈ P ∪ {0},

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 247

showing x+z ≤ y +z. The proof is completed by noting (4.3b) is precisely the statement
that P ∪ {0} is closed under multiplication. 

D.3 Integers
As compared to our goal, the set of real numbers R, the set N0 still has three defi-
ciencies, namely the lack of inverse elements for addition, the lack of inverse elements
for multiplication, and that the order ≤ lacks completeness. The construction of the
integers will remedy (only) the first of the three deficiencies by providing the inverse
elements of addition.
Definition and Remark D.9. The relation ∼ on N0 × N0 defined by
(a, b) ∼ (c, d) :⇔ a + d = b + c, (D.15)
constitutes an equivalence relation on N0 × N0 (cf. Def. 2.23): Indeed, if a, b ∈ N0 , then
a + b = b + a shows (a, b) ∼ (a, b), proving ∼ to be reflexive. If a, b, c, d ∈ N0 , then
(a, b) ∼ (c, d) ⇒ a+d=b+c ⇒ c+b=d+a ⇒ (c, d) ∼ (a, b),
proving ∼ to be symmetric. If a, b, c, d, e, f ∈ N0 , then
(a, b) ∼ (c, d) ∧ (c, d) ∼ (e, f ) ⇒ a+d=b+c ∧ c+f =d+e
(D.12a)
⇒ a+d+c+f =b+c+d+e ⇒ a+f =b+e ⇒ (a, b) ∼ (e, f ),
proving ∼ to be transitive and an equivalence relation.
Definition D.10. (a) Define the set of integers Z as the set of equivalence classes of
the equivalence relation ∼ defined in (D.15), i.e.

Z := (N0 × N0 )/ ∼ = [(a, b)] : (a, b) ∈ N0 × N0 (D.16)
is the quotient set of N0 × N0 with respect to ∼ (cf. Ex. 2.24(c)). To simplify
notation, in the following, we will write
[a, b] := [(a, b)] (D.17)
for the equivalence class of (a, b) with respect to ∼.
(b) Addition on Z is defined by

+ : Z × Z −→ Z, [a, b], [c, d] →7 [a, b] + [c, d] := [a + c, b + d]. (D.18)
Subtraction on Z is defined by

− : Z × Z −→ Z, [a, b], [c, d] →7 [a, b] − [c, d] := [a, b] + [d, c]. (D.19)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 248

For the definitions in Def. D.10(b) to make sense, one needs to check that they do not
depend on the chosen representatives of the equivalence classes. Moreover, one needs to
convince oneself that these definitions yield the desired familiar operations of addition
and subtraction. Let us start by verifying the independence of the representatives is the
following Lem. D.11.

Lemma D.11. The definitions in Def. D.10(b) do not depend on the chosen represen-
tatives, i.e.
 
∀ ˜ ⇒ [a + c, b + d] = [ã + c̃, b̃ + d]
[a, b] = [ã, b̃] ∧ [c, d] = [c̃, d] ˜ (D.20)
˜ 0
a,b,c,d,ã,b̃,c̃,d∈N

and
 
∀ ˜
[a, b] = [ã, b̃]∧[c, d] = [c̃, d] ⇒ ˜ . (D.21)
[a, b]−[c, d] = [ã, b̃]−[c̃, d]
˜ 0
a,b,c,d,ã,b̃,c̃,d∈N

˜ means c + d˜ = d + c̃,
Proof. (D.20): [a, b] = [ã, b̃] means a + b̃ = b + ã, [c, d] = [c̃, d]
implying a + c + b̃ + d˜ = b + ã + d + c̃, i.e. [a + c, b + d] = [ã + c̃, b̃ + d].
˜
(D.21) is just (D.19) combined with (D.20). 

Theorem D.12. The set of integers Z forms a commutative group with respect to ad-
dition as defined in Def. D.10(b), where [0, 0] is the neutral element, [b, a] is the inverse
element of [a, b] for each a, b ∈ N0 , and, denoting the inverse element of [a, b] by −[a, b]
in the usual way, [a, b] − [c, d] = [a, b] + (−[c, d]) for each a, b, c, d ∈ N0 .

Proof. To verify commutativity and associativity of addition on Z, let a, b, c, d, e, f ∈ N0 .


Then
[a, b] + [c, d] = [a + c, b + d] = [c + a, b + d] = [c, d] + [a, b],
proving commutativity, and

[a, b] + [c, d] + [e, f ] = [a, b] + [c + e, d + f ] = [a + (c + e), b + (d + f )]
= [(a + c) + e, (b + d) + f ] = [a + c, b + d] + [e, f ]

= [a, b] + [c, d] + [e, f ],

proving associativity. For every a, b ∈ N0 , one obtains [a, b]+[0, 0] = [a+0, b+0] = [a, b],
proving neutrality of [0, 0], whereas [a, b] + [b, a] = [a + b, b + a] = [a + b, a + b] = [0, 0]
(since (a + b, a + b) ∼ (0, 0)) shows [b, a] = −[a, b]. Now [a, b] − [c, d] = [a, b] + (−[c, d])
is immediate from (D.19). 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 249

Remark D.13. The map

ι : N0 −→ Z, ι(n) := [n, 0], (D.22)

is a monomorphism, i.e. it is injective (since ι(m) = [m, 0] = ι(n) = [n, 0] implies


m + 0 = 0 + n, i.e. m = n) and satisfies

∀ ι(m + n) = [m + n, 0] = [m, 0] + [n, 0] = ι(m) + ι(n). (D.23)


m,n∈N0

It is customary to identify N0 with ι(N0 ), as it usually does not cause any confusion.
One then just writes n instead of [n, 0] and −n instead of [0, n] = −[n, 0].
Lemma D.14. We have the disjoint decomposition
˙
Z = N ∪{0} ∪˙ Z− , Z− := −N = {n ∈ Z : −n ∈ N}. (D.24)

Proof. Note that, due to (D.15), an equivalence class remains the same if a natural
number is added or subtracted in both components: [a, b] = [a + m, b + m]. Thus, for
each x = [a, b] ∈ Z, if a > b, then x = [a − b, 0] ∈ N; if a = b, then x = [0, 0] = 0; if
a < b, then x = [0, b − a] = −[b − a, 0] ∈ Z− . It just remains to verify that the union in
(D.24) is disjoint. However, if [n, 0] = [0, m] with m, n ∈ N0 , then n + m = 0, proving
n = m = 0, completing the proof. 
Remark D.15. In the above construction, we obtained the commutative group (Z, +)
from the commutative semigroup (N0 , +). It is worth pointing out that the same con-
struction always works when, instead of with N0 , one starts with any commutative
semigroup (H, +) that satisfies the cancellation law a + c = b + c ⇒ a = b, to obtain a
commutative group (G, +) and a monomorphism ι : H −→ G.

To obtain the expected laws of arithmetic, multiplication on Z needs to be defined such


that (a − b) · (c − d) = (ac + bd) − (ad + bc), which leads to the following definition.
Definition D.16. Multiplication on Z is defined by

· : Z × Z −→ Z, [a, b], [c, d] 7→ [a, b] · [c, d] := [ac + bd, ad + bc]. (D.25)

Lemma D.17. The definition in Def. D.16 does not depend on the chosen representa-
tives, i.e.
 
∀ ˜ ˜ ˜
[a, b] = [ã, b̃] ∧ [c, d] = [c̃, d] ⇒ [ac + bd, ad + bc] = [ãc̃ + b̃d, ãd + b̃c̃] .
˜ 0
a,b,c,d,ã,b̃,c̃,d∈N
(D.26)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 250

Proof. As mentioned before, due to (D.15), an equivalence class remains the same if a
natural number is added or subtracted in both components. Thus, one computes
(D.15)
[ac + bd, ad + bc] = [ac + bd + b̃c, ad + bc + b̃c] = [(a + b̃)c + bd, ad + bc + b̃c]
(D.15)
= [(ã + b)c + bd, ad + bc + b̃c] = [ãd˜ + ãc + bd, ãd˜ + ad + b̃c]
= [ã(d˜ + c) + bd, ãd˜ + ad + b̃c] = [ã(d + c̃) + bd, ãd˜ + ad + b̃c]
= [ãc̃ + (ã + b)d, ãd˜ + ad + b̃c] = [ãc̃ + (a + b̃)d, ãd˜ + ad + b̃c]
(D.15)
= [ãc̃ + b̃d + b̃c̃, ãd˜ + b̃c + b̃c̃] = [ãc̃ + b̃(d + c̃), ãd˜ + b̃c + b̃c̃]
= [ãc̃ + b̃(d˜ + c), ãd˜ + b̃c + b̃c̃] = [ãc̃ + b̃d,
˜ ãd˜ + b̃c̃], (D.27)
completing the proof. 
Theorem D.18. The set of integers Z is associative and commutative with respect to
the multiplication defined in Def. D.16. Moreover, distributivity, i.e. Def. C.7(iii) is
satisfied, [1, 0] is the neutral element of multiplication, and there are no zero divisors,
i.e.
 
∀ [a, b] · [c, d] = [ac + bd, ad + bc] = [0, 0] ⇒ [a, b] = [0, 0] ∨ [c, d] = [0, 0] .
a,b,c,d∈N0
(D.28)
Algebraically, the theorem can be summarized by saying that (Z, +, ·) constitutes a prin-
cipal ideal domain.

Proof. Let a, b, c, d, e, f ∈ N0 . Then, using commutativity of addition and multiplication


on N0 ,
[a, b] · [c, d] = [ac + bd, ad + bc] = [ca + db, cb + da] = [c, d] · [a, b],
proving commutativity of multiplication on Z; using distributivity and commutativity
of addition on N0 ,

[a, b] · [c, d] · [e, f ] = [a, b] · [ce + df, cf + de]
= [ace + adf + bcf + bde, acf + ade + bce + bdf ]
= [ace + bde + adf + bcf, acf + bdf + ade + bce]

= [ac + bd, ad + bc] · [e, f ] = [a, b] · [c, d] · [e, f ],
proving associativity of multiplication on Z; again using distributivity and commutativ-
ity of addition on N0 ,

[a, b] · [c, d] + [e, f ] = [a, b] · [c + e, d + f ] = [ac + ae + bd + bf, ad + af + bc + be]
= [ac + bd + ae + bf, ad + bc + af + be]
= [ac + bd, ad + bc] + [ae + bf, af + be]
= [a, b] · [c, d] + [a, b] · [e, f ],

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 251

proving distributivity on Z. Next, [a, b] · [1, 0] = [a · 1 + b · 0, a · 0 + b · 1] = [a, b] proves


neutrality of [1, 0]. It remains to prove (D.28). Note that, due to (D.15), the conclusion
is equivalent to a = b or c = d. We assume 0 ≤ a < b and have to prove c = d. According
to Def. D.5, a < b means b = a + k for some k ∈ N. Thus, [ac + bd, ad + bc] = [0, 0]
implies
(D.12b)
ac+(a+k)d = ac+bd = ad+bc = ad+(a+k)c ⇒ kd = kc ⇒ c = d, (D.29)

establishing the case. 

Definition D.19. For each k, l ∈ Z, let

l≤k :⇔ k − l ∈ N0 . (D.30)

Theorem D.20. (a) The relation defined in (D.30) constitutes a total order on Z that
is compatible with addition and multiplication, i.e. it satisfies (4.3).

(b) The map ι from (D.22) is strictly increasing.

Proof. (a) follows from (D.30), (D.24), and Th. D.8 since N0 is closed under addition
and multiplication.
(b): According to Def. D.5, if m, n ∈ N with n < m, then m = n + k for some k ∈ N.
In consequence ι(m) = ι(n) + ι(k) by (D.23), i.e. ι(m) − ι(n) = ι(k) ∈ N, proving
ι(n) < ι(m). 

D.4 Rational Numbers


The remaining two deficiencies of the set of integers Z (as compared with R) are the
lack of inverse elements for multiplication and that the order ≤ lacks completeness.
We proceed to the construction of the rational numbers, which will provide the inverse
elements for multiplication. The completion of the order will then be achieved in the
last step in the next section.

Definition and Remark D.21. The relation ∼ on Z × (Z \ {0}) defined by

(a, b) ∼ (c, d) :⇔ ad = bc, (D.31)

constitutes an equivalence relation on Z × (Z \ {0}) (cf. Def. 2.23): Indeed, noting that
(D.31) is precisely the same as (D.15) if + is replaced by ·, the proof from Def. and Rem.
D.9 also shows that (D.31) does, indeed, define an equivalence relation on Z × (Z \ {0}):
One merely replaces each + with · and each N0 with Z or Z \ {0}), respectively. The

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 252

only modification needed occurs for 0 ∈ {a, c, e} in the proof of transitivity (in this case,
the proof of Def. and Rem. D.9 yields adcf = 0 = bcde, which does not imply af = be),
where one now argues, for a = 0,

(a, b) ∼ (c, d) ∧ (c, d) ∼ (e, f ) ⇒ ad = 0 = bc ∧ cf = de


b6=0 d6=0
⇒ c=0 ⇒ e=0 ⇒ af = 0 = be ⇒ (a, b) ∼ (e, f ),

for c = 0,

(a, b) ∼ (c, d) ∧ (c, d) ∼ (e, f ) ⇒ ad = 0 = bc ∧ cf = 0 = de


d6=0
⇒ a = e = 0 af = 0 = be ⇒ (a, b) ∼ (e, f ),

and, for e = 0,

(a, b) ∼ (c, d) ∧ (c, d) ∼ (e, f ) ⇒ ad = bc ∧ cf = 0 = de


f 6=0 d6=0
⇒ c=0 ⇒ a=0 ⇒ af = 0 = be ⇒ (a, b) ∼ (e, f ).

Definition D.22. (a) Define the set of rational numbers Q as the set of equivalence
classes of the equivalence relation ∼ defined in (D.31), i.e.
 
Q := Z × (Z \ {0}) / ∼ = [(a, b)] : (a, b) ∈ Z × (Z \ {0}) (D.32)

is the quotient set of Z×(Z\{0}) with respect to ∼ (cf. Ex. 2.24(c)). As is common,
we will write
a
:= a/b := [(a, b)] (D.33)
b
for the equivalence class of (a, b) with respect to ∼.
(b) Addition on Q is defined by
a c  a c ad + bc
+ : Q × Q −→ Q, , 7→ + := . (D.34)
b d b d bd
Multiplication on Q is defined by
a c  a c ac
· : Q × Q −→ Q, , 7→ · := . (D.35)
b d b d bd

For the definitions in Def. D.22(b) to make sense, one needs to check that they do not
depend on the chosen representatives of the equivalence classes, and that the results of
both addition and multiplication are always elements of Q. All this is provided by the
following lemma.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 253

Lemma D.23. The definitions in Def. D.22(b) do not depend on the chosen represen-
tatives, i.e.
!
a ã c c̃ ad + bc ãd˜ + b̃c̃
∀ ∀ = ∧ = ⇒ = (D.36)
a,c,ã,c̃∈Z ˜
b,d,b̃,d∈Z\{0} b b̃ d d˜ bd b̃d˜
and  
a ã c c̃ ac ãc̃
∀ ∀ = ∧ = ⇒ = . (D.37)
a,c,ã,c̃∈Z ˜
b,d,b̃,d∈Z\{0} b b̃ d d˜ bd b̃d˜
Furthermore, the results of both addition and multiplication are always elements of Q.

Proof. (D.36) and (D.37): a/b = ã/b̃ means ab̃ = ãb, c/d = c̃/d˜ means cd˜ = c̃d, implying
ad + bc ãd˜ + b̃c̃
(ad + bc)b̃d˜ = bd(ãd˜ + b̃c̃), i.e. = (D.38)
bd b̃d˜
and
ac ãc̃
acb̃d˜ = bdãc̃,
= . i.e. (D.39)
bd b̃d˜
That the results of both addition and multiplication are always elements of Q follows
from (D.28), i.e. from the fact that Z has no zero divisors. In particular, if b, d 6= 0,
then bd 6= 0, showing (ad + bc)/(bd) ∈ Q and (ac)/(bd) ∈ Q. 
Theorem D.24. (a) The set of rational numbers Q with addition and multiplication as
defined in Def. D.22 forms a field, where 0/1 and 1/1 are the neutral elements with
respect to addition and multiplication, respectively, (−a/b) is the additive inverse
to a/b, whereas b/a is the multiplicative inverse to a/b with a 6= 0.
(b) Defining subtraction and division in the usual way, for each r, s ∈ Q, by s − r :=
s + (−r) and s/r := sr−1 , respectively, with −r denoting the additive inverse of r
and r−1 denoting the multiplicative inverse of r 6= 0, all the rules stated in Th. C.10
are valid in Q.
(c) The map
k
ι : Z −→ Q, ι(k) := , (D.40)
1
is a monomorphism, i.e. it is injective and satisfies
∀ ι(k + l) = ι(k) + ι(l), (D.41a)
k,l∈Z

∀ ι(kl) = ι(k) · ι(l). (D.41b)


k,l∈Z

It is customary to identify Z with ι(Z), as it usually does not cause any confusion.
One then just writes k instead of k1 .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 254

Proof. (a): We verify + and · to be commutative and associative on Q: Let a, c, e ∈ Z


and b, d, f ∈ Z \ {0}. Then, using commutativity on Z, we compute
c a cb + da ad + bc a c c a ca ac a c
+ = = = + , · = = = · ,
d b db bd b d d b db bd b d
showing commutativity on Q. Using associativity and distributivity on Z, we compute
 
a c e a cf + de adf + b(cf + de) (ad + bc)f + bde
+ + = + = =
b d f b df bdf bdf
ad + bc e  a c  e
= + = + + ,
bd f b d f
   
a c e a(ce) (ac)e a c e
· · = = = · · ,
b d f b(df ) (bd)f b d f
showing associativity on Q. We proceed to checking distributivity on Q: Using commu-
tativity, associativity, and distributivity on Z, we compute
 
a c e a(cf + de) acf + dae acbf + bdae ac ae a c a e
· + = = = = + = · + · ,
b d f bdf bdf bdbf bd bf b d b f
proving distributivity on Q. We now check the claims regarding neutral and inverse
elements:
a 0 a·1+b·0 a
+ = = ,
b 1 b·1 b
a −a ab + b(−a) Def. C.7(iii) for Z (a − a)b 0 (D.31) 0
+ = 2
= 2
= 2 = ,
b b b b b 1
a 1 a·1 a
· = = ,
b 1 b·1 b
a b ab (D.31) 1
· = = .
b a ba 1
Thus, (Q, +, ·) is a ring and (Q \ {0}, ·) is a group, implying Q to be a field.
(b) is a consequence of (a), since Th. C.10 and its proof are valid in every field.
(c): The map ι is injective, as ι(k) = k/1 = ι(l) = l/1 implies k · 1 = l · 1, i.e. k = l.
Moreover,
k l k·1+1·l k+l
ι(k) + ι(l) = + = = = ι(k + l), (D.42a)
1 1 1 1
k l kl
ι(k) · ι(l) = · = = ι(kl), (D.42b)
1 1 1
completing the proof. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 255

Definition and Remark D.25. Define


 
+ a
Q := r ∈ Q : ∃ r= . (D.43)
a,b∈N b

We then have the decomposition

Q = Q+ ∪{0}
˙ ∪˙ Q− , Q− := −Q+ = {r ∈ Q : −r ∈ Q+ }, (D.44)

since

a/b ∈ Q+ ⇔ (a > 0 ∧ b > 0) ∨ (a < 0 ∧ b < 0) , (D.45a)
a/b = 0 ⇔ a = 0, (D.45b)

a/b ∈ Q− ⇔ (a > 0 ∧ b < 0) ∨ (a < 0 ∧ b > 0) . (D.45c)

Definition D.26. For each r, s ∈ Q, let

s≤r :⇔ r − s ∈ Q+ +
0 := Q ∪ {0}. (D.46)

Theorem D.27. (a) The relation defined in (D.46) constitutes a total order on Q that
is compatible with addition and multiplication, i.e. it satisfies (4.3); in other words
(Q, +, ·, ≤) constitutes a totally ordered field.

(b) All the rules stated in Th. 4.5 are valid in Q.

(c) The map ι from (D.40) is strictly increasing.

Proof. (a) follows from (D.46), (D.44), and Th. D.8, since it is immediate from (D.34)
and (D.35) that Q+ is closed under addition and multiplication.
(b) is a consequence of (a), since Th. 4.5 and its proof are valid in every totally ordered
field.
(c): According to Def. D.27, if k, l ∈ Z with l < k, then n := k − l ∈ N. In consequence
ι(k) = ι(l) + ι(n) by (D.41a), i.e. ι(k) − ι(l) = ι(n) = n/1 ∈ Q+ , proving ι(l) < ι(k). 

D.5 Real Numbers


In the previous section, the construction of the rational numbers Q yielded a totally
ordered field. However, the order on Q is not complete – for example, Rem. and Def.
7.62 shows that the set M := {r ∈ Q : r2 < 2}, which is bounded from above (for
example by 2), has no supremum in Q (otherwise, we had a rational number q = sup M
with q 2 = 2). Finally, in the present section, we will start out from Q to construct

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 256

the set of real numbers R such that it becomes a complete totally ordered field. There
are several different important constructions to obtain R from Q. We will describe
the construction that defines real numbers as equivalence classes of rational Cauchy
sequences following [EHH+ 95, Ch. 2.3]. The construction using so-called Dedekind cuts
can be found in [EHH+ 95, Ch. 2.2], the construction via nested intervals in [EHH+ 95,
Ch. 2.4].
Definition D.28. (a) Let S denote the set of all Cauchy sequences in Q, where we
call a sequence (rn )n∈N in Q a Cauchy sequence if, and only if,
∀ ∃ ∀ |rn − rm | < ǫ, (D.47)
ǫ∈Q+ N ∈N n,m>N

which defers from (7.25) in that ǫ has to be from Q+ rather than from R+ .
(b) Addition on S is defined by
+ : S × S −→ S, ((rn )n∈N , (sn )n∈N ) 7→ (rn )n∈N + (sn )n∈N := (rn + sn )n∈N . (D.48)
Multiplication on S is defined by
· : S × S −→ S, ((rn )n∈N , (sn )n∈N ) 7→ (rn )n∈N · (sn )n∈N := (rn sn )n∈N . (D.49)
As a consequence of the following Lem. D.29, addition and multiplication are well-
defined on S.
Lemma D.29. If (rn )n∈N and (sn )n∈N are Cauchy sequences in Q, so are (rn + sn )n∈N
and (rn sn )n∈N .

Proof. The proofs are analogous to the proofs of Th. 7.13(7.11b),(7.11c):


Given ǫ ∈ Q+ , there exists N ∈ N such that, for each n, m > N , |rn − rm | < ǫ/2 and
|sn − sm | < ǫ/2, implying
∀ |rn + sn − (rm + sm )| ≤ |rn − rm | + |sn − sm | < ǫ/2 + ǫ/2 = ǫ, (D.50)
n,m>N

proving (rn + sn )n∈N is Cauchy.


The proof of Th. 7.29 shows both (rn )n∈N and (sn )n∈N are bounded, i.e. there exists M ∈
Q+ that is an upper bound for the sets {|rn | : n ∈ N} and {|sn | : n ∈ N}. Moreover,
given ǫ ∈ Q+ , there exists N ∈ N such that, for each n, m > N , |rn − rm | < ǫ/(2M ) and
|sn − sm | < ǫ/(2M ), implying
 
|rn sn − rm sm | = (rn − rm )sn + rm (sn − sm )
 
∀  Mǫ Mǫ ,
n,m>N
≤ |sn | · |rn − rm | + |rm | · |sn − sm | < + =ǫ
2M 2M
(D.51)
completing the proof of the lemma. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 257

Theorem D.30. (S, +) is a group and, in addition, S is associative and commutative


with respect to multiplication. Moreover, distributivity also holds in S. In algebraic
terms, this can be summarized as the statement that (S, +, ·) constitutes a commutative
ring.

Proof. Note that, since the rational sequence (rn )n∈N is nothing but the function f :
N −→ Q, f (n) = rn , addition and multiplication as defined in Def. D.28(b) is analogous
to the definition of addition and multiplication of real-valued functions in (6.1a), (6.1c),
respectively. It is an easy exercise to verify that these function operations always inherit
associativity, commutativity, and distributivity if these rules hold for the operations
defined on the function’s codomain (i.e. for + and · on Q in our present situation of
rational sequences). The constant sequence (0, 0, . . . ) is the neutral element of addition
on S and −(rn )n∈N = (−rn )n∈N is the additive inverse of (rn )n∈N . 

The reason that we need another step in our construction of R is the fact that S is not
a field: As soon as 0 occurs, even just once, in the sequence (rn )n∈N ∈ S, the sequence
does not have a multiplicative inverse (where the neutral element of multiplication is
obviously the constant sequence (1, 1, . . . )). The solution to this problem consists of
factoring out all sequences converging to 0.
Definition and Remark D.31. Let
n o
N := (rn )n∈N ∈ S : lim rn = 0 . (D.52)
n→∞

be the set of rational sequences converging to zero. The relation ∼ on S defined by

(rn )n∈N ∼ (sn )n∈N ⇔ (rn )n∈N − (sn )n∈N ∈ N , (D.53)

constitutes an equivalence relation on S (cf. Def. 2.23): Indeed, ∼ is reflexive, as f ∈ S


implies f − f = 0 ∈ N ; ∼ is symmetric, since f, g ∈ S with f − g ∈ N implies
g − f ∈ N ; ∼ is transitive, as f, g, h ∈ S with f − g ∈ N and g − h ∈ N implies
f − h = f − g + g − h ∈ N.
Definition D.32. (a) Define the set of real numbers R as the set of equivalence classes
of the equivalence relation ∼ defined in (D.53), i.e.

R := S/ ∼ = {[(rn )n∈N ] : (rn )n∈N ∈ S} (D.54)

is the quotient set of S with respect to ∼ (cf. Ex. 2.24(c)).

(b) Addition on R is defined by



+ : R × R −→ R, [f ], [g] →7 [f ] + [g] := [f + g]. (D.55)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 258

Multiplication on R is defined by

· : R × R −→ R, [f ], [g] →7 [f ] · [g] := [f g]. (D.56)

Once again, for the definitions in Def. D.32(b) to make sense, one needs to check that
they do not depend on the chosen representatives of the equivalence classes, and once
again, we provide a lemma providing this check:

Lemma D.33. The definitions in Def. D.32(b) do not depend on the chosen represen-
tatives, i.e.

∀ f − f˜ ∈ N ∧ g − g̃ ∈ N ⇒ f + g − (f˜ + g̃) ∈ N (D.57)
f,g,f˜,g̃∈S

and 
∀ f − f˜ ∈ N ∧ g − g̃ ∈ N ⇒ f g − (f˜g̃) ∈ N . (D.58)
f,g,f˜,g̃∈S

Proof. Let f = (rn )n∈N , g = (sn )n∈N , f˜ = (r̃n )n∈N , g̃ = (s̃n )n∈N be elements of S such
that f − f˜ ∈ N and g − g̃ ∈ N , i.e. limn→∞ (rn − r̃n ) = limn→∞ (sn − s̃n ) = 0.

Then (7.11b) implies 0 = limn→∞ rn + sn − (r̃n + s̃n ) , proving (D.57).
To prove (D.58), one computes
 
lim rn sn − r̃n s̃n = lim rn (sn − s̃n ) − s̃n (rn − r̃n ) = 0, (D.59)
n→∞ n→∞

where the last equality follows from the boundedness of (rn )n∈N and (s̃n )n∈N together
with Prop. 7.11(b). 

We will also use the following auxiliary result:

Proposition D.34. If (rn )n∈N ∈ S, then precisely one of the following statements is
correct:

(rn )n∈N ∈ N , (D.60a)


∃ #{n ∈ N : rn ≤ ǫ} ∈ N0 , (D.60b)
ǫ∈Q+

∃ #{n ∈ N : rn ≥ −ǫ} ∈ N0 . (D.60c)


ǫ∈Q+

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 259

Proof. Let us first verify that the three statements in (D.60) are mutually exclusive. If
(D.60a) holds, then, for every ǫ ∈ Q+ , −ǫ < rn < ǫ holds for almost all (in particular,
for infinitely many) n ∈ N, i.e. (D.60b) and (D.60c) are both false. If (D.60b) holds,
then (D.60a) must be false as we have just seen. Moreover, if rn ≤ ǫ holds for at most
finitely many n ∈ N, then rn > ǫ > 0 must hold for infinitely many n ∈ N, i.e. (D.60c)
is false.
Now suppose (D.60a) and (D.60b) are false. We have to show that (D.60c) is true. Since
(D.60a) is false, there exists δ > 0 and an increasing sequence of indices (nk )k∈N with
|rnk | > δ for each k ∈ N. Since (D.60b) is false, there is an increasing sequence of indices
(mk )k∈N with rmk < 1/k. Thus, since (rn )n∈N is a Cauchy sequence, only finitely many
rnk > δ and infinitely many rnk < −δ. Now, if N ∈ N is such that |rn − rm | < δ/2 for
all n, m > N and k0 ∈ N such that nk0 > N , then rn < −δ/2 for each n > N (since
|rn − rnk0 | < δ/2). Thus, (D.60c) holds with ǫ := δ/2. 

Theorem D.35. (a) The set of real numbers R with addition and multiplication as
defined in Def. D.32 forms a field, where [(0, 0, . . . )] and [(1, 1, . . . )] are the neutral
elements with respect to addition and multiplication, respectively.

(b) The map  


ι : Q −→ R, ι(r) := (r, r, . . . ) , (D.61)
is a monomorphism, i.e. it is injective and satisfies

∀ ι(r + s) = ι(r) + ι(s), (D.62a)


r,s∈Q

∀ ι(rs) = ι(r) · ι(s). (D.62b)


r,s∈Q

It is customary to identify Q withι(Q), as it usually does not cause any confusion.
One then just writes r instead of (r, r, . . . ) .

Proof. (a): Clearly, Def. D.32(b) ensures the laws of associativity and commutativity
of addition and multiplication valid in S are preserved in R, and, likewise, the law of
distributivity. It is also immediate from (D.55) and (D.56), respectively, that [(0, 0, . . . )]
and [(1, 1, . . . )] are the respective neutral elements of addition and multiplication. More-
over, if −f is the additive inverse of f ∈ S, then [−f ] is the additive inverse of [f ] ∈ R.
It remains to show that each x = [(rn )n∈N ] 6= [(0, 0, . . . )] has a multiplicative inverse x−1
in R. We claim x−1 = [(sn )n∈N ], where
(
rn−1 for rn 6= 0,
∀ sn := (D.63)
n∈N 1 for rn = 0.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 260

We need to verify [(sn )n∈N ] ∈ R, i.e. (sn )n∈N is a Cauchy sequence. We know (rn )n∈N is
a Cauchy sequence that does not converge to 0. Thus, according to Prop. D.34, there
exists δ > 0 and M ∈ N such that, for each n > M , we have |rn | > δ (in particular,
rn 6= 0). Let ǫ > 0. As (rn )n∈N is a Cauchy sequence, there exists N ∈ N such that
N ≥ M and, for each n, m > N , |rn − rm | < ǫ δ 2 . Thus,

1 1 rn − rm ǫ δ 2
∀ |sn − sm | = − = < = ǫ, (D.64)
n,m>N rn rm rn rm δ2

proving (sn )n∈N is a Cauchy sequence. Moreover,


       
(rn )n∈N · (sn )n∈N = (rn sn )n∈N = (1, 1, . . . , ) , (D.65)

since rn sn = 1 for almost all n ∈ N, and the proof of (a) is complete.


(b): The map ι is injective, since ι(r) = [(r, r, . . . )] = ι(s) = [(s, s, . . . )] implies
limn→∞ (r − s) = 0, i.e. r = s. Moreover,
     
ι(r) + ι(s) = (r, r, . . . ) + (s, s, . . . ) = (r + s, r + s, . . . ) = ι(r + s), (D.66a)
     
ι(r) · ι(s) = (r, r, . . . ) · (s, s, . . . ) = (rs, rs, . . . ) = ι(rs), (D.66b)

completing the proof. 


Definition D.36. We define R+ to consist of all real numbers represented by sequences
(rn )n∈N such that there exists ǫ ∈ Q+ satisfying rn > ǫ for almost all n ∈ N, i.e.
 
+
 
R := (rn )n∈N ∈ R : ∃ + #{n ∈ N : rn ≤ ǫ} ∈ N0 . (D.67)
ǫ∈Q

Proposition D.37. (a) The definition in (D.67) does not depend on the chosen rep-
resentatives (rn )n∈N .

(b) We have the decomposition

R = R+ ∪{0}
˙ ∪˙ R− , R− := −R+ = {x ∈ R : −x ∈ R+ }. (D.68)

Proof. (a): If (sn )n∈N ∈ S with limn→∞ (rn − sn ) = 0, then |rn − sn | < ǫ/2 for almost all
n ∈ N. Thus, since |sn | ≥ |rn | − |rn − sn |, we obtain sn > ǫ/2 for almost all n ∈ N, i.e.
#{n ∈ N : sn ≤ 2ǫ } ∈ N0 .
(b) is an immediate consequence of Prop. D.34. 
Definition D.38. For each x, y ∈ R, let

y≤x :⇔ x − y ∈ R+ +
0 := R ∪ {0}. (D.69)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 261

Theorem D.39. (a) The relation defined in (D.69) constitutes a total order on R that
is compatible with addition and multiplication, i.e. it satisfies (4.3); in other words
(R, +, ·, ≤) constitutes a totally ordered field.

(b) The map ι from (D.61) is strictly increasing.

Proof. (a) follows from (D.69), (D.68), and Th. D.8, once we have shown that R+ is
closed under addition and multiplication. Let (rn )n∈N ∈ S, (sn )n∈N ∈ S. If rn > ǫ1 ∈ Q+
for almost all n ∈ N and sn > ǫ2 ∈ Q+ for almost all n ∈ N, then rn + sn > ǫ1 + ǫ2 ,
showing R+ is closed under addition. Moreover, rn sn > ǫ1 ǫ2 , showing R+ is closed under
multiplication.
(b): According to Def. D.39, if r, s ∈ Q with s < r, then q := r − s ∈ Q+ . In
consequence ι(r) = ι(s) + ι(q) by (D.62a), i.e. ι(r) − ι(s) = ι(q) = [(q, q, . . . )] ∈ R+ ,
proving ι(s) < ι(r). 

Finally, we will show in Th. D.41 below that the order ≤ on R is complete. However,
we first need some additional auxiliary results.

Proposition D.40. (a) For each x ∈ R, there is (rn )n∈N ∈ S satisfying limn→∞ rn = x.

(b) Every (rn )n∈N ∈ S converges in R – more precisely, limn→∞ rn = [(rn )n∈N ].

(c) Every Cauchy sequence in R converges in R.

Proof. (a) and (b): If x = [(rn )n∈N ] with (rn )n∈N ∈ S, then, given ǫ > 0, choose N ∈ N
such that, for each m, n > N , one has |rn − rm | < ǫ/2. Then, for each k > N , one has
|x − rk | = |[(rn − rk )n∈N ]| < ǫ, since |rn − rk | < ǫ/2 for all n ≥ k, showing limn→∞ rn = x.
(c): Let (xn )n∈N be a Cauchy sequence in R. According to (a), for each n ∈ N, there
exists rn ∈ Q such that |xn − rn | < n1 . Then (rn )n∈N is a Cauchy sequence: Given ǫ > 0,
choose k ∈ N such that k1 < 3ǫ and |xn − xm | < 3ǫ for each n, m > k. Then
ǫ ǫ ǫ
∀ |rn − rm | ≤ |rn − xn | + |xn − xm | + |xm − rm | < + + = ǫ, (D.70)
n,m>k 3 3 3

showing (rn )n∈N is Cauchy. Thus, from (b), we obtain x ∈ R with limn→∞ rn = x. We
can now show, limn→∞ xn = x as well: Given ǫ > 0, choose N ∈ N such that N1 < 2ǫ and
|x − rn | < 2ǫ for each n > N . Then
ǫ ǫ
∀ |x − xn | ≤ |x − rn | + |rn − xn | < + = ǫ, (D.71)
n>N 2 2
showing limn→∞ xn = x and completing the proof. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 262

Theorem D.41. The order ≤ on R is complete, i.e. (R, +, ·, ≤) constitutes a complete


totally ordered field.

Proof. Let ∅ 6= A ⊆ R and let M ∈ R be an upper bound for A. We have to show that
A has a supremum in R. To this end, we recursively construct two Cauchy sequences
(xn )n∈N and (yn )n∈N in R such that (xn )n∈N is increasing, (yn )n∈N is decreasing, xn < yn ,
and limn→∞ (yn − xn ) = 0. Let x1 ∈ A be arbitrary and y1 := M . Define
 ( 
(xn + yn )/2 if (xn + yn )/2 is not an upper bound for A,
xn+1 := 
 xn otherwise, 
∀ 
 (  (D.72)

n∈N
 (xn + yn )/2 if (xn + yn )/2 is an upper bound for A, 
yn+1 :=
yn otherwise.

Then, clearly, the xn are increasing, the yn are decreasing, and xn ≤ yn holds for each
n ∈ N. Moreover, letting d := M − x1 ≥ 0, a simple induction shows yn − xn = d/2n−1
and limn→∞ (yn − xn ) = 0. Also, for m > n,
m−1 m−1 m−1 m−1−n
X X d X −i+n d X −i 2d
xm − xn = (xi+1 − xi ) ≤ d 2−i = 2 = 2 ≤ n, (D.73)
i=n i=n
2n i=n 2n i=0 2

showing (xn )n∈N is a Cauchy sequence. Analogous, one sees that (yn )n∈N is a Cauchy
sequence. By Prop. D.40(c), we obtain s ∈ R such that s = limn→∞ xn = limn→∞ (yn −
xn + xn ) = limn→∞ yn . We claim s = sup A. If s < y, then there is n ∈ N with
s ≤ yn < y, showing y ∈/ A, i.e. s is an upper bound for A. If y < s, then there is n ∈ N
with y < xn ≤ s, showing y is not an upper bound for A. Thus, s is the smallest upper
bound for A, i.e. s = sup A. 

D.6 Uniqueness
We will show in Th. D.45 below that, up to a unique isomorphism, R is the only complete
totally ordered field.

Notation D.42. Let (A, +, ·, ≤) be a complete totally ordered field. The neutral el-
ements with respect to + and ·, we denote with 0A and 1A , respectively. We recur-
sively define (n + 1)A := nA + 1A for each n ∈ N. Then NA := {nA : n ∈ N},
ZA := NA ∪ {0A } ∪ {−nA : n ∈ N}, QA := {0A } ∪ { kl : k, l ∈ ZA \ {0A }}.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 263

Proposition D.43. Let (A, +, ·, ≤) and (B, +, ·, ≤) be complete totally ordered fields.
Moreover, let φ : A −→ B be a field isomorphism, i.e. a bijective map, satisfying

∀ φ(x + y) = φ(x) + φ(y), (D.74a)


x,y∈A

∀ φ(xy) = φ(x)φ(y). (D.74b)


x,y∈A

(a) φ(nA ) = nB holds for each n ∈ N.

(b) φ is strictly isotone, i.e.



∀ x < y ⇒ φ(x) < φ(y) . (D.75)
x,y∈A

Proof. (a): As (D.74a) and (D.74b) state φ to be a group homomorphism with respect
to addition and multiplication, respectively, Prop. C.4(a) yields φ(0A ) = 0B and φ(1A ) =
1B . If n ∈ N, then (D.74a) implies φ(nA + 1A ) = φ(nA ) + 1B and, thus, an induction
shows φ(nA ) = nB for each n ∈ N.
(b): If x, y ∈ A with x < y, then, by Rem. and Def. 7.61, there exists a unique z ∈ A
such that z 2 = y − x. Thus, (φ(z))2 = φ(y) − φ(x). By Th. 4.5(c), we have (φ(z))2 > 0
and, thus, φ(x) < φ(y), proving the strict isotonicity of φ. 

Proposition D.44. Let (A, +, ·, ≤) be a complete totally ordered field. If φ : A −→ A


is a (field) automorphism, i.e. a bijective map, satisfying (D.74), then φ is the identity
on A.

Proof. From Prop. D.43(a), we already know φ(n) = n for each n ∈ NA . Next, if
n ∈ NA , then, using Prop. C.4(b), we obtain φ(−n) = −φ(n) = −n, showing φ(k) = k
for each k ∈ ZA . If k, l ∈ ZA \ {0A }, then φ(k/l) = φ(k · l−1 ) = φ(k) · (φ(l))−1 = kl−1 ,
where Prop. C.4(b) was used again. Thus, we already have φ(q) = q for each q ∈ QA .
From Prop. D.43(b), we know φ to be strictly isotone. Finally, if x ∈ A, then, by Th.
7.68(c), there exist a sequences (rn )n∈N in QA and (sn )n∈N in QA such that (rn )n∈N is
strictly increasing, (sn )n∈N is strictly decreasing, and

lim rn = lim sn = x.
n→∞ n→∞

As φ is isotone, we obtain

∀ φ(rn ) = rn ≤ f (x) ≤ sn = φ(sn ).


n∈N

But then f (x) = x follows from Th. 7.16, proving φ = IdA . 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 264

Theorem D.45. Let (A, +, ·, ≤) and (B, +, ·, ≤) be complete totally ordered fields. Then
there exists a unique isomorphism φ : A −→ B, i.e. a unique bijective φ : A −→ B,
satisfying (D.74a), (D.74b), and (D.75).

Proof. Uniqueness: Suppose, φ : A −→ B and ψ : A −→ B are both isomorphisms.


Then, according to Prop. D.44, φ−1 ◦ ψ = IdA , where Prop. C.4(c) was used as well.
However, this already shows ψ = φ.
Existence: Due to Prop. D.43(b), it suffices to show there exists a bijective φ : A −→ B,
satisfying (D.74). We define φ and verify (D.74) in several steps. In the first step, set

∀ φ(nA ) := nB . (D.76)
n∈N0

Then φ : NA ∪ {0A } −→ NB ∪ {0B } is bijective with φ−1 : NB ∪ {0B } −→ NA ∪ {0A },


φ−1 (nB ) = nA . We first verify

∀ φ(1A + nA ) = φ(1A ) + φ(nA ) : (D.77)


n∈N0

Indeed,

φ(1A + nA ) = φ((n + 1)A ) = (n + 1)B = 1B + nB = φ(1A ) + φ(nA ).

Next, we verify
∀ φ(mA + nA ) = φ(mA ) + φ(nA ) (D.78)
m,n∈N0

via induction on m: The case m = 0 holds, due to φ(0A + nA ) = φ(nA ) = nB =


0B + nB = φ(0A ) + φ(nA ). The case m = 1 holds, due to (D.77). For the induction step,
we compute
ind. hyp.
φ((m + 1)A + nA ) = φ(mA + 1A + nA ) = φ(mA ) + φ(1A + nA )
(D.77) (D.77)
= φ(mA ) + φ(1A ) + φ(nA ) = φ((m + 1)A ) + φ(nA ),

proving (D.78). We now prove

∀ φ(mA nA ) = φ(mA )φ(nA ) (D.79)


m,n∈N0

via induction on m: The case m = 0 holds, due to

φ(0A nA ) = φ(0A ) = 0B = φ(0A )φ(nA ).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 265

For the induction step, we compute


(D.78)
φ((m + 1)A nA ) = φ((mA + 1A ) nA ) = φ(mA nA + nA ) = φ(mA nA ) + φ(nA )
ind. hyp. 
= φ(mA )φ(nA ) + φ(nA ) = φ(mA ) + 1B φ(nA )
(D.78)
= φ(mA + 1A ) φ(nA ) = φ((m + 1)A ) φ(nA ).

In particular, according to (D.78) and (D.79), (D.74) holds for each x, y ∈ NA .


In the second step, set
∀ φ(−nA ) := −nB . (D.80)
n∈N

Then φ : ZA −→ ZB is still bijective, where, for each m, n ∈ N, we have φ−1 (−nB ) =


−nA .
Let m, n ∈ N0 . If m ≤ n, then

φ(−mA + nA ) = φ((n − m)A ) = (n − m)B = −mB + nB = φ(−mA ) + φ(nA ).

If m > n, then

φ(−mA + nA ) = φ(−(m − n)A ) = −(m − n)B = −mB + nB = φ(−mA ) + φ(nA ).

Now, for arbitrary m, n ∈ N0 , φ(mA + (−nA )) = φ(−nA + mA ) = φ(−nA ) + φ(mA ) =


φ(mA ) + φ(−nA ) and φ(−mA + (−nA )) = φ(−(mA + nA ) = −φ(mA + nA ) = −(φ(mA ) +
φ(nA )) = −φ(mA ) − φ(nA ) = φ(−mA ) + φ(−nA ). We now consider multiplication, still
for m, n ∈ N0 :
(D.79)
φ((−mA ) nA ) = −φ(mA nA ) = −φ(mA )φ(nA ) = φ(−mA )φ(nA ).

Then φ(mA (−nA )) = φ(mA )φ(−nA ) also follows and


(D.79)
φ((−mA )(−nA ) = φ(mA nA ) = φ(mA )φ(nA ) = φ(−mA )φ(−nA ).

Thus, we have verified (D.74) for each x, y ∈ ZA .


In the third step, set
∀ φ(k/l) := φ(k)/φ(l). (D.81)
k,l∈ZA \{0A }

We verify that (D.81) well-defines φ for each q ∈ QA : If m, n, k, l ∈ ZA with n, l 6= 0A ,


then m/n = k/l implies ml = kn and φ(m)φ(l) = φ(ml) = φ(kn) = φ(k)φ(n). Thus,

φ(m/n) = φ(m)/φ(n) = φ(k)/φ(l) = φ(k/l).

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 266

We show φ : QA −→ QB to be bijective by providing the inverse map: Define ψ :


QB −→ QA by setting

∀ ψ(k/l) := φ−1 (k)/φ−1 (l).


k,l∈ZB \{0B }

We claim that ψ = φ−1 on QB : Indeed, for each k, l ∈ ZA \ {0A } and for each m, n ∈
ZB \ {0B }

ψ(φ(k/l)) = ψ φ(k)/φ(l) = φ−1 (φ(k))/φ−1 (φ(l)) = k/l,

φ(ψ(m/n)) = φ φ−1 (m)/φ−1 (n) = φ(φ−1 (m))/φ(φ−1 (n)) = m/n.

Moreover, we have
   
m k ml + kn φ(ml + kn) (D.74) for ZA φ(m)φ(l) + φ(k)φ(n)
φ + =φ = =
n l nl φ(nl) φ(n)φ(l)
   
φ(m) φ(k) m k
= + =φ +φ
φ(n) φ(l) n l

and
   
m k mk φ(mk) (D.74) for ZA φ(m)φ(k)
φ · =φ = =
n l nl φ(nl) φ(n)φ(l)
   
φ(m) φ(k) m k
= · =φ ·φ .
φ(n) φ(l) n l

Thus, we have verified (D.74) for each x, y ∈ QA .


We now show φ : QA −→ QB to be strictly isotone, i.e.
 
∀ r < s ⇒ φ(r) < φ(s) : (D.82)
r,s∈QA

Let r, s ∈ QA such that r < s. Then d := s − r > 0A , i.e. there are m, n ∈ NA satisfying
d= m n
. Then φ(s) − φ(r) = φ(d) = φ(m)
φ(n)
> 0B , proving φ(r) < φ(s).
In the fourth (and last) step, for each x ∈ A, we choose a sequence (rn )n∈N in QA such
that x = limn→∞ rn and set
φ(x) := lim φ(rn ). (D.83)
n→∞

To show that φ is well-defined by (D.83), we have to verify that (φ(rn ))n∈N does, indeed,
converge in B, and that φ(x) does not depend on the chosen sequence (rn )n∈N . As
(rn )n∈N converges to x, it has to be a Cauchy sequence by Th. 7.29. We show that

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
D CONSTRUCTION OF THE REAL NUMBERS 267

(φ(rn ))n∈N must be a Cauchy sequence as well: Let ǫ ∈ B, ǫ > 0B and choose ǫ̃ ∈ QB
such that 0B < ǫ̃ < ǫ. Then
∃ ∀ |rn − rm | < φ−1 (ǫ̃).
N ∈N n,m>N

As φ is strictly isotone, we obtain


∀ |φ(rn ) − φ(rm )| < ǫ̃ < ǫ,
n,m>N

proving (φ(rn ))n∈N to be a Cauchy sequence. Now Th. 7.29 implies the convergence of
(φ(rn ))n∈N . Next, we show limn→∞ rn = 0A implies limn→∞ φ(rn ) = 0B for each sequence
in QA : Indeed, as above, let ǫ > 0B and choose ǫ̃ ∈ QB such that 0B < ǫ̃ < ǫ. Then
∃ ∀ |rn | < φ−1 (ǫ̃).
N ∈N n>N

As φ is strictly isotone, we obtain


∀ |φ(rn )| < ǫ̃ < ǫ,
n>N

proving limn→∞ φ(rn ) = 0B . Thus, if (rn )n∈N and (sn )n∈N are sequences in QA such that
limn→∞ rn = x = limn→∞ sn , then
lim φ(rn ) = lim φ(rn − sn + sn ) = 0B + lim φ(sn ) = lim φ(sn ),
n→∞ n→∞ n→∞ n→∞

showing φ to be well-defined by (D.83). To see that φ is injective, let x, y ∈ A with


x < y and choose r, s ∈ QA such that x < r < s < y. If (rn )n∈N and (sn )n∈N are
sequences in QA such that x = limn→∞ rn and y = limn→∞ sn , then
 
∃ ∀ rn < r < s < s n .
N ∈N n>N

As φ is strictly isotone, we obtain


 
∀ φ(rn ) < φ(r) < φ(s) < φ(sn ) ,
n>N

showing φ(x) 6= φ(y) and the injectivity of φ. To see that φ is surjective, let b ∈ B and
let (rn )n∈N be an increasing sequence in QB such that b = limn→∞ rn . Then (φ−1 (rn ))n∈N
is an increasing sequence in QA that is bounded, i.e. it must converge to some a ∈ A.
Then
φ(a) = lim φ(φ−1 (rn )) = lim rn = b,
n→∞ n→∞
showing φ to be surjective. Finally, if x, y ∈ A, then let (rn )n∈N and (sn )n∈N be sequences
in QA such that x = limn→∞ rn and y = limn→∞ sn . Then
φ(x + y) = lim φ(rn + sn ) = lim φ(rn ) + lim φ(sn ) = φ(x) + φ(y)
n→∞ n→∞ n→∞
and
φ(xy) = lim φ(rn sn ) = lim φ(rn ) lim φ(sn ) = φ(x)φ(y).
n→∞ n→∞ n→∞
Thus, we have verified (D.74) for each x, y ∈ A, and, thereby completed the proof. 

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 268

E Series: Additional Material

E.1 Riemann Rearrangement Theorem


Here, we provide the details for the proof of the Riemann rearrangement Th. 7.93, that
was merely sketched in Sec. 7.3.3.

Proof of Th. 7.93. As already stated in the sketch, we define


 
−k for x = −∞,
 −k for y = −∞,

∀ xk := x for x ∈ R, yk := y for y ∈ R, (E.1)
k∈N 
 

k for x = ∞, k for y = ∞,

noting xk ≤ yk for almost all k ∈ N. Next, we observe

N = I + ∪˙ I − , where (E.2a)
I + := {j ∈ N : aj ≥ 0}, (E.2b)
I − := {j ∈ N : aj < 0}. (E.2c)

We have to define a suitable bijective map φ : N −→ N such that

∀ bj := aφ(j) , (E.3a)
j∈N
n
X
∀ tn := bj . (E.3b)
n∈N
j=1

The definition of φ will be recursive, and we will also need to recursively define an
auxiliary sequence (σj )j∈N taking values in {−1, 1}, serving as an accounting tool to
keep track if we are in the process of moving right (i.e. adding a+j ) or moving left (i.e.

subtracting aj ). Moreover, we need a recursively defined auxiliary function κ : N −→ N
to update the left and right boundaries xk and yk , respectively, to handle the first and
third case of (E.1) if need be. The recursion is initialized by

φ(1) := 1, (E.4a)
(
1 if t1 ≤ y1 ,
σ1 := (E.4b)
−1 if t1 > y1 ,
(
1 if t1 ≤ y1 ,
κ(1) := (E.4c)
2 if t1 > y1 ,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 269

and completed by
( 
min I + \ φ{1, . . . , j − 1} if σj−1 = 1,
∀ φ(j) :=  (E.5a)
j>1 min I − \ φ{1, . . . , j − 1} if σj−1 = −1,


 1 if σj−1 = 1 and tj ≤ yκ(j−1) ,

−1 if σ
j−1 = 1 and tj > yκ(j−1) ,
∀ σj := (E.5b)
j>1 

 −1 if σj−1 = −1 and tj ≥ xκ(j−1) ,

1 if σj−1 = −1 and tj < xκ(j−1) ,


 κ(j − 1) if σj−1 = 1 and tj ≤ yκ(j−1) ,

1 + κ(j − 1) if σ
j−1 = 1 and tj > yκ(j−1) ,
∀ κ(j) := (E.5c)
j>1 

 κ(j − 1) if σj−1 = −1 and tj ≥ xκ(j−1) ,

1 + κ(j − 1) if σj−1 = −1 and tj < xκ(j−1) .

We note that φ is well-defined, since, according to (7.86), both I + and I − must have
infinitely many elements. Moreover, φ is injective, since, for j1 < j2 , φ(j2 ) 6= φ(j1 )
is immediate from (E.5a). Finally, φ is also surjective: Otherwise, there is a smallest
n ∈ N \ {1} such that n ∈ / φ(N). Suppose n ∈ I + . Then, according to (E.5a), there
must be j0 ∈ N such that σj = −1 for every j > j0 , i.e., according to P∞(E.5b) and

(E.5c), tj ≥ xκ(j0 ) ∈ R for each j > j0 , which is in contradiction to the j=1 aj = ∞
P
part of (7.86). Analogously, n ∈ I − leads to a contradiction to the ∞ +
j=1 aj = P∞ part

of (7.86), completing theP∞proof of surjectivity of φ. So we have shown that P∞ j=1 bj
is a rearrangement of j=1 aj as desired. We still need to verify that j=1 bj (i.e.
(tn )n∈N ) has precisely all elements of [x, y] as cluster points. To this end, first note
that, due to (7.86) and (E.1), limj→∞ xκ(j) = −∞ holds if, and only if, x = −∞; and
limj→∞ xκ(j) = ∞ holds if, and only if, x = ∞; and likewise for the yκ(j) and y. If
x = −∞, then limj→∞ xκ(j) = −∞ and the bijectivity of φ together with (E.5b) and
(E.5c) implies
∀ ∃ tj < xκ(j−1) ≤ −N, (E.6)
N ∈N j∈N

showing −∞ is a cluster point of (tn )n∈N . Analogously, if y = ∞, then limj→∞ yκ(j) = ∞


and the bijectivity of φ together with (E.5b) and (E.5c) implies

∀ ∃ tj > yκ(j−1) ≥ N, (E.7)


N ∈N j∈N

showing ∞ is a cluster point of (tn )n∈N . Now let ξ ∈ [x, y] ∩ R and ǫ > 0. Due to

limj→∞ a+
j = limj→∞ aj = 0, we have

∃ ∀ tj − tj−1 < ǫ. (E.8)


N ∈N j>N

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 270

Due to the bijectivity of φ together with (E.5b) and (E.5c), for each j0 ∈ N, there exists
j > max{j0 , N } such that tj−1 ≤ ξ ≤ tj , showing ξ is a cluster point of (tn )n∈N . On the
other hand, if ξ ∈] − ∞, x[, then x 6= −∞. If x = ∞, then limj→∞ tj = ∞ and ξ is not
a cluster point of (tn )n∈N . If ξ < x < ∞, then let ǫ := (x − ξ)/2 and choose N as in
(E.8). Then, by (E.5b) and (E.5c), for each j > N , tj > x − ǫ = ξ + ǫ, showing ξ is
not a cluster point of (tn )n∈N . Analogously, one sees that ξ ∈]y, ∞[ can not be a cluster
point of (tn )n∈N . 

E.2 b-Adic Representations of Real Numbers


The main goal of this section is to provide a proof of Th. 7.99. We begin with some
preparatory lemmas.
Lemma E.1. Given a natural number b ≥ 2, consider the b-adic series given by (7.96).
Then ∞
X
dN −ν bN −ν ≤ bN +1 , (E.9)
ν=0

and, in particular, the b-adic series converges to some x ∈ R+ 0 . Moreover, equality in


(E.9) holds if, and only if, dn = b − 1 for every n ∈ {N, N − 1, N − 2, . . . }.

Proof. One estimates, using the formula for the value of a geometric series:
∞ ∞ ∞
X
N −ν
X
N −ν
X 1
dN −ν b ≤ (b − 1) b = (b − 1)b N
b−ν = (b − 1)bN 1 = bN +1 . (E.10)
ν=0 ν=0 ν=0
1− b

Note that (E.10) also shows that equality is achieved if all dn are equal to b − 1. Con-
versely, if there is n ∈ {N, N − 1, N − 2, . . . } such that dn < b − 1, then there is ñ ∈ N
such that dN −ñ < b − 1 and one estimates

X ñ−1
X ∞
X
N −ν N −ν N −ñ
dN −ν b < dN −ν b + (b − 1)b + dN −ν bN −ν ≤ bN +1 , (E.11)
ν=0 ν=0 ν=ñ+1

showing that the inequality in (E.9) is strict. 


Lemma E.2. Given a natural number b ≥ 2, consider two b-adic series

X ∞
X
N −ν
x := dN −ν b = eN −ν bN −ν , (E.12)
ν=0 ν=0

N ∈ Z and dn , en ∈ {0, . . . , b − 1} for each n ∈ {N, N − 1, N − 2, . . . }. If dN < eN , then


eN = dN + 1, dn = b − 1 for each n < N and en = 0 for each n < N .

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 271

Proof. By subtracting dN bN from both series, one can assume dN = 0 without loss of
generality. From Lem. E.1, we know

X ∞
X
N −ν
x= dN −ν b = dN −1−ν bN −1−ν ≤ bN . (E.13a)
ν=0 ν=0

On the other hand: ∞


X
x= eN −ν bN −ν ≥ bN . (E.13b)
ν=0

Combining (E.13a) and (E.13b) yields x = bN . Once again employing Lem. E.1, (E.13a)
also shows that dn = b − 1 for each n ≤ N − 1 as claimed. Since eN > 0 and en ≥ 0
for each n, equality in (E.13b) can only occur for eN = 1 and en = 0 for each n < N ,
thereby completing the proof of the lemma. 

Notation E.3. For each x ∈ R, we let

⌊x⌋ := max{k ∈ Z : k ≤ x} (E.14)

denote the integral part of x (also called floor of x or x rounded down).

Proof of Th. 7.99. We start by constructing numbers N and dn , n ∈ {N, N − 1, N −


2, . . . }, such that (7.97) holds. For x = 0, one chooses an arbitrary N ∈ Z and dn = 0
for each n ∈ {N, N − 1, N − 2, . . . }. Thus, for the remainder of the proof, fix x > 0. Let

N := max{n ∈ Z : bn ≤ x}. (E.15)

The numbers dN −n ∈ {0, . . . , b − 1} and xn ∈ R+ , n ∈ N0 , are defined inductively by


letting
jxk
dN := N , x0 := dN bN , (E.16a)
 b 
x − xn−1
dN −n := , xn := xn−1 + dN −n bN −n for n ≥ 1. (E.16b)
bN −n

Claim 3. One can verify by induction on n that the numbers dN −n and xn enjoy the
following properties for each n ∈ N0 :

dN −n ∈ {0, . . . , b − 1}, (E.17a)


Xn
0 < xn = dN −ν bN −ν ≤ x, (E.17b)
ν=0
N −n
x − xn < b . (E.17c)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 272

Proof. The induction is carried out for all three statements of (E.17) simultaneously.
From (E.15), we know bN ≤ x < bN +1 , i.e. 1 ≤ bxN < b. Using (E.16a), this yields
dN ∈ {1, . . . , b − 1} and 0 < x0 = dN bN = bN dN ≤ bN bxN = x as well as x − x0 =
x − dN bN = bN ( bxN − dN ) < bN . For n ≥ 1, by induction, one obtains 0 ≤ x − xn−1 <
b1+N −n , i.e. 0 ≤ x−x n−1
bN −n
< b. Using (E.16b), this yields dN −n ∈ {0, . . . , b − 1} and
x =x + dN −n b N −n
≤ xn−1 + bN −n x−x n−1
= x. Moreover, by induction, 0 < xn−1 =
Pn n−1 n−1 N −ν bN −n
N −n
ν=0 dN −ν b , such that (E.16b) implies P xn = xn−1 + dNP −n b ≥ xn−1 > 0 and
n−1
xn = xn−1 + dN −n bN −n = dN −n bN −n + ν=0 dN −ν bN −ν = nν=0 dN −ν bN −ν . Finally,
x − xn = x − xn−1 − dN −n bN −n = bN −n ( x−x n−1
bN −n
− dN −n ) ≤ bN −n , completing the proof
of the claim. N

Since, for each n ∈ N0 ,


(E.17b) (E.17c)
0 ≤ x − xn < bN −n , (E.18)

and limn→∞ bN −n = 0, we have limn→∞ xn = x, thereby establishing (7.97).


It remains to verify the equivalence of (i) – (iv).
(ii) ⇒ (i) is trivial.
“(iii) ⇒ (i)”: Assume (iii) holds. Without loss of generality, we can assume that n0
is the largest index such that dn = 0 for each n ≤ n0 . We distinguish two cases. If
n0 < N − 1 or dN 6= 1, then
N −n
X 0 −2 ∞
X
N −ν
dN −ν b + (dn0 +1 − 1)b n0 +1
+ (b − 1) bN −ν
ν=0 ν=N −n0

is a different b-adic representation of x and its first coefficient is nonzero. If n0 = N − 1


and dN = 1, then
X∞ ∞
X
N −ν
(b − 1) b = (b − 1) bN −1−ν
ν=1 ν=0

is a different b-adic representation of x and its first coefficient is nonzero.


“(iv) ⇒ (i)”: Assume (iv) holds. Without loss of generality, we can assume that n0 is
the largest index such that dn = b − 1 for each n ≤ n0 . Then
N −n
X 0 −2 ∞
X
dN −ν bN −ν + (dn0 +1 + 1)bn0 +1 + 0 bN −ν
ν=0 ν=N −n0

is a different b-adic representation of x and its first coefficient is nonzero.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 273

We will now show that, conversely, (i) implies (ii), (iii), and (iv). To that end, let x > 0
and suppose that x has two different b-adic representations

X ∞
X
N1 −ν
x= dN1 −ν b = eN2 −ν bN2 −ν (E.19)
ν=0 ν=0

with N1 , N2 ∈ Z; dn , en ∈ {0, . . . , b − 1}; and dN1 , eN2 > 0. This implies

x ≥ bN1 , x ≥ bN2 . (E.20a)

Moreover, Lem. E.1 yields


x ≤ bN1 +1 , x ≤ bN2 +1 . (E.20b)
If N2 > N1 , then (E.20) imply N2 = N1 + 1 and bN2 ≤ x ≤ bN1 +1 = bN2 , i.e. x = bN2 =
bN1 +1 . Since eN2 > 0, one must have eN2 = 1, and, in turn, en = 0 for each n < N2 .
Moreover, x = bN1 +1 and Lem. E.1 imply that dn = b − 1 for each n ∈ {N1 , N1 − 1, . . . }.
Thus, for N2 > N1 , the value of N1 is determined by N2 and the values of all dn and en are
also completely determined, showing that there are precisely two b-adic representations
of x. Moreover, the dn have the property required in (iv) and the en have the property
required in (iii). The argument also shows that, for N1 > N2 , one must have N1 = N2 +1
with the en taking the values of the dn and vice versa. Once again, there are precisely
two b-adic representations of x; now the dn have the property required in (iii) and the
en have the property required in (iv).
It remains to consider the case N := N1 = N2 . Since, by hypothesis, the two b-adic
representations of x in (E.19) are not identical, there must be a largest index n ≤ N
such that dn 6= en . Thus, (E.19) implies

X ∞
X
n−ν
y := dn−ν b = en−ν bn−ν . (E.21)
ν=0 ν=0

Now Lem. E.2 shows that there are precisely two b-adic representations of x, one having
the property required in (iii) and the other having property required in (iv).
Thus, in each case (N2 > N1 , N1 > N2 , and N1 = N2 ), we find that (i) implies (ii), (iii),
and (iv), thereby concluding the proof of the theorem. 

In most cases, it is understood that we work only with decimal representations such
that there is no confusion about the meaning of symbol strings like 101.01. However,
in general, 101.01 could also be meant with respect to any other base, and, the number
represented by the same string of symbols does obviously depend on the base used.
Thus, when working with different representations, one needs some notation to keep
track of the base.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
E SERIES: ADDITIONAL MATERIAL 274

Notation E.4. Given a natural number b ≥ 2 and finite sequences

(dN1 , dN1 −1 , . . . , d0 ) ∈ {0, . . . , b − 1}N1 +1 , (E.22a)


(e1 , e2 , . . . , eN2 ) ∈ {0, . . . , b − 1}N2 , (E.22b)
N3
(p1 , p2 , . . . , pN3 ) ∈ {0, . . . , b − 1} , (E.22c)

N1 , N2 , N3 ∈ N0 (where N2 = 0 or N3 = 0 is supposed to mean that the corresponding


sequence is empty), the respective string

(dN1 dN1 −1 . . .d0 )b for N2 = N3 = 0,


(E.23)
(dN1 dN1 −1 . . .d0 . e1 . . . eN2 p1 . . . pN3 )b for N2 + N3 > 0
represents the number
N1
X N2
X ∞ X
X N3
d ν bν + eν b−ν + pν b−N2 −αN3 −ν . (E.24)
ν=0 ν=1 α=0 ν=1

Example E.5. For the number from (7.95), we get

x = (131.6)10 = (10000011.10)2 = (83.A)16 (E.25)

(for the hexadecimal system, it is customary to use the symbols 0, 1, 2, 3, 4, 5, 6, 7, 8,


9, A, B, C, D, E, F).

One frequently needs to convert representations with respect to one base into represen-
tations with respect to another base. When working with digital computers, conversions
between bases 10 and 2 and vice versa are the most obvious ones that come up. Con-
verting representations is related to the following elementary remainder theorem and
the well-known long division algorithm.
Theorem E.6. For each pair of numbers (a, b) ∈ N2 , there exists a unique pair of
numbers (q, r) ∈ N20 satisfying the two conditions a = qb + r and 0 ≤ r < b.

Proof. Existence: Define

q := max{n ∈ N0 : nb ≤ a}, (E.26a)


r := a − qb. (E.26b)

Then q ∈ N0 by definition and (E.26b) immediately yields a = qb + r as well as r ∈ Z.


Moreover, from (E.26a), qb ≤ a = qb + r, i.e. 0 ≤ r, in particular, r ∈ N0 . Since (E.26a)
also implies (q + 1)b > a = qb + r, we also have b > r as required.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
F CARDINALITY OF R AND SOME RELATED SETS 275

Uniqueness: Suppose (q1 , r1 ) ∈ N0 , satisfying the two conditions a = q1 b + r1 and


0 ≤ r1 < b. Then q1 b = a − r1 ≤ a and (q1 + 1)b = a − r1 + b > a, showing
q1 = max{n ∈ N0 : nb ≤ a} = q. This, in turn, implies r1 = a − q1 b = a − qb = r,
thereby establishing the case. 

F Cardinality of R and Some Related Sets


Theorem F.1. (a) The set of natural numbers N is countable.

(b) The set of integers Z is countable: #Z = #N.

(c) The set of rational numbers Q is countable: #Q = #N.

Proof. (a): The identity Id : N −→ N shows N is countable.


(b): Using (D.24), the map

n/2
 if n is even,
φ : N −→ Z, φ(n) := 0 if n = 1, (F.1)


−(n − 1)/2 if n is odd,

is clearly bijective, proving #Z = #N.


(c): According to (b), Z and Z \ {0} are countable. Then Th. 3.16 implies that A :=
Z × (Z \ {0}) is countable and there is a bijective map f : N −→ A. It is then immediate
from Def. D.22(a) that the map
 
φ : N −→ Q, φ(n) := f (n) , (F.2)

where [f (n)] denotes the equivalence class of f (n) with respect to ∼ from (D.31), is
surjective. Thus, Q is countable by Prop. 3.15. 

In the following theorem and its two corollaries, we will see that the set R of real numbers
is not countable, but has the same cardinality as the power set of N. Moreover, the same
is true for every nontrivial interval of real numbers.

Theorem F.2. Let a, b ∈ R with a < b. Recalling the notations F N, {0, 1} = {0, 1}N
for the set of sequences in {0, 1}, we obtain the following equalities of cardinalities:

#R = #]a, b[= #{0, 1}N = #P(N). (F.3)

Proof. We divide the proof into the following steps:

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
F CARDINALITY OF R AND SOME RELATED SETS 276

(i) #{0, 1}N = #P(N).

(ii) #]0, 1[= #{0, 1}N .

(iii) #] − 1, 1[= #R.

(iv) #]a, b[= #]0, 1[.

(i): To prove #{0, 1}N = #P(N), we have to show the existence of a bijective map
f : {0, 1}N −→ P(N). Given σ ∈ {0, 1}N , i.e. σ is a function σ : N −→ {0, 1}, define

f (σ) := σ −1 {1} = {n ∈ N : σ(n) = 1}. (F.4)

Then, indeed, f : {0, 1}N −→ P(N). It remains to show f is bijective. To verify f is


injective, consider σ, τ ∈ {0, 1}N . If σ 6= τ , then there exists n ∈ N with σ(n) 6= τ (n). If
σ(n) = 1, then τ (n) = 0, i.e. n ∈ f (σ), but n ∈ / f (τ ), showing f (σ) 6= f (τ ). Analogously,
if σ(n) = 0, then τ (n) = 1, i.e. n ∈ f (τ ), but n ∈ / f (σ), again showing f (σ) 6= f (τ ),
concluding the proof that f is injective. To verify f is surjective, for each A ∈ P(N),
define (
1 if n ∈ A,
σA : N −→ {0, 1}, σA (n) := (F.5)
0 if n ∈
/ A.
Then σA ∈ {0, 1}N and f (σA ) = σA−1 {1} = A, proving f is surjective.
(ii): To prove #{0, 1}N = #]0, 1[, we have to show the existence of a bijective map
f : {0, 1}N −→]0, 1[. The map

X

N
g : {0, 1} −→ [0, 1], g (xi )i∈N := xi 2−i , (F.6)
i=1

is well-defined by Lem. E.1 (i.e. 0 ≤ g ≤ 1). Moreover, according to Th. 7.99, g is


surjective, but not injective, as there are numbers x ∈]0, 1[, that have two different dual
(i.e. 2-adic) representations. However, as there are only countably many such numbers,
we can use a modification to obtain our desired f . In preparation, we define, for each
n ∈ N, the sequences en := (eni )i∈N and fn := (fni )i∈N , where
(
1 for i = n,
∀ eni := (F.7a)
n,i∈N 0 for i 6= n,
(
1 for i > n,
∀ fni := (F.7b)
n,i∈N 0 for i ≤ n,

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
F CARDINALITY OF R AND SOME RELATED SETS 277

and we note

g (0, 0, . . . ) = 0, (F.8a)

g (1, 1, . . . ) = 1, (F.8b)
g(en ) = g(fn ) = 2−n for each n ∈ N. (F.8c)

We are now in a position to define




 2−1 if (xi )i∈N = (0, 0, . . . ),

 −2
2 if (xi )i∈N = (1, 1, . . . ),



f : {0, 1}N −→]0, 1[, f (xi )i∈N := 2−(2n+1) if xi = eni for each i ∈ N, (F.9)

2−(2n+2)



 if xi = fni for each i ∈ N,
 P
 ∞ x 2−i
i=1 i otherwise.

Introducing the auxiliary sets

A := {(0, 0, . . . ), (1, 1, . . . )} ∪ {en : n ∈ N} ∪ {fn : n ∈ N}, (F.10a)


B := {2−n : n ∈ N}, (F.10b)

it follows from Th. 7.99 that (the following restrictions of f which, to simplify notation,
we also denote by f )

f : {0, 1}N \ A −→]0, 1[ \B, (F.11a)

and

f : A −→ B (F.11b)

are bijective, i.e. the full f of (F.9) is itself bijective, completing the proof of (ii).
(iii): To prove #] − 1, 1[= #R, we have to show the existence of a bijective map f :
R −→] − 1, 1[. Since we know from Def. and Rem. 8.27 that arctan : R −→] − π/2, π/2[
is bijective, we can define
2 arctan x
f : R −→]0, 1[, f (x) := . (F.12)
π
However, even though this provides a valid proof, arctan is a somewhat complicated
function (as it is defined via sin and cos, which are defined via power series). Thus, it
might be desirable to see an alternative proof, using a more elementary f . We claim
that
x
f : R −→] − 1, 1[, f (x) := , (F.13)
|x| + 1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
F CARDINALITY OF R AND SOME RELATED SETS 278

is also bijective. Since f is clearly continuous, according to the intermediate value Th.
7.57, it suffices to show

∀ ∃ f (x1 ) < −1 + ǫ < 1 − ǫ < f (x2 ). (F.14)


ǫ∈]0,1[ x1 ,x2 ∈R

However, for each ǫ ∈]0, 1[,


−1 + ǫ x1
x1 < = −ǫ−1 + 1 ⇒ x1 < x1 − 1 − ǫ x1 + ǫ ⇒ f (x1 ) = < −1 + ǫ,
ǫ −x1 + 1
1−ǫ x
x2 > = ǫ−1 − 1 ⇒ x2 > 1 + x2 − ǫ − ǫ x2 ⇒ f (x2 ) = > 1 − ǫ,
ǫ x+1
proving (F.14) and the surjectivity of f . To verify f is injective, it suffices to show that
f is strictly increasing. Since
x1 x2
x1 ≤ 0 ≤ x2 ∧ x1 < x2 ⇒ f (x1 ) = ≤0≤ = f (x2 )
−x1 + 1 x2 + 1
∧ f (x1 ) < f (x2 ),
x1 < x2 ≤ 0 ⇒ −x1 x2 + x1 < −x1 x2 + x2
x1 x2
⇒ f (x1 ) = < = f (x2 ),
−x1 + 1 −x2 + 1
0 ≤ x1 < x2 ⇒ x1 x2 + x1 < x1 x2 + x2
x1 x2
⇒ f (x1 ) = < = f (x2 ),
x1 + 1 x2 + 1
showing f is strictly increasing and, hence, injective.
(iv): To prove #]a, b[= #]0, 1[, we have to show the existence of a bijective map f :
]a, b[−→]0, 1[. Such a bijective map is given by the (restriction of an) affine map
x−a
f : ]a, b[−→]0, 1[, f (x) := . (F.15)
b−a
The proof that f is bijective can be conducted analogous to (but much simpler than) the
proof in (iii), or one can use (for example, from Linear Algebra) that every nonconstant
affine map from R into R is bijective. 

Corollary F.3. #R = #P(N) – in particular, R is not countable.

Proof. #R = #P(N) was proved in Th. F.2 and P(N) is uncountable by Th. A.70. 

Corollary F.4. If a, b ∈ R with a < b, then #(Q∩]a, b[) = #N and #(]a, b[ \Q) = #R,
i.e. ]a, b[ contains countably many rational and uncountably many irrational numbers.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
G PARTIAL FRACTION DECOMPOSITION 279

Proof. Since Q∩]a, b[⊆ Q, the claim #(Q∩]a, b[) = #N follows from Th. F.1(c), Prop.
3.14, and Th. 7.68(a).
To prove #(]a, b[ \Q) = #R, a bijection between ]a, b[ \Q and R can be constructed
analogous to the construction of f in step (ii) of the proof of Th. F.2, making use of the
fact that #]a, b[= #R and #Q = #N. 
Theorem F.5. The set of complex numbers C = R × R has the same cardinality as R:
#(R × R) = #R = #P(N).

Proof. Let
A := {0, 1}N . (F.16)
By an application of Th. F.2, it suffices to prove #A = #(A×A), which is accomplished
by showing the existence of a bijective map f : A −→ A × A. We define
 
f : A −→ A × A, f (xj )j∈N := (yj )j∈N , (zj )j∈N , (F.17a)
where
∀ yj := x2j−1 , (F.17b)
j∈N

∀ zj := x2j , (F.17c)
j∈N

and 
g : A × A −→ A, g (yj )j∈N , (zj )j∈N := (xj )j∈N , (F.18a)
where (
y(j+1)/2 for j odd,
∀ xj := (F.18b)
j∈N zj/2 for j even.
Clearly, g = f −1 , proving that f is bijective as desired. 

G Partial Fraction Decomposition


We consider C-valued rational functions of the form
P (z)
z 7→ R(z) := , (G.1)
Q(z)
where P, Q : C −→ C are polynomials such that deg(P ) < deg(Q) =: n. Using Cor.
8.33 as well as Rem. 6.7, we write Q in the form
k
Y
Q(z) = c (z − λj )mj , (G.2)
j=1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
G PARTIAL FRACTION DECOMPOSITION 280

where c ∈ C, and λ1 , . . . , λk ∈ C, k ∈ {1, . . . , n}, are the distinct zeros of Q, mj ∈ N


P
with kj=1 mj = n being their respective multiplicities.
In can be useful to write R as a linear combination of the so-called partial fractions
1 1 1
, , ..., (j = 1, . . . , k) (G.3)
z − λj (z − λj ) 2 (z − λj )mj

(for example, for the computation of the antiderivative of R, cf. Ex. 10.22(b)). The
following Th. G.1 guarantees this is always possible:

Theorem G.1. Let P, Q : C −→ C be polynomials such that deg(P ) < deg(Q) =: n.


Moreover, let N (Q) denote the set of zeros of Q, and assume Q to have the form of
(G.2). Then there exists a unique family of coefficients

ajl ∈ C (j = 1, . . . , k, l = 1, . . . , mj ), (G.4)

such that
k mj
P (z) X X ajl
∀ R(z) = = . (G.5)
z∈C\N (Q) Q(z) j=1 l=1
(z − λj )l

Proof. We first prove the existence of the decomposition (G.5) via induction on n =
deg(Q). If n = 1, then P must be constant and there is nothing to prove. For the
induction step, consider n ≥ 2. Let ζ be a zero of Q with multiplicity m ∈ {1, . . . , n}.
Then, according to Rem. 6.7, there exists a polynomial S : C −→ C such that Q(z) =
(z − ζ)m S(z) and S(ζ) 6= 0. Noting

P (z) P (ζ) P (z)S(ζ) − S(z)P (ζ)


R̃(z) := − = (G.6)
S(z) S(ζ) S(z)S(ζ)

and that P (z)S(ζ)−S(z)P (ζ) vanishes for z = ζ, there exists a polynomial T : C −→ C,


deg T ≤ n − 2, such that
(z − ζ)T (z)
R̃(z) = . (G.7)
S(z)
Thus, for each z ∈ C \ N (Q), we have

P (ζ) R̃(z) (G.7) T (z)


R(z) − = = . (G.8)
(z − ζ)m S(ζ) (z − ζ)m (z − ζ)m−1 S(z)

We will now
m−1
 apply (G.8) with ζ = λk and m = mk . Since deg(T ) < n − 1 = deg (z −
ζ) S(z) < deg(Q), the induction hypothesis applies to the function in (G.8), yielding

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
G PARTIAL FRACTION DECOMPOSITION 281

coefficients ajl ∈ C, j = 1, . . . , k, l = 1, . . . , mj for j < k, l = 1, . . . , mj − 1 for j = k,


satisfying
mj
k−1 X m k −1
P (λk ) X ajl X akl
R(z) − m
= l
+ , (G.9)
(z − λk ) S(λk )
k
j=1 l=1
(z − λj ) l=1
(z − λk )l

thereby completing the induction for the existence proof.


It remains to prove the uniqueness of the coefficients ajl in (G.5). Thus, suppose one
has bjl ∈ C, j = 1, . . . , k, l = 1, . . . , mj , such that
mj
k X k mj
X ajl XX bjl
= . (G.10)
j=1 l=1
(z − λj )l j=1 l=1
(z − λ j ) l

We fix j0 and prove aj0 l = bj0 l via induction on l = 1, . . . , mj0 : Let l ∈ {1, . . . , mj0 } and
assume aj0 α = bj0 α has already been shown for each α > l (the induction does, indeed,
start at l = mj0 , working itself down to l = 1). Then (G.10) implies
l k mj l k mj
X aj 0 β XX ajβ X bj 0 β XX bjβ
β
+ β
= β
+ . (G.11)
β=1
(z − λj0 ) j=1, β=1
(z − λj ) β=1
(z − λj0 ) j=1, β=1
(z − λj )β
j6=j0 j6=j0

One now multiplies (G.11) by (z − λj0 )l . Then taking the limit for z → λj0 on both
sides yields aj0 l = bj0 l as desired. 

If, in (G.1), P, Q : R −→ R, then the partial fraction decomposition (G.5) of Th. G.1 is
not quite satisfactory, since, even though P and Q are both real, the ajl will typically
be nonreal elements of C. As the following Th. G.2 shows, if P, Q are real, then it is
always possible to obtain a partial fraction decomposition with only real coefficients,
however its form is somewhat more complicated.
We start by using the real factorization of Q : R −→ R, deg(Q) = n ∈ N, according to
(8.58), where, as in (G.2), we combine identical factors, obtaining
k1
Y k2
Y
mj
Q(x) = c (x − λj ) (x2 + αj x + βj )nj , (G.12)
j=1 j=1

where c ∈ R; λ1 , . . . , λk1 ∈ R, k1 ∈ {0, . . . , n}, are the distinct real zeros of Q (if
any), mj ∈ N being their respective multiplicities; and (α1 , β1 ), . . . , (αk2 , βk2 ) ∈ R2
k2 ∈ {0, . . . , n}, are distinct pairs of real numbers, each pair arising from combining

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
G PARTIAL FRACTION DECOMPOSITION 282

two conjugate nonreal zeros of Q according to (8.60), nj ∈ N being their respective


multiplicities;
Xk1 k2
X
mj + 2 nj = n. (G.13)
j=1 j=1

Theorem G.2. Let P, Q : R −→ R be polynomials such that deg(P ) < deg(Q) =: n.


Moreover, let N (Q) denote the set of zeros of Q, and assume Q to have the form of
(G.12). Then there exist families of coefficients

ajl ∈ R (j = 1, . . . , k1 , l = 1, . . . , mj ), (G.14a)
bjl ∈ R (j = 1, . . . , k2 , l = 1, . . . , nj ), (G.14b)
cjl ∈ R (j = 1, . . . , k2 , l = 1, . . . , nj ) (G.14c)

such that
1 X k mj 2 X k nj
P (x) X ajl X bjl x + cjl
∀ R(x) = = l
+ . (G.15)
x∈R\N (Q) Q(x) j=1 l=1
(x − λj ) j=1 l=1
(x + αj x + βj )l
2

Proof. We show that, if P, Q are real, where Q is as in (G.12), then (G.5) can be
rewritten in the form (G.15): First, consider λj0 ∈ R to be a real zero of Q. Then
all corresponding coefficients in (G.5) are real: We prove aj0 l ∈ R via induction on
l = 1, . . . , mj0 : Let l ∈ {1, . . . , mj0 } and assume aj0 α ∈ R has already been shown for
each α > l. Then (G.5) (with z replaced by x) implies
mj 0 l k mj
aj 0 β
X X aj 0 β XX ajβ
∀ S(x) := R(x) − β
= β
+ ∈ R.
x∈R\N (Q)
β=l+1
(x − λj0 ) β=1
(x − λj0 ) j=1, β=1
(x − λj )β
j6=j0
(G.16)
One now multiplies (G.16) by (x − λj0 )l . Then taking the limit for x → λj0 on both
sides yields aj0 l ∈ R as desired (the limit on the right-hand side is clearly aj0 l and all
values and, thus, the limit on the left-hand side are clearly in R).
Thus, the summands corresponding to real zeros of Q are identical in (G.5) and (G.15).
It remains to show that terms in (G.5), corresponding to conjugate nonreal zeros of
Q, can be combined to result in the summands involving the bjl and cjl in (G.15). To
this end, consider λj0 , λj1 ∈ C to be conjugate nonreal zeros of Q, λj1 = λj0 . Then all
corresponding coefficients in (G.5) are conjugate: We prove aj0 l = aj1 l via induction on
l = 1, . . . , mj0 = mj1 : Let l ∈ {1, . . . , mj0 } and assume aj0 α = aj1 α has already been
shown for each α > l. We once again have the formula (G.16) for S(x) (even for each

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
G PARTIAL FRACTION DECOMPOSITION 283

x ∈ C \ N (Q), but we can no longer expect S(x) ∈ R). As before, after multiplying
(G.16) by (x − λj0 )l , we obtain

lim S(x)(x − λj0 )l = aj0 l . (G.17)
x→λj0

Analogously, we also have


mj 1 l k mj
X aj 1 β X aj 1 β XX ajβ
∀ R(x) − β
= β
+ β
. (G.18)
x∈C\N (Q)
β=l+1
(x − λ j1 )
β=1
(x − λ j1 ) j=1, β=1
(x − λ j )
j6=j1

Taking complex conjugates in (G.16) and using the induction hypothesis as well as
R(x) = R(x) (since the coefficients of P, Q are real) yields
mj 1 l k mj
aj 1 β
X (G.18) X aj 1 β XX ajβ
∀ S(x) = R(x) − = + .
x∈C\N (Q)
β=l+1
(x − λj1 )β β=1
(x − λj1 )β j=1, β=1 (x − λj )β
j6=j1
(G.19)
l
If we multiply (G.19) by (x − λj1 ) , we obtain

lim S(x)(x − λj1 )l = aj1 l . (G.20)
x→λj1

Thus,
(G.17)   (G.20)
aj 0 l = lim S(x)(x − λj0 )l = lim S(x)(x − λj1 )l = aj 1 l , (G.21)
x→λj0 x→λj0

as needed.
We now combine two corresponding summands of (G.5) (for x ∈ R \ N (Q)):

aj 0 l aj 0 l aj0 l (x − λj0 )l + aj0 l (x − λj0 )l a(x − λ)l + a(x − λ)l


σl := + = = ,
(x − λj0 )l (x − λj0 )l (x2 − 2x Re λj0 + |λj0 |2 )l (x2 + bx + c)l
(G.22)
where we have set

a := aj0 l , λ := λj0 , b := −2 Re λj0 , c := |λj0 |2 , (G.23)

to simplify notation. To finish the proof of (G.15), it remains to show there are real
coefficients s1l , . . . , sll and t1l , . . . , tll such that
l
X sβl x + tβl
∀ σl = , (G.24)
l∈{1,...,mj0 }
β=1
(x2 + bx + c)β

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
G PARTIAL FRACTION DECOMPOSITION 284

which we prove via induction on l: For l = 1, we have


a(x − λ) + a(x − λ) (a + a)x − (aλ + aλ)
σ1 =
2
= , (G.25)
x + bx + c x2 + bx + c
which fits the requirements of (G.24). For the induction step, we consider, for l =
1, . . . , mj0 − 1,
a(x − λ)l+1 + a(x − λ)l+1
σl+1 = . (G.26)
(x2 + bx + c)l+1
The numerator can be rewritten as
a(x − λ)l+1 + a(x − λ)l+1 = a(x − λ)l+1 + a(x − λ)l (x − λ)
− a(x − λ)l (x − λ) − a(x − λ)l (x − λ)
+ a(x − λ)l (x − λ) + a(x − λ)l+1 . (G.27)
Thus, σl+1 = S1 + S2 + S3 , where

a(x − λ)l + a(x − λ)l (x − λ) σl (x − λ)
S1 = 2 l+1
= 2 , (G.28)
(x + bx + c) x + bx + c

a(x − λ)l−1 + a(x − λ)l−1 (x − λ)(x − λ) σl−1
S2 = − 2 l+1
= 2 , (G.29)
(x + bx + c) x + bx + c

a(x − λ)l + a(x − λ)l (x − λ) σl (x − λ)
S3 = = , (G.30)
(x2 + bx + c)l+1 x2 + bx + c
where, for (G.29) to hold for l = 1, we set σ0 := a + a ∈ R. Using the induction
hypothesis, S2 clearly has the form required by (G.24). Using the induction hypothesis
together with the elementary equality
sx2 sbx + sc
2
=s− 2 , (G.31)
x + bx + c x + bx + c
we also obtain
l
σl (2x + b) 2x + b X sβl x + tβl
S1 + S3 = =
x2 + bx + c x2 + bx + c β=1 (x2 + bx + c)β
l
X 2sβl x2 + (bsβl + 2tβl )x + btβl
=
β=1
(x2 + bx + c)β (x2 + bx + c)
l  
(G.31) X 1 2sβl (bx + c)
= 2sβl − 2
β=1
(x2 + bx + c)β x + bx + c
l
X (bsβl + 2tβl )x + btβl
+ (G.32)
β=1
(x2 + bx + c)β+1

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
H IRRATIONALITY OF e AND π 285

to have the form required by (G.24), thereby finishing the induction and the proof of
the theorem. 
Remark G.3. Given a rational function R = P/Q as in Th. G.1 (or Th. G.2), there
remains the question of how to actually compute the coefficients ajl of the partial fraction
decomposition (G.5) (or ajl , bjl , cjl of the partial fraction decomposition (G.15) in the real
case)? First, one always needs to obtain the zeros λj and their respective multiplicities
mj , which, for deg(Q) large, can be very difficult. Then there are basically three different
possibilities to proceed, where, in practice, the most efficient way in a concrete situation
might be to mix the three strategies:

(a) Linear System: To determine k unknown coefficients, one can plug k different values
for z into (G.5) (or for x into (G.15)) to obtain a linear system for the unknown
coefficients.
(b) One can multiply (G.5) (or (G.15)) by Q, obtaining a polynomial on both sides of
the equation. As the polynomials need to be equal, the coefficients of equal powers
need to be equal on both sides, yielding a system of equations for the unknown
coefficients.
(c) Multiplying (G.5) (or (G.15)) by (z − λj )mj and setting z = λj yields the coefficient
of (z−λ1j )mj , etc.

H Irrationality of e and π

H.1 Irrationality of e
The following Prop. H.1, which will then be used to prove the irrationality of e in
Th. H.2, shows, in particular, that the series (8.26) can be used to efficiently compute
accurate approximations of e.
Proposition H.1. Defining
n−1 j
z
X z
∀ ∀ Rn (z) := e − , (H.1)
n∈N z∈C
j=0
j!

we have  n

Rn (z) ≤ 2 |z|

∀ |z| ≤ 1 ⇒ , (H.2)
n∈N n!
i.e. the error made when approximating ez by the partial sum (for |z| ≤ 1) is at most as
large as twice the modulus of the first missing summand.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
H IRRATIONALITY OF e AND π 286

Proof. One estimates, for each n ∈ N and each z ∈ C with |z| ≤ 1,


∞  
(8.24),(7.81) X |z|j (7.73) |z|n |z| |z|2
Rn (z) ≤ = 1+ + + ...
j=n
j! n! n + 1 (n + 1)(n + 2)
 
|z|≤1 |z|n 1 1 (7.71) 2 |z|
n
≤ 1 + + 2 + ... = , (H.3)
n! 2 2 n!
which establishes the case. 
Theorem H.2. Euler’s number e is irrational.

Proof. Seeking a contradiction, we assume e to be rational. Then there exist m, n ∈ N


with n ≥ 2 such that e = mn
. Then n!e ∈ N and, thus,
n
(H.1) X 1
n! Rn+1 (1) = n! e − n! ∈ Z, (H.4)
j=0
j!

2
in contradiction to 0 < |n! Rn+1 (1)| < n+1
< 1, which holds according to (H.2) (recalling
n ≥ 2). 

H.2 Irrationality of π
Theorem H.3. π 2 is irrational (then, in particular, π must be irrational as well).

Proof. Seeking a contradiction, we assume π 2 to be rational. Then


a
∃ π2 = . (H.5)
a,b∈N b
We can then choose some even n ∈ N satisfying
π an
0< < 1. (H.6)
n!
We now consider the function
2n  
xn (1 − x)n (∗) 1 X k n
f : R −→ R, f (x) := = (−1) xk , (H.7)
n! n! k=n k−n

where the equality at (∗) is proved by


n   2n  
xn (1 − x)n (5.22) xn X k n k n even 1
X
k n
= (−1) x = (−1) xk . (H.8)
n! n! k=0 k n! k=n k−n

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
H IRRATIONALITY OF e AND π 287

Thus, for the polynomial f , we obtain the derivatives



0
 for 0 ≤ j < n,
(j) j! j n

f (0) = n! (−1) j−n for n ≤ j ≤ 2n, (H.9)


0 for 2n < j.
j! n

In consequence, since, for n ≤ j ≤ 2n, n! ∈ N and j−n ∈ N,

∀ f (j) (0) ∈ Z. (H.10)


j∈N0

Moreover, since f (1 − x) = f (x) for each x ∈ R, and, thus, f (j) (1 − x) = (−1)j f (j) (x)
for each x ∈ R, we also have
∀ f (j) (1) ∈ Z. (H.11)
j∈N0

Next, we consider another polynomial, namely


n
X
n
g : R −→ R, g(x) := b (−1)k π 2(n−k) f (2k) (x). (H.12)
k=0

Due to (H.5), (H.10), (H.11), and (H.12), we have


 
∀ g(0) ∈ Z ∧ g(1) ∈ Z . (H.13)
j∈N0

For each x ∈ R, one calculates


n
X n
X
′′ 2 n k 2(n−k) (2(k+1)) n
g (x) + π g(x) = b (−1) π f (x) + b (−1)k π 2(n−(k−1)) f (2k) (x)
k=0 k=0
n+1
X n
X
n k−1 2(n−(k−1)) (2k) n
=b (−1) π f (x) + b (−1)k π 2(n−(k−1)) f (2k) (x)
k=1 k=0

= bn (−1)n f (2n+2) (x) + bn π 2n+2 f (x) = bn π 2n+2 f (x), (H.14)

and, thus, for

h : R −→ R, h(x) := g ′ (x) sin(πx) − πg(x) cos(πx), (H.15)

one obtains, for each x ∈ R,

h′ (x) = g ′′ (x) sin(πx) + πg ′ (x) cos(πx) − πg ′ (x) cos(πx) + π 2 g(x) sin(πx)


 (H.14)
= g ′′ (x) + π 2 g(x) sin(πx) = bn π 2n+2 f (x) sin(πx)
(H.5)
= π 2 an f (x) sin(πx), (H.16)

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
I TRIGONOMETRIC FUNCTIONS 288

implying the function h is the antiderivative of the function x 7→ π 2 an f (x) sin(πx).


This, together with the fundamental theorem of calculus in the form Th. 10.20(b) implies
Z
π 2 an 1 h(1) − h(0) πg(1) + πg(0)
I := f (x) sin(πx) dx = = = g(1) + g(0) ∈ Z.
π 0 π π
(H.17)
On the other hand, the definition of f in (H.7) yields
1
∀ 0 < f (x) < , (H.18)
0<x<1 n!
and, thus, by (10.30) (i.e. by the monotonicity of the integral),
π an (H.6)
0<I< < 1. (H.19)
n!
The contradiction between (H.19) and (H.17) establishes the case. 

I Trigonometric Functions

I.1 Additional Trigonometric Formulas


Proposition I.1. We have the following identities:

∀ sin(2z) = 2 sin z cos z, (I.1a)


z∈C
∀ cos(2z) = (cos z)2 − (sin z)2 , (I.1b)
z∈C
1 − cos z  z 2
∀ = sin , (I.1c)
z∈C 2 2
z sin z
∀ tan = . (I.1d)
z∈C\{(2k+1)π: k∈Z} 2 cos z + 1
1 − (tan z2 )2
∀ cos z = . (I.1e)
z∈C\{(2k+1)π: k∈Z} 1 + (tan z2 )2

Proof. (I.1a) is immediate from (8.44c), (I.1b) is immediate from (8.44d).


(I.1c): For each z ∈ C, one computes

1 − cos z (I.1b) 1 − (cos z2 )2 + (sin z2 )2 (8.44e) 2 (sin z2 )2  z 2


= = = sin , (I.2)
2 2 2 2
thereby establishing the case.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
J DIFFERENTIAL CALCULUS 289

(I.1d): Note that, according to (8.47d), it is


z
cos =0 ⇔ ∃ z = (2k + 1) π. (I.3)
2 k∈Z

Thus, for each z ∈ C \ {(2k + 1)π : k ∈ Z}, one computes

z 2 sin z2 cos z2 (I.1a),(8.44e) sin z sin z


tan = = = , (I.4)
2 2 (cos z2 )2 (cos z 2
2
) z 2
− (sin 2 ) + 1 cos z + 1

thereby establishing the case.


(I.1e): Once again, using (I.3), one computes for each z ∈ C \ {(2k + 1)π : k ∈ Z}:

(I.1b),(8.44e) (cos z2 )2 − (sin z2 )2 1 − (tan z2 )2


cos z = = , (I.5)
(cos z2 )2 + (sin z2 )2 1 + (tan z2 )2

as claimed. 

J Differential Calculus

J.1 Continuous, But Nowhere Differentiable Functions


The following Ex. J.1 provides functions from f : R −→ R that are continuous, but
nowhere differentiable.

Example J.1. We start by defining the triangle wave function


(
x−k for k ≤ x ≤ k + 21 , k ∈ Z,
g : R −→ R, g(x) := (J.1)
−x + k + 1 for k + 21 ≤ x ≤ k + 1, k ∈ Z.

Then g is well-defined and continuous, since, clearly, g is piecewise affine, k + 21 − k =


1
2
= −k − 21 + k + 1, and −(k + 1) + k + 1 = 0 = k + 1 − (k + 1). Moreover, for each
k ∈ Z, g is clearly strictly increasing on [k, k + 12 ] and clearly strictly decreasing on
[k + 12 , k + 1], implying
1
∀ 0 ≤ g(x) ≤ . (J.2)
x∈R 2
Clearly, (J.1) implies g to be periodic with period 1, i.e.

∀ g(x + 1) = g(x). (J.3)


x∈R

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
J DIFFERENTIAL CALCULUS 290

Now fix q ∈ R, a ∈ N such that


0<q<1 ∧ a≥4 ∧ aq > 2 (J.4)
(clearly, q = 12 and a = 5 satisfy (J.4), and there are (uncountably) many other admis-
sible choices for a and q). We now claim that

X
f : R −→ R, f (x) := q n g(an x), (J.5)
n=0
P∞
is continuous and nowhere differentiable. We first note that, as n=0 q n converges and
(J.2) (J.2) qn
∀ ∀ 0 ≤ q n g(an x) ≤ < qn, (J.6)
n∈N x∈R 2
Cor. 8.7(b) implies the series in (J.5) to converge uniformly. Then, since each function
fn : R −→ R, fn (x) := q n g(an x), is continuous, f must be continuous by Cor. 8.7(c).
In preparation for showing f to be nowhere differentiable, we have to further investigate
the properties of g. We proceed by showing g to be Lipschitz continuous with Lipschitz
constant 1, i.e.
∀ |g(x) − g(y)| ≤ |x − y| : (J.7)
x,y∈R
1
If |x − y| ≥ 2
, then, using (J.2),
1
|g(x) − g(y)| ≤ ≤ |x − y|.
2
If |x − y| < 12 , then we distinguish four cases, where, without loss of generality, we let
x denote the smaller of the two points and y the larger, i.e. x ≤ y. Case (i): There is
k ∈ Z such that k ≤ x, y ≤ k + 12 . Then
|g(x) − g(y)| = |x − k − y + k| = |x − y|.
1
Case (ii): There is k ∈ Z such that k + 2
≤ x, y ≤ k + 1. Then
|g(x) − g(y)| = | − x + k + 1 + y − k − 1| = |x − y|.
Case (iii): There is k ∈ Z such that k ≤ x ≤ k + 12 and k + 12 ≤ y ≤ k + 1. Then
x − k − 12 ≤ 0 and y − k − 12 ≥ 0, implying

1 1
|g(x) − g(y)| = |x − k + y − k − 1| = x − k − + + y − k − 1
2 2

1 1
≤ x − k − − y + k + = |x − y|.
2 2

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
J DIFFERENTIAL CALCULUS 291

Case (iv): There is k ∈ Z such that k − 21 ≤ x ≤ k and k ≤ y ≤ k + 21 . Then −x + k ≥ 0


and −y + k ≤ 0, implying
|g(x) − g(y)| = | − x + k − y + k| ≤ | − x + k + y − k| = |x − y|,
finishing the proof of (J.7).
For each c ∈ R, we also consider the following modified versions of g:
(
cx − k for k ≤ cx ≤ k + 12 , k ∈ Z,
gc : R −→ R, gc (x) := g(cx) = (J.8)
−cx + k + 1 for k + 12 ≤ cx ≤ k + 1, k ∈ Z.

Then, for c 6= 0, gc is periodic with period c−1 :


(J.3)
∀ gc (x + c−1 ) = g(cx + 1) = g(cx) = gc (x). (J.9)
x∈R

Moreover, gc is Lipschitz continuous with Lipschitz constant |c|:


(J.7)
∀ |gc (x) − gc (y)| = |g(cx) − g(cy)| ≤ |cx − cy| = |c| |x − y|. (J.10)
x,y∈R

To show that f is nowhere differentiable, we will now study suitable difference quotients.
Let (hk )k∈N be a sequence such that
1
∀ hk = ± . (J.11)
k∈N ak+1
Then (J.4) implies limk→∞ hk = 0. Let x ∈ R be arbitrary. Define
g(an (x + hk )) − g(an x)
∀ δk g(an x) := . (J.12)
k,n∈N hk
Then
(J.10) an |hk |
∀ |δk g(an x)| ≤ = an , (J.13)
k,n∈N |hk |
and, recalling a ∈ N,
1 an−(k+1)
n gan (x ± ak+1
) − gan (x) gan (x ± an
) − gan (x) (J.9)
∀ δk g(a x) = = = 0. (J.14)
n>k∈N hk hk
Thus, for each k ∈ N, we obtain
P∞ n n n

f (x + hk ) − f (x) q g(a (x + h k )) − g(a x)
δk f := = n=0
hk hk
k−1
X
(J.14)
= q n δk g(an x) + q k δk g(ak x) (J.15)
n=0

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
REFERENCES 292

and estimate, recalling aq > 2,


k−1 k−1 k−1
X
n n
X
n n
(J.13) X 1 − q k ak q k ak
q n an =

q δk g(a x) ≤ q |δk g(a x)| ≤ < . (J.16)
n=0 n=0 n=0
1 − qa qa − 1

We rewrite (J.16) as

Xk−1
n n 1
< η q k ak ,

q δ k g(a x) where η := , 0 < η < 1. (J.17)

n=0
aq − 1
1
According to (J.8), gak is affine on intervals of length 2ak
. Thus, since a ≥ 4 implies
1 1
∀ ≤ , (J.18)
k∈N ak+1 4ak
we can always choose the sign of hk such that
|ak (x + hk − x)|
|δk g(ak x)| = = ak . (J.19)
|hk |
Using this choice for the hk , we combine our estimates to obtain
k−1
(J.17),(J.19)
(J.15) X n n k k
∀ |δk f | = q δk g(a x) + q δk g(a x) ≥ (1 − η) q k ak . (J.20)
k∈N n=0

Thus, as qa > 2, limk→∞ |δk f | = ∞, proving that f is not differentiable in x. As x ∈ R


was arbitrary, f is nowhere differentiable.

References
[Bla84] A. Blass. Existence of Bases Implies the Axiom of Choice. Contemporary
Mathematics 31 (1984), 31–33.
[EFT07] H.-D. Ebbinghaus, J. Flum, and W. Thomas. Einführung in die math-
ematische Logik, 5th ed. Spektrum Akademischer Verlag, Heidelberg, 2007
(German).
[EHH+ 95] H.-D. Ebbinghaus, H. Hermes, F. Hirzebruch, M. Koecher,
K. Mainzer, J. Neukirch, A. Prestel, and R. Remmert. Numbers.
Graduate Texts in Mathematics, Vol. 123, Springer-Verlag, New York, 1995,
corrected 3rd printing.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34
REFERENCES 293

[Hal17] Lorenz J. Halbeisen. Combinatorial Set Theory, 2nd ed. Springer Mono-
graphs in Mathematics, Springer, Cham, Switzerland, 2017.

[Jec73] T. Jech. The Axiom of Choice. North-Holland, Amsterdam, 1973.

[Kön04] Konrad Königsberger. Analysis 1, 6th ed. Springer-Verlag, Berlin, 2004


(German).

[Kun80] Kenneth Kunen. Set Theory. Studies in Logic and the Foundations of
Mathematics, Vol. 102, North-Holland, Amsterdam, 1980.

[Kun12] Kenneth Kunen. The Foundations of Mathematics. Studies in Logic,


Vol. 19, College Publications, London, 2012.

[Phi17] P. Philip. Analysis III: Measure and Integration Theory of Several Vari-
ables. Lecture Notes, LMU Munich, 2016/2017, AMS Open Math Notes Ref.
# OMN:202109.111308, available in PDF format at
https://www.ams.org/open-math-notes/omn-view-listing?listingId=111308.

[Phi19] P. Philip. Linear Algebra I. Lecture Notes, Ludwig-Maximilians-Universi-


tät, Germany, 2018/2019, available in PDF format at
http://www.math.lmu.de/~philip/publications/lectureNotes/philipPeter_LinearAlgebra1.pdf.

[Wal02] Wolfgang Walter. Analysis 2, 5th ed. Springer-Verlag, Berlin, 2002 (Ger-
man).

[Wal04] Wolfgang Walter. Analysis 1, 7th ed. Springer-Verlag, Berlin, 2004 (Ger-
man).

[Wik15a] Wikipedia. Coq — Wikipedia, The Free Encyclopedia. 2015, https://en.


wikipedia.org/wiki/Coq Online; accessed Sep-01-2015.

[Wik15b] Wikipedia. HOL Light — Wikipedia, The Free Encyclopedia. 2015, https:
//en.wikipedia.org/wiki/HOL_Light Online; accessed Sep-01-2015.

[Wik15c] Wikipedia. Isabelle (proof assistant) — Wikipedia, The Free En-


cyclopedia. 2015, https://en.wikipedia.org/wiki/Isabelle_(proof_
assistant) Online; accessed Sep-01-2015.

AMS Open Math Notes: Works in Progress; Reference # OMN:202109.111306; Last Revised: 2021-09-17 09:47:34

You might also like