Uploaded by Dave Lacquio

A first course in mathematical logic and set theory ( PDFDrive )

advertisement
A FIRST COURSE IN
MATHEMATICAL LOGIC
AND SET THEORY
A FIRST COURSE IN
MATHEMATICAL LOGIC
AND SET THEORY
Michael L. O’Leary
College of DuPage
c
Copyright 2016
by John Wiley & Sons, Inc. All rights reserved
Published by John Wiley & Sons, Inc., Hoboken, New Jersey
Published simultaneously in Canada
No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form
or by any means, electronic, mechanical, photocopying, recording, scanning, or otherwise, except as
permitted under Section 107 or 108 of the 1976 United States Copyright Act, without either the prior
written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to
the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, (978) 750-8400,
fax (978) 750-4470, or on the web at www.copyright.com. Requests to the Publisher for permission should
be addressed to the Permissions Department, John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ
07030, (201) 748-6011, fax (201) 748-6008, or online at http://www.wiley.com/go/permission.
Limit of Liability/Disclaimer of Warranty: While the publisher and author have used their best efforts in
preparing this book, they make no representations or warranties with respect to the accuracy or
completeness of the contents of this book and specifically disclaim any implied warranties of
merchantability or fitness for a particular purpose. No warranty may be created or extended by sales
representatives or written sales materials. The advice and strategies contained herein may not be
suitable for your situation. You should consult with a professional where appropriate. Neither the
publisher nor author shall be liable for any loss of profit or any other commercial damages, including
but not limited to special, incidental, consequential, or other damages.
For general information on our other products and services or for technical support, please contact our Customer Care
Department within the United States at (800) 762-2974, outside the United States at (317) 572-3993 or fax (317) 5724002.
Wiley also publishes its books in a variety of electronic formats. Some content that appears in print
may not be available in electronic formats. For more information about Wiley products, visit our web site at www.wiley.com.
Library of Congress Cataloging-in-Publication Data applied for.
ISBN: 9780470905883
Printed in the United States of America
10 9 8 7 6 5 4 3 2 1
For my parents
CONTENTS
Preface
xiii
Acknowledgments
List of Symbols
1
Propositional Logic
1.1
1.2
1.3
Symbolic Logic
Propositions
Propositional Forms
Interpreting Propositional Forms
Valuations and Truth Tables
Inference
Semantics
Syntactics
Replacement
Semantics
Syntactics
xv
xvii
1
1
2
5
7
10
19
21
23
31
31
34
vii
viii
CONTENTS
1.4
1.5
2
First-Order Logic
2.1
2.2
2.3
2.4
3
Proof Methods
Deduction Theorem
Direct Proof
Indirect Proof
The Three Properties
Consistency
Soundness
Completeness
Languages
Predicates
Alphabets
Terms
Formulas
Substitution
Terms
Free Variables
Formulas
Syntactics
Quantifier Negation
Proofs with Universal Formulas
Proofs with Existential Formulas
Proof Methods
Universal Proofs
Existential Proofs
Multiple Quantifiers
Counterexamples
Direct Proof
Existence and Uniqueness
Indirect Proof
Biconditional Proof
Proof of Disunctions
Proof by Cases
40
40
44
47
51
51
55
58
63
63
63
67
70
71
75
75
76
78
85
85
87
90
96
97
99
100
102
103
104
105
107
111
112
Set Theory
117
3.1
117
118
Sets and Elements
Rosters
CONTENTS
3.2
3.3
3.4
4
Famous Sets
Abstraction
Set Operations
Union and Intersection
Set Difference
Cartesian Products
Order of Operations
Sets within Sets
Subsets
Equality
Families of Sets
Power Set
Union and Intersection
Disjoint and Pairwise Disjoint
ix
119
121
126
126
127
130
132
135
135
137
148
151
151
155
Relations and Functions
161
4.1
161
163
165
168
171
172
177
180
181
183
189
194
195
196
197
203
205
208
211
212
216
4.2
4.3
4.4
4.5
4.6
Relations
Composition
Inverses
Equivalence Relations
Equivalence Classes
Partitions
Partial Orders
Bounds
Comparable and Compatible Elements
Well-Ordered Sets
Functions
Equality
Composition
Restrictions and Extensions
Binary Operations
Injections and Surjections
Injections
Surjections
Bijections
Order Isomorphims
Images and Inverse Images
x
5
CONTENTS
Axiomatic Set Theory
225
5.1
225
226
227
228
229
230
234
237
239
242
243
249
250
253
256
257
260
264
268
268
271
274
275
278
280
5.2
5.3
5.4
5.5
5.6
6
Axioms
Equality Axioms
Existence and Uniqueness Axioms
Construction Axioms
Replacement Axioms
Axiom of Choice
Axiom of Regularity
Natural Numbers
Order
Recursion
Arithmetic
Integers and Rational Numbers
Integers
Rational Numbers
Actual Numbers
Mathematical Induction
Combinatorics
Euclid’s Lemma
Strong Induction
Fibonacci Sequence
Unique Factorization
Real Numbers
Dedekind Cuts
Arithmetic
Complex Numbers
Ordinals and Cardinals
283
6.1
283
286
290
292
293
298
300
303
307
308
6.2
6.3
Ordinal Numbers
Ordinals
Classification
Burali-Forti and Hartogs
Transfinite Recursion
Equinumerosity
Order
Diagonalization
Cardinal Numbers
Finite Sets
CONTENTS
6.4
6.5
7
Countable Sets
Alephs
Arithmetic
Ordinals
Cardinals
Large Cardinals
Regular and Singular Cardinals
Inaccessible Cardinals
xi
310
313
316
316
322
327
328
331
Models
333
7.1
333
335
340
346
348
353
361
363
366
368
374
380
384
388
394
394
397
399
409
410
414
415
417
7.2
7.3
7.4
7.5
First-Order Semantics
Satisfaction
Groups
Consequence
Coincidence
Rings
Substructures
Subgroups
Subrings
Ideals
Homomorphisms
Isomorphisms
Elementary Equivalence
Elementary Substructures
The Three Properties Revisited
Consistency
Soundness
Completeness
Models of Different Cardinalities
Peano Arithmetic
Compactness Theorem
Löwenheim–Skolem Theorems
The von Neumann Hierarchy
Appendix: Alphabets
427
References
429
Index
435
PREFACE
This book is inspired by The Structure of Proof: With Logic and Set Theory published
by Prentice Hall in 2002. My motivation for that text was to use symbolic logic as a
means by which to learn how to write proofs. The purpose of this book is to present
mathematical logic and set theory to prepare the reader for more advanced courses that
deal with these subjects either directly or indirectly. It does this by starting with propositional logic and first-order logic with sections dedicated to the connection of logic
to proof-writing. Building on this, set theory is developed using first-order formulas.
Set operations, subsets, equality, and families of sets are covered followed by relations
and functions. The axioms of set theory are introduced next, and then sets of numbers are constructed. Finite numbers, such as the natural numbers and the integers, are
defined first. All of these numbers are actually sets constructed so that they resemble
the numbers that are their namesakes. Then, the infinite ordinal and cardinal numbers
appear. The last chapter of the book is an introduction to model theory, which includes
applications to abstract algebra and the proofs of the completeness and compactness
theorems. The text concludes with a note on Gödel’s incompleteness theorems.
MICHAEL L. O’LEARY
Glen Ellyn, Illinois
July 2015
xiii
ACKNOWLEDGMENTS
Thanks are due to Susanne Steitz-Filler, Senior Editor at Wiley, for her support of this
project. Thanks are also due to Sari Friedman and Katrina Maceda, both at Wiley, for
their work in producing this book. Lastly, I wish to thank the anonymous reviewer
whose comments proved beneficial.
On a personal note, I would like to express my gratitude to my parents for their
continued caring and support; to my brother and his wife, who will make sure my
niece learns her math; to my dissertation advisor, Paul Eklof, who taught me both set
theory and model theory; to Robert Meyer, who introduced me to symbol logic; to
David Elfman, who taught me about logic through programming on an Apple II; and
to my wife, Barb, whose love and patience supported me as I finished this book.
xv
SYMBOLS
Symbol
:=
¬
∧
∨
→
↔
๐—
โŠจ
ฬธโŠจ
⇒
โŠข
⇔
โŠข∗
โŠฌ
๐–ข๐—ˆ๐—‡
๐–ต๐– ๐–ฑ
∀
∃
Page(s)
7
7, 11
7, 11
7, 11
7, 12
7, 13
10
21, 336, 346, 352
23, 338
24
26
34
40
52
52, 395
67
68, 72
68, 72
Symbol
=
๐– 
๐–ฒ๐–ณ
๐–ญ๐–ณ
๐– ๐–ฑ
๐– ๐–ฑ′
๐–ฆ๐–ฑ
๐–ฑ๐–จ
๐–ฎ๐–ฅ
๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐– )
๐–ฒ
๐–ซ(๐–ฒ)
๐‘ก
๐‘ฅ
๐‘(๐‘ฅ)
๐‘Žฬ‚
โˆฃ
|๐‘|
Page(s)
68, 226
68
68
69
69
69
69
69
69
70
70
73
75
80
88
98
116, 307
xvii
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
xviii
LIST OF SYMBOLS
Symbol
∈
∉
∅
๐
๐™
๐
๐‘
๐‚
∞
[๐‘Ž, ๐‘]
(๐‘Ž, ๐‘)
{๐‘ฅ : ๐‘(๐‘ฅ)}
๐™+
๐™−
∪
∩
โงต
๐ด
×
๐‘2
๐‘๐‘›
⊆
*
⊂
gcd
P(๐ด)
โ‹ƒ
โ„ฑ
โ‹‚
โ„ฑ
(๐ด, ๐‘…)
๐ผ๐ด
dom(๐‘…)
ran(๐‘…)
โ—ฆ
๐‘…−1
๐‘Ž๐‘…๐‘
๐‘Ž๐‘…
ฬธ ๐‘
≡
mod ๐‘š
[๐‘Ž]๐‘…
[๐‘›]๐‘š
๐ดโˆ•๐‘…
๐™๐‘š
4
๐”ธ∗
Page(s)
118
118
118
119
119
119
119
119
120
120
120
122
123
123
126
126
127
128
130
130
131
135, 361
135
136
140
151
152
152
161
162
162
162
163, 195
166, 203
168
168
170, 387
170
171
171
172
172
178
178
Symbol
โŸ‚
๐‘“ (๐‘Ž)
๐‘“ :๐ท→๐ต
๐ท→๐ต
โŸฆ๐‘ฅโŸง
๐ด๐ต
๐œ–๐‘Ž
∗
≅
๐œ‘[๐ถ]
๐œ‘−1 [๐ถ]
๐–ข๐—(๐’œ )
๐—ญ
๐—ญ๐—™
๐—ญ๐—™๐—–
M
๐‘Ž+
๐œ”
๐‘›!
โ„ค
โ„š
๐‘›(๐‘ƒ)๐‘Ÿ
๐‘›
๐‘Ÿ
๐น๐‘›
๐œ
๐—Œ๐–พ๐—€4 (๐ด, ๐‘)
โ„
0
1
โ„‚
๐‘–
๐—Œ๐–พ๐—€(๐›ผ, ๐›ฝ)
<๐›ผ๐ด
≈
๐‘›โ„ค
โ‰‰
โชฏ
ฬธโชฏ
โ‰บ
๐œ’๐ต
ℵ0
ℵ
Page(s)
178
183
190
190
191
192
193
194
197, 352
198
212, 380
216
217
232
235
235
235
236
237, 290
238
242
250
254
262
263
269
271
275
276
278
278
281
281
286
294
298
298
298
300
300
300
304
313
313
LIST OF SYMBOLS
Symbol
๐–ข๐–ง
๐–ฆ๐–ข๐–ง
i
๐–ผ๐–ฟ
๐”„
(๐ด, ๐”ž)
๐ผ๐‘ฅ๐‘Ž
๐–ฒ๐–บ๐—๐—ฆ
M๐‘› (โ„)
M∗๐‘› (โ„)
GL(๐‘›, โ„)
๐–ฒ๐–บ๐—
๐”๐‘› (โ„)
โŸจ๐‘ŽโŸฉ
โ‹ƒ
๐›พ∈๐œ… ๐”„๐›พ
๐œ‘:๐”„→๐”…
Page(s)
313
314
316
328
333
333
336
340
345
345
345
352
355
363
372
375
Symbol
๐”„→๐”…
ker(๐œ“)
โชฏ
๐”„∼
๐–ณ๐—(๐”„)
๐‘†
๐—ฃ
๐”“
๐—ฃ๐—”
๐”“′
๐‘‰๐›ผ
๐—ฉ
TC(๐ด)
๐”™๐›ผ
Page(s)
375
379
389
393
404
408
410
410
410
411
411
417
419
419
420
xix
CHAPTER 1
PROPOSITIONAL LOGIC
1.1
SYMBOLIC LOGIC
Let us define mathematics as the study of number and space. Although representations
can be found in the physical world, the subject of mathematics is not physical. Instead,
mathematical objects are abstract, such as equations in algebra or points and lines in
geometry. They are found only as ideas in minds. These ideas sometimes lead to the
discovery of other ideas that do not manifest themselves in the physical world as when
studying various magnitudes of infinity, while others lead to the creation of tangible
objects, such as bridges or computers.
Let us define logic as the study of arguments. In other words, logic attempts to codify
what counts as legitimate means by which to draw conclusions from given information.
There are many variations of logic, but they all can be classified into one of two types.
There is inductive logic in which if the argument is good, the conclusion will probably
follow from the hypotheses. This is because inductive logic rests on evidence and
observation, so there can never be complete certainty whether the conclusions reached
do indeed describe the universe. An example of an inductive argument is:
1
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
2
Chapter 1 PROPOSITIONAL LOGIC
A red sky in the morning means that a storm is coming.
We see a red sky this morning.
Therefore, there will be a storm today.
Whether this is a trust-worthy argument or not rests on the strength of the predictive
abilities of a red sky, and we know about that by past observations. Thus, the argument
is inductive. The other type is deductive logic. Here the methods yield conclusions
with complete certainty, provided, of course, that no errors in reasoning were made.
An example of a deductive argument is:
All geometers are mathematicians.
Euclid is a geometer.
Therefore, Euclid is a mathematician.
Whether Euclid refers to the author of the Elements or is Mr. Euclid from down the
street is irrelevant. The argument works because the third sentence must follow from
the first two.
As anyone who has solved an equation or written a proof can attest, deductive logic
is the realm of the mathematician. This is not to say that there are not other aspects to
the discovery of mathematical results, such as drawing conclusions from diagrams or
patterns, using computational software, or simply making a lucky guess, but it is to say
that to accept a mathematical statement requires the production of a deductive proof of
that statement. For example, in elementary algebra, we know that given
2๐‘ฅ − 5 = 11,
we can conclude
2๐‘ฅ = 6
and then
๐‘ฅ = 3.
As each of the steps is legal, it is certain that the conclusion of ๐‘ฅ = 3 follows. In
geometry, we can write a two-column proof that shows that
∠๐ต ≅ ∠๐ท
is guaranteed to follow from
ABCD is a parallelogram.
The study of these types of arguments, those that are deductive and mathematical in
content, is called mathematical logic.
Propositions
To study arguments, one must first study sentences because they are the main parts of
arguments. However, not just any type of sentence will do. Consider
all squares are rectangles.
Section 1.1 SYMBOLIC LOGIC
3
The purpose of this sentence is to affirm that things called squares also belong to the
category of things called rectangles. In this case, the assertion made by the sentence is
correct. Also, consider,
circles are not round.
This sentence denies that things called circles have the property of being round. This
denial is incorrect. If a sentence asserts or denies accurately, the sentence is true, but if
it asserts or denies inaccurately, the sentence is false. These are the only truth values
that a sentence can have, and if a sentence has one, it does not have the other. As
arguments intend to draw true conclusions from presumably true given sentences, we
limit the sentences that we study to only those with a truth value. This leads us to our
first definition.
DEFINITION 1.1.1
A sentence that is either true or false is called a proposition.
Not all sentences are propositions, however. Questions, exclamations, commands,
or self-contradictory sentences like the following examples can neither be asserted nor
be denied.
โˆ™ Is mathematics logic?
โˆ™ Hey there!
โˆ™ Do not panic.
โˆ™ This sentence is false.
Sometimes it is unclear whether a sentence identifies a proposition. This can be
due to factors such as imprecision or poor sentence structure. Another example is the
sentence
it is a triangle.
Is this true or false? It is impossible to know because, unlike the other words of the
sentence, the meaning of the word it is not determined. In this sentence, the word it is
acting like a variable as in ๐‘ฅ + 2 = 5. As the value of ๐‘ฅ is undetermined, the sentence
๐‘ฅ + 2 = 5 is neither true nor false. However, if ๐‘ฅ represents a particular value, we could
make a determination. For example, if ๐‘ฅ = 3, the sentence is true, and if ๐‘ฅ = 10, the
sentence is false. Likewise, if it refers to a particular object, then it is a triangle would
identify a proposition.
There are two types of propositions. An atom is a proposition that is not comprised
of other propositions. Examples include
the angle sum of a triangle equals two right angles
and
some quadratic equations have real solutions.
4
Chapter 1 PROPOSITIONAL LOGIC
A proposition that is not an atom but is constructed using other propositions is called
a compound proposition. There are five types.
โˆ™ A negation of a given proposition is a proposition that denies the truth of the
given proposition. For example, the negation of 3 + 8 = 5 is 3 + 8 ≠ 5. In this
case, we say that 3 + 8 = 5 has been negated. Negating the proposition the sine
function is periodic yields the sine function is not periodic.
โˆ™ A conjunction is a proposition formed by combining two propositions (called
conjuncts) with the word and. For example,
the base angles of an isosceles triangle are congruent,
and a square has no right angles
is a conjunction with the base angles of an isosceles triangle are congruent and
a square has no right angles as conjuncts.
โˆ™ A disjunction is a proposition formed by combining two propositions (called
disjuncts) with the word or. The sentence
the base angles of an isosceles triangle are congruent,
or a square has no right angles
is a disjunction.
โˆ™ An implication is a proposition that claims a given proposition (called the antecedent) entails another proposition (called the consequent). Implications are
also known as conditional propositions. For example,
if rectangles have four sides, then squares have for sides
(1.1)
is a conditional proposition. Its antecedent is rectangles have four sides, and its
consequent is squares have four sides. This implication can also be written as
rectangles have four sides implies that squares have four sides,
squares have four sides if rectangles have four sides,
rectangles have four sides only if squares have four sides,
and
if rectangles have four sides, squares have four sides.
A conditional proposition can also be written using the words sufficient and necessary. The word sufficient means “adequate” or “enough,” and necessary means
“needed” or “required.” Thus, the sentence
Section 1.1 SYMBOLIC LOGIC
5
rectangles having four sides is sufficient for squares to have four sides
translates (1.1). In other words, the fact that rectangles have four sides is enough
for us to know that squares have four sides. Likewise,
squares having four sides is necessary for rectangles to have four sides
is another translation of the implication because it means that squares must have
four sides because rectangle have four sides. Summing up, the antecedent is
sufficient for the consequent, and the consequent is necessary for the antecedent.
โˆ™ A biconditional proposition is the conjunction of two implications formed by
exchanging their antecedents and consequents. For example,
if rectangles have four sides, then squares have four sides,
and if squares have four sides, then rectangles have four sides.
To remove the redundancy in this sentence, notice that the first conditional can
be written as
rectangles have four sides only if squares have four sides
and the second conditional can be written as
rectangles have four sides if squares have four sides,
resulting in the biconditional being written as
rectangles have four sides if and only if squares have four sides
or the equivalent
rectangles having four sides is necessary and sufficient
for squares to have four sides.
Propositional Forms
As a typical human language has many ways to express the same thought, it is beneficial
to study propositions by translating them into a notation that has a very limited collection of symbols yet is still able to express the basic logic of the propositions. Once this
is done, rules that determine the truth values of propositions using the new notation can
be developed. Any such system designed to concisely study human reasoning is called
a symbolic logic. Mathematical logic is an example of symbolic logic.
Let ๐‘ be a finite sequence of characters from a given collection of symbols. Call
the collection an alphabet. Call ๐‘ a string over the alphabet. The alphabet chosen so
that ๐‘ can represent a mathematical proposition is called the proposition alphabet and
consists of the following symbols.
6
Chapter 1 PROPOSITIONAL LOGIC
โˆ™ Propositional variables: Uppercase English letters, ๐‘ƒ , ๐‘„, ๐‘…, … , or uppercase
English letters with subscripts, ๐‘ƒ๐‘› , ๐‘„๐‘› , ๐‘…๐‘› , … , where ๐‘› = 0, 1, 2, …
โˆ™ Connectives: ¬, ∧, ∨, →, ↔
โˆ™ Grouping symbols: (, ), [, ].
The sequences ๐‘ƒ ∨ ๐‘„ and ๐‘ƒ1 ๐‘„1 ∧ ↔ ((( and the empty string, a string with no characters, are examples of strings over this alphabet, but only certain strings will be chosen
for our study. A string is selected because it is able to represent a proposition. These
strings will be determined by a method called a grammar. The grammar chosen for
our present purposes is given in the next definition. It is given recursively. That is, the
definition is first given for at least one special case, and then the definition is given for
other cases in terms of itself.
DEFINITION 1.1.2
A propositional form is a nonempty string over the proposition alphabet such
that
โˆ™ every propositional variable is a propositional form.
โˆ™ ¬๐‘ is a propositional form if ๐‘ is a propositional form.
โˆ™ (๐‘ ∧ ๐‘ž), (๐‘ ∨ ๐‘ž), (๐‘ → ๐‘ž), and (๐‘ ↔ ๐‘ž) are propositional forms if ๐‘ and ๐‘ž are
propositional forms.
We follow the convention that parentheses can be replaced with brackets and
outermost parenthesis or brackets can be omitted. As with propositions, a propositional form that consists only of a propositional variable is an atom. Otherwise,
it is compound.
The strings ๐‘ƒ , ๐‘„1 , ¬๐‘ƒ , (๐‘ƒ1 ∨ ๐‘ƒ2 ) ∧ ๐‘ƒ3 , and (๐‘ƒ → ๐‘„) ∧ (๐‘… ↔ ¬๐‘ƒ ) are examples of
propositional forms. To prove that the last string is a propositional form, proceed using
Definition 1.1.2 by noting that (๐‘ƒ → ๐‘„) ∧ (๐‘… ↔ ¬๐‘ƒ ) is the result of combining ๐‘ƒ → ๐‘„
and ๐‘… ↔ ¬๐‘ƒ with ∧. The propositional form ๐‘ƒ → ๐‘„ is from ๐‘ƒ and ๐‘„ combined with
→, and ๐‘… ↔ ¬๐‘ƒ is from ๐‘… and ¬๐‘ƒ combined with ↔. These and ¬๐‘ƒ are propositional
forms because ๐‘ƒ , ๐‘„, and ๐‘… are propositional variables. This derivation yields the
following parsing tree:
(P → Q) ∧ (R ↔ ¬P)
P→Q
P
R ↔ ¬P
Q
R
¬P
P
Section 1.1 SYMBOLIC LOGIC
7
The parsing tree yields the formation sequence of the propositional form:
๐‘ƒ , ๐‘„, ๐‘…, ¬๐‘ƒ , ๐‘ƒ → ๐‘„, ๐‘… ↔ ¬๐‘ƒ , (๐‘ƒ → ๐‘„) ∧ (๐‘… ↔ ¬๐‘ƒ ).
The sequence is formed by listing each distinct term of the tree starting at the bottom
row and moving upwards.
EXAMPLE 1.1.3
Make the following assignments:
๐‘ := ๐‘… ↔ (๐‘ƒ ∧ ๐‘„),
๐‘ž := (๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„.
The symbol := indicates that an assignment has been made. It means that the
propositional form on the right has been assigned to the lowercase letter on the
left. Using these designations, we can write new propositional forms using ๐‘ and
๐‘ž. The propositional form ๐‘ ∧ ๐‘ž is
[๐‘… ↔ (๐‘ƒ ∧ ๐‘„)] ∧ [(๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„]
with the formation sequence,
๐‘ƒ , ๐‘„, ๐‘…, ๐‘ƒ ∧ ๐‘„, ๐‘… ↔ ๐‘ƒ ,
๐‘… ↔ (๐‘ƒ ∧ ๐‘„), (๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„, [๐‘… ↔ (๐‘ƒ ∧ ๐‘„)] ∧ [(๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„],
and ¬๐‘ž → ๐‘ is
¬[(๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„] → [๐‘… ↔ (๐‘ƒ ∧ ๐‘„)]
with the formation sequence
๐‘ƒ , ๐‘„, ๐‘…, ๐‘… ↔ ๐‘ƒ , ๐‘ƒ ∧ ๐‘„, (๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„, ๐‘… ↔ (๐‘ƒ ∧ ๐‘„),
¬[(๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„], ¬[(๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„] → [๐‘… ↔ (๐‘ƒ ∧ ๐‘„)].
Interpreting Propositional Forms
Notice that determining whether a string is a propositional form is independent of the
meaning that we give the symbols. However, as we do want these symbols to convey meaning, we assume that the propositional variables represent atoms and set this
interpretation on the connectives:
¬
∧
∨
→
↔
not
and
or
implies
if and only if
Because of this interpretation, name the compound propositional forms as follows:
8
Chapter 1 PROPOSITIONAL LOGIC
¬๐‘
๐‘∧๐‘ž
๐‘∨๐‘ž
๐‘→๐‘ž
๐‘↔๐‘ž
negation
conjunction
disjunction
implication
biconditional
EXAMPLE 1.1.4
To see how this works, assign some propositions to some propositional variables:
๐‘ƒ := The sine function is not one-to-one.
๐‘„ := The square root function is one-to-one.
๐‘… := The absolute value function is not onto.
The following symbols represent the indicated propositions:
โˆ™ ¬๐‘…
The absolute value function is onto.
โˆ™ ¬๐‘ƒ ∨ ¬๐‘„
The sine function is one-to-one, or the square root function is not one-toone.
โˆ™ ๐‘„→๐‘…
If the square root function is one-to-one, the absolute function is not onto.
โˆ™ ๐‘…↔๐‘ƒ
The absolute value function is not onto if and only if the sine function is not
one-to-one.
โˆ™ ๐‘ƒ ∧๐‘„
The sine function is not one-to-one, and the square root function is one-toone.
โˆ™ ¬๐‘ƒ ∧ ๐‘„
The sine function is one-to-one, and the square root function is one-to-one.
โˆ™ ¬(๐‘ƒ ∧ ๐‘„)
It is not the case that the sine function is not one-to-one and the square root
function is one-to-one.
The proposition
the absolute value function is not onto if and only if
both the sine function is not one-to-one and the square root function is one-to-one
is translated as ๐‘… ↔ (๐‘ƒ ∧ ๐‘„) and
Section 1.1 SYMBOLIC LOGIC
9
the absolute value function is not onto if and only if the sine function is not
one-to-one, and the square root function is one-to-one
is translated as (๐‘… ↔ ๐‘ƒ ) ∧ ๐‘„. If the parenthesis are removed, the resulting string is
๐‘… ↔ ๐‘ƒ ∧ ๐‘„. It is simpler, but it is not clear how it should be interpreted. To eliminate
its ambiguity, we introduce an order of connectives as in algebra. In this way, certain
strings without parentheses can be read as propositional forms.
DEFINITION 1.1.5 [Order of Connectives]
To interpret a propositional form, read from left to right and use the following
precedence:
โˆ™ propositional forms within parentheses or brackets (innermost first),
โˆ™ negations,
โˆ™ conjunctions,
โˆ™ disjunctions,
โˆ™ conditionals,
โˆ™ biconditionals.
EXAMPLE 1.1.6
To write the propositional form ¬๐‘ƒ ∨ ๐‘„ ∧ ๐‘… with parentheses, we begin by interpreting ¬๐‘ƒ . According to the order of operations, the conjunction is next, so we
evaluate ๐‘„∧๐‘…. This is followed by the disjunction, and we have the propositional
form ¬๐‘ƒ ∨ (๐‘„ ∧ ๐‘…).
EXAMPLE 1.1.7
To interpret ๐‘ƒ ∧ ๐‘„ ∨ ๐‘… correctly, use the order of operations. We discover that
it has the same meaning as (๐‘ƒ ∧ ๐‘„) ∨ ๐‘…, but how is this distinguished from
๐‘ƒ ∧ (๐‘„ ∨ ๐‘…) in English? Parentheses are not appropriate because they are not
used as grouping symbols in sentences. Instead, use either. . . or. Then, using the
assignments from Example 1.1.4, (๐‘ƒ ∧ ๐‘„) ∨ ๐‘… can be translated as
either the sine function is not one-to-one
and the square root function is one-to-one,
or the absolute value function is not onto.
Notice that either. . . or works as a set of parentheses. We can use this to translate
๐‘ƒ ∧ (๐‘„ ∨ ๐‘…):
the sine function is not one-to-one,
and either the square root function is one-to-one
or the absolute value function is not onto.
10
Chapter 1 PROPOSITIONAL LOGIC
Be careful to note that the either-or phrasing is logically inclusive. For instance,
some colleges require their students to take either logic or mathematics. This
choice is meant to be exclusive in the sense that only one is needed for graduation.
However, it is not logically exclusive. A student can take logic to satisfy the
requirement yet still take a math class.
EXAMPLE 1.1.8
Let us interpret ¬(๐‘ƒ ∧ ๐‘„). We can try translating this as not ๐‘ƒ and ๐‘„, but this
represents ¬๐‘ƒ ∧๐‘„ according to the order of operations. To handle a propositional
form such as ¬(๐‘ƒ ∧ ๐‘„), use a phrase like it is not the case or it is false and the
word both. Therefore, ¬(๐‘ƒ ∧ ๐‘„) becomes
it is not the case that both ๐‘ƒ and ๐‘„
or
it is false that both ๐‘ƒ and ๐‘„.
For instance, make the assignments.
๐‘ƒ := quadratic equations have at most two real solutions,
๐‘„ := the discriminant can be negative.
Then,
quadratic equations do not have at most two real solutions,
and the discriminant can be negative
is a translation of ¬๐‘ƒ ∧ ๐‘„. On the other hand, ¬(๐‘ƒ ∧ ๐‘„) can be
it is not the case that both quadratic equations have at most two real solutions
and the discriminant can be negative.
To interpret ¬๐‘ƒ ∧ ¬๐‘„, use neither-nor:
neither do quadratic equations have at most two real solutions,
nor can the discriminant be negative.
Valuations and Truth Tables
Propositions have truth values, but propositional forms do not. This is because every
propositional form represents any one of infinitely many propositions. However, once a
propositional form is identified with a proposition, there should be a process by which
the truth value of the proposition is associated with the propositional form. This is
done with a rule ๐— called a valuation. The input of ๐— is a propositional form, and its
output is T or F. Suppose that ๐‘ƒ is a propositional variable. If ๐‘ƒ has been assigned a
proposition,
{
T if ๐‘ƒ is true,
๐—(๐‘ƒ ) =
F if ๐‘ƒ is false.
Section 1.1 SYMBOLIC LOGIC
11
For example, if ๐‘ƒ := 2 + 3 = 5, then ๐—(๐‘ƒ ) = T, and if ๐‘ƒ := 2 + 3 = 7, then ๐—(๐‘ƒ ) = F.
If ๐‘ƒ has not been assigned a proposition, then ๐—(๐‘ƒ ) can be defined arbitrarily as either
T or F.
The valuation of a compound propositional form is defined using truth tables. Let
๐‘ and ๐‘ž be given propositional forms. Along the top row write ๐‘ and, if needed, ๐‘ž.
Draw a vertical line. To its right identify the desired propositional form consisting
of ๐‘, possibly ๐‘ž, and a single connective. In the body of the table, on the left of the
vertical line are all combinations of T and F for ๐‘ and possibly ๐‘ž. On the right are the
results of applying the connective. Each connective will have its own truth table, and
we want to define these tables so that they match our understanding of the meaning of
each connective.
Since the truth value of the negation of a given proposition is the opposite of that
proposition’s truth value,
๐‘ ¬๐‘
T F
F T
This means that ๐—(¬๐‘) = F if ๐—(๐‘) = T and ๐—(¬๐‘) = T if ๐—(๐‘) = F.
The conjunction,
3 + 6 = 9, and all even integers are divisible by two,
is true, but
all integers are rational, and 4 is odd
is false because the second conjunct is false. The disjunction
3 + 7 = 9, or all even integers are divisible by three,
is false since both disjuncts are false. On the other hand,
3 + 7 = 9, or circles are round
is true. This illustrates that
โˆ™ a conjunction is true when both of its conjuncts are true, and false otherwise, and
โˆ™ a disjunction is true when at least one disjunct is true, and false otherwise.
We use these principles to define the truth tables for ๐‘ ∧ ๐‘ž and ๐‘ ∨ ๐‘ž:
๐‘
T
T
F
F
๐‘ž ๐‘∧๐‘ž
T
T
F
F
T
F
F
F
๐‘ ๐‘ž ๐‘∨๐‘ž
T T
T
T F
T
F T
T
F F
F
We must remember that only one disjunct needs to be true for the entire disjunction to
be true. For this reason, the logical disjunction is sometimes called an inclusive or.
The propositional form for the exclusive or is
(๐‘ ∨ ๐‘ž) ∧ ¬(๐‘ ∧ ๐‘ž).
12
Chapter 1 PROPOSITIONAL LOGIC
There are many ways to understand an implication. Sometimes it represents causation as in
if I score at least 70 on the exam, I will earn a passing grade.
Other times it indicates what would have been the case if some past event had gone
differently as in
if I had not slept late, I would not have missed the meeting.
Study of such conditional propositions is a very involved subject, one that need not
concern us here because in mathematics a simpler understanding of the implication
is enough. Suppose that ๐‘ƒ and ๐‘„ are assigned propositions so that ๐‘ƒ → ๐‘„ is a true
implication. In mathematics, this means that it is not the case that ๐‘ƒ is true but ๐‘„ is
false. This understanding of the conditional is known as material implication. For
example,
if rectangles have four sides, then squares have four sides,
if rectangles have three sides, then squares have four sides,
and
if rectangles have three sides, then squares have three sides
are all true, but
if rectangles have four sides, then squares have three sides
is false. Generalizing, in mathematics, ๐‘ → ๐‘ž means ¬(๐‘∧¬๐‘ž), which has the following
truth table:
๐‘
T
T
F
F
๐‘ž ¬๐‘ž
T F
F T
T F
F T
๐‘ ∧ ¬๐‘ž
F
T
F
F
¬(๐‘ ∧ ¬๐‘ž)
T
F
T
T
The truth table for ๐‘ → ๐‘ž is then defined as follows:
๐‘ ๐‘ž ๐‘→๐‘ž
T T
T
T F
F
F T
T
F F
T
The truth table for ๐‘ ↔ ๐‘ž is simpler because we understand ๐‘ ↔ ๐‘ž to mean
(๐‘ → ๐‘ž) ∧ (๐‘ž → ๐‘).
The truth table for this propositional form requires five columns:
Section 1.1 SYMBOLIC LOGIC
๐‘ ๐‘ž ๐‘→๐‘ž
T T
T
T F
F
F T
T
F F
T
13
๐‘ž → ๐‘ (๐‘ → ๐‘ž) ∧ (๐‘ž → ๐‘)
T
T
T
F
F
F
T
T
Therefore, define the truth table of ๐‘ ↔ ๐‘ž as:
๐‘ ๐‘ž ๐‘↔๐‘ž
T T
T
T F
F
F T
F
F F
T
This understanding of the biconditional is known as material equivalence.
Using these truth tables, the valuation of an arbitrary propositional form can be
defined.
DEFINITION 1.1.9
Let ๐‘ and ๐‘ž be propositional forms.
{
T if ๐—(๐‘) = F,
โˆ™ ๐—(¬๐‘) =
F if ๐—(๐‘) = T.
{
T if ๐—(๐‘) = T and ๐—(๐‘ž) = T,
โˆ™ ๐—(๐‘ ∧ ๐‘ž) =
F otherwise.
{
F if ๐—(๐‘) = F and ๐—(๐‘ž) = F,
โˆ™ ๐—(๐‘ ∨ ๐‘ž) =
T otherwise.
{
F if ๐—(๐‘) = T and ๐—(๐‘ž) = F,
โˆ™ ๐—(๐‘ → ๐‘ž) =
T otherwise.
{
T if ๐—(๐‘) = ๐—(๐‘ž),
โˆ™ ๐—(๐‘ ↔ ๐‘ž) =
F otherwise.
EXAMPLE 1.1.10
Consider the propositional form (๐‘ƒ ↔ ๐‘„) ∨ (๐‘… → ๐‘ƒ ) where
๐—(๐‘ƒ ) = F, ๐—(๐‘„) = T, and ๐—(๐‘…) = F.
Then, ๐—(๐‘ƒ ↔ ๐‘„) = F because ๐—(๐‘ƒ ) ≠ ๐—(๐‘„), and ๐—(๐‘… → ๐‘ƒ ) = T because
๐—(๐‘…) = F. Therefore, because ๐—(๐‘… → ๐‘ƒ ) = T,
๐—([๐‘ƒ ↔ ๐‘„] ∨ [๐‘… → ๐‘ƒ ]) = T,
14
Chapter 1 PROPOSITIONAL LOGIC
We now generalize the definition of a truth table to create truth tables for more complicated propositional forms and then use the tables to find the valuation of a propositional form given the valuations of its proposition variables.
EXAMPLE 1.1.11
To write the truth table of ๐‘ƒ → ๐‘„∧¬๐‘ƒ , identify the column headings by drawing
the parsing tree for this form:
P → Q ∧ ¬P
P
Q ∧ ¬P
Q
¬P
P
Reading from the bottom, we see that a formation sequence for the propositional
form is
๐‘ƒ , ๐‘„, ¬๐‘ƒ , ๐‘„ ∧ ¬๐‘ƒ , ๐‘ƒ → ๐‘„ ∧ ¬๐‘ƒ .
Hence, the truth table for this form is
๐‘ƒ
T
T
F
F
๐‘„ ¬๐‘ƒ
T F
F
F
T T
F T
๐‘„ ∧ ¬๐‘ƒ
F
F
T
F
๐‘ƒ → ๐‘„ ∧ ¬๐‘ƒ
F
F
T
T
So, if ๐—(๐‘ƒ ) = T and ๐—(๐‘„) = F,
๐—(๐‘ƒ → ๐‘„ ∧ ¬๐‘ƒ ) = F.
That is, any proposition represented by ๐‘ƒ → ๐‘„∧¬๐‘ƒ is false when the proposition
assigned to ๐‘ƒ is true and the proposition assigned to ๐‘„ is false.
The propositional form in the next example has three propositional variables. To
make clear the truth value pattern that is to the left of the vertical line, note that if there
are ๐‘› variables, the number of rows is twice the number of rows for ๐‘› − 1 variables.
To see this, start with one propositional variable. Such a truth table has only two rows.
Add a variable, we obtain four rows. The pattern is obtained by writing the one variable
case twice. For the first time, it has a T written in front of each row. The second copy
has an F in front of each row. To obtain the pattern for three variables, copy the twovariable pattern twice as in Figure 1.1. To generalize, if there are ๐‘› variables, there will
be 2๐‘› rows.
Section 1.1 SYMBOLIC LOGIC
Two
rows
๐‘ƒ
T
F
{
−→ Four
rows
โŽง
โŽช
โŽจ
โŽช
โŽฉ
๐‘ƒ
T
T
F
F
Figure 1.1
๐‘„
T
F
T
F
−→
Eight
rows
โŽง
โŽช
โŽช
โŽช
โŽช
โŽจ
โŽช
โŽช
โŽช
โŽช
โŽฉ
๐‘ƒ
T
T
T
T
F
F
F
F
๐‘„
T
T
F
F
T
T
F
F
๐‘…
T
F
T
F
T
F
T
F
Valuation patterns.
EXAMPLE 1.1.12
Use a truth table to find the truth value of
if the derivative of the sine function is the cosine function
and the second derivative of the sine function is the sine function,
then the third derivative of the sine function is the cosine function.
Define:
๐‘ƒ := the derivative of the sine function is the cosine function,
๐‘„ := the second derivative of the sine function is the sine function,
๐‘… := the third derivative of the sine function is the cosine function.
So ๐‘ƒ represents a true proposition, but ๐‘„ and ๐‘… represent false propositions. The
proposition is represented by
๐‘ƒ ∧๐‘„→๐‘…
with truth table:
๐‘ƒ
T
T
T
T
F
F
F
F
๐‘„
T
T
F
F
T
T
F
F
๐‘… ๐‘ƒ ∧๐‘„
T
T
F
T
T
F
F
F
T
F
F
F
T
F
F
F
๐‘ƒ ∧๐‘„→๐‘…
T
F
T
T
T
T
T
T
Notice that we could have determined the truth value by simply writing one line
from the truth table:
๐‘ƒ
T
๐‘„
F
๐‘… ๐‘ƒ ∧๐‘„
F
F
๐‘ƒ ∧๐‘„→๐‘…
T
We see that ๐—(๐‘ƒ ∧ ๐‘„ → ๐‘…) = T when ๐—(๐‘ƒ ) = T, ๐—(๐‘„) = F, and ๐—(๐‘…) = F.
Therefore, the proposition is true.
15
16
Chapter 1 PROPOSITIONAL LOGIC
EXAMPLE 1.1.13
Both ๐‘ƒ ∨ ¬๐‘ƒ and ๐‘ƒ → ๐‘ƒ share an important property. Their columns in their
truth tables are all T. For example, the truth table of ๐‘ƒ ∨ ¬๐‘ƒ is:
๐‘ƒ
T
F
¬๐‘ƒ
F
T
๐‘ƒ ∨ ¬๐‘ƒ
T
T
Therefore, ๐—(๐‘ƒ ∨ ¬๐‘ƒ ) always equals T, no matter the choice of ๐—.
However, the columns for ๐‘ƒ ∧ ¬๐‘ƒ and ๐‘ƒ ↔ ¬๐‘ƒ are all F. To check the first
one, examine its truth table:
๐‘ƒ
T
F
¬๐‘ƒ
F
T
๐‘ƒ ∧ ¬๐‘ƒ
F
F
This means that ๐—(๐‘ƒ ∧ ¬๐‘ƒ ) is always F for every valuation ๐—.
Based on the last example, we make the next definition.
DEFINITION 1.1.14
A propositional form ๐‘ is a tautology if ๐—(๐‘) always equals T for every valuation
๐—, and ๐‘ is a contradiction if ๐—(๐‘) always equals F for every ๐—. A propositional
form that is neither a tautology nor a contradiction is called a contingency.
Exercises
1. Identify each sentence as either a proposition or not a proposition. Explain.
(a) Trisect the angle.
(b) Some exponential functions are increasing.
(c) All exponential functions are increasing.
(d) 3 + 8 = 18
(e) 3 + ๐‘ฅ = 18
(f) Yea, logic!
(g) A triangle is a three-sided polygon.
(h) The function is differentiable.
(i) This proposition is true.
(j) This proposition is not true.
2. Identify the antecedent and the consequent for the given implications.
(a) If the triangle has two congruent sides, it is isosceles.
(b) The polynomial has at most two roots if it is a quadratic.
(c) The data is widely spread only if the standard deviation is large.
(d) The function being constant implies that its derivative is zero.
(e) The system of equations is consistent is necessary for it to have a solution.
Section 1.1 SYMBOLIC LOGIC
17
(f) A function is even is sufficient for its square to be even.
3. Give the truth value of each proposition.
(a) A system of equations always has a solution, or a quadratic equation always
has a real solution.
(b) It is false that every polynomial function in one variable is differentiable.
(c) Vertical lines have no slope, and lines through the origin have a positive
๐‘ฆ-intercept.
(d) Every integer is even, or every even natural number is an integer.
(e) If every parabola intersects the ๐‘ฅ-axis, then an ellipse has only one vertex.
(f) The sine function is periodic if and only if every exponential function is always nonnegative.
(g) It is not the case that 2 + 4 ≠ 6.
(h) The distance between two points is always positive if every line segment is
horizontal.
(i) The derivative of a constant function is zero is necessary for the product rule
to be true.
(j) The derivative of the sine function being cosine is sufficient for the derivative
of the cosine function being sine.
(k) Any real number is negative or positive, but not both.
4. For each sentence, fill in the blank using as many of the words and, or, if, and if and
only if as possible to make the proposition true.
(a) Triangles have three sides
3 + 5 = 6.
(b) 3 + 5 = 6
triangles have three sides.
(c) Ten is the largest integer
zero is the smallest integer.
(d) The derivative of a constant function is zero
tangent lines for increasing functions have positive slope.
5. Extend Figure 1.1 by writing the typical pattern of Ts and Fs for the truth table of
a propositional form with four propositional variables and then with five propositional
variables.
6. Use a parsing tree to show that the given string is a propositional form.
(a) ๐‘ƒ ∧ ๐‘„ ∨ ๐‘…
(b) ๐‘„ ↔ ๐‘… ∨ ¬๐‘„
(c) ๐‘ƒ → ๐‘„ → ๐‘… → ๐‘†
(d) ¬๐‘ƒ ∧ ๐‘„ ∨ (๐‘ƒ → ๐‘„) ∧ ¬๐‘†
(e) (๐‘ƒ ∧ ๐‘„ → ๐‘„) ∧ ๐‘ƒ → ๐‘„
(f) ¬¬๐‘ƒ ∨ ๐‘ƒ ∧ ๐‘† → ๐‘„ ∨ [๐‘… → ¬๐‘ƒ → ¬(๐‘„ ∨ ๐‘…)]
7. Define:
๐‘ƒ := the angle sum of a triangle is 180,
๐‘„ := 3 + 7 = 10,
๐‘… := the sine function is continuous.
18
Chapter 1 PROPOSITIONAL LOGIC
Translate the given propositional forms into English.
(a) ๐‘ƒ ∨ ๐‘„
(b) ๐‘ƒ ∧ ๐‘„
(c) ๐‘ƒ ∧ ¬๐‘„
(d) ๐‘„ ∨ ¬๐‘…
(e) ๐‘„ ↔ ¬๐‘…
(f) ๐‘… → ๐‘„
(g) ๐‘ƒ ∨ ๐‘… → ¬๐‘„
(h) ๐‘„ ↔ ๐‘… ∧ ¬๐‘„
(i) ¬(๐‘ƒ ∧ ๐‘„)
(j) ¬(๐‘ƒ ∨ ๐‘„)
(k) ๐‘ƒ ∨ ๐‘„ ∧ ๐‘…
(l) (๐‘ƒ ∨ ๐‘„) ∧ ๐‘…
8. Write the following sentences as propositional forms using the variables ๐‘ƒ , ๐‘„, and
๐‘… as defined in Exercise 7.
(a) The sine function is continuous, and 3 + 7 = 10.
(b) The angle sum of a triangle is 180, or the angle sum of a triangle is 180.
(c) If 3 + 7 = 10, then the sine function is not continuous.
(d) The angle sum of a triangle is 180 if and only if the sine function is continuous.
(e) The sine function is continuous if and only if 3 + 7 = 10 implies that the
angle sum of a triangle is not 180.
(f) It is not the case that 3 + 7 ≠ 10.
9. Let ๐—(๐‘ƒ ) = T, ๐—(๐‘„) = T, ๐—(๐‘…) = F, and ๐—(๐‘†) = F. Find the given valuations. (See
Exercise 6.)
(a) ๐‘ƒ ∧ ๐‘„ ∨ ๐‘…
(b) ๐‘„ ↔ ๐‘… ∨ ¬๐‘„
(c) ๐‘ƒ → ๐‘„ → ๐‘… → ๐‘†
(d) ¬๐‘ƒ ∧ ๐‘„ ∨ (๐‘ƒ → ๐‘„) ∧ ¬๐‘†
(e) (๐‘ƒ ∧ ๐‘„ → ๐‘„) ∧ ๐‘ƒ → ๐‘„
(f) ¬¬๐‘ƒ ∨ ๐‘ƒ ∧ ๐‘† → ๐‘„ ∨ [๐‘… → ¬๐‘ƒ → ¬(๐‘„ ∨ ๐‘…)]
10. Write the truth table for each of the given propositional forms.
(a) ¬๐‘ƒ → ๐‘ƒ
(b) ๐‘ƒ → ¬๐‘„
(c) (๐‘ƒ ∨ ๐‘„) ∧ ¬(๐‘ƒ ∧ ๐‘„)
(d) (๐‘ƒ → ๐‘„) ∨ (๐‘„ ↔ ๐‘ƒ )
(e) ๐‘ƒ ∧ (๐‘„ ∨ ๐‘…)
(f) ๐‘ƒ ∨ ๐‘„ → ๐‘…
(g) ๐‘ƒ → ๐‘„ ∧ ¬(๐‘… ∨ ๐‘ƒ )
(h) ๐‘ƒ → ๐‘„ ↔ ๐‘… → ๐‘†
(i) ๐‘ƒ ∨ (¬๐‘„ ↔ ๐‘…) ∧ ๐‘„
Section 1.2 INFERENCE
19
(j) (¬๐‘ƒ ∨ ๐‘„) ∧ ([๐‘ƒ → ๐‘„] ∨ ¬๐‘†)
11. Check the truth value of these propositions using truth tables as in Example 1.1.12.
(a) If 2 + 3 = 7, then 5 − 9 ≠ 0.
(b) If a square is round implies that some functions have a derivative at ๐‘ฅ = 2,
then every function has a derivative at ๐‘ฅ = 2.
(c) Either four is odd or two is even implies that three is even.
(d) Every even integer is divisible by 4 if and only if either 7 divides 21 or 9
divides 12.
(e) The graph of the tangent function has asymptotes, and if sine is an increasing
function, then cosine is a decreasing function.
12. If possible, find propositional forms ๐‘ and ๐‘ž such that
(a) ๐‘ ∧ ๐‘ž is a tautology.
(b) ๐‘ ∨ ๐‘ž is a contradiction.
(c) ¬๐‘ is a tautology.
(d) ๐‘ → ๐‘ž is a contradiction.
(e) ๐‘ ↔ ๐‘ž is a tautology.
(f) ๐‘ ↔ ๐‘ž is a contradiction.
1.2
INFERENCE
Now that we have a collection of propositional forms and a means by which to interpret them as either true or false, we want to define a system that expands these ideas to
include methods by which we can prove certain propositional forms from given propositional forms. What we will define is familiar because it is similar to what Euclid did
with his geometry. Take, for example, the familiar result,
opposite angles in a parallelogram are congruent.
In other words,
if ๐ด๐ต๐ถ๐ท is a parallelogram, then ∠๐ต ≅ ∠๐ท,
which translates to:
Given:
Prove:
๐ด๐ต๐ถ๐ท is a parallelogram,
∠๐ต ≅ ∠๐ท.
To demonstrate this, we draw a diagram,
A
D
and then write a proof:
B
C
20
Chapter 1 PROPOSITIONAL LOGIC
1.
2.
3.
4.
5.
6.
7.
๐ด๐ต๐ถ๐ท is a parallelogram
Join ๐ด๐ถ
∠๐ท๐ด๐ถ ≅ ∠๐ต๐ถ๐ด
∠๐ด๐ถ๐ท ≅ ∠๐ถ๐ด๐ต
๐ด๐ถ ≅ ๐ด๐ถ
โ–ณ๐ด๐ถ๐ท ≅ โ–ณ๐ถ๐ด๐ต
∠๐ต ≅ ∠๐ท
Given
Postulate
Alternate interior angles
Alternate interior angles
Reflexive
ASA
Corresponding parts
Euclid’s geometry consists of geometric propositions that are established by proofs
like the above. These proofs rely on rules of logic, previously proved propositions
(lemmas, theorems, and corollaries), and propositions that are assumed to be true (the
postulates). Using this system of thought, we can show which geometric propositions
follow from the postulates and conclude which propositions are true, whatever it means
for a geometric proposition to be true. Euclidean geometry serves as a model for the
following modern definition.
DEFINITION 1.2.1
A logical system consists of the following:
โˆ™ An alphabet
โˆ™ A grammar
โˆ™ Propositional forms that require no proof
โˆ™ Rules that determine truth
โˆ™ Rules that are used to write proofs.
Although Euclid did not provide an alphabet or a grammar specifically for his geometry,
his system did include the last three aspects of a logical system. In this chapter we
develop the logical system known as propositional logic. Its alphabet, grammar, and
rules that determine truth were defined in Section 1.1. The remainder of this chapter is
spent establishing the other two components.
Consider the following collection of propositions:
If squares are rectangles, then squares are quadrilaterals.
Squares are rectangles.
Therefore, squares are quadrilaterals.
This is an example of a deduction, a collection of propositions of which one is supposed
to follow necessarily from the others. In this particular case,
if squares are rectangles, then squares are quadrilaterals
and
Section 1.2 INFERENCE
21
squares are rectangles
are the premises, and
squares are quadrilaterals
is the conclusion. We recognize that in this case, the conclusion does follow from the
premises because whenever the premises are true, the conclusion must also be true.
When this is the case, the deduction is semantically valid, else it is semantically invalid.
Notice that not only do we see that the deduction works because of the meaning of
the propositions, but we also see that it is valid based on the forms of the sentences. In
other words, we also recognize this deduction as valid:
If Hausdorff spaces are preregular, their points can be separated.
Hausdorff spaces are preregular.
Therefore, their points can be separated.
Although we might not know the terms Hausdorff space, preregular, and separated, we
recognize the deduction as valid because it is of the same pattern as the first deduction:
๐‘→๐‘ž
๐‘
∴ ๐‘ž
(1.2)
When the deduction is found to work based on its form, the deduction is syntactically
valid, else it is syntactically invalid.
We study both types of validity by examining general patterns of deductions and
choosing rules that determine which forms correspond to deductions that are valid semantically and which forms correspond to deductions that are valid syntactically.
Semantics
The study of meaning is called semantics. We began this study when we wrote truth
tables. These are characterized as semantic because the truth value of a proposition is
based on its meaning. Our goal is to use truth tables to determine when an argument
form, an example being (1.2), corresponds to a deduction that is semantically valid.
We begin with a definition.
DEFINITION 1.2.2
Let ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 and ๐‘ž be propositional forms.
โˆ™ If ๐‘ž is a tautology, write โŠจ ๐‘ž.
โˆ™ Define ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 to logically imply ๐‘ž if
โŠจ ๐‘0 ∧ ๐‘1 ∧ · · · ∧ ๐‘๐‘›−1 → ๐‘ž.
22
Chapter 1 PROPOSITIONAL LOGIC
When ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 logically imply ๐‘ž, write
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠจ ๐‘ž.
and say that ๐‘ž is a consequence of ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 . Call the propositional
forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 the premises of the implication and ๐‘ž the conclusion.
Notice that if ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠจ ๐‘ž, then for any valuation ๐—, whenever ๐—(๐‘๐‘– ) = T for all
๐‘– = 0, 1, … , ๐‘› − 1, it must be the case that ๐—(๐‘ž) = T. Moreover, any deduction with
premises represented by ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 and conclusion by ๐‘ž is semantically valid if
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠจ ๐‘ž.
EXAMPLE 1.2.3
Because of Example 1.1.13, both โŠจ ๐‘ƒ → ๐‘ƒ and โŠจ ๐‘ƒ ∨ ¬๐‘ƒ .
EXAMPLE 1.2.4
Prove: ๐‘ƒ → ๐‘„, ๐‘ƒ โŠจ ๐‘„
To accomplish this, show that the propositional form
(๐‘ƒ → ๐‘„) ∧ ๐‘ƒ → ๐‘„,
with antecedent equal to the conjunction of the premises and consequent consisting of the conclusion is a tautology.
๐‘„ ๐‘ƒ → ๐‘„ (๐‘ƒ → ๐‘„) ∧ ๐‘ƒ
T
T
T
F
F
F
T
T
F
F
T
F
๐‘ƒ
T
T
F
F
(๐‘ƒ → ๐‘„) ∧ ๐‘ƒ → ๐‘„
T
T
T
T
Therefore, โŠจ (๐‘ƒ → ๐‘„) ∧ ๐‘ƒ → ๐‘„, so
๐‘ƒ → ๐‘„, ๐‘ƒ โŠจ ๐‘„,
and any deduction based on this form is semantically valid.
EXAMPLE 1.2.5
Prove: ๐‘ƒ ∨ ๐‘„ → ๐‘„, ๐‘ƒ โŠจ ๐‘„
๐‘ƒ
T
T
F
F
๐‘„
T
F
T
F
๐‘ƒ ∨๐‘„
T
T
T
F
๐‘ƒ ∨๐‘„→๐‘„
T
F
T
T
(๐‘ƒ ∨ ๐‘„ → ๐‘„) ∧ ๐‘ƒ
T
F
F
F
(๐‘ƒ ∨ ๐‘„ → ๐‘„) ∧ ๐‘ƒ → ๐‘„
T
T
T
T
If it is possible for ๐—(๐‘๐‘– ) = T for ๐‘– = 0, 1, … , ๐‘› − 1 yet ๐—(๐‘ž) = F, the propositional form
๐‘0 ∧ ๐‘1 ∧ · · · ∧ ๐‘๐‘›−1 → ๐‘ž
Section 1.2 INFERENCE
23
is not a tautology, so ๐‘ž is not a consequence of ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 . If this is the case, write
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ฬธโŠจ ๐‘ž.
EXAMPLE 1.2.6
Prove: ๐‘ƒ ∧ ๐‘„ → ๐‘„, ๐‘ƒ ฬธโŠจ ๐‘„
๐‘ƒ
T
T
F
F
๐‘„
T
F
T
F
๐‘ƒ ∧๐‘„
T
F
F
F
๐‘ƒ ∧๐‘„→๐‘„
T
T
T
T
(๐‘ƒ ∧ ๐‘„ → ๐‘„) ∧ ๐‘ƒ
T
T
F
F
(๐‘ƒ ∧ ๐‘„ → ๐‘„) ∧ ๐‘ƒ → ๐‘„
T
F
T
T
Notice that F appears for (๐‘ƒ ∧ ๐‘„ → ๐‘„) ∧ ๐‘ƒ → ๐‘„ on a line when
๐—(๐‘ƒ ∧ ๐‘„ → ๐‘„) = ๐—(๐‘ƒ ) = T
yet ๐—(๐‘„) = F. Because of this, we can shorten the procedure for showing that a propositional form is not a consequence of other propositional forms.
EXAMPLE 1.2.7
Prove: ๐‘ƒ ∧ ๐‘„ → ๐‘… ฬธโŠจ ๐‘ƒ → ๐‘…
๐‘ƒ
T
๐‘„
F
๐‘… ๐‘ƒ ∧๐‘„→๐‘…
F
T
๐‘ƒ →๐‘…
F
Observe that this shows that the valuation of ๐‘ƒ ∧ ๐‘„ → ๐‘… can be T at the same
time that the valuation of ๐‘ƒ → ๐‘… is F.
Syntactics
Although we will return to semantics, it is important to note that using truth tables to
check for logical implication has its limitations. If the argument form involves many
propositional forms or if the propositional forms are complicated, the truth table used
to show or disprove the logical implication can become unwieldy. Another issue is
that in practice, truth tables are not the method of choice when determining whether
a conclusion follows from the premises. What is typically done is to follow Euclid’s
example (page 20), basing the conclusions on the syntax of the argument form, namely,
based only on its pattern and structure.
Let us return to the deduction on page 20 and work with it differently. Start with the
two propositions,
if squares are rectangles, then squares are quadrilaterals
and
squares are rectangles.
Because of the combined structure of the two sentences, we know that we can write
24
Chapter 1 PROPOSITIONAL LOGIC
squares are quadrilaterals.
This act of writing (on paper or a blackboard or in the mind) means that we have the
proposition and that it follows from the first two. Similarly, if we start with
squares are triangles, or squares are rectangles
and
squares are not triangles,
we can write
squares are rectangles.
Determining a method that will model this reasoning requires us to find rules by
which propositional forms can be written from other propositional forms. Since every logical system requires a starting point, the first step in this process is to choose
which propositional forms can be written without any prior justification. Each such
propositional form is called an axiom. Playing the same role as that of a postulate in
Euclidean geometry, an axiom can be considered as a rule of the game. Certain propositional forms lend themselves as good candidates for axioms because they are regarded
as obvious. That is, they are self-evident. Other propositional forms are good candidates to be axioms, not because they are necessarily self-evident, but because they are
helpful. In either case, the number of axioms should be as few as possible so as to minimize the number of assumptions. For propositional logic, we choose only three. They
were first found in work of Gottlob Frege (1879) and later in that of Jan ลukasiewicz
(1930).
AXIOMS 1.2.8 [Frege–ลukasiewicz]
Let ๐‘, ๐‘ž, and ๐‘Ÿ be propositional forms.
โˆ™ [FL1] ๐‘ → (๐‘ž → ๐‘)
โˆ™ [FL2] ๐‘ → (๐‘ž → ๐‘Ÿ) → (๐‘ → ๐‘ž → [๐‘ → ๐‘Ÿ])
โˆ™ [FL3] ¬๐‘ → ¬๐‘ž → (๐‘ž → ๐‘).
The next step in defining propositional logic is to state when it is legal to write a
propositional form from given propositional forms.
DEFINITION 1.2.9
The propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 infer ๐‘ž if ๐‘ž can be written whenever
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 are written. Denote this by
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ⇒ ๐‘ž.
This is known as an inference.
To make rigorous which propositional forms can be inferred from given forms, we
establish some rules. These are chosen because they model basic reasoning. They are
also not proved, so they serve as postulates for our logic.
Section 1.2 INFERENCE
25
INFERENCE RULES 1.2.10
Let ๐‘, ๐‘ž, ๐‘Ÿ, and ๐‘  be propositional forms.
โˆ™ Modus Ponens [MP]
๐‘ → ๐‘ž, ๐‘ ⇒ ๐‘ž
โˆ™ Modus Tolens [MT]
๐‘ → ๐‘ž, ¬๐‘ž ⇒ ¬๐‘
โˆ™ Constructive Dilemma [CD]
(๐‘ → ๐‘ž) ∧ (๐‘Ÿ → ๐‘ ), ๐‘ ∨ ๐‘Ÿ ⇒ ๐‘ž ∨ ๐‘ 
โˆ™ Destructive Dilemma [DD]
(๐‘ → ๐‘ž) ∧ (๐‘Ÿ → ๐‘ ), ¬๐‘ž ∨ ¬๐‘  ⇒ ¬๐‘ ∨ ¬๐‘Ÿ
โˆ™ Disjunctive Syllogism [DS]
๐‘ ∨ ๐‘ž, ¬๐‘ ⇒ ๐‘ž
โˆ™ Hypothetical Syllogism [HS]
๐‘ → ๐‘ž, ๐‘ž → ๐‘Ÿ ⇒ ๐‘ → ๐‘Ÿ
โˆ™ Conjunction [Conj]
๐‘, ๐‘ž ⇒ ๐‘ ∧ ๐‘ž
โˆ™ Simplification [Simp]
๐‘∧๐‘ž ⇒๐‘
โˆ™ Addition [Add]
๐‘ ⇒ ๐‘ ∨ ๐‘ž.
To use Inference Rules 1.2.10, match the form exactly. For example, even though
๐‘ƒ → ๐‘… appears to follow from (๐‘ƒ ∧ ๐‘„) → ๐‘… as an application of simplification, it does
not. The problem is that simplification can only be applied to propositional forms with
the ๐‘ ∧ ๐‘ž pattern, but (๐‘ƒ ∧ ๐‘„) → ๐‘… is of the form ๐‘ → ๐‘ž. With this detail in mind, we
make some inferences.
EXAMPLE 1.2.11
Each inference is justified by the indicated rule.
โˆ™ Modus ponens
๐‘ƒ ∧ ๐‘„ → ¬๐‘…, ๐‘ƒ ∧ ๐‘„ ⇒ ¬๐‘…
โˆ™ Addition
๐‘ƒ ⇒๐‘ƒ ∨๐‘„∧๐‘…
โˆ™ Modus tolens
¬¬๐‘ƒ , ๐‘„ ∨ ๐‘… → ¬๐‘ƒ ⇒ ¬(๐‘„ ∨ ๐‘…).
26
Chapter 1 PROPOSITIONAL LOGIC
EXAMPLE 1.2.12
Since it is possible that some propositional forms are not needed for the inference,
we also have the following:
โˆ™ Modus ponens
๐‘ƒ ∧ ๐‘„ → ¬๐‘…, ๐‘ƒ ∨ ๐‘„, ๐‘ƒ ∧ ๐‘„, ๐‘„ ↔ ๐‘† ⇒ ¬๐‘…
โˆ™ Addition
๐‘ƒ , ๐‘…, ๐‘† → ๐‘‡ ⇒ ๐‘ƒ ∨ ๐‘„ ∧ ๐‘…
โˆ™ Modus tolens
¬๐‘†, ¬¬๐‘ƒ , ๐‘ƒ ∧ ๐‘‡ , ๐‘„ ∨ ๐‘… → ¬๐‘ƒ ⇒ ¬(๐‘„ ∨ ๐‘…).
Inference is a powerful tool, but it can only be used to check simple deductions.
Sometimes multiple inferences are needed to move from a collection of premises to a
conclusion. For example, if we write
๐‘ ∨ ๐‘ž, ¬๐‘, ๐‘ž → ๐‘Ÿ,
based on the first two propositional forms, we can write
๐‘ž
by DS, and then based on this propositional form and the third of the given propositional
forms, we can write
๐‘Ÿ
by MP. This is a simple example of the next definition.
DEFINITION 1.2.13
โˆ™ A formal proof of the propositional form ๐‘ž (the conclusion) from the propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 (the premises) is a sequence of propositional forms,
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘š−1 ,
such that ๐‘ž๐‘š−1 = ๐‘ž, and for all ๐‘– = 0, 1, … , ๐‘š − 1, either ๐‘ž๐‘– is an axiom,
if ๐‘– = 0, then ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ⇒ ๐‘ž๐‘– , or
if ๐‘– > 0, then ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘–−1 ⇒ ๐‘ž๐‘– .
If there exists a formal proof of ๐‘ž from ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , then ๐‘ž is proved or deduced from ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 and we write
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠข ๐‘ž.
โˆ™ If there are no premises, a formal proof of ๐‘ž is a sequence,
๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘š−1 ,
Section 1.2 INFERENCE
27
such that ๐‘ž0 is an axiom, ๐‘ž๐‘š−1 = ๐‘ž, and for all ๐‘– > 0, either ๐‘ž๐‘– is an axiom or
๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘–−1 ⇒ ๐‘ž๐‘– .
In this case, write โŠข ๐‘ž and call ๐‘ž a theorem.
Observe that any deduction with premises represented by ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 and conclusion by ๐‘ž is syntactically valid if ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠข ๐‘ž.
We should note that although ⇒ and โŠข have different meanings as syntactic symbols,
they are equivalent. If ๐‘ ⇒ ๐‘ž, then ๐‘ โŠข ๐‘ž using the proof ๐‘, ๐‘ž. Conversely, suppose
๐‘ โŠข ๐‘ž. This means that there exists a proof
๐‘, ๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘›−1 , ๐‘ž,
so every time we write down ๐‘, we can also write down ๐‘ž. That is, ๐‘ ⇒ ๐‘ž. We summarize this as follows.
THEOREM 1.2.14
For all propositional forms ๐‘ and ๐‘ž, ๐‘ ⇒ ๐‘ž if and only if ๐‘ โŠข ๐‘ž.
We use a particular style to write formal proofs. They will be in two-column format
with each line being numbered. In the first column will be the sequence of propositional
forms that make up the proof. In the second column will be the reasons that allowed
us to include each form. The only reasons that we will use are
โˆ™ Given (for premises),
โˆ™ FL1, FL2, or FL3 (for an axiom),
โˆ™ An inference rule.
An inference rule is cited by giving the line numbers used as the premises followed by
the abbreviation for the rule. Thus, the following proves ๐‘ƒ ∨ ๐‘„ → ๐‘„ ∧ ๐‘…, ๐‘ƒ โŠข ๐‘„:
1.
2.
3.
4.
5.
๐‘ƒ ∨๐‘„→๐‘„∧๐‘…
๐‘ƒ
๐‘ƒ ∨๐‘„
๐‘„∧๐‘…
๐‘„
Given
Given
2 Add
1, 3 MP
4 Simp
Despite the style, we should remember that a proof is a sequence of propositional forms
that satisfy Definition 1.2.13. In this case, the sequence is
๐‘ƒ ∨ ๐‘„ → ๐‘„ ∧ ๐‘…, ๐‘ƒ , ๐‘ƒ ∨ ๐‘„, ๐‘„ ∧ ๐‘…, ๐‘„.
The first two examples involve proofs that use the axioms.
28
Chapter 1 PROPOSITIONAL LOGIC
EXAMPLE 1.2.15
Prove: โŠข ๐‘ƒ → ๐‘„ → (๐‘ƒ → ๐‘ƒ )
1.
2.
3.
๐‘ƒ → (๐‘„ → ๐‘ƒ )
๐‘ƒ → (๐‘„ → ๐‘ƒ ) → (๐‘ƒ → ๐‘„ → [๐‘ƒ → ๐‘ƒ ])
๐‘ƒ → ๐‘„ → (๐‘ƒ → ๐‘ƒ )
FL1
FL2
MP
This proves that ๐‘ƒ → ๐‘„ → (๐‘ƒ → ๐‘ƒ ) is a theorem. Also, by adding ๐‘ƒ → ๐‘„ as a
given and an application of MP at the end, we can prove
๐‘ƒ → ๐‘„ โŠข ๐‘ƒ → ๐‘ƒ.
This result should not be surprising since ๐‘ƒ → ๐‘ƒ is a tautology. We would expect
any premise to be able to prove it.
EXAMPLE 1.2.16
Prove: ¬(๐‘„ → ๐‘ƒ ), ¬๐‘ƒ โŠข ¬๐‘„
1.
2.
3.
4.
5.
¬(๐‘„ → ๐‘ƒ )
¬๐‘ƒ
¬๐‘ƒ → ¬๐‘„ → (๐‘„ → ๐‘ƒ )
¬๐‘ƒ → ¬๐‘„
¬๐‘„
Given
Given
FL3
1, 3 MT
2, 4 MP
The next three examples do not use an axiom in their proofs.
EXAMPLE 1.2.17
Prove: ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘† ∨ ¬๐‘…, ¬๐‘† โŠข ¬๐‘ƒ
1.
2.
3.
4.
5.
6.
7.
๐‘ƒ →๐‘„
๐‘„→๐‘…
๐‘† ∨ ¬๐‘…
¬๐‘†
๐‘ƒ →๐‘…
¬๐‘…
¬๐‘ƒ
Given
Given
Given
Given
1, 2 HS
3, 4 DS
5, 6 MT
EXAMPLE 1.2.18
Prove: ๐‘ƒ → ๐‘„, ๐‘ƒ → ๐‘„ → (๐‘‡ → ๐‘†), ๐‘ƒ ∨ ๐‘‡ , ¬๐‘„ โŠข ๐‘†
1.
2.
3.
๐‘ƒ →๐‘„
๐‘ƒ → ๐‘„ → (๐‘‡ → ๐‘†)
๐‘ƒ ∨๐‘‡
Given
Given
Given
Section 1.2 INFERENCE
4.
5.
6.
7.
8.
¬๐‘„
๐‘‡ →๐‘†
(๐‘ƒ → ๐‘„) ∧ (๐‘‡ → ๐‘†)
๐‘„∨๐‘†
๐‘†
Given
1, 2 MP
1, 5 Conj
3, 6 CD
4, 7 DS
EXAMPLE 1.2.19
Prove: ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ¬๐‘… โŠข ¬๐‘„ ∨ ¬๐‘ƒ
1.
2.
3.
4.
5.
6.
๐‘ƒ →๐‘„
๐‘„→๐‘…
¬๐‘…
(๐‘„ → ๐‘…) ∧ (๐‘ƒ → ๐‘„)
¬๐‘… ∨ ¬๐‘„
¬๐‘„ ∨ ¬๐‘ƒ
Given
Given
Given
1, 2 Conj
3 Add
4, 5 DD
Exercises
1. Show using truth tables.
(a) ¬๐‘ƒ ∨ ๐‘„, ¬๐‘„ โŠจ ¬๐‘ƒ
(b) ¬(๐‘ƒ ∧ ๐‘„), ๐‘ƒ โŠจ ¬๐‘„
(c) ๐‘ƒ → ๐‘„, ๐‘ƒ โŠจ ๐‘„ ∨ ๐‘…
(d) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘ƒ โŠจ ๐‘…
(e) ๐‘ƒ ∨ ๐‘„ ∧ ๐‘…, ¬๐‘ƒ โŠจ ๐‘…
2. Show the following using truth tables.
(a) ¬(๐‘ƒ ∧ ๐‘„) ฬธโŠจ ¬๐‘ƒ
(b) ๐‘ƒ → ๐‘„ ∨ ๐‘…, ๐‘ƒ ฬธโŠจ ๐‘„
(c) ๐‘ƒ ∧ ๐‘„ → ๐‘… ฬธโŠจ ๐‘„ → ๐‘…
(d) (๐‘ƒ → ๐‘„) ∨ (๐‘… → ๐‘†), ๐‘ƒ ∨ ๐‘… ฬธโŠจ ๐‘„ ∨ ๐‘†
(e) ¬(๐‘ƒ ∧ ๐‘„) ∨ ๐‘…, ๐‘ƒ ∧ ๐‘„ ∨ ๐‘† ฬธโŠจ ๐‘… ∧ ๐‘†
(f) ๐‘ƒ ∨ ๐‘…, ๐‘„ ∨ ๐‘†, ๐‘… ↔ ๐‘† ฬธโŠจ ๐‘… ∧ ๐‘†
3. Identify the rule from Inference Rules 1.2.10.
(a) ๐‘ƒ → ๐‘„ → ๐‘ƒ , ๐‘ƒ → ๐‘„ ⇒ ๐‘ƒ
(b) ๐‘ƒ , ๐‘„ ∨ ๐‘… ⇒ ๐‘ƒ ∧ (๐‘„ ∨ ๐‘…)
(c) ๐‘ƒ ⇒ ๐‘ƒ ∨ (๐‘… ↔ ¬๐‘ƒ ∧ ¬[๐‘„ → ๐‘†])
(d) ๐‘ƒ , ๐‘ƒ → (๐‘„ ↔ ๐‘†) ⇒ ๐‘„ ↔ ๐‘†
(e) ๐‘ƒ ∨ ๐‘„ ∨ ๐‘„, (๐‘ƒ ∨ ๐‘„ → ๐‘„) ∧ (๐‘„ → ๐‘† ∧ ๐‘‡ ) ⇒ ๐‘„ ∨ ๐‘† ∧ ๐‘‡
(f) ๐‘ƒ ∨ (๐‘„ ∨ ๐‘†), ¬๐‘ƒ ⇒ ๐‘„ ∨ ๐‘†
(g) ๐‘ƒ → ¬๐‘„, ¬¬๐‘„ ⇒ ¬๐‘ƒ
(h) (๐‘ƒ → ๐‘„) ∧ (๐‘„ → ๐‘…), ¬๐‘„ ∨ ¬๐‘… ⇒ ¬๐‘ƒ ∨ ¬๐‘„
(i) (๐‘ƒ → ๐‘„) ∧ (๐‘„ → ๐‘…) ⇒ ๐‘ƒ → ๐‘„
29
30
Chapter 1 PROPOSITIONAL LOGIC
4. Arrange each collection of propositional forms into a proof for the given deductions
and supply the appropriate reasons.
(a) ๐‘ƒ → ๐‘„, ๐‘… → ๐‘†, ๐‘ƒ โŠข ๐‘„ ∨ ๐‘†
โˆ™ ๐‘ƒ
โˆ™ ๐‘„∨๐‘†
โˆ™ ๐‘…→๐‘†
โˆ™ (๐‘ƒ → ๐‘„) ∧ (๐‘… → ๐‘†)
โˆ™ ๐‘ƒ ∨๐‘…
โˆ™ ๐‘ƒ →๐‘„
(b) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘ƒ โŠข ๐‘… ∨ ๐‘„
โˆ™ ๐‘ƒ
โˆ™ ๐‘ƒ →๐‘…
โˆ™ ๐‘ƒ →๐‘„
โˆ™ ๐‘…
โˆ™ ๐‘„→๐‘…
โˆ™ ๐‘…∨๐‘„
(c) (๐‘ƒ → ๐‘„) ∨ (๐‘„ → ๐‘…), ¬(๐‘ƒ → ๐‘„), ¬๐‘…, ๐‘„ ∨ ๐‘† โŠข ๐‘†
โˆ™ ¬(๐‘ƒ → ๐‘„)
โˆ™ ๐‘†
โˆ™ (๐‘ƒ → ๐‘„) ∨ (๐‘„ → ๐‘…)
โˆ™ ๐‘„→๐‘…
โˆ™ ¬๐‘…
โˆ™ ¬๐‘„
โˆ™ ๐‘„∨๐‘†
(d) (๐‘ƒ ∨ ๐‘„) ∧ ๐‘…, ๐‘„ ∨ ๐‘† → ๐‘‡ , ¬๐‘ƒ โŠข ¬๐‘ƒ ∧ ๐‘‡
โˆ™ ๐‘ƒ ∨๐‘„
โˆ™ ๐‘„
โˆ™ ๐‘„∨๐‘†
โˆ™ ๐‘‡
โˆ™ (๐‘ƒ ∨ ๐‘„) ∧ ๐‘…
โˆ™ ๐‘„∨๐‘† →๐‘‡
โˆ™ ¬๐‘ƒ
โˆ™ ¬๐‘ƒ ∧ ๐‘‡
5. Prove using Axioms 1.2.8.
(a) โŠข ๐‘ƒ → ๐‘ƒ
(b) โŠข ¬¬๐‘ƒ → ๐‘ƒ
(c) โŠข ๐‘ƒ → (๐‘ƒ → [๐‘„ → ๐‘ƒ ])
(d) ¬๐‘ƒ โŠข ๐‘ƒ → ๐‘„
(e) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘… โŠข ๐‘ƒ → ๐‘… (Do not use HS.)
(f) ๐‘ƒ → ๐‘„, ¬๐‘„ โŠข ¬๐‘ƒ (Do not use MT.)
31
Section 1.3 REPLACEMENT
6. Prove. Axioms 1.2.8 are not required.
(a) ๐‘ƒ → ๐‘„, ๐‘ƒ ∨ (๐‘… → ๐‘†), ¬๐‘„ โŠข ๐‘… → ๐‘†
(b) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ¬๐‘… โŠข ¬๐‘ƒ
(c) ๐‘ƒ → ๐‘„, ๐‘… → ๐‘†, ¬๐‘„ ∨ ¬๐‘† โŠข ¬๐‘ƒ ∨ ¬๐‘…
(d) [๐‘ƒ → (๐‘„ → ๐‘…)] ∧ [๐‘„ → (๐‘… → ๐‘ƒ )], ๐‘ƒ ∨ ๐‘„, ¬(๐‘„ → ๐‘…), ¬๐‘ƒ โŠข ¬๐‘…
(e) ๐‘ƒ → ๐‘„ → (๐‘… → ๐‘†), ๐‘† → ๐‘‡ , ๐‘ƒ → ๐‘„, ๐‘… โŠข ๐‘‡
(f) ๐‘ƒ → ๐‘„ ∧ ๐‘…, ๐‘„ ∨ ๐‘† → ๐‘‡ ∧ ๐‘ˆ , ๐‘ƒ โŠข ๐‘‡
(g) ๐‘ƒ → ๐‘„ ∧ ๐‘…, ¬(๐‘„ ∧ ๐‘…), ๐‘„ ∧ ๐‘… ∨ (¬๐‘ƒ → ๐‘†) โŠข ๐‘†
(h) ๐‘ƒ ∨ ๐‘„ → ¬๐‘… ∧ ¬๐‘†, ๐‘„ → ๐‘…, ๐‘ƒ โŠข ¬๐‘„
(i) ๐‘ → ๐‘ƒ , ๐‘ƒ ∨ ๐‘„ ∨ ๐‘… → ๐‘† ∨ ๐‘‡ , ๐‘† ∨ ๐‘‡ → ๐‘‡ , ๐‘ โŠข ๐‘‡
(j) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘… → ๐‘†, ๐‘† → ๐‘‡ , ๐‘ƒ ∨ ๐‘…, ¬๐‘… โŠข ๐‘‡
(k) ๐‘ƒ ∨ ๐‘„ → ๐‘… ∨ ๐‘†, (๐‘… → ๐‘‡ ) ∧ (๐‘† → ๐‘ˆ ), ๐‘ƒ , ¬๐‘‡ โŠข ๐‘ˆ
(l) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘… → ๐‘†, (๐‘ƒ ∨ ๐‘„) ∧ (๐‘… ∨ ๐‘†) โŠข ๐‘„ ∨ ๐‘†
(m) ๐‘ƒ ∨ ¬๐‘„ ∨ ๐‘… → (๐‘† → ๐‘ƒ ), ๐‘ƒ ∨ ¬๐‘„ → (๐‘ƒ → ๐‘…), ๐‘ƒ โŠข ๐‘† → ๐‘…
1.3
REPLACEMENT
There are times when writing a formal proof that we want to substitute one propositional
form for another. This happens when two propositional forms have the same valuations.
It also happens when a particular sentence pattern should be able to replace another
sentence pattern. We give rules in this section that codify both ideas.
Semantics
Consider the propositional form ¬(๐‘ƒ ∨ ๐‘„). Its valuation equals T when it is not the
case that ๐—(๐‘ƒ ) = T or ๐—(๐‘„) = T (Definition 1.1.9). This implies that ๐—(๐‘ƒ ) = F and
๐—(๐‘„) = F, so ๐—(¬๐‘ƒ ∧ ¬๐‘„) = T. Conversely, the valuation of ¬๐‘ƒ ∧ ¬๐‘„ is T implies that
the valuation of ¬(๐‘ƒ ∨ ๐‘„) is T for similar reasons. Since no additional premises were
assumed in this discussion, we conclude that
๐—(¬[๐‘ƒ ∨ ๐‘„]) = ๐—(¬๐‘ƒ ∧ ¬๐‘„),
(1.3)
and this means that
¬(๐‘ƒ ∨ ๐‘„) ↔ ¬๐‘ƒ ∧ ¬๐‘„
is a tautology. There is a name for this.
DEFINITION 1.3.1
Two propositional forms ๐‘ and ๐‘ž are logically equivalent if โŠจ ๐‘ ↔ ๐‘ž.
Observe that Definition 1.3.1 implies the following result.
THEOREM 1.3.2
All tautologies are logically equivalent, and all contradictions are logically equivalent.
32
Chapter 1 PROPOSITIONAL LOGIC
Because of (1.3), we can use a truth table to prove logical equivalence.
EXAMPLE 1.3.3
Prove: โŠจ ¬(๐‘ƒ ∨ ๐‘„) ↔ ¬๐‘ƒ ∧ ¬๐‘„.
๐‘ƒ
T
T
F
F
๐‘„ ๐‘ƒ ∨ ๐‘„ ¬(๐‘ƒ ∨ ๐‘„)
T
T
F
F
T
F
T
T
F
F
F
T
¬๐‘ƒ
F
F
T
T
¬๐‘„ ¬๐‘ƒ ∧ ¬๐‘„
F
F
T
F
F
F
T
T
EXAMPLE 1.3.4
Because โŠจ ๐‘ƒ → ๐‘„ ↔ ¬๐‘ƒ ∨ ๐‘„, we can replace ๐‘ƒ → ๐‘„ with ¬๐‘ƒ ∨ ๐‘„, and vice
versa, at any time. When this is done, the resulting propositional form is logically
equivalent to the original. For example,
โŠจ ๐‘„ ∧ (๐‘ƒ → ๐‘„) ↔ ๐‘„ ∧ (¬๐‘ƒ ∨ ๐‘„).
To see this, examine the truth table
๐‘ƒ
T
T
F
F
๐‘„ ๐‘ƒ → ๐‘„ ๐‘ธ ∧ (๐‘ท → ๐‘ธ) ¬๐‘ƒ
T
T
T
F
F
F
F
F
T
T
T
T
F
T
F
T
¬๐‘ƒ ∨ ๐‘„ ๐‘ธ ∧ (¬๐‘ท ∨ ๐‘ธ)
T
T
F
F
T
T
T
F
EXAMPLE 1.3.5
Consider ๐‘… ∧ (๐‘ƒ → ๐‘„). Since โŠจ ๐‘ƒ → ๐‘„ ↔ ¬๐‘„ → ¬๐‘ƒ [Exercise 2(f)], we can
replace ๐‘ƒ → ๐‘„ with ¬๐‘„ → ¬๐‘ƒ giving
โŠจ ๐‘… ∧ (๐‘ƒ → ๐‘„) ↔ ๐‘… ∧ (¬๐‘„ → ¬๐‘ƒ ).
When studying an implication, we sometimes need to investigate the different ways
that its antecedent and consequent relate to each other.
DEFINITION 1.3.6
The converse of a given implication is the conditional proposition formed by
exchanging the antecedent and consequent of the implication (Figure 1.2).
DEFINITION 1.3.7
The contrapositive of a given implication is the conditional proposition formed
by exchanging the antecedent and consequent of the implication and then replacing them with their negations (Figure 1.3).
Section 1.3 REPLACEMENT
If
If
the antecedent,
the consequent,
Figure 1.2
If
If
then
the consequent.
then
the antecedent.
Writing the converse.
the antecedent,
then
not the consequent,
then
Figure 1.3
33
the consequent.
not the antecedent.
Writing the contrapositive.
For example, the converse of
if rectangles have four sides, squares have for sides
is
if squares have four sides, rectangles have four sides,
and its contrapositive is
if squares do not have four sides, rectangles do not have four sides.
Notice that a biconditional proposition is simply the conjunction of a conditional with
its converse.
EXAMPLE 1.3.8
The propositional form ๐‘ƒ → ๐‘„ has ๐‘„ → ๐‘ƒ as its converse and ¬๐‘„ → ¬๐‘ƒ as its
contrapositive. The first and fourth columns on the right of the next truth table
show โŠจ ๐‘ƒ → ๐‘„ ↔ ¬๐‘„ → ¬๐‘ƒ , while ฬธโŠจ ๐‘ƒ → ๐‘„ ↔ ๐‘„ → ๐‘ƒ is shown by the first
and last columns.
๐‘ƒ ๐‘„ ๐‘ƒ → ๐‘„ ¬๐‘„ ¬๐‘ƒ ¬๐‘„ → ¬๐‘ƒ ๐‘„ → ๐‘ƒ
T T
T
F
F
T
T
T F
F
T
F
F
T
F T
T
F
T
T
F
F F
T
T
T
T
T
34
Chapter 1 PROPOSITIONAL LOGIC
Syntactics
If we limit our proofs to Inference Rules 1.2.10, we quickly realize that there will be
little of interest that we can prove. We would have no reason on which to base such
clear inferences as
๐‘ƒ โŠข๐‘„∨๐‘ƒ
or
๐‘ƒ ∨ ๐‘„, ¬๐‘„ โŠข ๐‘ƒ .
To fix this, we expand our collection of inference rules with a new type.
Suppose that we know that the form ๐‘ ∧ ๐‘ž can replace ๐‘ž ∧ ๐‘ at any time and vice
versa. For example, in the propositional form
๐‘ƒ ∧ ๐‘„ → ๐‘…,
(1.4)
๐‘ƒ ∧ ๐‘„ can be replaced with ๐‘„ ∧ ๐‘ƒ so that we can write the new form
๐‘„ ∧ ๐‘ƒ → ๐‘….
(1.5)
This type of rule is called a replacement rule and is written using the ⇔ symbol. For
example, the replacement rule that allowed us to write (1.5) from (1.4) is
๐‘ ∧ ๐‘ž ⇔ ๐‘ž ∧ ๐‘.
Similarly, when the replacement rule
¬(๐‘ ∧ ๐‘ž) ⇔ ¬๐‘ ∨ ¬๐‘ž
is applied to ๐‘ƒ ∨ (¬๐‘„ ∨ ¬๐‘…), the result is ๐‘ƒ ∨ ¬(๐‘„ ∧ ๐‘…). We state without proof the
standard replacement rules.
REPLACEMENT RULES 1.3.9
Let ๐‘, ๐‘ž, and ๐‘Ÿ be propositional forms.
โˆ™ Associative Laws [Assoc]
๐‘ ∧ ๐‘ž ∧ ๐‘Ÿ ⇔ ๐‘ ∧ (๐‘ž ∧ ๐‘Ÿ)
๐‘ ∨ ๐‘ž ∨ ๐‘Ÿ ⇔ ๐‘ ∨ (๐‘ž ∨ ๐‘Ÿ)
โˆ™ Commutative Laws [Com]
๐‘∧๐‘ž ⇔๐‘ž∧๐‘
๐‘∨๐‘ž ⇔๐‘ž∨๐‘
โˆ™ Distributive Laws [Distr]
๐‘ ∧ (๐‘ž ∨ ๐‘Ÿ) ⇔ ๐‘ ∧ ๐‘ž ∨ ๐‘ ∧ ๐‘Ÿ
๐‘ ∨ ๐‘ž ∧ ๐‘Ÿ ⇔ (๐‘ ∨ ๐‘ž) ∧ (๐‘ ∨ ๐‘Ÿ)
โˆ™ Contrapositive Law [Contra]
๐‘ → ๐‘ž ⇔ ¬๐‘ž → ¬๐‘
Section 1.3 REPLACEMENT
35
โˆ™ Double Negation [DN]
๐‘ ⇔ ¬¬๐‘
โˆ™ De Morgan’s Laws [DeM]
¬(๐‘ ∧ ๐‘ž) ⇔ ¬๐‘ ∨ ¬๐‘ž
¬(๐‘ ∨ ๐‘ž) ⇔ ¬๐‘ ∧ ¬๐‘ž
โˆ™ Idempotency [Idem]
๐‘∧๐‘⇔๐‘
๐‘∨๐‘⇔๐‘
โˆ™ Material Equivalence [Equiv]
๐‘ ↔ ๐‘ž ⇔ (๐‘ → ๐‘ž) ∧ (๐‘ž → ๐‘)
๐‘ ↔ ๐‘ž ⇔ ๐‘ ∧ ๐‘ž ∨ ¬๐‘ ∧ ¬๐‘ž
โˆ™ Material Implication [Impl]
๐‘ → ๐‘ž ⇔ ¬๐‘ ∨ ๐‘ž
โˆ™ Exportation [Exp]
๐‘ ∧ ๐‘ž → ๐‘Ÿ ⇔ ๐‘ → (๐‘ž → ๐‘Ÿ).
A replacement rule is used in a formal proof by appealing to the next inference rule.
INFERENCE RULE 1.3.10
For all propositional forms ๐‘ and ๐‘ž, if ๐‘ if obtained from ๐‘ž using a replacement
rule, ๐‘ ⇒ ๐‘ž and ๐‘ž ⇒ ๐‘.
As with Inference Rules 1.2.10, the replacement rule used when applying Inference
Rule 1.3.10 must be used exactly as stated. This includes times when it seems unnecessary because it appears obvious. For example, ๐‘ƒ ∨ ๐‘„ does not follow directly from
¬๐‘ƒ → ๐‘„ using Impl. Instead, include a use of DN to give the correct sequence,
¬๐‘ƒ → ๐‘„, ¬¬๐‘ƒ ∨ ๐‘„, ๐‘ƒ ∨ ๐‘„.
Similarly, the inference rule Add does not allow for ๐‘„∨๐‘ƒ to be derived from ๐‘ƒ . Instead,
we derive ๐‘ƒ ∨ ๐‘„. Follow this by Com to conclude ๐‘„ ∨ ๐‘ƒ .
When writing formal proofs that appeal to Inference Rule 1.3.10, do not cite that
particular rule but reference the replacement rule’s abbreviation using Replacement
Rule 1.3.9 and the line which serves as a premise to the replacement. We use this
practice in the following examples.
EXAMPLE 1.3.11
Although it is common practice to move parentheses freely when solving equations, inferences such as
๐‘ƒ ∧ ๐‘„ ∧ (๐‘… ∧ ๐‘†) โŠข ๐‘ƒ ∧ (๐‘„ ∧ ๐‘…) ∧ ๐‘†
36
Chapter 1 PROPOSITIONAL LOGIC
must be carefully demonstrated. Fortunately, this example only requires two applications of the associative law. Using the boxes as a guide, notice that
๐‘ƒ ∧๐‘„ ∧( ๐‘… ∧ ๐‘† )
is of the same form as the right-hand side of the associative law. Hence, we can
remove the parentheses to obtain
๐‘ƒ ∧๐‘„ ∧ ๐‘… ∧ ๐‘† .
Next, view the propositional form as
๐‘ƒ ∧๐‘„∧๐‘… ∧ ๐‘† .
One more application within the first box yields the result,
๐‘ƒ ∧ (๐‘„ ∧ ๐‘…) ∧ ๐‘† .
We may, therefore, write a sequence of inferences by Inference Rule 1.3.10,
๐‘ƒ ∧ ๐‘„ ∧ (๐‘… ∧ ๐‘†) ⇒ (๐‘ƒ ∧ ๐‘„ ∧ ๐‘…) ∧ ๐‘† ⇒ [๐‘ƒ ∧ (๐‘„ ∧ ๐‘…)] ∧ ๐‘†,
and we have a proof of [๐‘ƒ ∧ (๐‘„ ∧ ๐‘…)] ∧ ๐‘† from ๐‘ƒ ∧ ๐‘„ ∧ (๐‘… ∧ ๐‘†):
1.
2.
3.
๐‘ƒ ∧ ๐‘„ ∧ (๐‘… ∧ ๐‘†)
๐‘ƒ ∧๐‘„∧๐‘…∧๐‘†
๐‘ƒ ∧ (๐‘„ ∧ ๐‘…) ∧ ๐‘†
Given
1 Assoc
2 Assoc
EXAMPLE 1.3.12
Prove: ๐‘… ∧ ๐‘† โŠข ¬๐‘… → ๐‘ƒ
1.
2.
3.
4.
5.
๐‘…∧๐‘†
๐‘…
๐‘…∨๐‘ƒ
¬¬๐‘… ∨ ๐‘ƒ
¬๐‘… → ๐‘ƒ
Given
1 Simp
2 Add
3 DN
4 Impl
EXAMPLE 1.3.13
Prove: ๐‘ƒ → ๐‘„, ๐‘… → ๐‘„ โŠข ๐‘ƒ ∨ ๐‘… → ๐‘„
1.
2.
3.
4.
5.
๐‘ƒ →๐‘„
๐‘…→๐‘„
(๐‘ƒ → ๐‘„) ∧ (๐‘… → ๐‘„)
(¬๐‘ƒ ∨ ๐‘„) ∧ (¬๐‘… ∨ ๐‘„)
(๐‘„ ∨ ¬๐‘ƒ ) ∧ (๐‘„ ∨ ¬๐‘…)
Given
Given
1, 2 Conj
3 Impl
4 Com
Section 1.3 REPLACEMENT
6.
7.
8.
9.
๐‘„ ∨ ¬๐‘ƒ ∧ ¬๐‘…
¬๐‘ƒ ∧ ¬๐‘… ∨ ๐‘„
¬(๐‘ƒ ∨ ๐‘…) ∨ ๐‘„
๐‘ƒ ∨๐‘…→๐‘„
37
5 Dist
6 Com
7 DeM
8 Impl
EXAMPLE 1.3.14
Prove: ๐‘ƒ ∧ ๐‘„ ∨ ๐‘… ∧ ๐‘† โŠข (๐‘ƒ ∨ ๐‘†) ∧ (๐‘„ ∨ ๐‘…)
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
๐‘ƒ ∧๐‘„∨๐‘…∧๐‘†
(๐‘ƒ ∧ ๐‘„ ∨ ๐‘…) ∧ (๐‘ƒ ∧ ๐‘„ ∨ ๐‘†)
(๐‘… ∨ ๐‘ƒ ∧ ๐‘„) ∧ (๐‘† ∨ ๐‘ƒ ∧ ๐‘„)
(๐‘… ∨ ๐‘ƒ ) ∧ (๐‘… ∨ ๐‘„) ∧ [(๐‘† ∨ ๐‘ƒ ) ∧ (๐‘† ∨ ๐‘„)]
(๐‘… ∨ ๐‘„) ∧ (๐‘… ∨ ๐‘ƒ ) ∧ [(๐‘† ∨ ๐‘ƒ ) ∧ (๐‘† ∨ ๐‘„)]
(๐‘… ∨ ๐‘„) ∧ ([๐‘… ∨ ๐‘ƒ ] ∧ [(๐‘† ∨ ๐‘ƒ ) ∧ (๐‘† ∨ ๐‘„)])
๐‘…∨๐‘„
(๐‘† ∨ ๐‘ƒ ) ∧ (๐‘† ∨ ๐‘„) ∧ [(๐‘… ∨ ๐‘„) ∧ (๐‘… ∨ ๐‘ƒ )]
(๐‘† ∨ ๐‘ƒ ) ∧ (๐‘† ∨ ๐‘„)
๐‘† ∨๐‘ƒ
(๐‘† ∨ ๐‘ƒ ) ∧ (๐‘… ∨ ๐‘„)
(๐‘ƒ ∨ ๐‘†) ∧ (๐‘„ ∨ ๐‘…)
Given
1 Dist
2 Com
3 Dist
4 Com
5 Assoc
6 Simp
5 Com
8 Simp
9 Simp
7, 10 Conj
11 Com
EXAMPLE 1.3.15
Both โŠข ๐‘ƒ → ๐‘ƒ and โŠข ๐‘ƒ ∨ ¬๐‘ƒ . Here is the proof for the second theorem:
1.
2.
3.
4.
5.
๐‘ƒ → (๐‘ƒ → ๐‘ƒ )
¬๐‘ƒ ∨ (¬๐‘ƒ ∨ ๐‘ƒ )
(¬๐‘ƒ ∨ ¬๐‘ƒ ) ∨ ๐‘ƒ
¬๐‘ƒ ∨ ๐‘ƒ
๐‘ƒ ∨ ¬๐‘ƒ
FL1
1 Impl
2 Assoc
3 Idem
4 Com
This proof can be generalized to any propositional form ๐‘, so that we also have
โŠข ๐‘ → ๐‘ and โŠข ๐‘ ∨ ¬๐‘.
The theorem ๐‘ ∨ ¬๐‘ is known as the law of the excluded middle, and the theorem
¬(๐‘ ∧ ¬๐‘) is the law of noncontradiction. Notice that by De Morgan’s law with Com
and DN, we have for all propositional forms ๐‘
โŠข (๐‘ ∨ ¬๐‘) ↔ ¬(๐‘ ∧ ¬๐‘).
Exercises
1. The inverse of a given implication is the contrapositive of the implication’s converse. Write the converse, contrapositive, and inverse for each conditional proposition
in Exercise 1.1.2.
38
Chapter 1 PROPOSITIONAL LOGIC
2. Prove using truth tables.
(a) โŠจ ๐‘ƒ ∨ ¬๐‘ƒ ↔ ๐‘ƒ → ๐‘ƒ
(b) โŠจ ๐‘ƒ ∨ ¬๐‘ƒ ↔ ๐‘ƒ ∨ ๐‘„ ∨ ¬(๐‘ƒ ∧ ๐‘„)
(c) โŠจ ๐‘ƒ ∧ ๐‘„ ↔ (๐‘ƒ ↔ ๐‘„) ∧ (๐‘ƒ ∨ ๐‘„)
(d) โŠจ ๐‘… ∧ (๐‘ƒ → ๐‘„) ↔ ๐‘… ∧ (¬๐‘„ → ¬๐‘ƒ )
(e) โŠจ ๐‘ƒ ∧ ๐‘„ → ๐‘… ↔ ๐‘ƒ ∧ ¬๐‘… → ¬๐‘„
(f) โŠจ ๐‘ƒ → ๐‘„ ↔ ¬๐‘„ → ¬๐‘ƒ
(g) โŠจ ๐‘ƒ → ๐‘„ ∧ ๐‘… ↔ (๐‘ƒ → ๐‘„) ∧ (๐‘ƒ → ๐‘…)
(h) โŠจ ๐‘ƒ → ๐‘„ → (๐‘† → ๐‘…) ↔ (๐‘ƒ → ๐‘„) ∧ ๐‘† → ๐‘…
3. A propositional form is in disjunctive normal form if it is a disjunction of conjunctions. For example, the propositional form ๐‘ƒ ∧ (¬๐‘„ ∨ ๐‘…) is logically equivalent
to
(๐‘ƒ ∧ ๐‘„ ∧ ๐‘…) ∨ (๐‘ƒ ∧ ¬๐‘„ ∧ ¬๐‘…) ∨ (¬๐‘ƒ ∧ ๐‘„ ∧ ¬๐‘…),
which is in disjunctive normal form. Find propositional forms in disjunctive normal
form that are logically equivalent to each of the following.
(a) ๐‘ƒ ∨ ๐‘„ ∧ (๐‘ƒ ∨ ¬๐‘…)
(b) (๐‘ƒ ∨ ๐‘„) ∧ (¬๐‘ƒ ∨ ¬๐‘„)
(c) (๐‘ƒ ∧ ¬๐‘„ ∨ ๐‘…) ∧ (๐‘„ ∧ ๐‘… ∨ ๐‘ƒ ∧ ¬๐‘…)
(d) ๐‘ƒ ∨ (¬๐‘„ ∨ [๐‘ƒ ∧ ¬๐‘… ∨ ๐‘ƒ ∧ ¬๐‘„])
4. Identify the rule from Replacement Rule 1.3.9.
(a) (๐‘ƒ → ๐‘„) ∨ (๐‘„ → ๐‘…) ∨ ๐‘† ⇔ (๐‘ƒ → ๐‘„) ∨ ([๐‘„ → ๐‘…] ∨ ๐‘†)
(b) ¬¬๐‘ƒ ↔ ๐‘„ ∧ ๐‘… ⇔ ๐‘ƒ ↔ ๐‘„ ∧ ๐‘…
(c) ๐‘ƒ ∨ ๐‘„ ∨ ๐‘… ⇔ ๐‘„ ∨ ๐‘ƒ ∨ ๐‘…
(d) ¬๐‘ƒ ∧ ¬(๐‘„ ∨ ๐‘…) ⇔ ¬(๐‘ƒ ∨ [๐‘„ ∨ ๐‘…])
(e) ¬(๐‘ƒ ∨ ๐‘„) ∨ ๐‘… ⇔ ๐‘ƒ ∨ ๐‘„ → ๐‘…
(f) ๐‘ƒ ↔ ๐‘„ → ๐‘… ⇔ ๐‘ƒ ∧ (๐‘„ → ๐‘…) ∨ ¬๐‘ƒ ∧ ¬(๐‘„ → ๐‘…)
(g) ๐‘ƒ ∨ ๐‘„ ↔ ๐‘„ ∧ ๐‘„ ⇔ ๐‘ƒ ∨ ๐‘„ ↔ ๐‘„
(h) ๐‘ƒ ∧ ๐‘„ ∧ ๐‘… ⇔ ๐‘… ∧ (๐‘ƒ ∧ ๐‘„)
(i) (๐‘ƒ ∨ ๐‘„) ∧ (๐‘„ ∨ ๐‘…) ⇔ (๐‘ƒ ∨ ๐‘„) ∧ ๐‘„ ∨ (๐‘ƒ ∨ ๐‘„) ∧ ๐‘…
(j) (๐‘ƒ ∧ [๐‘… → ๐‘„]) → ๐‘† ⇔ ๐‘ƒ → (๐‘… → ๐‘„ → ๐‘†)
5. For each given propositional form ๐‘, find another propositional form ๐‘ž such that
๐‘ ⇔ ๐‘ž using Replacement Rules 1.3.9.
(a) ¬¬๐‘ƒ
(b) ๐‘ƒ ∨ ๐‘„
(c) ๐‘ƒ → ๐‘„
(d) ¬(๐‘ƒ ∧ ๐‘„)
(e) (๐‘ƒ ↔ ๐‘„) ∧ (¬๐‘ƒ ↔ ๐‘„)
(f) (๐‘ƒ → ๐‘„) ∨ (๐‘„ → ๐‘†)
(g) (๐‘ƒ → ๐‘„) ∨ ๐‘ƒ
(h) ¬(๐‘ƒ → ๐‘„)
(i) (๐‘ƒ → ๐‘„) ∧ (๐‘„ ↔ ๐‘…)
Section 1.3 REPLACEMENT
39
(j) (๐‘ƒ ∨ ¬๐‘„) ↔ ๐‘‡ ∧ ๐‘„
(k) ๐‘ƒ ∨ (¬๐‘„ ↔ ๐‘‡ ) ∧ ๐‘„
6. Arrange each collection of propositional forms into a proof for the given deduction
and supply the appropriate reasons.
(a) ๐‘ƒ ∨ ๐‘„ → ๐‘… โŠข (๐‘ƒ → ๐‘…) ∧ (๐‘„ → ๐‘…)
โˆ™ ๐‘… ∨ ¬๐‘ƒ ∧ ¬๐‘„
โˆ™ (¬๐‘ƒ ∨ ๐‘…) ∧ (¬๐‘„ ∨ ๐‘…)
โˆ™ ¬(๐‘ƒ ∨ ๐‘„) ∨ ๐‘…
โˆ™ (๐‘ƒ → ๐‘…) ∧ (๐‘„ → ๐‘…)
โˆ™ (๐‘… ∨ ¬๐‘ƒ ) ∧ (๐‘… ∨ ¬๐‘„)
โˆ™ ¬๐‘ƒ ∧ ¬๐‘„ ∨ ๐‘…
โˆ™ ๐‘ƒ ∨๐‘„→๐‘…
(b) ¬(๐‘ƒ ∧ ๐‘„) → ๐‘… ∨ ๐‘†, ¬๐‘ƒ , ¬๐‘† โŠข ๐‘…
โˆ™ ¬(๐‘ƒ ∧ ๐‘„) → ๐‘… ∨ ๐‘†
โˆ™ ¬๐‘†
โˆ™ ๐‘† ∨๐‘…
โˆ™ ๐‘…
โˆ™ ¬(๐‘ƒ ∧ ๐‘„)
โˆ™ ¬๐‘ƒ
โˆ™ ๐‘…∨๐‘†
โˆ™ ¬๐‘ƒ ∨ ¬๐‘„
(c) ๐‘ƒ → (๐‘„ → ๐‘…), ¬๐‘ƒ → ๐‘†, ¬๐‘„ → ๐‘‡ , ๐‘… → ¬๐‘… โŠข ¬๐‘‡ → ๐‘†
โˆ™ ๐‘† ∨๐‘‡
โˆ™ ¬๐‘… ∨ ¬๐‘…
โˆ™ ๐‘… → ¬๐‘…
โˆ™ ¬๐‘…
โˆ™ ¬(๐‘ƒ ∧ ๐‘„)
โˆ™ ๐‘‡ ∨๐‘†
โˆ™ ¬¬๐‘‡ ∨ ๐‘†
โˆ™ ¬๐‘ƒ ∨ ¬๐‘„
โˆ™ ๐‘ƒ ∧๐‘„→๐‘…
โˆ™ ๐‘ƒ → (๐‘„ → ๐‘…)
โˆ™ ¬๐‘ƒ → ๐‘†
โˆ™ ¬๐‘„ → ๐‘‡
โˆ™ ¬๐‘‡ → ๐‘†
โˆ™ (¬๐‘ƒ → ๐‘†) ∧ (¬๐‘„ → ๐‘‡ )
7. Prove.
(a) ¬๐‘ƒ โŠข ๐‘ƒ → ๐‘„
(b) ๐‘ƒ โŠข ¬๐‘„ → ๐‘ƒ
(c) ¬๐‘„ ∨ (¬๐‘… ∨ ¬๐‘ƒ ) โŠข ๐‘ƒ → ¬(๐‘„ ∧ ๐‘…)
(d) ๐‘ƒ → ๐‘„ โŠข ๐‘ƒ ∧ ๐‘… → ๐‘„
40
Chapter 1 PROPOSITIONAL LOGIC
(e)
(f)
(g)
(h)
(i)
(j)
(k)
(l)
(m)
(n)
(o)
(p)
(q)
(r)
(s)
(t)
(u)
(v)
(w)
(x)
(y)
(z)
1.4
๐‘ƒ →๐‘„∧๐‘…โŠข๐‘ƒ →๐‘„
๐‘ƒ ∨ ๐‘„ → ๐‘… โŠข ¬๐‘… → ¬๐‘„
๐‘ƒ → (๐‘„ → ๐‘…) โŠข ๐‘„ ∧ ¬๐‘… → ¬๐‘ƒ
๐‘ƒ → (๐‘„ → ๐‘…) โŠข ๐‘„ → (๐‘ƒ → ๐‘…)
๐‘ƒ ∧ ๐‘„ ∨ ๐‘… ∧ ๐‘† โŠข ¬๐‘† → ๐‘ƒ ∧ ๐‘„
๐‘„ → ๐‘… โŠข ๐‘ƒ → (๐‘„ → ๐‘…)
๐‘ƒ → ¬(๐‘„ → ๐‘…) โŠข ๐‘ƒ → ¬๐‘…
๐‘ƒ ∨ (๐‘„ ∨ ๐‘… ∨ ๐‘†) โŠข (๐‘ƒ ∨ ๐‘„) ∨ (๐‘… ∨ ๐‘†)
๐‘„∨๐‘ƒ →๐‘…∧๐‘† โŠข๐‘„→๐‘…
๐‘ƒ ↔๐‘„∧๐‘…โŠข๐‘ƒ →๐‘„
๐‘ƒ ↔๐‘„∨๐‘…โŠข๐‘„→๐‘ƒ
๐‘ƒ ∨๐‘„∨๐‘…→๐‘† โŠข๐‘„→๐‘†
(๐‘ƒ ∨ ๐‘„) ∧ (๐‘… ∨ ๐‘†) โŠข ๐‘ƒ ∧ ๐‘… ∨ ๐‘ƒ ∧ ๐‘† ∨ (๐‘„ ∧ ๐‘… ∨ ๐‘„ ∧ ๐‘†)
๐‘ƒ ∧ (๐‘„ ∨ ๐‘…) → ๐‘„ ∧ ๐‘… โŠข ๐‘ƒ → (๐‘„ → ๐‘…)
๐‘ƒ ↔ ๐‘„, ¬๐‘ƒ โŠข ¬๐‘„
๐‘ƒ → ๐‘„ → ๐‘…, ¬๐‘… โŠข ¬๐‘„
๐‘ƒ → (๐‘„ → ๐‘…), ๐‘… → ๐‘† ∧ ๐‘‡ โŠข ๐‘ƒ → (๐‘„ → ๐‘‡ )
๐‘ƒ → (๐‘„ → ๐‘…), ๐‘… → ๐‘† ∨ ๐‘‡ โŠข (๐‘ƒ → ๐‘†) ∨ (๐‘„ → ๐‘‡ )
๐‘ƒ → ๐‘„, ๐‘ƒ → ๐‘… โŠข ๐‘ƒ → ๐‘„ ∧ ๐‘…
๐‘ƒ ∨ ๐‘„ → ๐‘… ∧ ๐‘†, ¬๐‘ƒ → (๐‘‡ → ¬๐‘‡ ), ¬๐‘… โŠข ¬๐‘‡
(๐‘ƒ → ๐‘„) ∧ (๐‘… → ๐‘†), ๐‘ƒ ∨ ๐‘…, (๐‘ƒ → ¬๐‘†) ∧ (๐‘… → ¬๐‘„) โŠข ๐‘„ ↔ ¬๐‘†
๐‘ƒ ∧ (๐‘„ ∧ ๐‘…), ๐‘ƒ ∧ ๐‘… → ๐‘† ∨ (๐‘‡ ∨ ๐‘€), ¬๐‘† ∧ ¬๐‘‡ โŠข ๐‘€
PROOF METHODS
The methods of Sections 1.2 and 1.3 provide a good start for writing formal proofs.
However, in practice we rarely limit ourselves to these rules. We often use inference
rules that give a straightforward way to prove conditional propositions and allow us to
prove a proposition when it is easier to disprove its negation. In both the cases, the new
inference rules will be justified using the rules we already know.
Deduction Theorem
Because of Axioms 1.2.8, not all of the inference rules are needed to write the proofs
found in Sections 1.2 and 1.3. This motivates the next definition.
DEFINITION 1.4.1
Let ๐‘ and ๐‘ž be propositional forms. The notation
๐‘ โŠข∗ ๐‘ž
means that there exists a formal proof of ๐‘ž from ๐‘ using only Axioms 1.2.8, MP,
and Inference Rule 1.3.10, and the notation
โŠข∗ ๐‘ž
Section 1.4 PROOF METHODS
41
means that there exists a formal proof of ๐‘ž from Axioms 1.2.8 using only MP and
Inference Rule 1.3.10
For example, ๐‘ƒ , ๐‘„, ¬๐‘ƒ ∨ (๐‘„ → ๐‘…) โŠข∗ ๐‘… because
1.
2.
3.
4.
5.
6.
๐‘ƒ
๐‘„
¬๐‘ƒ ∨ (๐‘„ → ๐‘…)
๐‘ƒ → (๐‘„ → ๐‘…)
๐‘„→๐‘…
๐‘…
Given
Given
Given
3 Impl
1, 4 MP
2, 5 MP
is a formal proof using MP and Impl as the only inference rules.
We now observe that the propositional forms that can be proved using the full collection of rules from Sections 1.2 and 1.3 are exactly the propositional forms that can
be proved when all rules are deleted from Inference Rules 1.2.10 except for MP.
THEOREM 1.4.2
For all propositional forms ๐‘ and ๐‘ž, ๐‘ โŠข∗ ๐‘ž if and only if ๐‘ โŠข ๐‘ž.
PROOF
Trivially, ๐‘ โŠข∗ ๐‘ž implies ๐‘ โŠข ๐‘ž, so suppose that ๐‘ โŠข ๐‘ž. We show that the remaining parts of Inference Rules 1.2.10 are equivalent to using only Axioms 1.2.8,
MP, and Replacement Rules 1.3.9. We show three examples and leave the proofs
of the remaining inference rules to Exercise 6. The proof
1.
2.
3.
4.
5.
6.
๐‘→๐‘ž
¬๐‘ž
¬¬๐‘ → ¬¬๐‘ž
¬¬๐‘ → ¬¬๐‘ž → (¬๐‘ž → ¬๐‘)
¬๐‘ž → ¬๐‘
¬๐‘
Given
Given
DN
FL3
3, 4 MP
2, 5 MP
shows that
๐‘ → ๐‘ž, ¬๐‘ž โŠข∗ ¬๐‘.
Thus, we do not need MT. The proof
1.
2.
3.
4.
5.
6.
๐‘
๐‘ → (¬๐‘ž → ๐‘)
¬๐‘ž → ๐‘
¬¬๐‘ž ∨ ๐‘
๐‘ž∨๐‘
๐‘∨๐‘ž
shows that
๐‘ โŠข∗ ๐‘ ∨ ๐‘ž,
Given
FL1
1, 2 MP
3 Impl
4 DN
5 Com
42
Chapter 1 PROPOSITIONAL LOGIC
so we do not need Add. This implies that we can use Add to demonstrate that
๐‘ → ๐‘ž, ๐‘ž → ๐‘Ÿ โŠข∗ ๐‘ → ๐‘Ÿ.
The proof is as follows:
1.
2.
3.
4.
5.
6.
7.
8.
9.
๐‘→๐‘ž
๐‘ž→๐‘Ÿ
¬๐‘ž ∨ ๐‘Ÿ
¬๐‘ž ∨ ๐‘Ÿ ∨ ¬๐‘
¬๐‘ ∨ (¬๐‘ž ∨ ๐‘Ÿ)
๐‘ → (๐‘ž → ๐‘Ÿ)
๐‘ → (๐‘ž → ๐‘Ÿ) → (๐‘ → ๐‘ž → [๐‘ → ๐‘Ÿ])
๐‘ → ๐‘ž → (๐‘ → ๐‘Ÿ)
๐‘→๐‘Ÿ
Given
Given
2 Impl
3 Add
4 Com
5 Impl
FL2
6, 7 MP
1, 8 MP
This implies that we do not need HS.
At this point, there is an obvious question: If only MP is needed from Inference
Rules 1.2.10, why were the other rules included? The answer is because the other
inference rules are examples of common reasoning and excluding them would introduce
unnecessary complications to the formal proofs. Try reproving some of the deductions
of Section 1.2 with only MP and the axioms to confirm this.
When a formal proof in Section 1.3 involved proving an implication, the replacement
rule Impl would often appear in the proof. However, as we know from geometry, this
is not the typical strategy used to prove an implication. What is usually done is that the
antecedent is assumed and then the consequent is shown to follow. That this procedure
justifies the given conditional is the next theorem. Its proof requires a lemma.
LEMMA 1.4.3
Let ๐‘ and ๐‘ž be propositional forms. If โŠข∗ ๐‘ž, then โŠข∗ ๐‘ → ๐‘ž.
PROOF
Let โŠข∗ ๐‘ž. By FL1, we have that โŠข∗ ๐‘ž → (๐‘ → ๐‘ž), so โŠข∗ ๐‘ → ๐‘ž follows by MP.
THEOREM 1.4.4 [Deduction]
For all propositional forms ๐‘ and ๐‘ž, ๐‘ โŠข ๐‘ž if and only if โŠข ๐‘ → ๐‘ž.
PROOF
Let ๐‘ and ๐‘ž be propositional forms. Assume that โŠข ๐‘ → ๐‘ž, so there exists propositional forms ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘›−1 such that
๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘›−1 , ๐‘ → ๐‘ž
is a proof, where ๐‘Ÿ0 is an axiom, ๐‘Ÿ๐‘– is an axiom or ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘–−1 ⇒ ๐‘Ÿ๐‘– for ๐‘– > 0,
and ๐‘ → ๐‘ž follows from ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘›−1 . Then,
Section 1.4 PROOF METHODS
๐‘, ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘›−1 , ๐‘ → ๐‘ž, ๐‘ž
is also a proof, where the last inference is due to MP. Therefore, ๐‘ โŠข ๐‘ž.
By Theorem 1.4.2, to prove the converse, we only need to prove that
if ๐‘ โŠข∗ ๐‘ž, then โŠข∗ ๐‘ → ๐‘ž.
Assume ๐‘ โŠข∗ ๐‘ž. First note that if ๐‘ž is an axiom, then โŠข∗ ๐‘ž, so โŠข∗ ๐‘ → ๐‘ž by
Lemma 1.4.3. Therefore, assume that ๐‘ž is not an axiom. We begin by checking
four cases.
โˆ™ Suppose that the proof has only one propositional form. In this case, we
have that ๐‘ = ๐‘ž, so the inference is of the form ๐‘ โŠข∗ ๐‘. By FL1,
โŠข∗ ๐‘ → (๐‘ → ๐‘).
Because
๐‘ → (๐‘ → ๐‘) ⇒ ¬๐‘ ∨ (¬๐‘ ∨ ๐‘)
⇒ (¬๐‘ ∨ ¬๐‘) ∨ ๐‘
⇒ ¬๐‘ ∨ ๐‘
⇒ ๐‘ → ๐‘,
we conclude that โŠข∗ ๐‘ → ๐‘.
โˆ™ Next, suppose the proof has two propositional forms and cannot be reduced
to the first case. This implies that ๐‘ โŠข∗ ๐‘ž by a single application of a
replacement rule. Thus,
๐‘ → (¬๐‘ž → ๐‘) โŠข∗ ๐‘ → (¬๐‘ž → ๐‘ž)
by a single application of the same replacement rule. Therefore,
๐‘ → (¬๐‘ž → ๐‘) ⇒ ๐‘ → (¬๐‘ž → ๐‘ž)
⇒ ๐‘ → (¬¬๐‘ž ∨ ๐‘ž)
⇒ ๐‘ → (๐‘ž ∨ ๐‘ž)
⇒ ๐‘ → ๐‘ž.
This implies that โŠข∗ ๐‘ → ๐‘ž.
โˆ™ We now consider the case when the proof of ๐‘ โŠข∗ ๐‘ž has three propositional
forms and ๐‘ž follows by a rule of replacement. Let ๐‘, ๐‘Ÿ, ๐‘ž be the proof. This
implies that ๐‘ โŠข∗ ๐‘Ÿ, which implies
โŠข∗ ๐‘ → ๐‘Ÿ
(1.6)
because either ๐‘Ÿ is an axiom and Lemma 1.4.3 applies or the previous two
cases apply. If ๐‘ž follows from ๐‘ by a rule of replacement, then โŠข ๐‘ → ๐‘ž by
43
44
Chapter 1 PROPOSITIONAL LOGIC
the previous case, so assume that ๐‘ž follows from ๐‘Ÿ by a rule of replacement.
Thus, โŠข∗ ๐‘Ÿ → ๐‘ž, which implies by Lemma 1.4.3 that
โŠข∗ ๐‘ → (๐‘Ÿ → ๐‘ž).
(1.7)
โŠข∗ ๐‘ → (๐‘Ÿ → ๐‘ž) → (๐‘ → ๐‘Ÿ → [๐‘ → ๐‘ž]).
(1.8)
By FL2,
Therefore, by (1.7) and (1.8) with MP,
โŠข∗ ๐‘ → ๐‘Ÿ → (๐‘ → ๐‘ž),
(1.9)
and by (1.6) and (1.9) with MP,
โŠข∗ ๐‘ → ๐‘ž.
โˆ™ Again, let the proof have three propositional forms and write it as ๐‘, ๐‘ , ๐‘ž.
Suppose that the inference that leads to ๐‘ž is MP. This means that ๐‘Ÿ and
๐‘Ÿ → ๐‘ž are in the proof. Because either ๐‘ = ๐‘Ÿ and ๐‘ โŠข∗ ๐‘Ÿ → ๐‘ž or ๐‘ = ๐‘Ÿ → ๐‘ž
and ๐‘ โŠข∗ ๐‘Ÿ,
โŠข∗ ๐‘ → ๐‘Ÿ
and
โŠข∗ ๐‘ → (๐‘Ÿ → ๐‘ž).
Thus, as in the previous case, using (1.8), we obtain โŠข∗ ๐‘ → ๐‘ž.
These four cases exhaust the ways by which ๐‘ž can be proved from ๐‘ with a
proof with at most three propositional forms. Therefore, since these cases can
be generalized to proofs of arbitrary length (Exercise 7), we conclude that ๐‘ โŠข ๐‘ž
implies โŠข ๐‘ → ๐‘ž.
The deduction theorem (1.4.4) yields the next result. Its proof is left to Exercise 8.
COROLLARY 1.4.5
For all propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž, ๐‘Ÿ,
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž โŠข ๐‘Ÿ
if and only if
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠข ๐‘ž → ๐‘Ÿ.
Direct Proof
Most propositions that mathematicians prove are implications. For example,
if a function is differentiable at a point, it is continuous at that same point.
As we know, this means that whenever the function ๐‘“ is differentiable at ๐‘ฅ = ๐‘Ž, it
must also be the case that ๐‘“ is continuous at ๐‘ฅ = ๐‘Ž. Proofs of conditionals like this
Section 1.4 PROOF METHODS
45
are typically very difficult if we are only allowed to use Inference Rules 1.2.10 and Replacement Rules 1.3.9. Fortunately, in practice another inference rule is used. To prove
the differentiability result, what is usually done is that ๐‘“ is assumed to be differential
at ๐‘ฅ = ๐‘Ž and then a series of steps that lead to the conclusion that ๐‘“ is continuous at
๐‘ฅ = ๐‘Ž are followed. We copy this strategy in our formal proofs using the next rule.
Sometimes known as conditional proof, this inference rule follows by Corollary 1.4.5
and Theorem 1.2.14.
INFERENCE RULE 1.4.6 [Direct Proof (DP)]
For propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž, ๐‘Ÿ,
if ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž โŠข ๐‘Ÿ, then ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ⇒ ๐‘ž → ๐‘Ÿ.
PROOF
Suppose ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž โŠข ๐‘Ÿ. Then, by Corollary 1.4.5,
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠข ๐‘ž → ๐‘Ÿ.
Therefore, by Simp and Com,
๐‘0 ∧ ๐‘1 ∧ · · · ∧ ๐‘๐‘›−1 โŠข ๐‘ž → ๐‘Ÿ,
so by Theorem 1.2.14,
๐‘0 ∧ ๐‘1 ∧ · · · ∧ ๐‘๐‘›−1 ⇒ ๐‘ž → ๐‘Ÿ.
Finally, we have by Conj and Theorem 1.2.14 that
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ⇒ ๐‘ž → ๐‘Ÿ.
To see how this works, let us use direct proof to prove
๐‘ƒ ∨ ๐‘„ → (๐‘… ∧ ๐‘†) โŠข ๐‘ƒ → ๐‘….
To do this, we first prove
๐‘ƒ ∨ ๐‘„ → (๐‘… ∧ ๐‘†), ๐‘ƒ โŠข ๐‘….
Here is the proof:
1.
2.
3.
4.
5.
๐‘ƒ ∨ ๐‘„ → (๐‘… ∧ ๐‘†)
๐‘ƒ
๐‘ƒ ∨๐‘„
๐‘…∧๐‘†
๐‘…
Given
Given
2 Add
1, 3 MP
4 Simp
46
Chapter 1 PROPOSITIONAL LOGIC
Therefore, by Inference Rule 1.4.6,
๐‘ƒ ∨ ๐‘„ → (๐‘… ∧ ๐‘†) ⇒ ๐‘ƒ → ๐‘….
A proof of the original deduction can now be written as
1.
2.
๐‘ƒ ∨ ๐‘„ → (๐‘… ∧ ๐‘†)
๐‘ƒ →๐‘…
Given
1 DP
However, instead of writing the first proof off to the side, it is typically incorporated
into the proof as follows:
1.
2.
3.
4.
5.
6.
๐‘ƒ ∨ ๐‘„ → (๐‘… ∧ ๐‘†)
→๐‘ƒ
๐‘ƒ ∨๐‘„
๐‘…∧๐‘†
๐‘…
๐‘ƒ →๐‘…
Given
Assumption
2 Add
1, 3 MP
4 Simp
2–5 DP
The proof that ๐‘ƒ infers ๐‘… is a subproof of the main proof. To separate the propositional forms of the subproof from the rest of the proof, they are indented with a vertical
line. The line begins with the assumption of ๐‘ƒ in line 2 as an additional premise.
Hence, its reason is Assumption. This assumption can only be used in the subproof.
Consider it a local hypothesis. It is only used to prove ๐‘ƒ → ๐‘…. If we were allowed to
use it in other places of the proof, we would be proving a theorem that had different
premises than those that were given. Similarly, all lines within the subproof cannot be
referenced from the outside. We use the indentation to isolate the assumption and the
propositional forms that follow from it. When we arrive at ๐‘…, we know that we have
proved ๐‘ƒ → ๐‘…. The next line is this propositional form. It is entered into the proof
with the reason DP. The lines that are referenced are the lines of the subproof.
EXAMPLE 1.4.7
Prove: ๐‘ƒ → ¬๐‘„, ¬๐‘… ∨ ๐‘† โŠข ๐‘… ∨ ๐‘„ → (๐‘ƒ → ๐‘†)
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
๐‘ƒ → ¬๐‘„
¬๐‘… ∨ ๐‘†
→๐‘… ∨ ๐‘„
→๐‘ƒ
¬๐‘„
๐‘„∨๐‘…
๐‘…
¬¬๐‘…
๐‘†
๐‘ƒ →๐‘†
(๐‘… ∨ ๐‘„) → (๐‘ƒ → ๐‘†)
Given
Given
Assumption
Assumption
1, 4 MP
3 Com
5, 6 DS
7 DN
2, 8 DS
4–9 DP
3–10 DP
Section 1.4 PROOF METHODS
47
EXAMPLE 1.4.8
Prove: ๐‘ƒ ∧ ๐‘„ → ๐‘… → ๐‘†, ¬๐‘„ ∨ ๐‘… โŠข ๐‘†
1.
2.
3.
4.
5.
6.
7.
8.
9.
Given
Given
Assumption
3 Com
4 Simp
5 DN
2, 6 DS
3–7 DP
1, 8 MP
๐‘ƒ ∧๐‘„→๐‘…→๐‘†
¬๐‘„ ∨ ๐‘…
→๐‘ƒ ∧๐‘„
๐‘„∧๐‘ƒ
๐‘„
¬¬๐‘„
๐‘…
๐‘ƒ ∧๐‘„→๐‘…
๐‘†
EXAMPLE 1.4.9
Prove: โŠข ๐‘ƒ ∨ ¬๐‘ƒ
1.
2.
3.
4.
→๐‘ƒ
๐‘ƒ →๐‘ƒ
¬๐‘ƒ ∨ ๐‘ƒ
๐‘ƒ ∨ ¬๐‘ƒ
Assumption
1 DP
2 Impl
3 Com
Note that lines 1–2 prove โŠข ๐‘ƒ → ๐‘ƒ .
Indirect Proof
When direct proof is either too difficult or not appropriate, there is another common
approach to writing formal proofs. Sometimes going by the name of proof by contradiction or reductio ad absurdum, this inference rule can also be used to prove propositional forms that are not implications.
INFERENCE RULE 1.4.10 [Indirect Proof (IP)]
For all propositional forms ๐‘ and ๐‘ž,
¬๐‘ž → (๐‘ ∧ ¬๐‘) ⇒ ๐‘ž.
PROOF
Notice that instead of repeating the argument from Example 1.4.9 in this proof,
the example is simply cited as the reason on line 2.
1.
2.
3.
4.
¬๐‘ž → (๐‘ ∧ ¬๐‘)
๐‘→๐‘
¬(๐‘ ∧ ¬๐‘) → ¬¬๐‘ž
¬(๐‘ ∧ ¬๐‘) → ๐‘ž
Given
Example 1.4.9
1 Contra
3 DN
48
Chapter 1 PROPOSITIONAL LOGIC
5.
6.
7.
8.
4 DeM
5 DN
6 Impl
2, 7 MP
¬๐‘ ∨ ¬¬๐‘ → ๐‘ž
¬๐‘ ∨ ๐‘ → ๐‘ž
๐‘→๐‘→๐‘ž
๐‘ž
The rule follows from Theorem 1.2.14.
To use indirect proof, assume each premise and assume the negation of the conclusion. Then, proceed with the proof until a contradiction is reached. (In Inference
Rule 1.4.10, the contradiction is represented by ๐‘ ∧ ¬๐‘.) At this point, deduce the
original conclusion.
EXAMPLE 1.4.11
Prove: ๐‘ƒ ∨ ๐‘„ → ๐‘…, ๐‘… ∨ ๐‘† → ¬๐‘ƒ ∧ ๐‘‡ โŠข ¬๐‘ƒ .
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
๐‘ƒ ∨๐‘„→๐‘…
๐‘… ∨ ๐‘† → ¬๐‘ƒ ∧ ๐‘‡
→¬¬๐‘ƒ
๐‘ƒ
๐‘ƒ ∨๐‘„
๐‘…
๐‘…∨๐‘†
¬๐‘ƒ ∧ ๐‘‡
¬๐‘ƒ
๐‘ƒ ∧ ¬๐‘ƒ
¬๐‘ƒ
Given
Given
Assumption
3 DN
4 Add
1, 5 MP
6 Add
2, 7 MP
8 Simp
4, 9 Conj
3–10 IP
Since IP involves proving an implication, the formal proof takes the same form
as a proof involving DP.
Indirect proof can also be nested within another indirect subproof. As with direct proof,
we cannot appeal to lines within a subproof from outside of it.
EXAMPLE 1.4.12
Prove: ๐‘ƒ → ๐‘„ ∧ ๐‘…, ๐‘„ → ๐‘†, ¬๐‘ƒ → ๐‘† โŠข ๐‘†.
1.
2.
3.
4.
5.
6.
7.
8.
9.
๐‘ƒ →๐‘„∧๐‘…
๐‘„→๐‘†
¬๐‘ƒ → ๐‘†
→¬๐‘†
¬๐‘„
→๐‘ƒ
๐‘„∧๐‘…
๐‘„
๐‘„ ∧ ¬๐‘„
Given
Given
Given
Assumption
2, 4 MT
Assumption
1, 6 MP
7 Simp
5, 8 Conj
Section 1.4 PROOF METHODS
10.
11.
12.
13.
6-9 IP
3, 10 MP
4, 11 Conj
4–12 IP
¬๐‘ƒ
๐‘†
๐‘† ∧ ¬๐‘†
๐‘†
Notice that line 11 was not the end of the proof since it was within the first subproof. It followed under the added hypothesis of ¬๐‘†.
EXAMPLE 1.4.13
Prove: ๐‘ƒ → ๐‘… โŠข ๐‘ƒ ∧ ๐‘„ → ๐‘… ∨ ๐‘†.
1.
2.
3.
4.
5.
6.
7.
8.
9.
Given
Assumption
Assumption
1, 3 MT
2 Simp
4, 5 Conj
3-6 IP
7 Add
2–8 DP
๐‘ƒ →๐‘…
→๐‘ƒ ∧ ๐‘„
→¬๐‘…
¬๐‘ƒ
๐‘ƒ
๐‘ƒ ∧ ¬๐‘ƒ
๐‘…
๐‘…∨๐‘†
๐‘ƒ ∧๐‘„→๐‘…∨๐‘†
Exercises
1. Find all mistakes in the given proofs.
(a) “๐‘ƒ ∨ ๐‘„ → ¬๐‘…, ๐‘… → ¬๐‘„ → ๐‘† ∨ ๐‘„ โŠข ๐‘†”
Attempted Proof
1. ๐‘ƒ ∨ ๐‘„ → ¬๐‘…
2. ๐‘… → ¬๐‘„ → ๐‘† ∨ ๐‘„
3. →๐‘…
4.
¬¬๐‘…
5.
¬(๐‘ƒ ∨ ๐‘„)
6.
¬๐‘ƒ ∧ ¬๐‘„
7.
¬๐‘„
8. ๐‘… → ¬๐‘„
9. ๐‘† ∨ ๐‘„
10. ๐‘„ ∨ ๐‘†
11. ๐‘†
Given
Given
Assumption
Assumption
1, 4 MT
5 DeM
6 Simp
3–7 DP
2, 8 MP
9 Com
7, 10 DS
(b) “¬๐‘ƒ ∨ ๐‘„ โŠข ๐‘ƒ → ๐‘„ → ๐‘…”
Attempted Proof
1. ¬๐‘ƒ ∨ ๐‘„
2. →๐‘ƒ
Given
Assumption
49
50
Chapter 1 PROPOSITIONAL LOGIC
3.
4.
5.
6.
7.
¬¬๐‘ƒ
๐‘„
๐‘ƒ →๐‘„
๐‘…
๐‘ƒ →๐‘„→๐‘…
2 DN
1, 3 DS
2–4 DP
MP
2–6 DP
(c) “¬๐‘… ∧ ๐‘†, ¬๐‘ƒ ∨ ๐‘„ → ๐‘… โŠข ¬๐‘ƒ ∨ ๐‘„ → ๐‘„”
Attempted Proof
1. ¬๐‘… ∧ ๐‘†
2. ¬๐‘ƒ ∨ ๐‘„ → ๐‘…
→ ¬๐‘ƒ
3.
→¬๐‘ƒ ∨ ๐‘„
4.
5.
๐‘…
¬๐‘…
6.
7.
๐‘… ∧ ¬๐‘…
8.
๐‘ƒ
9.
¬¬๐‘ƒ
๐‘„
10.
11.
¬๐‘ƒ ∨ ๐‘„ → ๐‘„
Given
Given
Assumption
Assumption
2, 3 MP
1 Simp
5, 6 Conj
4–7 IP
8 DN
4, 9 DS
3–10 DP
2. Prove using direct proof.
(a) ๐‘ƒ → ๐‘„ ∧ ๐‘… โŠข ๐‘ƒ → ๐‘„
(b) ๐‘ƒ ∨ ๐‘„ → ๐‘… โŠข ๐‘ƒ → ๐‘…
(c) ๐‘ƒ ∨ (๐‘„ ∨ ๐‘…) → ๐‘† โŠข ๐‘„ → ๐‘†
(d) ๐‘ƒ → ๐‘„, ๐‘… → ๐‘„ โŠข ๐‘ƒ ∨ ๐‘… → ๐‘„
(e) ๐‘ƒ → ๐‘„, ๐‘ƒ → ๐‘… โŠข ๐‘ƒ → ๐‘„ ∧ ๐‘…
(f) ๐‘ƒ → (๐‘„ → ๐‘…) โŠข ๐‘„ ∧ ¬๐‘… → ¬๐‘ƒ
(g) ๐‘… → ¬๐‘† โŠข ๐‘ƒ ∧ ๐‘„ → (๐‘… → ¬๐‘†)
(h) ๐‘ƒ → (๐‘„ → ๐‘…) โŠข ๐‘„ → (๐‘ƒ → ๐‘…)
(i) ๐‘ƒ → (๐‘„ → ๐‘…), ๐‘„ → (๐‘… → ๐‘†) โŠข ๐‘ƒ → (๐‘„ → ๐‘†)
(j) ๐‘ƒ → (๐‘„ → ๐‘…), ๐‘… → ๐‘† ∧ ๐‘‡ โŠข ๐‘ƒ → (๐‘„ → ๐‘‡ )
(k) ๐‘ƒ ↔ ๐‘„ ∧ ๐‘… โŠข ๐‘ƒ → ๐‘„
(l) ๐‘ƒ ∧ ๐‘„ ∨ ๐‘… → ๐‘„ ∧ ๐‘… โŠข ๐‘ƒ → (๐‘„ → ๐‘…)
3. Prove using indirect proof.
(a) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ¬๐‘… โŠข ¬๐‘ƒ
(b) ๐‘ƒ ∨ ๐‘„ ∧ ๐‘…, ๐‘ƒ → ๐‘†, ๐‘„ → ๐‘† โŠข ๐‘†
(c) ๐‘ƒ ∨ ๐‘„ ∧ ¬๐‘…, ๐‘ƒ → ๐‘†, ๐‘„ → ๐‘… โŠข ๐‘†
(d) ๐‘ƒ ⇔ ๐‘„, ¬๐‘ƒ โŠข ¬๐‘„
(e) ๐‘ƒ ∨ ๐‘„ ∨ ๐‘… → ๐‘„ ∧ ๐‘… โŠข ¬๐‘ƒ ∨ ๐‘„ ∧ ๐‘…
(f) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘† → ๐‘‡ , ๐‘ƒ ∨ ๐‘…, ¬๐‘… โŠข ๐‘‡
(g) ๐‘ƒ → ¬๐‘„, ๐‘… → ¬๐‘†, ๐‘‡ → ๐‘„, ๐‘ˆ → ๐‘†, ๐‘ƒ ∨ ๐‘… โŠข ¬๐‘‡ ∨ ¬๐‘ˆ
Section 1.5 THE THREE PROPERTIES
51
(h) ๐‘ƒ → ๐‘„ ∧ ๐‘…, ๐‘„ ∨ ๐‘† → ๐‘‡ ∧ ๐‘ˆ , ๐‘ƒ โŠข ๐‘‡
4. Prove using both direct and indirect proof.
(a) ๐‘ƒ → ๐‘„, ๐‘ƒ ∨ (๐‘… → ๐‘†), ¬๐‘„ โŠข ๐‘… → ๐‘†
(b) ๐‘ƒ → ¬(๐‘„ → ¬๐‘…) โŠข ๐‘ƒ → ๐‘…
(c) ๐‘ƒ → ๐‘„ โŠข ๐‘ƒ ∧ ๐‘… → ๐‘„
(d) ๐‘ƒ ⇔ ๐‘„ ∨ ๐‘… โŠข ๐‘„ → ๐‘ƒ
5. Prove by using direct proof to prove the contrapositive.
(a) ๐‘ƒ → ๐‘„, ๐‘… → ๐‘†, ๐‘† → ๐‘‡ , ¬๐‘„ โŠข ¬๐‘‡ → ¬(๐‘ƒ ∨ ๐‘…)
(b) ๐‘ƒ ∧ ๐‘„ ∨ ๐‘… ∧ ๐‘† โŠข ¬๐‘† → ๐‘ƒ ∧ ๐‘„
(c) ๐‘ƒ ∨ ๐‘„ → ¬๐‘…, ๐‘† → ๐‘… โŠข ๐‘ƒ → ¬๐‘†
(d) ¬๐‘ƒ → ¬๐‘„, (¬๐‘… ∨ ๐‘†) ∧ (๐‘… ∨ ๐‘„) โŠข ¬๐‘† → ๐‘ƒ
6. Prove to complete the proof of Theorem 1.4.2.
(a) ๐‘ ∨ ๐‘ž, ¬๐‘ โŠข∗ ๐‘ž
(b) ๐‘, ๐‘ž โŠข∗ ๐‘ ∧ ๐‘ž
(c) (๐‘ → ๐‘ž) ∧ (๐‘Ÿ → ๐‘ ), ๐‘ ∨ ๐‘Ÿ โŠข∗ ๐‘ž ∨ ๐‘ 
(d) (๐‘ → ๐‘ž) ∧ (๐‘Ÿ → ๐‘ ), ¬๐‘ž ∨ ¬๐‘  โŠข∗ ¬๐‘ ∨ ¬๐‘Ÿ
(e) ๐‘ ∧ ๐‘ž โŠข∗ ๐‘
7. Given there is a proof of ๐‘ž from ๐‘ with four propositional forms, prove โŠข ๐‘ → ๐‘ž.
Generalize the proof for ๐‘› propositional forms.
8. Prove Corollary 1.4.5.
9. Can MP be replaced with another inference rule in Definition 1.4.1 and still have
Theorem 1.4.2 hold true? If so, find the inference rules.
10. Can any of the replacement rules be removed from Definition 1.4.1 and still have
Theorem 1.4.2 hold true? If so, how many can be removed and which ones?
1.5
THE THREE PROPERTIES
We finish our introduction to propositional logic by showing that this logical system has
three important properties. These are properties that are shared with Euclid’s geometry,
but they are not common to all logical systems.
Consistency
Since we can need to consider infinitely many propositional forms, we now write our
lists of propositional forms as
๐‘0 , ๐‘1 , ๐‘2 , … ,
allowing this sequence to be finite or infinite. Since proofs are finite, the notation
๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž
52
Chapter 1 PROPOSITIONAL LOGIC
means that there exists a subsequence ๐‘–0 , ๐‘–1 , … , ๐‘–๐‘›−1 of 0, 1, 2, … such that
๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 โŠข ๐‘ž.
The notation
๐‘0 , ๐‘1 , ๐‘2 , … โŠฌ ๐‘ž
means that no such subsequence exists.
DEFINITION 1.5.1
โˆ™ The propositional forms ๐‘0 , ๐‘1 , ๐‘2 , … are consistent if for every propositional
form ๐‘ž,
๐‘0 , ๐‘1 , ๐‘2 , … โŠฌ ๐‘ž ∧ ¬๐‘ž,
and we write ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ). Otherwise, ๐‘0 , ๐‘1 , ๐‘2 , … is inconsistent.
โˆ™ A logical system is consistent if no contradiction is a theorem.
We have two goals. The first is to show that propositional logic is consistent. The
second is to discover properties of sequences of consistent propositional forms that will
aid in proving other properties of propositional logic. The next theorem is important
to meet both of these goals. The equivalence of the first two parts is known as the
compactness theorem
THEOREM 1.5.2
If ๐‘0 , ๐‘1 , ๐‘2 , … are propositional forms, the following are equivalent in propositional logic.
โˆ™ ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ).
โˆ™ Every finite subsequence of ๐‘0 , ๐‘1 , ๐‘2 , … is consistent.
โˆ™ There exists a propositional form ๐‘ such that ๐‘0 , ๐‘1 , ๐‘2 , … โŠฌ ๐‘.
PROOF
We have three implications to prove.
โˆ™ Suppose there is a finite subsequence ๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 that proves ๐‘ž ∧ ¬๐‘ž
for some propositional form ๐‘ž. This implies that there is a formal proof of
๐‘ž ∧ ¬๐‘ž from ๐‘0 , ๐‘1 , ๐‘2 , … , therefore, not ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ).
โˆ™ Assume that ๐‘0 , ๐‘1 , ๐‘2 , … proves every propositional form. In particular,
if we take a propositional form ๐‘ž, we have that ๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž ∧ ¬๐‘ž. This
means that there is a finite subsequence ๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 that proves ๐‘ž ∧¬๐‘ž.
โˆ™ Lastly, assume that there exists a propositional form ๐‘ž such that
๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž ∧ ¬๐‘ž.
Section 1.5 THE THREE PROPERTIES
53
This means that there exist subscripts ๐‘–0 , ๐‘–1 , … , ๐‘–๐‘›−1 and propositional forms
๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘š−1 such that
๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 , ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘š−1 , ๐‘ž ∧ ¬๐‘ž
is a proof. Take any propositional form ๐‘. Then,
๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 , ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘š−1 , ¬๐‘, ๐‘ž ∧ ¬๐‘ž, ๐‘
is a proof of ๐‘ by IP. Therefore, ๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘.
A sequence of propositional forms, such as ๐‘ƒ → ๐‘„, ๐‘ƒ , ๐‘„, although consistent, has
the property that there are propositional forms that can be added to the sequence so that
the resulting list remains consistent. When the sequence can no longer take new forms
and remain consistent, we have arrived at a sequence that satisfies the next definition.
DEFINITION 1.5.3
A sequence of propositional forms ๐‘0 , ๐‘1 , ๐‘2 , … is maximally consistent whenever ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ) and for all propositional forms ๐‘, ๐–ข๐—ˆ๐—‡(๐‘, ๐‘0 , ๐‘1 , ๐‘2 , … )
implies that ๐‘ = ๐‘๐‘– for some ๐‘–.
It is a convenient result of propositional logic that every consistent sequence of
propositional forms can be extended to a maximally consistent sequence. This is possible because all possible propositional forms can be put into a list. Following Definition 1.1.2, we first list the propositional variables:
๐ด, ๐ต, ๐ถ,
๐ด0 , ๐ด1 , ๐ด2 ,
๐ต0 , ๐ต1 , ๐ต2 ,
โ‹ฎ
โ‹ฎ
โ‹ฎ
๐‘0 , ๐‘1 , ๐‘2 ,
…
…
…
๐‘‹,
๐‘Œ,
๐‘,
…
Then, we list all propositional forms with only one propositional variable:
¬๐ด,
¬๐ด0 ,
¬๐ต0 ,
โ‹ฎ
¬๐‘0 ,
¬๐ต,
¬๐ด1 ,
¬๐ต1 ,
โ‹ฎ
¬๐‘1 ,
¬๐ถ,
¬๐ด2 ,
¬๐ต2 ,
โ‹ฎ
¬๐‘2 ,
…
…
…
¬๐‘‹, ¬๐‘Œ , ¬๐‘,
…
Next, we list all propositional forms with exactly two propositional variables starting
by writing ๐ด on the right:
๐ด ∨ ๐ด, ๐ต ∨ ๐ด, ๐ถ ∨ ๐ด,
๐ด0 ∨ ๐ด, ๐ด1 ∨ ๐ด, ๐ด2 ∨ ๐ด,
๐ต0 ∨ ๐ด, ๐ต1 ∨ ๐ด, ๐ต2 ∨ ๐ด,
โ‹ฎ
โ‹ฎ
โ‹ฎ
๐‘0 ∨ ๐ด, ๐‘1 ∨ ๐ด, ๐‘2 ∨ ๐ด,
…
…
…
…
๐‘‹ ∨ ๐ด,
๐‘Œ ∨ ๐ด,
๐‘ ∨ ๐ด,
54
Chapter 1 PROPOSITIONAL LOGIC
๐ด ∧ ๐ด, ๐ต ∧ ๐ด, ๐ถ ∧ ๐ด,
๐ด0 ∧ ๐ด, ๐ด1 ∧ ๐ด, ๐ด2 ∧ ๐ด,
๐ต0 ∧ ๐ด, ๐ต1 ∧ ๐ด, ๐ต2 ∧ ๐ด,
โ‹ฎ
โ‹ฎ
โ‹ฎ
๐‘0 ∧ ๐ด, ๐‘1 ∧ ๐ด, ๐‘2 ∧ ๐ด,
…
…
…
๐ด → ๐ด, ๐ต → ๐ด, ๐ถ → ๐ด,
๐ด0 → ๐ด, ๐ด1 → ๐ด, ๐ด2 → ๐ด,
๐ต0 → ๐ด, ๐ต1 → ๐ด, ๐ต2 → ๐ด,
โ‹ฎ
โ‹ฎ
โ‹ฎ
๐‘0 → ๐ด, ๐‘1 → ๐ด, ๐‘2 → ๐ด,
…
…
…
๐ด ↔ ๐ด, ๐ต ↔ ๐ด, ๐ถ ↔ ๐ด,
๐ด0 ↔ ๐ด, ๐ด1 ↔ ๐ด, ๐ด2 ↔ ๐ด,
๐ต0 ↔ ๐ด, ๐ต1 ↔ ๐ด, ๐ต2 ↔ ๐ด,
โ‹ฎ
โ‹ฎ
โ‹ฎ
๐‘0 ↔ ๐ด, ๐‘1 ↔ ๐ด, ๐‘2 ↔ ๐ด,
…
…
…
๐‘‹ ∧ ๐ด,
๐‘Œ ∧ ๐ด,
๐‘ ∧ ๐ด,
…
๐‘‹ → ๐ด,
๐‘Œ → ๐ด,
๐‘ → ๐ด,
๐‘‹ ↔ ๐ด,
๐‘Œ ↔ ๐ด,
๐‘ ↔ ๐ด,
…
…
After following the same pattern by attaching ๐ด on the left, we continue by writing
¬๐ด on the right, and then on the left, and then we adjoin ๐ต and ¬๐ต, etc., and then
use 3 propositional variables, and then four, etc. Following a careful path through this
infinite list, we arrive at a sequence
๐‘ž0 , ๐‘ž1 , ๐‘ž2 , …
of all propositional forms. We are now ready for the theorem.
THEOREM 1.5.4
A consistent sequence of propositional forms is a subsequence of a maximally
consistent sequence of propositional forms.
PROOF
Let ๐‘0 , ๐‘1 , ๐‘2 , … be consistent and ๐‘ž0 , ๐‘ž1 , ๐‘ž2 , … be a sequence of all propositional
forms. Define the sequence ๐‘Ÿ๐‘– as follows:
โˆ™ Let ๐‘Ÿ0 = ๐‘0 and
{
๐‘ž0
๐‘Ÿ1 =
๐‘0
if ๐–ข๐—ˆ๐—‡(๐‘ž0 , ๐‘0 , ๐‘1 , ๐‘2 , … ),
otherwise,
so the sequence at this stage is ๐‘0 , ๐‘ž0 or ๐‘0 , ๐‘0 . Both of these are consistent.
โˆ™ Let ๐‘Ÿ2 = ๐‘1 and
{
๐‘ž1
๐‘Ÿ3 =
๐‘1
if ๐–ข๐—ˆ๐—‡(๐‘ž1 , ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿ2 , ๐‘0 , ๐‘1 , ๐‘2 , … ),
otherwise.
Section 1.5 THE THREE PROPERTIES
55
At this stage, the sequence is still consistent and of the form ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘1 , ๐‘ž1
or ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘1 , ๐‘1 . The first sequence is consistent by the definition of ๐‘Ÿ3 , and
the second sequence is consistent because ๐–ข๐—ˆ๐—‡(๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘1 ).
โˆ™ Generalizing, let ๐‘Ÿ2๐‘˜ = ๐‘๐‘˜ and
๐‘Ÿ2๐‘˜+1
{
๐‘ž๐‘˜
=
๐‘๐‘˜
if ๐–ข๐—ˆ๐—‡(๐‘ž๐‘˜ , ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ2๐‘˜ , ๐‘0 , ๐‘1 , ๐‘2 , … ),
otherwise,
resulting in a consistent sequence of the form
๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿ2 , ๐‘Ÿ3 , … , ๐‘๐‘˜ , ๐‘ž๐‘˜
or
๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿ2 , ๐‘Ÿ3 , … , ๐‘๐‘˜ , ๐‘๐‘˜ .
Since ๐‘0 , ๐‘1 , ๐‘3 , … is a subsequence of ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿ2 , … , it only remains to show that
the new sequence is maximally consistent.
โˆ™ Let ๐‘  be a propositional form such that ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿ2 , … โŠข ๐‘  ∧ ¬๐‘ . This implies
that there exists a sequence ๐‘–๐‘— such that ๐‘–0 < ๐‘–1 < · · · < ๐‘–๐‘˜ and
๐‘Ÿ๐‘–0 , ๐‘Ÿ๐‘–1 , … , ๐‘Ÿ๐‘–๐‘˜ โŠข ๐‘  ∧ ¬๐‘ ,
but by Theorem 1.5.2, this is impossible because ๐–ข๐—ˆ๐—‡(๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘–๐‘˜ ).
โˆ™ Suppose that ๐‘  is a propositional form so that ๐–ข๐—ˆ๐—‡(๐‘ , ๐‘Ÿ0 , ๐‘Ÿ1 , ๐‘Ÿ2 , … ). Write
๐‘  = ๐‘ž๐‘– for some ๐‘–. Therefore,
๐–ข๐—ˆ๐—‡(๐‘ž๐‘– , ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ2๐‘– , ๐‘0 , ๐‘1 , ๐‘2 , … ),
which means that ๐‘  is a term of the sequence ๐‘Ÿ๐‘– because ๐‘ž๐‘– was added at
step 2๐‘– + 1.
Soundness
We have defined two separate tracks in propositional logic. One track is used to assign
T or F to a propositional form, and thus it can be used to determine the truth value of
a proposition. The other track focused on developing methods by which one propositional form can be proved from other propositional forms. These methods are used to
write proofs in various fields of mathematics. The question arises whether these two
tracks have been defined in such a way that they get along with each other. In other
words, we want the propositional forms that we prove always to be assigned T, and
we want the propositional forms that we always assign T to be provable. This means
that we want semantic methods to yield syntactic results and syntactic methods to yield
semantic results.
56
Chapter 1 PROPOSITIONAL LOGIC
Sound
The statement form
is a tautology.
The statement form
is a theorem.
Complete
Figure 1.4
Sound and complete logics.
DEFINITION 1.5.5
โˆ™ A logic is sound if every theorem is a tautology.
โˆ™ A logic is complete if every tautology is a theorem.
There is no guarantee that the construction of the two tracks for a logic will have these
two properties (Figure 1.4), but it does in the case of propositional logic.
To prove that propositional logic is sound, we need three lemmas. The proof of the
first is left to Exercise 1.
LEMMA 1.5.6
The propositional forms of Axioms 1.2.8 are tautologies.
LEMMA 1.5.7
Let ๐‘, ๐‘ž, and ๐‘Ÿ be propositional forms.
โˆ™ If ๐‘ ⇒ ๐‘Ÿ, then ๐‘ → ๐‘Ÿ is a tautology.
โˆ™ If ๐‘, ๐‘ž ⇒ ๐‘Ÿ, then ๐‘ ∧ ๐‘ž → ๐‘Ÿ is a tautology.
PROOF
This is simply a matter of checking Inference Rules 1.2.10 and 1.3.10.
For example, to check the theorem for De Morgan’s law, we must show that
¬(๐‘ ∧ ๐‘ž) → ¬๐‘ ∨ ¬๐‘ž
and
¬๐‘ ∨ ¬๐‘ž → ¬(๐‘ ∧ ๐‘ž)
are tautologies. To do this, examine the truth table
๐‘
T
T
F
F
๐‘ž
T
F
T
F
๐‘∧๐‘ž
T
F
F
F
¬(๐‘ ∧ ๐‘ž)
F
T
T
T
¬๐‘
F
F
T
T
¬๐‘ž
F
T
F
T
¬๐‘ ∨ ¬๐‘ž
F
T
T
T
¬(๐‘ ∧ ๐‘ž) → ¬๐‘ ∨ ¬๐‘ž
T
T
T
T
¬๐‘ ∨ ¬๐‘ž → ¬(๐‘ ∧ ๐‘ž)
T
T
T
T
Section 1.5 THE THREE PROPERTIES
We also have to show that
¬(๐‘ ∨ ๐‘ž) → ¬๐‘ ∧ ¬๐‘ž
and
¬๐‘ ∧ ¬๐‘ž → ¬(๐‘ ∨ ๐‘ž)
are tautologies.
As another example, to check that the disjunctive syllogism leads to an implication that is a tautology, examining the truth table:
๐‘
T
T
F
F
๐‘ž ๐‘∨๐‘ž
T
T
F
T
T
T
F
F
¬๐‘ (๐‘ ∨ ๐‘ž) ∧ ¬๐‘
F
F
F
F
T
T
T
F
(๐‘ ∨ ๐‘ž) ∧ ¬๐‘ → ๐‘ž
T
T
T
T
The other rules are checked similarly (Exercise 2).
LEMMA 1.5.8
If ๐‘ → ๐‘ž and ๐‘ are tautologies, then ๐‘ž is a tautology.
PROOF
This is done by examining the truth table of ๐‘ → ๐‘ž (page 12) where we see that
๐—(๐‘ž) is constant and equal to T because ๐—(๐‘) and ๐—(๐‘ → ๐‘ž) are constant and both
equal to T.
We now prove the first important property of propositional logic.
THEOREM 1.5.9 [Soundness]
Every theorem of propositional logic is a tautology.
PROOF
Let ๐‘ be a theorem. Let ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 be propositional forms such that
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1
is a proof for ๐‘ such that ๐‘1 is an axiom, ๐‘๐‘– with ๐‘– > 0 is an axiom or follows
by a rule of inference, and ๐‘๐‘›−1 = ๐‘ (Definition 1.2.13). We now examine the
propositional forms of the proof.
โˆ™ By Lemma 1.5.6, ๐‘0 is a tautology.
โˆ™ The propositional form ๐‘1 is a tautology for one of two reasons. If ๐‘1 is
an axiom, it is a tautology (Lemma 1.5.6). If it follows from ๐‘0 because
๐‘0 ⇒ ๐‘1 , then ๐‘0 → ๐‘1 is a tautology (Lemma 1.5.7), so ๐‘1 is a tautology
by Lemma 1.5.8.
57
58
Chapter 1 PROPOSITIONAL LOGIC
โˆ™ If ๐‘2 follows from ๐‘0 or ๐‘1 , reason as in the previous case. Suppose ๐‘0 , ๐‘1 ⇒
๐‘2 . Then, by Lemma 1.5.7, (๐‘0 ∧ ๐‘1 ) → ๐‘2 is a tautology. Because
๐‘0 , ๐‘1 ⇒ ๐‘0 ∧ ๐‘1 by Conj, ๐‘0 ∧ ๐‘1 is a tautology. Thus, ๐‘2 is a tautology by
Lemma 1.5.8.
Since every ๐‘๐‘– with ๐‘– > 0 is an axiom, follows from some ๐‘๐‘— with ๐‘— < ๐‘–, or follows
from some ๐‘๐‘— , ๐‘๐‘˜ with ๐‘—, ๐‘˜ < ๐‘–, continuing in this manner, we find after finitely
many steps that ๐‘ is a tautology.
COROLLARY 1.5.10
For all propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž,
if ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠข ๐‘ž, then ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠจ ๐‘ž.
The Law of Noncontradiction being a theorem of propositional logic (page 37) suggests
that we have the following result, which is the second important property of propositional logic.
COROLLARY 1.5.11
Propositional logic is consistent.
PROOF
Let ๐‘ be a propositional form. Suppose that ๐‘ ∧ ¬๐‘ is a theorem. This implies that
it is a tautology by the soundness theorem (1.5.9), but ๐‘ ∧ ¬๐‘ is a contradiction.
Completeness
We use the consistency of propositional logic to prove that propositional logic is complete. For this we need a few lemmas.
LEMMA 1.5.12
If not ๐–ข๐—ˆ๐—‡(¬๐‘ž, ๐‘0 , ๐‘1 , ๐‘2 , … ), then ๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž.
PROOF
If not ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ), then ๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž by Theorem 1.5.2, so suppose
๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ). Assume that there exists a propositional form ๐‘Ÿ such that
¬๐‘ž, ๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘Ÿ ∧ ¬๐‘Ÿ.
This implies that there exists a formal proof
๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 , ¬๐‘ž, ๐‘ 0 , ๐‘ 1 , … , ๐‘ ๐‘š−1 , ๐‘Ÿ ∧ ¬๐‘Ÿ,
where ¬๐‘ž is in the proof because ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ). Then,
๐‘๐‘–0 , ๐‘๐‘–1 , … , ๐‘๐‘–๐‘›−1 , ๐‘ž
(1.10)
59
Section 1.5 THE THREE PROPERTIES
is a proof, where ๐‘ž follows by IP with (1.10) as the subproof. Therefore,
๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž.
LEMMA 1.5.13
If ๐‘0 , ๐‘1 , ๐‘2 , … are maximally consistent, then for every propositional form ๐‘ž,
either ๐‘ž = ๐‘๐‘– or ¬๐‘ž = ๐‘๐‘– for some ๐‘–.
PROOF
Since ๐‘0 , ๐‘1 , ๐‘2 , … are consistent, both ๐‘ž and ¬๐‘ž cannot be terms of the sequence.
Suppose that it is ¬๐‘ž that is not in the list. By the definition of maximal consistency, we conclude that ¬๐‘ž, ๐‘0 , ๐‘1 , ๐‘2 , … are not consistent. Therefore, by
Lemma 1.5.12, we conclude that ๐‘0 , ๐‘1 , ๐‘2 , … โŠข ๐‘ž, and since the sequence is
maximally consistent, ๐‘ž = ๐‘๐‘– for some ๐‘–.
To prove the next lemma, we use a technique called induction on propositional
forms. It states that a property will hold true for all propositional forms if two conditions are met:
โˆ™ The property holds for all propositional variables.
โˆ™ If the property holds for ๐‘ and ๐‘ž, the property holds for ¬๐‘, ๐‘ ∧ ๐‘ž, ๐‘ ∨ ๐‘ž, ๐‘ → ๐‘ž,
and ๐‘ ↔ ๐‘ž.
In proving the second condition, we first assume that
the property holds for the propositional forms ๐‘ and ๐‘ž.
(1.11)
This assumption (1.11) is known as an induction hypothesis. Because
๐‘ ∨ ๐‘ž ⇔ ¬๐‘ → ๐‘ž,
๐‘ ∧ ๐‘ž ⇔ ¬(๐‘ → ¬๐‘ž),
and
๐‘ ↔ ๐‘ž ⇔ (๐‘ → ¬๐‘ž) → ¬(¬๐‘ → ๐‘ž),
we need only to show that the induction hypothesis implies that the property holds for
¬๐‘ and ๐‘ → ๐‘ž.
LEMMA 1.5.14
If ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ), there exists a valuation ๐— such that
๐—(๐‘) = T if and only if ๐‘ = ๐‘๐‘–
for some ๐‘– = 0, 1, 2, … .
60
Chapter 1 PROPOSITIONAL LOGIC
PROOF
Since ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ), we know that
๐‘0 , ๐‘1 , ๐‘2 , …
(1.12)
can be extended to a maximally consistent sequence of propositional forms by
Theorem 1.5.4. If we find the desired valuation for the extended sequence, that
valuation will also work for the original sequence, so assume that (1.12) is maximally consistent. Let ๐‘‹0 , ๐‘‹1 , ๐‘‹2 , … represent all of the possible propositional
variables (page 53). Define
{
T if ๐‘‹๐‘— is a propositional variable of ๐‘๐‘– for some ๐‘–,
๐—(๐‘‹๐‘— ) =
F otherwise.
We prove that this is the desired valuation by induction on propositional forms.
We first claim that for all ๐‘—,
๐—(๐‘‹๐‘— ) = T if and only if ๐‘‹๐‘— = ๐‘๐‘– for some ๐‘–.
To prove this, first note that if ๐‘‹๐‘— is a term of (1.12), then ๐—(๐‘‹๐‘— ) = T by definition
of ๐—. To show the converse, suppose that ๐—(๐‘‹๐‘— ) = T but ๐‘‹๐‘— is not a term of
(1.12). By Lemma 1.5.13, there exists ๐‘– such that ¬๐‘‹๐‘— = ๐‘๐‘– . This implies that
๐—(¬๐‘‹๐‘— ) = T, and then ๐—(๐‘‹๐‘— ) = F (Definition 1.1.9), a contradiction.
Now assume that
๐—(๐‘ž) = T if and only if ๐‘ž = ๐‘๐‘– for some ๐‘–
and
๐—(๐‘Ÿ) = T if and only if ๐‘Ÿ = ๐‘๐‘– for some ๐‘–.
We first prove that
๐—(¬๐‘ž) = T if and only if ¬๐‘ž = ๐‘๐‘– for some ๐‘–.
โˆ™ Suppose that ๐—(¬๐‘ž) = T. Then, ๐—(๐‘ž) = F, and by induction, ๐‘ž is not
in (1.12). Since ๐‘0 , ๐‘1 , ๐‘2 , … is maximally consistent, ¬๐‘ž is in the list
(Lemma 1.5.13).
โˆ™ Conversely, let ¬๐‘ž = ๐‘๐‘– for some ๐‘–. By consistency, ๐‘ž is not in the sequence. Therefore, by the induction hypothesis, ๐—(๐‘ž) = F, which implies
that ๐—(¬๐‘ž) = T.
We next prove that
๐—(๐‘ž → ๐‘Ÿ) = T if and only if ๐‘ž → ๐‘Ÿ = ๐‘๐‘– for some ๐‘–.
โˆ™ Assume that ๐—(๐‘ž → ๐‘Ÿ) = T. We have two cases to check. First, let
๐—(๐‘ž) = ๐—(๐‘Ÿ) = T, so both ๐‘ž and ๐‘Ÿ are in (1.12) by the induction hypothesis. Suppose ๐‘ž → ๐‘Ÿ is not a term of the sequence. This implies that its
Section 1.5 THE THREE PROPERTIES
61
negation ¬(๐‘ž → ๐‘Ÿ) is a term of the sequence by Lemma 1.5.13. Therefore,
๐‘ž ∧ ¬๐‘Ÿ is in the sequence by maximal consistency, so ¬๐‘Ÿ is also in the sequence, a contradiction. Second, let ๐—(๐‘ž) = F. By induction, ๐‘ž is not a term
of the sequence, which implies that ¬๐‘ž is a term. Hence, ¬๐‘ž ∨ ๐‘Ÿ is in the
sequence, which implies that ๐‘ž → ๐‘Ÿ is also in the sequence.
โˆ™ To prove the converse, suppose that ๐‘ž → ๐‘Ÿ = ๐‘๐‘– for some ๐‘–. Assume
๐—(๐‘ž → ๐‘Ÿ) = F. This means that ๐—(๐‘ž) = T and ๐—(๐‘Ÿ) = F. Therefore, by
induction, ๐‘ž is a term of the sequence but ๐‘Ÿ is not. This implies that ¬๐‘Ÿ
is in the sequence, so ¬๐‘ž is in the sequence by MT. This contradicts the
consistency of (1.12).
THEOREM 1.5.15 [Completeness]
Every tautology of propositional logic is a theorem.
PROOF
Let ๐‘ be a propositional form such that FL1, FL2, FL3 โŠฌ ๐‘. By Lemma 1.5.12,
๐–ข๐—ˆ๐—‡(FL1, FL2, FL3, ¬๐‘). Therefore, by Lemma 1.5.14, there exists a valuation
such that ๐—(๐‘) = F, which implies that ฬธโŠจ ๐‘.
COROLLARY 1.5.16
If ๐‘0 , ๐‘1 , ๐‘2 , … โŠจ ๐‘ž, then ๐‘0 , ๐‘1 , ๐‘2 โŠข ๐‘ž for all propositional forms ๐‘0 , ๐‘1 , ๐‘2 , … .
We conclude by Theorems 1.5.9 and 1.5.15 and their corollaries that the notions
of semantically valid and syntactically valid coincide for deductions in propositional
logic.
Exercises
1. Prove Lemma 1.5.6.
2. Provide the remaining parts of the proof of Lemma 1.5.7.
3. Let ๐‘, ๐‘0 , ๐‘1 , ๐‘2 , … be propositional forms. Prove the following.
(a) If ๐‘0 , ๐‘1 , ๐‘2 , · · · โŠฌ ๐‘, then ๐–ข๐—ˆ๐—‡(¬๐‘, ๐‘0 , ๐‘1 , ๐‘2 , … ).
(b) If ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ) and ๐‘0 , ๐‘1 , ๐‘2 , · · · โŠข ๐‘, then ๐–ข๐—ˆ๐—‡(๐‘, ๐‘0 , ๐‘1 , ๐‘2 , … ).
(c) If ๐–ข๐—ˆ๐—‡(๐‘0 , ๐‘1 , ๐‘2 , … ), then ๐–ข๐—ˆ๐—‡(๐‘, ๐‘0 , ๐‘1 , ๐‘2 , … ) or ๐–ข๐—ˆ๐—‡(¬๐‘, ๐‘0 , ๐‘1 , ๐‘2 , … ).
4. Let ๐‘0 , ๐‘1 , ๐‘2 , … be a maximally consistent sequence of propositional forms. Let ๐‘
and ๐‘ž be propositional forms. Prove the following.
(a) If ๐‘0 , ๐‘1 , ๐‘2 , · · · โŠข ๐‘, then ๐‘ = ๐‘๐‘˜ for some ๐‘˜.
(b) ๐‘ ∧ ๐‘ž = ๐‘๐‘˜ for some ๐‘˜ if and only if ๐‘ = ๐‘๐‘– and ๐‘ž = ๐‘๐‘— for some ๐‘– and ๐‘—.
(c) If (๐‘ → ๐‘ž) = ๐‘๐‘– and ๐‘ = ๐‘๐‘— for some ๐‘– and ๐‘—, then ๐‘ž = ๐‘๐‘˜ for some ๐‘˜.
5. Use truth tables to prove the following. Explain why this is a legitimate technique.
(a) ¬๐‘„ ∨ (¬๐‘… ∨ ¬๐‘ƒ ) โŠข ๐‘ƒ → ¬(๐‘„ ∧ ๐‘…)
(b) ๐‘ƒ ∨ ๐‘„ → ๐‘… โŠข ¬๐‘… → ¬๐‘„
62
Chapter 1 PROPOSITIONAL LOGIC
(c) ๐‘ƒ → ¬(๐‘„ → ๐‘…) โŠข ๐‘ƒ → ¬๐‘…
(d) ๐‘ƒ ↔ ๐‘„ ∨ ๐‘… โŠข ๐‘„ → ๐‘ƒ
(e) ๐‘ƒ ∨ ๐‘„ → ๐‘… ∧ ๐‘†, ¬๐‘ƒ → (๐‘‡ → ¬๐‘‡ ), ¬๐‘… โŠข ¬๐‘‡
6. Write a formal proof to show the following. Explain why this is a legitimate technique.
(a) ¬๐‘ƒ ∨ ๐‘„, ¬๐‘„ โŠจ ¬๐‘ƒ
(b) ¬(๐‘ƒ ∧ ๐‘„), ๐‘ƒ โŠจ ¬๐‘„
(c) ๐‘ƒ → ๐‘„, ๐‘ƒ โŠจ ๐‘„ ∨ ๐‘…
(d) ๐‘ƒ → ๐‘„, ๐‘„ → ๐‘…, ๐‘ƒ โŠจ ๐‘…
(e) ๐‘ƒ ∨ ๐‘„ ∧ ๐‘…, ¬๐‘ƒ โŠจ ๐‘…
7. Write a formal proof to show the following.
(a) ¬(๐‘ƒ ∧ ๐‘„) ฬธโŠจ ¬๐‘ƒ
(b) (๐‘ƒ → ๐‘„) ∨ (๐‘… → ๐‘†), ๐‘ƒ ∨ ๐‘… ฬธโŠจ ๐‘„ ∨ ๐‘†
(c) ๐‘ƒ ∨ ๐‘…, ๐‘„ ∨ ๐‘†, ๐‘… ↔ ๐‘† ฬธโŠจ ๐‘… ∧ ๐‘†
8. Write a formal proof to demonstrate the following.
(a) ๐‘ ∨ ¬๐‘ is a tautology.
(b) ๐‘ ∧ ¬๐‘ is a contradiction.
9. Modify propositional logic by removing all replacement rules (1.3.9). Is the resulting logic consistent? sound? complete?
10. Modify propositional logic by removing all inference rules (1.2.10) except for Inference Rule 1.3.10. Is the resulting logic consistent? sound? complete?
CHAPTER 2
FIRST-ORDER LOGIC
2.1
LANGUAGES
We developed propositional logic to model basic proof and truth. We did so by using propositional forms to represent sentences that were either true or false. We saw
that propositional logic is consistent, sound, and complete. However, the sentences
of mathematics involve ideas that cannot be fully represented in propositional logic.
These sentences are able to characterize objects, such as numbers or geometric figures,
by describing properties of the objects, such as being even or being a rectangle, and
the relationships between them, such as equality or congruence. Since propositional
logic is not well suited to handle these ideas, we extend propositional logic to a richer
system.
Predicates
Consider the sentence it is a real number. This sentence has no truth value because the
meaning of the word it is undetermined. As noted on page 3, the word it is like a gap
in the sentence. It is as if the sentence was written as
63
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
64
Chapter 2 FIRST-ORDER LOGIC
is a real number.
However, that gap can be filled. Let us make some substitutions for it:
5 is a real number.
๐œ‹โˆ•7 is a real number.
Fido is a real number.
My niece’s toy is a real number.
With each replacement, the word that is undetermined is given meaning, and then the
sentence has a truth value. In the examples, the first two propositions are true, and the
last two are false.
Notice that
changes, whereas is a real number remains fixed. This is because
these two parts of the sentence have different purposes. The first is a reference to an
object, and the second is a property of the object. What we did in the examples was to
choose a property and then test whether various objects have that property:
It
↑
5
is a real number.
is a real number.
True
๐œ‹โˆ•7
is a real number.
True
Fido
is a real number.
False
My niece’s toy
is a real number.
False
Depending on the result, these two parts are put together to form a sentence that either
accurately or inaccurately affirms that a particular property is an attribute of an object.
For example, the first sentence states that being a real number is a characteristic of 5,
which is correct, and the last sentence states that being a real number is a characteristic
of my niece’s toy, which is not correct. We have terminology for all of this.
โˆ™ The subject of a sentence is the expression that refers to an object.
โˆ™ The predicate of a sentence is the expression that ascribes a property to the object
identified by the subject.
Thus, 5, ๐œ‹โˆ•7, Fido, and my niece’s toy are subjects that are substituted for it, and is a
real number is a predicate.
Substituting for the subject but not substituting for the predicate is a restriction that
we make. We limit the extension of propositional logic this way because it is to be part
of mathematical logic and this is what we do in mathematics. For example, take the
sentence,
๐‘ฅ + 2 = 7.
On its own it has no truth value, but when we substitute ๐‘ฅ = 5 or ๐‘ฅ = 10, the sentence
becomes a proposition. In this sense, ๐‘ฅ + 2 = 7 is like it is a real number. Both
the sentences have a gap that is assigned a meaning giving the sentence a truth value.
65
Section 2.1 LANGUAGES
However, the mathematical sentence is different from the English sentence in that it is
unclear as to what part is the subject. Is it ๐‘ฅ or ๐‘ฅ + 2? For our purposes, the answer is
irrelevant. This is because in mathematical logic, the subject is replaced with a variable
or, sometimes, with multiple variables. This change leads to a modification of what a
predicate is.
DEFINITION 2.1.1
A predicate is an expression that ascribes a property to the objects identified by
the variables of the sentence.
Therefore, the sentence ๐‘ฅ + 2 = 7 is a predicate. It describes a characteristic of ๐‘ฅ.
When expressions are substituted for ๐‘ฅ, the resulting sentence will be either a proposition that affirms that the value added to 2 equals 5 or another predicate. If ๐‘ฅ = 5 is
substituted, the result is the true proposition 5 + 2 = 7. That is, ๐‘ฅ + 2 = 7 is satisfied by
5. If ๐‘ฅ = 10 and the substitution is made, the resulting proposition 10+2 = 7 is false. In
mathematics, it is also common to substitute with undetermined values. For example,
if the substitution is ๐‘ฅ = ๐‘ฆ, the result is the predicate ๐‘ฆ + 2 = 7, and if the substitution
is ๐‘ฅ = sin ๐œƒ, the result is the predicate sin ๐œƒ + 2 = 7. Substituting ๐‘ฅ = ๐‘ฆ + 2๐‘ง2 yields
๐‘ฆ + 2๐‘ง2 + 2 = 7, a predicate with multiple variables. If we substitute ๐‘ฅ = ๐‘ฆ2 − 7๐‘ฆ, the
result is ๐‘ฆ2 − 7๐‘ฆ + 2 = 7, which is a predicate with multiple occurrences of the same
variable.
Assume that ๐‘ฅ represents a real number and consider the sentence
there exists x such that ๐‘ฅ + 2 = 7.
(2.1)
Although there is a variable with 2 occurrences, the sentence is a proposition, so in
propositional logic, it is an atom and would be represented by ๐‘ƒ . This does not tell us
much about the sentence, so we instead break the sentence into two parts:
Quantifier
↓
there exists ๐‘ฅ
such that
Predicate
↓
๐‘ฅ+2=7
A quantifier indicates how many objects satisfy the sentence. In this case, the quantifier is the existential quantifier, making the sentence an existential proposition. Such
a proposition claims that there is at least one object that satisfies the predicate. In particular, although other sentences mean the same as (2.1) assuming that ๐‘ฅ represents a
real number, such as
there exists a real number ๐‘ฅ such that ๐‘ฅ + 2 = 7,
๐‘ฅ + 2 = 7 for some real number x,
and
some real number ๐‘ฅ satisfies ๐‘ฅ + 2 = 7,
they all claim along with (2.1) that there is a real number ๐‘ฅ such that ๐‘ฅ + 2 = 7. This
is true because 5 satisfies ๐‘ฅ + 2 = 7 (Figure 2.1).
66
Chapter 2 FIRST-ORDER LOGIC
The real numbers
1
3
Satisfies x + 2 = 7
2
5
4
There is at least one object that satisfies ๐‘ฅ + 2 = 7. Therefore,
there exists a real number ๐‘ฅ such that ๐‘ฅ + 2 = 7 is true.
Figure 2.1
Again, assume that ๐‘ฅ is a real number. The sentence
for all ๐‘ฅ, ๐‘ฅ + 5 = 5 + ๐‘ฅ
(2.2)
claims that ๐‘ฅ + 5 = 5 + ๐‘ฅ is true for every real number ๐‘ฅ. To see this, break (2.2) like
(2.1):
Quantifier
↓
for all ๐‘ฅ
,
Predicate
↓
๐‘ฅ+5=5+๐‘ฅ
This quantifier is called the universal quantifier. It states that the predicate is satisfied
by each and every object. Including propositions that have the same meaning as (2.2),
such as
for all real numbers ๐‘ฅ, ๐‘ฅ + 5 = 5 + ๐‘ฅ,
and
๐‘ฅ + 5 = 5 + ๐‘ฅ for every real number x,
(2.2) is true because each substitution of a real number for ๐‘ฅ satisfies the predicate
(Figure 2.2). These are examples of universal propositions,.
A proposition can have multiple quantifiers. Take the equation ๐‘ฆ = 2๐‘ฅ2 + 1. Before
we learned the various techniques that make graphing this equation simple, we graphed
it by writing a table with one column holding the ๐‘ฅ values and another holding the ๐‘ฆ
values. We then chose numbers to substitute for ๐‘ฅ and calculated the corresponding ๐‘ฆ.
Although we did not explicitly write it this way, we learned that
for every real number ๐‘ฅ,
there exists a real number ๐‘ฆ such that ๐‘ฆ = 2๐‘ฅ2 + 1.
This is a universal proposition, and its predicate is
there exists a real number ๐‘ฆ such that ๐‘ฆ = 2๐‘ฅ2 + 1.
(2.3)
Section 2.1 LANGUAGES
The real numbers
1
3
Satisfies x + 5 = 5 + x
2
4
67
5
Every object satisfies ๐‘ฅ + 5 = 5 + ๐‘ฅ. Therefore,
for all real numbers ๐‘ฅ, ๐‘ฅ + 5 = 5 + ๐‘ฅ is true.
Figure 2.2
We conclude that (2.3) is true because whenever ๐‘ฅ is replaced with any real number,
there is a real number ๐‘ฆ that satisfies ๐‘ฆ = 2๐‘ฅ2 + 1.
The symbolic logic that we define in this chapter is intended to model propositions
that have a predicate and possibly a quantifier. As with propositional logic, an alphabet
and a grammar will be chosen that enable us to write down the appropriate symbols
to represent these sentences. Unlike propositional logic where we worked on both its
syntax and semantics at the same time, this logic starts with a study only of its syntax.
Alphabets
No matter what mathematical subject we study, whether it is algebra, number theory, or
something else, we can write our conclusions as propositions. These sentences usually
involve mathematical symbols particular to the subject being studied. For example,
both (2.1) and (2.2) are algebraic propositions. We know this because we recognize
the symbols from algebra class. For this reason, the symbolic logic that we define will
consist of two types of symbols.
DEFINITION 2.1.2
Logic symbols consist of the following:
โˆ™ Variable symbols: Sometimes simply referred to as variables, these symbols serve as placeholders. On their own, they are without meaning but
can be replaced with symbols that do have meaning. A common example
are the variable symbols in an algebraic equation. Variable symbols can
be lowercase English letters ๐‘ฅ, ๐‘ฆ, and ๐‘ง, or lowercase English letters with
subscripts, ๐‘ฅ๐‘› , ๐‘ฆ๐‘› , and ๐‘ง๐‘› . Depending on the context, variable symbols are
sometimes expanded to include uppercase English letters, with or without
subscripts, as well as Greek and Hebrew letters. Denote the collection of
variable symbols by ๐–ต๐– ๐–ฑ.
โˆ™ Connectives: ¬, ∧, ∨, →, ↔
68
Chapter 2 FIRST-ORDER LOGIC
โˆ™ Quantifier symbols: ∀, ∃
โˆ™ Equality symbol: =
โˆ™ Grouping symbols: (, ), [, ], {, }
Theory symbols consist of the following:
โˆ™ Constant symbols: These are used to represent important objects in the
subject that do not change. Common constant symbols are 0 and ๐‘’.
โˆ™ N-ary function symbols: The term n-ary refers to the number of arguments. For example, these symbols can represent unary functions that
take one argument such as cosine or binary functions such as multiplication that take two arguments.
โˆ™ N-ary relation symbols: These symbols are used to represent relations.
For example, < represents the binary relation of less than and ๐‘… can represent the unary relation is an even number.
It is not necessary to choose any theory symbols. However, if there are any, theory symbols must be chosen so that they are not connectives, quantifier symbols,
the equality symbol, or grouping symbols, and the selection of theory symbols
has precedence over the selection of variable symbols. This means that these two
collections must have no common symbols. Moreover, although this is not the
case in general, we assume that the logic symbols that are not variable symbols
will be the same for all applications. On the other hand, theory symbols (if there
are any) vary depending on the current subject of study. A collection of all logic
symbols and any theory symbols is called a first-order alphabet and is denoted
by ๐– .
The term theory refers to a collection of propositions all surrounding a particular
subject. Since different theories have different notation (think about how algebraic
notation differs from geometric notation), alphabets change depending on the subject
matter. Let us then consider the alphabets for a number of theories that will be introduced later in the text. The foundational theory is the first example.
EXAMPLE 2.1.3
Set theory is the study of collections of objects. The ∈ is the only relation symbol, and it is binary. It has no other theory symbols. The theory symbols of set
theory are denoted by ๐–ฒ๐–ณ and are summarized in the following table.
๐–ฒ๐–ณ
Constant symbols
Function symbols
Relation symbol
∈
Section 2.1 LANGUAGES
69
EXAMPLE 2.1.4
Number theory is the study of the natural numbers. The symbols + and ⋅ represent regular addition and multiplication, respectively. As such they are binary
function symbols. These and its other theory symbols are indicated in the following table and are denoted by ๐–ญ๐–ณ.
๐–ญ๐–ณ
Constant symbols
0 1
Function symbols
+ ⋅
Relation symbols
Another approach to number theory is called Peano arithmetic (Peano 1889).
It is the study of the natural numbers using Peano’s axioms. It has a constant
symbol 0 and a unary function symbol ๐‘†. Denote these symbols by ๐– ๐–ฑ.
๐– ๐–ฑ
Constant symbol
0
Function symbol
๐‘†
Relation symbols
The Peano arithmetic symbols are sometimes extended to include symbols for
the operations of addition and multiplication and the less-than relation. Denote
this extended collection by ๐– ๐–ฑ′ .
๐– ๐–ฑ′
Constant symbol
0
Function symbols
๐‘† + ⋅
Relation symbol
<
EXAMPLE 2.1.5
Group theory is the study of groups. A group is a set with an operation that satisfies certain properties. Typically, the operation is denoted by the binary function
symbol โ—ฆ. There is also a constant represented by ๐‘’. Ring theory is the study
of rings. A ring is a set with two operations that satisfy certain properties. The
operations are usually denoted by the binary function symbols ⊕ and ⊗. The
constant symbol is โ—‹. These collections of theory symbols are denoted by ๐–ฆ๐–ฑ
and ๐–ฑ๐–จ, respectively, and are summarized in the following tables.
๐–ฆ๐–ฑ
Constant symbol
๐‘’
Function symbol
โ—ฆ
๐–ฑ๐–จ
Constant symbols
โ—‹
Function symbols
⊕ ⊗
Relation symbols
Relation symbol
Field theory is the study of fields. A field is a type of ring with extra properties,
so the theory symbols ๐–ฑ๐–จ can be used to write about fields. However, if an order is
defined on a field, a binary relation symbol is needed, and the result is the theory
of ordered fields. Denote these symbols by ๐–ฎ๐–ฅ.
๐–ฎ๐–ฅ
Constant symbol
โ—‹
Function symbols
⊕ ⊗
Relation symbol
<
Notice that the collections of theory symbols in the previous examples had at most
two constant symbols. This is typical since subjects of study usually have only a few
70
Chapter 2 FIRST-ORDER LOGIC
objects that require special recognition. However, there will be times when some extra
constants are needed to reference objects that may or may not be named by the constant
theory symbols. To handle these situations, we expand the given theory symbols by
adding new constant symbols.
DEFINITION 2.1.6
Let ๐–  be a first-order alphabet with theory symbols ๐–ฒ. When constant symbols
not in ๐–ฒ are combined with the symbols of ๐–ฒ, the resulting collection of theory
symbols is denoted by ๐–ฒ. The number of new constant symbols varies depending
on need.
For example, suppose that we are working in a situation where we need four constant
symbols in addition to the ones in ๐–ฎ๐–ฅ. Denote these new symbols by ๐‘1 , ๐‘2 , ๐‘3 , and ๐‘4 .
Then, ๐–ฎ๐–ฅ consists of these four constant symbols and the symbols from ๐–ฎ๐–ฅ.
Terms
For a string to represent a proposition or a predicate from a particular theory, each
nonlogic symbol of the string must be a theory symbol of that subject. For example,
∀๐‘ฅ(∈ ๐‘ฅ๐ด)
and
∨∃ ∈ ๐‘Ž๐‘4 {} ∧ ๐‘ฅ¬๐‘ฆ
are strings for set theory. However, some strings have a reasonable chance of being
given meaning, others do not. As with propositional logic, we need a grammar that
will determine which strings are legal for the logic. Because a predicate might have
variables, the types of representations that we want to make are more complicated than
those of propositional logic. Hence, the grammar also will be more complicated. We
begin with the next inductive definition (compare page 59).
DEFINITION 2.1.7
Let ๐–  be a first-order alphabet with theory symbols ๐–ฒ. An ๐—ฆ-term is a string over
๐–  such that
โˆ™ a variable symbol from ๐–  is an ๐–ฒ-term,
โˆ™ a constant symbol from ๐–ฒ is an ๐–ฒ-term,
โˆ™ ๐‘“ ๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1 is an ๐–ฒ-term if ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 are ๐–ฒ-terms and ๐‘“ is a function
symbol from ๐–ฒ.
Denote the collection of strings over ๐–  that are ๐–ฒ-terms by ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐– ).
The string ๐‘“ ๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1 is often written as ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) because it is common to
write functions using this notation. We must remember, however, that this notation can
Section 2.1 LANGUAGES
71
always be replaced with the notation of Definition 2.1.7. Furthermore, when writing
about a general ๐–ฒ-term where it is not important to mention ๐–ฒ, we often simply write
using the word term without the ๐–ฒ. We will follow this convention when writing about
similar definitions that require the theory symbols ๐–ฒ.
EXAMPLE 2.1.8
Here are examples of terms for each of the indicated theories. Assume that ๐‘ฅ, ๐‘ฆ
and ๐‘ง1 are variable symbols.
โˆ™ 0 [Peano arithmetic]
โˆ™ +๐‘ฅ๐‘ฆ [number theory]
โˆ™ โ—ฆ0๐‘ง1 [group theory].
The string +๐‘ฅ๐‘ฆ is typically written as ๐‘ฅ+๐‘ฆ, and the string โ—ฆ0๐‘Ž1 is typically written
as 0 โ—ฆ ๐‘Ž1 . If ๐–ญ๐–ณ is expanded to ๐–ญ๐–ณ by adding the constants ๐‘ and ๐‘‘, then +๐‘๐‘‘ is
an ๐–ญ๐–ณ-term.
As suggested by Example 2.1.8, the purpose of a term is to represent an object of study.
A variable symbol represents an undetermined object. A constant symbol represents an
object that does not change, such as a number. A function symbol is used to represent
a particular object given another object. For example, the ๐–ญ๐–ณ-term +๐‘ฅ2 represents the
number that is obtained when ๐‘ฅ is added to 2.
Formulas
As propositional forms are used to symbolize propositions, the next definition is the
grammar used to represent propositional forms and predicates. The definition is given
inductively and resembles the parentheses-less prefix notation invented by ลukasiewicz
(1951) in the early 1920s.
DEFINITION 2.1.9
Let ๐–ฒ be the theory symbols from a first-order alphabet ๐– . An ๐—ฆ-formula is a
string over ๐–  such that
โˆ™ ๐‘ก0 = ๐‘ก1 is an ๐–ฒ-formula if ๐‘ก0 and ๐‘ก1 are ๐–ฒ-terms.
โˆ™ ๐‘…๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1 is an ๐–ฒ-formula if ๐‘… is an ๐‘›-ary relation symbol from ๐–ฒ and
๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 are ๐–ฒ-terms.
โˆ™ ¬๐‘ is an ๐–ฒ-formula if ๐‘ is an ๐–ฒ-formula.
โˆ™ ๐‘ ∧ ๐‘ž, ๐‘ ∨ ๐‘ž, ๐‘ → ๐‘ž, and ๐‘ ↔ ๐‘ž are ๐–ฒ-formulas if ๐‘ and ๐‘ž are ๐–ฒ-formulas.
โˆ™ ∀๐‘ฅ๐‘ and ∃๐‘ฅ๐‘ are ๐–ฒ-formulas if ๐‘ is an ๐–ฒ-formula and ๐‘ฅ is a variable symbol
from ๐– .
72
Chapter 2 FIRST-ORDER LOGIC
The string ๐‘…๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1 is often represented as ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ). A formula of
the form ∀๐‘ฅ๐‘ is called a universal formula, and a formula of the form ∃๐‘ฅ๐‘ is an
existential formula. Parentheses can be used around an ๐–ฒ-formula for readability, especially when quantifier symbols are involved. For example, ∀๐‘ฅ∃๐‘ฆ๐‘ is the
same formula as ∀๐‘ฅ(∃๐‘ฆ๐‘) and ∀๐‘ฅ∃๐‘ฆ(๐‘).
EXAMPLE 2.1.10
Let ๐‘ฅ and ๐‘ฆ be variable symbols. Let ๐‘ be a constant symbol; ๐‘“ , ๐‘”, and โ„Ž be unary
function symbols; and ๐‘… be a binary relation symbol from ๐–ฒ. The following are
๐–ฒ-formulas.
โˆ™ ๐‘ฅ=๐‘
โˆ™ ๐‘…๐‘๐‘“ ๐‘ฆ
โˆ™ ¬(๐‘ฆ = ๐‘”๐‘)
โˆ™ ๐‘…๐‘ฅ๐‘“ ๐‘ฅ → ๐‘…๐‘“ ๐‘ฅ๐‘ฅ
โˆ™ ∀๐‘ฅ¬(๐‘“ ๐‘ฅ = ๐‘“ ๐‘)
โˆ™ ∃๐‘ฅ∀๐‘ฆ(๐‘…๐‘“ ๐‘ฅ๐‘”๐‘ฆ ∧ ๐‘…๐‘“ ๐‘ฅโ„Ž๐‘ฆ).
In practice, ๐‘…๐‘๐‘“ ๐‘ฆ is usually written as ๐‘…(๐‘, ๐‘“ (๐‘ฆ)), ¬(๐‘ฆ = ๐‘”๐‘) as ๐‘ฆ ≠ ๐‘”(๐‘), ๐‘…๐‘ฅ๐‘“ ๐‘ฅ
as ๐‘…(๐‘ฅ, ๐‘“ (๐‘ฅ)), ๐‘…๐‘“ ๐‘ฅ๐‘ฅ as ๐‘…(๐‘“ (๐‘ฅ), ๐‘ฅ), and ๐‘…๐‘“ ๐‘ฅ๐‘”๐‘ฆ as ๐‘…(๐‘“ (๐‘ฅ), ๐‘”(๐‘ฆ)).
EXAMPLE 2.1.11
These are some ๐–ฒ๐–ณ-formulas with their standard translations.
โˆ™ ¬ ∈ ๐‘ฅ{ }
๐‘ฅ is not an element of { }.
โˆ™ ∀๐‘ฅ(∈ ๐‘ฅ๐ด → ∈๐‘ฅ๐ต)
For all ๐‘ฅ, if ๐‘ฅ ∈ ๐ด, then ๐‘ฅ ∈ ๐ต.
โˆ™ ¬∃๐‘ฅ∀๐‘ฆ(∈๐‘ฆ๐‘ฅ)
It is not the case that there exists ๐‘ฅ such that for all ๐‘ฆ, ๐‘ฆ ∈ ๐‘ฅ.
โˆ™ ๐‘ฅ = ๐‘ฆ ∨ ∈๐‘ฆ๐‘ฅ ∨ ∈๐‘ฅ๐‘ฆ
๐‘ฅ = ๐‘ฆ or ๐‘ฆ ∈ ๐‘ฅ or ๐‘ฅ ∈ ๐‘ฆ.
EXAMPLE 2.1.12
Here are some ๐–ญ๐–ณ-formulas with their corresponding predicates.
โˆ™ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง [++๐‘ฅ๐‘ฆ๐‘ง = +๐‘ฅ+๐‘ฆ๐‘ง]
(๐‘ฅ + ๐‘ฆ) + ๐‘ง = ๐‘ฅ + (๐‘ฆ + ๐‘ง) for all ๐‘ฅ, ๐‘ฆ, and ๐‘ง.
โˆ™ ¬(๐‘ฅ = 0) → ∀๐‘ฆ∃๐‘ง(๐‘ง = ⋅๐‘ฅ๐‘ฆ)
If ๐‘ฅ ≠ 0, then for all ๐‘ฆ, there exists ๐‘ง such that ๐‘ง = ๐‘ฅ๐‘ฆ.
Section 2.1 LANGUAGES
73
โˆ™ ∀๐‘ฅ(⋅1๐‘ฅ = ๐‘ฅ)
For all ๐‘ฅ, 1๐‘ฅ = ๐‘ฅ.
We now name the system just developed.
DEFINITION 2.1.13
An alphabet combined with a grammar is called a language. The language given
by Definitions 2.1.2 and 2.1.9 is known as a first-order language. A formula
of a first-order language is a first-order formula, and all of the formulas of the
first-order language with theory symbols ๐–ฒ is denoted by ๐–ซ(๐–ฒ).
The first-order language developed to represent predicates, either with quantifiers
or without them, is summarized in Figure 2.3. An alphabet that has both logic and
theory symbols is chosen. Using a grammar, terms are defined, and then by extending
the grammar, formulas are defined. A natural question to ask regarding this system is
what makes it first-order? Look at the last part of Definition 2.1.9. It gives the rule that
allows the addition of a quantifier symbol in a formula. Only ∀๐‘ฅ or ∃๐‘ฅ are permitted,
where ๐‘ฅ is a variable symbol representing an object of study. Thus, only propositions
that begin as for all x . . . or there exists x such that . . . can be represented as a firstorder formula. To quantify over function and relation symbols, we need to define a
second-order formula. Augment the alphabet ๐–  with function and relation symbols,
which creates what is know as a second-order alphabet, and then add
∀๐‘“ ๐‘ and ∃๐‘“ ๐‘ are ๐–ฒ-formulas if ๐‘ is an ๐–ฒ-formula
and ๐‘“ is a function symbol from ๐– 
and
∀๐‘…๐‘ and ∃๐‘…๐‘ are ๐–ฒ-formulas if ๐‘ is an ๐–ฒ-formula
and ๐‘… is a relation symbol from ๐– 
to Definition 2.1.9. For example, if the first-order formula
∀๐‘ฅ(∈ ๐‘ฅ๐ด →∈ ๐‘ฅ๐ต)
is intended to be true for any ๐ด and ๐ต, it can be written as the second-order formula
∀๐ด∀๐ต∀๐‘ฅ(๐ด๐‘ฅ → ๐ต๐‘ฅ),
where ๐ด๐‘ฅ represents the relation ๐‘ฅ ∈ ๐ด and ๐ต๐‘ฅ represents the relation ๐‘ฅ ∈ ๐ต.
Logic symbols
Grammar
Theory symbols
Terms
Grammar
Represent
objects
Alphabet
Figure 2.3
A first-order language.
Formulas
Properties
of objects
74
Chapter 2 FIRST-ORDER LOGIC
Exercises
1. Determine whether the given strings are ๐–ฆ๐–ฑ-terms. If a string is not a ๐–ฆ๐–ฑ-term, find
all issues that prevent it from being one.
(a) 2๐‘’
(b) ๐‘ฅ๐‘ฆโ—ฆ
(c) โ—ฆ
(d) <๐‘ฅ๐‘’
(e) โ—ฆโ—ฆ๐‘ฅ๐‘ฆ๐‘’
2. Write the ๐–ฆ๐–ฑ-terms from Exercise 1 in their usual form (Example 2.1.10).
3. Extend ๐–ฒ๐–ณ to ๐–ฒ๐–ณ′ by adding the constant symbol ∅ and the unary function symbol
P. Determine whether the given strings are ๐–ฒ๐–ณ′ -terms. If a string is not a ๐–ฒ๐–ณ′ -term,
find all issues that prevent it from being one.
(a) ∈๐‘ฅ∅
(b) ∅
(c) ๐‘ฅ
(d) P∈๐‘ฅ๐‘ฆ
(e) P๐‘ฅ
4. Determine whether the given strings are ๐–ฎ๐–ฅ-formulas. If a given string is not a
๐–ฎ๐–ฅ-formula, find all issues that prevent it from being one.
(a) < ⊕๐‘ฅ๐‘ฆ⊗๐‘ฅ๐‘ฆ
(b) ⊕๐‘ฅ = ⊗๐‘ฆ
(c) ∀๐‘ฅ∃๐‘ฆ(+๐‘ฅ๐‘ฆ = 0)
(d) ∀๐‘ฅ∀๐‘ฆ(⊗๐‘ฅ๐‘ฆ = ⊕๐‘ฆ๐‘ฅ) → <๐‘ฅ๐‘ฆ
(e) ¬(<⊗๐‘ฅ๐‘ฆ ∧ ⊕๐‘ฅ๐‘ฆ) ↔ ∀๐‘ข∀๐‘ฃ(<⊗๐‘ฅ๐‘ฆ⊕๐‘ข๐‘ฃ)
5. Write the ๐–ฑ๐–จ-formulas from Exercise 4 in their usual form.
6. Extend ๐–ฒ๐–ณ to ๐–ฒ๐–ณ′′ by adding the constant symbol ∅ and the binary relation symbol
๐‘…. Determine whether the given strings are ๐–ฒ๐–ณ′′ -formulas. If a given string is not a
๐–ฒ๐–ณ′′ -formula, find all issues that prevent it from being one.
(a) ๐‘…๐‘ฅ∅ ∨ ∈๐‘ฅ๐‘ฆ
(b) ๐‘…๐‘ฅ∅ → ∈+๐‘ฅ๐‘ฆ๐‘ง
(c) ∀๐‘ฅ∃∅(¬∈๐‘ฅ∅)
(d) ∀๐‘ฅ ∨ ∀๐‘ฆ ∨ ∀๐‘ง(∈∅๐‘ฅ๐‘ฆ๐‘ง) ∨ ∈∅∅
(e) [∃๐‘ฅ(๐‘…๐‘ฅ∅) → ∃๐‘ข∃๐‘ฃ∃๐‘ค(¬∈๐‘ข๐‘ฃ ↔ ¬∈๐‘ข๐‘ค)] ∨ ∅ = ๐‘ฅ1
7. Write the ๐–ฒ๐–ณ′′ -formulas from Exercise 6 in their usual form.
8. Translate the given sentence to an ๐–ฒ-formula for the given theory symbol.
(a) For all ๐‘ฅ, ๐‘†(๐‘ฅ) ≠ 0. (๐–ฒ = ๐– ๐–ฑ)
(b) For every number ๐‘ฅ, there is a number ๐‘ฆ such that ๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’. (๐–ฒ = ๐–ฆ๐–ฑ)
(c) If ๐‘ฅ < ๐‘ฆ and ๐‘ฆ < ๐‘ง, then ๐‘ฅ < ๐‘ง. (๐–ฒ = ๐–ฎ๐–ฅ)
(d) It is false that ๐‘ฅ ∈ ๐‘ฆ, ๐‘ฆ ∈ ๐‘ง, and ๐‘ง ∈ ๐‘ฅ for all ๐‘ฅ, ๐‘ฆ, and ๐‘ง. (๐–ฒ = ๐–ฒ๐–ณ)
Section 2.2 SUBSTITUTION
75
(e) For every ๐‘ข and ๐‘ฃ, there exists ๐‘ค such that if ๐‘ข = ๐‘ฃ, then ๐‘ข + ๐‘ค = ๐‘ฃ + ๐‘ค.
(๐–ฒ = ๐–ฑ๐–จ and + should be translated as ⊕ in the ๐–ฑ๐–จ-formula.)
9. Extend ๐–ญ๐–ณ to ๐–ญ๐–ณ′ by adding the numerals 2, 3, … , 9. Answer the following questions.
(a) Is +34 = 7 a ๐–ญ๐–ณ′ -term? Explain.
(b) Is ⋅4+39 a ๐–ญ๐–ณ′ -term? Explain.
(c) Is ∃๐‘ฅ(+⋅4๐‘ฅ8 = +3๐‘ฅ) a ๐–ญ๐–ณ′ -formula? If it is, find ๐‘ฅ.
(d) If possible, give an example of a ๐–ญ๐–ณ-formula that is not a ๐–ญ๐–ณ′ -formula.
(e) If possible, give an example of a ๐–ญ๐–ณ′ -formula that is not a ๐–ญ๐–ณ-formula.
2.2
SUBSTITUTION
As noted in the beginning of Section 2.1, there are times when a substitution will be
made for a predicate’s variable. For example, in algebra, if ๐‘“ (๐‘ฅ) = 3๐‘ฅ2 + ๐‘ฅ + 1, then
๐‘“ (๐‘ฆ) = 3๐‘ฆ2 + ๐‘ฆ + 1, ๐‘“ (2) = 3(2)2 + 2 + 1, and ๐‘“ (sin ๐‘ฅ) = 3 sin2 ๐‘ฅ + sin ๐‘ฅ + 1. That is,
we can substitute with variables, constants, and the results from functions. Therefore,
to represent this in formulas, we can replace a variable with a term.
Terms
We begin by defining what it means to substitute for a variable in a term. We use ⇔
because one string can be replaced with the other string.
DEFINITION 2.2.1 [Substitution in Terms]
Let ๐–ฒ be theory symbols from a first-order alphabet ๐– . Let ๐‘ฅ be a variable symbol
from ๐–  and ๐‘ก be an ๐–ฒ-term. The notation
๐‘ก
๐‘ฅ
means that ๐‘ฅ is replaced with ๐‘ก at every appropriate occurrence of ๐‘ฅ. For terms,
this means the following.
โˆ™ If ๐‘ฆ is a variable symbol from ๐– ,
๐‘ก
๐‘ฆ ⇔
๐‘ฅ
{
๐‘ก if ๐‘ฅ = ๐‘ฆ,
๐‘ฆ if ๐‘ฅ ≠ ๐‘ฆ.
โˆ™ If ๐‘ is a constant symbol from ๐–ฒ,
๐‘
๐‘ก
⇔ ๐‘.
๐‘ฅ
โˆ™ If ๐‘“ is an ๐‘›-ary function symbol from ๐–ฒ and ๐‘ 0 , ๐‘ 1 , … , ๐‘ ๐‘›−1 are ๐–ฒ-terms,
(
)
๐‘ก
๐‘ก ๐‘ก
๐‘ก
(๐‘“ ๐‘ 0 ๐‘ 1 · · · ๐‘ ๐‘›−1 ) ⇔ ๐‘“ ๐‘ 0 ๐‘ 1 · · · ๐‘ ๐‘›−1
.
๐‘ฅ
๐‘ฅ ๐‘ฅ
๐‘ฅ
76
Chapter 2 FIRST-ORDER LOGIC
Observe that when a substitution of a term is made into a term, the result is another
term.
EXAMPLE 2.2.2
Let ๐‘ฅ, ๐‘ฆ, and ๐‘ง be distinct variable symbols; ๐‘ and ๐‘‘ be constant symbols; ๐‘“ be
a binary function symbol; and ๐‘” and โ„Ž be unary function symbols. This means
that ๐‘“ ๐‘๐‘”๐‘”๐‘‘ is typically written as ๐‘“ (๐‘, ๐‘”(๐‘”(๐‘‘))).
๐‘ฆ
โˆ™ ๐‘ฅ ⇔๐‘ฆ
๐‘ฅ
๐‘
โˆ™ ๐‘ฅ ⇔๐‘
๐‘ฅ
๐‘“ ๐‘ฅ๐‘ง
⇔๐‘ฆ
๐‘ฅ
๐‘ฅ
๐‘ ⇔๐‘
๐‘ฅ
๐‘
๐‘‘ ⇔๐‘‘
๐‘ฅ
๐‘”๐‘ฅ
๐‘
⇔๐‘
๐‘ฅ
๐‘
(๐‘“ ๐‘ฅ๐‘ฆ) ⇔ ๐‘“ ๐‘ฅ๐‘
๐‘ฆ
(
)
โ„Ž๐‘ ๐‘”๐‘ฆ
(๐‘“ ๐‘ฆ๐‘ง)
⇔ ๐‘“ โ„Ž๐‘๐‘ง
๐‘ฆ
๐‘ฅ
([( ) ] )
๐‘ ๐‘‘ ๐‘ ๐‘‘
๐‘ง
⇔๐‘
๐‘ฅ ๐‘ฆ ๐‘ง ๐‘ฅ
) ] )
([(
๐‘ ๐‘”๐‘ฅ ๐‘ฅ ๐‘”๐‘‘
⇔ ๐‘“ ๐‘๐‘”๐‘”๐‘‘.
๐‘“ ๐‘ฅ๐‘ฆ
๐‘ฅ ๐‘ฆ ๐‘ง ๐‘ฅ
โˆ™ ๐‘ฆ
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
Free Variables
Substitution for a variable in a formula is a bit more involved. This is because of the
influence of any possible quantifiers. For example, take the formula
๐‘ฅ = ๐‘ฆ ∨ ๐‘ฅ = ๐‘“ ๐‘ฆ.
(2.4)
By Definition 2.2.1, we know that we can substitute constants ๐‘ for ๐‘ฅ and ๐‘‘ for ๐‘ฆ in the
terms ๐‘ฅ, ๐‘ฆ, and ๐‘“ ๐‘ฆ. We expect that we should also be able to make this substitution
into the formula resulting in
๐‘ = ๐‘‘ ∨ ๐‘ = ๐‘“ ๐‘‘.
However, in the formula
∀๐‘ฅ∃๐‘ฆ(๐‘ฅ = ๐‘ฆ ∨ ๐‘ฅ = ๐‘“ ๐‘ฆ),
(2.5)
Section 2.2 SUBSTITUTION
77
the situation is different because of the quantifiers. Consider the corresponding occurrences in (2.4) and its quantified (2.5):
x=y∨x=fy
∀x∃y(x = y ∨ x = f y)
Even though each occurrence of ๐‘ฅ and ๐‘ฆ in (2.4) can receive a substitution, each corresponding occurrence in (2.5) cannot because of the quantifiers.
DEFINITION 2.2.3
Let ๐–ฒ be theory symbols. Let ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 be ๐–ฒ-terms, ๐‘… be an ๐‘›-ary relation
symbol from ๐–ฒ, and ๐‘ and ๐‘ž be ๐–ฒ-formulas. An occurrence of a variable in a
formula is free or not free only according to the following rules:
โˆ™ A variable occurrence in ๐‘ก0 = ๐‘ก1 and ๐‘…๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1 is free.
โˆ™ A variable occurrence in ¬๐‘ is free if and only if the corresponding occurrence in ๐‘ is free.
โˆ™ A variable occurrence in ๐‘ ∧ ๐‘ž, ๐‘ ∨ ๐‘ž, ๐‘ → ๐‘ž, and ๐‘ ↔ ๐‘ž is free if and only
if the corresponding occurrence in ๐‘ or ๐‘ž is free.
โˆ™ Any occurrence of ๐‘ฅ in ∀๐‘ฅ๐‘ and ∃๐‘ฅ๐‘ is not free.
โˆ™ Any occurrence of ๐‘ฆ ≠ ๐‘ฅ is free in ∀๐‘ฅ๐‘ and ∃๐‘ฅ๐‘ if and only if the corresponding occurrence of ๐‘ฆ is free in ๐‘.
If an occurrence of a variable is not free, it is bound. All free occurrences of
๐‘ฅ in ๐‘ are within the scope of the universal quantifier in ∀๐‘ฅ๐‘ and the existential
quantifier in ∃๐‘ฅ๐‘.
EXAMPLE 2.2.4
Let ๐‘“ be a unary function symbol and ๐‘… be a 3-ary relation symbol. In the
formula
∀๐‘ฅ∃๐‘ฆ(๐‘ฅ = ๐‘ฆ ∨ ๐‘“ ๐‘ฅ = ๐‘ฆ → ๐‘…๐‘ฅ๐‘ฆ๐‘ง),
all occurrences of ๐‘ฅ and ๐‘ฆ are bound, but the occurrence of ๐‘ง is free. In the
formula
∃๐‘ฆ(๐‘ฅ = ๐‘ฆ ∨ ๐‘“ ๐‘ฅ = ๐‘ฆ → ๐‘…๐‘ฅ๐‘ฆ๐‘ง),
all occurrences of ๐‘ฅ and ๐‘ง are free, but the occurrences of ๐‘ฆ are bound. In
๐‘ฅ = ๐‘ฆ ∨ ๐‘“ ๐‘ฅ = ๐‘ฆ → ๐‘…๐‘ฅ๐‘ฆ๐‘ง,
all occurrences are free because all occurrences are free in ๐‘ฅ = ๐‘ฆ, ๐‘“ ๐‘ฅ = ๐‘ฆ, and
๐‘…๐‘ฅ๐‘ฆ๐‘ง.
We need to know whether a formula has a free occurrence of a variable, so we make
the next definition.
78
Chapter 2 FIRST-ORDER LOGIC
DEFINITION 2.2.5
A free variable of the ๐–ฒ-formula ๐‘ is a variable that has a free occurrence in ๐‘.
EXAMPLE 2.2.6
Using the ๐‘“ and ๐‘… from Example 2.2.4, both ๐‘ฅ and ๐‘ฆ are free variables of the
formula
๐‘…๐‘ฅ๐‘ฆ๐‘ → ∃๐‘ฅ(๐‘“ ๐‘ฆ = ๐‘),
even though ๐‘ฅ has both free and bound occurrences.
Formulas
We can now define what it means to make a substitution into a formula.
DEFINITION 2.2.7 [Substitution in Formulas]
Let ๐–ฒ be theory symbols from a first-order alphabet ๐– . Let ๐‘ฅ and ๐‘ฆ be variable
symbols of ๐–  and ๐‘… be an ๐‘›-ary relation symbol from ๐–ฒ. Suppose that ๐‘ and ๐‘ž
are ๐–ฒ-formulas and ๐‘ก, ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 are ๐–ฒ-terms.
๐‘ก
๐‘ก
๐‘ก
⇔ ๐‘ก0 = ๐‘ก1
๐‘ฅ
๐‘ฅ
๐‘ฅ
๐‘ก ๐‘ก
๐‘ก
๐‘ก
(๐‘…๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1 ) ⇔ ๐‘…๐‘ก0 ๐‘ก1 · · · ๐‘ก๐‘›−1
๐‘ฅ
๐‘ฅ ๐‘ฅ
๐‘ฅ
( )
๐‘ก
๐‘ก
(¬๐‘) ⇔ ¬ ๐‘
๐‘ฅ
๐‘ฅ
๐‘ก
๐‘ก
๐‘ก
(๐‘ ∧ ๐‘ž) ⇔ ๐‘ ∧ ๐‘ž
๐‘ฅ
๐‘ฅ
๐‘ฅ
๐‘ก
๐‘ก
๐‘ก
(๐‘ ∨ ๐‘ž) ⇔ ๐‘ ∨ ๐‘ž
๐‘ฅ
๐‘ฅ
๐‘ฅ
๐‘ก
๐‘ก
๐‘ก
(๐‘ → ๐‘ž) ⇔ ๐‘ → ๐‘ž
๐‘ฅ
๐‘ฅ
๐‘ฅ
๐‘ก
๐‘ก
๐‘ก
(๐‘ ↔ ๐‘ž) ⇔ ๐‘ ↔ ๐‘ž
๐‘ฅ
๐‘ฅ
๐‘ฅ
{ ( ๐‘ก)
∀๐‘ฆ ๐‘
if ๐‘ฅ ≠ ๐‘ฆ and ๐‘ฆ is not a symbol of ๐‘ก
๐‘ก
๐‘ฅ
(∀๐‘ฆ๐‘) ⇔
๐‘ฅ
∀๐‘ฆ๐‘
otherwise
(
)
{
๐‘ก
∃๐‘ฆ ๐‘
if ๐‘ฅ ≠ ๐‘ฆ and ๐‘ฆ is not a symbol of ๐‘ก
๐‘ก
๐‘ฅ
(∃๐‘ฆ๐‘) ⇔
๐‘ฅ
∃๐‘ฆ๐‘
otherwise.
โˆ™ (๐‘ก0 = ๐‘ก1 )
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
โˆ™
The condition on the term ๐‘ก in the last two parts of Definition 2.2.7 is important.
Consider the ๐–ฑ๐–จ-formula
๐‘ := ∃๐‘ฆ(๐‘ฅ ⊕ ๐‘ฆ = โ—‹).
79
Section 2.2 SUBSTITUTION
The usual interpretation of ๐‘ is that given ๐‘ฅ, there exists an number ๐‘ฆ such that ๐‘ฅ+๐‘ฆ = 0.
Let ๐‘“ be a unary function symbol and ๐‘ง be a variable symbol. Since ๐‘ฆ is not a symbol
of ๐‘ง,
๐‘ง
๐‘ ⇔ ∃๐‘ฆ(๐‘ง ⊕ ๐‘ฆ = โ—‹),
๐‘ฅ
which has the same standard interpretation as ๐‘. Since ๐‘ฆ is not a symbol of ๐‘“ ๐‘ง, we can
substitute to find that
๐‘“๐‘ง
⇔ ∃๐‘ฆ(๐‘“ ๐‘ง ⊕ ๐‘ฆ = โ—‹).
๐‘
๐‘ฅ
This is a reasonable substitution because it states that for the number given by ๐‘“ ๐‘ง, there
is a number ๐‘ฆ that when added to ๐‘“ ๐‘ง gives 0. This is very similar to the standard interpretation of ๐‘. Both of these substitutions work because the number of free occurrences
is unchanged by the substitution. However, if we allow the term to include ๐‘ฆ among its
๐‘ฆ
symbols, the substitution ๐‘ would yield
๐‘ฅ
∃๐‘ฆ(๐‘ฆ ⊕ ๐‘ฆ = โ—‹).
(2.6)
The typical interpretation of (2.6) is that there exists a number ๐‘ฆ such that ๐‘ฆ + ๐‘ฆ = 0.
This is a reasonable proposition, but not in the spirit of the original formula ๐‘. The
change of interpretation is due to the change in the number of free occurrences. The
formula ๐‘ has one free occurrence, while (2.6) has none. Therefore, when making a
substitution, the number of free occurrences should not change, and for this reason,
∃๐‘ฆ(๐‘ฅ ⊕ ๐‘ฆ = โ—‹)
๐‘ฆ
⇔ ∃๐‘ฆ(๐‘ฅ ⊕ ๐‘ฆ = โ—‹).
๐‘ฅ
EXAMPLE 2.2.8
Let ๐‘ be the ๐–ญ๐–ณ formula
∀๐‘ฅ0 ∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + ๐‘ฅ2 = ๐‘ฅ1 + ๐‘ฅ2 ).
Notice that ๐‘ฅ2 is a free variable of ๐‘. However, all occurrences of ๐‘ฅ0 and ๐‘ฅ1 are
bound. Therefore,
๐‘
0
⇔ ∀๐‘ฅ0 ∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + ๐‘ฅ2 = ๐‘ฅ1 + ๐‘ฅ2 ),
๐‘ฅ0
and then letting ๐‘ฆ be a variable symbol,
(
)
๐‘ฆ
0
๐‘
⇔ ∀๐‘ฅ0 ∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + ๐‘ฅ2 = ๐‘ฅ1 + ๐‘ฅ2 ),
๐‘ฅ0 ๐‘ฅ1
and finally,
[(
) ]
๐‘ฆ 1
0
๐‘
⇔ ∀๐‘ฅ0 ∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + 1 = ๐‘ฅ1 + 1).
๐‘ฅ0 ๐‘ฅ1 ๐‘ฅ2
This last formula has no free variables. A standard interpretation of this formula
is
80
Chapter 2 FIRST-ORDER LOGIC
for every integer ๐‘ฅ0 and ๐‘ฅ1 , if ๐‘ฅ0 = ๐‘ฅ1 , then ๐‘ฅ0 + 1 = ๐‘ฅ1 + 1.
Furthermore,
๐‘
๐‘ฅ
๐‘ฅ1
⇔ [∀๐‘ฅ0 ∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + ๐‘ฅ2 = ๐‘ฅ1 + ๐‘ฅ2 )] 1
๐‘ฅ2
๐‘ฅ2
๐‘ฅ1
⇔ ∀๐‘ฅ0 [∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + ๐‘ฅ2 = ๐‘ฅ1 + ๐‘ฅ2 )]
๐‘ฅ2
⇔ ∀๐‘ฅ0 ∀๐‘ฅ1 (๐‘ฅ0 = ๐‘ฅ1 → ๐‘ฅ0 + ๐‘ฅ2 = ๐‘ฅ1 + ๐‘ฅ2 ).
Let ๐‘ represent the ๐–ญ๐–ณ-formula +๐‘ฆ2 = 7. Observe that ๐‘ has ๐‘ฆ as a free variable. To
emphasize this, instead of writing
๐‘ := +๐‘ฆ2 = 7,
we often denote the formula by ๐‘(๐‘ฆ) and write
๐‘(๐‘ฆ) := +๐‘ฆ2 = 7.
Although ๐‘ฅ is not a variable in the equation, we can also write
๐‘ž(๐‘ฅ, ๐‘ฆ) := +๐‘ฆ2 = 7.
For example, if we wanted to interpret the formula as an equation with one variable,
we would use ๐‘(๐‘ฆ). If we wanted to view it as the horizontal line ๐‘ฆ = 5, we would use
๐‘ž(๐‘ฅ, ๐‘ฆ).
DEFINITION 2.2.9
Let ๐–  be a first-order alphabet with theory symbols ๐–ฒ. Let ๐‘ be an ๐–ฒ-formula. If
๐‘ has no free variables or the free variables of ๐‘ are among the distinct variables
๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 from ๐– , define
๐‘(๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 ) ⇔ ๐‘.
The notation of Definition 2.2.9 can also be used to represent substitutions. Consider
the formula ๐‘ := ๐‘ฅ + ๐‘ฆ = 0. Observe that
( )
๐‘ฅ ๐‘ฆ
๐‘
⇔๐‘ฆ+๐‘ฆ=0
๐‘ฆ ๐‘ฅ
and
( ๐‘ฆ)
๐‘ฅ
๐‘
⇔ ๐‘ฅ + ๐‘ฅ = 0.
๐‘ฅ ๐‘ฆ
However, suppose that we want to substitute into ๐‘ so that the result is ๐‘ฆ + ๐‘ฅ = 0. To
accomplish this, we need two new and distinct variable symbols, ๐‘ข and ๐‘ฃ. Then,
81
Section 2.2 SUBSTITUTION
( )
๐‘ข ๐‘ฃ
๐‘
⇔ ๐‘ข + ๐‘ฃ = 0,
๐‘ฅ ๐‘ฆ
(2.7)
( ๐‘ฆ)
๐‘ฅ
⇔ ๐‘ฆ + ๐‘ฅ = 0.
๐‘
๐‘ข ๐‘ฃ
(2.8)
and then,
Therefore,
([( ) ] )
๐‘ข ๐‘ฃ ๐‘ฆ ๐‘ฅ
๐‘
⇔ ๐‘ฆ + ๐‘ฅ = 0.
๐‘ฅ ๐‘ฆ ๐‘ข ๐‘ฃ
This works because in (2.7), each variable symbol is replaced by a new symbol in such
a way that the resulting formula has the same meaning as the original. In this way, when
the original variable symbols are brought back in (2.8), all of the substitutions are made
into distinct variables so that there are no conflicts and the switch can be made. This
process can be generalized to any number of variable symbols in any order.
DEFINITION 2.2.10
Let ๐‘(๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 ) be an ๐–ฒ-formula and ๐‘ข0 , ๐‘ข1 , … , ๐‘ข๐‘›−1 be distinct variable
symbols not among ๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 . Define
๐‘(๐‘ข0 , ๐‘ข1 , … , ๐‘ข๐‘›−1 ) ⇔
) ]
)
( [(
๐‘ข๐‘›−1
๐‘ข
๐‘ข1
···
··· ๐‘ 0
.
๐‘ฅ0 ๐‘ฅ1
๐‘ฅ๐‘›−1
Then, for all ๐–ฒ-terms ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 , define
) ]
)
( [(
๐‘ก๐‘›−1
๐‘ก0 ๐‘ก1
๐‘(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ⇔ · · · ๐‘
···
.
๐‘ข0 ๐‘ข1
๐‘ข๐‘›−1
This is called a simultaneous substitution and is equivalent to replacing all free
occurrences of ๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 in ๐‘ with ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 , respectively.
Observe that if ๐‘ does not have free variables, then ๐‘(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ⇔ ๐‘.
EXAMPLE 2.2.11
To illustrate Definition 2.2.10, let ๐–ฑ๐–จ′ be the theory symbols of ๐–ฑ๐–จ combined with
constant symbols 1 and 2. Let ๐‘ be the ๐–ฑ๐–จ′ -formula
๐‘ฅ ⊗ (๐‘ฆ ⊕ ๐‘ง) = ๐‘ฅ ⊗ ๐‘ฆ ⊕ ๐‘ฅ ⊗ ๐‘ง.
82
Chapter 2 FIRST-ORDER LOGIC
Since ๐‘ฅ, ๐‘ฆ, and ๐‘ง are free variables, represent ๐‘ by ๐‘(๐‘ฅ, ๐‘ฆ, ๐‘ง, ๐‘ค). Then,
([(
) ] )
1
๐‘ฅ ๐‘ฆ
2
๐‘(1, ๐‘ฅ, ๐‘ฆ, 2) ⇔
๐‘
๐‘ข1 ๐‘ข2 ๐‘ข3 ๐‘ข4
) ] )
([(
[
] 1
๐‘ฅ ๐‘ฆ 2
⇔
๐‘ข1 ⊗ (๐‘ข2 ⊕ ๐‘ข3 ) = ๐‘ข1 ⊗ ๐‘ข2 ⊕ ๐‘ข1 ⊗ ๐‘ข3
๐‘ข1 ๐‘ข2 ๐‘ข3 ๐‘ค
([
] )
(
) ๐‘ฅ ๐‘ฆ 2
⇔
1 ⊗ (๐‘ข2 ⊕ ๐‘ข3 ) = 1 ⊗ ๐‘ข2 ⊕ 1 ⊗ ๐‘ข3
๐‘ข ๐‘ข
๐‘ค
)2 3
(
[
] ๐‘ฆ 2
⇔ 1 ⊗ (๐‘ฅ ⊕ ๐‘ข3 ) = 1 ⊗ ๐‘ฅ ⊕ 1 ⊗ ๐‘ข3
๐‘ข3 ๐‘ค
2
⇔ (1 ⊗ (๐‘ฅ ⊕ ๐‘ฆ) = 1 ⊗ ๐‘ฅ ⊕ 1 ⊗ ๐‘ฆ)
๐‘ค
⇔ 1 ⊗ (๐‘ฅ ⊕ ๐‘ฆ) = 1 ⊗ ๐‘ฅ ⊕ 1 ⊗ ๐‘ฆ.
EXAMPLE 2.2.12
Let ๐–ฒ have constant symbols 5 and 9. Define ๐‘(๐‘ฅ, ๐‘ฆ) to be the ๐–ฒ-formula
∀๐‘ฅ∃๐‘ง [๐‘ž(๐‘ฅ, ๐‘ฆ) ∧ ๐‘Ÿ(๐‘ง)] ∨ ∃๐‘ฆ [๐‘Ÿ(๐‘ฆ) → ๐‘ (๐‘ฅ)] .
The first occurrence of ๐‘ฆ is free, and since ∀๐‘ฅ applies only to variables within the
brackets on the left, the last occurrence of ๐‘ฅ is also free. Therefore, ๐‘(๐‘ฅ, ๐‘ฆ) is the
disjunction of two formulas that we call ๐‘ข(๐‘ฆ) and ๐‘ฃ(๐‘ฅ):
๐‘ข(๐‘ฆ)
๐‘ฃ(๐‘ฅ)
โžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโž โžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโž
∀๐‘ฅ∃๐‘ง [๐‘ž(๐‘ฅ, ๐‘ฆ) ∧ ๐‘Ÿ(๐‘ง)] ∨ ∃๐‘ฆ [๐‘Ÿ(๐‘ฆ) → ๐‘ (๐‘ฅ)],
from which we derive
๐‘(9, 5) ⇔ ๐‘ข(5) ∨ ๐‘ฃ(9) ⇔ ∀๐‘ฅ∃๐‘ง[๐‘ž(๐‘ฅ, 5) ∧ ๐‘Ÿ(๐‘ง)] ∨ ∃๐‘ฆ[๐‘Ÿ(๐‘ฆ) → ๐‘ (9)]
As in Example 2.2.11, finding ๐‘(9, 5) is equivalent to simply replacing the free
occurrences of ๐‘ฅ with 9 and the occurrences of ๐‘ฆ with 5.
EXAMPLE 2.2.13
Consider the ๐–ญ๐–ณ-formula,
∀๐‘ฅ∀๐‘ฆ∀๐‘ง [++๐‘ฅ๐‘ฆ๐‘ง = +๐‘ฅ+๐‘ฆ๐‘ง] ,
(2.9)
∀๐‘ฅ∀๐‘ฆ∀๐‘ง [(๐‘ฅ + ๐‘ฆ) + ๐‘ง = ๐‘ฅ + (๐‘ฆ + ๐‘ง)] .
(2.10)
which is often written as
The formula within the scope of the first quantifier symbol in (2.10) is
๐‘(๐‘ฅ) := ∀๐‘ฆ∀๐‘ง [(๐‘ฅ + ๐‘ฆ) + ๐‘ง = ๐‘ฅ + (๐‘ฆ + ๐‘ง)] .
(2.11)
Section 2.2 SUBSTITUTION
83
Notice that the occurrences of ๐‘ฅ are free in (2.11), but the occurrences of ๐‘ฆ and
๐‘ง are bound. For example, we can make the substitution
๐‘(2) ⇔ ∀๐‘ฆ∀๐‘ง[(2 + ๐‘ฆ) + ๐‘ง = 2 + (๐‘ฆ + ๐‘ง)].
Now, letting
๐‘ž(๐‘ฅ, ๐‘ฆ) := ∀๐‘ง[(๐‘ฅ + ๐‘ฆ) + ๐‘ง = ๐‘ฅ + (๐‘ฆ + ๐‘ง)],
we have that ๐‘(๐‘ฅ) ⇔ ∀๐‘ฆ๐‘ž(๐‘ฅ, ๐‘ฆ). We can also define
๐‘Ÿ(๐‘ฅ, ๐‘ฆ, ๐‘ง) := (๐‘ฅ + ๐‘ฆ) + ๐‘ง = ๐‘ฅ + (๐‘ฆ + ๐‘ง)
so that ๐‘ž(๐‘ฅ, ๐‘ฆ) ⇔ ∀๐‘ง ๐‘Ÿ(๐‘ฅ, ๐‘ฆ, ๐‘ง). What we have done is to break apart an ๐–ญ๐–ณformula that contains multiple quantifiers into a sequence of formulas:
๐‘(๐‘ฅ)
โžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโž
∀๐‘ฅ ∀๐‘ฆ ∀๐‘ง [(๐‘ฅ + ๐‘ฆ) + ๐‘ง = ๐‘ฅ + (๐‘ฆ + ๐‘ง)] .
โŸโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโŸโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโŸ
๐‘Ÿ(๐‘ฅ, ๐‘ฆ, ๐‘ง)
โŸโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโŸโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโŸ
๐‘ž(๐‘ฅ, ๐‘ฆ)
We conclude that (2.10) can be written as
∀๐‘ฅ ๐‘(๐‘ฅ),
∀๐‘ฅ∀๐‘ฆ ๐‘ž(๐‘ฅ, ๐‘ฆ),
or
∀๐‘ฅ∀๐‘ฆ∀๐‘ง ๐‘Ÿ(๐‘ฅ, ๐‘ฆ, ๐‘ง).
Observe that (2.9) has no free variables. It is a type of formula of particular importance
because it represents a proposition. In other words, (2.9) is a propositional form.
DEFINITION 2.2.14
An ๐–ฒ-formula with no free variables is called an ๐—ฆ-sentence.
Exercises
1. Given a term, make the indicated substitution.
( )
5 6
(a) 4
๐‘ฆ
[(๐‘ฅ )
]
๐‘Ž ๐‘“ (๐‘ฅ) ๐‘
(b)
๐‘ฅ
๐‘ฆ
๐‘ฅ
๐‘ฅ
[
]
๐‘Ž ๐‘”(8)
(c) (๐‘ฅ + ๐‘ฆ)
๐‘ฅ ๐‘ฆ
[([
] ) ]
๐‘Ÿ ๐‘ ๐‘ 5
(d)
(๐‘ฅ + ๐‘ฆ + ๐‘ง)
๐‘ฅ ๐‘ฆ ๐‘ง ๐‘ฅ
84
Chapter 2 FIRST-ORDER LOGIC
2. Given a formula, make the indicated substitution.
[
๐‘ฆ] 4
(a) (๐‘ฅ < 2)
๐‘ฅ ๐‘ฅ
)
([
๐‘ฆ] ๐‘ฅ 3
(b)
(๐‘ฅ < 5 → ๐‘ฅ + 4 < 9)
๐‘ฅ ๐‘ฆ ๐‘ฅ
([([
] ) ] )
5 ๐‘ข ๐‘ฃ 6 ๐‘
(c)
(๐‘ฅ = 5 ∧ ๐‘ฅ = ๐‘ฆ)
๐‘ฅ ๐‘ฆ ๐‘ฅ ๐‘ข ๐‘ฃ
(
)
๐‘ฅ ๐‘ฅ
(d) [∃๐‘ฅ(๐‘ฅ − 4 = ๐‘ฆ) ∨ ∀๐‘ฆ(๐‘ฆ + ๐‘ง = ๐‘ฅ + 3)]
๐‘ฆ ๐‘ง
3. Identify all free occurrences of ๐‘ฅ in the given formulas.
(a) ๐‘ฅ + 4 < 10
(b) ∃๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0)
(c) ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = ๐‘ฆ + ๐‘ฅ) ∨ ∀๐‘ง(๐‘ฅ + ๐‘ฆ = ๐‘ง − 3)
(d) ∀๐‘ฅ[∀๐‘ง∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = 2 ⋅ ๐‘ง) ↔ ๐‘ฅ + ๐‘ฅ = 2 ⋅ ๐‘ฅ]
4. Identify all bound occurrences of ๐‘ฅ in the given formulas.
(a) (∀๐‘ฅ) [๐‘(๐‘ฅ) → (∃๐‘ฆ)๐‘ž(๐‘ฆ)]
(b) (∃๐‘ฅ)๐‘(๐‘ฅ, ๐‘ฆ) → (∀๐‘ฆ)๐‘ž(๐‘ฅ)
(c) (∃๐‘ฅ)(๐‘ฅ > 4) ∧ ๐‘ฅ < 10
(d) [(∀๐‘ฅ)(๐‘ฅ + 3 = 1) ∧ ๐‘ฅ = 9] ∨ (∃๐‘ฆ)(๐‘ฅ < 0)
5. Make the simultaneous substitution ๐‘(๐‘Ž) for each of the given formulas.
(a) ๐‘(๐‘ฅ) := 2๐‘ฅ + ๐‘Ž = 9
(b) ๐‘(๐‘ฅ) := ∀๐‘ฅ(๐‘ฅ + ๐‘ฆ = ๐‘ฆ + ๐‘ฅ)
(c) ๐‘(๐‘ฅ) := ∃๐‘ฅ๐‘ž(๐‘ฅ) → ∀๐‘ฆ๐‘Ÿ(๐‘ฅ, ๐‘ฆ)
(d) ๐‘(๐‘ฅ) := ๐‘ฅ + 4 = 0 ∧ ∃๐‘ฅ(๐‘ฆ + ๐‘ง = ๐‘ฅ) ∨ ∃๐‘ง(๐‘ฅ + ๐‘ง = 3)
6. Make the simultaneous substitution ๐‘ž(1, 2, 3) for each of the given formulas.
(a) ๐‘ž(๐‘ฅ, ๐‘ฆ, ๐‘ง) := ๐‘ข + ๐‘ฃ + ๐‘ค
(b) ๐‘ž(๐‘ฅ, ๐‘ฆ, ๐‘ง) := ๐‘Ÿ(๐‘ฅ, ๐‘ฆ) → (∀๐‘ฅ) [๐‘(๐‘ฅ) ∧ (∃๐‘ฆ)๐‘ก(๐‘ฆ, ๐‘ง)] .
(c) ๐‘ž(๐‘ฅ, ๐‘ฆ, ๐‘ง) := ∃๐‘ฅ(๐‘ฅ + ๐‘ฆ = ๐‘ง) ∧ ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ + ๐‘ง) ∨ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ + ๐‘ง)
(d) ๐‘ž(๐‘ฅ, ๐‘ฆ, ๐‘ง) := ∃๐‘ฅ[๐‘(๐‘ฅ) ∧ ∀๐‘ฆ(๐‘(๐‘ฅ) ∧ ๐‘(๐‘ฆ) → ๐‘ฅ = ๐‘ฆ ∨ ๐‘ฆ = ๐‘ง)]
7. Identify the formula to the right of each quantifiers in the given formulas.
(a) (∀๐‘ฅ) [๐‘(๐‘ฅ) → (∃๐‘ฆ)๐‘ž(๐‘ฆ)]
(b) (∃๐‘ฅ)๐‘(๐‘ฅ, ๐‘ฆ) → (∀๐‘ฆ)๐‘ž(๐‘ฆ)
(c) (∀๐‘ฅ)(∃๐‘ฆ)(∃๐‘ง)๐‘(๐‘ฅ, ๐‘ฆ, ๐‘ง) ∧ ๐‘Ÿ(๐‘ค)
(d) ๐‘(๐‘ฅ) ∧ (∃๐‘ฅ)(∀๐‘ฆ)๐‘ž(๐‘ฅ, ๐‘ฆ) ∨ (∀๐‘ง)๐‘Ÿ(๐‘ง)
8. Which of the following are sentences.
(a) ∀๐‘ฅ∀๐‘ฆ[๐‘ž(๐‘ฅ) ∨ ๐‘Ÿ(๐‘ฆ)] ⇒ ∀๐‘ฆ[๐‘ž(๐‘Ž) ∨ ๐‘Ÿ(๐‘ฆ)]
(b) ๐‘ž(2) ∧ ๐‘ก(3) ⇒ ∃๐‘ฆ[๐‘ž(2) ∧ ๐‘ก(๐‘ฆ)]
(c) ∃๐‘ค[∃๐‘ฅ[๐‘(๐‘ฅ) ∨ ∃๐‘ง๐‘ž(๐‘ง) ↔ ∃๐‘ฆ[๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฆ)] → ∀๐‘ฅ๐‘Ÿ(๐‘ฅ)] → ∀๐‘ฆ๐‘Ÿ(๐‘ฅ)]
(d) ∀๐‘ฅ∀๐‘ฆ(๐‘(๐‘ฅ) → [๐‘ž(๐‘ฆ) ∧ ๐‘Ÿ(๐‘ง)]) → ∃๐‘ฅ¬๐‘ž(๐‘ฅ)
Section 2.3 SYNTACTICS
2.3
85
SYNTACTICS
Since formulas without free variables are propositional forms, we can write proofs
involving them using the rules from Sections 1.2 through 1.4. However, since the inference rules did not involve quantification, we need new rules to deal with universal
and existential formulas. We need rules covering not only negations but also rules that
enable the removal (instantiation) and adjoining (generalization) of quantifiers. We
add these rules to Definition 1.2.13 to obtain a stronger notion of proof. Furthermore,
since every sentence is a propositional form, every reference to a propositional form in
Definition 1.2.13 can be considered to be a reference to a sentence. This allows us to
write formal proofs using first-order languages.
Quantifier Negation
Consider the proposition
all rectangles are squares.
This sentence is false because there is a rectangle in which one side is twice the length
of the adjacent side, so
not all rectangles are squares.
That is,
some rectangle is not a square
is true. Generalizing, we conclude that
the negation of all P are Q is some P are not Q.
This should be translated as an inference rule for formulas, so we assume
¬∀๐‘ฅ๐‘ ⇒ ∃๐‘ฅ¬๐‘.
(2.12)
Now consider the proposition
some rectangles are round.
This is false because there are no round rectangles, so
all rectangles are not round
is true. Generalizing, we conclude that
the negation of some P are Q is all P are not Q.
Again, this should be translated as an inference rule for formulas, so we assume
¬∃๐‘ฅ๐‘ ⇒ ∀๐‘ฅ¬๐‘.
(2.13)
86
Chapter 2 FIRST-ORDER LOGIC
∀x p
∀x ¬p
Negations
∃x p
Figure 2.4
∃x ¬p
The modern square of opposition.
Furthermore, by DN and (2.13),
∃๐‘ฅ¬๐‘ ⇒ ¬¬∃๐‘ฅ¬๐‘ ⇒ ¬∀๐‘ฅ¬¬๐‘ ⇒ ¬∀๐‘ฅ๐‘,
and by DN with (2.12),
∀๐‘ฅ¬๐‘ ⇒ ¬¬∀๐‘ฅ¬๐‘ ⇒ ¬∃๐‘ฅ¬¬๐‘ ⇒ ¬∃๐‘ฅ๐‘.
We summarize assumptions (2.12) and (2.13) and the two conclusions in the following
replacement rule.
REPLACEMENT RULES 2.3.1 [Quantifier Negation (QN)]
Let ๐–ฒ be theory symbols and ๐‘ be an ๐–ฒ-formula.
โˆ™ ¬∀๐‘ฅ๐‘ ⇔ ∃๐‘ฅ¬๐‘
โˆ™ ¬∃๐‘ฅ๐‘ ⇔ ∀๐‘ฅ¬๐‘.
QN is illustrated with the modern square of opposition (Figure 2.4). Negations of
quantified formulas are found at opposite corners. A version of the Square is found in
Aristotle’s De Interpretatione, dating around 350 BC (Aristotle 1984).
Whenever we negate formulas of the form ∀๐‘ฅ๐‘ or ∃๐‘ฅ๐‘, to make it easier to read,
the final form should not have a negation immediately to the left of any quantifier, and
using the replacement rules, the negation should be as far into the formula ๐‘ as possible.
We say that such negations are in positive form.
EXAMPLE 2.3.2
Find the negation of ∀๐‘ฅ(๐‘ ∧ ๐‘ž) and put the final answer in positive form.
¬∀๐‘ฅ(๐‘ ∧ ๐‘ž) ⇔ ∃๐‘ฅ¬(๐‘ ∧ ๐‘ž) ⇔ ∃๐‘ฅ(¬๐‘ ∨ ¬๐‘ž).
The next example will use De Morgan’s law as the last one did. It also needs material
implication and double negation.
Section 2.3 SYNTACTICS
87
EXAMPLE 2.3.3
Find the negation of ∀๐‘ฅ∃๐‘ฆ[๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)] and put it into positive form.
¬∀๐‘ฅ∃๐‘ฆ[๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)] ⇔ ∃๐‘ฅ¬∃๐‘ฆ[๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)]
⇔ ∃๐‘ฅ∀๐‘ฆ¬[๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)]
⇔ ∃๐‘ฅ∀๐‘ฆ¬[¬๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฆ)]
⇔ ∃๐‘ฅ∀๐‘ฆ[๐‘(๐‘ฅ) ∧ ¬๐‘ž(๐‘ฆ)].
Proofs with Universal Formulas
Consider the sentence all multiples of 4 are even. This implies, for instance, that 8,
100, and −16 are even. To generalize this reasoning to formulas means that whenever
we have ∀๐‘ฅ๐‘(๐‘ฅ), we also have ๐‘(๐‘Ž), where ๐‘Ž might be either a constant symbol such as
8, 100, and −16 or a randomly chosen constant symbol. This gives the first inference
rule that involves a quantifier. We do not prove it, but take it to be an axiom. Observe
the use of Definition 2.1.6
INFERENCE RULE 2.3.4 [Universal Instantiation (UI)]
If ๐‘(๐‘ฅ) is an ๐–ฒ-formula, then for every constant symbol ๐‘Ž from ๐–ฒ,
∀๐‘ฅ๐‘(๐‘ฅ) ⇒ ๐‘(๐‘Ž),
We make two observations about UI. First, since the resulting formula is to be part
of a proof, the substitution must yield a sentence, so ๐‘Ž must be a constant symbol.
Second, we use the notation ๐‘(๐‘ฅ) to represent the formula instead of ๐‘ because ∀๐‘ฅ๐‘(๐‘ฅ)
will be part of a proof, which means, again, that it must be a sentence. Writing ๐‘(๐‘ฅ)
limits the formula to have only ๐‘ฅ as a possible free variable, so ∀๐‘ฅ๐‘(๐‘ฅ) is a sentence.
If the formula ๐‘ had free variables other than ๐‘ฅ, then ∀๐‘ฅ๐‘ would not be a sentence and
not suitable for a proof.
EXAMPLE 2.3.5
Let ๐‘, ๐‘ž, and ๐‘Ÿ be ๐–ฒ-formulas and ๐‘Ž and ๐‘ be constant symbols from ๐–ฒ. The
following are legitimate uses of UI. Notice that each of the inferences results in
an ๐–ฒ-formula.
โˆ™ ∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)] ⇒ ๐‘(๐‘Ž) → ๐‘ž(๐‘Ž)
โˆ™ ∀๐‘ฅ [๐‘(๐‘ฅ) ∨ ∀๐‘ฆ๐‘ž(๐‘ฆ)] ⇒ ๐‘(๐‘Ž) ∨ ∀๐‘ฆ๐‘ž(๐‘ฆ)
โˆ™ ∀๐‘ฅ∀๐‘ฆ [๐‘ž(๐‘ฅ) ∨ ๐‘Ÿ(๐‘ฆ)] ⇒ ∀๐‘ฆ [๐‘ž(๐‘Ž) ∨ ๐‘Ÿ(๐‘ฆ)]
โˆ™ ∀๐‘ฆ [๐‘ž(๐‘Ž) ∨ ๐‘Ÿ(๐‘ฆ)] ⇒ ๐‘ž(๐‘Ž) ∧ ๐‘Ÿ(๐‘Ž)
โˆ™ ∀๐‘ฆ [๐‘ž(๐‘Ž) ∨ ๐‘Ÿ(๐‘ฆ)] ⇒ ๐‘ž(๐‘Ž) ∧ ๐‘Ÿ(๐‘).
88
Chapter 2 FIRST-ORDER LOGIC
Before we can write formal proofs, we need a rule that will attach a universal quantifier. It will be different from universal instantiation, for it requires a criterion on the
constant.
DEFINITION 2.3.6
Let ๐‘Ž be a constant symbol introduced into a formal proof by a substitution. If
the first occurrence of ๐‘Ž is in a sentence that follows by UI, then ๐‘Ž is arbitrary.
Otherwise, ๐‘Ž is particular and can be denoted by ๐‘Žฬ‚ to serve as a reminder that
the symbol is not arbitrary.
The idea behind Definition 2.3.6 is that if a constant symbol ๐‘Ž is arbitrary, it represents a randomly selected object, but if ๐‘Ž is particular, then ๐‘Ž represents an object with
at least one known property. This property can be identified by a formula ๐‘(๐‘ฅ) so that
we have ๐‘(๐‘Ž).
EXAMPLE 2.3.7
The constant symbol ๐‘Ž in the following is not arbitrary because its first occurrence
of ๐‘Ž in line 1 is not the result of UI.
1.
2.
๐‘(๐‘Ž)
๐‘(๐‘Ž) ∨ ๐‘ž(๐‘Ž)
Given
Add
Therefore, these two lines should be written using ๐‘Žฬ‚ instead of ๐‘Ž.
1.
2.
๐‘(๐‘Ž)
ฬ‚
๐‘(๐‘Ž)
ฬ‚ ∨ ๐‘ž(๐‘Ž)
ฬ‚
Given
Add
However, ๐‘Ž in the following is arbitrary because its first occurrence is in line 2,
and line 2 follows from line 1 by UI.
1.
2.
3.
∀๐‘ฅ๐‘(๐‘ฅ)
๐‘(๐‘Ž)
๐‘(๐‘Ž) ∨ ๐‘ž(๐‘Ž)
Given
1 UI
Add
We can now introduce the rule of inference that allows us to attach universal quantifiers to formulas. We state it as an axiom.
INFERENCE RULE 2.3.8 [Universal Generalization (UG)]
If ๐‘(๐‘ฅ) is an ๐–ฒ-formula with no particular constant symbols and ๐‘Ž is an arbitrary
constant symbol from ๐–ฒ,
๐‘(๐‘Ž) ⇒ ∀๐‘ฅ๐‘(๐‘ฅ).
Consider the following argument:
All squares are rectangles.
All rectangles are quadrilaterals.
Therefore, all squares are quadrilaterals.
Section 2.3 SYNTACTICS
89
Representing the premises by ∀๐‘ฅ[๐‘ (๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)] and ∀๐‘ฅ[๐‘Ÿ(๐‘ฅ) → ๐‘ž(๐‘ฅ)], a formal proof of
this includes UG:
1.
2.
3.
4.
5.
6.
∀๐‘ฅ[๐‘ (๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)]
∀๐‘ฅ[๐‘Ÿ(๐‘ฅ) → ๐‘ž(๐‘ฅ)]
๐‘ (๐‘Ž) → ๐‘Ÿ(๐‘Ž)
๐‘Ÿ(๐‘Ž) → ๐‘ž(๐‘Ž)
๐‘ (๐‘Ž) → ๐‘ž(๐‘Ž)
∀๐‘ฅ[๐‘ (๐‘ฅ) → ๐‘ž(๐‘ฅ)]
Given
Given
1 UI
2 UI
3, 4 HS
5 UG
Since ๐‘Ž was introduced in line 3 by UI, it is an arbitrary constant symbol. In addition,
๐‘ (๐‘Ž) → ๐‘ž(๐‘Ž) contains no particular constant symbols that appeared in the proof by
substitution. Hence, the application of UG in line 6 is legal.
EXAMPLE 2.3.9
These are illegal uses of universal generalization.
โˆ™ Let ๐‘(๐‘ฅ) be an ๐–ฒ-formula with ๐‘ a constant symbol.
1.
2.
๐‘(๐‘)
∀๐‘ฅ๐‘(๐‘ฅ)
Given
1 UG [error]
The constant symbol ๐‘ in line 1 is particular, even without being written as
๐‘.
ฬ‚ It was not introduced to the proof by UI. Therefore, universal generalization does not apply.
โˆ™ Suppose that ๐‘Ž is an arbitrary constant symbol.
1.
2.
๐‘Ž + ๐‘ฬ‚ = 0
∀๐‘ฅ(๐‘ฅ + ๐‘ฬ‚ = 0)
1 UG [error]
The restriction against ๐‘(๐‘ฅ) containing particular symbols prevents the errant conclusion in line 2.
โˆ™ The following is an attempt to prove ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = 2 ⋅ ๐‘ฅ) from the formula
∀๐‘ฅ(๐‘ฅ + ๐‘ฅ = 2 ⋅ ๐‘ฅ).
1.
2.
3.
4.
∀๐‘ฅ(๐‘ฅ + ๐‘ฅ = 2 ⋅ ๐‘ฅ)
๐‘Ž+๐‘Ž=2⋅๐‘Ž
∀๐‘ฆ(๐‘Ž + ๐‘ฆ = 2 ⋅ ๐‘Ž)
∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = 2 ⋅ ๐‘ฅ)
Given
1 UI
2 UG [error]
3 UG
Although the constant symbol ๐‘Ž in line 2 is arbitrary, the proof is not valid.
The reason is that an illegal substitution was made in line 3. To see this, let
๐‘(๐‘ฅ) := ๐‘ฅ + ๐‘ฅ = 2 ⋅ ๐‘ฅ.
90
Chapter 2 FIRST-ORDER LOGIC
Applying universal generalization gives
∀๐‘ฆ(๐‘ฆ + ๐‘ฆ = 2 ⋅ ๐‘ฆ)
because ๐‘(๐‘ฆ) ⇔ ๐‘ฆ + ๐‘ฆ = 2 ⋅ ๐‘ฆ, but this is not what was written in line 3.
Now for some formal proofs.
EXAMPLE 2.3.10
Prove: ∀๐‘ฅ∀๐‘ฆ๐‘(๐‘ฅ, ๐‘ฆ) โŠข ∀๐‘ฆ∀๐‘ฅ๐‘(๐‘ฅ, ๐‘ฆ)
1.
2.
3.
4.
5.
∀๐‘ฅ∀๐‘ฆ๐‘(๐‘ฅ, ๐‘ฆ)
∀๐‘ฆ๐‘(๐‘Ž, ๐‘ฆ)
๐‘(๐‘Ž, ๐‘)
∀๐‘ฅ๐‘(๐‘ฅ, ๐‘)
∀๐‘ฆ∀๐‘ฅ๐‘(๐‘ฅ, ๐‘ฆ)
Given
1 UI
2 UI
3 UG
4 UG
Since both ๐‘Ž and ๐‘ first appear because of UI, they are arbitrary and universal
generalization can be applied to both constant symbols.
EXAMPLE 2.3.11
Prove: ∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)] , ∀๐‘ฅ¬ [๐‘ž(๐‘ฅ) ∨ ๐‘Ÿ(๐‘ฅ)] โŠข ∀๐‘ฅ¬๐‘(๐‘ฅ)
1.
2.
3.
4.
5.
6.
7.
8.
∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)]
∀๐‘ฅ¬ [๐‘ž(๐‘ฅ) ∨ ๐‘Ÿ(๐‘ฅ)]
๐‘(๐‘Ž) → ๐‘ž(๐‘Ž)
¬ [๐‘ž(๐‘Ž) ∨ ๐‘Ÿ(๐‘Ž)]
¬๐‘ž(๐‘Ž) ∧ ¬๐‘Ÿ(๐‘Ž)
¬๐‘ž(๐‘Ž)
¬๐‘(๐‘Ž)
∀๐‘ฅ¬๐‘(๐‘ฅ)
Given
Given
1 UI
2 UI
4 DeM
5 Simp
3, 6 MT
7 UG
Notice that the ๐‘Ž in line 3 was introduced because of UI. Hence, ๐‘Ž is arbitrary
throughout the proof.
Proofs with Existential Formulas
Since 4 + 5 = 9, we conclude that there exists ๐‘ฅ such that ๐‘ฅ + 5 = 9. Since we
can construct an isosceles triangle, we conclude that isosceles triangles exist. This
motivates the next inference rule.
91
Section 2.3 SYNTACTICS
THEOREM 2.3.12 [Existential Generalization (EG)]
If ๐‘(๐‘ฅ) is an ๐–ฒ-formula and ๐‘Ž is a constant symbol from ๐–ฒ,
๐‘(๐‘Ž) ⇒ ∃๐‘ฅ๐‘(๐‘ฅ).
PROOF
Assume ๐‘(๐‘Ž). Suppose that ∀๐‘ฅ¬๐‘(๐‘ฅ). By UI we have that ¬๐‘(๐‘Ž). Therefore,
¬∀๐‘ฅ¬๐‘(๐‘ฅ) by IP, from which ∃๐‘ฅ๐‘(๐‘ฅ) follows by QN.
EXAMPLE 2.3.13
Each of the following is a valid use of existential generalization.
โˆ™ ๐‘(๐‘Ž) ∧ ¬๐‘Ÿ(๐‘Ž) ⇒ ∃๐‘ฅ [๐‘(๐‘ฅ) ∧ ¬๐‘Ÿ(๐‘ฅ)]
โˆ™ ๐‘ž(๐‘Ž) ∧ ๐‘ก(๐‘) ⇒ ∃๐‘ฆ [๐‘ž(๐‘Ž) ∧ ๐‘ก(๐‘ฆ)].
Before we write some proofs, here is the inference rule that allows us to detach
existential quantifiers.
THEOREM 2.3.14 [Existential Instantiation (EI)]
If ๐‘(๐‘ฅ) is an ๐–ฒ-formula,
∃๐‘ฅ๐‘(๐‘ฅ) ⇒ ๐‘(๐‘Ž),
ฬ‚
where ๐‘Ž is a constant symbol from ๐–ฒ that has no occurrence in the formal proof
prior to ๐‘(๐‘Ž).
ฬ‚
PROOF
Assume that ∃๐‘ฅ๐‘(๐‘ฅ) does not infer ๐‘(๐‘) for any constant symbol ๐‘. Suppose
∃๐‘ฅ๐‘(๐‘ฅ). Combined with the assumption, this implies that ¬๐‘(๐‘) for every constant
symbol ๐‘ by the law of the excluded middle (page 37) and DS. That is, ¬๐‘(๐‘Ž) for
an arbitrary constant symbol ๐‘Ž. Therefore, ∀๐‘ฅ¬๐‘(๐‘ฅ) by UG, so ¬∃๐‘ฅ๐‘(๐‘ฅ) by QN,
which is a contradiction. Therefore, ∃๐‘ฅ๐‘(๐‘ฅ) ⇒ ๐‘(๐‘Ž) for some constant symbol ๐‘Ž,
and ๐‘Ž is particular by Definition 2.3.6.
The constant symbol ๐‘Žฬ‚ obtained by EI is called a witness of ∃๐‘ฅ๐‘(๐‘ฅ).
The reason that the constant symbol ๐‘Ž must have no prior occurrence in a proof
when applying EI is because a used symbol already represents some object, so if ๐‘Ž had
appeared in an earlier line, we would have no reason to assume that we could write ๐‘(๐‘Ž)
from ∃๐‘ฅ๐‘(๐‘ฅ). For example, given
∃๐‘ฅ(๐‘ฅ + 4 = 5)
(2.14)
∃๐‘ฅ(๐‘ฅ + 2 = 13),
(2.15)
and
we write ๐‘Ž + 4 = 5 for some constant symbol ๐‘Ž by EI and (2.14). By EI and (2.15),
we conclude that ๐‘ + 2 = 13 for some constant symbol ๐‘, but inferring ๐‘Ž + 2 = 13 is
invalid because ๐‘Ž ≠ ๐‘.
92
Chapter 2 FIRST-ORDER LOGIC
EXAMPLE 2.3.15
The following are legal uses of existential instantiation assuming that ๐‘Ž and ๐‘
have no prior occurrences.
โˆ™ ∃๐‘ฅ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฅ)] ⇒ ๐‘(๐‘Ž)
ฬ‚ ∧ ๐‘ž(๐‘Ž)
ฬ‚
ฬ‚ ๐‘) → ๐‘Ÿ(๐‘Ž, ๐‘,
ฬ‚ ๐‘)
โˆ™ ∃๐‘ฆ [๐‘Ÿ(๐‘Ž, ๐‘ฆ, ๐‘) → ๐‘Ÿ(๐‘Ž, ๐‘ฆ, ๐‘)] ⇒ ๐‘Ÿ(๐‘Ž, ๐‘,
โˆ™ ∃๐‘ฅ∀๐‘ฆ∃๐‘ง๐‘ž(๐‘ฅ, ๐‘ฆ, ๐‘ง) ⇒ ∀๐‘ฆ∃๐‘ง๐‘ž(๐‘Ž,
ฬ‚ ๐‘ฆ, ๐‘ง).
EXAMPLE 2.3.16
Existential Instantiation cannot be used to justify either of the following.
ฬ‚ ∨ ๐‘ž(๐‘ง) by EI because the substitution
โˆ™ ∃๐‘ง [๐‘(๐‘ง) ∨ ๐‘ž(๐‘ง)] does not imply ๐‘(๐‘)
ฬ‚ ∨ ๐‘ž(๐‘).
ฬ‚
was not made correctly. The result should have been ๐‘(๐‘)
โˆ™ ∃๐‘ฅ∃๐‘ฆ๐‘(๐‘Ž, ๐‘ฅ, ๐‘ฆ) does not imply ∃๐‘ฆ๐‘(๐‘Ž, ๐‘Ž, ๐‘ฆ) by EI because ๐‘Ž has a prior occurrence. Also, notice that the hat notation was not used.
EXAMPLE 2.3.17
Assume that ๐‘ฅ is a real number and consider the following.
1.
2.
3.
4.
5.
∀๐‘ฅ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0)
∃๐‘ฆ(๐‘Ž + ๐‘ฆ = 0)
๐‘Ž + ๐‘ฬ‚ = 0
∀๐‘ฅ(๐‘ฅ + ๐‘ฬ‚ = 0)
∃๐‘ฆ∀๐‘ฅ(๐‘ฅ + ๐‘ฆ = 0)
Given
1 UI
2 EG
3 UG [error]
4 EG
The conclusion in line 5 is incorrect. The problem lies in line 4 where UG was
applied despite the particular constant symbol in line 3 (compare Example 2.3.9).
This example makes clear why there is a restriction on particular elements in UG
(Inference Rule 2.3.8). Since ๐‘ represents a particular real number, line 3 cannot
be used to conclude that all real numbers plus that particular ๐‘ equals 0. We
know that ๐‘ is the witness to line 2, but that is all we know about it. To correct
the argument, we can essentially reverse the steps to arrive back at the premise.
1.
2.
3.
4.
5.
∀๐‘ฅ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0)
∃๐‘ฆ(๐‘Ž + ๐‘ฆ = 0)
๐‘Ž + ๐‘ฬ‚ = 0
∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0)
∀๐‘ฅ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0)
Given
1 UI
2 EG
3 EG
4 UG
Notice that there is no particular constant symbol in line 4, so line 5 does legally
follow by UG.
Here are some formal proofs that use existential instantiation and generalization.
EXAMPLE 2.3.18
Prove: ∃๐‘ฅ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฅ)] , ∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)] โŠข ∃๐‘ฅ๐‘Ÿ(๐‘ฅ)
Section 2.3 SYNTACTICS
1.
2.
3.
4.
5.
6.
7.
∃๐‘ฅ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฅ)]
∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)]
๐‘(๐‘)
ฬ‚ ∧ ๐‘ž(๐‘)
ฬ‚
๐‘(๐‘)
ฬ‚ → ๐‘Ÿ(๐‘)
ฬ‚
๐‘(๐‘)
ฬ‚
๐‘Ÿ(๐‘)
ฬ‚
∃๐‘ฅ๐‘Ÿ(๐‘ฅ)
93
Given
Given
1 EI
2 UI
3 Simp
4, 5 MP
6 EG
EXAMPLE 2.3.19
Prove: ∀๐‘ฅ∃๐‘ฆ [๐‘ž(๐‘ฅ) ∧ ๐‘ก(๐‘ฆ)] โŠข ∀๐‘ฅ [๐‘ž(๐‘ฅ) ∧ (∃๐‘ฆ)๐‘ก(๐‘ฆ)]
1.
2.
3.
4.
5.
6.
7.
8.
9.
∀๐‘ฅ∃๐‘ฆ [๐‘ž(๐‘ฅ) ∧ ๐‘ก(๐‘ฆ)]
∃๐‘ฆ [๐‘ž(๐‘Ž) ∧ ๐‘ก(๐‘ฆ)]
๐‘ž(๐‘Ž) ∧ ๐‘ก(๐‘)
ฬ‚
๐‘ก(๐‘)
ฬ‚ ∧ ๐‘ž(๐‘Ž)
๐‘ก(๐‘)
ฬ‚
∃๐‘ฆ๐‘ก(๐‘ฆ)
๐‘ž(๐‘Ž)
๐‘ž(๐‘Ž) ∧ (∃๐‘ฆ)๐‘ก(๐‘ฆ)
∀๐‘ฅ [๐‘ž(๐‘ฅ) ∧ (∃๐‘ฆ)๐‘ก(๐‘ฆ)]
Given
1 UI
2 EI
3 Com
4 Simp
5 EG
3 Simp
6, 7 Conj
8 UG
EXAMPLE 2.3.20
Prove : ๐‘(๐‘Ž) → ∃๐‘ฅ [๐‘ž(๐‘ฅ) ∧ ๐‘Ÿ(๐‘ฅ)] , ๐‘(๐‘Ž) โŠข ∃๐‘ฅ๐‘Ÿ(๐‘ฅ)
1.
2.
3.
4.
5.
6.
7.
๐‘(๐‘Ž)
ฬ‚ → ∃๐‘ฅ [๐‘ž(๐‘ฅ) ∧ ๐‘Ÿ(๐‘ฅ)]
๐‘(๐‘Ž)
ฬ‚
∃๐‘ฅ [๐‘ž(๐‘ฅ) ∧ ๐‘Ÿ(๐‘ฅ)]
ฬ‚ ∧ ๐‘Ÿ(๐‘)
ฬ‚
๐‘ž(๐‘)
ฬ‚
ฬ‚
๐‘Ÿ(๐‘) ∧ ๐‘ž(๐‘)
ฬ‚
๐‘Ÿ(๐‘)
∃๐‘ฅ๐‘Ÿ(๐‘ฅ)
Given
Given
1, 2 MP
3 EI
4 Com
5 Simp
6 EG
Exercises
1. Use QN and other replacement rules to determine whether the following are legal
replacements.
(a) ¬∀๐‘ฅ๐‘(๐‘ฅ) ⇔ ∀๐‘ฅ¬๐‘(๐‘ฅ)
(b) ¬∃๐‘ฅ๐‘(๐‘ฅ) ⇔ ¬∀๐‘ฅ๐‘(๐‘ฅ)
(c) ∀๐‘ฅ¬๐‘(๐‘ฅ) ⇔ ∃๐‘ฅ๐‘(๐‘ฅ)
(d) ∃๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)] ⇔ ¬∀๐‘ฅ [¬๐‘(๐‘ฅ) → ¬๐‘ž(๐‘ฅ)]
(e) ¬∀๐‘ฅ∃๐‘ฆ๐‘(๐‘ฅ, ๐‘ฆ) ⇔ ∃๐‘ฅ∀๐‘ฆ¬๐‘(๐‘ฅ, ๐‘ฆ)
(f) ¬∀๐‘ฅ∃๐‘ฆ๐‘(๐‘ฅ, ๐‘ฆ) ⇔ ∃๐‘ฆ∀๐‘ฅ๐‘(๐‘ฅ, ๐‘ฆ)
94
Chapter 2 FIRST-ORDER LOGIC
2. Negate and put into positive form.
(a) ∃๐‘ฅ [๐‘ž(๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)]
(b) ∀๐‘ฅ∃๐‘ฆ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฆ)]
(c) ∃๐‘ฅ∃๐‘ฆ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ, ๐‘ฆ)]
(d) ∀๐‘ฅ∀๐‘ฆ [๐‘(๐‘ฅ) ∨ (∃๐‘ง)๐‘ž(๐‘ฆ, ๐‘ง)]
(e) ∃๐‘ฅ¬๐‘Ÿ(๐‘ฅ) ∨ ∀๐‘ฅ [๐‘ž(๐‘ฅ) ↔ ¬๐‘(๐‘ฅ)]
(f) ∀๐‘ฅ∀๐‘ฆ∃๐‘ง(๐‘(๐‘ฅ) → [๐‘ž(๐‘ฆ) ∧ ๐‘Ÿ(๐‘ง)]) → ∃๐‘ฅ¬๐‘ž(๐‘ฅ)
3. Determine whether each pair of propositions are negations. If they are not, write
the negation of both.
(a) Every real number has a square root.
Every real number does not have a square root.
(b) Every multiple of four is a multiple of two.
Some multiples of two are multiples of four.
(c) For all ๐‘ฅ, if ๐‘ฅ is odd, then ๐‘ฅ2 is odd.
There exists ๐‘ฅ such that if ๐‘ฅ is odd, then ๐‘ฅ2 is even.
(d) There exists an integer ๐‘ฅ such that ๐‘ฅ + 1 = 10.
For all integers ๐‘ฅ, ๐‘ฅ + 1 ≠ 10.
4. Write the negation of the following propositions in positive form and in English.
(a) For all ๐‘ฅ, there exists ๐‘ฆ such that ๐‘ฆโˆ•๐‘ฅ = 9.
(b) There exists ๐‘ฅ so that ๐‘ฅ๐‘ฆ = 1 for all real numbers ๐‘ฆ.
(c) Every multiple of ten is a multiple of five.
(d) No interval contains a rational number.
(e) There is an interval that contains a rational number.
5. Let ๐‘“ be a function and ๐‘ be a real number in the open interval ๐ผ. Then, ๐‘“ is
continuous at ๐‘ if for every ๐œ– > 0, there exists ๐›ฟ > 0 such that for all ๐‘ฅ in ๐ผ, if
0 < |๐‘ฅ − ๐‘| < ๐›ฟ, then |๐‘“ (๐‘ฅ) − ๐‘“ (๐‘)| < ๐œ–.
(a) Write what it means for ๐‘“ to be not continuous at ๐‘.
(b) The function ๐‘“ is continuous on an open interval if it is continuous at every
point of the interval. Write what it means for ๐‘“ to be not continuous on an
open interval.
6. Let ๐‘“ be a function and ๐‘ a real number in the open interval ๐ผ. Then, ๐‘“ is uniformly
continuous on ๐ผ means that for every ๐œ– > 0, there exists ๐›ฟ > 0 such that for all ๐‘ and
๐‘ฅ in ๐ผ, if 0 < |๐‘ฅ − ๐‘| < ๐›ฟ, then |๐‘“ (๐‘ฅ) − ๐‘“ (๐‘)| < ๐œ–.
(a) Write what it means for ๐‘“ to be not uniformly continuous on ๐ผ.
(b) How does ๐‘“ being continuous on ๐ผ differ from ๐‘“ being uniformly continuous
on ๐ผ?
7. Prove using QN.
(a) ∃๐‘ฅ๐‘(๐‘ฅ) โŠข ∀๐‘ฅ¬๐‘(๐‘ฅ) → ∀๐‘ฅ๐‘ž(๐‘ฅ)
(b) ∀๐‘ฅ[๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)], ∀๐‘ฅ[๐‘ž(๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)], ¬∀๐‘ฅ๐‘Ÿ(๐‘ฅ) โŠข ∃๐‘ฅ¬๐‘(๐‘ฅ)
(c) ∀๐‘ฅ๐‘(๐‘ฅ) → ∀๐‘ฆ[๐‘ž(๐‘ฆ) → ๐‘Ÿ(๐‘ฆ)], ∃๐‘ฅ[๐‘ž(๐‘ฅ) ∧ ¬๐‘Ÿ(๐‘ฅ)] โŠข ∃๐‘ฅ¬๐‘(๐‘ฅ)
(d) ∃๐‘ฅ๐‘(๐‘ฅ) → ∃๐‘ฆ๐‘ž(๐‘ฆ), ∀๐‘ฅ¬๐‘ž(๐‘ฅ) โŠข ∀๐‘ฅ¬๐‘(๐‘ฅ)
Section 2.3 SYNTACTICS
8. Find all errors in the given proofs.
(a) “∃๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)] , ∃๐‘ฅ¬๐‘ž(๐‘ฅ) โŠข ∃๐‘ฅ๐‘(๐‘ฅ)”
Attempted Proof
1. ∃๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)]
2. ∃๐‘ฅ¬๐‘ž(๐‘ฅ)
3. ๐‘(๐‘) ∨ ๐‘ž(๐‘)
4. ¬๐‘ž(๐‘)
5. ¬๐‘(๐‘)
6. ∃๐‘ฅ¬๐‘(๐‘ฅ)
Given
Given
1 EI
2 EI
3, 4 DS
5 EG
(b) “∀๐‘ฅ๐‘(๐‘ฅ) โŠข ∃๐‘ฅ∀๐‘ฆ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฆ)]”
Attempted Proof
1. ∀๐‘ฅ๐‘(๐‘ฅ)
2. ๐‘(๐‘)
ฬ‚
3. ๐‘(๐‘)
ฬ‚ ∨ ๐‘ž(๐‘Ž)
4. ∀๐‘ฆ [๐‘(๐‘)
ฬ‚ ∨ ๐‘ž(๐‘ฆ)]
5. ∃๐‘ฆ∀๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฆ)]
Given
1 UI
2 Add
3 UG
4 EG
(c) “∃๐‘ฅ๐‘(๐‘ฅ), ∃๐‘ฅ๐‘ž(๐‘ฅ) โŠข ∀๐‘ฅ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฅ)]”
Attempted Proof
1. ∃๐‘ฅ๐‘(๐‘ฅ)
2. ∃๐‘ฅ๐‘ž(๐‘ฅ)
3. ๐‘(๐‘)
ฬ‚
4. ๐‘ž(๐‘)
ฬ‚
5. ๐‘(๐‘)
ฬ‚ ∧ ๐‘ž(๐‘)
6. ∀๐‘ฅ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฅ)]
Given
Given
1 EI
2 EI
3, 4 Conj
5 UG
9. Prove.
(a) ∀๐‘ฅ๐‘(๐‘ฅ) โŠข ∀๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)]
(b) ∀๐‘ฅ๐‘(๐‘ฅ), ∀๐‘ฅ [๐‘ž(๐‘ฅ) → ¬๐‘(๐‘ฅ)] โŠข ∀๐‘ฅ¬๐‘ž(๐‘ฅ)
(c) ∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)] , ∀๐‘ฅ๐‘(๐‘ฅ) โŠข ∀๐‘ฅ๐‘ž(๐‘ฅ)
(d) ∀๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)] , ∀๐‘ฅ¬๐‘ž(๐‘ฅ) โŠข ∀๐‘ฅ๐‘(๐‘ฅ)
(e) ∃๐‘ฅ๐‘(๐‘ฅ) โŠข ∃๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)]
(f) ∃๐‘ฅ∃๐‘ฆ๐‘(๐‘ฅ, ๐‘ฆ) โŠข ∃๐‘ฆ∃๐‘ฅ๐‘(๐‘ฅ, ๐‘ฆ)
(g) ∀๐‘ฅ∀๐‘ฆ๐‘(๐‘ฅ, ๐‘ฆ) โŠข ∀๐‘ฆ∀๐‘ฅ๐‘(๐‘ฅ, ๐‘ฆ)
(h) ∀๐‘ฅ¬๐‘(๐‘ฅ), ∃๐‘ฅ [๐‘ž(๐‘ฅ) → ๐‘(๐‘ฅ)] โŠข ∃๐‘ฅ¬๐‘ž(๐‘ฅ)
(i) ∃๐‘ฅ๐‘(๐‘ฅ), ∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)] โŠข ∃๐‘ฅ๐‘ž(๐‘ฅ)
(j) ∀๐‘ฅ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฅ)] , ∀๐‘ฅ [๐‘Ÿ(๐‘ฅ) → ๐‘ (๐‘ฅ)] , ∃๐‘ฅ [๐‘(๐‘ฅ) ∨ ๐‘Ÿ(๐‘ฅ)] โŠข ∃๐‘ฅ [๐‘ž(๐‘ฅ) ∨ ๐‘ (๐‘ฅ)]
(k) ∃๐‘ฅ [๐‘(๐‘ฅ) ∧ ¬๐‘Ÿ(๐‘ฅ)] , ∀๐‘ฅ [๐‘ž(๐‘ฅ) → ๐‘Ÿ(๐‘ฅ)] โŠข ∃๐‘ฅ¬๐‘ž(๐‘ฅ)
(l) ∀๐‘ฅ๐‘(๐‘ฅ), ∀๐‘ฅ([๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)] → [๐‘Ÿ(๐‘ฅ) ∧ ๐‘ (๐‘ฅ)]) โŠข ∃๐‘ฅ๐‘ (๐‘ฅ)
(m) ∃๐‘ฅ๐‘(๐‘ฅ), ∃๐‘ฅ๐‘(๐‘ฅ) → ∀๐‘ฅ∃๐‘ฆ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)] โŠข ∀๐‘ฅ๐‘ž(๐‘ฅ) ∨ ∃๐‘ฅ๐‘Ÿ(๐‘ฅ)
95
96
Chapter 2 FIRST-ORDER LOGIC
(n) ∃๐‘ฅ๐‘(๐‘ฅ), ∃๐‘ฅ๐‘ž(๐‘ฅ), ∃๐‘ฅ∃๐‘ฆ [๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฆ)] → ∀๐‘ฅ๐‘Ÿ(๐‘ฅ) โŠข ∀๐‘ฅ๐‘Ÿ(๐‘ฅ)
10. Prove the following:
(a) ๐‘(๐‘ฅ) ∧ ∃๐‘ฆ๐‘ž(๐‘ฆ) ⇔ ∃๐‘ฆ[๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฆ)]
(b) ๐‘(๐‘ฅ) ∨ ∃๐‘ฆ๐‘ž(๐‘ฆ) ⇔ ∃๐‘ฆ[๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฆ)]
(c) ๐‘(๐‘ฅ) ∧ ∀๐‘ฆ๐‘ž(๐‘ฆ) ⇔ ∀๐‘ฆ[๐‘(๐‘ฅ) ∧ ๐‘ž(๐‘ฆ)]
(d) ๐‘(๐‘ฅ) ∨ ∀๐‘ฆ๐‘ž(๐‘ฆ) ⇔ ∀๐‘ฆ[๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฆ)]
(e) ∃๐‘ฅ๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ) ⇔ ∀๐‘ฅ[๐‘(๐‘ฅ) → ๐‘(๐‘ฆ)]
(f) ∀๐‘ฅ๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ) ⇔ ∃๐‘ฅ[๐‘(๐‘ฅ) → ๐‘(๐‘ฆ)]
(g) ๐‘(๐‘ฅ) → ∃๐‘ฆ๐‘ž(๐‘ž) ⇔ ∃๐‘ฆ[๐‘(๐‘ฅ) → ๐‘(๐‘ฆ)]
(h) ๐‘(๐‘ฅ) → ∀๐‘ฆ๐‘ž(๐‘ž) ⇔ ∀๐‘ฆ[๐‘(๐‘ฅ) → ๐‘(๐‘ฆ)]
11. An ๐–ฒ-formula is in prenex normal form if it is of the form
๐‘„0 ๐‘ฅ0 ๐‘„1 ๐‘ฅ1 … ๐‘„๐‘›−1 ๐‘ฅ๐‘›−1 ๐‘ž,
where ๐‘„0 , ๐‘„1 , … , ๐‘„๐‘›−1 are quantifier symbols and ๐‘ž is a ๐–ฒ-formula. Every ๐–ฒ-formula
can be replaced with an ๐–ฒ-formula in prenex normal form, although variables might
need to be renamed. Use Exercise 10 and QN to put the given formulas in prenex
normal form.
(a) [๐‘(๐‘ฅ) ∨ ๐‘ž(๐‘ฅ)] → ∃๐‘ฆ[๐‘ž(๐‘ฆ) → ๐‘Ÿ(๐‘ฆ)]
(b) ¬∀๐‘ฅ[๐‘(๐‘ฅ) → ¬∃๐‘ฆ๐‘ž(๐‘ฆ)]
(c) ∃๐‘ฅ๐‘(๐‘ฅ) → ∀๐‘ฅ∃๐‘ฆ [๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)]
(d) ∀๐‘ฅ[๐‘(๐‘ฅ) → ๐‘ž(๐‘ฆ)] ∧ ¬∃๐‘ฆ∀๐‘ง[๐‘Ÿ(๐‘ฆ) → ๐‘ (๐‘ง)]
2.4
PROOF METHODS
The purpose of the propositional logic of Chapter 1 is to model the basic reasoning that
one does in mathematics. Rules that determine truth (semantics) and establish valid
forms of proof (syntactics) were developed. The logic developed in this chapter is an
extension of propositional logic. It is called first-order logic. As with propositional
logic, first-order logic provides a working model of a type of deductive reasoning that
happens when studying mathematics, but with a greater emphasis on what a particular
proposition is communicating about its subject. Geometry can serve as an example.
To solve geometric problems and prove geometric propositions means to work with the
axioms and theorems of geometry. What steps are legal when doing this are dictated by
the choice of the logic. Because the subjects of geometric propositions are mathematical objects, such as points, lines, and planes, first-order logic is a good choice. However,
sometimes other logical systems are used. An example of such an alternative is secondorder logic, which allows quantification over function and relation symbols (page 73)
in addition to quantification over variable symbols. Whichever logic is chosen, that
logic provides the general rules of reasoning for the mathematical theory. Since it has
its own axioms and theorems, a logic itself is a theory, but because it is intended to
provide rules for other theories, it is sometimes called a metatheory (Figure 2.5).
Although all mathematicians use logic, they usually do not use symbolic logic. Instead, their proofs are written using sentences in English or some other human language
Section 2.4 PROOF METHODS
97
First-order logic (metatheory)
Mathematics
Geometry
(theory)
Figure 2.5
Calculus
(theory)
First-order logic is a metatheory of mathematical theories.
and usually do not provide all of the details. Call these paragraph proofs. Their intention is to lead the reader from the premises to the conclusion, making the result
convincing. In many instances, a proof could be translated into the first-order logic,
but this is not needed to meet the need of the mathematician. However, that it could
be translated means that we can use first-order logic to help us write paragraph proofs,
and in this section, we make that connection.
Universal Proofs
Our first paragraph proofs will be for propositions with universal quantifiers. To prove
∀๐‘ฅ๐‘(๐‘ฅ) from a given set of premises, we show that every object satisfies ๐‘(๐‘ฅ) assuming
those premises. Since the proofs are mathematical, we can restrict the objects to a
particular universe. A proposition of the form ∀๐‘ฅ๐‘(๐‘ฅ) is then interpreted to mean that
๐‘(๐‘ฅ) holds for all x from a given universe. This restriction is reasonable because we are
studying mathematical things, not airplanes or puppies. To indicate that we have made
a restriction to a universe, we randomly select an object of the universe by writing an
introduction. This is a proposition that declares the type of object represented by the
variable. The following are examples of introductions:
Let ๐‘Ž be a real number.
Take ๐‘Ž to be an integer.
Suppose ๐‘Ž is a positive integer.
From here we show that ๐‘(๐‘Ž) is true. This process is exemplified by the next diagram:
98
Chapter 2 FIRST-ORDER LOGIC
∀x p(x)
Let a be an object in the universe.
Prove p(a).
These types of proofs are called universal proofs.
EXAMPLE 2.4.1
To prove that for all real numbers ๐‘ฅ,
(๐‘ฅ − 1)3 = ๐‘ฅ3 − 3๐‘ฅ2 + 3๐‘ฅ − 1,
we introduce a real number and then check the equation.
PROOF
Let ๐‘Ž be a real number. Then,
(๐‘Ž − 1)3 = (๐‘Ž − 1)(๐‘Ž − 1)2
= (๐‘Ž − 1)(๐‘Ž2 − 2๐‘Ž + 1)
= ๐‘Ž3 − 3๐‘Ž2 + 3๐‘Ž − 1.
For our next example, we need some terminology.
DEFINITION 2.4.2
For all integers ๐‘Ž and ๐‘, ๐‘Ž divides ๐‘ (written as ๐‘Ž โˆฃ ๐‘) if ๐‘Ž ≠ 0 and there exists an
integer ๐‘˜ such that ๐‘ = ๐‘Ž๐‘˜.
Therefore, 6 divides 18, but 8 does not divide 18. In this case, write 6 โˆฃ 18 and 8 - 18.
If ๐‘Ž โˆฃ ๐‘, we can also write that ๐‘ is divisible by ๐‘Ž, ๐‘Ž is a divisor or a factor of ๐‘, or ๐‘ is
a multiple of ๐‘Ž. For this reason, to translate a predicate like 4 divides ๐‘Ž, write
๐‘Ž = 4๐‘˜ for some integer ๐‘˜.
A common usage of divisibility is to check whether an integer is divisible by 2 or not.
DEFINITION 2.4.3
An integer ๐‘› is even if 2 โˆฃ ๐‘›, and ๐‘› is odd if 2 โˆฃ (๐‘› − 1).
We are now ready for the example.
EXAMPLE 2.4.4
Let us prove the proposition
Section 2.4 PROOF METHODS
99
the square of every even integer is even.
This can be rewritten using a variable:
for all even integers ๐‘›, ๐‘›2 is even.
The proof goes like this.
PROOF
Let ๐‘› be an even integer. This means that there exists an integer ๐‘˜ such
that ๐‘› = 2๐‘˜, so we can calculate:
๐‘›2 = (2๐‘˜)2 = 4๐‘˜2 = 2(2๐‘˜2 ).
Since 2๐‘˜2 is an integer, ๐‘›2 is even.
Notice how the definition was used in the proof. After the even number was
introduced, a proposition that translated the introduction into a form that was
easier to use was written. This was done using Definitions 2.4.2 and 2.4.3.
Existential Proofs
Suppose that we want to write a paragraph proof for ∃๐‘ฅ๐‘(๐‘ฅ). This means that we must
show that there exists at least one object of the universe that satisfies the formula ๐‘(๐‘ฅ).
It will be our job to find that object. To do this directly, we pick an object that we
think will satisfy ๐‘(๐‘ฅ). This object is called a candidate. We then check that it does
satisfy ๐‘(๐‘ฅ). This type of a proof is called a direct existential proof, and its structure
is illustrated as follows:
∃x p(x)
Choose a candidate from the universe.
Check that the candidate satisfies p(x).
EXAMPLE 2.4.5
To prove that there exists an integer ๐‘ฅ such that ๐‘ฅ2 + 2๐‘ฅ − 3 = 0, we find an
integer that satisfies the equation. A basic factorization yields
(๐‘ฅ + 3)(๐‘ฅ − 1) = 0.
Since ๐‘ฅ = −3 or ๐‘ฅ = 1 will work, we choose (arbitrarily) ๐‘ฅ = 1. Therefore,
(1)2 + 2(1) − 3 = 0,
proving the existence of the desired integer.
100
Chapter 2 FIRST-ORDER LOGIC
Suppose that we want to prove that there is a function ๐‘“ such that the derivative of ๐‘“
is 2๐‘ฅ. After a quick mental calculation, we choose ๐‘“ (๐‘ฅ) = ๐‘ฅ2 as a candidate and check
to find that
๐‘‘ 2
๐‘ฅ = 2๐‘ฅ.
(2.16)
๐‘‘๐‘ฅ
Notice that ๐‘‘โˆ•๐‘‘๐‘ฅ is a function that has functions as its inputs and outputs. Let ๐‘‘ represent this function. That is, ๐‘‘ is a function symbol such that
๐‘‘(๐‘“ )(๐‘ฅ) =
๐‘‘
๐‘“ (๐‘ฅ).
๐‘‘๐‘ฅ
A formula that represents the proposition that was just proved is
∃๐‘“ (๐‘‘(๐‘“ )(๐‘ฅ) = 2๐‘ฅ).
(2.17)
This is a second-order formula (page 73) because the variable symbol ๐‘ฅ represents a
real number (an object of the universe) and ๐‘“ is a function symbol taking real numbers
as arguments. Although this kind of reasoning is common to mathematics, it cannot
be written as a first-order formula. This shows that there is a purpose to second-order
logic. It is not a novelty.
EXAMPLE 2.4.6
To represent (2.16) as a first-order formula, let ๐‘‘ be a unary function symbol
representing the derivative and write
∀๐‘ฅ[๐‘‘(๐‘ฅ2 ) = 2๐‘ฅ].
Notice, however, that this formula does not convey the same meaning as (2.17).
Multiple Quantifiers
Let us take what we have learned concerning the universal and existential quantifiers
and write paragraph proofs involving both. The first example is a simple one from
algebra but will nicely illustrate our method.
EXAMPLE 2.4.7
Prove that for every real number ๐‘ฅ, there exists a real number ๐‘ฆ so that ๐‘ฅ + ๐‘ฆ = 2.
This translates to
∀๐‘ฅ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 2)
with the universe equal to the real numbers. Remembering that a universal quantifier must apply to all objects of the universe and an existential quantifier means
that we must find the desired object, we have the following:
Section 2.4 PROOF METHODS
Let x be a real number.
101
Find a real number y.
∀x ∃y (x + y = 2)
Check that x + y = 2.
After taking an arbitrary ๐‘ฅ, our candidate will be ๐‘ฆ = 2 − ๐‘ฅ.
PROOF
Let ๐‘ฅ be a real number. Choose ๐‘ฆ = 2 − ๐‘ฅ and calculate:
๐‘ฅ + ๐‘ฆ = ๐‘ฅ + (2 − ๐‘ฅ) = 2.
Now let us switch the order of the quantifiers.
EXAMPLE 2.4.8
To see that there exists an integer ๐‘ฅ such that for all integers ๐‘ฆ, ๐‘ฅ + ๐‘ฆ = ๐‘ฆ, we use
the following:
Find an integer x.
Let y be an integer.
∃x ∀y (x + y = y)
Show x + y = y.
In the proof, the first goal is to identify a candidate. Then, we must show that it
works with every real number.
PROOF
We claim that 0 is the sought after object. To see this, let ๐‘ฆ be an integer.
Then 0 + ๐‘ฆ = ๐‘ฆ.
The next example will involve two existential quantifiers. Therefore, we have to find
two candidates.
EXAMPLE 2.4.9
We prove that there exist real numbers ๐‘Ž and ๐‘ so that for every real number ๐‘ฅ,
(๐‘Ž + 2๐‘)๐‘ฅ + (2๐‘Ž − ๐‘) = 2๐‘ฅ − 6.
Translating we arrive at
102
Chapter 2 FIRST-ORDER LOGIC
Find a real number a.
Find a real number b.
∃a ∃b ∀x[(a + 2b)x + (2a − b) = 2x − 6]
Let x be a real number.
Show (a + 2b)x + (2a − b) = 2x − 6.
We have to choose two candidates, one for ๐‘Ž and one for ๐‘, and then check by
taking an arbitrary ๐‘ฅ.
PROOF
Solving the system,
๐‘Ž + 2๐‘ = 2,
2๐‘Ž − ๐‘ = −6,
we choose ๐‘Ž = −2 and ๐‘ = 2. Let ๐‘ฅ be a real number. Then,
(๐‘Ž + 2๐‘)๐‘ฅ + (2๐‘Ž − ๐‘) = (−2 + 2 ⋅ 2)๐‘ฅ + (2 ⋅ (−2) − 2) = 2๐‘ฅ − 6.
Counterexamples
There are many times in mathematics when we must show that a proposition of the
form ∀๐‘ฅ๐‘(๐‘ฅ) is false. This can be accomplished by proving ∃๐‘ฅ¬๐‘(๐‘ฅ) is true, and this
is done by showing that an object ๐‘Ž exists in the universe such that ๐‘(๐‘Ž) is false. This
object is called a counterexample to ∀๐‘ฅ๐‘(๐‘ฅ).
EXAMPLE 2.4.10
Show false:
๐‘ฅ + 2 = 7 for all real numbers ๐‘ฅ.
To do this, we show that ∃๐‘ฅ(๐‘ฅ + 2 ≠ 7) is true by noting that the real number 0
satisfies ๐‘ฅ + 2 ≠ 7. Hence, 0 is a counterexample to ∀๐‘ฅ(๐‘ฅ + 2 = 7).
The idea of a counterexample can be generalized to cases with multiple universal quantifiers.
EXAMPLE 2.4.11
We know from algebra that every quadratic polynomial with real coefficients has
a real zero is false. This can be symbolized as
∀๐‘Ž∀๐‘∀๐‘∃๐‘ฅ(๐‘Ž๐‘ฅ2 + ๐‘๐‘ฅ + ๐‘ = 0),
where the universe is the collection of real numbers. The counterexample is found
by demonstrating
∃๐‘Ž∃๐‘∃๐‘∀๐‘ฅ(๐‘Ž๐‘ฅ2 + ๐‘๐‘ฅ + ๐‘ ≠ 0).
Section 2.4 PROOF METHODS
103
Many polynomials could be chosen, but we select ๐‘ฅ2 + 1. Its only zeros are ๐‘– and
−๐‘–. This means that ๐‘Ž = 1, ๐‘ = 0, and ๐‘ = 1 is our counterexample.
Direct Proof
Direct Proof is the preferred method for proving implications. To use Direct Proof to
write a paragraph proof identify the antecedent and consequent of the implication and
then follow these steps:
โˆ™ assume the antecedent,
โˆ™ translate the antecedent,
โˆ™ translate the consequent so that the goal of the proof is known,
โˆ™ deduce the consequent.
Our first example will use Definitions 2.4.2 and 2.4.3. Notice the introductions in the
proof.
EXAMPLE 2.4.12
We use Direct Proof to write a paragraph proof of the proposition
for all integers ๐‘ฅ, if 4 divides ๐‘ฅ, then ๐‘ฅ is even.
First, randomly choose an integer ๐‘Ž and then identify the antecedent and consequent:
if 4 divides ๐‘ฅ , then ๐‘ฅ is even .
Use these to identify the structure of the proof:
4 divides a.
This means a = 4k.
…
Show a = 2l.
2 divides a.
Assume the antecedent.
Translate the antecedent.
Translate the consequent.
Deduce the consequent.
Now write the final version from this structure.
PROOF
Assume that 4 divides the integer ๐‘Ž. This means ๐‘Ž = 4๐‘˜ for some integer
๐‘˜. We must show that ๐‘Ž = 2๐‘™ for some integer ๐‘™, but we know that ๐‘Ž =
4๐‘˜ = 2(2๐‘˜). Hence, let ๐‘™ = 2๐‘˜.
Sometimes it is difficult to prove a conditional directly. An alternative is to prove
the contrapositive. This is sometimes easier or simply requires fewer lines. The next
example shows this method in a paragraph proof.
104
Chapter 2 FIRST-ORDER LOGIC
EXAMPLE 2.4.13
Let us show that
for all integers ๐‘›, if ๐‘›2 is odd, then ๐‘› is odd.
A direct proof of this is a problem. Instead, we prove its contrapositive,
if ๐‘› is not odd, then ๐‘›2 is not odd.
In other words, we prove that
if ๐‘› is even, then ๐‘›2 is even.
This will be done using Direct Proof:
Let n be even.
This means n = 2k.
…
Show n2 = 2l.
2 divides n 2 .
Assume the antecedent.
Translate the antecedent.
Translate the consequent.
Deduce the consequent.
This leads to the final proof.
PROOF
Let ๐‘› be an even integer. This means that ๐‘› = 2๐‘˜ for some integer ๐‘˜. To
see that ๐‘›2 is even, calculate to find
๐‘›2 = (2๐‘˜)2 = 4๐‘˜2 = 2(2๐‘˜2 ).
Notice that this proof is basically the same as the proof for the square of every
even integer is even on page 99. This illustrates that there is a connection between
universal proofs and direct proofs (Exercise 4).
Existence and Uniqueness
There will be times when we want to show that there exists exactly one object that
satisfies a given predicate ๐‘(๐‘ฅ). In other words, there exists a unique object that satisfies
๐‘(๐‘ฅ). This is a two-step process.
โˆ™ Existence: Show that there is at least one object that satisfies ๐‘(๐‘ฅ).
โˆ™ Uniqueness: Show that there is at most one object that satisfies ๐‘(๐‘ฅ). This is
usually done by assuming both ๐‘Ž and ๐‘ satisfy ๐‘(๐‘ฅ) and then proving ๐‘Ž = ๐‘.
This means to prove that there exists a unique ๐‘ฅ such that ๐‘(๐‘ฅ), we prove
∃๐‘ฅ๐‘(๐‘ฅ) ∧ ∀๐‘ฅ∀๐‘ฆ(๐‘(๐‘ฅ) ∧ ๐‘(๐‘ฆ) → ๐‘ฅ = ๐‘ฆ).
Use direct or indirect existential proof to demonstrate that an object exists. The next
example illustrates proving uniqueness.
Section 2.4 PROOF METHODS
105
EXAMPLE 2.4.14
Let ๐‘š and ๐‘› be nonnegative integers with ๐‘š ≠ 0. To show that there is at most
one pair of nonnegative integers ๐‘Ÿ and ๐‘ž such that
๐‘Ÿ < ๐‘š and ๐‘› = ๐‘š๐‘ž + ๐‘Ÿ,
suppose in addition to ๐‘Ÿ and ๐‘ž that there exists nonnegative integers ๐‘Ÿ′ and ๐‘ž ′ such
that
๐‘Ÿ′ < ๐‘š and ๐‘› = ๐‘š๐‘ž ′ + ๐‘Ÿ′ .
Assume that ๐‘Ÿ′ > ๐‘Ÿ. By Exercise 13, ๐‘ž > ๐‘ž ′ , so there exists ๐‘ข, ๐‘ฃ > 0 such that
๐‘Ÿ′ = ๐‘Ÿ + ๐‘ข and ๐‘ž = ๐‘ž ′ + ๐‘ฃ.
Therefore,
๐‘š(๐‘ž ′ + ๐‘ฃ) + ๐‘Ÿ = ๐‘š๐‘ž ′ + ๐‘Ÿ + ๐‘ข,
๐‘š๐‘ž ′ + ๐‘š๐‘ฃ + ๐‘Ÿ = ๐‘š๐‘ž ′ + ๐‘Ÿ + ๐‘ข,
๐‘š๐‘ฃ = ๐‘ข.
Since ๐‘ฃ > 0, there exists ๐‘ค such that ๐‘ฃ = ๐‘ค + 1. Hence,
๐‘š๐‘ค + ๐‘š = ๐‘š(๐‘ค + 1) = ๐‘š๐‘ฃ = ๐‘ข,
so ๐‘š ≤ ๐‘ข. However, since ๐‘Ÿ < ๐‘š and ๐‘Ÿ′ < ๐‘š, we have that ๐‘ข < ๐‘š (Exercise 13),
which is impossible. Lastly, the assumption ๐‘Ÿ > ๐‘Ÿ′ leads to a similar contradiction. Therefore, ๐‘Ÿ = ๐‘Ÿ′ , which implies that ๐‘ž = ๐‘ž ′ .
EXAMPLE 2.4.15
To prove 2๐‘ฅ + 1 = 5 has a unique real solution, we show
∃๐‘ฅ(2๐‘ฅ + 1 = 5)
and
∀๐‘ฅ∀๐‘ฆ(2๐‘ฅ + 1 = 5 ∧ 2๐‘ฆ + 1 = 5 → ๐‘ฅ = ๐‘ฆ).
โˆ™ We know that ๐‘ฅ = 2 is a solution since 2(2) + 1 = 5.
โˆ™ Suppose that ๐‘Ž and ๐‘ are solutions. We know that both 2๐‘Ž + 1 = 5 and
2๐‘ + 1 = 5, so we calculate:
2๐‘Ž + 1 = 2๐‘ + 1,
2๐‘Ž = 2๐‘,
๐‘Ž = ๐‘.
Indirect Proof
To use indirect proof, assume each premise and then assume the negation of the conclusion. Then proceed with the proof until a contradiction is reached. At this point,
deduce the original conclusion.
106
Chapter 2 FIRST-ORDER LOGIC
EXAMPLE 2.4.16
Earlier we proved
for all integers ๐‘›, if ๐‘›2 is odd, then ๐‘› is odd
by showing its contrapositive. Here we nest an Indirect Proof within a Direct
Proof:
Assume is odd.
Suppose n is even.
Assume the antecedent.
Assume the negation.
…
n2
Contradiction
n is odd.
Deduce a contradiction.
Conclude the consequent.
We use this structure to write the paragraph proof.
PROOF
Take an integer ๐‘› and let ๐‘›2 be odd. In order to obtain a contradiction,
assume that ๐‘› is even. So, ๐‘› = 2๐‘˜ for some integer ๐‘˜. Substituting, we
have
๐‘›2 = (2๐‘˜)2 = 2(2๐‘˜2 ),
showing that ๐‘›2 is even. This is a contradiction. Therefore, ๐‘› is an odd
integer.
Indirect proof has been used to prove many famous mathematical results including the
next example.
EXAMPLE 2.4.17
We show that
√
2 is an irrational number. Suppose instead that
√
๐‘Ž
2= ,
๐‘
where ๐‘Ž and ๐‘ are integers, ๐‘ ≠ 0, and the fraction ๐‘Žโˆ•๐‘ has been reduced. Then,
2=
๐‘Ž2
,
๐‘2
so ๐‘Ž2 = 2๐‘2 . Therefore, ๐‘Ž = 2๐‘˜ for some integer ๐‘˜ [Exercise 11(c)]. We conclude
that ๐‘2 = 2๐‘˜2 , which implies that ๐‘ is also even. However, this is impossible
because the fraction was reduced.
There are times when it is difficult or impossible to find a candidate for an existential
proof. When this happens, an indirect existential proof can sometimes help.
Section 2.4 PROOF METHODS
107
EXAMPLE 2.4.18
If ๐‘› is an integer such that ๐‘› > 1, then ๐‘› is prime means that the only positive
divisors of ๐‘› are 1 and ๐‘›; else ๐‘› is composite. From the definition, if ๐‘› is composite, there exist ๐‘Ž and ๐‘ such that ๐‘› = ๐‘Ž๐‘ and 1 < ๐‘Ž ≤ ๐‘ < ๐‘›. For example, 2,
11, and 97 are prime, and 4 and 87 are composite. Euclid proved that there are
infinitely many prime numbers (Elements IX.20). Since it is impossible to find
all of these numbers, we prove this theorem indirectly. Suppose that there are
finitely many prime numbers and list them as
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 .
Consider
๐‘ž = ๐‘0 ๐‘1 · · · ๐‘๐‘›−1 + 1.
If ๐‘ž is prime, ๐‘ž = ๐‘๐‘– for some ๐‘– = 0, 1, … , ๐‘› − 1, which would imply that ๐‘๐‘–
divides 1. Therefore, ๐‘ž is composite, but this is also impossible because ๐‘ž must
have a prime factor, which again means that 1 would be divisible by a prime.
Biconditional Proof
The last three types of proofs that we will examine rely on direct proof. The first of
these takes three forms, and they provide the usual method of proving biconditionals.
The first form follows because ๐‘ → ๐‘ž and ๐‘ž → ๐‘ imply (๐‘ → ๐‘ž) ∧ (๐‘ž → ๐‘) by Conj,
which implies ๐‘ ↔ ๐‘ž by Equiv.
INFERENCE RULE 2.4.19 [Biconditional Proof (BP)]
๐‘ → ๐‘ž, ๐‘ž → ๐‘ ⇒ ๐‘ ↔ ๐‘ž.
The rule states that a biconditional can be proved by showing both implications. Each is
usually proved with direct proof. As is seen in the next example, the ๐‘ → ๐‘ž subproof is
introduced by (→) and its converse with (←). The conclusions of the two applications
of direct proof are combined in line 7.
EXAMPLE 2.4.20
Prove: ๐‘ → ๐‘ž โŠข ๐‘ ∧ ๐‘ž ↔ ๐‘
1.
(→) 2.
3.
(←) 4.
5.
6.
7.
๐‘→๐‘ž
→๐‘ ∧ ๐‘ž
๐‘
→๐‘
๐‘ž
๐‘∧๐‘ž
๐‘∧๐‘ž ↔๐‘
Given
Assumption
2 Simp
Assumption
1, 4 MP
4, 5 Conj
2–6 BP
108
Chapter 2 FIRST-ORDER LOGIC
Sometimes the steps for one part are simply the steps for the other in reverse. When
this happens, our work is cut in half, and we can use the short rule of biconditional
proof. These proofs are simply a sequence of replacements with or without the reasons.
This is a good method when only rules of replacement are used as in the next example.
(Notice that there are no hypotheses to assume.)
EXAMPLE 2.4.21
Prove: โŠข ๐‘ → ๐‘ž ↔ ๐‘ ∧ ¬๐‘ž → ¬๐‘
๐‘ → ๐‘ž ⇔ ¬๐‘ ∨ ๐‘ž
⇔ ¬๐‘ ∨ ¬๐‘ ∨ ๐‘ž
⇔ ¬๐‘ ∨ (¬๐‘ ∨ ๐‘ž)
⇔ ¬๐‘ ∨ ๐‘ž ∨ ¬๐‘
⇔ ¬๐‘ ∨ ¬¬๐‘ž ∨ ¬๐‘
⇔ ¬(๐‘ ∧ ¬๐‘ž) ∨ ¬๐‘
⇔ ๐‘ ∧ ¬๐‘ž → ¬๐‘
Impl
Idem
Assoc
Com
DN
DM
Impl
EXAMPLE 2.4.22
Let us use biconditional proof to show that
for all integers ๐‘›, ๐‘› is even if and only if ๐‘›3 is even.
Since this is a biconditional, we must show both implications:
if ๐‘› is even , then ๐‘›3 is even ,
and
if ๐‘›3 is even , then ๐‘› is even .
To prove the second conditional, we prove its contrapositive. Therefore, using
the pattern of the previous example, the structure is
(→)
(←)
→Assume ๐‘› is even
Then, ๐‘› = 2๐‘˜ for some integer ๐‘˜
โ‹ฎ
๐‘›3 = 2๐‘™, with ๐‘™ being an integer
๐‘›3 is even
→Assume ๐‘› is odd
Then, ๐‘› = 2๐‘˜ + 1 for some integer ๐‘˜
โ‹ฎ
๐‘›3 = 2๐‘™ + 1, with ๐‘™ being an integer
๐‘›3 is odd
Section 2.4 PROOF METHODS
109
PROOF
Let ๐‘› be an integer.
โˆ™ Assume ๐‘› is even. Then, ๐‘› = 2๐‘˜ for some integer ๐‘˜. We show that
๐‘›3 is even. To do this, we calculate:
๐‘›3 = (2๐‘˜)3 = 2(4๐‘˜3 ),
which means that ๐‘›3 is even.
โˆ™ Now suppose that ๐‘› is odd. This means that ๐‘› = 2๐‘˜ + 1 for some
integer ๐‘˜. To show that ๐‘›3 is odd, we again calculate:
๐‘›3 = (2๐‘˜ + 1)3 = 8๐‘˜3 + 12๐‘˜2 + 6๐‘˜ + 1 = 2(4๐‘˜3 + 6๐‘˜2 + 3๐‘˜) + 1.
Hence, ๐‘›3 is odd.
Notice that the words were chosen carefully to make the proof more readable. Furthermore, the example could have been written with the words necessary and sufficient
introducing the two subproofs. The (→) step could have been introduced with a phrase
like
to show sufficiency,
and the (←) could have opened with
as for necessity.
There will be times when we need to prove a sequence of biconditionals. The propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 are pairwise equivalent if for all ๐‘–, ๐‘—,
๐‘๐‘– ↔ ๐‘๐‘— .
In other words,
๐‘0 ↔ ๐‘1 , ๐‘0 ↔ ๐‘2 , … , ๐‘1 ↔ ๐‘2 , ๐‘1 ↔ ๐‘3 , … , ๐‘๐‘›−2 ↔ ๐‘๐‘›−1 .
To prove all of these, we make use of the Hypothetical Syllogism. For example, if we
know that ๐‘0 → ๐‘1 , ๐‘1 → ๐‘2 , and ๐‘2 → ๐‘0 , then
๐‘0 → ๐‘2 (because ๐‘0 → ๐‘1 ∧ ๐‘1 → ๐‘2 ),
๐‘1 → ๐‘0 (because ๐‘1 → ๐‘2 ∧ ๐‘2 → ๐‘0 ),
๐‘2 → ๐‘1 (because ๐‘2 → ๐‘0 ∧ ๐‘0 → ๐‘1 ).
The result is the equivalence rule.
110
Chapter 2 FIRST-ORDER LOGIC
INFERENCE RULE 2.4.23 [Equivalence Rule]
To prove that the propositional forms ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 are pairwise equivalent,
prove:
๐‘0 → ๐‘1 , ๐‘1 → ๐‘2 , … , ๐‘๐‘›−2 → ๐‘๐‘›−1 , ๐‘๐‘›−1 → ๐‘0 .
In practice, the equivalence rule will typically be used to prove propositions that
include the phrase
the following are equivalent.
EXAMPLE 2.4.24
Let
๐‘“ (๐‘ฅ) = ๐‘Ž๐‘› ๐‘ฅ๐‘› + ๐‘Ž๐‘›−1 ๐‘ฅ๐‘›−1 + · · · + ๐‘Ž1 ๐‘ฅ + ๐‘Ž0
be a polynomial with real coefficients. That is, each ๐‘Ž๐‘– is a real number and ๐‘› is
a nonnegative integer. An integer ๐‘Ÿ is a zero of ๐‘“ (๐‘ฅ) if
๐‘Ž๐‘› ๐‘Ÿ๐‘› + ๐‘Ž๐‘›−1 ๐‘Ÿ๐‘›−1 + · · · + ๐‘Ž1 ๐‘Ÿ + ๐‘Ž0 = 0,
written as ๐‘“ (๐‘Ÿ) = 0, and a polynomial ๐‘”(๐‘ฅ) is a factor of ๐‘“ (๐‘ฅ) if there is a polynomial โ„Ž(๐‘ฅ) such that ๐‘“ (๐‘ฅ) = ๐‘”(๐‘ฅ)โ„Ž(๐‘ฅ). Whether ๐‘”(๐‘ฅ) is a factor of ๐‘“ (๐‘ฅ) or not,
there exist unique polynomials ๐‘ž(๐‘ฅ) and ๐‘Ÿ(๐‘ฅ) such that
๐‘“ (๐‘ฅ) = ๐‘”(๐‘ฅ)๐‘ž(๐‘ฅ) + ๐‘Ÿ(๐‘ฅ)
(2.18)
the degree of ๐‘Ÿ(๐‘ฅ) is less than the degree of ๐‘”(๐‘ฅ).
(2.19)
and
This result is called the polynomial division algorithm, The polynomial ๐‘ž(๐‘ฅ) is
the quotient and ๐‘Ÿ(๐‘ฅ) is the remainder. We prove that the following are equivalent:
โˆ™ ๐‘Ÿ is a zero of ๐‘“ (๐‘ฅ).
โˆ™ ๐‘Ÿ is a solution to ๐‘“ (๐‘ฅ) = 0.
โˆ™ ๐‘ฅ − ๐‘Ÿ is a factor of ๐‘“ (๐‘ฅ).
To do this, we prove three conditionals:
if ๐‘Ÿ is a zero of ๐‘“ (๐‘ฅ) , then ๐‘Ÿ is a solution to ๐‘“ (๐‘ฅ) = 0 ,
if ๐‘Ÿ is a solution to ๐‘“ (๐‘ฅ) = 0 , then ๐‘ฅ − ๐‘Ÿ is a factor of ๐‘“ (๐‘ฅ) ,
and
if ๐‘ฅ − ๐‘Ÿ is a factor of ๐‘“ (๐‘ฅ) , then ๐‘Ÿ is a zero of ๐‘“ (๐‘ฅ) .
Section 2.4 PROOF METHODS
111
We use direct proof on each.
PROOF
Let ๐‘Ž๐‘› ๐‘ฅ๐‘› + ๐‘Ž๐‘›−1 ๐‘ฅ๐‘›−1 + · · · + ๐‘Ž1 ๐‘ฅ + ๐‘Ž0 be a polynomial and assume that the
coefficients are real numbers. Denote the polynomial by ๐‘“ (๐‘ฅ).
โˆ™ Let ๐‘Ÿ be a zero of ๐‘“ (๐‘ฅ). By definition, this means ๐‘“ (๐‘Ÿ) = 0, so ๐‘Ÿ is a
solution to ๐‘“ (๐‘ฅ) = 0.
โˆ™ Suppose ๐‘Ÿ is a solution to ๐‘“ (๐‘ฅ) = 0. The polynomial division algorithm (2.18) gives polynomials ๐‘ž(๐‘ฅ) and ๐‘Ÿ(๐‘ฅ) such that
๐‘“ (๐‘ฅ) = ๐‘ž(๐‘ฅ)(๐‘ฅ − ๐‘Ÿ) + ๐‘Ÿ(๐‘ฅ)
and the degree of ๐‘Ÿ(๐‘ฅ) is less than 1 (2.19). Hence, ๐‘Ÿ(๐‘ฅ) is a constant
that we simply write as ๐‘. Now,
0 = ๐‘“ (๐‘Ÿ) = ๐‘ž(๐‘Ÿ)(๐‘Ÿ − ๐‘Ÿ) + ๐‘ = 0 + ๐‘ = ๐‘.
Therefore, ๐‘“ (๐‘ฅ) = ๐‘ž(๐‘ฅ)(๐‘ฅ − ๐‘Ÿ), so ๐‘ฅ − ๐‘Ÿ is a factor of ๐‘“ (๐‘ฅ).
โˆ™ Lastly, assume ๐‘ฅ − ๐‘Ÿ is a factor of ๐‘“ (๐‘ฅ). This means that there exists
a polynomial ๐‘ž(๐‘ฅ) so that
๐‘“ (๐‘ฅ) = (๐‘ฅ − ๐‘Ÿ)๐‘ž(๐‘ฅ).
Thus,
๐‘“ (๐‘Ÿ) = (๐‘Ÿ − ๐‘Ÿ)๐‘ž(๐‘Ÿ) = 0,
which means ๐‘Ÿ is a zero of ๐‘“ (๐‘ฅ).
Proof of Disunctions
The second type of proof that relies on direct proof is the proof of a disjunction. To
prove ๐‘ ∨ ๐‘ž, it is standard to assume ¬๐‘ and show ๐‘ž. This means that we would be using
direct proof to show ¬๐‘ → ๐‘ž. This is what we want because
¬๐‘ → ๐‘ž ⇔ ¬¬๐‘ ∨ ๐‘ž ⇔ ๐‘ ∨ ๐‘ž.
The intuition behind the strategy goes like this. If we need to prove ๐‘ ∨ ๐‘ž from some
hypotheses, it is not reasonable to believe that we can simply prove ๐‘ and then use
Addition to conclude ๐‘ ∨ ๐‘ž. Indeed, if we could simply prove ๐‘, we would expect the
conclusion to be stated as ๐‘ and not ๐‘ ∨ ๐‘ž. Hence, we need to incorporate both disjuncts
into the proof. We do this by assuming the negation of one of the disjuncts. If we can
prove the other, the disjunction must be true.
112
Chapter 2 FIRST-ORDER LOGIC
EXAMPLE 2.4.25
To prove
for all integers ๐‘Ž and ๐‘, if ๐‘Ž๐‘ = 0, then ๐‘Ž = 0 or ๐‘ = 0,
we assume ๐‘Ž๐‘ = 0 and show that ๐‘Ž ≠ 0 implies ๐‘ = 0.
PROOF
Let ๐‘Ž and ๐‘ be integers. Let ๐‘Ž๐‘ = 0 and suppose ๐‘Ž ≠ 0. Then, ๐‘Ž−1 exists.
Multiplying both sides of the equation by ๐‘Ž−1 gives
๐‘Ž−1 ๐‘Ž๐‘ = ๐‘Ž−1 ⋅ 0,
so ๐‘ = 0.
Proof by Cases
The last type of proof that relies on direct proof is proof by cases. Suppose that we
want to prove ๐‘ → ๐‘ž and this is difficult for some reason. We notice, however, that ๐‘
can be broken into cases. Namely, there exist ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 such that
๐‘ ⇔ ๐‘0 ∨ ๐‘1 ∨ · · · ∨ ๐‘๐‘›−1 .
If we can prove ๐‘๐‘– → ๐‘ž for each ๐‘–, we have proved ๐‘ → ๐‘ž. If ๐‘› = 2, then ๐‘ ⇔ ๐‘0 ∨ ๐‘1 ,
and the justification of this is as follows:
(๐‘0 → ๐‘ž) ∧ (๐‘1 → ๐‘ž) ⇔ (¬๐‘0 ∨ ๐‘ž) ∧ (¬๐‘1 ∨ ๐‘ž)
⇔ (๐‘ž ∨ ¬๐‘0 ) ∧ (๐‘ž ∨ ¬๐‘1 )
⇔ ๐‘ž ∨ ¬๐‘0 ∧ ¬๐‘1
⇔ ¬๐‘0 ∧ ¬๐‘1 ∨ ๐‘ž
⇔ ¬(๐‘0 ∨ ๐‘1 ) ∨ ๐‘ž
⇔ ๐‘0 ∨ ๐‘1 → ๐‘ž
⇔ ๐‘ → ๐‘ž.
This generalizes to the next rule.
INFERENCE RULE 2.4.26 [Proof by Cases (CP)]
For every positive integer ๐‘›, if ๐‘ ⇔ ๐‘0 ∨ ๐‘1 ∨ · · · ∨ ๐‘๐‘›−1 , then
๐‘0 → ๐‘ž, ๐‘1 → ๐‘ž, … , ๐‘๐‘›−1 → ๐‘ž ⇒ ๐‘ → ๐‘ž.
For example, since ๐‘Ž is a real number if and only if ๐‘Ž > 0, ๐‘Ž = 0, or ๐‘Ž < 0, if we
needed to prove a proposition about an arbitrary real number, it would suffice to prove
the result individually for ๐‘Ž > 0, ๐‘Ž = 0, and ๐‘Ž < 0.
Section 2.4 PROOF METHODS
113
EXAMPLE 2.4.27
Our example of a proof by cases is a well-known one:
for all integers ๐‘Ž and ๐‘, if ๐‘Ž = ±๐‘, then ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž.
The antecedent means ๐‘Ž = ๐‘ or ๐‘Ž = −๐‘, which are the two cases, so we have to
show both
if ๐‘Ž = ๐‘, then ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž
and
if ๐‘Ž = −๐‘, then ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž.
This leads to the following structure:
(Case 1)
(Case 2)
→Suppose ๐‘Ž = ±๐‘
Thus, ๐‘Ž = ๐‘ or ๐‘Ž = −๐‘
→Assume ๐‘Ž = ๐‘
โ‹ฎ
๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž
→Assume ๐‘Ž = −๐‘
โ‹ฎ
๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž
∴ ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž
The final proof looks something like this.
PROOF
Let ๐‘Ž and ๐‘ be integers and suppose ๐‘Ž = ±๐‘. To show ๐‘Ž divides ๐‘ and ๐‘
divides ๐‘Ž, we have two cases to prove.
โˆ™ Assume ๐‘Ž = ๐‘. Then, ๐‘Ž = ๐‘ ⋅ 1 and ๐‘ = ๐‘Ž ⋅ 1.
โˆ™ Next assume ๐‘Ž = −๐‘. This means that ๐‘Ž = ๐‘ ⋅ (−1) and ๐‘ = ๐‘Ž ⋅ (−1).
In both the cases, we have proved that ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘Ž.
Exercises
1. Let ๐‘› be an integer. Demonstrate that each of the following are divisible by 6.
(a) 18
(b) −24
(c) 0
(d) 6๐‘› + 12
(e) 23 ⋅ 34 ⋅ 75
(f) (2๐‘› + 2)(3๐‘› + 6)
2. Let ๐‘Ž be a nonzero integer. Write paragraph proofs for each proposition.
114
Chapter 2 FIRST-ORDER LOGIC
(a) 1 divides ๐‘Ž.
(b) ๐‘Ž divides 0.
(c) ๐‘Ž divides ๐‘Ž.
3. Write a universal proof for each proposition.
(a) For all real numbers ๐‘ฅ, (๐‘ฅ + 2)2 = ๐‘ฅ2 + 4๐‘ฅ + 4.
(b) For all integers ๐‘ฅ, ๐‘ฅ − 1 divides ๐‘ฅ3 − 1.
(c) The square of every even integer is even.
4. Show why a proof of (∀๐‘ฅ)๐‘(๐‘ฅ) is an application of direct proof.
5. Write existential proofs for each proposition.
(a) There exists a real number ๐‘ฅ such that ๐‘ฅ − ๐œ‹ = 9.
(b) There exists an integer ๐‘ฅ such that ๐‘ฅ2 + 2๐‘ฅ − 3 = 0.
(c) The square of some integer is odd.
6. Prove each of the following by writing a paragraph proof.
(a) For all real numbers ๐‘ฅ, ๐‘ฆ, and ๐‘ง, ๐‘ฅ2 ๐‘ง + 2๐‘ฅ๐‘ฆ๐‘ง + ๐‘ฆ2 ๐‘ง = ๐‘ง(๐‘ฅ + ๐‘ฆ)2 .
(b) There exist real numbers ๐‘ข and ๐‘ฃ such that 2๐‘ข + 5๐‘ฃ = −29.
(c) For all real numbers ๐‘ฅ, there exists a real number ๐‘ฆ so that ๐‘ฅ − ๐‘ฆ = 10.
(d) There exists an integer ๐‘ฅ such that for all integers ๐‘ฆ, ๐‘ฆ๐‘ฅ = ๐‘ฅ.
(e) For all real numbers ๐‘Ž, ๐‘, and ๐‘, there exists a complex number ๐‘ฅ such that
๐‘Ž๐‘ฅ2 + ๐‘๐‘ฅ + ๐‘ = 0.
(f) There exists an integer that divides every integer.
7. Provide counterexamples for each of the following false propositions.
(a) Every integer is a solution to ๐‘ฅ + 1 = 0.
(b) For every integer ๐‘ฅ, there exists an integer ๐‘ฆ such that ๐‘ฅ๐‘ฆ = 1.
(c) The product of any two integers is even.
(d) For every integer ๐‘›, if ๐‘› is even, then ๐‘›2 is a multiple of eight.
8. Assuming that ๐‘Ž, ๐‘, ๐‘, and ๐‘‘ are integers with ๐‘Ž ≠ 0 and ๐‘ ≠ 0, give paragraph
proofs using direct proof for the following divisibility results.
(a) If ๐‘Ž divides ๐‘, then ๐‘Ž divides ๐‘๐‘‘.
(b) If ๐‘Ž divides ๐‘ and ๐‘Ž divides ๐‘‘, then ๐‘Ž2 divides ๐‘๐‘‘.
(c) If ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘‘, then ๐‘Ž๐‘ divides ๐‘๐‘‘.
(d) If ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘, then ๐‘Ž divides ๐‘.
9. Write paragraph proofs using direct proof.
(a) The sum of two even integers is even.
(b) The sum of two odd integers is even.
(c) The sum of an even and an odd is odd.
(d) The product of two even integers is even.
(e) The product of two odd integers is odd.
(f) The product of an even and an odd is even.
10. Let ๐‘Ž and ๐‘ be integers. Write paragraph proofs.
Section 2.4 PROOF METHODS
115
(a) If ๐‘Ž and ๐‘ are even, then ๐‘Ž4 + ๐‘4 + 32 is divisible by 8.
(b) If ๐‘Ž and ๐‘ are odd, then 4 divides ๐‘Ž4 + ๐‘4 + 6.
11. Prove the following by using direct proof to prove the contrapositive.
(a) For every integer ๐‘›, if ๐‘›4 is even, then ๐‘› is even.
(b) For all integers ๐‘›, if ๐‘›3 + ๐‘›2 is odd, then ๐‘› is odd.
(c) For all integers ๐‘Ž and ๐‘, if ๐‘Ž๐‘ is even, then ๐‘Ž is even or ๐‘ is even.
12. Write paragraph proofs.
(a) The equation ๐‘ฅ − 10 = 23 has a unique solution.
√
(b) The equation 2๐‘ฅ − 5 = 2 has a unique solution.
(c) For every real number ๐‘ฆ, the equation 2๐‘ฅ + 5๐‘ฆ = 10 has a unique solution.
(d) The equation ๐‘ฅ2 + 5๐‘ฅ + 6 = 0 has at most two integer solutions.
13. From the proof of Example 2.4.14, prove that ๐‘ž > ๐‘ž ′ and ๐‘ข < ๐‘š.
14. Prove the results of Exercise 9 indirectly.
15. Prove using the method of biconditional proof.
(a) ๐‘ ∨ ๐‘ž → ¬๐‘Ÿ, ๐‘  → ๐‘Ÿ, ¬๐‘ ∨ ๐‘ž → ๐‘  โŠข ๐‘ ↔ ¬๐‘ 
(b) ๐‘ ∨ (¬๐‘ž ∨ ๐‘), ๐‘ž ∨ (¬๐‘ ∨ ๐‘ž) โŠข ๐‘ ↔ ๐‘ž
(c) (๐‘ → ๐‘ž) ∧ (๐‘Ÿ → ๐‘ ), (๐‘ → ¬๐‘ ) ∧ (๐‘Ÿ → ¬๐‘ž), ๐‘ ∨ ๐‘Ÿ โŠข ๐‘ž ↔ ¬๐‘ 
(d) ๐‘ ∧ ๐‘Ÿ → ¬(๐‘  ∨ ๐‘ก), ¬๐‘  ∨ ¬๐‘ก → ๐‘ ∧ ๐‘Ÿ โŠข ๐‘  ↔ ๐‘ก
16. Prove using the short rule of biconditional proof.
(a) ๐‘ → ๐‘ž ∧ ๐‘Ÿ ⇔ (๐‘ → ๐‘ž) ∧ (๐‘ → ๐‘Ÿ)
(b) ๐‘ → ๐‘ž ∨ ๐‘Ÿ ⇔ ๐‘ ∧ ¬๐‘ž → ๐‘Ÿ
(c) ๐‘ ∨ ๐‘ž → ๐‘Ÿ ⇔ (๐‘ → ๐‘Ÿ) ∧ (๐‘ž → ๐‘Ÿ)
(d) ๐‘ ∧ ๐‘ž → ๐‘Ÿ ⇔ (๐‘ → ๐‘Ÿ) ∨ (๐‘ž → ๐‘Ÿ)
17. Let ๐‘Ž, ๐‘, ๐‘, and ๐‘‘ be integers. Write paragraph proofs using biconditional proof.
(a) ๐‘Ž is even if and only if ๐‘Ž2 is even.
(b) ๐‘Ž is odd if and only if ๐‘Ž + 1 is even.
(c) ๐‘Ž is even if and only if ๐‘Ž + 2 is even.
(d) ๐‘Ž3 + ๐‘Ž2 + ๐‘Ž is even if and only if ๐‘Ž is even.
(e) If ๐‘ ≠ 0, then ๐‘Ž divides ๐‘ if and only if ๐‘Ž๐‘ divides ๐‘๐‘.
18. Suppose that ๐‘Ž and ๐‘ are integers. Prove that the following are equivalent.
โˆ™ ๐‘Ž divides ๐‘.
โˆ™ ๐‘Ž divides −๐‘.
โˆ™ −๐‘Ž divides ๐‘.
โˆ™ −๐‘Ž divides −๐‘.
19. Let ๐‘Ž be an integer. Prove that the following are equivalent.
โˆ™ ๐‘Ž is divisible by 3.
โˆ™ 3๐‘Ž is divisible by 9.
โˆ™ ๐‘Ž + 3 is divisible by 3.
116
Chapter 2 FIRST-ORDER LOGIC
20. Prove by using direct proof but do not use the contrapositive: for all integers ๐‘Ž and
๐‘, if ๐‘Ž๐‘ is even, then ๐‘Ž is even or ๐‘ is even.
21. Prove using proof by cases (Inference Rule 2.4.26).
(a) For all integers ๐‘Ž, if ๐‘Ž = 0 or ๐‘ = 0, then ๐‘Ž๐‘ = 0.
(b) The square of every odd integer is of the form 8๐‘˜ + 1 for some integer ๐‘˜.
(Hint: Square 2๐‘™ + 1. Then consider two cases: ๐‘™ is even and ๐‘™ is odd.)
(c) ๐‘Ž2 + ๐‘Ž + 1 is odd for every integer ๐‘Ž.
(d) The fourth power of every odd integer is of the form 16๐‘˜+1 for some integer
๐‘˜.
(e) Every nonhorizontal line intersects the ๐‘ฅ-axis.
22. Let ๐‘Ž be an integer. Prove by cases.
(a) 2 divides ๐‘Ž(๐‘Ž + 1)
(b) 3 divides ๐‘Ž(๐‘Ž + 1)(๐‘Ž + 2)
23. For any real number ๐‘, the absolute value of ๐‘ is defined as
{
๐‘
if ๐‘ ≥ 0,
|๐‘| =
−๐‘ if ๐‘ < 0.
Let ๐‘Ž be a positive real number. Prove the following about absolute value for every real
number ๐‘ฅ.
(a) | − ๐‘ฅ| = |๐‘ฅ|
(b) |๐‘ฅ2 | = |๐‘ฅ|2
(c) ๐‘ฅ ≤ |๐‘ฅ|
(d) |๐‘ฅ๐‘ฆ| = |๐‘ฅ| |๐‘ฆ|
(e) |๐‘ฅ| < ๐‘Ž if and only if −๐‘Ž < ๐‘ฅ < ๐‘Ž.
(f) |๐‘ฅ| > ๐‘Ž if and only if ๐‘ฅ > ๐‘Ž or ๐‘ฅ < −๐‘Ž.
24. Take ๐‘Ž and ๐‘ to be nonzero integers and prove the given propositions.
(a) ๐‘Ž divides 1 if and only if ๐‘Ž = ±1.
(b) If ๐‘Ž = ±๐‘, then |๐‘Ž| = |๐‘|.
CHAPTER 3
SET THEORY
3.1
SETS AND ELEMENTS
The development of logic that resulted in the work of Chapters 1 and 2 went through
many stages and benefited from the work of various mathematicians and logicians
through the centuries. Although modern logic can trace its roots to Descartes with
his mathesis universalis and Gottfried Leibniz’s De Arte Combinatoria (1666), the
beginnings of modern symbolic logic is generally attributed to Augustus De Morgan
[Formal Logic (1847)], George Boole [Mathematical Analysis of Logic (1847) and An
Investigation of the Laws of Thought (1847)], and Frege [Begriffsschrift (1879), Die
Grundlagen der Arithmetik (1884), and Grundgesetze der Arithmetik (1893)]. However, when it comes to set theory, it was Georg Cantor who, with his first paper, “Ueber
eine Eigenschaft des Inbegriffs aller reellen algebraischen Zahlen” (1874), and over a
decade of research, is the founder of the subject. For the next four chapters, Cantor’s
set theory will be our focus.
A set is a collection of objects known as elements. An element can be almost anything, such as numbers, functions, or lines. A set is a single object that can contain
many elements. Think of it as a box with things inside. The box is the set, and the
things are the elements. We use uppercase letters to label sets, and elements will usu117
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
118
Chapter 3 SET THEORY
ally be represented by lowercase letters. The symbol ∈ (fashioned after the Greek letter
epsilon) is used to mean “element of,” so if ๐ด is a set and ๐‘Ž is an element of ๐ด, write
∈ ๐‘Ž๐ด or, the more standard, ๐‘Ž ∈ ๐ด. The notation ๐‘Ž, ๐‘ ∈ ๐ด means ๐‘Ž ∈ ๐ด and ๐‘ ∈ ๐ด. If
๐‘ is not an element of ๐ด, write ๐‘ ∉ ๐ด. If ๐ด contains no elements, it is the empty set. It
is represented by the symbol ∅. Think of the empty set as a box with no things inside.
Rosters
Since the elements are those that distinguish one set from another, one method that
is used to write a set is to list its elements and surround them with braces. This is
called the roster method of writing a set, and the list is known as a roster. The braces
signify that a set has been defined. For example, the set of all integers between 1 and
10 inclusive is
{1, 2, 3, 4, 5, 6, 7, 8, 9, 10}.
Read this as “the set containing 1, 2, 3, 4, 5, 6, 7, 8, 9, and 10.” The set of all integers
between 1 and 10 exclusive is
{2, 3, 4, 5, 6, 7, 8, 9}.
If the roster is too long, use ellipses (… ). When there is a pattern to the elements of
the set, write down enough members so that the pattern is clear. Then use the ellipses
to represent the continuing pattern. For example, the set of all integers inclusively
between 1 and 1,000,000 can be written as
{1, 2, 3, … , 999,999, 1,000,000}.
Follow this strategy to write infinite sets as rosters. For instance, the set of even integers
can be written as
{… , −4, −2, 0, 2, 4, … }.
EXAMPLE 3.1.1
โˆ™ As a roster, { } denotes the empty set. Warning: Never write {∅} for the empty
set. This set has one element in it.
โˆ™ A set that contains exactly one element is called a singleton. Hence, the sets {1},
{๐‘“ }, and {∅} are singletons written in roster form. Also, 1 ∈ {1}, ๐‘“ ∈ {๐‘“ },
and ∅ ∈ {∅}.
โˆ™ The set of linear functions that intersect the origin with an integer slope can be
written as:
{… , −2๐‘ฅ, −๐‘ฅ, 0, ๐‘ฅ, 2๐‘ฅ, … }.
(Note: Here 0 represents the function ๐‘“ (๐‘ฅ) = 0.)
Let ๐ด and ๐ต be sets. These are equal if they contain exactly the same elements. The
notation for this is ๐ด = ๐ต. What this means is if any element is in ๐ด, it is also in ๐ต, and
Section 3.1 SETS AND ELEMENTS
119
conversely, if an element is in ๐ต, it is in ๐ด. To fully understand set equality, consider
again the analogy between sets and boxes. Suppose that we have a box containing a
carrot and a rabbit. We could describe it with the phrase the box that contains the
carrot and the rabbit. Alternately, it could be referred to as the box that contains the
orange vegetable and the furry, cotton-tailed animal with long ears. Although these
are different descriptions, they do not refer to different boxes. Similarly, the set {1, 3}
and the set containing the solutions of (๐‘ฅ − 1)(๐‘ฅ − 3) = 0 are equal because they contain
the same elements. Furthermore, the order in which the elements are listed does not
matter. The box can just as easily be described as the box with the rabbit and the carrot.
Likewise, {1, 3} = {3, 1}. Lastly, suppose that the box is described as containing the
carrot, the rabbit, and the carrot, forgetting that the carrot had already been mentioned.
This should not be confusing, for one understands that such mistakes are possible. It is
similar with sets. A repeated element does not add to the set. Hence, {1, 3} = {1, 3, 1}.
Famous Sets
Although sets can contain many different types of elements, numbers are probably the
most common for mathematics. For this reason particular important sets of numbers
have been given their own symbols.
Symbol
๐
๐™
๐
๐‘
๐‚
Name
The set of natural numbers
The set of integers
The set of rational numbers
The set of real numbers
The set of complex numbers
As rosters,
๐ = {0, 1, 2, … }
and
๐™ = {… , −2, −1, 0, 1, 2, … }.
Notice that we define the set of natural numbers to include zero and do not make a
distinction between counting numbers and whole numbers. Instead, write
๐™+ = {1, 2, 3, … }
and
๐™− = {… , −3, −2, −1}.
EXAMPLE 3.1.2
โˆ™ 10 ∈ ๐™+ , but 0 ∉ ๐™+ .
โˆ™ 4 ∈ ๐, but −5 ∉ ๐.
โˆ™ −5 ∈ ๐™, but .65 ∉ ๐™.
120
Chapter 3 SET THEORY
โˆ™ .65 ∈ ๐ and 1โˆ•2 ∈ ๐, but ๐œ‹ ∉ ๐.
โˆ™ ๐œ‹ ∈ ๐‘, but 3 − 2๐‘– ∉ ๐‘.
โˆ™ 3 − 2๐‘– ∈ ๐‚.
Of the sets mentioned above, the real numbers are probably the most familiar. It
is the set of numbers most frequently used in calculus and is often represented by a
number line. The line can be subdivided into intervals. Given two endpoints, an
interval includes all real numbers between the endpoints and possibly the endpoints
themselves. Interval notation is used to name these sets. A parenthesis next to an
endpoint means that the endpoint is not included in the set, while a bracket means that
the endpoint is included. If the endpoints are included, the interval is closed. If they
are excluded, the interval is open. If one endpoint is included and the other is not, the
interval is half-open. If the interval has only one endpoint, then the set is called a ray
and is defined using the infinity symbol (∞), with or without the negative sign.
DEFINITION 3.1.3
Let ๐‘Ž, ๐‘ ∈ ๐‘ such that ๐‘Ž < ๐‘.
closed interval
open interval
half-open interval
half-open interval
[๐‘Ž, ๐‘]
(๐‘Ž, ๐‘)
[๐‘Ž, ๐‘)
(๐‘Ž, ๐‘]
closed ray
closed ray
open ray
open ray
[๐‘Ž, ∞)
(−∞, ๐‘Ž]
(๐‘Ž, ∞)
(−∞, ๐‘Ž)
EXAMPLE 3.1.4
We can describe (4, 7) as
the interval of real numbers between 4 and 7 exclusive
and [4, 7] as
all real numbers between 4 and 7 inclusive.
There is not a straightforward way to name the half-open intervals. For (4, 7], we
can try
the set of all real numbers ๐‘ฅ such that 4 < ๐‘ฅ ≤ 7
or
the set of all real numbers greater than 4 and less than or equal to 7.
The infinity symbol does not represent a real number, so a parenthesis must be used
with it. Furthermore, the interval (−∞, ∞) can be used to denote ๐‘.
Section 3.1 SETS AND ELEMENTS
−4 −3 −2 −1
0
1
2
3
4
−4 −3 −2 −1
(a) (−1, 3]
0
1
2
3
121
4
(b) (−∞, 2]
Figure 3.1
A half-open interval and a closed ray.
EXAMPLE 3.1.5
The interval (−1, 3] contains all real numbers that are greater than −1 but less
than or equal to 3 [Figure 3.1(a)]. A common mistake is to equate (−1, 3] with
{0, 1, 2, 3}. It is important to remember that (−1, 3] includes all real numbers
between −1 and 3. Hence, this set is infinite, as is (−∞, 2). It contains all real
numbers less than 2 [Figure 3.1(b)].
EXAMPLE 3.1.6
Let ๐‘(๐‘ฅ) := ๐‘ฅ + 2 = 7. Since ๐‘(5) and there is no other real number ๐‘Ž such that
๐‘(๐‘Ž), there exists a unique ๐‘ฅ ∈ ๐‘ such that ๐‘(๐‘ฅ). However, if ๐‘Ž is an element of
๐™− or (−∞, 5), then ¬๐‘(๐‘Ž).
EXAMPLE 3.1.7
If ๐‘ž(๐‘ฅ) := ๐‘ฅ ≥ 10, then ๐‘ž(๐‘ฅ) for all ๐‘ฅ ∈ [20, 100], there exists ๐‘ฅ ∈ ๐ such that
¬๐‘ž(๐‘ฅ), and there is no element ๐‘Ž of {1, 2, 3} such that ๐‘ž(๐‘Ž).
Abstraction
When trying to write sets as rosters, we quickly discover issues with the technique.
Since the rational numbers are defined using integers, we suspect that ๐ can be written
as a roster, but when we try to begin a list, such as
1 1
2
1, , , … , 2, , … ,
2 3
3
we realize that there are complications with the pattern and are not quite sure that the
we will exhaust them all. When considering ๐‘, we know immediately that a roster is
out of the question. We conclude that we need another method.
Fix a first-order alphabet with theory symbols ๐–ฒ. Let ๐ด be a set and ๐–ฒ-formula ๐‘(๐‘ฅ)
have the property that for every ๐‘Ž,
๐‘Ž ∈ ๐ด ⇔ ๐‘(๐‘Ž).
Notice that ๐‘(๐‘ฅ) completely describes the members of ๐ด. Namely, whenever we write
๐‘Ž ∈ ๐ด, we can also write ๐‘(๐‘Ž), and, conversely, whenever we write ๐‘(๐‘Ž), we can also
write ๐‘Ž ∈ ๐ด. For example, let ๐ธ be the set of even integers. As a roster,
๐ธ = {… , −4, −2, 0, 2, 4, … }.
122
Chapter 3 SET THEORY
Let ๐‘(๐‘ฅ) denote the formula
∃๐‘›(๐‘› ∈ ๐™ ∧ ๐‘ฅ = 2๐‘›).
(3.1)
The even integers are exactly those numbers ๐‘ฅ such that ๐‘(๐‘ฅ). In particular, we have
๐‘(2) and ๐‘(−4) but not ๐‘(5). Therefore, 2 and −4 are elements of ๐ธ, but 5 is not.
DEFINITION 3.1.8
Let ๐–  be a first-order alphabet with theory symbols ๐–ฒ. Let ๐‘(๐‘ฅ) be an ๐–ฒ-formula
and ๐ด be a set. Write ๐ด = {๐‘ฅ : ๐‘(๐‘ฅ)} to mean
๐‘Ž ∈ ๐ด ⇔ ๐‘(๐‘Ž).
Using {๐‘ฅ : ๐‘(๐‘ฅ)} to identify ๐ด is called the method of abstraction.
Using Definition 3.1.8 and (3.1), write ๐ธ using abstraction as
๐ธ = {๐‘ฅ : ∃๐‘›(๐‘› ∈ ๐™ ∧ ๐‘ฅ = 2๐‘›)}
or
๐ธ = {๐‘ฅ : ๐‘(๐‘ฅ)}.
Read this as “the set of ๐‘ฅ such that ๐‘(๐‘ฅ).” Because ๐‘ฅ = 2๐‘›, it is customary to remove ๐‘ฅ
from the definition of sets like ๐ธ and write
๐ธ = {2๐‘› : ∈ ๐‘›๐™},
or
๐ธ = {2๐‘› : ๐‘› ∈ ๐™}.
Read this as “the set of all 2๐‘› such that ๐‘› is an integer.” This simplified notation is still
considered abstraction. Its form can be summarized as
{elements : condition}.
That is, what the elements look like come before the colon, and the condition that must
be satisfied to be an element of the set comes after the colon.
EXAMPLE 3.1.9
Given the quadratic equation ๐‘ฅ2 − ๐‘ฅ − 2 = 0, we know that its solutions are −1
and 2. Thus, its solution set is ๐ด = {−1, 2}. Using the method of abstraction,
this can be written as
๐ด = {๐‘ฅ : − − ⋅ ๐‘ฅ๐‘ฅ๐‘ฅ2 = 0 ∧ ∈๐‘ฅ๐‘}.
However, as we have seen, it is customary to write the formula so that it is easier
to read, so
๐ด : ๐‘ฅ2 − ๐‘ฅ − 2 = 0 ∧ ๐‘ฅ ∈ ๐‘},
Section 3.1 SETS AND ELEMENTS
123
or, using a common notation,
๐ด{๐‘ฅ ∈ ๐‘ : ๐‘ฅ2 − ๐‘ฅ − 2 = 0}.
Therefore, given an arbitrary polynomial ๐‘“ (๐‘ฅ), its solution set over the real numbers is
{๐‘ฅ ∈ ๐‘ : ๐‘“ (๐‘ฅ) = 0}.
EXAMPLE 3.1.10
Since ๐‘ฅ ∉ ∅ is always true, to write the empty set using abstraction, we use a
formula like ๐‘ฅ ≠ ๐‘ฅ or a contradiction like ๐‘ƒ ∧ ¬๐‘ƒ , where ๐‘ƒ is a propositional
form. Then,
∅ = {๐‘ฅ ∈ ๐‘ : ๐‘ฅ ≠ ๐‘ฅ} = {๐‘ฅ : ๐‘ƒ ∧ ¬๐‘ƒ }.
EXAMPLE 3.1.11
Using the natural numbers as the starting point, ๐™ and ๐ can be defined using
the abstraction method by writing
๐™ = {๐‘› : ๐‘› ∈ ๐ ∨ −๐‘› ∈ ๐}
and
๐=
{
}
๐‘Ž
: ๐‘Ž, ๐‘ ∈ ๐™ ∧ ๐‘ ≠ 0 .
๐‘
Notice the redundancy in the definition of ๐. The fraction 1โˆ•2 is named multiple
times like 2โˆ•4 or 9โˆ•18, but remember that this does not mean that the numbers
appear infinitely many times in the set. They appear only once.
EXAMPLE 3.1.12
The open intervals can be defined using abstraction.
(๐‘Ž, ๐‘) = {๐‘ฅ ∈ ๐‘ : ๐‘Ž < ๐‘ฅ < ๐‘},
(๐‘Ž, ∞) = {๐‘ฅ ∈ ๐‘ : ๐‘Ž < ๐‘ฅ},
(−∞, ๐‘) = {๐‘ฅ ∈ ๐‘ : ๐‘ฅ < ๐‘Ž}.
See Exercise 9 for the closed and half-open intervals. Also, as with ๐™+ and ๐™− ,
the superscript + or − is always used to denote the positive or negative numbers,
respectively, of a set. For example,
๐‘+ = {๐‘ฅ ∈ ๐‘ : ๐‘ฅ > 0}
and
๐‘− = {๐‘ฅ ∈ ๐‘ : ๐‘ฅ < 0}.
124
Chapter 3 SET THEORY
EXAMPLE 3.1.13
Each of the following are written using the abstraction method and, where appropriate, as a roster.
โˆ™ The set of all rational numbers with denominator equal to 3 in roster form
is
}
{
3 2 1 0 1 2 3
…,− ,− ,− , , , , ,… .
3 3 3 3 3 3 3
Using abstraction, it is
{
}
๐‘›
:๐‘›∈๐™ .
3
โˆ™ The set of all linear polynomials with integer coefficients and leading coefficient equal to 1 is
{… , ๐‘ฅ − 2, ๐‘ฅ − 1, ๐‘ฅ, ๐‘ฅ + 1, ๐‘ฅ + 2, … }.
Using abstraction it is
{๐‘ฅ + ๐‘› : ๐‘› ∈ ๐™}.
โˆ™ The set of all polynomials of degree at most 5 can be written as
{๐‘Ž5 ๐‘ฅ5 + · · · + ๐‘Ž1 ๐‘ฅ + ๐‘Ž0 : ๐‘Ž๐‘– ∈ ๐‘ ∧ ๐‘– = 0, 1, … , 5}.
Exercises
1. Determine whether the given propositions are true or false.
(a) 0 ∈ ๐
(b) 1โˆ•2 ∈ ๐™
(c) −4 ∈ ๐
(d) 4 + ๐œ‹ ∈ ๐‘
(e) 4.34534 ∈ ๐‚
(f) {1, 2} = {2, 1}
(g) {1, 2} = {1, 2, 1}
(h) [1, 2] = {1, 2}
(i) (1, 3) = {2}
(j) −1 ∈ (−∞, −1)
(k) −1 ∈ [−1, ∞)
(l) ∅ ∈ (−2, 2)
(m) ∅ ∈ ∅
(n) 0 ∈ ∅
2. Write the given sets as rosters.
(a) The set of all integers between 1 and 5 inclusive
(b) The set of all odd integers
(c) The set of all nonnegative integers
Section 3.1 SETS AND ELEMENTS
125
(d) The set of integers in the interval (−3, 7]
(e) The set of rational numbers in the interval (0, 1) that can be represented with
exactly two decimal places
3. If possible, find an element ๐‘Ž in the given set such that ๐‘Ž + 3.14 = 0.
(a) ๐™
(b) ๐‘
(c) ๐‘+
(d) ๐‘−
(e) ๐
(f) ๐‚
(g) (0, 6)
(h) (−∞, −1)
4. Determine whether the following are true or false.
(a) 1 ∈ {๐‘ฅ : ๐‘(๐‘ฅ)} when ¬๐‘(1)
(b) 7 ∈ {๐‘ฅ ∈ ๐‘ : ๐‘ฅ2 − 5๐‘ฅ − 14 = 0}
(c) 7๐‘ฅ2 − 0.5๐‘ฅ ∈ {๐‘Ž2 ๐‘ฅ2 + ๐‘Ž1 ๐‘ฅ + ๐‘Ž0 : ๐‘Ž๐‘– ∈ ๐}
(d) ๐‘ฅ๐‘ฆ ∈ {2๐‘˜ : ๐‘˜ ∈ ๐™} if ๐‘ฅ is even and ๐‘ฆ is odd.
(e) cos ๐œƒ ∈ {๐‘Ž cos ๐œƒ + ๐‘ sin ๐œƒ : ๐‘Ž, ๐‘ ∈ ๐‘}
(f) {1, 3} = {๐‘ฅ : (๐‘ฅ − 1)(๐‘ฅ − 3) = 0}
(g) {1, 3} = {๐‘ฅ : (๐‘ฅ − 1)(๐‘ฅ − 3)2 = 0}
}
{[ ๐‘Ž 0 ]
} {[ ๐‘ฅ 0 ]
(h)
0 ๐‘ฆ : ๐‘ฅ ∈ ๐‘ and ๐‘ฆ = 0
00 :๐‘Ž∈๐‘ =
5. Write the following given sets in roster form.
(a) {−3๐‘› : ๐‘› ∈ ๐™}
(b) {0 ⋅ ๐‘› : ๐‘› ∈ ๐‘}
(c) {๐‘› cos ๐‘ฅ : ๐‘› ∈ ๐™}
(d) {๐‘Ž๐‘ฅ2 + ๐‘Ž๐‘ฅ + ๐‘Ž : ๐‘Ž ∈ ๐}
{[ ๐‘› 0 ]
}
(e)
00 :๐‘›∈๐™
6. Use a formula to uniquely describe the elements in the following sets. For example,
๐‘ฅ ∈ ๐ if and only if ๐‘ฅ ∈ ๐™ ∧ ๐‘ฅ ≥ 0.
(a) (0, 1)
(b) (−3, 3]
(c) [0, ∞)
(d) ๐™+
(e) {… , −2, −1, 0, 1, 2, … }
(f) {2๐‘Ž, 4๐‘Ž, 6๐‘Ž, … }
7. Write each set using the method of abstraction.
(a) All odd integers
(b) All positive rational numbers
(c) All integer multiples of 7
(d) All integers that have a remainder of 1 when divided by 3
126
Chapter 3 SET THEORY
(e) All ordered pairs of real numbers in which the ๐‘ฅ-coordinate is positive and
the ๐‘ฆ-coordinate is negative
(f) All complex numbers whose real part is 0
(g) All closed intervals that contain ๐œ‹
(h) All open intervals that do not contain a rational number
(i) All closed rays that contain no numbers in (−∞, 3]
(j) All 2 × 2 matrices with real entries that have a diagonal sum of 0
(k) All polynomials of degree at most 3 with real coefficients
8. Write the given sets of real numbers using interval notation.
(a) The set of real numbers greater than 4
(b) The set of real numbers between −6 and −5 inclusive
(c) The set of real numbers ๐‘ฅ so that ๐‘ฅ < 5
(d) The set of real numbers ๐‘ฅ such that 10 < ๐‘ฅ ≤ 14
9. Let ๐‘Ž, ๐‘ ∈ ๐‘ with ๐‘Ž < ๐‘. Write the given intervals using the abstraction method.
(a) [๐‘Ž, ๐‘]
(b) [๐‘Ž, ∞)
(c) (−∞, ๐‘]
(d) (๐‘Ž, ๐‘]
3.2
SET OPERATIONS
We now use connectives to define the set operations. These allow us to build new sets
from given ones.
Union and Intersection
The first set operation is defined using the ∨ connective.
DEFINITION 3.2.1
The union of ๐ด and ๐ต is
๐ด ∪ ๐ต = {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∨ ๐‘ฅ ∈ ๐ต}.
The union of sets can be viewed as the combination of all elements from the sets. On
the other hand, the next set operation is defined with ∧ and can be considered as the
overlap between the given sets.
DEFINITION 3.2.2
The intersection of ๐ด and ๐ต is
๐ด ∩ ๐ต = {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∧ ๐‘ฅ ∈ ๐ต}.
Section 3.2 SET OPERATIONS
U
A
B
U
A
(a) ๐ด ∪ ๐ต
Figure 3.2
127
B
(b) ๐ด ∩ ๐ต
Venn diagrams for union and intersection.
For example, if ๐ด = {1, 2, 3, 4} and ๐ต = {3, 4, 5, 6}, then
๐ด ∪ ๐ต = {1, 2, 3, 4, 5, 6}
and
๐ด ∩ ๐ต = {3, 4}.
The operations of union and intersection can be illustrated with pictures called Venn
diagrams, named after the logician John Venn who used a variation of these drawings
in his text on symbolic logic (Venn 1894). First, assume that all elements are members
of a fixed universe ๐‘ˆ (page 97). In set theory, the universe is considered to be the
set of all possible elements in a given situation. Use circles to represent sets and a
rectangle to represent ๐‘ˆ . The space inside these shapes represent where elements might
exist. Shading is used to represent where elements might exist after applying some set
operations. The Venn diagram for union is in Figure 3.2(a) and the one for intersection
is in Figure 3.2(b). If sets have no elements in their intersection, we can use the next
definition to name them.
DEFINITION 3.2.3
The sets ๐ด and ๐ต are disjoint or mutually exclusive when ๐ด ∩ ๐ต = ∅.
The sets {1, 2, 3} and {6, 7} are disjoint. A Venn diagram for two disjoint sets is given
in Figure 3.3.
Set Difference
The next two set operations take all of the elements of one set that are not in another.
They are defined using the not connective.
DEFINITION 3.2.4
The set difference of ๐ต from ๐ด is
๐ด โงต ๐ต = {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∧ ๐‘ฅ ∉ ๐ต}.
128
Chapter 3 SET THEORY
U
A
B
Figure 3.3
A Venn diagram for disjoint sets.
The complement of ๐ด is defined as
๐ด = ๐‘ˆ โงต ๐ด = {๐‘ฅ : ๐‘ฅ ∈ ๐‘ˆ ∧ ๐‘ฅ ∉ ๐ด}.
Read ๐ด โงต ๐ต as “๐ด minus ๐ต” or “๐ด without ๐ต.” See Figure 3.4(a) for the Venn diagram
of the set difference of sets and Figure 3.4(b) for the complement of a set.
EXAMPLE 3.2.5
The following equalities use set difference.
โˆ™ ๐ = ๐™ โงต ๐™−
โˆ™ The set of irrational numbers is ๐‘ โงต ๐.
โˆ™ ๐‘ = ๐‚ โงต {๐‘Ž + ๐‘๐‘– : ๐‘Ž, ๐‘ ∈ ๐‘ ∧ ๐‘ ≠ 0}
U
A
B
(a) ๐ด โงต ๐ต
Figure 3.4
U
A
(b) ๐ด
Venn diagrams for set difference and complement.
Section 3.2 SET OPERATIONS
129
EXAMPLE 3.2.6
Let ๐‘ˆ = {1, 2, … , 10}. Use a Venn diagram to find the results of the set operations on ๐ด = {1, 2, 3, 4, 5} and ๐ต = {3, 4, 5, 6, 7, 8}. Each element will be
represented as a point and labeled with a number.
U
A
1
2
3
4
5
6
B
7
8
9
10
โˆ™ ๐ด ∪ ๐ต = {1, 2, 3, 4, 5, 6, 7, 8}
โˆ™ ๐ด ∩ ๐ต = {3, 4, 5}
โˆ™ ๐ด โงต ๐ต = {1, 2}
โˆ™ ๐ด = {6, 7, 8, 9, 10}
EXAMPLE 3.2.7
Let ๐ถ = (−4, 2), ๐ท = [−1, 3], and ๐‘ˆ = ๐‘. Use the diagram to perform the set
operations.
C∪D
−4 −3 −2 −1
0
1
2
3
4
2
3
4
C∩D
−4 −3 −2 −1
C๎œD
โˆ™ ๐ถ ∪ ๐ท = (−4, 3]
โˆ™ ๐ถ ∩ ๐ท = [−1, 2)
โˆ™ ๐ถ โงต ๐ท = (−4, −1)
โˆ™ ๐ถ = (−∞, −4] ∪ [2, ∞)
0
1
130
Chapter 3 SET THEORY
Cartesian Products
The last set operation is not related to the logic connectives as the others, but it is
nonetheless very important to mathematics. Let ๐ด and ๐ต be sets. Given elements
๐‘Ž ∈ ๐ด and ๐‘ ∈ ๐ต, we call (๐‘Ž, ๐‘) an ordered pair. In this context, ๐‘Ž and ๐‘ are called
coordinates. It is similar to the set {๐‘Ž, ๐‘} except that the order matters. The definition
is due to Kazimierz Kuratowski (1921).
DEFINITION 3.2.8
If ๐‘Ž ∈ ๐ด and ๐‘ ∈ ๐ต,
(๐‘Ž, ๐‘) = {{๐‘Ž}, {๐‘Ž, ๐‘}}.
Notice that (๐‘Ž, ๐‘) = (๐‘Ž′ , ๐‘′ ) means that
{{๐‘Ž}, {๐‘Ž, ๐‘}} = {{๐‘Ž′ }, {๐‘Ž′ , ๐‘′ }},
which implies that ๐‘Ž = ๐‘Ž′ and ๐‘ = ๐‘′ . Therefore,
(๐‘Ž, ๐‘) = (๐‘Ž′ , ๐‘′ ) if and only if ๐‘Ž = ๐‘Ž′ and ๐‘ = ๐‘′ .
The set of all ordered pairs with the first coordinate from ๐ด and the second from ๐ต
is named after René Descartes.
DEFINITION 3.2.9
The Cartesian product of ๐ด and ๐ต is
๐ด × ๐ต = {(๐‘Ž, ๐‘) : ๐‘Ž ∈ ๐ด ∧ ๐‘ ∈ ๐ต}.
The product ๐‘2 = ๐‘ × ๐‘ is the set of ordered pairs of the Cartesian plane.
EXAMPLE 3.2.10
โˆ™ Since (1, 2) ≠ (2, 1),
{1, 2} × {0, 1, 2} = {(1, 0), (1, 1), (1, 2), (2, 0), (2, 1), (2, 2)}.
Even though we have a set definition for an ordered pair, we can still visually
represent this set on a grid as in Figure 3.5(a).
โˆ™ If ๐ด = {1, 2, 7} and ๐ต = {∅, {1, 5}},
๐ด × ๐ต = {(1, ∅), (1, {1, 5}), (2, ∅), (2, {1, 5}), (7, ∅), (7, {1, 5})}.
See Figure 3.5(b).
โˆ™ For any set ๐ด, ∅ × ๐ด = ๐ด × ∅ = ∅ because ¬∃๐‘ฅ(๐‘ฅ ∈ ∅ ∧ ๐‘ฆ ∈ ๐ด).
Section 3.2 SET OPERATIONS
(1, 2)
(2, {1, 5})
2
{1, 5}
1
∅
0
1
(7, ∅)
2
1
(a) {1, 2} × {0, 1, 2}
Figure 3.5
2
7
(b) {1, 2, 7} × {∅, {1, 5}}
Two Cartesian products.
We generalize Definition 3.2.8 by defining
(๐‘Ž, ๐‘, ๐‘) = {{๐‘Ž}, {๐‘Ž, ๐‘}, {๐‘Ž, ๐‘, ๐‘}},
(๐‘Ž, ๐‘, ๐‘, ๐‘‘) = {{๐‘Ž}, {๐‘Ž, ๐‘}, {๐‘Ž, ๐‘, ๐‘}, {๐‘Ž, ๐‘, ๐‘, ๐‘‘}},
and for ๐‘› ∈ ๐,
(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 ) = {{๐‘Ž0 }, {๐‘Ž0 , ๐‘Ž1 }, … , {๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 }},
which is called an ordered n-tuple. Then,
๐ด × ๐ต × ๐ถ = {(๐‘Ž, ๐‘, ๐‘) : ๐‘Ž ∈ ๐ด ∧ ๐‘ ∈ ๐ต ∧ ๐‘ ∈ ๐ถ},
๐ด × ๐ต × ๐ถ × ๐ท = {(๐‘Ž, ๐‘, ๐‘, ๐‘‘) : ๐‘Ž ∈ ๐ด ∧ ๐‘ ∈ ๐ต ∧ ๐‘ ∈ ๐ถ ∧ ๐‘‘ ∈ ๐ท},
and
๐ด๐‘› = {(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 ) : ๐‘Ž๐‘– ∈ ๐ด ∧ ๐‘– = 0, 1, … , ๐‘› − 1}.
Specifically, ๐‘3 = ๐‘ × ๐‘ × ๐‘, and
๐‘๐‘› = ๐‘ × ๐‘ × · · · × ๐‘,
โŸโžโžโžโžโžโžโžโžโŸโžโžโžโžโžโžโžโžโŸ
๐‘› times
which is known as Cartesian n-space.
EXAMPLE 3.2.11
Let ๐ด = {1}, ๐ต = {2}, ๐ถ = {3}, and ๐ท = {4}. Then,
๐ด × ๐ต × ๐ถ × ๐ท = {(1, 2, 3, 4)}
is a singleton containing an ordered 4-tuple. Also,
(๐ด × ๐ต) ∪ (๐ถ × ๐ท) = {(1, 2), (3, 4)},
but
๐ด × (๐ต ∪ ๐ถ) × ๐ท = {(1, 2, 4), (1, 3, 4)}.
131
132
Chapter 3 SET THEORY
Order of Operations
As with the logical connectives, we need an order to make sense of expressions that
involve many operations. To do this, we note the association between the set operations
and certain logical connectives.
}
๐ด
¬
๐ดโงต๐ต
๐ด∩๐ต
∧
๐ด∪๐ต
∨
From this we derive the order for the set operations.
DEFINITION 3.2.12 [Order of Operations]
To find a set determined by set operations, read from left to right and use the
following precedence.
โˆ™ sets within parentheses (innermost first),
โˆ™ complements,
โˆ™ set differences,
โˆ™ intersections,
โˆ™ unions.
EXAMPLE 3.2.13
If the universe is taken to be {1, 2, 3, 4, 5}, then
{5} ∪ {1, 2} ∩ {2, 3} = {5} ∪ {3, 4, 5} ∩ {2, 3} = {5} ∪ {3} = {3, 5}
This set can be written using parentheses as
{5} ∪ ({1, 2} ∩ {2, 3}).
However,
({5} ∪ {1, 2}) ∩ {2, 3} = ({5} ∪ {3, 4, 5}) ∩ {2, 3} = {3, 4, 5} ∩ {2, 3} = {3},
showing that
{5} ∪ {1, 2} ∩ {2, 3} ≠ ({5} ∪ {1, 2}) ∩ {2, 3}.
Section 3.2 SET OPERATIONS
133
EXAMPLE 3.2.14
Define ๐ด = {1}, ๐ต = {2}, ๐ถ = {3}, and ๐ท = {4}. Then, we have
๐ด ∪ ๐ต ∪ ๐ถ ∪ ๐ท = {1, 2, 3, 4}.
Written with parentheses and brackets,
๐ด ∪ ๐ต ∪ ๐ถ ∪ ๐ท = ([(๐ด ∪ ๐ต) ∪ ๐ถ] ∪ ๐ท).
With the given assignments, ๐ด ∩ ๐ต ∩ ๐ถ ∩ ๐ท is empty.
Exercises
1. Each of the given propositions are false. Replace the underlined word with another
word to make the proposition true.
(a) Intersection is defined using a disjunction.
(b) Set diagrams are used to illustrate set operations.
(c) ๐‘ โงต (๐‘ โงต ๐) is the set of irrational numbers.
(d) Set difference has a higher order of precedence than complements.
(e) The complement of ๐ด is equal to ๐‘ set minus ๐ด.
(f) The intersection of two intervals is always an interval.
(g) The union of two intervals is never an interval.
(h) ๐ด × ๐ต is equal to ∅ if ๐ต does not contain ordered pairs.
2. Let ๐ด = {0, 2, 4, 6}, ๐ต = {3, 4, 5, 6}, ๐ถ = {0, 1, 2}, and ๐‘ˆ = {0, 1, … , 9, 10}.
Write the given sets in roster notation.
(a) ๐ด ∪ ๐ต
(b) ๐ด ∩ ๐ต
(c) ๐ด โงต ๐ต
(d) ๐ต โงต ๐ด
(e) ๐ด
(f) ๐ด × ๐ต
(g) ๐ด ∪ ๐ต ∩ ๐ถ
(h) ๐ด ∩ ๐ต ∪ ๐ถ
(i) ๐ด ∩ ๐ต
(j) ๐ด ∪ ๐ด ∩ ๐ต
(k) ๐ด โงต ๐ต โงต ๐ถ
(l) ๐ด ∪ ๐ต โงต ๐ด ∩ ๐ถ
3. Write each of the given sets using interval notation.
]
(a) [2, 3] ∩ (5โˆ•2, 7
(b) (−∞, 4) ∪ (−6, ∞)
(c) (−2, 4) ∩ [−6, ∞)
(d) ∅ ∪ (4, 12]
134
Chapter 3 SET THEORY
4. Identify each of the the given sets.
(a) [6, 17] ∩ [17, 32)
(b) [6, 17) ∩ [17, 32)
(c) [6, 17] ∪ (17, 32)
(d) [6, 17) ∪ (17, 32)
5. Draw Venn diagrams.
(a) ๐ด ∩ ๐ต
(b) ๐ด ∪ ๐ต
(c) ๐ด ∩ ๐ต ∪ ๐ถ
(d) (๐ด ∪ ๐ต) โงต ๐ถ
(e) ๐ด ∩ ๐ถ ∩ ๐ต
(f) ๐ด โงต ๐ต ∩ ๐ถ
6. Match each Venn diagram to as many sets as possible.
(A)
(B)
(C)
(D)
(E)
(F)
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
(i)
(j)
๐ด∪๐ต
๐ดโงต๐ต
๐ด∩๐ต
(๐ด ∪ ๐ต) โงต (๐ด ∩ ๐ต)
๐ด∩๐ต∪๐ด∩๐ถ ∪๐ต∩๐ถ
๐ด∩๐ต∪๐ดโงต๐ต
(๐ด ∪ ๐ต) ∩ (๐ด ∪ ๐ต)
[(๐ด ∪ ๐ต) ∩ ๐ถ] ∪ [(๐ด ∪ ๐ต) ∩ ๐ถ]
[๐ด ∩ ๐ต ∩ ๐ถ] ∪ [๐ด ∩ ๐ต ∩ ๐ถ]
๐ด โงต (๐ต ∪ ๐ถ) ∩ ๐ต โงต (๐ด ∪ ๐ถ) ∩ ๐ถ โงต (๐ด ∪ ๐ต)
7. The ordered pair (1, 2) is paired with the ordered pair (2, 1) using a set operation.
Write each resulting set as a roster.
(a) (1, 2) ∪ (2, 1)
(b) (1, 2) ∩ (2, 1)
(c) (1, 2) โงต (2, 1)
8. Let ๐‘(๐‘ฅ) be a formula. Prove the following.
(a) ∀๐‘ฅ[๐‘ฅ ∈ ๐ด ∪ ๐ต → ๐‘(๐‘ฅ)] ⇒ ∀๐‘ฅ[๐‘ฅ ∈ ๐ด ∩ ๐ต → ๐‘(๐‘ฅ)]
Section 3.3 SETS WITHIN SETS
135
(b) ∀๐‘ฅ[๐‘ฅ ∈ ๐ด ∪ ๐ต → ๐‘(๐‘ฅ)] ⇔ ∀๐‘ฅ[๐‘ฅ ∈ ๐ด → ๐‘(๐‘ฅ)] ∧ ∀๐‘ฅ[๐‘ฅ ∈ ๐ต → ๐‘(๐‘ฅ)]
(c) ∃๐‘ฅ[๐‘ฅ ∈ ๐ด ∩ ๐ต ∧ ๐‘(๐‘ฅ)] ⇒ ∃๐‘ฅ[๐‘ฅ ∈ ๐ด ∧ ๐‘(๐‘ฅ)] ∧ ∃๐‘ฅ[๐‘ฅ ∈ ๐ต ∧ ๐‘(๐‘ฅ)]
(d) ∃๐‘ฅ[๐‘ฅ ∈ ๐ด ∪ ๐ต ∧ ๐‘(๐‘ฅ)] ⇔ ∃๐‘ฅ[๐‘ฅ ∈ ๐ด ∧ ๐‘(๐‘ฅ)] ∨ ∃๐‘ฅ[๐‘ฅ ∈ ๐ต ∧ ๐‘(๐‘ฅ)]
9. Find a formula ๐‘(๐‘ฅ) and sets ๐ด and ๐ต to show that the following are false.
(a) ∀๐‘ฅ[๐‘ฅ ∈ ๐ด → ๐‘(๐‘ฅ)] ∨ ∀๐‘ฅ[๐‘ฅ ∈ ๐ต → ๐‘(๐‘ฅ)] ⇒ ∀๐‘ฅ[๐‘ฅ ∈ ๐ด ∪ ๐ต → ๐‘(๐‘ฅ)]
(b) ∃๐‘ฅ[๐‘ฅ ∈ ๐ด ∧ ๐‘(๐‘ฅ)] ∧ ∃๐‘ฅ[๐‘ฅ ∈ ๐ต ∧ ๐‘(๐‘ฅ)] ⇒ ∃๐‘ฅ[๐‘ฅ ∈ ๐ด ∩ ๐ต ∧ ๐‘(๐‘ฅ)]
3.3
SETS WITHIN SETS
An important relation between any two sets is when one is contained within another.
Subsets
Let ๐ด and ๐ต be sets. ๐ด is a subset of ๐ต exactly when every element of ๐ด is also an
element of ๐ต, in symbols ๐ด ⊆ ๐ต. This is represented in a Venn diagram by the circle
for ๐ด being within the circle for ๐ต [Figure 3.6(a)].
DEFINITION 3.3.1
For all sets ๐ด and ๐ต,
๐ด ⊆ ๐ต ⇔ ∀๐‘ฅ(๐‘ฅ ∈ ๐ด → ๐‘ฅ ∈ ๐ต).
If ๐ด is not a subset of ๐ต, write ๐ด * ๐ต. This is represented in a Venn diagram by ๐ด
overlapping ๐ต with a point in ๐ด but not within ๐ต [Figure 3.6(b)]. Logically, this means
๐ด * ๐ต ⇔ ¬∀๐‘ฅ(๐‘ฅ ∈ ๐ด → ๐‘ฅ ∈ ๐ต) ⇔ ∃๐‘ฅ(๐‘ฅ ∈ ๐ด ∧ ๐‘ฅ ∉ ๐ต).
Thus, to show ๐ด * ๐ต find an element in ๐ด that is not in ๐ต. For example, if we let
๐ด = {1, 2, 3} and ๐ต = {1, 2, 5}, then ๐ด * ๐ต because 3 ∈ ๐ด but 3 ∉ ๐ต.
U
B
A
U
B
A
a
(a) ๐ด ⊆ ๐ต
Figure 3.6
(b) ๐ด * ๐ต
Venn diagrams for a subset and a nonsubset.
136
Chapter 3 SET THEORY
EXAMPLE 3.3.2
The proposition for all sets ๐ด and ๐ต, ๐ด ∪ ๐ต ⊆ ๐ด is false. To see this, we must
prove that there exists ๐ด and ๐ต such that ๐ด ∪ ๐ต * ๐ด. We take ๐ด = {1} and
๐ต = {2} as our candidates. Since ๐ด∪๐ต = {1, 2} and 2 ∉ ๐ด, we have ๐ด∪๐ต * ๐ด.
Every set is the improper subset of itself. The notation ๐ด ⊂ ๐ต means ๐ด ⊆ ๐ต but
๐ด ≠ ๐ต. In this case, ๐ด is a proper subset of ๐ต.
EXAMPLE 3.3.3
โˆ™ {1, 2, 3} ⊂ {1, 2, 3, 4, 5} and {1, 2, 3} ⊆ {1, 2, 3}
โˆ™ ๐⊂๐™⊂๐⊂๐‘⊂๐‚
โˆ™ (4, 5) ⊂ (4, 5] ⊂ [4, 5]
โˆ™ {1, 2, 3} is not a subset of {1, 2, 4}.
Our study of subsets begins with a fundamental theorem. It states that the empty set
is a subset of every set.
THEOREM 3.3.4
∅ ⊆ ๐ด.
PROOF
Let ๐ด be a set. Since ๐‘ฅ ∉ ∅, we have
∀๐‘ฅ(๐‘ฅ ∉ ∅ ∨ ๐‘ฅ ∈ ๐ด) ⇔ ∀๐‘ฅ(๐‘ฅ ∈ ∅ → ๐‘ฅ ∈ ๐ด) ⇔ ∅ ⊆ ๐ด.
Proving that one set is a subset of another always means proving an implication, so
Direct Proof is the primary tool in these proofs. That is, usually to prove that a set ๐ด is
a subset of a set ๐ต, take an element ๐‘ฅ ∈ ๐ด and show ๐‘ฅ ∈ ๐ต.
EXAMPLE 3.3.5
โˆ™ Let ๐‘ฅ ∈ ๐™ โงต ๐. Then, ๐‘ฅ ∈ ๐™ but ๐‘ฅ ∉ ๐, so ๐‘ฅ ∈ ๐™ by Simp. Thus, ๐™ โงต ๐ ⊆ ๐™.
โˆ™ Let
๐ด = {๐‘ฅ : ∃๐‘› ∈ ๐™ [(๐‘ฅ − 2๐‘›)(๐‘ฅ − 2๐‘› − 2) = 0]}
and
๐ต = {2๐‘› : ๐‘› ∈ ๐™}.
Take ๐‘ฅ ∈ ๐ด. This means that
∃๐‘› ∈ ๐™ [(๐‘ฅ − 2๐‘›)(๐‘ฅ − 2๐‘› − 2) = 0] .
In other words,
(๐‘ฅ − 2๐‘›)(๐‘ฅ − 2๐‘› − 2) = 0
Section 3.3 SETS WITHIN SETS
137
for some ๐‘› ∈ ๐™. Hence,
๐‘ฅ − 2๐‘› = 0 or ๐‘ฅ − 2๐‘› − 2 = 0.
We have two cases to check. If ๐‘ฅ − 2๐‘› = 0, then ๐‘ฅ = 2๐‘›. This means ๐‘ฅ ∈ ๐ต. If
๐‘ฅ − 2๐‘› − 2 = 0, then ๐‘ฅ = 2๐‘› + 2 = 2(๐‘› + 1), which also means ๐‘ฅ ∈ ๐ต since ๐‘› + 1
is an integer. Hence, ๐ด ⊆ ๐ต.
โˆ™ Fix ๐‘Ž, ๐‘ ∈ ๐™. Let ๐‘ฅ = ๐‘›๐‘Ž, some ๐‘› ∈ ๐™. This means ๐‘ฅ = ๐‘›๐‘Ž + 0๐‘, which implies
that {๐‘›๐‘Ž : ๐‘› ∈ ๐™} ⊆ {๐‘›๐‘Ž + ๐‘š๐‘ : ๐‘›, ๐‘š ∈ ๐™}.
This next result is based only on the definitions and Inference Rules 1.2.10. As with
Section 3.2, we see the close ties between set theory and logic.
THEOREM 3.3.6
โˆ™ ๐ด ⊆ ๐ด.
โˆ™ If ๐ด ⊆ ๐ต and ๐ต ⊆ ๐ถ, then ๐ด ⊆ ๐ถ.
โˆ™ If ๐ด ⊆ ๐ต and ๐‘ฅ ∈ ๐ด, then ๐‘ฅ ∈ ๐ต.
โˆ™ If ๐ด ⊆ ๐ต and ๐‘ฅ ∉ ๐ต, then ๐‘ฅ ∉ ๐ด.
โˆ™ Let ๐ด ⊆ ๐ต and ๐ถ ⊆ ๐ท. If ๐‘ฅ ∈ ๐ด or ๐‘ฅ ∈ ๐ถ, then ๐‘ฅ ∈ ๐ต or ๐‘ฅ ∈ ๐ท.
โˆ™ Let ๐ด ⊆ ๐ต and ๐ถ ⊆ ๐ท. If ๐‘ฅ ∉ ๐ต or ๐‘ฅ ∉ ๐ท, then ๐‘ฅ ∉ ๐ด or ๐‘ฅ ∉ ๐ถ.
PROOF
Assume ๐ด ⊆ ๐ต and ๐ต ⊆ ๐ถ. By definition, ๐‘ฅ ∈ ๐ด implies ๐‘ฅ ∈ ๐ต and ๐‘ฅ ∈ ๐ต
implies ๐‘ฅ ∈ ๐ถ. Therefore, by HS, if ๐‘ฅ ∈ ๐ด, then ๐‘ฅ ∈ ๐ถ. In other words, ๐ด ⊆ ๐ถ.
The remaining parts are left to Exercise 4.
Note that it is not the case that for all sets ๐ด and ๐ต, ๐ด ⊆ ๐ต implies that ๐ต ⊆ ๐ด. To
prove this, choose ๐ด = ∅ and ๐ต = {0}. Then, ๐ด ⊆ ๐ต yet ๐ต * ๐ด because 0 ∈ ๐ต and
0 ∉ ∅.
Equality
For two sets to be equal, they must contain exactly the same elements (page 118 of
Section 3.1). This can be stated more precisely using the idea of a subset.
DEFINITION 3.3.7
๐ด = ๐ต means ๐ด ⊆ ๐ต and ๐ต ⊆ ๐ด.
To prove that two sets are equal, we show both inclusions. Let us do this to prove
๐ด ∪ ๐ต = ๐ด ∩ ๐ต. This is one of De Morgan’s laws. To prove it, we demonstrate both
๐ด∪๐ต ⊆๐ด∩๐ต
138
Chapter 3 SET THEORY
and
๐ด ∩ ๐ต ⊆ ๐ด ∪ ๐ต.
This amounts to proving a biconditional, which means we will use the rule of biconditional proof. Look at the first direction:
๐‘ฅ∈๐ด∪๐ต
¬(๐‘ฅ ∈ ๐ด ∪ ๐ต)
¬(๐‘ฅ ∈ ๐ด ∨ ๐‘ฅ ∈ ๐ต)
๐‘ฅ∉๐ด∧๐‘ฅ∉๐ต
๐‘ฅ∈๐ด∧๐‘ฅ∈๐ต
๐‘ฅ∈๐ด∩๐ต
Given
Definition of complement
Definition of union
De Morgan’s law
Definition of complement
Definition of intersection
Now read backward through those steps. Each follows logically when read in this order
because only definitions and replacement rules were used. This means that the steps
are reversible. Hence, we have a series of biconditionals, and we can use the short rule
of biconditional proof (page 108):
๐‘ฅ ∈ ๐ด ∪ ๐ต ⇔ ¬(๐‘ฅ ∈ ๐ด ∪ ๐ต)
⇔ ¬(๐‘ฅ ∈ ๐ด ∨ ๐‘ฅ ∈ ๐ต)
⇔๐‘ฅ∉๐ด∧๐‘ฅ∉๐ต
⇔๐‘ฅ∈๐ด∧๐‘ฅ∈๐ต
⇔ ๐‘ฅ ∈ ๐ด ∩ ๐ต.
Hence, ๐ด ∪ ๐ต = ๐ด ∩ ๐ต.
We must be careful when writing these types of proofs since it is easy to confuse
the notation.
โˆ™ The correct translation for ๐‘ฅ ∈ ๐ด ∩ ๐ต is ๐‘ฅ ∈ ๐ด ∧ ๐‘ฅ ∈ ๐ต. Common mistakes for
this translation include using formulas with set operations:
x∈A ∩ x∈B
Incorrect. A set should be on
either side of a set operation.
. . . and sets with connectives:
x∈A ∧ B
Correct.
Incorrect. A formula should
be present here.
Section 3.3 SETS WITHIN SETS
139
Remember that connectives connect formulas and set operations connect sets.
โˆ™ Negations also pose problems. If a complement is used, first translate using
๐‘ฅ∈๐ด⇔๐‘ฅ∈๐‘ˆ ∧๐‘ฅ∉๐ด⇔๐‘ฅ∉๐ด
and then proceed with the proof. Similarly, the formula ๐‘ฅ ∉ ๐ด∪๐ต can be written
as neither ๐‘ฅ ∉ ๐ด ∪ ๐‘ฅ ∉ ๐ต nor ๐‘ฅ ∉ ๐ด ∨ ๐‘ฅ ∉ ๐ต. Instead, use DeM:
๐‘ฅ ∉ ๐ด ∪ ๐ต ⇔ ¬(๐‘ฅ ∈ ๐ด ∪ ๐ต)
⇔ ¬(๐‘ฅ ∈ ๐ด ∨ ๐‘ฅ ∈ ๐ต)
⇔๐‘ฅ∉๐ด∧๐‘ฅ∉๐ต
⇔ ๐‘ฅ ∈ ๐ด ∩ ๐ต.
We can now prove many basic properties about set operations. Notice how the following are closely related to their corresponding replacement rules (1.3.9).
THEOREM 3.3.8
โˆ™ Associative Laws
๐ด ∩ ๐ต ∩ ๐ถ = ๐ด ∩ (๐ต ∩ ๐ถ)
๐ด ∪ ๐ต ∪ ๐ถ = ๐ด ∪ (๐ต ∪ ๐ถ)
โˆ™ Commutative Laws
๐ด∩๐ต =๐ต∩๐ด
๐ด∪๐ต =๐ต∪๐ด
โˆ™ De Morgan’s Laws
๐ด∪๐ต =๐ด∩๐ต
๐ด∩๐ต =๐ด∪๐ต
โˆ™ Distributive Laws
๐ด ∩ (๐ต ∪ ๐ถ) = ๐ด ∩ ๐ต ∪ ๐ด ∩ ๐ถ
๐ด ∪ ๐ต ∩ ๐ถ = (๐ด ∪ ๐ต) ∩ (๐ด ∪ ๐ถ)
โˆ™ Idempotent Laws
๐ด∩๐ด=๐ด
๐ด ∪ ๐ด = ๐ด.
EXAMPLE 3.3.9
We use the short rule of biconditional proof (page 108) to prove the equality
๐ด ∩ ๐ต ∩ ๐ถ = ๐ด ∩ (๐ต ∩ ๐ถ).
๐‘ฅ∈๐ด∩๐ต∩๐ถ ⇔๐‘ฅ∈๐ด∩๐ต∧๐‘ฅ∈๐ถ
⇔๐‘ฅ∈๐ด∧๐‘ฅ∈๐ต∧๐‘ฅ∈๐ถ
⇔ ๐‘ฅ ∈ ๐ด ∧ (๐‘ฅ ∈ ๐ต ∧ ๐‘ฅ ∈ ๐ถ)
⇔ ๐‘ฅ ∈ ๐ด ∧ (๐‘ฅ ∈ ๐ต ∩ ๐ถ)
⇔ ๐‘ฅ ∈ ๐ด ∩ (๐ต ∩ ๐ถ).
140
Chapter 3 SET THEORY
Another way to prove it is to use a chain of equal signs.
๐ด ∩ ๐ต ∩ ๐ถ = {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∩ ๐ต ∧ ๐‘ฅ ∈ ๐ถ}
= {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∧ ๐‘ฅ ∈ ๐ต ∧ ๐‘ฅ ∈ ๐ถ}
= {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∧ (๐‘ฅ ∈ ๐ต ∧ ๐‘ฅ ∈ ๐ถ)}
= {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∧ ๐‘ฅ ∈ ๐ต ∩ ๐ถ}
= ๐ด ∩ (๐ต ∩ ๐ถ).
We have to be careful when proving equality. If two sets are equal, there are always
proofs for both inclusions. However, the steps needed for the one implication might
not simply be the steps for the converse in reverse. The next example illustrates this.
It is always true that ๐ด ∩ ๐ต ⊆ ๐ด. However, the premise is needed to show the other
inclusion.
EXAMPLE 3.3.10
Let ๐ด ⊆ ๐ต. Prove ๐ด ∩ ๐ต = ๐ด.
โˆ™ Let ๐‘ฅ ∈ ๐ด ∩ ๐ต. This means that ๐‘ฅ ∈ ๐ด and ๐‘ฅ ∈ ๐ต. Then, ๐‘ฅ ∈ ๐ด (Simp).
โˆ™ Assume that ๐‘ฅ is an element of ๐ด. Since ๐ด ⊆ ๐ต, ๐‘ฅ is also an element of ๐ต.
Hence, ๐‘ฅ ∈ ๐ด ∩ ๐ต.
A more involved example of this uses the concept of divisibility. Let ๐‘Ž, ๐‘ ∈ ๐™, not
both equal to zero. A common divisor of ๐‘Ž and ๐‘ is ๐‘ when ๐‘ โˆฃ ๐‘Ž and ๐‘ โˆฃ ๐‘. For
example, 4 is a common divisor of 48 and 36, but it is not the largest.
DEFINITION 3.3.11
Let ๐‘Ž, ๐‘ ∈ ๐™ with ๐‘Ž and ๐‘ not both zero. The integer ๐‘” is the greatest common
divisor of ๐‘Ž and ๐‘ if ๐‘” is a common divisor of ๐‘Ž and ๐‘ and ๐‘’ ≤ ๐‘” for every common
divisor ๐‘’ ∈ ๐™. In this case, write ๐‘” = gcd(๐‘Ž, ๐‘).
For example,
12 = gcd(48, 36)
and
7 = gcd(0, 7).
Notice that it is important that at least one of the integers is not zero. The gcd(0, 0) is
undefined since all ๐‘Ž such that ๐‘Ž ≠ 0 divide 0. Further notice that the greatest common
divisor is positive.
EXAMPLE 3.3.12
Let ๐‘Ž, ๐‘ ∈ ๐™. Prove for all ๐‘› ∈ ๐™,
gcd(๐‘Ž + ๐‘›๐‘, ๐‘) = gcd(๐‘Ž, ๐‘).
Section 3.3 SETS WITHIN SETS
141
If both pairs of numbers have the same common divisors, their greatest common
divisors must be equal. So, define
๐‘† = {๐‘˜ ∈ ๐™ : ๐‘˜ โˆฃ ๐‘Ž + ๐‘›๐‘ ∧ ๐‘˜ โˆฃ ๐‘}
and
๐‘‡ = {๐‘˜ ∈ ๐™ : ๐‘˜ โˆฃ ๐‘Ž ∧ ๐‘˜ โˆฃ ๐‘}.
To show that the greatest common divisors are equal, prove ๐‘† = ๐‘‡ .
โˆ™ Let ๐‘‘ ∈ ๐‘†. Then ๐‘‘ โˆฃ ๐‘Ž + ๐‘›๐‘ and ๐‘‘ โˆฃ ๐‘. This means ๐‘Ž + ๐‘›๐‘ = ๐‘‘๐‘™ and ๐‘ = ๐‘‘๐‘˜
for some ๐‘™, ๐‘˜ ∈ ๐™. We are left to show ๐‘‘ โˆฃ ๐‘Ž. By substitution, ๐‘Ž+๐‘›๐‘‘๐‘˜ = ๐‘‘๐‘™.
Hence, ๐‘‘ ∈ ๐‘‡ because
๐‘Ž = ๐‘‘๐‘™ − ๐‘›๐‘‘๐‘˜ = ๐‘‘(๐‘™ − ๐‘›๐‘˜).
โˆ™ Now take ๐‘‘ ∈ ๐‘‡ . This means ๐‘‘ โˆฃ ๐‘Ž and ๐‘‘ โˆฃ ๐‘. Thus, there exists ๐‘™, ๐‘˜ ∈ ๐™
such that ๐‘Ž = ๐‘‘๐‘™ and ๐‘ = ๐‘‘๐‘˜. Then,
๐‘Ž + ๐‘›๐‘ = ๐‘‘๐‘™ + ๐‘›๐‘‘๐‘˜ = ๐‘‘(๐‘™ + ๐‘›๐‘˜).
Therefore, ๐‘‘ โˆฃ ๐‘Ž + ๐‘›๐‘ and ๐‘‘ ∈ ๐‘†.
As with subsets, let us now prove some results concerning the empty set and the
universe. We use two strategies.
โˆ™ Let ๐ด be a set. We know that
๐ด = ∅ if and only if ∀๐‘ฅ(๐‘ฅ ∉ ๐ด).
Therefore, to prove that ๐ด is empty, take an arbitrary ๐‘Ž and show ๐‘Ž ∉ ๐ด. This can
sometimes be done directly, but more often an indirect proof is better. That is,
assume ๐‘Ž ∈ ๐ด and derive a contradiction. Since the contradiction arose simply
by assuming ๐‘Ž ∈ ๐ด, this formula must be the problem. Hence, ๐ด can have no
elements.
โˆ™ Let ๐‘ˆ be a universe. To prove ๐ด = ๐‘ˆ , we must show that ๐ด ⊆ ๐‘ˆ and ๐‘ˆ ⊆ ๐ด.
The first subset relation is always true. To prove the second, take an arbitrary element and show that it belongs to ๐ด. This works because ๐‘ˆ contains all possible
elements.
EXAMPLE 3.3.13
โˆ™ Suppose ๐‘ฅ ∈ ๐ด ∩ ∅. Then ๐‘ฅ ∈ ∅, which is impossible. Therefore, ๐ด ∩ ∅ = ∅.
โˆ™ Certainly, ๐ด ⊆ ๐ด ∪ ∅, so to show the opposite inclusion take ๐‘ฅ ∈ ๐ด ∪ ∅. Since
๐‘ฅ ∉ ∅, it must be the case that ๐‘ฅ ∈ ๐ด. Thus, ๐ด ∪ ∅ ⊆ ๐ด, and we have that
๐ด ∪ ∅ = ๐ด.
โˆ™ From Exercise 5(b), we know ๐ด ∩ ๐‘ˆ ⊆ ๐ด, so let ๐‘ฅ ∈ ๐ด. This means that ๐‘ฅ must
also belong to the universe. Hence, ๐‘ฅ ∈ ๐ด ∩ ๐‘ˆ , so ๐ด ∩ ๐‘ˆ = ๐ด.
โˆ™ Certainly, ๐ด ∪ ๐‘ˆ ⊆ ๐‘ˆ . Moreover, by Exercise 5(c), we have the other inclusion.
Thus, ๐ด ∪ ๐‘ˆ = ๐‘ˆ .
142
Chapter 3 SET THEORY
EXAMPLE 3.3.14
To prove that a set ๐ด is not equal to the empty set, show ¬∀๐‘ฅ(๐‘ฅ ∉ ๐ด), but this is
equivalent to ∃๐‘ฅ(๐‘ฅ ∈ ๐ด). For instance, let
๐ด = {๐‘ฅ ∈ ๐‘ : ๐‘ฅ2 + 6๐‘ฅ + 5 = 0}.
We know that ๐ด is nonempty since −1 ∈ ๐ด.
EXAMPLE 3.3.15
Let
๐ด = {(๐‘Ž, ๐‘) ∈ ๐‘ × ๐‘ : ๐‘Ž + ๐‘ = 0}
and
๐ต = {(0, ๐‘) : ๐‘ ∈ ๐‘}.
These sets are not disjoint since (0, 0) is an element of both ๐ด and ๐ต. However,
๐ด ≠ ๐ต because (1, −1) ∈ ๐ด but (1, −1) ∉ ๐ต.
Let us combine the two strategies to show a relationship between ∅ and ๐‘ˆ .
THEOREM 3.3.16
For all sets ๐ด and ๐ต and universe ๐‘ˆ , the following are equivalent.
โˆ™ ๐ด⊆๐ต
โˆ™ ๐ด∪๐ต =๐‘ˆ
โˆ™ ๐ด ∩ ๐ต = ∅.
PROOF
โˆ™ Assume ๐ด ⊆ ๐ต. Suppose ๐‘ฅ ∉ ๐ด. Then, ๐‘ฅ ∈ ๐ด, which implies that ๐‘ฅ ∈ ๐ต.
Hence, for every element ๐‘ฅ, we have that ๐‘ฅ ∈ ๐ด or ๐‘ฅ ∈ ๐ต, and we conclude that
๐ด ∪ ๐ต = ๐‘ˆ.
โˆ™ Suppose ๐ด ∪ ๐ต = ๐‘ˆ . In order to obtain a contradiction, take ๐‘ฅ ∈ ๐ด ∩ ๐ต.
Then, ๐‘ฅ ∈ ๐ต. Since ๐‘ฅ ∈ ๐ด, the supposition also gives ๐‘ฅ ∈ ๐ต, a contradiction.
Therefore, ๐ด ∩ ๐ต = ∅.
โˆ™ Let ๐ด ∩ ๐ต = ∅. Assume ๐‘ฅ ∈ ๐ด. By hypothesis, ๐‘ฅ cannot be a member of ๐ต,
otherwise the intersection would be nonempty. Hence, ๐‘ฅ ∈ ๐ต.
The following theorem is a generalization of the corresponding result concerning
subsets. The proof could have been written using the short rule of biconditional proof
or by appealing to Lemma 3.3.6 (Exercise 21).
THEOREM 3.3.17
If ๐ด = ๐ต and ๐ต = ๐ถ, then ๐ด = ๐ถ.
Section 3.3 SETS WITHIN SETS
143
PROOF
Assume ๐ด = ๐ต and ๐ต = ๐ถ. This means ๐ด ⊆ ๐ต, ๐ต ⊆ ๐ด, ๐ต ⊆ ๐ถ, and ๐ถ ⊆ ๐ต.
We must show that ๐ด = ๐ถ.
โˆ™ Let ๐‘ฅ ∈ ๐ด. By hypothesis, ๐‘ฅ is then an element of ๐ต, which implies that
๐‘ฅ ∈ ๐ถ.
โˆ™ Let ๐‘ฅ ∈ ๐ถ. Then, ๐‘ฅ ∈ ๐ต, from which ๐‘ฅ ∈ ๐ด follows.
The last result of the section involves the Cartesian product. The first part is illustrated in Figure 3.7. The sets ๐ต and ๐ถ are illustrated along the vertical axis and ๐ด is
illustrated along the horizontal axis. The Cartesian products are represented as boxes.
The other parts of the theorem can be similarly visualized.
THEOREM 3.3.18
โˆ™ ๐ด × (๐ต ∩ ๐ถ) = (๐ด × ๐ต) ∩ (๐ด × ๐ถ).
โˆ™ ๐ด × (๐ต ∪ ๐ถ) = (๐ด × ๐ต) ∪ (๐ด × ๐ถ).
โˆ™ ๐ด × (๐ต โงต ๐ถ) = (๐ด × ๐ต) โงต (๐ด × ๐ถ).
โˆ™ (๐ด × ๐ต) ∩ (๐ถ × ๐ท) = (๐ด ∩ ๐ถ) × (๐ต ∩ ๐ท).
PROOF
We prove the first equation. The last three are left to Exercise 19. Take three sets
๐ด, ๐ต, and ๐ถ. Then,
(๐‘Ž, ๐‘) ∈ ๐ด × (๐ต ∩ ๐ถ) ⇔ ๐‘Ž ∈ ๐ด ∧ ๐‘ ∈ ๐ต ∩ ๐ถ
⇔๐‘Ž∈๐ด∧๐‘∈๐ต∧๐‘Ž∈๐ด∧๐‘∈๐ถ
⇔ (๐‘Ž, ๐‘) ∈ ๐ด × ๐ต ∧ (๐‘Ž, ๐‘) ∈ ๐ด × ๐ถ
⇔ (๐‘Ž, ๐‘) ∈ (๐ด × ๐ต) ∩ (๐ด × ๐ถ).
A×B
B
A × (B ∩ C)
C
A×C
A
Figure 3.7
๐ด × (๐ต ∩ ๐ถ) = (๐ด × ๐ต) ∩ (๐ด × ๐ถ).
144
Chapter 3 SET THEORY
Inspired by the last result, we might try proving that (๐ด × ๐ต) ∪ (๐ถ × ๐ท) is always
equal to (๐ด ∪ ๐ถ) × (๐ต ∪ ๐ท), but no such proof exists. To show this, take ๐ด = ๐ต = {1}
and ๐ถ = ๐ท = {2}. Then
(๐ด ∪ ๐ถ) × (๐ต ∪ ๐ท) = {1, 2} × {1, 2} = {(1, 1), (1, 2), (2, 1), (2, 2)}.
but
(๐ด × ๐ต) ∪ (๐ถ × ๐ท) = {(1, 1)} ∪ {(2, 2)} = {(1, 1), (2, 2)},
Hence, (๐ด ∪ ๐ถ) × (๐ต ∪ ๐ท) * (๐ด × ๐ต) ∪ (๐ถ × ๐ท). Notice, however, that the opposite
inclusion is always true (Exercise 3.3.8).
Exercises
1. Answer true or false.
(a) ∅ ∈ ∅
(b) ∅ ⊆ {1}
(c) 1 ∈ ๐™
(d) 1 ⊆ ๐™
(e) 1 ∈ ∅
(f) {1} ⊆ ∅
(g) 0 ∈ ∅
(h) {1} ∈ ๐™
(i) ∅ ⊆ ∅
2. Answer true or false. For each false proposition, find one element that is in the first
set but is not in the second.
(a) ๐™+ ⊆ ๐‚
(b) ๐+ ⊆ ๐™+
(c) ๐ โงต ๐‘ ⊆ ๐™
(d) ๐‘ โงต ๐ ⊆ ๐™
(e) ๐™ ∩ (−1, 1) ⊆ ๐
(f) (0, 1) ⊆ ๐+
(g) (0, 1) ⊆ {0, 1, 2}
(h) (0, 1) ⊆ (0, 1]
3. Prove.
(a) {๐‘ฅ ∈ ๐‘ : ๐‘ฅ2 − 3๐‘ฅ + 2 = 0} ⊆ ๐
(b) (0, 1) ⊆ [0, 1]
(c) [0, 1] * (0, 1)
(d) ๐™ × ๐™ * ๐™ × ๐
(e) (0, 1) ∩ ๐ * [0, 1] ∩ ๐™
(f) ๐‘ ⊆ ๐‚
(g) {๐‘๐‘– : ๐‘ ∈ ๐‘} ⊆ ๐‚
(h) ๐‚ * ๐‘
Section 3.3 SETS WITHIN SETS
145
4. Prove the remaining parts of Theorem 3.3.6.
5. Prove.
(a) ๐ด ⊆ ๐‘ˆ
(b) ๐ด ∩ ๐ต ⊆ ๐ด
(c) ๐ด ⊆ ๐ด ∪ ๐ต
(d) ๐ด โงต ๐ต ⊆ ๐ด
(e) If ๐ด ⊆ ๐ต, then ๐ด ∪ ๐ถ ⊆ ๐ต ∪ ๐ถ.
(f) If ๐ด ⊆ ๐ต, then ๐ด ∩ ๐ถ ⊆ ๐ต ∩ ๐ถ.
(g) If ๐ด ⊆ ๐ต, then ๐ถ โงต ๐ต ⊆ ๐ถ โงต ๐ด.
(h) If ๐ด ≠ ∅, then ๐ด * ๐ด.
(i) If ๐ด ⊆ ๐ต, then ๐ต ⊆ ๐ด.
(j) If ๐ด ⊆ ๐ต, then {1} × ๐ด ⊆ {1} × ๐ต.
(k) If ๐ด ⊆ ๐ถ and ๐ต ⊆ ๐ท, then ๐ด × ๐ต ⊆ ๐ถ × ๐ท.
(l) If ๐ด ⊆ ๐ถ and ๐ต ⊆ ๐ท, then ๐ถ × ๐ท ⊆ ๐ด × ๐ต.
6. Prove that ๐ด ⊆ ๐ต if and only if ๐ต ⊆ ๐ด.
7. Show that the given proposition is false:
for all sets ๐ด and ๐ต, if ๐ด ∩ ๐ต ≠ ∅, then ๐ด * ๐ด ∩ ๐ต.
8. Prove: (๐ด × ๐ต) ∪ (๐ถ × ๐ท) ⊆ (๐ด ∪ ๐ถ) × (๐ต ∪ ๐ท).
9. Take ๐‘Ž, ๐‘, ๐‘ ∈ ๐. Let ๐ด = {๐‘› ∈ ๐ : ๐‘› โˆฃ ๐‘Ž} and ๐ถ = {๐‘› ∈ ๐ : ๐‘› โˆฃ ๐‘}. Suppose ๐‘Ž โˆฃ ๐‘
and ๐‘ โˆฃ ๐‘. Prove ๐ด ⊆ ๐ถ.
10. Prove.
(a) ๐ต ⊆ ๐ด and ๐ถ ⊆ ๐ด if and only if ๐ต ∪ ๐ถ ⊆ ๐ด.
(b) ๐ด ⊆ ๐ต and ๐ด ⊆ ๐ถ if and only if ๐ด ⊆ ๐ต ∩ ๐ถ.
11. Prove the unproven parts of Theorem 3.3.8.
12. Prove each equality.
(a) ∅ = ๐‘ˆ
(b) ๐‘ˆ = ∅
(c) ๐ด ∩ ๐ด = ∅
(d) ๐ด ∪ ๐ด = ๐‘ˆ
(e)
(f)
(g)
(h)
(i)
(j)
(k)
๐ด=๐ด
๐ดโงต๐ด=∅
๐ดโงต∅=๐ด
๐ด∩๐ต =๐ตโงต๐ด
๐ดโงต๐ต =๐ด∩๐ต
๐ด∪๐ต∩๐ต =∅
๐ด∩๐ตโงต๐ด=∅
146
Chapter 3 SET THEORY
13. Sketch a Venn diagram for each problem and then write a proof.
(a) ๐ด = ๐ด ∩ ๐ต ∪ ๐ด ∩ ๐ต
(b) ๐ด ∪ ๐ต = ๐ด ∪ ๐ด ∩ ๐ต
(c) ๐ด โงต (๐ต โงต ๐ถ) = ๐ด ∩ (๐ต ∪ ๐ถ)
(d) ๐ด โงต (๐ด ∩ ๐ต) = ๐ด โงต ๐ต
(e) ๐ด ∩ ๐ต ∪ ๐ด ∩ ๐ต ∪ ๐ด ∩ ๐ต = ๐ด ∪ ๐ต
(f) (๐ด ∪ ๐ต) โงต ๐ถ = ๐ด โงต ๐ถ ∪ ๐ต โงต ๐ถ
(g) ๐ด โงต (๐ต โงต ๐ถ) = ๐ด โงต ๐ต ∪ ๐ด ∩ ๐ถ
(h) ๐ด โงต ๐ต โงต ๐ถ = ๐ด โงต (๐ต ∪ ๐ถ)
14. Prove.
(a) If ๐ด ⊆ ๐ต, then ๐ด โงต ๐ต = ∅.
(b) If ๐ด ⊆ ∅, then ๐ด = ∅.
(c) Let ๐‘ˆ be a universe. If ๐‘ˆ ⊆ ๐ด, then ๐ด = ๐‘ˆ .
(d) If ๐ด ⊆ ๐ต, then ๐ต โงต (๐ต โงต ๐ด) = ๐ด.
(e) ๐ด × ๐ต = ∅ if and only if ๐ด = ∅ or
15. Let ๐‘Ž, ๐‘, ๐‘š ∈ ๐™ and define ๐ด = {๐‘Ž + ๐‘š๐‘˜ : ๐‘˜ ∈ ๐™} and ๐ต = {๐‘Ž + ๐‘š(๐‘ + ๐‘˜) : ๐‘˜ ∈ ๐™}.
Show ๐ด = ๐ต.
16. Prove ๐ด = ๐ต, where
๐ด=
and
๐ต=
{[ ๐‘Ž ๐‘ ]
๐‘ ๐‘‘
{[ ๐‘Ž 0 ]
0๐‘
}
: ๐‘Ž + ๐‘ = 0 ∧ ๐‘Ž, ๐‘ ∈ ๐‘
}
: ๐‘Ž = −๐‘‘ ∧ ๐‘2 + ๐‘ 2 = 0 ∧ ๐‘Ž, ๐‘, ๐‘, ๐‘‘ ∈ ๐‘ .
17. Prove.
(a) ๐ ≠ ๐™
(b) ๐‚ ≠ ๐‘
(c) {0} × ๐™ ≠ ๐™
(d) ๐‘ × ๐™ ≠ ๐™ × ๐‘
(e) If ๐ด = {๐‘Ž๐‘ฅ3 + ๐‘ : ๐‘Ž, ๐‘ ∈ ๐‘} and ๐ต = {๐‘ฅ3 + ๐‘ : ๐‘ ∈ ๐‘}, then ๐ด ≠ ๐ต.
(f) If ๐ด = {๐‘Ž๐‘ฅ3 + ๐‘ : ๐‘Ž, ๐‘ ∈ ๐™} and ๐ต = {๐‘Ž๐‘ฅ3 + ๐‘ : ๐‘Ž, ๐‘ ∈ ๐‚}, then ๐ด ≠ ๐ต.
(g) ๐ด ≠ ๐ต, where
{[ ]
}
๐ด = ๐‘Ž0 0๐‘ : ๐‘Ž, ๐‘ ∈ ๐‘
and
๐ต=
{[ ๐‘Ž ๐‘ ]
๐‘ ๐‘‘
}
: ๐‘Ž, ๐‘, ๐‘, ๐‘‘ ∈ ๐‘ .
18. Find ๐ด and ๐ต to illustrate the given inequalities.
(a) ๐ด โงต ๐ต ≠ ๐ต โงต ๐ด
(b) (๐ด × ๐ต) × ๐ถ ≠ ๐ด × (๐ต × ๐ถ)
(c) ๐ด × ๐ต ≠ ๐ต × ๐ด
19. For the remaining parts of Theorem 3.3.18, draw diagrams as in Figure 3.7 and
prove the results.
Section 3.3 SETS WITHIN SETS
147
20. Is it possible for ๐ด = ๐ด? Explain.
21. Prove Theorem 3.3.17 by first using the short rule of biconditional proof and then
by directly appealing to Theorem 3.3.6.
22. Prove that the following are equivalent.
โˆ™ ๐ด⊆๐ต
โˆ™ ๐ด∪๐ต =๐ต
โˆ™ ๐ดโงต๐ต =∅
โˆ™ ๐ด∩๐ต =๐ด
23. Prove that the following are equivalent.
โˆ™ ๐ด∩๐ต =∅
โˆ™ ๐ดโงต๐ต =∅
โˆ™ ๐ด⊆๐ต
24. Find an example of sets ๐ด, ๐ต, and ๐ถ such that ๐ด ∩ ๐ต = ๐ด ∩ ๐ถ but ๐ต ≠ ๐ถ.
25. Does ๐ด ∪ ๐ต = ๐ด ∪ ๐ถ imply ๐ต = ๐ถ for all sets ๐ด and ๐ต? Explain.
26. Prove.
(a) If ๐ด ∪ ๐ต ⊆ ๐ด ∩ ๐ต, then ๐ด = ๐ต.
(b) If ๐ด ∩ ๐ต = ๐ด ∩ ๐ถ and ๐ด ∪ ๐ต = ๐ด ∪ ๐ถ, then ๐ต = ๐ถ.
27. Prove that there is a cancellation law with the Cartesian product. Namely, if ๐ด ≠ ∅
and ๐ด × ๐ต = ๐ด × ๐ถ, then ๐ต = ๐ถ.
28. When does ๐ด × ๐ต = ๐ถ × ๐ท imply that ๐ด = ๐ถ and ๐ต = ๐ท?
29. Prove.
(a) If ๐ด ∪ ๐ต ≠ ∅, then ๐ด ≠ ∅ or ๐ต ≠ ∅.
(b) If ๐ด ∩ ๐ต ≠ ∅, then ๐ด ≠ ∅ and ๐ต ≠ ∅.
30. Find the greatest common divisors of each pair.
(a) 12 and 18
(b) 3 and 9
(c) 14 and 0
(d) 7 and 15
31. Let ๐‘Ž be a positive integer. Find the following and prove the result.
(a) gcd(๐‘Ž, ๐‘Ž + 1)
(b) gcd(๐‘Ž, 2๐‘Ž)
(c) gcd(๐‘Ž, ๐‘Ž2 )
(d) gcd(๐‘Ž, 0)
148
Chapter 3 SET THEORY
3.4
FAMILIES OF SETS
The elements of a set can be sets themselves. We call such a collection a family of sets
and often use capital script letters to name them. For example, let
โ„ฐ = {{1, 2, 3}, {2, 3, 4}, {3, 4, 5}}.
(3.2)
The set โ„ฐ has three elements: {1, 2, 3}, {2, 3, 4}, and {3, 4, 5}.
EXAMPLE 3.4.1
โˆ™ {1, 2, 3} ∈ {{1, 2, 3}, {1, 4, 9}}
โˆ™ 1 ∉ {{1, 2, 3}, {1, 4, 9}}
โˆ™ {1, 2, 3} * {{1, 2, 3}, {1, 4, 9}}
โˆ™ {{1, 2, 3}} ⊆ {{1, 2, 3}, {1, 4, 9}}.
EXAMPLE 3.4.2
โˆ™ ∅ ⊆ {∅, {∅}} by Theorem 3.3.4.
โˆ™ {∅} ⊆ {∅, {∅}} because ∅ ∈ {∅, {∅}}.
โˆ™ {{∅}} ⊆ {∅, {∅}} because {∅} ∈ {∅, {∅}}.
Families of sets can have infinitely many elements. For example, let
โ„ฑ = {[๐‘›, ๐‘› + 1] : ๐‘› ∈ ๐™}.
(3.3)
In roster notation,
โ„ฑ = {… , [−2, −1] , [−1, 0] , [0, 1] , [1, 2] , … }.
Notice that in this case, abstraction is more convenient. For each integer ๐‘›, the closed
interval [๐‘›, ๐‘› + 1] is in โ„ฑ . The set ๐™ plays the role of an index set, a set whose only
purpose is to enumerate the elements of the family. Each element of an index set is
called an index. If we let ๐ผ = ๐™ and ๐ด๐‘– = [๐‘–, ๐‘– + 1], the family can be written as
โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ}.
To write the family โ„ฐ (3.2) using an index set, let ๐ผ = {0, 1, 2} and define
๐ด0 = {1, 2, 3},
๐ด1 = {2, 3, 4},
๐ด2 = {3, 4, 5}.
Then, the family illustrated in Figure 3.8 is
โ„ฐ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} = {๐ด1 , ๐ด2 , ๐ด3 }.
Section 3.4 FAMILIES OF SETS
A0
1
โ„ฐ
2
4
A2
Figure 3.8
3
149
2
3
3
4
A1
5
The family of sets โ„ฐ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} with ๐ผ = {1, 2, 3}.
There is no reason why ๐ผ must be {0, 1, 2}. Any three-element set will do. The order
in which the sets are defined is also irrelevant. For instance, we could have defined
๐ผ = {๐‘ค, ๐œ‹, 99} and
๐ด๐‘ค = {3, 4, 5},
๐ด๐œ‹ = {2, 3, 4},
๐ด99 = {1, 2, 3}.
The goal is to have each set in the family referenced or indexed by at least one element
of the set. We will still have โ„ฐ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} with a similar diagram (Figure 3.9).
EXAMPLE 3.4.3
Write โ„ฑ (3.3) using ๐ as the index set instead of ๐™. Use the even natural numbers to index the intervals with a nonnegative integer left-hand endpoint. The
odd natural numbers will index the intervals with a negative integer left-hand
endpoint. To do this, define the sets ๐ต๐‘– as follows:
A99
Aw
Figure 3.9
1
3
โ„ฐ
2
4
2
3
3
4
Aπ
5
The family of sets โ„ฐ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} with ๐ผ = {๐‘ค, ๐œ‹, 99}.
150
Chapter 3 SET THEORY
๐ต5
๐ต3
๐ต1
๐ต0
๐ต2
๐ต4
โ‹ฎ
= [−3, −2]
= [−2, −1]
= [−1, 0]
= [0, 1]
= [1, 2]
= [2, 3]
โ‹ฎ
Use 2๐‘› + 1 to represent the odd natural numbers and 2๐‘› to represent the even
natural numbers (๐‘› ∈ ๐). Then,
๐ต2(2)+1
โ‹ฎ
= [−2 − 1, −2]
๐ต2(1)+1 = [−1 − 1, −1]
๐ต2(0)+1 = [−0 − 1, −0]
and
๐ต2(0) = [0, 0 + 1]
๐ต2(1) = [1, 1 + 1]
๐ต2(2) = [2, 2 + 1]
โ‹ฎ
Therefore, define for all natural numbers ๐‘›,
๐ต2๐‘›+1 = [−๐‘› − 1, −๐‘›]
and
๐ต2๐‘› = [๐‘›, ๐‘› + 1] .
We have indexed the elements of โ„ฑ as
๐ต0
๐ต1
๐ต2
๐ต3
๐ต4
๐ต5
= [0, 1]
= [−1, 0]
= [1, 2]
= [−2, −1]
= [2, 3]
= [−3, −2]
โ‹ฎ
So under this definition, โ„ฑ = {๐ต๐‘– : ๐‘– ∈ ๐}.
Section 3.4 FAMILIES OF SETS
151
Power Set
There is a natural way to form a family of sets. Take a set ๐ด. The collection of all
subsets of ๐ด is called the power set of ๐ด. It is represented by P(๐ด).
DEFINITION 3.4.4
For any set ๐ด,
P(๐ด) = {๐ต : ๐ต ⊆ ๐ด}.
Notice that ∅ ∈ P(๐ด) by Theorem 3.3.4.
EXAMPLE 3.4.5
โˆ™ P({1, 2, 3}) = {∅, {1}, {2}, {3}, {1, 2}, {1, 3}, {2, 3}, {1, 2, 3}}
โˆ™ P({∅, {∅}}) = {∅, {∅}, {{∅}}, {∅, {∅}}}
โˆ™ P(๐) = {∅, {0}, {1}, … , {0, 1}, {0, 2}, … }.
Consider ๐ด = {2, 6} and ๐ต = {2, 6, 10}, so ๐ด ⊆ ๐ต. Examining the power sets of each,
we find that
P(๐ด) = {∅, {2}, {6}, {2, 6}}
and
P(๐ต) = {∅, {2}, {6}, {10}, {2, 6}, {2, 10}, {6, 10}, {2, 6, 10}}.
Hence, P(๐ด) ⊆ P(๐ต). This result is generalized in the next lemma.
LEMMA 3.4.6
๐ด ⊆ ๐ต if and only if P(๐ด) ⊆ P(๐ต).
PROOF
โˆ™ Let ๐ด ⊆ ๐ต. Assume ๐‘‹ ∈ P(๐ด). Then, ๐‘‹ ⊆ ๐ด, which gives ๐‘‹ ⊆ ๐ต by
Theorem 3.3.6. Hence, ๐‘‹ ∈ P(๐ต).
โˆ™ Assume P(๐ด) ⊆ P(๐ต). Let ๐‘ฅ ∈ ๐ด. In other words, {๐‘ฅ} ⊆ ๐ด, but this means that
{๐‘ฅ} ∈ P(๐ด). Hence, {๐‘ฅ} ∈ P(๐ต) by hypothesis, so ๐‘ฅ ∈ ๐ต.
The definition of set equality and Lemma 3.4.6 are used to prove the next theorem. Its
proof is left to Exercise 9.
THEOREM 3.4.7
๐ด = ๐ต if and only if P(๐ด) = P(๐ต).
Union and Intersection
We now generalize the set operations. Define the union of a family of sets to be the set
of all elements that belong to some member of the family. This union is denoted by the
same notation as in Definition 3.2.1 and can be defined using the abstraction method.
152
Chapter 3 SET THEORY
DEFINITION 3.4.8
Let โ„ฑ be a family of sets. Define
โ‹ƒ
โ„ฑ = {๐‘ฅ : ∃๐ด(๐ด ∈ โ„ฑ ∧ ๐‘ฅ ∈ ๐ด)}.
If the family is indexed so that โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ}, define
โ‹ƒ
๐ด๐‘– = {๐‘ฅ : ∃๐‘–(๐‘– ∈ ๐ผ ∧ ๐‘ฅ ∈ ๐ด๐‘– )}.
๐‘–∈๐ผ
Observe that
โ‹ƒ
โ„ฑ =
โ‹ƒ
๐‘–∈๐ผ
๐ด๐‘– [Exercise 16(a)].
We generalize the notion of intersection similarly. The intersection of a family of
sets is the set of all elements that belong to each member of the family.
DEFINITION 3.4.9
Let โ„ฑ be a family of sets.
โ‹‚
โ„ฑ = {๐‘ฅ : ∀๐ด(๐ด ∈ โ„ฑ → ๐‘ฅ ∈ ๐ด)}.
If the family is indexed so that โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ}, define
โ‹‚
๐ด๐‘– = {๐‘ฅ : ∀๐‘–(๐‘– ∈ ๐ผ → ๐‘ฅ ∈ ๐ด๐‘– )}.
๐‘–∈๐ผ
Observe that
โ‹‚
โ„ฑ =
โ‹‚
๐‘–∈๐ผ
๐ด๐‘– [Exercise 16(b)].
Furthermore, notice that both definitions are indeed generalizations of the operations
of Section 3.2 because as noted in Exercise 17,
โ‹ƒ
{๐ด, ๐ต} = ๐ด ∪ ๐ต,
and
โ‹‚
{๐ด, ๐ต} = ๐ด ∩ ๐ต.
EXAMPLE 3.4.10
Define โ„ฐ = {[๐‘›, ๐‘› + 1] : ๐‘› ∈ ๐™}.
โˆ™ When all of these intervals are combined, the result is the real line. This
means that
โ‹ƒ
โ„ฐ = ๐‘.
โˆ™ There is not one element that is common to all of the intervals. Hence,
โ‹‚
โ„ฐ = ∅.
The next example illustrates how to write the union or intersection of a family of sets
as a roster.
Section 3.4 FAMILIES OF SETS
153
EXAMPLE 3.4.11
Let โ„ฑ = {{1, 2, 3}, {2, 3, 4}, {3, 4, 5}}.
โ‹ƒ
โˆ™ Since 1 is in the first set of โ„ฑ , 1 ∈
โ„ฑ . The others can be explained
similarly, so
โ‹ƒ
โ„ฑ = {1, 2, 3, 4, 5}.
Notice that mechanically this amounts to removing the braces around the
sets of the family and setting the union to the resulting set:
Remove braces
{{1, 2, 3}, {2, 3, 4}, {3, 4, 5}}
โ„ฑ
{1, 2, 3, 2, 3, 4, 3, 4, 5}
{1, 2, 3, 4, 5}
โ‹ƒ
โ„ฑ
โˆ™ The generalized intersection
โ‹‚ is simply the overlap of all of the sets. Hence,
3 is the only element of โ„ฑ . That is,
โ‹‚
โ„ฑ = {3},
and this is illustrated by the following diagram:
{{1, 2, 3 }, {2, 3 , 4}, { 3 , 4, 5}}
โ„ฑ
{3}
โ‹ƒโ„ฑ
EXAMPLE 3.4.12
Since a family of sets can be empty, we must
โ‹ƒ be able to take the
โ‹ƒunion and intersection of the empty set. To prove that ∅ = ∅, take ๐‘ฅ ∈ ∅. This means
that there exists ๐ด ∈ ∅ such that ๐‘ฅ ∈ ๐ด, but this is impossible. We leave the fact
โ‹‚
that ∅ is equal to the universe to Exercise 24.
The next theorem generalizes the Distributive Laws. Exercise 2.4.10 plays an important role in its proof. It allows us to move the quantifier.
154
Chapter 3 SET THEORY
THEOREM 3.4.13
Let ๐ด๐‘– (๐‘– ∈ ๐ผ) and ๐ต be sets.
โ‹‚
โ‹‚
๐ด๐‘– =
(๐ต ∪ ๐ด๐‘– )
โˆ™ ๐ต∪
๐‘–∈๐ผ
๐‘–∈๐ผ
โˆ™ ๐ต∩
โ‹ƒ
โ‹ƒ
๐ด๐‘– =
(๐ต ∩ ๐ด๐‘– ).
๐‘–∈๐ผ
๐‘–∈๐ผ
PROOF
We leave the second part to Exercise 19. The first part is demonstrated by the
following biconditional proof:
โ‹‚
โ‹‚
๐‘ฅ∈๐ต∪
๐ด๐‘– ⇔ ๐‘ฅ ∈ ๐ต ∨ ๐‘ฅ ∈
๐ด๐‘–
๐‘–∈๐ผ
๐‘–∈๐ผ
⇔ ๐‘ฅ ∈ ๐ต ∨ ∀๐‘–(๐‘– ∈ ๐ผ → ๐‘ฅ ∈ ๐ด๐‘– )
⇔ ๐‘ฅ ∈ ๐ต ∨ ∀๐‘–(๐‘– ∉ ๐ผ ∨ ๐‘ฅ ∈ ๐ด๐‘– )
⇔ ∀๐‘–(๐‘ฅ ∈ ๐ต ∨ ๐‘– ∉ ๐ผ ∨ ๐‘ฅ ∈ ๐ด๐‘– )
⇔ ∀๐‘–(๐‘– ∉ ๐ผ ∨ ๐‘ฅ ∈ ๐ต ∨ ๐‘ฅ ∈ ๐ด๐‘– )
⇔ ∀๐‘–(๐‘– ∈ ๐ผ → ๐‘ฅ ∈ ๐ต ∨ ๐‘ฅ ∈ ๐ด๐‘– )
⇔ ∀๐‘–(๐‘– ∈ ๐ผ → ๐‘ฅ ∈ ๐ต ∪ ๐ด๐‘– )
โ‹‚
⇔๐‘ฅ∈
(๐ต ∪ ๐ด๐‘– ).
๐‘–∈๐ผ
To understand the next result, consider the following example. Choose the universe
to be {1, 2, 3, 4, 5} and perform some set operations on the family
{{1, 2, 3}, {2, 3, 4}, {2, 3, 5}}.
First,
โ‹‚
{{1, 2, 3}, {2, 3, 4}, {2, 3, 5}} = {2, 3} = {1, 4, 5},
and second,
โ‹ƒ
โ‹ƒ
{{1, 2, 3}, {2, 3, 4}, {2, 3, 5}} =
{{4, 5}, {1, 5}, {1, 4}} = {1, 4, 5}.
This leads us to the next generalization of De Morgan’s laws. Its proof is left to Exercise 23.
THEOREM 3.4.14
Let {๐ด๐‘– : ๐‘– ∈ ๐ผ} be a family of sets.
โ‹‚
โ‹ƒ
โˆ™
๐ด๐‘– =
๐ด๐‘–
โˆ™
๐‘–∈๐ผ
๐‘–∈๐ผ
โ‹ƒ
โ‹‚
๐‘–∈๐ผ
๐ด๐‘– =
๐‘–∈๐ผ
๐ด๐‘– .
Section 3.4 FAMILIES OF SETS
155
Disjoint and Pairwise Disjoint
What it means for two sets to be disjoint was defined in Section 3.2 (Definition 3.2.3).
The next definition generalizes that notion to families of sets. Because a family can
have more than two elements, it is appropriate to expand the concept of disjointness.
DEFINITION 3.4.15
Let โ„ฑ be a family of sets.
โˆ™ โ„ฑ is disjoint when
โ‹‚
โ„ฑ = ∅.
โˆ™ โ„ฑ is pairwise disjoint when for all ๐ด, ๐ต ∈ โ„ฑ , if ๐ด ≠ ๐ต, then ๐ด ∩ ๐ต = ∅.
Observe that {{1, 2}, {3, 4}, {5, 6}} is both disjoint and pairwise disjoint because its
elements have no common members.
EXAMPLE 3.4.16
Let ๐ด be a set. We see that
โ‹ƒ
P(๐ด) = {๐‘ฅ : ∃๐ต[๐ต ∈ P(๐ด) ∧ ๐‘ฅ ∈ ๐ต]} = ๐ด.
Although P(๐ด) is not a pairwise disjoint family of sets, it is a disjoint because
∅ ∈ P(๐ด).
If the family is indexed, we can use another test to determine if it is pairwise disjoint.
Let โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} be a family of sets. If for all ๐‘–, ๐‘— ∈ ๐ผ,
๐‘– ≠ ๐‘— implies ๐ด๐‘– ∩ ๐ด๐‘— = ∅,
(3.4)
then โ„ฑ is pairwise disjoint. To prove this, let โ„ฑ be a family that satisfies (3.4). Take
๐ด๐‘– , ๐ด๐‘— ∈ โ„ฑ for some ๐‘–, ๐‘— ∈ ๐ผ and assume ๐ด๐‘– ≠ ๐ด๐‘— . Therefore, ๐‘– ≠ ๐‘—, for otherwise
they would be the equal. Hence, ๐ด๐‘– ∩ ๐ด๐‘— = ∅ by Condition 3.4.
The next result illustrates the relationship between these two terms. One must be
careful, though. The converse is false (Exercise 14).
THEOREM 3.4.17
Let โ„ฑ be a family of sets with at least two elements. If โ„ฑ is pairwise disjoint,
then โ„ฑ is disjoint.
PROOF
Assume โ„ฑ is pairwise disjoint. Since โ„ฑ contains
at least two sets, let ๐ด, ๐ต ∈ โ„ฑ
โ‹‚
such that ๐ด ≠ ๐ต. Then, using Exercise 27, โ„ฑ ⊆ ๐ด ∩ ๐ต = ∅.
Exercises
1. Let ๐ผ = {1, 2, 3, 4, 5}, ๐ด1 = {1, 2}, ๐ด2 = {3, 4}, ๐ด3 = {1, 4}, ๐ด4 = {3, 4}, and
๐ด5 = {1, 3}. Write the given families of sets as a rosters.
156
Chapter 3 SET THEORY
(a)
(b)
(c)
(d)
(e)
(f)
{๐ด๐‘–
{๐ด๐‘–
{๐ด๐‘–
{๐ด๐‘–
{๐ด๐‘–
{๐ด๐‘–
: ๐‘– ∈ ๐ผ}
: ๐‘– ∈ {2, 4}}
: ๐‘– = 1}
: ๐‘– = 1, 2}
: ๐‘– ∈ ∅}
: ๐‘– ∈ ๐ด5 }
2. Answer true or false.
(a) 1 ∈ {{1}, {2}, {1, 2}}
(b) {1} ∈ {{1}, {2}, {1, 2}}
(c) {1} ⊆ {{1}, {2}, {1, 2}}
(d) {1, 2} ∈ {{{1, 2}, {3, 4}}, {1, 2}}
(e) {1, 2} ⊆ {{{1, 2}, {3, 4}}, {1, 2}}
(f) {3, 4} ∈ {{{1, 2}, {3, 4}}, {1, 2}}
(g) {3, 4} ⊆ {{{1, 2}, {3, 4}}, {1, 2}}
(h) ∅ ∈ {{{1, 2}, {3, 4}}, {1, 2}}
(i) ∅ ⊆ {{{1, 2}, {3, 4}}, {1, 2}}
(j) {∅} ∈ {∅, {∅, {∅}}}
(k) {∅} ⊆ {∅, {∅, {∅}}}
(l) ∅ ∈ {∅, {∅, {∅}}}
(m) ∅ ⊆ {∅, {∅, {∅}}}
(n) {∅} ⊆ ∅
(o) {∅} ⊆ {∅, {∅}}
(p) {∅} ⊆ {{∅, {∅}}}
(q) {{∅}} ⊆ {∅, {∅}}
(r) {1} ∈ P(๐™)
(s) {1} ⊆ P(๐™)
(t) ∅ ∈ P(∅)
(u) {∅} ∈ P(∅)
(v) {∅} ⊆ P(∅)
3. Find sets such that๐ผ ∩ ๐ฝ = ∅ but {๐ด๐‘– : ๐‘– ∈ ๐ผ} and {๐ด๐‘— : ๐‘— ∈ ๐ฝ } are not disjoint.
4. Let {๐ด๐‘– : ๐‘– ∈ ๐พ} be a family of sets and let ๐ผ and ๐ฝ be subsets of ๐พ. Define
โ„ฐ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} and โ„ฑ = {๐ด๐‘— : ๐‘— ∈ ๐ฝ } and prove the following.
(a) If ๐ผ ⊆ ๐ฝ , then โ„ฐ ⊆ โ„ฑ .
(b) โ„ฐ ∪ โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ ∪ ๐ฝ }
(c) {๐ด๐‘– : ๐‘– ∈ ๐ผ ∩ ๐ฝ } ⊆ โ„ฐ ∩ โ„ฑ
5. Using the same notation as in the previous problem, find a family {๐ด๐‘– : ๐‘– ∈ ๐พ} and
subsets ๐ผ and ๐ฝ of ๐พ such that:
(a) โ„ฐ ∩ โ„ฑ * {๐ด๐‘– : ๐‘– ∈ ๐ผ ∩ ๐ฝ }
(b) {๐ด๐‘– : ๐‘– ∈ ๐ผ โงต ๐ฝ } * โ„ฐ โงต โ„ฑ
6. Show {๐ด๐‘– : ๐‘– ∈ ๐ผ} = ∅ if and only if ๐ผ = ∅.
Section 3.4 FAMILIES OF SETS
157
7. Let ๐ด be a finite set. How many elements are in P(๐ด) if ๐ด has ๐‘› elements? Explain.
8. Find the given power sets.
(a) P({1, 2})
(b) P(P({1, 2}))
(c) P(∅)
(d) P(P(∅))
(e) P({{∅}})
(f) P({∅, {∅}, {∅, {∅}}})
9. Prove Theorem 3.4.7.
10. For each of the given equalities, prove or show false. If one is false, prove any true
inclusion.
(a) P(๐ด ∪ ๐ต) = P(๐ด) ∪ P(๐ต)
(b) P(๐ด ∩ ๐ต) = P(๐ด) ∩ P(๐ต)
(c) P(๐ด โงต ๐ต) = P(๐ด) โงต P(๐ต)
(d) P(๐ด × ๐ต) = P(๐ด) × P(๐ต)
11. Prove P(๐ด) ⊆ P(๐ต) implies ๐ด ⊆ ๐ต by using the fact that ๐ด ∈ P(๐ด).
12. Write the following sets in roster form.
โ‹ƒ
(a)
{{1, 2}, {1, 2}, {1, 3}, {1, 4}}
โ‹‚
(b)
{{1, 2}, {1, 2}, {1, 3}, {1, 4}}
โ‹‚
(c)
P(∅)
โ‹ƒ
(d)
P(∅)
โ‹ƒโ‹ƒ
(e)
{{{1}}, {{1, 2}}, {{1, 3}}, {{1, 4}}}
โ‹ƒโ‹‚
(f)
{{{1}}, {{1, 2}}, {{1, 3}}, {{1, 4}}}
โ‹‚โ‹ƒ
(g)
{{{1}}, {{1, 2}}, {{1, 3}}, {{1, 4}}}
โ‹‚โ‹‚
(h)
{{{1}}, {{1, 2}}, {{1, 3}}, {{1, 4}}}
โ‹ƒโ‹ƒ
(i)
∅
โ‹‚โ‹ƒ
(j)
∅
13. Draw Venn diagrams for a disjoint family of sets that is not pairwise disjoint and
for a pairwise disjoint family of sets.
14. Show by example that a disjoint family of sets might not be pairwise disjoint.
15. Given a family of sets {๐ด๐‘– : ๐‘– ∈ ๐ผ}, find a family โ„ฌ = {๐ต๐‘– : ๐‘– ∈ ๐ผ} such that
{๐ด๐‘– × ๐ต๐‘– : ๐‘– ∈ ๐ผ} is pairwise disjoint.
16. Let โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} be a family of sets. Prove the given equations.
โ‹ƒ
โ‹ƒ
(a)
โ„ฑ = ๐‘–∈๐ผ ๐ด๐‘–
โ‹‚
โ‹‚
(b)
โ„ฑ = ๐‘–∈๐ผ ๐ด๐‘–
โ‹ƒ
โ‹‚
17. Prove for any sets ๐ด and ๐ต, {๐ด, ๐ต} = ๐ด ∪ ๐ต and {๐ด, ๐ต} = ๐ด ∩ ๐ต.
โ‹ƒ
โ‹‚
18. Let ๐’ž = {(0, ๐‘›) : ๐‘› ∈ ๐™+ }. Prove that ๐’ž = (0, ∞) and ๐’ž = (0, 1).
158
Chapter 3 SET THEORY
19. Prove the second part of Theorem 3.4.13.
20. Prove Theorem 3.4.17 indirectly.
21. Is Theorem 3.4.17 still true if the family of sets โ„ฑ has at most one element? Explain.
22. Let {๐ด๐‘– : ๐‘– ∈ ๐ผ} be a family of sets and prove the following.
โ‹ƒ
(a) If ๐ต ⊆ ๐ด๐‘– for some ๐‘– ∈ ๐ผ, then ๐ต ⊆ ๐‘–∈๐ผ ๐ด๐‘– .
โ‹‚
(b) If ๐ด๐‘– ⊆ ๐ต for all ๐‘– ∈ ๐ผ, ๐‘–∈๐ผ ๐ด๐‘– ⊆ ๐ต.
โ‹‚
(c) If ๐ต ⊆ ๐‘–∈๐ผ ๐ด๐‘– , then ๐ต ⊆ ๐ด๐‘– for all ๐‘– ∈ ๐ผ.
23. Prove Theorem 3.4.14.
โ‹‚
24. Show ∅ = ๐‘ˆ where ๐‘ˆ is a universe.
โ‹‚
25. Let โ„ฑ be a family of sets such that ∅ ∈ โ„ฑ . Prove โ„ฑ = ∅.
โ‹ƒ
โ‹ƒ
26. Find families of sets โ„ฐ and โ„ฑ so that โ„ฐ =
โ„ฑ but โ„ฐ ≠ โ„ฑ . Can this be
repeated by replacing union with intersection?
โ‹‚
โ‹ƒ
27. Let โ„ฑ be a family of sets, and let ๐ด ∈ โ„ฑ . Prove โ„ฑ ⊆ ๐ด ⊆ โ„ฑ .
28. Let โ„ฐ and โ„ฑ be families of sets. Show the following.
โ‹ƒ
(a)
{โ„ฑ } = โ„ฑ
โ‹‚
(b)
{โ„ฑ } = โ„ฑ
โ‹ƒ
โ‹ƒ
(c) If โ„ฐ ⊆ โ„ฑ , then โ„ฐ ⊆ โ„ฑ .
โ‹‚
โ‹‚
(d) If โ„ฐ ⊆ โ„ฑ , then โ„ฑ ⊆ โ„ฐ .
โ‹ƒ
โ‹ƒ
โ‹ƒ
(e)
(โ„ฐ ∪ โ„ฑ ) = โ„ฐ ∪ โ„ฑ
โ‹‚
โ‹‚
โ‹‚
(f)
(โ„ฐ ∪ โ„ฑ ) = โ„ฐ ∩ โ„ฑ
29. Find families of sets โ„ฐ and โ„ฑ that make the following false.
โ‹‚
โ‹‚
โ‹‚
(a)
(โ„ฐ ∩ โ„ฑ ) = โ„ฐ ∩ โ„ฑ
โ‹ƒ
โ‹ƒ
โ‹ƒ
(b)
(โ„ฐ ∩ โ„ฑ ) = โ„ฐ ∩ โ„ฑ
30. Let โ„ฑ be a family of sets. For each of the given equalities, prove true or show that
it is false by finding a counter-example.
โ‹ƒ
(a)
P(โ„ฑ ) = โ„ฑ
โ‹‚
(b)
P(โ„ฑ ) = โ„ฑ
โ‹ƒ
(c) P( โ„ฑ ) = โ„ฑ
โ‹‚
(d) P( โ„ฑ ) = โ„ฑ
31. Prove these alternate forms of De Morgan’s laws.
โ‹ƒ
โ‹‚
(a) ๐ด โงต ๐‘–∈๐ผ ๐ต๐‘– = ๐‘–∈๐ผ ๐ด โงต ๐ต๐‘–
โ‹‚
โ‹ƒ
(b) ๐ด โงต ๐‘–∈๐ผ ๐ต๐‘– = ๐‘–∈๐ผ ๐ด โงต ๐ต๐‘–
32. Assume that the universe ๐‘ˆ contains only sets such that for all ๐ด ∈ ๐‘ˆ , ๐ด ⊆ ๐‘ˆ and
P(๐ด) ∈ ๐‘ˆ . Prove the following equalities.
โ‹ƒ
(a)
๐‘ˆ =๐‘ˆ
โ‹‚
(b)
๐‘ˆ =∅
Section 3.4 FAMILIES OF SETS
33. Let ๐ผ๐‘› = [๐‘›, ๐‘› + 1] and ๐ฝ๐‘› = [๐‘›, ๐‘› + 1], ๐‘› ∈ ๐™. Define
โ„ฑ = {๐ผ๐‘› × ๐ฝ๐‘š : ๐‘›, ๐‘š ∈ ๐™}.
Show
โ‹ƒ
โ„ฑ = ๐‘2 and
โ‹‚
โ„ฑ = ∅. Is โ„ฑ pairwise disjoint?
159
CHAPTER 4
RELATIONS AND FUNCTIONS
4.1
RELATIONS
A relation is an association between objects. A book on a table is an example of the
relation of one object being on another. It is especially common to speak of relations
among people. For example, one person could be the niece of another. In mathematics,
there are many relations such as equals and less-than that describe associations between
numbers. To formalize this idea, we make the next definition.
DEFINITION 4.1.1
A set ๐‘… is an (n-ary) relation if there exist sets ๐ด0 , ๐ด1 , … , ๐ด๐‘›−1 such that
๐‘… ⊆ ๐ด0 × ๐ด1 × · · · × ๐ด๐‘›−1 .
In particular, ๐‘… is a unary relation if ๐‘› = 1 and a binary relation if ๐‘› = 2. If
๐‘… ⊆ ๐ด × ๐ด for some set ๐ด, then ๐‘… is a relation on ๐ด and we write (๐ด, ๐‘…).
The relation on can be represented as a subset of the Cartesian product of the set of all
books and the set of all tables. We could then write (dictionary, desk) to mean that the
161
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
162
Chapter 4 RELATIONS AND FUNCTIONS
dictionary is on the desk. Similarly, the set {(2, 4), (7, 3), (0, 0)} is a relation because
it is a subset of ๐™ × ๐™. The ordered pair (2, 4) means that 2 is related to 4. Likewise,
๐‘ × ๐ is a relation where every real number is related to every rational number, and
according to Definition 4.1.1, the empty set is also a relation because ∅ = ∅ × ∅.
EXAMPLE 4.1.2
For any set ๐ด, define
๐ผ๐ด = {(๐‘Ž, ๐‘Ž) : ๐‘Ž ∈ ๐ด}.
Call this set the identity on ๐ด. In particular, the identity on ๐‘ is
๐ผ๐‘ = {(๐‘ฅ, ๐‘ฅ) : ๐‘ฅ ∈ ๐‘},
and the identity on ๐™ is
๐ผ๐™ = {(๐‘ฅ, ๐‘ฅ) : ๐‘ฅ ∈ ๐™}.
Notice that ∅ is the identity on ∅.
EXAMPLE 4.1.3
The less-than relation on ๐™ is defined as
๐ฟ = {(๐‘Ž, ๐‘) : ๐‘Ž, ๐‘ ∈ ๐™ ∧ ๐‘Ž < ๐‘}.
Another approach is to use membership in the set of positive integers as our condition. That is,
๐ฟ = {(๐‘Ž, ๐‘) : ๐‘Ž, ๐‘ ∈ ๐™ ∧ ๐‘ − ๐‘Ž ∈ ๐™+ }.
Hence, (4, 7) ∈ ๐ฟ because 7 − 4 ∈ ๐™+ . See Exercise 1 for another definition of
๐ฟ.
When a relation ๐‘… ⊆ ๐ด × ๐ต is defined, all elements in ๐ด or ๐ต might not be used.
For this reason, it is important to identify the sets that comprise all possible values for
the two coordinates of the relation.
DEFINITION 4.1.4
Let ๐‘… ⊆ ๐ด × ๐ต. The domain of ๐‘… is the set
dom(๐‘…) = {๐‘ฅ ∈ ๐ด : ∃๐‘ฆ(๐‘ฆ ∈ ๐ต ∧ (๐‘ฅ, ๐‘ฆ) ∈ ๐‘…)},
and the range of ๐‘… is the set
ran(๐‘…) = {๐‘ฆ ∈ ๐ต : ∃๐‘ฅ(๐‘ฅ ∈ ๐ด ∧ (๐‘ฅ, ๐‘ฆ) ∈ ๐‘…)}.
EXAMPLE 4.1.5
If ๐‘… = {(1, 3), (2, 4), (2, 5)}, then dom(๐‘…) = {1, 2} and ran(๐‘…) = {3, 4, 5}. We
represent this in Figure 4.1 where ๐ด and ๐ต are sets so that {1, 2} ⊆ ๐ด and
{3, 4, 5} ⊆ ๐ต. The ordered pair (1, 3) is denoted by the arrow pointing from
1 to 3. Also, both dom(๐‘…) and ran(๐‘…) are shaded.
Section 4.1 RELATIONS
163
R
A
1
2
Figure 4.1
3
B
4
5
๐‘… = {(1, 3), (2, 4), (2, 5)}.
EXAMPLE 4.1.6
Let ๐‘† = {(๐‘ฅ, ๐‘ฆ) : |๐‘ฅ| = ๐‘ฆ ∧ ๐‘ฅ, ๐‘ฆ ∈ ๐‘}. Notice that both (2, 2) and (−2, 2) are
elements of ๐‘†. Furthermore, dom(๐‘†) = ๐‘ and ran(๐‘†) = [0, ∞).
The domain and range of a relation can be the same set as in the next two examples.
EXAMPLE 4.1.7
If ๐‘… = {(0, 1), (0, 2), (1, 0), (2, 0)}, then ๐‘… is a relation on the set {0, 1, 2} with
dom(๐‘…) = ran(๐‘…) = {0, 1, 2}.
EXAMPLE 4.1.8
For the relation
}
{
√
๐‘† = (๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 ≥ 1 ,
both the domain and range equal ๐‘. First, note that it is clear by the definition
of ๐‘† that both dom(๐‘†) and ran(๐‘†) are subsets of ๐‘. For the other inclusion, take
๐‘ฅ ∈ ๐‘. Since (๐‘ฅ, 1) ∈ ๐‘†, ๐‘ฅ ∈ dom(๐‘†), and since (1, ๐‘ฆ) ∈ ๐‘†, ๐‘ฆ ∈ ran(๐‘†).
Composition
Given the relations ๐‘… and ๐‘†, let us define a new relation. Suppose that (๐‘Ž, ๐‘) ∈ ๐‘† and
(๐‘, ๐‘) ∈ ๐‘…. Therefore, ๐‘Ž is related to ๐‘ through ๐‘. The new relation will contain the
ordered pair (๐‘Ž, ๐‘) to represent this relationship.
DEFINITION 4.1.9
Let ๐‘… ⊆ ๐ด × ๐ต and ๐‘† ⊆ ๐ต × ๐ถ. The composition of ๐‘† and ๐‘… is the subset of
๐ด × ๐ถ defined as
๐‘† โ—ฆ ๐‘… = {(๐‘ฅ, ๐‘ง) : ∃๐‘ฆ(๐‘ฆ ∈ ๐ต ∧ (๐‘ฅ, ๐‘ฆ) ∈ ๐‘… ∧ (๐‘ฆ, ๐‘ง) ∈ ๐‘†)}.
As illustrated in Figure 4.2, the reason that (๐‘Ž, ๐‘) ∈ ๐‘… โ—ฆ ๐‘† is because (๐‘Ž, ๐‘) ∈ ๐‘† and
(๐‘, ๐‘) ∈ ๐‘…. That is, ๐‘Ž is related to ๐‘ via ๐‘… โ—ฆ ๐‘† because ๐‘Ž is related to ๐‘ via ๐‘† and then ๐‘
164
Chapter 4 RELATIONS AND FUNCTIONS
A
C
Rโ—ฆS
c
a
S
R
B
b
Figure 4.2
A composition of relations.
is related to ๐‘ via ๐‘…. The composition can be viewed as the direct path from ๐‘Ž to ๐‘ that
does not require the intermediary ๐‘.
EXAMPLE 4.1.10
To clarify the definition, let
๐‘… = {(2, 4), (1, 3), (2, 5)}
and
๐‘† = {(0, 1), (1, 0), (0, 2), (2, 0)}.
We have that ๐‘… โ—ฆ ๐‘† = {(0, 3), (0, 4), (0, 5)}. Notice that (0, 3) ∈ ๐‘… โ—ฆ ๐‘† because
(0, 1) ∈ ๐‘† and (1, 3) ∈ ๐‘…. However, ๐‘† โ—ฆ ๐‘… is empty since ran(๐‘…) and dom(๐‘†)
are disjoint.
EXAMPLE 4.1.11
Define
and
๐‘… = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 = 1}
๐‘† = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฆ = ๐‘ฅ + 1}.
Notice that ๐‘… is the unit circle and ๐‘† is the line with slope of 1 and ๐‘ฆ-intercept
of (0, 1). Let us find ๐‘… โ—ฆ ๐‘†.
๐‘… โ—ฆ ๐‘† = {(๐‘ฅ, ๐‘ง) ∈ ๐‘2 : ∃๐‘ฆ(๐‘ฆ ∈ ๐‘ ∧ (๐‘ฅ, ๐‘ฆ) ∈ ๐‘† ∧ (๐‘ฆ, ๐‘ง) ∈ ๐‘…)}
= {(๐‘ฅ, ๐‘ง) ∈ ๐‘2 : ∃๐‘ฆ(๐‘ฆ ∈ ๐‘ ∧ ๐‘ฆ = ๐‘ฅ + 1 ∧ ๐‘ฆ2 + ๐‘ง2 = 1)}
= {(๐‘ฅ, ๐‘ง) ∈ ๐‘2 : (๐‘ฅ + 1)2 + ๐‘ง2 = 1}.
Section 4.1 RELATIONS
165
Therefore, ๐‘… โ—ฆ ๐‘† is the circle with center (−1, 0) and radius 1.
Example 4.1.10 shows that it is possible that ๐‘† โ—ฆ ๐‘… ≠ ๐‘… โ—ฆ ๐‘†. However, we can
change the order of the composition.
THEOREM 4.1.12
If ๐‘… ⊆ ๐ด × ๐ต, ๐‘† ⊆ ๐ต × ๐ถ, and ๐‘‡ ⊆ ๐ถ × ๐ท, then ๐‘‡ โ—ฆ (๐‘† โ—ฆ ๐‘…) = (๐‘‡ โ—ฆ ๐‘†) โ—ฆ ๐‘….
PROOF
Assume that ๐‘… ⊆ ๐ด × ๐ต, ๐‘† ⊆ ๐ต × ๐ถ, and ๐‘‡ ⊆ ๐ถ × ๐ท. Then,
(๐‘Ž, ๐‘‘) ∈ ๐‘‡ โ—ฆ (๐‘† โ—ฆ ๐‘…)
⇔ ∃๐‘(๐‘ ∈ ๐ถ ∧ (๐‘Ž, ๐‘) ∈ ๐‘† โ—ฆ ๐‘… ∧ (๐‘, ๐‘‘) ∈ ๐‘‡ )
⇔ ∃๐‘(๐‘ ∈ ๐ถ ∧ ∃๐‘(๐‘ ∈ ๐ต ∧ (๐‘Ž, ๐‘) ∈ ๐‘… ∧ (๐‘, ๐‘) ∈ ๐‘†) ∧ (๐‘, ๐‘‘) ∈ ๐‘‡ )
⇔ ∃๐‘∃๐‘(๐‘ ∈ ๐ถ ∧ ๐‘ ∈ ๐ต ∧ (๐‘Ž, ๐‘) ∈ ๐‘… ∧ (๐‘, ๐‘) ∈ ๐‘† ∧ (๐‘, ๐‘‘) ∈ ๐‘‡ )
⇔ ∃๐‘∃๐‘(๐‘ ∈ ๐ต ∧ (๐‘Ž, ๐‘) ∈ ๐‘… ∧ ๐‘ ∈ ๐ถ ∧ (๐‘, ๐‘) ∈ ๐‘† ∧ (๐‘, ๐‘‘) ∈ ๐‘‡ )
⇔ ∃๐‘(๐‘ ∈ ๐ต ∧ (๐‘Ž, ๐‘) ∈ ๐‘… ∧ ∃๐‘[๐‘ ∈ ๐ถ ∧ (๐‘, ๐‘) ∈ ๐‘† ∧ (๐‘, ๐‘‘) ∈ ๐‘‡ ])
⇔ ∃๐‘(๐‘ ∈ ๐ต ∧ (๐‘Ž, ๐‘) ∈ ๐‘… ∧ (๐‘, ๐‘‘) ∈ ๐‘‡ โ—ฆ ๐‘†)
⇔ (๐‘Ž, ๐‘‘) ∈ (๐‘‡ โ—ฆ ๐‘†) โ—ฆ ๐‘….
Inverses
Let ๐‘… ⊆ ๐ด × ๐ต. We know that ๐‘… โ—ฆ ๐ผ๐ด = ๐‘… and ๐ผ๐ต โ—ฆ ๐‘… = ๐‘… [Exercise 10(a)]. If we
want ๐ผ๐ด = ๐ผ๐ต , we need ๐ด = ๐ต so that ๐‘… is a relation on ๐ด. Then,
๐‘… โ—ฆ ๐ผ๐ด = ๐ผ๐ด โ—ฆ ๐‘… = ๐‘….
For example, if we again define ๐‘… = {(2, 4), (1, 3), (2, 5)} and view it as a relation on
๐™, then ๐‘… composed on either side by ๐ผ๐™ yields ๐‘…. To illustrate, consider the ordered
pair (1, 3). It is an element of ๐‘… โ—ฆ ๐ผ๐™ because
(1, 1) ∈ ๐ผ๐™ ∧ (1, 3) ∈ ๐‘…,
and it is also an element of ๐ผ๐™ โ—ฆ ๐‘… because
(1, 3) ∈ ๐‘… ∧ (3, 3) ∈ ๐ผ๐™ .
Notice that not every identity relation will have this property. Using the same definition
of ๐‘… as above,
๐‘… โ—ฆ ๐ผ{1,2} = ๐‘…,
but
๐ผ{1,2} โ—ฆ ๐‘… = ∅.
Now let us change the problem. Given a relation ๐‘… on ๐ด, can we find a relation ๐‘†
on ๐ด such that ๐‘… โ—ฆ ๐‘† = ๐‘† โ—ฆ ๐‘… = ๐ผ๐ด ? The next definition is used to try to answer this
question.
166
Chapter 4 RELATIONS AND FUNCTIONS
DEFINITION 4.1.13
Let ๐‘… be a binary relation. The inverse of ๐‘… is the set
๐‘…−1 = {(๐‘ฆ, ๐‘ฅ) : (๐‘ฅ, ๐‘ฆ) ∈ ๐‘…}.
For a relation ๐‘†, we say that ๐‘… and ๐‘† are inverse relations if ๐‘…−1 = ๐‘†.
EXAMPLE 4.1.14
Let ๐ฟ be the less-than relation on ๐‘ (Example 4.1.3). Then,
๐ฟ−1
= {(๐‘ฆ, ๐‘ฅ) ∈ ๐‘2 : (๐‘ฅ, ๐‘ฆ) ∈ ๐ฟ}
= {(๐‘ฆ, ๐‘ฅ) ∈ ๐‘2 : ๐‘ฅ < ๐‘ฆ}
= {(๐‘ฆ, ๐‘ฅ) ∈ ๐‘2 : ๐‘ฆ > ๐‘ฅ}.
This shows that less-than and greater-than are inverse relations.
We now check whether ๐‘… โ—ฆ ๐‘…−1 = ๐‘…−1 โ—ฆ ๐‘… = ๐ผ๐ด for any relation ๐‘… on ๐ด. Consider
๐‘… = {(2, 1), (4, 3)}, which is a relation on {1, 2, 3, 4}. Then,
๐‘…−1 = {(1, 2), (3, 4)},
and we see that composing does not yield the identity on {1, 2, 3, 4} because
๐‘… โ—ฆ ๐‘…−1 = {(1, 1), (3, 3)} = ๐ผ{1,3}
and
๐‘…−1 โ—ฆ ๐‘… = {(2, 2), (4, 4)} = ๐ผ{2,4} .
The situation is worse when we define ๐‘† = {(2, 1), (2, 3), (4, 3)}. In this case, we have
that
๐‘† −1 = {(1, 2), (3, 2), (3, 4)}
but
๐‘† โ—ฆ ๐‘† −1 = {(1, 1), (1, 3), (3, 1), (3, 3)}
and
๐‘† −1 โ—ฆ ๐‘† = {(2, 2), (2, 4), (4, 4)}.
Neither of these compositions leads to an identity, but at least we have that
{(1, 1), (3, 3)} ⊆ ๐‘† โ—ฆ ๐‘† −1
and
{(2, 2), (4, 4)} ⊆ ๐‘† −1 โ—ฆ ๐‘†.
This can be generalized.
THEOREM 4.1.15
๐ผran(๐‘…) ⊆ ๐‘… โ—ฆ ๐‘…−1 and ๐ผdom(๐‘…) ⊆ ๐‘…−1 โ—ฆ ๐‘… for any binary relation ๐‘….
Section 4.1 RELATIONS
167
PROOF
The first inclusion is proved in Exercise 12. To see the second inclusion, let
๐‘ฅ ∈ dom(๐‘…). By definition, there exists ๐‘ฆ ∈ ran(๐‘…) so that (๐‘ฅ, ๐‘ฆ) ∈ ๐‘…. Hence,
(๐‘ฆ, ๐‘ฅ) ∈ ๐‘…−1 , which implies that (๐‘ฅ, ๐‘ฅ) ∈ ๐‘…−1 โ—ฆ ๐‘….
Exercises
1. Let ๐ฟ ⊆ ๐™ × ๐™ be the less-than relation as defined in Example 4.1.3. Prove that
๐ฟ = {(๐‘Ž, ๐‘) ∈ ๐™ × ๐™ : ๐‘Ž − ๐‘ ∈ ๐™− }.
2. Find the domain and range of the given relations.
(a) {(0, 1), (2, 3), (4, 5), (6, 7)}
(b) {((๐‘Ž, ๐‘), 1), ((๐‘Ž, ๐‘), 2), ((๐‘Ž, ๐‘‘), 3)}
(c) ๐‘ × ๐™
(d) ∅ × ∅
(e) ๐ × ∅
(f) {(๐‘ฅ, ๐‘ฆ) : ๐‘ฅ, ๐‘ฆ ∈ [0, 1] ∧ ๐‘ฅ < ๐‘ฆ}
(g) {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฆ = 3}
(h) {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฆ = |๐‘ฅ|}
(i) {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 +√
๐‘ฆ2 = 4}
2
(j) {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘ : ๐‘ฆ ≤ ๐‘ฅ ∧ ๐‘ฅ ≥ 0}
(k) {(๐‘“ , ๐‘”) : ∃๐‘Ž ∈ ๐‘(๐‘“ (๐‘ฅ) = ๐‘’๐‘ฅ ∧ ๐‘”(๐‘ฅ) = ๐‘Ž๐‘ฅ)}
(l) {((๐‘Ž, ๐‘), ๐‘Ž + ๐‘) : ๐‘Ž, ๐‘ ∈ ๐™}
3. Write ๐‘… โ—ฆ ๐‘† as a roster.
(a) ๐‘… = {(1, 0), (2, 3), (4, 6)}, ๐‘†
(b) ๐‘… = {(1, 3), (2, 5), (3, 1)}, ๐‘†
(c) ๐‘… = {(1, 2), (3, 4), (5, 6)}, ๐‘†
(d) ๐‘… = {(1, 2), (3, 4), (5, 6)}, ๐‘†
= {(1, 2), (2, 3), (3, 4)}
= {(1, 3), (3, 1), (5, 2)}
= {(1, 2), (3, 4), (5, 6)}
= {(2, 1), (3, 5), (5, 7)}
4. Write ๐‘… โ—ฆ ๐‘† using abstraction.
(a) ๐‘… = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 = 1}
๐‘† = ๐‘2
(b) ๐‘… = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 = 1}
๐‘† =๐™×๐™
(c) ๐‘… = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 = 1}
๐‘† = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : (๐‘ฅ − 2)2 + ๐‘ฆ2 = 1}
(d) ๐‘… = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 = 1}
๐‘† = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฆ = 2๐‘ฅ − 1}
5. Write the inverse of each relation. Use the abstraction method where appropriate.
(a) ∅
(b) ๐ผ๐™
(c) {(1, 0), (2, 3), (4, 6)}
168
Chapter 4 RELATIONS AND FUNCTIONS
(d)
(e)
(f)
(g)
(h)
{(1, 0), (1, 1), (2, 1)}
๐™×๐‘
{(๐‘ฅ, sin ๐‘ฅ) : ๐‘ฅ ∈ ๐‘}
{(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ + ๐‘ฆ = 1}
{(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : ๐‘ฅ2 + ๐‘ฆ2 = 1}
6. Let ๐‘… ⊆ ๐ด × ๐ต and ๐‘† ⊆ ๐ต × ๐ถ. Show the following.
(a) ๐‘…−1 โ—ฆ ๐‘† −1 ⊆ ๐ถ × ๐ด.
(b) dom(๐‘…) = ran(๐‘…−1 ).
(c) ran(๐‘…) = dom(๐‘…−1 ).
7. Prove that if ๐‘… is a binary relation, (๐‘…−1 )−1 = ๐‘….
8. Prove that (๐‘† โ—ฆ ๐‘…)−1 = ๐‘…−1 โ—ฆ ๐‘† −1 if ๐‘… ⊆ ๐ด × ๐ต and ๐‘† ⊆ ๐ต × ๐ถ.
9. Let ๐‘…, ๐‘† ⊆ ๐ด × ๐ต. Prove the following.
(a) If ๐‘… ⊆ ๐‘†, then ๐‘…−1 ⊆ ๐‘† −1 .
(b) (๐‘… ∪ ๐‘†)−1 = ๐‘…−1 ∪ ๐‘† −1 .
(c) (๐‘… ∩ ๐‘†)−1 = ๐‘…−1 ∩ ๐‘† −1 .
10. Let ๐‘… ⊆ ๐ด × ๐ต.
(a) Prove ๐‘… โ—ฆ ๐ผ๐ด = ๐‘… and ๐ผ๐ต โ—ฆ ๐‘… = ๐‘….
(b) Show that if there exists a set ๐ถ such that ๐ด and ๐ต are subsets of ๐ถ, then
๐‘… โ—ฆ ๐ผ๐ถ = ๐ผ๐ถ โ—ฆ ๐‘… = ๐‘….
11. Let ๐‘… ⊆ ๐ด × ๐ต and ๐‘† ⊆ ๐ต × ๐ถ. Show that ๐‘† โ—ฆ ๐‘… = ∅ if and only if dom(๐‘†) and
ran(๐‘…) are disjoint.
12. For any relation ๐‘…, prove ๐ผran(๐‘…) ⊆ ๐‘… โ—ฆ ๐‘…−1 .
13. Let ๐‘… ⊆ ๐ด × ๐ต. Prove.
โ‹ƒ
(a)
{๐‘ฅ ∈ ๐ด : (๐‘ฅ, ๐‘) ∈ ๐‘…} = dom(๐‘…).
โ‹ƒ๐‘∈๐ต
(b)
๐‘Ž∈๐ด {๐‘ฆ ∈ ๐ต : (๐‘Ž, ๐‘ฆ) ∈ ๐‘…} = ran(๐‘…).
4.2
EQUIVALENCE RELATIONS
In practice we usually do not write relations as sets of ordered pairs. We instead write
propositions like 4 = 4 or 3 < 9. To copy this, we will introduce an alternate notation.
DEFINITION 4.2.1
Let ๐‘… be a relation on ๐ด. For all ๐‘Ž, ๐‘ ∈ ๐ด,
๐‘Ž ๐‘… ๐‘ if and only if (๐‘Ž, ๐‘) ∈ ๐‘…,
and
๐‘Ž๐‘…
ฬธ ๐‘ if and only if (๐‘Ž, ๐‘) ∉ ๐‘….
For example, the less-than relation ๐ฟ (Example 4.1.3) is usually denoted by <, and we
write 2 < 3 instead of (2, 3) ∈ ๐ฟ or (2, 3) ∈ <.
Section 4.2 EQUIVALENCE RELATIONS
169
EXAMPLE 4.2.2
Define the relation ๐‘… on ๐™ by
๐‘… = {(๐‘Ž, ๐‘) ∈ ๐™ × ๐™ : ∃๐‘ ∈ ๐™(๐‘ = ๐‘Ž๐‘ ∧ ๐‘Ž ≠ 0)}.
Therefore, for all ๐‘Ž, ๐‘ ∈ ๐™, ๐‘Ž ๐‘… ๐‘ if and only if ๐‘Ž divides ๐‘. Therefore, 4 ๐‘… 8 but
8๐‘…
ฬธ 4.
Relations can have different properties depending on their definitions. Here are three
important examples using relations on ๐ด = {1, 2, 3}.
โˆ™ {(1, 1), (2, 2), (3, 3)} has the property that every element of ๐ด is related to itself.
โˆ™ {(1, 2), (2, 1), (2, 3), (3, 2)} has the property that if ๐‘Ž is related to ๐‘, then ๐‘ is related to ๐‘Ž.
โˆ™ {(1, 2), (2, 3), (1, 3)} has the property that if ๐‘Ž is related to ๐‘ and ๐‘ is related to ๐‘,
then ๐‘Ž is related to ๐‘.
These examples lead to the following definitions.
DEFINITION 4.2.3
Let ๐‘… be a relation on ๐ด.
โˆ™ ๐‘… is reflexive if ๐‘Ž ๐‘… ๐‘Ž for all ๐‘Ž ∈ ๐ด.
โˆ™ ๐‘… is symmetric when for all ๐‘Ž, ๐‘ ∈ ๐ด, if ๐‘Ž ๐‘… ๐‘, then ๐‘ ๐‘… ๐‘Ž.
โˆ™ ๐‘… is transitive means that for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐ด, if ๐‘Ž ๐‘… ๐‘ and ๐‘ ๐‘… ๐‘, then ๐‘Ž ๐‘… ๐‘.
Notice that the relation in Example 4.2.2 is not reflexive because 0 does not divide 0
and is not symmetric because 4 divides 8 but 8 does not divide 4, but it is transitive
because if ๐‘Ž divides ๐‘ and ๐‘ divides ๐‘, then ๐‘Ž divides ๐‘.
When a relation is reflexive, symmetric, and transitive, it behaves very much like an
identity relation (Example 4.1.2). Such relations play an important role in mathematics,
so we name them.
DEFINITION 4.2.4
A relation ๐‘… on ๐ด is an equivalence relation if ๐‘… is reflexive, symmetric, and
transitive.
Observe that the relation in Example 4.2.2 is not an equivalence relation. However, any
identity relation is an equivalence relation. We see this assumption at work in the next
example.
170
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.2.5
Let ๐‘… be a relation on ๐™ × (๐™ โงต {0}) so that for all ๐‘Ž, ๐‘ ∈ ๐™ and ๐‘, ๐‘‘ ∈ ๐™ โงต {0},
(๐‘Ž, ๐‘) ๐‘… (๐‘, ๐‘‘) if and only if ๐‘Ž๐‘‘ = ๐‘๐‘.
To see that this is an equivalence relation, let (๐‘Ž, ๐‘), (๐‘, ๐‘‘), and (๐‘’, ๐‘“ ) be elements
of ๐™ × (๐™ โงต {0}).
โˆ™ (๐‘Ž, ๐‘) ๐‘… (๐‘Ž, ๐‘) since ๐‘Ž๐‘ = ๐‘Ž๐‘.
โˆ™ Assume (๐‘Ž, ๐‘) ๐‘… (๐‘, ๐‘‘). Then, ๐‘Ž๐‘‘ = ๐‘๐‘. This implies that ๐‘๐‘ = ๐‘‘๐‘Ž, so
(๐‘, ๐‘‘) ๐‘… (๐‘Ž, ๐‘).
โˆ™ Let (๐‘Ž, ๐‘) ๐‘… (๐‘, ๐‘‘) and (๐‘, ๐‘‘) ๐‘… (๐‘’, ๐‘“ ). This gives ๐‘Ž๐‘‘ = ๐‘๐‘ and ๐‘๐‘“ = ๐‘‘๐‘’.
Therefore, (๐‘Ž, ๐‘) ๐‘… (๐‘’, ๐‘“ ) because
๐‘Ž๐‘“ =
๐‘๐‘๐‘“
= ๐‘๐‘’.
๐‘‘
EXAMPLE 4.2.6
Take ๐‘š ∈ ๐™+ and let ๐‘Ž, ๐‘, and ๐‘ be integers. Define ๐‘Ž to be congruent to ๐‘
modulo ๐‘š and write
๐‘Ž ≡ ๐‘ (mod ๐‘š) if and only if ๐‘š โˆฃ ๐‘Ž − ๐‘.
That is,
๐‘Ž ≡ ๐‘ (mod ๐‘š) if and only if ๐‘Ž = ๐‘ + ๐‘š๐‘˜ for some ๐‘˜ ∈ ๐™.
For example, we have that 7 ≡ 1 (mod 3), 1 ≡ 13 (mod 3), and 27 ≡ 0 (mod 3),
but 2 โ‰ข 9 (mod 3) and 25 โ‰ข 0 (mod 3). Congruence modulo ๐‘š defines the
relation
๐‘…๐‘š = {(๐‘Ž, ๐‘) : ๐‘Ž ≡ ๐‘ (mod ๐‘š)}.
Observe that
๐‘…๐‘š = {(๐‘Ž, ๐‘) : ๐‘š โˆฃ ๐‘Ž − ๐‘} = {(๐‘Ž, ๐‘) : ∃๐‘˜(๐‘˜ ∈ ๐™ ∧ ๐‘Ž = ๐‘ + ๐‘š๐‘˜)}.
Prove that ๐‘…๐‘š is an equivalence relation.
โˆ™ (๐‘Ž, ๐‘Ž) ∈ ๐‘…๐‘š because ๐‘Ž ≡ ๐‘Ž (mod ๐‘š).
โˆ™ Assume (๐‘Ž, ๐‘) ∈ ๐‘…๐‘š . This implies that ๐‘š โˆฃ ๐‘Ž − ๐‘. By Exercise 2.4.18,
๐‘š โˆฃ ๐‘ − ๐‘Ž. Hence, (๐‘, ๐‘Ž) ∈ ๐‘…๐‘š .
โˆ™ Let (๐‘Ž, ๐‘), (๐‘, ๐‘) ∈ ๐‘…๐‘š . Then ๐‘Ž = ๐‘ + ๐‘š๐‘˜ and ๐‘ = ๐‘ + ๐‘š๐‘™ for some ๐‘˜, ๐‘™ ∈ ๐™.
Substitution yields
๐‘Ž = ๐‘ + ๐‘š๐‘™ + ๐‘š๐‘˜ = ๐‘ + ๐‘š(๐‘™ + ๐‘˜).
Since the sum of two integers is an integer, (๐‘Ž, ๐‘) ∈ ๐‘…๐‘š .
171
Section 4.2 EQUIVALENCE RELATIONS
Equivalence Classes
Let ๐‘… be a relation on {1, 2, 3, 4} such that
๐‘… = {(1, 2), (1, 3), (2, 4)}.
(4.1)
Observe that 1 is related to 2 and 3, 2 is related to 4, and 3 and 4 are not related to any
number. Combining the elements that are related to a particular element results in a set
named by the next definition.
DEFINITION 4.2.7
Let ๐‘… be a relation on ๐ด with ๐‘Ž ∈ ๐ด. The class of ๐‘Ž with respect to ๐‘… is the set
[๐‘Ž]๐‘… = {๐‘ฅ ∈ ๐ด : ๐‘Ž ๐‘… ๐‘ฅ}.
If ๐‘… is an equivalence relation, [๐‘Ž]๐‘… is called an equivalence class. We often
denote [๐‘Ž]๐‘… by [๐‘Ž] if the relation is clear from context.
Using ๐‘… as defined in (4.1),
[1]๐‘… = {2, 3}, [2]๐‘… = {4}, and [3]๐‘… = [4]๐‘… = ∅.
If ๐‘… had been an equivalence relation on a set ๐ด, then [๐‘Ž]๐‘… would be nonempty for all
๐‘Ž ∈ ๐ด because ๐‘Ž would be an element of [๐‘Ž]๐‘… (Exercise 17).
EXAMPLE 4.2.8
Let ๐‘… be the equivalence relation from Example 4.2.5. We prove that
[(1, 3)] = {(๐‘›, 3๐‘›) : ๐‘› ∈ ๐™ โงต {0}}.
To see this, take (๐‘Ž, ๐‘) ∈ [(1, 3)]. This means that (1, 3) ๐‘… (๐‘Ž, ๐‘), so ๐‘ = 3๐‘Ž and
๐‘Ž ≠ 0 because ๐‘ ≠ 0. Hence, (๐‘Ž, ๐‘) = (๐‘Ž, 3๐‘Ž). Conversely, let ๐‘› ≠ 0. Then,
(1, 3) ๐‘… (๐‘›, 3๐‘›) because 1 ⋅ 3๐‘› = 3 ⋅ ๐‘›. Thus, (๐‘›, 3๐‘›) ∈ [(1, 3)].
EXAMPLE 4.2.9
Using the notation of Example 4.2.6, let ๐‘…๐‘š be the relation defined by congruence
modulo ๐‘š. For all ๐‘› ∈ ๐™, define
[๐‘›]๐‘š = ๐‘…๐‘š (๐‘›).
Therefore, when ๐‘š = 5, the equivalence classes are:
[0]5
[1]5
[2]5
[3]5
[4]5
= {… , −10, −5, 0, 5, 10, … },
= {… , −9, −4, 1, 6, 11, … },
= {… , −8, −3, 2, 7, 12, … },
= {… , −7, −2, 3, 8, 13, … },
= {… , −6, −1, 4, 9, 14, … }.
172
Chapter 4 RELATIONS AND FUNCTIONS
In addition,
[๐‘Ž]5 = [๐‘]5 if and only if ๐‘Ž ≡ ๐‘ (mod 5).
The collection of all equivalence classes of a relation is a set named by the next
definition.
DEFINITION 4.2.10
Let ๐‘… be an equivalence relation on ๐ด. The quotient set of ๐ด modulo ๐‘… is
๐ดโˆ•๐‘… = {[๐‘Ž]๐‘… : ๐‘Ž ∈ ๐ด}.
Observe by Exercise 3 that it is always the case that
โ‹ƒ
๐ด=
๐ดโˆ•๐‘….
(4.2)
EXAMPLE 4.2.11
Let ๐‘š ∈ ๐™+ . The quotient set ๐™โˆ•๐‘…๐‘š is denoted by ๐™๐‘š . That is,
๐™๐‘š = {[0]๐‘š , [1]๐‘š , … , [๐‘š − 1]๐‘š }.
EXAMPLE 4.2.12
Define the relation ๐‘… on ๐‘2 by
(๐‘Ž, ๐‘) ๐‘… (๐‘, ๐‘‘) if and only if ๐‘ − ๐‘Ž = ๐‘‘ − ๐‘.
๐‘… is an equivalence relation by Exercise 2. We note that for any (๐‘Ž, ๐‘) ∈ ๐‘2 ,
[(๐‘Ž, ๐‘)] = {(๐‘ฅ, ๐‘ฆ) : (๐‘ฅ, ๐‘ฆ) ๐‘… (๐‘Ž, ๐‘)}
= {(๐‘ฅ, ๐‘ฆ) : ๐‘ฆ − ๐‘ฅ = ๐‘ − ๐‘Ž}
= {(๐‘ฅ, ๐‘ฆ) : ๐‘ฆ = ๐‘ฅ + (๐‘ − ๐‘Ž)}.
Therefore, the equivalence class of (๐‘Ž, ๐‘) is the line with a slope of 1 and a ๐‘ฆintercept equal to (0, ๐‘ − ๐‘Ž). The equivalence classes of (0, 1.5) and (0, −1) are
illustrated in the graph in Figure 4.3. The quotient set ๐‘2 โˆ•๐‘… is the collection of
all such lines. Notice that
โ‹ƒ
๐‘2 =
๐‘2 โˆ•๐‘….
Partitions
In Example 4.2.12, we saw that ๐‘2 is the union of all the lines with slope equal to 1, and
since the lines are parallel, they form a pairwise disjoint set. These properties can be
observed in the other equivalence relations that we have seen. Each set is equal to the
union of the equivalence classes, and the quotient set is pairwise disjoint. Generalizing
these two properties leads to the next definition.
Section 4.2 EQUIVALENCE RELATIONS
173
R(0, 1.5)
R(0, −1)
Figure 4.3
Two equivalence classes in ๐‘2 when (๐‘Ž, ๐‘) ๐‘… (๐‘, ๐‘‘) if and only if ๐‘ − ๐‘Ž = ๐‘‘ − ๐‘.
DEFINITION 4.2.13
Let ๐ด be a nonempty set. The family ๐’ซ is a partition of ๐ด if and only if
โˆ™ ๐’ซ ⊆ P(๐ด),
โˆ™
โ‹ƒ
๐’ซ = ๐ด,
โˆ™ ๐’ซ is pairwise disjoint.
To illustrate the definition, let ๐ด = {1, 2, 3, 4, 5, 6, 7} and define the elements of the
partition to be ๐ด0 = {1, 2, 5}, ๐ด1 = {3}, and ๐ด2 = {4, 6, 7}. The family
๐’ซ = {๐ด1 , ๐ด2 , ๐ด3 }
is a subset of P(๐ด), ๐ด = ๐ด0 ∪ ๐ด1 ∪ ๐ด2 , and ๐’ซ is pairwise disjoint. Therefore, ๐’ซ is a
partition of ๐ด. This is illustrated in Figure 4.4.
A1
A2
1
2
3
5
6
4
7
A3
Figure 4.4
A partition of the set ๐ด = {1, 2, 3, 4, 5, 6, 7}.
174
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.2.14
For each real number ๐‘Ÿ ≥ 0, define ๐ถ๐‘Ÿ to be the circle with radius ๐‘Ÿ centered at
the origin. Namely,
{
}
√
๐ถ๐‘Ÿ = (๐‘ฅ, ๐‘ฆ) : ๐‘ฅ2 + ๐‘ฆ2 = ๐‘Ÿ .
Let ๐’ž = {๐ถ๐‘Ÿ : ๐‘Ÿ ∈ [0, ∞)}. We claim that ๐’ž is a partition of ๐‘2 .
โˆ™ ๐’ž ⊆ P(๐‘2 ) because ๐ถ๐‘Ÿ ⊆ ๐‘2 for all ๐‘Ÿ ≥ 0.
โ‹ƒ
โ‹ƒ
โˆ™ To prove that ๐‘2 = ๐‘Ÿ∈[0,∞) ๐ถ๐‘Ÿ , it suffices to show that ๐‘2 ⊆ ๐‘Ÿ∈[0,∞) ๐ถ๐‘Ÿ ,
but this follows because if (๐‘Ž, ๐‘) ∈ ๐‘2 , then
(๐‘Ž, ๐‘) ∈ ๐ถ√
๐‘Ž2 +๐‘2
.
โˆ™ To see that ๐’ž is pairwise disjoint, let ๐‘Ÿ, ๐‘  ≥ 0 and assume that (๐‘Ž, ๐‘) is an
element of ๐ถ๐‘Ÿ ∩ ๐ถ๐‘  . Then,
√
๐‘Ÿ = ๐‘Ž2 + ๐‘2 = ๐‘ ,
which implies that ๐ถ๐‘Ÿ = ๐ถ๐‘  .
โ‹ƒ
The set ๐™5 is a family of subsets of ๐™, has the property that ๐™5 = ๐™, and is
pairwise disjoint. Hence, ๐™5 is a partition for ๐™. We generalize this result to the next
theorem. It uses an arbitrary equivalence relation on a given set to define a partition
for that set. In this case, we say that the equivalence relation induces the partition.
THEOREM 4.2.15
If ๐‘… is an equivalence relation on ๐ด, then ๐ดโˆ•๐‘… is a partition of ๐ด.
PROOF
Take a set ๐ด with an equivalence relation ๐‘….
โˆ™ Since an equivalence class is a subset of ๐ด, we have ๐ดโˆ•๐‘… ⊆ P(๐ด).
โ‹ƒ
โˆ™
๐ดโˆ•๐‘… = ๐ด is (4.2).
โˆ™ Let [๐‘Ž], [๐‘] ∈ ๐ดโˆ•๐‘… and assume that there exists ๐‘ฆ ∈ [๐‘Ž] ∩ [๐‘]. In other
words, ๐‘Ž ๐‘… ๐‘ฆ and ๐‘ ๐‘… ๐‘ฆ. Now take ๐‘ฅ ∈ [๐‘Ž]. This means that ๐‘Ž ๐‘… ๐‘ฅ. Since
๐‘ฅ ๐‘… ๐‘Ž and ๐‘ฆ ๐‘… ๐‘ by symmetry, we have that ๐‘ฅ ๐‘… ๐‘ฆ, and then ๐‘ฅ ๐‘… ๐‘ by transitivity. Thus, ๐‘ฅ ∈ [๐‘], which shows [๐‘Ž] ⊆ [๐‘]. Similarly, [๐‘] ⊆ [๐‘Ž], so
[๐‘Ž] = [๐‘].
The collection of equivalence relations forms a partition of a set. Conversely, if we
have a partition of a set, the partition gives rise to an equivalence relation on the set.
To see this, take any set ๐ด and a partition ๐’ซ of ๐ด. For all ๐‘Ž, ๐‘ ∈ ๐ด, define
๐‘Ž ๐‘… ๐‘ if and only if there exists ๐ถ ∈ ๐’ซ such that ๐‘Ž, ๐‘ ∈ ๐ถ.
Section 4.2 EQUIVALENCE RELATIONS
175
To show that ๐‘… is an equivalence relation, take ๐‘Ž, ๐‘, and ๐‘ in ๐ด.
โ‹ƒ
โˆ™ Since ๐‘Ž ∈ ๐ด and ๐ด = ๐’ซ , there exists ๐ถ ∈ ๐’ซ such that ๐‘Ž ∈ ๐ถ. Therefore,
๐‘Ž ๐‘… ๐‘Ž.
โˆ™ Assume ๐‘Ž ๐‘… ๐‘. This means that ๐‘Ž, ๐‘ ∈ ๐ถ for some ๐ถ ∈ ๐’ซ . This, of course, is
the same as ๐‘, ๐‘Ž ∈ ๐ถ. Hence, ๐‘ ๐‘… ๐‘Ž.
โˆ™ Suppose ๐‘Ž ๐‘… ๐‘ and ๐‘ ๐‘… ๐‘. Then, there are sets ๐ถ and ๐ท in ๐’ซ so that ๐‘Ž, ๐‘ ∈ ๐ถ and
๐‘, ๐‘ ∈ ๐ท. This means that ๐ถ ∩ ๐ท ≠ ∅. Since ๐’ซ is pairwise disjoint, ๐ถ = ๐ท. So,
๐‘Ž and ๐‘ are elements of ๐ถ, and we have ๐‘Ž ๐‘… ๐‘.
This equivalence relation is said to be induced from the partition.
EXAMPLE 4.2.16
The sets
{… , −10, −5, 0, 5, 10, … },
{… , −9, −4, 1, 6, 11, … },
{… , −8, −3, 2, 7, 12, … },
{… , −7, −2, 3, 8, 13, … },
{… , −6, −1, 4, 9, 14, … },
form a collection that is a partition of ๐™. The equivalence relation that is induced
from this partition is congruence modulo 5 (Example 4.2.9).
Exercises
1. For all ๐‘Ž, ๐‘ ∈ ๐‘ โงต {0}, let ๐‘Ž ๐‘… ๐‘ if and only if ๐‘Ž๐‘ > 0.
(a) Show that ๐‘… is an equivalence relation on ๐‘ โงต {0}.
(b) Find [1] and [−3].
2. Define the relation ๐‘† on ๐‘2 by (๐‘Ž, ๐‘) ๐‘† (๐‘, ๐‘‘) if and only if ๐‘ − ๐‘Ž = ๐‘‘ − ๐‘. Prove that
๐‘† is an equivalence relation.
โ‹ƒ
3. Let ๐‘† be an equivalence relation on ๐ด. Prove that ๐ด = ๐‘Ž∈๐ด [๐‘Ž].
4. Prove that if ๐ถ is an equivalence class for some equivalence relation ๐‘… and ๐‘Ž ∈ ๐ถ,
then ๐ถ = [๐‘Ž].
5. For all ๐‘Ž, ๐‘ ∈ ๐™, let ๐‘Ž ๐‘… ๐‘ if and only if |๐‘Ž| = |๐‘|.
(a) Prove ๐‘… is an equivalence relation on ๐™.
(b) Sketch the partition of ๐™ induced by this equivalence relation.
6. For all (๐‘Ž, ๐‘), (๐‘, ๐‘‘) ∈ ๐™ × ๐™, define (๐‘Ž, ๐‘) ๐‘† (๐‘, ๐‘‘) if and only if ๐‘Ž๐‘ = ๐‘๐‘‘.
(a) Show that ๐‘† is an equivalence relation on ๐™ × ๐™.
(b) What is the equivalence class of (1, 2)?
(c) Sketch the partition of ๐™ × ๐™ induced by this equivalence relation.
7. Let ๐ด be a set and ๐‘Ž ∈ ๐ด. Show that the given relations are not equivalence relations
on P(๐ด).
176
Chapter 4 RELATIONS AND FUNCTIONS
(a) For all ๐ถ, ๐ท ⊆ ๐ด, define ๐ถ ๐‘… ๐ท if and only if ๐ถ ∩ ๐ท ≠ ∅.
(b) For all ๐ถ, ๐ท ⊆ ๐ด, define ๐ถ ๐‘† ๐ท if and only if ๐‘Ž ∈ ๐ถ ∩ ๐ท.
8. Let ๐‘ง, ๐‘ง′ ∈ ๐‚ and write ๐‘ง = ๐‘Ž + ๐‘๐‘– and ๐‘ง′ = ๐‘Ž′ + ๐‘′ ๐‘–. Define ๐‘ง = ๐‘ง′ to mean that
๐‘Ž = ๐‘Ž′ and ๐‘ = ๐‘′ . Prove that ๐‘… is an equivalence relation.
9. Find.
(a)
(b)
(c)
(d)
[3]5
[12]6
[2]5 ∪ [27]5
[4]7 ∩ [5]7
10. Let ๐‘Ÿ be the remainder obtained when ๐‘› is divided by ๐‘š. Prove [๐‘›]๐‘š = [๐‘Ÿ]๐‘š .
11. Let ๐‘, ๐‘š ∈ ๐™ and suppose that gcd(๐‘, ๐‘š) = 1. Prove.
(a) There exists ๐‘ such that ๐‘๐‘ ≡ 1 (mod ๐‘š). (Notice that ๐‘ is that multiplicative
inverse of ๐‘ modulo ๐‘š.)
(b) If ๐‘๐‘Ž ≡ ๐‘๐‘ (mod ๐‘š), then ๐‘Ž ≡ ๐‘ (mod ๐‘š).
(c) Prove that the previous implication is false if gcd(๐‘, ๐‘š) ≠ 1.
12. Prove that {(๐‘›, ๐‘› + 1] : ๐‘› ∈ ๐™} is a partition of ๐‘.
13. Prove that the following are partitions of ๐‘2 .
(a) ๐’ซ = {{(๐‘Ž, ๐‘)} : ๐‘Ž, ๐‘ ∈ ๐‘}.
(b) ๐’ฑ = {{(๐‘Ÿ, ๐‘ฆ) : ๐‘ฆ ∈ ๐‘} : ๐‘Ÿ ∈ ๐‘}.
(c) โ„‹ = {๐‘ × (๐‘›, ๐‘› + 1] : ๐‘› ∈ ๐™}.
14. Is {[๐‘›, ๐‘› + 1] × (๐‘›, ๐‘› + 1) : ๐‘› ∈ ๐™} a partition of ๐‘2 ? Explain.
15. Let ๐‘… be a relation on ๐ด and show the following:
(a) ๐‘… is reflexive if and only if ๐‘…−1 is reflexive.
(b) ๐‘… is symmetric if and only if ๐‘… = ๐‘…−1 .
(c) ๐‘… is symmetric if and only if (๐ด × ๐ด) โงต ๐‘… is symmetric.
16. Let ๐‘… and ๐‘† be equivalence relations on ๐ด. Prove or show false.
(a) ๐‘… ∪ ๐‘† is an equivalence relation on ๐ด.
(b) ๐‘… ∩ ๐‘† is an equivalence relation on ๐ด.
17. Prove for all relations ๐‘… on ๐ด.
(a) ๐‘… is reflexive if and only if ∀๐‘Ž(๐‘Ž ∈ [๐‘Ž]).
(b) ๐‘… is symmetric if and only if ∀๐‘Ž∀๐‘(๐‘Ž ∈ [๐‘] ↔ ๐‘ ∈ [๐‘Ž]).
(c) ๐‘… is transitive if and only if ∀๐‘Ž∀๐‘∀๐‘([๐‘ ∈ [๐‘Ž] ∧ ๐‘ ∈ [๐‘]] → ๐‘ ∈ [๐‘Ž]).
(d) ๐‘… is an equivalence relation if and only if ∀๐‘Ž∀๐‘((๐‘Ž, ๐‘) ∈ ๐‘… ↔ [๐‘Ž] = [๐‘]).
18. Let ๐‘… be a relation on ๐ด with the property that if ๐‘Ž ๐‘… ๐‘ and ๐‘ ๐‘… ๐‘, then ๐‘ ๐‘… ๐‘Ž. Prove
that if ๐‘… is also reflexive, ๐‘… is an equivalence relation.
19. Define the relation ๐‘… on ๐‚ by ๐‘Ž + ๐‘๐‘– ๐‘… ๐‘ + ๐‘‘๐‘– if and only if
√
√
๐‘Ž2 + ๐‘2 = ๐‘ 2 + ๐‘‘ 2 .
Section 4.3 PARTIAL ORDERS
177
(a) Prove ๐‘… is an equivalence relation on ๐‚.
(b) Graph [1 + ๐‘–] in the complex plane.
(c) Describe the partition that ๐‘… induces on ๐‚.
20. Let ๐‘… and ๐‘† be relations on ๐ด. The symmetric closure of ๐‘… is ๐‘† if ๐‘… ⊆ ๐‘† and
for all symmetric relations ๐‘‡ on ๐ด such that ๐‘… ⊆ ๐‘‡ , then ๐‘† ⊆ ๐‘‡ . Prove the following.
(a) ๐‘… ∪ ๐‘…−1 is the symmetric closure of ๐‘….
(b) A symmetric closure is unique.
4.3
PARTIAL ORDERS
While equivalence relations resemble equality, there are other common relations in
mathematics that we can model. To study some of their attributes, we expand Definition 4.2.3 with three more properties.
DEFINITION 4.3.1
Let ๐‘… be a relation on ๐ด.
โˆ™ ๐‘… is irreflexive if ๐‘Ž ๐‘…
ฬธ ๐‘Ž for all ๐‘Ž ∈ ๐ด.
โˆ™ ๐‘… is asymmetric when for all ๐‘Ž, ๐‘ ∈ ๐ด, if ๐‘Ž ๐‘… ๐‘, then ๐‘ ๐‘…
ฬธ ๐‘Ž.
โˆ™ ๐‘… is antisymmetric means that for all ๐‘Ž, ๐‘ ∈ ๐ด, if ๐‘Ž ๐‘… ๐‘ and ๐‘ ๐‘… ๐‘Ž, then
๐‘Ž = ๐‘.
Notice that a relation on a nonempty set cannot be both reflexive and irreflexive. However, many relations have neither property. For example, consider the relation ๐‘… =
{(1, 1)} on {1, 2}. Since (1, 1) ∈ ๐‘…, the relation ๐‘… is not irreflexive, and ๐‘… is not
reflexive because (2, 2) ∉ ๐‘…. Likewise, a relation on a nonempty set cannot be both
symmetric and asymmetric.
EXAMPLE 4.3.2
The less-than relation on ๐™ is irreflexive and asymmetric. It is also antisymmetric. To see this, let ๐‘Ž, ๐‘ ∈ ๐™. Since ๐‘Ž < ๐‘ and ๐‘ < ๐‘Ž is false, the implication
if ๐‘Ž < ๐‘ and ๐‘ < ๐‘Ž, then ๐‘Ž = ๐‘
is true. The ≤ relation is also antisymmetric. However, ≤ is neither irreflexive
nor asymmetric since 3 ≤ 3.
EXAMPLE 4.3.3
Let ๐‘… = {(1, 2)} and ๐‘† = {(1, 2), (2, 1)}. Both are relations on {1, 2}. The first
relation is asymmetric since 1 ๐‘… 2 but 2 ๐‘…
ฬธ 1. It is also antisymmetric, but ๐‘† is
not antisymmetric because 2 ๐‘† 1 and 1 ๐‘† 2. Both the relations are irreflexive.
178
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.3.4
Let ๐‘… be a relation on a set ๐ด. We prove that
๐‘… is antisymmetric if and only if ๐‘… ∩ ๐‘…−1 ⊆ ๐ผ๐ด .
โˆ™ Assume that ๐‘… is antisymmetric and take (๐‘Ž, ๐‘) ∈ ๐‘… ∩ ๐‘…−1 . This means
that (๐‘Ž, ๐‘) ∈ ๐‘… and (๐‘Ž, ๐‘) ∈ ๐‘…−1 . Therefore, (๐‘, ๐‘Ž) ∈ ๐‘…, and since ๐‘… is
antisymmetric, ๐‘Ž = ๐‘.
โˆ™ Now suppose ๐‘… ∩ ๐‘…−1 ⊆ ๐ผ๐ด . Let (๐‘Ž, ๐‘), (๐‘, ๐‘Ž) ∈ ๐‘…. We conclude that
(๐‘Ž, ๐‘) ∈ ๐‘…−1 , which implies that (๐‘Ž, ๐‘) ∈ ๐‘… ∩ ๐‘…−1 . Then, (๐‘Ž, ๐‘) ∈ ๐ผ๐ด .
Hence, ๐‘Ž = ๐‘.
As an equivalence relation is a generalization of an identity relation, the following
relation is a generalization of ≤ on ๐. For this reason, instead of naming the relation
๐‘…, it is denoted by the symbol 4.
DEFINITION 4.3.5
If a relation 4 on a set ๐ด is reflexive, antisymmetric, and transitive, 4 is a partial
order on ๐ด and the ordered pair (๐ด, 4) is called a partially ordered set (or
simply a poset). Furthermore, for all ๐‘Ž, ๐‘ ∈ ๐ด, the notation ๐‘Ž โ‰บ ๐‘ means ๐‘Ž 4 ๐‘
but ๐‘Ž ≠ ๐‘.
For example, ≤ and = are partial orders on ๐‘, but < is not a partial order on ๐‘ because
the relation < is not reflexive. Although = is a partial order on any set, in general an
equivalence relation is not a partial order (Example 4.2.6).
EXAMPLE 4.3.6
Divisibility (Definition 2.4.2) is a partial order on ๐™+ . To prove this, let ๐‘Ž, ๐‘, and
๐‘ be positive integers.
โˆ™ ๐‘Ž โˆฃ ๐‘Ž since ๐‘Ž = ๐‘Ž ⋅ 1 and ๐‘Ž ≠ 0.
โˆ™ Suppose that ๐‘Ž โˆฃ ๐‘ and ๐‘ โˆฃ ๐‘Ž. This means that ๐‘ = ๐‘Ž๐‘˜ and ๐‘Ž = ๐‘๐‘™, for some
๐‘˜, ๐‘™ ∈ ๐™+ . Hence, ๐‘ = ๐‘๐‘™๐‘˜, so ๐‘™๐‘˜ = 1. Since ๐‘˜ and ๐‘™ are positive integers,
๐‘™ = ๐‘˜ = 1. That is, ๐‘Ž = ๐‘.
โˆ™ Assume that we have ๐‘Ž โˆฃ ๐‘ and ๐‘ โˆฃ ๐‘. This means that ๐‘ = ๐‘Ž๐‘™ and ๐‘ = ๐‘๐‘˜
for some ๐‘™, ๐‘˜ ∈ ๐™+ . By substitution, ๐‘ = (๐‘Ž๐‘™)๐‘˜ = ๐‘Ž(๐‘™๐‘˜). Hence, ๐‘Ž โˆฃ ๐‘.
EXAMPLE 4.3.7
Let ๐”ธ be a collection of symbols and let ๐”ธ∗ denote the set of all strings over ๐”ธ
(as on page 5). Use the symbol to denote the empty string, the string of length
zero. As with the empty set, the empty string is always an element of ๐”ธ∗ . For
example, if ๐”ธ = {๐‘Ž, ๐‘, ๐‘}, then ๐‘Ž๐‘๐‘, ๐‘Ž๐‘Ž๐‘Ž๐‘๐‘๐‘, ๐‘, and are elements of ๐”ธ∗ . Now,
take ๐œŽ, ๐œ ∈ ๐”ธ∗ . The concatenation of ๐œŽ and ๐œ is denoted by ๐œŽ โŒข ๐œ and is the
Section 4.3 PARTIAL ORDERS
000
001
010
00
011
100
01
101
110
10
0
179
111
11
1
โ–ก
Figure 4.5
A partial order defined on {0, 1}∗ .
string consisting of the elements of ๐œŽ followed by those of ๐œ. For example, if
๐œŽ = 011 and ๐œ = 1010, then ๐œŽ โŒข ๐œ = 0111010. Finally, for all ๐œŽ, ๐œ ∈ ๐”ธ∗ , define
๐œŽ 4 ๐œ if and only if there exists ๐œˆ ∈ ๐”ธ∗ such that ๐œ = ๐œŽ โŒข ๐œˆ.
It can be shown that 4 is a partial order on ๐”ธ∗ (Exercise 8) with the structure
seen in Figure 4.5 for ๐”ธ = {0, 1}.
The partial order ≤ on ๐‘ has the property that for all ๐‘Ž, ๐‘ ∈ ๐‘, either ๐‘Ž < ๐‘, ๐‘ < ๐‘Ž,
or ๐‘Ž = ๐‘. As we see in Figure 4.5, this is not the case for every partially ordered set.
We do, however, have the following slightly weaker property.
THEOREM 4.3.8 [Weak-Trichotomy Law]
If 4 is a partial order on ๐ด, for all ๐‘Ž, ๐‘ ∈ ๐ด, at most one of the following are true:
๐‘Ž โ‰บ ๐‘, ๐‘ โ‰บ ๐‘Ž, or ๐‘Ž = ๐‘.
PROOF
Let ๐‘Ž, ๐‘ ∈ ๐ด. We have three cases to consider.
โˆ™ Suppose ๐‘Ž โ‰บ ๐‘. This means that ๐‘Ž 4 ๐‘ and ๐‘Ž ≠ ๐‘. If in addition ๐‘ โ‰บ ๐‘Ž, by
transitivity ๐‘Ž โ‰บ ๐‘Ž, which is a contradiction.
โˆ™ That ๐‘ โ‰บ ๐‘Ž precludes both ๐‘Ž โ‰บ ๐‘ and ๐‘Ž = ๐‘ is proved like the first case.
โˆ™ If ๐‘Ž = ๐‘, then by definition of โ‰บ it is impossible for ๐‘Ž โ‰บ ๐‘ or ๐‘ โ‰บ ๐‘Ž to be
true.
Technically, subset is not a relation, but in a natural way, it can be considered as one.
Let โ„ฑ be a family of sets. Define
๐‘† = {(๐ด, ๐ต) : ๐ด ⊆ ๐ต ∧ ๐ด, ๐ต ∈ โ„ฑ }.
Associate ⊆ with the relation ๐‘†.
EXAMPLE 4.3.9
Let ๐ด be a nonempty set. We show that (P(๐ด), ⊆) is a partially ordered set. Let
๐ต, ๐ถ, and ๐ท be subsets of ๐ด.
180
Chapter 4 RELATIONS AND FUNCTIONS
โˆ™ Since ๐ต ⊆ ๐ต, the relation ⊆ is reflexive.
โˆ™ Since ๐ต ⊆ ๐ถ and ๐ถ ⊆ ๐ต implies that ๐ต = ๐ถ (Definition 3.3.7), ⊆ is
antisymmetric.
โˆ™ Since ๐ต ⊆ ๐ถ and ๐ถ ⊆ ๐ท implies ๐ต ⊆ ๐ท (Theorem 3.3.6), we see that ⊆ is
transitive.
This example is in line with what we know about subsets. For instance, if ๐ด ⊂ ๐ต,
then we conclude that ๐ต ⊄ ๐ด and ๐ด ≠ ๐ต, which is what we expect from the
weak-trichotomy law (4.3.8).
Bounds
Let 4 be a partial order on ๐ด with elements ๐‘š and ๐‘š′ such that ๐‘Ž 4 ๐‘š and ๐‘Ž 4 ๐‘š′
for all ๐‘Ž ∈ ๐ด. In particular, this implies that ๐‘š 4 ๐‘š′ and ๐‘š′ 4 ๐‘š, so since 4 is
antisymmetric, we conclude that ๐‘š = ๐‘š′ . Similarly, if ๐‘š 4 ๐‘Ž and ๐‘š′ 4 ๐‘Ž for all ๐‘Ž ∈ ๐ด,
then ๐‘š = ๐‘š′ . This argument justifies the use of the word the in the next definition.
DEFINITION 4.3.10
Let (๐ด, 4) be a poset and ๐‘š ∈ ๐ด.
โˆ™ ๐‘š is the least element of ๐ด (with respect to 4) if ๐‘š 4 ๐‘Ž for all ๐‘Ž ∈ ๐ด.
โˆ™ ๐‘š is the greatest element of ๐ด (with respect to 4) if ๐‘Ž 4 ๐‘š for all ๐‘Ž ∈ ๐ด.
There is no guarantee that a partially ordered set will have a least or a greatest element.
In Example 4.3.9, the greatest element of P(๐ด) with respect to ⊆ is ๐ด and the least
element is ∅. However, in Example 4.3.7 (Figure 4.5), the least element of ๐”ธ∗ is ,
but there is no greatest element.
EXAMPLE 4.3.11
If ๐ด is a finite set of real numbers, ๐ด has a least and a greatest element with
respect to ≤. Under the partial order ≤, the set ๐™+ also has a least element but no
greatest element, and both {5๐‘› : ๐‘› ∈ ๐™− } and ๐™− have greatest elements but no
least elements. To show that ๐ต = {5๐‘› : ๐‘› ∈ ๐™+ } has a least element with respect
to ≤, use the fact that the least element of ๐™+ is 1. Therefore, the least element
of ๐ต is 5 because 5 ∈ ๐ต and 5(1) ≤ 5๐‘› for all ๐‘› ∈ ๐™+ .
Some sets will not have a least or greatest element with a given partial order but
there will still be elements that are considered greater or lesser than every element of
the set.
DEFINITION 4.3.12
Let 4 be a partial order on ๐ด and ๐ต ⊆ ๐ด.
โˆ™ ๐‘ข ∈ ๐ด is an upper bound of ๐ต if ๐‘ 4 ๐‘ข for all ๐‘ ∈ ๐ต. The element ๐‘ข is the
least upper bound of ๐ต if it is an upper bound and for all upper bounds ๐‘ข′
of ๐ต, ๐‘ข 4 ๐‘ข′ .
181
Section 4.3 PARTIAL ORDERS
โˆ™ ๐‘™ ∈ ๐ด is a lower bound of ๐ต if ๐‘™ 4 ๐‘ for all ๐‘ ∈ ๐ต. The element ๐‘™ is the
greatest lower bound of ๐ต if it is a lower bound and for all lower bounds
๐‘™′ of ๐ต, ๐‘™′ 4 ๐‘™.
In the definition we can write the word the because the order is antisymmetric.
EXAMPLE 4.3.13
The interval (3, 5) is a subset of ๐‘. Under the partial order ≤, both 5 and 10 are
upper bounds of this interval, while 5 is a least upper bound. Also, 3 and −๐œ‹ are
lower bounds, but 3 is the greatest lower bound.
EXAMPLE 4.3.14
Assume
โ‹ƒ that ๐’ž ⊆ P(๐™). Since the elements of ๐’ž are subsets
โ‹ƒ of ๐™, we conclude
โ‹ƒ
that ๐’ž ∈ P(๐™) (Definition 3.4.8). If ๐ด ∈ ๐’ž , then ๐ด ⊆ ๐’ž . Therefore, ๐’ž
is an upper bound of ๐’ž with respect to ⊆. To see that it is the least upper bound,
โ‹ƒ
let ๐‘ˆ be any upper bound of ๐’ž . Take ๐‘ฅ ∈
๐’ž . This means that there exists
๐ท ∈ ๐’ž such that ๐‘ฅ ∈ ๐ท. Since
๐‘ˆ
is
an
upper
bound of ๐’ž , ๐ท ⊆ ๐‘ˆ . Hence,
โ‹ƒ
๐‘ฅ ∈ ๐‘ˆ , and we conclude that ๐’ž ⊆ ๐‘ˆ .
Comparable and Compatible Elements
In Figure 4.5, we see that 01 4 010 and 01 4 011. However, 010 โ‹  011 and 011 โ‹ 
010. This means that in the poset of Example 4.3.7, there are pairs of elements that are
related to each other and there are other pairs that are not.
DEFINITION 4.3.15
Let (๐ด, 4) be a poset and ๐‘Ž, ๐‘ ∈ ๐ด. If ๐‘Ž 4 ๐‘ or ๐‘ 4 ๐‘Ž, then ๐‘Ž and ๐‘ are comparable with respect to 4. Elements of ๐ด are incomparable if they are not
comparable.
Continuing our review of the partially ordered set of Example 4.3.7, we note that
the element has the property that no element is less than it, but as seen in Figure 4.5,
for every element of ๐”ธ∗ , there exists an element of ๐”ธ∗ that is greater. However, that
same relation defined on
๐ด = {๐œŽ ∈ ๐”ธ∗ : ๐œŽ has at most 3 characters}
(4.3)
has the property that 000, 001, 010, 011, 100, 101, 110, 111 have no elements greater
than them. This leads to the next definition.
DEFINITION 4.3.16
Let (๐ด, 4) be a poset and ๐‘š ∈ ๐ด.
โˆ™ ๐‘š is a minimal element of ๐ด (with respect to 4) if ๐‘Ž โŠ€ ๐‘š for all ๐‘Ž ∈ ๐ด.
โˆ™ ๐‘š is a maximal element of ๐ด (with respect to 4) if ๐‘š โŠ€ ๐‘Ž for all ๐‘Ž ∈ ๐ด.
182
Chapter 4 RELATIONS AND FUNCTIONS
Therefore, the empty string is a minimal element of ๐ด (4.3), and 000, 001, 010, 011,
100, 101, 110, 111 are maximal. Notice that every least element is minimal and every
greatest element is maximal.
Although not every pair of elements is comparable in the partially ordered set of
Example 4.3.7, there are infinite sequences of comparable elements, such as
4 0 4 01 4 001 4 0001 4 · · · .
DEFINITION 4.3.17
A subset ๐ถ of the poset (๐ด, 4) is a chain with respect to 4 if ๐‘Ž is comparable to
๐‘ for all ๐‘Ž, ๐‘ ∈ ๐ถ.
When ๐™ is partially ordered by ≤, the sets {0, 1, 2, 3, … }, {… , −3, −2, −1, 0}, and
{… , −2, 0, 2, 4, … } are chains. In P(๐™), both
{{1, 2}, {1, 2, 3, 4}, {1, 2, 3, 4, 5, 6}}
and
{∅, {0}, {0, 1}, {0, 1, 2}, … }
are chains with respect to ⊆.
EXAMPLE 4.3.18
To see that {๐ด๐‘˜ : ๐‘˜ ∈ ๐™+ } where ๐ด๐‘˜ = {๐‘ฅ ∈ ๐™ : (๐‘ฅ − 1)(๐‘ฅ − 2) · · · (๐‘ฅ − ๐‘˜) = 0}
is a chain with respect to ⊆, take ๐‘š, ๐‘› ∈ ๐™+ . By definition,
๐ด๐‘š = {1, 2, … , ๐‘š}
and
๐ด๐‘› = {1, 2, … , ๐‘›}.
If ๐‘š ≤ ๐‘›, then ๐ด๐‘š ⊆ ๐ด๐‘› , otherwise ๐ด๐‘› ⊆ ๐ด๐‘š .
EXAMPLE 4.3.19
Let ๐ถ0 and ๐ถ1 be chains of ๐ด with respect to 4. Take ๐‘Ž, ๐‘ ∈ ๐ถ0 ∩ ๐ถ1 . Then,
๐‘Ž, ๐‘ ∈ ๐ถ0 , so ๐‘Ž 4 ๐‘ or ๐‘ 4 ๐‘Ž. Therefore, ๐ถ0 ∩ ๐ถ1 is a chain. However, the union
of two chains might not be a chain. For example, {{1}, {1, 2}} and {{1}, {1, 3}}
are chains in (P(๐™), ⊆), but {{1}, {1, 2}, {1, 3}} is not a chain because {1, 2} *
{1, 3} and {1, 3} * {1, 2}.
The sets ๐™, ๐, and ๐‘ are chains of ๐‘ with respect to ≤. In fact, any subset of ๐‘ is a
chain of ๐‘ because subsets of chains are chains. This motivates the next definition.
DEFINITION 4.3.20
The poset (๐ด, 4) is a linearly ordered set and 4 is a linear order if ๐ด is a chain
with respect to 4.
Section 4.3 PARTIAL ORDERS
183
Since every subset ๐ด of ๐‘ is a chain with respect to ≤, the relation ≤ is a linear order
on ๐ด. Furthermore, since every pair of elements in a linear order are comparable,
Theorem 4.3.8 can be strengthened.
THEOREM 4.3.21 [Trichotomy Law]
If 4 is a linear order on ๐ด, for all ๐‘Ž, ๐‘ ∈ ๐ด, exactly one of the following are true:
๐‘Ž โ‰บ ๐‘, ๐‘ โ‰บ ๐‘Ž, or ๐‘Ž = ๐‘.
Although it is not the case that every pair of elements of the poset defined in Example 4.3.7 is comparable, it is the case that for any given pair of elements, there exists
another element that is related to the given elements. For example, for the pair 100 and
110, 1 4 100 and 1 4 110. Also, for the pair 101 and 10, 10 4 101 and 10 4 10.
DEFINITION 4.3.22
Let (๐ด, 4) be a poset. The elements ๐‘Ž, ๐‘ ∈ ๐ด are compatible if there exists ๐‘ ∈ ๐ด
such that ๐‘ 4 ๐‘Ž and ๐‘ 4 ๐‘. If ๐‘Ž and ๐‘ are not compatible, they are incompatible
and we write ๐‘Ž โŸ‚ ๐‘.
Observe that if ๐‘Ž and ๐‘ are comparable, they are also compatible. On the other hand, it
takes some work to define a relation in which every pair of elements is incompatible.
DEFINITION 4.3.23
A subset ๐ท of a poset (๐ด, 4) is an antichain with respect to 4 when for all
๐‘Ž, ๐‘ ∈ ๐ท, if ๐‘Ž ≠ ๐‘, then ๐‘Ž โŸ‚ ๐‘.
The sets
{{๐‘›} : ๐‘› ∈ ๐™}
and
{{1}, {2, 3}, {4, 5, 6}, {7, 8, 9, 10}, … }
are antichains of P(๐™) with respect to ⊆.
Well-Ordered Sets
The notion of a linear order incorporates many of the properties of ≤ on ๐ since ≤ is
reflexive, antisymmetric, and transitive, and ๐ is a chain with respect to ≤. However,
there is one important property of (๐, ≤) that is not included among those of a linear
order. Because of the nature of the natural numbers, every subset of ๐ that is nonempty
has a least element. For example, 5 is the least element of {5, 7, 32, 99} and 2 is the
least element of {2, 4, 6, 8, … }. We want to be able to identify those partial orders that
also have this property.
DEFINITION 4.3.24
The linearly ordered set (๐ด, 4) is a well-ordered set and 4 is a well-order if
every nonempty subset of ๐ด has a least element with respect to 4.
184
Chapter 4 RELATIONS AND FUNCTIONS
According to Definition 4.3.24, (๐, ≤) is a well-ordered set, but this fact about the
natural numbers cannot be proved without making an assumption. Therefore, so that
we have at least one well-ordered set with which to work, we assume the following.
AXIOM 4.3.25
(๐, ≤) is a well-ordered set.
Coupling Axiom 4.3.25 with the next theorem will yield infinitely many well-ordered
sets.
THEOREM 4.3.26
If (๐ด, 4) is well-ordered and ๐ต is a nonempty subset of ๐ด, then (๐ต, 4) is wellordered.
PROOF
Let ๐ต ⊆ ๐ด and ๐ต ≠ ∅. To prove that ๐ต is well-ordered, let ๐ถ ⊆ ๐ต and ๐ถ ≠ ∅.
Then, ๐ถ ⊆ ๐ด. Since 4 well-orders ๐ด, we know that ๐ถ has a least element with
respect to 4.
Because ๐™ ∩ [5, ∞) ⊆ ๐™+ ⊆ ๐, by Axiom 4.3.25 and Theorem 4.3.26, both (๐™+ , ≤)
and (๐™ ∩ [5, ∞) , ≤) are well-ordered sets.
EXAMPLE 4.3.27
Let ๐ด = {๐‘›๐œ‹ : ๐‘› ∈ ๐}. To prove that ๐ด is well-ordered by ≤, let ๐ต ⊆ ๐ด such
that ๐ต ≠ ∅. This means that there exists a nonempty subset ๐ผ of ๐ such that
๐ต = {๐‘›๐œ‹ : ๐‘› ∈ ๐ผ}. Since ๐ is well-ordered, ๐ผ has least element ๐‘š. We claim
that ๐‘š๐œ‹ is the least element of ๐ต. To see this, take ๐‘ ∈ ๐ต. Then, ๐‘ = ๐‘–๐œ‹ for some
๐‘– ∈ ๐ผ. Since ๐‘š is the least element of ๐ผ, ๐‘š ≤ ๐‘–. Therefore, ๐‘š๐œ‹ ≤ ๐‘–๐œ‹ = ๐‘.
To prove that a set ๐ด is not well-ordered by 4, we must find a nonempty subset ๐ต
of ๐ด that does not have a least element. This means that for every ๐‘ ∈ ๐ต, there exists
๐‘ ∈ ๐ต such that ๐‘ โ‰บ ๐‘. That is, there are elements ๐‘๐‘› ∈ ๐ต (๐‘› ∈ ๐) such that
· · · โ‰บ ๐‘2 โ‰บ ๐‘1 โ‰บ ๐‘0 .
This informs the next definition.
DEFINITION 4.3.28
Let 4 be a partial order on ๐ด and ๐ต = {๐‘Ž๐‘– : ๐‘– ∈ ๐} be a subset of ๐ด.
โˆ™ ๐ต is increasing means ๐‘– < ๐‘— implies ๐‘Ž๐‘– โ‰บ ๐‘Ž๐‘— for all ๐‘–, ๐‘— ∈ ๐.
โˆ™ ๐ต is decreasing means ๐‘– < ๐‘— implies ๐‘Ž๐‘— โ‰บ ๐‘Ž๐‘– for all ๐‘–, ๐‘— ∈ ๐.
If a set is well-ordered, it has a least element, but the converse is not true. To see
this, consider ๐ด = {0, 1โˆ•2, 1โˆ•3, 1โˆ•4, … }. It has a least element, namely, 0, but ๐ด also
Section 4.3 PARTIAL ORDERS
contains the decreasing set
185
{
}
1 1 1 1
, , , ,… .
(4.4)
2 3 4 5
Therefore, ๐ด has a subset without a least element, so ๐ด is not well-ordered. We summarize this observation with the following theorem, and leave its proof to Exercise 23.
THEOREM 4.3.29
(๐ด, 4) is not a well-ordered set if and only if (๐ด, 4) does not have a decreasing
subset.
Theorem 4.3.29 implies that any finite linear order is well-ordered.
EXAMPLE 4.3.30
The decreasing sequence (4.4) with Theorem 4.3.29 shows that the sets (0, 1),
[0, 1], ๐, and {1โˆ•๐‘› : ๐‘› ∈ ๐™+ } are not well-ordered by ≤.
We close this section by proving two important results from number theory. Their
proofs use Axiom 4.3.25. The strategy is to define a nonempty subset of a well-ordered
set. Its least element ๐‘Ÿ will be a number that we want. This least element also needs
to have a particular property, say ๐‘(๐‘Ÿ). To show that it has the property, assume ¬๐‘(๐‘Ÿ)
and use this to find another element of the set that is less than ๐‘Ÿ. This contradicts the
minimality of ๐‘Ÿ allowing us to conclude ๐‘(๐‘Ÿ).
THEOREM 4.3.31 [Division Algorithm]
If ๐‘š, ๐‘› ∈ ๐ with ๐‘š ≠ 0, there exist unique ๐‘ž, ๐‘Ÿ ∈ ๐ such that ๐‘Ÿ < ๐‘š and
๐‘› = ๐‘š๐‘ž + ๐‘Ÿ.
PROOF
Uniqueness is proved in Example 2.4.14. To prove existence, take ๐‘š, ๐‘› ∈ ๐ and
define
๐‘† = {๐‘˜ ∈ ๐ : ∃๐‘™(๐‘™ ∈ ๐ ∧ ๐‘› = ๐‘š๐‘™ + ๐‘˜)}.
Notice that ๐‘› ∈ ๐‘†, so ๐‘† ≠ ∅. Therefore, ๐‘† has a least element by Axiom 4.3.25.
Call it ๐‘Ÿ and write ๐‘› = ๐‘š๐‘ž + ๐‘Ÿ for some natural number ๐‘ž. Assume ๐‘Ÿ ≥ ๐‘š, which
implies that ๐‘Ÿ − ๐‘š ≥ 0. Also,
๐‘› = ๐‘š(๐‘ž − 1) + ๐‘Ÿ − ๐‘š
because
๐‘Ÿ − ๐‘š = ๐‘› − ๐‘š๐‘ž − ๐‘š = ๐‘› − ๐‘š(๐‘ž + 1),
so ๐‘Ÿ − ๐‘š ∈ ๐‘†. Since ๐‘Ÿ > ๐‘Ÿ − ๐‘š because ๐‘š is positive, ๐‘Ÿ cannot be the minimum
of ๐‘†, a contradiction.
The value ๐‘ž of the division algorithm (Theorem 4.3.31) is called the quotient and ๐‘Ÿ is
the remainder. For example, if we divide 5 into 17, the division algorithm returns a
quotient of 3 and a remainder of 2, so we can write that 17 = 5(3) + 2. Notice that
2 < 3.
186
Chapter 4 RELATIONS AND FUNCTIONS
Call ๐‘› ∈ ๐™ a linear combination of the integers ๐‘Ž and ๐‘ if ๐‘› = ๐‘ข๐‘Ž + ๐‘ฃ๐‘ for some
๐‘ข, ๐‘ฃ ∈ ๐™. Since 37 = 5(2) + 3(9), we see that 37 is a linear combination of 2 and 9.
Furthermore, if ๐‘‘ โˆฃ ๐‘Ž and ๐‘‘ โˆฃ ๐‘, then ๐‘‘ โˆฃ ๐‘ข๐‘Ž + ๐‘ฃ๐‘. To see this, write ๐‘Ž = ๐‘‘๐‘™ and ๐‘ = ๐‘‘๐‘˜
for some ๐‘™, ๐‘˜ ∈ ๐™. Then,
๐‘ข๐‘Ž + ๐‘ฃ๐‘ = ๐‘ข๐‘‘๐‘™ + ๐‘ฃ๐‘‘๐‘˜ = ๐‘‘(๐‘ข๐‘™ + ๐‘ฃ๐‘˜),
and this means ๐‘‘ โˆฃ ๐‘ข๐‘Ž + ๐‘ฃ๐‘.
THEOREM 4.3.32
Let ๐‘Ž, ๐‘ ∈ ๐™ with not both equal to 0. If ๐‘ = gcd(๐‘Ž, ๐‘), there exists ๐‘š, ๐‘› ∈ ๐™ such
that ๐‘ = ๐‘š๐‘Ž + ๐‘›๐‘.
PROOF
Define
๐‘‡ = {๐‘ง ∈ ๐™+ : ∃๐‘ฅ∃๐‘ฆ(๐‘ฅ, ๐‘ฆ ∈ ๐™ ∧ ๐‘ง = ๐‘ฅ๐‘Ž + ๐‘ฆ๐‘)}.
Notice that ๐‘‡ is not empty because ๐‘Ž2 + ๐‘2 ∈ ๐‘‡ . By Axiom 4.3.25, ๐‘‡ has a least
element ๐‘‘, so write ๐‘‘ = ๐‘š๐‘Ž + ๐‘›๐‘ for some ๐‘š, ๐‘› ∈ ๐™.
โˆ™ Since ๐‘‘ > 0, the division algorithm (4.3.31) yields ๐‘Ž = ๐‘‘๐‘ž + ๐‘Ÿ for some
natural numbers ๐‘ž and ๐‘Ÿ with ๐‘Ÿ < ๐‘‘. Then,
๐‘Ÿ = ๐‘Ž − ๐‘‘๐‘ž = ๐‘Ž − (๐‘š๐‘Ž − ๐‘›๐‘)๐‘ž = (1 − ๐‘š๐‘ž)๐‘Ž + (๐‘›๐‘ž)๐‘.
If ๐‘Ÿ > 0, then ๐‘Ÿ ∈ ๐‘‡ , which is impossible because ๐‘‘ is the least element of
๐‘‡ . Therefore, ๐‘Ÿ = 0 and ๐‘‘ โˆฃ ๐‘Ž. Similarly, ๐‘‘ โˆฃ ๐‘.
โˆ™ To show that ๐‘‘ is the greatest of the common divisors, suppose ๐‘  โˆฃ ๐‘Ž and
๐‘  โˆฃ ๐‘ with ๐‘  ∈ ๐™+ . By definition, ๐‘Ž = ๐‘ ๐‘˜ and ๐‘ = ๐‘ ๐‘™ for some ๐‘˜, ๐‘™ ∈ ๐™.
Hence,
๐‘‘ = ๐‘š(๐‘ ๐‘˜) + ๐‘›(๐‘ ๐‘™) = ๐‘ (๐‘š๐‘˜ − ๐‘›๐‘™).
Thus, ๐‘  ≤ ๐‘‘ because ๐‘  is nonzero and ๐‘  divides ๐‘‘ (Exercise 25).
Exercises
1. Is (∅, ⊆) a partial order, linear order, or well order? Explain.
2. For each relation on {1, 2}, determine if it is reflexive, irreflexive, symmetric, asymmetric, antisymmetric, or transitive.
(a) {(1, 2)}
(b) {(1, 2), (2, 1)}
(c) {(1, 1), (1, 2), (2, 1)}
(d) {(1, 1), (1, 2), (2, 2)}
(e) {(1, 1), (1, 2), (2, 1), (2, 2)}
(f) ∅
3. Give an example of a relation that is neither symmetric nor asymmetric.
Section 4.3 PARTIAL ORDERS
187
4. Let ๐‘… be a relation on ๐ด. Prove that ๐‘… is reflexive if and only if (๐ด × ๐ด) โงต ๐‘… is
irreflexive.
5. Show that a relation ๐‘… on ๐ด is asymmetric if and only if ๐‘… ∩ ๐‘…−1 = ∅.
6. Let (๐ด, 4) and (๐ต, ≤) be posets. Define ∼ on ๐ด × ๐ต by
(๐‘Ž, ๐‘) ∼ (๐‘Ž′ , ๐‘′ ) if and only if ๐‘Ž 4 ๐‘Ž′ and ๐‘ ≤ ๐‘′ .
Show that ∼ is a partial order on ๐ด × ๐ต.
7. For any alphabet ๐”ธ, prove the following.
(a) For all ๐œŽ, ๐œ, ๐œˆ ∈ ๐”ธ∗ , ๐œŽ โŒข (๐œ โŒข ๐œˆ) = (๐œŽ โŒข ๐œ)โŒข ๐œˆ.
(b) There exists ๐œŽ, ๐œ ∈ ๐”ธ∗ such that ๐œŽ โŒข ๐œ ≠ ๐œ โŒข ๐œŽ.
(c) ๐”ธ∗ has an identity with respect to โŒข, but for all ๐œŽ ∈ ๐”ธ∗ , there is no inverse
for ๐œŽ if ๐œŽ ≠ .
8. Prove that ๐”ธ∗ from Example 4.3.7 is partially ordered by 4.
9. Prove that (P(๐ด), ⊆) is not a linear order if ๐ด has at least three elements.
10. Show that (๐”ธ∗ , 4) is not a linear order if ๐”ธ has at least two elements.
11. Prove that the following families of sets are chains with respect to ⊆.
(a) {[0, ๐‘›] : ๐‘› ∈ ๐™+ }
(b) {(2๐‘› )๐™ : ๐‘› ∈ ๐} where (2๐‘› )๐™ = {2๐‘› ⋅ ๐‘˜ : ๐‘˜ ∈ ๐™}
โ‹ƒ
(c) {๐ต๐‘› : ๐‘› ∈ ๐} where ๐ต๐‘› = {๐ด๐‘– : ๐‘– ∈ ๐ ∧ ๐‘– ≤ ๐‘›} and ๐ด๐‘– is a set for all
๐‘–∈๐
12. Can a chain be disjoint or pairwise disjoint? Explain.
13. Suppose that {๐ด๐‘– : ๐‘– ∈ ๐} is a chain of sets such that for all ๐‘– ≤ ๐‘—, ๐ด๐‘– ⊆ ๐ด๐‘— . Prove
for all ๐‘˜ ∈โ‹ƒ
๐.
(a)
{๐ด โˆถ ๐‘– ∈ ๐ ∧ ๐‘– ≤ ๐‘˜} = ๐ด๐‘˜
โ‹‚ ๐‘–
(b)
{๐ด๐‘– โˆถ ๐‘– ∈ ๐ ∧ ๐‘– ≤ ๐‘˜} = ๐ด0
โ‹ƒ
14. Let {๐ด๐‘› : ๐‘› ∈ ๐} be a family of sets. For every ๐‘š ∈ ๐, define ๐ต๐‘š = ๐‘š
๐‘–=0 ๐ด๐‘– .
Show that {๐ต๐‘š : ๐‘š ∈ ๐} is a chain.
โ‹‚
15. Let ๐’ž๐‘– be a chain of the poset (๐ด, 4) for all ๐‘– ∈ ๐ผ. Prove that ๐‘–∈๐ผ ๐’ž๐‘– is a chain.
โ‹ƒ
Is ๐‘–∈๐ผ ๐’ž๐‘– necessarily a chain? Explain.
16. Let (๐ด, 4) and (๐ต, ≤) be linear orders. Define ∼ on ๐ด × ๐ต by
(๐‘Ž, ๐‘) ∼ (๐‘Ž′ , ๐‘′ )
if and only if
๐‘Ž โ‰บ ๐‘Ž′ or ๐‘Ž = ๐‘Ž′ and ๐‘ ≤ ๐‘′ .
This relation is called a lexicographical order since it copies the order of a dictionary.
(a) Prove that (๐ด × ๐ต, ∼) is a linear order.
(b) Suppose that (๐‘Ž, ๐‘) is a maximal element of ๐ด × ๐ต with respect to ∼. Show
that ๐‘Ž is a maximal element of ๐ด with respect to 4.
188
Chapter 4 RELATIONS AND FUNCTIONS
17. Let ๐ด be a set. Prove that ๐ต ⊆ P(๐ด) โงต {∅} is an antichain with respect to ⊆ if and
only if ๐ต is pairwise disjoint.
18. Prove the following true or false.
(a) Every well-ordered set contains a least element.
(b) Every well-ordered set contains a greatest element.
(c) Every subset of a well-ordered set contains a least element.
(d) Every subset of a well-ordered set contains a greatest element.
(e) Every well-ordered set has a decreasing subset.
(f) Every well-ordered set has a increasing subset.
19. For each of the given sets, indicate whether or not it is well-ordered by ≤. If it is,
prove it. If it is not, find a decreasing sequence of elements of the set.
√
(a) { 2, 5, 6, 10.56, 17, −100}
(b) {2๐‘› : ๐‘› ∈ ๐}
(c) {๐œ‹โˆ•๐‘› : ๐‘› ∈ ๐™+ }
(d) {๐œ‹โˆ•๐‘› : ๐‘› ∈ ๐™− }
(e) {−4, −3, −2, −1, … }
(f) ๐™ ∩ (๐œ‹, ∞)
(g) ๐™ ∩ (−7, ∞)
20. Prove that (๐™ ∩ (๐‘ฅ, ∞), ≤) is a well-ordered set for all ๐‘ฅ ∈ ๐‘.
21. Show that a well-ordered set has a unique least element.
22. Let (๐ด, 4) be a well-ordered set. If ๐ต ⊆ ๐ด and there is an upper bound for ๐ต in ๐ด,
then ๐ต has a greatest element.
23. Prove Theorem 4.3.29.
24. Where does the proof of Theorem 4.3.31 go wrong if ๐‘š = 0?
25. Prove that if ๐‘Ž โˆฃ ๐‘, then ๐‘Ž ≤ ๐‘ for all ๐‘Ž, ๐‘ ∈ ๐™.
26. Let ๐‘Ž, ๐‘, ๐‘ ∈ ๐™. Show that if gcd(๐‘Ž, ๐‘) = gcd(๐‘, ๐‘) = 1, then gcd(๐‘Ž, ๐‘๐‘) = 1.
27. Let ๐‘Ž, ๐‘ ∈ ๐™ and assume that gcd(๐‘Ž, ๐‘) = 1. Prove.
(a) If ๐‘Ž โˆฃ ๐‘› and ๐‘ โˆฃ ๐‘›, then ๐‘Ž๐‘ โˆฃ ๐‘›.
(b) gcd(๐‘Ž + ๐‘, ๐‘) = gcd(๐‘Ž + ๐‘, ๐‘Ž) = 1.
(c) gcd(๐‘Ž + ๐‘, ๐‘Ž − ๐‘) = 1 or gcd(๐‘Ž + ๐‘, ๐‘Ž − ๐‘) = 2.
(d) If ๐‘ โˆฃ ๐‘Ž, then gcd(๐‘, ๐‘) = 1.
(e) If ๐‘ โˆฃ ๐‘Ž + ๐‘, then gcd(๐‘Ž, ๐‘) = gcd(๐‘, ๐‘) = 1.
(f) If ๐‘‘ โˆฃ ๐‘Ž๐‘ and ๐‘‘ โˆฃ ๐‘๐‘, then ๐‘‘ โˆฃ ๐‘.
28. Let ๐‘Ž, ๐‘ ∈ ๐™, where at least one is nonzero. Prove that ๐‘† = ๐‘‡ if
๐‘† = {๐‘ฅ : ∃๐‘™[๐‘™ ∈ ๐™ ∧ ๐‘ฅ = ๐‘™ gcd(๐‘Ž, ๐‘)]}
and
๐‘‡ = {๐‘ฅ : ๐‘ฅ > 0 ∧ ∃๐‘ข∃๐‘ฃ(๐‘ข, ๐‘ฃ ∈ ๐™ ∧ ๐‘ฅ = ๐‘ข๐‘Ž + ๐‘ฃ๐‘)}.
Section 4.4 FUNCTIONS
189
29. Prove that if ๐‘‘ โˆฃ ๐‘Ž and ๐‘‘ โˆฃ ๐‘, then ๐‘‘ โˆฃ gcd(๐‘Ž, ๐‘) for all ๐‘Ž, ๐‘, ๐‘‘ ∈ ๐™.
4.4
FUNCTIONS
From algebra and calculus, we know what a function is. It is a rule that assigns to each
possible input value a unique output value. The common picture is that of a machine
that when a certain button is pushed, the same result always happens. In basic algebra
a relation can be graphed in the Cartesian plane. Such a relation will be a function if
and only if every vertical line intersects its graphs at most once. This is known as the
vertical line test (Figure 4.6). This criteria is generalized in the next definition.
DEFINITION 4.4.1
Let ๐ด and ๐ต be sets. A relation ๐‘“ ⊆ ๐ด × ๐ต is a function means that for all
(๐‘ฅ, ๐‘ฆ), (๐‘ฅ′ , ๐‘ฆ′ ) ∈ ๐ด × ๐ต,
if ๐‘ฅ = ๐‘ฅ′ , then ๐‘ฆ = ๐‘ฆ′ .
The function ๐‘“ is an n-ary function if there exists sets ๐ด0 , ๐ด1 , … , ๐ด๐‘›−1 such that
๐ด = ๐ด0 × ๐ด1 × · · · × ๐ด๐‘›−1 . If ๐‘› = 1, then ๐‘“ is a unary function, and if ๐‘› = 2,
then ๐‘“ is a binary function,
EXAMPLE 4.4.2
The set {(1, 2), (4, 5), (6, 5)} is a function, but {(1, 2), (1, 5), (6, 5)} is not since it
contains (1, 2) and (1, 5). Also, ∅ is a function (Exercise 16).
Intersects at most once
Figure 4.6
Passing the vertical line test.
190
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.4.3
Define ๐‘“ = {(๐‘ข, 1 + 2 cos ๐œ‹๐‘ข) : ๐‘ข ∈ ๐‘}. Assume that both (๐‘ฅ, 1 + 2 cos ๐œ‹๐‘ฅ) and
(๐‘ฅ′ , 1 + 2 cos ๐œ‹๐‘ฅ′ ) are elements of ๐‘“ . Then, because cosine is a function,
๐‘ฅ = ๐‘ฅ′ ⇒ ๐œ‹๐‘ฅ = ๐œ‹๐‘ฅ′
⇒ cos ๐œ‹๐‘ฅ = cos ๐œ‹๐‘ฅ′
⇒ 2 cos ๐œ‹๐‘ฅ = 2 cos ๐œ‹๐‘ฅ′
⇒ 1 + 2 cos ๐œ‹๐‘ฅ = 1 + 2 cos ๐œ‹๐‘ฅ′ .
Therefore, ๐‘“ is a function.
EXAMPLE 4.4.4
The standard arithmetic operations are functions. For example, taking a square
root is a unary function, while addition, subtraction, multiplication, and division
are binary functions. To illustrate this, addition on ๐™ is the set
๐ด = {((๐‘Ž, ๐‘), ๐‘Ž + ๐‘) : ๐‘Ž, ๐‘ ∈ ๐™}.
(4.5)
Let ๐‘“ ⊆ ๐ด × ๐ต be a function. This implies that for all ๐‘Ž ∈ ๐ด, either [๐‘Ž]๐‘“ = ∅ or
[๐‘Ž]๐‘“ is a singleton (Definition 4.2.7). For example, if ๐‘“ = {(1, 2), (4, 5), (6, 5)}, then
[1]๐‘“ = {2} and [3]๐‘“ = ∅. Because [๐‘Ž]๐‘“ can contain at most one element, we typically
simplify the notation.
DEFINITION 4.4.5
Let ๐‘“ be a function. For all ๐‘Ž ∈ ๐ด, define
๐‘“ (๐‘Ž) = ๐‘ if and only if [๐‘Ž]๐‘“ = {๐‘}
and write that ๐‘“ (๐‘Ž) is undefined if [๐‘Ž]๐‘“ = ∅.
For example, using (4.5),
๐ด((5, 7)) = 12,
but ๐ด(5, ๐œ‹) is undefined. With ordered ๐‘›-tuples, the outer parentheses are usually eliminated so that we write
๐ด(5, 7) = 12.
Moreover, if ๐ท ⊆ ๐ด is the domain of the function ๐‘“ , write the function notation
๐‘“ :๐ท→๐ต
and call ๐ต a codomain of ๐‘“ (Figure 4.7). Because functions are often represented by
arrows that “send” one element to another, a function can be called a map. If ๐‘“ (๐‘ฅ) = ๐‘ฆ,
we can say that “๐‘“ maps ๐‘ฅ to ๐‘ฆ.” We can also say that ๐‘ฆ is the image of ๐‘ฅ under ๐‘“ and
๐‘ฅ is a pre-image of ๐‘ฆ. For example, if
๐‘“ = {(3, 1), (4, 2), (5, 2)},
Section 4.4 FUNCTIONS
191
Domain
f:D→B
Codomain
Function name
Figure 4.7
Function notation.
๐‘“ maps 3 to 1, 2 is the image of 4, and 5 is a pre-image of 2 (Figure 4.8). If ๐‘” is also a
function with domain ๐ท and codomain ๐ต, we can use the abbreviation
๐‘“, ๐‘” : ๐ท → ๐ต
to represent both functions. An alternate choice of notation involves referring to the
functions as ๐ท → ๐ต.
EXAMPLE 4.4.6
If ๐‘“ (๐‘ฅ) = cos
√๐‘ฅ, then ๐‘“ maps ๐œ‹ to −1, 0 is the image of ๐œ‹โˆ•2, and ๐œ‹โˆ•4 is a
pre-image of 2โˆ•2.
EXAMPLE 4.4.7
Let ๐ด be any set. The identity relation on ๐ด (Definition 4.1.2) is a function, so
call ๐ผ๐ด the identity map and write
๐ผ๐ด (๐‘ฅ) = ๐‘ฅ.
Sometimes a function ๐‘“ is defined using a rule that pairs an element of the function’s
domain with an element of its codomain. When this is done successfully, ๐‘“ is said to
be well-defined. Observe that proving that ๐‘“ is well-defined is the same as proving it
to be a function.
f
3
4
5
Figure 4.8
1
2
The map ๐‘“ = {(3, 1), (4, 2), (5, 2)}.
192
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.4.8
Let ๐‘“ : ๐‘ → ๐‘ be defined by ๐‘“ (๐‘ฅ) = 1 + 2 cos ๐œ‹๐‘ฅ. This is the function
๐‘“ = {(๐‘ฅ, 1 + 2 cos ๐œ‹๐‘ฅ) : ๐‘ฅ ∈ ๐‘}.
The work of Example 4.4.3 shows that ๐‘“ is well-defined.
Before we examine another example, let us set a convention on naming functions.
It is partly for aesthetics, but it does help in organizing functions based on the type of
elements in their domains and ranges.
โˆ™ Use English letters (usually, ๐‘“ , ๐‘”, and โ„Ž) for naming functions that involve numbers. Typically, these will be lowercase, but there are occasions when we will
choose them to be uppercase.
โˆ™ Use Greek letters (often ๐œ‘ or ๐œ“) for general functions or those with domains not
consisting of numbers. They are also usually lowercase, but uppercase Greek
letters like Φ and Ψ are sometimes appropriate. (See the appendix for the Greek
alphabet.)
EXAMPLE 4.4.9
Let ๐‘›, ๐‘š ∈ ๐™+ such that ๐‘š โˆฃ ๐‘›. Define ๐œ‘([๐‘Ž]๐‘› ) = [๐‘Ž]๐‘š for all ๐‘Ž ∈ ๐™. This means
that
๐œ‘ = {([๐‘Ž]๐‘› , [๐‘Ž]๐‘š ) : ๐‘Ž ∈ ๐™}.
It is not clear that ๐œ‘ is well-defined since an equivalence class can have many
representatives, so assume that [๐‘Ž]๐‘› = [๐‘]๐‘› for ๐‘Ž, ๐‘ ∈ ๐™. Therefore, ๐‘› โˆฃ ๐‘Ž − ๐‘.
Then, by hypothesis, ๐‘š โˆฃ ๐‘Ž − ๐‘, and this yields
๐œ‘([๐‘Ž]๐‘› ) = [๐‘Ž]๐‘š = [๐‘]๐‘š = ๐œ‘([๐‘]๐‘› ).
EXAMPLE 4.4.10
Let ๐‘ฅ ∈ ๐‘. Define the greatest integer function as
โŸฆ๐‘ฅโŸง = the greatest integer ≤ ๐‘ฅ.
For example, โŸฆ5โŸง = 5, โŸฆ1.4โŸง = 1, and โŸฆ−3.4โŸง = −4. The greatest integer
function is a function ๐‘ → ๐™. It is well-defined because the relation ≤ wellorders ๐™ (Exercise 4.3.22).
If a relation on ๐‘ is not a function, it will fail the vertical line test (Figure 4.9). To
generalize, let ๐œ‘ ⊆ ๐ด × ๐ต. To show that ๐œ‘ is not a function, we must show that there
exists (๐‘ฅ, ๐‘ฆ1 ), (๐‘ฅ, ๐‘ฆ2 ) ∈ ๐œ‘ such that ๐‘ฆ1 ≠ ๐‘ฆ2 .
EXAMPLE 4.4.11
The relation ๐‘“ = ๐ผ๐‘ ∪ {(๐‘ฅ, −๐‘ฅ) : ๐‘ฅ ∈ ๐‘} is not a function since (4, 4) ∈ ๐‘“ and
(4, −4) ∈ ๐‘“ .
Section 4.4 FUNCTIONS
y1
193
(x, y1)
x
y2
Figure 4.9
(x, y2)
The relation is not a function.
EXAMPLE 4.4.12
Define ๐œ‘ ⊆ ๐™2 × ๐™3 by ๐œ‘([๐‘Ž]2 ) = [๐‘Ž]3 for all ๐‘Ž ∈ ๐™. Since 3 does not divide 2,
๐œ‘ is not well-defined. This is proved by noting that [0]2 = [2]2 but [0]3 ≠ [2]3 .
There will be times when we want to examine sets of functions. If each function is
to have the same domain and codomain, we use the following notation.
DEFINITION 4.4.13
If ๐ด and ๐ต are sets,
๐ด
๐ต = {๐œ‘ : ๐œ‘ is a function ๐ด → ๐ต}.
For instance, ๐ด ๐ต is a set of real-valued functions if ๐ด, ๐ต ⊆ ๐‘.
EXAMPLE 4.4.14
If ๐‘Ž๐‘› is a sequence of real numbers with ๐‘› = 0, 1, 2, … , the sequence ๐‘Ž๐‘› is an
element of ๐ ๐‘. Illustrating this, the sequence
๐‘Ž๐‘› = (−1โˆ•2)๐‘›
is a function and can be graphed as in Figure 4.10.
EXAMPLE 4.4.15
Let ๐ด ⊆ ๐‘ and fix ๐‘Ž ∈ ๐ด. The evaluation map,
๐œ–๐‘Ž : ๐ด๐ด → ๐ด,
is defined as
๐œ–๐‘Ž (๐‘“ ) = ๐‘“ (๐‘Ž).
194
Chapter 4 RELATIONS AND FUNCTIONS
Figure 4.10
๐‘Ž๐‘› = (−1โˆ•2)๐‘› is a function.
For example, if ๐‘”(๐‘ฅ) = ๐‘ฅ2 , then ๐œ–3 (๐‘”) = 9. Observe that the evaluation map is an
๐ด
element of ( ๐ด)๐ด.
Equality
Since functions are sets, we already know that two functions are equal when they contain the same ordered pairs. However, there is a common test to determine function
equality other than a direct appeal to Definition 3.3.7.
THEOREM 4.4.16
Functions ๐œ‘, ๐œ“ : ๐ด → ๐ต are equal if and only if ๐œ‘(๐‘ฅ) = ๐œ“(๐‘ฅ) for all ๐‘ฅ ∈ ๐ด.
PROOF
Sufficiency is clear, so to prove necessity suppose that ๐œ‘(๐‘ฅ) = ๐œ“(๐‘ฅ) for every
๐‘ฅ ∈ ๐ด. Take (๐‘Ž, ๐‘) ∈ ๐œ‘. This means that ๐œ‘(๐‘Ž) = ๐‘. By hypothesis, ๐œ“(๐‘Ž) = ๐‘,
which implies that (๐‘Ž, ๐‘) ∈ ๐œ“. Hence, ๐œ‘ ⊆ ๐‘”. The proof of ๐œ“ ⊆ ๐œ‘ is similar, so
๐œ‘ = ๐‘”.
We use Theorem 4.4.16 in the next examples.
EXAMPLE 4.4.17
Let ๐‘“ and ๐‘” be functions ๐‘ → ๐‘ defined by
๐‘“ (๐‘ฅ) = (๐‘ฅ − 3)2 + 2
and
๐‘”(๐‘ฅ) = ๐‘ฅ2 − 6๐‘ฅ + 11.
We show that ๐‘“ = ๐‘” by taking ๐‘ฅ ∈ ๐‘ and calculating
๐‘“ (๐‘ฅ) = (๐‘ฅ − 3)2 + 2 = (๐‘ฅ2 − 6๐‘ฅ + 9) + 2 = ๐‘ฅ2 − 6๐‘ฅ + 11 = ๐‘”(๐‘ฅ).
Section 4.4 FUNCTIONS
195
EXAMPLE 4.4.18
Let ๐œ‘, ๐œ“ : ๐™ → ๐™6 be functions such that
๐œ‘(๐‘›) = [๐‘›]6
and
๐œ“(๐‘›) = [๐‘› + 12]6 .
Take ๐‘› ∈ ๐™. We show that [๐‘› + 12]6 = [๐‘›]6 by proceeding as follows:
๐‘ฅ ∈ [๐‘› + 12]6 ⇔ ∃๐‘˜ [๐‘˜ ∈ ๐™ ∧ ๐‘ฅ = ๐‘› + 12 + 6๐‘˜]
⇔ ∃๐‘˜ [๐‘˜ ∈ ๐™ ∧ ๐‘ฅ = ๐‘› + 6(2 + ๐‘˜)]
⇔ ๐‘ฅ ∈ [๐‘›]6 .
Therefore, ๐œ‘ = ๐œ“.
EXAMPLE 4.4.19
Define
๐œ“ :๐‘→
(๐‘ )
๐‘
๐‘
by ๐œ“(๐‘ฅ) = ๐œ–๐‘ฅ for all ๐‘ฅ ∈ ๐‘ (Example 4.4.15). We show that ๐œ“ is well-defined.
Take ๐‘Ž, ๐‘ ∈ ๐‘ and assume that ๐‘Ž = ๐‘. To show that ๐œ“(๐‘Ž) = ๐œ“(๐‘), we prove
๐œ–๐‘Ž = ๐œ–๐‘ . Therefore, let ๐‘“ ∈ ๐‘ ๐‘. Since ๐‘“ is a function and ๐‘Ž, ๐‘ ∈ dom(๐‘“ ), we
have that ๐‘“ (๐‘Ž) = ๐‘“ (๐‘). Thus,
๐œ–๐‘Ž (๐‘“ ) = ๐‘“ (๐‘Ž) = ๐‘“ (๐‘) = ๐œ–๐‘ (๐‘“ ).
By Theorem 4.4.16, we see that two functions ๐‘“ and ๐‘” are not equal when either
dom(๐‘“ ) ≠ dom(๐‘”) or ๐‘“ (๐‘ฅ) ≠ ๐‘”(๐‘ฅ) for some ๐‘ฅ in their common domain. For example,
๐‘“ (๐‘ฅ) = ๐‘ฅ2 and ๐‘”(๐‘ฅ) = 2๐‘ฅ are not equal because ๐‘“ (3) = 9 and ๐‘”(3) = 6. Although
these two functions differ for every ๐‘ฅ ≠ 0 and ๐‘ฅ ≠ 2, it only takes one inequality to
prove that the functions are not equal. For example, if we define
{
๐‘ฅ2 if ๐‘ฅ ≠ 0,
โ„Ž(๐‘ฅ) =
7 if ๐‘ฅ = 0,
then ๐‘“ ≠ โ„Ž since ๐‘“ (0) = 0 and โ„Ž(0) = 7.
Composition
We now consider the composition of relations when those relations are functions.
THEOREM 4.4.20
If ๐œ‘ : ๐ด → ๐ต and ๐œ“ : ๐ถ → ๐ท are functions such that ran(๐œ‘) ⊆ ๐ถ, then ๐œ“ โ—ฆ ๐œ‘ is
a function ๐ด → ๐ท and (๐œ“ โ—ฆ ๐œ‘)(๐‘ฅ) = ๐œ“(๐œ‘(๐‘ฅ)).
196
Chapter 4 RELATIONS AND FUNCTIONS
PROOF
Because the range of ๐œ‘ is a subset of ๐ถ, we know that ๐œ‘ ⊆ ๐ด× ๐ถ and ๐œ“ ⊆ ๐ถ × ๐ท.
Let (๐‘Ž, ๐‘‘1 ), (๐‘Ž, ๐‘‘2 ) ∈ ๐œ“ โ—ฆ ๐œ‘. This means by Definition 4.1.9 that there exists
๐‘1 , ๐‘2 ∈ ๐ถ such that (๐‘Ž, ๐‘1 ), (๐‘Ž, ๐‘2 ) ∈ ๐œ‘ and (๐‘1 , ๐‘‘1 ), (๐‘2 , ๐‘‘2 ) ∈ ๐œ“. Since ๐œ‘ is a
function, ๐‘1 = ๐‘2 , and then since ๐œ“ is a function, ๐‘‘1 = ๐‘‘2 . Therefore, ๐œ“ โ—ฆ ๐œ‘ is a
function, which is clearly ๐ด → ๐ท. Furthermore,
๐œ“ โ—ฆ ๐œ‘ = {(๐‘ฅ, ๐‘ง) : ∃๐‘ฆ[๐‘ฆ ∈ ๐ถ ∧ (๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘ ∧ (๐‘ฆ, ๐‘ง) ∈ ๐œ“]}
= {(๐‘ฅ, ๐‘ง) : ∃๐‘ฆ[๐‘ฆ ∈ ๐ถ ∧ ๐œ‘(๐‘ฅ) = ๐‘ฆ ∧ ๐œ“(๐‘ฆ) = ๐‘ง]}
= {(๐‘ฅ, ๐‘ง) : ๐‘ฅ ∈ ๐ด ∧ ๐œ“(๐œ‘(๐‘ฅ)) = ๐‘ง}
= {(๐‘ฅ, ๐œ“(๐œ‘(๐‘ฅ))) : ๐‘ฅ ∈ ๐ด}.
Hence, (๐œ“ โ—ฆ ๐œ‘)(๐‘ฅ) = ๐œ“(๐œ‘(๐‘ฅ)).
The ran(๐œ‘) ⊆ ๐ถ condition is important to√Theorem 4.4.20. For example, take the realvalued functions ๐‘“ (๐‘ฅ) = ๐‘ฅ and ๐‘”(๐‘ฅ) = ๐‘ฅ. Since ๐‘“ (−1) = −1 but ๐‘”(−1) ∉ ๐‘, we
conclude that (๐‘” โ—ฆ ๐‘“ )(−1) is undefined.
EXAMPLE 4.4.21
Define the two functions ๐‘“ : ๐‘ → ๐™ and ๐‘” : ๐‘ โงต {0} → ๐‘ by ๐‘“ (๐‘ฅ) = โŸฆ๐‘ฅโŸง and
๐‘”(๐‘ฅ) = 1โˆ•๐‘ฅ. Since ran(๐‘“ ) = ๐™ * dom(๐‘”), there are elements of ๐‘ for which
๐‘” โ—ฆ ๐‘“ is undefined. However,
ran(๐‘”) = ๐‘ โงต {0} ⊆ ๐‘ = dom(๐‘“ ),
so ๐‘“ โ—ฆ ๐‘” is defined and for all ๐‘ฅ ∈ ๐‘,
(๐‘“ โ—ฆ ๐‘”)(๐‘ฅ) = ๐‘“ (๐‘”(๐‘ฅ)) = ๐‘“ (1โˆ•๐‘ฅ) = โŸฆ1โˆ•๐‘ฅโŸง.
EXAMPLE 4.4.22
Let ๐œ“ : ๐™ ๐™ → ๐™ be defined by ๐œ“(๐‘“ ) = ๐œ–3 (๐‘“ ) and also let ๐œ‘ : ๐™ → ๐™7 be
๐œ‘(๐‘›) = [๐‘›]7 . Since ran(๐œ“) ⊆ dom(๐œ‘), ๐œ‘ โ—ฆ ๐œ“ is defined. Thus, if ๐‘” : ๐™ → ๐™ is
defined as ๐‘”(๐‘›) = 3๐‘›,
(๐œ‘ โ—ฆ ๐œ“)(๐‘”) = ๐œ‘(๐œ“(๐‘”)) = ๐œ‘(๐‘”(3)) = ๐œ‘(9) = [9]7 = [2]7 .
We should note that function composition is not a binary operation unless both functions are ๐ด → ๐ด for some set ๐ด. In this case, function composition is a binary operation
on ๐ด๐ด.
Restrictions and Extensions
There are times when a subset of a given function is required. For example, consider
๐‘“ = {(๐‘ฅ, ๐‘ฅ2 ) : ๐‘ฅ ∈ ๐‘}.
Section 4.4 FUNCTIONS
197
If only positive values of ๐‘ฅ are required, we can define
๐‘” = {(๐‘ฅ, ๐‘ฅ2 ) : ๐‘ฅ ∈ (0, ∞)}
so that ๐‘” ⊆ ๐‘“ . We have notation for this.
DEFINITION 4.4.23
Let ๐œ‘ : ๐ด → ๐ต be a function and ๐ถ ⊆ ๐ด.
โˆ™ The restriction of ๐œ‘ to ๐ถ is the function ๐œ‘ ๐ถ : ๐ถ → ๐ต so that
(๐œ‘ ๐ถ)(๐‘ฅ) = ๐œ‘(๐‘ฅ) for all ๐‘ฅ ∈ ๐ถ.
โˆ™ The function ๐œ“ : ๐ท → ๐ธ is an extension of ๐œ‘ if ๐ด ⊆ ๐ท, ๐ต ⊆ ๐ธ, and
๐œ“ ๐ด = ๐œ‘.
EXAMPLE 4.4.24
Let ๐‘“ = {(1, 2), (2, 3), (3, 4), (4, 1)} and ๐‘” = {(1, 2), (2, 3)}. We conclude that
๐‘” = ๐‘“ {1, 2}, and ๐‘“ is an extension of ๐‘”.
EXAMPLE 4.4.25
Let ๐œ‘ : ๐‘ˆ → ๐‘‰ be a function and ๐ด, ๐ต ⊆ ๐‘ˆ . We conclude that
๐œ‘ (๐ด ∪ ๐ต) = (๐œ‘ ๐ด) ∪ (๐œ‘ ๐ต)
because
(๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘ (๐ด ∪ ๐ต) ⇔ ๐‘ฆ = ๐œ‘(๐‘ฅ) ∧ ๐‘ฅ ∈ ๐ด ∪ ๐ต
⇔ ๐‘ฆ = ๐œ‘(๐‘ฅ) ∧ (๐‘ฅ ∈ ๐ด ∨ ๐‘ฅ ∈ ๐ต)
⇔ ๐‘ฆ = ๐œ‘(๐‘ฅ) ∧ ๐‘ฅ ∈ ๐ด ∨ ๐‘ฆ = ๐œ‘(๐‘ฅ) ∧ ๐‘ฅ ∈ ๐ต
⇔ (๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘ ๐ด ∨ (๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘ ๐ต
⇔ (๐‘ฅ, ๐‘ฆ) ∈ (๐œ‘ ๐ด) ∪ (๐œ‘ ๐ต).
Binary Operations
Standard addition and multiplication of real numbers are functions ๐‘ × ๐‘ → ๐‘ (Example 4.4.4). This means two things. First, given any two real numbers, their sum or
product will always be the same number. For instance, 3+5 is 8 and never another number. Second, given any two real numbers, their sum or product is also a real number.
Notice that subtraction also has these two properties when it is considered an operation
involving real numbers, but when we restrict substraction to ๐™+ , it no longer has the
second property because the difference of two positive integers might not be a positive
integer. That is, subtraction is not a function ๐™+ × ๐™+ → ๐™+ .
198
Chapter 4 RELATIONS AND FUNCTIONS
DEFINITION 4.4.26
A binary operation ∗ on the nonempty set ๐ด is a function ๐ด × ๐ด → ๐ด.
The symbol that represents the addition function is +. It can be viewed as a function
๐‘ → ๐‘. Therefore, using function notation, +(3, 5) = 8. However, we usually write
this as 3 + 5 = 8. Similarly, since ∗ represents an operation like addition, instead of
writing ∗ (๐‘Ž, ๐‘), we usually write ๐‘Ž ∗ ๐‘,
To prove that a relation ∗ is a binary operation on ๐ด, we must show that it satisfies
Definition 4.4.26. To do this, take ๐‘Ž, ๐‘Ž′ , ๐‘, ๐‘′ ∈ ๐ด, and prove:
โˆ™ ๐‘Ž = ๐‘Ž′ and ๐‘ = ๐‘′ implies ๐‘Ž ∗ ๐‘ = ๐‘Ž′ ∗ ๐‘′ ,
โˆ™ ๐ด is closed under ∗, that is ๐‘Ž ∗ ๐‘ ∈ ๐ด.
EXAMPLE 4.4.27
Define ๐‘ฅ ∗ ๐‘ฆ = 2๐‘ฅ − ๐‘ฆ and take ๐‘Ž, ๐‘Ž′ , ๐‘, ๐‘′ ∈ ๐™.
โˆ™ Assume ๐‘Ž = ๐‘Ž′ and ๐‘ = ๐‘′ . Then,
๐‘Ž ∗ ๐‘ = 2๐‘Ž − ๐‘ = 2๐‘Ž′ − ๐‘′ = ๐‘Ž′ ∗ ๐‘′ .
The second equality holds because multiplication and subtraction are binary operations on ๐™.
โˆ™ Because the product and difference of two integers is an integer, we have
that ๐‘Ž ∗ ๐‘ ∈ ๐™, so ๐™ is closed under ∗.
Thus, ∗ is a binary operation on ๐™.
EXAMPLE 4.4.28
Let ๐‘† = {๐‘’, ๐‘Ž, ๐‘, ๐‘} and define ∗ by the following table:
∗ ๐‘’
๐‘’ ๐‘’
๐‘Ž ๐‘Ž
๐‘ ๐‘
๐‘ ๐‘
๐‘Ž ๐‘ ๐‘
๐‘Ž ๐‘ ๐‘
๐‘ ๐‘ ๐‘’
๐‘ ๐‘’ ๐‘Ž
๐‘’ ๐‘Ž ๐‘
The table is read from left to right, so ๐‘ ∗ ๐‘ = ๐‘Ž. The table makes ∗ into a binary
operation since every pair of elements of ๐‘† is assigned a unique element of ๐‘†.
EXAMPLE 4.4.29
Fix a set ๐ด. For any ๐‘‹, ๐‘Œ ∈ P(๐ด), define ๐‘‹ ∗ ๐‘Œ = ๐‘‹ ∪ ๐‘Œ .
โˆ™ Let ๐‘‹1 , ๐‘‹2 , ๐‘Œ1 , ๐‘Œ2 ∈ P(๐ด). If we assume that ๐‘‹1 = ๐‘‹2 and ๐‘Œ1 = ๐‘Œ2 , we
have ๐‘‹1 ∪ ๐‘Œ1 = ๐‘‹2 ∪ ๐‘Œ2 . Hence, ∗ is well-defined.
Section 4.4 FUNCTIONS
199
โˆ™ To show that P(๐ด) is closed under ∗, let ๐ต and ๐ถ be subsets of ๐ด. Then
๐ต ∗ ๐ถ = ๐ต ∪ ๐ถ ∈ P(๐ด) because ๐ต ∪ ๐ถ ⊆ ๐ด [Exercise 3.3.10(a)].
This shows that ∗ is a binary operation on P(๐ด). Notice that ∗ is a subset of
[P(๐ด) × P(๐ด)] × P(๐ด).
EXAMPLE 4.4.30
Let ๐‘š ∈ ๐™+ and define [๐‘Ž]๐‘š + [๐‘]๐‘š = [๐‘Ž + ๐‘]๐‘š . We show that this is a binary
operation on ๐™๐‘š .
โˆ™ Let ๐‘Ž1 , ๐‘Ž2 , ๐‘1 , ๐‘2 ∈ ๐™. Suppose [๐‘Ž1 ]๐‘š = [๐‘Ž2 ]๐‘š and [๐‘1 ]๐‘š = [๐‘2 ]๐‘š . This
means that ๐‘Ž1 = ๐‘Ž2 + ๐‘›๐‘˜ and ๐‘1 = ๐‘2 + ๐‘›๐‘™ for some ๐‘˜, ๐‘™ ∈ ๐™. Hence,
๐‘Ž1 + ๐‘1 = ๐‘Ž2 + ๐‘2 + ๐‘›(๐‘˜ + ๐‘™), and we have [๐‘Ž1 + ๐‘1 ]๐‘š = [๐‘Ž2 + ๐‘2 ]๐‘š .
โˆ™ For closure, let [๐‘Ž]๐‘š , [๐‘]๐‘š ∈ ๐™๐‘š where ๐‘Ž and ๐‘ are integers. Then, we have
that [๐‘Ž]๐‘š + [๐‘]๐‘š = [๐‘Ž + ๐‘]๐‘š ∈ ๐™๐‘š since ๐‘Ž + ๐‘ is an integer.
Many binary operations share similar properties with the operations of + and × on ๐‘.
The next definition gives four of these properties.
DEFINITION 4.4.31
Let ∗ be a binary operation on ๐ด.
โˆ™ ∗ is associative means that (๐‘Ž ∗ ๐‘) ∗ ๐‘ = ๐‘Ž ∗ (๐‘ ∗ ๐‘) for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐ด.
โˆ™ ∗ is commutative means that ๐‘Ž ∗ ๐‘ = ๐‘ ∗ ๐‘Ž for all ๐‘Ž, ๐‘ ∈ ๐ด.
โˆ™ The element ๐‘’ is an identity of ๐ด with respect to ∗ when ๐‘’ ∈ ๐ด and ๐‘’ ∗
๐‘Ž = ๐‘Ž ∗ ๐‘’ = ๐‘Ž for all ๐‘Ž ∈ ๐ด.
โˆ™ Suppose that ๐ด has an identity ๐‘’ with respect to ∗ and let ๐‘Ž ∈ ๐ด. The
element ๐‘Ž′ ∈ ๐ด is an inverse of ๐‘Ž with respect to ∗ if ๐‘Ž ∗ ๐‘Ž′ = ๐‘Ž′ ∗ ๐‘Ž = ๐‘’.
Notice that the identity, if it exists, must be unique. To prove this, suppose that
both ๐‘’ and ๐‘’′ are identities. These must be equal because ๐‘’ = ๐‘’ ∗ ๐‘’′ = ๐‘’′ . So, if
a set has an identity with respect to an operation, we can refer to it as the identity of
the set. Similarly, we can write the inverse if it exists for associative binary operations
(Exercise 20).
EXAMPLE 4.4.32
We assume that + and × are both associative and commutative on ๐‚ and all subsets of ๐‚, that 0 is the identity with respect to + (the additive identity) and 1 is the
identity with respect to × (the multiplicative identity), and that every complex
number has an inverse with respect to + (an additive inverse) and every nonzero
complex number has an inverse with respect to × (a multiplicative inverse).
200
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.4.33
The binary operation defined in Example 4.4.30 is both associative and commutative. To see that it is commutative, let ๐‘Ž, ๐‘ ∈ ๐™. Then,
[๐‘Ž]๐‘š + [๐‘]๐‘š = [๐‘Ž + ๐‘]๐‘š = [๐‘ + ๐‘Ž]๐‘š = [๐‘]๐‘š + [๐‘Ž]๐‘š ,
(4.6)
where the second equality holds because + is commutative on ๐™. Its identity is
[0]๐‘š [let ๐‘ = 0 in (4.6)] and the additive inverse of [๐‘Ž]๐‘š is [−๐‘Ž]๐‘š .
EXAMPLE 4.4.34
Since ๐ด ∪ ๐ต = ๐ต ∪ ๐ด and (๐ด ∪ ๐ต) ∪ ๐ถ = ๐ด ∪ (๐ต ∪ ๐ถ) for all sets ๐ด, ๐ต, and ๐ถ,
the binary operation in Example 4.4.29 is both associative and commutative. Its
identity is ∅, and only ∅ has an inverse.
Exercises
1. Indicate whether each of the given relations are functions. If a relation is not a
function, find an element of its domain that is paired with two elements of its range.
(a) {(1, 2), (2, 3), (3, 4), (4, 5), (5, 1)}
(b) {(1, 1),
√ (1, 2), (1, 3), (1, 4), (1, 5)}
(c) {(๐‘ฅ, |๐‘ฅ| : ๐‘ฅ ∈ ๐‘}
√
(d) {(๐‘ฅ, ± |๐‘ฅ|) : ๐‘ฅ ∈ ๐‘}
(e) {(๐‘ฅ, ๐‘ฅ2 ) : ๐‘ฅ ∈ ๐‘}
(f) {([๐‘Ž]5 , ๐‘) : ∃๐‘˜ ∈ ๐™(๐‘Ž = ๐‘ + 5๐‘˜)}
(g) ๐œ“ : ๐™ → ๐™5 if ๐œ“(๐‘Ž) = [๐‘Ž]5
2. Prove that the given relations are functions.
(a) {(๐‘ฅ, 1โˆ•๐‘ฅ) : ๐‘ฅ ∈ ๐‘ โงต {0}}
(b) {(๐‘ฅ, ๐‘ฅ + 1) : ๐‘ฅ ∈ ๐™}
(c) {(๐‘ฅ, |๐‘ฅ|)
√ : ๐‘ฅ ∈ ๐‘}
(d) {(๐‘ฅ, ๐‘ฅ) : ๐‘ฅ ∈ [0, ∞)}
3. Let ๐‘“ = {(๐‘ฅ, ๐‘ฆ) ∈ ๐‘2 : 2๐‘ฅ + ๐‘ฆ = 1}. Show that ๐‘“ is a function with domain and
codomain equal to the set of real numbers.
4. Let ๐‘“ , ๐‘” : ๐‘ → ๐‘ be functions. Prove that ๐œ‘(๐‘ฅ, ๐‘ฆ) = (๐‘“ (๐‘ฅ), ๐‘”(๐‘ฆ)) is a function with
domain and codomain equal to ๐‘ × ๐‘.
5. Let ๐ด be a set and define ๐œ“(๐ด) = P(๐ด). Show that ๐œ“ is a function.
6. Define
{
๐‘ฅ2
๐‘“ (๐‘ฅ) =
5
if ๐‘ฅ ≥ 0,
if ๐‘ฅ < 0.
Show that ๐‘“ is well-defined with domain equal to ๐‘. What is ran(๐‘“ )?
Section 4.4 FUNCTIONS
201
7. Let ๐‘ฅ be in the domain and ๐‘ฆ in the range of each relation. Explain why each of the
given equations does not describe a function.
(a) ๐‘ฆ = 5 ± ๐‘ฅ
(b) ๐‘ฅ2 + ๐‘ฆ2 = 1
(c) ๐‘ฅ = 4๐‘ฆ2 − 1
(d) ๐‘ฆ2 − ๐‘ฅ2 = 9
8. Let ๐œ‘ : ๐™ → ๐™7 be defined by ๐œ‘(๐‘Ž) = [๐‘Ž]7 . Write the given images as rosters.
(a) ๐œ‘(0)
(b) ๐œ‘(7)
(c) ๐œ‘(3)
(d) ๐œ‘(−3)
9. Define ๐œ‘([๐‘Ž]๐‘› ) = [๐‘Ž]๐‘š for all ๐‘Ž ∈ ๐™. Is ๐œ‘ being function sufficient for ๐‘š โˆฃ ๐‘›?
Explain.
10. Give an example of a function that is an element of the given sets.
(a) ๐‘ ๐‘
(b) ๐‘ ๐™
(c) ๐ ๐‘
(d) ๐‘ [0, ∞)
(e) ๐™ (๐™5 )
11. Evaluate the indicated expressions.
(a) ๐œ–4 (๐‘“ ) if ๐‘“ (๐‘ฅ) = 9๐‘ฅ + 2
(b) ๐œ–๐œ‹ (๐‘”) if ๐‘”(๐œƒ) = sin ๐œƒ
12. Since functions are sets, we can perform set operations on them. Let ๐‘“ (๐‘ฅ) = ๐‘ฅ2
and ๐‘”(๐‘ฅ) = −๐‘ฅ. Find the following.
(a) ๐‘“ ∪ ๐‘”
(b) ๐‘“ ∩ ๐‘”
(c) ๐‘“ โงต ๐‘”
(d) ๐‘” โงต ๐‘“
โ‹ƒ
13. Let ๐’ž be a chain of functions with respect to ⊆. Prove that ๐’ž is a function.
14. Let ๐‘“ and ๐‘” be functions. Prove the following.
(a) If ๐‘“ and ๐‘” are functions, ๐‘“ ∩ ๐‘” is a function.
(b) ๐‘“ ∪ ๐‘” is a function if and only if ๐‘“ (๐‘ฅ) = ๐‘”(๐‘ฅ) for all ๐‘ฅ ∈ dom(๐‘“ ) ∩ dom(๐‘”).
15. Let ๐‘“ : ๐ด → ๐ต be a function. Define a relation ๐‘† on ๐ด by ๐‘Ž ๐‘† ๐‘ if and only if
๐‘“ (๐‘Ž) = ๐‘“ (๐‘).
(a) Show ๐‘† is an equivalence relation.
(b) Find [3]๐‘† if ๐‘“ : ๐™ → ๐™ is defined by ๐‘“ (๐‘›) = 2๐‘›.
(c) Find [2]๐‘† if ๐‘“ : ๐™ → ๐™5 is given by ๐‘“ (๐‘›) = [๐‘›]5 .
16. Show that ∅ is a function and find its domain and range.
202
Chapter 4 RELATIONS AND FUNCTIONS
17. Define ∗ by ๐‘ฅ ∗ ๐‘ฆ = ๐‘ฅ + ๐‘ฆ + 2 for all ๐‘ฅ, ๐‘ฆ ∈ ๐™.
(a) Show that ∗ is a binary operation on ๐™.
(b) Prove that −2 is the identity of ๐™ with respect to ∗.
(c) For every ๐‘› ∈ ๐™, show that −๐‘› − 4 is the inverse of ๐‘› with respect to ∗.
18. Define the binary operation ∗ by ๐‘ฅ ∗ ๐‘ฆ = 2๐‘ฅ − ๐‘ฆ for all ๐‘ฅ, ๐‘ฆ ∈ ๐™.
(a) Is there an integer that serves as an identity with respect to ∗?
(b) Does every integer have an inverse with respect to ∗?
19. Let ๐‘“ , ๐‘” : ๐‘ → ๐‘ be functions. Prove that ๐‘“ โ—ฆ ๐‘” is well-defined.
20. For an associative binary operation, prove that the inverse of an element is unique
if it exists. Show that this might not be the case if the binary operation is not associative.
21. Prove that the given pairs of functions are equal.
(a) ๐‘“ (๐‘ฅ) = (๐‘ฅ − 1)(๐‘ฅ − 2)(๐‘ฅ + 3) and ๐‘”(๐‘ฅ) = ๐‘ฅ3 − 7๐‘ฅ + 6
where ๐‘“ , ๐‘” : ๐‘ → ๐‘
(b) ๐œ‘(๐‘Ž, ๐‘) = ๐‘Ž + ๐‘ and ๐œ“(๐‘Ž, ๐‘) = ๐‘ + ๐‘Ž
where ๐œ‘, ๐œ“ : ๐™ × ๐™ → ๐™
(c) ๐œ‘(๐‘Ž, ๐‘) = ([๐‘Ž]5 , [๐‘ + 7]5 ) and ๐œ“(๐‘Ž, ๐‘) = ([๐‘Ž + 5]5 , [๐‘ − 3]5 )
where ๐œ‘, ๐œ“ : ๐™ × ๐™ → ๐™5 × ๐™5
(d) ๐œ‘(๐‘“ ) = ๐‘“ ๐™ and ๐œ“(๐‘“ ) = {(๐‘›, ๐‘“ (๐‘›)) : ๐‘› ∈ ๐™}
where ๐œ‘, ๐œ“ : ๐‘ ๐‘ → ๐™ ๐‘
22. Show that the given pairs of functions are not equal.
(a) ๐‘“ (๐‘ฅ) = ๐‘ฅ and ๐‘”(๐‘ฅ) = 2๐‘ฅ where ๐‘“ , ๐‘” : ๐‘ → ๐‘
(b) ๐‘“ (๐‘ฅ) = ๐‘ฅ − 3 and ๐‘”(๐‘ฅ) = ๐‘ฅ + 3 where ๐‘“ , ๐‘” : ๐‘ → ๐‘
(c) ๐œ‘(๐‘Ž) = [๐‘Ž]5 and ๐œ“(๐‘Ž) = [๐‘Ž]4 where ๐œ‘, ๐œ“ : ๐™ → ๐™4 ∪ ๐™5
(d) ๐œ‘(๐ด) = ๐ด โงต {0} and ๐œ“(๐ด) = ๐ด ∩ {1, 2, 3} where ๐œ‘, ๐œ“ : P(๐™) → P(๐™)
23. Let ๐œ“ : ๐‘ → ๐‘ ๐‘ be defined by ๐œ“(๐‘Ž) = ๐‘“๐‘Ž where ๐‘“๐‘Ž is the function ๐‘“๐‘Ž : ๐‘ → ๐‘
with ๐‘“๐‘Ž (๐‘ฅ) = ๐‘Ž๐‘ฅ. Prove that ๐œ“ is well-defined.
24. For each pair of functions, find the indicated values when possible.
(a) ๐‘“ : ๐‘ → ๐‘ and ๐‘“ (๐‘ฅ) = 2๐‘ฅ3
๐‘” : ๐‘ → ๐‘ and ๐‘”(๐‘ฅ) = ๐‘ฅ + 1
(๐‘“ โ—ฆ ๐‘”)(2)
(๐‘” โ—ฆ ๐‘“ )(0)
√
(b) ๐‘“ : [0, ∞) → ๐‘ and ๐‘“ (๐‘ฅ) = ๐‘ฅ
๐‘” : ๐‘ → ๐‘ and ๐‘”(๐‘ฅ) = |๐‘ฅ| − 1
(๐‘“ โ—ฆ ๐‘”)(0)
(๐‘” โ—ฆ ๐‘“ )(4)
(c) ๐œ‘ : ๐™ → ๐™5 and ๐œ‘(๐‘Ž) = [๐‘Ž]5
๐œ“ : ๐‘ ๐‘ → ๐‘ and ๐œ“(๐‘“ ) = ๐‘“ (0)
(๐œ‘ โ—ฆ ๐œ“)(.5๐‘ฅ + 1)
(๐œ“ โ—ฆ ๐œ‘)(2)
Section 4.5 INJECTIONS AND SURJECTIONS
203
25. For each of the given functions, find the composition of the function with itself.
For example, find ๐‘“ โ—ฆ ๐‘“ for part (a).
(a) ๐‘“ : ๐‘ → ๐‘ with ๐‘“ (๐‘ฅ) = ๐‘ฅ2
(b) ๐‘” : ๐‘ → ๐‘ with ๐‘”(๐‘ฅ) = 3๐‘ฅ + 1
(c) ๐œ‘ : ๐™ × ๐™ → ๐™ × ๐™ with ๐œ‘(๐‘ฅ, ๐‘ฆ) = (2๐‘ฆ, 5๐‘ฅ − ๐‘ฆ)
(d) ๐œ“ : ๐™๐‘š → ๐™๐‘š with ๐œ“([๐‘›]๐‘š ) = [๐‘› + 2]๐‘š
26. Let ๐œ‘ : ๐ด → ๐ต be a function and ๐œ“ = ๐œ‘ ๐ถ where ๐ถ ⊆ ๐ด. Prove that if ๐œ„ : ๐ถ → ๐ด
is defined by ๐œ„(๐‘) = ๐‘ (known as the inclusion map), then ๐œ“ = ๐œ‘ โ—ฆ ๐œ„.
27. Write the given restrictions as rosters.
(a) {(1, 2), (2, 2), (3, 4), (4, 7)} {1, 3}
(b) ๐‘“ {0, 1, 2, 3} where ๐‘“ (๐‘ฅ) = 7๐‘ฅ − 1 and dom(๐‘“ ) = ๐‘
(c) (๐‘” + โ„Ž) {−3.3, 1.2, 7} where ๐‘”(๐‘ฅ) = โŸฆ๐‘ฅโŸง, โ„Ž(๐‘ฅ) = ๐‘ฅ + 1, and both dom(๐‘”)
and dom(โ„Ž) equal ๐‘
28. For functions ๐‘“ and ๐‘” such that ๐ด, ๐ต ⊆ dom(๐‘“ ), prove the following.
(a) ๐‘“ ๐ด = ๐‘“ ∩ [๐ด × ran(๐‘“ )]
(b) ๐‘“ (๐ด ∩ ๐ต) = (๐‘“ ๐ด) ∩ (๐‘“ ๐ต)
(c) ๐‘“ (๐ด โงต ๐ต) = (๐‘“ ๐ด) โงต (๐‘“ ๐ต)
(d) (๐‘” โ—ฆ ๐‘“ ) ๐ด = ๐‘” โ—ฆ (๐‘“ ๐ด)
29. Let ๐‘“ : ๐‘ˆ → ๐‘‰ be a function. Prove that if ๐ด ⊆ ๐‘ˆ , then ๐‘“ ๐ด = ๐‘“ โ—ฆ ๐ผ๐ด .
30. Let ๐œ‘ : ๐ด ๐ถ → ๐ต๐ถ be defined by ๐œ‘(๐‘“ ) = ๐‘“ ๐ต. Prove that ๐œ‘ is well-defined.
31. A real-valued function ๐‘“ is periodic if there exists ๐‘˜ > 0 so that ๐‘“ (๐‘ฅ) = ๐‘“ (๐‘ฅ + ๐‘˜)
for all ๐‘ฅ ∈ dom(๐‘“ ). Let ๐‘”, โ„Ž : ๐‘ → ๐‘ be functions with period ๐‘˜. Prove that ๐‘” โ—ฆ โ„Ž is
periodic with period ๐‘˜.
32. Let (๐ด, 4) be a poset. A function ๐œ‘ : ๐ด → ๐ด is increasing means for all ๐‘ฅ, ๐‘ฆ ∈ ๐ด,
if ๐‘ฅ โ‰บ ๐‘ฆ, then ๐œ‘(๐‘ฅ) โ‰บ ๐œ‘(๐‘ฆ). A decreasing function is defined similarly. Suppose that
๐œŽ and ๐œ are increasing. Prove that ๐œŽ โ—ฆ ๐œ is increasing.
4.5
INJECTIONS AND SURJECTIONS
When looking at relations, we studied the concept of an inverse relation. Given ๐‘…, obtain ๐‘…−1 by exchanging the ๐‘ฅ- and ๐‘ฆ-coordinates. The same can be done with functions,
but the inverse might not be a function. For example, given
๐‘“ = {(1, 2), (2, 3), (3, 2)},
its inverse is
๐‘“ −1 = {(2, 1), (3, 2), (2, 3)}.
However, if the original relation is a function, we often want the inverse also to be a
function. This leads to the next definition.
204
Chapter 4 RELATIONS AND FUNCTIONS
DEFINITION 4.5.1
๐œ‘ : ๐ด → ๐ต is invertible means that ๐œ‘−1 is a function ๐ต → ๐ด.
An immediate consequence of the definition is the next result.
LEMMA 4.5.2
Let ๐œ‘ be invertible. Then, ๐œ‘(๐‘ฅ) = ๐‘ฆ if and only if ๐œ‘−1 (๐‘ฆ) = ๐‘ฅ for all ๐‘ฅ ∈ dom(๐œ‘).
PROOF
Suppose that ๐œ‘(๐‘ฅ) = ๐‘ฆ. This means that (๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘, so (๐‘ฆ, ๐‘ฅ) ∈ ๐œ‘−1 . Since
๐œ‘ is invertible, ๐œ‘−1 is a function, so write ๐œ‘−1 (๐‘ฆ) = ๐‘ฅ. The converse is proved
similarly.
We use Lemma 4.5.2 in the proof of the next theorem, which gives conditions for
when a function is invertible.
THEOREM 4.5.3
๐œ‘ : ๐ด → ๐ต is invertible if and only if ๐œ‘−1 โ—ฆ ๐œ‘ = ๐ผ๐ด and ๐œ‘ โ—ฆ ๐œ‘−1 = ๐ผ๐ต .
PROOF
Take a function ๐œ‘ : ๐ด → ๐ต. Then, ๐œ‘−1 ⊆ ๐ต × ๐ด.
โˆ™ Assume that ๐œ‘−1 is a function ๐ต → ๐ด. Let ๐‘ฅ ∈ ๐ด and ๐‘ฆ ∈ ๐ต. By assumption, we have ๐‘ฅ0 ∈ ๐ด and ๐‘ฆ0 ∈ ๐ต such that ๐œ‘(๐‘ฅ) = ๐‘ฆ0 and ๐œ‘−1 (๐‘ฆ) = ๐‘ฅ0 .
This implies that ๐œ‘−1 (๐‘ฆ0 ) = ๐‘ฅ and ๐œ‘(๐‘ฅ0 ) = ๐‘ฆ by Lemma 4.5.2. Therefore,
(๐œ‘−1 โ—ฆ ๐œ‘)(๐‘ฅ) = ๐œ‘−1 (๐œ‘(๐‘ฅ)) = ๐œ‘−1 (๐‘ฆ0 ) = ๐‘ฅ,
and
(๐œ‘ โ—ฆ ๐œ‘−1 )(๐‘ฆ) = ๐œ‘(๐œ‘−1 (๐‘ฆ)) = ๐œ‘(๐‘ฅ0 ) = ๐‘ฆ.
โˆ™ Now assume ๐œ‘−1 โ—ฆ ๐œ‘ = ๐ผ๐ด and ๐œ‘ โ—ฆ ๐œ‘−1 = ๐ผ๐ต . To show that ๐œ‘−1 is a
function, take (๐‘ฆ, ๐‘ฅ), (๐‘ฆ, ๐‘ฅ′ ) ∈ ๐œ‘−1 . From this, we know that (๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘.
Therefore,
(๐‘ฅ, ๐‘ฅ′ ) ∈ ๐œ‘−1 โ—ฆ ๐œ‘ = ๐ผ๐ด ,
so ๐‘ฅ = ๐‘ฅ′ . In addition, we know that dom(๐œ‘−1 ) ⊆ ๐ต, so to prove equality,
let ๐‘ฆ ∈ ๐ต. Then,
(๐‘ฆ, ๐‘ฆ) ∈ ๐ผ๐ต = ๐œ‘ โ—ฆ ๐œ‘−1 .
Thus, there exists ๐‘ฅ ∈ ๐ด such that (๐‘ฆ, ๐‘ฅ) ∈ ๐œ‘−1 , so ๐‘ฆ ∈ dom(๐œ‘−1 ).
EXAMPLE 4.5.4
โˆ™ Let ๐‘“ : ๐‘ → ๐‘ be the function given by ๐‘“ (๐‘ฅ) = ๐‘ฅ + 2. Its inverse is ๐‘”(๐‘ฅ) = ๐‘ฅ − 2
by Theorem 4.5.3. This is because
(๐‘” โ—ฆ ๐‘“ )(๐‘ฅ) = ๐‘”(๐‘ฅ + 2) = (๐‘ฅ + 2) − 2 = ๐‘ฅ
Section 4.5 INJECTIONS AND SURJECTIONS
205
and
(๐‘“ โ—ฆ ๐‘”)(๐‘ฅ) = ๐‘“ (๐‘ฅ − 2) = (๐‘ฅ − 2) + 2 = ๐‘ฅ.
โˆ™ If the function ๐‘” : [0, ∞) → [0, ∞) is defined by √
๐‘”(๐‘ฅ) = ๐‘ฅ2 , then ๐‘” −1 is a
−1
function [0, ∞) → [0, ∞) and is defined by ๐‘” (๐‘ฅ) = ๐‘ฅ.
โˆ™ Let โ„Ž : ๐‘ → (0, ∞) be defined as โ„Ž(๐‘ฅ) = ๐‘’๐‘ฅ . By Theorem 4.5.3, we know that
โ„Ž−1 (๐‘ฅ) = ln ๐‘ฅ because ๐‘’ln ๐‘ฅ = ๐‘ฅ for all ๐‘ฅ ∈ [0, ∞) and ln ๐‘’๐‘ฅ = ๐‘ฅ for all ๐‘ฅ ∈ ๐‘.
Injections
Theorem 4.5.3 can be improved by finding a condition for the invertibility of a function
based only on the given function. Consider the following. In order for a relation to be a
function, it cannot look like Figure 4.9. Since the inverse exchanges the roles of the two
coordinates, in order for an inverse to be a function, the original function cannot look
like the graph in Figure 4.11. In other words, if ๐‘“ −1 is to be a function, there cannot
exist ๐‘ฅ1 and ๐‘ฅ2 so that ๐‘ฅ1 ≠ ๐‘ฅ2 and ๐‘“ (๐‘ฅ1 ) = ๐‘“ (๐‘ฅ2 ). But then,
¬∃๐‘ฅ1 ∃๐‘ฅ2 [๐‘ฅ1 ≠ ๐‘ฅ2 ∧ ๐‘“ (๐‘ฅ1 ) = ๐‘“ (๐‘ฅ2 )]
is equivalent to
∀๐‘ฅ1 ∀๐‘ฅ2 [๐‘ฅ1 = ๐‘ฅ2 ∨ ๐‘“ (๐‘ฅ1 ) ≠ ๐‘“ (๐‘ฅ2 )],
which in turn is equivalent to
∀๐‘ฅ1 ∀๐‘ฅ2 [๐‘“ (๐‘ฅ1 ) = ๐‘“ (๐‘ฅ2 ) → ๐‘ฅ1 = ๐‘ฅ2 ].
Hence, ๐‘“ being an invertible function implies that for every ๐‘ฅ1 , ๐‘ฅ2 ∈ dom(๐‘“ ),
๐‘ฅ1 = ๐‘ฅ2 if and only if ๐‘“ (๐‘ฅ1 ) = ๐‘“ (๐‘ฅ2 ).
f
(x 1, y)
x1
Figure 4.11
y
(x 2, y)
x2
The inverse of a function might not be a function.
(4.7)
206
Chapter 4 RELATIONS AND FUNCTIONS
๐œ‘
A
Figure 4.12
B
๐œ‘ is a one-to-one function.
This means that the elements of the domain of ๐‘“ and the elements of the range of ๐‘“ form
pairs of elements as illustrated in Figure 4.12. The sufficiency of (4.7) is the definition
of a function, while necessity is the next definition.
DEFINITION 4.5.5
The function ๐œ‘ : ๐ด → ๐ต is one-to-one if and only if for all ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐ด,
if ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ2 ), then ๐‘ฅ1 = ๐‘ฅ2 .
A one-to-one function is sometimes called an injection.
EXAMPLE 4.5.6
Define ๐‘“ : ๐‘ → ๐‘ by ๐‘“ (๐‘ฅ) = 5๐‘ฅ+1. To show that ๐‘“ is one-to-one, let ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐‘
and assume ๐‘“ (๐‘ฅ1 ) = ๐‘“ (๐‘ฅ2 ). Then,
5๐‘ฅ1 + 1 = 5๐‘ฅ2 + 1,
5๐‘ฅ1 = 5๐‘ฅ2 ,
๐‘ฅ1 = ๐‘ฅ2 .
EXAMPLE 4.5.7
Let ๐œ‘ : ๐™ × ๐™ → ๐™ × ๐™ × ๐™ be the function
๐œ‘(๐‘Ž, ๐‘) = (๐‘Ž, ๐‘, 0).
For any (๐‘Ž1 , ๐‘1 ), (๐‘Ž2 , ๐‘2 ) ∈ ๐™ × ๐™, assume
๐œ‘(๐‘Ž1 , ๐‘1 ) = ๐œ‘(๐‘Ž2 , ๐‘2 ).
This means that
(๐‘Ž1 , ๐‘1 , 0) = (๐‘Ž2 , ๐‘2 , 0).
Hence, ๐‘Ž1 = ๐‘Ž2 and ๐‘1 = ๐‘2 , and this yields (๐‘Ž1 , ๐‘1 ) = (๐‘Ž2 , ๐‘2 ).
Section 4.5 INJECTIONS AND SURJECTIONS
๐œ‘
A
Figure 4.13
207
B
๐œ‘ is not a one-to-one function.
If a function is not one-to-one, there must be an element of the range that has at
least two pre-images (Figure 4.13). An example of a function that is not one-to-one is
๐‘“ (๐‘ฅ) = ๐‘ฅ2 where both the domain and codomain of ๐‘“ are ๐‘. This is because ๐‘“ (2) = 4
and ๐‘“ (−2) = 4. Another example is ๐‘” : ๐‘ → ๐‘ defined by ๐‘”(๐œƒ) = cos ๐œƒ. It is not an
injection because ๐‘”(0) = ๐‘”(2๐œ‹) = 1.
Although the original function might not be one-to-one, we can always restrict the
function to a subset of its domain so that the resulting function is one-to-one. This is
illustrated in the next two examples.
EXAMPLE 4.5.8
Let ๐‘“ be the function {(1, 5), (2, 8), (3, 8), (4, 6)}. We observe that ๐‘“ is not oneto-one, but both
๐‘“ {1, 2} = {(1, 5), (2, 8)}
and
๐‘“ {3, 4} = {(3, 8), (4, 6)}
are one-to-one as in Figure 4.14.
A
B
1
f โ†พA
5
2
8
3
6
4
f โ†พB
Figure 4.14
Restrictions of ๐‘“ to ๐ด and ๐ต are one-to-one.
208
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.5.9
Let ๐‘” : ๐‘ → ๐‘ be the function ๐‘”(๐‘ฅ) = ๐‘ฅ2 . This function is not one-to-one, but
๐‘” [0, ∞) and ๐‘” (−10, −5) are one-to-one.
√
Let ๐‘“ (๐‘ฅ) = 3๐‘ฅ + 6 and ๐‘”(๐‘ฅ) = 5๐‘ฅ − 8. Both are injections. Notice that
√
(๐‘“ โ—ฆ ๐‘”)(๐‘ฅ) = 3 5๐‘ฅ − 8 + 6
and
(๐‘” โ—ฆ ๐‘“ )(๐‘ฅ) =
√
15๐‘ฅ + 22
are also injections. We can generalize this result to the next theorem.
THEOREM 4.5.10
If ๐œ‘ : ๐ด → ๐ต and ๐œ“ : ๐ต → ๐ถ are injections, ๐œ“ โ—ฆ ๐œ‘ is an injection.
PROOF
Assume that ๐œ‘ : ๐ด → ๐ต and ๐œ“ : ๐ต → ๐ถ are one-to-one. Let ๐‘Ž1 , ๐‘Ž2 ∈ ๐ด and
assume (๐œ“ โ—ฆ ๐œ‘)(๐‘Ž1 ) = (๐œ“ โ—ฆ ๐œ‘)(๐‘Ž2 ). Then,
๐œ“(๐œ‘(๐‘Ž1 )) = ๐œ“(๐œ‘(๐‘Ž2 )).
Since ๐œ“ is one-to-one,
๐œ‘(๐‘Ž1 ) = ๐œ‘(๐‘Ž2 ),
and since ๐œ‘ is one-to-one, ๐‘Ž1 = ๐‘Ž2 .
Surjections
The function being an injection is not sufficient for it to be invertible since it is possible that not every element of the codomain will have a pre-image. In this case, the
codomain cannot be the domain of the inverse. To prevent this situation, we will need
the function to satisfy the next definition.
DEFINITION 4.5.11
A function ๐œ‘ : ๐ด → ๐ต is onto if and only if for every ๐‘ฆ ∈ ๐ต, there exists ๐‘ฅ ∈ ๐ด
such that ๐œ‘(๐‘ฅ) = ๐‘ฆ. An onto function is also called a surjection.
This definition is related to the range (or image) of the function. The range of the
function ๐œ‘ : ๐ด → ๐ต is
ran(๐œ‘) = {๐‘ฆ : ∃๐‘ฅ(๐‘ฅ ∈ ๐ด ∧ (๐‘ฅ, ๐‘ฆ) ∈ ๐œ‘)} = {๐œ‘(๐‘ฅ) : ๐‘ฅ ∈ dom(๐œ‘)}
as illustrated in Figure 4.15. Thus, ๐œ‘ is onto if and only if ran(๐œ‘) = ๐ต.
EXAMPLE 4.5.12
Define ๐‘“ : [0, ∞) → [0, ∞) by ๐‘“ (๐‘ฅ) =
√
๐‘ฅ. Its range is also [0, ∞), so it is onto.
Section 4.5 INJECTIONS AND SURJECTIONS
209
ran( ๐œ‘ )
A
B
x
Figure 4.15
๐œ‘
y
The range of ๐œ‘ : ๐ด → ๐ต with ๐‘ฆ = ๐œ‘(๐‘ฅ).
EXAMPLE 4.5.13
The ranges of the following functions are different from their codomains, so they
are not onto.
โˆ™ Let ๐‘” : ๐‘ → ๐‘ be defined by ๐‘”(๐‘ฅ) = |๐‘ฅ|. Then, ran(๐‘”) = [0, ∞).
โˆ™ Define โ„Ž : ๐™ → ๐™ by โ„Ž(๐‘›) = 2๐‘›. Here, ran(โ„Ž) = {2๐‘› : ๐‘› ∈ ๐™}.
The functions illustrated in Figures 4.12, 4.13, and 4.14 are onto functions as are
those in the next examples.
EXAMPLE 4.5.14
Any linear function ๐‘“ : ๐‘ → ๐‘ that is not a horizontal line is a surjection. To
see this, let ๐‘“ (๐‘ฅ) = ๐‘Ž๐‘ฅ + ๐‘ for some ๐‘Ž ≠ 0. Take ๐‘ฆ ∈ ๐‘. We need to find ๐‘ฅ ∈ ๐‘
so that ๐‘Ž๐‘ฅ + ๐‘ = ๐‘ฆ. Choose
๐‘ฆ−๐‘
๐‘ฅ=
.
๐‘Ž
Then,
(
)
๐‘ฆ−๐‘
๐‘“ (๐‘ฅ) = ๐‘Ž
+ ๐‘ = ๐‘ฆ.
๐‘Ž
The approach in the example is typical. To show that a function is onto, take an
arbitrary element of the codomain and search for a candidate to serve as its pre-image.
When found, check it.
EXAMPLE 4.5.15
Take a positive integer ๐‘š and let ๐œ‘ : ๐™ → ๐™๐‘š be defined as ๐œ‘(๐‘˜) = [๐‘˜]๐‘š . To see
that ๐œ‘ is onto, take [๐‘™]๐‘š ∈ ๐™๐‘š for some ๐‘™ ∈ ๐™. We then find that ๐œ‘(๐‘™) = [๐‘™]๐‘š .
EXAMPLE 4.5.16
Let ๐‘š, ๐‘› ∈ ๐ with ๐‘š > ๐‘›. A function ๐œ‹ : ๐‘๐‘š → ๐‘๐‘› defined by
๐œ‹(๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 , ๐‘ฅ๐‘› , … , ๐‘ฅ๐‘š−1 ) = (๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 )
210
Chapter 4 RELATIONS AND FUNCTIONS
is called a projection.. Such functions are not one-to-one, but they are onto. For
instance, define ๐œ‹ : ๐‘ × ๐‘ × ๐‘ → ๐‘ × ๐‘ by
๐œ‹(๐‘ฅ, ๐‘ฆ, ๐‘ง) = (๐‘ฅ, ๐‘ฆ).
It is not one-to-one because ๐œ‹(1, 2, 3) = ๐œ‹(1, 2, 4) = (1, 2). However, if we take
(๐‘Ž, ๐‘) ∈ ๐‘ × ๐‘, then ๐œ‹(๐‘Ž, ๐‘, 0) = (๐‘Ž, ๐‘), so ๐œ‹ is onto.
If a function is not onto, it has a diagram like that of Figure 4.16. Therefore, to show
that a function is not a surjection, we must find an element of the codomain that does
not have a pre-image.
EXAMPLE 4.5.17
Define ๐‘“ : ๐™ → ๐™ by ๐‘“ (๐‘›) = 3๐‘›. This function is not onto because 5 does not
have a pre-image in ๐™.
EXAMPLE 4.5.18
The function
๐œ‘:๐™×๐™→๐™×๐™×๐™
defined by ๐œ‘(๐‘Ž, ๐‘) = (๐‘Ž, ๐‘, 0) is not onto because (1, 1, 1) does not have a preimage in ๐™ × ๐™.
We have the following analog of Theorem 4.5.10 for surjections.
THEOREM 4.5.19
If ๐œ‘ : ๐ด → ๐ต and ๐œ“ : ๐ต → ๐ถ are surjections, ๐œ“ โ—ฆ ๐œ‘ is a surjection.
PROOF
Assume that ๐œ‘ : ๐ด → ๐ต and ๐œ“ : ๐ต → ๐ถ are surjections. Take ๐‘ ∈ ๐ถ. Then,
there exists ๐‘ ∈ ๐ต so that ๐œ“(๐‘) = ๐‘ and ๐‘Ž ∈ ๐ด such that ๐œ‘(๐‘Ž) = ๐‘. Therefore,
(๐œ“ โ—ฆ ๐œ‘)(๐‘Ž) = ๐œ“(๐œ‘(๐‘Ž)) = ๐œ“(๐‘) = ๐‘.
๐œ‘
A
Figure 4.16
๐œ‘ is not a onto function.
B
Section 4.5 INJECTIONS AND SURJECTIONS
211
Bijections
If we have a function that is both one-to-one and onto, then it is called a bijection or a
one-to-one correspondence. Observe that ∅ is a bijection.
EXAMPLE 4.5.20
As illustrated in Examples 4.5.6 and 4.5.14, every linear ๐‘“ : ๐‘ → ๐‘ with nonzero
slope is a bijection.
EXAMPLE 4.5.21
Both ๐‘” : (−๐œ‹โˆ•2, ๐œ‹โˆ•2) → ๐‘ such that ๐‘”(๐œƒ) = tan ๐œƒ and โ„Ž : ๐‘ → (0, ∞) where
โ„Ž(๐‘ฅ) = ๐‘’๐‘ฅ are bijections.
We are now ready to give the standard test for invertibility. Its proof requires both
Lemma 4.5.2 and Theorem 4.5.3. The benefit of this theorem is that it provides a test
for invertibility in which the given function is examined instead of its inverse.
THEOREM 4.5.22
A function is invertible if and only if it is a bijection.
PROOF
Let ๐œ‘ : ๐ด → ๐ต be a function.
โˆ™ Suppose that ๐œ‘ is invertible. To show that ๐œ‘ is one-to-one, let ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐ด
and assume that ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ2 ). Then, by Theorem 4.5.3,
๐‘ฅ1 = ๐œ‘−1 (๐œ‘(๐‘ฅ1 )) = ๐œ‘−1 (๐œ‘(๐‘ฅ2 )) = ๐‘ฅ2 .
To see that ๐œ‘ is onto, take ๐‘ฆ ∈ ๐ต. Then, there exists ๐‘ฅ ∈ ๐ด such that
๐œ‘−1 (๐‘ฆ) = ๐‘ฅ. Hence, ๐œ‘(๐‘ฅ) = ๐‘ฆ by Lemma 4.5.2.
โˆ™ Assume that ๐œ‘ is both one-to-one and onto. To show that ๐œ‘−1 is a function,
let (๐‘ฆ, ๐‘ฅ), (๐‘ฆ, ๐‘ฅ′ ) ∈ ๐œ‘−1 . This implies that ๐œ‘(๐‘ฅ) = ๐‘ฆ = ๐œ‘(๐‘ฅ′ ). Since ๐œ‘ is
one-to-one, ๐‘ฅ = ๐‘ฅ′ . To prove that the domain of ๐œ‘−1 is ๐ต, take ๐‘ฆ ∈ ๐ต.
Since ๐œ‘ is onto, there exists ๐‘ฅ ∈ ๐ด such that ๐œ‘(๐‘ฅ) = ๐‘ฆ. By Lemma 4.5.2,
we have that ๐œ‘−1 (๐‘ฆ) = ๐‘ฅ, so ๐‘ฆ ∈ dom(๐œ‘−1 ).
By Theorem 4.5.22 the functions of Examples 4.5.20 and 4.5.21 are invertible.
THEOREM 4.5.23
If ๐œ‘ : ๐ด → ๐ต and ๐œ“ : ๐ต → ๐ถ are bijections, then ๐œ‘−1 and ๐œ“ โ—ฆ ๐œ‘ are bijections.
PROOF
Suppose ๐œ‘ is a bijection. By Theorem 4.5.22, it is invertible, so ๐œ‘−1 is a function
that has ๐œ‘ as its inverse. Therefore, ๐œ‘−1 is a bijection by Theorem 4.5.22. Combining the proofs of Theorems 4.5.10 and 4.5.19 show that ๐œ“ โ—ฆ ๐œ‘ is a bijection
when ๐œ“ is a bijection.
212
Chapter 4 RELATIONS AND FUNCTIONS
Using the functions ๐‘” and โ„Ž from Example 4.5.21, we conclude from Theorem 4.5.23
that โ„Ž โ—ฆ ๐‘” is a bijection with domain (−๐œ‹โˆ•2, ๐œ‹โˆ•2) and range (0, ∞).
Order Isomorphims
Consider the function
๐œ‘ : ๐™ × {0} → {0} × ๐™
(4.8)
defined by ๐œ‘(๐‘š, 0) = (0, ๐‘š). It can be shown that ๐œ‘ is a bijection (compare Exercise 12).
Define the linear orders 4 on ๐™ × {0} and 4′ on {0} × ๐™ by
(๐‘š, 0) 4 (๐‘›, 0) if and only if ๐‘š ≤ ๐‘›
and
(0, ๐‘š) 4′ (0, ๐‘›) if and only if ๐‘š ≤ ๐‘›.
Notice that (3, 0) 4 (5, 0) and (0, 3) 4′ (0, 5) because 3 ≤ 5. We can generalize this to
conclude that
(๐‘š, 0) 4 (๐‘›, 0) if and only if (0, ๐‘š) 4′ (0, ๐‘›),
and this implies that
(๐‘š, 0) 4 (๐‘›, 0) if and only if ๐œ‘(๐‘š, 0) 4′ ๐œ‘(๐‘›, 0).
This leads to the next definition.
DEFINITION 4.5.24
Let ๐‘… be a relation on ๐ด and ๐‘† be a relation on ๐ต.
โˆ™ ๐œ‘ : ๐ด → ๐ต is an order-preserving function if for all ๐‘Ž1 , ๐‘Ž2 ∈ ๐ด,
(๐‘Ž1 , ๐‘Ž2 ) ∈ ๐‘… if and only if (๐œ‘(๐‘Ž1 ), ๐œ‘(๐‘Ž2 )) ∈ ๐‘†
and we say that ๐œ‘ preserves ๐‘… with ๐‘†.
โˆ™ An order-preserving bijection is an order isomorphism.
โˆ™ (๐ด, ๐‘…) is order isomorphic to (๐ต, ๐‘†) and we write (๐ด, ๐‘…) ≅ (๐ต, ๐‘†) if
there exists an order isomorphism ๐œ‘ : ๐ด → ๐ต preserving ๐‘… with ๐‘†. If
(๐ด, ๐‘…) ≅ (๐ต, ๐‘†) and the relations ๐‘… and ๐‘† are clear from context, we can
write ๐ด ≅ ๐ต. Sometimes (๐ด, ๐‘…) and (๐ต, ๐‘†) are said to have the same order
type when they are order isomorphic.
An isomorphism pairs elements from two sets in such a way that the orders on the two
sets appear to be the same.
Section 4.5 INJECTIONS AND SURJECTIONS
213
EXAMPLE 4.5.25
Define ๐‘“ : ๐‘+ → ๐‘− by ๐‘“ (๐‘ฅ) = −๐‘ฅ (Example 3.1.12). Clearly, ๐‘“ is a bijection.
Moreover, ๐‘“ preserves ≤ with ≥. To prove this, let ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐‘+ . Then,
๐‘ฅ1 ≤ ๐‘ฅ1 ⇔ −๐‘ฅ1 ≥ −๐‘ฅ2 ⇔ ๐‘“ (๐‘ฅ1 ) ≥ ๐‘“ (๐‘ฅ2 ).
Therefore, (๐‘+ , ≤) ≅ (๐‘− , ≥).
Observe that the inverse of (4.8) is the function
๐œ‘−1 : {0} × ๐™ → ๐™ × {0}
such that ๐œ‘−1 (0, ๐‘š) = (๐‘š, 0). This function preserves 4′ with 4. This result is generalized and proved in the next theorem.
THEOREM 4.5.26
The inverse of an order isomorphism preserving ๐‘… with ๐‘† is an order isomorphism preserving ๐‘† with ๐‘….
PROOF
Let ๐œ‘ : ๐ด → ๐ต be an order isomorphism preserving ๐‘… with ๐‘†. By Theorem 4.5.23, ๐œ‘−1 is a bijection. Suppose that (๐‘1 , ๐‘2 ) ∈ ๐‘†. Since ๐œ‘ is onto,
there exists ๐‘Ž1 , ๐‘Ž2 ∈ ๐ด such that ๐œ‘(๐‘Ž1 ) = ๐‘1 and ๐œ‘(๐‘Ž2 ) = ๐‘2 . This implies
that (๐œ‘(๐‘Ž1 ), ๐œ‘(๐‘Ž2 )) ∈ ๐‘†. Since ๐œ‘ is an isomorphism preserving ๐‘… with ๐‘†,
we have that (๐‘Ž1 , ๐‘Ž2 ) ∈ ๐‘…. However, ๐‘Ž1 = ๐œ‘−1 (๐‘1 ) and ๐‘Ž2 = ๐œ‘−1 (๐‘2 ), so
(๐œ‘−1 (๐‘1 ), ๐œ‘−1 (๐‘2 )) ∈ ๐‘…. Therefore, ๐œ‘−1 : ๐ต → ๐ด is an isomorphism preserving
๐‘† with ๐‘….
Also, observe that ๐‘” : ๐‘ → (0, ∞) defined by ๐‘”(๐‘ฅ) = ๐‘’๐‘ฅ is an order isomorphism
preserving ≤ with ≤. Using ๐‘“ from Example 4.5.25, the composition ๐‘“ โ—ฆ ๐‘” is an order
isomorphism ๐‘ → (−∞, 0) such that (๐‘“ โ—ฆ ๐‘”)(๐‘ฅ) = −๐‘’๐‘ฅ . That this happens in general is
the next theorem. Its proof is left to Exercise 25.
THEOREM 4.5.27
If ๐œ‘ : ๐ด → ๐ต is an isomorphism preserving ๐‘… with ๐‘† and ๐œ“ : ๐ต → ๐ถ is an
isomorphism preserving ๐‘† with ๐‘‡ , then ๐œ“ โ—ฆ ๐œ‘ : ๐ด → ๐ถ is an isomorphism
preserving ๐‘… with ๐‘‡ .
EXAMPLE 4.5.28
Let ๐‘… be a relation on ๐ด, ๐‘† be a relation on ๐ต, and ๐‘‡ be a relation on ๐ถ.
โˆ™ Since the identity map is an order isomorphism, (๐ด, ๐‘…) ≅ (๐ด, ๐‘…).
โˆ™ Suppose (๐ด, ๐‘…) ≅ (๐ต, ๐‘†). This means that there exists an order isomorphism ๐œ‘ : ๐ด → ๐ต preserving ๐‘… with ๐‘†. By Theorem 4.5.26, ๐œ‘−1 is an
order isomorphism preserving ๐‘† with ๐‘…. Therefore, (๐ต, ๐‘†) ≅ (๐ด, ๐‘…).
214
Chapter 4 RELATIONS AND FUNCTIONS
โˆ™ Let (๐ด, ๐‘…) ≅ (๐ต, ๐‘†) and (๐ต, ๐‘†) ≅ (๐ถ, ๐‘‡ ). By Theorem 4.5.27, we conclude
that (๐ด, ๐‘…) ≅ (๐ถ, ๐‘‡ ).
If an order-preserving function is one-to-one, even it is not a surjection, the function
still provides an order isomorphism between its domain and range. This concept is
named by the next definition.
DEFINITION 4.5.29
๐œ‘ : ๐ด → ๐ต is an embedding if ๐œ‘ is an order isomorphism ๐ด → ran(๐œ‘).
For example, ๐‘“ : ๐™ → ๐ such that ๐‘“ (๐‘›) = ๐‘› is an embedding preserving ≤ and
๐œ‹ : ๐‘2 → ๐‘3 such that ๐œ“(๐‘ฅ, ๐‘ฆ) = (๐‘ฅ, ๐‘ฆ, 0) is an embedding preserving the lexicographical order (Exercise 4.3.16). Although ๐‘2 is not a subset of ๐‘3 , we view the image of
๐œ“ as a copy of ๐‘2 in ๐‘3 that preserves the orders.
Exercises
1. Show that the given pairs of functions are inverses.
(a) ๐‘“ (๐‘ฅ) = 3๐‘ฅ + 2 and ๐‘”(๐‘ฅ) = 31 ๐‘ฅ − 23
(b) ๐œ‘(๐‘Ž, ๐‘) = (2๐‘Ž, ๐‘ + 2) and ๐œ“(๐‘Ž, ๐‘) = ( 12 ๐‘Ž, ๐‘ − 2)
(c) ๐‘“ (๐‘ฅ) = ๐‘Ž๐‘ฅ and ๐‘”(๐‘ฅ) = log๐‘Ž ๐‘ฅ, where ๐‘Ž > 0
2. For each function, graph the indicated restriction.
(a) ๐‘“ (0, ∞), ๐‘“ (๐‘ฅ) = ๐‘ฅ2
(b) ๐‘” [−5, −2], ๐‘”(๐‘ฅ) = |๐‘ฅ|
[
]
(c) โ„Ž 0, ๐œ‹โˆ•2 , โ„Ž(๐‘ฅ) = cos ๐‘ฅ
3. Prove that the given functions are one-to-one.
(a) ๐‘“ : ๐‘ → ๐‘, ๐‘“ (๐‘ฅ) = 2๐‘ฅ + 1
(b) ๐‘” : ๐‘2 → ๐‘2 , ๐‘”(๐‘ฅ, ๐‘ฆ) = (3๐‘ฆ, 2๐‘ฅ)
(c) โ„Ž : ๐‘ โงต {9} → ๐‘ โงต {0}, โ„Ž(๐‘ฅ) = 1โˆ•(๐‘ฅ − 9)
(d) ๐œ‘ : ๐™ × ๐‘ → ๐™ × (0, ∞), ๐œ‘(๐‘›, ๐‘ฅ) = (3๐‘›, ๐‘’๐‘ฅ )
(e) ๐œ“ : P(๐ด) → P(๐ต), ๐œ“(๐ถ) = ๐ถ ∪ {๐‘} where ๐ด ⊂ ๐ต and ๐‘ ∈ ๐ต โงต ๐ด
4. Let ๐‘“ : (๐‘Ž, ๐‘) → (๐‘, ๐‘‘) be defined by
๐‘“ (๐‘ฅ) =
๐‘‘−๐‘
(๐‘ฅ − ๐‘Ž) + ๐‘.
๐‘−๐‘Ž
Graph ๐‘“ and show that it is a bijection.
5. Let ๐‘“ and ๐‘” be functions such that ran(๐‘”) ⊆ dom(๐‘“ ).
(a) Prove that if ๐‘“ โ—ฆ ๐‘” is one-to-one, ๐‘” is one-to-one.
(b) Give an example of functions ๐‘“ and ๐‘” such that ๐‘“ โ—ฆ ๐‘” is one-to-one, but ๐‘“ is
not one-to-one.
6. Define ๐œ‘ : ๐™ → ๐™๐‘š by ๐œ‘(๐‘˜) = [๐‘˜]๐‘š . Show that ๐œ‘ is not one-to-one.
Section 4.5 INJECTIONS AND SURJECTIONS
215
7. Show that the given functions are not one-to-one.
(a) ๐‘“ : ๐‘ → ๐‘, ๐‘“ (๐‘ฅ) = ๐‘ฅ4 + 3
(b) ๐‘” : ๐‘ → ๐‘, ๐‘”(๐‘ฅ) = |๐‘ฅ − 2| + 4
(c) ๐œ‘ : P(๐ด) → {{๐‘Ž}, ∅}, ๐œ‘(๐ต) = ๐ต ∩ {๐‘Ž}, where ๐‘Ž ∈ ๐ด and ๐ด has at least two
elements
(d) ๐œ–5 : ๐‘ ๐‘ → ๐‘, ๐œ–5 (๐‘“ ) = ๐‘“ (5)
8. Let ๐‘“ : ๐‘ → ๐‘ be periodic (Exercise 4.4.31). Prove that ๐‘“ is not one-to-one.
9. Show that the given functions are onto.
(a) ๐‘“ : ๐‘ → ๐‘, ๐‘“ (๐‘ฅ) = 2๐‘ฅ + 1
(b) ๐‘” : ๐‘ → (0, ∞), ๐‘”(๐‘ฅ) = ๐‘’๐‘ฅ
(c) โ„Ž : ๐‘ โงต {0} → ๐‘ โงต {0}, โ„Ž(๐‘ฅ) = 1โˆ•๐‘ฅ
(d) ๐œ‘ : ๐™ × ๐™ → ๐™, ๐œ‘(๐‘Ž, ๐‘) = ๐‘Ž + ๐‘
(e) ๐œ–5 : ๐‘ ๐‘ → ๐‘, ๐œ–5 (๐‘“ ) = ๐‘“ (5)
10. Show that the given functions are not onto.
(a) ๐‘“ : ๐‘ → ๐‘, ๐‘“ (๐‘ฅ) = ๐‘’๐‘ฅ
(b) ๐‘” : ๐‘ → ๐‘, ๐‘”(๐‘ฅ) = |๐‘ฅ|
(c) ๐œ‘ : ๐™ × ๐™ → ๐™ × ๐™, ๐œ‘(๐‘Ž, ๐‘) = (3๐‘Ž, ๐‘2 )
(d) ๐œ“ : ๐‘ → ๐‘ ๐‘, ๐œ“(๐‘Ž) = ๐‘“ , where ๐‘“ (๐‘ฅ) = ๐‘Ž for all ๐‘ฅ ∈ ๐‘
11. Let ๐‘“ and ๐‘” be functions such that ran(๐‘”) ⊆ dom(๐‘“ ).
(a) Prove that if ๐‘“ โ—ฆ ๐‘” is onto, then ๐‘“ is onto.
(b) Give an example of functions ๐‘“ and ๐‘” such that ๐‘“ โ—ฆ ๐‘” is onto, but ๐‘” is not
onto.
12. Define ๐œ‘ : ๐ × ๐™ → ๐™ × ๐ by ๐œ‘(๐‘ฅ, ๐‘ฆ) = (๐‘ฆ, ๐‘ฅ). Show that ๐œ‘ is a bijection.
13. Show that the function ๐›พ : ๐ด × ๐ต → ๐ถ × ๐ท defined by
๐›พ(๐‘Ž, ๐‘) = (๐œ‘(๐‘Ž), ๐œ“(๐‘))
is a bijection if both ๐œ‘ : ๐ด → ๐ถ and ๐œ“ : ๐ต → ๐ท are bijections.
14. Define ๐›พ : ๐ด × (๐ต × ๐ถ) → (๐ด × ๐ต) × ๐ถ by
๐›พ(๐‘Ž, (๐‘, ๐‘)) = ((๐‘Ž, ๐‘), ๐‘).
Prove ๐›พ is a bijection.
15. Demonstrate that the inverse of a bijection is a bijection.
16. Prove that the empty set is a bijection with domain and range equal to ∅.
17. Let ๐ด ⊆ ๐‘ and define ๐œ‘ : ๐‘ ๐‘ → ๐ด ๐‘ by ๐œ‘(๐‘“ ) = ๐‘“ ๐ด. Is ๐œ‘ always one-to-one? Is
it always onto? Explain.
18. A function ๐‘“ : ๐ด → ๐ต has a left inverse if there exists a function ๐‘” : ๐ต → ๐ด such
that ๐‘” โ—ฆ ๐‘“ = ๐ผ๐ด . Prove that a function is one-to-one if and only if it has a left inverse.
216
Chapter 4 RELATIONS AND FUNCTIONS
19. A function ๐‘“ : ๐ด → ๐ต has a right inverse if there exists a function ๐‘” : ๐ต → ๐ด so
that ๐‘“ โ—ฆ ๐‘” = ๐ผ๐ต . Prove a function is onto if and only if it has a right inverse.
20. Let (๐ด, 4) be a poset. For every bijection ๐œ‘ : ๐ด → ๐ด, ๐œ‘ is increasing (Exercise 4.4.32) if and only if ๐œ‘−1 is increasing.
21. Show that if a real-valued function is increasing, it is one-to-one.
22. Prove or show false this modification of Theorem 4.5.23: If ๐œ‘ : ๐ด → ๐ต and
๐œ“ : ๐ถ → ๐ท are bijections with ran(๐œ‘) ⊆ ๐ถ, then ๐œ“ โ—ฆ ๐œ‘ is a bijection.
23. Let ๐ด, ๐ต ⊆ ๐‘ be two sets ordered by ≤. Let ๐ด be well-ordered by ≤. Prove that if
๐‘“ : ๐ด → ๐ต is an order-preserving surjection, ๐ต is well-ordered by ≤.
24. Let ๐œ‘ : ๐ด → ๐ต be an isomorphism preserving ๐‘… with ๐‘…′ . Let ๐ถ ⊆ ๐ด and ๐ท ⊆ ๐ต.
Prove the following.
(a) (๐ถ, ๐‘… ∩ [๐ถ × ๐ถ]) ≅ (๐œ‘ [๐ถ] , ๐‘…′ ∩ [๐œ‘ [๐ถ] × ๐œ‘ [๐ถ]])
(b) (๐ท, ๐‘…′ ∩ [๐ท × ๐ท]) ≅ (๐œ‘−1 [๐ท] , ๐‘… ∩ [๐œ‘−1 [๐ท] × ๐œ‘−1 [๐ท]])
25. Prove Theorem 4.5.27.
26. Define ๐‘“ : ๐‘ → ๐‘ by ๐‘“ (๐‘ฅ) = 2๐‘ฅ + 1. Prove that ๐‘“ is an order isomorphism
preserving < with <.
27. Suppose that (๐ด, ๐‘…) and (๐ต, ๐‘†) are posets. Let ๐œ‘ : ๐ด → ๐ต be an order isomorphism preserving ๐‘… with ๐‘† and ๐ถ ⊆ ๐ด. Prove that if ๐‘š is the least element of ๐ถ with
respect to ๐‘…, then ๐œ‘(๐‘š) is the least element of ๐œ‘ [๐ถ] with respect to ๐‘†.
28. Find linear orders (๐ด, 4) and (๐ต, 4′ ) such that each is isomorphic to a subset of
the other but (๐ด, 4) is not isomorphic to (๐ต, 4′ ).
29. Let (๐ด, 4) be a poset. Prove that there exists ๐ต ⊆ P(๐ด) such that (๐ด, 4) ≅ (๐ต, ⊆).
4.6
IMAGES AND INVERSE IMAGES
So far we have focused on the image of single element in the domain of a function.
Sometimes we will need to examine a set of images.
DEFINITION 4.6.1
Let ๐œ‘ : ๐ด → ๐ต be a function and ๐ถ ⊆ ๐ด. The image of ๐ถ (under ๐œ‘) is
๐œ‘ [๐ถ] = {๐œ‘(๐‘ฅ) : ๐‘ฅ ∈ ๐ถ}.
Notice that ๐œ‘ [๐ถ] ⊆ ๐ต (Figure 4.17) and ๐œ‘ [๐ด] = ran(๐œ‘).
A similar definition can be made with subsets of the codomain.
DEFINITION 4.6.2
Let ๐œ‘ : ๐ด → ๐ต be a function and ๐ท ⊆ ๐ต. The inverse image of ๐ท (under ๐œ‘) is
๐œ‘−1 [๐ท] = {๐‘ฅ ∈ ๐ด : ๐œ‘(๐‘ฅ) ∈ ๐ท}.
Section 4.6 IMAGES AND INVERSE IMAGES
217
๐œ‘[C]
A
B
๐œ‘
x
y
C
Figure 4.17
The image of ๐ถ under ๐œ‘.
D
A
B
๐œ‘
x
y
๐œ‘−1 [D]
Figure 4.18
The inverse image of ๐ท under ๐œ‘.
Observe that ๐œ‘−1 [๐ท] ⊆ ๐ด (Figure 4.18) and ๐œ‘−1 [ran(๐œ‘)] = ๐ด.
EXAMPLE 4.6.3
Let ๐‘“ = {(1, 2), (2, 4), (3, 5), (4, 5)}. This set is a function. Its domain is {1, 2, 3, 4},
and its range is {2, 4, 5}. Then,
๐‘“ [{1, 3}] = {2, 5}
and
๐‘“ −1 [{5}] = {3, 4}.
EXAMPLE 4.6.4
Define ๐‘“ : ๐‘ → ๐‘ by ๐‘“ (๐‘ฅ) = ๐‘ฅ2 + 1.
โˆ™ To prove ๐‘“ [(1, 2)] = (2, 5), we show both inclusions. Let ๐‘ฆ ∈ ๐‘“ [(1, 2)].
Then, ๐‘ฆ = ๐‘ฅ2 + 1 for some ๐‘ฅ ∈ (1, 2). By a little algebra, we see that
2 < ๐‘ฅ2 + 1 < 5. Hence, ๐‘ฆ ∈ (2, 5). Conversely, let ๐‘ฆ ∈ (2, 5). Because
√
2 < ๐‘ฆ < 5 ⇔ 1 < ๐‘ฆ − 1 < 4 ⇔ 1 < ๐‘ฆ − 1 < 2,
218
Chapter 4 RELATIONS AND FUNCTIONS
(a) ๐‘“ [(1, 2)] = (2, 5)
Figure 4.19
√
(b) ๐‘“ −1 [(2, 5)] = (−2, −1) ∪ (1, 2)
The image and inverse image of a set under ๐‘“ .
๐‘ฆ − 1 ∈ (1, 2). Furthermore,
√
√
√
๐‘“ ( ๐‘ฆ − 1) = ๐‘“ ( ๐‘ฆ − 1) = ( ๐‘ฆ − 1)2 + 1 = ๐‘ฆ.
Therefore, ๐‘ฆ ∈ ๐‘“ [(1, 2)]. This is illustrated in Figure 4.19(a).
โˆ™ Simply because ๐‘“ [(1, 2)] = (2, 5), we cannot conclude ๐‘“ −1 [(2, 5)] equals
(1, 2). Instead, ๐‘“ −1 [(2, 5)] = (−2, −1) ∪ (1, 2), as seen in Figure 4.19(b),
because
๐‘ฅ ∈ (−2, −1) ∪ (1, 2) ⇔ −2 < ๐‘ฅ < −1 or 1 < ๐‘ฅ < 2
⇔ 1 < ๐‘ฅ2 < 4
⇔ 2 < ๐‘ฅ2 + 1 < 5
⇔ ๐‘“ (๐‘ฅ) ∈ (2, 5)
⇔ ๐‘ฅ ∈ ๐‘“ −1 [(2, 5)] .
โˆ™ To show that ๐‘“ −1 [(−2, −1)] is empty, take ๐‘ฅ ∈ ๐‘“ −1 [(−2, −1)]. This
means that −2 < ๐‘“ (๐‘ฅ) < −1, but this is impossible because ๐‘“ (๐‘ฅ) is positive.
Let ๐‘“ = {(1, 3), (2, 3), (3, 4), (4, 5)}. This is a function {1, 2, 3, 4} → {3, 4, 5}.
Notice that
๐‘“ [{1} ∪ {2, 3}] = {3, 4} = ๐‘“ [{1}] ∪ ๐‘“ [{2, 3}],
๐‘“ [{1, 2} ∩ {3}] = ๐‘“ [∅] = ∅ ⊆ {3} = ๐‘“ [{1}] ∩ ๐‘“ [{2, 3}],
๐‘“ −1 [{3, 4} ∪ {5}] = {1, 2, 3, 4} = ๐‘“ −1 [{3, 4}] ∪ ๐‘“ −1 [{5}],
Section 4.6 IMAGES AND INVERSE IMAGES
219
and
๐‘“ −1 [{3, 4} ∩ {5}] = ∅ = ๐‘“ −1 [{3, 4}] ∩ ๐‘“ −1 [{5}].
This result concerning the interaction of images and inverse images with union and
intersection is generalized in the next theorem.
THEOREM 4.6.5
Let ๐œ‘ : ๐ด → ๐ต be a function with ๐ถ, ๐ท ⊆ ๐ด and ๐ธ, ๐น ⊆ ๐ต.
โˆ™ ๐œ‘ [๐ถ ∪ ๐ท] = ๐œ‘ [๐ถ] ∪ ๐œ‘ [๐ท].
โˆ™ ๐œ‘ [๐ถ ∩ ๐ท] ⊆ ๐œ‘ [๐ถ] ∩ ๐œ‘ [๐ท].
โˆ™ ๐œ‘−1 [๐ธ ∪ ๐น ] = ๐œ‘−1 [๐ธ] ∪ ๐œ‘−1 [๐น ].
โˆ™ ๐œ‘−1 [๐ธ ∩ ๐น ] = ๐œ‘−1 [๐ธ] ∩ ๐œ‘−1 [๐น ].
PROOF
We prove the first and third parts, leaving the others for Exercise 7. By Exercise 3.2.8(d),
๐‘ฆ ∈ ๐œ‘ [๐ถ ∪ ๐ท] ⇔ ∃๐‘ฅ(๐‘ฅ ∈ ๐ถ ∪ ๐ท ∧ ๐œ‘(๐‘ฅ) = ๐‘ฆ)
⇔ ∃๐‘ฅ(๐‘ฅ ∈ ๐ถ ∧ ๐œ‘(๐‘ฅ) = ๐‘ฆ) ∨ ∃๐‘ฅ(๐‘ฅ ∈ ๐ท ∧ ๐œ‘(๐‘ฅ) = ๐‘ฆ)
⇔ ๐‘ฆ ∈ ๐œ‘ [๐ถ] ∨ ๐‘ฆ ∈ ๐œ‘ [๐ท]
⇔ ๐‘ฆ ∈ ๐œ‘ [๐ถ] ∪ ๐œ‘ [๐ท] .
In addition,
๐‘ฅ ∈ ๐œ‘−1 [๐ธ ∪ ๐น ] ⇔ ๐œ‘(๐‘ฅ) ∈ ๐ธ ∪ ๐น
⇔ ๐œ‘(๐‘ฅ) ∈ ๐ธ ∨ ๐œ‘(๐‘ฅ) ∈ ๐น
⇔ ๐‘ฅ ∈ ๐œ‘−1 [๐ธ] ∨ ๐‘ฅ ∈ ๐œ‘−1 [๐น ]
⇔ ๐‘ฅ ∈ ๐œ‘−1 [๐ธ] ∪ ๐œ‘−1 [๐น ] .
It might seem surprising that we only have an inclusion in the second part of Theorem 4.6.5. To see that the other inclusion is false, let ๐‘“ = {(1, 3), (2, 3)}. Then,
๐‘“ [{1} ∩ {2}] = ๐‘“ [∅] = ∅,
but
๐‘“ [{1}] ∩ ๐‘“ [{2}] = {3} ∩ {3} = {3}.
Hence, ๐‘“ [{1}] ∩ ๐‘“ [{2}] * ๐‘“ [{1} ∩ {2}]. However, if ๐‘“ had been a bijection, the
inclusion would hold (Exercise 13).
220
Chapter 4 RELATIONS AND FUNCTIONS
EXAMPLE 4.6.6
Let ๐‘“ (๐‘ฅ) = ๐‘ฅ2 + 1. We check the union results of Theorem 4.6.5.
โˆ™ We have already seen in Example 4.6.4 that ๐‘“ [(1, 2)] = (2, 5). Since
(1, 2) = (1, 1.5] ∪ [1.5, 2), apply ๐‘“ to both of these intervals and find that
๐‘“ [(1, 1.5]] = (2, 3.25],
๐‘“ [[1.5, 2)] = [3.25, 5).
Therefore, ๐‘“ [(1, 2)] = ๐‘“ [(1, 1.5]] ∪ ๐‘“ [[1.5, 2)].
โˆ™ Also, from Example 4.6.4, ๐‘“ −1 [(2, 5)] = (−2, −1) ∪ (1, 2). We can write
(2, 5) as the union of (2, 4) and (3, 5). Since
) ( √ )
( √
๐‘“ −1 [(2, 4)] = − 3, −1 ∪ 1, 3 ,
(
√ ) (√ )
๐‘“ −1 [(3, 5)] = −2, − 2 ∪
2, 2 ,
we have that
๐‘“ −1 [(2, 5)] = ๐‘“ −1 [(2, 4)] ∪ ๐‘“ −1 [(3, 5)].
Theorem 4.6.5 can be modified to arbitrary unions and intersections. The proof is
left to Excercise 17.
THEOREM 4.6.7
Let {๐ด๐‘– : ๐‘– ∈ ๐ผ} be a family of sets.
[โ‹ƒ ] โ‹ƒ
โˆ™ ๐œ‘
๐ด๐‘– =
๐œ‘[๐ด๐‘– ]
๐‘–∈๐ผ
๐‘–∈๐ผ
[โ‹‚ ] โ‹‚
โˆ™ ๐œ‘
๐ด๐‘– ⊆
๐œ‘[๐ด๐‘– ]
๐‘–∈๐ผ
โˆ™ ๐œ‘−1
[โ‹ƒ
๐‘–∈๐ผ
โˆ™ ๐œ‘−1
[โ‹‚
๐‘–∈๐ผ
๐‘–∈๐ผ
] โ‹ƒ
๐ด๐‘– =
๐œ‘−1 [๐ด๐‘– ]
๐‘–∈๐ผ
] โ‹‚
๐ด๐‘– =
๐œ‘−1 [๐ด๐‘– ].
๐‘–∈๐ผ
We know that the composition of a function and its inverse equals the identity map
(Theorem 4.5.3). The last two results of the section show a similar result involving the
image and inverse image using functions that are either one-to-one or onto.
Section 4.6 IMAGES AND INVERSE IMAGES
221
THEOREM 4.6.8
Let ๐œ‘ : ๐ด → ๐ต be a function. Suppose ๐ถ ⊆ ๐ด and ๐ท ⊆ ๐ต.
โˆ™ If ๐œ‘ is one-to-one, then ๐œ‘−1 [๐œ‘[๐ถ]] = ๐ถ.
โˆ™ If ๐œ‘ is onto, then ๐œ‘[๐œ‘−1 [๐ท]] = ๐ท.
PROOF
We prove the first part and leave the second to Exercise 8. Suppose that ๐œ‘ is an
injection. We show that ๐œ‘−1 [๐œ‘ [๐ถ]] = ๐ถ.
โˆ™ Let ๐‘ฅ ∈ ๐œ‘−1 [๐œ‘[๐ถ]]. This means that there exists ๐‘ฆ ∈ ๐œ‘[๐ถ] such that
๐œ‘(๐‘ฅ) = ๐‘ฆ. Furthermore, there is a ๐‘ง ∈ ๐ถ so that ๐œ‘(๐‘ง) = ๐‘ฆ. Therefore, since
๐œ‘ is one-to-one, ๐‘ฅ = ๐‘ง, which means that ๐‘ฅ ∈ ๐ถ.
โˆ™ This step will work for any function. Take ๐‘ฅ ∈ ๐ถ, so ๐œ‘(๐‘ฅ) ∈ ๐œ‘[๐ถ]. By
definition,
๐œ‘−1 [๐œ‘[๐ถ]] = {๐‘ง ∈ ๐ด : ๐œ‘(๐‘ง) ∈ ๐œ‘[๐ถ]}.
Since ๐‘ฅ is also an element of ๐ด, we conclude that ๐‘ฅ ∈ ๐œ‘−1 [๐œ‘[๐ถ]].
Because we used the one-to-one condition only to prove ๐œ‘−1 [๐œ‘[๐ถ]] ⊆ ๐ถ, we suspect
that this inclusion is false if the function is not one-to-one. To see this, let ๐‘“ : ๐‘ → ๐‘
be defined as ๐‘“ (๐‘ฅ) = ๐‘ฅ2 . We know that ๐‘“ is not one-to-one. Choose ๐ถ = {1}. Then,
๐‘“ −1 [๐‘“ [๐ถ]] = ๐‘“ −1 [{1}] = {−1, 1}.
Therefore, ๐‘“ −1 [๐‘“ [๐ถ]] * ๐ถ.
When examining the example, we might conjecture that the function being one-toone is necessary for equality. That this is the case is the following theorem.
THEOREM 4.6.9
Let ๐œ‘ : ๐ด → ๐ต be a function.
โˆ™ If ๐œ‘−1 [๐œ‘ [๐ถ]] = ๐ถ for all ๐ถ ⊆ ๐ด, then ๐œ‘ is one-to-one.
[
]
โˆ™ If ๐œ‘ ๐œ‘−1 [๐ท] = ๐ท for all ๐ท ⊆ ๐ต, then ๐œ‘ is onto.
PROOF
As with the previous theorem, we prove only the first part. The second part is
Exercise 8. Assume
๐œ‘−1 [๐œ‘[๐ถ]] = ๐ถ for all ๐ถ ⊆ ๐ด.
Take ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐ด and let ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ2 ). Now,
{๐‘ฅ1 } = ๐œ‘−1 [๐œ‘[{๐‘ฅ1 }]] = {๐‘ง ∈ ๐ด : ๐œ‘(๐‘ง) = ๐œ‘(๐‘ฅ1 )}.
Because ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ2 ), we have that {๐‘ฅ1 , ๐‘ฅ2 } ⊆ {๐‘ง ∈ ๐ด : ๐œ‘(๐‘ง) = ๐œ‘(๐‘ฅ1 )}, so
we must have ๐‘ฅ1 = ๐‘ฅ2 .
222
Chapter 4 RELATIONS AND FUNCTIONS
Exercises
1. Let ๐‘“
(a)
(b)
(c)
(d)
: ๐‘ → ๐‘ be defined by ๐‘“ (๐‘ฅ) = 2๐‘ฅ + 1. Find the images and inverse images.
๐‘“ [(1, 3]]
๐‘“ [(−∞, 0)]
๐‘“ −1 [(−1, 1)]
๐‘“ −1 [(0, 2) ∪ (5, 8)]
2. Let ๐‘”
(a)
(b)
(c)
(d)
: ๐‘ → ๐‘ be the function ๐‘”(๐‘ฅ) = ๐‘ฅ4 − 1. Find the images and inverse images.
๐‘”[{0}]
๐‘”[๐™]
๐‘” −1 [{0, 15}]
๐‘” −1 [[−9, −5] ∪ [0, 5]]
3. Define ๐œ‘ : ๐‘ × ๐‘ → ๐™ by ๐œ‘(๐‘Ž, ๐‘) = โŸฆ๐‘ŽโŸง + โŸฆ๐‘โŸง. Find the images and inverse images.
(a) ๐œ‘[{0} × ๐‘]
(b) ๐œ‘[(0, 1) × (0, 1)]
(c) ๐œ‘−1 [{2, 4}]
(d) ๐œ‘−1 [๐]
4. Let ๐œ“
(a)
(b)
(c)
(d)
: ๐™ → ๐™ be a function and define ๐›พ : P(๐™) → P(๐™) by ๐›พ(๐ถ) = ๐œ“[๐ถ].
Prove that ๐›พ is well-defined.
Let ๐œ“(๐‘›) = 2๐‘›. Find ๐›พ[{1, 2, 3}], ๐›พ[๐™], ๐›พ −1 [{1, 2, 3}], and ๐›พ −1 [๐™].
Under what conditions is ๐›พ one-to-one?
Under what conditions is ๐›พ onto?
5. For any function ๐œ“, show that ๐œ“[∅] = ∅ and ๐œ“ −1 [∅] = ∅.
6. Prove for every ๐ต ⊆ ๐ด, ๐ผ๐ด [๐ต] = ๐ต and (๐ผ๐ด )−1 [๐ต] = ๐ต.
7. Prove the remaining parts of Theorem 4.6.5.
8. Prove the unproven parts of Theorems 4.6.8 and 4.6.9.
9. Let ๐œ‘ : ๐ด → ๐ต be an injection and ๐ถ ⊆ ๐ด.
(a) Prove ๐œ‘(๐‘ฅ) ∈ ๐œ‘[๐ถ] if and only if ๐‘ฅ ∈ ๐ถ.
(b) Show that Exercise 9(a) is false if the function is not one-to-one.
10. If ๐œ“ is a function, ๐ด ⊆ ๐ต ⊆ dom(๐œ“), and ๐ถ ⊆ ๐ท ⊆ ran(๐œ“), show that both
๐œ“[๐ด] ⊆ ๐œ“[๐ต] and ๐œ“ −1 [๐ถ] ⊆ ๐œ“ −1 [๐ท].
11. Let ๐œ‘ : ๐ด → ๐ต be a function and take disjoint sets ๐‘ˆ and ๐‘‰ .
(a) Prove false: If ๐‘ˆ , ๐‘‰ ⊆ ๐ด, then ๐œ‘[๐‘ˆ ] ∩ ๐œ‘[๐‘‰ ] = ∅.
(b) Prove false: If ๐‘ˆ , ๐‘‰ ⊆ ๐ต, then ๐œ‘−1 [๐‘ˆ ] ∩ ๐œ‘−1 [๐‘‰ ] = ∅.
(c) What additional assumption is needed to prove both of the implications?
12. Assume that ๐œ‘ and ๐œ“ are functions such that ran(๐œ“) ⊆ dom(๐œ‘). Let ๐ด be a subset
of dom(๐œ“). Prove or show false with a counterexample: (๐œ‘ โ—ฆ ๐œ“)[๐ด] = ๐œ‘[๐œ“[๐ด]].
13. Let ๐œ‘ : ๐ด → ๐ต be one-to-one. Prove the following.
Section 4.6 IMAGES AND INVERSE IMAGES
223
(a) ๐œ‘[๐ด] ∩ ๐œ‘[๐ต] ⊆ ๐œ‘[๐ด ∩ ๐ต].
(b) If ๐ถ ⊆ ๐ด and ๐ท ⊆ ๐ต, then ๐œ‘[๐ถ] = ๐ท if and only if ๐œ‘−1 [๐ท] = ๐ถ.
14. Prove that if ๐œ‘ : ๐ด → ๐ต is a bijection and ๐ถ ⊆ ๐ด, then ๐œ‘[๐ด โงต ๐ถ] = ๐ต โงต ๐œ‘[๐ถ].
15. Find a function ๐œ‘ : ๐ด → ๐ต and a set ๐ท ⊆ ๐ต such that ๐ท * ๐œ‘[๐œ‘−1 [๐ท]].
16. Let ๐œ“ be an injection with ๐ด ⊆ dom(๐œ“) and ๐ต ⊆ ran(๐œ“). Prove that
๐ต ⊆ ๐œ“[๐ด] if and only if ๐œ“ −1 [๐ต] ⊆ ๐ด.
17. Prove Theorem 4.6.7.
CHAPTER 5
AXIOMATIC SET THEORY
5.1
AXIOMS
When we began studying set theory in Chapter 3, we made several assumptions regarding which things are sets. For example, we assumed that collections of numbers,
like ๐, ๐‘, or (0, ∞) are sets. We supposed that operating with given sets to form new
collections, as with union or intersection, resulted in sets. We also assumed that formulas could be used to describe certain sets. All of this seemed perfectly reasonable,
but since all of these assumptions were made without a carefully thought-out system,
we would be wise to pause and investigate if we have introduced any problems.
Consider the following question. Given a formula ๐‘(๐‘ฅ), is there a set of the form
{๐‘ฅ : ๐‘(๐‘ฅ)}? Consistent with the attitude of our previous work, we might quickly answer
in the affirmative. Mathematicians, including Cantor, also initially thought that this was
the case. However, it was shown independently by Bertrand Russell and Ernst Zermelo
that not every formula can be used to define a set. For example, let ๐‘(๐‘ฅ) := ๐‘ฅ ∉ ๐‘ฅ and
๐ด = {๐‘ฅ : ๐‘(๐‘ฅ)} and consider whether ๐ด is an element of itself. If ๐ด ∈ ๐ด, then due
to the definition of ๐‘(๐‘ฅ), ๐ด ∉ ๐ด, and if ๐ด ∉ ๐ด, then ๐ด ∈ ๐ด. Because of this built-in
contradiction, it is impossible for ๐ด to be a set. This is known as Russell’s paradox,
and it was a serious challenge to set theory. One solution would have been to dismiss
225
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
226
Chapter 5 AXIOMATIC SET THEORY
set theory altogether. The problem was that this new subject combined with advances
in logic appeared to promise a framework in which to study foundational questions of
mathematics. David Hilbert famously supported set theory by remarking that “no one
will drive us out of this paradise that Cantor has created for us,” so dismissal was not
an option. In order to prevent contradictions such as Russell’s paradox from appearing,
mathematicians settled on the method of Euclid as the solution, but instead of assuming
geometric postulates, over a period of time, certain set-theoretic axioms were chosen.
Their purpose was to define a system by which one could determine whether a given
collection should be considered a set in such a manner that prevented any contradictions
from arising. In this chapter we identify the axioms and then redefine ๐, ๐‘, and other
collections so that we are confident in our previous assumptions regarding them being
sets.
Equality Axioms
We begin with the basics. Although not officially among the set axioms, = is always
assumed to satisfy the following rules. They are defined to replicate the standard reasoning of equality that we have been using in previous chapters.
AXIOMS 5.1.1 [Equality]
Let ๐‘ฅ, ๐‘ฆ, and ๐‘ง be variable symbols from theory symbols ๐–ฒ.
โˆ™ [E1] ๐‘ฅ = ๐‘ฅ.
โˆ™ [E2] ๐‘ฅ = ๐‘ฆ ⇔ ๐‘ฆ = ๐‘ฅ.
โˆ™ [E3] ๐‘ฅ = ๐‘ฆ, ๐‘ฆ = ๐‘ง ⇒ ๐‘ฅ = ๐‘ง.
Let ๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 and ๐‘ฆ0 , ๐‘ฆ1 , … , ๐‘ฆ๐‘›−1 be variable symbols from ๐–ฒ.
โˆ™ [E4] For any ๐‘›-ary function symbol ๐‘“ of ๐–ฒ,
๐‘ฅ0 = ๐‘ฆ0 ∧ ๐‘ฅ1 = ๐‘ฆ1 ∧ · · · ∧ ๐‘ฅ๐‘›−1 = ๐‘ฆ๐‘›−1
⇒ ๐‘“ (๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 ) = ๐‘“ (๐‘ฆ0 , ๐‘ฆ1 , … , ๐‘ฆ๐‘›−1 ).
โˆ™ [E5] For any ๐–ฒ-formula ๐‘(๐‘ข0 , ๐‘ข1 , … , ๐‘ข๐‘›−1 ),
๐‘ฅ0 = ๐‘ฆ0 ∧ ๐‘ฅ1 = ๐‘ฆ1 ∧ · · · ∧ ๐‘ฅ๐‘›−1 = ๐‘ฆ๐‘›−1
⇒ ๐‘(๐‘ฅ0 , ๐‘ฅ1 , … , ๐‘ฅ๐‘›−1 ) ↔ ๐‘(๐‘ฆ0 , ๐‘ฆ1 , … , ๐‘ฆ๐‘›−1 ).
Axioms E1, E2, and E3 give = the behavior of an equivalence relation (Definition 4.2.4).
For example, we can use E2 to prove that for any constant symbols ๐‘0 and ๐‘1 ,
โŠข ๐‘0 = ๐‘1 ↔ ๐‘1 = ๐‘0 .
Section 5.1 AXIOMS
The proof goes as follows:
227
]
[
๐‘ ๐‘1
๐‘0 = ๐‘1 ⇔ (๐‘ฅ0 = ๐‘ฅ1 ) 0
๐‘ฅ0 ๐‘ฅ1
[
]
๐‘ ๐‘1
⇔ (๐‘ฅ1 = ๐‘ฅ0 ) 0
๐‘ฅ0 ๐‘ฅ1
⇔ ๐‘1 = ๐‘0 .
Axiom E4 allows a function symbol to be used in a proof as a function, and E5 allows
equal terms to be substituted into formulas with the result being equivalent formulas.
For example, given the ๐–ญ๐–ณ-term ๐‘ข + ๐‘ฃ, by E4,
๐‘ฅ0 = ๐‘ฆ0 ∧ ๐‘ฅ1 = ๐‘ฆ1 ⇒ ๐‘ฅ0 + ๐‘ฅ1 = ๐‘ฆ0 + ๐‘ฆ1 ,
and given the ๐–ญ๐–ณ-formula ๐‘ข + ๐‘ฃ = ๐‘ฃ + ๐‘ข, by E5,
๐‘ฅ0 = ๐‘ฆ0 ∧ ๐‘ฅ1 = ๐‘ฆ1 ⇒ ๐‘ฅ0 + ๐‘ฅ1 = ๐‘ฅ1 + ๐‘ฅ0 ↔ ๐‘ฆ0 + ๐‘ฆ1 = ๐‘ฆ1 + ๐‘ฆ0 .
Formal proofs that require a deduction on an equality need to reference one of the
equality axioms from Axioms 5.1.1.
Existence and Uniqueness Axioms
The axioms will be ๐–ฒ๐–ณ-formulas (Example 2.1.3), where all terms represent sets. We
begin by assuming the existence of two sets.
AXIOM 5.1.2 [Empty Set]
∃๐‘ฅ∀๐‘ฆ(¬ ๐‘ฆ ∈ ๐‘ฅ).
In ๐–ฒ๐–ณ-formulas, a witness for the empty set axiom is denoted by { }, although it is
usually written as ∅.
AXIOM 5.1.3 [Infinity]
∃๐‘ฅ({ } ∈ ๐‘ฅ ∧ ∀๐‘ข[๐‘ข ∈ ๐‘ฅ → ∃๐‘ฆ(๐‘ฆ ∈ ๐‘ฅ ∧ ๐‘ข ∈ ๐‘ฆ ∧ ∀๐‘ฃ[๐‘ฃ ∈ ๐‘ข → ๐‘ฃ ∈ ๐‘ฆ])]).
The standard interpretation of the infinity axiom is that there exists a set that contains
∅ and ๐‘Ž ∪ {๐‘Ž} is an element of the set if ๐‘Ž is an element of the set.
In Chapter 3, we noted that the elements of a set determine the set. For example,
{1, 2} = {1, 2, 2}. This principle is codified by the next axiom.
AXIOM 5.1.4 [Extensionality]
∀๐‘ฅ∀๐‘ฆ(∀๐‘ข[๐‘ข ∈ ๐‘ฅ ↔ ๐‘ข ∈ ๐‘ฆ] → ๐‘ฅ = ๐‘ฆ).
Suppose ๐ด1 and ๐ด2 are witnesses to the empty set axiom (5.1.2). Since ๐‘ฅ ∉ ๐ด1 and
๐‘ฅ ∉ ๐ด2 for every ๐‘ฅ, we conclude that
๐‘ฅ ∈ ๐ด1 ↔ ๐‘ฅ ∈ ๐ด2 .
228
Chapter 5 AXIOMATIC SET THEORY
Therefore, ๐ด1 = ๐ด2 by extensionality (Axiom 5.1.4), which means that ∅ is the witness
to the empty set axiom. This uniqueness result does not appear to extend to the infinity
axiom because both
{∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}, … }
and
{∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}},
… , {∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}}, … }
are witnesses, provided that they are sets.
Construction Axioms
Now to build some sets. The next four axioms allow us to do this.
AXIOM 5.1.5 [Pairing]
∀๐‘ข ∀๐‘ฃ ∃๐‘ฅ ∀๐‘ค (๐‘ค ∈ ๐‘ฅ ↔ ๐‘ค = ๐‘ข ∨ ๐‘ค = ๐‘ฃ).
Suppose that ๐‘€ and ๐‘ are sets. Since {๐‘€, ๐‘} is a witness of
∃๐‘ฅ ∀๐‘ค (๐‘ค ∈ ๐‘ฅ ↔ ๐‘ค = ๐‘€ ∨ ๐‘ค = ๐‘),
{๐‘€, ๐‘} is a set by the pairing axiom, and from this, we conclude that {๐‘€, {๐‘€, ๐‘}},
{๐‘, {๐‘€, ๐‘}}, and {{๐‘€, ๐‘}} = {{๐‘€, ๐‘}, {๐‘€, ๐‘}} are sets. Because {๐‘€, ๐‘€}
equals {๐‘€}, pairing along with extensionality prove the existence of singletons. For
example, if ๐‘Š is a witness to the Infinity Axiom, {∅, ๐‘Š }, {∅}, and {∅, {∅, ๐‘Š }} are
sets.
AXIOM 5.1.6 [Union]
∀๐‘ฅ∃๐‘ฆ∀๐‘ข(๐‘ข ∈ ๐‘ฆ ↔ ∃๐‘ฃ[๐‘ฃ ∈ ๐‘ฅ ∧ ๐‘ข ∈ ๐‘ฃ]).
โ‹ƒ
โ‹ƒ
By the union axiom, ๐‘€ is a set, and since ๐‘€ ∪ ๐‘ = {๐‘€, ๐‘}, we conclude
that ๐‘€ ∪ ๐‘ is a set. Furthermore, the empty set, union, and pairing axioms can be used
to prove that for any ๐‘› ∈ ๐, there exists a set of the form {๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘› } (Exercise 3).
AXIOM 5.1.7 [Power Set]
∀๐‘ฅ∃๐‘ฆ∀๐‘ข(๐‘ข ∈ ๐‘ฆ ↔ ∀๐‘ฃ[๐‘ฃ ∈ ๐‘ข → ๐‘ฃ ∈ ๐‘ฅ]).
Because of Definition 3.3.1, the power set axiom can be written as
∀๐‘ฅ∃๐‘ฆ∀๐‘ข(๐‘ข ∈ ๐‘ฆ ↔ ๐‘ข ⊆ ๐‘ฅ).
We conclude that for every set ๐‘€, P(๐‘€) is a set by the power set axiom, and by extensionality, P(๐‘€) is the unique set of subsets of ๐‘€.
The next axiom is actually what is called an axiom scheme, infinitely many axioms,
one for every formula. They are sometimes called the separation axioms.
Section 5.1 AXIOMS
229
AXIOMS 5.1.8 [Subset]
For every ๐–ฒ๐–ณ-formula ๐‘(๐‘ข) not containing the symbol ๐‘ฆ, the following is an axiom:
∀๐‘ฅ∃๐‘ฆ∀๐‘ข [๐‘ข ∈ ๐‘ฆ ↔ ๐‘ข ∈ ๐‘ฅ ∧ ๐‘(๐‘ข)] .
The formula ๐‘(๐‘ข) in the subset axioms cannot contain the symbol ๐‘ฆ because the axioms
yield the existence of this set. If ๐‘ฆ was among its symbols, the existence of ๐‘ฆ would
depend on ๐‘ฆ.
The subset axioms yield many familiar sets.
โˆ™ Let โ„ฑ be a set. By a subset axiom, there exists a set ๐ถ such that
โ‹ƒ
๐‘ฅ∈๐ถ↔๐‘ฅ∈
โ„ฑ ∧ ∀๐‘(๐‘ ∈ โ„ฑ → ๐‘ฅ ∈ ๐‘).
Observe that the symbol ๐ถ does not appear in the formula
∀๐‘(๐‘ ∈ ๐ด → ๐‘ฅ ∈ ๐‘).
โ‹‚
Also, observe that
โ‹‚the set ๐ถ is the intersection of โ„ฑ . Hence, โ„ฑ is a set, and
since ๐‘€ ∩ ๐‘ = {๐‘€, ๐‘}, we conclude that ๐‘€ ∩ ๐‘ is a set.
โˆ™ By a subset axiom, there exists a set ๐ท such that
๐‘ฅ ∈ ๐ท ↔ ๐‘ฅ ∈ ๐‘€ ∧ ๐‘ฅ ∉ ๐‘,
so ๐‘€ โงต ๐‘ is a set.
Replacement Axioms
Given sets ๐ด and ๐ต, the function ๐œ‘ : ๐ด → ๐ต is a set because ๐œ‘ ⊆ ๐ด × ๐ต and ๐ด × ๐ต
is a set. Suppose, instead, that the function is defined using a formula ๐‘(๐‘ฅ, ๐‘ฆ) and that
its domain is given by a set ๐ด. It cannot be concluded from Axioms 5.1.4–5.1.10 that
the range {๐‘ฆ : ๐‘ฅ ∈ ๐ด ∧ ๐‘(๐‘ฅ, ๐‘ฆ)} is a set. However, it appears reasonable that it is. For
example, define
๐‘(๐‘ฅ, ๐‘ฆ) := ๐‘ฆ ∈ ๐™ ∧ ๐‘ฆ ≤ ๐‘ฅ ∧ ∀๐‘ง(๐‘ง ∈ ๐™ ∧ ๐‘ฆ ≤ ๐‘ง → ๐‘ฅ < ๐‘ง).
An examination of the formula shows ๐‘(3.4, 4) and ๐‘(−7.1, −8). Since
๐‘(๐‘ฅ, ๐‘ฆ1 ) ∧ ๐‘(๐‘ฅ, ๐‘ฆ2 ) → ๐‘ฆ1 = ๐‘ฆ2 ,
the formula ๐‘(๐‘ฅ, ๐‘ฆ) defines a function (Definition 4.4.10). If the domain is given to be
[0, ∞), its range is the set [0, ∞) ∩ ๐™. Generalizing to an arbitrary ๐‘(๐‘ฅ, ๐‘ฆ), it is expected
that the range would be a set if the domain is a set. The next axiom scheme guarantees
this. It was first found in correspondence between Cantor and Richard Dedekind (Cantor 1932) and Dmitry Mirimanoff (1917) with formal versions by Abraham Fraenkel
(1922) and Thoralf Skolem (1922).
230
Chapter 5 AXIOMATIC SET THEORY
AXIOMS 5.1.9 [Replacement]
For every ๐–ฒ๐–ณ-formula ๐‘(๐‘ก, ๐‘ค) not containing the symbol ๐‘ฆ, the following is an
axiom:
∀๐‘ฅ[∀๐‘ข∀๐‘ฃ1 ∀๐‘ฃ2 (๐‘ข ∈ ๐‘ฅ ∧ ๐‘(๐‘ข, ๐‘ฃ1 ) ∧ ๐‘(๐‘ข, ๐‘ฃ2 ) → ๐‘ฃ1 = ๐‘ฃ2 )
→ ∃๐‘ฆ∀๐‘ค(๐‘ค ∈ ๐‘ฆ ↔ ∃๐‘ก[๐‘ก ∈ ๐‘ฅ ∧ ๐‘(๐‘ก, ๐‘ค)])].
As an example, every indexed family of sets is a set. To prove this, let ๐ผ be a set and
๐ด๐‘– be a set for all ๐‘– ∈ ๐ผ. Define
๐‘(๐‘–, ๐‘ฆ) := ๐‘ฆ = ๐ด๐‘– .
Observe that by E2 and E3 (Axioms 5.1.1),
๐‘ฆ1 = ๐ด๐‘– ∧ ๐‘ฆ2 = ๐ด๐‘– ⇒ ๐‘ฆ1 = ๐‘ฆ2 .
Therefore, by a replacement axiom (5.1.9), there exists a set โ„ฑ such that
๐‘ค ∈ โ„ฑ ↔ ∃๐‘–(๐‘– ∈ ๐ผ ∧ ๐‘ค = ๐ด๐‘– ),
so โ„ฑ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} is a set, which implies that the union and intersection of families
of sets are sets.
Suppose ๐‘Ž ∈ ๐‘€ and ๐‘ ∈ ๐‘. By the subset and pairing axioms, we conclude that
{๐‘Ž, ๐‘}, (๐‘Ž, ๐‘) = {{๐‘Ž}, {๐‘Ž, ๐‘}} (Definition 3.2.8), and {(๐‘Ž, ๐‘)} are sets . Therefore, fixing
๐‘,
{{(๐‘Ž, ๐‘)} : ๐‘Ž ∈ ๐‘€}
is a family of sets, which implies that it is a set. Likewise,
{{{(๐‘Ž, ๐‘)} : ๐‘Ž ∈ ๐‘€} : ๐‘ ∈ ๐‘}
is a set. Hence, by the union axiom (5.1.6),
โ‹ƒโ‹ƒ
๐‘€ ×๐‘ =
{{{(๐‘Ž, ๐‘)} : ๐‘Ž ∈ ๐‘€} : ๐‘ ∈ ๐‘}
is a set. This implies, using a subset axiom (5.1.8), that any binary relation ๐‘… such that
dom(๐‘…) ⊆ ๐‘€ and ran(๐‘…) ⊆ ๐‘ is a set.
Axiom of Choice
Suppose that we are given the pairwise disjoint family of sets
โ„ฑ = {{1, 3, 5}, {2, 9, 11}, {7, 8, 13}}.
It is easy to find a set ๐‘† such that
๐‘† ∩ ๐ด is a singleton for every ๐ด ∈ โ„ฑ .
(5.1)
Section 5.1 AXIOMS
231
Simply run through the elements of โ„ฑ and choose an element from each set and put it
in ๐‘†. Since โ„ฑ is pairwise disjoint, each choice will differ from the others. For example,
it might be that
๐‘† = {1, 9, 13}.
However, what happens if โ„ฑ is an infinite set? If there was not a systematic way where
elements could be chosen from the sets of โ„ฑ , we would be left with making infinitely
many choices, which is something that we cannot do. Nonetheless, it appears reasonable that there is a set ๐‘† that intersects each member of โ„ฑ exactly once. Such a set
cannot be proved to exist from Axioms 5.1.2 to 5.1.9, so we need another axiom. It is
called the axiom of choice. We will have to use it every time that an infinite number
of arbitrary choices need to be made.
AXIOM 5.1.10 [Choice]
If โ„ฑ is a family of pairwise disjoint, nonempty sets, there exists ๐‘† ⊆
that ๐‘† ∩ ๐ด is a singleton for all ๐ด ∈ โ„ฑ .
โ‹ƒ
โ„ฑ such
The statement of the axiom of choice can be written as an ๐–ฒ๐–ณ-formula (Exercise 12).
Also, notice that ๐‘† in Axiom 5.1.10 is a function (Exercise 13). It is called a selector.
The next follows quickly from the axiom of choice. In fact, the proposition is equivalent to the axiom (Exercise 15), so this corollary is often used as a replacement for
it.
COROLLARY 5.1.11
For every binary relation ๐‘…, there exists a function ๐œ‘ such that ๐œ‘ ⊆ ๐‘… and
dom(๐œ‘) = dom(๐‘…).
PROOF
Let ๐‘… ⊆ ๐ด×๐ต. Define โ„ฑ = {{๐‘Ž}×[๐‘Ž]๐‘… : ๐‘Ž ∈ ๐ด}, which is a set by a replacement
axiom (Exercise 14). Since โ„ฑ is pairwise disjoint and the set {๐‘Ž} × [๐‘Ž]๐‘… ≠ ∅ for
all ๐‘Ž ∈ dom(๐‘…), the axiom of choice (5.1.10) implies that there exists a selector
โ‹ƒ
๐‘† such that ๐‘† ⊆ โ„ฑ and ๐‘† ∩ ({๐‘Ž} × [๐‘Ž]๐‘… ) is a singleton for all ๐‘Ž ∈ dom(๐‘…).
Thus, ๐‘† ⊆ ๐‘…, and for all ๐‘Ž ∈ dom(๐‘…), there exists a unique ๐‘ ∈ ๐ต such that
๐‘Ž ๐‘… ๐‘. This implies that ๐‘† is the desired function.
โ‹ƒ
Given a family of sets โ„ฑ , define a relation ๐‘… ⊆ โ„ฑ × โ„ฑ by
๐‘… = {(๐ด, ๐‘Ž) : ๐ด ∈ โ„ฑ ∧ ๐‘Ž ∈ ๐ด}.
โ‹ƒ
By Corollary 5.1.11, there exists a function ๐œ‰ : โ„ฑ → โ„ฑ such that ๐œ‰(๐ด) ∈ ๐ด for all
๐ด ∈ โ„ฑ . The function ๐œ‰ is called a choice function.
EXAMPLE 5.1.12
Let ๐’œ = {๐ด๐‘– : ๐‘– ∈ ๐ผ} be a family of nonempty sets. We want to define a family
of singletons โ„ฌ such that for all ๐‘– ∈ ๐ผ,
if {๐‘Ž๐‘– } ∈ โ„ฌ, then ๐‘Ž๐‘– ∈ ๐ด๐‘– .
232
Chapter 5 AXIOMATIC SET THEORY
โ‹ƒ
By Corollary 5.1.11, there exists a choice function ๐œ‰ : ๐’œ → ๐’œ . The family
โ„ฌ = {{๐œ‰(๐ด๐‘– )} : ๐‘– ∈ ๐ผ} is the desired set because ๐œ‰(๐ด๐‘– ) ∈ ๐ด๐‘– for all ๐‘– ∈ ๐ผ.
There are many theorems equivalent to the axiom of choice (5.1.10). One such result
involves families of sets. Take ๐‘› ∈ ๐ and define
๐’œ๐‘› = {{0}, {0, 1}, … , {0, 1, … , ๐‘›}}.
Observe that ๐’œ๐‘› is a chain with respect to ⊆ and contains a maximal element (Definition 4.3.16), which is {0, 1, … , ๐‘›}. However, the chain
๐’œ = {{0}, {0, 1}, … , {0, 1, … , ๐‘›}, … }
has no maximal element. There are many sets that can be added to ๐’œ to give it a
maximal element, but the natural choice is to add the union of ๐’œ to the family giving
๐’œ ′ = {{0}, {0, 1}, … , {0, 1, … , ๐‘›}, … , ๐}.
๐’œ ′ has a maximal element, namely, ๐. The generalization of this result to any family
of sets was first proved by Kuratowski (1922) and then independently by Zorn (1935)
for whom the theorem is named despite Kuratowski’s priority. The proof given is essentially due to Zermelo (Halmos 1960).
THEOREM 5.1.13 [Zorn’s Lemma]
โ‹ƒ
Let ๐’œ be a family of sets. If ๐’ž ∈ ๐’œ for every chain ๐’ž of ๐’œ with respect to
⊆, there exists ๐‘€ ∈ ๐’œ such that ๐‘€ ⊄ ๐ด for all ๐ด ∈ ๐’œ .
PROOF
Let ๐œ‰ : P(๐’œ ) โงต {∅} → ๐’œ be a choice function (Corollary 5.1.11). For every
chain ๐’ž of ๐’œ , define
๐’žฬ„ = {๐ด ∈ ๐’œ : ๐’ž ∪ {๐ด} is a chain}.
Notice that ๐’žฬ„ is the set of elements of ๐’œ that when added to ๐’ž yields a chain.
Let ๐–ข๐—(๐’œ ) be defined as the set of all chains of ๐’œ . That both ๐’žฬ„ and ๐–ข๐—(๐’œ ) are
sets is left to Exercise 17(a). Define
X : ๐–ข๐—(๐’œ ) → ๐–ข๐—(๐’œ )
by
{
๐’ž ∪ {๐œ‰(๐’žฬ„ โงต ๐’ž )} if ๐’žฬ„ โงต ๐’ž ≠ ∅,
X(๐’ž ) =
๐’ž
if ๐’žฬ„ โงต ๐’ž = ∅.
Next, suppose that ๐’žโ‹ƒ0 is a chain of ๐’œ such thatโ‹ƒX(๐’ž0 ) = ๐’ž0 . By the assumption
on ๐’œ , we have that ๐’ž
โ‹ƒ0 ∈ ๐’œ . To prove that ๐’ž0 is a maximal element of ๐’œ ,
take ๐ด ∈ ๐’œ such that ๐’ž0 ⊂ ๐ด. Since ๐ด has an element that is not in any of
the elements of ๐’ž0 , the union ๐’ž0 ∪ {๐ด} is a chain properly containing ๐’ž0 , which
is impossible. We conclude that the theorem is proved if ๐’ž0 is shown to exist.
Section 5.1 AXIOMS
233
To accomplish this, we begin with a definition. A subset ๐’ฏ of ๐–ข๐—(๐’œ ) is called a
tower when
โˆ™ ∅∈๐’ฏ,
โˆ™ X(๐’ž ) ∈ ๐’ฏ for all ๐’ž ∈ ๐’ฏ ,
โ‹ƒ
โˆ™
๐’Ÿ ∈ ๐’ฏ for every chain ๐’Ÿ ⊆ ๐’ฏ .
โ‹‚
Define ๐–ณ๐—ˆ(๐’œ ) to be the set of all towers of ๐’œ . Observe that ๐–ข๐—(๐’ž ) and ๐–ณ๐—ˆ(๐’œ )
are both towers [Exercise 17(b)].
โ‹‚
Take ๐ถ ∈
๐–ณ๐—ˆ(๐’œ
โ‹‚ ) such that ๐ถ is comparable with respect to inclusion to
every element of ๐–ณ๐—ˆ(๐’œ ). Such a set
element of ๐–ณ๐—ˆ(๐’œ )
โ‹‚ ๐ถ exists since ∅ is anโ‹‚
and comparable to โ‹‚
every element of ๐–ณ๐—ˆ(๐’œ ). Suppose
๐ด
∈
๐–ณ๐—ˆ(๐’œ ) such that
โ‹‚
๐ด ⊂ ๐ถ. Because ๐–ณ๐—ˆ(๐’œ ) is a tower, X(๐ด) ∈
๐–ณ๐—ˆ(๐’œ ). If ๐ถ ⊂ X(๐ด), then
X(๐ด) โงต ๐ด has at least two elements, which is impossible. Therefore,
โ‹‚
(5.2)
if ๐ด ∈
๐–ณ๐—ˆ(๐’œ ) and ๐ด ⊂ ๐ถ, then X(๐ด) ⊆ ๐ถ.
Define
๐‘‡ = {๐ด ∈
โ‹‚
๐–ณ๐—ˆ(๐’œ ) : ๐ด ⊆ ๐ถ ∨ X(๐ถ) ⊆ ๐ด}.
If ๐ด ⊆ ๐ถ, then ๐ด ⊆ X(๐ถ) since ๐ถ ⊆ X(๐ถ). Hence, every element of ๐‘‡ is
comparable to X(๐ถ). Also, ๐‘‡ is a tower because of the following:
โ‹‚
โˆ™ ∅ ∈ ๐‘‡ because ∅ ∈ ๐–ณ๐—ˆ(๐’œ ) and ∅ ⊆ ๐ถ.
โ‹‚
โ‹‚
โˆ™ Let ๐ต ∈ ๐‘‡ . Since ๐–ณ๐—ˆ(๐’œ ) is a tower, X(๐ต) ∈
๐–ณ๐—ˆ(๐’œ ). If ๐ต ⊂ ๐ถ,
then X(๐ต) ⊆ ๐ถ by (5.2). If ๐ถ ⊆ ๐ต, then X(๐ถ) ⊆ X(๐ต). In both cases,
X(๐ต) ∈ ๐‘‡ .
โ‹ƒ
โ‹‚
โˆ™ Let ๐’ž ⊆ ๐‘‡ be a chain. Then, ๐’ž ∈
๐–ณ๐—ˆ(๐’œ ). Suppose there exists
๐ถ0 ∈ ๐’ž such that ๐ถ0 is not a subset of ๐ถ. This implies that X(๐ถ) ⊆ ๐ถ0 .
โ‹ƒ
โ‹ƒ
Thus, X(๐ถ) ⊆ ๐’ž , so ๐’ž ∈ ๐‘‡ .
โ‹‚
โ‹‚
โ‹‚
Hence, ๐‘‡ = ๐–ณ๐—ˆ(๐’œ ) because ๐‘‡ ⊆ ๐–ณ๐—ˆ(๐’œ ).โ‹‚ Therefore, ๐–ณ๐—ˆ(๐’œ ) is a chain
because ๐ถ is arbitrary [Exercise 17(c)]. Since ๐–ณ๐—ˆ(๐’œ ) is a tower,
โ‹ƒโ‹‚
โ‹‚
๐–ณ๐—ˆ(๐’œ ) ∈
๐–ณ๐—ˆ(๐’œ )
and then
) โ‹‚
๐–ณ๐—ˆ(๐’œ ) ∈
๐–ณ๐—ˆ(๐’œ ).
โ‹ƒโ‹‚
โ‹‚
Hence, since
๐–ณ๐—ˆ(๐’œ ) is an upper bound of ๐–ณ๐—ˆ(๐’œ ),
(โ‹ƒ โ‹‚
) โ‹ƒโ‹‚
X
๐–ณ๐—ˆ(๐’œ ) ⊆
๐–ณ๐—ˆ(๐’œ ),
X
which yields
X
(โ‹ƒ โ‹‚
(โ‹ƒ โ‹‚
) โ‹ƒโ‹‚
๐–ณ๐—ˆ(๐’œ ) =
๐–ณ๐—ˆ(๐’œ ).
There are many equivalents to the axiom of choice (Rubin and Rubin 1985, Jech
1973). One of them is Zorn’s lemma.
234
Chapter 5 AXIOMATIC SET THEORY
THEOREM 5.1.14
The axiom of choice is equivalent to Zorn’s lemma.
PROOF
Since we have already proved that the axiom of choice (5.1.10) implies Zorn’s
lemma (Theorem 5.1.13), we only need to prove the converse. We must be careful
to only use Axioms 5.1.2–5.1.9 and not Axiom 5.1.10.
Assume Zorn’s lemma and let ๐‘… be a relation. Define
๐’œ = {๐œ‘ : ๐œ‘ ⊆ ๐‘… ∧ ๐œ‘ is a function}.
Let ๐’ž be a chain of elements of ๐’œ .
โ‹ƒ
โˆ™ Take ๐œ“ ∈ ๐’ž , so ๐œ“ ∈ ๐ถ for some ๐ถ โ‹ƒ
∈ ๐’ž . This implies that ๐ถ is a subset
of ๐‘…. Therefore, ๐œ“ ∈ ๐‘…, proving that ๐’ž ⊆ ๐‘….
โˆ™
โ‹ƒ
๐’ž is a function by Exercise 4.4.13.
โ‹ƒ
We conclude that ๐’ž ∈ ๐’œ . Thus, by Zorn’s lemma, there exists a maximal
element Φ ∈ ๐’œ . This means that Φ ⊆ ๐‘… and dom(Φ) ⊆ dom(๐‘…). To prove that
dom(๐‘…) ⊆ dom(Φ), let (๐‘ฅ, ๐‘ฆ) ∈ ๐‘… and suppose that ๐‘ฅ ∉ dom(Φ). This implies
that Φ ∪ {(๐‘ฅ, ๐‘ฆ)} ⊆ ๐‘… is a function. Hence, Φ ∪ {(๐‘ฅ, ๐‘ฆ)} ∈ ๐’œ , contradicting the
maximality of Φ. We conclude that Φ is the desired function, and the axiom of
choice follows (Exercise 15).
There was a controversy regarding the axiom of choice when Zermelo first proposed
it. Despite mathematicians having previously used it implicitly, some objected to its
nonconstructive nature. The other axioms yield distinct results, but the axiom of choice
results in a set with elements that are not clearly identified. Over time, however, most
objections have faded. This is because the majority of mathematicians regard it as reasonable and generally those who question the axiom of choice realize that eliminating
it would lead to serious problems because many proofs in various fields of mathematics
rely on the axiom.
Axiom of Regularity
The ideas that led to the next axiom (also known as the axiom of foundation) can be
found in Mirimanoff (1917), while the statement of the axiom is credited to Skolem
(1922) and John von Neumann (1923).
AXIOM 5.1.15 [Regularity]
∀๐‘ฅ(๐‘ฅ ≠ { } → ∃๐‘ฆ[๐‘ฆ ∈ ๐‘ฅ ∧ ¬∃๐‘ข(๐‘ข ∈ ๐‘ฆ ∧ ๐‘ข ∈ ๐‘ฅ)]).
The main result of the regularity axiom is that it prevents sets from being elements of
themselves. Suppose there exists a set ๐ด such that ๐ด ∈ ๐ด. Then, ๐ด ∩ {๐ด} ≠ ∅, but the
regularity axiom implies that ๐ด ∩ {๐ด} should be empty.
Section 5.1 AXIOMS
235
THEOREM 5.1.16
No set is an element of itself.
If ๐‘‰ = {๐‘ฅ : ๐‘ฅ is a set} is a set, ๐‘‰ ∈ ๐‘‰ , contradicting Theorem 5.1.16. Thus, the
theorem’s corollary quickly follows.
COROLLARY 5.1.17
There is no set of all sets.
Because of the regularity axiom (5.1.15), ๐ด ∉ ๐ด for all sets ๐ด. Therefore, {๐‘ฅ : ๐‘ฅ ∉ ๐‘ฅ}
is not a set by, which prevents Russell’s paradox from being deduced from the axioms,
provided that they are consistent (Theorem 1.5.2).
The axiom of regularity is the final axiom of our chosen collection of axioms. It
is believed that they do not prove any contradictions, which implies that the axioms
prevent the construction of {๐‘ฅ : ๐‘ฅ ∉ ๐‘ฅ} as a set. Therefore, we write the following definition. The collections of axioms are named after the mathematician who was
primarily responsible for their selection (Zermelo 1908).
DEFINITION 5.1.18
โˆ™ Axioms 5.1.2–5.1.8 are the Zermelo axioms. This collection of sentences is
denoted by ๐—ญ.
โˆ™ Axioms 5.1.2–5.1.8 combined with replacement and regularity (Axioms 5.1.9
and 5.1.15) are the Zermelo–Fraenkel axioms, denoted by ๐—ญ๐—™.
โˆ™ The Zermelo–Fraenkel axioms with the axiom of choice is denoted by ๐—ญ๐—™๐—–.
The nonempty sets that follow from ๐—ญ๐—™๐—– have the property that all of their elements
are sets. Sets with this property are called hereditary or pure. Assuming ๐—ญ๐—™๐—– does
not prevent us from working with different types of sets, such as sets of symbols or
formulas, but we must remember that such nonhereditary sets are not products of ๐—ญ๐—™๐—–,
so they must be handled with care because we do not want to fall into a paradox.
Exercises
1. Let ๐–ฒ be a set of theory symbols. Let ๐‘1 , ๐‘2 , ๐‘3 , ๐‘4 ∈ ๐–ฒ be constant symbols and
๐‘“ ∈ ๐–ฒ be a binary function symbol. Suppose that ๐‘(๐‘ฅ, ๐‘ฆ) is an ๐–ฒ-formula. Use the
Equality Axioms (5.1.1) to prove the following.
(a) โŠข ๐‘1 = ๐‘1
(b) โŠข ๐‘1 = ๐‘2 ∧ ๐‘2 = ๐‘3 → ๐‘1 = ๐‘3
(c) โŠข ๐‘1 = ๐‘2 ∧ ๐‘3 = ๐‘4 → ๐‘“ (๐‘1 , ๐‘3 ) = ๐‘“ (๐‘2 , ๐‘4 )
(d) โŠข ๐‘1 = ๐‘2 ∧ ๐‘3 = ๐‘4 → [๐‘(๐‘“ (๐‘1 ), ๐‘3 ) ↔ ๐‘(๐‘“ (๐‘2 ), ๐‘4 )]
2. Prove ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ = ๐‘ฆ → ∀๐‘ข[๐‘ข ∈ ๐‘ฅ ↔ ๐‘ข ∈ ๐‘ฆ]).
3. For any ๐‘› ∈ ๐ with ๐‘› > 0, prove that there exists a set of the form {๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 }.
โ‹ƒโ‹‚
4. Let โ„ฑ be a family of sets. Prove that P(
โ„ฑ ) is a set.
236
Chapter 5 AXIOMATIC SET THEORY
5. Use a subset axiom (5.1.8) to prove that there is no set of all sets. This proves that
Russell’s paradox does not follow from the axiom of ๐—ญ.
6. Prove that there is no set that has every singleton as an element.
7. Let ๐ด and ๐ต be sets. Define the symmetric difference of ๐ด and ๐ต by
๐ด M ๐ต = ๐ด โงต ๐ต ∪ ๐ต โงต ๐ด.
Prove that ๐ด M ๐ต is a set.
8. Prove the given equations for all sets ๐ด, ๐ต, and ๐ถ.
(a) ๐ด M ∅ = ๐ด
(b) ๐ด M ๐‘ˆ = ๐ด
(c) ๐ด M ๐ต = ๐ต M ๐ด
(d) ๐ด M (๐ต M ๐ถ) = (๐ด M ๐ต) M ๐ถ
(e) ๐ด M ๐ต = (๐ด ∪ ๐ต) โงต (๐ต ∪ ๐ด)
(f) (๐ด M ๐ต) ∩ ๐ถ = (๐ด ∩ ๐ถ) M (๐ต ∩ ๐ถ)
โ‹ƒ
โ‹‚
9. Given sets ๐ผ and ๐ด๐‘– for all ๐‘– ∈ ๐ผ, show that ๐‘–∈๐ผ ๐ด๐‘– and ๐‘–∈๐ผ ๐ด๐‘– are sets.
10. Prove that ๐ด0 × ๐ด1 × · · · × ๐ด๐‘›−1 is a set if ๐ด0 , ๐ด1 , … , ๐ด๐‘›−1 are sets.
11. Show that the Cartesian product of a nonempty family of nonempty sets is not
empty.
12. Write the Axiom of Choice (5.1.10) as an ๐–ฒ๐–ณ-formula.
13. Demonstrate that a selector is a function.
14. In the proof of Corollary 5.1.11, prove that {{๐‘Ž} × [๐‘Ž]๐‘… : ๐‘Ž ∈ ๐ด} is a set.
15. Prove that Corollary 5.1.11 implies Axiom 5.1.10.
16. Let ๐‘… = ๐‘ × {0, 1}. Find a function ๐น : ๐‘ → {0, 1} such that ๐น (๐‘Ž) = 0 for all
๐‘Ž ∈ ๐‘.
17. Prove the following parts of the proof of Zorn’s lemma (5.1.13).
(a) ๐’žฬ„ and ๐–ข๐—(๐’œ ) are sets.
(b) ๐–ข๐—(๐’ž ) and ∩๐–ณ๐—ˆ(๐’œ ) are towers.
โ‹‚
(c)
๐–ณ๐—ˆ(๐’œ ) is a chain.
18. Prove that ๐—ญ๐—™ and Zorn’s lemma imply the axiom of choice.
19. Use Zorn’s lemma to prove that for every function ๐œ‘ : ๐ด → ๐ต, there exists a
maximal ๐ถ ⊆ ๐ด such that ๐œ‘ ๐ถ is one-to-one.
โ‹ƒ
20. Prove that if ๐’œ is a collection of sets such that ๐’ž ∈ ๐’œ for every chain ๐’ž ⊆ ๐’œ ,
then ๐’œ contains a maximal element.
21. Let (๐ด, ๐‘…) be a partial order. Prove that there exists an order ๐‘† on ๐ด such that
๐‘… ⊆ ๐‘† and ๐‘† is linear.
Section 5.2 NATURAL NUMBERS
237
22. Use the regularity axiom (5.1.15) to prove that if {๐‘Ž, {๐‘Ž, ๐‘}} = {๐‘, {๐‘, ๐‘‘}, then
๐‘Ž = ๐‘ and ๐‘ = ๐‘‘. This gives an alternative to Kuratowski’s definition of an ordered
pair (Definition 3.2.8).
23. Prove that there does not exist an infinite sequence of sets ๐ด0 , ๐ด1 , ๐ด2 , … such that
๐ด๐‘–+1 ∈ ๐ด๐‘– for all ๐‘– = 0, 1, 2, … .
24. Prove that the empty set axiom (5.1.2) can be proved from the other axioms of
๐—ญ๐—™๐—–.
25. Use Axiom 5.1.9 to prove that the subset axioms (5.1.8) can be proved from the
other axioms of ๐—ญ๐—™๐—–.
26. Let (๐ด, ๐‘…) be a poset. Prove that every chain of ๐ด is contained in a maximal chain
with respect to ⊆. This is called the Hausdorff maximal princple.
27. Prove that the Hausdorff maximal principle implies Zorn’s lemma, so is equivalent
to the axiom of choice.
5.2
NATURAL NUMBERS
In order to study mathematics itself, as opposed to studying the contents of mathematics as when we study calculus or Euclidean geometry, we need to first develop a system
in which all of mathematics can be interpreted. In such a system, we should be able to
precisely define mathematical concepts, like functions and relations; construct examples of them; and write statements about them using a very precise language. ๐—ญ๐—™๐—– with
first-order logic seems like a natural choice for such an endeavor. However, in order
to be a success, this system must have the ability to represent the most basic objects
of mathematical study. Namely, it must be able to model numbers. Therefore, within
๐—ญ๐—™๐—– families of sets that copy the properties of ๐, ๐™, ๐, ๐‘, and ๐‚ will be constructed.
Since we can do this, we conclude that ๐, ๐™, ๐, ๐‘, and ๐‚ are themselves sets, and we
also conclude that what we discover about their analogs are true about them. We begin
with the natural numbers.
DEFINITION 5.2.1
For every set ๐‘Ž, the successor of ๐‘Ž is the set ๐‘Ž+ = ๐‘Ž ∪ {๐‘Ž}. If ๐‘Ž is the successor
of ๐‘, then ๐‘ is the predecessor of ๐‘Ž and we write ๐‘ = ๐‘Ž− .
For example, we have that {3, 5, 7}+ = {3, 5, 7, {3, 5, 7}} and ∅+ = {∅}. For convenience, write ๐‘Ž++ for (๐‘Ž+ )+ , so ∅++ = {∅, {∅}}. Also, the predecessor of the set
{∅, {∅}, {∅, {∅}}} is {∅, {∅}}, so write {∅, {∅}, {∅, {∅}}}− = {∅, {∅}}. Furthermore, since ∅ contains no elements, we have the following.
THEOREM 5.2.2
∅ does not have a predecessor.
238
Chapter 5 AXIOMATIC SET THEORY
Although every set has a successor, we are primarily concerned with certain sets
that have a particular property.
DEFINITION 5.2.3
The set ๐ด is inductive if ∅ ∈ ๐ด and ๐‘Ž+ ∈ ๐ด for all ๐‘Ž ∈ ๐ด.
Definition 5.2.3 implies that if ๐ด is inductive, ๐ด contains the sets
∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}, …
because
∅+
{∅}+
{∅, {∅}}+
{∅, {∅}, {∅, {∅}}}+
= {∅},
= {∅, {∅}},
= {∅, {∅}, {∅, {∅}}},
= {∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}},
โ‹ฎ
Of course, Definition 5.2.3 does not guarantee the existence of an inductive set. The
infinity axiom (5.1.3) does that. Then, by a subset axiom (5.1.8),
∀๐‘ฅ∃๐‘ฆ∀๐‘ข(๐‘ข ∈ ๐‘ฆ ↔ ๐‘ข ∈ ๐‘ฅ ∧ ๐‘ฅ is inductive ∧ ∀๐‘ค[๐‘ค is inductive → ๐‘ข ∈ ๐‘ค]),
where w is inductive can be written as
∅ ∈ ๐‘ค ∧ ∀๐‘Ž(๐‘Ž ∈ ๐‘ค → ๐‘Ž+ ∈ ๐‘ค).
Therefore, the collection that contains the elements that are common to all inductive
sets is a set, and we can make the next definition (von Neumann 1923).
DEFINITION 5.2.4
An element that is a member of every inductive set is called a natural number.
Let ๐œ” denote the set of natural numbers. That is,
๐œ” = {∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}, … }.
Definition 5.2.4 suggests that the elements of ๐œ” will be interpreted to represent the
elements of ๐, so represent each natural number with the appropriate element of ๐:
0 = ∅,
1 = {∅},
2 = {∅, {∅}},
3 = {∅, {∅}, {∅, {∅}}},
4 = {∅, {∅}, {∅, {∅}}, {∅, {∅}, {∅, {∅}}}},
โ‹ฎ
Section 5.2 NATURAL NUMBERS
239
Although the choice of which sets should represent the numbers of ๐ is arbitrary, our
choice does have some fortunate properties. For example, the number of elements in
each natural number and the number of ๐ that it represents are the same. Moreover,
notice that each natural number can be written as
0 = { },
1 = {0},
2 = {0, 1},
3 = {0, 1, 2},
4 = {0, 1, 2, 3},
โ‹ฎ
and also note that
โ‹ƒ
โ‹ƒ
โ‹ƒ
โ‹ƒ
1 = 0,
2 = 1,
3 = 2,
4 = 3,
โ‹ฎ
The empty set is an element of ๐œ” by definition, so suppose ๐‘› ∈ ๐œ” and take ๐ด to be
an inductive set. Since ๐‘› is a natural number, ๐‘› ∈ ๐ด, and since ๐ด is inductive, ๐‘›+ ∈ ๐ด.
Because ๐ด is arbitrary, we conclude that ๐‘›+ is an element of every inductive set, so
๐‘›+ ∈ ๐œ” (Definition 5.2.4). This proves the next theorem.
THEOREM 5.2.5
๐œ” is inductive.
The proof of the next theorem is left to Exercise 5.
THEOREM 5.2.6
If ๐ด is an inductive set and ๐ด ⊆ ๐œ”, then ๐ด = ๐œ”.
Order
Because we want to interpret ๐œ” so that it represents ๐, we should be able to order ๐œ” as
๐ is ordered by ≤. That is, we need to find a partial order on ๐œ” that is also a well-order.
To do this, we start with a definition.
DEFINITION 5.2.7
A set ๐ด is transitive means for all ๐‘Ž and ๐‘, if ๐‘ ∈ ๐‘Ž and ๐‘Ž ∈ ๐ด, then ๐‘ ∈ ๐ด.
Observe that ∅ is transitive because ๐‘Ž ∈ ∅ is always false. More transitive sets are
found in the next theorem.
240
Chapter 5 AXIOMATIC SET THEORY
THEOREM 5.2.8
โˆ™ Every natural number is transitive.
โˆ™ ๐œ” is transitive.
PROOF
The proof of the second part is left to Exercise 4. We prove the first by defining
๐ด = {๐‘› ∈ ๐œ” : ๐‘› is transitive}.
We have already noted that 0 ∈ ๐ด, so assume that ๐‘› ∈ ๐ด and let ๐‘ ∈ ๐‘Ž and
๐‘Ž ∈ ๐‘› ∪ {๐‘›}. If ๐‘Ž ∈ {๐‘›}, then ๐‘ ∈ ๐‘›. If ๐‘Ž ∈ ๐‘›, then by hypothesis, ๐‘ ∈ ๐‘›. In
either case, ๐‘ ∈ ๐‘›+ since ๐‘› ⊆ ๐‘›+ , so ๐‘›+ ∈ ๐ด. Hence, ๐ด is inductive, so ๐ด = ๐œ”
by Theorem 5.2.6.
In addition to being transitive, each element of ๐œ” has another useful property that
provides another reason why their choice was a wise one.
LEMMA 5.2.9
If ๐‘š, ๐‘› ∈ ๐œ”, then ๐‘š ⊂ ๐‘› if and only if ๐‘š ∈ ๐‘›.
PROOF
Take ๐‘š, ๐‘› ∈ ๐œ”. If ๐‘š ∈ ๐‘›, then ๐‘š ⊂ ๐‘› since ๐‘› is transitive (Exercise 3). To prove
the converse, define
๐ด = {๐‘˜ ∈ ๐œ” : ∀๐‘™(๐‘™ ∈ ๐œ” ∧ [๐‘™ ⊂ ๐‘˜ → ๐‘™ ∈ ๐‘˜])}.
We show that ๐ด is inductive. Trivially, 0 ∈ ๐ด. Now take ๐‘› ∈ ๐ด and let ๐‘š ∈ ๐œ”
such that ๐‘š ⊂ ๐‘›+ . We have two cases to consider.
โˆ™ Suppose ๐‘› ∉ ๐‘š. Then, ๐‘š ⊆ ๐‘›. If ๐‘š ⊂ ๐‘›, then ๐‘š ∈ ๐‘› by hypothesis, which
implies that ๐‘š ∈ ๐‘›+ . If ๐‘š = ๐‘›, then ๐‘š ∈ ๐‘›+ .
โˆ™ Next assume that ๐‘› ∈ ๐‘š. Take ๐‘ฅ ∈ ๐‘›. Since ๐‘š is transitive by Theorem 5.2.8, we have that ๐‘ฅ ∈ ๐‘š. This implies that ๐‘› ∪ {๐‘›} ⊆ ๐‘š, but this is
impossible since ๐‘š ⊂ ๐‘›+ , so ๐‘› ∉ ๐‘š.
For example, we see that 2 ∈ 3 and 2 ⊂ 3 because
{∅, {∅}} ∈ {∅, {∅}, {∅, {∅}}}
and
{∅, {∅}} ⊂ {∅, {∅}, {∅, {∅}}}.
Now use Lemma 5.2.9 to define an order on ๐œ”. Instead of using 4, we use ≤ to copy
the order on ๐.
Section 5.2 NATURAL NUMBERS
241
DEFINITION 5.2.10
For all ๐‘š, ๐‘› ∈ ๐œ”, let
๐‘š ≤ ๐‘› if and only if ๐‘š ⊆ ๐‘›.
Define ๐‘š < ๐‘› to mean ๐‘š ≤ ๐‘› but ๐‘š ≠ ๐‘›.
Lemma 5.2.9 in conjunction with Definition 5.2.10 implies that for all ๐‘š, ๐‘› ∈ ๐œ”, we
have that
๐‘š < ๐‘› if and only if ๐‘š ⊂ ๐‘› if and only if ๐‘š ∈ ๐‘›.
The order of Definition 5.2.10 makes ๐œ” a chain with ∅ as its least element.
THEOREM 5.2.11
(๐œ”, ≤) is a linear order.
PROOF
That (๐œ”, ≤) is a poset follows as in Example 4.3.9. To show that ๐œ” is a chain
under ≤, define
๐ด = {๐‘˜ ∈ ๐œ” : ∀๐‘™(๐‘™ ∈ ๐œ” ∧ [๐‘˜ ≤ ๐‘™ ∨ ๐‘™ ≤ ๐‘˜])}.
We prove that ๐ด is inductive.
โˆ™ Since ∅ is a subset of every set, 0 ∈ ๐ด.
โˆ™ Suppose that ๐‘› ∈ ๐ด and let ๐‘š ∈ ๐œ”. We have two cases to check. First,
assume that ๐‘› ≤ ๐‘š. If ๐‘› < ๐‘š, then ๐‘›+ ≤ ๐‘š, while if ๐‘› = ๐‘š, then ๐‘š < ๐‘›+ .
Now suppose that ๐‘š ≤ ๐‘›, but this implies that ๐‘š < ๐‘›+ . In either case,
๐‘›+ ∈ ๐ด.
We can now quickly prove the following.
COROLLARY 5.2.12
For all ๐‘š, ๐‘› ∈ ๐œ”, if ๐‘š+ = ๐‘›+ , then ๐‘š = ๐‘›.
PROOF
Let ๐‘š and ๐‘› be natural numbers and assume that
๐‘š ∪ {๐‘š} = ๐‘› ∪ {๐‘›}.
(5.3)
Take ๐‘ฅ ∈ ๐‘š. Then, ๐‘ฅ ∈ ๐‘› or ๐‘ฅ = ๐‘›. If ๐‘ฅ ∈ ๐‘›, then ๐‘š ⊆ ๐‘›, so suppose that
๐‘ฅ = ๐‘›. This implies that ๐‘› ∈ ๐‘š, so ๐‘š ≠ ๐‘› by Theorem 5.1.16. However, we then
have ๐‘š ∈ ๐‘› by (5.3), which contradicts the trichotomy law (Theorem 4.3.21).
Similarly, we can prove that ๐‘› ⊆ ๐‘š.
Since the successor defines a function ๐‘† : ๐œ” → ๐œ” where ๐‘†(๐‘›) = ๐‘›+ , Corollary 5.2.12
implies that ๐‘† is one-to-one.
Thereom 5.2.11 shows that (๐œ”, ≤) is a linear order as (๐, ≤) is a linear order. The
next theorem shows that the similarity goes further.
242
Chapter 5 AXIOMATIC SET THEORY
THEOREM 5.2.13
(๐œ”, ≤) is a well-ordered set.
PROOF
By Theorem 5.2.11, (๐œ”, ≤) is a linear order. To prove that it is well-ordered,
suppose that ๐ด ⊆ ๐œ” such that ๐ด does not have a least element. Define
๐ต = {๐‘˜ ∈ ๐œ” : {0, 1, … , ๐‘˜} ∩ ๐ด = ∅}.
We prove that ๐ต is inductive.
โˆ™ If 0 ∈ ๐ด, then ๐ด has a least element because 0 is the least element of ๐œ”.
Hence, {0} ∩ ๐ด = ∅, so 0 ∈ ๐ต.
โˆ™ Let ๐‘› ∈ ๐ต. This implies that {0, 1, … , ๐‘›} ∩ ๐ด is empty. Thus, ๐‘›+ cannot
be an element of ๐ด for then it would be the least element of ๐ด. Hence,
{0, 1, … , ๐‘›, ๐‘›+ } ∩ ๐ด = ∅ proving that ๐‘›+ ∈ ๐ต.
Therefore, ๐ต = ๐œ”, which implies that ๐ด is empty.
Recursion
The familiar factorial is defined recursively as
0! = 1,
(๐‘› + 1)! = (๐‘› + 1)๐‘›! (๐‘› ∈ ๐).
(5.4)
(5.5)
A recursive definition is one that is given in terms of itself. This is illustrated in
(5.5) where the factorial is defined using the factorial. It appears that the factorial is a
function ๐ → ๐, but a function is a set, so why do (5.4) and (5.5) define a set? Such a
definition is not found among the axioms of Section 5.1 or the methods of Section 3.1.
That they do define a function requires an important theorem.
THEOREM 5.2.14 [Recursion]
Let ๐ด be a set and ๐‘Ž ∈ ๐ด. If ๐‘” is a function ๐ด → ๐ด, there exists a unique function
๐‘“ : ๐œ” → ๐ด such that
โˆ™ ๐‘“ (0) = ๐‘Ž,
โˆ™ ๐‘“ (๐‘›+ ) = ๐‘”(๐‘“ (๐‘›)) for all ๐‘› ∈ ๐œ”.
PROOF
Let ๐‘” : ๐ด → ๐ด be a function. Define
โ„ฑ = {โ„Ž : โ„Ž ⊆ ๐œ” × ๐ด ∧ (0, ๐‘Ž) ∈ โ„Ž
∧ ∀๐‘›∀๐‘ฆ[(๐‘›, ๐‘ฆ) ∈ โ„Ž → (๐‘›+ , ๐‘”(๐‘ฆ)) ∈ โ„Ž])}.
(5.6)
Note that โ„ฑ is a set by a subset axiom (5.1.8) and โ„ฑ is nonempty because ๐œ”×๐ด ∈
โ„ฑ . Let
โ‹‚
๐‘“=
โ„ฑ.
Section 5.2 NATURAL NUMBERS
243
Observe that ๐‘“ ∈ โ„ฑ (Exercise 8). Define
๐ท = {๐‘› ∈ ๐œ” : ∃๐‘ง[(๐‘›, ๐‘ง) ∈ ๐‘“ ] ∧ ∀๐‘ฆ∀๐‘ฆ′ [(๐‘›, ๐‘ฆ) ∈ ๐‘“ ∧ (๐‘›, ๐‘ฆ′ ) ∈ ๐‘“ → ๐‘ฆ = ๐‘ฆ′ ]}.
We prove that ๐ท is inductive.
โˆ™ Since โ„ฑ ≠ ∅, we know that (0, ๐‘Ž) ∈ ๐‘“ by (5.6), so let (0, ๐‘) also be an
element of ๐‘“ . If we assume that ๐‘Ž ≠ ๐‘, then
โ‹‚ ๐‘“ โงต {(0, ๐‘)} ∈ โ„ฑ , which is
impossible because it implies that (0, ๐‘) ∉ โ„ฑ . Hence, 0 ∈ ๐ท.
โˆ™ Suppose that ๐‘› ∈ ๐ท. This means that (๐‘›, ๐‘ง) ∈ ๐‘“ for some ๐‘ง ∈ ๐ด. Thus,
(๐‘›+ , ๐‘”(๐‘ง)) ∈ ๐‘“ by (5.6). Assume that (๐‘›+ , ๐‘ฆ) ∈ ๐‘“ . If ๐‘ฆ ≠ ๐‘”(๐‘ง),โ‹‚then
๐‘“ โงต {(๐‘›+ , ๐‘ฆ)} ∈ โ„ฑ , which again leads to the contradictory (๐‘›+ , ๐‘ฆ) ∉ โ„ฑ .
Hence, ๐‘“ is a function ๐œ” → ๐ด (Theorem 5.2.6). We confirm that ๐‘“ has the
desired properties.
โ‹‚
โˆ™ ๐‘“ (0) = ๐‘Ž because (0, ๐‘Ž) ∈ โ„ฑ .
โˆ™ Take ๐‘› ∈ ๐œ” and write ๐‘ฆ = ๐‘“ (๐‘›). This implies that (๐‘›, ๐‘ฆ) ∈ ๐‘“ , so we have
that (๐‘›+ , ๐‘”(๐‘ฆ)) ∈ ๐‘“ . Therefore, ๐‘“ (๐‘›+ ) = ๐‘”(๐‘ฆ) = ๐‘”(๐‘“ (๐‘›)).
To prove that ๐‘“ is unique, let ๐‘“ ′ : ๐œ” → ๐ด be a function such that ๐‘“ ′ (0) = ๐‘Ž
and ๐‘“ ′ (๐‘›+ ) = ๐‘”(๐‘“ ′ (๐‘›)) for all ๐‘› ∈ ๐œ”. Let
๐ธ = {๐‘› ∈ ๐œ” : ๐‘“ (๐‘›) = ๐‘“ ′ (๐‘›)}.
The set ๐ธ is inductive because ๐‘“ (0) = ๐‘Ž = ๐‘“ ′ (0) and assuming ๐‘› ∈ ๐œ” we have
that ๐‘“ (๐‘›+ ) = ๐‘”(๐‘“ (๐‘›)) = ๐‘”(๐‘“ ′ (๐‘›)) = ๐‘“ ′ (๐‘›+ ).
The factorial function has domain ๐ but the recursion theorem (5.2.14) gives a function with domain ๐œ” and uses the successor of Definition 5.2.1. We need a connection
between ๐ and ๐œ”. We do this by defining two operations on ๐œ” that we designate by +
and ⋅ and then showing that the basic properties of ๐œ” under these two operations are
the same as the basic properties of ๐ under standard addition and multiplication.
Arithmetic
We begin with addition. Let ๐‘” : ๐œ” → ๐œ” be defined by ๐‘”(๐‘›) = ๐‘›+ . For every ๐‘š ∈ ๐œ”, by
Theorem 5.2.14, there exists a unique function ๐‘“๐‘š : ๐œ” → ๐œ” such that
๐‘“๐‘š (0) = ๐‘š
and for all ๐‘› ∈ ๐œ”,
Define
๐‘“๐‘š (๐‘›+ ) = ๐‘”(๐‘“๐‘š (๐‘›)) = [๐‘“๐‘š (๐‘›)]+ .
๐‘Ž = {((๐‘š, ๐‘›), ๐‘“๐‘š (๐‘›)) : ๐‘š, ๐‘› ∈ ๐œ”}.
Since ๐‘“๐‘š is a function for every ๐‘› ∈ ๐œ”, the set ๐‘Ž is a binary operation (Definition 4.4.26).
Observe that for all ๐‘š, ๐‘› ∈ ๐œ”,
244
Chapter 5 AXIOMATIC SET THEORY
โˆ™ ๐‘Ž(๐‘š, 0) = ๐‘š,
โˆ™ ๐‘Ž(๐‘š, ๐‘›+ ) = ๐‘Ž(๐‘š, ๐‘›)+ .
We know that for every ๐‘˜, ๐‘™ ∈ ๐, we have that ๐‘˜ + 0 = ๐‘˜ and to add ๐‘˜ + ๐‘™, one simply
adds 1 a total of ๐‘™ times to ๐‘˜. This is essentially what ๐‘Ž does to the natural numbers.
For example, for 1, 3, 4 ∈ ๐,
1 + 3 = ([(1 + 1) + 1] + 1) = 4
and for 1, 2, 3, 4 ∈ ๐œ” (page 238),
๐‘Ž(1, 3) = ๐‘Ž(1, 2+ )
= ๐‘Ž(1, 2)+
= ๐‘Ž(1, 1+ )+
= ๐‘Ž(1, 1)++
= ๐‘Ž(1, 0+ )++
= ๐‘Ž(1, 0)+++
= 1+++
= 2++
= 3+
= 4.
Therefore, we choose ๐‘Ž to be addition on ๐œ”. To define the addition, we only need to cite
the two properties given in Theorem 5.2.14. Since each ๐‘“๐‘š is unique, there is no other
function to which the definition could be referring. Therefore, we can define addition
recursively.
DEFINITION 5.2.15
For all ๐‘š, ๐‘› ∈ ๐œ”,
โˆ™ ๐‘š + 0 = ๐‘š,
โˆ™ ๐‘š + ๐‘›+ = (๐‘š + ๐‘›)+ .
Notice that ๐‘š+ = (๐‘š + 0)+ = ๐‘š + 0+ = ๐‘š + 1. Furthermore, using the notation from
page 238, we conclude that 1 + 3 = 4 because ๐‘Ž(1, 3) = 4.
The following lemma shows that the addition given in Definition 5.2.15 has a property similar to that of commutativity.
LEMMA 5.2.16
For all ๐‘š, ๐‘› ∈ ๐œ”,
โˆ™ 0 + ๐‘› = ๐‘›.
โˆ™ ๐‘š+ + ๐‘› = (๐‘š + ๐‘›)+ .
Section 5.2 NATURAL NUMBERS
245
PROOF
Let ๐‘š, ๐‘› ∈ ๐œ”. To prove that 0 + ๐‘› = ๐‘›, we show that
๐ด = {๐‘˜ ∈ ๐œ” : 0 + ๐‘˜ = ๐‘˜}
is inductive.
โˆ™ 0 ∈ ๐ด because 0 + 0 = 0.
โˆ™ Let ๐‘› ∈ ๐ด. Then, 0 + ๐‘›+ = (0 + ๐‘›)+ = ๐‘›+ , so ๐‘›+ ∈ ๐ด.
To prove that ๐‘š+ + ๐‘› = (๐‘š + ๐‘›)+ , we show that
๐ต = {๐‘˜ ∈ ๐œ” : ๐‘š+ + ๐‘˜ = (๐‘š + ๐‘˜)+ }
is inductive.
โˆ™ Again, 0 ∈ ๐ต because 0 + 0 = 0.
โˆ™ Suppose that ๐‘› ∈ ๐ต. We have that
๐‘š+ + ๐‘›+ = (๐‘š+ + ๐‘›)+ = (๐‘š + ๐‘›)++ = (๐‘š + ๐‘›+ )+ ,
where the first and third equality follow by Definition 5.2.15 and the second
follows because ๐‘› ∈ ๐ต. Thus, ๐‘›+ ∈ ๐ต.
Now to see that + behaves on ๐œ” as + behaves on ๐, we use Definition 4.4.31.
THEOREM 5.2.17
โˆ™ The binary operation + on ๐œ” is associative and commutative.
โˆ™ 0 is the additive identity for ๐œ”.
PROOF
0 is the additive identity by Definition 5.2.15 and Lemma 5.2.16, and that + is
associative on ๐œ” is Exercise 14. To show that + is commutative, let ๐‘š ∈ ๐œ” and
define
๐ด = {๐‘˜ ∈ ๐œ” : ๐‘š + ๐‘˜ = ๐‘˜ + ๐‘š}.
As has been our strategy, we show that ๐ด is inductive.
โˆ™ ๐‘š + 0 = ๐‘š by Definition 5.2.15, and 0 + ๐‘š = ๐‘š by Lemma 5.2.16, so
0 ∈ ๐ด.
โˆ™ Let ๐‘› ∈ ๐ด. Therefore, ๐‘š + ๐‘› = ๐‘› + ๐‘š, which implies that
๐‘š + ๐‘›+ = (๐‘š + ๐‘›)+ = (๐‘› + ๐‘š)+ = ๐‘›+ + ๐‘š.
Hence, ๐‘›+ ∈ ๐ด.
Multiplication on ๐ can be viewed as iterated addition. For example,
3 ⋅ 4 = 3 + 3 + 3 + 3,
so we define multiplication recursively along these lines. As with addition, the result
is a binary operation by Theorem 5.2.14.
246
Chapter 5 AXIOMATIC SET THEORY
DEFINITION 5.2.18
For all ๐‘š, ๐‘› ∈ ๐œ”,
โˆ™ ๐‘š ⋅ 0 = 0,
โˆ™ ๐‘š ⋅ ๐‘›+ = ๐‘š ⋅ ๐‘› + ๐‘š.
For example, 3 ⋅ 4 = 12 because
3⋅4=3⋅3+3
= (3 ⋅ 2 + 3) + 3
= ([3 ⋅ 1 + 3] + 3) + 3
= ([(3 ⋅ 0 + 3) + 3] + 3) + 3
= ([3 + 3] + 3) + 3
= 12,
where ([3 + 3] + 3) + 3 = 12 is left to Exercise 13.
The next result is analogous to Lemma 5.2.16. Its proof is left to Exercise 9.
LEMMA 5.2.19
For all ๐‘š, ๐‘› ∈ ๐œ”,
โˆ™ 0 ⋅ ๐‘š = 0,
โˆ™ ๐‘›+ ⋅ ๐‘š = ๐‘› ⋅ ๐‘š + ๐‘š.
We now prove that ⋅ on ๐œ” behaves as ⋅ on ๐ and that + and ⋅ on ๐œ” interact with each
other via the distributive law as the two operations on ๐. For the proof, we introduce
two common conventions for these two operations.
โˆ™ So that we lessen the use of parentheses, define ⋅ to have precedence over +, and
read from left to right. That is,
๐‘š ⋅ ๐‘› + ๐‘œ = (๐‘š ⋅ ๐‘›) + ๐‘œ and ๐‘š + ๐‘› ⋅ ๐‘œ = ๐‘š + (๐‘› ⋅ ๐‘œ),
and
๐‘š + ๐‘› + ๐‘œ = (๐‘š + ๐‘›) + ๐‘œ and ๐‘š ⋅ ๐‘› ⋅ ๐‘œ = (๐‘š ⋅ ๐‘›) ⋅ ๐‘œ.
โˆ™ Define ๐‘š๐‘› = ๐‘š ⋅ ๐‘›.
THEOREM 5.2.20
โˆ™ The binary operation ⋅ on ๐œ” is associative and commutative.
โˆ™ 1 is the multiplicative identity for ๐œ”.
โˆ™ The distributive law holds for ๐œ”. This means that for all ๐‘š, ๐‘›, ๐‘œ ∈ ๐œ”,
๐‘š(๐‘› + ๐‘œ) = ๐‘š๐‘› + ๐‘š๐‘œ.
Section 5.2 NATURAL NUMBERS
247
PROOF
That the associative and commutative properties hold is Exercise 12. To prove
the other parts of the theorem, let ๐‘š, ๐‘›, ๐‘œ ∈ ๐œ”. Since 0+ = 1, by Definition 5.2.18
and Lemma 5.2.16,
๐‘š1 = ๐‘š0+ = ๐‘š0 + ๐‘š = 0 + ๐‘š = ๐‘š,
and by Lemma 5.2.19 and Definition 5.2.15, we have that 1๐‘š = ๐‘š. To show that
the distributive law holds, define
๐ด = {๐‘˜ ∈ ๐œ” : ๐‘˜(๐‘› + ๐‘œ) = ๐‘˜๐‘› + ๐‘˜๐‘œ}.
Since 0(๐‘› + ๐‘œ) = 0 and 0๐‘› + 0๐‘œ = 0 + 0 = 0, we have that 0 ∈ ๐ด. Now suppose
that ๐‘š ∈ ๐ด. That is,
๐‘š(๐‘› + ๐‘œ) = ๐‘š๐‘› + ๐‘š๐‘œ.
Therefore, ๐‘š+ ∈ ๐ด because
๐‘š+ (๐‘› + ๐‘œ) = ๐‘š(๐‘› + ๐‘œ) + (๐‘› + ๐‘œ)
= (๐‘š๐‘› + ๐‘š๐‘œ) + (๐‘› + ๐‘œ)
= (๐‘š๐‘› + ๐‘›) + (๐‘š๐‘œ + ๐‘œ)
= ๐‘š+ ๐‘› + ๐‘š+ ๐‘œ.
EXAMPLE 5.2.21
We revisit the factorial function. Since addition and multiplication are now defined on ๐œ”, let ๐‘” : ๐œ” × ๐œ” → ๐œ” × ๐œ” be the function
๐‘”(๐‘˜, ๐‘™) = (๐‘˜ + 1, (๐‘˜ + 1)๐‘™).
Theorem 5.2.14 implies that there exists a unique function ๐‘“ : ๐œ” → ๐œ” × ๐œ” such
that
๐‘“ (0) = (0, 1)
and
๐‘“ (๐‘› + 1) = ๐‘”(๐‘“ (๐‘›)).
Let ๐œ‹ be the projection map ๐œ‹(๐‘˜, ๐‘™) = ๐‘™. By Theorem 5.2.6 and Exercise 10, we
conclude that for all ๐‘› ∈ ๐œ”,
๐‘“ (๐‘› + 1) = (๐‘› + 1, (๐‘› + 1)(๐œ‹ โ—ฆ ๐‘“ )(๐‘›)).
Therefore, the factorial function ๐‘›!, which is defined recursively by
0! = 1,
(๐‘› + 1)! = (๐‘› + 1)๐‘›! (๐‘› ∈ ๐œ”),
is ๐œ‹ โ—ฆ ๐‘“ by the uniqueness of ๐‘“ .
(5.7)
248
Chapter 5 AXIOMATIC SET THEORY
To solve an equation like 6 + 2๐‘ฅ = 14 where the coefficients are from ๐, we can use
the cancellation law and write
6 + 2๐‘ฅ = 14,
6 + 2๐‘ฅ = 6 + 8,
2๐‘ฅ = 8,
2๐‘ฅ = 2 ⋅ 4,
๐‘ฅ = 4.
For equations with coefficients in ๐œ”, we need a similar law.
THEOREM 5.2.22 [Cancellation]
Let ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ”.
โˆ™ If ๐‘Ž + ๐‘ = ๐‘Ž + ๐‘, then ๐‘ = ๐‘.
โˆ™ If ๐‘Ž๐‘ = ๐‘Ž๐‘ and ๐‘Ž ≠ 0, then ๐‘ = ๐‘.
PROOF
The proof for multiplication is Exercise 15. For addition, define
๐ด = {๐‘˜ ∈ ๐œ” : ∀๐‘™∀๐‘š(๐‘™ ∈ ๐œ” ∧ ๐‘š ∈ ๐œ” ∧ [๐‘˜ + ๐‘™ = ๐‘˜ + ๐‘š → ๐‘™ = ๐‘š])}.
Clearly, 0 ∈ ๐ด, so assume ๐‘› ∈ ๐ด. To prove that ๐‘›+ ∈ ๐ด, let ๐‘, ๐‘ ∈ ๐œ” and suppose
that ๐‘›+ + ๐‘ = ๐‘›+ + ๐‘. Then, (๐‘› + ๐‘)+ = (๐‘› + ๐‘)+ by Lemma 5.2.16, so ๐‘› + ๐‘ = ๐‘› + ๐‘
by Corollary 5.2.12. Hence, ๐‘ = ๐‘ because ๐‘› ∈ ๐ด.
Exercises
1. For every set ๐ด, show that ๐ด+ is a set.
2. For every ๐‘› ∈ ๐œ”, show that ๐‘›+++++ = ๐‘› + 5.
3. Show that the following are equivalent.
โˆ™ ๐ด is transitive.
โˆ™ If ๐‘Ž ∈ ๐ด, then ๐‘Ž ⊂ ๐ด.
โˆ™ ๐ด ⊆ P(๐ด).
โ‹ƒ
โˆ™
๐ด ⊆ ๐ด.
โ‹ƒ +
โˆ™
(๐ด ) = ๐ด.
4. Prove that ๐œ” is transitive.
5. Prove Theorem 5.2.6.
6. Let ๐ด ⊆ ๐œ”. Show that if
โ‹ƒ
๐ด = ๐ด, then ๐ด = ๐œ”.
7. Take ๐‘ข, ๐‘ฃ, ๐‘ฅ, ๐‘ฆ ∈ ๐œ” and assume that ๐‘ข + ๐‘ฅ = ๐‘ฃ + ๐‘ฆ. Prove that ๐‘ข ∈ ๐‘ฃ if and only if
๐‘ฆ ∈ ๐‘ฅ.
Section 5.3 INTEGERS AND RATIONAL NUMBERS
249
8. Prove that ๐‘“ ∈ โ„ฑ in the proof of Theorem 5.2.14.
9. Prove Lemma 5.2.19.
10. Prove (5.7) from Example 5.2.21.
11. Let ๐ด be a set and ๐œ‘ : ๐ด → ๐ด be a one-to-one function. Take ๐‘Ž ∈ ๐ด โงต ran(๐œ‘).
Recursively define ๐‘“ : ๐œ” → ๐ด such that
๐‘“ (0) = ๐‘Ž,
๐‘“ (๐‘›+ ) = ๐œ‘(๐‘“ (๐‘›)).
Prove that ๐‘“ is one-to-one.
12. Let ๐‘š, ๐‘›, ๐‘œ ∈ ๐œ”. Prove the given equations.
(a) ๐‘š๐‘› = ๐‘›๐‘š.
(b) ๐‘š(๐‘› + ๐‘œ) = ๐‘š๐‘› + ๐‘š๐‘œ.
13. Show that ([3 + 3] + 3) + 3 = 12.
14. Prove that addition on ๐œ” is associative.
15. For all ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ” with ๐‘Ž ≠ 0, prove that if ๐‘Ž๐‘ = ๐‘Ž๐‘, then ๐‘ = ๐‘.
16. Show that for all ๐‘š, ๐‘› ∈ ๐œ”, if ๐‘š๐‘› = 0, then ๐‘š = 0 or ๐‘› = 0.
17. We define exponentiation on ๐œ”. For all ๐‘› ∈ ๐œ”,
๐‘›0 = 1,
+
๐‘›๐‘˜ = ๐‘›๐‘˜ ⋅ ๐‘›.
(a) Use the recursion (Theorem 5.2.14) to prove that this defines a function ๐œ” ×
๐œ” → ๐œ”.
(b) Show that exponentiation on ๐œ” is one-to-one.
18. Let ๐‘ฅ, ๐‘ฆ, ๐‘ฅ ∈ ๐œ”. Use the definition of Exercise 17 to prove the given equations.
(a) ๐‘ฅ๐‘ฆ+๐‘ง = ๐‘ฅ๐‘ฆ ๐‘ฅ๐‘ง .
(b) (๐‘ฅ๐‘ฆ)๐‘ง = ๐‘ฅ๐‘ง ๐‘ฆ๐‘ง .
(c) (๐‘ฅ๐‘ฆ )๐‘ง = ๐‘ฅ๐‘ฆ๐‘ง .
19. Let ๐‘š, ๐‘›, ๐‘˜ ∈ ๐œ”. Assume that ๐‘š ≤ ๐‘› and 0 ≤ ๐‘˜. Demonstrate the following.
(a) ๐‘š + ๐‘˜ ≤ ๐‘› + ๐‘˜.
(b) ๐‘š๐‘˜ ≤ ๐‘›๐‘˜.
(c) ๐‘š๐‘˜ ≤ ๐‘›๐‘˜ .
5.3
INTEGERS AND RATIONAL NUMBERS
Now that we have defined within set theory the set of natural numbers and confirmed
its basic properties, we wish to continue this with other sets of numbers. We begin with
the integers.
250
Chapter 5 AXIOMATIC SET THEORY
Integers
We build the integers using the natural numbers. The problem is how to define the
negative integers. We need to decide how to represent the adjoining of the negative
sign to a natural number. One option that might work is to use ordered pairs. These are
always good options when extra information needs to be included with each element
of a set. For example, (4, 0) could represent 4 because 4 − 0 = 4, and (0, 4) could
represent −4 because 0 − 4 = −4. However, this is a problem because we did not
define subtraction on ๐œ”, so we need another solution. Our decision is to generalize
this idea of subtracting coordinates to the set ๐œ” × ๐œ”, but use addition to do it. Since
there are infinitely many pairs (๐‘š, ๐‘›) such that ๐‘š − ๐‘› = 4, we equate them by using the
equivalence relation of Exercise 4.2.2.
DEFINITION 5.3.1
Let ๐‘… be the equivalence relation on ๐œ” × ๐œ” defined by
(๐‘š, ๐‘›) ๐‘… (๐‘š′ , ๐‘›′ ) if and only if ๐‘š + ๐‘›′ = ๐‘š′ + ๐‘›.
Define โ„ค = (๐œ” × ๐œ”)โˆ•๐‘… to be the set of integers.
Using Definition 5.3.1 we associate elements of ๐™ with elements of โ„ค:
๐™
โ‹ฎ
−2
−1
0
1
2
โ‹ฎ
โ„ค
โ‹ฎ
[(0, 2)]
[(0, 1)]
[(0, 0)]
[(1, 0)]
[(2, 0)]
โ‹ฎ
Notice that because 0 and 2 are elements of ๐œ”, the equivalence class [(0, 2)] is the name
for [(∅, {∅, {∅}})]. Also,
[(0, 2)] = [(1, 3)] = [(2, 4)] = · · · .
The ordering of โ„ค is defined in terms of the ordering on ๐œ”. Be careful to note that the
symbol ≤ will represent two different orders, one on โ„ค and one on ๐œ” (Definition 5.2.10).
This overuse of the symbol will not lead to confusion because the order will be clear
from context. For example, in the next definition, the first ≤ is the order on โ„ค, and the
second ≤ is the order on ๐œ”.
DEFINITION 5.3.2
For all [(๐‘š, ๐‘›)], [(๐‘š′ , ๐‘›′ )] ∈ โ„ค, define
[(๐‘š, ๐‘›)] ≤ [(๐‘š′ , ๐‘›′ )] if and only if ๐‘š + ๐‘›′ ≤ ๐‘š′ + ๐‘›
and
Section 5.3 INTEGERS AND RATIONAL NUMBERS
251
[(๐‘š, ๐‘›)] < [(๐‘š′ , ๐‘›′ )] if and only if ๐‘š + ๐‘›′ < ๐‘š′ + ๐‘›.
Since −4 = [(0, 4)] and 3 = [(5, 2)], we have that −4 < 3 because 0 + 2 < 5 + 4.
Since (โ„ค, ≤) has no least element (Exercise 3), it is not well-ordered. However, it is
a chain. Note that in the proof the symbol ≤ is again overused.
THEOREM 5.3.3
(โ„ค, ≤) is a linear order.
PROOF
We use the fact that the order on ๐œ” is a linear order (Theorem 5.2.11). Let ๐‘Ž, ๐‘ ∈
โ„ค, so ๐‘Ž = (๐‘š, ๐‘›) and ๐‘ = (๐‘š′ , ๐‘›′ ) for some ๐‘š, ๐‘›, ๐‘š′ , ๐‘›′ ∈ ๐œ”.
โˆ™ Because ๐‘š + ๐‘› ≤ ๐‘š + ๐‘›, we have ๐‘Ž ≤ ๐‘Ž.
โˆ™ Suppose that ๐‘Ž ≤ ๐‘ and ๐‘ ≤ ๐‘Ž. This implies that ๐‘š + ๐‘›′ ≤ ๐‘š′ + ๐‘› and
๐‘š′ + ๐‘› ≤ ๐‘š + ๐‘›′ . Since ≤ is antisymmetric on ๐œ”, ๐‘š + ๐‘›′ = ๐‘š′ + ๐‘›, which
implies that ๐‘Ž = ๐‘, so ≤ is antisymmetric on โ„ค.
โˆ™ Using a similar strategy, it can be shown that ≤ is transitive on โ„ค (Exercise 4). Thus, (โ„ค, ≤) is a poset.
โˆ™ Since (๐œ”, ≤) is a linear order, ๐‘š + ๐‘›′ ≤ ๐‘š′ + ๐‘› or ๐‘š′ + ๐‘› ≤ ๐‘š + ๐‘›′ . This
implies that ๐‘Ž ≤ ๐‘ or ๐‘ ≤ ๐‘Ž, so โ„ค is a chain under ≤.
Although ๐œ” is not a subset of โ„ค like ๐ is a subset of ๐™, the set ๐œ” can be embedded
in โ„ค (Definition 4.5.29). This is shown by the function ๐œ‘ : ๐œ” → โ„ค defined by
๐œ‘(๐‘›) = [(๐‘›, 0)].
(5.8)
We see that ๐œ‘ is one-to-one because [(๐‘š, 0)] = [(๐‘›, 0)] implies that ๐‘š = ๐‘›. The function
๐œ‘ is then an order isomorphism using the relation from Definition 5.3.2 (Exercise 1).
As with ๐œ”, we define what it means to add and multiply two integers.
DEFINITION 5.3.4
Let [(๐‘š, ๐‘›)], [(๐‘ข, ๐‘ฃ)] ∈ โ„ค.
โˆ™ [(๐‘š, ๐‘›)] + [(๐‘ข, ๐‘ฃ)] = [(๐‘š + ๐‘ข, ๐‘› + ๐‘ฃ)].
โˆ™ [(๐‘š, ๐‘›)] ⋅ [(๐‘ข, ๐‘ฃ)] = [(๐‘š๐‘ข + ๐‘›๐‘ฃ, ๐‘š๐‘ฃ + ๐‘ข๐‘›)].
For example, the equation in โ„ค,
[(5, 2)] + [(7, 1)] = [(12, 3)],
corresponds to the equation in ๐™,
3 + 6 = 9,
252
Chapter 5 AXIOMATIC SET THEORY
and the equation
[(5, 2)] ⋅ [(7, 1)] = [(35 + 2, 5 + 14)] = [(37, 19)]
corresponds to the equation
3 ⋅ 6 = 18.
Before we check the properties of + and ⋅ on โ„ค, we should confirm that these are
well-defined. Exercise 5 is for addition. To prove that multiplication is well-defined, let
[(๐‘š, ๐‘›)] = [(๐‘š′ , ๐‘›′ )] and [(๐‘ข, ๐‘ฃ)] = [(๐‘ข′ , ๐‘ฃ′ )]. Then, ๐‘š + ๐‘›′ = ๐‘š′ + ๐‘› and ๐‘ข + ๐‘ฃ′ = ๐‘ข′ + ๐‘ฃ.
Hence, by Theorem 5.2.20 in ๐œ” we have that
๐‘š๐‘ข + ๐‘›′ ๐‘ข = ๐‘š′ ๐‘ข + ๐‘›๐‘ข,
๐‘š๐‘ฃ + ๐‘›′ ๐‘ฃ = ๐‘š′ ๐‘ฃ + ๐‘›๐‘ฃ,
๐‘š′ ๐‘ข + ๐‘š′ ๐‘ฃ′ = ๐‘š′ ๐‘ข′ + ๐‘š′ ๐‘ฃ,
๐‘›′ ๐‘ข + ๐‘›′ ๐‘ฃ′ = ๐‘›′ ๐‘ข′ + ๐‘›′ ๐‘ฃ.
Therefore,
๐‘š๐‘ข + ๐‘›′ ๐‘ข + ๐‘š′ ๐‘ฃ + ๐‘›๐‘ฃ + ๐‘š′ ๐‘ข + ๐‘š′ ๐‘ฃ′ + ๐‘›′ ๐‘ข′ + ๐‘›′ ๐‘ฃ
equals
๐‘š′ ๐‘ข + ๐‘›๐‘ข + ๐‘š๐‘ฃ + ๐‘›′ ๐‘ฃ + ๐‘š′ ๐‘ข′ + ๐‘š′ ๐‘ฃ + ๐‘›′ ๐‘ข + ๐‘›′ ๐‘ฃ′ ,
so by Cancellation (Theorem 5.2.22),
๐‘š๐‘ข + ๐‘›๐‘ฃ + ๐‘š′ ๐‘ฃ′ + ๐‘›′ ๐‘ข′ = ๐‘š๐‘ฃ + ๐‘›๐‘ข + ๐‘š′ ๐‘ข′ + ๐‘›′ ๐‘ฃ′ .
This implies by Definition 5.3.1 that
[(๐‘š๐‘ข + ๐‘›๐‘ฃ, ๐‘š๐‘ฃ + ๐‘›๐‘ข)] = [(๐‘š′ ๐‘ข′ + ๐‘›′ ๐‘ฃ′ , ๐‘š′ ๐‘ฃ′ + ๐‘›′ ๐‘ข′ )],
and this by Definition 5.3.4 implies that
[(๐‘š, ๐‘›)] ⋅ [(๐‘ข, ๐‘ฃ)] = [(๐‘š′ , ๐‘›′ )] ⋅ [(๐‘ข′ , ๐‘ฃ′ )].
We follow the same order of operations with + and ⋅ on โ„ค as on ๐œ” and also write
๐‘š๐‘› for ๐‘š ⋅ ๐‘›.
THEOREM 5.3.5
โˆ™ The binary operations + and ⋅ on โ„ค are associative and commutative.
โˆ™ [(0, 0)] is the additive identity for โ„ค, and [(1, 0)] is its multiplicative identity.
โˆ™ For every ๐‘Ž ∈ โ„ค, there exists an additive inverse of ๐‘Ž.
โˆ™ For all ๐‘Ž, ๐‘, ๐‘ ∈ โ„ค, the distributive law holds.
Section 5.3 INTEGERS AND RATIONAL NUMBERS
253
PROOF
We will prove parts of the first and third properties and leave the remaining properties to Exercise 6. Let ๐‘Ž, ๐‘, ๐‘ ∈ โ„ค. This means that there exist natural numbers
๐‘š, ๐‘› , ๐‘Ÿ, ๐‘ , ๐‘ข, and ๐‘ฃ such that ๐‘Ž = [(๐‘š, ๐‘›)], ๐‘ = [(๐‘Ÿ, ๐‘ )], and ๐‘ = [(๐‘ข, ๐‘ฃ)]. Then,
๐‘Ž + ๐‘ + ๐‘ = [(๐‘š, ๐‘›)] + [(๐‘Ÿ, ๐‘ )] + [(๐‘ข, ๐‘ฃ)]
= [(๐‘š + ๐‘Ÿ, ๐‘› + ๐‘ )] + [(๐‘ข, ๐‘ฃ)]
= [(๐‘š + ๐‘Ÿ + ๐‘ข, ๐‘› + ๐‘  + ๐‘ฃ)]
= [(๐‘š + [๐‘Ÿ + ๐‘ข] , ๐‘› + [๐‘  + ๐‘ฃ])]
= [(๐‘š, ๐‘›)] + [(๐‘Ÿ + ๐‘ข, ๐‘  + ๐‘ฃ)]
= [(๐‘š, ๐‘›)] + ([(๐‘Ÿ, ๐‘ )] + [(๐‘ข, ๐‘ฃ)])
= ๐‘Ž + (๐‘ + ๐‘)
and
๐‘Ž๐‘๐‘ = [(๐‘š, ๐‘›)] ⋅ [(๐‘Ÿ, ๐‘ )] ⋅ [(๐‘ข, ๐‘ฃ)]
= [(๐‘š๐‘Ÿ + ๐‘›๐‘ , ๐‘š๐‘  + ๐‘›๐‘Ÿ)] ⋅ [(๐‘ข, ๐‘ฃ)]
= [(๐‘ข [๐‘š๐‘Ÿ + ๐‘›๐‘ ] + ๐‘ฃ [๐‘š๐‘  + ๐‘›๐‘Ÿ] , ๐‘ฃ [๐‘š๐‘Ÿ + ๐‘›๐‘ ] + ๐‘ข [๐‘š๐‘  + ๐‘›๐‘Ÿ])]
= [(๐‘ข๐‘š๐‘Ÿ + ๐‘ข๐‘›๐‘  + ๐‘ฃ๐‘š๐‘  + ๐‘ฃ๐‘›๐‘Ÿ, ๐‘ฃ๐‘š๐‘Ÿ + ๐‘ฃ๐‘›๐‘  + ๐‘ข๐‘š๐‘  + ๐‘ข๐‘›๐‘Ÿ)]
= [(๐‘š [๐‘Ÿ๐‘ข + ๐‘ ๐‘ฃ] + ๐‘› [๐‘ ๐‘ข + ๐‘Ÿ๐‘ฃ] , ๐‘š [๐‘ ๐‘ข + ๐‘Ÿ๐‘ฃ] + ๐‘› [๐‘Ÿ๐‘ข + ๐‘ ๐‘ฃ])]
= [(๐‘š, ๐‘›)] ⋅ [(๐‘Ÿ๐‘ข + ๐‘ ๐‘ฃ, ๐‘ ๐‘ข + ๐‘Ÿ๐‘ฃ)]
= [(๐‘š, ๐‘›)] ⋅ ([(๐‘Ÿ, ๐‘ )] ⋅ [(๐‘ข, ๐‘ฃ)])
= ๐‘Ž(๐‘๐‘).
To prove that every element of โ„ค has an additive inverse, notice that
[(๐‘š, ๐‘›)] + [(๐‘›, ๐‘š)] = [(๐‘š + ๐‘›, ๐‘› + ๐‘š)] = [(0, 0)].
Therefore, if ๐‘Ž = [(๐‘š, ๐‘›)], the additive inverse of ๐‘Ž is [(๐‘›, ๐‘š)].
For all ๐‘› ∈ โ„ค, denote the additive inverse of ๐‘› by −๐‘›, and for all ๐‘š, ๐‘›, ๐‘Ÿ, ๐‘  ∈ ๐œ”, define
[(๐‘š, ๐‘›)] − [(๐‘Ÿ, ๐‘ )] = [(๐‘š, ๐‘›)] + [(๐‘ , ๐‘Ÿ)].
Rational Numbers
As the integers were built using ๐œ”, so the set of rational numbers will be built using โ„ค.
Its definition is motivated by the behavior of fractions in ๐. For instance,
2
8
=
3 12
because 2 ⋅ 12 = 3 ⋅ 8. Imagining that the ordered pair (๐‘š, ๐‘›) represents the fraction
๐‘šโˆ•๐‘›, we define an equivalence relation. Notice that this is essentially the relation from
Example 4.2.5. Notice that โ„ค is defined using addition (Definition 5.3.1) while the
rational numbers are defined using multiplication.
254
Chapter 5 AXIOMATIC SET THEORY
DEFINITION 5.3.6
Let ๐‘† be the equivalence relation on โ„ค × (โ„ค โงต {0}) defined by
(๐‘š, ๐‘›) ๐‘† (๐‘š′ , ๐‘›′ ) if and only if ๐‘š๐‘›′ = ๐‘š′ ๐‘›.
Define โ„š = [โ„ค × (โ„ค โงต {0})]โˆ•๐‘† to be the set of rational numbers.
Using Definition 5.3.6, we associate elements of ๐ with elements of โ„š:
๐
โ‹ฎ
−2
−1
0
1
2
โ‹ฎ
๐
โ‹ฎ
1โˆ•2
2โˆ•3
3โˆ•4
4โˆ•5
5โˆ•6
โ‹ฎ
โ„š
โ‹ฎ
[(−2, 1)]
[(−1, 1)]
[(0, 1)]
[(1, 1)]
[(2, 1)]
โ‹ฎ
โ„š
โ‹ฎ
[(1, 2)]
[(2, 3)]
[(3, 4)]
[(4, 5)]
[(5, 6)]
โ‹ฎ
Notice that since 1, 2 ∈ โ„ค, the equivalence class [(1, 2)] is the name for
[([(1, 0)]๐‘… , [(2, 0)]๐‘… ]๐‘† = [[({∅}, ∅)]๐‘… , [({∅, {∅}}, ∅)]๐‘… ]๐‘† ,
where ๐‘… is the relation of Definition 5.3.1. Also,
[(1, 2)] = [(2, 4)] = [(3, 6)] = · · · .
We next define a partial order on โ„š.
DEFINITION 5.3.7
For all [(๐‘š, ๐‘›)], [(๐‘š′ , ๐‘›′ )] ∈ โ„š, define
[(๐‘š, ๐‘›)] ≤ [(๐‘š′ , ๐‘›′ )] if and only if ๐‘š๐‘›′ ≤ ๐‘›๐‘š′
and
[(๐‘š, ๐‘›)] < [(๐‘š′ , ๐‘›′ )] if and only if ๐‘š๐‘›′ < ๐‘›๐‘š′ .
When working with โ„š, we often denote [(๐‘š, ๐‘›)] by ๐‘šโˆ•๐‘› or ๐‘š๐‘› and [(๐‘š, 1)] by ๐‘š. Thus,
since 2โˆ•3 = [(2, 3)] and 7โˆ•8 = [(7, 8)], we conclude that 2โˆ•3 < 7โˆ•8 because 2⋅8 < 3⋅7.
As ๐œ” can be embedded in โ„ค, so โ„ค can be embedded in โ„š. The function that can be
used is
[
]
๐œ“ : โ„ค → โ„ค × (โ„ค โงต {0}) โˆ•๐‘†
defined by
๐œ“(๐‘›) = [(๐‘›, 1)].
(5.9)
(See Exercise 2) Using the function ๐œ‘ (5.8), we see that ๐œ” can be embedded in โ„š via
the order isomorphism ๐œ“ โ—ฆ ๐œ‘.
Since (โ„š, ≤) has no least element (Exercise 3), it is not well-ordered. However, it is
a chain (Exercise 11).
Section 5.3 INTEGERS AND RATIONAL NUMBERS
255
THEOREM 5.3.8
(โ„š, ≤) is a linear order.
Lastly, we define two operations on โ„š that represent the standard operations of +
and ⋅ on ๐.
DEFINITION 5.3.9
Let [(๐‘š, ๐‘›)], [(๐‘š′ , ๐‘›′ )] ∈ โ„š.
โˆ™ [(๐‘š, ๐‘›)] + [(๐‘š′ , ๐‘›′ )] = [(๐‘š๐‘›′ + ๐‘š′ ๐‘›, ๐‘›๐‘›′ )].
โˆ™ [(๐‘š, ๐‘›)] ⋅ [(๐‘š′ , ๐‘›′ )] = [(๐‘š๐‘š′ , ๐‘›๐‘›′ )].
That + and ⋅ as given in Definition 5.3.9 are binary operations is left to Exercise 13.
As examples of these two operations, the equation in โ„š
[(1, 3)] + [(3, 4)] = [(1 ⋅ 4 + 3 ⋅ 3, 3 ⋅ 4)] = [(13, 12)]
corresponds to the equation in ๐
1 3
4
9
13
+ =
+
=
3 4 12 12 12
and the equation in โ„š
[(1, 3)] ⋅ [(3, 4)] = [(1 ⋅ 3, 3 ⋅ 4)] = [(3, 12)] = [(1, 4)]
corresponds to the equation in ๐
3
1
1 3
⋅ =
= .
3 4 12 4
We follow the same order of operations with + and ⋅ on โ„š as on โ„ค and also write
๐‘š๐‘› for ๐‘š ⋅ ๐‘›.
THEOREM 5.3.10
โˆ™ The binary operations + and ⋅ on โ„š are associative and commutative.
โˆ™ [(0, 1)] is the additive identity, and [(1, 1)] is the multiplicative identity.
โˆ™ Every rational number has an additive inverse, and every element of โ„šโงต{[(0, 1)]}
has a multiplicative inverse.
โˆ™ The distributive law holds.
256
Chapter 5 AXIOMATIC SET THEORY
PROOF
Since [(๐‘š, ๐‘›)] ⋅ [(1, 1)] = [(๐‘š ⋅ 1, ๐‘› ⋅ 1)] = [(๐‘š, ๐‘›)] for all [(๐‘š, ๐‘›)] ∈ โ„š and
multiplication is commutative, the multiplicative identity of โ„š is [(1, 1)]. Also,
because ๐‘š ≠ 0 and ๐‘› ≠ 0 implies that
[(๐‘š, ๐‘›)] ⋅ [(๐‘›, ๐‘š)] = [(๐‘š๐‘›, ๐‘›๐‘š)] = [(1, 1)],
every element of โ„š โงต [(0, 1)] has a multiplicative inverse. The other properties
are left to Exercise 14.
For all ๐‘Ž, ๐‘ ∈ โ„š with ๐‘ ≠ 0, denote the additive inverse of ๐‘Ž by −๐‘Ž and the multiplicative
inverse of ๐‘ by ๐‘−1 .
Actual Numbers
We now use the elements of ๐œ”, โ„ค, and โ„š as if they were the actual elements of ๐, ๐™,
and ๐. For example, we understand the formula ๐‘› ∈ โ„ค to mean that ๐‘› is an integer
of Definition 5.3.1 with all of the properties of ๐‘› ∈ ๐—ญ. Also, when we write that the
formula ๐‘(๐‘›) is satisfied by some rational number ๐‘Ž, we interpret this to mean that ๐‘(๐‘Ž)
with ๐‘Ž ∈ โ„š where ๐‘Ž has all of the properties of ๐‘Ž ∈ ๐. This allows us to use the set
properties of a natural number, integer, or rational number when needed yet also use
the results we know concerning the actual numbers. That the partial orders and binary
operations defined on ๐œ”, โ„ค, and โ„š have essentially the same properties as those on ๐,
๐™, and ๐ allows for this association of properties to be legitimate.
Exercises
1. Prove that the function ๐œ‘ (5.8) is an order isomorphism using the order of Definition 5.3.2.
2. Prove that the function ๐œ“ (5.9) is an order isomorphism using the order of Definition 5.3.7.
3. Show that (โ„ค, ≤) and (โ„š, ≤) do not have least elements.
4. Prove that ≤ is transitive on โ„ค.
5. Prove that + is well-defined on โ„ค.
6. Complete the remaining proofs of the properties of Theorem 5.3.5.
7. Show that every nonempty subset of โ„ค− = {๐‘› : ๐‘› ∈ โ„ค ∧ ๐‘› < 0} has a greatest
element.
8. Let ๐‘š, ๐‘› ∈ โ„ค and ๐‘˜ ∈ โ„ค− (Exercise 7) Prove that if ๐‘š ≤ ๐‘›, then ๐‘›๐‘˜ ≤ ๐‘š๐‘˜.
9. Prove that for every ๐‘š, ๐‘› ∈ โ„ค, if ๐‘š๐‘› = 0, then ๐‘š = 0 or ๐‘› = 0. Show that this same
result also holds for โ„š.
10. The cancellation law for the integers states that for all ๐‘š, ๐‘›, ๐‘˜ ∈ โ„ค,
โˆ™ if ๐‘š + ๐‘˜ = ๐‘› + ๐‘˜, then ๐‘š = ๐‘›,
โˆ™ if ๐‘š๐‘˜ = ๐‘›๐‘˜ and ๐‘˜ ≠ 0, then ๐‘š = ๐‘›.
Section 5.4 MATHEMATICAL INDUCTION
257
Prove that the cancellation law holds for โ„ค. Prove that a similar law holds for โ„š.
11. Prove that (โ„š, ≤) is a chain.
12. Demonstrate that every nonempty set of integers with a lower bound with respect
to ≤ has a least element.
13. Show that + and ⋅ as given in Definition 5.3.9 are binary operations on โ„š.
14. Prove the remaining parts of Theorem 5.3.10.
15. Let ๐‘Ž, ๐‘, ๐‘, ๐‘‘ ∈ โ„š. Prove.
(a) If 0 ≤ ๐‘Ž ≤ ๐‘ and 0 ≤ ๐‘ ≤ ๐‘‘, then ๐‘Ž๐‘ ≤ ๐‘๐‘‘.
(b) If 0 ≤ ๐‘Ž < ๐‘ and 0 ≤ ๐‘ < ๐‘‘, then ๐‘Ž๐‘ < ๐‘๐‘‘.
16. Let ๐‘Ž, ๐‘ ∈ โ„š. Prove that if ๐‘ < 0, then ๐‘Ž + ๐‘ < ๐‘Ž.
17. Let ๐‘Ž, ๐‘ ∈ โ„š. Prove that if ๐‘Ž ≥ 0 and ๐‘ < 1, then ๐‘Ž๐‘ ≤ ๐‘Ž.
18. Prove that between any two rational numbers is a rational number. This means that
โ„š is dense. Show that this is not the case for โ„ค.
19. Generalize the definition of exponentiation given in Exercise 5.2.17 to โ„ค. Prove
that this defines a one-to-one function โ„ค × (โ„ค+ ∪ {0}) → โ„ค. Does the proof require
recursion (Theorem 5.2.14)?
20. Let ๐‘š, ๐‘›, ๐‘˜ ∈ โ„ค. Assume that ๐‘š ≤ ๐‘›. Demonstrate the following.
(a) ๐‘š + ๐‘˜ ≤ ๐‘› + ๐‘˜.
(b) ๐‘š๐‘˜ ≤ ๐‘›๐‘˜ if 0 ≤ ๐‘˜.
(c) ๐‘›๐‘˜ ≤ ๐‘š๐‘˜ if ๐‘˜ < 0.
(d) ๐‘š๐‘˜ ≤ ๐‘›๐‘˜ if ๐‘˜ ≥ 0.
21. Let ๐œ‘ : ๐œ” → โ„ค be the embedding defined by (5.8). Define ๐‘“ : ๐œ” × ๐œ” → ๐œ” by
๐‘“ (๐‘š, ๐‘›) = ๐‘š๐‘› and ๐‘” : โ„ค × (โ„ค+ ∪ {0}) → โ„ค by ๐‘”(๐‘ข, ๐‘ฃ) = ๐‘ข๐‘ฃ . Prove that for all ๐‘š, ๐‘› ∈ ๐œ”,
we have that ๐‘”(๐œ‘(๐‘š), ๐œ‘(๐‘›)) = ๐œ‘(๐‘“ (๐‘š, ๐‘›)). Explain the significance of this result.
5.4
MATHEMATICAL INDUCTION
Suppose that we want to prove ๐‘(๐‘›) for all integers ๐‘› greater than or equal to some ๐‘›0 .
Our previous method (Section 2.4) is to take an arbitrary ๐‘› ≥ ๐‘›0 and try to prove ๐‘(๐‘›).
If this is not possible, we might be tempted to try proving ๐‘(๐‘›) for each ๐‘› individually.
This is impossible because it would take infinitely many steps. Instead, we combine
the results of Sections 5.2 and 5.3 and use the next theorem.
THEOREM 5.4.1 [Mathematical Induction 1]
Let ๐‘(๐‘˜) be a formula. For any ๐‘›0 ∈ โ„ค, if
๐‘(๐‘›0 ) ∧ ∀๐‘› ∈ โ„ค[๐‘› ≥ ๐‘›0 ∧ ๐‘(๐‘›) → ๐‘(๐‘› + 1)],
then
∀๐‘› ∈ โ„ค[๐‘› ≥ ๐‘›0 → ๐‘(๐‘›)].
258
Chapter 5 AXIOMATIC SET THEORY
PROOF
Assume ๐‘(๐‘›0 ) and that ๐‘(๐‘˜) implies ๐‘(๐‘˜ + 1) for all integers ๐‘˜ ≥ ๐‘›0 . Define
๐ด = {๐‘˜ ∈ ๐œ” : ๐‘(๐‘›0 + ๐‘˜)}.
Notice that 0 ∈ ๐ด ⇔ ๐‘(๐‘›0 ), 1 ∈ ๐ด ⇔ ๐‘(๐‘›0 + 1), and so on. To prove ๐‘(๐‘˜) for all
integers ๐‘˜ ≥ ๐‘›0 , we show that ๐ด is inductive.
โˆ™ By hypothesis, 0 ∈ ๐ด because ๐‘(๐‘›0 ).
โˆ™ Assume ๐‘› ∈ ๐ด. This implies that ๐‘(๐‘›0 + ๐‘›). Therefore, ๐‘(๐‘›0 + ๐‘› + 1), so
๐‘› + 1 ∈ ๐ด.
Theorem 5.4.1 gives rise to a standard proof technique known as mathematical
induction. First, prove ๐‘(๐‘›0 ). Then, show that ๐‘(๐‘›) implies ๐‘(๐‘› + 1) for every integer
๐‘› ≥ ๐‘›0 . Often an analogy of dominoes is used to explain this. Proving ๐‘(๐‘›0 ) is like
tipping over the first domino, and then proving the implication shows that the dominoes
have been set properly. This means that by modus ponens, if ๐‘(๐‘›0 + 1) is true, then
๐‘(๐‘›0 + 2) is true, and so forth, each falling like dominoes:
๐‘(๐‘›0 )
๐‘(๐‘›0 ) → ๐‘(๐‘›0 + 1)
∴ ๐‘(๐‘›0 + 1)
๐‘(๐‘›0 + 1) → ๐‘(๐‘›0 + 2)
∴ ๐‘(๐‘›0 + 2) · · ·
This two-step process is characteristic of proofs by mathematical induction. So
much so, the two stages have their own terminology.
โˆ™ Proving ๐‘(๐‘›0 ) is called the basis case. It is typically the easiest part of the proof
but should, nonetheless, be explicitly shown.
โˆ™ Proving that ๐‘(๐‘›) implies ๐‘(๐‘› + 1) is the induction step. For this, we typically
use direct proof, assuming ๐‘(๐‘›) to show ๐‘(๐‘› + 1). The assumption is called the
induction hypothesis.
Often induction is performed to prove a formula for all positive integers, so to represent
this set, define
โ„ค+ = {1, 2, 3, … }.
EXAMPLE 5.4.2
Prove ๐‘(๐‘˜) for all ๐‘˜ ∈ โ„ค+ , where
๐‘(๐‘˜) := 12 + 22 + · · · + ๐‘˜2 =
๐‘˜(๐‘˜ + 1)(2๐‘˜ + 1)
.
6
Proceed by mathematical induction.
โˆ™ The basis case ๐‘(1) holds because
1(1 + 1)(2 ⋅ 1 + 1) 1(2)(3)
=
= 12 .
6
6
Section 5.4 MATHEMATICAL INDUCTION
259
โˆ™ Now for the induction step, assume
12 + 22 + · · · + ๐‘›2 =
๐‘›(๐‘› + 1)(2๐‘› + 1)
.
6
(5.10)
This is the induction hypothesis. We must show that the equation holds for
๐‘› + 1. Adding (๐‘› + 1)2 to both sides of (5.10) gives
12 + 22 + · · · + ๐‘›2 + (๐‘› + 1)2 =
=
=
=
=
๐‘›(๐‘› + 1)(2๐‘› + 1)
+ (๐‘› + 1)2
6
(๐‘› + 1) [๐‘›(2๐‘› + 1) + 6(๐‘› + 1)]
6
(๐‘› + 1)(2๐‘›2 + 7๐‘› + 6)
6
(๐‘› + 1)(๐‘› + 2)(2๐‘› + 3)
6
(๐‘› + 1)([๐‘› + 1] + 1)(2[๐‘› + 1] + 1)
.
6
EXAMPLE 5.4.3
Let ๐‘ฅ be a positive rational number. To prove that for any ๐‘˜ ∈ โ„ค+ ,
(๐‘ฅ + 1)๐‘˜ ≥ ๐‘ฅ๐‘˜ + 1,
we proceed by mathematical induction.
โˆ™ (๐‘ฅ + 1)1 = ๐‘ฅ + 1 = ๐‘ฅ1 + 1.
โˆ™ Let ๐‘› ∈ โ„ค+ and assume that (๐‘ฅ + 1)๐‘› ≥ ๐‘ฅ๐‘› + 1. By multiplying both sides
of the given inequality by ๐‘ฅ + 1, we have
(๐‘ฅ + 1)๐‘› (๐‘ฅ + 1) ≥ (๐‘ฅ๐‘› + 1)(๐‘ฅ + 1)
= ๐‘ฅ๐‘›+1 + ๐‘ฅ๐‘› + ๐‘ฅ + 1
≥ ๐‘ฅ๐‘›+1 + 1.
The first inequality is true by induction (that is, by appealing to the induction hypothesis) and since ๐‘ฅ + 1 is positive, and the last one holds because
๐‘ฅ ≥ 0.
EXAMPLE 5.4.4
Recursively define a sequence of numbers,
๐‘Ž1 = 3,
๐‘Ž๐‘› = 2๐‘Ž๐‘›−1 for all ๐‘› ∈ โ„ค such that ๐‘› > 1.
260
Chapter 5 AXIOMATIC SET THEORY
The sequence is
๐‘Ž1 = 3, ๐‘Ž2 = 6, ๐‘Ž3 = 12, ๐‘Ž4 = 24, ๐‘Ž5 = 48, … ,
and we conjecture that ๐‘Ž๐‘˜ = 3 ⋅ 2๐‘˜−1 for all positive integers ๐‘˜. We prove this by
mathematical induction.
โˆ™ For the basis case, ๐‘Ž1 = 3 ⋅ 20 = 3.
โˆ™ Let ๐‘› > 1 and assume ๐‘Ž๐‘› = 3 ⋅ 2๐‘›−1 . Then,
๐‘Ž๐‘›+1 = 2๐‘Ž๐‘› = 2 ⋅ 3 ⋅ 2๐‘›−1 = 3 ⋅ 2๐‘› .
EXAMPLE 5.4.5
Use mathematical induction to prove that ๐‘˜3 < ๐‘˜! for all integers, ๐‘˜ ≥ 6.
โˆ™ First, show that the inequality holds for ๐‘› = 6:
63 = 216 < 720 = 6!.
โˆ™ Assume ๐‘›3 < ๐‘›! with ๐‘› ≥ 6. The induction hypothesis yields three inequalities. Namely,
3๐‘›2 < ๐‘› ⋅ ๐‘›2 ≤ ๐‘›!,
3๐‘› < ๐‘› ⋅ ๐‘› < ๐‘›3 ≤ ๐‘›!,
and
1 < ๐‘›!.
Therefore,
(๐‘› + 1)3 = ๐‘›3 + 3๐‘›2 + 3๐‘› + 1
< ๐‘›! + ๐‘›! + ๐‘›! + ๐‘›!
= 4๐‘›!
< (๐‘› + 1)๐‘›!
= (๐‘› + 1)!.
Combinatorics
We now use mathematical induction to prove some basic results from two areas of
mathematics. The first is combinatorics, the study of the properties that sets have
based purely on their size.
A permutation of a given set is an arrangement of the elements of the set. For
example, the number of permutations of {๐‘Ž, ๐‘, ๐‘, ๐‘‘, ๐‘’, ๐‘“ } is 720. If we were to write all
of the permutations in a list, it would look like the following:
Section 5.4 MATHEMATICAL INDUCTION
๐‘Ž
๐‘Ž
๐‘Ž
๐‘
๐‘
๐‘
๐‘
๐‘
๐‘
๐‘“
๐‘“
๐‘’
๐‘’
๐‘‘
๐‘‘
๐‘‘
๐‘‘
๐‘’
๐‘’
๐‘“
๐‘‘
๐‘“
๐‘’
๐‘“
๐‘
๐‘
๐‘Ž
๐‘
๐‘
๐‘Ž
261
โ‹ฎ
We observe that 6! = 720 and hypothesize that the number of permutations of a set
with ๐‘˜ elements is ๐‘˜!. To prove it, we use mathematical induction.
โˆ™ There is only one way to write the elements of a singleton. Since 1! = 1, we have
proved the basis case.
โˆ™ Assume that the number of permutations of a set with ๐‘› ≥ 1 elements is ๐‘›!. Let
๐ด = {๐‘Ž1 , ๐‘Ž2 , … , ๐‘Ž๐‘›+1 } be a set with ๐‘› + 1 elements. By induction, there are ๐‘›!
permutations of the set {๐‘Ž1 , ๐‘Ž2 , … , ๐‘Ž๐‘› }. After writing the permutations in a list,
notice that there are ๐‘› + 1 columns before, between, and after each element of
the permutations:
๐‘Ž1
๐‘Ž1
โ‹ฎ
๐‘Ž๐‘›
…
…
๐‘Ž2
๐‘Ž2
โ‹ฎ
๐‘Ž๐‘›−1
…
๐‘Ž๐‘›
๐‘Ž๐‘›−1
โ‹ฎ
๐‘Ž1
To form the permutations of ๐ด, place ๐‘Ž๐‘›+1 into the positions of each empty column. For example, if ๐‘Ž๐‘›+1 is put into the first column, the following permutations
are obtained:
๐‘Ž๐‘›+1
๐‘Ž๐‘›+1
โ‹ฎ
๐‘Ž๐‘›+1
๐‘Ž1
๐‘Ž1
โ‹ฎ
๐‘Ž๐‘›
๐‘Ž2
๐‘Ž2
โ‹ฎ
๐‘Ž๐‘›−1
…
…
…
๐‘Ž๐‘›
๐‘Ž๐‘›−1
โ‹ฎ
๐‘Ž1
Since there are ๐‘›! rows with ๐‘› + 1 ways to add ๐‘Ž๐‘›+1 to each row, we conclude
that there are (๐‘› + 1)๐‘›! = (๐‘› + 1)! permutations of ๐ด.
This argument proves the first theorem.
THEOREM 5.4.6
Let ๐‘› ∈ โ„ค+ . The number of permutations of a set with ๐‘› elements is ๐‘›!.
Suppose that we do not want to rearrange the entire set but only subsets of it. For
example, let ๐ด = {๐‘Ž, ๐‘, ๐‘, ๐‘‘, ๐‘’}. To see all three-element permutations of ๐ด, look at the
262
Chapter 5 AXIOMATIC SET THEORY
following list:
๐‘Ž๐‘๐‘
๐‘Ž๐‘๐‘‘
๐‘Ž๐‘๐‘’
๐‘Ž๐‘๐‘‘
๐‘Ž๐‘๐‘’
๐‘Ž๐‘‘๐‘’
๐‘๐‘๐‘‘
๐‘๐‘๐‘’
๐‘๐‘‘๐‘’
๐‘๐‘‘๐‘’
๐‘Ž๐‘๐‘
๐‘Ž๐‘‘๐‘
๐‘Ž๐‘’๐‘
๐‘Ž๐‘‘๐‘
๐‘Ž๐‘’๐‘
๐‘Ž๐‘’๐‘‘
๐‘๐‘‘๐‘
๐‘๐‘’๐‘
๐‘๐‘’๐‘‘
๐‘๐‘’๐‘‘
๐‘๐‘Ž๐‘
๐‘๐‘Ž๐‘‘
๐‘๐‘Ž๐‘’
๐‘๐‘Ž๐‘‘
๐‘๐‘Ž๐‘’
๐‘‘๐‘Ž๐‘’
๐‘๐‘๐‘‘
๐‘๐‘’๐‘
๐‘‘๐‘๐‘’
๐‘‘๐‘๐‘’
๐‘๐‘๐‘Ž
๐‘๐‘‘๐‘Ž
๐‘๐‘’๐‘Ž
๐‘๐‘‘๐‘Ž
๐‘๐‘’๐‘Ž
๐‘‘๐‘’๐‘Ž
๐‘๐‘‘๐‘
๐‘๐‘๐‘’
๐‘‘๐‘’๐‘
๐‘‘๐‘’๐‘
๐‘๐‘Ž๐‘
๐‘‘๐‘Ž๐‘
๐‘’๐‘Ž๐‘
๐‘‘๐‘Ž๐‘
๐‘’๐‘Ž๐‘
๐‘’๐‘Ž๐‘‘
๐‘‘๐‘๐‘
๐‘’๐‘๐‘
๐‘’๐‘๐‘‘
๐‘’๐‘๐‘‘
๐‘๐‘๐‘Ž
๐‘‘๐‘๐‘Ž
๐‘’๐‘๐‘Ž
๐‘‘๐‘๐‘Ž
๐‘’๐‘๐‘Ž
๐‘’๐‘‘๐‘Ž
๐‘‘๐‘๐‘
๐‘’๐‘๐‘
๐‘’๐‘‘๐‘
๐‘’๐‘‘๐‘
There are 60 arrangements because there are 5 choices for the first entry. Once that is
chosen, there are only 4 left for the second, and then 3 for the last. We calculate that as
60 = 5 ⋅ 4 ⋅ 3 =
5!
5⋅4⋅3⋅2⋅1
=
.
2⋅1
(5 − 3)!
Generalizing, we define for all ๐‘›, ๐‘Ÿ ∈ ๐œ”,
๐‘› ๐‘ƒ๐‘Ÿ
=
๐‘›!
,
(๐‘› − ๐‘Ÿ)!
and we conclude the following theorem.
THEOREM 5.4.7
Let ๐‘Ÿ, ๐‘› ∈ โ„ค+ . The number of permutations of ๐‘Ÿ elements from a set with ๐‘›
elements is ๐‘› ๐‘ƒ๐‘Ÿ .
Now suppose that we only want to count subsets. For example, ๐ด = {๐‘Ž, ๐‘, ๐‘, ๐‘‘, ๐‘’}
has 10 subsets of three elements. They are the following:
{๐‘Ž, ๐‘, ๐‘}
{๐‘Ž, ๐‘‘, ๐‘’}
{๐‘Ž, ๐‘, ๐‘‘} {๐‘Ž, ๐‘, ๐‘’} {๐‘Ž, ๐‘, ๐‘‘} {๐‘Ž, ๐‘, ๐‘’}
{๐‘, ๐‘, ๐‘‘} {๐‘, ๐‘, ๐‘’} {๐‘, ๐‘‘, ๐‘’} {๐‘, ๐‘‘, ๐‘’}
The number of subsets can be calculated by considering the next grid.
๐‘Ž๐‘๐‘ ๐‘Ž๐‘๐‘ ๐‘๐‘Ž๐‘ ๐‘๐‘๐‘Ž ๐‘๐‘Ž๐‘ ๐‘๐‘๐‘Ž
๐‘Ž๐‘๐‘‘ ๐‘Ž๐‘‘๐‘ ๐‘๐‘Ž๐‘‘ ๐‘๐‘‘๐‘Ž ๐‘‘๐‘Ž๐‘ ๐‘‘๐‘๐‘Ž
๐‘Ž๐‘๐‘’ ๐‘Ž๐‘’๐‘ ๐‘๐‘Ž๐‘’ ๐‘๐‘’๐‘Ž ๐‘’๐‘Ž๐‘ ๐‘’๐‘๐‘Ž
๐‘Ž๐‘๐‘‘ ๐‘Ž๐‘‘๐‘ ๐‘๐‘Ž๐‘‘ ๐‘๐‘‘๐‘Ž ๐‘‘๐‘Ž๐‘ ๐‘‘๐‘๐‘Ž
๐‘Ž๐‘๐‘’ ๐‘Ž๐‘’๐‘ ๐‘๐‘Ž๐‘’ ๐‘๐‘’๐‘Ž ๐‘’๐‘Ž๐‘ ๐‘’๐‘๐‘Ž
๐‘Ž๐‘‘๐‘’ ๐‘Ž๐‘’๐‘‘ ๐‘‘๐‘Ž๐‘’ ๐‘‘๐‘’๐‘Ž ๐‘’๐‘Ž๐‘‘ ๐‘’๐‘‘๐‘Ž
๐‘๐‘๐‘‘ ๐‘๐‘‘๐‘ ๐‘๐‘๐‘‘ ๐‘๐‘‘๐‘ ๐‘‘๐‘๐‘ ๐‘‘๐‘๐‘
๐‘๐‘๐‘’ ๐‘๐‘’๐‘ ๐‘๐‘’๐‘ ๐‘๐‘๐‘’ ๐‘’๐‘๐‘ ๐‘’๐‘๐‘
๐‘๐‘‘๐‘’ ๐‘๐‘’๐‘‘ ๐‘‘๐‘๐‘’ ๐‘‘๐‘’๐‘ ๐‘’๐‘๐‘‘ ๐‘’๐‘‘๐‘
๐‘๐‘‘๐‘’ ๐‘๐‘’๐‘‘ ๐‘‘๐‘๐‘’ ๐‘‘๐‘’๐‘ ๐‘’๐‘๐‘‘ ๐‘’๐‘‘๐‘
โŸโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโŸโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโžโŸ
3! columns
โŽซ
โŽช
โŽช
โŽช
โŽช
โŽช
โŽฌ 10 rows
โŽช
โŽช
โŽช
โŽช
โŽช
โŽญ
Section 5.4 MATHEMATICAL INDUCTION
263
There are 5 ๐‘ƒ3 permutations with three elements from ๐ด. They are found as the entries
in the grid. However, since we are looking at subsets, we do not want to count ๐‘Ž๐‘๐‘
as different from ๐‘Ž๐‘๐‘ because {๐‘Ž, ๐‘, ๐‘} = {๐‘Ž, ๐‘, ๐‘}. For this reason, all elements in any
given row of the grid are considered as one subset. Each row has 6 = 3! entries because
that is the number of permutations of a set with three elements. Hence, multiplying the
number of rows by the number of columns gives
5 ๐‘ƒ3
= 10(3!).
Therefore,
5 ๐‘ƒ3
3!
=
5!
5!
=
= 10.
3!(5 − 3)! 3!2!
A generalization of this calculation leads to the formula for the arbitrary binomial
coefficient,
( )
๐‘›
๐‘›!
=
,
๐‘Ÿ
๐‘Ÿ!(๐‘› − ๐‘Ÿ)!
()
where ๐‘›, ๐‘Ÿ ∈ ๐œ”. Read ๐‘›๐‘Ÿ as “๐‘› choose ๐‘Ÿ.” A generalization of the argument leads to
the next theorem.
THEOREM 5.4.8
Let( ๐‘›,
๐‘Ÿ ∈ โ„ค+ . The number of subsets of ๐‘Ÿ elements from a set with ๐‘› elements
๐‘›)
is ๐‘Ÿ .
When we expand (๐‘ฅ + 1)3 , we find that
(๐‘ฅ + 1)3 = ๐‘ฅ3 + 3๐‘ฅ2 + 3๐‘ฅ2 + 1 =
3 ( )
∑
3 3−๐‘Ÿ ๐‘Ÿ
๐‘ฅ 1.
๐‘Ÿ
๐‘Ÿ=0
To prove this for any binomial (๐‘ฅ + ๐‘ฆ)๐‘› , we need the following equation. It was proved
by Blaise Pascal (1653). The proof is Exercise 7.
LEMMA 5.4.9 [Pascal’s Identity]
If ๐‘›, ๐‘Ÿ ∈ ๐œ” so that ๐‘› ≥ ๐‘Ÿ,
( ) (
) (
)
๐‘›
๐‘›
๐‘›+1
+
=
.
๐‘Ÿ
๐‘Ÿ−1
๐‘Ÿ
THEOREM 5.4.10 [Binomial Theorem]
Let ๐‘› ∈ โ„ค+ . Then,
๐‘› ( )
∑
๐‘› ๐‘›−๐‘Ÿ ๐‘Ÿ
(๐‘ฅ + ๐‘ฆ) =
๐‘ฅ ๐‘ฆ.
๐‘Ÿ
๐‘Ÿ=0
๐‘›
264
Chapter 5 AXIOMATIC SET THEORY
PROOF
โˆ™ Since
( 1)
0
=
(1)
1
= 1,
( )
( )
1 ( )
∑
1 1−๐‘Ÿ ๐‘Ÿ
1
1
๐‘ฅ ๐‘ฆ.
๐‘ฆ=
๐‘ฅ+
(๐‘ฅ + ๐‘ฆ) =
๐‘Ÿ
1
0
๐‘Ÿ=0
1
โˆ™ Assume for ๐‘˜ ∈ โ„ค+ ,
๐‘˜ ( )
∑
๐‘˜ ๐‘˜−๐‘Ÿ ๐‘Ÿ
๐‘ฅ ๐‘ฆ.
(๐‘ฅ + ๐‘ฆ) =
๐‘Ÿ
๐‘Ÿ=0
๐‘˜
Then,
(๐‘ฅ + ๐‘ฆ)๐‘˜+1 = (๐‘ฅ + ๐‘ฆ)(๐‘ฅ + ๐‘ฆ)๐‘˜ = (๐‘ฅ + ๐‘ฆ)
๐‘˜ ( )
∑
๐‘˜ ๐‘˜−๐‘Ÿ ๐‘Ÿ
๐‘ฅ ๐‘ฆ.
๐‘Ÿ
๐‘Ÿ=0
Multiplying the (๐‘ฅ + ๐‘ฆ) term through the summation yields
๐‘˜ ( )
๐‘˜ ( )
∑
๐‘˜ ๐‘˜−๐‘Ÿ+1 ๐‘Ÿ ∑ ๐‘˜ ๐‘˜−๐‘Ÿ ๐‘Ÿ+1
๐‘ฅ
๐‘ฆ +
๐‘ฅ ๐‘ฆ .
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ=0
๐‘Ÿ=0
Taking out the (๐‘˜ + 1)-degree terms and shifting the index on the second summation gives
๐‘ฅ
๐‘˜+1
)
๐‘˜ ( )
๐‘˜ (
∑
๐‘˜ ๐‘˜−๐‘Ÿ+1 ๐‘Ÿ ∑
๐‘˜
+
๐‘ฅ
๐‘ฆ +
๐‘ฅ๐‘˜−๐‘Ÿ+1 ๐‘ฆ๐‘Ÿ + ๐‘ฆ๐‘›+1 ,
๐‘Ÿ
๐‘Ÿ
−
1
๐‘Ÿ=1
๐‘Ÿ=1
which using Pascal’s identity (Lemma 5.4.9) equals
๐‘ฅ๐‘˜+1 +
and this is
)
๐‘˜ (
∑
๐‘˜ + 1 ๐‘˜−๐‘Ÿ+1 ๐‘Ÿ
๐‘ฅ
๐‘ฆ + ๐‘ฆ๐‘˜+1 ,
๐‘Ÿ
๐‘Ÿ=1
(
๐‘˜+1
∑
๐‘Ÿ=0
)
๐‘˜ + 1 ๐‘˜+1−๐‘Ÿ ๐‘Ÿ
๐‘ฅ
๐‘ฆ.
๐‘Ÿ
Euclid’s Lemma
Our second application of mathematical induction comes from number theory. It is the
study of the greatest common divisor (Definition 3.3.11). We begin with a lemma.
LEMMA 5.4.11
Let ๐‘Ž, ๐‘, ๐‘ ∈ โ„ค such that ๐‘Ž ≠ 0 or ๐‘ ≠ 0. If ๐‘Ž โˆฃ ๐‘๐‘ and gcd(๐‘Ž, ๐‘) = 1, then ๐‘Ž โˆฃ ๐‘.
Section 5.4 MATHEMATICAL INDUCTION
265
PROOF
Assume ๐‘Ž โˆฃ ๐‘๐‘ and gcd(๐‘Ž, ๐‘) = 1. Then, ๐‘๐‘ = ๐‘Ž๐‘˜ for some ๐‘˜ ∈ โ„ค. By Theorem 4.3.32, there exist ๐‘š, ๐‘› ∈ โ„ค such that
1 = ๐‘š๐‘Ž + ๐‘›๐‘.
Therefore, ๐‘Ž โˆฃ ๐‘ because
๐‘ = ๐‘๐‘š๐‘Ž + ๐‘๐‘›๐‘ = ๐‘๐‘š๐‘Ž + ๐‘›๐‘Ž๐‘˜ = ๐‘Ž(๐‘๐‘š + ๐‘›๐‘˜).
Suppose that ๐‘ ∈ โ„ค is a prime (Example 2.4.18) that does not divide ๐‘Ž. We show
that gcd(๐‘Ž, ๐‘) = 1. Take ๐‘‘ > 0 and assume ๐‘‘ โˆฃ ๐‘Ž and ๐‘‘ โˆฃ ๐‘. Since ๐‘ is prime, ๐‘‘ = 1 or
๐‘‘ = ๐‘. Since ๐‘ - ๐‘Ž, we conclude that ๐‘‘ must equal 1, which means gcd(๐‘Ž, ๐‘) = 1. Use
this to prove the next result attributed to Euclid (Elements VII.30).
THEOREM 5.4.12 [Euclid’s Lemma]
An integer ๐‘ > 1 is prime if and only if ๐‘ โˆฃ ๐‘Ž๐‘ implies ๐‘ โˆฃ ๐‘Ž or ๐‘ โˆฃ ๐‘ for all
๐‘Ž, ๐‘ ∈ โ„ค.
PROOF
โˆ™ Let ๐‘ be prime. Suppose ๐‘ โˆฃ ๐‘Ž๐‘ but ๐‘ - ๐‘Ž. Then, gcd(๐‘Ž, ๐‘) = 1. Therefore, ๐‘ โˆฃ ๐‘
by Theorem 5.4.11.
โˆ™ Let ๐‘ > 1. Suppose ๐‘ satisfies the condition,
∀๐‘Ž∀๐‘(๐‘ โˆฃ ๐‘Ž๐‘ → ๐‘ โˆฃ ๐‘Ž ∨ ๐‘ โˆฃ ๐‘).
Assume ๐‘ is not prime. This means that there are integers ๐‘ and ๐‘‘ so that ๐‘ = ๐‘๐‘‘
with 1 < ๐‘ ≤ ๐‘‘ < ๐‘. Hence, ๐‘ โˆฃ ๐‘๐‘‘. By hypothesis, ๐‘ โˆฃ ๐‘ or ๐‘ โˆฃ ๐‘‘. However,
since ๐‘, ๐‘‘ < ๐‘, ๐‘ can divide neither ๐‘ nor ๐‘‘. This is a contradiction. Hence, ๐‘
must be prime.
Since 6 divides 3 ⋅ 4 but 6 - 3 and 6 - 4, the lemma tells us that 6 is not prime. On
the other hand, if ๐‘ is a prime that divides 12, then ๐‘ divides 4 or 3. This means that
๐‘ = 2 or ๐‘ = 3.
The next theorem is a generalization of Euclid’s lemma. Its proof uses mathematical
induction.
THEOREM 5.4.13
Let ๐‘ be prime and ๐‘Ž๐‘– ∈ โ„ค for ๐‘– = 0, 1, … , ๐‘› − 1. If ๐‘ โˆฃ ๐‘Ž0 ๐‘Ž1 · · · ๐‘Ž๐‘›−1 , then ๐‘ โˆฃ ๐‘Ž๐‘—
for some ๐‘— = 0, 1, … , ๐‘› − 1.
PROOF
โˆ™ The case when ๐‘› = 1 is trivial because ๐‘ divides ๐‘Ž0 by definition of the product.
โˆ™ Assume if ๐‘ โˆฃ ๐‘Ž0 ๐‘Ž1 · · · ๐‘Ž๐‘›−1 , then ๐‘ โˆฃ ๐‘Ž๐‘— for some ๐‘— = 0, 1, … , ๐‘› − 1. Suppose
๐‘ โˆฃ ๐‘Ž0 ๐‘Ž1 · · · ๐‘Ž๐‘› . Then, by Lemma 5.4.12,
๐‘ โˆฃ ๐‘Ž0 ๐‘Ž1 · · · ๐‘Ž๐‘›−1 or ๐‘ โˆฃ ๐‘Ž๐‘› .
266
Chapter 5 AXIOMATIC SET THEORY
If ๐‘ โˆฃ ๐‘Ž๐‘› , we are done. Otherwise, ๐‘ divides ๐‘Ž0 ๐‘Ž1 · · · ๐‘Ž๐‘›−1 . Hence, ๐‘ divides one
of the ๐‘Ž๐‘– by induction.
Exercises
1. Let ๐‘› ∈ โ„ค+ . Prove.
๐‘›(๐‘› + 1)
2
(b) 1 + 3 + 5 + · · · + (2๐‘› − 1) = ๐‘›2
(a) 1 + 2 + 3 + · · · + ๐‘› =
๐‘›(2๐‘› − 1)(2๐‘› + 1)
(c) 12 + 32 + 52 + · · · + (2๐‘› − 1)2 =
3
[
]2
๐‘›(๐‘› + 1)
3
3
3
3
(d) 1 + 2 + 3 + · · · + ๐‘› =
2
1 − ๐‘Ÿ๐‘›+1
(๐‘Ÿ ≠ 1)
1−๐‘Ÿ
(f) 1 ⋅ 1! + 2 ⋅ 2! + · · · + ๐‘› ⋅ ๐‘›! = (๐‘› + 1)! − 1
2
๐‘›
1
1
(g)
+
+···+
=1−
2! 3!
(๐‘› + 1)!
(๐‘› + 1)!
(2๐‘›)!
(h) 2 ⋅ 6 ⋅ 10 ⋅ 14 ⋅ · · · ⋅ (4๐‘› − 2) =
๐‘›!
2. Prove for all positive integers ๐‘›.
๐‘›
∑
๐‘›(๐‘› + 1)(๐‘› + 2)
(a)
๐‘–(๐‘– + 1) =
3
๐‘–=1
(e) 1 + ๐‘Ÿ + ๐‘Ÿ2 + · · · + ๐‘Ÿ๐‘› =
(b)
๐‘›
∑
๐‘–=1
1
๐‘›
=
(2๐‘– − 1)(2๐‘– + 1) 2๐‘› + 1
3. Let ๐‘› ∈ โ„ค+ . Prove.
(a) ๐‘› < 2๐‘›
(b) ๐‘›! ≤ ๐‘›๐‘›
๐‘›
∑
1
1
(c)
≤2−
2
๐‘›
๐‘–
๐‘–=1
(d)
1
2
3
๐‘›
๐‘›
+
+
+···+ ๐‘› ≤2− ๐‘›
2 22 23
2
2
4. Let ๐‘› ∈ โ„ค. Prove.
(a) ๐‘›2 < 2๐‘› for all ๐‘› ≥ 5.
(b) 2๐‘› < ๐‘›! for all ๐‘› ≥ 4.
(c) ๐‘›2 < ๐‘›! for all ๐‘› ≥ 4.
5. For ๐‘› ∈ โ„ค+ , prove that if ๐ด has ๐‘› elements, P(๐ด) has 2๐‘› elements.
6. Prove that the number of lines in a truth table with ๐‘› propositional variables is 2๐‘› .
7. Demonstrate Pascal’s identity (Lemma 5.4.9).
Section 5.4 MATHEMATICAL INDUCTION
267
8. For all integers ๐‘› ≥ ๐‘Ÿ ≥ 0, prove the given equations.
( ) ( )
๐‘›
๐‘›
(a)
= 1.
=
๐‘›
0
)
( ) (
๐‘›
๐‘›
(b)
.
=
๐‘›−๐‘Ÿ
๐‘Ÿ
9. Let ๐‘›, ๐‘Ÿ ∈ โ„ค+ with ๐‘› ≥ ๐‘Ÿ. Prove.
( ) (
)
( ) (
)
๐‘Ÿ
๐‘Ÿ+1
๐‘›
๐‘›+1
(a)
+
+···+
=
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ+1
(
)
2๐‘› + 1
2
2
2
2
(b) 1 + 3 + 5 + · · · + (2๐‘› − 1) =
3
10. Let ๐‘› ≥ 2 be an integer and prove the given equations.
๐‘› ( )
∑
๐‘›
(a)
๐‘Ÿ
= ๐‘›2๐‘›−1
๐‘Ÿ
๐‘Ÿ=1
( )
๐‘›
∑
๐‘›
๐‘Ÿ−1
(b)
(−1) ๐‘Ÿ
=0
๐‘Ÿ
๐‘Ÿ=1
11. Let ๐‘› and ๐‘Ÿ be positive integers and ๐‘› ≥ ๐‘Ÿ. Use induction to show the given
equations.
( ) (
)
( ) (
)
๐‘Ÿ
๐‘Ÿ+1
๐‘›
๐‘›+1
(a)
+
+···+
=
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ+1
(
)
2๐‘› + 1
2
2
2
2
(b) 1 + 3 + 5 + · · · + (2๐‘› − 1) =
3
12. Prove the following for all ๐‘› ∈ ๐œ”.
(a) 5 โˆฃ ๐‘›5 − ๐‘›
(b) 9 โˆฃ ๐‘›3 + (๐‘› + 1)3 + (๐‘› + 2)3
(c) 8 โˆฃ 52๐‘› + 7
(d) 5 โˆฃ 33๐‘›+1 + 2๐‘›+1
13. If ๐‘ is prime and ๐‘Ž and ๐‘ are positive integers such that ๐‘Ž + ๐‘ = ๐‘, prove that
gcd(๐‘Ž, ๐‘) = 1.
14. Prove for all ๐‘› ∈ โ„ค+ , there exist ๐‘› consecutive composite integers (Example 2.4.18)
by showing that (๐‘› + 1)! + 2, (๐‘› + 1)! + 3, … , (๐‘› + 1)! + ๐‘› + 1 are composite.
15. Prove.
(a) If ๐‘Ž ≠ 0, then ๐‘Ž ⋅ gcd(๐‘, ๐‘) = gcd(๐‘Ž๐‘, ๐‘Ž๐‘).
(b) Prove if gcd(๐‘Ž๐‘– , ๐‘) = 1 for ๐‘– = 1, … , ๐‘›, then gcd(๐‘Ž1 ⋅ ๐‘Ž2 · · · ๐‘Ž๐‘› , ๐‘) = 1).
16. For ๐‘˜ ∈ โ„ค+ , let ๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘˜−1 ∈ ๐œ”, not all equal to zero. Define
๐‘” = gcd(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘˜−1 )
to mean that ๐‘” is the greatest integer such that ๐‘” โˆฃ ๐‘Ž๐‘– for all ๐‘– = 0, 1, … , ๐‘˜−1. Assuming
๐‘˜ ≥ 3, prove the given equations.
268
Chapter 5 AXIOMATIC SET THEORY
(a) gcd(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘˜−1 ) = gcd(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘˜−3 , gcd(๐‘Ž๐‘˜−2 , ๐‘Ž๐‘˜−1 ))
(b) gcd(๐‘๐‘Ž0 , ๐‘๐‘Ž1 , … , ๐‘๐‘Ž๐‘˜−1 ) = ๐‘ gcd(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘˜−1 ) for all integers ๐‘ ≠ 0
5.5
STRONG INDUCTION
Suppose we want to find an equation for the terms of a sequence defined recursively in
which each term is based on two or more previous terms. To prove that such an equation
is correct, we modify mathematical induction. Remember the domino picture that we
used to explain how mathematical induction works (page 258). The first domino is
tipped causing the second to fall, which in turn causes the third to fall. By the time the
sequence of falls reaches the ๐‘› + 1 domino, ๐‘› dominoes have fallen. This means that
sentences ๐‘(1) through ๐‘(๐‘›) have been proved true. It is at this point that ๐‘(๐‘› + 1) is
proved. This is the intuition behind the next theorem. It is sometimes called strong
induction.
THEOREM 5.5.1 [Mathematical Induction 2]
Let ๐‘(๐‘˜) be a formula. For any ๐‘›0 ∈ โ„ค, if
๐‘(๐‘›0 ) ∧ ∀๐‘˜(๐‘˜ ∈ ๐œ” ∧ [∀๐‘™(๐‘™ ∈ ๐œ” ∧ [๐‘™ ≤ ๐‘˜ → ๐‘(๐‘›0 + ๐‘™)]) → ๐‘(๐‘›0 + ๐‘˜ + 1)]),
then
∀๐‘˜[๐‘˜ ∈ โ„ค ∧ ๐‘˜ ≥ ๐‘›0 → ๐‘(๐‘˜)].
PROOF
Assume ๐‘(๐‘›0 ) and
๐‘(๐‘›0 ) ∧ ๐‘(๐‘›0 + 1) ∧ · · · ∧ ๐‘(๐‘›0 + ๐‘˜) → ๐‘(๐‘›0 + ๐‘˜ + 1)
(5.11)
for ๐‘˜ ∈ ๐œ”. Define
๐‘ž(๐‘˜) := ๐‘(๐‘›0 ) ∧ ๐‘(๐‘›0 + 1) ∧ · · · ∧ ๐‘(๐‘›0 + ๐‘˜).
We proceed with the induction.
โˆ™ Since ๐‘(๐‘›0 ) holds, we have ๐‘ž(0).
โˆ™ Assume ๐‘ž(๐‘›) with ๐‘› ≥ 0. By definition of ๐‘ž(๐‘›), we have ๐‘(๐‘›0 ) through
๐‘(๐‘›0 + ๐‘›). Thus, ๐‘(๐‘›0 + ๐‘› + 1) by (5.11) from which ๐‘ž(๐‘› + 1) follows.
Therefore, ๐‘ž(๐‘›) is true for all ๐‘› ∈ ๐œ” (Theorem 5.4.1). Hence, ๐‘(๐‘›) for all integers
๐‘› ≥ ๐‘›0 .
Fibonacci Sequence
Leonardo of Pisa (known as Fibonacci) in his 1202 work Liber abaci posed a problem
about how a certain population of rabbits increases with time (Fibonacci and Sigler
2002). Each rabbit that is at least 2 months old is considered an adult. It is a young
Section 5.5 STRONG INDUCTION
269
rabbit if it is a month old. Otherwise, it is a baby. The rules that govern the population
are as follows:
โˆ™ No rabbits die.
โˆ™ The population starts with a pair of adult rabbits.
โˆ™ Each pair of adult rabbits will bear a new pair each month.
The population then grows according to the following table:
Month
1
2
3
4
5
6
Adult Pairs
1
1
2
3
5
8
Young Pairs
0
1
1
2
3
5
Baby Pairs
1
1
2
3
5
8
It appears that the number of adult (or baby) pairs at month ๐‘› is given by the sequence,
1, 1, 2, 3, 5, 8, 13, 21, 34, … .
This is known as the Fibonacci sequence, and each term of the sequence is called a
Fibonacci number. Let ๐น๐‘› denote the ๐‘›th term of the sequence. So
๐น1 = 1, ๐น2 = 1, ๐น3 = 2, ๐น4 = 3, ๐น5 = 5, ๐น6 = 8, … .
Each term of the sequence can be calculated recursively by
๐น1 = 1,
๐น2 = 1,
๐น๐‘› = ๐น๐‘›−1 + ๐น๐‘›−2 (๐‘› > 2).
(5.12)
Since we have only checked a few terms, we have not proved that ๐น๐‘› is equal to the
number of adult pairs in the ๐‘›th month. To show this, we use strong induction. Since
the recursive definition starts by explicitly defining ๐น1 and ๐น2 , the basis case for the
induction will prove that the formula holds for ๐‘› = 1 and ๐‘› = 2.
โˆ™ From the table, in each of the first 2 months, there is exactly one adult pair of
rabbits. This coincides with ๐น1 = 1 and ๐น2 = 1 in (5.12).
โˆ™ Let ๐‘› > 2, and assume that ๐น๐‘˜ equals the number of adult pairs in the ๐‘˜th month
for all ๐‘˜ ≤ ๐‘›. Because of the third rule, the number of pairs of adults in any
month is the same as the number of adult pairs in the previous month plus the
270
Chapter 5 AXIOMATIC SET THEORY
number of baby pairs 2 months prior. Therefore,
# of adult pairs
in month ๐‘› + 1
=
# of adult pairs
in month ๐‘›
+
# of young pairs
in month ๐‘›
=
# of adult pairs
in month ๐‘›
+
# of baby pairs
in month ๐‘› − 1
=
# of adult pairs
in month ๐‘›
+
# of adult pairs
in month ๐‘› − 1
= ๐น๐‘› + ๐น๐‘›−1
= ๐น๐‘›+1 .
It turns out that the Fibonacci sequence is closely related to another famous object
of study in the history of mathematics. Letting ๐‘› ≥ 1, define
๐‘Ž๐‘› =
๐น๐‘›+1
.
๐น๐‘›
The first seven terms of this sequence are
๐‘Ž1
๐‘Ž2
๐‘Ž3
๐‘Ž4
๐‘Ž5
๐‘Ž6
๐‘Ž7
= 1โˆ•1 = 1,
= 2โˆ•1 = 2,
= 3โˆ•2 = 1.5,
= 5โˆ•3 ≈ 1.667,
= 8โˆ•5 = 1.6,
= 13โˆ•8 = 1.625,
= 21โˆ•13 ≈ 1.615.
This sequence has a limit that we call ๐œ. To find this limit, notice that
๐น๐‘›+1
๐น + ๐น๐‘›−1
๐น
= ๐‘›
= 1 + ๐‘›−1 .
๐น๐‘›
๐น๐‘›
๐น๐‘›
Because ๐‘Ž๐‘›−1 = ๐น๐‘› โˆ•๐น๐‘›−1 when ๐‘› > 1,
๐‘Ž๐‘› = 1 +
and therefore,
๐‘Ž๐‘› − 1 −
1
,
๐‘Ž๐‘›−1
1
= 0.
๐‘Ž๐‘›−1
Because
lim ๐‘Ž
๐‘›→∞ ๐‘›
we conclude that
= lim ๐‘Ž๐‘›−1 = ๐œ,
๐‘›→∞
๐œ 2 − ๐œ − 1 = 0.
(5.13)
Section 5.5 STRONG INDUCTION
271
√
Therefore, (1 ± 5)โˆ•2 are the solutions to (5.13), but since ๐น๐‘›+1 โˆ•๐น๐‘› > 0, we take the
positive value and find that
√
1+ 5
๐œ=
.
2
The number ๐œ is called the golden ratio. It was considered by the ancient Greeks to
represent the ratio of the sides of the most beautiful rectangle.
EXAMPLE 5.5.2
Prove ๐น๐‘› ≤ ๐œ ๐‘›−1 when ๐‘› ≥ 2 using strong induction.
โˆ™ Since ๐น2 = 1 < ๐œ 1 ≈ 1.618, the inequality holds for ๐‘› = 2.
โˆ™ Let ๐‘› ≥ 3, and √
assume that ๐น๐‘˜ ≤ ๐œ ๐‘˜−1 for√all ๐‘˜ such that 2 ≤ ๐‘˜ ≤ ๐‘›.
−1
Because ๐œ = ( 5 − 1)โˆ•2 and ๐œ −2 = (3 − 5)โˆ•2,
๐œ −1 + ๐œ −2 = 1.
Therefore, the induction hypothesis gives
๐น๐‘›+1 = ๐น๐‘› + ๐น๐‘›−1 ≤ ๐œ ๐‘›−1 + ๐œ ๐‘›−2 = ๐œ ๐‘› (๐œ −1 + ๐œ −2 ) = ๐œ ๐‘› .
Unique Factorization
Theorem 5.4.13 states that if a prime divides an integer, it divides one of the factors of
the integer. It appears reasonable that any integer can then be written as a product that
includes all of its prime divisors. For example, we can write 126 = 2 ⋅ 3 ⋅ 3 ⋅ 7, and this
is essentially the only way in which we can write 126 as a product of primes. All of
this is summarized in the next theorem. It is also known as the fundamental theorem
of arithmetic. It is the reason the primes are important. They are the building blocks
of the integers.
THEOREM 5.5.3 [Unique Factorization]
If ๐‘› > 1, there exists a unique sequence of primes ๐‘0 ≤ ๐‘1 ≤ · · · ≤ ๐‘๐‘˜ (๐‘˜ ∈ ๐œ”)
such that ๐‘› = ๐‘0 ๐‘1 · · · ๐‘๐‘˜ .
PROOF
Prove existence with strong induction on ๐‘›.
โˆ™ When ๐‘› = 2, we are done since 2 is prime.
โˆ™ Assume that ๐‘˜ can be written as the product of primes as described above
for all ๐‘˜ such that 2 ≤ ๐‘˜ < ๐‘›. If ๐‘› is prime, we are done as in the basis
case. So suppose ๐‘› is composite. Then, there exist integers ๐‘Ž and ๐‘ such
that ๐‘› = ๐‘Ž๐‘ and 1 < ๐‘Ž ≤ ๐‘ < ๐‘›. By the induction hypothesis, we can write
๐‘Ž = ๐‘ž0 ๐‘ž1 · · · ๐‘ž๐‘ข
272
Chapter 5 AXIOMATIC SET THEORY
and
๐‘ = ๐‘Ÿ0 ๐‘Ÿ1 · · · ๐‘Ÿ๐‘ฃ ,
where the ๐‘ž๐‘– and ๐‘Ÿ๐‘— are primes. Now place these primes together in increasing order and relabel them as
๐‘0 ≤ ๐‘1 ≤ · · · ≤ ๐‘๐‘˜
with ๐‘˜ = ๐‘ข + ๐‘ฃ. Then, ๐‘› = ๐‘0 ๐‘1 · · · ๐‘๐‘˜ as desired.
For uniqueness, suppose that there are two sets of primes
๐‘0 ≤ ๐‘1 ≤ · · · ≤ ๐‘๐‘˜ and ๐‘ž0 ≤ ๐‘ž1 ≤ · · · ≤ ๐‘ž๐‘™
so that
๐‘› = ๐‘0 ๐‘1 · · · ๐‘๐‘˜ = ๐‘ž0 ๐‘ž1 · · · ๐‘ž๐‘™ .
By canceling, if necessary, we can assume the sides have no common primes. If
the cancellation yields 1 = 1, the sets of primes are the same. In order to obtain a
contradiction, assume that there is at least one prime remaining on the left-hand
side. Suppose it is ๐‘0 . If the product on the right equals 1, then ๐‘0 โˆฃ 1, which is
impossible. If there are primes remaining on the right, ๐‘0 divides one of them
by Lemma 5.4.12. This is also a contradiction, since the sides have no common
prime factors because of the cancellation. Hence, the two sequences must be the
same.
Unique Factorization allows us to make the following definition.
DEFINITION 5.5.4
Let ๐‘› ∈ โ„ค+ . If ๐‘0 , ๐‘1 , … , ๐‘๐‘˜−1 are distinct primes and ๐‘Ÿ0 , ๐‘Ÿ1 , … , ๐‘Ÿ๐‘˜−1 are natural
numbers such that
๐‘Ÿ๐‘˜−1
๐‘Ÿ ๐‘Ÿ
๐‘› = ๐‘00 ๐‘11 · · · ๐‘๐‘˜−1
,
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ
๐‘˜−1
is called a prime power decomposition of ๐‘›.
then ๐‘00 ๐‘11 · · · ๐‘๐‘˜−1
EXAMPLE 5.5.5
Consider the integer 360. It has 23 ⋅32 ⋅51 as a prime power decomposition. If the
exponents are limited to positive integers, the expression is unique. In this sense,
we can say that 23 ⋅ 32 ⋅ 51 is the prime power decomposition of 360. However,
there are times when primes need to be included in the product that are not factors
of the integer. By setting the exponent to zero, these primes can be included. For
example, we can also write 360 as 23 ⋅ 32 ⋅ 51 ⋅ 70 .
EXAMPLE 5.5.6
Suppose ๐‘› ∈ โ„ค such that ๐‘› > 1. Use unique factorization (Theorem 5.5.3)
to prove that ๐‘› is a perfect square if and only if all powers in a prime power
decomposition of ๐‘› are even.
Section 5.5 STRONG INDUCTION
273
โˆ™ Let ๐‘› be a perfect square. This means that ๐‘› = ๐‘˜2 for some integer ๐‘˜ > 1.
Write a prime power decomposition of ๐‘˜,
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ
๐‘™−1
๐‘˜ = ๐‘00 ๐‘11 · · · ๐‘๐‘™−1
.
Therefore,
2๐‘Ÿ
2๐‘Ÿ
2๐‘Ÿ
๐‘› = ๐‘˜2 = ๐‘0 0 ๐‘1 1 · · · ๐‘๐‘™−1๐‘™−1 .
โˆ™ Assume all the powers are even in a prime power decomposition of ๐‘›.
Namely,
๐‘Ÿ๐‘™−1
๐‘Ÿ ๐‘Ÿ
๐‘› = ๐‘00 ๐‘11 · · · ๐‘๐‘™−1
,
where there exists ๐‘ข๐‘– ∈ โ„ค so that ๐‘Ÿ๐‘– = 2๐‘ข๐‘– for ๐‘– = 0, 1, … , ๐‘™ − 1. Thus,
( ๐‘ข ๐‘ข
2๐‘ข
๐‘ข๐‘™−1 )2
2๐‘ข 2๐‘ข
,
๐‘› = ๐‘0 0 ๐‘1 1 · · · ๐‘๐‘™−1๐‘™−1 = ๐‘00 ๐‘11 · · · ๐‘๐‘™−1
a perfect square.
Exercises
1. Given each recursive definition, prove the formula for ๐‘Ž๐‘› holds for all positive integers ๐‘›.
(a) If ๐‘Ž1 = −1 and ๐‘Ž๐‘› = −๐‘Ž๐‘›−1 , then ๐‘Ž๐‘› = (−1)๐‘› .
(b) If ๐‘Ž1 = 1 and ๐‘Ž๐‘› = 1โˆ•3๐‘Ž๐‘›−1 , then ๐‘Ž๐‘› = (1โˆ•3)๐‘›−1 .
(c) If ๐‘Ž1 = 0, ๐‘Ž2 = −6, and ๐‘Ž๐‘› = 5๐‘Ž๐‘›−1 − 6๐‘Ž๐‘›−2 , then
๐‘Ž๐‘› = 3 ⋅ 2๐‘› − 2 ⋅ 3๐‘› .
(d) If ๐‘Ž1 = 4, ๐‘Ž2 = 12, and ๐‘Ž๐‘› = 4๐‘Ž๐‘›−1 − 2๐‘Ž๐‘›−2 , then
√
√
๐‘Ž๐‘› = (2 + 2)๐‘› + (2 − 2)๐‘› .
(e) If ๐‘Ž1 = 1, ๐‘Ž2 = 5, and ๐‘Ž๐‘›1 = ๐‘Ž๐‘› + 2๐‘Ž๐‘›−1 for all ๐‘› > 2, then
๐‘Ž๐‘› = 2๐‘› + (−1)๐‘› .
(f) If ๐‘Ž1 = 3, ๐‘Ž2 = −3, ๐‘Ž3 = 9, and ๐‘Ž๐‘› = ๐‘Ž๐‘›−1 + 4๐‘Ž๐‘›−2 − 4๐‘Ž๐‘›−3 , then
๐‘Ž๐‘› = 1 − (−2)๐‘› .
(g) If ๐‘Ž1 = 3, ๐‘Ž2 = 10, ๐‘Ž3 = 21, and ๐‘Ž๐‘› = 3๐‘Ž๐‘›−1 − 3๐‘Ž๐‘›−2 + ๐‘Ž๐‘›−3 , then
๐‘Ž๐‘› = ๐‘› + 2๐‘›2 .
2. Let ๐‘”1 = ๐‘Ž, ๐‘”2 = ๐‘, and ๐‘”๐‘› = ๐‘”๐‘›−1 + ๐‘”๐‘›−2 for all ๐‘› > 2. This sequence is called the
generalized Fibonacci sequence. Show that ๐‘”๐‘› = ๐‘Ž๐‘“๐‘›−2 + ๐‘๐‘“๐‘›−1 for all ๐‘› > 2.
274
Chapter 5 AXIOMATIC SET THEORY
3. Let ๐‘› > 0 be an integer. Prove.
(a) ๐น๐‘›+2 > ๐œ ๐‘›
๐‘›
∑
(b)
๐น๐‘– = ๐น๐‘›+2 − 1
๐‘–=1
4. Prove that Theorem 5.5.1 implies Theorem 5.4.1.
√
๐œ ๐‘› − ๐œŽ๐‘›
5. Let ๐œŽ = (1 − 5)โˆ•2 and demonstrate that ๐น๐‘› = √ .
5
6. Let ๐‘› ≥ 1 and ๐‘Ž ∈ โ„ค. Prove.
(a) ๐‘Ž๐‘›+1 − 1 = (๐‘Ž + 1)(๐‘Ž๐‘› − 1) − ๐‘Ž(๐‘Ž๐‘›−1 − 1).
(b) ๐‘Ž๐‘› − 1 = (๐‘Ž − 1)(๐‘Ž๐‘›−1 + ๐‘Ž๐‘›−2 + · · · + ๐‘Ž + 1).
7. For all ๐‘› ∈ ๐œ”, prove that 12 divides ๐‘›4 − ๐‘›2 (Definition 2.4.2).
8. Assume ๐‘’ โˆฃ ๐‘Ž and ๐‘’ โˆฃ ๐‘. Write prime power decompositions for ๐‘Ž and ๐‘:
๐‘Ÿ
๐‘Ÿ
๐‘Ÿ
๐‘ 
๐‘ 
๐‘ 
๐‘˜−1
๐‘Ž = ๐‘00 ๐‘11 · · · ๐‘๐‘˜−1
and
๐‘˜−1
.
๐‘ = ๐‘00 ๐‘11 · · · ๐‘๐‘˜−1
Prove that there exist ๐‘ก0 , ๐‘ก1 , … ๐‘ก๐‘˜−1 ∈ ๐œ” such that
๐‘ก
๐‘ก
๐‘ก
๐‘˜−1
,
๐‘’ = ๐‘00 ๐‘11 · · · ๐‘๐‘˜−1
๐‘ก๐‘– ≤ ๐‘Ÿ๐‘– , and ๐‘ก๐‘– ≤ ๐‘ ๐‘– for all ๐‘– = 0, 1, … , ๐‘˜ − 1.
9. Prove that ๐‘Ž3 โˆฃ ๐‘2 implies ๐‘Ž โˆฃ ๐‘ for all ๐‘Ž, ๐‘ ∈ โ„ค.
10. Let ๐‘Ž ∈ โ„ค+ . Let ๐‘Ž have the property that for all primes ๐‘, if ๐‘ โˆฃ ๐‘Ž, then ๐‘2 โˆฃ ๐‘Ž.
Prove that ๐‘Ž is the product of a perfect square and a perfect cube.
11. Prove that gcd(๐น๐‘› , ๐น๐‘›+2 ) = 1 for all ๐‘› ∈ โ„ค+ .
5.6
REAL NUMBERS
As โ„ค is defined using ๐œ” and โ„š is defined using โ„ค, the set analog to ๐‘ is defined using
โ„š. We start with a definition.
DEFINITION 5.6.1
Let (๐ด, 4) be a poset. The set ๐ต is an initial segment of ๐ด when ๐ต ⊆ ๐ด and
for all ๐‘Ž, ๐‘ ∈ ๐ด, if ๐‘Ž 4 ๐‘ and ๐‘ ∈ ๐ต, then ๐‘Ž ∈ ๐ต.
(5.14)
An initial segment ๐ต of ๐ด is proper if ๐ต ≠ ๐ด.
The condition (5.14) is called downward closed. Notice that for all ๐‘Ž ∈ ๐‘, both
(−∞, ๐‘Ž) and (−∞, ๐‘Ž] are initial segments of (๐‘, ≤). A poset is an initial segment of
itself, but it is not proper.
Section 5.6 REAL NUMBERS
275
DEFINITION 5.6.2
Let (๐ด, 4) be a poset with ๐‘ ∈ ๐ด. Define
๐—Œ๐–พ๐—€4 (๐ด, ๐‘) = {๐‘Ž ∈ ๐ด : ๐‘Ž โ‰บ ๐‘}.
For example,
๐—Œ๐–พ๐—€≤ (๐‘, 5) = (−∞, 5)
in (๐‘, ≤), and
๐—Œ๐–พ๐—€≤ (โ„ค, 5) = {… , 0, 1, 2, 3, 4}
in (โ„ค, ≤). Both of these are proper initial segments. Notice that every initial segment
of โ„ค is of the form ๐—Œ๐–พ๐—€≤ (โ„ค, ๐‘›) for some ๐‘› ∈ โ„ค.
Neither โ„ค nor โ„š are well-ordered by ≤. If a poset is well-ordered, its initial segments
have a particular form.
LEMMA 5.6.3
If (๐ด, 4) is a well-ordered set and ๐ต is a proper initial segment of ๐ด, there exists
a unique ๐‘š ∈ ๐ด such that ๐ต = ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š).
PROOF
Suppose that (๐ด, 4) is well-ordered and ๐ต ⊆ ๐ด is a proper initial segment. First,
note that ๐ด โงต ๐ต is not empty since ๐ต is proper. Thus, ๐ด โงต ๐ต contains a least
element ๐‘š because 4 well-orders ๐ด.
โˆ™ Let ๐‘Ž ∈ ๐ต, which implies that ๐‘Ž โ‰บ ๐‘š. Otherwise, ๐‘š would be an element of
๐ต because ๐ต is downward closed. Hence, ๐‘Ž ∈ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š), which implies
that ๐ต ⊆ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š).
โˆ™ Conversely, take ๐‘Ž ∈ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š), which means ๐‘Ž โ‰บ ๐‘š. If ๐‘Ž ∈ ๐ด โงต ๐ต, then
๐‘š 4 ๐‘Ž because ๐‘š is the least element of ๐ด โงต ๐ต, so ๐‘Ž must be an element of
๐ต by the trichotomy law (Theorem 4.3.21). Thus, ๐—Œ๐–พ๐—€4 (๐ด, ๐ต) ⊆ ๐ต.
To prove uniqueness, let ๐‘š′ ∈ ๐ด such that ๐‘š ≠ ๐‘š′ and ๐ต = ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š′ ). If
๐‘š โ‰บ ๐‘š′ , then ๐‘š ∈ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š′ ), and if ๐‘š′ โ‰บ ๐‘š, then ๐‘š′ ∈ ๐ต = ๐—Œ๐–พ๐—€4 (๐ด, ๐‘š). Both
cases are impossible.
Dedekind Cuts
A basic property of ๐‘ is that it is complete. This means that
every nonempty set of real numbers with an upper bound
has a real least upper bound
(Definition 4.3.12). For example, the set
{
}
1
๐ด = 1 − : ๐‘› ∈ ๐™+
๐‘›
276
Chapter 5 AXIOMATIC SET THEORY
is bounded from above, and its least upper bound is 1. Also,
๐ต = {3, 3.1, 3.14, 3.141, 3.1415, 3.14159, … }
is bounded from above, and its least upper bound is ๐œ‹. Observe that both ๐ด and ๐ต are
sets of rational numbers. The set ๐ด has a rational least upper bound, but ๐ต does not.
This shows that the rational numbers are not complete. Intuitively, the picture is that
of a number line. If ๐‘ is graphed, there are no holes because of completeness, but if
๐ is graphed, there are holes. These holes represent the irrational numbers that when
filled, complete the rational numbers resulting in the set of reals. We use this idea to
construct a model of the real numbers from โ„š (Dedekind 1901).
DEFINITION 5.6.4
A set x of rational numbers is a Dedekind cut (or a real number) if x is a subset
of (โ„š, ≤) such that
โˆ™ x is nonempty,
โˆ™ x is a proper initial segment of โ„š ,
โˆ™ x does not have a greatest element.
Denote the set of Dedekind cuts by โ„.
By Lemma 5.6.3, some Dedekind cuts are of the form ๐—Œ๐–พ๐—€≤ (โ„š, ๐‘Ž) for some ๐‘Ž ∈ โ„š. In
this case, write a = ๐—Œ๐–พ๐—€≤ (โ„š, ๐‘Ž). Therefore, โ„š can be embedded in โ„ using the function
๐‘“ : โ„š → โ„ defined by ๐‘“ (๐‘Ž) = a (Exercise 14). The elements of โ„ โงต ran(๐‘“ ) are the
irrational numbers.
EXAMPLE 5.6.5
โˆ™ The Dedekind cut that corresponds to the integer 7 is
7 = ๐—Œ๐–พ๐—€≤ (โ„š, 7).
Note that there is no gap between 7 and โ„š โงต 7 = {๐‘Ž ∈ โ„š : ๐‘Ž ≥ 7}.
โˆ™ The Dedekind cut x that corresponds to ๐œ‹ includes
{3, 3.1, 3.14, 3.141, 3.1415, 3.14159, … }
as a subset. Notice that ๐œ‹ ∉ x and ๐œ‹ ∉ โ„š โงต x because ๐œ‹ is not rational. This
means that we imagine a gap between x and โ„šโงตx. This gap is where ๐œ‹ is located.
Imagine that only the rational numbers have been placed on the number line (Section 5.3). For every Dedekind cut x, call the point on the number line where x and
โ„š โงต x meet a cut. Following our intuition, if the least upper bound of x is an element of
โ„š โงต x, there is a point at the cut [Figure 5.1(a)]. This means that x represents a rational
number, a point already on the number line. However, if the least upper bound of x is
Section 5.6 REAL NUMBERS
x
โ„š๎œx
โ„š
x
277
โ„š๎œx
โ„š
The cut
The cut
(a) A rational number
Figure 5.1
(b) An irrational number
The cuts of two types of numbers.
not an element of โ„š โงต x, there is no point at the cut [Figure 5.1(b)]. This means that
x represents an irrational number. To obtain all real numbers, a point must be placed
at each cut without a point, filling the entire number line. Therefore, the first step in
showing that โ„ is a suitable model for ๐‘ is to prove that every set of Dedekind cuts
with an upper bound must must have a least upper bound that is a Dedekind cut. That
is, โ„ must be shown to be complete. To accomplish this, we first define on order on โ„.
DEFINITION 5.6.6
Let x, y ∈ โ„. Define x ≤ y if and only if x ⊆ y, and define x < y to mean x ≤ y
and x ≠ y.
For example, 3 ≤ 4 in โ„ because 3 = ๐—Œ๐–พ๐—€≤ (โ„š, 3) and 4 = ๐—Œ๐–พ๐—€≤ (โ„š, 4). Because the order
on โ„š is linear (Theorem 5.3.8), it is left to Exercise 5 to prove that we have defined a
linear order on โ„.
THEOREM 5.6.7
(โ„, ≤) is a linear order but it is not a well-order.
Now to prove that โ„ is complete using the order of Definition 5.6.6.
THEOREM 5.6.8
Every nonempty subset of โ„ with an upper bound has a real least upper bound.
PROOF
Let โ„ฑ ≠ ∅ and โ„ฑ ⊆ โ„. Let m ∈ โ„ be an upper bound
โ‹ƒ
โ‹ƒ of โ„ฑ . By Example 4.3.14,
โ„ฑ is the least upper bound of โ„ฑ . We show that โ„ฑ ∈ โ„.
โˆ™ Takeโ‹ƒx ∈ โ„ฑ . Since Dedekind cuts are nonempty, x ≠ ∅, which implies
that โ„ฑ ≠ ∅.
โˆ™ By hypothesis, x ⊆ m for all x ∈ โ„ฑ . Hence,
โ‹ƒ
have that m ≠ โ„š, so โ„ฑ ⊂ โ„š.
โ‹ƒ
โ„ฑ ⊆ m. Since m ∈ โ„, we
โ‹ƒ
โˆ™ Let ๐‘ฅ ∈
โ„ฑ and ๐‘ฆ ≤ ๐‘ฅ. Thus, ๐‘ฅ ∈ a for
โ‹ƒ some Dedekind cut a ∈ โ„ฑ .
Sinceโ‹ƒ
a is downward closed, ๐‘ฆ ∈ a, so ๐‘ฆ ∈ โ„ฑ . Hence, with the previous
part, โ„ฑ is a proper initial segment of โ„š.
278
Chapter 5 AXIOMATIC SET THEORY
โˆ™ Let ๐‘ฅ ∈ a ∈ โ„ฑ . Since a is a Dedekind cut, it has no greatest element.
โ‹ƒ
Thus,
there exists ๐‘ฆ ∈ a such that ๐‘ฅ < ๐‘ฆ. Because ๐‘ฆ ∈ โ„ฑ , we see that
โ‹ƒ
โ„ฑ has no greatest element.
Since every nonempty bounded subset of real numbers has a least upper bound in โ„, the
set of Dedekind cuts does not have the same issue with gaps as โ„š does. For example,
the least upper bound of the set of real numbers
{2, 2.7, 2.71, 2.718, 2.7182, 2.71828, … }
is the Dedekind cut that corresponds to ๐‘’ ∈ ๐‘.
Arithmetic
As with the other sets of numbers that have been constructed using the axioms of ๐—ญ๐—™๐—–,
we now define addition and multiplication on โ„. First, define the following Dedekind
cuts:
โˆ™ 0 = ๐—Œ๐–พ๐—€≤ (โ„š, 0)
โˆ™ 1 = ๐—Œ๐–พ๐—€≤ (โ„š, 1).
Let x, y ∈ โ„. That
๐‘† = {๐‘Ž + ๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y}
is a Dedekind cut is Exercise 3. Assume that 0 ≤ x and 0 ≤ y. We claim that the set
๐‘ƒ1 = {๐‘Ž๐‘ : 0 ≤ ๐‘Ž ∈ x ∧ 0 ≤ ๐‘ ∈ y} ∪ 0
is a Dedekind cut.
โˆ™ ๐‘ƒ1 ≠ ∅ because 0 ≠ ∅.
โˆ™ Let ๐‘ข ∈ โ„š โงต x and ๐‘ฃ ∈ โ„š โงต y. Let
{
๐‘ข if ๐‘ข ≥ ๐‘ฃ,
๐‘š=
๐‘ฃ if ๐‘ฃ > ๐‘ข.
Then, ๐‘š2 ∉ ๐‘ƒ1 , so ๐‘ƒ1 ≠ โ„š.
โˆ™ Let 0 ≤ ๐‘Ž ∈ x and 0 ≤ ๐‘ ∈ y, and suppose that ๐‘ค ∈ โ„š such that ๐‘ค < ๐‘Ž๐‘. Hence,
๐‘ค๐‘−1 < ๐‘Ž, which implies that ๐‘ค๐‘−1 ∈ x since x is downward closed. Therefore,
๐‘ค ∈ ๐‘ƒ1 because
๐‘ค = (๐‘ค๐‘−1 )๐‘.
โˆ™ Again, let 0 ≤ ๐‘Ž ∈ x and 0 ≤ ๐‘ ∈ y. Since x and y do not have greatest
elements, there exists ๐‘ข ∈ x and ๐‘ฃ ∈ y such that ๐‘Ž < ๐‘ข and ๐‘ < ๐‘ฃ. Then, by
Exercise 5.3.15(b)
๐‘Ž๐‘ < ๐‘ข๐‘ฃ ∈ ๐‘ƒ1 .
279
Section 5.6 REAL NUMBERS
Furthermore, if x < 0 and y < 0, define
๐‘ƒ2 = {๐‘Ž๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y ∧ 0 ≤ −๐‘Ž ∧ 0 ≤ −๐‘} ∪ 0,
(5.15)
if x < 0 and 0 ≤ y, define
๐‘ƒ3 = {๐‘Ž๐‘ : ๐‘Ž ∈ x ∧ 0 ≤ ๐‘ ∈ y},
(5.16)
๐‘ƒ4 = {๐‘Ž๐‘ : 0 ≤ ๐‘Ž ∈ x ∧ ๐‘ ∈ y},
(5.17)
or 0 ≤ x and y < 0, define
Since ๐‘ƒ1 , ๐‘ƒ2 , ๐‘ƒ3 , ๐‘ƒ4 , and ๐‘† are Dedekind cuts (Exercise 4), we can use them to
define the two standard operations on โ„.
DEFINITION 5.6.9
Let x, y ∈ โ„. Define
x + y = {๐‘Ž + ๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y}
and
โŽง{๐‘Ž๐‘ : 0 ≤ ๐‘Ž ∈ x ∧ 0 ≤ ๐‘ ∈ y} ∪ 0
โŽช
โŽช{๐‘Ž๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y ∧ 0 ≤ −๐‘Ž ∧ 0 ≤ −๐‘} ∪ 0
x⋅y=โŽจ
โŽช{๐‘Ž๐‘ : ๐‘Ž ∈ x ∧ 0 ≤ ๐‘ ∈ y}
โŽช{๐‘Ž๐‘ : 0 ≤ ๐‘Ž ∈ x ∧ ๐‘ ∈ y}
โŽฉ
if 0 ≤ x ∧ 0 ≤ y,
if x < 0 ∧ y < 0,
if x < 0 ∧ 0 ≤ y,
if 0 ≤ x ∧ y < 0.
Since addition and multiplication on โ„š are associative and commutative, addition
and multiplication are associative and commutative on โ„.
THEOREM 5.6.10
Addition and multiplication of real numbers are associative and commutative.
PROOF
Let x, y, z ∈ โ„. We prove that addition is associative, leaving the rest to Exercise 7.
x + (y + z) = x + {๐‘ + ๐‘ : ๐‘ ∈ y ∧ ๐‘ ∈ z}
= {๐‘Ž + ๐‘ฃ : ๐‘Ž ∈ x ∧ ๐‘ฃ ∈ {๐‘ + ๐‘ : ๐‘ ∈ y ∧ ๐‘ ∈ z}}
= {๐‘Ž + (๐‘ + ๐‘) : ๐‘Ž ∈ x ∧ (๐‘ ∈ y ∧ ๐‘ ∈ z)}
= {(๐‘Ž + ๐‘) + ๐‘ : (๐‘Ž ∈ x ∧ ๐‘ ∈ y) ∧ ๐‘ ∈ z}
= {๐‘ข + ๐‘ : ๐‘ข ∈ {๐‘Ž + ๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y} ∧ ๐‘ ∈ z}
= {๐‘Ž + ๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y} + z
= (x + y) + z.
280
Chapter 5 AXIOMATIC SET THEORY
The Dedekind cuts 0 and 1 behave as expected. For example,
0 + 4 = {๐‘Ž + ๐‘ : ๐‘Ž ∈ 0 ∧ ๐‘ ∈ 4} = ๐—Œ๐–พ๐—€≤ (โ„š, 4) = 4
and
1 ⋅ 4 = {๐‘Ž๐‘ : 0 ≤ ๐‘Ž ∈ 1 ∧ 0 ≤ ๐‘ ∈ 4} ∪ 0 = ๐—Œ๐–พ๐—€≤ (โ„š, 4) = 4.
These equations suggest the following.
THEOREM 5.6.11
โ„ has additive and multiplicative identities.
PROOF
Let x ∈ โ„. We first show that
x = {๐‘Ž + ๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ 0} = x + 0.
Take ๐‘ข ∈ x. Since x has no greatest element, there exists ๐‘ฃ ∈ x such that ๐‘ข < ๐‘ฃ.
Write ๐‘ข = ๐‘ฃ + (๐‘ข − ๐‘ฃ). Since ๐‘ข − ๐‘ฃ < 0, we have that ๐‘ข ∈ x + 0. Conversely, let
๐‘ ∈ 0. Since ๐‘ ∈ 0 implies that ๐‘ < 0, we have that ๐‘ข + ๐‘ < ๐‘ข (Exercise 5.3.16),
and because x is downward closed, ๐‘ข + ๐‘ ∈ x.
We next show that
x = x ⋅ 1.
We have two cases to consider.
โˆ™ Let 0 ≤ x. By Definition 5.6.9,
x ⋅ 1 = {๐‘Ž ⋅ ๐‘ : 0 ≤ ๐‘Ž ∈ x ∧ 0 ≤ ๐‘ ∈ 1} ∪ 0.
Let 0 ≤ ๐‘Ž ∈ x and 0 ≤ ๐‘ < 1. Then, ๐‘Ž๐‘ ≤ ๐‘Ž (Exercise 5.3.17), so ๐‘Ž๐‘ ∈ x
since x is downward closed. Conversely, take ๐‘Ž ∈ x. If ๐‘Ž < 0, then ๐‘Ž ∈ 0,
and if ๐‘Ž = 0, then ๐‘Ž = 0 ⋅ 0, so suppose ๐‘Ž > 0. Since x has no greatest
element, there exists ๐‘ข ∈ x such that ๐‘Ž < ๐‘ข. This implies that ๐‘Ž๐‘ข−1 < 1, so
๐‘Ž ∈ x ⋅ 1 because ๐‘Ž = ๐‘ข(๐‘Ž๐‘ข−1 ).
โˆ™ Let x < 0 and proceed like in the previous case.
We leave the proof of the last result to Exercise 10.
THEOREM 5.6.12
โˆ™ Every element of โ„ has an additive inverse.
โˆ™ Every nonzero element of โ„ has a multiplicative inverse.
โˆ™ The distributive law holds for โ„.
Complex Numbers
The last set of numbers that we define are the complex numbers.
Section 5.6 REAL NUMBERS
281
DEFINITION 5.6.13
Define โ„‚ = โ„ × โ„ to be the set of complex numbers. Denote (a, b) ∈ โ„‚ by
a + b๐‘–.
Observe that the standard embedding ๐‘“ : โ„ → โ„‚ defined by ๐‘“ (x) = (x, 0) allows us
to consider โ„ as a subset of โ„‚. We will not define an order on โ„‚ but will define the
standard two operations.
DEFINITION 5.6.14
Let a + b๐‘–, c + d๐‘– ∈ โ„‚.
โˆ™ (a + b๐‘–) + (c + d๐‘–) = (a + c) + (b + d)๐‘–.
โˆ™ (a + b๐‘–) ⋅ (c + d๐‘–) = (ac − bd) + (ad + bc)๐‘–.
We leave the proof of the following to Exercise 11.
THEOREM 5.6.15
โˆ™ ๐‘–2 = –1 + 0๐‘–.
โˆ™ Addition and multiplication are associative and commutative.
โˆ™ โ„‚ has additive and multiplicative identities.
โˆ™ Every element of โ„‚ has an additive inverse.
โˆ™ Every nonzero element of โ„‚ has a multiplicative inverse.
โˆ™ The distributive law holds in โ„‚.
Exercises
1. Find the given initial segments in the indicate posets.
(a) ๐—Œ๐–พ๐—€โˆฃ (โ„ค, 50) from Example 4.3.6
(b) ๐—Œ๐–พ๐—€4 ({0, 1}∗ , 101010) from Example 4.3.7
(c) ๐—Œ๐–พ๐—€⊆ (P(โ„ค), {1, 2, 3, 4}) from Example 4.3.9
2. Let (๐ด, 4) be a poset. Assume that ๐ต and ๐ถ are initial segments of ๐ด. Prove that
๐ต ∩ ๐ถ is an initial segment of ๐ด. Is ๐ต ∪ ๐ถ also an initial segment of ๐ด?
3. Let x, y ∈ โ„. Prove that ๐‘† = {๐‘Ž + ๐‘ : ๐‘Ž ∈ x ∧ ๐‘ ∈ y} is a Dedekind cut.
4. Prove that ๐‘ƒ2 (5.15), ๐‘ƒ3 (5.16), and ๐‘ƒ4 (5.17) are Dedekind cuts.
5. Prove Theorem 5.6.7.
6. Show that every Dedekind cut has an upper bound in โ„.
7. Finish the proof of Theorem 5.6.10.
282
Chapter 5 AXIOMATIC SET THEORY
8. Prove that 0 and 1 are Dedekind cuts.
9. Let ๐‘Ž, ๐‘ ∈ โ„ and ๐‘Ž๐‘ = 0. Prove that ๐‘Ž = 0 or ๐‘ = 0.
10. Prove Theorem 5.6.12.
11. Prove Theorem 5.6.15.
12. Prove that between any two real numbers is another real number.
13. Prove that between any two real numbers is a rational number.
14. Show that โ„š can be embedded in โ„ by proving that ๐‘“ : โ„š → โ„ defined by
๐‘“ (๐‘Ž) = ๐—Œ๐–พ๐—€≤ (โ„š, ๐‘Ž)
is an order isomorphism preserving ≤ with ≤. Furthermore, show that both ๐œ” and โ„ค
can be embedded into โ„.
15. Let ๐‘“ be the function defined in Exercise 14. Show that ๐‘“ (๐‘Ž + 1) is an upper bound
of ๐‘“ (๐‘Ž) for all ๐‘Ž ∈ โ„š.
16. The absolute value function (Exercise 2.4.23) can be defined so that for all x ∈ โ„,
|x| = x ∪ −x, where −x refers to the additive inverse of x (Theorem 5.6.12). Let a be
a positive real number. Prove the following for every x, y ∈ โ„.
(a) | − x| = |x|.
(b) |x2 | = |x|2 .
(c) x ≤ |x|.
(d) |xy| = |x| |y|.
(e) |x| < a if and only if −a < x < a.
(f) a < |x| if and only if a < x or x < −a.
CHAPTER 6
ORDINALS AND CARDINALS
6.1
ORDINAL NUMBERS
In Chapter 5, we defined certain sets to represent collections of numbers. Despite being
sets themselves, the elements of those sets were called numbers. We continue this association with sets as numbers but for a different purpose. While before we defined ๐œ”, โ„ค,
โ„š, โ„, and โ„‚ to represent ๐, ๐™, ๐, ๐‘, and ๐‚, the definitions of this chapter are intended
to be a means by which all sets can be classified according to a particular criterion.
Specifically, in the later part of the chapter, we will define sets for the purpose of identifying the size of a given set, and we begin the chapter by defining sets that are used
to identify whether two well-ordered sets have the same order type (Definition 4.5.24).
A crucial tool in this pursuit is the following generalization of Theorem 5.5.1 to wellordered infinite sets.
THEOREM 6.1.1 [Transfinite Induction 1]
Let (๐ด, 4) be a well-ordered set. If ๐ต ⊆ ๐ด and ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ) ⊆ ๐ต implies ๐‘ฅ ∈ ๐ต
for all ๐‘ฅ ∈ ๐ด, then ๐ด = ๐ต.
283
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
284
Chapter 6 ORDINALS AND CARDINALS
PROOF
To show that ๐ด is a subset of ๐ต, suppose that ๐ด โงต ๐ต is nonempty. Since ๐ด
is well-ordered by 4, let ๐‘š be the least element of ๐ด โงต ๐ต. This implies that
๐—Œ๐–พ๐—€4 (๐ด, ๐‘š) ⊆ ๐ต, so ๐‘š ∈ ๐ต by hypothesis, a contradiction.
Note that transfinite induction restricted to ๐œ” is simply strong induction (Theorem 5.5.1).
To see this, let the well-ordered set (๐ด, 4) of Theorem 6.1.1 be (๐œ”, ≤). Define the set
๐ต = {๐‘˜ : ๐‘(๐‘˜)} ⊆ ๐œ” for some formula ๐‘(๐‘˜). The conditional
๐—Œ๐–พ๐—€≤ (๐œ”, ๐‘›) ⊆ ๐ต → ๐‘› ∈ ๐ต
implies ๐‘(0) when ๐‘› = 0 because ๐—Œ๐–พ๐—€≤ (๐œ”, 0) = ∅ and implies
๐‘(0) ∧ ๐‘(1) ∧ · · · ∧ ๐‘(๐‘› − 1) → ๐‘(๐‘›)
when ๐‘› > 0 because ๐—Œ๐–พ๐—€≤ (๐œ”, ๐‘›) = ๐‘›.
Our first use of transfinite induction is the following lemma. It uses the terminology
of Exercise 4.4.32 and is the first of a sequence of lemmas that will play a critical role.
LEMMA 6.1.2
Let (๐ด, 4) be well-ordered. If ๐œ‘ : ๐ด → ๐ด is increasing, then ๐‘Ž 4 ๐œ‘(๐‘Ž) for all
๐‘Ž ∈ ๐ด.
PROOF
Define ๐ต = {๐‘ฅ ∈ ๐ด : ๐‘ฅ 4 ๐œ‘(๐‘ฅ)}, where ๐œ‘ is an increasing function ๐ด → ๐ด.
Let ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž) ⊆ ๐ต. We note that ๐‘Ž is the least element of ๐ด โงต ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž). Let
๐‘ฆ ∈ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž). This implies that ๐‘ฆ 4 ๐œ‘(๐‘ฆ) โ‰บ ๐œ‘(๐‘Ž) by definition of ๐ต and
because ๐‘ฆ โ‰บ ๐‘Ž. Hence, ๐œ‘(๐‘Ž) ∈ ๐ด โงต ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž). Thus, ๐‘Ž 4 ๐œ‘(๐‘Ž) and ๐ด = ๐ต by
transfinite induction (Theorem 6.1.1).
LEMMA 6.1.3
For all well-ordered sets (๐ด, 4) and (๐ด′ , 4′ ), there exists at most one order isomorphism ๐œ‘ : ๐ด → ๐ด′ .
PROOF
Let ๐œ‘ : ๐ด → ๐ด′ and ๐œ“ : ๐ด → ๐ด′ be order isomorphisms. Since both ๐œ‘−1 and
๐œ“ −1 are order isomorphisms ๐ด′ → ๐ด (Theorem 4.5.26), ๐œ“ −1 โ—ฆ ๐œ‘ and ๐œ‘−1 โ—ฆ ๐œ“ are
order isomorphisms ๐ด → ๐ด (Theorem 4.5.27). We note that for every ๐‘, ๐‘ ∈ ๐ด,
if ๐‘ โ‰บ ๐‘, then ๐œ‘(๐‘) โ‰บ′ ๐œ‘(๐‘) and then ๐œ“ −1 (๐œ‘(๐‘)) โ‰บ ๐œ“ −1 (๐œ‘(๐‘)). This means that
๐œ“ −1 โ—ฆ ๐œ‘ is increasing. A similar argument proves that ๐œ‘−1 โ—ฆ ๐œ“ is increasing. To
show that ๐œ‘ = ๐œ“, let ๐‘Ž ∈ ๐ด. By Lemma 6.1.2,
๐‘Ž 4 (๐œ“ −1 โ—ฆ ๐œ‘)(๐‘Ž)
and
๐‘Ž 4 (๐œ‘−1 โ—ฆ ๐œ“)(๐‘Ž).
Therefore, ๐œ“(๐‘Ž) 4′ ๐œ‘(๐‘Ž) and ๐œ‘(๐‘Ž) 4′ ๐œ“(๐‘Ž). Since 4′ is antisymmetric, we have
that ๐œ‘(๐‘Ž) = ๐œ“(๐‘Ž).
Section 6.1 ORDINAL NUMBERS
285
LEMMA 6.1.4
No well-ordered set (๐ด, 4) is order isomorphic to any of its proper initial segments.
PROOF
Let (๐ด, 4) be a well-ordered set. Suppose that ๐‘† is a proper initial segment
of ๐ด. In order to obtain a contradiction, assume that ๐œ‘ : ๐ด → ๐‘† is an order
isomorphism. Take ๐‘Ž ∈ ๐ด โงต ๐‘†. Since ๐œ‘(๐‘Ž) ∈ ๐‘† and ๐œ‘ is increasing, we have that
๐‘Ž 4 ๐œ‘(๐‘Ž) โ‰บ ๐‘Ž by Lemma 6.1.2.
The next result follows from Lemma 6.1.4 (Exercise 1).
LEMMA 6.1.5
Distinct initial segments of a well-ordered set are not order isomorphic.
The lemmas lead to the following theorem.
THEOREM 6.1.6
If (๐ด, 4) and (๐ต, 4′ ) are well-ordered sets, there exists an order isomorphism
such that exactly one of the following holds.
โˆ™ ๐ด ≅ ๐ต.
โˆ™ ๐ด is order isomorphic to a proper initial segment of ๐ต.
โˆ™ ๐ต is order isomorphic to a proper initial segment of ๐ด.
PROOF
Let (๐ด, 4) and (๐ต, 4′ ) be well-ordered sets. Appealing to Lemma 6.1.5, if ๐‘ฅ ∈ ๐ด,
there is at most one ๐‘ฆ ∈ ๐ต such that ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ), so define the
function
๐œ‘ = {(๐‘ฅ, ๐‘ฆ) ∈ ๐ด × ๐ต : ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ)}.
We have a number of facts to prove.
โˆ™ Let ๐‘ฆ1 , ๐‘ฆ2 ∈ ran(๐œ‘) such that ๐‘ฆ1 = ๐‘ฆ2 . Take ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐ด such that
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ1 ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ1 )
and
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ2 ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ2 ).
Then, we have ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ1 ) ≅ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ2 ), and ๐‘ฅ1 = ๐‘ฅ2 by Lemma 6.1.5.
Therefore, ๐œ‘ is one-to-one.
โˆ™ Take ๐‘ฅ1 , ๐‘ฅ2 ∈ dom(๐œ‘) and assume that ๐‘ฅ1 4 ๐‘ฅ2 . This implies that
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ1 ) ⊆ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ2 ).
286
Chapter 6 ORDINALS AND CARDINALS
Then, by definition of ๐œ‘, we have
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ1 ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐œ‘(๐‘ฅ1 ))
and
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ2 ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐œ‘(๐‘ฅ2 )).
Hence, ๐—Œ๐–พ๐—€4′ (๐ต, ๐œ‘(๐‘ฅ1 )) is order isomorphic to an initial segment ๐‘† of
๐—Œ๐–พ๐—€4′ (๐ต, ๐œ‘(๐‘ฅ2 )) (Exercise 17). If ๐‘† ≠ ๐—Œ๐–พ๐—€4′ (๐ต, ๐œ‘(๐‘ฅ1 )), then ๐ต has two
distinct isomorphic initial segments, contradicting Lemma 6.1.5. This implies that ๐œ‘(๐‘ฅ1 ) 4′ ๐œ‘(๐‘ฅ2 ), so ๐œ‘ is order-preserving.
โˆ™ Let ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐ด. Suppose that ๐‘ฅ1 4 ๐‘ฅ2 and ๐‘ฅ2 ∈ dom(๐œ‘). This means that
there exists ๐‘ฆ2 ∈ ๐ต such that
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ2 ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ2 ).
If ๐‘ฅ1 = ๐‘ฅ2 , then ๐‘ฅ1 ∈ dom(๐œ‘), so assume that ๐‘ฅ1 ≠ ๐‘ฅ2 . Since ๐‘ฅ1 โ‰บ ๐‘ฅ2 , we
have that ๐‘ฅ1 ∈ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ2 ). Because ๐œ‘ is order-preserving,
๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ1 ) ≅ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ1 )
for some ๐‘ฆ1 ∈ ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘ฆ2 ) (Exercise 17). Therefore, (๐‘ฅ1 , ๐‘ฆ1 ) ∈ ๐œ‘, so
๐‘ฅ1 ∈ dom(๐œ‘), proving that the domain of ๐œ‘ is an initial segment of ๐ด.
โˆ™ That the range of ๐œ‘ is an initial segment of ๐ต is proved like the previous
case.
If ๐œ‘ is a surjection and dom(๐œ‘) = ๐ด, then ๐œ‘ is an order isomorphism ๐ด → ๐ต,
else ๐œ‘−1 [๐ต] is a proper initial segment of ๐ด. If ๐œ‘ is not a surjection, ๐œ‘[๐ด] is a
proper initial segment of ๐ต.
Ordinals
Theorem 6.1.6 is a sort of trichotomy law for well-ordered sets. Two well-ordered sets
look alike, or one has a copy of itself in the other. This suggests that we should be
able to choose certain well-ordered sets to serve as representatives of all the different
types of well-ordered sets. No two of the chosen sets should be order isomorphic, but
it should be the case that every well-ordered set is order isomorphic to exactly one of
them. That is, we should be able to classify all of the well-ordered sets. This will be
our immediate goal and is the purpose behind the next definition.
DEFINITION 6.1.7
The set ๐›ผ is an ordinal number (or simply an ordinal) if (๐›ผ, ⊆) is a well-ordered
set and ๐›ฝ = ๐—Œ๐–พ๐—€⊆ (๐›ผ, ๐›ฝ) for all ๐›ฝ ∈ ๐›ผ. For ordinals, define
๐—Œ๐–พ๐—€(๐›ผ, ๐›ฝ) = ๐—Œ๐–พ๐—€⊆ (๐›ผ, ๐›ฝ).
Section 6.1 ORDINAL NUMBERS
287
Definition 6.1.7 implies that ๐œ” and every natural number is an ordinal because they are
well-ordered by ⊆ and for all ๐‘› ∈ ๐œ” โงต {0},
๐‘› = {0, 1, 2, … , ๐‘› − 1} = ๐—Œ๐–พ๐—€(๐œ”, ๐‘›),
and for all ๐‘˜ ∈ ๐‘›,
๐‘˜ = {0, 1, 2, … , ๐‘˜ − 1} = ๐—Œ๐–พ๐—€(๐‘›, ๐‘˜).
For example, 5 ∈ 7 and
5 = {0, 1, 2, 3, 4} = ๐—Œ๐–พ๐—€(7, 5).
We now prove a sequence of basic results about ordinals. The first is similar to
Theorem 5.2.8, so its proof is left to Exercise 5.
THEOREM 6.1.8
Ordinals are transitive sets.
THEOREM 6.1.9
The elements of ordinals are transitive sets.
PROOF
Let ๐›ผ be an ordinal and ๐›ฝ ∈ ๐›ผ. Take ๐›พ ∈ ๐›ฝ and ๐›ฟ ∈ ๐›พ. Since ๐›ผ is transitive
(Theorem 6.1.8), we have that ๐›พ ∈ ๐›ผ. Therefore, ๐›พ = ๐—Œ๐–พ๐—€(๐›ผ, ๐›พ) and ๐›ฝ = ๐—Œ๐–พ๐—€(๐›ผ, ๐›ฝ),
so
๐›ฟ ∈ ๐—Œ๐–พ๐—€(๐›ผ, ๐›พ) ⊆ ๐—Œ๐–พ๐—€(๐›ผ, ๐›ฝ) = ๐›ฝ,
which implies that ๐›ฝ is transitive (Definition 5.2.7).
THEOREM 6.1.10
Every element of an ordinal is an ordinal.
PROOF
Let ๐›ผ be an ordinal and ๐›ฝ ∈ ๐›ผ. Notice that this implies that ๐›ฝ is transitive (Theorem 6.1.9). Since ๐›ฝ ⊆ ๐›ผ, we have that (๐›ฝ, ⊆) is a well-ordered set by a subset
axiom (5.1.8) and Theorem 4.3.26. Now take ๐›ฟ ∈ ๐›ฝ. Since ๐›ฟ ∈ ๐›ผ, we have that
๐›ฟ is transitive. Therefore, by Exercise 5.2.3,
๐›ฟ = {๐›พ : ๐›พ ∈ ๐›ฟ}
= {๐›พ : ๐›พ ∈ ๐›ฝ ∧ ๐›พ ∈ ๐›ฟ}
⊆ {๐›พ : ๐›พ ∈ ๐›ฝ ∧ ๐›พ ⊂ ๐›ฟ}
= ๐—Œ๐–พ๐—€(๐›ฝ, ๐›ฟ)
⊆ ๐—Œ๐–พ๐—€(๐›ผ, ๐›ฟ)
= ๐›ฟ.
From this, we conclude that ๐›ฟ = ๐—Œ๐–พ๐—€(๐›ฝ, ๐›ฟ).
288
Chapter 6 ORDINALS AND CARDINALS
THEOREM 6.1.11
Let ๐›ผ and ๐›ฝ be ordinals. Then, ๐›ผ ⊂ ๐›ฝ if and only if ๐›ผ ∈ ๐›ฝ.
PROOF
If ๐›ผ ∈ ๐›ฝ, then ๐›ผ ⊂ ๐›ฝ because ๐›ฝ is transitive (Theorem 6.1.8 and Exercise 5.2.3).
Conversely, suppose that ๐›ผ ⊂ ๐›ฝ. Let ๐›พ ∈ ๐›ผ and ๐›ฟ ⊂ ๐›พ with ๐›ฟ ∈ ๐›ฝ. Since ๐›ฝ is an
ordinal, ๐›ฟ = ๐—Œ๐–พ๐—€(๐›ฝ, ๐›ฟ). Hence, ๐›ฟ = ๐—Œ๐–พ๐—€(๐›พ, ๐›ฟ), which implies that ๐›ฟ ∈ ๐›พ because
๐›พ is an ordinal by Theorem 6.1.10. Therefore, ๐›ฟ ∈ ๐›ผ because ๐›ผ is transitive
(Theorem 6.1.8). This shows that ๐›ผ is a proper initial segment of ๐›ฝ with respect
to ⊆ (Definition 5.6.1). From this, it follows by Lemma 5.6.3 that ๐›ผ = ๐—Œ๐–พ๐—€(๐›ฝ, ๐œ)
for some ๐œ ∈ ๐›ฝ. Hence, ๐›ผ ∈ ๐›ฝ.
THEOREM 6.1.12
Every ordinal is well-ordered by ∈.
The next theorem is an important part of the process of showing that the ordinals
are the sets that classify all well-ordered sets according to their order types. It states
that distinct ordinals are not order isomorphic with respect to ⊆.
THEOREM 6.1.13
For all ordinals ๐›ผ and ๐›ฝ, if (๐›ผ, ⊆) ≅ (๐›ฝ, ⊆), then ๐›ผ = ๐›ฝ.
PROOF
Let ๐œ‘ : ๐›ผ → ๐›ฝ be an order isomorphism preserving ⊆. Define
๐ด = {๐›พ ∈ ๐›ผ : ๐œ‘(๐›พ) = ๐›พ}.
Take ๐›ฟ ∈ ๐›ผ and assume that ๐—Œ๐–พ๐—€(๐›ผ, ๐›ฟ) ⊆ ๐ด. Then,
๐œ‘(๐›ฟ) = ๐—Œ๐–พ๐—€(๐›ฝ, ๐œ‘(๐›ฟ))
= ๐œ‘[๐—Œ๐–พ๐—€(๐›ผ, ๐›ฟ)]
= {๐œ‘(๐›พ) : ๐›พ ∈ ๐›ผ ∧ ๐›พ ⊂ ๐›ฟ}
= {๐›พ : ๐›พ ⊂ ๐›ฟ}
= ๐›ฟ.
The first equality follows because ๐œ‘(๐›ฟ) is an ordinal in ๐›ฝ, the second follows
because ๐œ‘ is an order isomorphism, and the fourth equation follows by the assumption. Therefore, by transfinite induction (Theorem 6.1.1), ๐ด = ๐›ผ, so ๐œ‘ is
the identity map and ๐›ผ = ๐›ฝ.
Because of Theorem 6.1.13, we are able to prove that there is a trichotomy law for
the ordinals with respect to ⊆.
THEOREM 6.1.14 [Trichotomy]
For all ordinals ๐›ผ and ๐›ฝ, exactly one of the following holds: ๐›ผ = ๐›ฝ, ๐›ผ ⊂ ๐›ฝ, or
๐›ผ ⊂ ๐›ฝ.
Section 6.1 ORDINAL NUMBERS
289
PROOF
Since (๐›ผ, ⊆) and (๐›ฝ, ⊆) are well-ordered, by Theorem 6.1.6, exactly one of the
following holds.
โˆ™ ๐›ผ ≅ ๐›ฝ, which implies that ๐›ผ = ๐›ฝ by Theorem 6.1.13.
โˆ™ There exists ๐›ฟ ∈ ๐›ฝ such that ๐›ผ ≅ ๐—Œ๐–พ๐—€(๐›ฝ, ๐›ฟ). Since ๐—Œ๐–พ๐—€(๐›ฝ, ๐›ฟ) is an ordinal
(Theorem 6.1.10), ๐›ผ = ๐—Œ๐–พ๐—€(๐›ฝ, ๐›ฟ), again by Theorem 6.1.13. Therefore,
๐›ผ ⊂ ๐›ฝ.
โˆ™ There exists ๐›พ ∈ ๐›ผ such that ๐›ฝ ≅ ๐—Œ๐–พ๐—€(๐›ผ, ๐›พ). As in the previous case, we
have that ๐›ฝ ⊂ ๐›ผ.
Because of Theorem 6.1.11, we can quickly conclude the following.
COROLLARY 6.1.15
For all ordinals ๐›ผ and ๐›ฝ, exactly one of the following holds: ๐›ผ = ๐›ฝ, ๐›ผ ∈ ๐›ฝ, or
๐›ผ ∈ ๐›ฝ.
In addition to the ordinals having a trichotomy law, the least upper bound with respect
to ⊆ of a set of ordinals is also an ordinal (compare Example 4.3.14).
THEOREM 6.1.16
If โ„ฑ is a set of ordinals,
โ‹ƒ
โ„ฑ is an ordinal.
PROOF
โ‹ƒ
We show that โ„ฑ satisfies
โ‹ƒ the conditions of Definition 6.1.7. Since the elements
of ordinals
are
ordinals,
โ„ฑ is a set of ordinals,โ‹ƒ
and by Theorem 6.1.14, we see
โ‹ƒ
that ( โ„ฑ , ⊆) is a linearly ordered set. Let ๐ต ⊆ โ„ฑ and take ๐›ผ ∈ ๐ต. We have
two cases to consider.
โˆ™ Suppose ๐›ผ ∩ ๐ต = ∅. Let ๐›ฝ ∈ ๐ต. Then, ๐›ฝ ∉ ๐›ผ, so by Theorem 6.1.11 and
Theorem 6.1.14, ๐›ผ ⊆ ๐›ฝ. Hence, ๐›ผ is the least element of ๐ต.
โˆ™ Let ๐›ผ ∩ ๐ต be nonempty. Since ๐›ผ is an ordinal, there exists an ordinal ๐›ฟ that
is the least element of ๐›ผ ∩ ๐ต with respect to ⊆. Let ๐›ฝ ∈ ๐ต. If ๐›ฝ ⊂ ๐›ผ, then
๐›ฝ ∈ ๐›ผ ∩ ๐ต, which implies that ๐›ฟ ⊆ ๐›ฝ. Also, if ๐›ผ ⊆ ๐›ฝ, then ๐›ฟ ⊂ ๐›ฝ. Since
these are the only two options (Theorem 6.1.14), this implies that ๐›ฟ is the
least element of ๐ต.
โ‹ƒ
We conclude that ( โ„ฑ , ⊆) is a well-ordered set.
โ‹ƒ
Next, let ๐›ฝ ∈
โ„ฑ. โ‹ƒ
This means that there exists an ordinal ๐›ผ ∈ ๐ด such
that ๐›ฝ ∈ ๐›ผ. Since ๐—Œ๐–พ๐—€( โ„ฑ , ๐›ฝ) ⊆ ๐›ฝ by definition, โ‹ƒ
take ๐›ฟ ∈ ๐›ฝ. Since ๐›ผ is
transitiveโ‹ƒ(Theorem 6.1.8), ๐›ฟ ∈ ๐›ผ. Therefore, ๐›ฟ ∈
โ„ฑ , which implies that
๐›ฝ ⊆ ๐—Œ๐–พ๐—€( โ„ฑ , ๐›ฝ).
290
Chapter 6 ORDINALS AND CARDINALS
Classification
Let ๐›ผ be an ordinal number. We check the two conditions of Definition 6.1.7 to show
that ๐›ผ + is an ordinal.
โˆ™ Let ๐ต be a nonempty subset of ๐›ผ ∪ {๐›ผ}. If ๐ต ∩ ๐›ผ ≠ ∅, then ๐ต has a least element
with respect to ⊆ since (๐›ผ, ⊆) is well-ordered. If ๐ต = {๐›ผ}, then ๐›ผ is the least
element of ๐ต.
โˆ™ Let ๐›ฝ ∈ ๐›ผ ∪ {๐›ผ}. If ๐›ฝ ∈ ๐›ผ, then ๐›ฝ = ๐—Œ๐–พ๐—€(๐›ผ, ๐›ฝ) = ๐—Œ๐–พ๐—€(๐›ผ + , ๐›ฝ) because ๐›ผ is an
ordinal number. Otherwise, ๐›ฝ = ๐›ผ = ๐—Œ๐–พ๐—€(๐›ผ + , ๐›ผ).
The ordinal ๐›ผ + is called a successor ordinal because it has a predecessor. For example,
every positive natural number is a successor ordinal.
Now assume that ๐›ผ ≠ ∅ and ๐›ผ is an ordinal that is not a successor.
โˆ™ Let ๐›ฟโ‹ƒ∈ ๐›ฝ ∈ ๐›ผ. Since ๐›ผ is transitive (Theorem 6.1.8), ๐›ฟ ∈ ๐›ผ. Thus, we conclude
that {๐›ฝ : ๐›ฝ ∈ ๐›ผ} ⊆ ๐›ผ.
โˆ™ Now take ๐›ฟ ∈ ๐›ผ. This implies that ๐›ฟ is an ordinal (Theorem 6.1.10), so ๐›ฟ ⊂ ๐›ผ by
Theorem 6.1.11. Therefore, ๐›ฟ + ⊆ ๐›ผ, so ๐›ฟ + ⊂ ๐›ผ since ๐›ผ is not a successor. Thus,
๐›ฟ + ∈โ‹ƒ๐›ผ, again by appealing to Theorem 6.1.11. Because ๐›ฟ ∈ ๐›ฟ + , we have that
๐›ผ ⊆ {๐›ฝ : ๐›ฝ ∈ ๐›ผ}.
We conclude that for every nonempty ordinal ๐›ผ that is not a successor,
โ‹ƒ
๐›ผ=
๐›ฝ.
๐›ฝ∈๐›ผ
Such an ordinal number is called a limit ordinal. For example, since every natural numโ‹ƒ
ber is an ordinal, ๐œ” = {๐‘› : ๐‘› ∈ ๐œ”} is a limit ordinal. Therefore, ๐œ”+ , ๐œ”++ , ๐œ”+++ , …
are also ordinals, but they are successors.
All of this proves the following.
THEOREM 6.1.17
A nonempty ordinal is either a successor or a limit ordinal.
Therefore, by Theorem 6.1.14 and Corollary 6.1.15, we can view the ordinals as sorted
by ⊂ giving
0 ⊂ 1 ⊂ 2 ⊂ · · · ⊂ ๐œ” ⊂ ๐œ”+ ⊂ ๐œ”++ ⊂ ๐œ”+++ ⊂ · · ·
and as sorted by ∈ giving
0 ∈ 1 ∈ 2 ∈ · · · ∈ ๐œ” ∈ ๐œ”+ ∈ ๐œ”++ ∈ ๐œ”+++ ∈ · · · .
Characterizing every ordinal as being equal to 0, a successor ordinal, or a limit
ordinal allows us to restate Theorem 6.1.1. The form of the theorem generalizes Theorem 5.4.1 to infinite ordinals. Its proof is left to Exercise 8.
Section 6.1 ORDINAL NUMBERS
291
THEOREM 6.1.18 [Transfinite Induction 2]
If ๐›ผ is an ordinal and ๐ด ⊆ ๐›ผ, then ๐ด = ๐›ผ if the following hold:
โˆ™ 0 ∈ ๐ด.
โˆ™ If ๐›ฝ ∈ ๐ด, then ๐›ฝ + ∈ ๐ด.
โˆ™ If ๐›ฝ is a limit ordinal such that ๐›ฟ ∈ ๐ด for all ๐›ฟ ∈ ๐›ฝ, then ๐›ฝ ∈ ๐ด.
We use this second form of transfinite induction to prove the sought-after classification
theorem for well-ordered sets.
THEOREM 6.1.19
Let (๐ด, 4) be a well-ordered set. Then, (๐ด, 4) ≅ (๐›ผ, ⊆) for some ordinal ๐›ผ.
PROOF
Define
๐‘(๐‘ฅ, ๐‘ฆ) := ๐‘ฆ is an ordinal ∧ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘ฅ) ≅ ๐‘ฆ.
Let ๐ต = {๐‘ฆ : ∃๐‘ฅ [๐‘ฅ ∈ ๐ด ∧ ๐‘(๐‘ฅ, ๐‘ฆ)]}. By Theorem 6.1.13, ๐‘(๐‘ฅ, ๐‘ฆ) defines a function, so by a replacement axiom (5.1.9), we conclude that ๐ต is a set. We have a
number of items to prove.
Let ๐ท ⊆ ๐ต and ๐ท ≠ ∅. Let ๐ถ = {๐‘Ž ∈ ๐ด : ∃๐›ผ[๐›ผ ∈ ๐ท ∧ ๐‘(๐‘Ž, ๐›ผ)]}. Observe
that ๐ถ is not empty. Therefore, there exists a least element ๐‘š ∈ ๐ถ with respect
to 4. Take an ordinal ๐›ฟ0 ∈ ๐ท such that
๐—Œ๐–พ๐—€4 (๐ด, ๐‘š) ≅ ๐›ฟ0 .
Let ๐›ฟ ∈ ๐ท. This means that ๐›ฟ is an ordinal and
๐—Œ๐–พ๐—€4 (๐ด, ๐‘) ≅ ๐›ฟ
for some ๐‘ ∈ ๐ถ. Since ๐‘š 4 ๐‘, we have that
๐—Œ๐–พ๐—€4 (๐ด, ๐‘š) ⊆ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘).
Hence, ๐›ฟ0 is isomorphic to a subset of ๐›ฟ, which implies that ๐›ฟ0 ⊆ ๐›ฟ (Theorem 6.1.13). We conclude that (๐ต, ⊆) is a well-ordered set.
Let ๐ธ = {๐›ฝ ∈ ๐ต : ๐—Œ๐–พ๐—€(๐ต, ๐›ฝ) = ๐›ฝ}. Let ๐—Œ๐–พ๐—€(๐ต, ๐œ–) ⊆ ๐ธ for ๐œ– ∈ ๐ต.
โˆ™ First, suppose that ๐œ– = ๐›พ + for some ordinal ๐›พ. Then, ๐—Œ๐–พ๐—€(๐ต, ๐›พ) = ๐›พ, so
๐—Œ๐–พ๐—€(๐ต, ๐›พ) ∪ {๐›พ} = ๐›พ + .
Also, ๐›พ + ≅ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž) for some ๐‘Ž ∈ ๐ด. Let ๐‘š be the greatest element of
๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž) (Exercise 9). This implies that ๐›พ ≅ ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž)โงต{๐‘š}, so ๐›พ ∈ ๐ต.
Hence,
๐—Œ๐–พ๐—€(๐ต, ๐›พ) ∪ {๐›พ} = ๐—Œ๐–พ๐—€(๐ต, ๐›พ + ),
292
Chapter 6 ORDINALS AND CARDINALS
and we have ๐œ– ∈ ๐ธ.
โ‹ƒ
โˆ™ Second, let ๐œ– = {๐›พ : ๐›พ ∈ ๐œ–}. This means that ๐—Œ๐–พ๐—€(๐ต, ๐›พ) = ๐›พ for all ๐›พ ∈ ๐œ–.
Therefore,
โ‹ƒ
โ‹ƒ
๐—Œ๐–พ๐—€(๐ต, ๐œ–) =
๐—Œ๐–พ๐—€(๐ต, ๐›พ) =
๐›พ = ๐œ–,
๐›พ∈๐œ–
๐›พ∈๐œ–
and ๐œ– is again an element of ๐ธ.
By transfinite induction (Theorem 6.1.18), ๐ธ = ๐ต. This combined with (๐ต, ⊆)
being a well-ordered set means that ๐ต is an ordinal.
Define ๐œ‘ : ๐ด → ๐ต by ๐œ‘(๐‘ฅ) = ๐‘ฆ ⇔ ๐‘(๐‘ฅ, ๐‘ฆ). Since ๐œ‘ is an order isomorphism
(Exercise 10), ๐ต is an ordinal that is order isomorphic to (๐ด, 4), and because of
Theorem 6.1.13, it is the only one.
For any well-ordered set (๐ด, โชฏ), the unique ordinal ๐›ผ such that ๐ด ≅ ๐›ผ guaranteed
by Theorem 6.1.19 is called the order type of ๐ด. Compare this definition with Definition 4.5.24. For example, the order type of ({2๐‘› : ๐‘› > 5 ∧ ๐‘› ∈ โ„ค}, ≤) is ๐œ”.
Burali-Forti and Hartogs
โ‹ƒ
Suppose ๐’œ = {0, 4, 6, 9}. Then, ๐’œ equals the ordinal
โ‹ƒ 9, which is the least upper
bound of ๐’œ . Also, assume that โ„ฌ = {5, 100, ๐œ”}. Then, โ„ฌ = ๐œ”. However, the least
upper bound of ๐’ž = {๐‘› ∈ ๐œ” :โ‹ƒ∃๐‘˜(๐‘˜ ∈ ๐œ” ∧ ๐‘› = 2๐‘˜)} is not an element of ๐œ”. Instead,
the least upper bound of ๐’ž is ๐’ž = ๐œ”. Moreover, notice that ๐’œ ⊆ 10, โ„ฌ ⊆ ๐œ”+ , and
๐’ž ⊆ ๐œ”+ . We generalize this to the next theorem.
THEOREM 6.1.20
If โ„ฑ is a set of ordinals, there exists an ordinal ๐›ผ such that โ„ฑ ⊆ ๐›ผ.
PROOF
โ‹ƒ
โ‹ƒ
Take โ„ฑ to be a set of ordinals and โ‹ƒ
let ๐›ผ ∈ โ„ฑ . Then,
โ‹ƒ ๐›ผ ⊆ โ„ฑ and โ„ฑ is an
ordinal by Theorem 6.1.16. If ๐›ผ ⊂ โ„ฑ , then ๐›ผ ∈ โ„ฑ by Theorem 6.1.11. If
(โ‹ƒ )+
โ‹ƒ
โ‹ƒ
๐›ผ = โ„ฑ , then ๐›ผ ∈ { โ„ฑ }. Thus, โ„ฑ ⊆
โ„ฑ .
Although every set of ordinals is a subset of an ordinal, there is no set of all ordinals,
otherwise a contradiction would arise, as was first discovered by Cesare Burali-Forti
(1897). This is why when we noted that ⊆ gives the ordinals a linear order, we did not
claim that ⊆ is used to define a linearly ordered set containing all ordinals.
THEOREM 6.1.21 [Burali-Forti]
There is no set that has every ordinal as an element.
PROOF
โ‹ƒ
Suppose โ„ฑ = {๐›ผ : ๐›ผ is an ordinal} is a set. This implies that โ„ฑ is an ordinal
by Theorem 6.1.16.โ‹ƒHowever, for
∈ ๐›ผ + ∈ ๐ด,
โ‹ƒ every ๐›ผ ∈ โ„ฑ , we haveโ‹ƒthat ๐›ผ โ‹ƒ
showing that โ„ฑ ⊆ โ„ฑ . Since โ„ฑ ∈ โ„ฑ , we also have โ„ฑ ∈ โ„ฑ , which
contradicts Theorem 5.1.16.
Section 6.1 ORDINAL NUMBERS
293
The Burali-Forti theorem places a limit on what can be done with ordinals. One such
example is a theorem of Friedrich Hartogs.
THEOREM 6.1.22 [Hartogs]
For every set ๐ด, there exists an ordinal ๐›ผ such that there are no injections of ๐›ผ
into ๐ด.
PROOF
Let ๐ด be a set. Define
โ„ฐ = {๐›ผ : ๐›ผ is an ordinal ∧ ∃๐œ“(๐œ“ is an injection ๐›ผ → ๐ด)}.
Notice that for every ๐›ผ ∈ โ„ฐ , there exists a bijection ๐œ‘๐›ผ such that
๐œ‘๐›ผ : ๐›ผ → ๐ต๐›ผ
for some ๐ต๐›ผ ⊆ ๐ด. Define a well-order 4๐›ผ on ๐ต๐›ผ by
๐œ‘๐›ผ (๐›ฝ1 ) 4๐›ผ ๐œ‘๐›ผ (๐›ฝ2 ) if and only if ๐›ฝ1 ⊆ ๐›ฝ2 for all ๐›ฝ1 , ๐›ฝ2 ∈ ๐›ผ.
Then, ๐œ‘๐›ผ is an order isomorphism preserving ⊆ with 4๐›ผ . Next, define
โ„ฑ = {(๐ต, ≤) : ๐ต ⊆ ๐ด ∧ ≤ is a well-ordering of ๐ต}.
Since โ„ฑ ⊆ P(๐ด) × P(๐ด × ๐ด), we have that โ„ฑ is a set by the Power Set Axiom
(5.1.7) and a Subset Axiom (5.1.8). Let
๐‘(๐‘ฅ, ๐‘ฆ) := ๐‘ฅ ∈ โ„ฑ ∧ ∃๐›พ(๐›พ is an order isomorphism ๐‘ฅ → ๐‘ฆ
preserving the order on ๐‘ฅ with ⊆}.
Suppose that ๐‘((๐ต, ≤), ๐›ผ1 ) and ๐‘((๐ต, ≤), ๐›ผ2 ). By Theorem 6.1.13, we have that
๐›ผ1 = ๐›ผ2 , so ๐‘(๐‘ฅ, ๐‘ฆ) defines a function with domain โ„ฑ . Moreover, โ„ฐ is a subset
of the range of this function because ๐‘((๐ต๐›ผ , 4๐›ผ ), ๐›ผ) due to ๐œ‘๐›ผ . Therefore, โ„ฐ is a
set by a replacement axiom (5.1.9) and a subset axiom, and โ„ฐ cannot contain all
ordinals by the Burali-Forti theorem (6.1.21).
Transfinite Recursion
Theorem 6.1.19 only applies to well-ordered sets, so, for example, it does not apply
to (โ„ค, ≤) or (โ„, ≤). However, if we change the order on โ„ค from the standard ≤ to 4
defined so as to put โ„ค into this order,
0, 1, −1, 2, −2, 3, −3, … ,
then (โ„ค, 4) is a well-ordered set of order type ๐œ”. That this can be done even with
sets like โ„ is due to a theorem first proved by Zermelo, which is often called the wellordering theorem. Its proof requires some preliminary work.
Let ๐ด be a set and ๐›ผ an ordinal. By Definition 4.4.13, ๐›ผ๐ด is the set of all functions
๐›ผ → ๐ด. Along these lines, define
๐ด = {๐œ‘ : ∃๐›ฝ(๐›ฝ ∈ ๐›ผ ∧ ๐œ‘ is a function ๐›ฝ → ๐ด)}.
<๐›ผ
294
Chapter 6 ORDINALS AND CARDINALS
For example, ๐‘“ , ๐‘” ∈ <5 โ„ค, where
๐‘“ = {(0, 1), (1, 2), (2, 3), (3, 4)}
and
๐‘” = {(0, −4), (1, 14)}.
Also, ๐‘“ , ๐‘” ∈ <๐œ” โ„ค, but the identity function on ๐œ” is not an element of <๐œ” โ„ค because its
domain is ๐œ”. We should also note that for any set ๐ด,
<∅
๐ด = ∅.
We use this notation in the following generalization of recursion to infinite ordinals.
THEOREM 6.1.23 [Transfinite Recursion]
Let ๐›ผ be an ordinal. For every function ๐œ“ :
function ๐œ‘ : ๐›ผ → ๐ด such that for every ๐›ฝ ∈ ๐›ผ,
<๐›ผ๐ด
→ ๐ด, there exists a unique
๐œ‘(๐›ฝ) = ๐œ“(๐œ‘ ๐›ฝ).
PROOF
To prove uniqueness, in addition to ๐œ‘, let ๐œ‘′ be a function ๐›ผ → ๐ด such that for
all ๐›ฝ ∈ ๐›ผ,
๐œ‘′ (๐›ฝ) = ๐œ“(๐œ‘′ ๐›ฝ).
Define ๐ต = {๐›ฝ ∈ ๐›ผ : ๐œ‘(๐›ฝ) = ๐œ‘′ (๐›ฝ)}. We use transfinite induction (Theorem 6.1.1) to show that ๐ต = ๐›ผ. Suppose seg(๐›ผ, ๐›ฟ) ⊆ ๐ต with ๐›ฟ ∈ ๐›ผ. That is,
∀๐›ฝ[๐›ฝ ∈ ๐›ฟ → ๐œ‘(๐›ฝ) = ๐œ‘′ (๐›ฝ)].
This implies that ๐œ‘ ๐›ฟ = ๐œ‘′ ๐›ฟ. Therefore,
๐œ‘(๐›ฟ) = ๐œ“(๐œ‘ ๐›ฟ) = ๐œ“(๐œ‘′ ๐›ฟ) = ๐œ‘′ (๐›ฟ),
so ๐›ฟ ∈ ๐ต, and we conclude that ๐œ‘ = ๐œ‘′ .
We prove existence indirectly. Suppose that ๐›ผ is the least ordinal (Exercise 2)
such that
there exists a function ๐œ“0 : <๐›ผ๐ด → ๐ด such that
for every ๐œ‘ : ๐›ผ → ๐ด, there exists ๐›ฝ ∈ ๐›ผ
such that ๐œ‘(๐›ฝ) ≠ ๐œ“0 (๐œ‘ ๐›ฝ).
Since the theorem is trivially true for ๐›ผ = 0, we have two cases to consider.
โˆ™ Let ๐›ผ = ๐›ฟ + for some ordinal ๐›ฟ. By minimality of ๐›ผ, we have a function
๐œ‘๐›ฟ : ๐›ฟ → ๐ด such that for all ๐›ฝ ∈ ๐›ฟ,
๐œ‘๐›ฟ (๐›ฝ) = ๐œ“0 (๐œ‘๐›ฟ ๐›ฝ).
Section 6.1 ORDINAL NUMBERS
295
Extend ๐œ‘๐›ฟ to ๐œ‘ฬ„ ๐›ฟ : ๐›ผ → ๐ด by defining ๐œ‘ฬ„ ๐›ฟ (๐›ฟ) = ๐œ“0 (๐œ‘๐›ฟ ), so
๐œ‘ฬ„ ๐›ฟ (๐›ฟ) = ๐œ“0 (๐œ‘ฬ„ ๐›ฟ ๐›ฟ).
This contradicts the minimality of ๐›ผ.
โˆ™ Let ๐›ผ be a limit ordinal. For each ๐›ฟ ∈ ๐›ผ, there exists a unique ๐œ‘๐›ฟ : ๐›ฟ → ๐ด
such that
๐œ‘๐›ฟ (๐›ฟ) = ๐œ“(๐œ‘๐›ฟ ๐›ฟ).
Notice that ๐›ฟ ∈ ๐›พ implies that ๐œ‘๐›พ is an extension of ๐œ‘๐›ฟ , otherwise ๐œ‘๐›พ ๐›ฟ
would have the property
(๐œ‘๐›พ ๐›ฟ)(๐›ฝ) = ๐œ“0 ([๐œ‘๐›พ ๐›ฟ] ๐›ฝ)
for all ๐›ฝ ∈ ๐›ฟ yet ๐œ‘๐›ฟ ≠ ๐œ‘๐›พ ๐›ฟ. This contradicts the uniqueness of ๐œ‘๐›ฟ .
Therefore, {๐œ‘๐›ฟ : ๐›ฟ ∈ ๐›ผ} is a chain, so, as in Exercise 4.4.13, define the
function ๐œ‘ : ๐›ผ → ๐ด by
โ‹ƒ
๐œ‘๐›ฟ .
๐œ‘=
๐›ฟ∈๐›ผ
To check that ๐œ‘ is the function given by the theorem, take ๐›ฝ ∈ ๐›ผ. Since ๐›ผ
is a limit ordinal, ๐›ฝ + ∈ ๐›ผ and
๐œ‘(๐›ฝ) = ๐œ‘๐›ฝ + (๐›ฝ) = ๐œ“0 (๐œ‘๐›ฝ + ๐›ฝ) = ๐œ“0 (๐œ‘ ๐›ฝ),
again contradicting the minimality of ๐›ผ.
Theorem 6.1.23 has a corollary that can be viewed as an extension of Theorem 5.2.14.
Its proof is left to Exercise 18.
COROLLARY 6.1.24
Let ๐ด be a set and ๐‘Ž ∈ ๐ด. For every ordinal ๐›ผ, if ๐œ“ is a function ๐ด → ๐ด, there
exists a unique function ๐œ‘ : ๐›ผ → ๐ด such that
โˆ™ ๐œ‘(0) = ๐‘Ž,
โˆ™ ๐œ‘(๐›ฝ + ) = ๐œ“(๐œ‘(๐›ผ)) for all ๐›ฝ ∈ ๐›ผ,
โ‹ƒ
โˆ™ ๐œ‘(๐›พ) = {๐œ‘(๐›ฝ) : ๐›ฝ ∈ ๐›พ} for all limit ordinals ๐›พ ∈ ๐›ผ.
We are now ready to prove that every set can be well-ordered. The following theorem
is equivalent to the axiom of choice (Exercise 20).
THEOREM 6.1.25 [Zermelo]
For any set ๐ด, there exists a relation ๐‘… on ๐ด such that (๐ด, ๐‘…) is a well-ordered
set.
296
Chapter 6 ORDINALS AND CARDINALS
PROOF
Take a set ๐ด and let ๐œ‰ : P(๐ด) → ๐ด be a choice function (Corollary 5.1.11). By
Theorem 6.1.22, there is an ordinal ๐›ผ such that no injection ๐›ผ → ๐ด exists. Hence,
we have ๐œ‘ : ๐ด → ๐›ผ that is one-to-one. Let ๐ต ⊆ ๐ด. Since every element of ๐›ผ is
an ordinal, there exists an ordinal ๐›ฟ ⊆ ๐›ผ such that ๐œ‘ [๐ต] = ๐›ฟ (Theorem 6.1.16).
Define
๐œ“๐ต = ๐œ‘ ๐ต.
Then, ๐œ“๐ต−1 ∈ <๐›ผ ๐ด. Let
๐‘ƒ = {๐œ“๐ต−1 : ๐ต ∈ P(๐ด)},
and define
โ„Ž1 : ๐‘ƒ → P(๐ด)
by โ„Ž1 (๐œ“๐ต−1 ) = ๐ต. Also, define
โ„Ž2 : ๐‘ƒ → ๐ด
by โ„Ž2 = ๐œ‰ โ—ฆ โ„Ž1 . Since ๐‘ƒ ⊆ <๐›ผ ๐ด, extend โ„Ž2 to some
โ„Ž : <๐›ผ ๐ด → ๐ด
such that โ„Ž ๐‘ƒ = โ„Ž2 . By transfinite recursion (Theorem 6.1.23), there exists a
function ๐‘“ : ๐›ผ → ๐ด such that for all ๐›ฝ ∈ ๐›ผ,
๐‘“ (๐›ฝ) = โ„Ž(๐‘“ ๐›ฝ).
Define Φ so that for all ๐›ฝ ∈ ๐›ผ,
{
โ„Ž(๐ด โงต ๐‘“ [๐›ฝ]) if ๐ด โงต ๐‘“ [๐›ฝ] ≠ ∅,
Φ(๐›ฝ) =
๐ด
if ๐ด โงต ๐‘“ [๐›ฝ] = ∅.
Notice that ๐ด ∉ ๐ด by Theorem 5.1.16. Let ๐›ฝ0 be the least ordinal such that
Φ(๐›ฝ0 ) = ๐ด. Then, Φ ๐›ฝ0 is a bijection ๐›ฝ0 → ๐ด [Exercise 19(a)]. Lastly, define
the relation ๐‘… on ๐ด by
๐‘Ž ๐‘… ๐‘ if and only if Φ−1 (๐‘Ž) ⊆ Φ−1 (๐‘),
for all ๐‘Ž, ๐‘ ∈ ๐ด. Since ⊆ is a well-order on ๐›ฝ0 , ๐‘… is a well-order on ๐ด [Exercise 19(b)].
Exercises
1. Prove Lemma 6.1.5.
โ‹ƒ
โ‹ƒ
2. Does ๐—Œ๐–พ๐—€( ๐ด, ๐›ฝ) = ๐›ผ∈๐ด ๐—Œ๐–พ๐—€(๐›ผ, ๐›ฝ) for all sets of ordinals ๐ด with ๐›ฝ ∈ ๐ด? Explain.
3. Explain why {0, 2, 3, 4, 5} is not an ordinal.
4. Prove that ∅ is an element of every ordinal.
Section 6.1 ORDINAL NUMBERS
297
5. Prove that an ordinal is a transitive set (Theorem 6.1.8).
6. Let ๐ด be a set of ordinals. Prove.
โ‹‚
(a)
๐ด is an ordinal.
(b) ๐ด has a least element with respect to ⊆.
7. Let ๐ต be a nonempty subset of the ordinal ๐›ผ. Prove that there exists ๐›ฝ ∈ ๐ต such
that ๐›ฝ and ๐ต are disjoint.
8. Prove Theorem 6.1.18.
9. From the proof of Theorem 6.1.19, prove that ๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž) has a greatest element.
10. Prove that ๐ด is order isomorphic to ๐ต in the proof of Theorem 6.1.19.
11. The proof of Theorem 6.1.19 contains many isomorphisms without explicitly identifying the isomorphism. Find these functions and prove that they are order isomorphisms.
12. Let ๐‘… be a well-ordering on ๐ด and suppose that ๐ด has no greatest element. Show
the the order type of (๐ด, ๐‘…) is a limit ordinal.
13. Find a transitive set that is not an ordinal.
14. Theorem 6.1.21 comes from the Burali-Forti paradox. Like Russell’s paradox
(page 225), it arises when any formula is allowed to define a set. In this case, suppose
that ๐ด = {๐›ผ : ๐›ผ is an ordinal} and assume that ๐ด is a set. Prove that ๐ด is an ordinal
that must include all ordinals as its elements.
15. Prove that there exists a function ๐น such that ๐น (๐‘›) is the ๐‘›th Fibonacci number.
16. Prove that for every function โ„Ž : <๐œ” ๐ด → ๐ด, there is a unique function ๐‘“ : ๐œ” → ๐ด
such that for all ๐‘› ∈ ๐œ”, ๐‘“ (๐‘›) = โ„Ž(๐‘“ ๐‘›).
17. Let (๐ด, 4) and (๐ต, 4′ ) be well-ordered sets and ๐œ‘ : ๐ด → ๐ต be an order-preserving
surjection. Prove that for every ๐‘Ž ∈ ๐ด, there exists ๐‘ ∈ ๐ต such that
๐œ‘[๐—Œ๐–พ๐—€4 (๐ด, ๐‘Ž)] = ๐—Œ๐–พ๐—€4′ (๐ต, ๐‘).
18. Prove Corollary 6.1.24.
19. Prove the following from the proof of Zermelo’s theorem (6.1.25).
(a) Φ ๐›ฝ0 is a bijection.
(b) ๐‘… is a well-order on ๐ด.
20. Prove that Theorem 6.1.25 implies Axiom 5.1.10.
21. Prove that Zorn’s lemma (5.1.13) implies Theorem 6.1.25 without using the axiom
of choice.
298
6.2
Chapter 6 ORDINALS AND CARDINALS
EQUINUMEROSITY
How can we determine whether two sets are of the same size? One possibility is to
count their elements. What happens, however, if the sets are infinite? We need another
method. Suppose ๐ด = {12, 47, 84} and ๐ต = {17, 101, 200}. We can see that these
two sets are the same size without counting. Define a function ๐‘“ : ๐ด → ๐ต so that
๐‘“ (12) = 17, ๐‘“ (47) = 101, and ๐‘“ (84) = 200. This function is a bijection. Since each
element is paired with exactly one element of the opposite set, ๐ด and ๐ต must be the
same size. This is the motivation behind our first definition.
DEFINITION 6.2.1
The sets ๐ด and ๐ต are equinumerous (written as ๐ด ≈ ๐ต) if there exists a bijection
๐œ‘ : ๐ด → ๐ต. If ๐ด and ๐ต are not equinumerous, write ๐ด โ‰‰ ๐ต.
EXAMPLE 6.2.2
Take ๐‘› ∈ โ„ค such that ๐‘› ≠ 0 and define
๐‘›โ„ค = {๐‘›๐‘˜ : ๐‘˜ ∈ โ„ค}.
We prove that โ„ค ≈ ๐‘›โ„ค. To show this, we must find a bijection ๐‘“ : โ„ค → ๐‘›โ„ค.
Define ๐‘“ (๐‘˜) = ๐‘›๐‘˜.
โˆ™ Assume ๐‘ฅ1 , ๐‘ฅ2 ∈ โ„ค, and let ๐‘“ (๐‘ฅ1 ) = ๐‘“ (๐‘ฅ2 ). Then ๐‘›๐‘ฅ1 = ๐‘›๐‘ฅ2 , which yields
๐‘ฅ1 = ๐‘ฅ2 since ๐‘› ≠ 0. Thus, ๐‘“ is one-to-one.
โˆ™ Let ๐‘ฆ ∈ ๐‘›โ„ค. This means that ๐‘ฆ = ๐‘›๐‘˜ for some ๐‘˜ ∈ โ„ค, so ๐‘ฆ = ๐‘“ (๐‘˜). This
shows that ๐‘“ is onto and, hence, a bijection.
EXAMPLE 6.2.3
To see โ„ค+ ≈ โ„ค, define a one-to-one correspondence so that each even integer is
paired with a nonnegative integer and every odd integer is paired with a negative
integer (Figure 6.1). Let ๐‘” : โ„ค+ → โ„ค be defined by
{
๐‘˜−1
๐‘”(๐‘›) =
−๐‘˜
if ๐‘› = 2๐‘˜ for some ๐‘˜ ∈ โ„ค+ ,
if ๐‘› = 2๐‘˜ − 1 for some ๐‘˜ ∈ โ„ค+ .
Notice that ๐‘”(4) = 1 since 4 = 2(2), and ๐‘”(5) = −3 because 5 = 2(3) − 1. This
function is a bijection (Exercise 12).
Equinumerosity plays a role similar to that of equality of integers. This is seen in the
next theorem. In fact, the theorem resembles Definition 4.2.4. Despite this, it does not
demonstrate the existence of an equivalence relation. This is because an equivalence
relation is a relation on a set, so, to define an equivalence relation, the next result would
require a set of all sets, contradicting Corollary 5.1.17.
Section 6.2 EQUINUMEROSITY
...
−4
1
2
3
4
5
6
7
−3
−2
−1
0
1
2
3
Figure 6.1
8
299
...
...
โ„ค ≈ โ„ค+ .
THEOREM 6.2.4
Let ๐ด, ๐ต, and ๐ถ be sets.
โˆ™ ๐ด ≈ ๐ด. (Reflexive)
โˆ™ If ๐ด ≈ ๐ต, then ๐ต ≈ ๐ด. (Symmetric)
โˆ™ If ๐ด ≈ ๐ต and ๐ต ≈ ๐ถ, then ๐ด ≈ ๐ถ. (Transitive)
PROOF
โˆ™ ๐ด ≈ ๐ด since the identity map is a bijection.
โˆ™ Assume ๐ด ≈ ๐ต. Then, there exists a bijection ๐œ‘ : ๐ด → ๐ต. Therefore, ๐œ‘−1 is a
bijection. Hence, ๐ต ≈ ๐ด.
โˆ™ By Theorem 4.5.23, the composition of two bijections is a bijection. Therefore,
๐ด ≈ ๐ต and ๐ต ≈ ๐ถ implies ๐ด ≈ ๐ถ.
The symmetric property allows us to conclude that ๐‘›โ„ค ≈ โ„ค and โ„ค ≈ โ„ค+ from
Examples 6.2.2 and 6.2.3. The transitivity part of Theorem 6.2.4 allows us to conclude
from this that ๐‘›โ„ค ≈ โ„ค+ .
EXAMPLE 6.2.5
Show (0, 1) ≈ โ„. We do this in two parts. First, let ๐‘“ : (0, 1) → (−๐œ‹โˆ•2, ๐œ‹โˆ•2) be
defined by ๐‘“ (๐‘ฅ) = ๐œ‹๐‘ฅ − ๐œ‹โˆ•2. This function is a one-to-one correspondence since
its graph is a nonvertical, nonhorizontal line (Exercise 4.5.4). Second, define
๐‘” : (−๐œ‹โˆ•2, ๐œ‹โˆ•2) → โ„ to be the function ๐‘”(๐‘ฅ) = tan ๐‘ฅ. From trigonometry, we
know that tangent is a bijection on (−๐œ‹โˆ•2, ๐œ‹โˆ•2). Hence,
(0, 1) ≈ (−๐œ‹โˆ•2, ๐œ‹โˆ•2)
and
(−๐œ‹โˆ•2, ๐œ‹โˆ•2) ≈ โ„.
Therefore, (0, 1) ≈ โ„ by Theorem 6.2.4.
300
Chapter 6 ORDINALS AND CARDINALS
Order
If ≈ resembles equality, the following resembles the ≤ relation.
DEFINITION 6.2.6
The set ๐ต dominates the set ๐ด (written as ๐ด โชฏ ๐ต) if there exists an injection
๐œ‘ : ๐ด → ๐ต. If ๐ต does not dominate ๐ด, write ๐ด ฬธโชฏ ๐ต. Furthermore, define ๐ด โ‰บ ๐ต
to mean ๐ด โชฏ ๐ต but ๐ด โ‰‰ ๐ต.
EXAMPLE 6.2.7
If ๐ด ⊆ ๐ต, then ๐ด โชฏ ๐ต. This is proved using the inclusion map (Exercise 4.4.26).
For instance, let ๐œ„ : โ„ค+ → โ„ค be the inclusion map and ๐‘“ : โ„ค → โ„ be defined as
๐‘“ (๐‘›) = ๐—Œ๐–พ๐—€≤ (โ„š, ๐‘›).
Then, โ„ค+ โชฏ โ„ because ๐‘“ โ—ฆ ๐œ„ is an injection. Similarly, โ„š โชฏ โ„‚. However, ๐ด ⊂ ๐ต
does not imply ๐ด โ‰‰ ๐ต. As an example, ๐‘›โ„ค ≈ โ„ค, but ๐‘›โ„ค ⊂ โ„ค when ๐‘› ≠ ±1.
Another method used to prove that ๐ด โชฏ ๐ต is to find a surjection ๐ต → ๐ด. Consider
the sets ๐ด = {1, 2} and ๐ต = {3, 4, 5}. Define ๐‘“ : ๐ต → ๐ด to be the surjection given by
๐‘“ (3) = 1, ๐‘“ (4) = 2, and ๐‘“ (5) = 2. This is the inverse of the relation ๐‘… in Figure 4.1.
To show that ๐ต dominates ๐ด, we must find an injection ๐ด → ๐ต. To do this, modify ๐‘“ −1
by deleting (2, 4) and call the resulting function ๐‘”. Observe that ๐‘”(1) = 3 and ๐‘”(2) = 5,
which is an injection, so ๐ด โชฏ ๐ต.
THEOREM 6.2.8
If there exists a surjection ๐œ‘ : ๐ด → ๐ต, then ๐ต โชฏ ๐ด.
PROOF
Let ๐œ‘ : ๐ด → ๐ต be onto. Define a relation ๐‘… ⊆ ๐ต × ๐ด by
๐‘… = {(๐‘, ๐‘Ž) : ๐œ‘(๐‘Ž) = ๐‘}.
Since ๐œ‘ is onto,
dom(๐‘…) = ran(๐œ‘) = ๐ต.
Corollary 5.1.11 yields a function ๐‘“ so that dom(๐‘“ ) = dom(๐‘…) and ๐‘“ ⊆ ๐‘….
We claim that ๐‘“ is one-to-one. Indeed, let ๐‘1 , ๐‘2 ∈ ๐ต. Assume that we have
๐‘“ (๐‘1 ) = ๐‘“ (๐‘2 ). Let ๐‘Ž1 = ๐‘“ (๐‘1 ) and ๐‘Ž2 = ๐‘“ (๐‘2 ) where ๐‘Ž1 , ๐‘Ž2 ∈ ๐ด. This means
๐‘Ž1 = ๐‘Ž2 . Also, ๐œ‘(๐‘Ž1 ) = ๐‘1 and ๐œ‘(๐‘Ž2 ) = ๐‘2 because ๐‘“ ⊆ ๐‘…. Since ๐œ‘ is a function,
๐‘1 = ๐‘2 .
EXAMPLE 6.2.9
Let ๐‘… be an equivalence relation on a set ๐ด. The map ๐œ‘ : ๐ด → ๐ดโˆ•๐‘… defined by
๐œ‘(๐‘Ž) = [๐‘Ž]๐‘… is a surjection. Therefore, ๐ดโˆ•๐‘… โชฏ ๐ด.
Section 6.2 EQUINUMEROSITY
301
EXAMPLE 6.2.10
We know that โ„ค+ โชฏ โ„ by Example 6.2.7. We can also prove this by using the
function ๐‘“ : โ„ → โ„ค+ defined by ๐‘“ (๐‘ฅ) = |โŸฆ๐‘ฅโŸง| + 1 and appealing to Theorem 6.2.8.
The next theorem states that โชฏ closely resembles an antisymmetric relation (Definition 4.3.1). Cantor was the first to publish a statement of it (1888). He proved it using
the axiom of choice, but it was later shown that it can be proved in ๐—ญ๐—™. It was proved
independently by Ernst Schröder and Felix Bernstein around 1890. The proof given
here follows that of Julius König (1906).
THEOREM 6.2.11 [Cantor–Schröder–Bernstein]
If ๐ด โชฏ ๐ต and ๐ต โชฏ ๐ด, then ๐ด ≈ ๐ต.
PROOF
Let ๐‘“ : ๐ด → ๐ต and ๐‘” : ๐ต → ๐ด be injections. To prove that ๐ด is equinumerous to
๐ต, we define a one-to-one correspondence โ„Ž : ๐ด → ๐ต. To do this, we recursively
define two sequences of sets by first letting
๐ถ0 = ๐ด โงต ran(๐‘”)
and
๐ท0 = ๐‘“ [๐ถ0 ].
Then, for ๐‘› ∈ ๐œ”,
๐ถ๐‘›+1 = ๐‘”[๐ท๐‘› ]
and
๐ท๐‘› = ๐‘“ [๐ถ๐‘› ].
This is illustrated in Figure 6.2. Note that both {๐ถ๐‘› : ๐‘› ∈ ๐œ”} and {๐ท๐‘› : ๐‘› ∈ ๐œ”}
are pairwise disjoint because ๐‘“ and ๐‘” are one-to-one (Exercise 11). Define โ„Ž by
(
)
(
)
โ‹ƒ
โ‹ƒ
−1
โ„Ž=๐‘“
๐ถ๐‘› ∪ ๐‘” ran(๐‘”) โงต
๐ถ๐‘› .
๐‘›∈๐œ”
๐‘›∈๐œ”
We show that โ„Ž is the desired function.
โˆ™ Let ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐ด such that ๐‘ฅ1 ≠ ๐‘ฅ2 . Since both ๐‘“ and ๐‘” are one-to-one, we
only need to check the case when ๐‘ฅ1 ∈ ๐ถ๐‘˜ for some ๐‘˜ ∈ ๐œ” and ๐‘ฅ2 ∉ ๐ถ๐‘›
for all ๐‘› ∈ ๐œ”. Then, ๐‘“ (๐‘ฅ1 ) ∈ ๐ท๐‘˜ but ๐‘” −1 (๐‘ฅ2 ) ∉ ๐ท๐‘˜ , so ๐‘“ (๐‘ฅ1 ) ≠ ๐‘” −1 (๐‘ฅ2 ).
That is, โ„Ž(๐‘ฅ1 ) ≠ โ„Ž(๐‘ฅ2 ).
โˆ™ Take ๐‘ฆ ∈ ๐ต. If ๐‘ฆ ∈ ๐ท๐‘˜ for some ๐‘˜ ∈ ๐œ”, then
โ‹ƒ ๐‘ฆ = ๐‘“ (๐‘ฅ) for some ๐‘ฅ ∈ ๐ถ๐‘˜ .
That is, ๐‘ฆ = โ„Ž(๐‘ฅ). Now suppose ๐‘ฆ ∉ ๐‘›∈๐œ” ๐ท๐‘› . Clearly, ๐‘”(๐‘ฆ) ∉ ๐ถ0 .
If ๐‘”(๐‘ฆ) ∈ ๐ถ๐‘˜ for some ๐‘˜ > 0, thenโ‹ƒ๐‘ฆ ∈ ๐ท๐‘˜−1 , a contradiction. Hence,
๐‘”(๐‘ฆ) = ๐‘ฅ for some ๐‘ฅ ∈ ran(๐‘”) โงต ๐‘›∈๐œ” ๐ถ๐‘› . This implies that we have
โ„Ž(๐‘ฅ) = ๐‘” −1 (๐‘ฅ) = ๐‘ฆ.
302
Chapter 6 ORDINALS AND CARDINALS
๎”
A ๎œ ran(g)
๎”
A
B
f โ†พC0
C0
D0
gโ†พD0
C1
f โ†พC1
D1
gโ†พD1
C2
ran(g)
. . .
. . .
ran(g) ๎œ
๎ƒ
Cn
n∈๐œ”
Figure 6.2
๎€˜
g −1โ†พ ran(g) ๎œ
๎ƒ
Cn
๎€™
n∈๐œ”
The Cantor–Schröder–Bernstein theorem.
B๎œ
๎ƒ
Dn
n∈๐œ”
Section 6.2 EQUINUMEROSITY
303
EXAMPLE 6.2.12
Because (0, 1) ⊆ [0, 1], we have that (0, 1) โชฏ [0, 1], and because the function
๐‘“ : [0, 1] → (0, 1) defined by
๐‘“ (๐‘ฅ) =
1
1
๐‘ฅ+
2
4
is an injection, we have that [0, 1] โชฏ (0, 1), so we conclude by the Cantor–
Schröder–Bernstein theorem (6.2.11) that (0, 1) ≈ [0, 1].
Diagonalization
The strict inequality ๐ด โ‰บ ๐ต is sometimes difficult to prove because we must show that
there does not exist a bijection from ๐ด onto ๐ต. The next method was developed by
Cantor (1891) to accomplish this for infinite sets. It is called diagonalization.
Let ๐‘€ be the set of all functions ๐‘“ : ๐œ” → {๐‘š, ๐‘ค}, where ๐‘š ≠ ๐‘ค. To show that
๐œ” โ‰บ ๐‘€, we prove two facts:
โˆ™ There exists an injection ๐œ” → ๐‘€.
โˆ™ There is no one-to-one correspondence between ๐œ” and ๐‘€.
Cantor’s method does both of these at once. Let ๐œ‘ : ๐œ” → ๐‘€ be a function. Writing
the functions of the range of ๐œ‘ as infinite tuples, let
๐‘“๐‘– = (๐‘Ž๐‘–0 , ๐‘Ž๐‘–1 , ๐‘Ž๐‘–2 , … , ๐‘Ž๐‘–๐‘— , … ),
where ๐‘Ž๐‘–๐‘— ∈ {๐‘š, ๐‘ค} for all ๐‘–, ๐‘— ∈ ๐œ”. For example,
๐œ‘(4)(3) = ๐‘“4 (3) = ๐‘Ž4,3 .
Now, write the functions in order:
๐‘“0 = (a00 , ๐‘Ž01 , ๐‘Ž02 , … , ๐‘Ž0๐‘— , … ),
๐‘“1 = (๐‘Ž10 , a11 , ๐‘Ž12 , … , ๐‘Ž1๐‘— , … ),
๐‘“2 = (๐‘Ž20 , ๐‘Ž21 , a22 , … , ๐‘Ž2๐‘— , … ),
โ‹ฎ
๐‘“๐‘– = (๐‘Ž๐‘–0 , ๐‘Ž๐‘–1 , ๐‘Ž๐‘–2 , … , aii , … , ๐‘Ž๐‘–๐‘— , … ),
โ‹ฎ
From this, a function ๐‘“ ∈ ๐‘€ that is not in the list can be found by identifying the
elements on the diagonal and defining ๐‘“ (๐‘›) to be the opposite of ๐‘Ž๐‘›๐‘› . In other words,
define for all ๐‘– ∈ ๐œ”,
{
๐‘š if ๐‘Ž๐‘–๐‘– = ๐‘ค,
๐‘๐‘– =
๐‘ค if ๐‘Ž๐‘–๐‘– = ๐‘š,
304
Chapter 6 ORDINALS AND CARDINALS
and ๐‘“ (๐‘›) = ๐‘๐‘› is an element of ๐‘€ not in the list because for all ๐‘› ∈ ๐œ”,
๐‘“ (๐‘›) ≠ ๐‘Ž๐‘›๐‘› .
Since the function ๐œ‘ mapping ๐œ” to ๐‘€ is arbitrary, there are injections ๐œ” → ๐‘€ but
none of them are onto. Therefore,
๐œ” โ‰บ ๐‘€.
Furthermore, note that the elements of [0, 1] can be uniquely represented as binary
numbers of the form
0 . ๐‘Ž0 ๐‘Ž1 ๐‘Ž2 … ๐‘Ž๐‘– … ,
where ๐‘Ž๐‘– ∈ {0, 1} for each ๐‘– ∈ ๐œ”. For example
1 = 0.1111111 …
and
1โˆ•2 = 0.1000000 … .
Therefore,
๐‘€ ≈ [0, 1] .
Hence, we can conclude like Cantor that since [0, 1] ≈ โ„ (Examples 6.2.5 and 6.2.12),
๐œ” โ‰บ โ„.
Cantor’s diagonalization argument can be generalized, but we first need a definition.
Let ๐ด be a set and ๐ต ⊆ ๐ด. The function
๐œ’๐ต : ๐ด → {0, 1}
is called a characteristic function and is defined by
{
1 if ๐‘Ž ∈ ๐ต,
๐œ’๐ต (๐‘Ž) =
0 if ๐‘Ž ∉ ๐ต.
For example, if ๐ด = โ„ค and ๐ต = {0, 1, 3, 5}, then ๐œ’๐ต (1) = 1 but ๐œ’๐ต (2) = 0. Moreover,
for every set ๐ด,
๐ด
2 = {๐œ’๐ต : ๐ต ⊆ ๐ด}.
The characteristic function plays an important role in the proof of the next theorem.
THEOREM 6.2.13
If ๐ด is a set, ๐ด โ‰บ ๐ด 2.
Section 6.2 EQUINUMEROSITY
305
PROOF
Since the function ๐œ“ : ๐ด → ๐ด 2 defined by
๐œ“(๐‘Ž) = ๐œ’{๐‘Ž}
is an injection, ๐ด 2 dominates ๐ด. To show that ๐ด is not equinumerous with ๐ด 2,
we show that there is no surjection ๐ด → ๐ด 2. Let ๐œ‘ โˆถ ๐ด → ๐ด 2 be a function, and
for all ๐‘Ž ∈ ๐ด, write ๐œ‘(๐‘Ž) = ๐œ’๐ต๐‘Ž for some ๐ต๐‘Ž ⊆ ๐ด. Define ๐œ’ so that
{
1 if ๐œ’๐ต๐‘Ž (๐‘Ž) = 0,
๐œ’(๐‘Ž) =
0 if ๐œ’๐ต๐‘Ž (๐‘Ž) = 1.
Therefore, ๐œ’ ∉ ran(๐œ‘) because ๐œ’๐ต๐‘Ž (๐‘Ž) ≠ ๐œ’(๐‘Ž) for all ๐‘Ž ∈ ๐ด. However, ๐œ’ ∈ ๐ด 2.
To prove this, define
๐ต = {๐‘Ž ∈ ๐ด : ๐œ’๐ต๐‘Ž (๐‘Ž) = 0}.
We conclude that ๐œ’(๐‘Ž) = ๐œ’๐ต (๐‘Ž) because if ๐œ’๐ต๐‘Ž (๐‘Ž) = 0, then ๐‘Ž ∈ ๐ต, so ๐œ’(๐‘Ž) = 1
and ๐œ’๐ต (๐‘Ž) = 1, but if ๐œ’๐ต๐‘Ž (๐‘Ž) = 1, then ๐‘Ž ∉ ๐ต, so ๐œ’(๐‘Ž) = 0 and ๐œ’๐ต (๐‘Ž) = 0.
Therefore, ๐œ’ = ๐œ’๐ต , and ๐œ‘ is not onto.
By Exercise 14,
P(๐ด) ≈ ๐ด 2.
This result combined with Theorem 6.2.13 quickly yield the following.
COROLLARY 6.2.14
If ๐ด is a set, ๐ด โ‰บ P(๐ด).
From the theorem, we conclude that there exists a sequence of sets
๐œ” โ‰บ P(๐œ”) โ‰บ P(P(๐œ”)) โ‰บ P(P(P(๐œ”))) โ‰บ · · · .
Thus, there are larger and larger magnitudes of infinity.
Exercises
1. Given ๐œ’๐ต : โ„ค → {0, 1} with ๐ต = {2, 5, 19, 23}, find
(a) ๐œ’๐ต (1)
(b) ๐œ’๐ต (2)
(c) ๐œ’๐ต (−10)
(d) ๐œ’๐ต (19)
2. Find a surjection ๐œ‘ : ๐œ” × ๐œ” → ๐œ” showing that ๐œ” โชฏ ๐œ” × ๐œ”.
3. Prove the given equations.
(a) ๐œ’๐ด∪๐ต = ๐œ’๐ด + ๐œ’๐ต − ๐œ’๐ด ๐œ’๐ต
(b) ๐œ’๐ด∩๐ต = ๐œ’๐ด ๐œ’๐ต
4. Prove that there exists a bijection between the given pairs of sets.
306
Chapter 6 ORDINALS AND CARDINALS
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
๐œ‹] and [−1, 1]
[[0,
]
−๐œ‹โˆ•2, ๐œ‹โˆ•2 and [−1, 1]
(0, ∞) and โ„
๐œ” and โ„ค
โ„ค+ and โ„ค−
{(๐‘ฅ, 0) : ๐‘ฅ ∈ โ„} and โ„
โ„ค and โ„ค × โ„ค
{(๐‘ฅ, ๐‘ฆ) ∈ โ„ × โ„ : ๐‘ฆ = 2๐‘ฅ + 4} and โ„
5. Let ๐‘Ž < ๐‘ and ๐‘ < ๐‘‘, where ๐‘Ž, ๐‘, ๐‘, ๐‘‘ ∈ โ„. Prove.
(a) (๐‘Ž, ๐‘) ≈ (๐‘, ๐‘‘)
(b) [๐‘Ž, ๐‘] ≈ [๐‘, ๐‘‘]
(c) (๐‘Ž, ๐‘) ≈ [๐‘, ๐‘‘]
(d) (๐‘Ž, ๐‘) ≈ (๐‘, ๐‘‘]
6. Prove โ„ ≈ โ„‚.
7. Let ๐ด, ๐ต, ๐ถ, and ๐ท be nonempty sets. Prove.
(a) โ„• โชฏ โ„ค−
(b) ๐ด ∩ ๐ต โชฏ P(๐ด)
(c) ๐ด โชฏ ๐ด × ๐ต
(d) [0, 2] โชฏ [5, 7]
(e) ๐ด ∩ ๐ต โชฏ ๐ด
(f) ๐ด๐ต โชฏ ๐ด × ๐ต
(g) ๐ด × {0} โชฏ (๐ด ∪ ๐ต) × {1}
(h) ๐ด × ๐ต × ๐ถ โชฏ ๐ด × ๐ต × ๐ถ × ๐ท
(i) ๐ด โงต ๐ต โชฏ ๐ถ × ๐ด
8. Prove.
(a) If ๐ด โชฏ ๐ต and ๐ต ≈ ๐ถ, then ๐ด โชฏ ๐ถ.
(b) If ๐ด ≈ ๐ต and ๐ต โชฏ ๐ถ, then ๐ด โชฏ ๐ถ.
(c) If ๐ด โชฏ ๐ต and ๐ต โชฏ ๐ถ, then ๐ด โชฏ ๐ถ.
(d) If ๐ด โชฏ ๐ต and ๐ถ โชฏ ๐ท, then ๐ด ๐ถ โชฏ ๐ต ๐ท.
(e) If ๐ด ≈ ๐ต, then P(๐ด) ≈ P(๐ต).
(f) If ๐ด ≈ ๐ต and ๐ถ ≈ ๐ท, then ๐ด × ๐ถ ≈ ๐ต × ๐ท.
(g) If ๐ด ≈ ๐ต, ๐‘Ž ∈ ๐ด, and ๐‘ ∈ ๐ต, then ๐ด โงต {๐‘Ž} ≈ ๐ต โงต {๐‘}.
(h) If ๐ด โงต ๐ต ≈ ๐ต โงต ๐ด, then ๐ด ≈ ๐ต.
(i) If ๐ด ⊆ ๐ต and ๐ด ≈ ๐ด ∪ ๐ถ, then ๐ต ≈ ๐ด ∪ ๐ต ∪ ๐ถ.
(j) If ๐ถ ⊆ ๐ด, ๐ต ⊆ ๐ท, and ๐ด ∪ ๐ต ≈ ๐ต, then ๐ถ ∪ ๐ท ≈ ๐ท.
9. Given the function ๐œ‘ : ๐ด → ๐ต, prove that ๐œ‘ ≈ ๐ด and ๐œ‘ โชฏ ran(๐œ‘).
10. Without appealing to the Cantor–Schröder–Bernstein theorem (6.2.11), prove that
(0, 1) ≈ [0, 1].
Section 6.3 CARDINAL NUMBERS
307
11. Prove that the sets {๐ถ๐‘› : ๐‘› ∈ ๐œ”} and {๐ท๐‘› : ๐‘› ∈ ๐œ”} from the proof of Theorem 6.2.11 are pairwise disjoint.
12. Show that ๐‘” is a bijection, where ๐‘” : โ„ค+ → โ„ค is defined by
{
๐‘˜−1
๐‘”(๐‘›) =
−๐‘˜
if ๐‘› = 2๐‘˜ for some ๐‘˜ ∈ ๐œ”,
if ๐‘› = 2๐‘˜ + 1 for some ๐‘˜ ∈ ๐œ”.
13. Let ๐ด be an infinite set and {๐‘Ž1 , ๐‘Ž2 , … } be a set of distinct elements from ๐ด. Prove
that ๐ด → ๐ด โงต {๐‘Ž1 } is a bijection, where
{
๐‘Ž๐‘›+1
๐œ‘(๐‘ฅ) =
๐‘ฅ
if ๐‘ฅ = ๐‘Ž๐‘› ,
otherwise.
14. For any set ๐ด, prove that P(๐ด) ≈ ๐ด 2.
15. Prove that if ๐‘“ : ๐ด → ๐ต is a surjection, there exists a function ๐‘” : ๐ต → ๐ด such
that ๐‘“ โ—ฆ ๐‘” = ๐ผ๐ต .
16. Use the power set to prove that there is no set of all sets.
6.3
CARDINAL NUMBERS
Let ๐ด be a set. Define
๐ต = {๐›ฝ : ๐›ฝ is an ordinal ∧ ๐›ฝ ≈ ๐ด}.
By Zermelo’s theorem (6.1.25), ๐ด can be well-ordered, so by Theorem 6.1.19, ๐ด is
order isomorphic to some ordinal, so ๐ต is nonempty. Moreover, ๐ต is subset of an
ordinal (Theorem 6.1.20). Therefore, (๐ต, ⊆) has a least element, which has the property
that it is not equinumerous to any of its elements. This allows us to define the second
of our new types of number (page 283). This type will be used to denote the size of a
set.
DEFINITION 6.3.1
An ordinal ๐œ… is a cardinal number (or simply a cardinal) if ๐›ผ โ‰‰ ๐œ… for every
๐›ผ ∈ ๐œ….
Observe that every infinite cardinal is a limit ordinal. This is because ๐›ผ ≈ ๐›ผ + for every
infinite ordinal ๐›ผ. However, a limit ordinal might not be a cardinal.
Let ๐œ… and ๐œ† be cardinals. Suppose that ๐ด ≈ ๐œ… and ๐ด ≈ ๐œ†. By Theorem 6.2.4,
we have that ๐œ… ≈ ๐œ†. If ๐œ… ∈ ๐œ† or ๐œ† ∈ ๐œ…, this would contradict the definition of a
cardinal number. Therefore, ๐œ… = ๐œ† (Corollary 6.1.15), and we conclude that every set
is equinumerous to exactly one cardinal.
308
Chapter 6 ORDINALS AND CARDINALS
DEFINITION 6.3.2
The cardinality of a set ๐ด is denoted by |๐ด| and defined as the unique cardinal
equinumerous to ๐ด.
Observe that the cardinality of a cardinal ๐œ… is ๐œ….
EXAMPLE 6.3.3
โ‹ƒ
Let ๐’œ be a set of cardinals. By Theorem 6.1.16, we know that ๐’œ is an ordinal.
Now we show that it is also a cardinal. Suppose that ๐›ผ is an ordinal such that
โ‹ƒ
โ‹ƒ
๐›ผ ≈
๐’œ . By Definition
โ‹ƒ 6.3.1, we must show that ๐’œ ⊆ ๐›ผ. Suppose there
exists an ordinal ๐›ฝ ∈
๐’œ such that ๐›ฝ ∉ ๐›ผ. This means that there exists a
cardinal ๐œ… ∈ ๐’œ such that ๐›ฝ ∈ ๐œ…. This is impossible because
โ‹ƒ
๐›ผโชฏ๐›ฝโ‰บ๐œ…โชฏ
๐’œ ≈ ๐›ผ.
Finite Sets
Intuitively, we know what a finite set is. Both of the sets ๐ด = {0, 2, 3, 5, 8, 10} and
๐ต = {๐‘› ∈ โ„ค : (๐‘› − 1)(๐‘› + 3) = 0} are examples because we can count all of their
elements and find that there is only one ordinal equinumerous to ๐ด and only one ordinal equinumerous to ๐ต. That is, |๐ด| = 6 and |๐ต| = 2. This suggests the following
definition.
DEFINITION 6.3.4
For every set ๐ด, if there exists ๐‘› ∈ ๐œ” such that ๐ด ≈ ๐‘›, then ๐ด is finite. If ๐ด is
not finite, it is infinite.
As we will see, finite sets are fundamentally different from infinite sets. There are
properties that finite sets have in addition to the number of their elements that infinite
sets do not have. Let us consider some of those properties of finite sets.
LEMMA 6.3.5
Let ๐‘› be a positive natural number. If ๐‘ฆ ∈ ๐‘›, then ๐‘› โงต {๐‘ฆ} ≈ ๐‘›− .
PROOF
We proceed by mathematical induction.
โˆ™ When ๐‘› = 1, it must be the case that ๐‘ฆ = 0, so ๐‘› โงต {๐‘ฆ} = 0 = 1− .
โˆ™ Take ๐‘ฆ ∈ ๐‘› + 1. If ๐‘ฆ = ๐‘›, then (๐‘› + 1) โงต {๐‘›} = ๐‘›, so suppose that ๐‘ฆ < ๐‘›.
By induction, there exists a bijection ๐‘” : ๐‘› โงต {๐‘ฆ} → ๐‘›− . Then, define
๐‘“ : (๐‘› + 1) โงต {๐‘ฆ} → ๐‘› by ๐‘“ (๐‘š) = ๐‘”(๐‘š) for all ๐‘š < ๐‘› and ๐‘“ (๐‘›) = ๐‘›. The
function ๐‘“ is a bijection (Exercise 3).
Lemma 6.3.5 is used to prove a characteristic property of finite cardinals.
Section 6.3 CARDINAL NUMBERS
309
THEOREM 6.3.6
No natural number is equinumerous to a proper subset of itself.
PROOF
Let ๐‘› ∈ ๐œ” be minimal such that there exists ๐ด ⊂ ๐‘› and ๐‘› ≈ ๐ด. Since ∅ has no
proper subsets and the only proper subset of 1 is 0, we can assume that ๐‘› ≥ 2.
Let ๐‘“ : ๐ด → ๐‘› be a bijection and ๐‘ฅ ∈ ๐ด and ๐‘ฆ ∈ ๐‘› โงต ๐ด such that ๐‘“ (๐‘ฅ) = ๐‘ฆ. We
check the following.
โˆ™ Let ๐‘Ž ∈ ๐ด โงต {๐‘ฅ}. If ๐‘“ (๐‘Ž) = ๐‘ฆ, then we contradict the hypothesis that ๐‘“ is
one-to-one because ๐‘“ (๐‘ฅ) = ๐‘ฆ and ๐‘ฅ ≠ ๐‘Ž. Thus, ๐‘“ (๐‘Ž) ∈ ๐‘› โงต {๐‘ฆ}.
โˆ™ Take ๐‘ ∈ ๐‘› โงต {๐‘ฆ}. Since ๐‘“ is a surjection, there exists ๐‘Ž ∈ ๐ด such that
๐‘“ (๐‘Ž) = ๐‘. If ๐‘Ž = ๐‘ฅ, then {๐‘, ๐‘ฆ} ⊆ [๐‘ฅ]๐‘“ (Definition 4.2.7), which is impossible because ๐‘“ is a function. Thus, ๐‘Ž ∈ ๐ด โงต {๐‘ฅ}, and we conclude that
๐‘“ (๐ด โงต {๐‘ฅ}) is onto ๐‘› โงต {๐‘ฆ}.
โˆ™ Since the restriction of a one-to-one function is one-to-one, ๐‘“ (๐ด โงต {๐‘ฅ})
is one-to-one.
Hence, ๐‘“0 = ๐‘“ (๐ด โงต {๐‘ฅ}) is a bijection with range ๐‘› โงต {๐‘ฆ}. We have two cases
to consider.
โˆ™ Suppose ๐‘›− ∉ ๐ด. This implies that ๐ด โงต {๐‘ฅ} ⊆ ๐‘›− . Since ๐‘ฅ ∈ ๐ด, we have
that ๐‘ฅ ≠ ๐‘›− , so ๐‘ฅ ∈ ๐‘›− . Thus, ๐ด โงต {๐‘ฅ} ⊂ ๐‘›− . By Lemma 6.3.5, there exists
a bijection ๐‘” : ๐‘› โงต {๐‘ฆ} → ๐‘›− , so we have ๐ด โงต {๐‘ฅ} ≈ ๐‘›− because ๐‘” โ—ฆ ๐‘“0 is a
bijection, contradicting the minimality of ๐‘›.
โˆ™ Assume ๐‘›− ∈ ๐ด. Define ๐ด′ = ๐ด โงต {๐‘›− } ∪ {๐‘ฆ}. Since ๐‘ฆ ∉ ๐ด, ๐ด′ ≈ ๐ด,
which implies that ๐ด′ ≈ ๐‘›. Replace ๐ด with ๐ด′ in the previous argument
and use ๐‘“ (๐ด′ โงต {๐‘ฆ}) to contradict the minimality of ๐‘›.
COROLLARY 6.3.7
Every finite set is equinumerous to exactly one natural number.
PROOF
Let ๐ด be finite. This means that ๐ด ≈ ๐‘› for some ๐‘› ∈ ๐œ”. Let ๐‘š ∈ ๐œ” also have
the property that ๐ด ≈ ๐‘š. This implies that ๐‘› ≈ ๐‘š. Hence, by Theorem 6.3.6, we
conclude that ๐‘› = ๐‘š because ๐‘› ⊆ ๐‘š or ๐‘š ⊆ ๐‘›.
There are many results that follow directly from Theorem 6.3.6. The following six
corollaries are among them.
COROLLARY 6.3.8 [Pigeonhole Principle]
Let ๐ด and ๐ต be finite sets with ๐ต โ‰บ ๐ด. There is no one-to-one function ๐ด → ๐ต.
310
Chapter 6 ORDINALS AND CARDINALS
PROOF
There exists unique ๐‘š, ๐‘› ∈ ๐œ” such that ๐ด ≈ ๐‘š and ๐ต ≈ ๐‘› by Corollary 6.3.7.
Assume that ๐‘“ : ๐ด → ๐ต is one-to-one. Then,
๐‘š ≈ ๐ด โชฏ ๐ต ≈ ๐‘›,
so ๐‘š is equinumerous to a subset of ๐‘›. This implies that ๐‘š ⊆ ๐‘›. However, ๐‘› ⊂ ๐‘š
because ๐ต โ‰บ ๐ด, which contradicts Theorem 6.1.14.
COROLLARY 6.3.9
No finite set is equinumerous to a proper subset of itself.
COROLLARY 6.3.10
A set equinumerous to a proper subset of itself is infinite.
Because ๐‘“ : ๐œ” → ๐œ” โงต {0} defined by ๐‘“ (๐‘›) = ๐‘› + 1 is a bijection, we have the next
result by Corollary 6.3.10.
COROLLARY 6.3.11
๐œ” is infinite.
The proofs of the last two corollaries are left to Exercise 5.
COROLLARY 6.3.12
If ๐ด is a proper subset of a natural number ๐‘›, there exists ๐‘š < ๐‘› such that ๐ด ≈ ๐‘š.
COROLLARY 6.3.13
Let ๐ด ⊆ ๐ต. If ๐ต is finite, ๐ด is finite, and if ๐ด is infinite, ๐ต is infinite.
Countable Sets
Since ๐œ” is the first infinite ordinal, it is also a cardinal. Therefore,
|๐ด| = ๐œ” if and only if ๐ด ≈ ๐œ”.
As sets go, finite sets and those equinumerous with ๐œ” are small, so we classify them
together using the next definition.
DEFINITION 6.3.14
A set ๐ด is countable if ๐ด โชฏ ๐œ”.
Sometimes countable sets are called discrete or denumerable. For example, the bijection ๐‘“ : โ„ค+ → ๐œ” defined by ๐‘“ (๐‘›) = ๐‘›− shows that โ„ค+ is countable. Moreover, a
nonempty finite set is countable and can be written as
{๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 },
Section 6.3 CARDINAL NUMBERS
311
for some positive integer ๐‘›. A countably infinite set can be written as
{๐‘Ž0 , ๐‘Ž1 , ๐‘Ž2 , … },
where there are infinitely many distinct elements of the set.
EXAMPLE 6.3.15
The set of rational numbers is a countable set. To prove this, define a bijection
๐‘“ : ๐œ” → โ„š by first mapping the even natural numbers to the nonnegative rational
numbers. The function is defined along the path indicated in Figure 6.3. When
a rational number that previously has been used is encountered, it is skipped.
To complete the definition, associate the odd naturals with the negative rational numbers using a path as in the diagram. This function is a bijection, so we
conclude that โ„š is countable.
We have defined countability in terms of bijections. Now let us identify a condition
for countability using surjections.
THEOREM 6.3.16
A set ๐ด is countable if and only if there exists a function from ๐œ” onto ๐ด.
PROOF
If ๐œ‘ : ๐œ” → ๐ด is a surjection, by Theorem 6.2.8, ๐ด โชฏ ๐œ”. Conversely, suppose ๐ด
is countable. We have two cases to check.
โˆ™ Suppose ๐ด ≈ ๐œ”. Then, there is a surjection from the set of natural numbers
to ๐ด.
f(6) =
2
1
f(4) =
1
1
f(0) =
0
1
Figure 6.3
2
2
f(2) =
0
2
1
2
f(8) =
2
3
f(10) =
1
3
0
3
The rational numbers are countable.
...
312
Chapter 6 ORDINALS AND CARDINALS
โˆ™ Now let ๐ด ≈ ๐‘› for some ๐‘› ∈ ๐œ”. If ๐ด = ∅, then ∅ is a surjection ๐œ” → ๐ด.
Thus, assume ๐ด ≠ ∅. This means that we can write
๐ด = {๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 }.
Define ๐œ‘ : ๐œ” → ๐ด by
{
๐‘Ž๐‘–
๐œ‘(๐‘–) =
๐‘Ž๐‘›−1
if ๐‘– = 0, 1, … , ๐‘› − 1,
otherwise.
This function is certainly onto.
If ๐ด๐‘– is countable for all ๐‘– = 0, 1, … , ๐‘› − 1, then ๐ด0 × ๐ด1 × · · · × ๐ด๐‘›−1 is countable
(Exercise 15). In particular, ๐œ” × ๐œ” and โ„ค × โ„ค are countable. We also have the next
theorem.
THEOREM 6.3.17
The union of a countable family of countable sets is countable.
PROOF
Let {๐ด
โ‹ƒ ๐›ผ : ๐›ผ ∈ ๐ผ} be a family of countable sets with ๐ผ countable. Since we have
that ∅ = ∅ is countable (Example 3.4.12), we can assume that ๐ผ is nonempty.
For each ๐›ผ ∈ ๐ผ, there exists a surjection in ๐œ” (๐ด๐›ผ ) by Theorem 6.3.16. Therefore,
by Corollary 5.1.11, there exists
๐‘“ : ๐ผ → ๐œ” (๐ด๐›ผ )
such that ๐‘“ (๐›ผ) is a surjection ๐œ” → ๐ด๐›ผ for all ๐›ผ ∈ ๐ผ. Because ๐ผ is countable,
we have a surjection ๐‘” : ๐œ” → ๐ผ, and since ๐œ” × ๐œ” is countable, we have another
surjection โ„Ž : ๐œ” → ๐œ” × ๐œ”. We now define
โ‹ƒ
๐œ“ :๐œ”×๐œ”→
๐ด๐›ผ
๐›ผ∈๐ผ
by ๐œ“(๐‘š, ๐‘›) = ๐‘“ (๐‘”(๐‘š))(๐‘›) and let
๐œ‘:๐œ”→
โ‹ƒ
๐ด๐›ผ
๐›ผ∈๐ผ
be defined by ๐œ‘ = ๐œ“ โ—ฆ โ„Ž. To check that ๐œ‘ is onto, let ๐‘Ž ∈ ๐ด๐›ผ , some ๐›ผ ∈ ๐ผ. Since
๐‘” is onto, there exists ๐‘– ∈ ๐œ” so that ๐‘”(๐‘–) = ๐›ผ. Furthermore, since ๐‘“ (๐›ผ) is onto,
we have ๐‘— ∈ ๐œ” such that ๐‘“ (๐›ผ)(๐‘—) = ๐‘Ž, and since โ„Ž is onto, there exists ๐‘˜ ∈ ๐œ” so
that โ„Ž(๐‘˜) = (๐‘–, ๐‘—). Therefore,
๐œ‘(๐‘˜) = (๐œ“ โ—ฆ โ„Ž)(๐‘˜) = ๐œ“(๐‘–, ๐‘—) = ๐‘“ (๐‘”(๐‘–))(๐‘—) = ๐‘“ (๐›ผ)(๐‘—) = ๐‘Ž.
For example,
โ‹ƒ
โ‹ƒ
{๐‘›โ„ค : ๐‘› ∈ ๐œ”} and {โ„š × {๐‘›} : ๐‘› ∈ ๐œ”} are countable.
Section 6.3 CARDINAL NUMBERS
313
Alephs
Cantor denoted the first infinite cardinal ๐œ” by ℵ0 . The symbol ℵ (aleph) is the first
letter of the Hebrew alphabet. The next magnitude of infinity is ℵ1 , which seems to
exist by Theorem 6.2.13. This continues and gives an increasing sequence of infinite
cardinals, and since natural numbers must be less than any infinite cardinal, we have
0 โ‰บ 1 โ‰บ 2 โ‰บ · · · โ‰บ ℵ0 โ‰บ ℵ1 โ‰บ ℵ2 โ‰บ · · · .
For instance, 4 โ‰บ ℵ1 , ℵ0 โชฏ ℵ0 , and ℵ3 โ‰บ ℵ7 .
EXAMPLE 6.3.18
Although he was unable to prove it, Cantor suspected that ℵ1 = |โ„|. This conjecture is called the continuum hypothesis (๐–ข๐–ง). However, it is possible that
ℵ1 โ‰บ |โ„|. It is also possible that ℵ1 = |โ„|. Cantor was unable to prove ๐–ข๐–ง
because it is undecidable assuming the axioms of ๐—ญ๐—™๐—–. In other words, it is an
๐–ฒ๐–ณ-sentence that can be neither proved nor disproved from ๐—ญ๐—™๐—–.
Now to define the other alephs. Pick an ordinal ๐›ผ. Define the function โ„Ž by
โ„Ž(๐‘”) = least infinite cardinal not in ran(๐‘”),
where ๐‘” is a function with dom(๐‘”) ∈ ๐›ผ. For example, if ๐›ผ = 5 and
๐‘” = {(0, ๐œ…0 ), (1, ๐œ…1 ), (2, ๐œ…2 ), (3, ๐œ…3 ), (4, ๐œ…4 )},
then โ„Ž(๐‘”) is the least infinite cardinal not in {๐œ…0 , ๐œ…1 , ๐œ…2 , ๐œ…3 , ๐œ…4 }. By Transfinite Recursion (6.1.23), there exists a unique function ๐‘“ with domain ๐›ผ and
๐‘“ (๐›ฝ) = least infinite cardinal not in ran(๐‘“ ๐›ฝ)
for all ๐›ฝ ∈ ๐›ผ. Define
ℵ๐›ฝ = ๐‘“ (๐›ฝ).
Since โ„Ž(0) = ๐œ”, we have that ℵ0 = ๐œ”. Moreover, the definition of ๐‘“ implies that
ℵ1 = least infinite cardinal not in {ℵ0 },
ℵ2 = least infinite cardinal not in {ℵ0 , ℵ1 },
ℵ3 = least infinite cardinal not in {ℵ0 , ℵ1 , ℵ2 },
โ‹ฎ
ℵ๐œ” = least infinite cardinal not in {ℵ๐‘› : ๐‘› ∈ ๐œ”},
โ‹ฎ
ℵ๐›ผ = least infinite cardinal not in {ℵ๐›ฝ : ๐›ฝ ∈ ๐›ผ}
โ‹ฎ
The question at this point is whether the alephs name all of the infinite cardinals.
The next theorem answers the question.
314
Chapter 6 ORDINALS AND CARDINALS
THEOREM 6.3.19
For every infinite cardinal ๐œ…, there exists an ordinal ๐›ผ such that ๐œ… = ℵ๐›ผ .
PROOF
Suppose ๐œ… ≠ ℵ๐›ผ for every ordinal ๐›ผ. We can assume that ๐œ… is the minimal
such cardinal. Thus, for all cardinals ๐œ† ∈ ๐œ…, there exists an ordinal ๐›ฝ๐œ† such that
๐œ† = ℵ๐›ฝ๐œ† . Therefore, the least infinite cardinal not an element of
{ℵ๐›ฝ๐œ† : ๐œ† ∈ ๐œ… ∧ ๐œ† is a cardinal}
is the next aleph, but this is ๐œ….
EXAMPLE 6.3.20
The generalized continuum hypothesis (๐–ฆ๐–ข๐–ง) states that for every ordinal ๐›ผ,
ℵ๐›ผ+ = |P(ℵ๐›ผ )|.
When ๐›ผ = 0, ๐–ฆ๐–ข๐–ง implies that
ℵ1 = |P(ℵ0 )| = |โ„|,
which is ๐–ข๐–ง (Example 6.3.18). Like ๐–ข๐–ง, ๐–ฆ๐–ข๐–ง is undecidable in ๐—ญ๐—™๐—–.
Like the ordinals, the cardinals can be divided into two classes.
DEFINITION 6.3.21
Let ๐œ… be a nonzero cardinal number. If ๐œ… ∈ ๐œ” or there exists an ordinal ๐›ผ such
that ๐œ… = ℵ๐›ผ+ , then ๐œ… is a successor cardinal. Otherwise, ๐œ… is a limit cardinal.
For example, the positive natural numbers and ℵ1 are successor cardinals, while ℵ0
and ℵ๐œ” are limit cardinals. Notice that if ๐œ… is a limit cardinal,
โ‹ƒ
๐œ…=
{๐œ† : ๐œ† ∈ ๐œ… ∧ ๐œ† is a cardinal}.
Exercises
1. Prove that |๐›ผ| = |๐›ผ + | for every ordinal ๐›ผ.
2. Let ๐‘(๐‘ฅ) be a formula. Prove that if ๐‘(๐›ผ) is false for some ordinal ๐›ผ, then there exists
a least ordinal ๐›ฝ such that ๐‘(๐›ฝ) is false.
3. Prove that the function ๐‘“ in the proof of Lemma 6.3.5 is a bijection.
4. Show that the following attempted generalization of Lemma 6.3.5 is false: Let ๐›ผ be
an ordinal and ๐‘Ž ∈ ๐ด. If |๐ด| = ℵ๐›ผ + , then |๐ด โงต {๐‘Ž}| = ℵ๐›ผ .
5. Demonstrate Corollaries 6.3.12 and 6.3.13.
6. Let ๐ด and ๐ต be finite sets. Prove the following.
Section 6.3 CARDINAL NUMBERS
315
(a) |๐ด ∪ ๐ต| = |๐ด| + |๐ต| − |๐ด ∩ ๐ต|
(b) |๐ด ∩ ๐ต| = |๐ด| − |๐ด โงต ๐ต|
(c) |๐ด × ๐ต| = |๐ด| ⋅ |๐ต|
7. Prove that the intersection and union of finite sets is finite is finite.
8. Prove that every finite set has a choice function without using the axiom of choice.
9. Let ๐‘… and ๐‘…−1 be well-orderings of a set ๐ด. Prove that ๐ด is finite.
10. Show that โ„ค is countable.
11. Let ๐ด be infinite. Find infinite sets ๐ต and ๐ถ such that ๐ด = ๐ต ∪ ๐ถ.
12. If ๐ต is countable, prove that |๐ด × ๐ต| = |๐ด|.
13. Let ๐ด and ๐ต be sets and ๐ด is countable. Prove that ๐ต is countable when ๐ด ≈ ๐ต.
14. Let ๐ด and ๐ต be countable sets. Show that the given sets are countable.
(a) ๐ด ∪ ๐ต
(b) ๐ด ∩ ๐ต
(c) ๐ด × ๐ต
(d) ๐ด โงต ๐ต
15. Let ๐ด0 , ๐ด1 , … , ๐ด๐‘›−1 be countable sets. Prove that the given sets are countable.
(a) ๐ด0 ∪ ๐ด1 ∪ · · · ∪ ๐ด๐‘›−1
(b) ๐ด0 ∩ ๐ด1 ∩ · · · ∩ ๐ด๐‘›−1
(c) ๐ด0 × ๐ด1 × · · · × ๐ด๐‘›−1
16. Prove.
(a) If ๐ด ∪ ๐ต is countable, ๐ด and ๐ต are countable.
(b) If ๐ด is countable, 2 ๐œ” ≈ ๐ด ๐œ”
โ‹ƒ
17. Let โ„ฑ be a set of cardinals. Prove that โ„ฑ is a cardinal.
18. A real number is algebraic if it is a root of a nonzero polynomial with integer
coefficients. A real number that is not algebraic is transcendental. Prove that the set
of algebraic numbers is countable and the set of transcendental numbers is uncountable.
19. Take a set ๐ด and define ๐ต = {๐›ผ : ๐›ผ is an ordinal ∧ ๐›ผ โชฏ ๐ด}. [See Hartogs’ theorem
(6.1.22).] Prove the following.
(a) ๐ต is a cardinal.
(b) |๐ด| โ‰บ ๐ต.
(c) ๐ต is the least cardinal such that |๐ด| โ‰บ ๐ต.
20. Assuming ๐–ฆ๐–ข๐–ง, find |P(P(P(P(P(โ„)))))|.
21. Prove that for all ordinals ๐›ผ and ๐›ฝ, if ๐›ผ ∈ ๐›ฝ, then ℵ๐›ผ ∈ ℵ๐›ฝ .
316
Chapter 6 ORDINALS AND CARDINALS
22. Recursively define the following function using i (beth), the second letter of the
Hebrew alphabet. Let ๐›ผ be an ordinal.
i0 = ℵ0 ,
i๐›ผ+ = 2i๐›ผ ,
โ‹ƒ
i๐›ผ =
i๐›ฝ if ๐›ผ is a limit.
๐›ฝ∈๐›ผ
(a) Use i to restate ๐–ฆ๐–ข๐–ง.
(b) Use transfinite recursion (Corollary 6.1.24) to prove that i defines a function.
6.4
ARITHMETIC
Since every natural number is both an ordinal and a cardinal, we want to extend the
operations on ๐œ” to all of the ordinals and all of the cardinals. Since the purpose of the
ordinals is to characterize well-ordered sets but the purpose of the cardinals is to count,
we expect the two extensions to be different.
Ordinals
Definitions 5.2.15 and 5.2.18 define what it means to add and multiply finite ordinals.
When generalizing these two definitions to the infinite ordinals, we must take care
because addition and multiplication should be binary operations, but if we defined these
operations on all ordinals, their domains would not be sets by the Burali-Forti Theroem
(6.1.21), resulting in the operations not being sets. Therefore, we choose an ordinal and
define addition and multiplication on it. Since 1 + 1 ∉ 2, the ordinal must be a limit
ordinal.
DEFINITION 6.4.1
Let ๐œ be a limit ordinal. For all ๐›ผ, ๐›ฝ ∈ ๐œ,
โˆ™ ๐›ผ + 0 = ๐›ผ,
โˆ™ ๐›ผ + ๐›ฝ + = (๐›ผ + ๐›ฝ)+ ,
โ‹ƒ
โˆ™ ๐›ผ + ๐›ฝ = {๐›ผ + ๐›ฟ : ๐›ฟ ∈ ๐›ฝ} if ๐›ฝ is a limit ordinal.
As with addition of natural numbers (Definition 5.2.15), to prove that Definition 6.4.1
gives a binary operation, let ๐œ“ : ๐œ → ๐œ be the successor function. By transfinite
recursion (Corollary 6.1.24), there exist unique functions ๐œ‘๐›ผ : ๐›ผ → ๐œ for all ๐›ผ ∈ ๐œ
such that
โˆ™ ๐œ‘๐›ผ (0) = ๐›ผ
โˆ™ ๐œ‘๐›ผ (๐›ฝ + ) = ๐œ“(๐œ‘๐›ผ (๐›ฝ)) = ๐œ‘๐›ผ (๐›ฝ)+
Section 6.4 ARITHMETIC
โˆ™ for all limit ordinals ๐›ฝ ∈ ๐œ,
๐œ‘๐›ผ (๐›ฝ) =
โ‹ƒ
317
๐œ‘๐›ผ (๐›ฟ).
๐›ฟ∈๐›ฝ
Define A : ๐œ × ๐œ → ๐œ by A(๐›ผ, ๐›ฝ) = ๐œ‘๐›ผ (๐›ฝ). By the uniqueness of each ๐œ‘๐›ผ , the binary
operation A is the function of Definition 6.4.1, so
๐›ผ + ๐›ฝ = A(๐›ผ, ๐›ฝ).
Furthermore, take ๐œ ′ to be a limit ordinal such that ๐œ ⊆ ๐œ ′ and define ๐œ“ ′ : ๐œ ′ → ๐œ ′ to
be the successor function. As above, there are unique functions ๐œ‘′๐›ผ for all ๐›ผ ∈ ๐œ ′ with
the same properties as ๐œ‘๐›ผ . Notice that
๐œ‘๐›ผ = ๐œ‘′๐›ผ ๐œ .
Otherwise, ๐œ‘′๐›ผ ๐œ would have the same properties as ๐œ‘๐›ผ yet be a different function,
contradicting the uniqueness given by transfinite recursion. Next, define the binary
operation A′ : ๐œ ′ × ๐œ ′ → ๐œ ′ by A′ (๐›ผ, ๐›ฝ) = ๐œ‘′๐›ผ (๐›ฝ). Therefore,
A = A′ (๐œ × ๐œ).
This implies that although addition is not defined as a binary operation on all ordinals,
we can add any two ordinals and obtain the same sum independent of the ordinal on
which the addition is defined.
Consider ๐‘š ∈ ๐œ”. Since ๐œ” is a limit ordinal,
โ‹ƒ
๐‘š+๐œ”=
{๐‘š + ๐‘› : ๐‘› ∈ ๐œ”} = ๐œ”.
However,
๐œ” + 1 = ๐œ” + 0+ = (๐œ” + 0)+ = ๐œ”+ = ๐œ” ∪ {๐œ”},
and
๐œ” + 2 = ๐œ” + 1+ = (๐œ” + 1)+ = (๐œ” ∪ {๐œ”})+ = ๐œ” ∪ {๐œ”} ∪ {๐œ” ∪ {๐œ”}} = ๐œ”++ .
Therefore, addition of infinite ordinals is not commutative. Moreover, an order isomorphism can be defined between ๐œ” + ๐‘› and ({0} × ๐œ”) ∪ ({1} × ๐‘›) ordered lexicographically
(Exercise 4.3.16) as
0
โ†•
(0, 0)
1
โ†•
(0, 1)
2
โ†•
(0, 2)
…
…
๐‘›
โ†•
(0, ๐‘›)
…
…
๐œ”
โ†•
(1, 0)
๐œ”+
โ†•
(1, 1)
๐œ”++
โ†•
(1, 2)
…
…
This means that ๐œ” + ๐‘› looks like ๐œ” followed by a copy of ๐‘›. Generalizing, the ordinal
๐œ” + ๐œ” looks like ๐œ” followed by a copy of ๐œ”. In particular,
โ‹ƒ
๐œ”+๐œ”=
๐œ” + ๐‘›,
๐‘›∈๐œ”
which means that the proof its existence requires a replacement axiom (5.1.9). All of
this suggests the next result.
318
Chapter 6 ORDINALS AND CARDINALS
LEMMA 6.4.2
Let ๐œ be a limit ordinal and ๐›ผ, ๐›ฝ ∈ ๐œ. If ๐‘ฅ ∈ ๐›ผ + ๐›ฝ, then ๐‘ฅ ∈ ๐›ผ or ๐‘ฅ ∈ ๐›ผ + ๐‘ for
some ๐‘ ∈ ๐›ฝ.
PROOF
Define
๐ด = {๐‘ง ∈ ๐œ : ∀๐‘ฅ(๐‘ฅ ∈ ๐›ผ + ๐‘ง → ๐‘ฅ ∈ ๐›ผ ∨ ∃๐‘” ∈ ๐‘ง[๐‘ฅ ∈ ๐›ผ + ๐‘”])}.
Clearly, 0 ∈ ๐ด, so take ๐›พ ∈ ๐œ such that ๐—Œ๐–พ๐—€(๐œ , ๐›พ) ⊆ ๐ด.
โˆ™ Let ๐›พ = ๐›ฟ + for some ordinal ๐›ฟ. Take ๐‘ฅ ∈ ๐›ผ + ๐›ฟ + , which implies that
๐‘ฅ ∈ (๐›ผ + ๐›ฟ)+ . This means that ๐‘ฅ ∈ ๐›ผ + ๐›ฟ or ๐‘ฅ = ๐›ผ + ๐›ฟ. If ๐‘ฅ ∈ ๐›ผ + ๐›ฟ, then
we are done. If ๐‘ฅ = ๐›ผ + ๐›ฟ, then ๐‘ฅ ∈ (๐›ผ + ๐›ฟ)+ = ๐›ผ + ๐›ฟ + .
โˆ™ Let ๐›พ be a limit ordinal. Take ๐‘ฅ ∈ ๐›ผ + ๐›พ. This means that ๐‘ฅ ∈ ๐›ผ + ๐›ฟ for
some ๐›ฟ ∈ ๐›พ. Therefore, ๐‘ฅ ∈ ๐›ผ or ๐‘ฅ ∈ ๐›ผ + ๐‘‘ for some ๐‘‘ ∈ ๐›ฟ ⊂ ๐›พ.
Using Lemma 6.4.2, we can prove the next useful result.
LEMMA 6.4.3
Let ๐œ be a limit ordinal. If ๐›ผ, ๐›ฝ ∈ ๐œ, then ๐›ผ + ๐›ฝ = ๐›ผ ∪ {๐›ผ + ๐‘ : ๐‘ ∈ ๐›ฝ}.
PROOF
Define
๐ด = {๐‘ง ∈ ๐œ : ๐›ผ + ๐‘ง = ๐›ผ ∪ {๐›ผ + ๐‘” : ๐‘” ∈ ๐‘ง}}.
We have that 0 ∈ ๐ด, so assume ๐—Œ๐–พ๐—€(๐œ , ๐›พ) ⊆ ๐ด, where ๐›พ ∈ ๐œ.
โˆ™ Suppose ๐›พ = ๐›ฟ + . Then,
๐›ผ ∪ {๐›ผ + ๐‘‘ : ๐‘‘ ∈ ๐›ฟ + } = ๐›ผ ∪ {๐›ผ + ๐‘‘ : ๐‘‘ ∈ ๐›ฟ} ∪ {๐›ผ + ๐›ฟ}
= (๐›ผ + ๐›ฟ) ∪ {๐›ผ + ๐›ฟ}
= (๐›ผ + ๐›ฟ)+
= ๐›ผ + ๐›ฟ+.
โˆ™ Let ๐›พ be a limit ordinal. Take ๐‘ฅ ∈ ๐›ผ + ๐›พ. By Lemma 6.4.2, we have ๐‘ฅ ∈ ๐›ผ
or ๐‘ฅ ∈ ๐›ผ + ๐‘” for some ๐‘” ∈ ๐›พ. If the former, we are done, so suppose the
latter. In this case, the assumption gives
๐›ผ + ๐‘” = ๐›ผ ∪ {๐›ผ + ๐‘” ′ : ๐‘” ′ ∈ ๐‘”}.
Thus, ๐‘ฅ = ๐›ผ +๐‘” ′ for some ๐‘” ′ ∈ ๐‘” ∈ ๐›พ. Conversely, if ๐‘ฅ ∈ ๐›ผ, then ๐‘ฅ ∈ ๐›ผ +0,
so ๐‘ฅ ∈ ๐›ผ + ๐›พ. For the other case, let ๐‘ฅ ∈ {๐›ผ + ๐‘” : ๐‘” ∈ ๐›พ}. This implies that
๐‘ฅ ∈ (๐›ผ + ๐‘”)+ = ๐›ผ + ๐‘” + for some ๐‘” ∈ ๐›พ. Since ๐‘” + ∈ ๐›พ,
โ‹ƒ
๐‘ฅ∈
{๐›ผ + ๐‘” ′ : ๐‘” ′ ∈ ๐›พ} = ๐›ผ + ๐›พ.
Although ordinal addition is not commutative, it does have other familiar properties
as noted in the next result.
Section 6.4 ARITHMETIC
319
THEOREM 6.4.4
Addition of ordinals is associative, and 0 is the additive identity.
PROOF
Let ๐œ be a limit ordinal. Define
๐ด = {๐›ฝ ∈ ๐œ : 0 + ๐›ฝ = ๐›ฝ}.
Suppose ๐›พ is an ordinal such that ๐—Œ๐–พ๐—€(๐œ†, ๐›พ) ⊆ ๐ด. We have two cases to check.
โˆ™ Let ๐›พ = ๐›ฟ + for some ordinal ๐›ฟ. Then,
0 + ๐›พ = 0 + ๐›ฟ + = (0 + ๐›ฟ)+ = ๐›ฟ + = ๐›พ.
โˆ™ Let ๐›พ be a limit ordinal. We then have
โ‹ƒ
โ‹ƒ
0+๐›พ =
(0 + ๐›ฟ) =
๐›ฟ = ๐›พ.
๐›ฟ∈๐›พ
๐›ฟ∈๐›พ
In both cases, ๐›พ ∈ ๐ด, so by transfinite induction, ๐ด = ๐œ. Since we can also prove
that
๐œ = {๐›ฝ ∈ ๐œ : ๐›ฝ + 0 = ๐›ฝ},
we conclude that 0 is the additive identity.
To prove that ordinal addition is associative, we proceed by transfinite induction. Let ๐›ผ, ๐›ฝ, ๐›ฟ ∈ ๐œ. Define
๐ต = {๐‘ง ∈ ๐œ : ๐›ผ + (๐›ฝ + ๐‘ง) = (๐›ผ + ๐›ฝ) + ๐‘ง}.
Assume that ๐—Œ๐–พ๐—€(๐œ, ๐›พ) ⊆ ๐ต. Then, ๐›พ ∈ ๐ต because by Lemma 6.4.3, we have
โ‹ƒ
๐›ผ + (๐›ฝ + ๐›พ) = ๐›ผ ∪ {๐›ผ + ๐‘ฅ : ๐‘ฅ ∈ ๐›ฝ + ๐›พ}
โ‹ƒ
= ๐›ผ ∪ {๐›ผ + ๐‘ฅ : ๐‘ฅ ∈ ๐›ฝ ∨ ∃๐‘”(๐‘” ∈ ๐›พ → ๐‘ฅ = ๐›ฝ + ๐‘”)}
โ‹ƒ
= ๐›ผ ∪ ({๐›ผ + ๐‘ฅ : ๐‘ฅ ∈ ๐›ฝ} ∪ {๐›ผ + (๐›ฝ + ๐‘”) : ๐‘” ∈ ๐›พ})
โ‹ƒ
โ‹ƒ
= ๐›ผ ∪ {๐›ผ + ๐‘ฅ : ๐‘ฅ ∈ ๐›ฝ} ∪ {(๐›ผ + ๐›ฝ) + ๐‘” : ๐‘” ∈ ๐›พ}
โ‹ƒ
= (๐›ผ ∪ ๐›ฝ) ∪ {๐›ผ + ๐›ฝ + ๐‘” : ๐‘” ∈ ๐›พ}
= (๐›ผ + ๐›ฝ) + ๐›พ.
Note that Exercise 3.4.28(e) is used on the fourth equality.
DEFINITION 6.4.5
Let ๐œ be a limit ordinal. For all ๐›ผ, ๐›ฝ ∈ ๐œ,
โˆ™ ๐›ผ ⋅ 0 = 0,
320
Chapter 6 ORDINALS AND CARDINALS
โˆ™ ๐›ผ ⋅ ๐›ฝ + = ๐›ผ ⋅ ๐›ฝ + ๐›ผ,
โ‹ƒ
โˆ™ ๐›ผ ⋅ ๐›ฝ = {๐›ผ ⋅ ๐›ฟ : ๐›ฟ ∈ ๐›ฝ} if ๐›ฝ is a limit ordinal.
As with ordinal addition, ordinal multiplication is well-defined by transfinite recursion
(Exercise 3). Also, as with addition of ordinals, certain expected properties hold, while
others do not. The next two results are the analog of Lemma 6.4.3 and Theorem 6.4.4.
Their proofs are Exercise 4.
LEMMA 6.4.6
If ๐›ผ and ๐›ฝ are ordinals, ๐›ผ ⋅ ๐›ฝ = {๐›ผ ⋅ ๐‘ + ๐‘Ž : ๐‘ ∈ ๐›ฝ ∧ ๐‘Ž ∈ ๐›ผ}.
THEOREM 6.4.7
Multiplication of ordinals is associative, and 1 is the multiplicative identity.
Observe that
0⋅๐œ”=
โ‹ƒ
{0 ⋅ ๐‘› : ๐‘› ∈ ๐œ”} = 0.
Also, ๐œ” ⋅ 1 = 1 ⋅ ๐œ” = ๐œ” (Theorem 6.4.7), so multiplication on the right by a natural
number behaves as we would expect in that
๐œ”⋅2=๐œ”⋅1+๐œ”=๐œ”+๐œ”
(6.1)
and
๐œ” ⋅ 3 = ๐œ” ⋅ 2 + ๐œ” = ๐œ” + ๐œ” + ๐œ”,
but
2⋅๐œ”=
โ‹ƒ
{2 ⋅ ๐‘› : ๐‘› ∈ ๐œ”} = ๐œ”
3⋅๐œ”=
โ‹ƒ
{3 ⋅ ๐‘› : ๐‘› ∈ ๐œ”} = ๐œ”.
and
Hence, multiplication of ordinals is not commutative. Because of this, it is not surprising that there are issues with the distributive law. For ordinals, there is a left distributive
law but not a right distributive law (Exercise 9).
THEOREM 6.4.8 [Left Distributive Law]
๐›ผ ⋅ (๐›ฝ + ๐›ฟ) = ๐›ผ ⋅ ๐›ฝ + ๐›ผ ⋅ ๐›ฟ for all ordinals ๐›ผ, ๐›ฝ, and ๐›ฟ.
Since addition of ordinals is an operation on a limit ordinal ๐œ, we know that for all
ordinals ๐›ผ, ๐›ฝ, ๐›ฟ ∈ ๐œ,
๐›ผ =๐›ฝ ⇒๐›ผ+๐›ฟ =๐›ฝ+๐›ฟ
and
๐›ผ = ๐›ฝ ⇒ ๐›ผ ⋅ ๐›ฟ = ๐›ฝ ⋅ ๐›ฟ.
The next result gives information regarding how ordinal multiplication behaves with
an inequality.
Section 6.4 ARITHMETIC
321
THEOREM 6.4.9
Let ๐›ผ, ๐›ฝ, and ๐›ฟ be ordinals.
โˆ™ If ๐›ผ ⊂ ๐›ฝ, then ๐›ฟ + ๐›ผ ⊂ ๐›ฟ + ๐›ฝ.
โˆ™ If ๐›ผ ⊆ ๐›ฝ, then ๐›ผ + ๐›ฟ ⊆ ๐›ฝ + ๐›ฟ.
โˆ™ If ๐›ผ ⊂ ๐›ฝ and ๐›ฟ ≠ 0, then ๐›ฟ ⋅ ๐›ผ ⊂ ๐›ฟ ⋅ ๐›ฝ.
โˆ™ If ๐›ผ ⊆ ๐›ฝ, then ๐›ผ ⋅ ๐›ฟ ⊆ ๐›ฝ ⋅ ๐›ฟ.
PROOF
We prove the third part, leaving the others to Exercise 7. Let ๐›ผ ⊂ ๐›ฝ and ๐›ฟ ≠ 0.
Then, by Lemma 6.4.6,
๐›ฟ ⋅ ๐›ผ = {๐›ฟ ⋅ ๐‘Ž + ๐‘‘ : ๐‘Ž ∈ ๐›ผ ∧ ๐‘‘ ∈ ๐›ฟ} ⊂ {๐›ฟ ⋅ ๐‘ + ๐‘‘ : ๐‘ ∈ ๐›ฝ ∧ ๐‘‘ ∈ ๐›ฟ} = ๐›ฟ ⋅ ๐›ฝ.
Finally, we define exponentiation so that it generalizes exponentiation on ๐œ” (Exercise 5.2.17). It is a binary operation (Exercise 3).
DEFINITION 6.4.10
Let ๐œ be a limit ordinal and ๐›ผ, ๐›ฝ ∈ ๐œ.
โˆ™ ๐›ผ0 = 1
+
โˆ™ ๐›ผ๐›ฝ = ๐›ผ๐›ฝ ⋅ ๐›ผ
โˆ™ ๐›ผ๐›ฝ =
โ‹ƒ ๐›ฟ
{๐›ผ : ๐›ฟ ∈ ๐›ฝ} if ๐›ฝ is a limit ordinal.
For example,
+
๐œ”1 = ๐œ”0 = ๐œ”0 ⋅ ๐œ” = 1 ⋅ ๐œ” = ๐œ”,
and
+
๐œ”2 = ๐œ”1 = ๐œ”1 ⋅ ๐œ” = ๐œ” ⋅ ๐œ”,
so raising an ordinal to a natural number appears to behave as expected. Also,
1๐œ” =
โ‹ƒ
โ‹ƒ
{1๐‘› : ๐‘› ∈ ๐œ”} =
{1} = 1
and
2๐œ” =
โ‹ƒ
{2๐‘› : ๐‘› ∈ ๐œ”} = ๐œ”.
We leave the proof of the following properties of ordinal exponentiation to Exercise 11.
322
Chapter 6 ORDINALS AND CARDINALS
THEOREM 6.4.11
Let ๐›ผ, ๐›ฝ, ๐›ฟ be ordinals.
โˆ™ ๐›ผ ๐›ฝ+๐›ฟ = ๐›ผ ๐›ฝ ⋅ ๐›ผ ๐›ฟ .
โˆ™ (๐›ผ ๐›ฝ )๐›ฟ = ๐›ผ ๐›ฝ⋅๐›ฟ .
Cardinals
Even though the finite cardinals are the same sets as the finite ordinals and every infinite
cardinal is a limit ordinal, the arithmetic defined on the cardinals will only apply when
the given sets are viewed as cardinals. The definitions for addition, multiplication, and
exponentiation for cardinals are not given recursively.
DEFINITION 6.4.12
Let ๐œ… and ๐œ† be cardinals.
โˆ™ ๐œ… + ๐œ† = |(๐œ… × {0}) ∪ (๐œ† × {1})|
โˆ™ ๐œ… ⋅ ๐œ† = |๐œ… × ๐œ†|
โˆ™ ๐œ… ๐œ† = |๐œ† ๐œ…|.
Since ordinal arithmetic was simply a generalization of the arithmetic of natural numbers, that the addition worked in the finite case was not checked. Here this is not the
case, so let us check
2 + 3 = |(2 × {0}) ∪ (3 × {1})| = |{(0, 0), (1, 0), (0, 1), (1, 1), (2, 1)}| = 5.
Also,
3 + 2 = |(3 × {0}) ∪ (2 × {1})| = |{(0, 0), (1, 0), (2, 0), (0, 1), (1, 1)}| = 5.
This suggests that cardinal addition is commutative. This and other basic results are
given in the next theorem. Some details of the proof are left to Exercise 14.
THEOREM 6.4.13
Addition of cardinals is associative and commutative, and 0 is the additive identity.
PROOF
Let ๐œ…, ๐œ†, and ๐œ‡ be cardinals. Addition is associative because
(๐œ… × {0}) ∪ [([๐œ† × {0}] ∪ [๐œ‡ × {1}]) × {1}]
is equinumerous to
[([๐œ… × {0}] ∪ [๐œ† × {1}]) × {0}] ∪ (๐œ‡ × {1}),
Section 6.4 ARITHMETIC
323
it is commutative because
(๐œ… × {0}) ∪ (๐œ† × {1}) ≈ (๐œ† × {1}) ∪ (๐œ… × {0}),
and 0 is the additive identity because
(๐œ… × {0}) ∪ (0 × {1}) = (๐œ… × {0}).
Now let us multiply
2 ⋅ 3 = |{0, 1} × {0, 1, 2}| = |{(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2)}| = 6
and
3 ⋅ 2 = |{0, 1, 2} × {0, 1}| = |{(0, 0), (0, 1), (1, 0), (1, 1), (2, 0), (2, 1)}| = 6.
As with cardinal addition, it seems that cardinal multiplication is commutative. This
and other results are stated in the next theorem. Its proof is left to Exercise 15.
THEOREM 6.4.14
Multiplication of cardinals is associative and commutative, and 1 is the multiplicative identity.
As with ordinal arithmetic (Theorem 6.4.8), cardinal arithmetic has a left distribution
law, but since cardinal multiplication is commutative, cardinal arithmetic also has a
right distribution law.
THEOREM 6.4.15 [Distributive Law]
Let ๐œ…, ๐œ†, and ๐œ‡ be cardinals.
โˆ™ ๐œ… ⋅ (๐œ† + ๐œ‡) = ๐œ… ⋅ ๐œ† + ๐œ… ⋅ ๐œ‡.
โˆ™ (๐œ… + ๐œ†) ⋅ ๐œ‡ = ๐œ… ⋅ ๐œ‡ + ๐œ† ⋅ ๐œ‡.
PROOF
The left distribution law holds because
๐œ… × ([๐œ† × {0}] ∪ [๐œ‡ × {1}]) ≈ ([๐œ… × ๐œ†] × {0}) ∪ ([๐œ… × ๐œ‡] × {1}).
The remaining details of the proof are left to Exercise 16.
The last of the operations of Definition 6.4.12 is exponentiation. Let ๐œ… be a cardinal.
Observe that since there is exactly one function 0 → ๐œ… (Exercise 4.4.16),
๐œ… 0 = |0 ๐œ…| = 1
and if ๐œ… ≠ 0,
0๐œ… = |๐œ… 0| = 0
because there are no functions ๐œ… → 0, and by Theorem 6.2.13,
๐œ… โ‰บ 2๐œ… .
In addition, cardinal exponentiation follows other expected rules.
324
Chapter 6 ORDINALS AND CARDINALS
THEOREM 6.4.16
Let ๐œ…, ๐œ†, and ๐œ‡ be cardinals.
โˆ™ ๐œ… ๐œ†+๐œ‡ = ๐œ… ๐œ† ⋅ ๐œ… ๐œ‡ .
โˆ™ (๐œ… ⋅ ๐œ†)๐œ‡ = ๐œ… ๐œ‡ ⋅ ๐œ†๐œ‡ .
โˆ™ (๐œ… ๐œ† )๐œ‡ = ๐œ… ๐œ†⋅๐œ‡ .
PROOF
We prove the last part and leave the rest to Exercise 17. Define
( )
๐œ‘ : ๐œ‡ ๐œ† ๐œ… → ๐œ†×๐œ‡ ๐œ…
such that for all ๐œ“ ∈
(
๐œ‡ ๐œ†๐œ…
)
, ๐›ผ ∈ ๐œ† and ๐›ฝ ∈ ๐œ‡,
๐œ‘(๐œ“)(๐›ผ, ๐›ฝ) = ๐œ“(๐›ผ)(๐›ฝ).
We claim that ๐œ‘ is a bijection.
( )
โˆ™ Let ๐œ“1 , ๐œ“2 ∈ ๐œ‡ ๐œ† ๐œ… . Assume that ๐œ‘(๐œ“1 ) = ๐œ‘(๐œ“2 ). Take ๐›ผ ∈ ๐œ† and ๐›ฝ ∈ ๐œ‡.
Then,
๐œ“1 (๐›ผ)(๐›ฝ) = ๐œ‘(๐œ“1 )(๐›ผ, ๐›ฝ) = ๐œ‘(๐œ“2 )(๐›ผ, ๐›ฝ) = ๐œ“2 (๐›ผ)(๐›ฝ).
Therefore, ๐œ“1 = ๐œ“2 , and ๐œ‘ is one-to-one.
โˆ™ Let ๐œ“ ∈ ๐œ†×๐œ‡ ๐œ…. For ๐›ผ ∈ ๐œ† and ๐›ฝ ∈ ๐œ‡, define ๐œ“ ′ (๐›ผ)(๐›ฝ) = ๐œ“(๐›ผ, ๐›ฝ). This
implies that ๐œ‘(๐œ“ ′ ) = ๐œ“, so ๐œ‘ is onto.
Since every infinite cardinal number can be represented by an ℵ, let us determine
how to calculate using this notation. We begin with a lemma.
LEMMA 6.4.17
If ๐‘› ∈ ๐œ” and ๐œ… an infinite cardinal, ๐‘› + ๐œ… = ๐‘› ⋅ ๐œ… = ๐œ….
PROOF
Let ๐‘› be a natural number. Define ๐œ‘ : (๐‘› × {0}) ∪ (๐œ… × {1}) → ๐œ… by
๐œ‘(๐‘–, 0) = ๐‘– for all ๐‘– < ๐‘›,
๐œ‘(๐›ผ, 1) = ๐‘› + ๐›ผ for all ๐›ผ ∈ ๐œ….
For example, if ๐‘› = 5, then ๐œ‘(4, 0) = 4, ๐œ‘(0, 1) = 5, ๐œ‘(6, 1) = 11, ๐œ‘(๐œ”, 1) = ๐œ”,
and ๐œ‘(๐œ” + 1, 1) = ๐œ” + 1. Therefore, ๐‘› + ๐œ… = ๐œ… because ๐œ‘ is a bijection. That
๐‘› ⋅ ๐œ… = ๐œ… is left to Exercise 18.
Lemma 6.4.17 allows us to compute with alephs.
Section 6.4 ARITHMETIC
325
THEOREM 6.4.18
Let ๐›ผ and ๐›ฝ be ordinals and ๐‘› ∈ ๐œ”.
โˆ™ ๐‘› + ℵ๐›ผ = ๐‘› ⋅ ℵ๐›ผ = ℵ๐›ผ .
{
ℵ๐›ผ
โˆ™ ℵ๐›ผ + ℵ๐›ฝ = ℵ๐›ผ ⋅ ℵ๐›ฝ =
ℵ๐›ฝ
if ๐›ผ ≥ ๐›ฝ,
otherwise.
PROOF
The first part follows by Lemma 6.4.17. To prove the addition equation from
the second part, let ๐›ผ and ๐›ฝ be ordinals. Without loss of generality assume that
๐›ผ ⊆ ๐›ฝ. By Definition 6.4.12,
ℵ๐›ผ + ℵ๐›ฝ = |(ℵ๐›ผ × {0}) ∪ (ℵ๐›ฝ × {1})|.
Since ℵ๐›ฝ = |ℵ๐›ฝ × {1}|,
ℵ๐›ฝ โชฏ |(ℵ๐›ผ × {0}) ∪ (ℵ๐›ฝ × {1})|.
Furthermore, because ℵ๐›ผ ⊆ ℵ๐›ฝ ,
|(ℵ๐›ผ × {0}) ∪ (ℵ๐›ฝ × {1})| โชฏ |(ℵ๐›ฝ × {0}) ∪ (ℵ๐›ฝ × {1})|
= |ℵ๐›ฝ × {0, 1}|
= ℵ๐›ฝ .
Because of Lemma 6.4.17, the last equality holds since ℵ๐›ฝ is infinite and {0, 1}
is finite. Hence,
ℵ๐›ผ + ℵ๐›ฝ ≈ ℵ๐›ฝ
by the Cantor–Schröder–Bernstein theorem (6.2.11). Since both ℵ๐›ผ + ℵ๐›ฝ and ℵ๐›ฝ
are cardinals, ℵ๐›ผ + ℵ๐›ฝ = ℵ๐›ฝ .
For example, ℵ5 + ℵ9 = ℵ5 ⋅ ℵ9 = ℵ9 . More generally, we quickly have the following
corollary by Theorem 6.3.19.
COROLLARY 6.4.19
For every infinite cardinal ๐œ…, both ๐œ… + ๐œ… = ๐œ… and ๐œ… ⋅ ๐œ… = ๐œ….
Exercises
1. Let ๐’œ = {๐ด0 , ๐ด1 , … , ๐ด๐‘›−1 } be a pairwise disjoint
family of sets. Assuming that
โ‹ƒ
the sets are distinct, prove that the cardinality of ๐’œ is equal to the sum
|๐ด0 | + |๐ด1 | + · · · + |๐ด๐‘›−1 |.
2. Let ๐›ผ and ๐›ฝ be ordinals. Let ๐œ‘ : ๐›ผ → ๐›ฝ be a function such that ๐œ‘(๐›ฟ) ∈ ๐œ‘(๐›พ) for all
๐›ฟ ∈ ๐›พ ∈ ๐›ผ (Compare Lemma 6.1.2). Prove the following.
326
Chapter 6 ORDINALS AND CARDINALS
(a) ๐›ผ ⊆ ๐›ฝ.
(b) ๐›ฟ ⊆ ๐œ‘(๐›ฟ) for all ๐›ฟ ∈ ๐›ผ.
3. Let ๐œ be a limit ordinal. Use transfinite recursion to prove that ordinal multiplication
(Definition 6.4.5) and ordinal exponentiation (Definition 6.4.10) are binary operations
on ๐œ.
4. Prove Lemma 6.4.6 and Theorem 6.4.7.
5. For every ordinal ๐›ผ, prove that 0 ⋅ ๐›ผ = 0.
6. Prove that for all ๐‘› ∈ ๐œ”, ๐‘› + ๐œ” = ๐‘› ⋅ ๐œ” as ordinals. Can this be generalized to
๐›ผ + ๐›ฝ = ๐›ผ ⋅ ๐›ฝ for ordinals ๐›ผ ∈ ๐›ฝ with ๐›ฝ being infinite? If so, is the ๐›ผ ∈ ๐›ฝ required?
7. Prove the remaining parts of Theorem 6.4.9.
8. Find ordinals ๐›ผ, ๐›ฝ, and ๐›ฟ such that the following properties hold.
(a) ๐›ผ ⊂ ๐›ฝ but ๐›ฝ + ๐›ฟ ⊆ ๐›ผ + ๐›ฟ.
(b) ๐›ผ ⊂ ๐›ฝ but ๐›ฝ ⋅ ๐›ฟ ⊆ ๐›ผ ⋅ ๐›ฟ.
9. Let ๐›ผ, ๐›ฝ, and ๐›ฟ be ordinals.
(a) Prove that ๐›ผ ⋅ (๐›ฝ + ๐›ฟ) = ๐›ผ ⋅ ๐›ฝ + ๐›ผ ⋅ ๐›ฟ.
(b) Show that it might be the case that (๐›ผ + ๐›ฝ) ⋅ ๐›ฟ ฬธ≠ ๐›ผ ⋅ ๐›ฟ + ๐›ฝ ⋅ ๐›ฟ.
(c) For which ordinals does the right distribution law hold?
10. Let ๐›ผ, ๐›ฝ, and ๐›ฟ be ordinals. Prove the following.
(a) ๐›ผ + ๐›ฝ ∈ ๐›ผ ⋅ ๐›ฟ if and only if ๐›ฝ ∈ ๐›ฟ.
(b) ๐›ผ + ๐›ฝ = ๐›ผ + ๐›ฟ if and only if ๐›ฝ = ๐›ฟ.
(c) If ๐›ผ + ๐›ฟ ∈ ๐›ฝ + ๐›ฟ, then ๐›ผ ∈ ๐›ฝ.
(d) ๐›ผ ⋅ ๐›ฝ ∈ ๐›ผ ⋅ ๐›ฟ if and only if ๐›ฝ ∈ ๐›ฟ and ๐›ผ ≠ 0.
(e) If ๐›ผ ⋅ ๐›ฝ = ๐›ผ ⋅ ๐›ฟ, then ๐›ฝ = ๐›ฟ or ๐›ผ = 0.
11. Prove Theorem 6.4.11.
12. Let ๐›ผ, ๐›ฝ, and ๐›ฟ be ordinals. Prove the following.
(a) ๐›ผ ๐›ฝ ∈ ๐›ผ ๐›ฟ if and only if ๐›ฝ ∈ ๐›ผ and 1 ∈ ๐›ผ.
(b) If ๐›ผ ∈ ๐›ฝ, then ๐›ผ ๐›ฟ ⊆ ๐›ฝ ๐›ฟ .
(c) If ๐›ผ ๐›ฟ ∈ ๐›ฝ ๐›ฟ , then ๐›ผ ∈ ๐›ฝ.
(d) If 1 ∈ ๐›ผ, then ๐›ฝ ⊆ ๐›ผ ๐›ฝ .
(e) If ๐›ผ ∈ ๐›ฝ, there exists a unique ordinal ๐›พ such that ๐›ผ + ๐›พ = ๐›ฝ.
13. Prove that the ordinal ๐œ” + ๐œ” is not a cardinal.
14. Provide the details for the proof of Theorem 6.4.13.
15. Prove Theorem 6.4.14.
16. Provide the details to the proof of Theorem 6.4.15.
17. Prove the remaining parts of Theorem 6.4.16.
18. Prove that for any natural number ๐‘› and infinite cardinal ๐œ…, ๐‘› ⋅ ๐œ… = ๐œ….
327
Section 6.5 LARGE CARDINALS
19. Prove that if ๐œ… is inifinte, (๐œ… + )๐œ… = 2๐œ… .
20. Is there a cancellation law for ordinals or for cardinals?
21. Prove that ๐œ… +๐œ† = ๐œ… ⋅๐œ† = ๐œ†, given that ๐œ… is a countable cardinal and ๐œ† is an infinite
cardinal.
22. Generalize Exercise 21 by showing that if ๐œ… and ๐œ† are cardinals with ๐œ† infinite
such that ๐œ… โชฏ ๐œ†, then ๐œ… + ๐œ† = ๐œ… ⋅ ๐œ† = ๐œ†.
23. Let ๐œ… and ๐œ† be cardinals with ℵ0 โชฏ ๐œ†. Show that if 2 โชฏ ๐œ… โ‰บ ๐œ†, then ๐œ… ๐œ† = 2๐œ† .
24. Prove for all ordinals ๐›ผ that |๐›ผ| โ‰บ ℵ๐›ผ .
25. For all ordinals ๐›ผ, define Hartogs’ function by
Γ(๐›ผ) = {๐›ฝ : ๐›ฝ is an ordinal ∧ ๐›ฝ โชฏ ๐›ผ}.
Prove that Γ(๐›ผ) is an ordinal and ๐›ผ โ‰บ Γ(๐›ผ) for all ordinals ๐›ผ.
26. Using Exercise 25, define an initial number ๐œ”๐›ผ as follows.
๐œ”0 = ๐œ”,
๐œ”๐›พ + = Γ(๐œ”๐›พ ),
โ‹ƒ
๐œ”๐›พ =
๐œ”๐›ฟ if ๐›พ is a limit ordinal.
๐›ฟ∈๐›พ
Prove that initial numbers are limit ordinals and ๐œ”1 is the first uncountable ordinal.
27. Prove that there is no greatest initial number.
28. For all ordinals ๐›ผ, show that ℵ๐›ผ = |๐œ”๐›ผ |. Can we write ℵ๐›ผ = ๐œ”๐›ผ ?
29. For all countable ordinals ๐›ผ, show that 2ℵ๐›ผ = ℵ๐›ผ+ implies that 2ℵ๐œ”1 = ℵ๐œ”1 +1 .
6.5
LARGE CARDINALS
Since every cardinal is a limit ordinal, every cardinal ๐œ… can be written in the form
โ‹ƒ
๐œ…=
{๐›ผ : ๐›ผ ∈ ๐œ…}.
In particular, for the limit cardinal ℵ๐œ”+๐œ” , we have that
โ‹ƒ
ℵ๐œ”+๐œ” =
{๐›ผ : ๐›ผ ∈ ℵ๐œ”+๐œ” }.
(6.2)
Notice that (6.2) is the union of a set with ℵ๐œ”+๐œ” elements. However, we also have
โ‹ƒ
ℵ๐œ”+๐œ” =
{ℵ๐›ผ : ๐›ผ ∈ ๐œ” + ๐œ”},
(6.3)
and
ℵ๐œ”+๐œ” =
โ‹ƒ
{ℵ๐œ”+๐‘› : ๐‘› ∈ ๐œ”}.
(6.4)
Both (6.3) and (6.4) are unions of sets with ℵ0 elements. The next definition is introduced to handle these differences. Since infinite cardinals are limit ordinals, the
definition is given for limit ordinals.
328
Chapter 6 ORDINALS AND CARDINALS
DEFINITION 6.5.1
The cofinality of a limit ordinal ๐›ผ is denoted by ๐–ผ๐–ฟ (๐›ผ) and defined as the least
โ‹ƒ
cardinal ๐œ† such that there exists โ„ฑ ⊆ ๐›ผ with |โ„ฑ | = ๐œ† and ๐›ผ = โ„ฑ .
โ‹ƒ
Observe
๐›ผ. There can be other sets ๐ต such that
โ‹ƒ that ๐–ผ๐–ฟ(๐›ผ) โชฏ |๐›ผ| because ๐›ผ =
๐›ผ = ๐ต, and they might have different cardinalities. Any set of ordinals ๐ต with this
property is said to be cofinal in ๐›ผ. Moreover, we can write ๐ต = {๐›ฝ๐›ฟ : ๐›ฟ ∈ ๐œ…} for some
cardinal ๐œ…, so
โ‹ƒ
๐›ผ=
๐›ฝ๐›ฟ ,
๐›ฟ∈๐œ…
and when ๐œ… = ๐–ผ๐–ฟ (๐›ผ),
โ‹ƒ
๐›ผ=
๐›ฝ๐›ฟ .
๐›ฟ∈๐–ผ๐–ฟ(๐›ผ)
EXAMPLE 6.5.2
Since the finite union of a finite set is finite, the cofinality of any infinite set must
be infinite. Therefore, because
โ‹ƒ
๐œ”=
๐‘›,
๐‘›∈๐œ”
we see that ๐–ผ๐–ฟ (๐œ”) = ℵ0 . Also, since
ℵ๐œ” =
โ‹ƒ
ℵ๐‘› ,
๐‘›∈๐œ”
we conclude that ๐–ผ๐–ฟ (ℵ๐œ” ) = ℵ0 and {ℵ๐‘› : ๐‘› ∈ ๐œ”} is cofinal in ℵ๐œ” . However,
๐–ผ๐–ฟ (ℵ1 ) = ℵ1 because the countable union of countable sets is countable (Theorem 6.3.17).
Regular and Singular Cardinals
Since infinite cardinals are limit ordinals, we can classify the cardinals based on their
cofinalities. We make the following definition.
DEFINITION 6.5.3
A cardinal ๐œ… is regular if ๐œ… = ๐–ผ๐–ฟ (๐œ…), else it is singular.
Notice that Example 6.5.2 shows that ℵ0 and ℵ1 are regular but ℵ๐œ” is singular. This
implies a direction to follow to characterize the cardinal numbers. We begin with the
successors.
THEOREM 6.5.4
Successor cardinals are regular.
Section 6.5 LARGE CARDINALS
PROOF
Let ๐›ผ be an ordinal and let โ„ฑ ⊆ ℵ๐›ผ+1 such that ℵ๐›ผ+1 =
|๐›ฝ| โชฏ ℵ๐›ผ for all ๐›ฝ ∈ โ„ฑ . Thus,
โ‹ƒ
โ„ฑ โชฏ |โ„ฑ | ⋅ ℵ๐›ผ .
โ‹ƒ
329
โ„ฑ . This implies that
By Theorem 6.4.18, we conclude that ℵ๐›ผ+1 โชฏ |โ„ฑ |, so ๐–ผ๐–ฟ (ℵ๐›ผ+1 ) = ℵ๐›ผ+1 by the
Cantor–Schröder–Bernstein theorem (6.2.11).
Since ℵ0 is regular, Theorem 6.5.4 tells us where to find the singular cardinals.
COROLLARY 6.5.5
A singular cardinal is an uncountable limit cardinal.
Because we have not proved the converse of Corollary 6.5.5, we investigate the cofinality of certain limit cardinals in an attempt to determine which limit cardinals are
singular.
THEOREM 6.5.6
If ๐›ผ is a limit ordinal, ๐–ผ๐–ฟ (ℵ๐›ผ ) = ๐–ผ๐–ฟ(๐›ผ).
PROOF
The proof uses the Cantor–Schröder–Bernstein theorem (6.2.11). Let ๐›ผ be a limit
ordinal.
โˆ™ We first show that ๐–ผ๐–ฟ (ℵ๐›ผ ) โชฏ ๐–ผ๐–ฟ(๐›ผ). Let ๐ด be a cofinal subset of ๐›ผ such that
|๐ด| = ๐–ผ๐–ฟ(๐›ผ). Notice that if ๐›ฝ ∈ ๐ด, then ℵ๐›ฝ ∈ ℵ๐›ผ . On the other hand, take
๐›ฟ ∈ ℵ๐›ผ , which implies that |๐›ฟ| โ‰บ ℵ๐›ผ . Therefore, there exists ๐›พ ∈ ๐›ผ such
that |๐›ฟ| = ℵ๐›พ . Since ๐ด is cofinal in ๐›ผ, there
โ‹ƒ exists ๐œ ∈ ๐ด such that ๐›ฟ ∈ ๐œ.
Hence, ๐›ฟ ∈ ℵ๐œ , which implies that ๐›ฟ ∈ {ℵ๐›ฝ : ๐›ฝ ∈ ๐ด}. Therefore,
โ‹ƒ
ℵ๐›ผ =
{ℵ๐›ฝ : ๐›ฝ ∈ ๐ด},
from which follows,
๐–ผ๐–ฟ(ℵ๐›ผ ) โชฏ |๐ด| = ๐–ผ๐–ฟ(๐›ผ).
โˆ™ We now show ๐–ผ๐–ฟ (๐›ผ) โชฏ ๐–ผ๐–ฟ (ℵ๐›ผ ). Let ๐ด ⊆ ℵ๐›ผ so that ℵ๐›ผ =
๐–ผ๐–ฟ(ℵ๐›ผ ). Define
โ‹ƒ
โ‹ƒ
๐ด and |๐ด| =
โ„ฑ = {๐›ฟ ∈ ๐›ผ : ∃๐›พ(๐›พ ∈ ๐ด ∧ |๐›พ| = ℵ๐›ฟ )}.
Then, ๐›ฝ = โ„ฑ is an ordinal by Theorem 6.1.16. For all ๐œ ∈ ๐ด, we have
that ๐œ ∈ ℵ๐›ฝ+1 because |๐œ| โชฏ ℵ๐›ฝ . Hence,
โ‹ƒ
๐ด ⊆ ℵ๐›ฝ ,
โ‹ƒ
from which follows that ๐›ผ ∈ ๐›ฝ, which means that ๐›ผ ⊆ โ„ฑ . Therefore,
โ‹ƒ
since the elements of โ„ฑ are ordinals of ๐›ผ, we have that ๐›ผ = โ„ฑ , so
๐–ผ๐–ฟ(๐›ผ) โชฏ |โ„ฑ | = |๐ด| = ๐–ผ๐–ฟ (ℵ๐›ผ ).
330
Chapter 6 ORDINALS AND CARDINALS
Theorem 6.5.6 confirms the result of Example 6.5.2 because
๐–ผ๐–ฟ(ℵ๐œ” ) = ๐–ผ๐–ฟ(๐œ”) = ℵ0 .
(6.5)
๐–ผ๐–ฟ (ℵ๐œ”+๐œ” ) = ๐–ผ๐–ฟ(๐œ” + ๐œ”) = ℵ0 ,
(6.6)
Also,
so ℵ๐œ”+๐œ” is singular. Observe that by (6.5) and (6.6),
๐–ผ๐–ฟ (๐–ผ๐–ฟ(ℵ๐œ” )) = ๐–ผ๐–ฟ (๐–ผ๐–ฟ(๐œ”)) = ๐–ผ๐–ฟ (ℵ0 ) = ℵ0
and
๐–ผ๐–ฟ (๐–ผ๐–ฟ(ℵ๐œ”+๐œ” )) = ๐–ผ๐–ฟ (๐–ผ๐–ฟ(๐œ” + ๐œ”)) = ๐–ผ๐–ฟ(ℵ0 ) = ℵ0 .
The next result generalizes this and proves that ๐–ผ๐–ฟ(๐–ผ๐–ฟ (๐›ผ)) = ๐–ผ๐–ฟ (๐›ผ) for every limit ordinal
๐›ผ.
THEOREM 6.5.7
For any limit ordinal ๐›ผ, ๐–ผ๐–ฟ (๐›ผ) is a regular cardinal.
PROOF
โ‹ƒ
Let ๐›ผ be a limit ordinal โ‹ƒ
and write ๐›ผ = {๐ด๐›พ : ๐›พ ∈ ๐–ผ๐–ฟ (๐›ผ)}. For every ordinal
๐›พ ∈ ๐–ผ๐–ฟ (๐›ผ), define ๐›ผ๐›พ = ๐›ฟ∈๐›พ ๐ด๐›ฟ . Then, {๐›ผ๐›พ : ๐›พ ∈ ๐–ผ๐–ฟ (๐›ผ)} is a chain of ordinals
(Theorem 6.1.16) and
โ‹ƒ
๐›ผ=
{๐›ผ๐›พ : ๐›พ ∈ ๐–ผ๐–ฟ (๐›ผ)}.
Now write
๐–ผ๐–ฟ (๐›ผ) =
Define
โ‹ƒ
{๐›ฝ๐›พ : ๐›พ ∈ ๐–ผ๐–ฟ(๐–ผ๐–ฟ(๐›ผ))}.
๐ต = {๐›ผ๐›ฝ๐›พ : ๐›พ ∈ ๐–ผ๐–ฟ (๐–ผ๐–ฟ(๐›ผ))}.
Let ๐œ ∈ ๐›ผ. This implies that ๐œ ∈ ๐›ผ๐›พ0 for some ๐›พ0 ∈ ๐–ผ๐–ฟ (๐›ผ). Then, ๐›พ0 ∈ ๐›ฝ๐›พ1 for
some ๐›พ1 ∈ ๐–ผ๐–ฟ (๐–ผ๐–ฟ(๐›ผ)). Hence,
๐œ ∈ ๐›ผ๐›พ0 ⊆ ๐›ผ๐›ฝ๐›พ ∈ ๐ต,
1
โ‹ƒ
โ‹ƒ
so ๐œ ∈ ๐ต. Therefore, ๐›ผ = ๐ต, and this implies that ๐–ผ๐–ฟ(๐›ผ) โชฏ ๐–ผ๐–ฟ (๐–ผ๐–ฟ(๐›ผ)). Since
the opposite inequality always holds, by the Cantor–Schröder–Bernstein theorem
(6.2.11), ๐–ผ๐–ฟ (๐›ผ) = ๐–ผ๐–ฟ (๐–ผ๐–ฟ(๐›ผ)), which means that ๐–ผ๐–ฟ (๐›ผ) is regular.
Although the Continuum Hypothesis cannot be proved, it is possible to discover
some information about the value of 2ℵ0 . Notice how its proof resembles Cantor’s
diagonalization (page 303). It is due to König (1905).
THEOREM 6.5.8 [König]
If ๐œ… is an infinite cardinal and ๐–ผ๐–ฟ (๐œ…) โชฏ ๐œ†, then ๐œ… โ‰บ ๐œ… ๐œ† .
Section 6.5 LARGE CARDINALS
331
PROOF
Suppose that ๐œ… is am infinite cardinal number and ๐–ผ๐–ฟ (๐œ…) = ๐œ†. Write
โ‹ƒ
๐œ…=
{๐›ฟ๐›ผ : ๐›ผ ∈ ๐œ†}.
Let โ„ฑ = {๐‘“๐›ฝ : ๐›ฝ ∈ ๐œ…} be a subset of ๐œ† ๐œ…. Define ๐‘” : ๐œ† → ๐œ… such that
๐‘”(๐›ผ) = least element of ๐œ… โงต {๐‘“๐›ฝ (๐›ฟ๐›ผ ) : ๐›ฝ ∈ ๐›ฟ๐›ผ }.
For any ๐›ผ ∈ ๐œ†,
๐‘”(๐›ผ) ≠ ๐‘“๐›ฝ (๐›ฟ๐›ผ ) for all ๐›ฝ ∈ ๐›ฟ๐›ผ .
Therefore,
๐‘” ≠ ๐‘“๐›ฝ for all ๐›ฝ ∈ ๐›ฟ๐›ผ .
(6.7)
Since (6.7) is true for all ๐›ผ ∈ ๐œ† and {๐›ฟ๐›ผ : ๐›ผ ∈ ๐œ†} is cofinal in ๐œ…, we conclude
that ๐‘” ≠ ๐‘“๐›ฝ for all ๐›ฝ ∈ ๐œ…. Therefore, ๐‘” ∉ โ„ฑ , so ๐œ… โ‰บ ๐œ… ๐œ† . Note that the same
argument leads to this conclusion if ๐–ผ๐–ฟ (๐œ…) โ‰บ ๐œ† (Exercise 4).
COROLLARY 6.5.9
Let ๐œ… be an infinite cardinal. Then, ๐œ… โ‰บ ๐–ผ๐–ฟ(2๐œ… ).
PROOF
Suppose that ๐–ผ๐–ฟ(2๐œ… ) โชฏ ๐œ…. By König’s theorem (6.5.8), Theorem 6.4.16, and
Corollary 6.4.19,
๐œ…
2๐œ… โ‰บ (2๐œ… )๐–ผ๐–ฟ(2 ) โชฏ (2๐œ… )๐œ… = 2๐œ…⋅๐œ… = 2๐œ… .
By Corollary 6.5.9,
ℵ0 โ‰บ ๐–ผ๐–ฟ(2ℵ0 ),
but by Example 6.5.2, we know that ๐–ผ๐–ฟ (ℵ๐œ” ) = ℵ0 , so
๐–ผ๐–ฟ(ℵ๐œ” ) โ‰บ ๐–ผ๐–ฟ(2ℵ0 ).
Hence, even though we cannot prove what the cardinality of 2ℵ0 is, we do know that it
is not ℵ๐œ” .
Inaccessible Cardinals
As we have noted, ℵ0 is both a regular and a limit cardinal. Are there any others with
this property?
DEFINITION 6.5.10
A regular limit cardinal that is uncountable is called weakly inaccessible.
It is not possible using the axioms of ๐—ญ๐—™๐—– to prove the existence of a weakly inaccessible cardinal. Here is another class of cardinals “beyond” the weakly inaccessible
cardinals.
332
Chapter 6 ORDINALS AND CARDINALS
DEFINITION 6.5.11
The cardinal ๐œ… is strongly inaccessible if it is an uncountable regular cardinal
such that 2๐œ† โ‰บ ๐œ… for all ๐œ† โ‰บ ๐œ….
Since every strongly inaccessible cardinal is weakly inaccessible (Exercise 1), it is
not possible to prove from the axioms of ๐—ญ๐—™๐—– that a strongly inaccessible cardinal
exists. However, it is apparent that assuming ๐–ฆ๐–ข๐–ง, ๐œ… is a weakly inaccessible cardinal
if and only if ๐œ… is a strongly inaccessible cardinal (Exercise 10). Cardinal numbers such
as these are known as large cardinals because assumptions beyond the axioms of ๐—ญ๐—™๐—–
are required to “reach” them.
Exercises
1. Prove that every strongly inaccessible cardinal is weakly inaccessible.
2. Let ๐‘› ∈ ๐œ”. Show that ๐–ผ๐–ฟ(ℵ๐‘› ) = ℵ๐‘› .
3. For all limit ordinals ๐›ผ, ๐›ฝ, and ๐›ฟ, show that if ๐›ผ is cofinal in ๐›ฝ and ๐›ฝ is cofinal in ๐›ฟ,
then ๐›ผ is cofinal in ๐›ฟ.
4. Rewrite the proof of König’s theorem (6.5.8) assuming that ๐–ผ๐–ฟ (๐œ…) โ‰บ ๐œ†.
5. Let ๐›ผ and ๐›ฝ be limit ordinals. Prove that ๐–ผ๐–ฟ (๐›ผ) = ๐–ผ๐–ฟ (๐›ผ) if and only if (๐›ผ, ⊆) and
(๐›ฝ, ⊆) have order isomorphic cofinal subsets.
6. Let ๐›ผ be a countable limit ordinal. Show that ๐–ผ๐–ฟ (๐›ผ) = ℵ0 .
7. Let ๐œ… and ๐œ† be cardinals such that ๐œ… is infinite and 2 โชฏ ๐œ†. Show the following.
(a) ๐œ… โ‰บ ๐–ผ๐–ฟ(๐œ†๐œ… ).
(b) ๐œ… โ‰บ ๐œ… ๐–ผ๐–ฟ(๐œ…) .
8. Assume ๐–ฆ๐–ข๐–ง and let ๐›ผ and ๐›ฝ be ordinals. Prove.
ℵ
(a) If ℵ๐›ฝ โ‰บ ๐–ผ๐–ฟ(ℵ๐›ผ ), then ℵ๐›ผ ๐›ฝ = ℵ๐›ผ .
ℵ
(b) If ๐–ผ๐–ฟ(ℵ๐›ผ ) โชฏ ℵ๐›ฝ โ‰บ ℵ๐›ผ , then ℵ๐›ผ ๐›ฝ = ℵ๐›ผ+ .
9. Let ๐›ผ be a limit ordinal. Show that ๐–ผ๐–ฟ (i๐›ผ ) = ๐–ผ๐–ฟ (๐›ผ). (See Exercise 6.3.22.)
10. Prove that ๐–ฆ๐–ข๐–ง implies that a cardinal is weakly inaccessible if and only if it is
strongly inaccessible.
11. Let ๐œ… be a cardinal. Prove the following biconditionals.
(a) ๐œ… is weakly inaccessible if and only if ๐œ… is regular and ℵ๐œ… = ๐œ….
(b) ๐œ… is strongly inaccessible if and only if ๐œ… is regular and i๐œ… = ๐œ….
CHAPTER 7
MODELS
7.1
FIRST-ORDER SEMANTICS
We now return to logic. In Section 1.5, we proved that propositional logic is both sound
and complete (Theorems 1.5.9 and 1.5.15). We now do the same for first-order logic.
We have an added complication in that this logic involves formulas with variables.
Sometimes the variables are all bound resulting in a sentence (Definition 2.2.14), but
other times the formula will have free occurrences. We need additional machinery to
handle this. Throughout this chapter, let ๐–  be a first-order alphabet and ๐–ฒ its set of
theory symbols. We start with the fundamental definition (compare Definition 4.1.1).
DEFINITION 7.1.1
The pair ๐”„ = (๐ด, ๐”ž) is an ๐—ฆ-structure if ๐ด ≠ ∅ and ๐”ž is a function with domain
๐–ฒ such that
โˆ™ ๐”ž(๐‘) is an element of ๐ด for every constant ๐‘ ∈ ๐–ฒ,
โˆ™ ๐”ž(๐‘…) is an ๐‘›-ary relation on ๐ด for every ๐‘›-ary relation symbol ๐‘… ∈ ๐–ฒ,
โˆ™ ๐”ž(๐‘“ ) is an ๐‘›-ary function on ๐ด for every ๐‘›-ary function symbol ๐‘“ ∈ ๐–ฒ.
333
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
334
Chapter 7 MODELS
The set ๐ด is the domain of ๐”„ and is denoted by dom(๐”„). The domain of the
function ๐”ž is the signature of the structure. If ๐–ฒ = {๐‘ 0 , ๐‘ 1 , ๐‘ 2 , … }, we often
write the structure as
(๐ด, ๐”ž(๐‘ 0 ), ๐”ž(๐‘ 1 ), ๐”ž(๐‘ 2 ), … )
or
(๐ด, ๐‘ ๐”„
, ๐‘ ๐”„ , ๐‘ ๐”„ , … )
0 1 2
The font used to identify a structure and its function is the traditional one. It is
called Fraktur and can be found in the appendix.
The purpose of the function ๐”ž is to associate a symbol with a particular object in
the domain of the structure. For this reason, if ๐‘  is an element of the signature of the
structure, the object ๐”ž(๐‘ ) is called the meaning of ๐‘  and the symbol ๐‘  is the name of
๐”ž(๐‘ ).
EXAMPLE 7.1.2
We are familiar with the constant, function, and relation symbols of ๐–ญ๐–ณ (Example 2.1.4). We define an ๐–ญ๐–ณ-structure ๐”„ with domain ๐œ”. To do this, specify the
function ๐”ž:
๐”ž(0) = ∅,
๐”ž(1) = {∅},
๐”ž(+) = {((๐‘š, ๐‘›), ๐‘š + ๐‘›) : ๐‘š, ๐‘› ∈ ๐œ”},
๐”ž(⋅) = {((๐‘š, ๐‘›), ๐‘š ⋅ ๐‘›) : ๐‘š, ๐‘› ∈ ๐œ”}.
Notice that ∅ is the meaning of 0 and 1 is the name of {∅}. Also, observe
that ๐”ž(+) and ๐”ž(⋅) are the addition and multiplication functions on ๐œ” (Definitions 5.2.15 and 5.2.18). These operations are usually represented by + and ⋅,
but these symbols already appear in ๐–ญ๐–ณ. Therefore, for the structure ๐”„,
0๐”„ = ๐”ž(0),
1๐”„ = ๐”ž(1),
+๐”„ = ๐”ž(+),
⋅๐”„ = ๐”ž(⋅),
so ๐”„ = (๐œ”, ๐”ž) is an ๐–ญ๐–ณ-structure with signature {0, 1, +, ⋅} that can be written as
(๐œ”, ∅, {∅}, {((๐‘š, ๐‘›), ๐‘š + ๐‘›) : ๐‘š, ๐‘› ∈ ๐œ”}, {((๐‘š, ๐‘›), ๐‘š ⋅ ๐‘›) : ๐‘š, ๐‘› ∈ ๐œ”})
or, more compactly,
(๐œ”, 0๐”„ , 1๐”„ , +๐”„ , ⋅๐”„ ).
Section 7.1 FIRST-ORDER SEMANTICS
335
EXAMPLE 7.1.3
When the symbols of a structure’s signature are not needed to represent the meanings of the symbols, the notation of Example 7.1.2 is not needed. This is common
for ๐–ฆ๐–ฑ-structures. For example, if ๐”Ÿ(๐‘’) = 0 and ๐”Ÿ(โ—ฆ) = +, then
(โ„ค, ๐”Ÿ) = (โ„ค, 0, +)
is a ๐–ฆ๐–ฑ-structure with signature {๐‘’, โ—ฆ} (Example 2.1.5), and if ๐” (๐‘’) = 1 and
๐” (โ—ฆ) = ⋅, then
(โ„ โงต {0}, ๐” ) = (โ„ โงต {0}, 1, ⋅)
is also a ๐–ฆ๐–ฑ-structure with signature {๐‘’, โ—ฆ}.
Satisfaction
The purpose of a structure is to serve as a universe for a given language. Recall that
the terms of a language represent objects and the formulas of a language describe the
properties of objects (Figure 2.3). With this in mind, we now make sense of the terms
and formulas within structures. The first step in doing this is to define how to give
meaning to terms. This is done so that the terms represent elements of the domain of
a structure. The second step is to develop a method by which it can be determined
which formulas hold true in a structure and which do not. We begin by defining the
interpretation of terms.
DEFINITION 7.1.4
Let ๐”„ = (๐ด, ๐”ž) be an ๐–ฒ-structure. Define an ๐—ฆ-interpretation of ๐”„ to be a function ๐ผ : ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐– ) → ๐ด that has the following properties:
โˆ™ If ๐‘ฅ is a variable symbol, then ๐ผ(๐‘ฅ) ∈ ๐ด, thus assigning a value to ๐‘ฅ.
โˆ™ If ๐‘ is a constant symbol, ๐ผ(๐‘) = ๐”ž(๐‘).
โˆ™ If ๐‘“ is a function symbol and ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 are ๐–ฒ-terms,
๐ผ(๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )) = ๐”ž(๐‘“ )(๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 )).
Definition 7.1.4 is an example of a definition by induction on terms. This means
that first the definition was made for variable and constant symbols, and then assuming
that it was made for terms in general, the definition is made for functions applied to
terms. Induction on terms simply follows Definition 2.1.7. A proof done by induction
on terms is one that uses the same process to prove a result about terms.
EXAMPLE 7.1.5
Let (๐œ”, 0๐”„ , 1๐”„ , +๐”„ , ⋅๐”„ ) be the ๐–ญ๐–ณ-structure from Example 7.1.2. Let ๐ผ be an
๐–ญ๐–ณ-interpretation such that
336
Chapter 7 MODELS
๐ผ(๐‘ฅ) = 5,
๐ผ(๐‘ฆ) = 7,
๐ผ(0) = 0๐”„ ,
๐ผ(1) = 1๐”„ .
Then, ๐ผ(๐‘ฅ + ๐‘ฆ) = 12 is the interpretation of ๐‘ฅ + ๐‘ฆ because
๐ผ(๐‘ฅ + ๐‘ฆ) = ๐”ž(+)(๐ผ(๐‘ฅ), ๐ผ(๐‘ฆ)) = ๐”ž(+)(5, 7) = 5 +๐”„ 7,
and ๐ผ(0 ⋅ 1) = 0๐”„ because
๐ผ(0 ⋅ 1) = ๐”ž(⋅)(๐ผ(0), ๐ผ(1)) = ๐”ž(⋅)(0, 1) = 0๐”„ ⋅๐”„ 1๐”„ .
We are now ready for the main definition. It describes what it means for a formula
to be interpreted as true. The definition, which is foundational to model theory, is
generally attributed to Alfred Tarski. This contribution is found in two papers, “Der
wahrheitsbegriff in den formalisierten sprachen” (1935) and “Arithmetical extensions
of relational systems” with Robert Vaught (1957).
DEFINITION 7.1.6
Let ๐”„ = (๐ด, ๐”ž) be an ๐–ฒ-structure and ๐ผ be an ๐–ฒ-interpretation of ๐”„. Assume that
๐‘ and ๐‘ž are ๐–ฒ-formulas and ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 are ๐–ฒ-terms. Define โŠจ as follows:
โˆ™ ๐”„ โŠจ ๐‘ก0 = ๐‘ก1 [๐ผ] ⇔ ๐ผ(๐‘ก0 ) = ๐ผ(๐‘ก1 ).
โˆ™ ๐”„ โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) [๐ผ] ⇔ (๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 )) ∈ ๐”ž(๐‘…).
โˆ™ ๐”„ โŠจ ¬๐‘ [๐ผ] ⇔ (not ๐”„ โŠจ ๐‘ [๐ผ]).
โˆ™ ๐”„ โŠจ ๐‘ → ๐‘ž [๐ผ] ⇔ (if ๐”„ โŠจ ๐‘ [๐ผ] then ๐”„ โŠจ ๐‘ž [๐ผ]).
โˆ™ ๐”„ โŠจ ∃๐‘ฅ๐‘ [๐ผ] ⇔ (๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด), where for every ๐‘ข ∈ ๐ด, the
function ๐ผ๐‘ฅ๐‘ข is the ๐–ฒ-interpretation of ๐”„ such that if ๐‘ฆ is a variable symbol,
{
๐‘ข
if ๐‘ฆ = ๐‘ฅ,
๐‘ข
๐ผ๐‘ฅ (๐‘ฆ) =
๐ผ(๐‘ฆ) if ๐‘ฆ ≠ ๐‘ฅ,
and if ๐‘ is a constant symbol, ๐ผ๐‘ฅ๐‘Ž (๐‘) = ๐ผ(๐‘).
Definition 7.1.6 is an example of a definition by induction on formulas. This means
that first the definition was made for the basic formulas involving equality and relation
symbols, and then assuming that it was made for formulas in general, the definition
is made for formulas written using connectives or quantifiers. Induction on formulas
simply follows Definition 2.1.9. A proof done by induction on formulas is one that uses
the same process to prove a result about terms.
Definition 7.1.6 can be extended to the other connectives and the existential quantifier using the next theorem.
Section 7.1 FIRST-ORDER SEMANTICS
337
THEOREM 7.1.7
Let ๐”„ = (๐ด, ๐”ž) be an ๐–ฒ-structure and ๐ผ be an ๐–ฒ-interpretation of ๐”„. Assume that
๐‘ and ๐‘ž are ๐–ฒ-formulas.
โˆ™ ๐”„ โŠจ ๐‘ ∧ ๐‘ž [๐ผ] ⇔ (๐”„ โŠจ ๐‘ [๐ผ] and ๐”„ โŠจ ๐‘ž [๐ผ]).
โˆ™ ๐”„ โŠจ ๐‘ ∨ ๐‘ž [๐ผ] ⇔ (๐”„ โŠจ ๐‘ [๐ผ] or ๐”„ โŠจ ๐‘ž [๐ผ]).
โˆ™ ๐”„ โŠจ ๐‘ [๐ผ] ↔ ๐‘ž ⇔ (๐”„ โŠจ ๐‘ [๐ผ] if and only if ๐”„ โŠจ ๐‘ž [๐ผ]).
โˆ™ ๐”„ โŠจ ∀๐‘ฅ๐‘ [๐ผ] ⇔ (๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for all ๐‘Ž ∈ ๐ด).
PROOF
Since ๐‘ ∨ ๐‘ž ⇔ ¬๐‘ → ๐‘ž (page 59), we have the following:
๐”„ โŠจ ๐‘ ∨ ๐‘ž [๐ผ] ⇔ ๐”„ โŠจ ¬๐‘ → ๐‘ž [๐ผ]
⇔ if ๐”„ โŠจ ¬๐‘ [๐ผ], then ๐”„ โŠจ ๐‘ž [๐ผ]
⇔ if not ๐”„ โŠจ ๐‘ [๐ผ], then ๐”„ โŠจ ๐‘ž [๐ผ]
⇔ ๐”„ โŠจ ๐‘ [๐ผ] or ๐”„ โŠจ ๐‘ž [๐ผ].
Since ∀๐‘ฅ๐‘ ⇔ ¬∃๐‘ฅ¬๐‘ by QN, we have the following:
๐”„ โŠจ ∀๐‘ฅ๐‘ [๐ผ] ⇔ ๐”„ โŠจ ¬∃๐‘ฅ¬๐‘ [๐ผ]
⇔ not ๐”„ โŠจ ∃๐‘ฅ¬๐‘ [๐ผ]
⇔ not (๐”„ โŠจ ¬๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด)
⇔ not (not ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด)
⇔ ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for all ๐‘Ž ∈ ๐ด.
The remaining parts are left to Exercise 1.
EXAMPLE 7.1.8
Let the ๐–ฒ๐–ณ-structure ๐”„ be defined as (๐œ”, ∈). Since ๐–ฒ๐–ณ has only relation symbols,
๐”„ is called a purely relational structure. Although we usually make a notational
distinction between a name and its meaning, we do not do this with the ∈ symbol. Likewise, the equality symbol = can be used in a formula and during any
interpretation of the formula. To see how this works, define
๐‘ := ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(๐‘ฅ ∈ ๐‘ฆ ∧ ๐‘ฆ ∈ ๐‘ง → ๐‘ฅ ∈ ๐‘ง).
We show that ๐”„ โŠจ ๐‘ for any ๐–ฒ๐–ณ-interpretation ๐ผ. To determine what needs
to be done to accomplish this, we work backwards using Definition 7.1.6 and
Theorem 7.1.7. Let ๐ผ be an ๐–ฒ๐–ณ-interpretation and assume that
๐”„ โŠจ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(๐‘ฅ ∈ ๐‘ฆ ∧ ๐‘ฆ ∈ ๐‘ง → ๐‘ฅ ∈ ๐‘ง) [๐ผ].
This implies that
338
Chapter 7 MODELS
๐”„ โŠจ ๐‘ฅ ∈ ๐‘ฆ ∧ ๐‘ฆ ∈ ๐‘ง → ๐‘ฅ ∈ ๐‘ง [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ] for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ”.
How this formula is interpreted is still based on the order of connectives found
in Definition 1.1.5. Therefore, since the conjunction has precedence, we apply
Definition 7.1.6 to find that
for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ”, if ๐”„ โŠจ ๐‘ฅ ∈ ๐‘ฆ ∧ ๐‘ฆ ∈ ๐‘ง [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ],
then ๐”„ โŠจ ๐‘ฅ ∈ ๐‘ง [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ].
Hence, by Theorem 7.1.7,
for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ”, if ๐”„ โŠจ ๐‘ฅ ∈ ๐‘ฆ [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ] and ๐”„ โŠจ ๐‘ฆ ∈ ๐‘ง [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ],
then ๐”„ โŠจ ๐‘ฅ ∈ ๐‘ง [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ].
We now apply the interpretation using Definition 7.1.4 to find that
for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ”, if ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ) ∈ ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฆ) and ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฆ) ∈ ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ง),
then ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ) ∈ ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ง).
Therefore,
for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐œ”, if ๐‘Ž ∈ ๐‘ and ๐‘ ∈ ๐‘, then ๐‘Ž ∈ ๐‘,
which follows by Theorem 5.2.8. This means that we can back through the steps
to prove that ๐”„ โŠจ ๐‘ [๐ผ].
Typically, formulas cannot be understood as true or false on their own. They have to
be examined against a given universe. The structure is that universe, and the examining
is done by the interpretation. For this reason, the structure-interpretation pair forms the
basis for our work in first-order logic, so we give it a name.
DEFINITION 7.1.9
Let ๐ผ be an ๐–ฒ-interpretation of the ๐–ฒ-structure ๐”„. Let ๐‘ be an ๐–ฒ-formula. The pair
(๐”„, ๐ผ) is called an ๐—ฆ-model. Additionally, if ๐”„ โŠจ ๐‘ [๐ผ], then (๐”„, ๐ผ) is a model
of ๐‘ and satisfies ๐‘. If (๐”„, ๐ผ) is not a model of ๐‘, write ๐”„ ฬธโŠจ ๐‘ [๐ผ].
EXAMPLE 7.1.10
Let ๐”„ be the ๐–ญ๐–ณ-structure with domain ๐œ” and signature {0, 1, +, ⋅} from Example 7.1.2. Let ๐ผ be an ๐–ญ๐–ณ-interpretation of ๐”„ such that
๐ผ(๐‘ฅ) = 0๐”„ , ๐ผ(๐‘ฅ๐‘– ) = ๐‘– +๐”„ 1๐”„ , and ๐ผ(1) = 1๐”„ .
โˆ™ (๐”„, ๐ผ) is a model of ๐‘ฅ4 + ๐‘ฅ7 = ๐‘ฅ7 + ๐‘ฅ4 because by Theorem 5.2.17,
๐ผ(๐‘ฅ4 + ๐‘ฅ7 ) = ๐ผ(๐‘ฅ4 ) +๐”„ ๐ผ(๐‘ฅ7 )
= 5 +๐”„ 8
= 8 +๐”„ 5
= ๐ผ(๐‘ฅ7 ) +๐”„ ๐ผ(๐‘ฅ4 )
= ๐ผ(๐‘ฅ7 + ๐‘ฅ4 ).
Section 7.1 FIRST-ORDER SEMANTICS
339
โˆ™ (๐”„, ๐ผ) is a model of ∀๐‘ฅ(๐‘ฅ = ๐‘ฅ2 → ๐‘ฅ + 1 = ๐‘ฅ2 + 1). To see this, take ๐‘› ∈ ๐œ”
and assume that ๐”„ โŠจ ๐‘ฅ = ๐‘ฅ2 [๐ผ๐‘ฅ๐‘› ]. This implies that ๐‘› = ๐ผ(๐‘ฅ2 ). Hence,
๐‘› +๐”„ 0๐”„ = ๐ผ(๐‘ฅ2 ) +๐”„ 0๐”„ ,
๐‘› +๐”„ ๐ผ(1) = ๐ผ(๐‘ฅ2 ) +๐”„ ๐ผ(1),
๐ผ๐‘ฅ๐‘› (๐‘ฅ) +๐”„ ๐ผ๐‘ฅ๐‘› (1) = ๐ผ๐‘ฅ๐‘› (๐‘ฅ2 ) +๐”„ ๐ผ๐‘ฅ๐‘› (1),
๐ผ๐‘ฅ๐‘› (๐‘ฅ + 1) = ๐ผ๐‘ฅ๐‘› (๐‘ฅ2 + 1).
Therefore, ๐”„ โŠจ ๐‘ฅ + 1 = ๐‘ฅ2 + 1 [๐ผ๐‘ฅ๐‘› ]. We conclude that for all ๐‘› ∈ ๐œ”,
if ๐”„ โŠจ ๐‘ฅ = ๐‘ฅ2 [๐ผ๐‘ฅ๐‘› ], then ๐”„ โŠจ ๐‘ฅ + 1 = ๐‘ฅ2 + 1 [๐ผ๐‘ฅ๐‘› ],
so for all ๐‘› ∈ ๐œ”,
๐”„ โŠจ ๐‘ฅ = ๐‘ฅ2 → ๐‘ฅ + 1 = ๐‘ฅ2 + 1 [๐ผ๐‘ฅ๐‘› ].
This implies that
๐”„ โŠจ ∀๐‘ฅ(๐‘ฅ = ๐‘ฅ2 → ๐‘ฅ + 1 = ๐‘ฅ2 + 1) [๐ผ].
Definition 7.1.9 can be generalized to sets of formulas.
DEFINITION 7.1.11
If (๐”„, ๐ผ) is an ๐–ฒ-model and โ„ฑ is a set of ๐–ฒ-formulas,
๐”„ โŠจ โ„ฑ [๐ผ] if and only if ๐”„ โŠจ ๐‘ [๐ผ] for all ๐‘ ∈ โ„ฑ .
If ๐”„ โŠจ โ„ฑ [๐ผ], then (๐”„, ๐ผ) is a model of โ„ฑ and satisfies โ„ฑ .
For example, using the ๐–ญ๐–ณ-model of Example 7.1.10, we see that
๐”„ โŠจ {๐‘ฅ4 + ๐‘ฅ7 = ๐‘ฅ7 + ๐‘ฅ4 , ∀๐‘ฅ(๐‘ฅ = ๐‘ฅ2 → ๐‘ฅ + 1 = ๐‘ฅ2 + 1)} [๐ผ],
(7.1)
so (๐”„, ๐ผ) is a model of {๐‘ฅ4 + ๐‘ฅ7 = ๐‘ฅ7 + ๐‘ฅ4 , ∀๐‘ฅ(๐‘ฅ = ๐‘ฅ2 → ๐‘ฅ + 1 = ๐‘ฅ2 + 1)}. Here is a
more involved example.
EXAMPLE 7.1.12
Let ๐”… be the ๐–ฆ๐–ฑ-structure (โ„ค, ๐”Ÿ) so that ๐”Ÿ is defined by ๐”Ÿ(๐‘’) = 0 and ๐”Ÿ(โ—ฆ) = +.
Let the interpretation ๐ฝ have the property that
{
๐‘–โˆ•2
if ๐‘– is even,
๐ฝ (๐‘ฅ๐‘– ) =
(๐‘– + 1)โˆ•2 if ๐‘– is odd.
โˆ™ ๐”… โŠจ ∃๐‘ฅ(๐‘ฅ7 โ—ฆ ๐‘ฅ = ๐‘’) [๐ฝ ]. This is because
๐”… โŠจ ๐‘ฅ7 โ—ฆ ๐‘ฅ = ๐‘’ [๐ฝ๐‘ฅ−4 ],
340
Chapter 7 MODELS
and this holds since 4 + (−4) = 0.
โˆ™ ๐”… โŠจ ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’) [๐ฝ ]. To prove this, take ๐‘› ∈ โ„ค. Then,
๐”… โŠจ ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’) [๐ฝ๐‘ฅ๐‘› ]
because
๐”… โŠจ ๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ [(๐ฝ๐‘ฅ๐‘› )−๐‘›
๐‘ฆ ].
This last satisfaction holds because −๐‘› ∈ โ„ค and ๐‘› + −๐‘› = 0.
โˆ™ The previous two satisfactions demonstrate that
๐”… โŠจ {∃๐‘ฅ(๐‘ฅ7 โ—ฆ ๐‘ฅ = ๐‘’), ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’)} [๐ฝ ].
(7.2)
The previous work with models leads to the following definition.
DEFINITION 7.1.13
The ๐–ฒ-formula ๐‘ is ๐—ฆ-satisfiable (denoted by ๐–ฒ๐–บ๐—๐—ฆ ๐‘) if there is an ๐–ฒ-structure ๐”„
and ๐–ฒ-interpretation ๐ผ such that (๐”„, ๐ผ) is a model for ๐‘. The set of ๐–ฒ-formulas โ„ฑ
is ๐—ฆ-satisfiable (denoted by ๐–ฒ๐–บ๐—๐—ฆ โ„ฑ ) if there is a model for โ„ฑ .
By (7.1), we have that
๐–ฒ๐–บ๐—๐—ก๐—ง {๐‘ฅ4 + ๐‘ฅ7 = ๐‘ฅ7 + ๐‘ฅ4 , ∀๐‘ฅ[(๐‘ฅ + ๐‘ฅ5 ) + ๐‘ฅ8 = ๐‘ฅ + (๐‘ฅ5 + ๐‘ฅ8 )]},
and by (7.2), we have that ๐–ฒ๐–บ๐—๐—š๐—ฅ {∃๐‘ฅ(๐‘ฅ7 โ—ฆ ๐‘ฅ = ๐‘’), ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’)}.
Groups
As noted in Example 2.1.5, the ๐–ฆ๐–ฑ symbols are intended for the study of a set with a
single binary operation defined on it. The basic example is โ„ค with addition. Although
other properties will be added later, the language developed for this is designed to
handle just the basic properties of this pair. Namely, the operation should be associative,
the set should have an identity, and every element should have an inverse. This means
that the axioms that will define this theory will be ๐–ฆ๐–ฑ-sentences, and we need only
three.
AXIOMS 7.1.14 [Group]
โˆ™ G1. ∀๐‘ฅ∀๐‘ฆ∀๐‘ง [๐‘ฅ โ—ฆ (๐‘ฆ โ—ฆ ๐‘ง) = (๐‘ฅ โ—ฆ ๐‘ฆ) โ—ฆ ๐‘ง].
โˆ™ G2. ∀๐‘ฅ(๐‘’ โ—ฆ ๐‘ฅ = ๐‘ฅ ∧ ๐‘ฅ โ—ฆ ๐‘’ = ๐‘ฅ).
โˆ™ G3. ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ ∧ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’).
Section 7.1 FIRST-ORDER SEMANTICS
341
EXAMPLE 7.1.15
Define a ๐–ฆ๐–ฑ-structure ๐”Š = (๐บ, ๐”ค) by letting the domain ๐บ = {0}, ๐”ค(๐‘’) = 0, and
๐”ค(โ—ฆ) = +. The binary operation + is addition of integers and 0 ∈ โ„ค (Section 5.3).
Let ๐ผ be an interpretation of ๐”Š. We show that ๐”Š โŠจ {G1, G2, G3} [๐ผ].
โˆ™ Let ๐‘Ž, ๐‘, ๐‘ ∈ ๐บ. Since 0 is the only element of ๐บ, observe that by Definition 7.1.4,
((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ โ—ฆ [๐‘ฆ โ—ฆ ๐‘ง]) = ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ) + ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฆ โ—ฆ ๐‘ง)
(7.3)
= ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ) + [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฆ) + ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ง)]
(7.4)
= 0 + (0 + 0)
=0+0+0
(7.5)
(7.6)
= [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ) + ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฆ)] + ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ง)
(7.7)
=
=
((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ฅ โ—ฆ ๐‘ฆ) + ((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง (๐‘ง)
((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ([๐‘ฅ โ—ฆ ๐‘ฆ] โ—ฆ ๐‘ง).
(7.8)
(7.9)
We see that (7.6) follows from (7.5) by Theorem 5.3.5. Therefore, by Theorem 7.1.7,
๐”Š โŠจ ๐‘ฅ โ—ฆ (๐‘ฆ โ—ฆ ๐‘ง) = (๐‘ฅ โ—ฆ ๐‘ฆ) โ—ฆ ๐‘ง [((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ )๐‘๐‘ง ] for all ๐‘Ž, ๐‘, ๐‘ ∈ ๐บ,
which implies that
๐”Š โŠจ ∀๐‘ง[๐‘ฅ โ—ฆ (๐‘ฆ โ—ฆ ๐‘ง) = (๐‘ฅ โ—ฆ ๐‘ฆ) โ—ฆ ๐‘ง] [(๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ ] for all ๐‘Ž, ๐‘ ∈ ๐บ.
Therefore,
๐”Š โŠจ ∀๐‘ฆ∀๐‘ง[๐‘ฅ โ—ฆ (๐‘ฆ โ—ฆ ๐‘ง) = (๐‘ฅ โ—ฆ ๐‘ฆ) โ—ฆ ๐‘ง] [๐ผ๐‘ฅ๐‘Ž ] for all ๐‘Ž ∈ ๐บ,
so
๐”Š โŠจ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง[๐‘ฅ โ—ฆ (๐‘ฆ โ—ฆ ๐‘ง) = (๐‘ฅ โ—ฆ ๐‘ฆ) โ—ฆ ๐‘ง] [๐ผ].
โˆ™ Let ๐‘Ž ∈ ๐บ. Again, by Definition 7.1.4,
๐ผ๐‘ฅ๐‘Ž (๐‘’ โ—ฆ ๐‘ฅ) = ๐ผ๐‘ฅ๐‘Ž (๐‘’) + ๐ผ๐‘ฅ๐‘Ž (๐‘ฅ) = 0 + 0 = 0 = ๐ผ๐‘ฅ๐‘Ž (๐‘ฅ),
so by Definition 7.1.6,
๐”Š โŠจ ๐‘’ โ—ฆ ๐‘ฅ = ๐‘ฅ [๐ผ๐‘ฅ๐‘Ž ].
Also,
so
๐ผ๐‘ฅ๐‘Ž (๐‘ฅ โ—ฆ ๐‘’) = ๐ผ๐‘ฅ๐‘Ž (๐‘ฅ) + ๐ผ๐‘ฅ๐‘Ž (๐‘’) = 0 + 0 = 0 = ๐ผ๐‘ฅ๐‘Ž (๐‘ฅ),
๐”Š โŠจ ๐‘ฅ โ—ฆ ๐‘’ = ๐‘ฅ [๐ผ๐‘ฅ๐‘Ž ].
342
Chapter 7 MODELS
Therefore, by Theorem 7.1.7,
๐”Š โŠจ ๐‘’ โ—ฆ ๐‘ฅ = ๐‘ฅ ∧ ๐‘ฅ โ—ฆ ๐‘’ = ๐‘ฅ [๐ผ๐‘ฅ๐‘Ž ].
Since ๐‘Ž was arbitrarily chosen,
๐”Š โŠจ ∀๐‘ฅ(๐‘’ โ—ฆ ๐‘ฅ = ๐‘ฅ ∧ ๐‘ฅ โ—ฆ ๐‘’ = ๐‘ฅ) [๐ผ].
โˆ™ Take ๐‘Ž ∈ ๐บ. Because
(๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ (๐‘ฅ โ—ฆ ๐‘ฆ) = (๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ (๐‘ฅ) + (๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ (๐‘ฆ) = 0 + 0 = 0 = (๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ (๐‘’),
we have
๐”Š โŠจ ๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ [(๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ ].
Similarly,
๐”Š โŠจ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’ [(๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ ],
so
๐”Š โŠจ ๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ ∧ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’ [(๐ผ๐‘ฅ๐‘Ž )0๐‘ฆ ].
Therefore, since 0 ∈ ๐บ,
there exists ๐‘ ∈ ๐บ such that ๐”Š โŠจ ๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ ∧ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’ [(๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ ],
and since ๐‘Ž was arbitrarily chosen, we have that
for all ๐‘Ž ∈ ๐บ, there exists ๐‘ ∈ ๐บ such that
๐”Š โŠจ ๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ ∧ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’ [(๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ ].
Then, by Definition 7.1.6, we conclude that
for all ๐‘Ž ∈ ๐บ, ๐”Š โŠจ ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ ∧ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ผ๐‘ฅ๐‘Ž ],
so by Theorem 7.1.7,
๐”Š โŠจ ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘’ ∧ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ผ].
Based on Example 7.1.15, we conclude that {G1, G2, G3} is ๐–ฆ๐–ฑ-satisfiable (Definition 7.1.13). That is, there exists a model in which G1, G2, and G3 are interpreted as
true. The name of the model was first used by Évariste Galois in the early 1830s.
DEFINITION 7.1.16
A ๐–ฆ๐–ฑ-structure that models the group axioms is called a group.
A group with a commutative binary operation, one that satisfies
∀๐‘ฅ∀๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘ฆ โ—ฆ ๐‘ฅ),
is known as an abelian group. It is named after the Norwegian mathematician Niels
Abel. Using the ๐–ฆ๐–ฑ-structure ๐”Š and interpretation ๐ผ of Example 7.1.15, the set of
๐–ฆ๐–ฑ-sentences {G1, G2, G3, ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘ฆ โ—ฆ ๐‘ฅ)} is shown to be ๐–ฆ๐–ฑ-satisfiable.
Section 7.1 FIRST-ORDER SEMANTICS
343
EXAMPLE 7.1.17
Using definitions of Examples 4.2.6 and 4.2.9 for โ„ค, define
โ„ค๐‘› = {[๐‘Ž]๐‘› : ๐‘Ž ∈ โ„ค},
and on this set, specify + by
[๐‘Ž]๐‘› + [๐‘]๐‘› = [๐‘Ž + ๐‘]๐‘› .
As we have seen, the meaning of the symbol + is determined by context. The
+ on the left is the new definition, but the + on the right is standard addition
(Definition 5.3.4). With this definition, a generalization of Example 4.4.30 shows
that + is a binary operation. We check that ๐”Š = (โ„ค๐‘› , [0]๐‘› , +) is a group. Let ๐ผ be
a ๐–ฆ๐–ฑ-interpretation of ๐”Š such that ๐ผ(๐‘’) = [0]๐‘› . We have three axioms to check.
โˆ™ Let [๐‘Ž]๐‘› , [๐‘]๐‘› , [๐‘]๐‘› ∈ โ„ค๐‘› , where ๐‘Ž, ๐‘, ๐‘ ∈ โ„ค. Observe that by Definition 7.1.4,
[๐‘Ž] [๐‘] [๐‘]
((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฅ โ—ฆ [๐‘ฆ โ—ฆ ๐‘ง])
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
(7.10)
= ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฅ) + ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฆ โ—ฆ ๐‘ง)
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
= ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฅ) + [((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฆ) + ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ง)]
(7.11)
= [๐‘Ž]๐‘› + ([๐‘]๐‘› + [๐‘]๐‘› )
= [๐‘Ž]๐‘› + ([๐‘ + ๐‘]๐‘› )
= [๐‘Ž + (๐‘ + ๐‘)]๐‘›
= [๐‘Ž + ๐‘ + ๐‘]๐‘›
= [๐‘Ž + ๐‘]๐‘› + [๐‘]๐‘›
= ([๐‘Ž]๐‘› + [๐‘]๐‘› ) + [๐‘]๐‘›
(7.12)
(7.13)
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
= [((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฅ) + ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฆ)] + ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ง)
=
=
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฅ โ—ฆ ๐‘ฆ) + ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ง)
[๐‘Ž] [๐‘] [๐‘]
((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› ([๐‘ฅ โ—ฆ ๐‘ฆ] โ—ฆ ๐‘ง).
(7.14)
(7.15)
(7.16)
Therefore,
[๐‘Ž] [๐‘] [๐‘]
[๐‘Ž] [๐‘] [๐‘]
((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› (๐‘ฅ โ—ฆ [๐‘ฆ โ—ฆ ๐‘ง]) = ((๐ผ๐‘ฅ ๐‘› )๐‘ฆ ๐‘› )๐‘ง ๐‘› ([๐‘ฅ โ—ฆ ๐‘ฆ] โ—ฆ ๐‘ง)
for all [๐‘Ž]๐‘› , [๐‘]๐‘› , [๐‘]๐‘› ∈ โ„ค๐‘› ,
which implies that
๐”Š โŠจ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง [๐‘ฅ โ—ฆ (๐‘ฆ โ—ฆ ๐‘ง) = (๐‘ฅ โ—ฆ ๐‘ฆ) โ—ฆ ๐‘ง] [๐ผ].
Notice that (7.13) follows from (7.12) by Theorem 5.3.5. More importantly, notice that (7.10), (7.11), (7.14), (7.15), and (7.16) mimic (7.3),
(7.4), (7.7), (7.8), and (7.9) of Example 7.1.15. The other equalities are
344
Chapter 7 MODELS
∗
๐‘’
๐‘Ž
๐‘
๐‘
๐‘’
๐‘’
๐‘Ž
๐‘
๐‘
Figure 7.1
๐‘Ž
๐‘Ž
๐‘’
๐‘
๐‘
๐‘
๐‘
๐‘
๐‘’
๐‘Ž
๐‘
๐‘
๐‘
๐‘Ž
๐‘’
The Klein-4 group.
specific to this example. We conclude that in order to prove that ๐”Š is a
model of G1, we only need to show the work specific to the structure ๐”Š.
We use this to prove ๐”Š โŠจ {G2, G3} [๐ผ].
โˆ™ Let ๐‘Ž ∈ โ„ค. Then,
[0]๐‘› + [๐‘Ž]๐‘› = [0 + ๐‘Ž]๐‘› = [๐‘Ž]๐‘› ,
and
[๐‘Ž]๐‘› + [0]๐‘› = [๐‘Ž + 0]๐‘› = [๐‘Ž]๐‘› .
Therefore, ๐”Š โŠจ G2 [๐ผ].
โˆ™ Let ๐‘ ∈ โ„ค. Then,
[๐‘]๐‘› + [−๐‘]๐‘› = [๐‘ + (−๐‘)]๐‘› = [0]๐‘›
and
[−๐‘]๐‘› + [๐‘]๐‘› = [(−๐‘) + ๐‘]๐‘› = [0]๐‘› .
Hence, ๐”Š โŠจ G3 [๐ผ].
Therefore, ๐”Š is a group. Moreover, because for all ๐‘Ž, ๐‘ ∈ โ„ค,
[๐‘Ž]๐‘› + [๐‘]๐‘› = [๐‘Ž + ๐‘]๐‘› = [๐‘ + ๐‘Ž]๐‘› = [๐‘]๐‘› + [๐‘Ž]๐‘› ,
๐”Š is an abelian group.
EXAMPLE 7.1.18
The structures (๐œ”+ , ∅, +), (โ„ค, 1, ⋅), and (โ„, 1, ⋅) are not groups, but (โ„ค, 0, +),
(โ„š, 0, +), (โ„, 0, +), (โ„š โงต {0}, 1, ⋅), and (โ„ โงต {0}, 1, ⋅) are. These are examples
of infinite groups. When the group’s set is finite, we use the term order to refer
to its cardinality. The group ๐”Š = ({๐œ–}, ๐œ–, ∗), where ∗ = {((๐œ–, ๐œ–), ๐œ–)}, is the only
group of order 1 (Example 7.1.15). This means that any other group with one
element, such as ({0}, 0, +) or ({1}, 1, ⋅), has the same structure as ๐”Š. We say
that these three groups are isomorphic. Any two groups of order 2 will be isomorphic, and any groups of order 3 will also be isomorphic. There are essentially
two groups of order 4, one being the Klein-4 group (Figure 7.1) and the other
being the group of Example 7.1.17 when ๐‘› = 4.
Section 7.1 FIRST-ORDER SEMANTICS
345
EXAMPLE 7.1.19
All of the examples of groups given so far have been abelian. Here is one that is
not. For all ๐‘› ∈ โ„ค+ , define M๐‘› (โ„) as the set of ๐‘› × ๐‘› matrices with real entries.
In other words, each matrix has ๐‘› rows, ๐‘› columns, and looks like
โŽก ๐‘Ž1,1
โŽข โ‹ฎ
โŽข
โŽฃ๐‘Ž๐‘š,1
๐‘Ž1,๐‘› โŽค
โ‹ฎ โŽฅ,
โŽฅ
๐‘Ž๐‘š,๐‘› โŽฆ
···
โ‹ฑ
···
where ๐‘Ž๐‘–,๐‘— ∈ โ„ for ๐‘– = 1, … , ๐‘› and ๐‘— = 1, … , ๐‘›. As an example,
โŽก1
โŽข4
โŽข
โŽฃ7
3โŽค
6โŽฅ ∈ M3 (โ„).
โŽฅ
9โŽฆ
2
5
8
Define matrix multiplication for 2 × 2 matrices by
] [
] [
[
๐‘Ž ๐‘ + ๐‘Ž1,2 ๐‘2,1
๐‘
๐‘1,2
๐‘Ž1,1 ๐‘Ž1,2
= 1,1 1,1
⋅ 1,1
๐‘Ž2,1 ๐‘1,1 + ๐‘Ž2,2 ๐‘2,1
๐‘2,1 ๐‘2,2
๐‘Ž2,1 ๐‘Ž2,2
For example,
[
1
3
] [
2
0
⋅
4
2
] [
−1
4
=
1
8
This multiplication is not commutative because
[
] [
] [
0 −1
1 2
−3
⋅
=
2 1
3 4
5
]
๐‘Ž1,1 ๐‘1,2 + ๐‘Ž1,2 ๐‘2,2
.
๐‘Ž2,1 ๐‘1,2 + ๐‘Ž2,2 ๐‘2,2
]
1
.
1
(7.17)
]
−4
,
8
(7.18)
but it is associative [Exercise 12(a)]. The identity matrix for M2 (โ„) is
[
]
1 0
I2 =
.
0 1
Notice that I2 is the multiplicative identity. If ๐ด ∈ M๐‘› (โ„), then ๐ด is invertible
if there exists ๐ต ∈ M๐‘› (โ„) such that ๐ด๐ต = ๐ต๐ด = I๐‘› . For ๐‘› = 2,
[
]
2 1
0 −3
is invertible, but
[
1
5
]
0
0
is not. All of this can be generalized to any ๐‘› × ๐‘› matrix. Finally, define
M∗๐‘› (โ„) = {๐ด ∈ M๐‘› (โ„) : ๐ด is invertible}.
Let GL(๐‘›, โ„) denote the group (M∗๐‘› (โ„), I๐‘› , ⋅). This is called the general linear
group of degree ๐‘›.
346
Chapter 7 MODELS
Consequence
We now generalize the notion of logical implication (Definition 1.2.2) to the theory of
models.
DEFINITION 7.1.20
Let โ„ฑ be a set of ๐–ฒ-formulas. An ๐–ฒ-formula ๐‘ is an ๐—ฆ-consequence of โ„ฑ (written
as โ„ฑ โŠจ ๐‘) when ๐”„ โŠจ โ„ฑ [๐ผ] implies ๐”„ โŠจ ๐‘ [๐ผ] for any ๐–ฒ-model (๐”„, ๐ผ). If
{๐‘} โŠจ ๐‘ž, simply write ๐‘ โŠจ ๐‘ž.
Definition 7.1.20 implies that if the ๐–ฒ-formula ๐‘ž is not an ๐–ฒ-consequence of โ„ฑ , there
exists an ๐–ฒ-structure ๐”„ with interpretation ๐ผ such that ๐”„ โŠจ โ„ฑ [๐ผ] but ๐”„ ฬธโŠจ ๐‘ž [๐ผ]. For
example, define
๐‘ := ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = ๐‘ฆ + ๐‘ฅ).
We know that GL(๐‘›, โ„) is a group, but under any ๐–ฆ๐–ฑ-interpretation ๐ผ of GL(๐‘›, โ„),
GL(๐‘›, โ„) ฬธโŠจ ๐‘ [๐ผ]
because of (7.17) and (7.18). Therefore, ๐‘ is not an ๐–ฒ-consequence of the group axioms
(7.1.14). In other words, not all groups are abelian groups.
EXAMPLE 7.1.21
Suppose that
โ„ฑ = {∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = ๐‘ฆ + ๐‘ฅ), ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0)}.
Let (๐”„, ๐ผ) be an ๐–ญ๐–ณ-model of โ„ฑ . Let ๐ด = dom(๐”„). Then,
๐”„ โŠจ ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0) [๐ผ],
so by Theorem 7.1.7,
for all ๐‘ข ∈ ๐ด, ๐”„ โŠจ ∃๐‘ฆ(๐‘ฅ + ๐‘ฆ = 0) [๐ผ๐‘ฅ๐‘ข ],
which by Definition 7.1.6 implies that
there exists ๐‘ฃ ∈ ๐ด such that for all ๐‘ข ∈ ๐ด, ๐”„ โŠจ ๐‘ฅ + ๐‘ฆ = 0 [(๐ผ๐‘ฅ๐‘ข )๐‘ฃ๐‘ฆ ].
Hence, for arbitrary ๐‘Ž (UI) and particular ๐‘ (EI) in ๐ด,
(๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (+)((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฅ), (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฆ)) = (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (0).
However, since
๐”„ โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = ๐‘ฆ + ๐‘ฅ) [๐ผ],
we find that
(๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (+)((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฅ), (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฆ)) = (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (+)((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฆ), (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฅ)),
Section 7.1 FIRST-ORDER SEMANTICS
347
so,
(๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (+)((๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฆ), (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (๐‘ฅ)) = (๐ผ๐‘ฅ๐‘Ž )๐‘๐‘ฆ (0).
Therefore, by EG and UG,
there exists ๐‘ฃ ∈ ๐ด such that for all ๐‘ข ∈ ๐ด, ๐ผ(+)(๐ผ(๐‘ฃ), ๐ผ(๐‘ข)) = ๐ผ(0),
and we can reverse the steps above to find that
๐”„ โŠจ ∃๐‘ฅ∀๐‘ฆ(๐‘ฆ + ๐‘ฅ = 0) [๐ผ],
so we conclude that
โ„ฑ โŠจ ∃๐‘ฅ∀๐‘ฆ(๐‘ฆ + ๐‘ฅ = 0).
We say that an ๐–ฒ-sentence ๐‘ is valid if ∅ โŠจ ๐‘ and write โŠจ ๐‘. This means that if an ๐–ฒsentence ๐‘ is valid, every ๐–ฒ-structure is a model of ๐‘ since every ๐–ฒ-structure is a model
of the empty set using any ๐–ฒ-interpretation (Exercise 7). For example, ∀๐‘ฅ(๐‘ฅ = ๐‘ฅ) and
๐‘ƒ ∨ ¬๐‘ƒ are valid.
We now connect the notions of consequence and satisfaction.
THEOREM 7.1.22
Let โ„ฑ be a set of ๐–ฒ-formulas and ๐‘ be an ๐–ฒ-formula. Then,
โ„ฑ โŠจ ๐‘ if and only if not ๐–ฒ๐–บ๐—๐—ฆ โ„ฑ ∪ {¬๐‘}.
PROOF
The following are equivalent:
โˆ™ โ„ฑ โŠจ ๐‘.
โˆ™ For every ๐–ฒ-model (๐”„, ๐ผ), if ๐”„ โŠจ โ„ฑ [๐ผ], then ๐”„ โŠจ ๐‘ [๐ผ].
โˆ™ There does not exist an ๐–ฒ-model (๐”„, ๐ผ) so that ๐”„ โŠจ โ„ฑ [๐ผ] and ๐”„ ฬธโŠจ ๐‘ [๐ผ].
โˆ™ There does not exist an ๐–ฒ-model (๐”„, ๐ผ) so that ๐”„ โŠจ โ„ฑ [๐ผ] and ๐”„ โŠจ ¬๐‘ [๐ผ].
โˆ™ There does not exist an ๐–ฒ-model (๐”„, ๐ผ) such that ๐”„ โŠจ โ„ฑ ∪ {¬๐‘} [๐ผ].
โˆ™ Not ๐–ฒ๐–บ๐—๐—ฆ โ„ฑ ∪ {¬๐‘}.
EXAMPLE 7.1.23
Since Zorn’s lemma was proved from the axioms of ๐—ญ๐—™๐—– (Theorem 5.1.13), as
in Example 7.1.17, we can use the work specific to the proof of Zorn’s lemma to
conclude that ๐—ญ๐—™๐—– โŠจ Zorn’s lemma, so by Theorem 7.1.22, there is no ๐–ฒ๐–ณ-model
that satisfies ๐—ญ๐—™๐—– and the negation of Zorn’s lemma.
348
Chapter 7 MODELS
EXAMPLE 7.1.24
The ๐–ฒ-formula ๐‘ is valid if ¬๐‘ is not ๐–ฒ-satisfiable. To see this, suppose that ๐‘ is
not valid. This means that there is an ๐–ฒ-model (๐”„, ๐ผ) so that ๐”„ ฬธโŠจ ๐‘ [๐ผ]. Hence,
by Definition 7.1.6, ๐”„ โŠจ ¬๐‘ [๐ผ]. Therefore, ๐–ฒ๐–บ๐—{¬๐‘}. That the converse is true
is Exercise 19.
Compare the next definition with Definition 1.3.1.
DEFINITION 7.1.25
Let ๐‘ and ๐‘ž be ๐–ฒ-formulas. Then, ๐‘ is logically equivalent to ๐‘ž means ๐‘ โŠจ ๐‘ž if
and only if ๐‘ž โŠจ ๐‘.
Notice Definition 7.1.25 implies that the formulas ๐‘ and ๐‘ž are logically equivalent if and
only if โŠจ (๐‘ ↔ ๐‘ž). For example, by De Morgan’s law, ¬(๐‘ ∧ ๐‘ž) is logically equivalent
to ¬๐‘ ∨ ¬๐‘ž, and by QN, we conclude that ¬∀๐‘ฅ๐‘(๐‘ฅ) is logically equivalent to ∃๐‘ฅ¬๐‘(๐‘ฅ).
Coincidence
Let ๐”„ = (โ„ค, ๐”ž) and ๐”… = (โ„ค, ๐”Ÿ) be ๐–ฆ๐–ฑ-structures such that
๐”ž(๐‘’) = ๐”Ÿ(๐‘’) = 0,
๐”ž(โ—ฆ) = ๐”Ÿ(โ—ฆ) = +.
Let ๐ผ be an ๐–ฆ๐–ฑ-interpretation of ๐”„ and ๐ฝ be a ๐–ฆ๐–ฑ-interpretation of ๐”… such that
๐ผ(๐‘ฅ) = ๐ฝ (๐‘ฅ) = 3.
Other assignments of these functions are not identified. Consider the following deduction:
−3 + 3 = 0.
๐‘› + 3 = 0 for some ๐‘› ∈ โ„ค.
๐ผ๐‘ฆ๐‘› (๐‘ฆ) + ๐ผ๐‘ฆ๐‘› (๐‘ฅ) = ๐ผ๐‘ฆ๐‘› (๐‘’) for some ๐‘› ∈ โ„ค.
๐ผ๐‘ฆ๐‘› (๐‘ฆ โ—ฆ ๐‘ฅ) = ๐ผ๐‘ฆ๐‘› (๐‘’) for some ๐‘› ∈ โ„ค.
๐”„ โŠจ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’ [๐ผ๐‘ฆ๐‘› ] for some ๐‘› ∈ โ„ค.
Therefore,
๐”„ โŠจ ∃๐‘ฆ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ผ].
By replacing ๐ผ with ๐ฝ and ๐”„ with ๐”… in the deduction, we conclude that
๐”… โŠจ ∃๐‘ฆ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ฝ ].
Since ๐ผ and ๐ฝ agree on their interpretations of ๐‘’, โ—ฆ, and ๐‘ฅ, it is not surprising that
they should agree on their interpretation of any {๐‘’, โ—ฆ}-formula with ๐‘ฅ as its only free
variable. The generalization of this to terms and formulas is the next two results.
Section 7.1 FIRST-ORDER SEMANTICS
349
LEMMA 7.1.26 [Coincidence for Terms]
Let ๐–ฒ and ๐–ณ be sets of theory symbols. Let ๐”„ = (๐ด, ๐”ž) be an ๐–ฒ-structure and
๐”… = (๐ต, ๐”Ÿ) be a ๐–ณ-structure such that ๐ด = ๐ต. Let ๐ผ be an interpretation of ๐”„
and ๐ฝ be an interpretation of ๐”…. If ๐ผ ๐–ต๐– ๐–ฑ = ๐ฝ ๐–ต๐– ๐–ฑ and ๐”ž(๐‘ข) = ๐”Ÿ(๐‘ข) for all
๐‘ข ∈ ๐–ฒ ∩ ๐–ณ, then ๐ผ(๐‘ก) = ๐ฝ (๐‘ก) for every (๐–ฒ ∩ ๐–ณ)-term ๐‘ก.
PROOF
By induction on (๐–ฒ ∩ ๐–ณ)-terms.
โˆ™ Let ๐‘ฅ be a variable symbol. Then, ๐ผ(๐‘ฅ) = ๐ฝ (๐‘ฅ) by hypothesis.
โˆ™ Let ๐‘ be a constant symbol in ๐–ฒ ∩ ๐–ณ. We have
๐ผ(๐‘) = ๐”ž(๐‘) = ๐”Ÿ(๐‘) = ๐ฝ (๐‘).
โˆ™ Suppose ๐ผ(๐‘ก๐‘– ) = ๐ฝ (๐‘ก๐‘– ) for all (๐–ฒ ∩ ๐–ณ)-terms ๐‘ก๐‘– with ๐‘– = 0, 1, … , ๐‘› − 1. Then,
๐ผ(๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )) = ๐”ž(๐‘“ )(๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 ))
= ๐”ž(๐‘“ )(๐ฝ (๐‘ก0 ), ๐ฝ (๐‘ก1 ), … , ๐ฝ (๐‘ก๐‘›−1 ))
= ๐”Ÿ(๐‘“ )(๐ฝ (๐‘ก0 ), ๐ฝ (๐‘ก1 ), … , ๐ฝ (๐‘ก๐‘›−1 ))
= ๐ฝ (๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )).
LEMMA 7.1.27 [Coincidence for Formulas]
Let ๐–ฒ and ๐–ณ be sets of theory symbols. Let ๐”„ = (๐ด, ๐”ž) be an ๐–ฒ-structure and
๐”… = (๐ต, ๐”Ÿ) be a ๐–ณ-structure such that ๐ด = ๐ต. Let ๐ผ be an interpretation of ๐”„ and
๐ฝ be an interpretation of ๐”…. If ๐ผ ๐–ต๐– ๐–ฑ = ๐ฝ ๐–ต๐– ๐–ฑ and ๐”ž(๐‘ข) = ๐”Ÿ(๐‘ข) for every
๐‘ข ∈ ๐–ฒ ∩ ๐–ณ, then ๐”„ โŠจ ๐‘ [๐ผ] if and only if ๐”… โŠจ ๐‘ [๐ฝ ] for all (๐–ฒ ∩ ๐–ณ)-formulas ๐‘.
PROOF
By induction on (๐–ฒ ∩ ๐–ณ)-formulas.
โˆ™ Let ๐‘ก0 and ๐‘ก1 be (๐–ฒ ∩ ๐–ณ)-terms. Then, by Lemma 7.1.26,
๐”„ โŠจ ๐‘ก0 = ๐‘ก1 [๐ผ] ⇔ ๐ผ(๐‘ก0 ) = ๐ผ(๐‘ก1 ) ⇔ ๐ฝ (๐‘ก0 ) = ๐ฝ (๐‘ก1 ) ⇔ ๐”… โŠจ ๐‘ก0 = ๐‘ก1 [๐ฝ ].
โˆ™ Let ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 be (๐–ฒ ∩ ๐–ณ)-terms and ๐‘… a relation symbol of ๐–ฒ ∩ ๐–ณ. Then,
by Lemma 7.1.26,
๐”„ โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) [๐ผ] ⇔ (๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 )) ∈ ๐”ž(๐‘…)
⇔ (๐ฝ (๐‘ก0 ), ๐ฝ (๐‘ก1 ), … , ๐ฝ (๐‘ก๐‘›−1 )) ∈ ๐”ž(๐‘…)
⇔ (๐ฝ (๐‘ก0 ), ๐ฝ (๐‘ก1 ), … , ๐ฝ (๐‘ก๐‘›−1 )) ∈ ๐”Ÿ(๐‘…)
⇔ ๐”… โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) [๐ฝ ].
Now let ๐‘ be an (๐–ฒ ∩ ๐–ณ)-formula.
โˆ™ ๐”„ โŠจ ¬๐‘ [๐ผ] ⇔ ๐”„ ฬธโŠจ ๐‘ [๐ผ] ⇔ ๐”… ฬธโŠจ ๐‘ [๐ฝ ] ⇔ ๐”… โŠจ ¬๐‘ [๐ฝ ].
350
Chapter 7 MODELS
โˆ™ Assume that ๐”„ โŠจ ๐‘ [๐ผ] implies ๐”„ โŠจ ๐‘ž [๐ผ]. Also, suppose that ๐”… โŠจ ๐‘ [๐ฝ ].
Then, ๐”„ โŠจ ๐‘ [๐ผ] by induction, so ๐”„ โŠจ ๐‘ž [๐ผ]. Thus, ๐”… โŠจ ๐‘ž [๐ฝ ] by
induction. The converse is proved similarly, so we have
๐”„ โŠจ ๐‘ → ๐‘ž [๐ผ] ⇔ if ๐”„ โŠจ ๐‘ [๐ผ] then ๐”„ โŠจ ๐‘ž [๐ผ]
⇔ if ๐”… โŠจ ๐‘ [๐ฝ ] then ๐”… โŠจ ๐‘ž [๐ฝ ]
⇔ ๐”… โŠจ ๐‘ → ๐‘ž [๐ฝ ].
โˆ™ Note that for all ๐‘ ∈ ๐ด,
๐ผ๐‘ฅ๐‘ ๐–ต๐– ๐–ฑ = ๐ฝ๐‘ฅ๐‘ ๐–ต๐– ๐–ฑ
because
and if ๐‘ฆ ≠ ๐‘ฅ,
๐ผ๐‘ฅ๐‘ (๐‘ฅ) = ๐‘ข = ๐ฝ๐‘ฅ๐‘ (๐‘ฅ)
๐ผ๐‘ฅ๐‘ (๐‘ฆ) = ๐ผ(๐‘ฆ) = ๐ฝ (๐‘ฆ) = ๐ฝ๐‘ฅ๐‘ (๐‘ฆ).
Therefore, by induction and since ๐ด = ๐ต,
๐”„ โŠจ ∃๐‘ฅ๐‘ [๐ผ] ⇔ ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด
⇔ ๐”„ โŠจ ๐‘ [๐ฝ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ต
⇔ ๐”„ โŠจ ∃๐‘ฅ๐‘ [๐ฝ ].
EXAMPLE 7.1.28
Define the sets of theory symbols ๐–ฒ = {0, 1, +, ⋅, ≤} and ๐–ณ = {0, 1, +, ∗, ≥}. Let
๐”„ = (๐œ”, ๐”ž) be an ๐–ฒ-structure and ๐”… = (๐œ”, ๐”Ÿ) be a ๐–ณ-structure where
๐”ž(0) = ๐”Ÿ(0) = ∅,
๐”ž(1) = ๐”Ÿ(1) = {∅},
๐”ž(+) = ๐”Ÿ(+).
Notice, for example, that under the right interpretation, ๐”„ could be a model of
∀๐‘ฅ(๐‘ฅ ⋅ 1 = ๐‘ฅ) since it is an ๐–ฒ-formula, but it does not make sense for ๐”… to be
a model of the same sentence because ∀๐‘ฅ(๐‘ฅ ⋅ 1 = ๐‘ฅ) is not a ๐–ณ-formula. Now,
let ๐ผ be an ๐–ฒ-interpretation of ๐”„ and ๐ฝ be a ๐–ณ-interpretation of ๐”… such that they
agree on all variable symbols. Since we have the hypotheses of Lemma 7.1.27
satisfied, let us confirm the lemma. Consider the (๐–ฒ ∩ ๐–ณ)-formula ∀๐‘ฅ(๐‘ฅ + 0 = ๐‘ฅ).
Assume
๐”„ โŠจ ∀๐‘ฅ(๐‘ฅ + 0 = ๐‘ฅ) [๐ผ],
so
๐”„ โŠจ (๐‘ฅ + 0 = ๐‘ฅ) [๐ผ๐‘ฅ๐‘› ] for all ๐‘› ∈ ๐œ”.
This implies that
Section 7.1 FIRST-ORDER SEMANTICS
351
๐”ž(+)(๐ผ๐‘ฅ๐‘› (๐‘ฅ), ๐ผ๐‘ฅ๐‘› (0)) = ๐ผ๐‘ฅ๐‘› (๐‘ฅ) for all ๐‘› ∈ ๐œ”.
Since ๐ผ and ๐ฝ agree on all variable symbols, ๐”ž(0) = ๐”Ÿ(0), and ๐”ž(+) = ๐”Ÿ(+),
๐”Ÿ(+)(๐ฝ๐‘ฅ๐‘› (๐‘ฅ), ๐ฝ๐‘ฅ๐‘› (0)) = ๐ฝ๐‘ฅ๐‘› (๐‘ฅ) for all ๐‘› ∈ ๐œ”.
Therefore,
๐”… โŠจ (๐‘› + 0 = ๐‘Ž) [๐ฝ๐‘ฅ๐‘› ] for all ๐‘› ∈ ๐œ”,
which gives
๐”… โŠจ ∀๐‘ฅ(๐‘ฅ + 0 = ๐‘ฅ) [๐ฝ ].
The purpose of the coincidence lemmas (7.1.26 and 7.1.27) is to minimize the use
of interpretation functions, especially when modeling sentences.
LEMMA 7.1.29
Let ๐”„ be an ๐–ฒ-structure. If ๐‘ is an ๐–ฒ-sentence,
๐”„ โŠจ ๐‘ [๐ผ] if and only if ๐”„ โŠจ ๐‘ [๐ฝ ]
for all ๐–ฒ-interpretations ๐ผ and ๐ฝ of ๐”„.
PROOF
Suppose that ๐ผ and ๐ฝ are ๐–ฒ-interpretations of ๐”„. Let ๐”„ โŠจ ๐‘ [๐ผ]. Since ๐‘ has no
free variables, ๐”„ โŠจ ๐‘ [๐ฝ ] by the proof of Lemma 7.1.27.
Lemma 7.1.29 implies that any interpretation will do when modeling sentences. Therefore, we make the next definition.
DEFINITION 7.1.30
For any ๐–ฒ-sentence ๐‘ and ๐–ฒ-structure ๐”„ = (๐ด, ๐”ž), write ๐”„ โŠจ ๐‘ if ๐”„ โŠจ ๐‘ [๐ผ] for
all ๐–ฒ-interpretations ๐ผ of ๐”„.
Lemma 7.1.29 can be used to quickly prove the next result.
THEOREM 7.1.31
For any ๐–ฒ-structure ๐”„ and ๐–ฒ-sentence ๐‘,
๐”„ โŠจ ๐‘ if and only if ๐”„ โŠจ ๐‘ [๐ผ] for some ๐–ฒ-interpretation ๐ผ of ๐”„.
Therefore, letting ๐”… be the ๐–ฆ๐–ฑ-structure of Example 7.1.12, by (7.2), we have that
๐”… โŠจ ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = 0).
The coincidence lemmas (7.1.26 and 7.1.27) also minimize the use of the sets of
theory symbols. Consider the following.
352
Chapter 7 MODELS
THEOREM 7.1.32
Let ๐–ฒ ⊆ ๐–ณ be theory symbol sets. If โ„ฑ is a set of ๐–ฒ-formulas, โ„ฑ is ๐–ฒ-satisfiable
if and only if โ„ฑ is ๐–ณ-satisfiable.
PROOF
Let โ„ฑ be a set of of ๐–ฒ-formulas. First, suppose that (๐ด, ๐”ž) โŠจ โ„ฑ [๐ผ], where
dom(๐”ž) = ๐–ฒ and dom(๐ผ) = ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐–ฒ). Let ๐”ž′ be an extension of ๐”ž to ๐–ณ and ๐ผ ′
be an extension of ๐ผ such that dom(๐ผ ′ ) = ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐–ณ). Notice that this implies
that ๐”ž and ๐”ž′ agree on ๐–ฒ = ๐–ฒ ∩ ๐–ณ. Therefore, (๐ด, ๐”ž′ ) โŠจ โ„ฑ [๐ผ ′ ] by Lemma 7.1.27.
Conversely, assume that (๐ด, ๐”ž) โŠจ โ„ฑ [๐ผ] such that both dom(๐”ž) = ๐–ณ and
dom(๐ผ) = ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐–ณ). Let ๐”ž′ = ๐”ž ๐–ฒ and ๐ผ ′ = ๐ผ ∩ ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐–ฒ). This implies that
(๐ด, ๐”ž′ ) โŠจ โ„ฑ [๐ผ ′ ] by Lemma 7.1.27.
There is terminology to name the relationship between the structures found in the
proof of Theorem 7.1.32. Let the theory symbols ๐–ฒ be a subset of the theory symbols
๐–ณ. Let ๐”„ = (๐ด, ๐”ž) be an ๐–ฒ-structure and ๐”„′ (๐ด′ , ๐”ž′ ) be an ๐–ณ-structure. If ๐ด = ๐ด′ and
๐”ž = ๐”ž′ ๐–ฒ, we call ๐”„′ an expansion of ๐”„ and ๐”„ a reduct of ๐”„′ . Hence, in the first
part of the proof, we started with a structure and then moved to an expansion, and in
the second part, we started with a structure and then moved to a reduct.
Theorems 7.1.31 and 7.1.32 motivate the next two definitions.
DEFINITION 7.1.33
Let ๐–ฒ ⊆ ๐–ณ be sets of theory symbols. An ๐–ฒ-formula ๐‘ is satisfiable (denoted by
๐–ฒ๐–บ๐— ๐‘) if there exists an ๐–ณ-structure that is a model for ๐‘. The set of ๐–ฒ-formulas
โ„ฑ is satisfiable (denoted by ๐–ฒ๐–บ๐— โ„ฑ ) if there exists an ๐–ณ-structure that is a model
for โ„ฑ .
DEFINITION 7.1.34
Let ๐–ฒ ⊆ ๐–ณ be sets of theory symbols. Assume that ๐’ฏ is a set of ๐–ฒ-sentences. An
๐–ฒ-sentence ๐‘ is a consequence of ๐’ฏ (denoted by ๐’ฏ โŠจ ๐‘) when ๐”„ โŠจ ๐’ฏ implies
๐”„ โŠจ ๐‘ for every ๐–ณ-structure ๐”„.
EXAMPLE 7.1.35
The group axioms state that in a group there is an identity and there are inverses.
Based on what we know about the integers, we should be able to prove more
about these elements. For example, we expect that in a group, both
there is exactly one identity
and
every element has a unique inverse
are true. The uniqueness of the identity is left to Exercise 20. To show the uniqueness of inverses, let ๐”Š = (๐บ, ๐‘’, โ—ฆ) be a group, and take ๐‘Ž ∈ ๐บ. Suppose that
Section 7.1 FIRST-ORDER SEMANTICS
353
๐‘Ž′ , ๐‘Ž′′ ∈ ๐บ and are inverses of ๐‘Ž. Then,
๐‘Ž′ = ๐‘Ž′ โ—ฆ ๐‘’ = ๐‘Ž′ โ—ฆ (๐‘Ž โ—ฆ ๐‘Ž′′ ) = (๐‘Ž′ โ—ฆ ๐‘Ž) โ—ฆ ๐‘Ž′′ = ๐‘’ โ—ฆ ๐‘Ž′′ = ๐‘Ž′′ .
Therefore, the uniqueness of inverses is a consequence (Definition 7.1.34) of the
group axioms.
Rings
Consider the equation 2๐‘ฅ + 1 = 0. The exact steps needed to find its solution are
(2๐‘ฅ + 1) + −1 = 0 + −1,
2๐‘ฅ + (1 + −1) = 0 + −1,
2๐‘ฅ + 0 = 0 + −1,
2๐‘ฅ = −1,
1โˆ•2(2๐‘ฅ) = 1โˆ•2(−1),
(1โˆ•2 ⋅ 2)๐‘ฅ = −1โˆ•2,
1๐‘ฅ = −1โˆ•2,
๐‘ฅ = −1โˆ•2.
Now examine the steps. There are two operations, addition and multiplication. We
used inverses and identities. The associative law was also used. When studying these
steps, we realize that they cannot be performed within (โ„ค, 0, +) even though the initial equation had only integer coefficients. This means that the group idea needs to be
expanded. This is done by including two symbols to represent addition and multiplication. Since these two operations can have their own identities, replace ๐‘’ with โ—‹ to
represent the additive identity. The ideas behind the group axioms are then extended
using ๐–ฑ๐–จ-sentences.
AXIOMS 7.1.36 [Ring]
โˆ™ R1. ∀๐‘ฅ∀๐‘ฆ∀๐‘ง [๐‘ฅ ⊕ (๐‘ฆ ⊕ ๐‘ง) = (๐‘ฅ ⊕ ๐‘ฆ) ⊕ ๐‘ง]
∀๐‘ฅ∀๐‘ฆ∀๐‘ง [๐‘ฅ ⊗ (๐‘ฆ ⊗ ๐‘ง) = (๐‘ฅ ⊗ ๐‘ฆ) ⊗ ๐‘ง]
โˆ™ R2. ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⊕ ๐‘ฆ = ๐‘ฆ ⊕ ๐‘ฅ)
โˆ™ R3. ∀๐‘ฅ(0 ⊕ ๐‘ฅ = ๐‘ฅ)
โˆ™ R4. ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ ⊕ ๐‘ฆ = โ—‹)
โˆ™ R5. ∀๐‘ฅ∀๐‘ฆ∀๐‘ง [๐‘ฅ ⊗ (๐‘ฆ ⊕ ๐‘ง) = ๐‘ฅ ⊗ ๐‘ฆ ⊕ ๐‘ฅ ⊗ ๐‘ง]
∀๐‘ฅ∀๐‘ฆ∀๐‘ง [(๐‘ฅ ⊕ ๐‘ฆ) ⊗ ๐‘ง = ๐‘ฅ ⊗ ๐‘ง ⊕ ๐‘ฆ ⊗ ๐‘ง]
354
Chapter 7 MODELS
DEFINITION 7.1.37
An ๐–ฑ๐–จ-structure ℜ = (๐‘…, 0, +, ⋅) that models the ring axioms is called a ring. If
there exists an multiplicative identity in ๐‘…, then ℜ is a ring with unity.
The additive inverse of ๐‘Ž is −๐‘Ž, and the multiplicative inverse of ๐‘Ž is ๐‘Ž−1 assuming that
๐‘Ž ≠ 0. We usually write ๐‘Ž−๐‘ instead of ๐‘Ž+(−๐‘). Notice that if ℜ = (๐‘…, 0, +, ⋅) is a ring,
its reduct (๐‘…, 0, +) is a group. Also, letting + and ⋅ denote addition and multiplication
on โ„ค,
โ„จ = (โ„ค, 0, +, ⋅)
is a ring with unity. Also, โ„š, โ„, and โ„‚ are the domains of rings with unity using the
typical operations of addition and multiplication.
EXAMPLE 7.1.38
Axioms 7.1.36 do not require that the ring multiplication be commutative. Let
ℜ = (๐‘…, 0, +, ⋅). Then, ℜ is a commutative ring. if
ℜ โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฆ ⊗ ๐‘ฅ).
โˆ™ Let + and ⋅ denote standard addition and multiplication on โ„ค. Let ๐‘› ∈ โ„ค.
We conclude that ๐”– = (๐‘›โ„ค, 0, +, ⋅) is a commutative ring. It is without
unity if ๐‘› ≠ ±1.
โˆ™ Take [๐‘Ž]๐‘› , [๐‘]๐‘› ∈ โ„ค๐‘› and define + as in Example 7.1.17 and multiplication
defined by
[๐‘Ž]๐‘› ⋅ [๐‘]๐‘› = [๐‘Ž๐‘]๐‘› .
Then, ๐”— = (โ„ค๐‘› , [0]๐‘› , +, ⋅) is a commutative ring (Exercise 29).
Axioms 7.1.36 also do not state that when the additive identity is multiplied by
any element of the ring, the result is the additive identity. It is not among the
axioms because it can be proved. Take ๐‘Ž ∈ ๐‘…. By R3, 0 + 0 = 0, so by R5,
0 ⋅ ๐‘Ž = (0 + 0) ⋅ ๐‘Ž = 0 ⋅ ๐‘Ž + 0 ⋅ ๐‘Ž.
By R4 and since + is a binary operation,
0 ⋅ ๐‘Ž + −(0 ⋅ ๐‘Ž) = (0 ⋅ ๐‘Ž + 0 ⋅ ๐‘Ž) + −(0 ⋅ ๐‘Ž).
Because of R1,
0 ⋅ ๐‘Ž + −(0 ⋅ ๐‘Ž) = 0 ⋅ ๐‘Ž + [0 ⋅ ๐‘Ž + −(0 ⋅ ๐‘Ž)].
Hence, 0 = 0 ⋅ ๐‘Ž + 0, which implies that 0 = 0 ⋅ ๐‘Ž. Therefore,
ℜ โŠจ ∀๐‘ฅ(โ—‹ = โ—‹ ⊗ ๐‘ฅ),
and ∀๐‘ฅ(โ—‹ = โ—‹ ⊗ ๐‘ฅ) is a consequence (Definition 7.1.34) of the ring axioms.
Section 7.1 FIRST-ORDER SEMANTICS
355
EXAMPLE 7.1.39
Let ๐‘› ∈ โ„ค+ , define matrix addition on M๐‘› (โ„) entrywise. For instance,
โŽก1
โŽข3
โŽข
โŽฃ5
2
4
6
1 โŽค โŽก1
−4โŽฅ + โŽข2
โŽฅ โŽข
0 โŽฆ โŽฃ0
0 8โŽค โŽก2
−5 0โŽฅ = โŽข5
โŽฅ โŽข
−2 3โŽฆ โŽฃ5
2
−1
4
9โŽค
−4โŽฅ .
โŽฅ
3โŽฆ
Let 0๐‘› be the zero matrix. It is the ๐‘› × ๐‘› matrix with all of its entries equal to
0. As in Example 7.1.19, let ⋅ represent matrix multiplication and ๐ผ๐‘› the identity
matrix. Prove that
๐”๐‘› (โ„) = (M๐‘› (โ„), 0๐‘› , +, ⋅)
is a ring.
โˆ™ To see that matrix addition is associative, we rely on the fact that standard
addition of real numbers is associative. Take three matrices from M2 (โ„)
and add:
])
] [
] ([
[
๐‘
๐‘
๐‘1,1 ๐‘1,2
๐‘Ž1,1 ๐‘Ž1,2
+ 1,1 1,2
+
๐‘2,1 ๐‘2,2
๐‘2,1 ๐‘2,2
๐‘Ž2,1 ๐‘Ž2,2
]
] [
[
๐‘1,1 + ๐‘1,1 ๐‘1,2 + ๐‘1,2
๐‘Ž1,1 ๐‘Ž1,2
+
=
๐‘2,1 + ๐‘2,1 ๐‘2,2 + ๐‘2,2
๐‘Ž2,1 ๐‘Ž2,2
[
]
๐‘Ž1,1 + (๐‘1,1 + ๐‘1,1 ) ๐‘Ž1,2 + (๐‘1,2 + ๐‘1,2 )
=
๐‘Ž2,1 + (๐‘2,1 + ๐‘2,1 ) ๐‘Ž2,2 + (๐‘2,2 + ๐‘2,2 )
]
[
(๐‘Ž1,1 + ๐‘1,1 ) + ๐‘1,1 (๐‘Ž1,2 + ๐‘1,2 ) + ๐‘1,2
=
(๐‘Ž2,1 + ๐‘2,1 ) + ๐‘2,1 (๐‘Ž2,2 + ๐‘2,2 ) + ๐‘2,2
]
] [
[
๐‘1,1 ๐‘1,2
๐‘Ž1,1 + ๐‘1,1 ๐‘Ž1,2 + ๐‘1,2
+
=
๐‘2,1 ๐‘2,2
๐‘Ž2,1 + ๐‘2,1 ๐‘Ž2,2 + ๐‘2,2
] [
]) [
]
([
๐‘1,1 ๐‘1,2
๐‘1,1 ๐‘1,2
๐‘Ž1,1 ๐‘Ž1,2
.
+
+
=
๐‘Ž2,1 ๐‘Ž2,2
๐‘2,1 ๐‘2,2
๐‘2,1 ๐‘2,2
โˆ™ Since
[
0
0
] [
0
๐‘Ž
+ 1,1
0
๐‘Ž2,1
] [
]
๐‘Ž1,2
0 + ๐‘Ž1,1 0 + ๐‘Ž1,2
=
๐‘Ž2,2
0 + ๐‘Ž2,1 0 + ๐‘Ž2,2
]
[
๐‘Ž
๐‘Ž1,2
,
= 1,1
๐‘Ž2,1 ๐‘Ž2,2
the zero matrix is the additive identity.
โˆ™ To prove that every element of M2 (โ„) has an additive inverse, take
[
๐‘Ž
๐ด = 1,1
๐‘Ž2,1
]
๐‘Ž1,2
∈ M2 (โ„).
๐‘Ž2,2
356
Chapter 7 MODELS
Then,
[
−๐‘Ž1,1
−๐ด =
−๐‘Ž2,1
−๐‘Ž1,2
−๐‘Ž2,2
]
because ๐ด + (−๐ด) = 02 . Generalizing, conclude that
๐”๐‘› (โ„) โŠจ {G1, G2, G3},
making the ๐–ฆ๐–ฑ-structure (M๐‘› (โ„), 0๐‘› , +) a group.
โˆ™ Matrix addition is commutative because addition on โ„ is commutative.
Therefore, (M๐‘› (โ„), 0๐‘› , +) is an abelian group.
โˆ™ Matrix multiplication is associative.
โˆ™ Lastly, to show that the operations are distributive, we must show for all
๐ด, ๐ต, ๐ถ ∈ M2 (โ„),
๐ด(๐ต + ๐ถ) = ๐ด๐ต + ๐ด๐ถ
and
(๐ด + ๐ต)๐ถ = ๐ด๐ถ + ๐ต๐ถ.
Therefore, ๐”๐‘› (โ„) โŠจ {R1, R2, R3, R4, R5}, so ๐”๐‘› (โ„) is a ring.
โˆ™ Since I๐‘› is the multiplicative identity, ℜ is a ring with unity, and since
matrix multiplication is not commutative, ℜ is a noncommutative ring.
This proves that ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฆ ⊗ ๐‘ฅ) is not a consequence of the ring
axioms.
EXAMPLE 7.1.40
If ๐‘… is the domain of a ring, ๐‘Ž, ๐‘ ∈ ๐‘… โงต {0} are zero divisors of the ring means
that ๐‘Ž ⋅ ๐‘ = 0. Defining addition and multiplication coordinatewise (Exercise 18),
the ring (โ„ค × โ„ค, (0, 0), +, ⋅) has zero divisors such as
(1, 0) ⋅ (0, 1) = (0, 0).
Other examples can be found in M2 (โ„) where
[
] [
] [
1 1
1
1
0
⋅
=
0 0
−1 −1
0
However,
[
1
−1
] [
1
1
⋅
−1
0
] [
1
1
=
0
−1
]
0
.
0
]
1
,
−1
showing that an element can be a left zero divisor but not a right zero divisor.
This situation is common for rings where multiplication is not commutative. We
do, however, have many rings that do not have zero divisors. An integral domain
is a commutative ring with unity that does not have zero divisors. The rings
(โ„ค, 0, +, ⋅), (โ„š, 0, +, ⋅), (โ„, 0, +, ⋅), and (โ„‚, 0, +, ⋅) are integral domains.
Section 7.1 FIRST-ORDER SEMANTICS
357
The equation 2๐‘ฅ + 1 = 0 is written with elements of โ„ค and the operations of regular
addition and multiplication. Although โ„ค has no zero divisors, there is no integer that is a
solution to this equation. To solve the equation, we need the existence of multiplicative
inverses. Let (๐‘…, 0, +, ⋅) be a ring with unity. If ๐‘ข ∈ ๐‘… has the property that there
exists ๐‘ฃ ∈ ๐‘… such that ๐‘ข ⋅ ๐‘ฃ = ๐‘ฃ ⋅ ๐‘ข = 1, then ๐‘ข is called a unit. Notice that units
are multiplicative inverses of each other. With this terminology, we make the next
definition.
DEFINITION 7.1.41
Let (๐‘…, 0, +, ⋅) be a ring with unity.
โˆ™ If all nonzero elements of ๐‘… are units, ๐‘… is called a division ring or sometimes a skew field.
โˆ™ A commutative division ring is called a field.
The reason that the equation on page 353 can be solved the way it was is that โ„ with
addition and multiplication form a field.
EXAMPLE 7.1.42
While โ„ค is not a field with standard addition and multiplication, โ„š, โ„, and โ„‚ are.
A more interesting structure is (โ„ค๐‘ , [0]๐‘ , +, ⋅) when ๐‘ is a prime. To prove that it
is a field, let [๐‘Ž]๐‘ ∈ โ„ค๐‘ so that [๐‘Ž]๐‘ ≠ [0]๐‘ . We must find an element of โ„ค๐‘ so
that when it is multiplied with [๐‘Ž]๐‘ the result is [1]๐‘ . Since [๐‘Ž]๐‘ ≠ [0]๐‘ , ๐‘ does
not divide ๐‘Ž. Hence, ๐‘ and ๐‘Ž are relatively prime, so there are integers ๐‘ข and ๐‘ฃ
such that ๐‘ข๐‘Ž + ๐‘ฃ๐‘ = 1. We are then able to calculate:
[๐‘ข]๐‘ ⋅ [๐‘Ž]๐‘ = [๐‘ข๐‘Ž]๐‘
= [1 − ๐‘ฃ๐‘]๐‘
= [1]๐‘ + [−๐‘ฃ๐‘]๐‘
= [1]๐‘ + [0]๐‘
= [1]๐‘ .
EXAMPLE 7.1.43
Let ℜ be a division ring and take ๐‘ข and ๐‘ฃ to be elements of the domain of ℜ. Let
1 be unity. Assume ๐‘ข๐‘ฃ = 0 and ๐‘ข ≠ 0. Then, ๐‘ข−1 exists, and we can calculate
๐‘ข−1 (๐‘ข๐‘ฃ) = ๐‘ข−1 0,
(๐‘ข−1 ๐‘ข)๐‘ฃ = 0,
1๐‘ฃ = 0,
๐‘ฃ = 0.
Therefore, ℜ has no zero divisors.
358
Chapter 7 MODELS
Exercises
1. Prove the remaining parts of Theorem 7.1.7.
2. Let 4 be a linear order on a nonempty set ๐ด. Let ๐‘… be a binary relation symbol.
Define the {๐‘…}-structure ๐”„ = (๐ด, 4). Let ๐ผ be an ๐–ฒ-interpretation of ๐”„ such that
๐ผ(๐‘…) = 4. Prove the following.
(a) ๐”„ โŠจ ∃๐‘ฅ(๐‘ฅ๐‘…๐‘ฅ) [๐ผ].
(b) ๐”„ โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ๐‘…๐‘ฆ ∨ ๐‘ฆ๐‘…๐‘ฅ) [๐ผ].
(c) ๐”„ โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ๐‘…๐‘ฆ ∧ ๐‘ฆ๐‘…๐‘ฅ → ๐‘ฅ = ๐‘ฆ) [๐ผ].
3. Let ๐”… be the ๐–ญ๐–ณ-structure (โ„ค, 0๐”… , 1๐”… , +๐”… , ⋅๐”… ), where 0๐”… and 1๐”… are the numbers 0
and 1 in โ„ค while +๐”… and ⋅๐”… are the standard operations of addition and multiplication
of integers. Let ๐ผ be a ๐–ญ๐–ณ-interpretation such that ๐ผ(๐‘ฅ) = 2 and ๐ผ(๐‘ฆ) = −2. Prove the
following.
(a) ๐”… โŠจ ๐‘ฅ + ๐‘ฆ = 0 [๐ผ].
(b) ๐”… โŠจ ∃๐‘ฅ([๐‘ฅ + 1] + 1 = 0) [๐ผ].
(c) ๐”… โŠจ ∃๐‘ฅ∀๐‘ฆ(๐‘ฅ ⋅ ๐‘ฆ = ๐‘ฆ) [๐ผ].
(d) ๐”… โŠจ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(¬๐‘ง = 0 ∧ ๐‘ฅ ⋅ ๐‘ง = ๐‘ฆ ⋅ ๐‘ง → ๐‘ฅ = ๐‘ฆ) [๐ผ].
4. Show that ๐”„ โŠจ ∀๐‘ฅ[(๐‘ฅ + ๐‘ฅ5 ) + ๐‘ฅ8 = ๐‘ฅ + (๐‘ฅ5 + ๐‘ฅ8 )], where ๐”„ is the ๐–ญ๐–ณ-structure of
Example 7.1.10.
5. Find a set of theory symbols ๐–ฒ, an ๐–ฒ-structure ๐”„, and an ๐–ฒ-interpretation ๐ผ such that
๐”„ โŠจ ๐‘ [๐ผ] for each given formula ๐‘.
(a) ๐‘ฅ + ๐‘ฆ = ([1 + 1] + 1) + 1
(b) ๐‘ฅโˆ•๐‘ฆ + ๐‘ง = 10
(c) ∃๐‘ฅ∃๐‘ฆ(๐‘ฅ < ๐‘ฆ ∧ ๐‘ฅ + 1 = ๐‘ฆ)
(d) ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(๐‘ฅ๐‘…๐‘ฆ ∧ ๐‘ฆ๐‘…๐‘ง → ๐‘ง๐‘…๐‘ฅ)
(e) ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⋅ ๐‘ฆ = 0 → ๐‘ฅ = 0 ∨ ๐‘ฆ = 0)
6. For each formula in Exercise 5, find a model (๐”„, ๐ผ) such that ๐”„ ฬธโŠจ ๐‘ [๐ผ] for each
given formula ๐‘.
7. Prove that every ๐–ฒ-structure is a model of the empty set.
8. Let ๐ด be a set. Is (P(๐ด), ∅, ∩) a group? Explain.
9. Explain why (โ„ค+ , 0, +), (โ„ค, 1, ⋅), and (โ„, 1, ⋅) are not groups, where the operations
are the standard ones.
10. Suppose that ∗ is an operation on โ„ค defined by ๐‘ฅ ∗ ๐‘ฆ = ๐‘ฅ + ๐‘ฆ + 2.
(a) Identify the identity ๐œ– and the inverses with respect to ∗.
(b) Prove that (โ„ค, ๐œ–, ∗) is a group, where ๐œ– is the identity found in Exercise 10(a).
(c) Solve 8 ∗ ๐‘ฅ = 10.
11. Let 0 represent the zero function โ„ → โ„ and + be function addition. That is, For
all ๐‘ฅ ∈ โ„,
(๐‘“ + ๐‘”)(๐‘ฅ) = ๐‘“ (๐‘ฅ) + ๐‘”(๐‘ฅ).
Section 7.1 FIRST-ORDER SEMANTICS
359
(a) Prove that (โ„ โ„, 0, +) is a group.
(b) Is โ„ โ„ the domain of a group where the binary operation is function division?
If so, what is its identity?
(c) Is โ„ โ„ the domain of a group where the binary operation is composition? If
so, what is identity?
12. Let ๐‘› be a positive integer.
(a) Prove that matrix multiplication is associative.
(b) Solve the equation in ๐‘€2 (โ„):
[
] [
] [
]
1 4
๐‘Ž ๐‘
−3
8
+
=
.
−3 0
๐‘ ๐‘‘
0 −6
(c) Show that ๐‘€๐‘› (โ„) is not the domain of a group under matrix multiplication.
13. Let (๐บ, ๐œ–, ∗) and (๐บ′ , ๐œ– ′ , ∗′ ) be two groups. For all ๐‘Ž, ๐‘ ∈ ๐บ and ๐‘Ž′ , ๐‘′ ∈ ๐บ′ define
(๐‘Ž, ๐‘Ž′ ) ⋅ (๐‘, ๐‘′ ) = (๐‘Ž ∗ ๐‘, ๐‘Ž′ ∗′ ๐‘′ ).
(a) Confirm that ⋅ is a binary operation on ๐บ × ๐บ′ .
(b) Show that (๐บ × ๐บ′ , (๐œ–, ๐œ– ′ ), ⋅) is a group. Prove that it is abelian if and only if
both of the given groups are abelian.
14. Let ๐‘› be an integer. Prove that (๐‘›โ„ค, 0, +, ⋅) is a commutative ring.
15. Why is
{[
๐‘Ž
๐‘
๐‘
๐‘‘
]
: ๐‘Ž, ๐‘, ๐‘, ๐‘‘, ∈ โ„ค
+
}
not the domain of a ring under the standard matrix operations?
16. Prove that the set
{[
]
}
๐‘Ž 0
: ๐‘Ž, ๐‘ ∈ โ„
0 ๐‘
is the domain of a ring with the standard matrix operations.
17. Both + (function addition) and โ—ฆ (composition) are binary operations on โ„ โ„, but
(โ„ โ„, 0, +, โ—ฆ) is not a ring. Identify which ring axioms fail.
18. Let (๐‘…, 0, +, ⋅) and (๐‘…′ , 0′ , +′ , ⋅′ ) be rings. Define addition and multiplication on
๐‘… × ๐‘…′ so that for all (๐‘Ž, ๐‘), (๐‘, ๐‘‘) ∈ ๐‘… × ๐‘…′ ,
(๐‘Ž, ๐‘) + (๐‘, ๐‘‘) = (๐‘Ž + ๐‘, ๐‘ +′ ๐‘‘),
and
(๐‘Ž, ๐‘) ⋅ (๐‘, ๐‘‘) = (๐‘Ž ⋅ ๐‘, ๐‘ ⋅′ ๐‘‘).
Prove that (๐‘… × ๐‘…′ , (0, 0), +, ⋅) is a ring.
19. Prove the converse of Example 7.1.24.
20. Prove that
{G1, G2, G3} โŠจ ∀๐‘ฅ∀๐‘ฆ[∀๐‘ง(๐‘ฅ โ—ฆ ๐‘ง = ๐‘ง ∧ ๐‘ง โ—ฆ ๐‘ฅ = ๐‘ฅ)
∧ ∀๐‘ง(๐‘ฆ โ—ฆ ๐‘ง = ๐‘ง ∧ ๐‘ง โ—ฆ ๐‘ฆ = ๐‘ง) → ๐‘ฅ = ๐‘ฆ].
360
Chapter 7 MODELS
21. Let ๐”Š = (๐บ, ๐œ–, ∗) be a group so that
๐”Š โŠจ ∀๐‘ฅ∀๐‘ฆ[(๐‘Ž โ—ฆ ๐‘) โ—ฆ (๐‘Ž โ—ฆ ๐‘) = ๐‘Ž โ—ฆ ๐‘Ž โ—ฆ ๐‘ โ—ฆ ๐‘].
Show that ๐”Š is abelian.
22. Prove that ∀๐‘ฅ(โ—‹ ⊗ ๐‘ฅ = โ—‹) is a consequence of the ring axioms.
23. Let − be a unary function symbol. Define ๐–ฑ๐–จ′ = ๐–ฑ๐–จ ∪ {−}. Show that the given
sentences are consequences of the ring axioms and ∀๐‘ฅ(−๐‘ฅ ⊕ ๐‘ฅ = โ—‹).
(a) ∀๐‘ฅ∀๐‘ฆ[−(๐‘ฅ ⊗ ๐‘ฆ) = −๐‘ฅ ⊗ ๐‘ฆ ∧ −๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฅ ⊗ −๐‘ฆ]
(b) ∀๐‘ฅ∀๐‘ฆ(−๐‘Ž ⊗ −๐‘ = ๐‘Ž ⊗ ๐‘)
(c) ∀๐‘ฅ∀[−(๐‘Ž ⊕ ๐‘) = −๐‘Ž ⊕ −๐‘]
(d) −โ—‹ = โ—‹
24. This exercise uses the notation of Exercise 23. Let ℜ be a ring with unity. Let ℜ′
be the expansion of ℜ to ๐–ฑ๐–จ′ = ๐–ฑ๐–จ ∪ {−}. Assume that for all ๐‘Ÿ ∈ dom(ℜ′ ),
−ℜ (๐‘Ÿ) ⊕ℜ ๐‘Ÿ = โ—‹ℜ .
Prove that ℜ′ โŠจ ∀๐‘ฅ[∀๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฆ ∧ ๐‘ฆ ⊗ ๐‘ฅ = ๐‘ฆ) → ∀๐‘ง(−๐‘ง = −๐‘ฅ ⊗ ๐‘ง)].
25. Let ๐‘ and ๐‘ž be ๐–ฒ-formulas. Prove that the given ๐–ฒ-sentences are valid.
(a) ๐‘ ∨ ¬๐‘
(b) ๐‘ → ๐‘ž ↔ ¬๐‘ ∨ ๐‘ž
(c) ∃๐‘ฅ(๐‘ ∨ ๐‘ž) ↔ ∃๐‘ฅ๐‘ ∨ ∃๐‘ฅ๐‘ž
(d) ∀๐‘ฅ(๐‘ ∧ ๐‘ž) ↔ ∀๐‘ฅ๐‘ ∧ ∀๐‘ฅ๐‘ž
26. Let ๐‘… be a binary relation symbols and ๐‘“ be a binary function symbol. Show that
the given sentences are satisfiable.
(a) ∃๐‘ฅ(๐‘ฅ = ๐‘ฅ)
(b) ∃๐‘ฅ∃๐‘ฆ∃๐‘ง(¬๐‘ฅ = ๐‘ฆ ∧ ¬๐‘ฅ = ๐‘ง ∧ ¬๐‘ฆ = ๐‘ง)
(c) ∃๐‘ฅ∀๐‘ฆ(๐‘…๐‘ฅ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ)
(d) ∀๐‘ฅ∀๐‘ฆ(๐‘“ ๐‘ฅ๐‘ฆ = 1)
(e) ∀๐‘ฅ∀๐‘ฆ[๐‘…๐‘ฅ๐‘ฆ → ∃๐‘ง(๐‘…๐‘ฅ๐‘ง ∧ ๐‘…๐‘ง๐‘ฆ)]
27. Let ๐–ฒ and ๐–ณ be sets of theory symbols such that ๐–ฒ ⊆ ๐–ณ. Let ๐”„ be a reduct to ๐–ฒ of
the ๐–ณ-structure ๐”…. Prove that ๐”„ โŠจ ๐‘ if and only if ๐”… โŠจ ๐‘ for all ๐–ฒ-sentences ๐‘.
28. Suppose that ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 are ๐–ฒ-sentences. For every ๐–ฒ-structure ๐”„, prove that
๐”„ โŠจ ๐‘0 ∧ ๐‘1 ∧ · · · ∧ ๐‘๐‘›−1 if and only if ๐”„ โŠจ ๐‘๐‘– for all ๐‘– = 0, 1, … , ๐‘› − 1.
29. Answer the following about (โ„ค๐‘› , [0]๐‘› , +, ⋅):
(a) Prove that addition and multiplication of congruence classes is well-defined.
(b) Show that the additive identity is [0]๐‘› .
(c) For all ๐‘Ž ∈ โ„ค, show that − [๐‘Ž]๐‘› = [๐‘› − ๐‘Ž]๐‘› .
(d) Show that [1]๐‘› is the multiplicative identity.
(e) Prove that (โ„ค๐‘› , [0]๐‘› , +, ⋅) is a commutative ring.
(f) Prove that the ring contains zero divisors when ๐‘› is not prime.
Section 7.2 SUBSTRUCTURES
361
30. Prove that ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(๐‘ฅ ⊕ ๐‘ฆ = ๐‘ฅ ⊕ ๐‘ง → ๐‘ฆ = ๐‘ง) is a consequence of the ring axioms.
31. Let ℜ be an integral domain. Prove the following.
(a) ℜ โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = โ—‹ → ๐‘ฅ = โ—‹ ∨ ๐‘ฆ = 0).
(b) ℜ โŠจ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฅ ⊗ ๐‘ง ∧ ๐‘ฅ ≠ โ—‹ → ๐‘ฆ = ๐‘ง).
32. Suppose that ℜ = (๐‘…, 0, +, ⋅) is a commutative ring with unity. Show that if
ℜ โŠจ ∀๐‘ฅ∀๐‘ฆ∃๐‘ง(๐‘ฅ ⊗ ๐‘ง ⊕ ๐‘ฆ = โ—‹), then ℜ a field.
33. Is ({0}, 0, +, ⋅) a field? Explain.
7.2
SUBSTRUCTURES
When looking for examples of groups, the ๐–ฆ๐–ฑ-structure (โ„ค, 0, +) is often the first to
come to mind. The benefit of this example is that not only are we familiar with the
integers but it has the property that many of its subsets also form groups. Let ๐‘› ∈ โ„ค.
Addition on โ„ค restricted to ๐‘›โ„ค × ๐‘›โ„ค is an associative binary operation on ๐‘›โ„ค, every
element of ๐‘›โ„ค has an additive inverse in ๐‘›โ„ค, and 0 ∈ ๐‘›โ„ค, so the ๐–ฆ๐–ฑ-structure (๐‘›โ„ค, 0, +)
is a group. Since ๐‘› ≠ ±1 implies that ๐‘›โ„ค ⊂ โ„ค, there are infinitely many different
examples of ๐–ฆ๐–ฑ-structures, all within (โ„ค, 0, +). We generalize this idea to arbitrary
structures.
DEFINITION 7.2.1
If ๐”„ = (๐ด, ๐”ž) and ๐”… = (๐ต, ๐”Ÿ) are ๐–ฒ-structures, ๐”„ is a substructure of ๐”… (written
as ๐”„ ⊆ ๐”…) means that ๐ด ⊆ ๐ต and the following properties hold.
โˆ™ ๐”ž(๐‘) = ๐”Ÿ(๐‘) for all constant symbols ๐‘.
โˆ™ ๐”ž(๐‘…) = ๐”Ÿ(๐‘…) ∩ ๐ด๐‘› for every ๐‘›-ary relation symbol ๐‘….
โˆ™ ๐”ž(๐‘“ ) = ๐”Ÿ(๐‘“ ) ๐ด๐‘› for every ๐‘›-ary function symbol ๐‘“ .
If ๐”„ is a substructure of ๐”…, then ๐”… is an extension of ๐”„.
Note the difference between a substructure and a reduct and between an extension
and an expansion (page 352). For all ๐‘› ∈ โ„ค, the group (๐‘›โ„ค, 0, +) is a substructure
of (โ„ค, 0, +), and (โ„ค, 0, +) is an extension of (๐‘›โ„ค, 0, +). Here both structures have the
same set of theory symbols, and the domain of one is a subset of the other. However,
(โ„ค, 0, +) is a reduct of (โ„ค, 0, 1, +, ⋅), and (โ„ค, 0, 1, +, ⋅) is an expansion of (โ„ค, 0, +). In
this case, the domains are the same, but the theory symbol set of the one is a subset of
the theory symbol set of the other.
EXAMPLE 7.2.2
Let ๐‘… be a binary relation symbol. Let ๐”„ = ([0, 1] , ๐”ž) and ๐”… = ([0, 2] , ๐”Ÿ) be
{๐‘…}-structures such that ๐”ž(๐‘…) and ๐”Ÿ(๐‘…) are both standard less-than. That is,
๐”ž(๐‘…) = {(๐‘ฅ, ๐‘ฆ) ∈ โ„ × โ„ : 0 ≤ ๐‘ฅ < ๐‘ฆ ≤ 1}
362
Chapter 7 MODELS
and
๐”Ÿ(๐‘…) = {(๐‘ฅ, ๐‘ฆ) ∈ โ„ × โ„ : 0 ≤ ๐‘ฅ < ๐‘ฆ ≤ 2}.
We conclude that ๐”„ ⊆ ๐”… because of the following:
โˆ™ [0, 1] ⊆ [0, 2].
โˆ™ There are no constant symbols.
โˆ™ ๐”ž(๐‘…) = ๐”Ÿ(๐‘…) ∩ ([0, 1] × [0, 1]).
โˆ™ There are no function symbols.
EXAMPLE 7.2.3
Let ๐‘› ∈ โ„ค. Define the ๐–ญ๐–ณ-structure ๐”… = (โ„ค, ๐”Ÿ), where ๐”Ÿ(0) is the additive
identity of โ„ค, ๐”Ÿ(1) is the multiplicative identity of โ„ค, ๐”Ÿ(+) is standard addition
on โ„ค, and ๐”Ÿ(⋅) is standard multiplication on โ„ค. Let ๐”„๐‘› = (๐‘›โ„ค, ๐”ž) such that
๐”ž(0) = ๐”Ÿ(0),
๐”ž(1) = ๐”Ÿ(1),
๐”ž(+) = ๐”Ÿ(+) (๐‘›โ„ค × ๐‘›โ„ค),
๐”ž(⋅) = ๐”Ÿ(⋅) (๐‘›โ„ค × ๐‘›โ„ค).
Then, ๐”„ is a substructure of ๐”….
In particular, Example 7.2.3 gives
๐”„8 ⊆ ๐”„4 ⊆ ๐”„2 ,
which implies that ๐”„8 ⊆ ๐”„2 . This is a special case of the next theorem.
THEOREM 7.2.4
Let ๐”„, ๐”…, and โ„ญ be ๐–ฒ-structures.
โˆ™ ๐”„ ⊆ ๐”„.
โˆ™ If ๐”„ ⊆ ๐”… and ๐”… ⊆ โ„ญ, then ๐”„ ⊆ โ„ญ.
PROOF
That ๐”„ is a substructure of itself is clear, so suppose that ๐”„ is a substructure of
๐”… and ๐”… is a substructure of โ„ญ. Write ๐”„ = (๐ด, ๐”ž), ๐”… = (๐ต, ๐”Ÿ), and โ„ญ = (๐ถ, ๐” ).
Then, for all constant symbols ๐‘,
๐”ž(๐‘) = ๐”Ÿ(๐‘) = ๐” (๐‘).
Since ๐”„ ⊆ ๐”…, ๐”ž(๐‘…) = ๐”Ÿ(๐‘…) ∩ ๐ด๐‘› , and since ๐”… ⊆ โ„ญ, ๐”Ÿ(๐‘…) = ๐” (๐‘…) ∩ ๐ต ๐‘› for every
๐‘›-ary relation symbol ๐‘…, so
๐”ž(๐‘…) = ๐” (๐‘…) ∩ ๐ต ๐‘› ∩ ๐ด๐‘› = ๐” (๐‘…) ∩ ๐ด๐‘› .
Section 7.2 SUBSTRUCTURES
363
Also, ๐”ž(๐‘“ ) = ๐”Ÿ(๐‘“ ) ๐ด๐‘› and ๐”Ÿ(๐‘“ ) = ๐” (๐‘“ ) ๐ต ๐‘› for all ๐‘›-ary function symbols ๐‘“ ,
so
๐”ž(๐‘“ ) = (๐” (๐‘“ ) ๐ต ๐‘› ) ๐ด๐‘› = ๐” (๐‘“ ) ๐ด๐‘› .
Therefore, ๐”„ is a substructure of โ„ญ.
Subgroups
Let ๐‘Ž be an element of a group (๐บ, ๐œ–, ∗). For all positive integers ๐‘›, define ๐‘Ž๐‘› to be the
result of operating ๐‘Ž with itself ๐‘› times. That is,
๐‘Ž1 = ๐‘Ž, ๐‘Ž2 = ๐‘Ž ∗ ๐‘Ž, ๐‘Ž3 = ๐‘Ž ∗ ๐‘Ž ∗ ๐‘Ž, …
and
๐‘Ž๐‘š ∗ ๐‘Ž๐‘› = ๐‘Ž๐‘š+๐‘› .
Further, define ๐‘Ž0 = ๐‘’ and ๐‘Ž−1 to be the inverse of ๐‘Ž. With this notation, we observe
that
(๐‘Ž ∗ ๐‘)−1 = ๐‘−1 ∗ ๐‘Ž−1
and
๐‘Ž−๐‘› = (๐‘Ž๐‘› )−1 = (๐‘Ž−1 )๐‘› .
We then gather all of these elements into a set,
โŸจ๐‘ŽโŸฉ = {๐‘Ž๐‘› : ๐‘› ∈ โ„ค},
and define the following.
DEFINITION 7.2.5
A group ๐”Š is cyclic if there exists ๐‘Ž ∈ dom(๐”Š) such that dom(๐”Š) = โŸจ๐‘ŽโŸฉ. The
element ๐‘Ž is called a generator of ๐”Š.
For example, (โ„ค, 0, +) is a cyclic group. Both 1 and −1 are generators. However, โ„š
and โ„ paired with addition do not form cyclic groups. As for finite groups, each โ„ค๐‘› is
cyclic, generated by [1]๐‘› , but the Klein-4 group (Example 7.1.18) is not cyclic because
๐‘Ž2 = ๐‘’ for all ๐‘Ž in the group.
An element ๐‘Ž of a group might not generate the entire group, but since ๐‘’ ∈ โŸจ๐‘ŽโŸฉ and
both ๐‘Ž๐‘› and ๐‘Ž−๐‘› are elements of โŸจ๐‘ŽโŸฉ, the set generated by ๐‘Ž forms a group using the
operation from ๐”Š.
DEFINITION 7.2.6
A substructure โ„Œ of a group ๐”Š that is a group is called a subgroup of ๐”Š.
Every group with at least two elements has at least two subgroups, itself (the improper subgroup) and the subgroup with domain {๐œ–} (the trivial subgroup). A group
that has at most these two subgroups is called simple. For example, (โ„ค2 , 0, +) and
(โ„ค3 , 0, +) form simple groups, but (โ„ค4 , 0, +) does not because it has a subgroup with domain {[0]4 , [2]4 }. Other examples of nonsimple groups are (โ„, 0, +) because (โ„ค, 0, +)
364
Chapter 7 MODELS
is one of its subgroups and (โŸจ2โŸฉ, 0, +) because (โŸจ6โŸฉ, 0, +) is one of its subgroups. These
subgroups that are not improper are called proper.
It is tempting to define a subgroup simply as a substructure of a group, but this would
not work if the subgroup is to be a group. For example, viewing ๐œ” as a subset of โ„ค via
(5.8) allows (๐œ”, 0, +) to be a ๐–ฆ๐–ฑ-substructure of (โ„ค, 0, +), but
(๐œ”, 0, +) โŠจ {G1, G2}
yet
(๐œ”, 0, +) ฬธโŠจ G3.
This example suggests the following.
THEOREM 7.2.7
A substructure โ„Œ of a group ๐”Š is a subgroup of ๐”Š if and only if โ„Œ โŠจ G3.
PROOF
Write ๐”Š = (๐บ, ๐”ค) and โ„Œ = (๐ป, ๐”ฅ) and let โ„Œ ⊆ ๐”Š. If โ„Œ is a subgroup, then
โ„Œ โŠจ G3. To prove the converse, assume โ„Œ โŠจ G3.
โˆ™ Let ๐‘ฅ, ๐‘ฆ, ๐‘ง ∈ ๐ป. Since โ„Ž(โ—ฆ) = ๐‘”(โ—ฆ) (๐ป × ๐ป),
๐”ฅ(โ—ฆ)(๐‘ฅ, ๐”ฅ(โ—ฆ)(๐‘ฆ, ๐‘ง)) = ๐”ค(โ—ฆ)(๐‘ฅ, ๐”ค(โ—ฆ)(๐‘ฆ, ๐‘ง))
= ๐”ค(โ—ฆ)(๐”ค(โ—ฆ)(๐‘ฅ, ๐‘ฆ), ๐‘ง)
= ๐”ฅ(โ—ฆ)(๐”ฅ(โ—ฆ)(๐‘ฅ, ๐‘ฆ), ๐‘ง).
The second equality holds because the interpretation of โ—ฆ in ๐”Š is associative.
โˆ™ Let ๐‘ฅ ∈ ๐ป. Because ๐”ฅ(๐‘’) = ๐”ค(๐‘’),
๐”ฅ(โ—ฆ)(๐”ฅ(๐‘’), ๐‘ฅ) = ๐”ค(โ—ฆ)(๐”ค(๐‘’), ๐‘ฅ) = ๐‘ฅ
and
๐”ฅ(โ—ฆ)(๐‘ฅ, ๐”ฅ(๐‘’)) = ๐”ค(โ—ฆ)(๐‘ฅ, ๐”ค(๐‘’)) = ๐‘ฅ.
Therefore, โ„Œ is a group and, thus, a subgroup of ๐”Š.
The standard way to show that a subset of a group forms a subgroup is not to show
directly that the set satisfies the three group axioms or to appeal to Theorem 7.2.7.
Instead, what is typically done in algebra is to check that the conditions of the next
theorem are satisfied by the set.
THEOREM 7.2.8
If ๐”Š = (๐บ, ๐œ–, ∗) is a group and ๐ป ⊆ ๐บ, there exists a subgroup of ๐”Š with domain
๐ป if
โˆ™ ๐ป is closed under ∗,
โˆ™ ๐œ– ∈ ๐ป,
โˆ™ ๐‘Ž−1 ∈ ๐ป for all ๐‘Ž ∈ ๐ป.
Section 7.2 SUBSTRUCTURES
365
PROOF
Suppose that the three hypotheses of the theorem hold.
โˆ™ Let ๐‘Ž, ๐‘ ∈ ๐ป. By the first hypothesis, ๐‘Ž ∗ ๐‘ ∈ ๐ป. Therefore, ∗ (๐ป × ๐ป)
is a binary operation on ๐ป.
โˆ™ Since ∗ is associative on ๐บ, the restriction of ∗ to ๐ป must be associative.
โˆ™ The second hypothesis gives ๐ป an identity element.
โˆ™ Every element of ๐ป has its inverse in ๐ป by the third hypothesis.
Therefore, (๐ป, ๐œ–, ∗ [๐ป × ๐ป]) is a group. Since ๐ป ⊆ dom(๐”Š), we conclude that
(๐ป, ๐œ–, ∗ [๐ป × ๐ป]) is a subgroup of ๐”Š.
EXAMPLE 7.2.9
To illustrate the theorem, take a group ๐”Š = (๐บ, ๐œ–, ∗) and a family of subgroups
(๐ป๐‘– , ๐œ–, ∗ [๐ป๐‘– × ๐ป๐‘– ]) for all ๐‘– ∈ ๐ผ. Although the union of subgroups might not
be a subgroup (Exercise 9), we can show that
(โ‹‚
๐ป๐‘– , ๐œ–, ∗ ๐‘–∈๐ผ
[โ‹‚
๐ป๐‘– ×
๐‘–∈๐ผ
โ‹‚
๐ป๐‘–
])
๐‘–∈๐ผ
is a subgroup of ๐”Š.
โˆ™ By Exercise 3.4.22(b),
โ‹‚
๐‘–∈๐ผ
๐ป๐‘– ⊆ ๐บ.
โ‹‚
โˆ™ Let ๐‘Ž, ๐‘ ∈ ๐‘–∈๐ผ ๐ป๐‘– . This means that ๐‘Ž, ๐‘ ∈ ๐ป๐‘– for all ๐‘– ∈ ๐ผ. Since each
๐ป๐‘– is closed under the operation of ๐”Š, ๐‘Ž ∗ ๐‘ ∈ ๐ป๐‘– for all ๐‘– ∈ ๐ผ. Hence,
๐‘Ž∗๐‘∈
โ‹‚
๐ป๐‘– .
๐‘–∈๐ผ
โˆ™ Since ๐œ– ∈ ๐ป๐‘– for every ๐‘– ∈ ๐ผ, we must have ๐œ– ∈
โˆ™ Take ๐‘Ž to be an element of
โ‹‚
๐‘Ž−1 ∈ ๐‘–∈๐ผ ๐ป๐‘– .
โ‹‚
๐‘–∈๐ผ
Now we return to cyclic groups.
THEOREM 7.2.10
A subgroup of a cyclic group is cyclic.
โ‹‚
๐‘–∈๐ผ
๐ป๐‘– .
๐ป๐‘– . Then, ๐‘Ž−1 ∈ ๐ป๐‘– for all ๐‘– ∈ ๐ผ, so
366
Chapter 7 MODELS
PROOF
Let ๐”Š = (๐บ, ๐œ–, ∗) be a cyclic group with generator ๐‘Ž. Let โ„Œ = (๐ป, ๐œ–, ∗) be a
subgroup of ๐”Š. If โ„Œ is the trivial subgroup, the subgroup is cyclic with generator
๐œ–. So suppose that โ„Œ is not the trivial subgroup. Because ๐”Š is cyclic, there exists
a least natural number ๐‘› > 0 such that ๐‘Ž๐‘› ∈ ๐ป (Theorem 5.2.13). Suppose that
๐‘Ž๐‘› is not a generator of โ„Œ. This means that there exists ๐‘š ∈ ๐œ” with ๐‘š > ๐‘›
such that ๐‘Ž๐‘š ∈ ๐ป but ๐‘Ž๐‘š ∉ โŸจ๐‘Ž๐‘› โŸฉ. This combined with the division algorithm
(Theorem 4.3.31) yields unique natural numbers ๐‘ž and ๐‘Ÿ such that ๐‘š = ๐‘›๐‘ž + ๐‘Ÿ
and 0 < ๐‘Ÿ < ๐‘›. Therefore,
๐‘Ž๐‘š = ๐‘Ž๐‘›๐‘ž+๐‘Ÿ = ๐‘Ž๐‘›๐‘ž ∗ ๐‘Ž๐‘Ÿ ,
and from this, we conclude that
๐‘Ž๐‘Ÿ = ๐‘Ž−๐‘›๐‘ž ∗ ๐‘Ž๐‘š .
Since ๐‘Ž−๐‘›๐‘ž , ๐‘Ž๐‘š ∈ ๐ป, ๐‘Ž๐‘Ÿ is an element of ๐ป. This contradicts the minimality of ๐‘›
because ๐‘Ÿ < ๐‘›. Thus, ๐‘Ž๐‘› is a generator of โ„Œ.
Subrings
Some of the examples of rings had domains that were subsets of other rings. For example, ๐‘›โ„ค is a subset of โ„ค, and โ„š is a subset of โ„. Generalizing leads to the next
definition.
DEFINITION 7.2.11
A substructure ๐”– of a ring ℜ that is a ring is called a subring of ℜ.
A subring of ℜ such that its domain is a proper subset of the domain of ℜ is called
a proper subring. The ring itself is called the improper subring. The subring with
domain {0} is the trivial subring.
EXAMPLE 7.2.12
โˆ™ ({[0]9 , [3]9 , [6]9 }, [0]9 , +, ⋅) is a subring of (โ„ค9 , [0]9 , +, ⋅).
โˆ™ (โ„ค, 0, +, ⋅) is a subring of (โ„, 0, +, ⋅).
โˆ™ (M2 (โ„), 02 , +, ⋅) is a subring of (M2 (โ„‚), 02 , +, ⋅).
As with subgroups, a substructure of a ring is not necessarily a subring, but we do
have results similar to those for groups found in Theorems 7.2.8 and 7.2.7. They are
stated without proof since they follow quickly from Definition 7.2.11.
THEOREM 7.2.13
A substructure ๐”– of a ring ℜ is a subring of ℜ if and only if ๐”– โŠจ R4.
We follow the convention that if ℜ represents an arbitrary ring, ℜ = (๐‘…, 0, +, ⋅) and
if ℜ′ also represents an arbitrary ring, ℜ′ = (๐‘…′ , 0′ , +′ , ⋅′ ). This will help us with our
notation.
Section 7.2 SUBSTRUCTURES
367
THEOREM 7.2.14
If ℜ is a ring and ๐‘† ⊆ ๐‘…, there exists a subring of ℜ with domain ๐‘† if
โˆ™ ๐‘† is closed under + and ⋅,
โˆ™ 0 ∈ ๐‘†,
โˆ™ −๐‘Ž ∈ ๐‘† for all ๐‘Ž ∈ ๐‘….
The subring found while proving Theorem 7.2.14 is (๐‘†, 0, + [๐‘† × ๐‘†], ⋅ [๐‘† × ๐‘†]).
EXAMPLE 7.2.15
We use Theorem 7.2.14 to show that ๐”– = (๐‘†, 02 , + [๐‘† × ๐‘†], ⋅ [๐‘† × ๐‘†]) is a
subring of ๐”2 (โ„) (Example 7.1.39), where
{[
๐‘†=
]
}
๐‘Ž 0
: ๐‘Ž, ๐‘ ∈ โ„ .
0 ๐‘
โˆ™ Let ๐‘Ž, ๐‘, ๐‘Ž′ , ๐‘′ ∈ โ„, and assume that
๐ด=
Then,
[
]
[ ′
๐‘Ž 0
๐‘Ž
and ๐ต =
0 ๐‘
0
[
๐‘Ž + ๐‘Ž′
๐ด+๐ต =
0
and
๐ด๐ต =
[ ′
๐‘Ž๐‘Ž
0
0
๐‘ + ๐‘′
]
0
.
๐‘๐‘′
These are elements of ๐‘†.
โˆ™ Clearly, the zero matrix is in ๐‘†. (Let ๐‘Ž = ๐‘ = 0.)
โˆ™ Take ๐‘Ž, ๐‘ ∈ โ„ and write
Hence,
is an element of ๐‘†.
[
]
๐‘Ž 0
๐ด=
.
0 ๐‘
[
−๐‘Ž
−๐ด =
0
]
0
−๐‘
]
0
.
๐‘′
]
368
Chapter 7 MODELS
EXAMPLE 7.2.16
Let ๐”– and ๐”— be subrings of a ring ℜ. Let ๐‘† be the domain of ๐”– and ๐‘‡ be the
domain of ๐”—. Check the conditions of Theorem 7.2.14 to show that there exists
a subring of ℜ with domain ๐‘† ∩ ๐‘‡ .
โˆ™ To prove closure, let ๐‘ฅ, ๐‘ฆ ∈ ๐‘† ∩๐‘‡ . This means that ๐‘ฅ+๐‘ฆ ∈ ๐‘† and ๐‘ฅ+๐‘ฆ ∈ ๐‘‡ .
Hence, ๐‘ฅ + ๐‘ฆ ∈ ๐‘† ∩ ๐‘‡ . Similarly, ๐‘ฅ๐‘ฆ ∈ ๐‘† ∩ ๐‘‡ .
โˆ™ Since 0 ∈ ๐‘† and 0 ∈ ๐‘‡ , 0 ∈ ๐‘† ∩ ๐‘‡ .
โˆ™ Suppose ๐‘ฅ ∈ ๐‘† ∩ ๐‘‡ . Then ๐‘ฅ ∈ ๐‘† and ๐‘ฅ ∈ ๐‘‡ . Since these are subrings,
−๐‘ฅ ∈ ๐‘† and −๐‘ฅ ∈ ๐‘‡ . Thus, −๐‘ฅ ∈ ๐‘† ∩ ๐‘‡ .
Ideals
The subring ๐”– of the ring ℜ in Example 7.2.15 lacks a property that is often desirable
to have in a subring. Observe that
[
] [
] [
]
1 0
1 2
1 2
⋅
=
∉ ๐‘†,
0 1
3 4
3 4
so in general it is false that ๐ด๐ต ∈ ๐‘† and ๐ต๐ด ∈ ๐‘† for all ๐ด ∈ ๐‘† and ๐ต ∈ M2 (โ„).
DEFINITION 7.2.17
Let ℜ be a ring with domain ๐‘… and ℑ be a subring of ℜ with domain ๐ผ.
โˆ™ If ๐‘Ÿ๐‘Ž ∈ ๐ผ and ๐‘Ž๐‘Ÿ ∈ ๐ผ for all ๐‘Ÿ ∈ ๐‘… and ๐‘Ž ∈ ๐ผ, then ℑ is an ideal of ℜ.
โˆ™ If ๐‘Ÿ๐‘Ž ∈ ๐ผ for all ๐‘Ÿ ∈ ๐‘… and ๐‘Ž ∈ ๐ผ, then ℑ is a left ideal.
โˆ™ If ๐‘Ž๐‘Ÿ ∈ ๐ผ for all ๐‘Ÿ ∈ ๐‘… and ๐‘Ž ∈ ๐ผ, then ℑ is a right ideal.
A ring ℜ is an ideal of itself, called the improper ideal of ℜ. All other ideals of ℜ are
proper, including the ideal formed by {0}. Furthermore, in a commutative ring, there
is no difference between a left and right ideal. However, if the ring is not commutative,
a left ideal might not be a right ideal.
EXAMPLE 7.2.18
Define
{[
]
}
๐‘Ž 0
๐ผ=
: ๐‘Ž, ๐‘ ∈ โ„ .
๐‘ 0
Using matrix multiplication,
[
] [
] [
]
๐‘ฅ ๐‘ฆ
๐‘Ž 0
๐‘ฅ๐‘Ž + ๐‘ฆ๐‘ 0
⋅
=
∈ ๐ผ,
๐‘ง ๐‘ค
๐‘ 0
๐‘ง๐‘Ž + ๐‘ค๐‘ 0
Section 7.2 SUBSTRUCTURES
but
] [
[
] [
1
1 0
1 1
=
⋅
1 1
1
1 0
369
]
1
∉ ๐ผ.
1
Therefore, the subring (๐ผ, 02 , + [๐ผ × ๐ผ], ⋅ [๐ผ × ๐ผ]) of ๐”2 (โ„) is a left ideal but
not a right ideal.
EXAMPLE 7.2.19
Let ๐ผ = {[0]4 , [2]4 }. Then, ℑ = (๐ผ, [0]4 , + [๐ผ × ๐ผ], ⋅ [๐ผ × ๐ผ]) is a subring
of the ring ℜ = (โ„ค4 [0]4 , +, ⋅). It also forms an ideal. To see this, check the
calculations:
[0]4 ⋅ [0]4
[1]4 ⋅ [0]4
[2]4 ⋅ [0]4
[3]4 ⋅ [0]4
[0]4 ⋅ [2]4
[1]4 ⋅ [2]4
[2]4 ⋅ [2]4
[3]4 ⋅ [2]4
= [0]4
= [0]4
= [0]4
= [0]4
= [0]4
= [2]4
= [0]4
= [2]4
[0]4 ⋅ [0]4
[0]4 ⋅ [1]4
[0]4 ⋅ [2]4
[0]4 ⋅ [3]4
[2]4 ⋅ [0]4
[2]4 ⋅ [1]4
[2]4 ⋅ [2]4
[2]4 ⋅ [3]4
= [0]4 ,
= [0]4 ,
= [0]4 ,
= [0]4 ,
= [0]4 ,
= [2]4 ,
= [0]4 ,
= [2]4 .
When we multiply any element of โ„ค4 by an element of ๐ผ on either side, the result
is an element of ๐ผ.
EXAMPLE 7.2.20
Let ℜ be a ring. It is left to Exercise 18 to show that the ring
๐”– = (๐‘… × {0}, (0, 0), +′ , ⋅′ )
is a subring of
๐”— = (๐‘… × ๐‘…, (0, 0), +, ⋅)
with + and ⋅ defined coordinatewise and +′ and ⋅′ being the restrictions of + and ⋅
to ๐‘…×{0}. To show that ๐”– is an ideal of ๐”—, let (๐‘Ÿ, ๐‘ ) ∈ ๐‘…×๐‘… and (๐‘Ž, 0) ∈ ๐‘…×{0}.
We then calculate
(๐‘Ÿ, ๐‘ ) ⋅ (๐‘Ž, 0) = (๐‘Ÿ๐‘Ž, 0) ∈ ๐‘… × {0}
and
(๐‘Ž, 0) ⋅ (๐‘Ÿ, ๐‘ ) = (๐‘Ž๐‘Ÿ, 0) ∈ ๐‘… × {0}.
EXAMPLE 7.2.21
Let ℜ be a ring. Let ๐‘† and ๐‘‡ be subsets of ๐‘… such that
๐”– = (๐‘†, 0, + [๐‘† × ๐‘†], ⋅ [๐‘† × ๐‘†])
370
Chapter 7 MODELS
and
๐”— = (๐‘‡ , 0, + [๐‘‡ × ๐‘‡ ], ⋅ [๐‘‡ × ๐‘‡ ])
are ideals of ℜ. Define
๐‘† + ๐‘‡ = {๐‘  + ๐‘ก : ๐‘  ∈ ๐‘† and ๐‘ก ∈ ๐‘‡ }.
We prove that ๐‘† + ๐‘‡ is the domain of an ideal of ℜ.
โˆ™ Take ๐‘ฅ, ๐‘ฆ ∈ ๐‘† + ๐‘‡ . This means that ๐‘ฅ = ๐‘  + ๐‘ก and ๐‘ฆ = ๐‘ ′ + ๐‘ก′ for some
๐‘ , ๐‘ ′ ∈ ๐‘† and ๐‘ก, ๐‘ก′ ∈ ๐‘‡ . Then,
๐‘ฅ + ๐‘ฆ = (๐‘  + ๐‘ก) + (๐‘ ′ + ๐‘ก′ )
= ๐‘  + (๐‘ก + ๐‘ ′ ) + ๐‘ก′
= ๐‘  + (๐‘ ′ + ๐‘ก) + ๐‘ก′
= (๐‘  + ๐‘ ′ ) + (๐‘ก + ๐‘ก′ ).
Thus, ๐‘ฅ + ๐‘ฆ ∈ ๐‘† + ๐‘‡ since ๐‘  + ๐‘ ′ ∈ ๐‘† and ๐‘ก + ๐‘ก′ ∈ ๐‘‡ . Also,
๐‘ฅ๐‘ฆ = (๐‘  + ๐‘ก)(๐‘ ′ + ๐‘ก′ ) = (๐‘  + ๐‘ก)๐‘ ′ + (๐‘  + ๐‘ก)๐‘ก′ .
Since ๐‘  + ๐‘ก ∈ ๐‘… and ๐”– is an ideal, (๐‘  + ๐‘ก)๐‘ ′ ∈ ๐‘†. Likewise, (๐‘  + ๐‘ก)๐‘ก′ is an
element of ๐‘‡ . Hence, ๐‘ฅ๐‘ฆ ∈ ๐‘† + ๐‘‡ .
โˆ™ We know that 0 ∈ ๐‘† + ๐‘‡ because 0 = 0 + 0 and 0 ∈ ๐‘† and 0 ∈ ๐‘‡ .
โˆ™ Let ๐‘  ∈ ๐‘† and ๐‘ก ∈ ๐‘‡ . Then, since −๐‘  ∈ ๐‘† and −๐‘ก ∈ ๐‘ก,
−(๐‘  + ๐‘ก) = −๐‘  + −๐‘ก ∈ ๐‘† + ๐‘‡ .
โˆ™ Let ๐‘Ÿ ∈ ๐‘…, ๐‘  ∈ ๐‘†, and ๐‘ก ∈ ๐‘‡ . Since ๐‘Ÿ๐‘  ∈ ๐‘† and ๐‘Ÿ๐‘ก ∈ ๐‘‡ ,
๐‘Ÿ(๐‘  + ๐‘ก) = ๐‘Ÿ๐‘  + ๐‘Ÿ๐‘ก ∈ ๐‘† + ๐‘‡ ,
and since ๐‘ ๐‘Ÿ ∈ ๐‘† and ๐‘ก๐‘Ÿ ∈ ๐‘‡ ,
(๐‘  + ๐‘ก)๐‘Ÿ = ๐‘ ๐‘Ÿ + ๐‘ก๐‘Ÿ ∈ ๐‘† + ๐‘‡ .
Cyclic subgroups (Definition 7.2.5) are generated by a single element. The corresponding notion in rings is the following definition.
DEFINITION 7.2.22
Let ℜ be a ring. For every ๐‘Ž ∈ ๐‘…, define
โŸจ๐‘ŽโŸฉ = {๐‘Ÿ๐‘Ž : ๐‘Ÿ ∈ ๐‘…}.
If ℑ is a left ideal of ℜ such that dom(ℑ) = โŸจ๐‘ŽโŸฉ for some ๐‘Ž ∈ ๐‘…, then ℑ is called
a principal ideal left ideal and ๐‘Ž is a generator of ℑ. If ℜ is commutative, ℑ
is an ideal of ℜ called a principal ideal.
The set ๐‘›โ„ค is the domain of an ideal of the ring (โ„ค, 0, +, ⋅). It is a principal ideal because
๐‘›โ„ค = โŸจ๐‘›โŸฉ. (See Exercises 16 and 21.) In fact, every element of a ring generates a left
ideal of the ring.
Section 7.2 SUBSTRUCTURES
371
THEOREM 7.2.23
For every ring ℜ and ๐‘Ž ∈ ๐‘…, the ring ℑ = (โŸจ๐‘ŽโŸฉ, 0, + [โŸจ๐‘ŽโŸฉ × โŸจ๐‘ŽโŸฉ], ⋅ [โŸจ๐‘ŽโŸฉ × โŸจ๐‘ŽโŸฉ])
is a left ideal of ℜ.
PROOF
First, show that ℑ is a subring.
โˆ™ Let ๐‘Ÿ, ๐‘  ∈ ๐‘…. Then, because ๐‘Ÿ + ๐‘  and (๐‘Ÿ๐‘Ž)๐‘  are elements of dom(๐‘…),
๐‘Ÿ๐‘Ž + ๐‘ ๐‘Ž = (๐‘Ÿ + ๐‘ )๐‘Ž ∈ โŸจ๐‘ŽโŸฉ
and
(๐‘Ÿ๐‘Ž)(๐‘ ๐‘Ž) = [(๐‘Ÿ๐‘Ž)๐‘ ] ๐‘Ž ∈ โŸจ๐‘ŽโŸฉ.
โˆ™ Since 0๐‘Ž = 0 [Exercise 7.1.22], we have that 0 ∈ โŸจ๐‘ŽโŸฉ.
โˆ™ If ๐‘Ÿ ∈ ๐‘…, then (−๐‘Ÿ)๐‘Ž ∈ โŸจ๐‘ŽโŸฉ and −๐‘Ÿ๐‘Ž + ๐‘Ÿ๐‘Ž = (−๐‘Ÿ + ๐‘Ÿ)๐‘Ž = 0๐‘Ž = 0.
To prove that ℑ is a left ideal of ℜ, take ๐‘Ÿ, ๐‘  ∈ ๐‘…. Then, ๐‘Ÿ(๐‘ ๐‘Ž) ∈ โŸจ๐‘ŽโŸฉ because
๐‘Ÿ๐‘  ∈ ๐‘… and ๐‘Ÿ(๐‘ ๐‘Ž) = (๐‘Ÿ๐‘ )๐‘Ž.
Notice that a principal left ideal might not be a two-sided ideal. Example 7.2.18 combined with the next example illustrates this fact.
EXAMPLE 7.2.24
Let ℑ = (๐ผ, 02 , + [๐ผ × ๐ผ], ⋅ [๐ผ × ๐ผ]) be a left ideal of ๐”2 (โ„), where
๐ผ=
{[
]
}
๐‘Ž 0
: ๐‘Ž, ๐‘ ∈ โ„ .
๐‘ 0
By Theorem 7.2.23, it is a principal left ideal of ๐”2 (โ„) because
โŸจ[
๐ผ=
1
0
]โŸฉ
0
.
0
To prove this, it suffices to take ๐‘Ž, ๐‘ ∈ โ„ and observe that
[
] [
] [
๐‘Ž 0
๐‘Ž 0
1
=
⋅
๐‘ 0
๐‘ 0
0
implies
[
] โŸจ[
๐‘Ž 0
1
∈
๐‘ 0
0
]
0
0
]โŸฉ
0
.
0
372
Chapter 7 MODELS
EXAMPLE 7.2.25
Every ideal in the ring โ„จ = (โ„ค, 0, +, ⋅) is principal. To see this, let ℑ be an ideal
of โ„จ. We must find an element of โ„ค that generates ๐ผ = dom(ℑ). We have two
cases to consider:
โˆ™ If ๐ผ = {0}, then ๐ผ = โŸจ0โŸฉ.
โˆ™ Suppose ๐ผ ≠ {0}. This means that ๐ผ ∩ โ„ค+ ≠ ∅. By Theorem 5.2.13 and
Exercise 5.3.1, ๐ผ must contain a minimal positive integer. Call it ๐‘š. We
claim that ๐ผ = โŸจ๐‘šโŸฉ. It is clear that โŸจ๐‘šโŸฉ ⊆ ๐ผ, so take ๐‘Ž ∈ ๐ผ and divide it by
๐‘š. The division algorithm (Theorem 4.3.31) gives ๐‘ž, ๐‘Ÿ ∈ ๐œ” so that
๐‘Ž = ๐‘š๐‘ž + ๐‘Ÿ
with 0 ≤ ๐‘Ÿ < ๐‘š. Then, ๐‘Ÿ ∈ ๐ผ because ๐‘Ÿ = ๐‘Ž − ๐‘š๐‘ž and ๐‘Ž, ๐‘š๐‘ž ∈ ๐ผ.
If ๐‘Ÿ > 0, then we have a contradiction of the fact that ๐‘š is the smallest
positive integer in ๐ผ. Hence, ๐‘Ÿ = 0 and ๐‘Ž = ๐‘š๐‘ž. This means that ๐‘Ž ∈ โŸจ๐‘šโŸฉ.
โ„จ is an example of a principal ideal domain, an integral domain in which every
ideal is principal.
Exercises
1. Let ๐‘… be a binary relation symbol. Define the {๐‘…}-structures ๐”„ = (โ„š, <โ„š ) and
๐”… = (โ„, <โ„ ), where <โ„š refers to standard less-than on โ„š and <โ„ refers to standard
less-than on โ„. Prove that ๐”„ is a substructure of ๐”….
2. Let ๐”„ = (๐ด, ๐”ž), ๐”… = (๐ต, ๐”Ÿ), and โ„ญ = (๐ถ, ๐” ) be ๐–ฒ-structures such that ๐”… ⊆ ๐”„ and
โ„ญ ⊆ ๐”„. Define ๐”‡ = (๐ท, ๐”ก) such that ๐ท = ๐ต ∩ ๐ถ, ๐”ก(๐‘) = ๐”ž(๐‘) for all constant symbols
๐‘ ∈ ๐–ฒ, ๐”ก(๐‘…) = ๐”Ÿ(๐‘…) ∩ ๐” (๐‘…) for all relation symbols ๐‘… ∈ ๐–ฒ, and ๐”ก(๐‘“ ) = ๐”Ÿ(๐‘“ ) ∩ ๐” (๐‘“ ) for
all function symbols ๐‘“ ∈ ๐–ฒ. Prove that ๐”‡ ⊆ ๐”„.
3. Let ๐œ… be a cardinal. Define ๐”„๐›ผ = (๐ด๐›ผ , ๐”ž๐›ผ ) to be an ๐–ฒ-structure for all ๐›ผ ∈ ๐œ…. The
family {๐”„๐›พ : ๐›พ ∈ ๐œ…} is called a chain of ๐–ฒ-structures if for all ๐›ผ ∈ ๐›ฝ ∈ ๐œ…, we have that
โ‹ƒ
โ‹ƒ
๐”„๐›ผ ⊆ ๐”„๐›ฝ . Define the ๐–ฒ-structure ๐›พ∈๐œ… ๐”„๐›พ = ( ๐›พ∈๐œ… ๐ด๐›พ , ๐”ž) so that for every relation
symbol ๐‘… ∈ ๐–ฒ,
โ‹ƒ
๐”ž(๐‘…) =
๐”ž๐›ผ (๐‘…),
๐›พ∈๐œ…
and for every function symbol ๐‘“ ∈ ๐–ฒ,
๐”ž(๐‘“ ) =
โ‹ƒ
๐”ž๐›ผ (๐‘“ ).
๐›พ∈๐œ…
Prove the following.
(a) {๐”ž๐›พ (๐‘…) : ๐›พ ∈ ๐œ…} is a chain with respect to ⊆ for all relation symbols ๐‘… ∈ ๐–ฒ.
โ‹ƒ
(b)
๐›พ∈๐œ… ๐”ž๐›พ (๐‘“ ) is a function.
โ‹ƒ
(c) ๐”„๐›ผ ⊆ ๐›พ∈๐œ… ๐”„๐›พ for all ๐›ผ ∈ ๐œ….
Section 7.2 SUBSTRUCTURES
373
4. Let ๐–ฒ ⊆ ๐–ณ be sets of subject symbols. Let ๐”„ and ๐”… be ๐–ณ-structures. Prove that if
๐”„ ⊆ ๐”…, then the reduct of ๐”„ to ๐–ฒ is a substructure of the reduct of ๐”… to ๐–ฒ.
5. Find ๐–ฒ-substructures of the given ๐–ฒ-structures. If possible, find a ๐–ฒ-sentence that is
true in the given structure but not true in the substructure.
(a) (โ„ค ∪ P(โ„ค), ∈), ๐–ฒ = ๐–ฒ๐–ณ
(b) (๐œ”, ∅, {∅}, +, ⋅), ๐–ฒ = ๐–ญ๐–ณ
(c) (โ„‚, 0, +), ๐–ฒ = ๐–ฆ๐–ฑ
(d) (โ„, 0, +, ⋅, <), ๐–ฒ = ๐–ฎ๐–ฅ
6. Let ๐”Š be a group with domain ๐บ. Find all subgroups of ๐”Š given the following and
assuming the standard operations for each set.
(a) ๐บ is the Klein-4 group
(b) ๐บ = โ„ค5
(c) ๐บ = โ„ค8
(d) ๐บ = โ„ค2 × โ„ค3
(e) ๐บ = โ„ค2 × โ„ค6
7. Show that {๐‘“ ∈
(โ„ โ„, 0, +).
โ„โ„
: ๐‘“ (0) = 0} is the domain of a subgroup of the group
8. Define ๐ป = {๐ด ∈ M๐‘› (โ„) : ๐‘Ž1,1 + ๐‘Ž2,2 + · · · + ๐‘Ž๐‘›,๐‘› = 0}. Prove that (๐ป, 0๐‘› , +) is a
subgroup of the group (M๐‘› (โ„), 0๐‘› , +).
9. Let
โ‹ƒ {โ„Œ๐‘– : ๐‘– ∈ ๐ผ} be a family of subgroups of the group ๐”Š. Give an example to show
that ๐‘–∈๐ผ โ„Œ๐‘– is not necessarily the domain of a subgroup of ๐”Š.
10. Let (๐ป, ๐œ–, ∗ [๐ป × ๐ป]) and (๐พ, ๐œ–, ∗ [๐พ × ๐พ]) be two subgroups of an abelian
group ๐”Š = (๐บ, ๐œ–, ∗). Define
๐ป๐พ = {๐‘Ž ∗ ๐‘ : ๐‘Ž ∈ ๐ป ∧ ๐‘ ∈ ๐พ}.
Prove that (๐ป๐พ, ๐œ–, ∗ [๐ป๐พ × ๐ป๐พ]) is a subgroup of ๐”Š.
11. For any group ๐”Š = (๐บ, ๐œ–, ∗), let ๐‘† ⊆ ๐บ and define
๐ป = {๐‘Ž ∈ ๐บ : ๐‘Ž ∗ ๐‘ = ๐‘ ∗ ๐‘Ž for all ๐‘ ∈ ๐‘†}.
Prove that (๐ป, ๐œ–, ∗ [๐ป × ๐ป]) is a subgroup of ๐”Š.
12. Demonstrate that simple groups are cyclic.
13. Prove Theorem 7.2.14.
14. Prove that the subrings of a ring with unity are rings with unity.
15. Let ℜ be a ring with domain ๐‘…. Find all ideals of ℜ given the following and
assuming the standard operations for each set.
(a) ๐‘… = โ„ค2
(b) ๐‘… = โ„ค6
(c) ๐‘… = โ„ค7
374
Chapter 7 MODELS
(d) ๐‘… = โ„ค12
16. Let ๐‘› ∈ โ„ค. Prove that ๐‘›โ„ค is the domain of an ideal of the ring (โ„ค, 0, +, ⋅).
17. Let ℜ = (โ„ค × โ„ค, (0, 0), +, ⋅).
(a) Prove that {(2๐‘š, 2๐‘›) : ๐‘š, ๐‘› ∈ โ„ค} is the domain of an ideal of ℜ.
(b) Prove that {(2๐‘›, 2๐‘›) : ๐‘› ∈ โ„ค} is not the domain of an ideal of ℜ.
(c) Prove that โ„ค × 3โ„ค is the domain of a principal ideal of ℜ.
(d) Is {(2๐‘š, 2๐‘›) : ๐‘š, ๐‘› ∈ โ„ค} the domain of a principal ideal of ℜ?
18. Prove the ๐”– is a subring of ๐”— in Example 7.2.20.
19. Let ℜ be a ring with unity and ℑ an ideal of ℜ. Let ๐‘… be the domain of ℜ and ๐ผ
be the domain of ℑ. Prove that if ๐‘ข is a unit and ๐‘ข ∈ ๐ผ, then ๐ผ = ๐‘….
20. Prove that a field has no proper, nontrivial ideals.
21. Prove that ๐‘›โ„ค is the domain of a principal ideal of the ring (โ„ค, 0, +, ⋅) for any
integer ๐‘›.
22. Show that for any ideal ℑ, if ๐‘Ž ∈ dom(ℑ), then โŸจ๐‘ŽโŸฉ ⊆ dom(ℑ).
23. Take a ring ℜ and let ๐‘Ž ∈ dom(ℜ). Show โŸจ๐‘ŽโŸฉ is the domain of a left ideal of ℜ
but not necessarily a right ideal if ℜ is not commutative.
24. Let ๐‘ข be a unit of a ring ℜ. Show for all ๐‘Ž ∈ dom(ℜ), โŸจ๐‘ŽโŸฉ = โŸจ๐‘ข๐‘ŽโŸฉ.
25. An ideal ๐”“ = (๐‘ƒ , 0, +, ⋅) of a commutative ring ℜ = (๐‘…, 0, +, ⋅) is prime means
for all ๐‘Ž, ๐‘ ∈ ๐‘…, if ๐‘Ž๐‘ ∈ ๐‘ƒ , then ๐‘Ž ∈ ๐‘ƒ or ๐‘ ∈ ๐‘ƒ . Let ๐‘ ∈ โ„ค+ be a prime number
(Example 2.4.18). Prove that (๐‘โ„ค, 0, + ๐‘โ„ค, ⋅ ๐‘โ„ค) is a prime ideal of (โ„ค, 0, +, ⋅).
26. Prove that the trivial subring of an integral domain is a prime ideal.
27. Let ℜ be a commutative ring and ๐” a proper ideal of ℜ. If no proper ideal of ℜ
has ๐” as a proper ideal, ๐” is a maximal ideal of ℜ. Assume that ๐‘ is a prime number.
Prove that ๐‘โ„ค is the domain of a maximal ideal of (โ„ค, 0, +, ⋅).
28. Prove that the trivial ideal is the maximal ideal of a field.
29. Let ๐‘ ∈ โ„ค be prime. Prove that {(๐‘๐‘Ž, ๐‘) : ๐‘Ž, ๐‘ ∈ โ„ค} is the domain of a maximal
ideal of the ring with domain โ„ค × โ„ค (Example 7.2.20).
30. Use Zorn’s lemma (Theorem 5.1.13) to prove that every commutative ring with
unity has a maximal ideal.
7.3
HOMOMORPHISMS
The function ๐‘“ : [0, ∞) → (−∞, 0] defined by ๐‘“ (๐‘ฅ) = −๐‘ฅ preserves ≤ with ≥ (Example 4.5.25). Define the {4}-structures ๐”„ = ([0, ∞) , ๐”ž) and ๐”… = ((−∞, 0] , ๐”Ÿ), where
๐”ž(4) = ≤ and ๐”Ÿ(4) = ≥. When ๐‘ฅ0 , ๐‘ฅ1 ∈ [0, ∞), observe that
(๐‘ฅ0 , ๐‘ฅ1 ) ∈ ๐”ž(4) if and only if (๐‘“ (๐‘ฅ0 ), ๐‘“ (๐‘ฅ1 )) ∈ ๐”Ÿ(4).
Section 7.3 HOMOMORPHISMS
375
because
๐‘ฅ0 ≤ ๐‘ฅ1 if and only if −๐‘ฅ0 ≥ −๐‘ฅ1 .
Therefore, we know that 4 behaves in ๐”… as it behaves in ๐”„ because of ๐‘“ . We generalize
this notion to all ๐–ฒ-structures in the next definition.
DEFINITION 7.3.1
Let ๐”„ = (๐ด, ๐”ž) and ๐”… = (๐ต, ๐”Ÿ) be ๐–ฒ-structures. A function ๐œ‘ : ๐ด → ๐ต is a
homomorphism ๐”„ → ๐”… if it preserves the structure of ๐”„ in ๐”…. This means
that ๐œ‘ satisfies the following conditions.
โˆ™ ๐œ‘(๐”ž(๐‘)) = ๐”Ÿ(๐‘) for all constant symbols ๐‘ ∈ ๐–ฒ.
โˆ™ If ๐‘… is an ๐‘›-ary relation symbol in ๐–ฒ, then for all ๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 ∈ ๐ด,
(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 ) ∈ ๐”ž(๐‘…)
if and only if
(๐œ‘(๐‘Ž0 ), ๐œ‘(๐‘Ž1 ), … , ๐œ‘(๐‘Ž๐‘›−1 )) ∈ ๐”Ÿ(๐‘…).
โˆ™ If ๐‘“ is an ๐‘›-ary function symbol in ๐–ฒ, for all ๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 ∈ ๐ด,
๐œ‘(๐”ž(๐‘“ )(๐‘Ž0 , ๐‘Ž1 , … , ๐‘Ž๐‘›−1 )) = ๐”Ÿ(๐‘“ )(๐œ‘(๐‘Ž0 ), ๐œ‘(๐‘Ž1 ), … , ๐œ‘(๐‘Ž๐‘›−1 )).
When ๐œ‘ : ๐ด → ๐ต is a homomorphism, write ๐œ‘ : ๐”„ → ๐”….
There are many important examples of homomorphisms that can be found in algebra.
For instance, let ℜ and ℜ′ be rings and ๐œ‘ : ℜ → ℜ′ be a homomorphism. By
Definition 7.3.1, this means that
๐œ‘(0) = 0′ .
Because ⊕ and ⊗ are the only function symbols in ๐–ฑ๐–จ, for all ๐‘Ž, ๐‘ ∈ ๐‘…,
๐œ‘(๐‘Ž + ๐‘) = ๐œ‘(๐‘Ž) +′ ๐œ‘(๐‘)
and
๐œ‘(๐‘Ž ⋅ ๐‘) = ๐œ‘(๐‘Ž) ⋅′ ๐œ‘(๐‘).
This together with a similar analysis of homomorphisms between groups motivates the
next definition.
DEFINITION 7.3.2
โˆ™ A group homomorphism is a homomorphism ๐”Š → ๐”Š′ , where ๐”Š and ๐”Š′ are
groups.
โˆ™ A ring homomorphism is an homomorphism ℜ → ℜ′ , where ℜ and ℜ′ are
rings.
Throughout this section the focus will be on ring homomorphisms.
376
Chapter 7 MODELS
EXAMPLE 7.3.3
Let ℜ and ℜ′ be rings. The function ๐œ“ : ๐‘… → ๐‘…′ so that ๐œ“(๐‘Ÿ) = 0′ for all ๐‘Ÿ ∈ ๐‘…
is a ring homomorphism. This function is called a zero map.
EXAMPLE 7.3.4
Define the function ๐œ‘ : โ„ค → โ„ค × โ„ค by ๐œ‘(๐‘›) = (๐‘›, 0). Assume that + and ⋅ are
the standard operations on โ„ค while +′ and ⋅′ are the coordinatewise operations
on โ„ค × โ„ค. Let ๐‘š, ๐‘› ∈ โ„ค. Then,
โˆ™ ๐œ‘(0) = (0, 0)
โˆ™ ๐œ‘(๐‘š + ๐‘›) = (๐‘š + ๐‘›, 0) = (๐‘š, 0) +′ (๐‘›, 0) = ๐œ‘(๐‘š) +′ ๐œ‘(๐‘›)
โˆ™ ๐œ‘(๐‘š๐‘›) = (๐‘š๐‘›, 0) = (๐‘š, 0) ⋅′ (๐‘›, 0) = ๐œ‘(๐‘š) ⋅′ ๐œ‘(๐‘›).
Therefore, ๐œ‘ is a ring homomorphism (โ„ค, 0, +, ⋅) → (โ„ค × โ„ค, (0, 0), +′ , ⋅′ ).
The homomorphism of Example 7.3.4 provides a good opportunity to clarify what
a ring homomorphism does. Take the integers 1 and 4. Adding them together in โ„ค
yields 5. The images of these integers under ๐œ‘ are ๐œ‘(1) = (1, 0), ๐œ‘(4) = (4, 0), and
๐œ‘(5) = (5, 0). Observe that (1, 0) + (4, 0) = (5, 0) in โ„ค × โ„ค, illustrating that the ring
homomorphism ๐œ‘ preserves the addition structure of โ„ค in โ„ค × โ„ค. That is, with respect
to addition, both sets behave the same way. Multiplication also has the property. For
example, when 3 is multiplied with 6 the result is 18 in โ„ค, and their images yield
๐œ‘(3)๐œ‘(6) = ๐œ‘(18) in โ„ค × โ„ค. This is illustrated in Figure 7.2.
Although ring homomorphisms will always preserve the additive identity by definition, this is not a condition that needs to be checked.
๐œ‘
โ„ค
1
(1, 0)
4
(4, 0)
5
โ„ค
3
(5, 0)
๐œ‘
6
18
Figure 7.2
โ„ค×โ„ค
(3, 0)
โ„ค×โ„ค
(6, 0)
(18, 0)
The ring homomorphism ๐œ‘ : โ„ค → โ„ค × โ„ค defined by ๐œ‘(๐‘›) = (๐‘›, 0).
Section 7.3 HOMOMORPHISMS
377
THEOREM 7.3.5
Let ℜ and ℜ′ be rings. Then, ๐œ‘ : ๐‘… → ๐‘…′ is a ring homomorphism if and only
if for all ๐‘ฅ, ๐‘ฆ ∈ ๐‘…,
โˆ™ ๐œ‘(๐‘ฅ + ๐‘ฆ) = ๐œ‘(๐‘ฅ) +′ ๐œ‘(๐‘ฆ), and
โˆ™ ๐œ‘(๐‘ฅ ⋅ ๐‘ฆ) = ๐œ‘(๐‘ฅ) ⋅′ ๐œ‘(๐‘ฆ).
PROOF
Since sufficiency is clear, assume that ๐œ‘ preserves addition and multiplication.
Then,
๐œ‘(0) +′ ๐œ‘(0) = ๐œ‘(0 + 0) = ๐œ‘(0).
Adding −๐œ‘(0) to both sides yields ๐œ‘(0) = 0′ .
In addition to preserving operations and the additive identity, ring homomorphisms
also preserve inverses. Remember, if ๐œ‘ : ๐‘… → ๐‘…′ is a function and ๐‘Ž ∈ ๐‘…, then −๐‘Ž is
the additive inverse of ๐‘Ž in ๐‘… and −๐œ‘(๐‘Ž) is the additive inverse of ๐œ‘(๐‘Ž) inside of ๐‘…′ .
THEOREM 7.3.6
Let ℜ and ℜ′ be rings. If ๐œ‘ : ℜ → ℜ′ is a ring homomorphism, for all ๐‘Ž ∈ ๐‘…,
๐œ‘(−๐‘Ž) = −๐œ‘(๐‘Ž).
PROOF
If ๐œ‘ : ℜ → ℜ′ is a ring homomorphism, then for all ๐‘ฅ ∈ ๐‘…,
๐œ‘(๐‘ฅ) +′ ๐œ‘(−๐‘ฅ) = ๐œ‘(๐‘ฅ + −๐‘ฅ) = ๐œ‘(0) = 0′ .
EXAMPLE 7.3.7
Check Theorem 7.3.6 using the function ๐œ‘ as defined in the Example 7.3.4. Since
5 and −5 are additive inverses in โ„ค, we have that
(5, 0) + (−5, 0) = (0, 0),
so ๐œ‘(5) and ๐œ‘(−5) are additive inverses in โ„ค × โ„ค.
Ring homomorphisms also preserve subrings and ideals.
THEOREM 7.3.8
Let ℜ and ℜ′ be rings and ๐œ‘ : ℜ → ℜ′ be a ring homomorphism. Let ℑ be an
ideal of ℜ with domain ๐ผ and ℑ′ be an ideal of ℜ′ with domain ๐ผ ′ .
โˆ™ If ๐œ‘ is onto, (๐œ‘[๐ผ], 0′ , +′ ๐œ‘[๐ผ], ⋅′ ๐œ‘[๐ผ]) is an ideal of ℜ′ .
โˆ™ ๐” = (๐œ‘−1 [๐ผ ′ ], 0, + ๐œ‘−1 [๐ผ ′ ], ⋅ ๐œ‘−1 [๐ผ ′ ]) is an ideal of ℜ.
378
Chapter 7 MODELS
PROOF
The first part is left to Exercise 6. To prove the second, first prove that ๐” is a
subring of ℜ.
โˆ™ To show closure, let ๐‘ฅ1 , ๐‘ฅ2 ∈ ๐œ‘−1 [๐ผ ′ ]. This means that ๐œ‘(๐‘ฅ1 ) and ๐œ‘(๐‘ฅ2 )
are elements of ๐ผ ′ . Hence,
๐œ‘(๐‘ฅ1 + ๐‘ฅ2 ) = ๐œ‘(๐‘ฅ1 ) +′ ๐œ‘(๐‘ฅ2 ) ∈ ๐ผ ′
and
๐œ‘(๐‘ฅ1 ⋅ ๐‘ฅ2 ) = ๐œ‘(๐‘ฅ1 ) ⋅′ ๐œ‘(๐‘ฅ2 ) ∈ ๐ผ ′ .
Therefore, ๐‘ฅ1 + ๐‘ฅ2 , ๐‘ฅ1 ⋅ ๐‘ฅ2 ∈ ๐œ‘−1 [๐ผ ′ ].
โˆ™ Since 0′ ∈ ๐ผ ′ and ๐œ‘(0) = 0′ , it follows that 0 ∈ ๐œ‘−1 [๐ผ ′ ].
โˆ™ Let ๐‘ฅ ∈ ๐œ‘−1 [๐ผ ′ ]. Then, ๐œ‘(๐‘ฅ) ∈ ๐ผ ′ , and we find that
๐œ‘(−๐‘ฅ) = −๐œ‘(๐‘ฅ) ∈ ๐ผ ′ .
Thus, −๐‘ฅ ∈ ๐œ‘−1 [๐ผ ′ ].
To see that ๐” is an ideal of ℜ, take ๐‘Ÿ ∈ ๐‘… and ๐‘Ž ∈ ๐œ‘−1 [๐ผ ′ ]. This means that
๐œ‘(๐‘Ž) ∈ ๐ผ ′ . Then,
๐œ‘(๐‘Ÿ๐‘Ž) = ๐œ‘(๐‘Ÿ)๐œ‘(๐‘Ž) ∈ ๐ผ ′
since ℑ′ is an ideal. Thus, ๐‘Ÿ๐‘Ž ∈ ๐œ‘−1 [๐ผ ′ ]. Similarly, ๐‘Ž๐‘Ÿ ∈ ๐œ‘−1 [๐ผ ′ ].
EXAMPLE 7.3.9
The function ๐œ‘ : โ„ค → โ„คโˆ•6โ„ค defined by ๐œ‘(๐‘›) = ๐‘› + 6โ„ค is an onto homomorphism. (See Exercise 7.) The image of 2โ„ค under this map is
๐œ‘[2โ„ค] = {๐‘› + 6โ„ค : ๐‘› ∈ 2โ„ค} = {0 + 6โ„ค, 2 + 6โ„ค, 4 + 6โ„ค}.
The pre-image of ๐ผ = {0 + 6โ„ค, 3 + 6โ„ค} is
๐œ‘−1 [๐ผ] = {๐‘› ∈ โ„ค : ∃๐‘˜ ∈ โ„ค(๐‘› = 6๐‘˜ ∨ ๐‘› = 3 + 6๐‘˜)} = 3โ„ค.
Notice that both ๐œ‘[2โ„ค] and ๐œ‘−1 [๐ผ] are domains of ideals of their respective rings.
Take a ring ℜ and let ๐‘… be its domain such that |๐‘…| ≥ 2. Let 0 be the additive
identity of ℜ. We specify two functions. First, define ๐œ‹ : ๐‘… × ๐‘… × ๐‘… → ๐‘… × ๐‘… by
๐œ‹(๐‘ฅ, ๐‘ฆ, ๐‘ง) = (๐‘ฅ, ๐‘ฆ). This is a function similar to that found in Example 4.5.16. Notice
that ran(๐œ‹) = ๐‘… × ๐‘…, which means that ๐œ‹ is onto, and
๐ด = {(๐‘ฅ, ๐‘ฆ, ๐‘ง) : ๐œ‹(๐‘ฅ, ๐‘ฆ, ๐‘ง) = (0, 0)} = {(0, 0, ๐‘ง) : ๐‘ง ∈ ๐‘…}.
Because |๐ด| > 1, we conclude that ๐œ‹ is not one-to-one. Second, define the function
๐œ“ : ๐‘… × ๐‘… → ๐‘… × ๐‘… × ๐‘… so that ๐œ“(๐‘ฅ, ๐‘ฆ) = (๐‘ฅ, ๐‘ฆ, 0). Compare ๐œ“ with Example 4.5.7.
Section 7.3 HOMOMORPHISMS
379
This function is not onto because ran(๐œ“) = ๐‘… × ๐‘… × {0}, but it does appear to be
one-to-one because
๐ต = {(๐‘ฅ, ๐‘ฆ) : ๐œ“(๐‘ฅ, ๐‘ฆ) = (0, 0, 0)} = {(0, 0)}.
The sets ๐ด and ๐ต that contain the elements of the domain that are mapped to the identity
of the range appear important to determining whether a function is one-to-one. For this
reason, such sets are named.
DEFINITION 7.3.10
Let ℜ and ℜ′ be rings. Let ๐œ‘ : ๐‘… → ๐‘…′ be a function. The kernel of ๐œ‘ is
ker(๐œ‘) = {๐‘ฅ ∈ ๐‘… : ๐œ‘(๐‘ฅ) = 0′ }.
Notice that Definition 7.3.10 implies that ker(๐œ‘) = ๐œ‘−1 [{0′ }]. Therefore, ker(๐œ‘) is
an ideal of ℜ (Theorem 7.3.8). Similarly, because ran(๐œ‘) = ๐œ‘[๐‘…], we conclude that
ran(๐œ‘) is an ideal of ℜ′ .
EXAMPLE 7.3.11
Let ๐œ‘ : โ„ค → โ„ค5 defined by ๐œ‘(๐‘›) = [๐‘›]5 .
โˆ™ To find the kernel, assume ๐œ‘(๐‘›) = [0]5 . By the assumption, [๐‘›]5 = [0]5 .
So, ๐‘› ∈ [0]5 , which means 5 โˆฃ ๐‘›. Hence,
ker(๐œ‘) ⊆ 5โ„ค.
Since the steps are reversible, ker(๐œ‘) = 5โ„ค.
โˆ™ Because ran(๐œ‘) = {[๐‘›]5 โˆถ ๐‘› ∈ โ„ค} encompasses all congruence classes
modulo 5, ๐œ‘ is onto and ran(๐œ‘) = โ„ค5 .
We now prove that the kernel does provide a test for whether a ring homomorphism
is an injection. Since any ring homomorphism ๐œ‘ : ℜ → ℜ′ maps the additive identity
of ℜ to the additive identity of ℜ′ , it is always the case that 0 ∈ ker(๐œ‘). Therefore, to
show that ker(๐œ‘) = {0}, we only need to prove ker(๐œ‘) ⊆ {0}.
THEOREM 7.3.12
Let ℜ and ℜ′ be rings such that 0 is the additive identity of ℜ. Suppose that the
function ๐œ‘ : ℜ → ℜ′ is a ring homomorphism. Then, ๐œ‘ is one-to-one if and
only if ker(๐œ‘) = {0}.
PROOF
โˆ™ Suppose that ๐œ‘ is one-to-one. Take ๐‘ฅ ∈ ker(๐œ‘). This means that ๐œ‘(๐‘ฅ) = 0′ ,
where 0′ is the additive identity of ℜ′ . Since ๐œ‘ is an injection, ๐‘ฅ = 0. This
implies that ker(๐œ‘) = {0}.
380
Chapter 7 MODELS
โˆ™ Now let ker(๐œ‘) = {0} and assume ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ2 ), where ๐‘ฅ1 , ๐‘ฅ2 ∈ dom(ℜ). We
then have ๐œ‘(๐‘ฅ1 ) − ๐œ‘(๐‘ฅ2 ) = 0. Since ๐œ‘ is a homomorphism, we have
๐œ‘(๐‘ฅ1 − ๐‘ฅ2 ) = ๐œ‘(๐‘ฅ1 ) − ๐œ‘(๐‘ฅ2 )
by Theorem 7.3.6. Hence, ๐‘ฅ1 − ๐‘ฅ2 ∈ ker(๐œ‘), which means ๐‘ฅ1 − ๐‘ฅ2 = 0 by
hypothesis. Therefore, ๐‘ฅ1 = ๐‘ฅ2 .
A similar proof can be used to demonstrate that same result for group homomorphisms.
THEOREM 7.3.13
Let ๐”Š and ๐”Š′ be groups such that ๐œ– is the additive identity of ๐”Š. Suppose that
the function ๐œ‘ : ๐”Š → ๐”Š′ is a group homomorphism. Then, ๐œ‘ is one-to-one if
and only if ker(๐œ‘) = {๐œ–}.
Isomorphisms
Define ๐œ“ : โ„ค5 × โ„ค6 → โ„ค6 × โ„ค5 by ๐œ“([๐‘Ž]5 , [๐‘]6 ) = ([๐‘]6 , [๐‘Ž]5 ). Let + be coordinatewise
addition on โ„ค5 × โ„ค6 and +′ be coordinatewise addition on โ„ค6 × โ„ค5 . We prove that this
function preserves addition and is a bijection.
โˆ™ Let ๐‘Ž, ๐‘, ๐‘, ๐‘‘ ∈ โ„ค. Then,
๐œ“(([๐‘Ž]5 , [๐‘]6 ) + ([๐‘]5 , [๐‘‘]6 )) = ๐œ“([๐‘Ž]5 + [๐‘]5 , [๐‘]6 + [๐‘‘]6 )
= ๐œ“([๐‘Ž + ๐‘]5 , [๐‘ + ๐‘‘]6 )
= ([๐‘ + ๐‘‘]6 , [๐‘Ž + ๐‘]5 )
= ([๐‘]6 + [๐‘‘]6 , [๐‘Ž]5 + [๐‘]5 )
= ([๐‘]6 , [๐‘Ž]5 ) +′ ([๐‘‘]6 , [๐‘]5 )
= ๐œ“([๐‘Ž]5 , [๐‘]6 ) +′ ๐œ“([๐‘]5 , [๐‘‘]6 ).
โˆ™ Let ([๐‘Ž]5 , [๐‘]6 ) ∈ ker(๐œ“). In other words, ([๐‘]6 , [๐‘Ž]5 ) = ([0]6 , [0]5 ). Hence,
5 โˆฃ ๐‘Ž and 6 โˆฃ ๐‘, which implies that [๐‘Ž]5 = [0]5 and [๐‘]6 = [0]6 (Example 4.2.6).
Thus, ๐œ“ is an injection by Theorem 7.3.13.
โˆ™ To see that ๐œ“ is onto, take ([๐‘]6 , [๐‘‘]5 ) ∈ โ„ค6 × โ„ค5 . Then,
๐œ“([๐‘‘]5 , [๐‘]6 ) = ([๐‘]6 , [๐‘‘]5 ).
The function ๐œ“ generalizes to the next definition. Note the similarity between this
definition and that of an order isomorphism (Definition 4.5.24).
DEFINITION 7.3.14
Let ๐”„ and ๐”… be ๐–ฒ-structures. An isomorphism ๐”„ → ๐”… is a homomorphism
๐”„ → ๐”… that is a bijection. An isomorphism ๐”„ → ๐”„ is called an automorphism.
If there is an isomorphism ๐œ‘ : ๐”„ → ๐”…, the ๐–ฒ-structures are isomorphic and we
write ๐”„ ≅ ๐”….
Section 7.3 HOMOMORPHISMS
381
If ๐”„ = (โ„ค5 × โ„ค6 , ([0]5 , [0]6 ), +) and ๐”… = (โ„ค6 × โ„ค5 , ([0]6 , [0]5 ), +′ ), then ๐œ“ shows that
๐”„ ≅ ๐”….
Like order isomorphisms (Example 4.5.28), there are three basic isomorphism results. The proof of this is left to Exercise 8.
THEOREM 7.3.15
Let ๐”„, ๐”…, and โ„ญ be ๐–ฒ-structures.
โˆ™ ๐”„ ≅ ๐”„.
โˆ™ If ๐”„ ≅ ๐”…, then ๐”… ≅ ๐”„.
โˆ™ If ๐”„ ≅ ๐”… and ๐”… ≅ โ„ญ, then ๐”„ ≅ โ„ญ.
EXAMPLE 7.3.16
Let ๐–ฒ = {0, <, ⋅}. Let ๐”„ = (2 ๐œ”, ๐”ž), where
๐”ž(0)(0) = 0 and ๐”ž(0)(1) = 0,
and for all ๐‘“ , ๐‘” ∈ 2 ๐œ”,
(๐‘“ , ๐‘”) ∈ ๐”ž(<) if and only if ๐‘“ (0) < ๐‘”(0) ∨ [๐‘“ (0) = ๐‘”(0) ∧ ๐‘“ (1) < ๐‘”(1)],
that is, < is interpreted as a lexicographical order (compare Exercise 4.3.16), and
๐”ž(⋅) is function multiplication.
This means that ๐”ž(⋅)(๐‘“ , ๐‘”)(๐‘ฅ) = ๐‘“ (๐‘ฅ)๐‘”(๐‘ฅ). Also, let ๐”… = (๐œ” × ๐œ” × {0}, ๐”Ÿ), where
๐”Ÿ(0) = (0, 0, 0),
and for all ๐‘š, ๐‘›, ๐‘š′ , ๐‘›′ ∈ ๐œ”,
((๐‘š, ๐‘›, 0), (๐‘š′ , ๐‘›′ , 0)) ∈ ๐”Ÿ(<) if and only if ๐‘› < ๐‘›′ ∨ [๐‘› = ๐‘›′ ∧ ๐‘š < ๐‘š′ ],
and
๐”Ÿ(⋅) is coordinate-wise multiplication.
That is,
๐”Ÿ(⋅)((๐‘š, ๐‘›, 0), (๐‘š′ , ๐‘›′ , 0)) = (๐‘š๐‘š′ , ๐‘›๐‘›′ , 0).
Show that ๐”„ ≅ ๐”… by showing that ๐œ‘ : ๐”„ → ๐”… defined by
๐œ‘(๐œ“) = (๐œ“(1), ๐œ“(0), 0)
is an ๐–ฒ-isomorphism.
โˆ™ ๐œ‘(๐”ž(0)) = (๐”ž(0)(1), ๐”ž(0)(0), 0) = (0, 0, 0) = ๐”Ÿ(0).
382
Chapter 7 MODELS
โˆ™ Let ๐‘“ , ๐‘” ∈ 2 ๐œ”. Then,
(๐‘“ , ๐‘”) ∈ ๐”ž(<) ⇔ ๐‘“ (0) < ๐‘”(0) ∨ [๐‘“ (0) = ๐‘”(0) ∧ ๐‘“ (1) < ๐‘”(1)]
⇔ ((๐‘“ (1), ๐‘“ (0), 0), (๐‘”(1), ๐‘”(0), 0)) ∈ ๐”Ÿ(<)
⇔ (๐œ‘(๐‘“ ), ๐œ‘(๐‘”)) ∈ ๐”Ÿ(<).
โˆ™ Let ๐‘“ , ๐‘” ∈ 2 ๐œ”. Then,
๐œ‘(๐”ž(⋅)(๐‘“ , ๐‘”)) = ๐œ‘({(0, ๐‘“ (0)๐‘”(0)), (1, ๐‘“ (1)๐‘”(1))})
= (๐‘“ (1)๐‘”(1), ๐‘“ (0)๐‘”(0), 0)
= ๐”Ÿ(⋅)((๐‘“ (1), ๐‘“ (0), 0), (๐‘”(1), ๐‘”(0), 0))
= ๐”Ÿ(⋅)(๐œ‘(๐‘“ ), ๐œ‘(๐‘”)).
Therefore, ๐œ‘ is an homomorphism. That ๐œ‘ is a bijection is Exercise 9. Hence,
๐”„ ≅ ๐”…, and by Theorem 7.3.15, we have that ๐”… ≅ ๐”„.
Because there is a bijection between โ„ค5 × โ„ค6 and โ„ค6 × โ„ค5 , they have the same
cardinality. Since the bijection that is typically chosen is also a homomorphism, the
algebraic structure of the two rings are the same. Putting this together, we conclude
that the two rings “look” the same as rings. The only difference is in the labeling of
their elements. To formalize this notion, we make the next definition.
DEFINITION 7.3.17
โˆ™ A group homomorphism that is a bijection is a group isomorphism.
โˆ™ A ring homomorphism that is a bijection is a ring isomorphism.
We say that two rings ๐‘… and ๐‘…′ are isomorphic if there is an isomorphism ๐œ“ : ๐‘… → ๐‘…′ .
If two rings are isomorphic, write ๐‘… ≅ ๐‘…′ . For example, we saw that
โ„ค5 × โ„ค6 ≅ โ„ค6 × โ„ค5 .
The next example illustrates what we mean by “looking” the same.
EXAMPLE 7.3.18
Suppose that ๐œ‘ : ℜ → ℜ′ is a ring isomorphism.
โˆ™ If ℜ is an integral domain, ℜ′ is an integral domain.
โˆ™ If ℜ is a division ring, ℜ′ is a division ring.
โˆ™ If ℜ is a field, ℜ′ is a field.
The proofs of the first two are left to Exercise 13. For example, suppose that ℜ
is a field. We show that ℜ′ is also a field.
โˆ™ Let 1 be unity from ℜ. Then, ๐œ‘(1) is unity from ℜ′ (Exercise 14).
Section 7.3 HOMOMORPHISMS
383
โˆ™ Let ๐‘ฆ0 , ๐‘ฆ1 ∈ ๐‘…′ . Since ๐œ‘ is onto, there exists ๐‘ฅ0 , ๐‘ฅ1 ∈ ๐‘… so that ๐œ‘(๐‘ฅ0 ) = ๐‘ฆ0
and ๐œ‘(๐‘ฅ1 ) = ๐‘ฆ1 . Hence,
๐‘ฆ0 ⋅′ ๐‘ฆ1 = ๐œ‘(๐‘ฅ0 ) ⋅′ ๐œ‘(๐‘ฅ1 )
= ๐œ‘(๐‘ฅ0 ⋅ ๐‘ฅ1 )
= ๐œ‘(๐‘ฅ1 ⋅ ๐‘ฅ0 )
= ๐œ‘(๐‘ฅ1 ) ⋅′ ๐œ‘(๐‘ฅ0 )
= ๐‘ฆ1 ⋅′ ๐‘ฆ0 .
โˆ™ Let ๐‘ฆ0 ∈ ๐‘…′ โงต {0′ }. There exists ๐‘ฅ0 ∈ ๐‘… such that ๐œ‘(๐‘ฅ0 ) = ๐‘ฆ0 . Since ๐œ‘ is
one-to-one, ๐‘ฅ0 ≠ 0. Thus, because ℜ is a field, there exists ๐‘ฅ1 ∈ ๐‘… โงต {0}
such that ๐‘ฅ0 ⋅ ๐‘ฅ1 = 1, so
๐‘ฆ0 ⋅′ ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ0 ) ⋅′ ๐œ‘(๐‘ฅ1 ) = ๐œ‘(๐‘ฅ0 ⋅ ๐‘ฅ1 ) = ๐œ‘(1).
The result of the next example is known as the fundamental homomorphism theorem.
EXAMPLE 7.3.19
Let ℜ and ℜ′ be rings. Let ๐œ‘ : ℜ → ℜ′ be a surjective ring homomorphism.
Define ๐œ“ : ๐‘…โˆ• ker(๐œ‘) → ๐‘…′ by
๐œ“(๐‘Ž + ker(๐œ‘)) = ๐œ‘(๐‘Ž).
To prove that ๐œ“ is well-defined, let ๐‘Ž, ๐‘ ∈ ๐‘… such that ๐‘Ž + ker(๐œ‘) = ๐‘ + ker(๐œ‘).
This implies that ๐‘Ž − ๐‘ ∈ ker(๐œ‘), so
0 = ๐œ‘(๐‘Ž − ๐‘) = ๐œ‘(๐‘Ž) −′ ๐œ‘(๐‘).
That is, ๐œ‘(๐‘Ž) = ๐œ‘(๐‘). Now to show that ๐œ“ is a ring homomorphism.
โˆ™ Take ๐‘Ž, ๐‘ ∈ ๐‘…. Then,
๐œ“(๐‘Ž + ๐‘ + ker(๐œ‘)) = ๐œ‘(๐‘Ž + ๐‘)
= ๐œ‘(๐‘Ž) +′ ๐œ‘(๐‘)
= ๐œ“(๐‘Ž + ker(๐œ‘)) +′ ๐œ“(๐‘ + ker(๐œ‘)),
and
๐œ“(๐‘Ž ⋅ ๐‘ + ker(๐œ‘)) = ๐œ‘(๐‘Ž ⋅ ๐‘)
= ๐œ‘(๐‘Ž) ⋅′ ๐œ‘(๐‘)
= ๐œ“(๐‘Ž + ker(๐œ‘)) ⋅′ ๐œ“(๐‘ + ker(๐œ‘)).
โˆ™ To prove that ๐œ“ is onto, take ๐‘ ∈ ๐‘…′ . Since ๐œ‘ is onto, there exists ๐‘Ž ∈ ๐‘…
such that ๐œ‘(๐‘Ž) = ๐‘. Therefore,
๐œ“(๐‘Ž + ker(๐œ‘)) = ๐œ‘(๐‘Ž) = ๐‘.
384
Chapter 7 MODELS
โˆ™ To prove that ๐œ“ is one-to-one, let ๐‘Ž, ๐‘ ∈ ๐‘… and assume that ๐œ“(๐‘Ž) = ๐œ“(๐‘).
Since ๐œ“ is a ring homomorphism,
๐œ“(๐‘Ž − ๐‘) = ๐œ“(๐‘Ž) −′ ๐œ“(๐‘) = 0.
Hence, ๐‘Ž − ๐‘ ∈ ker(๐œ‘), which implies that ๐‘Ž + ker(๐œ‘) = ๐‘ + ker(๐œ‘).
Elementary Equivalence
Let ๐–  be a first-order language with theory symbols ๐–ฒ. Suppose that ๐”„ = (๐ด, ๐”ž) and
๐”… = (๐ต, ๐”Ÿ) are ๐–ฒ-structures and ๐ผ is an interpretation of ๐”„. Let ๐œ‘ : ๐”„ → ๐”… be an
isomorphism. Define
๐ผ๐œ‘ : ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐– ) → ๐ต
by ๐ผ๐œ‘ (๐‘ก) = (๐œ‘ โ—ฆ ๐ผ)(๐‘ก) for all ๐‘ก ∈ ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐– ). We confirm that ๐ผ๐œ‘ is an interpretation of
๐”….
โˆ™ If ๐‘ฅ is a variable symbol, ๐ผ๐œ‘ (๐‘ฅ) = (๐œ‘ โ—ฆ ๐ผ)(๐‘ฅ) ∈ ๐ต.
โˆ™ Let ๐‘ be a constant symbol. Then, ๐ผ๐œ‘ (๐‘) = (๐œ‘ โ—ฆ ๐ผ)(๐‘) = ๐œ‘(๐”ž(๐‘)) = ๐”Ÿ(๐‘).
โˆ™ Let ๐‘“ be an ๐‘›-ary function symbol and ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ∈ ๐–ณ๐–ค๐–ฑ๐–ฌ๐–ฒ(๐– ). Then,
๐ผ๐œ‘ (๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )) = (๐œ‘ โ—ฆ ๐ผ)(๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ))
= ๐œ‘(๐”ž(๐‘“ )(๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 )))
= ๐”Ÿ(๐‘“ )(๐œ‘(๐ผ(๐‘ก0 )), ๐œ‘(๐ผ(๐‘ก1 )), … , ๐œ‘(๐ผ(๐‘ก๐‘›−1 )))
= ๐”Ÿ(๐‘“ )(๐ผ๐œ‘ (๐‘ก0 ), ๐ผ๐œ‘ (๐‘ก1 ), … , ๐ผ๐œ‘ (๐‘ก๐‘›−1 )).
An example will clarify the use of the interpretation ๐ผ๐œ‘ .
EXAMPLE 7.3.20
Let ๐”Š0 = (โ„ค, 0, +) and ๐”Š1 = (2โ„ค, 0, + 2โ„ค). Assume that ๐ผ is an interpretation
of ๐”Š0 . Define the group isomorphism ๐‘“ : โ„ค → 2โ„ค by ๐‘“ (๐‘›) = 2๐‘›. We consider
the ๐–ฆ๐–ฑ-formula ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’.
๐”„ โŠจ ∃๐‘ฅ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ผ] ⇔ ๐”„ โŠจ (๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ โ„ค
⇔ ๐ผ๐‘ฅ๐‘Ž (๐‘ฆ โ—ฆ ๐‘ฅ) = ๐ผ๐‘ฅ๐‘Ž (๐‘’) for some ๐‘Ž ∈ โ„ค
⇔ ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘ฆ โ—ฆ ๐‘ฅ)) = ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘’)) for some ๐‘Ž ∈ โ„ค
⇔ ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘ฆ)) + ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘ฅ)) = ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘’)) for some ๐‘Ž ∈ โ„ค
⇔ ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘ฆ)) + ๐‘“ (๐‘Ž) = ๐‘“ (๐ผ๐‘ฅ๐‘Ž (๐‘’)) for some ๐‘Ž ∈ โ„ค
⇔ ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘ฆ)) + ๐‘ = ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘’)) for some ๐‘ ∈ 2โ„ค
⇔ ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘ฆ)) + ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘ฅ)) = ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘’)) for some ๐‘ ∈ 2โ„ค
⇔ ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘ฆ โ—ฆ ๐‘ฅ)) = ๐‘“ (๐ผ๐‘ฅ๐‘ (๐‘’)) for some ๐‘ ∈ 2โ„ค
⇔ ๐”… โŠจ ๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’ [(๐ผ๐‘“ )๐‘๐‘ฅ ] for some ๐‘ ∈ 2โ„ค
⇔ ๐”… โŠจ ∃๐‘ฅ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) [๐ผ๐‘“ ].
Section 7.3 HOMOMORPHISMS
385
Using ๐‘“ , we see that the interpretation of ๐”„ gives rise to an interpretation of ๐”….
For example, notice that if ๐ผ(๐‘ฆ) = 3, then ๐‘“ (๐ผ(๐‘ฆ)) = 6. Hence, in ๐”„,
∃๐‘ฅ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) is interpreted as 3 + ๐‘ฅ = 0 for some ๐‘ฅ ∈ ๐ด,
and the witness of ∃๐‘ฅ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) is −3. In ๐”…,
∃๐‘ฅ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) is interpreted as 6 + ๐‘ฅ = 0 for some ๐‘ฅ ∈ ๐ต
and the witness of ∃๐‘ฅ(๐‘ฆ โ—ฆ ๐‘ฅ = ๐‘’) is −6.
LEMMA 7.3.21
Let ๐”„ = (๐ด, ๐”ž) and ๐”… = (๐ต, ๐”Ÿ) be ๐–ฒ-structures. Assume that ๐œ‘ : ๐”„ → ๐”… is an
for all
isomorphism. Let ๐ผ be an ๐–ฒ-interpretation of ๐”„. Then, (๐ผ๐‘ฅ๐‘Ž )๐œ‘ = (๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ
๐‘Ž ∈ ๐ด.
PROOF
We proceed by induction on terms, relying on Definitions 7.1.4 and 7.3.1.
โˆ™ Let ๐‘ฅ and ๐‘ฆ be variable symbols. Then,
(๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘ฅ) = ๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘ฅ)) = ๐œ‘(๐‘Ž) = (๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ (๐‘ฅ),
and if ๐‘ฆ ≠ ๐‘ฅ,
(๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘ฆ) = ๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘ฆ)) = ๐œ‘(๐ผ(๐‘ฆ)) = (๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ (๐‘ฆ).
โˆ™ If ๐‘ is a constant symbol, then
(๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘) = ๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘)) = ๐œ‘(๐ผ(๐‘)) = (๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ (๐‘).
โˆ™ Let ๐‘“ be an ๐‘›-ary function symbol and ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 be ๐–ฒ-terms. Take
๐‘Ž ∈ ๐ด. Then,
(๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ))
= ๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )))
= ๐œ‘(๐”ž(๐‘“ )(๐ผ๐‘ฅ๐‘Ž (๐‘ก0 ), ๐ผ๐‘ฅ๐‘Ž (๐‘ก1 ), … , ๐ผ๐‘ฅ๐‘Ž (๐‘ก๐‘›−1 )))
= ๐”Ÿ(๐‘“ )(๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘ก0 )), ๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘ก1 )), … , ๐œ‘(๐ผ๐‘ฅ๐‘Ž (๐‘ก๐‘›−1 )))
= ๐”Ÿ(๐‘“ )((๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘ก0 ), (๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘ก1 ), … , (๐ผ๐‘ฅ๐‘Ž )๐œ‘ (๐‘ก๐‘›−1 ))
๐œ‘(๐‘Ž)
๐œ‘(๐‘Ž)
= ๐”Ÿ(๐‘“ )(((๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ )(๐‘ก0 ), ((๐ผ๐œ‘ )๐‘ฅ )(๐‘ก1 ), … , ((๐ผ๐œ‘ )๐‘ฅ )(๐‘ก๐‘›−1 ))
= (๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ (๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )).
The fifth equality follows by induction.
If ๐œ‘ : ๐”„ → ๐”… is an isomorphism, the structures look the same. Thus, formulas
should be interpreted in ๐”… essentially in the same way that they are interpreted in ๐”„.
In other words, given an interpretation ๐ผ of ๐”„, there should be an interpretation of ๐”…
that is like ๐ผ. This interpretation is ๐ผ๐œ‘ .
386
Chapter 7 MODELS
LEMMA 7.3.22
Let ๐”„ and ๐”… be ๐–ฒ-structures with isomorphism ๐œ‘ : ๐”„ → ๐”…. Assume that ๐ผ is
an ๐–ฒ-interpretation of ๐”„. Then, for every ๐–ฒ-formula ๐‘, ๐”„ โŠจ ๐‘ [๐ผ] if and only if
๐”… โŠจ ๐‘ [๐ผ๐œ‘ ].
PROOF
We proceed by induction on formulas, relying on Theorem 7.1.7.
โˆ™ Suppose that ๐‘ก0 and ๐‘ก1 are ๐–ฒ-terms. Since ๐œ‘ is one-to-one,
๐”„ โŠจ ๐‘ก0 = ๐‘ก1 [๐ผ] ⇔ ๐ผ(๐‘ก0 ) = ๐ผ(๐‘ก1 )
⇔ ๐œ‘(๐ผ(๐‘ก0 )) = ๐œ‘(๐ผ(๐‘ก1 ))
⇔ ๐ผ๐œ‘ (๐‘ก0 ) = ๐ผ๐œ‘ (๐‘ก1 )
⇔ ๐”… โŠจ ๐‘ก0 = ๐‘ก1 [๐ผ๐œ‘ ].
โˆ™ Let ๐‘… be an ๐‘›-ary relation symbol and ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 be ๐–ฒ-terms. Since ๐œ‘
is an isomorphism,
๐”„ โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) [๐ผ]
⇔ (๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 )) ∈ ๐ผ(๐‘…)
⇔ (๐œ‘(๐ผ(๐‘ก0 )), ๐œ‘(๐ผ(๐‘ก1 )), … , ๐œ‘(๐ผ(๐‘ก๐‘›−1 ))) ∈ ๐œ‘(๐ผ(๐‘…))
⇔ (๐ผ๐œ‘ (๐‘ก0 ), ๐ผ๐œ‘ (๐‘ก1 ), … , ๐ผ๐œ‘ (๐‘ก๐‘›−1 )) ∈ ๐ผ๐œ‘ (๐‘…)
⇔ ๐”… โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) [๐ผ๐œ‘ ].
โˆ™ Assume that ๐‘ž is an ๐–ฒ-formula. Then,
๐”„ โŠจ ¬๐‘ž [๐ผ] ⇔ ๐”„ ฬธโŠจ ๐‘ž [๐ผ] ⇔ ๐”… ฬธโŠจ ๐‘ž [๐ผ๐œ‘ ] ⇔ ๐”… โŠจ ¬๐‘ž [๐ผ๐œ‘ ],
where the middle equivalence holds by induction.
โˆ™ Let ๐‘ž and ๐‘Ÿ be ๐–ฒ-formulas. Suppose that ๐”„ โŠจ ๐‘ž → ๐‘Ÿ [๐ผ]. This means that
๐”„ โŠจ ๐‘ž [๐ผ] implies ๐”„ โŠจ ๐‘Ÿ [๐ผ]. Assume ๐”… โŠจ ๐‘ž [๐ผ๐œ‘ ]. By induction, we
have that ๐”„ โŠจ ๐‘ž [๐ผ], whence ๐”„ โŠจ ๐‘Ÿ [๐ผ]. Again, by induction, ๐”… โŠจ ๐‘Ÿ [๐ผ๐œ‘ ].
Therefore, ๐”… โŠจ ๐‘ž → ๐‘Ÿ [๐ผ๐œ‘ ]. The converse is proved similarly.
โˆ™ Let ๐‘ž be an ๐–ฒ-formula. Then, by Lemma 7.3.21 and since ๐œ‘ is an isomorphism,
๐”„ โŠจ ∃๐‘ฅ๐‘ž [๐ผ] ⇔ ๐”„ โŠจ ๐‘ž [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด
⇔ ๐”… โŠจ ๐‘ž [(๐ผ๐‘ฅ๐‘Ž )๐œ‘ ] for some ๐‘Ž ∈ ๐ด
⇔ ๐”… โŠจ ๐‘ž [(๐ผ๐œ‘ )๐œ‘(๐‘Ž)
๐‘ฅ ] for some ๐‘Ž ∈ ๐ด
⇔ ๐”… โŠจ ๐‘ž [(๐ผ๐œ‘ )๐‘๐‘ฅ ] for some ๐‘ ∈ ๐ต
⇔ ๐”… โŠจ ∃๐‘ฅ๐‘ž [๐ผ๐œ‘ ].
Section 7.3 HOMOMORPHISMS
387
Let ๐”Š1 = (2โ„ค, 0, +) and ๐”Š2 = (3โ„ค, 0, +). These isomorphic groups are infinite
and cyclic, each with exactly two generators. The group ๐”Š1 is generated by 2 and
−2, and ๐”Š2 is generated by 3 and −3. When Lemma 7.3.22 is restricted to sentences,
we conclude that isomorphic structures model the same sentences. Hence, the ๐–ฆ๐–ฑsentences satisfied by ๐”Š1 are exactly those ๐–ฆ๐–ฑ-sentences satisfied by ๐”Š2 . For example,
๐”Š1 โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘ฆ โ—ฆ ๐‘ฅ),
and
๐”Š2 โŠจ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ โ—ฆ ๐‘ฆ = ๐‘ฆ โ—ฆ ๐‘ฅ).
Also,
๐”Š1 โŠจ ∃๐‘ฅ∀๐‘ฆ∃๐‘ง(๐‘ฆ = ๐‘ฅ โ—ฆ ๐‘ง),
and
๐”Š2 โŠจ ∃๐‘ฅ∀๐‘ฆ∃๐‘ง(๐‘ฆ = ๐‘ฅ โ—ฆ ๐‘ง).
Therefore, we make the next definition and follow it immediately with the theorem that
follows from Lemma 7.3.22 and Theorem 7.1.31.
DEFINITION 7.3.23
The ๐–ฒ-structures ๐”„ and ๐”… are elementary equivalent (denoted by ๐”„ ≡ ๐”…) if for
all ๐–ฒ-sentences ๐‘, ๐”„ โŠจ ๐‘ if and only if ๐”… โŠจ ๐‘.
THEOREM 7.3.24
For all ๐–ฒ-structures ๐”„ and ๐”…, if ๐”„ ≅ ๐”…, then ๐”„ ≡ ๐”….
Theorem 7.3.24 implies that if structures are isomorphic, there is no first-order sentence that can be used to distinguish between the two. That is, there is no sentence ๐‘
such that ๐”„ โŠจ ๐‘ but ๐”… ฬธโŠจ ๐‘ when ๐”„ ≅ ๐”…. Conversely, if there is a sentence that is
satisfied by one structure but not the other, the structures cannot be isomorphic.
EXAMPLE 7.3.25
Let ๐”„ and ๐”… be ๐–ฒ๐–ณ-structures such that ๐”„ = (๐œ”+ , ∈) and ๐”… = (๐œ”, ∈). Observe
that
๐”„ โŠจ ∃๐‘ฅ∀๐‘ฆ(๐‘ฆ ∈ ๐‘ฅ ∨ ๐‘ฆ = ๐‘ฅ)
but
๐”… ฬธโŠจ ∃๐‘ฅ∀๐‘ฆ(๐‘ฆ ∈ ๐‘ฅ ∨ ๐‘ฆ = ๐‘ฅ).
Therefore, by Theorem 7.3.24, we have that ๐”„ is not isomorphic to ๐”….
EXAMPLE 7.3.26
Let ๐”„ = (๐ด, ๐”ž) and ๐”… = (๐ต, ๐”Ÿ) be {๐‘…}-structures, where ๐‘… is a binary relation
symbol. Let ๐ด = {๐‘Ž0 , ๐‘Ž1 } with ๐‘Ž0 ≠ ๐‘Ž1 . Assume that the structures are not
isomorphic yet ๐”„ ≡ ๐”…. Define
๐‘ := ∃๐‘ฅ(๐‘ฅ = ๐‘ฅ),
388
Chapter 7 MODELS
๐‘ž := ∀๐‘ฅ∃๐‘ฆ(๐‘ฅ ≠ ๐‘ฆ),
and
๐‘Ÿ := ∀๐‘ฅ∀๐‘ฆ∀๐‘ง(๐‘ฅ ≠ ๐‘ฆ ∧ ๐‘ฅ ≠ ๐‘ง → ๐‘ฆ = ๐‘ง).
Then, ๐”„ โŠจ ๐‘ because ๐ด is nonempty, ๐”„ โŠจ ๐‘ž because ๐ด has at least two elements,
and ๐”„ โŠจ ๐‘Ÿ because ๐ด has at most two elements. Hence,
๐”„ โŠจ ๐‘ ∧ ๐‘ž ∧ ๐‘Ÿ.
Since ๐”„ ≡ ๐”…, we have that ๐”… โŠจ ๐‘ ∧ ๐‘ž ∧ ๐‘Ÿ, so |๐ต| = 2. Write ๐ต = {๐‘0 , ๐‘1 } with
๐‘0 ≠ ๐‘1 . This implies that there are only two bijections ๐ด → ๐ต. Call them ๐œ‘0
and ๐œ‘1 , and write
๐œ‘0 = {(๐‘Ž0 , ๐‘0 ), (๐‘Ž1 , ๐‘1 )}
and
๐œ‘1 = {(๐‘Ž0 , ๐‘1 ), (๐‘Ž1 , ๐‘0 )}.
By assumption, neither of these functions are {๐‘…}-homomorphisms, so they fail
to preserve ๐‘…. Suppose that (๐‘Ž0 , ๐‘Ž1 ) ∈ ๐”ž(๐‘…) is the particular element that causes
the second condition of Definition 7.3.1 to fail. Therefore,
๐”„ โŠจ ∃๐‘ฅ∃๐‘ฆ[๐‘ฅ ≠ ๐‘ฆ ∧ ๐‘…(๐‘ฅ, ๐‘ฆ)].
Since ๐œ‘0 does not preserve ๐‘…, (๐‘0 , ๐‘1 ) ∉ ๐”Ÿ(๐‘…), and since ๐œ‘1 does not preserve
๐‘…, (๐‘1 , ๐‘0 ) ∉ ๐”Ÿ(๐‘…). Hence,
๐”… ฬธโŠจ ∃๐‘ฅ∃๐‘ฆ[๐‘ฅ ≠ ๐‘ฆ ∧ ๐‘…(๐‘ฅ, ๐‘ฆ)],
which is impossible because ๐”„ ≡ ๐”…, so ๐œ‘0 or ๐œ‘1 must be an {๐‘…}-isomorphism.
The argument generalizes for all structures with a domain of cardinality 2 (Exercise 17) and then to all structures with a finite domain (Exercise 18). Therefore,
the converse of Theorem 7.3.24 holds, provided that the domains of the structures
are finite.
Elementary Substructures
Let ๐”„ = (๐ด, ๐”ž) and ๐”… = (๐ต, ๐”Ÿ) be ๐–ฒ-structures such that ๐”„ ⊆ ๐”…. Let ๐ผ be an interpretation of ๐”„. By Definition 7.1.4, we have the following:
โˆ™ For all variable symbols ๐‘ฅ, ๐ผ(๐‘ฅ) ∈ ๐ด ⊆ ๐ต.
โˆ™ For all constant symbols ๐‘, ๐ผ(๐‘) ∈ ๐ด ⊆ ๐ต.
โˆ™ For all ๐‘›-ary function symbols ๐‘“ and ๐–ฒ-terms ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 , because ๐ผ(๐‘ก๐‘˜ ) ∈ ๐ด
for ๐‘˜ = 0, 1, … , ๐‘› − 1 and ๐”ž(๐‘“ ) = ๐”Ÿ(๐‘“ ) ๐ด๐‘› ,
๐ผ(๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 )) = ๐”ž(๐‘“ )(๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 ))
= ๐”Ÿ(๐‘“ )(๐ผ(๐‘ก0 ), ๐ผ(๐‘ก1 ), … , ๐ผ(๐‘ก๐‘›−1 )).
This implies that ๐ผ is also an interpretation of ๐”…. Moreover, when ๐‘ is a quantifier-free
๐–ฒ-formula, the interpretation of ๐‘ in ๐”… requires no element of ๐ต โงต ๐ด. This implies the
next theorem. Its proof is left to Exercise 21.
Section 7.3 HOMOMORPHISMS
389
THEOREM 7.3.27
Let ๐”„ and ๐”… be ๐–ฒ-structures such that ๐”„ ⊆ ๐”…. For all ๐–ฒ-interpretations ๐ผ of ๐”„
and all quantifier-free ๐–ฒ-formulas ๐‘, ๐”„ โŠจ ๐‘ [๐ผ] if and only if ๐”… โŠจ ๐‘ [๐ผ].
Next, consider ๐”… โŠจ ∃๐‘ฅ๐‘ [๐ผ]. This means that ๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ] for some ๐‘ ∈ ๐ต. If ๐‘ is
also in ๐ด, then ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ], but this conclusion is not guaranteed if ๐‘ ∈ ๐ต โงต ๐ด. For
example, let ๐”„ = (โ„ค, <) and ๐”… = (โ„š, <) be {๐‘…}-structures, where ๐‘… is a binary relation
symbol. Since โ„ค ⊆ โ„š, we have that ๐”„ is a substructure of ๐”…. However, because of
differences between โ„š and โ„ค,
๐”… โŠจ ∀๐‘ฅ1 ∀๐‘ฅ2 ∃๐‘ฆ(๐‘ฅ1 < ๐‘ฆ < ๐‘ฅ2 )
but
๐”„ ฬธโŠจ ∀๐‘ฅ1 ∀๐‘ฅ2 ∃๐‘ฆ(๐‘ฅ1 < ๐‘ฆ < ๐‘ฅ2 ).
Hence, in order for ๐”„ โŠจ ๐‘ [๐ผ] if and only if ๐”… โŠจ ๐‘ [๐ผ] for even quantified formulas ๐‘,
a stronger condition is required.
DEFINITION 7.3.28
๐”„ is an elementary substructure of ๐”… (written as ๐”„ โชฏ ๐”…) if ๐”„ ⊆ ๐”… and for all
interpretations ๐ผ of ๐”„ and ๐–ฒ-formulas ๐‘,
๐”„ โŠจ ๐‘ [๐ผ] if and only if ๐”… โŠจ ๐‘ [๐ผ].
If ๐”„ โชฏ ๐”…, then ๐”… is an elementary extension of ๐”„.
Suppose that ๐”„ and ๐”… are ๐–ฒ-structures such that ๐”„ is an elementary substructure of
๐”…. Let ๐‘ be an ๐–ฒ-sentence. Since ๐‘ is also an ๐–ฒ-formula, ๐”„ โŠจ ๐‘ if and only if ๐”… โŠจ ๐‘
by Definition 7.3.28. Therefore, ๐”„ and ๐”… are elementary equivalent, proving the next
result.
THEOREM 7.3.29
Let ๐”„ and ๐”… be ๐–ฒ-structures. If ๐”„ โชฏ ๐”…, then ๐”„ ≡ ๐”….
However, the converse of Theorem 7.3.29 does not hold. It is possible for ๐”„ ⊆ ๐”…
and ๐”„ ≡ ๐”… yet ๐”„ not be an elementary substructure of ๐”…. If this is to be the case,
there must be a formula ๐‘ with at least one free variable that is satisfied by one of the
structures but not the other.
EXAMPLE 7.3.30
Let ๐”„ = ([0, 1] , ๐”ž) and ๐”… = ([0, 2] , ๐”Ÿ) be the {๐‘…}-structures of Example 7.2.2,
where ๐”ž(๐‘…) is standard < on [0, 1] and ๐”Ÿ(๐‘…) is standard < on [0, 2]. Recall that
๐”„ is a substructure of ๐”…. Since ๐‘“ defined by ๐‘“ (๐‘ฅ) = 2๐‘ฅ is an {๐‘…}-isomorphism
[0, 1] → [0, 2], we conclude that ๐”„ ≡ ๐”… (Theorem 7.3.24). However, define the
formula
๐‘(๐‘ฅ) := ∃๐‘ฆ๐‘…(๐‘ฅ, ๐‘ฆ)
390
Chapter 7 MODELS
and let ๐ผ be an interpretation such that ๐ผ(๐‘ฅ) = 1. Then,
๐”„ ฬธโŠจ ๐‘(๐‘ฅ) [๐ผ]
and
๐”… โŠจ ๐‘(๐‘ฅ) [๐ผ].
Therefore, ๐”„ is not an elementary substructure of ๐”….
Because of Theorem 7.3.27, proving that one structure is an elementary substructure
of another reduces to a check of one particular condition. This condition is found in
the next theorem, which is due to Tarski and Vaught (1957).
THEOREM 7.3.31 [Tarski–Vaught]
Let ๐”„ and ๐”… be ๐–ฒ-structures such that ๐”„ ⊆ ๐”…. Let ๐ด be the domain of ๐”„. The
following are equivalent.
โˆ™ ๐”„ is an elementary substructure of ๐”….
โˆ™ For every ๐–ฒ-formula ๐‘ and ๐–ฒ-interpretation ๐ผ of ๐”„,
if ๐”… โŠจ ∃๐‘ฅ๐‘ [๐ผ] then ๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด.
(7.19)
PROOF
Suppose that ๐”„ โชฏ ๐”…. Let ๐‘ be an ๐–ฒ-formula and ๐ผ an ๐–ฒ-interpretation of ๐”„. Let
๐”… โŠจ ∃๐‘ฅ๐‘ [๐ผ]. By hypothesis, we have that ๐”„ โŠจ ∃๐‘ฅ๐‘ [๐ผ]. Thus, ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for
some ๐‘Ž ∈ ๐ด, so again by hypothesis, ๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด.
Conversely, let ๐‘ be an ๐–ฒ-formula and ๐ผ an ๐–ฒ-interpretation of ๐”„. We proceed
by induction on formulas to show that ๐”„ โชฏ ๐”….
โˆ™ If ๐‘ is ๐‘ก0 = ๐‘ก1 or ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) for ๐–ฒ-terms ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 , then ๐‘ has no
quantifiers, and the result follows by Theorem 7.3.27.
โˆ™ Let ๐‘ be ¬๐‘ž. Then,
๐”„ โŠจ ๐‘ [๐ผ] ⇔ ๐”„ โŠจ ¬๐‘ž [๐ผ]
⇔ ๐”„ ฬธโŠจ ๐‘ž [๐ผ]
⇔ ๐”… ฬธโŠจ ๐‘ž [๐ผ]
⇔ ๐”… โŠจ ¬๐‘ž [๐ผ]
⇔ ๐”… โŠจ ๐‘ [๐ผ].
โˆ™ Suppose ๐‘ is ๐‘ž → ๐‘Ÿ. Assume that ๐”„ โŠจ ๐‘ž → ๐‘Ÿ [๐ผ] and let ๐”… โŠจ ๐‘ž [๐ผ]. Then,
๐”„ โŠจ ๐‘ž [๐ผ] by induction, so ๐”„ โŠจ ๐‘Ÿ [๐ผ]. Again, by induction, ๐”… โŠจ ๐‘Ÿ [๐ผ].
Therefore, ๐”… โŠจ ๐‘Ÿ → ๐‘ž [๐ผ]. The converse is proved similarly.
โˆ™ Now let ๐‘ be ∃๐‘ฅ๐‘ž. First, because ๐”„ ⊆ ๐”…,
๐”„ โŠจ ∃๐‘ฅ๐‘ž [๐ผ] ⇒ ๐”„ โŠจ ๐‘ž [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด
⇒ ๐”… โŠจ ๐‘ž [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ต
⇒ ๐”… โŠจ ∃๐‘ฅ๐‘ž [๐ผ].
Section 7.3 HOMOMORPHISMS
391
Second, by (7.19),
๐”… โŠจ ∃๐‘ฅ๐‘ž [๐ผ] ⇒ ๐”… โŠจ ๐‘ž [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด
⇒ ๐”„ โŠจ ๐‘ž [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด
⇒ ๐”„ โŠจ ∃๐‘ฅ๐‘ž [๐ผ].
EXAMPLE 7.3.32
Prove that ๐”„ = (โ„š, ≤) is an elementary substructure of ๐”… = (โ„, ≤). To do this,
follow the test given by Theorem 7.3.31. Let ๐‘ be an ๐–ฒ-formula and ๐ผ be an
๐–ฒ-interpretation of ๐”„. Assume that ๐”… โŠจ ∃๐‘ฅ๐‘ [๐ผ]. This implies that
๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ] for some ๐‘ ∈ โ„.
Take ๐‘Ž0 , ๐‘Ž1 , ๐‘Ž ∈ โ„š such that ๐‘, ๐‘Ž ∈ (๐‘Ž0 , ๐‘Ž1 ) and define the functions
๐‘“0
๐‘“1
๐‘“2
๐‘“3
: (−∞, ๐‘Ž0 ] → (−∞, ๐‘Ž0 ],
: (๐‘Ž0 , ๐‘Ž) → (๐‘Ž0 , ๐‘),
: [๐‘, ๐‘Ž1 ) → [๐‘Ž, ๐‘Ž1 ),
: [๐‘Ž1 , ∞) → [๐‘Ž1 , ∞),
so that ๐‘“0 is the identity on (−∞, ๐‘Ž0 ], ๐‘“3 is the identity on [๐‘Ž1 , ∞), and ๐‘“1 and
๐‘“2 are order isomorphisms (Exercise 22). Let ๐‘“ = ๐‘“0 ∪ ๐‘“1 ∪ ๐‘“2 ∪ ๐‘“3 . Then,
๐‘“ is an order isomorphism โ„ → โ„ such that ๐‘“ (๐‘) = ๐‘Ž, ๐‘“ (๐‘Ž0 ) = ๐‘“ (๐‘Ž0 ), and
๐‘“ (๐‘Ž1 ) = ๐‘“ (๐‘Ž1 ). Hence, ๐‘“ is an automorphism (Exercise 23). Therefore, by
Lemmas 7.3.21 and 7.3.22,
๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ] ⇔ ๐”… โŠจ ๐‘ [(๐ผ๐‘ฅ๐‘ )๐‘“ ] ⇔ ๐”… โŠจ ๐‘ [(๐ผ๐‘“ )๐‘“๐‘ฅ (๐‘) ],
which implies that
๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ] for some ๐‘ ∈ โ„š.
Thus, ๐”„ โชฏ ๐”…, which also implies that ๐”„ ≡ ๐”… (Theorem 7.3.29).
Notice that this example implies that there is no first-order sentence that can distinguish between (โ„š, <) and (โ„, <) even though they are not isomorphic. Also, this
example shows that the converse of Theorem 7.3.24 is false if the domains of the structures are infinite.
Certainly, for all ๐–ฒ-structures ๐”„, ๐”…, and โ„ญ, we have the following by Exercise 19:
โˆ™ ๐”„ ≡ ๐”„.
โˆ™ If ๐”„ ≡ ๐”…, then ๐”… ≡ ๐”„.
โˆ™ If ๐”„ ≡ ๐”… and ๐”… ≡ โ„ญ, then ๐”„ ≡ โ„ญ.
Similar to this and Theorem 7.2.4, we have the following result for elementary substructures.
392
Chapter 7 MODELS
THEOREM 7.3.33
Let ๐”„, ๐”…, and โ„ญ be ๐–ฒ-structures.
โˆ™ ๐”„ โชฏ ๐”„.
โˆ™ If ๐”„ โชฏ ๐”… and ๐”… โชฏ โ„ญ, then ๐”„ โชฏ โ„ญ.
โˆ™ If ๐”„ โชฏ โ„ญ, ๐”… โชฏ โ„ญ, and ๐”„ ⊆ ๐”…, then ๐”„ โชฏ ๐”….
PROOF
We prove the second part and leave the others to Exercise 25. Let ๐”„ โชฏ ๐”… and
๐”… โชฏ โ„ญ. This first means that ๐”„ ⊆ ๐”… and ๐”… ⊆ โ„ญ. Thus, by Theorem 7.2.4,
๐”„ ⊆ โ„ญ. Let ๐ผ be an interpretation of ๐”„. The proof is completed for quantifierfree formulas using Theorem 7.3.27 (Exercise 24), so let ๐‘ be an ๐–ฒ-formula with
a quantifier.
โˆ™ Suppose ๐”„ โŠจ ∃๐‘ฅ๐‘ [๐ผ]. This implies that ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] for some ๐‘Ž ∈ ๐ด.
Then, ๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ] from which it follows that โ„ญ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ]. Since ๐ด ⊆ ๐ถ, we
conclude that โ„ญ โŠจ ∃๐‘ฅ๐‘ [๐ผ].
โˆ™ Conversely, let โ„ญ โŠจ ∃๐‘ฅ๐‘ [๐ผ]. By Theorem 7.3.31, there exists ๐‘ ∈ ๐ต such
that โ„ญ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ]. By hypothesis, ๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘ ], so ๐”… โŠจ ∃๐‘ฅ๐‘ [๐ผ]. Again,
by Theorem 7.3.31, there exists ๐‘Ž ∈ ๐ด such that ๐”… โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ]. From this
follows that ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐‘Ž ], whence ๐”„ โŠจ ∃๐‘ฅ๐‘ [๐ผ].
Exercises
1. Let ๐”„, ๐”…, and โ„ญ be ๐–ฒ-structures such that ๐”„ ⊆ ๐”…. Let ๐œ‘ : ๐”… → โ„ญ be a homomorphism. Prove that ๐œ‘[dom(๐”„)] is the domain of a substructure of โ„ญ.
2. Prove that the following are group homomorphisms. Assume in each instance that
โ—ฆ is interpreted as the standard addition on the set.
(a) ๐œ‘ : โ„ค × โ„ค → โ„ค where ๐œ‘(๐‘Ž, ๐‘) = ๐‘Ž
(b) ๐œ‘ : โ„ค12 → โ„ค6 where ๐œ‘([๐‘Ž]12 ) = [๐‘Ž]6
(c) ๐œ“ : โ„ค × โ„ค → M2, 2 (โ„) where
[
]
๐‘Ž 0
๐œ“(๐‘Ž, ๐‘) =
0 ๐‘
3. Prove that the zero map is a group homomorphism.
4. Let โ„จ = (โ„ค, +) and โ„จ๐‘› = (โ„ค๐‘› , +) with ๐‘› ∈ โ„ค+ . Prove that ๐œ‘ : โ„ค → โ„ค๐‘› defined by
๐œ‘(๐‘ฅ) = [๐‘ฅ]๐‘› is a ring homomorphism โ„จ → โ„จ๐‘› .
5. Let โ„ญ = (โ„‚, 0 + 0๐‘–, +, ⋅). Define ๐œ“ : โ„‚ → โ„‚ by ๐œ“(๐‘Ž + ๐‘๐‘–) = ๐‘Ž − ๐‘๐‘– for all ๐‘Ž, ๐‘ ∈ โ„.
Prove the following.
(a) ๐œ“ is a ring homomorphism โ„ญ → โ„ญ.
(b) ๐œ“ is one-to-one and onto.
Section 7.3 HOMOMORPHISMS
393
6. Prove the first part of Theorem 7.3.8.
7. Let ℜ = (๐‘…, 0, +, ⋅) be a ring and ℑ be an ideal of ℜ with domain ๐ผ. Let the
function ๐œ‘ : ๐‘… → ๐‘…โˆ•๐ผ be defined as ๐œ‘(๐‘Ž) = ๐‘Ž + ๐ผ for all ๐‘Ž ∈ ๐‘….
(a) Prove that ๐œ‘ is a homomorphism.
(b) Show that ker(๐œ‘) = ๐ผ.
(c) Prove that ๐œ‘ is onto.
8. Prove Theorem 7.3.15.
9. Prove that the function ๐œ‘ of Example 7.3.16 is a bijection.
10. Let ๐”„ and ๐”… be ๐–ฒ-structures. Write ๐”„ - ๐”… if there exists a one-to-one homomorphism ๐œ‘ : ๐”„ → ๐”…. Prove the following.
(a) ๐”„ - ๐”„
(b) If ๐”„ - ๐”… and ๐”… - โ„ญ, then ๐”„ - โ„ญ.
(c) It is possible for ๐”„ - ๐”… and ๐”… - โ„ญ yet ๐”„ and ๐”… are different ๐–ฒ-structures.
11. Let ๐”„ and ๐”„′ be {๐‘…}-structures with equal domains that are finite, where ๐‘… is a
binary relation symbol. Assume that
๐”„ โŠจ X has a least element with respect to R
and
๐”„′ โŠจ X has a least element with respect to R
for every ๐‘‹ ⊆ dom(๐”„). Prove that ๐”„ ≅ ๐”„′ .
12. Assume that ๐–ฒ is finite and that the domain of the ๐–ฒ-structure ๐”„ is finite. Find a
๐–ฒ-sentence ๐‘ such that ๐”… โŠจ ๐‘ if and only if ๐”… ≅ ๐”„, for all ๐–ฒ-structures ๐”… .
13. Suppose that ๐œ‘ : ℜ → ℜ′ is a ring isomorphism as in Example 7.3.18. Prove.
(a) If ℜ is an integral domain, ℜ′ is an integral domain.
(b) If ℜ is a division ring, ℜ′ is a division ring.
14. Let ๐œ‘ : ℜ → ℜ′ be a ring isomorphism. Prove that 1 is unity from ℜ if and only
if ๐œ‘(1) is unity from ℜ′ .
15. Prove or show false the given elementary equivalences.
(a) (โ„, 0, +) ≡ (โ„‚, 0, +).
(b) (โ„ค, 0, +) ≡ (โ„š, 0, +).
(c) (โ„ค, ≤) ≡ (โ„š, ≤).
(d) (๐œ”, ≤) ≡ (โ„ค+ , ≤).
(e) (๐œ”, ≤) ≡ (โ„ค− , ≤).
16. Let ๐ด and ๐ต be sets and assume that ๐–ฒ = ∅. Prove that (๐ด) ≡ (๐ต).
17. Generalize the argument of Example 7.3.26 to prove that ๐”„ ≡ ๐”… implies that
๐”„ ≅ ๐”… if the cardinality of the domain of ๐”„ equals 2.
18. Generalize the argument of Example 7.3.26 to prove that ๐”„ ≡ ๐”… implies that
๐”„ ≅ ๐”… if the domain of ๐”„ is finite.
394
Chapter 7 MODELS
19. Let ๐”„, ๐”…, and โ„ญ be ๐–ฒ-structures. Prove the following.
(a) ๐”„ ≡ ๐”„.
(b) If ๐”„ ≡ ๐”…, then ๐”… ≡ ๐”„.
(c) If ๐”„ ≡ ๐”… and ๐”… ≡ โ„ญ, then ๐”„ ≡ โ„ญ.
20. Let ๐”„ and ๐”… be elementary equivalent ๐–ฒ-structures. Prove that there exists an
๐–ฒ-structure โ„ญ such that ๐”„ โชฏ โ„ญ and ๐”… โชฏ โ„ญ.
21. Prove Theorem 7.3.27.
22. Find the order isomorphisms ๐‘“1 and ๐‘“2 in Example 7.3.32.
23. Prove that the function ๐‘“ from Example 7.3.32 is a {≤}-isomorphism.
24. Assume that for all ๐–ฒ-structures ๐”„, ๐”…, and โ„ญ, if ๐”„ โชฏ ๐”… and ๐”… โชฏ โ„ญ, then ๐”„ โŠจ ๐‘ if
and only if โ„ญ โŠจ ๐‘ for all quantifier-free ๐‘.
25. Prove the first and third parts of Theorem 7.3.33.
26. Let โ„ฑ = {๐”„๐›พ : ๐›พ ∈ ๐œ…} be a chain of ๐–ฒ-structures for some cardinal ๐œ… (Exercise 7.2.3). Assume that ๐”„๐›ผ โชฏ ๐”„๐›ฝ when ๐›ผ ∈ ๐›ฝ ∈ ๐œ…, making โ„ฑ an elementary chain.
โ‹ƒ
Prove that ๐”„๐›ผ โชฏ ๐›พ∈๐œ… ๐”„๐›พ for all ๐›ผ ∈ ๐œ….
27. Find a set of subject symbols ๐–ฒ and a chain of ๐–ฒ-structures {๐”„๐‘› : ๐‘› ∈ ๐œ”} such that
โ‹ƒ
๐”„๐‘– ≡ ๐”„๐‘— for all ๐‘–, ๐‘— ∈ ๐œ” but ๐”„0 โ‰ข ๐‘›∈๐œ” ๐”„๐‘› .
7.4
THE THREE PROPERTIES REVISITED
This section is an extension of Section 1.5 to first-order logic. First, we define what it
means to be consistent with the ultimate goal to show that any consistent system has a
model. Next, we show that first-order logic is sound in that every sentence that can be
proved is true, provided we have the correct understanding of what it means to prove
and what it means to be true. Lastly, we show that first-order logic is complete in that
every true sentence can be proved.
Consistency
Consider the following propositions representing Euclid’s axioms for his geometry (Euclid 1925, I Postulates).
โˆ™ Eu1. A line can be drawn through two distinct points.
โˆ™ Eu2. A line segment can be drawn between two distinct points.
โˆ™ Eu3. A circle can be drawn given any center and radius.
โˆ™ Eu4. All right angles are congruent to each another.
โˆ™ Eu5. If a line falling on two straight lines make the interior angles on the same
side less than two right angles, the two lines intersect on the side on which the
angles are less than two right angles.
Section 7.4 THE THREE PROPERTIES REVISITED
395
There are three sets of proposition here. First, is the list itself,
โ„ฐ = {Eu1, Eu2, Eu3, Eu4, Eu5}.
Second, is the set of consequences (Definition 7.1.20) of โ„ฐ ,
โ„ฐ1 = {๐‘ : โ„ฐ โŠจ ๐‘}.
Third, is the set of propositions provable (Definition 1.2.13) from โ„ฐ ,
โ„ฐ2 = {๐‘ : โ„ฐ โŠข ๐‘}.
For Euclidean geometry, โ„ฐ1 = โ„ฐ2 . This means that both โ„ฐ1 and โ„ฐ2 can be said to
describe all of the propositions that are in the geometry. It is the job of the geometer to
discover what those propositions are.
We want to similarly analyze first-order logic. Given the rules of logic and a set
of theory symbols, we want the set of consequences to be equal to the set of provable
sentences. One way to have this happen is for a contradiction to follow from the axioms
(1.2.8). Then, by Theorem 1.5.2, every propositional form will have a proof and, in
turn, will also be a consequence, but we do not want such a system. Therefore, for all of
this to work effectively, it is a requirement that first-order logic is without contradiction.
To deal with this concept, we start with a definition.
DEFINITION 7.4.1
An ๐—ฆ-theory is a set of ๐–ฒ-sentences.
The set of all sentences provable in first-order logic given a set of theory symbols ๐–ฒ is
an example of an ๐–ฒ-theory. The theories that we study should have a familiar property
(compare Definition 1.5.1).
DEFINITION 7.4.2
An ๐–ฒ-theory ๐’ฏ is consistent [denoted by ๐–ข๐—ˆ๐—‡(๐น )] if ๐’ฏ โŠฌ ๐‘ž ∧ ¬๐‘ž for every
๐–ฒ-sentence ๐‘ž. Otherwise, ๐’ฏ is inconsistent.
We next generalize Theorem 1.5.2 to first-order logic. Its proof is left to Exercise 2.
THEOREM 7.4.3
Let ๐’ฏ be an ๐–ฒ-theory. The following are equivalent.
โˆ™ ๐’ฏ is consistent.
โˆ™ Every finite subset of ๐’ฏ is consistent.
โˆ™ There is an ๐–ฒ-sentence that is not provable from ๐’ฏ .
As a consequence of Theorem 7.4.3, a theory ๐’ฏ is inconsistent if either it has a finite
subset that is inconsistent or it is able to prove all sentences. This identifies the major
396
Chapter 7 MODELS
weakness of an inconsistent theory. Not only can a contradiction be proved, but any
sentence can also be proved. Such a system is certainly worthless.
If a theory is consistent, it can be shown to be a subset of a theory that is the greatest
possible consistent theory. We generalize Definition 1.5.3 to name this theory.
DEFINITION 7.4.4
The ๐–ฒ-theory ๐’ฏ is maximally consistent if ๐’ฏ is consistent and ๐–ข๐—ˆ๐—‡(๐’ฏ ∪ {๐‘})
implies ๐‘ ∈ ๐’ฏ for all ๐–ฒ-sentences ๐‘.
The next lemma gives some basic properties of maximally consistent theories.
LEMMA 7.4.5
Let ๐’ฏ be a maximally consistent ๐–ฒ-theory. Let ๐‘ and ๐‘ž be ๐–ฒ-sentences.
โˆ™ If ๐’ฏ โŠข ๐‘, then ๐‘ ∈ ๐’ฏ .
โˆ™ If โŠข ๐‘, then ๐‘ ∈ ๐’ฏ .
โˆ™ ๐‘ ∈ ๐’ฏ or ¬๐‘ ∈ ๐’ฏ , but not both.
โˆ™ If ๐‘ → ๐‘ž, ๐‘ ∈ ๐’ฏ , then ๐‘ž ∈ ๐’ฏ .
PROOF
โˆ™ Let ๐’ฏ โŠข ๐‘. Suppose that ๐’ฏ ∪ {๐‘} โŠข ๐‘ž ∧ ¬๐‘ž for some ๐–ฒ-sentence ๐‘ž. This
means that there exists ๐–ฒ-sentences ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 such that
๐‘, ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘ž ∧ ¬๐‘ž
is a proof of ๐‘ž ∧ ¬๐‘ž from ๐’ฏ ∪ {๐‘}. Since ๐’ฏ proves ๐‘, there are ๐–ฒ-sentences
๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘š−1 such that
๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘š−1 , ๐‘
is a proof of ๐‘ from ๐’ฏ . Therefore,
๐‘ž0 , ๐‘ž1 , … , ๐‘ž๐‘š−1 , ๐‘, ๐‘0 , … , ๐‘๐‘›−1 , ๐‘ž ∧ ¬๐‘ž
is a proof of ๐‘ž ∧ ¬๐‘ž from ๐’ฏ , a contradiction. This implies that ๐’ฏ ∪ {๐‘} is
consistent, so since ๐’ฏ is maximally consistent, ๐‘ ∈ ๐’ฏ .
โˆ™ If โŠข ๐‘, then ๐’ฏ โŠข ๐‘, so ๐‘ ∈ ๐’ฏ by the first property.
โˆ™ Since ๐’ฏ is consistent, ๐–ข๐—ˆ๐—‡(๐’ฏ ∪ {๐‘}) or ๐–ข๐—ˆ๐—‡(๐’ฏ ∪ {¬๐‘}). Because ๐’ฏ is
maximally consistent, ๐‘ ∈ ๐’ฏ or ¬๐‘ ∈ ๐’ฏ , but not both because ๐–ข๐—ˆ๐—‡(๐’ฏ ).
โˆ™ Let ๐‘ → ๐‘ž and ๐‘ be members of ๐’ฏ . By MP, we have that ๐’ฏ โŠข ๐‘ž, so again
by the first part, ๐‘ž ∈ ๐’ฏ .
Given a consistent theory, the construction of the maximally consistent theory that
contains it requires the use of a chain.
Section 7.4 THE THREE PROPERTIES REVISITED
397
LEMMA 7.4.6
If (๐’ž , ⊆) is a chain of consistent ๐–ฒ-theories,
โ‹ƒ
๐’ž is consistent.
PROOF
Let ๐’ž be a chain of consistent sets of ๐–ฒ-sentences with respect to ⊆. Suppose
โ‹ƒ
โ‹ƒ
๐’ž โŠข ๐‘ž ∧ ¬๐‘ž for some sentence ๐‘ž. Hence, there exists ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ∈ ๐’ž
such that
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 โŠข ๐‘ž ∧ ¬๐‘ž.
This implies that for each ๐‘– = 0, 1, … , ๐‘› − 1, there exists ๐ถ๐‘– ∈ ๐’ž such that
๐‘๐‘– ∈ ๐ถ๐‘– . Since {๐ถ๐‘– : ๐‘– = 0, 1, … , ๐‘› − 1} is a finite chain, arrange them so that ๐ถ๐‘›
is the greatest element with respect to ⊆. This gives
{๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 } ⊆ ๐ถ๐‘› ,
which implies that ๐ถ๐‘› โŠข ๐‘ž ∧ ¬๐‘ž, contradicting the consistency of ๐ถ๐‘› .
The next result proves the existence of a greatest consistent theory by generalizing
Theorem 1.5.4 to an arbitrary set of sentences.
THEOREM 7.4.7 [Lindenbaum]
Every consistent ๐–ฒ-theory is a subset of a maximally consistent ๐–ฒ-theory.
PROOF
Let ๐’ฏ be a consistent set of ๐–ฒ-sentences. Define
๐’œ = {๐ธ : ๐’ฏ ⊆ ๐ธ and ๐ธ is a consistent set of ๐–ฒ-formulas}.
Note that ๐’œ ≠ ∅ since ๐’ฏ ∈ ๐’œ . Let ๐’ž be a chain in ๐’œ . Then,
โ‹ƒ
โˆ™ ๐’ฏ ⊆ ๐’ž.
โˆ™
โ‹ƒ
๐’ž ∈ ๐’œ because
โ‹ƒ
๐’ž is consistent because each element of the chain ๐’ž is consistent by
Lemma 7.4.6.
We conclude by Zorn’s lemma (5.1.13) that there is a greatest element ๐‘€ ∈ ๐’œ
with respect to ⊆. By definition, ๐’ฏ ⊆ ๐‘€ and ๐–ข๐—ˆ๐—‡(๐‘€). Also, let ๐‘ be a formula
such that ๐–ข๐—ˆ๐—‡(๐‘€ ∪ {๐‘}). Then, ๐‘€ ∪ {๐‘} ∈ ๐’œ , but since ๐‘€ is maximal, we
conclude that ๐‘€ ∪ {๐‘} = ๐‘€. Thus, ๐‘ ∈ ๐‘€, showing that ๐‘€ is maximally
consistent.
Soundness
We previously defined a logic to be sound if every theorem is a tautology and complete
if every tautology is a theorem (Definition 1.5.5). We now give the corresponding
definition for first-order logic (Figure 7.3).
398
Chapter 7 MODELS
Sound
The sentence
is valid.
The sentence
is a theorem.
Complete
Figure 7.3
Sound and complete logics.
DEFINITION 7.4.8
โˆ™ A logic is sound if every theorem is valid.
โˆ™ A logic is complete if every valid sentence is a theorem.
A propositional form ๐‘ is a tautology if ๐—(๐‘) = T for every valuation ๐— (Definition 1.1.14).
This means that any interpretation of ๐‘ will always yield a true proposition. This, in
turn, implies that ๐‘ will hold in every model. That is,
tautologies are valid.
Therefore, it is evident that Definition 7.4.8 is a generalization of Definition 1.5.5 to
first-order logic.
As in Section 1.5, we begin with soundness.
THEOREM 7.4.9 [Soundness]
Let ๐’ฏ be a set of ๐–ฒ-sentences. For any ๐–ฒ-sentence ๐‘, if ๐’ฏ โŠข ๐‘, then ๐’ฏ โŠจ ๐‘.
PROOF
Using strong induction, we prove that for all ๐‘˜ ∈ โ„ค+ ,
if there exists a proof of ๐‘ from ๐’ฏ
consisting of ๐‘˜ sentences, then ๐’ฏ โŠจ ๐‘.
(7.20)
In the case that ๐‘› = 1, we see that โŠข ๐‘. This implies that ๐‘ is an axiom or ๐‘ ∈ ๐’ฏ
(Definition 1.2.13), so we have that ๐’ฏ โŠจ ๐‘.
Now suppose that ๐‘› > 1 and assume that (7.20) holds for all ๐‘˜ ≤ ๐‘›. This
means that there exists sentences ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 provable from ๐’ฏ such that
๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘
is a proof of ๐‘. Let ๐”„ be any model of ๐’ฏ . We have three cases to consider
(Theorem 1.4.2).
โˆ™ ๐‘ ∈ ๐’ฏ , which implies that ๐”„ โŠจ ๐‘.
Section 7.4 THE THREE PROPERTIES REVISITED
399
โˆ™ ๐‘ is derived using MP. This implies there exists ๐–ฒ-sentences ๐‘Ÿ and ๐‘ such
that ๐‘Ÿ, ๐‘Ÿ → ๐‘ ∈ {๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 , ๐‘}. Since a sentence of a proof is proved
from previous sentences of the proof, by induction, ๐”„ โŠจ ๐‘Ÿ → ๐‘ and ๐”„ โŠจ ๐‘Ÿ.
Therefore, ๐”„ โŠจ ๐‘.
โˆ™ ๐‘ is logically equivalent to ๐‘๐‘– for some ๐‘– = 0, 1, … , ๐‘› − 1. This means that
๐‘ ↔ ๐‘๐‘– is a tautology, so ๐”„ โŠจ ๐‘ ↔ ๐‘๐‘– . By Theorem 7.1.7, we conclude that
๐”„ โŠจ ๐‘ if and only if ๐”„ โŠจ ๐‘๐‘– , but by induction, ๐”„ โŠจ ๐‘๐‘– , so ๐”„ โŠจ ๐‘.
If ๐’ฏ = ∅, we can apply Theorem 7.4.9 to conclude the corollary.
COROLLARY 7.4.10
First-order logic is sound.
As with Corollary 1.5.11 for propositional logic, we can also prove the next result.
COROLLARY 7.4.11
First-order logic is consistent.
Completeness
It is now time to prove the completeness of first-order predicate logic. This result is
due to Kurt Gödel (1929) but the proof given here is due to Leon Henkin (1949). As
opposed to the soundness theorem (7.4.9), the proof is rather involved. We begin with
a definition (compare with EI, Theorem 2.3.14).
DEFINITION 7.4.12
Let ๐’ฏ be a set of ๐–ฒ-sentences and ๐ถ a set of constant symbols of ๐–ฒ. Define ๐ถ to
be a witness set for ๐’ฏ if for all ๐–ฒ-formulas ๐‘ = ๐‘(๐‘ฅ), there exists ๐‘ ∈ ๐ถ such that
๐‘
๐’ฏ โŠข ∃๐‘ฅ๐‘ → ๐‘ .
๐‘ฅ
For example, let ๐’ฏ = {๐‘ฅ + 1 = 0, ๐‘ฅ + (1 + 1) = 0, (๐‘ฅ + ๐‘ฆ) + 1 = 1} be a set of
๐–ญ๐–ณ-sentences. Extend ๐–ญ๐–ณ to ๐–ฒ = ๐–ญ๐–ณ ∪ ๐ถ, where ๐ถ = {−1, −2}. Then, ๐ถ is a witness
set for ๐’ฏ if
−1
๐’ฏ โŠข ∃๐‘ฅ(๐‘ฅ + 1 = 0) → (๐‘ฅ + 1 = 0)
๐‘ฅ
and
๐’ฏ โŠข ∃๐‘ฅ(๐‘ฅ + (1 + 1) = 0) → (๐‘ฅ + (1 + 1) = 0)
−2
.
๐‘ฅ
Since (๐‘ฅ + ๐‘ฆ) + 1 = 1 has two free variables, the witness set does not apply to it.
Recall that for a set of theory symbols ๐–ฒ, the notation ๐–ซ(๐–ฒ) refers to the set of all
๐–ฒ-formulas (Definition 2.1.13).
400
Chapter 7 MODELS
LEMMA 7.4.13
Let ๐’ฏ be a consistent set of ๐–ฒ-sentences. Let ๐ถ be a set of constant symbols
not in ๐–ฒ such that |๐ถ| = |๐–ซ(๐–ฒ)|. Then, ๐’ฏ can be extended to a consistent set of
(๐–ฒ ∪ ๐ถ)-sentences that has ๐ถ as a witness set.
PROOF
Let ๐œ… = |๐–ซ(๐–ฒ)| and suppose that ๐ถ = {๐‘๐›ผ : ๐›ผ ∈ ๐œ…} is a set of new constant
symbols. Assume that ๐‘๐›ผ ≠ ๐‘๐›ฝ for all ๐›ผ ∈ ๐›ฝ ∈ ๐œ…. Since |๐ฟ(๐–ฒ ∪ ๐ถ)| = ๐œ…, we
identify the formulas of ๐ฟ(๐–ฒ ∪ ๐ถ) with at most one free variable as
๐‘๐›ผ = ๐‘๐›ผ (๐‘ฅ๐›ผ )
with ๐›ผ ∈ ๐œ…. Now define a chain ๐’ž = {๐’ฏ๐›พ : ๐›พ ∈ ๐œ…} and a set {๐‘‘๐›พ : ๐›พ ∈ ๐œ…} by
transfinite recursion (Corollary 6.1.24) such that for all ๐›ฝ ∈ ๐œ…,
โˆ™ ๐’ฏ๐›ฝ is consistent,
โˆ™ ๐’ฏ๐›ฝ is a (๐–ฒ ∪ ๐ถ)-theory,
โˆ™ |๐’ฏ๐›ฝ+1 โงต ๐’ฏ๐›ฝ | = 1,
โˆ™ ๐‘‘๐›ฝ is not among the symbols of the sentences in ๐’ฏ๐›ฝ .
Let ๐’ฏ0 = ๐’ฏ and suppose that ๐’ฏ๐›ฟ and ๐‘‘๐›ฟ has been defined for all ๐›ฟ ∈ ๐›ผ with the
indicated properties.
โˆ™ Let ๐›ผ = ๐›ฝ + 1. Choose ๐‘‘๐›ฝ from ๐ถ such that ๐‘‘๐›ฝ is not among the symbols
of the formulas in ๐’ฏ๐›ฝ ∪ {๐‘๐›ฝ }. This can be done since there are less than ๐œ…
theory symbols in ๐’ฏ๐›ฝ ∪ {๐‘๐›ฝ }. Now define
๐’ฏ๐›ฝ+1
}
{
๐‘‘๐›ฝ
.
= ๐’ฏ๐›ฝ ∪ ∃๐‘ฅ๐›ฝ ๐‘๐›ฝ → ๐‘๐›ฝ
๐‘ฅ๐›ฝ
(7.21)
In order to obtain a contradiction, suppose that ๐’ฏ๐›ฝ+1 is inconsistent. Since
๐’ฏ๐›ฝ is consistent, it must be the case that
)
(
๐‘‘๐›ฝ
.
๐’ฏ๐›ฝ โŠข ¬ ∃๐‘ฅ๐›ฝ ๐‘๐›ฝ → ๐‘๐›ฝ
๐‘ฅ๐›ฝ
This implies that
๐’ฏ๐›ฝ โŠข ∃๐‘ฅ๐›ฝ ๐‘๐›ฝ ∧ ¬๐‘๐›ฝ
By Com and Simp,
๐’ฏ๐›ฝ โŠข ¬๐‘๐›ฝ
๐‘‘๐›ฝ
๐‘ฅ๐›ฝ
๐‘‘๐›ฝ
๐‘ฅ๐›ฝ
.
.
Since ๐‘‘๐›ฝ is not among the symbols of ๐‘๐›ฝ or ๐’ฏ๐›ฝ , it is arbitrary, so by UG,
๐’ฏ๐›ฝ โŠข ∀๐‘ฅ๐›ฝ ¬๐‘๐›ฝ .
401
Section 7.4 THE THREE PROPERTIES REVISITED
That is,
๐’ฏ๐›ฝ โŠข ¬∃๐‘ฅ๐›ฝ ๐‘๐›ฝ .
Using Simp and Conj, we conclude that
๐’ฏ๐›ฝ โŠข ∃๐‘ฅ๐›ฝ ๐‘๐›ฝ ∧ ¬∃๐‘ฅ๐›ฝ ๐‘๐›ฝ ,
contradicting the consistency of ๐’ฏ๐›ฝ .
โˆ™ If ๐›ผ is a limit ordinal, define
๐’ฏ๐›ผ =
โ‹ƒ
๐’ฏ๐›ฝ ,
๐›ฝ∈๐›ผ
which is consistent by Lemma 7.4.6.
We claim that the (๐–ฒ ∪ ๐ถ)-theory
โ‹ƒ
๐’ฏ๐›ฝ
๐›ฝ∈๐œ…
is the desired extension of ๐’ฏ . To see this, first note that this union is consistent by
Lemma 7.4.6. Also, let ๐‘(๐‘ฅ) be an ๐–ฒ ∪ ๐ถ-formula with at most one free variable.
This implies that ๐‘ = ๐‘๐›ผ and ๐‘ฅ = ๐‘ฅ๐›ผ for some ๐›ผ ∈ ๐œ…. Therefore,
∃๐‘ฅ๐›ผ ๐‘๐›ผ → ๐‘๐›ผ
so
โ‹ƒ
๐›พ∈๐œ…
๐‘‘๐›ผ
∈ ๐’ฏ๐›ผ+1 ,
๐‘ฅ๐›ผ
๐’ฏ๐›พ โŠข ∃๐‘ฅ๐›ผ ๐‘๐›ผ → ๐‘๐›ผ
๐‘‘๐›ผ
.
๐‘ฅ๐›ผ
Let ๐–ฒ be a set of theory symbols and ๐ถ be a set of constants from ๐–ฒ. Let ๐’ฏ be a
consistent ๐–ฒ-theory. We want to define an ๐–ฒ-structure that has a chance of being a model
of ๐’ฏ . One of the issues that must be overcome is whether any pair of constants should
be interpreted as being equal. If so, we want to view those constants as equivalent. To
do this, define a relation ∼ on ๐ถ by
๐‘0 ∼ ๐‘1 if and only if ๐‘0 = ๐‘1 ∈ ๐’ฏ
(7.22)
for all ๐‘0 , ๐‘1 ∈ ๐ถ. This defines an equivalence relation under the right conditions.
LEMMA 7.4.14
Let ๐ถ be a set of constants from ๐–ฒ. If ๐’ฏ is a maximally consistent ๐–ฒ-theory, then
∼ is an equivalence relation on ๐ถ.
402
Chapter 7 MODELS
PROOF
Let ๐‘0 , ๐‘1 , ๐‘2 ∈ ๐ถ.
โˆ™ Since ๐’ฏ ∪ {๐‘0 = ๐‘0 } is consistent by E1 (Axioms 5.1.1), by the maximal
consistency of ๐’ฏ , we have that ๐‘0 = ๐‘0 ∈ ๐’ฏ .
โˆ™ Suppose ๐‘0 = ๐‘1 ∈ ๐’ฏ . Then, ๐’ฏ โŠข ๐‘0 = ๐‘1 , so ๐’ฏ โŠข ๐‘1 = ๐‘0 by E2. Thus,
๐‘1 = ๐‘0 ∈ ๐’ฏ .
โˆ™ Let ๐‘0 = ๐‘1 ∈ ๐’ฏ and ๐‘1 = ๐‘2 ∈ ๐’ฏ . Hence, we have ๐’ฏ โŠข ๐‘0 = ๐‘1 and
๐’ฏ โŠข ๐‘1 = ๐‘2 . Then, ๐’ฏ โŠข ๐‘0 = ๐‘2 by E3, which yields ๐‘0 = ๐‘2 ∈ ๐’ฏ .
We define the domain of the desired structure to be ๐ถโˆ•∼, the set of equivalence classes
modulo ∼ (Definition 4.2.10). Before we define the function ๐”ž, we need some actual
functions and relations on ๐ถโˆ•∼. Use the following lemma to define them.
LEMMA 7.4.15
Let ๐’ฏ be a maximally consistent ๐–ฒ-theory and ๐ถ be a set of constant symbols
from ๐–ฒ. Let ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 , ๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 be ๐–ฒ-terms such that
๐‘ก0 ∼ ๐‘ก′0 , ๐‘ก1 ∼ ๐‘ก′1 , … , ๐‘ก๐‘›−1 ∼ ๐‘ก′๐‘›−1 .
โˆ™ For every ๐‘›-ary function symbol ๐‘“ ,
๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ∼ ๐‘“ (๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 ).
โˆ™ For every ๐‘›-ary relation symbol ๐‘…,
๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ∈ ๐’ฏ ⇔ ๐‘…(๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 ) ∈ ๐’ฏ .
PROOF
By (7.22) we have that
๐’ฏ โŠข ๐‘ก0 = ๐‘ก′0 , ๐’ฏ โŠข ๐‘ก1 = ๐‘ก′1 , … , ๐’ฏ โŠข ๐‘ก๐‘›−1 = ๐‘ก′๐‘›−1 .
Let ๐‘“ be an ๐‘›-ary function symbol. Then,
๐’ฏ โŠข ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) = ๐‘“ (๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 )
by E4 (Axioms 5.1.1), so
๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) = ๐‘“ (๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 ) ∈ ๐’ฏ
by Lemma 7.4.5, which implies that ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ∼ ๐‘“ (๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 ). Next,
let ๐‘… be an ๐‘›-ary relation symbol such that ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ∈ ๐’ฏ . This implies that ๐’ฏ โŠข ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ), so we have that ๐’ฏ โŠข ๐‘…(๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 ) by E5.
Again, by Lemma 7.4.5, ๐‘…(๐‘ก′0 , ๐‘ก′1 , … , ๐‘ก′๐‘›−1 ) ∈ ๐’ฏ .
Section 7.4 THE THREE PROPERTIES REVISITED
403
We now define the functions and relations on the domain ๐ถโˆ•∼ with ๐’ฏ being maximally
consistent.
โˆ™ Let ๐‘“ be an ๐‘›-ary function symbol from ๐–ฒ. Let ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ∈ ๐ถ. Define the
๐‘›-ary function ๐‘“∼ : ๐ถโˆ•∼ → ๐ถโˆ•∼ by
๐‘“∼ ([๐‘0 ], [๐‘1 ], … , [๐‘๐‘›−1 ]) = [๐‘] ⇔ ๐‘“ (๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ) = ๐‘ ∈ ๐’ฏ .
(7.23)
Since ๐‘“∼ is defined on a set of equivalence classes, we must confirm that ๐‘“∼ is
well-defined. Let ๐‘‘0 , ๐‘‘1 , … , ๐‘‘๐‘›−1 ∈ ๐ถ such that
๐‘0 ∼ ๐‘‘0 , ๐‘1 ∼ ๐‘‘1 , … , ๐‘๐‘›−1 ∼ ๐‘‘๐‘›−1 .
Then, [๐‘] = [๐‘‘] because by Lemma 7.4.15,
๐‘“ (๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ) ∼ ๐‘“ (๐‘‘0 , ๐‘‘1 , … , ๐‘‘๐‘›−1 ).
โˆ™ Let ๐‘… be an ๐‘›-ary relation symbol from ๐–ฒ and take ๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ∈ ๐ถ. Define
the ๐‘›-ary relation ๐‘…∼ on ๐ถโˆ•∼ by
([๐‘0 ], [๐‘1 ], … , [๐‘๐‘›−1 ]) ∈ ๐‘…∼ ⇔ ๐‘…(๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ) ∈ ๐’ฏ .
(7.24)
We must check that the relation holds if different constant symbols are used to
represent the classes [๐‘0 ], [๐‘1 ], … , [๐‘๐‘›−1 ]. To do this, let ๐‘‘0 , ๐‘‘1 , … , ๐‘‘๐‘›−1 ∈ ๐ถ
such that ๐‘0 ∼ ๐‘‘0 , ๐‘1 ∼ ๐‘‘1 , … , ๐‘๐‘›−1 ∼ ๐‘‘๐‘›−1 . Then, we have that
([๐‘‘0 ], [๐‘‘1 ], … , [๐‘‘๐‘›−1 ]) ∈ ๐‘…∼
since by Lemma 7.4.15,
๐‘…(๐‘‘0 , ๐‘‘1 , … , ๐‘‘๐‘›−1 ) ∈ ๐’ฏ .
The previous work makes the interpretation of the function and relation symbols of
๐–ฒ obvious. However, what about the constant symbols? To find out, take a constant
symbol ๐‘ of ๐–ฒ. If ๐‘ ∈ ๐ถ, then ๐‘ should be interpreted as [๐‘], so suppose that ๐‘ ∉ ๐ถ.
Because โŠข ∃๐‘ฅ(๐‘ฅ = ๐‘) and ๐’ฏ is maximally consistent,
∃๐‘ฅ(๐‘ฅ = ๐‘) ∈ ๐’ฏ .
At this point make the further assumption that ๐ถ is a witness set of ๐’ฏ . Then, there
exists ๐‘‘ ∈ ๐ถ such that
๐‘ =๐‘‘ ∈๐’ฏ.
This implies that [๐‘] = [๐‘‘]. Therefore, for every ๐‘ ∈ ๐–ฒ, there exists ๐‘∼ ∈ ๐ถ such that
[๐‘] = [๐‘∼ ]. We can now define the structure.
404
Chapter 7 MODELS
DEFINITION 7.4.16
Let ๐’ฏ be a maximally consistent set of ๐–ฒ-sentences with witness set ๐ถ. Let ∼
be the relation (7.22). Define ๐”„∼ to be the ๐–ฒ-structure with domain ๐ถโˆ•∼ and
function ๐”ž such that
โˆ™ ๐”ž(๐‘) = [๐‘∼ ] for all constant symbols ๐‘ ∈ ๐–ฒ,
โˆ™ ๐”ž(๐‘…) = ๐‘…∼ for all ๐‘›-ary relation symbols ๐‘… ∈ ๐–ฒ,
โˆ™ ๐”ž(๐‘“ ) = ๐‘“∼ for all ๐‘›-ary function symbols ๐‘“ ∈ ๐–ฒ.
Since we are investigating consistent sets of sentences, we do not need to involve a
particular interpretation for the structures in the proofs of the following results (Theorem 7.1.31). Using Definition 7.4.16 will suffice.
LEMMA 7.4.17
Suppose that ๐’ฏ is a maximally consistent set of ๐–ฒ-sentences with ๐ถ as a witness
set. Then, ๐”„∼ โŠจ ๐‘ก = ๐‘ if and only if ๐‘ก = ๐‘ ∈ ๐’ฏ , for every constant symbol ๐‘ ∈ ๐–ฒ
and ๐–ฒ-term ๐‘ก.
PROOF
Assume that ๐‘ ∈ ๐–ฒ is a constant symbol and let ๐‘ก be an ๐–ฒ-term. Since ๐’ฏ is a set
of sentences, ๐‘ก cannot have any free variables.
โˆ™ Let ๐‘ก = ๐‘‘ for some constant symbol ๐‘‘ from ๐–ฒ. Then,
๐”„∼ โŠจ ๐‘‘ = ๐‘ ⇔ [๐‘‘] = [๐‘] ⇔ ๐‘‘ ∼ ๐‘ ⇔ ๐‘‘ = ๐‘ ∈ ๐’ฏ .
โˆ™ Let ๐‘ก = ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) for some ๐‘›-ary function symbol ๐‘“ and ๐–ฒ-terms
๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 containing no free variables. Assume that
๐”„∼ โŠจ ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) = ๐‘.
Since ๐’ฏ is maximally consistent,
∃๐‘ฅ(๐‘ก๐‘– = ๐‘ฅ) ∈ ๐’ฏ
for ๐‘– = 0, 1, … , ๐‘› − 1. Since ๐ถ is a witness set, there exists ๐‘๐‘– ∈ ๐ถ such
that
๐‘ก๐‘– = ๐‘๐‘– ∈ ๐’ฏ ,
which is equivalent to
๐‘ก๐‘– ∼ ๐‘๐‘– .
Therefore, by induction,
๐”„∼ โŠจ ๐‘ก๐‘– = ๐‘๐‘– .
Hence,
๐”„∼ โŠจ ๐‘“ (๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ) = ๐‘.
(7.25)
Section 7.4 THE THREE PROPERTIES REVISITED
405
This implies that
๐‘“∼ ([๐‘0 ], [๐‘1 ], … , [๐‘๐‘›−1 ]) = [๐‘],
so by (7.23),
๐‘“ (๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ) = ๐‘ ∈ ๐’ฏ .
Therefore, by Lemma 7.4.15 and (7.25),
๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) = ๐‘ ∈ ๐’ฏ .
The converse is left to Exercise 4.
LEMMA 7.4.18
If ๐’ฏ is a consistent set of ๐–ฒ-sentences with ๐ถ as a witness set, then ๐‘ ∈ ๐’ฏ if and
only if ๐”„∼ โŠจ ๐‘ [๐ผ] for every ๐–ฒ-sentence ๐‘.
PROOF
We can assume that ๐’ฏ is maximally consistent (Exercise 5). Let ๐ถ be a witness
set of ๐’ฏ . We prove the theorem by induction on formulas.
โˆ™ Let ๐‘ก0 = ๐‘ก1 ∈ ๐’ฏ for ๐–ฒ-terms ๐‘ก0 and ๐‘ก1 with no free variables. Since ๐’ฏ is
maximally consistent,
∃๐‘ฅ(๐‘ก0 = ๐‘ฅ) ∈ ๐’ฏ
and
∃๐‘ฅ(๐‘ก1 = ๐‘ฅ) ∈ ๐’ฏ .
Because ๐ถ is a witness set for ๐’ฏ , there exists ๐‘, ๐‘‘ ∈ ๐ถ such that
๐‘ก0 = ๐‘ ∈ ๐’ฏ
and
๐‘ก1 = ๐‘‘ ∈ ๐’ฏ .
This implies that ๐‘ก0 ∼ ๐‘ and ๐‘ก1 ∼ ๐‘‘, so because ๐‘ก0 ∼ ๐‘ก1 ,
[๐‘] = [๐‘‘].
Therefore, ๐”„∼ โŠจ ๐‘ = ๐‘‘, but because ๐”„∼ โŠจ ๐‘ก0 = ๐‘ and ๐”„∼ โŠจ ๐‘ก1 = ๐‘‘
(Lemma 7.4.17), ๐”„∼ โŠจ ๐‘ก0 = ๐‘ก1 . The proof of the converse is Exercise 7(a).
โˆ™ Let ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ∈ ๐’ฏ for some relation symbol ๐‘… in ๐–ฒ and ๐–ฒ-terms
๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 with no free variables. As in the first part of this proof, there
exist constants ๐‘๐‘– ∈ ๐ถ (๐‘– = 0, 1, … , ๐‘› − 1) such that ๐‘ก๐‘– ∼ ๐‘๐‘– . Thus, by
Lemma 7.4.15, ๐‘…(๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ) ∈ ๐’ฏ , and then by the definition of ๐‘…∼ ,
we have that
([๐‘0 ], [๐‘1 ], … , [๐‘๐‘›−1 ]) ∈ ๐‘…∼ .
Therefore,
๐”„∼ โŠจ ๐‘…(๐‘0 , ๐‘1 , … , ๐‘๐‘›−1 ),
406
Chapter 7 MODELS
so ๐”„∼ โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 … , ๐‘ก๐‘›−1 ) [๐ผ] as in first part. The proof of the converse is
Exercise 7(b).
โˆ™ Let ๐‘ be an ๐–ฒ-sentence. Then, by induction and the maximal consistency
of ๐’ฏ ,
¬๐‘ ∈ ๐’ฏ ⇔ ๐‘ ∉ ๐’ฏ ⇔ ๐”„∼ ฬธโŠจ ๐‘ ⇔ ๐”„∼ โŠจ ¬๐‘ [๐ผ].
โˆ™ Let ๐‘ and ๐‘ž be ๐–ฒ-sentences. Assume ๐‘ → ๐‘ž ∈ ๐’ฏ and ๐”„∼ โŠจ ๐‘ [๐ผ]. By
induction, ๐‘ ∈ ๐’ฏ , so by Lemma 7.4.5 and MP, ๐‘ž ∈ ๐’ฏ . Again by induction,
๐”„∼ โŠจ ๐‘ž [๐ผ], so ๐”„∼ โŠจ ๐‘ → ๐‘ž [๐ผ]. See Exercise 7(c) for the proof of the
converse.
โˆ™ Let ๐‘ be an ๐–ฒ-formula with at most one free variable ๐‘ฅ. By induction and
the maximal consistency of ๐’ฏ ,
๐”„∼ โŠจ ∃๐‘ฅ๐‘ ⇔ ๐‘
๐‘
∈ ๐’ฏ for some ๐‘ ∈ ๐ถ ⇔ ∃๐‘ฅ๐‘ ∈ ๐’ฏ .
๐‘ฅ
The hard work of this section leads to the fundamental theorem of model theory.
THEOREM 7.4.19 [Henkin]
An ๐–ฒ-theory ๐’ฏ is consistent if and only if ๐’ฏ has a model.
PROOF
Let ๐’ฏ be a set of ๐–ฒ-sentences. Suppose ๐’ฏ is consistent and let ๐ถ be a set of
constant symbols not found in ๐–ฒ such that |๐ถ| = |๐–ซ(๐–ฒ)|. By Lemma 7.4.13, there
exists a consistent extension ๐’ฏ of ๐’ฏ such that the elements of ๐’ฏ are (๐–ฒ ∪ ๐ถ)sentences and ๐ถ is a witness set for ๐’ฏ . By Lemma 7.4.18, there exists a model
๐”„ of ๐’ฏ . Since the sentences of ๐’ฏ do not contain any of the constants of ๐ถ, the
reduct of ๐”„ to ๐–ฒ is a model of ๐’ฏ . To see this, let ๐”… be the indicated reduct and
proceed by induction on formulas.
โˆ™ Let ๐‘ก0 and ๐‘ก1 be ๐–ฒ-terms such that (๐‘ก0 = ๐‘ก1 ) ∈ ๐’ฏ . These terms are also
(๐–ฒ ∪ ๐ถ)-terms, so since ๐’ฏ ⊆ ๐’ฏ , we have that ๐”„ โŠจ ๐‘ก1 = ๐‘ก2 . Thus, there is
an (๐–ฒ∪๐ถ)-interpretation ๐ผ such that ๐”„ โŠจ ๐‘ก1 = ๐‘ก2 [๐ผ]. Hence, ๐ผ(๐‘ก1 ) = ๐ผ(๐‘ก2 ).
Let ๐ผ ′ be the restriction of ๐ผ to ๐–ฒ. Since ๐‘ก1 and ๐‘ก2 contain no symbols from
๐ถ, ๐ผ ′ (๐‘ก1 ) = ๐ผ ′ (๐‘ก2 ), so ๐”… โŠจ ๐‘ก1 = ๐‘ก2 [๐ผ ′ ] because ๐”… is the reduct of ๐”„ to ๐–ฒ.
By Theorem 7.1.31, we have that ๐”… โŠจ ๐‘ก1 = ๐‘ก2 .
โˆ™ That ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) ∈ ๐’ฏ implies ๐”… โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) for any ๐‘›-ary
relation symbol ๐‘… ∈ ๐–ฒ and ๐–ฒ-terms ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 is Exercise 8(a).
โˆ™ Let ¬๐‘ ∈ ๐’ฏ . This implies that ๐”„ โŠจ ¬๐‘. That is, not ๐”„ โŠจ ๐‘. Since ๐”… is the
reduct of ๐”„ to ๐–ฒ, not ๐”… โŠจ ๐‘. In other words, ๐”… โŠจ ¬๐‘.
โˆ™ That (๐‘ → ๐‘ž) ∈ ๐’ฏ implies ๐”… โŠจ ๐‘ → ๐‘ž for ๐–ฒ-sentences ๐‘ and ๐‘ž is Exercise 8(b).
Section 7.4 THE THREE PROPERTIES REVISITED
407
โˆ™ Suppose that ∃๐‘ฅ๐‘ ∈ ๐’ฏ , where ๐‘ = ๐‘(๐‘ฅ) is an ๐–ฒ-formula. Since ๐ถ is a
๐‘
witness set and ๐’ฏ is maximally consistent, ๐‘ ∈ ๐’ฏ for some ๐‘ ∈ ๐ถ. By
๐‘ฅ
induction,
๐‘
๐”„โŠจ๐‘ .
๐‘ฅ
Let ๐ผ be an interpretation of ๐”„. Then, ๐”„ โŠจ ๐‘ [๐ผ๐‘ฅ๐”ž(๐‘) ], so we conclude that
๐”„ โŠจ ∃๐‘ฅ๐‘.
To prove the converse, suppose that ๐’ฏ โŠข ๐‘ž ∧ ¬๐‘ž for some ๐–ฒ-sentence ๐‘ž. By
the soundness theorem (7.4.9), ๐’ฏ โŠจ ๐‘ž ∧ ¬๐‘ž. Hence, if ๐”„ is an ๐–ฒ-structure such
that ๐”„ โŠจ ๐’ฏ , then ๐”„ โŠจ ๐‘ž ∧ ¬๐‘ž, which implies that ๐”„ โŠจ ๐‘ž and not ๐”„ โŠจ ๐‘ž. Hence,
๐’ฏ does not have a model.
Henkin’s theorem (7.4.19) is used to prove the converse to the soundness theorem
(7.4.9).
COROLLARY 7.4.20 [Completeness]
Let ๐’ฏ be a consistent set of ๐–ฒ-sentences. For any ๐–ฒ-sentence ๐‘, if ๐’ฏ โŠจ ๐‘, then
๐’ฏ โŠข ๐‘.
PROOF
Suppose that ๐’ฏ โŠฌ ๐‘. This implies that ๐’ฏ ∪ {¬๐‘} is consistent, so by Henkin’s
theorem (7.4.19), there exists an ๐–ฒ-structure ๐”„ such that ๐”„ โŠจ ๐’ฏ ∪ {¬๐‘}. Hence,
๐’ฏ ฬธโŠจ ๐‘.
COROLLARY 7.4.21
First-order logic is complete.
The next corollary was first proved by Kurt Gödel (1929). It was his doctoral dissertation.
COROLLARY 7.4.22 [Gödel’s Completeness Theorem]
An ๐–ฒ-sentence is a theorem if and only if it is valid.
COROLLARY 7.4.23
For all sets of ๐–ฒ-sentences ๐’ฏ and ๐–ฒ-sentences ๐‘, ๐’ฏ โŠข ๐‘ if and only if ๐’ฏ โŠจ ๐‘.
The use of models to show consistency is now a common technique in mathematical
logic. An early example was Hilbert’s use of โ„ × โ„ to show that Euclidean geometry
is consistent (Hilbert 1899). That is, โ„ × โ„ serves as the domain of a structure that
models the postulates of Euclidean geometry. The proof relies on the consistency of
the properties of โ„ × โ„, a fact that is left unproved. A later usage is the model used by
Gödel that shows that both ๐–ข๐–ง and the axiom of choice (5.1.10) are consistent with ๐—ญ๐—™
(Gödel 1940). Both are examples of relative consistency proofs. Gödel assumed the
consistency of ๐—ญ๐—™, so he had a model for it. From this he defined another model that
408
Chapter 7 MODELS
satisfies ๐–ข๐–ง and, in addition, the axiom of choice. We represent this by writing
๐–ข๐—ˆ๐—‡(๐—ญ๐—™) → ๐–ข๐—ˆ๐—‡(๐—ญ๐—™๐—– + ๐–ข๐–ง).
This is a partial answer to one of the ten then unsolved problems that Hilbert presented at the International Congress of Mathematicians at Paris in 1900. An expanded
list was published that year in Germany with an English translation released two years
later (Hilbert 1902). Hilbert’s intention was to outline the important problems that
mathematicians should strive to prove in the twentieth century. The first problem involved the continuum hypothesis. Namely, it should be determined whether
as regards equivalence there are . . . only two assemblages of numbers, the
countable assemblage and the continuum.
Hilbert’s second problem dealt with the axioms of arithmetic. Specifically, he thought
that any of the chosen axioms should not be provable from the others. That is, they
should be independent. More importantly, mathematicians should seek
[t]o prove that they [the axioms] are not contradictory, that is, that a finite
number of logical steps based upon them can never lead to contradictory
results.
In this problem we see the notions of finite proof (Definition 1.2.13) and consistency
(Definitions 1.5.1 and 7.4.2), which would influence to the development of model theory some 20–30 years after Hilbert’s talk and lead to some interesting results.
Exercises
1. Is ∅ consistent? Explain.
2. Prove Theorem 7.4.3.
3. Prove that the group axioms (7.1.14) and the ring axioms (Axioms 7.1.36) are consistent.
4. From the proof of Lemma 7.4.17, prove that if ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) = ๐‘ ∈ ๐’ฏ , then
๐”„∼ โŠจ ๐‘“ (๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 ) = ๐‘ for all constant symbols ๐‘, ๐‘›-ary function symbols ๐‘“ , and
๐–ฒ-terms ๐‘ก0 , ๐‘ก1 , … , ๐‘ก๐‘›−1 .
5. In Lemma 7.4.18, why can ๐’ฏ be assumed to be maximally consistent?
6. Let ๐”„ be an ๐–ฒ-structure. Define ๐–ณ๐—(๐”„) to be the set of ๐–ฒ-sentences that are satisfied
by ๐”„. Prove that ๐–ณ๐—(๐”„) is maximally consistent.
7. Prove the given results from the proof of Lemma 7.4.18.
(a) If ๐”„∼ โŠจ ๐‘ก0 = ๐‘ก1 , then ๐‘ก0 = ๐‘ก1 ∈ ๐’ฏ for all ๐–ฒ-terms ๐‘ก0 and ๐‘ก1 with no free
variables.
(b) For all ๐–ฒ-terms ๐‘ก0 , ๐‘ก1 … , ๐‘ก๐‘›−1 with no free variables, ๐‘…(๐‘ก0 , ๐‘ก1 … , ๐‘ก๐‘›−1 ) ∈ ๐’ฏ
if ๐”„∼ โŠจ ๐‘…(๐‘ก0 , ๐‘ก1 … , ๐‘ก๐‘›−1 ).
(c) ๐”„∼ โŠจ ๐‘ → ๐‘ž implies that ๐‘ → ๐‘ž ∈ ๐’ฏ for all ๐–ฒ-sentences ๐‘ and ๐‘ž.
8. Show the following from the proof of Henkin’s theorem (7.4.19).
(a) If ๐‘…(๐‘ก1 , … , ๐‘ก๐‘› ) ∈ ๐’ฏ , then ๐”… โŠจ ๐‘…(๐‘ก1 , … , ๐‘ก๐‘› ) for any relation symbol ๐‘… ∈ ๐–ฒ
and ๐–ฒ-terms ๐‘ก1 , … , ๐‘ก๐‘› .
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
409
(b) If ๐‘ → ๐‘ž ∈ ๐’ฏ , then ๐”… โŠจ ๐‘ → ๐‘ž for all ๐–ฒ-sentences ๐‘ and ๐‘ž.
9. Prove Corollary 7.4.22.
10. Let ๐’ฏ and ๐’ฐ be consistent sets of ๐–ฒ-sentences. Prove that if ๐’ฏ ∪๐’ฐ is inconsistent,
then there exists an ๐–ฒ-sentence ๐‘ such that ๐’ฏ โŠจ ๐‘ but ๐’ฐ ฬธโŠจ ๐‘.
11. We say that an ๐–ฒ-theory ๐’ฏ is closed under deductions if for all ๐–ฒ-sentences ๐‘, if
๐’ฏ โŠข ๐‘, then ๐‘ ∈ ๐’ฏ .
(a) If a theory is closed under deductions, is the theory maximally consistent?
(b) Let ๐›ฝ be an ordinal and suppose that {๐’ฏ๐›ผ : ๐›ผ ∈ ๐›ฝ} is a family of ๐–ฒ-theories
such that for all ๐›พ, ๐›ฟ ∈ ๐›ฝ, ifโ‹ƒ๐›พ ∈ ๐›ฟ, then ๐’ฏ๐›พ ⊆ ๐’ฏ๐›ฟ . Prove that if ๐’ฏ๐›ผ is
consistent for all ๐›ผ ∈ ๐›ฝ, then ๐›ผ∈๐›ฝ ๐’ฏ๐›ผ is consistent.
12. A theory ๐’ฏ is finitely axiomatizable if there exists an ๐–ฒ-sentence ๐‘ such that for
all ๐–ฒ-sentences ๐‘ž, ๐’ฏ โŠจ ๐‘ž if and only if ๐‘ โŠจ ๐‘ž. Let {๐’ฏ๐‘› : ๐‘› ∈ ๐œ”} be a chain of finitely
axiomatizable ๐–ฒ-theories such that ๐‘š ≤ ๐‘› implies that ๐’ฏ๐‘š ⊆ ๐’ฏ๐‘› . Suppose that for all
โ‹ƒ
๐‘› ∈ ๐œ”, there exists a model of ๐’ฏ๐‘› that is not a model of ๐’ฏ๐‘›+1 . Prove that ๐‘›∈๐œ” ๐’ฏ๐‘› is
not finitely axiomatizable.
13. Suppose that ๐’ฏ1 and ๐’ฏ2 are ๐–ฒ-theories. Let ๐”„ and ๐”… be ๐–ฒ-structures. Prove the
following.
(a) If ๐’ฏ1 ⊆ ๐’ฏ2 , then every model of ๐’ฏ2 is a model of ๐’ฏ1 .
(b) If ๐”„ ⊆ ๐”…, then ๐–ณ๐—(๐”…) ⊆ ๐–ณ๐—(๐”„).
(c) If ๐”„ โŠจ ๐’ฏ1 ∪ ๐’ฏ2 , then ๐”„ โŠจ ๐’ฏ1 and ๐”„ โŠจ ๐’ฏ2 .
(d) ๐”„ โŠจ ๐’ฏ1 if and only if ๐’ฏ1 ⊆ ๐–ณ๐—(๐”„).
7.5
MODELS OF DIFFERENT CARDINALITIES
The natural numbers and their basic operations of addition and subtraction can be defined using the axioms of ๐—ญ๐—™๐—– (Section 5.2). It can then be proved that ๐œ” has basically
the same algebraic properties as ๐. Another approach to studying the natural numbers
is to break down the subject to its most basic parts. Think about addition of natural
numbers and how it is first explained. Adding 4 to 3 to obtain 7, for example, means
starting at 3 and adding 1 four times in sequence:
3 + 1 = 4,
4 + 1 = 5,
5 + 1 = 6,
6 + 1 = 7.
Adding 1 simply means moving to the next natural number, so to add any two natural
numbers, all one needs is to know the numerals and have the ability to count (compare
Definition 5.2.15). Although not efficient, this will do. Multiplication, the other operation, is based on addition. Multiplying 4 by 3 means writing 4 down 3 times and then
adding:
4 + 4 + 4 = 12.
410
Chapter 7 MODELS
This means that multiplication is also based on the ability to count (compare Definition 5.2.18). This ability is represented by the successor function,
๐‘†๐‘› = ๐‘› + 1,
which is the basis of arithmetic (compare Definition 5.2.1). Thus, to understand the
natural numbers and how they work, we simply need a successor function that satisfies
the right rules.
Peano Arithmetic
Using the theory symbols ๐– ๐–ฑ (Example 2.1.4), we make the following axioms. They
are a minor modification of the axioms for the system of arithmetic found in Arithmetices principia by Giuseppe Peano (1889).
AXIOMS 7.5.1
โˆ™ P1. ∀๐‘ฅ(¬ ๐‘†๐‘ฅ = 0)
โˆ™ P2. ∀๐‘ฅ∀๐‘ฆ(๐‘†๐‘ฅ = ๐‘†๐‘ฆ → ๐‘ฅ = ๐‘ฆ)
โˆ™ P3. For every ๐– ๐–ฑ-formula ๐‘ with free variable ๐‘ฅ,
๐‘(0) ∧ ∀๐‘ฅ[๐‘(๐‘ฅ) → ๐‘(๐‘†๐‘ฅ)] → ∀๐‘ฅ๐‘(๐‘ฅ).
Denote {P1, P2, P3} by ๐—ฃ. If we define
๐”ญ(๐‘†) = + ,
๐”ญ(0) = ∅,
(7.26)
then Theorem 5.2.2, Theorem 5.2.6, and Corollary 5.2.12 imply that ๐”“ = (๐œ”, ๐”ญ) is a
model of ๐—ฃ. Because of the existence of a model for these axioms, which was constructed using ๐—ญ๐—™๐—–, we conclude the following (Theorem 7.4.19).
THEOREM 7.5.2
The consistency of ๐—ฃ is a consequence of ๐—ญ๐—™๐—–.
Since the Peano axioms are intended to be the assumptions of number theory, ๐– ๐–ฑ is
often extended to ๐– ๐–ฑ′ (Example 2.1.4) and the sentences of Axioms 7.5.1 broadened to
include axioms involving +, ⋅, and <. These axioms serve as the foundation of number
theory.
AXIOMS 7.5.3 [Peano]
โˆ™ PA1. ∀๐‘ฅ(¬ ๐‘†๐‘ฅ = 0)
โˆ™ PA2. ∀๐‘ฅ∀๐‘ฆ(๐‘†๐‘ฅ = ๐‘†๐‘ฆ → ๐‘ฅ = ๐‘ฆ)
โˆ™ PA3. ∀๐‘ฅ(๐‘ฅ + 0 = ๐‘ฅ)
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
411
โˆ™ PA4. ∀๐‘ฅ∀๐‘ฆ[๐‘ฅ + ๐‘†๐‘ฆ = ๐‘†(๐‘ฅ + ๐‘ฆ)]
โˆ™ PA5. ∀๐‘ฅ(๐‘ฅ ⋅ 0 = 0)
โˆ™ PA6. ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⋅ ๐‘†๐‘ฆ = ๐‘ฅ ⋅ ๐‘ฆ + ๐‘ฅ)
โˆ™ PA7. ∀๐‘ฅ(¬ ๐‘ฅ < 0)
โˆ™ PA8. ∀๐‘ฅ∀๐‘ฆ[๐‘ฅ < ๐‘†๐‘ฆ ↔ (๐‘ฅ < ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ)]
โˆ™ PA9. ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ < ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ ∨ ๐‘ฆ < ๐‘ฅ)
โˆ™ PA10. For every ๐– ๐–ฑ′ -formula ๐‘ with free variable ๐‘ฅ,
๐‘(0) ∧ ∀๐‘ฅ[๐‘(๐‘ฅ) → ๐‘(๐‘†๐‘ฅ)] → ∀๐‘ฅ๐‘(๐‘ฅ).
Let ๐—ฃ๐—” denote the Peano axioms (7.5.3). We call the set of consequences of the Peano
axioms Peano arithmetic. To find a model for these axioms, extend the function ๐”ญ
(7.26) to ๐”ญ′ so that ๐”ญ′ (+) is the addition of Definition 5.2.15, ๐”ญ′ (⋅) is the multiplication
of Definition 5.2.19, and using the order of Definition 5.2.10,
๐”ญ′ (<) = {(๐‘š, ๐‘›) ∈ ๐œ” × ๐œ” : ๐‘š ∈ ๐‘›}.
That ๐”ญ′ (<) satisfies PA7–PA9 is left to Exercise 2. Then, as before, ๐”“′ = (๐œ”, ๐”ญ′ ) is a
model of the Peano axioms, which is an expansion of the model ๐”“, and we can again
apply Theorem 7.4.19 to obtain a consistency result.
THEOREM 7.5.4
The consistency of ๐—ฃ๐—” is a consequence of ๐—ญ๐—™๐—–.
We call ๐”“′ and any ๐– ๐–ฑ′ -structure isomorphic to it a standard model of Peano arithmetic. The the elements of the domain of any standard model are called standard
numbers. Any model of Peano arithmetic that is not isomorphic to ๐”“′ is called a
nonstandard model of Peano arithmetic.
Observe that ๐—ฃ and ๐—ฃ๐—” are sets of axioms separate from ๐—ญ๐—™๐—–. However, because
of the work of Section 5.2, both ๐—ฃ and ๐—ฃ๐—” can be viewed as consequences of ๐—ญ๐—™๐—–.
Specifically, if P1 is replaced by
P2 is replaced by
∀๐‘ฅ(๐‘ฅ+ ≠ ∅),
(7.27)
∀๐‘ฅ∀๐‘ฆ(๐‘ฅ+ = ๐‘ฆ+ → ๐‘ฅ = ๐‘ฆ),
(7.28)
๐‘(∅) ∧ ∀๐‘ฅ[๐‘(๐‘ฅ) → ๐‘(๐‘ฅ+ )] → ∀๐‘ฅ๐‘(๐‘ฅ)
(7.29)
and P3 is replaced by
for every ๐–ฒ๐–ณ-formula ๐‘ with free variable ๐‘ฅ, then
๐—ญ๐—™๐—– โŠจ {(7.27), (7.28), (7.27)}.
412
Chapter 7 MODELS
That is, a copy of ๐—ฃ is found among the consequences of ๐—ญ๐—™๐—–. Also, using Definitions 5.2.15 and 5.2.18, it can be shown that a copy of Peano arithmetic is found among
the consequences of ๐—ญ๐—™๐—–.
We will not repeat all our number theory work in Peano arithmetic, although it is
possible to do so. Instead, we give a sample of some basic results to illustrate that
the two systems are the same. Because of the work in Section 5.2, the addition and
multiplication of Peano arithmetic are commutative, associative, satisfy the distributive
and cancellation laws, and have identities in the standard model. That these properties
hold in every model of Peano arithmetic is a separate issue, yet their proofs are similar
to the work of Section 5.2.
THEOREM 7.5.5
The following ๐– ๐–ฑ′ -sentences are theorems of ๐—ฃ๐—”.
โˆ™ Associative Laws
∀๐‘ฅ∀๐‘ฆ∀๐‘ง[๐‘ฅ + (๐‘ฆ + ๐‘ง) = (๐‘ฅ + ๐‘ฆ) + ๐‘ง]
∀๐‘ฅ∀๐‘ฆ∀๐‘ง[๐‘ฅ ⋅ (๐‘ฆ ⋅ ๐‘ง) = (๐‘ฅ ⋅ ๐‘ฆ) ⋅ ๐‘ง]
โˆ™ Commutative Laws
∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ฆ = ๐‘ฆ + ๐‘ฅ)
∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⋅ ๐‘ฆ = ๐‘ฆ ⋅ ๐‘ฅ)
โˆ™ Additive Identity
∀๐‘ฅ(0 + ๐‘ฅ = ๐‘ฅ)
โˆ™ Multiplicative Identity
∀๐‘ฅ(๐‘†0 ⋅ ๐‘ฅ = ๐‘ฅ)
โˆ™ Distributive Law
∀๐‘ฅ∀๐‘ฆ∀๐‘ง[๐‘ฅ ⋅ (๐‘ฆ + ๐‘ง) = ๐‘ฅ ⋅ ๐‘ฆ + ๐‘ฅ ⋅ ๐‘ง]
PROOF
We prove the first associative law and the multiplicative identity property, leaving
the others to Exercise 9.
โˆ™ Define
๐‘(๐‘ฅ) := ∀๐‘ง[๐‘ฅ + (๐‘ฆ + ๐‘ง) = (๐‘ฅ + ๐‘ฆ) + ๐‘ง].
By PA3, we have ๐‘(0) because
๐‘ฅ + (๐‘ฆ + 0) = (๐‘ฅ + ๐‘ฆ) + 0.
Next, assume ๐‘(๐‘˜). Since ๐‘† is a function symbol, PA4 yields
๐‘† [๐‘ฅ + (๐‘ฆ + ๐‘˜)] = ๐‘† [(๐‘ฅ + ๐‘ฆ) + ๐‘˜] ,
๐‘ฅ + ๐‘†(๐‘ฆ + ๐‘˜) = (๐‘ฅ + ๐‘ฆ) + ๐‘†๐‘˜,
๐‘ฅ + (๐‘ฆ + ๐‘†๐‘˜) = (๐‘ฅ + ๐‘ฆ) + ๐‘†๐‘˜.
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
413
Therefore, ๐‘(๐‘†๐‘˜), so the associative law for multiplication holds by PA10.
โˆ™ Assuming that the commutative laws have already been proved, PA6, PA5,
and PA3 imply that
๐‘†0 ⋅ ๐‘ฅ = ๐‘ฅ ⋅ ๐‘†0 = ๐‘ฅ ⋅ 0 + ๐‘ฅ = 0 + ๐‘ฅ = ๐‘ฅ + 0 = ๐‘ฅ,
so ๐‘†0 ⋅ ๐‘ฅ = ๐‘ฅ by E3 (Axioms 5.1.1).
To ease our notation when dealing with numbers other than 1 = ๐‘†0, define
๐‘† 0๐‘ฅ = ๐‘ฅ
and for all ๐‘› ∈ ๐œ” โงต {0},
๐‘† ๐‘› = ๐‘†๐‘† … ๐‘† ๐‘ฅ.
โŸโžโŸโžโŸ
๐‘› times
We will use this notation in the next example.
(7.30)
(7.31)
EXAMPLE 7.5.6
In the formula,
(3 + 2) + 1 = (1 + 3) + 2,
(7.32)
two properties from Theorem 7.5.5 are being applied. First, we conclude that
(3 + 2) + 1 = 1 + (3 + 2) by the commutative law. The associative law is then
used to draw the final conclusion. Notice that (7.32) is the standard interpretation
of
++๐‘†๐‘†๐‘†0๐‘†๐‘†0๐‘†0 = +๐‘†0๐‘†๐‘†๐‘†0 + ๐‘†๐‘†0,
which, using (7.31), is equivalent to
++๐‘† 3 0๐‘† 2 0๐‘† 1 0 = +๐‘† 1 0๐‘† 3 0 + ๐‘† 2 0.
Also, the distributive law is applied in the formula
2 ⋅ ๐‘ฅ + 3 ⋅ ๐‘ฅ = 5 ⋅ ๐‘ฅ,
because
2 ⋅ ๐‘ฅ + 3 ⋅ ๐‘ฅ = ๐‘ฅ ⋅ 2 + ๐‘ฅ ⋅ 3 = ๐‘ฅ ⋅ (2 + 3) = ๐‘ฅ ⋅ 5 = 5 ⋅ ๐‘ฅ.
Given an equation like
5 ⋅ ๐‘ฅ + 1 = 11,
(7.33)
the standard routine is to solve it like this:
5 ⋅ ๐‘ฅ + 1 = 11,
5 ⋅ ๐‘ฅ = 10,
๐‘ฅ = 2.
In the first step, −1 was added to both sides, and in the second, 1โˆ•5 was multiplied by
both sides. However, only 0 has an additive inverse in Peano arithmetic, and only 1 has
a multiplicative inverse. The way around this problem are cancellation laws.
414
Chapter 7 MODELS
THEOREM 7.5.7 [Cancellation]
The following ๐– ๐–ฑ′ -sentences are theorems of ๐—ฃ๐—”.
โˆ™ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง[(๐‘ฅ + ๐‘ง = ๐‘ฆ + ๐‘ง) → ๐‘ฅ = ๐‘ฆ]
โˆ™ ∀๐‘ฅ∀๐‘ฆ∀๐‘ง[(๐‘ฅ ⋅ ๐‘ง = ๐‘ฆ ⋅ ๐‘ง ∧ ¬๐‘ง = 0) → ๐‘ฅ = ๐‘ฆ].
PROOF
The proof of the cancellation law for multiplication is Exercise 10. For addition,
define
๐‘(๐‘ง) := ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ + ๐‘ง = ๐‘ฆ + ๐‘ง → ๐‘ฅ = ๐‘ฆ).
First, notice that if ๐‘ฅ + 0 = ๐‘ฆ + 0,
๐‘ฅ = ๐‘ฅ + 0 = ๐‘ฆ + 0 = ๐‘ฆ,
where the first and last equality hold by PA3. Therefore, ๐‘(0). For the induction
step, suppose ๐‘(๐‘˜). Then, by PA4 and PA2,
๐‘ฅ + ๐‘†๐‘˜ = ๐‘ฆ + ๐‘†๐‘˜,
๐‘†(๐‘ฅ + ๐‘˜) = ๐‘†(๐‘ฆ + ๐‘˜),
๐‘ฅ + ๐‘˜ = ๐‘ฆ + ๐‘˜,
๐‘ฅ = ๐‘ฆ,
where the last step follows by ๐‘(๐‘˜). Therefore, ∀๐‘ง๐‘(๐‘ง) by PA10.
Therefore, in Peano arithmetic, solve (7.33) like this:
5 ⋅ ๐‘ฅ + 1 = 11,
5 ⋅ ๐‘ฅ + 1 = 10 + 1,
5 ⋅ ๐‘ฅ = 10,
5 ⋅ ๐‘ฅ = 5 ⋅ 2,
๐‘ฅ = 2.
The third and last equations follow by cancellation (Theorem 7.5.7).
Compactness Theorem
Having a model that satisfies ๐—ฃ or ๐—ฃ๐—” means that basic arithmetic can be done in this
model. We know that this happens in the standard model (assuming ๐—ญ๐—™๐—–), so we now
want to know whether there are any nonstandard models of Peano arithmetic. To be
successful in our search, we turn to some theorems that follow from Henkin’s theorem
(7.4.19). The first states that the existence of a model rests on the finite. It was first
proved by Gödel for countable first-order theories (1930) and by Anatolij Mal’tsev for
arbitrary first-order theories (1936).
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
415
THEOREM 7.5.8 [Compactness]
An ๐–ฒ-theory ๐’ฏ has a model if and only if every finite subset of ๐’ฏ has a model.
PROOF
Since sufficiency is clear, let ๐’ฏ have the property that every finite subset has
a model. This implies that every finite subset of ๐’ฏ is consistent by Henkin’s
theorem. Therefore, ๐’ฏ is consistent by Theorem 7.4.3, so ๐’ฏ also has a model
by Henkin’s Theorem.
We apply the compactness theorem to find another model of ๐—ฃ๐—”. Let ๐‘ be a constant
symbol other than 0. For every natural number ๐‘›, use (7.30) and (7.31) to define the
set of (๐– ๐–ฑ′ ∪ {๐‘})-sentences
โ„ฑ๐‘› = ๐—ฃ๐—” ∪ {๐‘† ๐‘– 0 < ๐‘ : ๐‘– ∈ ๐œ” ∧ ๐‘– ≤ ๐‘›},
and let ๐”„๐‘› = ({0, 1, … , ๐‘› + 1}, ๐”ž), where
๐”ž ๐– ๐–ฑ′ = ๐”ญ′
and
๐”ž(๐‘) = ๐‘› + 1.
Observe that
๐”„๐‘› โŠจ โ„ฑ๐‘› ,
so โ„ฑ๐‘› is consistent for all ๐‘› ∈ ๐œ”, which implies that it has a model by Henkin’s Theorem
(7.4.19). Therefore,
โ‹ƒ
โ„ฑ =
โ„ฑ๐‘›
๐‘›∈๐œ”
has a model by the compactness theorem (7.5.8). Let ๐”„ be the reduct of this model
to ๐– ๐–ฑ′ . Notice that ๐”„ โŠจ ๐—ฃ๐—”, but because it has an element that is interpreted to be
greater than every element of its domain, ๐”„ is a nonstandard model of Peano arithmetic
(Theorem 7.3.24).
Löwenheim–Skolem Theorems
Now that we know that a nonstandard model of Peano arithmetic exists under ๐—ญ๐—™๐—–,
we want to know if there are others. For this, we need a definition. The power of an
๐–ฒ-structure ๐”„ is denoted by |๐”„| and refers to the cardinality of the domain of ๐”„. This
means that a model is countable if and only if its power is countable. We introduce a
sequence of theorems related to this. They are due to Skolem (1922), Tarski and Vaught
(1957), and Leopold Löwenheim (1915).
THEOREM 7.5.9 [Downward Löwenheim–Skolem]
Every consistent ๐–ฒ-theory has a model with power of at most |๐–ซ(๐–ฒ)|.
416
Chapter 7 MODELS
PROOF
Recall these facts from the proof of Henkin’s theorem (7.4.19).
โˆ™ ๐ถ is a set of new constant symbols such that |๐ถ| = |๐–ซ(๐–ฒ)|.
โˆ™ ๐”„ is a (๐–ฒ ∪ ๐ถ)-structure.
โˆ™ ๐”… is the reduct of ๐”„ to ๐–ฒ.
We can assume that ๐”„ has the property that every element of the domain of ๐”„ is
the interpretation of a constant of ๐ถ. Then, since ๐”„ and ๐”… have the same power,
|๐”…| ≤ |๐–ซ(๐–ฒ ∪ ๐ถ)| = |๐–ซ(๐–ฒ)|.
Since the witness set of a single sentence is finite and there are only finitely many
symbols in a sentence, ๐–ฒ ∪ ๐ถ can be assumed to be finite in the proof of the downward
Löwenheim–Skolem theorem (7.5.9). If this is the case, |๐–ซ(๐–ฒ ∪ ๐ถ)| = |๐–ซ(๐–ฒ)| = ℵ0 , and
we have the next corollary.
COROLLARY 7.5.10 [Löwenheim]
If an ๐–ฒ-sentence has a model, it has a countable model.
Since ๐–ซ(๐– ๐–ฑ′ ) is countable, Theorem 7.5.9 also implies the following.
COROLLARY 7.5.11 [Skolem]
๐—ญ๐—™๐—– implies that there is a countable nonstandard model of Peano arithmetic.
In fact, although we do not prove it here, ๐—ญ๐—™๐—– implies that there are 2ℵ0 countable
nonstandard models of Peano arithmetic.
The title of Theorem 7.5.9 suggests the existence of the following theorem.
THEOREM 7.5.12 [Upward Löwenheim–Skolem]
If an ๐–ฒ-theory has an infinite model, it has a model of cardinality ๐œ… for every
๐œ… ≥ |๐–ซ(๐–ฒ)|.
PROOF
Let ๐’ฏ be an ๐–ฒ-theory with an infinite model ๐”„ = (๐ด, ๐”ž). Take ๐œ… ≥ |๐–ซ(๐–ฒ)|.
Choose a set ๐ถ = {๐‘๐›ผ : ๐›ผ ∈ ๐œ…} of distinct constant symbols not found in ๐–ฒ.
Extend ๐’ฏ to the (๐–ฒ ∪ ๐ถ)-theory ๐’ฏ by defining
๐’ฏ = ๐’ฏ ∪ {๐‘๐›ผ ≠ ๐‘๐‘ : ๐›ผ ∈ ๐›ฝ ∈ ๐œ…}.
Let ๐’ฎ be a finite subset of ๐’ฏ . This implies that there exists
๐ถ ′ = {๐‘๐›ผ0 , ๐‘๐›ผ1 , … , ๐‘๐›ผ๐‘›−1 } ⊆ ๐ถ
such that the constants of the sentences of ๐’ฎ are among the constants of ๐ถ ′ .
Expand the ๐–ฒ-structure ๐”„ to the (๐–ฒ ∪ {๐‘๐›ผ0 , ๐‘๐›ผ1 , … , ๐‘๐›ผ๐‘›−1 })-structure ๐”„′ = (๐ด, ๐”ž′ ),
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
where ๐”ž′ ๐–ฒ = ๐”ž and
417
๐”ž′ (๐‘๐›ผ๐‘– ) ∈ ๐ด
for all ๐‘– = 0, 1, … , ๐‘› − 1. Since ๐ด is infinite, we further assume that
๐”ž′ (๐‘๐›ผ0 ), ๐”ž′ (๐‘๐›ผ1 ), … , ๐”ž′ (๐‘๐›ผ๐‘›−1 )
are distinct. Therefore, ๐”„′ โŠจ ๐’ฎ , so by the compactness theorem (7.5.8), there
exists a model ๐”… of ๐’ฏ , and due to the interpretation of the new constants, the
power of ๐”… is ๐œ…. Thus, the reduct of ๐”… to ๐–ฒ is a model of ๐’ฏ and has power ๐œ….
The upward Löwenheim–Skolem theorem with ๐—ญ๐—™๐—– implies that the Peano axioms
have models of all infinite cardinalities, which are nonstandard by definition.
The von Neumann Hierarchy
We constructed a model of ๐—ฃ๐—” using ๐—ญ๐—™๐—–. If we can find models of ๐—ญ๐—™๐—–, we would
have other models of ๐—ฃ๐—”, plus prove the consistency of ๐—ญ๐—™๐—–. In order to find models
of ๐—ญ๐—™๐—–, we begin by searching for models of individual axioms using a definition due
to von Neumann (1929). The objective of the definition is to construct a sequence
of sets that have the property that every set is in one of the stages of the sequence.
However, since von Neumann’s definition used functions instead of sets, it was Zermelo
(1930) who gave it its more recognizable form. While von Neumann left his base stage
empty, Zermelo allowed the first set of the sequence to contain objects, which are called
urelements, that were not sets yet were allowed to be elements of sets (Exercise 20).
These two approaches are combined in the following definition, which is named after
von Neumann.
DEFINITION 7.5.13
Let ๐‘‰0 = ∅. Let ๐›ผ be an ordinal.
โˆ™ ๐‘‰๐›ผ+1 = P(๐‘‰๐›ผ ).
โ‹ƒ
โˆ™ ๐‘‰๐›ผ = ๐›ฝ<๐›ผ ๐‘‰๐›ฝ if ๐›ผ is a limit ordinal.
This is called the von Neumann hierarchy.
As proved in Exercise 22, ๐‘‰๐›ผ is a set for every ordinal ๐›ผ by the empty set, union, power
set, and replacement axioms (5.1.2, 5.1.6, 5.1.7, 5.1.9). Thus, every element of ๐‘‰๐›ผ is a
set. For example,
๐‘‰1 = {∅},
๐‘‰2 = {∅, {∅}},
๐‘‰3 = {∅, {∅}, {{∅}}, {∅, {∅}}}
โ‹ฎ
The sets in ๐‘‰๐œ” are finite and called the hereditarily finite sets. There are countably
many of these sets because ๐‘‰๐‘› is countable for each ๐‘› ∈ ๐œ” (Theorem 6.3.17), while
418
Chapter 7 MODELS
|๐‘‰๐œ”+1 | = 2ℵ0 . Observe that the cardinality of ๐‘‰๐‘› for each ๐‘› ∈ ๐œ” is finite but grows very
quickly.
|๐‘‰0 | = 0,
|๐‘‰1 | = 1,
|๐‘‰2 | = 2,
|๐‘‰3 | = 22 ,
2
|๐‘‰4 | = 22 ,
22
|๐‘‰5 | = 22 ,
2
22
|๐‘‰6 | = 22
โ‹ฎ
Before we can use individual stages of von Neumann’s hierarchy, we need to know
some of their key properties. For this, we start with some lemmas.
LEMMA 7.5.14
Let ๐›ผ and ๐›ฝ be ordinals.
โˆ™ ๐‘‰๐›ผ is a transitive set.
โˆ™ If ๐›ผ ⊆ ๐›ฝ, then ๐‘‰๐›ผ ⊆ ๐‘‰๐›ฝ .
PROOF
Let ๐œ be a limit ordinal containing ๐›ผ and ๐›ฝ. Define
๐ด = {๐œ‚ ∈ ๐œ : ๐‘‰๐œ‚ is transitive}.
Assume that ๐—Œ๐–พ๐—€(๐ด, ๐›ฟ) ⊆ ๐ด for the ordinal ๐›ฟ ∈ ๐œ. Let ๐ต ∈ ๐‘‰๐›ฟ .
โˆ™ ๐‘‰0 is transitive because ๐‘‰0 = ∅.
โˆ™ Suppose that ๐›ฟ = ๐›พ + for some ordinal ๐›พ. By definition, ๐ต ∈ P(๐‘‰๐›พ ), so
๐ต ⊆ ๐‘‰๐›พ . Take ๐‘ฅ ∈ ๐ต. This implies that ๐‘ฅ ∈ ๐‘‰๐›พ . Because ๐‘‰๐›พ is transitive,
we have that ๐‘ฅ ⊆ ๐‘‰๐›พ . Hence, ๐‘ฅ ∈ P(๐‘‰๐›พ ) = ๐‘‰๐›ฟ , so ๐ต ⊆ ๐‘‰๐›ฟ .
โˆ™ Let ๐›ฟ be a limit ordinal. Then, there exists ๐›พ ∈ ๐›ฟ such that ๐ต ∈ ๐‘‰๐›พ . Since
๐‘‰๐›พ is transitive by hypothesis, ๐ต ⊆ ๐‘‰๐›พ ⊆ ๐‘‰๐›ฟ .
We conclude that ๐›ฟ ∈ ๐ด. Thus, by transfinite induction (Theorem 6.1.18), ๐ด = ๐œ
and ๐‘‰๐›ผ is transitive.
Now, suppose ๐›ผ ⊆ ๐›ฝ and take ๐‘ฅ ∈ ๐‘‰๐›ผ . Since
๐‘‰๐›ผ ∈ P(๐‘‰๐›ผ ) ⊆ ๐‘‰๐›ฝ ,
๐‘‰๐›ผ ∈ ๐‘‰๐›ฝ . However, the first part shows that ๐‘‰๐›ฝ is transitive, so ๐‘‰๐›ผ ⊆ ๐‘‰๐›ฝ .
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
419
LEMMA 7.5.15
If ๐›ผ is an ordinal, then ๐›ผ ∈ ๐‘‰๐›ผ+ .
PROOF
Let ๐›ผ be an ordinal and take ๐œ to be a limit ordinal such that ๐›ผ ∈ ๐œ and proceed
by transfinite induction.
โˆ™ Since ๐‘‰0 = ∅, we have that ∅ ∈ ๐‘‰1 .
โˆ™ Suppose that ๐›ผ ∈ ๐‘‰๐›ผ+ . Since ๐‘‰๐›ผ+ is transitive (Lemma 7.5.14), ๐›ผ ⊂ ๐‘‰๐›ผ+ .
Also, {๐›ผ} ⊂ ๐‘‰๐›ผ+ . Therefore, ๐›ผ ∪ {๐›ผ} ⊂ ๐‘‰๐›ผ+ , so ๐›ผ + ∈ ๐‘‰๐›ผ ++ .
โˆ™ Let ๐›ผ be a limit ordinal and assume that ๐›ฝ ∈ ๐‘‰๐›ฝ + for all ๐›ฝ ∈ ๐›ผ. That is,
๐›ฝ ⊂ ๐‘‰๐›ฝ + for all ๐›ฝ ∈ ๐›ผ since each ๐‘‰๐›ฝ + is transitive. Hence,
๐›ผ=
โ‹ƒ
๐›ฝ∈๐›ผ
๐›ฝ⊆
โ‹ƒ
๐‘‰๐›ฝ + = ๐‘‰๐›ผ ,
๐›ฝ∈๐›ผ
which implies that ๐›ผ ∈ ๐‘‰๐›ผ+ .
The definition of the von Neumann hierarchy along with the fact that every ordinal
belongs to a member of the hierarchy suggests that many, if not all, sets also belong to
the hierarchy. For this reason, we define the following.
DEFINITION 7.5.16
Let ๐—ฉ denote the collection of all sets ๐ด such that ๐ด ∈ ๐‘‰๐›ผ for some ordinal ๐›ผ.
It is the case that all sets belong to the hierarchy. This is the next theorem. Its proof is
aided by the use of two terms and a lemma.
โˆ™ A set ๐ด is grounded if there exists an ordinal ๐›ผ such that ๐ด ⊆ ๐‘‰๐›ผ .
โˆ™ The transitive closure of ๐ด is
TC(๐ด) = {๐‘ข : ∀๐‘ฃ(๐ด ∈ ๐‘ฃ ∧ ๐‘ฃ is transitive → ๐‘ข ∈ ๐‘ฃ)}.
As the name implies, TC(๐ด) is a transitive set (Exercise 24).
The proof of the next lemma is left to Exercise 21.
LEMMA 7.5.17
Every element of ๐ด is grounded if and only if ๐ด is grounded.
THEOREM 7.5.18
For every set ๐ด, there exists an ordinal ๐›ผ such that ๐ด ∈ ๐‘‰๐›ผ .
420
Chapter 7 MODELS
PROOF
Suppose that ๐ด is a set such that ๐ด ∉ ๐‘‰๐›ผ for all ordinals ๐›ผ. If ๐ด ⊆ ๐‘‰๐›ฝ for some
ordinal ๐›ฝ, then ๐ด ∈ ๐‘‰๐›ฝ+ , which contradicts the hypothesis. Hence, ๐ด is not
grounded, which implies that {๐ด} is not grounded (Lemma 7.5.17). Define
๐ต = {๐‘ข ∈ TC({๐ด}) : ๐‘ข is not grounded}.
Since ๐ต ≠ ∅, the regularity axiom (5.1.15) implies that there exists ๐ถ ∈ ๐ต such
that ๐ถ ∩ ๐ต = ∅. Let ๐‘ฅ ∈ ๐ถ. Since transitive closures are transitive, ๐‘ฅ ∈ TC(๐ด).
However, ๐‘ฅ ∉ ๐ต, so ๐‘ฅ is grounded. From this we conclude that ๐ถ is grounded
by Lemma 7.5.17, a contradiction.
Therefore, every set is in ๐—ฉ, but we know by Corollary 5.1.17 that ๐—ฉ is not a set in that
it cannot be built using ๐—ญ๐—™๐—–. However, we sometimes want to refer to such collections
even though they are not sets. For this reason the term class was introduced, so we call
๐—ฉ the class of all sets.
We are now ready to use sets from the von Neumann hierarchy to serve as models
for axioms from ๐—ญ๐—™๐—– (Section 5.1).
DEFINITION 7.5.19
Let ๐›ผ be an ordinal. Define the ๐–ฒ๐–ณ-structure ๐”™๐›ผ = (๐‘‰๐›ผ , ∈).
Consider
๐”™3 = ({∅, {∅}, {{∅}}, {∅, {∅}}}, ∈).
(7.34)
The elements of ๐‘‰3 are the sets of the model. These elements are equal exactly when
they share the same elements (Definition 3.3.7), so
๐”™3 โŠจ extensionality axiom.
Because the union of any two elements of ๐‘‰3 is an element of ๐‘‰3 , such as
∅ ∪ {∅} = {∅}
and
{{∅}} ∪ {∅, {∅}} = {∅, {∅}},
we see that
๐”™3 โŠจ union axiom.
If we take a ๐–ฒ๐–ณ-formula ๐‘(๐‘ฅ) and ๐ด ∈ ๐‘‰3 , then {๐‘ฅ : ๐‘ฅ ∈ ๐ด ∧ ๐‘(๐‘ฅ)} ∈ ๐‘‰3 . Hence,
๐”™3 โŠจ subset axioms.
Because ๐‘‰3 is finite,
and we have that
๐”™3 โŠจ axiom of choice,
๐”™3 โŠจ axiom of regularity
since, for example,
{∅} ∩ {{∅}} = ∅.
These results are particular examples of the next general theorem.
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
421
THEOREM 7.5.20
If ๐›ผ is an ordinal,
โˆ™ ๐”™๐›ผ โŠจ extensionality axiom,
โˆ™ ๐”™๐›ผ โŠจ union axiom,
โˆ™ ๐”™๐›ผ โŠจ subset axioms,
โˆ™ ๐”™๐›ผ โŠจ axiom of choice,
โˆ™ ๐”™๐›ผ โŠจ axiom of regularity.
PROOF
Let ๐›ผ be an ordinal and ๐ผ be an ๐–ฒ๐–ณ-interpretation of ๐”™๐›ผ . We check that the
extensionality axiom (5.1.4) holds in the model and leave the other parts of the
proof to Exercise 25. Take ๐ด, ๐ต ∈ ๐‘‰๐›ผ . Assume that
๐”™๐›ผ โŠจ ∀๐‘ข(๐‘ข ∈ ๐‘ฅ ↔ ๐‘ข ∈ ๐‘ฆ) [(๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ ].
That is,
๐ด ๐ต ๐‘š
๐‘š
for all ๐‘š ∈ ๐‘‰๐›ผ , ๐”™๐›ผ โŠจ ๐‘ข ∈ ๐‘ฅ [((๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ )๐‘ข ] if and only if ๐”™๐›ผ โŠจ ๐‘ข ∈ ๐‘ฆ [((๐ผ๐‘ฅ )๐‘ฆ )๐‘ข ].
We want to show that ๐ด = ๐ต, so let ๐‘Ž ∈ ๐ด. Since ๐‘‰๐›ผ is transitive (Lemma 7.5.14),
๐‘Ž ∈ ๐‘‰๐›ผ , which implies that
๐‘Ž
๐”™๐›ผ โŠจ ๐‘ข ∈ ๐‘ฅ [((๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ )๐‘ข ].
Therefore,
๐‘Ž
๐”™๐›ผ โŠจ ๐‘ข ∈ ๐‘ฆ [((๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ )๐‘ข ],
which implies that ๐‘Ž ∈ ๐ต. This proves that ๐ด ⊆ ๐ต. A similar proof shows that
๐ต ⊆ ๐ด. Thus, ๐ด = ๐ต, from which follows that
๐”™๐›ผ โŠจ ๐‘ฅ = ๐‘ฆ [(๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ ].
Therefore,
๐ด ๐ต
if ๐”™๐›ผ โŠจ ∀๐‘ข(๐‘ข ∈ ๐‘ฅ ↔ ๐‘ข ∈ ๐‘ฆ) [(๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ ], then ๐”™๐›ผ โŠจ ๐‘ฅ = ๐‘ฆ [(๐ผ๐‘ฅ )๐‘ฆ ].
In other words,
๐”™๐›ผ โŠจ ∀๐‘ข(๐‘ข ∈ ๐‘ฅ ↔ ๐‘ข ∈ ๐‘ฆ) → ๐‘ฅ = ๐‘ฆ [(๐ผ๐‘ฅ๐ด )๐ต
๐‘ฆ ].
Since ๐ด and ๐ต were arbitrarily chosen,
๐”™๐›ผ โŠจ ∀๐‘ฅ∀๐‘ฆ(∀๐‘ข[๐‘ข ∈ ๐‘ฅ ↔ ๐‘ข ∈ ๐‘ฆ] → ๐‘ฅ = ๐‘ฆ).
Again, using the ๐–ฒ๐–ณ-structure ๐”™3 (7.34), we see that the result of pairing two arbitrarily chosen elements of ๐‘‰3 into a single set might not be an element of ๐‘‰3 . For
example,
{{∅}, {{∅}}} ∉ ๐‘‰3 .
422
Chapter 7 MODELS
The same is true regarding power sets. For example,
P({{∅}}) = {∅, {{∅}}} ∉ ๐‘‰3 .
However, both of these sets are elements of ๐‘‰๐œ” , which suggests that for the pairing
(5.1.5) and power set (5.1.7) axioms to hold in stage ๐‘‰๐›ผ of the von Neumann hierarchy,
๐›ผ needs to be a limit ordinal.
THEOREM 7.5.21
If ๐›ผ is a limit ordinal,
โˆ™ ๐”™๐›ผ โŠจ pairing axiom,
โˆ™ ๐”™๐›ผ โŠจ power set axiom.
PROOF
Let ๐›ผ be a limit ordinal. Take ๐‘Ž, ๐‘ ∈ ๐‘‰๐›ผ . Then, there exists ๐›ฝ1 , ๐›ฝ2 ∈ ๐›ผ such that
๐‘Ž ∈ ๐‘‰๐›ฝ1 and ๐‘ ∈ ๐‘‰๐›ฝ2 . Without loss of generality, we can assume that ๐›ฝ1 ⊆ ๐›ฝ2 ,
which implies that ๐‘Ž ∈ ๐‘‰๐›ฝ2 by Lemma 7.5.14. Therefore, {๐‘Ž, ๐‘} ⊆ ๐‘‰๐›ฝ2 , so
{๐‘Ž, ๐‘} ∈ P(๐‘‰๐›ฝ2 ) = ๐‘‰๐›ฝ + ⊆ ๐‘‰๐›ผ .
2
Hence,
๐”™๐›ผ โŠจ ∀๐‘ข ∀๐‘ฃ ∃๐‘ฅ ∀๐‘ค (๐‘ค ∈ ๐‘ฅ ↔ ๐‘ค = ๐‘ข ∨ ๐‘ค = ๐‘ฃ).
The proof of the second part of the theorem is left to Exercise 26.
Certainly, the empty set will be an element of ๐‘‰๐›ผ provided that ๐›ผ is not empty, so
the proof of the next theorem is left to Exercise 27. In addition, ๐‘‰๐›ผ needs to contain ๐œ”
to satisfy the infinity axiom.
THEOREM 7.5.22
If ๐›ผ is a nonempty ordinal, ๐”™๐›ผ โŠจ empty set axiom.
THEOREM 7.5.23
If ๐›ผ is an ordinal such that ๐œ” ∈ ๐›ผ, then ๐”™๐›ผ โŠจ infinity axiom.
PROOF
Since Lemma 7.5.15 implies that ๐œ” ∈ ๐‘‰๐œ”+ , by Lemma 7.5.14, we have that
๐”™๐›ผ โŠจ ∃๐‘ฅ({ } ∈ ๐‘ฅ ∧ ∀๐‘ข[๐‘ข ∈ ๐‘ฅ → ∃๐‘ฆ(๐‘ฆ ∈ ๐‘ฅ ∧ ๐‘ข ∈ ๐‘ฆ ∧ ∀๐‘ฃ[๐‘ฃ ∈ ๐‘ข → ๐‘ฃ ∈ ๐‘ฆ])]).
Since ๐œ” ⋅ 2 = ๐œ” + ๐œ” is a limit ordinal greater than ๐œ”, we conclude the following.
THEOREM 7.5.24
๐”™๐œ”⋅2 โŠจ ๐—ญ.
The ordinal ๐œ” + ๐œ” requires one of the replacement axioms (5.1.9) to prove its existence
(page 317). This means that the proof of the consistency of ๐—ญ relies on axioms not in
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
423
๐—ญ, so let us continue to examine the von Neumann hierarchy for a model of ๐—ญ๐—™๐—–. It
will not be ๐”™๐œ”⋅2 because not all replacement axioms are true in ๐”™๐œ”⋅2 . This is because
๐”™๐œ”⋅2 does not satisfy Theorem 6.1.19. To use a stage of the von Neumann hierarchy to
serve as a model for ๐—ญ๐—™๐—–, we need a strongly inaccessible cardinal (Definition 6.5.11),
which is the next theorem. We state it without proof.
THEOREM 7.5.25
๐”™๐œ… โŠจ ๐—ญ๐—™๐—– if and only if ๐œ… is a strongly inaccessible cardinal.
However, the sentence
there exists a strongly inaccessible cardinal
(7.35)
cannot be proved or disproved using ๐—ญ๐—™๐—–. This means that (7.35) is independent of
the axioms of set theory. It also means that we are ready for the next definition. Do not
confuse it with the notion of a complete logic (Definition 7.4.8).
DEFINITION 7.5.26
An ๐–ฒ-theory ๐’ฏ is complete if ๐’ฏ โŠข ๐‘ or ๐’ฏ โŠข ¬๐‘ for all ๐–ฒ-sentences ๐‘, else ๐’ฏ is
incomplete.
The definition means that ๐—ญ๐—™๐—– is not complete. It turns out that the underlying issue
is that ๐—ญ๐—™๐—– satisfies the Peano axioms (7.5.3). The ability to do basic arithmetic guarantees that there exists a sentence that is independent of ๐—ญ๐—™๐—–. This result generalizes
to the incompleteness theorems due to Gödel (1931).
THEOREM 7.5.27 [Gödel’s First Incompleteness Theorem]
If the Peano axioms are provable from a consistent theory, the theory is incomplete.
Since ๐—ญ๐—™๐—– is assumed to be consistent and it can be used to deduce the Peano axioms,
we conclude that ๐—ญ๐—™๐—– is incomplete. Now suppose that instead of assuming the consistency of ๐—ญ๐—™๐—–, we try to prove that ๐—ญ๐—™๐—– is consistent using its own axioms. Gödel’s
next theorem proves that we cannot do this, except under one condition.
THEOREM 7.5.28 [Gödel’s Second Incompleteness Theorem]
If a theory proves the Peano axioms and its own consistency, the theory is inconsistent.
Therefore, if ๐—ญ๐—™๐—– could be used to prove (7.35), ๐—ญ๐—™๐—– would prove that it has a model,
which would imply that ๐—ญ๐—™๐—– is consistent by Henkin’s theorem (7.4.19). Since we
believe that ๐—ญ๐—™๐—– is consistent, we conclude that ๐—ญ๐—™๐—– cannot prove the existence of a
strongly inaccessible cardinal. If ๐—ญ๐—™๐—– was extended with an axiom that would allow
such a proof, there would be another issue with the extension that would prevent it from
proving its own consistency, provided that the new theory was consistent.
424
Chapter 7 MODELS
The statement of the first incompleteness theorem does not explicitly give a mathematical statement that cannot be proved. It was left to later mathematicians to find
some. For example, when combined with Gödel’s proof of the relative consistency of
๐–ข๐–ง, the proof of Paul Cohen (1963) of
๐–ข๐—ˆ๐—‡(๐—ญ๐—™๐—–) → ๐–ข๐—ˆ๐—‡(๐—ญ๐—™๐—– + ¬๐–ข๐–ง),
shows that both ๐–ข๐–ง and its negation cannot be proved from ๐—ญ๐—™๐—–. This means that the
continuum hypothesis is independent of ๐—ญ๐—™๐—–. In general, to prove that a sentence ๐‘ is
independent of a theory ๐’ฏ , do two things.
โˆ™ Find a model ๐”„ such that ๐”„ โŠจ ๐’ฏ ∪ {๐‘}. Then, ๐’ฏ ฬธโŠจ ¬๐‘ by Definition 7.1.20.
This implies that ๐’ฏ โŠฌ ¬๐‘ by Corollary 7.4.23.
โˆ™ Find another model ๐”… such that ๐”… โŠจ ๐’ฏ ∪ {¬๐‘}. From this conclude that ๐’ฏ ฬธโŠจ ๐‘,
which implies ๐’ฏ โŠฌ ๐‘.
Using this strategy, other statements have been discovered to be independent of ๐—ญ๐—™๐—–.
Some of these are technical set theoretic statements such as Martin’s axiom (Martin and
Solovay 1970) or the diamond principle (Jensen 1972). There are independent statements in other branches of mathematics as well. An example from group theory is the
independence of the Whitehead problem (Shelah 1974), named after the mathematician
John H. C. Whitehead (1950).
Exercises
1. Write an ๐– ๐–ฑ′ -formula equivalent to the given English sentences.
(a) ๐‘ฅ is an even number.
(b) ๐‘ฅ is an odd number.
(c) ๐‘ฅ is a prime.
(d) ๐‘ฅ divides ๐‘ฆ.
2. Prove that the order of Definition 5.2.10 proves the given ๐– ๐–ฑ′ -sentences.
(a) ∀๐‘ฅ(¬ ๐‘ฅ < 0)
(b) ∀๐‘ฅ∀๐‘ฆ[๐‘ฅ < ๐‘†๐‘ฆ ↔ (๐‘ฅ < ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ)]
(c) ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ < ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ ∨ ๐‘ฆ < ๐‘ฅ).
3. Prove that the given ๐– ๐–ฑ′ -formulas can be proved from Axioms 7.5.3.
(a) ∀๐‘ฅ(0 < ๐‘ฅ ∨ ๐‘ฅ = 0)
(b) ∀๐‘ฅ[๐‘ฅ = 0 ∨ ∃๐‘ฆ(๐‘ฅ = ๐‘†๐‘ฆ)]
(c) ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⋅ ๐‘ฆ = 0 → ๐‘ฅ = 0 ∨ ๐‘ฆ = 0)
(d) ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ < ๐‘ฆ ↔ ๐‘†๐‘ฅ < ๐‘†๐‘ข)
(e) ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ < ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ ∨ ๐‘ฆ < ๐‘ฅ)
4. Prove ๐—ฃ๐—” โŠจ (๐‘ฅ + 2) ⋅ (๐‘ฅ + 3) = (๐‘ฅ ⋅ ๐‘ฅ + 5 ⋅ ๐‘ฅ) + 6, where 2, 3, 5, and 6 are understood
to mean the appropriate successors of 0.
Section 7.5 MODELS OF DIFFERENT CARDINALITIES
425
5. Find examples of the following if possible.
(a) An ๐–ฒ-theory ๐’ฏ such that every two finite models of ๐’ฏ are isomorphic, but
there exists two models of ๐’ฏ that are infinite and not isomorphic.
(b) An ๐–ฒ-theory ๐’ฏ such that every two countable models of ๐’ฏ are isomorphic,
but there exists two models of ๐’ฏ that are uncountable and not isomorphic.
6. Let ๐’ฏ be an ๐–ฒ-theory and ๐‘ an ๐–ฒ-sentence where ๐’ฏ โŠข ๐‘. Prove that there exists a
finite ๐’ฐ ⊆ ๐’ฏ such that ๐’ฐ โŠข ๐‘.
7. Let ๐’ฏ be an ๐–ฒ-theory and ๐‘ an ๐–ฒ-sentence such that ๐’ฏ โŠจ ๐‘ and ๐‘ โŠจ ๐’ฏ . Show that
there exists a finite subset ๐’ฐ of ๐’ฏ such that ๐’ฐ โŠจ ๐’ฏ .
8. Prove that the following are equivalent for an ๐–ฒ-theory ๐’ฏ .
โˆ™ ๐’ฏ is consistent.
โˆ™ ๐’ฏ has a model.
โˆ™ ๐’ฏ has a countable model.
โˆ™ Every finite subset of ๐’ฏ is consistent.
โˆ™ Every finite subset of ๐’ฏ has a model.
โˆ™ Every finite subset of ๐’ฏ has a countable model.
9. Finish the proof of Theorem 7.5.5.
10. Prove that the cancellation law holds for multiplication in Peano arithmetic. See
Exercise 7.5.7.
11. If possible, find an ๐– ๐–ฑ′ -sentence ๐‘ ∈ ๐–ณ๐—(๐”“′ ) that is not a consequence of ๐—ฃ๐—”
(Exercise 7.4.6).
12. Demonstrate that there is no finite model of ๐—ฃ or ๐—ฃ๐—”.
13. Prove that there exists a model of ๐—ฃ that is not isomorphic to ๐”“.
14. Prove that there is a countable nonstandard model of Peano arithmetic.
15. The axioms for an ordered field are the ring axioms (7.1.36) plus the following
๐–ฎ๐–ฅ-sentences:
โˆ™ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฆ ⊗ ๐‘ฅ),
โˆ™ ∃๐‘ฅ∀๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = ๐‘ฆ),
โˆ™ ∀๐‘ฅ[¬๐‘ฅ = 0 → ∃๐‘ฆ(๐‘ฅ ⊗ ๐‘ฆ = 1)],
โˆ™ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ < ๐‘ฆ ∨ ๐‘ฅ = ๐‘ฆ ∨ ๐‘ฆ < ๐‘ฅ).
Find a model for these axioms.
16. Show that there exists a model of the axioms for an ordered field (Exercise 15)
such that ๐œ” is a subset of the domain of the model and there exists ๐‘š in the domain so
that ๐‘› < ๐‘š for every natural number ๐‘›. In addition, prove that there are infinitely many
such ๐‘š. What does 1โˆ•๐‘š look like?
17. Let ℜ = (โ„, 0, +, ⋅, <) be a model of the axioms for an ordered field. Prove that
there is a countable model of ๐–ณ๐—(ℜ). What is the significance of this model?
426
Chapter 7 MODELS
18. Expand ๐–ฎ๐–ฅ to ๐–ฎ๐–ฅ ∪ {๐ธ}, where ๐ธ is a binary function symbol. Find a model of
the axioms for an ordered field including the following (๐–ฎ๐–ฅ ∪ {๐ธ})-sentences:
โˆ™ ∀๐‘ฅ(๐‘ฅ๐ธ0 = ๐‘†0)
โˆ™ ∀๐‘ฅ∀๐‘ฆ(๐‘ฅ๐ธ(๐‘†๐‘ฆ) = (๐‘ฅ๐ธ๐‘ฆ)๐‘ฅ
19. Let ๐’ฏ be a theory such that for every ๐‘› ∈ ๐œ”, there exists ๐‘š ∈ ๐œ” such that ๐‘š > ๐‘›
and ๐’ฏ has a model of power ๐‘š. Prove that ๐’ฏ has an infinite model.
20. This hierarchy is due to Zermelo. Let ๐‘‰0 be a set of atoms.โ‹ƒFor every ordinal ๐›ผ,
define ๐‘‰๐›ผ+ = ๐‘‰๐›ผ ∪ P(๐‘‰๐›ผ ), and if ๐›ผ is a limit ordinal, define ๐‘‰๐›ผ = ๐›ฝ<๐›ผ ๐‘‰๐›ฝ . Find ๐‘‰1 and
๐‘‰2 assuming the given sets of atoms.
(a) ๐‘‰0 = ∅
(b) ๐‘‰0 = {๐‘Ž}
(c) ๐‘‰0 = {0, 1, 2, 3}
21. Prove Lemma 7.5.17.
22. Prove that ๐‘‰๐›ผ is a set for every ordinal ๐›ผ.
23. Using Exercise 6.3.22, show that |๐‘‰๐œ”+๐›ผ | = i๐›ผ for every ordinal ๐›ผ.
24. Let ๐ด be a set. Prove that TC(๐ด) is a transitive set.
25. Prove the remaining parts of Theorem 7.5.20.
26. Let ๐›ผ be a limit ordinal. Prove that ๐”™๐›ผ โŠจ ∀๐‘ฅ∃๐‘ฆ∀๐‘ข(๐‘ข ∈ ๐‘ฆ ↔ ∀๐‘ฃ[๐‘ฃ ∈ ๐‘ข → ๐‘ฃ ∈ ๐‘ฅ]).
27. Prove that ๐”™๐›ผ โŠจ ∃๐‘ฅ∀๐‘ฆ¬(๐‘ฆ ∈ ๐‘ฅ) if ๐›ผ ≠ ∅.
28. Let ๐”„ be an ๐–ฒ-structure. Prove that ๐–ณ๐—(๐”„) is complete.
29. Let ๐œ… be a cardinal and {๐’ฏ๐›ผ : ๐›ผ ∈ ๐œ…} be a chain of complete ๐–ฒ-theories such that
โ‹ƒ
for all ๐›พ ∈ ๐›ฟ ∈ ๐œ… we have that ๐’ฏ๐›พ ⊆ ๐’ฏ๐›ฟ . Show that ๐›ผ∈๐œ… ๐’ฏ๐›ผ is complete.
30. Let ๐’ฎ ⊆ ๐’ฏ be ๐–ฒ-theories. Prove or show false: If ๐’ฏ is complete, then ๐’ฎ is
complete.
APPENDIX
ALPHABETS
Greek Alphabet
Upper
A
B
Γ
Δ
E
Z
H
Θ
I
K
Λ
M
Lower
๐›ผ
๐›ฝ
๐›พ
๐›ฟ
๐œ–
๐œ
๐œ‚
๐œƒ
๐œ„
๐œ…
๐œ†
๐œ‡
Name
alpha
beta
gamma
delta
epsilon
zeta
eta
theta
iota
kappa
lambda
mu
Upper
N
Ξ
O
Π
P
Σ
T
Υ
Φ
X
Ψ
Ω
Lower
๐œˆ
๐œ‰
๐‘œ
๐œ‹
๐œŒ
๐œŽ
๐œ
๐œ
๐œ‘
๐œ’
๐œ“
๐œ”
Name
nu
xi
omicron
pi
rho
sigma
tau
upsilon
phi
chi
psi
omega
427
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
428
ALPHABETS
English Alphabet in the Fraktur Font
Upper
๐”„
๐”…
โ„ญ
๐”‡
๐”ˆ
๐”‰
๐”Š
โ„Œ
ℑ
๐”
๐”Ž
๐”
๐”
Lower
๐”ž
๐”Ÿ
๐” 
๐”ก
๐”ข
๐”ฃ
๐”ค
๐”ฅ
๐”ฆ
๐”ง
๐”จ
๐”ฉ
๐”ช
Name
A
B
C
D
E
F
G
H
I
J
K
L
M
Upper
๐”‘
๐”’
๐”“
๐””
ℜ
๐”–
๐”—
๐”˜
๐”™
๐”š
๐”›
๐”œ
โ„จ
Lower
๐”ซ
๐”ฌ
๐”ญ
๐”ฎ
๐”ฏ
๐”ฐ
๐”ฑ
๐”ฒ
๐”ณ
๐”ด
๐”ต
๐”ถ
๐”ท
Name
N
O
P
Q
R
S
T
U
V
W
X
Y
Z
REFERENCES
Aristotle (1984). The Complete Works of Aristotle, ed. J. Barnes. Princeton, NJ: Princeton
University Press.
Boole, G. (1847). The Mathematical Analysis of Logic : Being an Essay Towards a Calculus of
Deductive Reasoning. Cambridge, UK: Macmillan, Barclay, & Macmillan.
Boole, G. (1854). An Investigation of the Laws of Thought, on Which Are Founded the Mathematical Theories of Logic and Probabilities. London: Walton and Maberly.
Boyer, C. B. and U. C. Merzbach (1991). A History of Mathematics (2nd ed.). New York: John
Wiley & Sons, Inc.
Burali-Forti, C. (1897). Una questione sui numeri transfiniti. Rendiconti del Circolo Matematico
di Palermo 11(1), 154–164.
Cantor, G. (1874). Ueber eine Eigenschaft des Inbegriffs aller reellen algebraischen Zahlen.
Journal Fur Die Reine Und Angewandte Mathematik 77, 258–262.
Cantor, G. (1888). Mitteilungen zur Lehre vom Transfiniten. Zeitschrift für Philosophie und
philosophische Kritik 91, 81–125.
Cantor, G. (1891). Über eine elementare Frage def Mannigfaltigkeitslehre. Jahresbericht der
Deutschen Mathematiker-Vereinigung 1, 75–78.
Cantor, G. (1932). Gesammelte Abhandlungen mathematischen und philosophischen Inhalts.
Berlin: Springer-Verlag.
429
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
430
REFERENCES
Chang, C. C. and H. J. Keisler (1990). Model Theory (3rd ed.). Studies in Logic and the Foundations of Mathematics. Amsterdam: North Holland.
Church, A. (1956). Introduction to Mathematical Logic. Princeton, NJ: Princeton University
Press.
Ciesielski, K. (1997). Set Theory for the Working Mathematician. London Mathematical Society
Student Texts. Cambridge, UK: Cambridge University Press.
Cohen, P. J. (1963). The independence of the continuum hypothesis. Proceedings of the National
Academy of Sciences of the United States of America 50(6), 1143–1148.
Copi, I. M. (1979). Symbolic Logic (5th ed.). New York: Macmillan Publishing.
Dauben, J. W. (1979). George Cantor: His Mathematics and Philosophy of the Infinite. Princeton, NJ: Princeton University Press.
De Morgan, A. (1847). Formal Logic: or, the Calculus of Inference, Necessary and Probable.
London: Taylor and Walton.
Dedekind, R. (1893). Was sind und was sollen die Zahlen? Braunschweig: F. Vieweg.
Dedekind, R. (1901). Essays on the Theory of Numbers, trans. W. W. Beman. Chicago: Open
Court.
Descartes, R. (1985). The Philosophical Writings of Descartes, eds. J. Cottingham, R. Stoothoff,
and D. Murdoch. Cambridge, UK: Cambridge University Press.
Doets, K. (1996). Basic Model Theory. Stanford: CSLI Publications.
Drake, F. R. (1974). Set Theory: An Introduction to Large Cardinals, Volume 76 of Studies in
Logic and Foundations of Mathematics. Amsterdam: North-Holland.
Ebbinghaus, H., J. Flum, and W. Thomas (1984). Mathematical Logic. Undergraduate Texts in
Mathematics. New York: Springer-Verlag.
Ebbinghaus, H. and V. Peckhaus (2007). Ernst Zermelo: An Approach to His Life and Work.
Berlin: Springer.
Eklof, P. C. (1976). Whitehead’s problem is undecidable. American Mathematical Monthly 83,
775–788.
Eklof, P. C. and A. H. Mekler (2002). Almost Free Modules: Set-theoretic Methods (revised ed.).
Amsterdam: North-Holland.
Enderton, H. B. (1977). Elements of Set Theory. San Diego: Academic Press.
Euclid (1925). The Elements, ed. T. Heath. Reprint, New York: Dover, 1956.
Ewald, W. B. (2007). From Kant to Hilbert: A Source Book in the Foundations of Mathematics.
Oxford: Oxford University Press.
Fibonacci and L. E. Sigler (2002). Fibonacci’s Liber Abaci: A Translation into Modern English of
Leonardo Pisano’s Book of Calculuation. Sources and Studies in the History of Mathematics
and Physical Sciences. New York: Springer.
Fraenkel, A. A. (1922). Zu den Grundlagen der Cantor-Zermeloschen Mengenlehre. Mathematische Annalen 86, 230–237.
Fraleigh, J. B. (1999). A First Course in Abstract Algebra (6th ed.). Reading, MA: AddisonWesley.
Frege, G. (1879). Begriffsschrift, Eine Der Arithmetischen Nachgebildete Formelsprache Des
Reinen Denkens. Halle a/S.
REFERENCES
431
Frege, G. (1884). Die Grundlagen Der Arithmetik: Eine Logisch Mathematische Untersuchung
Über Den Begriff Der Zahl. Breslau: W. Koebner.
Frege, G. (1893). Grundgesetze Der Arithmetik : Begriffsschriftlich Abgeleitet. Jena: H. Pohle.
Gödel, K. (1929). Über die Vollständigkeit des Logikkalküls. Ph. D. thesis, University of Vienna.
Gödel, K. (1930). Die Vollständigkeit der Axiome des logischen Functionenkalküls. Monatshefte für Mathematik und Physik 37, 349–360.
Gödel, K. (1931). Über formal unentscheidbare Sätze der Principia Mathematica und verwandter
Systeme, I. Monatshefte für Mathematik und Physik 38, 173–198.
Gödel, K. (1940). Consistency of the Continuum Hypothesis. Princeton, NJ: Princeton University
Press.
Halmos, P. R. (1960). Naive Set Theory. New York: Springer-Verlag.
Henkin, L. (1949). The Completeness of Formal Systems. Ph. D. thesis, Princeton University,
Princeton, NJ.
Herrlich, H. (2006). Axiom of Choice. Lecture Notes in Mathematics. Berlin: Springer.
Hilbert, D. (1899). Grundlagen der Geometrie. Leipzig: Teubner.
Hilbert, D. (1902). Mathematical problems, trans. M. F. W. Newson. Bulletin of the American
Mathematical Society 8, 437–479.
Hodges, W. (1993). Model Theory. Encyclopedia of Mathematics and its Applications. Cambridge, U.K.: Cambridge University Press.
Hofstadter, D. R. (1989). Gödel, Escher, Bach: an Eternal Golden Braid. Reprint, New York:
Vintage Books.
Jech, T. (1973). The Axiom of Choice, Volume 75 of Studies in Logic and the Foundations of
Mathematics. Amsterdam: North-Holland.
Jech, T. (2003). Set Theory: The Third Millennium Edition. Springer Monographs in Mathematics. Berlin: Springer.
Jensen, R. B. (1972). The fine structure of the constructible hierarchy. Annals of Mathematical
Logic 4(3), 229–308.
Kaye, R. (1991). Models of Peano Arithmetic, eds. D. S. Angus Macintyre, John Shepherdson.
Oxford, UK: Clarendon Press.
Kline, M. (1972). Mathematical Thought from Ancient to Modern Times. New York: Oxford
University Press.
Kneale, W. and M. Kneale (1964). The Development of Logic. Oxford, UK: Clarendon Press.
König, J. (1905). Zum Kontinuum-problem. Mathematische Annalen 60(2), 177–180.
König, J. (1906). Sur la théorie des ensembles. Comptes rendus hebdomadaires des séances de
l’Académie des sciences 143, 110–112.
Kunen, K. (1990). Set Theory: An Introduction to Independence Proofs, Volume 102 of Studies
in Logic and the Foundations of Mathematics. Amsterdam: North-Holland.
Kuratowski, K. (1921). Sur la notion de l’order dans la théorie des ensembles. Fundamenta
Mathematicae 2, 161–171.
Kuratowski, K. (1922). Sur l’opération a de l’analysis situs. Fundamenta Mathematicae 3, 182–
199.
432
REFERENCES
Leary, C. C. (2000). A Friendly Introduction to Mathematical Logic. Upper Saddle River, NJ:
Prentice Hall.
Leibniz, G. W. (1666). Dissertatio de arte combinatoria, in qua ex arithmeticae fundamentis
complicationum ac transpositionum doctrina novis praeceptis extruitur, & usus ambarum per
universum scientiarum orbem ostenditur; nova etiam artis meditandi, seu logicae inventionis
semina sparguntur. Lipsiae, apud Joh. Simon Fickium et Joh. Polycarp. Seuboldum, Literis
Spörelianis.
Levy, A. (1979). Basic Set Theory. Perspectives in Mathematical Logic. Berlin: Springer-Verlag.
Löwenheim, L. (1915). Über Möglichkeiten im Relativkalkül. Mathematische Annalen 76(4),
447–470.
ลukasiewicz, J. (1930). Elementy logiki matematycznej. Warsaw: s.n.
ลukasiewicz, J. (1951). Aristotle’s Syllogistic From the Standpoint of Modern Formal Logic.
Oxford, UK: Clarendon Press.
Mal’tsev, A. I. (1936). Untersuchungen aus dem Gebiete der mathematischen Logik. Matematicheskii Sbornik 1(43), 323–336.
Martin, D. A. and R. M. Solovay (1970). Internal Cohen extensions. Annals of Mathematical
Logic 2(2), 143–178.
Mirimanoff, D. (1917). Les antinomies de Russel et de Burali-Forti et le problème fondamental
de la théorie des ensembles. L’Enseignement Mathématique 19, 37–52.
Pascal, B. (1665). Traité du triangle arithmetique, avec quelques autres petits traitez sur la
mesme matrière. Paris: chez G. Desprez.
Peano, G. (1889). Arithmetices principia: nova methodo. Augustae Taurinorum [Torino]:
Fratres Bocca.
Reid, C. (1996). Hilbert. New York: Copernicus.
Rosen, K. H. (1993). Elementary Number Theory and Its Applications (3rd ed.). Reading, MA:
Addison-Wesley.
Rubin, H. and J. E. Rubin (1985). Equivalents of the Axiom of Choice, II, Volume 116 of Studies
in Logic and the Foundations of Mathematics. Amsterdam: North-Holland.
Rubin, J. E. (1973). The compactness theorem in mathematical logic. Mathematics Magazine 46(5), 261–265.
Shelah, S. (1974). Infinite abelian groups, Whitehead problem and some constructions. Isreal
Journal of Mathematics 18(3), 243–256.
Skolem, T. (1922). Einige Bemerkungen zur axiomatischen Begründung der Mengenlehre. In
Proceedings of the 5th Scandinavian Mathematicians’ Congress in Helsinki, pp. 217–32.
Suppes, P. (1972). Axiomatic Set Theory. New York: Dover Publications.
Tarski, A. (1935). Der wahrheitsbegriff in den formalisierten sprachen. Studia Philosophica 1,
261–405.
Tarski, A. (1983). Logic, Semantics, Metamathematics: Papers from 1923 to 1938, trans. J. H.
Woodger. Indianapolis: Hackett Publishing Company.
Tarski, A. and R. L. Vaught (1957). Arithmetical extensions of relational systems. Compositio
Mathematica 13, 81–102.
van Dalen, D. (1994). Logic and Structure (3rd ed.). Berlin: Springer-Verlag.
REFERENCES
433
van Dalen, D., H. C. Doets, and H. de Swart (1978). Sets: Naive, Axiomatic and Applied, Volume
106 of International Series in Pure and Applied Mathematics. Oxford, UK: Pergamon Press.
Van Heijenoort, J. (1971). Frege and Gödel: Two Fundamental Texts in Mathematical Logic.
Cambridge MA: Harvard University Press.
Van Heijenoort, J. (1977). From Frege to Gödel: A Source Book in Mathematical Logic, 1879–
1931. Source Books in History of Sciences. Cambridge, MA: Harvard University Press.
Venn, J. (1894). Symbolic Logic (2nd ed.). London: Macmillan.
von Neumann, J. (1923). Zur Einführung der transfiniten Zahlen. Acta literarum ac scientiarum Regiae Universitatis Hungaricae Francisco-Josephinae, Sectio scientiarum mathematicarum 1, 199–208.
von Neumann, J. (1928). Über die Definition durch transfinite Induktion und verwandte Fragen
der allgemeinen Mengenlehre. Mathematische Annalen 99, 373–391.
von Neumann, J. (1929). Über eine Widerspruchsfreiheitsfrage der axiomatischen Mengenlehre.
Journal für die reine und angewandte Mathematik 160, 227–241.
Whitehead, A. N. and B. Russell (1910). Principia Mathematica (2nd ed.). Cambridge, UK:
Cambridge University Press.
Whitehead, J. H. C. (1950). Simple homotopy types. American Journal of Mathematics 72(1),
1–57.
Wussing, H. (1984). The Genesis of the Abstract Group Concept, ed. H. Grant, trans. A. Shenitzer. Cambridge, MA: MIT Press.
Zermelo, E. (1908). Untersuchungen über die grundlagen der mengenlehre. Mathematische
Annalen 65, 261–281.
Zermelo, E. (1930). Über Grenzzahlen und Mengenbereiche. Neue Untersuchungen über die
Grundlagen der Mengenlehre. Fundamenta mathematicae 16, 29–47.
Zorn, M. (1935). A remark on method in transfinite algebra. Bulletin of the American Mathematical Society 41, 667–670.
INDEX
Abel, Niels, 342
Abelian group, 342
Absolute value, 116
Abstraction, 122
Addition, inference rule, 25
Addition, matrix, 355
Additive identity, 199, 412
Additive inverse, 199
Aleph, 313
Algebraic, 315
Alphabet, 5
first-order, 68
second-order, 73
And, 4
Antecedent, 4
Antichain, 183
Antisymmetric, 177
Arbitrary, 88
Argument form, 21
Assignment, 7
Associative, 34, 139, 199, 412
Assumption, 46
Asymmetric, 177
Atom, 3, 6
Automorphism, 380
Axiom scheme, 228
Axiom(s), 24
choice, 231, 235
empty set, 227
equality, 226
extensionality, 227
foundation, 234
Frege–ลukasiewicz, 24
group, 340
paring, 228
power set, 228
regularity, 234
replacement, 230
ring, 353
separation, 228
subset, 229
union, 228
Zermelo, 231
Axiomatizable, finitely, 409
Basis case, 258
Bernstein, Felix, 301
Beth, 316
Biconditional, 5
Biconditional proof, 107
435
A First Course in Mathematical Logic and Set Theory, First Edition. Michael L. O’Leary.
© 2016 John Wiley & Sons, Inc. Published 2016 by John Wiley & Sons, Inc.
436
INDEX
Bijection, 211
Binary, 68
function, 189
operation, 198
relation, 161
Binomial coefficient, 263
Binomial theorem, 263
Boole, George, 117
Bound
lower, 181
upper, 180
Bound occurrence, 77
Burali-Forti paradox, 297
Burali-Forti theorem, 292
Cancellation, 248, 256, 414
Candidate, 99
Cantor–Schröder–Bernstein theorem, 301
Cantor, Georg, 117, 226, 229, 303, 313
Cardinal, 307
large, 332
limit, 314
regular, 328
singular, 328
strongly inaccessible, 332
successor, 314
weakly inaccessible, 331
Cardinality, 308
Cartesian ๐‘›-space, 131
Cartesian plane, 130
Cartesian product, 130
Cases, 112
Chain, 182
elementary, 394
of structures, 372
Characteristic function, 304
Choice axiom, 231, 235
Choice function, 231
Class, 420
Class (relation), 171
Closed, 198
Closed interval, 120
Closed under deductions, 409
Codomain, 190
Cofinal, 328
Cofinality, 328
Cohen, Paul, 424
Coincidence, 349
Combinatorics, 260
Common divisor, 140
Commutative, 34, 139, 199, 412
Commutative ring, 354
Compactness theorem, 52, 415
Comparable, 181
Compatible, 183
Complement, 128
Complete, 56, 398
Complete theory, 423
Completeness theorem, 61, 407
Gödel’s, 407
Completeness, real numbers, 275
Complex number, 281
Composite number, 107
Composition, 163, 195
Compound, 4, 6
Concatenation, 178
Conclusion, 22, 26
Conditional, 4
Conditional proof, 45
Congruent, 170
Conjunct, 4
Conjunction, 4
Conjunction, inference rule, 25
Connective, 6
Consequence, 22, 346, 352
Consequent, 4
Consistency, relative, 407
Consistent, 52, 395
maximally, 53, 396
Constant symbol, 68
Constructive dilemma, 25
Contingency, 16
Continuous, 94
Continuum hypothesis, 313
generalized, 314
Contradiction, 16
Contradiction, proof by, 47
Contrapositive, 32, 103
Contrapositive law, 34
Converse, 32
Coordinates, 130
Coordinatewise, 356
Copy, 214
Corollary, 20
Corresponding occurrences, 77
Countable, 310
Counterexample, 102
Cyclic group, 363, 365
De Morgan, Augustus, 117
De Morgan’s laws, 35, 137, 139
Decreasing, 184, 203
Dedekind cut, 276
Dedekind, Richard, 229, 276
Deduce, 26
Deduction, 20
Deduction theorem, 42
Deductive logic, 2
Dense, 257
Denumerable, 310
INDEX
Descartes, René, 117
Destructive dilemma, 25
Diagonalization, 303
Diamond principle, 424
Direct existential proof, 99
Direct proof, 45
Discrete, 310
Disjoint, 127, 155
pairwise, 155
Disjunct, 4
Disjunction, 4
Disjunctive normal form, 38
Disjunctive syllogism, 25
Distributive, 34, 139, 246, 323, 412
left, 320
Divides, 98
Divisible, 98
Division algorithm, 185
polynomial, 110
Division ring, 357
Divisor, 98
common, 140
zero, 356
Domain, 162, 334
Dominate, 300
Double negation, 35
Downward closed, 274
Downward Löwenheim–Skolem theorem, 415
Element, 117
Elementary chain, 394
Elementary equivalent, 387
Elementary extension, 389
Elementary substructure, 389
Embedding, 214
Empty set, 118, 141, 227
Empty set axiom, 227
Empty string, 6
Endpoint, 120
Equal, 118
Equality, 194
Equality axioms, 226
Equality symbol, 68
Equinumerous, 298
Equivalence class, 171
Equivalence relation, 169
induced, 175
Equivalence rule, 110
Equivalence, logical, 31, 348
Equivalent
elementary, 387
pairwise, 109
Euclid, 19, 107, 265
Euclid’s lemma, 265
Evaluation map, 193
Even integer, 98
Excluded middle, 37
Exclusive or, 11
Existence, 104
Existential formula, 72
Existential generalization, 91
Existential instantiation, 91
Existential proof, 99, 106
Existential proposition, 65
Existential quantifier, 65
Expansion, 352
Exponentiation, 249, 321, 322
Exportation, 35
Extension, 197, 361
elementary, 389
Extensionality axiom, 227
Factor, 98, 110
Factorial, 242
False, 3
Family of sets, 148
Fibonacci, 268
Fibonacci number, 269
Fibonacci sequence, 269
generalized, 273
Field, 357
ordered, 69, 425
Field theory, 69
Finite, 308
hereditarily, 417
Finitely axiomatizable, 409
First-order
alphabet, 68
formula, 73
language, 73
logic, 96
Formal proof, 26
Formation sequence, 7
Formula, 71, 72
first-order, 73
second-order, 73
Foundation axiom, 234
Fraenkel, Abraham, 229
Fraktur, 334, 428
Free occurrence, 77
Free variable, 78
Frege, Gottlob, 24, 117
Function, 189
bijection, 211
binary, 189
characteristic, 304
choice, 231
continuous, 94
decreasing, 203
embedding, 214
437
438
INDEX
evaluation, 193
greatest integer, 192
homomorphism, 375
identity, 191
inclusion, 203
increasing, 203
injection, 206
inverse, 204
invertible, 204
isomorphism, 382
one-to-one, 206
one-to-one correspondence, 211
onto, 208
order-preserving, 212
periodic, 203
projection, 210
real-valued, 193
surjection, 208
unary, 189
uniformly continuous, 94
zero, 376
Function equality, 194
Function notation, 190
Function symbol, 68
Fundamental homomorphism theorem, 383
Fundamental theorem of arithmetic, 271
Galois, Évariste, 342
General linear group, 345
Generalization, 85
existential, 91
universal, 88
Generalized continuum hypothesis, 314
Generalized Fibonacci sequence, 273
Generator, 363, 370
Gödel, Kurt, 399, 407, 414, 423
Gödel’s completeness theorem, 407
Gödel’s incompleteness theorems, 423
Golden ratio, 271
Grammar, 6
Greatest common divisor, 140
Greatest element, 180
Greatest integer function, 192
Greatest lower bound, 181
Grounded, 419
Group, 342
abelian, 342
cyclic, 363
general linear, 345
Klein-4, 344
simple, 363
Group axioms, 340
Group theory, 69
Grouping symbol, 6
Half-open interval, 120
Hartogs’ function, 327
Hartogs’ theorem, 293
Hausdorff maximal principle, 237
Henkin, Leon, 399
Henkin’s theorem, 406
Hereditarily finite sets, 417
Hereditary set, 235
Hierarchy, von Neumann, 417
Hilbert, David, 226, 408
Hilbert’s problems, 408
Homomorphism, 375
group, 375
ring, 375
Hypothetical syllogism, 25
Ideal, 368
improper, 368
left, 368
maximal, 374
prime, 374
principal, 370
principal left, 370
proper, 368
right, 368
Idempotent laws, 139
Identity, 162, 199
additive, 199
multiplicative, 199
Identity map, 191
Image, 190, 208, 216
Implication, 4
Improper ideal, 368
Improper subgroup, 363
Improper subring, 366
Improper subset, 136
Inaccessible cardinal, 331
Inclusion map, 203
Inclusive or, 11
Incomparable, 181
Incompatible, 183
Incomplete theory, 423
Incompleteness theorems, 423
Inconsistent, 52, 395
Increasing, 184, 203
Independent, 408
Index, 148
Index set, 148
Indexed, 149
Indirect existential proof, 106
Indirect proof, 47
Induced equivalence relation, 175
Induced partition, 174
Induction
mathematical, 257
INDEX
on formulas, 336
on propositional forms, 59
on terms, 335
strong, 268
transfinite, 283, 291
Induction hypothesis, 59, 258
Induction step, 258
Inductive logic, 1
Inductive set, 238
Infer, 24
Inference, 24
Inference rule, 25
Infinite, 308, 344
Infinity axiom, 227
Infinity symbol, 120
Initial number, 327
Initial segment, 274
proper, 274
Injection, 206
Instantiation, 85
existential, 91
universal, 87
Integers, 250
Integral domain, 356
International Congress of Mathematicians, 408
Interpretation, 335
Intersection, 126, 152
Interval, 120
Interval notation, 120
Introduction, 97
Invalid
semantically, 21
syntactically, 21
Inverse, 166, 199
additive, 199
function, 204
image, 216
left, 215
multiplicative, 199
right, 216
Inverse image, 216
Inverse relation, 166
Inverse statement, 37
Invertible function, 204
Invertible matrix, 345
Irrational number, 128, 276
Irreflexive, 177
Isomorphic, 212, 344, 380, 382
Isomorphism, 212, 380
group, 382
ring, 382
Kernel, 379
Klein-4 group, 344, 363
König, Julius, 301, 330
König’s theorem, 330
Kuratowski, Kazimierz, 130, 232, 237
Language, 73
first order, 73
Large cardinal, 332
Law of noncontradiction, 37
Law of the excluded middle, 37
Least element, 180
Least upper bound, 180
Left distributive, 320
Left ideal, 368
Left inverse, 215
Left zero divisor, 356
Leibniz, Gottfried, 117
Lemma, 20
Leonardo of Pisa, 268
Lexicographical order, 187
Liber abaci, 268
Limit cardinal, 314
Limit ordinal, 290
Lindenbaum’s theorem, 397
Linear combination, 186
Linear order, 182
Linearly ordered set, 182
Logic, 1
deductive, 2
first-order, 96
inductive, 1
mathematical, 2
propositional, 20
second-order, 96
Logic symbol, 67
Logical implication, 21
Logical system, 20
Logically equivalent, 31, 348
Löwenheim, Leopold, 415
Löwenheim’s theorem, 416
Löwenheim–Skolem theorem
downward, 415
upward, 416
Lower bound, 181
greatest, 181
ลukasiewicz, Jan, 24, 71
Mal’tsev, Anatolij, 414
Map, 190
Martin’s axiom, 424
Material equivalence, 13, 35
Material implication, 12, 35
Mathematical induction, 257
strong, 268
Mathematical logic, 2
Mathematics, 1
Mathesis universalis, 117
439
440
INDEX
Matrix, 345
identity, 345
invertible, 345
zero, 355
Matrix addition, 355
Matrix multiplication, 345
Maximal element, 181
Maximal ideal, 374
Maximally consistent, 53, 396
Meaning, 334
Metatheory, 96
Minimal element, 181
Mirimanoff, Dmitry, 229, 234
Model, 338, 339
Modern square of opposition, 86
Modulo, 170, 172
Modus ponens, 25
Modus tolens, 25
Multiple, 98
Multiplication, matrix, 345
Multiplicative identity, 199, 412
Multiplicative inverse, 176, 199
Mutually exclusive, 127
๐‘-ary, 68, 161, 189
Name, 334
Natural number, 238
Necessary, 4, 109
Negated, 4
Negation, 4
Niece, 64, 161
Noncontradiction law, 37
Nonstandard model, 411
Norway, 342
Number, 256
standard, 411
Number theory, 69
Occurrence, 65
Odd integer, 98
One-to-one correspondence, 211
One-to-one function, 206
Onto function, 208
Open interval, 120
Operation, binary, 198
associative, 199
commutative, 199
Operation, set, 126
Or, 4
exclusive, 11
inclusive, 11
Order, 344
increasing, 184
lexicographical, 187
linear, 182
partial, 178
well, 183
Order isomorphism, 212
Order of connectives, 8, 9
Order of operations, 132
Order type, 212, 292
Order-preserving, 212
Ordered ๐‘›-tuple, 131
Ordered field, 69, 425
Ordered pair, 130
Ordinal, 286
limit, 290
successor, 290
Ordinal number, 286
Pairing axiom, 228
Pairwise disjoint, 155
Pairwise equivalent, 109
Paragraph proof, 97
Parsing tree, 6
Partial order, 178
Partially ordered set, 178
Particular, 88
Partition, 173
induced, 174
Pascal’s identity, 263
Peano arithmetic, 69, 411
Peano axioms, 410
Peano, Giuseppe, 410
Periodic, 203
Permutation, 260
Pigeonhole principle, 309
Polynomial division algorithm, 110
Poset, 178
Positive form, 86
Postulate, 20
Power, 415
Power set, 151
Power set axiom, 228
Pre-image, 190
Predecessor, 237
Predicate, 64, 65
Prefix notation, 71
Premise, 22, 26
Prenex normal form, 96
Preserves, 212, 375
Prime ideal, 374
Prime number, 107
Prime power decomposition, 272
Principal ideal, 370
left, 370
Principal ideal domain, 372
Projection, 210
Proof methods
biconditional, 107
INDEX
biconditional, short rule, 108
cases, 112
conditional, 45
contradiction, 47
counterexample, 102
direct, 45
disjunctions, 111
equivalence rule, 110
existence, 104
existential, direct, 99
existential, indirect, 106
indirect, 47
reductio ad absurdum, 47
relative consistency, 407
uniqueness, 104
universal, 98
Proof, formal, 26
Proof, paragraph, 97
Proof, two-column, 27
Proper ideal, 368
Proper initial segment, 274
Proper subgroup, 364
Proper subring, 366
Proper subset, 136
Proposition, 3
compound, 4
universal, 66
Proposition alphabet, 5
Propositional form, 6
Propositional logic, 20
Propositional variable, 6
Prove, 26
Pure set, 235
Purely relational, 337
Quantifier, 65, 66
Quantifier negation, 86
Quantifier symbol, 68
Quotient, 110, 185
Quotient set, 172
Range, 162, 208
Rational numbers, 254
Ray, 120
Real number, 276
Real-valued function, 193
Recursion, 242
transfinite, 294
Recursive definition, 6, 242
Reduct, 352
Reductio ad absurdum, 47
Reflexive, 169
Regular cardinal, 328
Regularity axiom, 234
Relation, 161
antisymmetric, 177
asymmetric, 177
binary, 161
inverse, 166
irreflexive, 177
reflexive, 169
symmetric, 169
transitive, 169
unary, 161
Relation on, 161
Relation symbol, 68
Relative consistency, 407
Remainder, 110, 185
Replacement axiom, 230
Replacement rule, 34
Restriction, 197
Right ideal, 368
Right inverse, 216
Right zero divisor, 356
Ring, 354
commutative, 354
division, 357
field, 357
integral domain, 356
principal ideal domain, 372
skew field, 357
with unity, 354
Ring axioms, 353
Ring theory, 69
Roster, 118
Roster method, 118
Russell’s paradox, 225
Russell, Bertrand, 225
Satisfiable, 340, 352
Satisfy, 65, 338, 339
Scheme, axiom, 228
Schröder, Ernst, 301
Scope, 77
Second-order
alphabet, 73
formula, 73
Second-order logic, 96
Selector, 231
Self-evident, 24
Semantically invalid, 21
Semantically valid, 21, 22
Semantics, 21
Sentence, 83
Separation axioms, 228
Set, 117
hereditary, 235
Set difference, 127
Set operation, 126
Set theory, 68
441
442
INDEX
Short rule of biconditional proof, 108
Signature, 334
Simple group, 363
Simplification, 25
Simultaneous substitution, 81
Singleton, 118
Singular cardinal, 328
Skew field, 357
Skolem’s theorem, 416
Skolem, Thoralf, 229, 234, 415
Sound, 56, 57, 398
Soundness theorem, 57, 398
Square of opposition, modern, 86
Standard model, 411
Standard number, 411
String, 5
Strong induction, 268
Strongly inaccessible, 332
Structure, 333
pureley relational, 337
Subgroup, 363
improper, 363
proper, 364
trivial, 363
Subject, 64
Subproof, 46
Subring, 366
improper, 366
proper, 366
trivial, 366
Subset, 135
Subset axioms, 229
Substitution, 64, 75, 78
simultaneous, 81
Substructure, 361
elementary, 389
Successor, 237
Successor cardinal, 314
Successor ordinal, 290
Sufficient, 4, 109
Surjection, 208
Symmetric, 169
Symmetric closure, 177
Symmetric difference, 236
Syntactically invalid, 21
Syntactically valid, 21, 27
Syntax, 23
Tarski, Alfred, 336, 415
Tarski–Vaught theorem, 390
Tautology, 16
Tautology rule, 35
Term, 70
Theorem, 20, 27
Theory, 395
Theory symbol, 68
Tower, 233
Transcendental, 315
Transfinite induction, 283, 291
Transfinite recursion, 294
Transitive, 169
Transitive closure, 419
Transitive set, 239
Trichotomy, 183, 288
weak, 179
Trichotomy law, 183
Trivial subgroup, 363
Trivial subring, 366
True, 3
Truth table, 11
Truth value, 3
Two-column proof, 27
Unary, 68
function, 189
relation, 161
Undecidable, 313
Undefined, 190
Uniformly continuous, 94
Union, 126, 151
Union axiom, 228
Unique, 104
Unique factorization theorem, 271
Unit, 357
Unity, 354
Universal formula, 72
Universal generalization, 88
Universal instantiation, 87
Universal proof, 98
Universal proposition, 66
Universal quantifier, 66
Universe, 97, 141
Upper bound, 180
least, 180
Upward Löwenheim–Skolem theorem, 416
Urelement, 417
Valid, 347
semantically, 21, 22
syntactically, 21, 27
Valuation, 10, 13
Valuation patterns, 15
Variable, 67
propositional, 6
Variable symbol, 67
Vaught, Robert, 336
Venn diagram, 127
Venn, John, 127
Vertical line test, 189
Von Neumann hierarchy, 417
INDEX
Von Neumann, John, 234, 238, 417
Weak-trichotomy law, 179
Weakly inaccessible, 331
Well-defined, 191
Well-ordered set, 183
Well-ordering theorem, 293
Whitehead, John H. C., 424
Whitehead problem, 424
Witness, 91
Witness set, 399
Write, 23
Zermelo axioms, 235
Zermelo’s axiom, 231
Zermelo’s theorem, 295
Zermelo, Ernst, 225, 234, 293, 417
Zermelo–Fraenkel axioms, 235
Zero, 102, 110
Zero divisor, 356
left, 356
right, 356
Zero map, 376
Zero matrix, 355
Zorn’s lemma, 232
Zorn, Max, 232
443
WILEY END USER LICENSE AGREEMENT
Go to www.wiley.com/go/eula to access Wiley’s ebook EULA.
Download