Logic, Mathematics, and Computer Science Modern Foundations With
Logic, Mathematics, and Computer Science Modern Foundations With
Logic, Mathematics,
and Computer
Science
Modern Foundations
with Practical Applications
Second Edition
www.pdfgrip.com
Yves Nievergelt
Logic, Mathematics,
and Computer Science
Modern Foundations with Practical
Applications
Second Edition
123
www.pdfgrip.com
Yves Nievergelt
Department of Mathematics
Eastern Washington University
Cheney, WA, USA
Cover art excerpted from triadic truth tables from Charles Sanders Peirce’s Logic Notebook.
Springer Science+Business Media LLC New York is part of Springer Science+Business Media (www.
springer.com)
www.pdfgrip.com
Preface
This second edition, entitled Logic, Mathematics, and Computer Science: Modern
Foundations with Practical Applications, has been adapted from Foundations of
Logic and Mathematics: Applications to Computer Science and Cryptography, ©
2002 by Birkhäuser, from which Chapters 1–5 have been retained but extensively
revised. Chapters 6 and 7 have been added.
This text discusses the foundations where logic, mathematics, and computer
science begin. The intended readership consists of undergraduate students majoring
in mathematics or computer science who must learn such foundations either for their
own interest or for further studies. For a motivated reader, there are no technical
prerequisites: you need not know any technical subject to start reading this text.
Although the text does not focus on the history and philosophy of the founda-
tions, the material cites copious references to the literature, where the reader may
find additional historical context. Consulting such references is neither suggested
nor necessary to study the theory or to work on the exercises, but individual citations
document the material by original sources, and all the citations together provide a
guide to the variations and chronological developments of logic, mathematics, and
computer science. For example, Chapter 1 traces the origin of Truth tables to Charles
Sanders Peirce’s unpublished 1909 Logic Notebook on philosophy and points out
their applications over one half of a century later to the design of computers for use
on Earth and on board the Apollo lunar spacecraft.
Along informal arguments, this text also shows the corresponding purely sym-
bolic manipulations of formulae, because they clarify the reasoning [11] and can
reveal hitherto hidden logical properties, such as the mutual independence of
different patterns of reasoning, or the impossibility of some proofs within some
logics:
As for algebra [of logic], the very idea of the art is that it presents formulae which can be
manipulated, and that by observing the effects of such manipulation we find properties not
to be otherwise discerned (Charles Sanders Peirce [104, p. 182]).
If professionals are unable to learn some topics by any means other than the pure
manipulation of symbols, then it would seem unfair to claim that all learning must be
intuitive and hide from students such purely manipulative but successful methods.
v
www.pdfgrip.com
vi Preface
The selection of topics also reflects major accomplishments from the twentieth
century: the foundation of all of mathematics, and later computer science, as well
as computer-assisted proofs of mathematical theorems, on a formal logic based on
only a few axioms, transformation rules, and postulates for set theory [47, 50, 54,
105, 139]. Also, while not written in formal logic, Nobel-Prize winning applications
to the social sciences rely on the same foundations, as shown in Chapter 7.
To introduce the foundations of logic, the provability theorem in Chapter 1
provides an algorithm to design proofs in propositional logic. Chapter 1 also
explains the concept of undecidability with multi-valued (“fuzzy”) logic and
presents a proof of unprovability. Chapter 2 introduces logical quantifiers.
A working knowledge of logical quantifiers is crucial for the study of basic concepts
in modern mathematical analysis and topology, such as the uniform convergence
of a sequence of equicontinuous functions. Continuing with the foundations of
mathematics, Chapter 3 presents a version of the Zermelo–Fraenkel set theory. At
the juncture of mathematics and computer science, Chapter 4 develops the concepts
of definition and proof by induction. Chapter 4 then uses induction with set theory
to define the integers and rational numbers and derive the associative, commutative,
and distributive laws, as well as algorithms, for their arithmetics. To give readers
some idea of topics at an intermediate level, Chapter 5 shows that in a well-formed
theory some paradoxes do not occur, while Chapter 6 completes the foundations of
set theory with the axiom of choice.
No extragalactic asteroid has yet been found with the universal laws of logic
engraved in it. Consequently, not just one logic but many different logics have
been invented. Different logics lead to different mathematics and different computer
sciences. However, the acid test for adopting a particular logic is its ability
to make predictions that are born by subsequent experiments. Formal logic is
thus a mathematical model of rational thought processes. In this aspect, logic,
mathematics, and computer science are experimental sciences. Only one logic has
passed all such tests, which is the one used throughout this text. Other logics
are outlined in Chapter 1 as a pedagogical device and to show some of their
shortcomings.
Acknowledgments
I thank Dr. Stephen P. Keeler at the Boeing Company for having corrected many
errors in the first edition. New and remaining errors are mine. I also thank Dr. Mary
Keeler for guiding me to Charles Sanders Peirce’s unpublished work on logic and
philosophy. I commend the editors at Birkhäuser and Springer, in particular, Ann
Kostant and Elizabeth Loew, for their hard work, patience, encouragements, and
positive attitude through the development and production of both texts.
Contents
vii
www.pdfgrip.com
viii Contents
Contents ix
x Contents
Contents xi
xii Contents
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381
www.pdfgrip.com
Chapter 1
Propositional Logic: Proofs from Axioms
and Inference Rules
1.1 Introduction
For the purpose of an introduction, the following example demonstrates how logic
can help in resolving practical issues in real life, how questions arise about the
validity of logical methods to reach conclusions, and eventually what thought
processes are acceptable or successful in explanations and predictions. Yet an
understanding of this example will not be necessary for any of the subsequent
material.
1.1 Example. The planetary status of Pluto has been debated in newspapers:
Is Pluto really a planet? Like all civil wars, this has even split families apart [67].
1.1 Introduction 3
The logical implication “if H, then C” (also worded as “H implies C”) is true by this
definition of planets. If the hypothesis H is also true, then the conclusion C is true,
too: C can be “detached from the implication” [80, p. 124].
With a logical implication (“if H, then C”) there are two other useful statements:
its converse (“if C, then H”) and its contraposition (“if not C, then not H”).
The hypothesis H can be tested by the contraposition “if not C, then not H”:
If a celestial object P is not bigger than every moon (if not C),
then P is not a planet (then not H).
In 1978 measurements revealed that Pluto was smaller than the Moon [69, p. 213];
consequently, Pluto would no longer be a planet, by the foregoing definition.
The definition “if H, then C” can also be tested in practice. For instance,
textbooks classify Mercury as a planet, but they also classify Ganymede as a moon
(of Jupiter), even though Mercury is smaller than Ganymede [69, p. 182 & 203].
Thus the statement “if Mercury is a planet, then Mercury is bigger than every moon”
is false. Therefore the foregoing definition “if H, then C” is false, and Pluto can
remain a planet. Logic has thus resolved the issue by revealing that the question
pertains not to the status of Pluto but to the definition of planets.
Other definitions and logical arguments have also been debated [73, 101]. Very
shortly thereafter, all existing definitions of planets were again put into question:
Scientifically, we are unable to define a planet in a sufficiently meaningful way such as to
include Pluto without including many other objects Œ: : : we are also unable to develop
a definition based on principles of astronomy and physics that excludes Pluto in any
nonarbitrary way [137, p. 219, summarizing chapter 14, p. 185–221].
The definition of planet will not be settled here, but some patterns of logical
reasoning will. For instance, the preceding arguments show that detachment and
contraposition are valid popular and scientific modes of reasoning; converse is not.
The preceding discussion contains one logical principle that has been so suc-
cessful that it remains widely accepted in theory and in practice: the law of
contraposition, which states that if an implication “if H, then C” holds, then its
contraposition “if not C, then not H” also holds.
The converse law of contraposition — which states that if the contraposition
“if not C, then not H” holds, then the implication “if H, then C” also holds — was
not used in the preceding discussion, and it is not a part of some logical systems.
In this example, the converse statement “if C, then H” (also worded as “C implies
H”) is false, because the Sun is bigger than every moon (C is true), but the Sun is
not a planet (H is false). Nevertheless, implications of the form “if H, then C” and
their converse “if C, then H” have been confused by professional scientists, so that
the difference between an implication and its converse bears being emphasized:
Leontovich explained to me why the paper could not be published in ZETP [“the main
Russian physics journal”]:
the paper claims that “A implies B” while every physicist knows examples showing
that B does not imply A; [...]
www.pdfgrip.com
An author, claiming that A implies B, must say whether the converse holds, otherwise the
reader who is not spoiled by the mathematical slang would interpret the claim as “A is
equivalent to B.” If mathematicians do not follow this rule, they are wrong [3, p. 619–620].
Yet such symbols may later denote various types of atomic formulae that apply
to such various fields as algebra, arithmetic, geometry, or set theory, and therein
lies the power of the pure propositional calculus. Furthermore, purely symbolic
manipulations of formulae can reveal hitherto hidden logical properties:
www.pdfgrip.com
As for algebra [of logic], the very idea of the art is that it presents formulae which can be
manipulated, and that by observing the effects of such manipulation we find properties not
to be otherwise discerned (Charles Sanders Peirce, [104, p. 182]).
Common instances of logical reasoning consist of sentences. For example, the first
axiom of Euclidean geometry is a sentence (paraphrased from [61, p. 3]):
For each pair of distinct points there exists exactly one line passing through both points.
Similarly, propositional logic starts with certain logical formulae called axioms.
The word “axiom” can mean “self-evident truth” but in the present context, which
focuses on patterns of reasoning, the word “axiom” means “initial” or “starting”
logical pattern [110, p. 55]. Different selections of axioms can lead to different
kinds of logic, but the present chapter focuses mainly on classical logic, which
has been successful for several millennia. Several choices for the initial axioms and
formulae lead to the same classical logic. Because the principal concepts of logic
consist of “negation” and “implication” several common choices of initial axioms
and formulae involve only the connectives for negation .:/ and implication .)/.
Also, to allow for applications in various areas, the pure propositional calculus
replaces the “atomic formulae” from algebra, arithmetic, geometry, or set theory
by general symbols called propositional variables or sentence symbols.
1.3 Definition (Well-formed formulae). Select two disjoint lists of symbols.
Every symbol from the first list of symbols, which may consist of one or more
letter(s) from a specified alphabet, P, Q, : : :, optionally with subscript(s) P[ , P[[ ; : : :,
superscript(s) P] , P]] , : : :, or “middlescript(s)” Pj, Pjj, : : :, is called a formulaic
letter. Such formulaic letters are not parts of the propositional calculus, but they
help in describing the following rules to define well-formed formulae.
Also, every symbol from the second list of symbols, which may consist of one
or more letter(s) from a specified alphabet, A, B, : : :, optionally with subscript(s)
A[ , A[[ , : : :, superscript(s) A] , A]] , : : :, or “middlescript(s)” Aj, Ajj, : : :, is called a
propositional variable or a sentence symbol [31, p. 17]. (Propositional variables
may later be replaced by atomic formulae specific to applications.)
Every propositional variable is a well-formed formula. For all well-formed
formulae P and Q, the following strings of symbols are also well-formed formulae:
By definition 1.3, a string of symbols such as .P/ ) .Q/ is not yet a well-
formed formula of the propositional calculus. It only becomes so after P and Q
have been replaced by propositional variables or well-formed formulae, for instance,
.A/ ) .B/. In this section, any capital letter may denote a propositional letter,
variable, or formula. In subsequent sections, however, the distinction may matter.
From definition 1.3, propositional letters or variables are “atoms” or “atomic”
propositional formulae in the sense that they are the simplest well-formed proposi-
tional formulae [72, p. 5], in contrast to more elaborate “composite” or “compound”
formulae, also called propositional forms, built from rules W1 and W2.
Several choices of well-formed propositional formulae can serve as axioms.
A system that remains concise and differentiates the roles of separate connectives
consists of the following three axioms. Subsection 1.3.10 explains their popularity
[18, 20, p. 119], [84, p. 49], [81, p. 31], [85, p. 165], [122, p. 165].
1.4 Definition (Jan Łukasiewicz’s axioms). A logical formula is an axiom of the
version of the classical propositional calculus considered here if and only if it is one
of the following three formulae, attributed to Łukasiewicz [62, p. 29], where P, Q,
and R may be any well-formed propositional formulae. The first two axioms are also
all the axioms of the Pure Positive Implicational Propositional Calculus:
Axiom P1 .P/ ) Œ.Q/ ) .P/.
Axiom P2 f.P/ ) Œ.Q/ ) .R/g ) fŒ.P/ ) .Q/ ) Œ.P/ ) .R/g.
Axiom P3 fŒ:.Q/ ) Œ:.P/g ) Œ.P/ ) .Q/.
The first two axioms are also in Frege’s work [38], [39, p. 137, eq. (1) & (2)],
where they reflect a common mathematical model of rational thinking:
• Axiom P1, .P/ ) Œ.Q/ ) .P/, is called the law of affirmation of the
consequent. In Frege’s (translated) words, axiom P1 states that
If a proposition ŒP holds, it also holds in case an arbitrary proposition ŒQ holds [39,
p. 137].
• Axiom P2, f.P/ ) Œ.Q/ ) .R/g ) fŒ.P/ ) .Q/ ) Œ.P/ ) .R/g, is called
the law of self-distributivity of implication. In Frege’s words, axiom P2 states
that
If a proposition ŒR is the necessary consequence of two propositions (ŒQ and ŒP),
that is, if Œ.P/ ) f.Q/ ) .R/g, and if the first term ŒQ is again the necessary
consequence of the other ŒP, then the proposition ŒR is the necessary consequence of
the last proposition ŒP alone [39, p. 139].
www.pdfgrip.com
• Axiom P3, fŒ:.Q/ ) Œ:.P/g ) Œ.P/ ) .Q/, called the converse law of
contraposition [18, 20, p. 119], states that a contraposition, Œ:.Q/ ) Œ:.P/
suffices to establish a classical logical implication .P/ ) .Q/.
The converse law of contraposition distinguishes classical logic from several
other systems of logic, for instance, Hilbert’s Positive Propositional Calculus,
Brouwer & Heyting’s Intuitionistic Logic, and Kolmogorov & Johansson’s Minimal
Logic [18, 26, p. 140–146]. Still, all these logical systems include Łukasiewicz’s
first two axioms, P1 and P2.
The propositional calculus includes the following concepts of theorem and proof.
1.5 Definition. A well-formed propositional form is a theorem of a propositional
calculus if and only if it is obtained by the following rules of inference:
1.6 Rule (Axioms).
Every axiom (of a logic) is a theorem (of the same logic).
The name of this rule will be printed here as “Detachment” to avoid unintended
awkward sentences. With Detachment, H is the minor premiss while .H/ ) .C/
is the major premiss (so spelled to distinguish its plural from “premises” [18, p. 1,
footnote 3]). Rule 1.7 states that if a hypothesis H and an implication .H/ ) .C/
hold, then the conclusion C may be “detached from the implication” (“von der
Implikation abgetrennt” in Łukasiewicz’s language [80, p. 124]).
1.8 Remark. Each use of the rule of Detachment requires two previously proved
well-formed formula: a proved hypothesis H and a proved implication .H/ ) .C/.
Definitions 1.3 and 1.4 allow P, Q, and R to denote any propositional letters
or well-formed propositional formulae, so that axioms P1–P3 are templates, or
schemas, to generate axioms. Alternatively, allowing only propositional variables
in axioms P1–P3 but introducing the substitution rule 1.9 gives equivalent axioms:
1.9 Rule (Substitution).
For each propositional variable K in a theorem R (which is a propositional form),
and for each well-formed propositional form L,
the propositional form obtained by replacing in R every occurrence of K by L is again a
theorem.
yields Œ:.C/ ) f.B/ ) Œ:.C/g, which is another instance of axiom P1, and
hence also a theorem. Because such substitutions in an axiom yield other axioms,
each axiom is also called an axiom schema. Thus both axioms .A/ ) Œ.B/ ) .A/
and Œ:.C/ ) f.B/ ) Œ:.C/g result from the axiom schema P1.
1.11 Definition. For every logical formula R, the symbol ` (called a “turnstyle”
and read “yield(s)” [110, p. 57]) and the notation
`R
means that there exists a proof of R. An alternative notation, P1, P2, P3 ` R, also
specifies the list of axioms, here P1, P2, P3, from which R is a theorem.
More generally, for all logical formulae P and R, the notation P ` R means that
with P added to the list of axioms, there exists a proof of R. The corresponding
alternative notation, P1, P2, P3, P ` R, again specifies the list of axioms. In other
words, R is a theorem for the logic with axioms P1, P2, P3, P. With a different
terminology, P ` R means that R is derivable from P and the axioms.
Yet more generally, for all logical formulae P; Q; : : : ; R, either notation
P; Q; : : : ` R or P1, P2, P3, P; Q; : : : ` R, means that with P; Q; : : : added to
the list of axioms, there exists a proof of R. In other words, R is a theorem for
the logic with axioms P1, P2, P3, P; Q; : : : The formula R is then derivable from
P; Q; : : :, if and only if P; Q; : : : ` R. In the notation of Smullyan [117, p. 17] and
Stolyar [123, p. 63], P; Q; : : : ` R is also denoted by
P; Q; : : :
:
R
Verifying a proof reduces to checking that each step conforms to the foregoing
definition of proof. In contrast, constructing a proof may require some creativity,
which may involve trying some rules and some axioms in various combinations,
some of which may fail whereas others may succeed [72, p. 55, lines 1–3], [114,
p. 31]. For the propositional calculus presented here, there is an algorithm (a recipe)
to design proofs, but its justification first requires most of the proofs presented here
[123, p. 193–197]. Moreover, the algorithm is cumbersome and would generate
proofs longer than the ones explained here. Nevertheless, the collection of all the
proofs shown here will demonstrate the steps that the algorithm would involve.
With such an understanding of the algorithm, a user might then automate the
algorithm with a computer [47, 50, 54, 139]. Furthermore, the following proofs also
provide some practice in creating proofs without using an algorithm, a practice that
corresponds more closely to the situation in mathematics, and for which there can
be no algorithm [46].
www.pdfgrip.com
With the two rules of inference (Detachment and substitution), the first two axioms
of classical propositional calculus (P1 and P2) pertain only to logical implications
.)/; they form the Pure Positive Implicational Propositional Calculus, which is
common to other logics [18, 29, p. 161]. In contrast, the concept of negation .:/
does not belong to the Positive Implicational Propositional Calculus. Axioms about
the negation, for instance, axiom P3, differentiate classical logic from other logics.
The following three theorems provide examples of proofs about or in the Pure
Positive Implicational Propositional Calculus, which is the propositional calculus
with implications but no negations. The first theorem is called a “derived rule”
(of inference) because it involves a hypothesis, T, which can be any axiom or
previous theorem. Such a derived rule of inference is a theorem about rather than
in the implicational calculus, but it provides a recipe to shorten subsequent proofs.
Specifically, theorem 1.12 shows a derivation of .S/ ) .T/ from T and axiom P1.
The proof of theorem 1.12 is also a building block of the Deduction Theorem 1.22.
1.12 Theorem (derived rule). For each well-formed formula S and for each
theorem T, the implication .S/ ) .T/ is a theorem: P1, T ` .S/ ) .T/.
Proof. Apply axiom P1 and Detachment as follows:
`T hypothesis (minor premiss),
` .T/ ) Œ.S/ ) .T/ substitution in axiom P1 (major premiss),
` .S/ ) .T/ Detachment and preceding two formulae.
Thus .S/ ) .T/ is a theorem derivable from axiom P1, the hypothesis T, and
Detachment. t
u
The second theorem uses axiom P2 and also involves a hypothesis, .H/ )
Œ.K/ ) .L/, which can be any formula of this form that has already been proved.
1.13 Theorem (derived rule). For all well-formed formulae H, K, L, if
.H/ ) Œ.K/ ) .L/
is a theorem, then
Œ.H/ ) .K/ ) Œ.H/ ) .L/
is also a theorem: P2, .H/ ) Œ.K/ ) .L/ ` Œ.H/ ) .K/ ) Œ.H/ ) .L/.
Proof. Apply axiom P2 and Detachment:
` .H/ ) Œ.K/ ) .L/ hypothesis,
` f.H/ ) Œ.K/ ) .L/g ) fŒ.H/ ) .K/ ) Œ.H/ ) .L/g axiom P2,
` Œ.H/ ) .K/ ) Œ.H/ ) .L/ Detachment.
t
u
www.pdfgrip.com
The third theorem (1.14) involves no hypotheses other than the axioms. Thus
theorem 1.14 is a theorem of the Pure Implicational Propositional Calculus. More
accurately, theorem 1.14 is a theorem schema representing a different theorem for
each different formula P. The proof of theorem 1.14 is also a building block of the
Deduction Theorem 1.22.
1.14 Theorem (reflexive law of implication). For each well-formed propositional
formula P, the formula
.P/ ) .P/
H K H L
‚…„ƒ ‚ …„ ƒ ‚…„ƒ ‚…„ƒ
2 ` f. P / ) Œ.P/ ) .P/g ) Œ. P / ) . P / theorem 1.13,
P Q R P Q P R
‚…„ƒ ‚ …„ ƒ ‚…„ƒ ‚…„ƒ ‚ …„ ƒ ‚…„ƒ ‚…„ƒ
2a ` . P /)fŒ.P/).P/). P /g ) f. P /)Œ.P/).P/g)Œ. P /). P / P2,
„ ƒ‚ … „ ƒ‚ …
S U
U
‚ …„ ƒ
2b ` f.P/)Œ.P/).P/g)Œ.P/).P/ 1, 2a, M.P.,
„ ƒ‚ … „ ƒ‚ …
V W
V
‚ …„ ƒ
4 ` .P/)Œ.P/).P/ P1,
Similarly, the following theorem simplifies the use of theorem 1.15 if one of the
components is already a theorem.
1.17 Theorem (derived rule). For all well-formed formulae H, K, L, if
.K/ and
.H/ ) Œ.K/ ) .L/
are theorems, then
.H/ ) .L/
is also a theorem. Thus, K; .H/ ) Œ.K/ ) .L/ ` .H/ ) .L/.
Proof. Apply theorems 1.12 and 1.15:
`K hypothesis,
` .H/ ) .K/ theorem 1.12,
` .H/ ) Œ.K/ ) .L/ hypothesis,
` .H/ ) .L/ theorem 1.15.
t
u
The following theorem demonstrates how the rule of Detachment extends to a
sequence of several logical implications.
1.18 Theorem (derived rule). For all well-formed formulae P, Q, R, S, if
.P/ ) .Q/,
.P/ ) Œ.Q/ ) .R/, and
.P/ ) Œ.R/ ) .S/
are theorems, then
.P/ ) .S/
is also a theorem. Thus,
.P/ ) .Q/; .P/ ) Œ.Q/ ) .R/; .P/ ) Œ.R/ ) .S/ ` .P/ ) .S/:
.H/ ) Œ.K/ ) .L/; .H/ ) f.K/ ) Œ.L/ ) .M/g ` .H/ ) Œ.K/ ) .M/:
In general the question arises, how to find a proof of a theorem. One guide to
design a proof of an implication .H/ ) .C/, where H and C denote well-formed
propositional formulae, begins by deriving a proof for a derivation H ` C of the
conclusion C from the hypothesis H and the axioms. For instance, all the proofs
of derived rules in subsection 1.3.2 are examples of such derivations. The method
for designing a proof then proceeds to “discharge” the hypothesis H to produce a
proof of .H/ ) .C/, as described in the proof of the Deduction Theorem 1.22,
which is not in but about the implicational calculus. More generally, from any proof
that a logical proposition S is derivable from proved hypotheses H; K; : : : ; M; N the
Deduction Theorem provides a recipe to turn that proof into a proof of
H; K; : : : ; M; N ` S
Proof (Outline). The Deduction Theorem removes the hypotheses one at a time, for
instance, beginning with the last one listed, here N, from all the steps in the proof.
www.pdfgrip.com
(D1) If the step ` P in the initial proof is the hypothesis N being removed, then
in the new proof the Deduction Theorem replaces the old step
` N (current hypothesis)
by a complete proof of .N/ ) .N/, for instance, that of theorem 1.14:
` .N/ ) fŒ.N/ ) .N/ ) .N/g axiom P1,
` .N/ ) fŒ.N/ ) .N/ ) .N/g axiom P2,
) f.N/ ) Œ.N/ ) .N/g ) Œ.N/ ) .N/
Still with the hypothesis N, after the completion of any operation (D1)–(D3) on step
P, the Deduction Theorem then performs the same operations (D1)–(D3) on each of
the following steps, Q; : : : ; R. After the completion of operations (D1)–(D3) on all
the steps P; Q; : : : ; R, for the hypothesis N, the Deduction Theorem gives a proof of
H; K; : : : ; M ` Œ.N/ ) .S/:
Then the Deduction Theorem repeats the whole process with the preceding hypothe-
ses, H; : : : ; M. The Deduction Theorem terminates with a proof of
The general case follows by several applications of the previous cases, in a way
that may be specified more explicitly after the availability of the Principle of
Mathematical Induction in chapter 4. t
u
Example 1.23 shows how to use theorem 1.22.
1.23 Example (Tarski’s axiom II). To prove Tarski’s axiom II, f.P/ ) Œ.P/ )
.Q/g ) Œ.P/ ) .Q/, define H and C by
Phase 1: a proof of H ` C.
To transform this derivation .H ` C/ into a proof of .H/ ) .C/, apply to each step
of phase 1 the procedure described in the Deduction Theorem (1.22):
` f.P/)Œ.P/).Q/g)f.P/)Œ.P/).Q/g (D1),
„ ƒ‚ … „ ƒ‚ …
H H
H
‚ …„ ƒ
` f.P/)Œ.P/).Q/g
) f.P/)Œ.P/).Q/g)fŒ.P/).P/)Œ.P/).Q/g (D2), P2,
„ ƒ‚ … „ ƒ‚ …
H L
H L
‚ …„ ƒ ‚ …„ ƒ
` f.P/)Œ.P/).Q/g)fŒ.P/).P/)Œ.P/).Q/g (D3),
„ ƒ‚ … „ ƒ‚ …
K C
H
‚ …„ ƒ
` f.P/)Œ.P/).Q/g)Œ.P/).P/ (D2), 1.12,
„ ƒ‚ …
K
H
‚ …„ ƒ
` f.P/)Œ.P/).Q/g)Œ.P/).Q/ (D3),
„ ƒ‚ …
C
which completes the proof of f.P/ ) Œ.P/ ) .Q/g ) Œ.P/ ) .Q/. A completely
formal proof would also expand each of the steps just listed into its own proof,
as specified in the proof of the Deduction Theorem 1.22, for instance, expanding
each use of the directive (D3), which uses theorem 1.15, into a complete proof of
theorem 1.15.
Although theorem 1.22 can produce lengthy proofs, the resulting long proofs can
also suggest shorter proofs, for instance, as in theorem 1.24.
1.24 Theorem. The formula f.P/ ) Œ.P/ ) .Q/g ) Œ.P/ ) .Q/ is a theorem.
Proof. Apply axiom P2 with theorems 1.14 and 1.17:
` .P/ ) .P/ theorem 1.14,
„ ƒ‚ …
K
The following proof shows the use of the Deduction Theorem in designing proofs.
1.26 Theorem. The law of assertion .A/ ) fŒ.A/ ) .B/ ) .B/g is a theorem.
Proof. A finished proof can proceed as follows:
` Œ.A/ ) .B/ ) Œ.A/ ) .B/ theorem 1.14,
` .A/ ) fŒ.A/ ) .B/ ) Œ.A/ ) .B/g theorem 1.12,
` .A/ ) fŒ.A/ ) .B/ ) .A/g axiom P1,
` .A/ ) fŒ.A/ ) .B/ ) .B/g theorem 1.20.
The following considerations explain how to design such a proof.
The formula .A/ ) fŒ.A/ ) .B/ ) .B/g has the pattern .H/ ) Œ.K/ ) .S/
of the Deduction Theorem, with A for H, .A/ ) .B/ for K, and B for S.
Step 1.
As in the Deduction Theorem, assume first that the hypotheses H and K are proved,
and from them derive the conclusion S by proving H; K ` S. Here, assume that the
hypotheses A and .A/ ) .B/ are both proved, and prove A; Œ.A/ ) .B/ ` B:
`A first temporary hypothesis,
` .A/ ) .B/ second temporary hypothesis,
`B Detachment.
www.pdfgrip.com
The foregoing derivation shows that if A and .A/ ) .B/ are proved, then B is
proved. Still under the first hypothesis A, the Deduction Theorem allows for the
removal of the second hypothesis, .A/ ) .B/, as follows.
Step 2.
Step 2.1
The first line in step 1 consists of the other hypothesis, A, which is assumed proved,
whence instructions (D2) in the Deduction Theorem replace A with a complete proof
of Œ.A/ ) .B/ ) .A/ as in theorem 1.12. In other words, replace the first line, ` A,
by the following three lines:
Proof of theorem 1.12:
`A temporary hypothesis,
` .A/ ) fŒ.A/ ) .B/ ) .A/g axiom P1,
„ ƒ‚ …
Q
Step 2.2
Similarly, the second line in step 1 consists of the hypothesis K being currently
removed, here .A/ ) .B/, which instructions (D1) in the Deduction Theorem
replace with a complete proof of .K/ ) .K/, here Œ.A/ ) .B/ ) Œ.A/ ) .B/,
as in theorem 1.14. Thus, replace the second line, ` .A/ ) .B/, by the following
lines:
Proof of theorem 1.14
with ŒK for Œ.A/).B/:
` ŒK) fŒK)ŒKg)ŒK axiom P1,
˚
` ŒK)
fŒK)ŒKg)ŒK
axiom P2 : : :
) ŒK)fŒK)ŒKg )fŒK)ŒKg : : : continued,
` ŒK)fŒK)ŒKg )fŒK)ŒKg Detachment,
` ŒK)fŒK)ŒKg axiom P1,
` ŒK)ŒK Detachment,
` Œ.A/).B/)Œ.A/).B/ substitution.
End of proof of theorem 1.14.
www.pdfgrip.com
Step 2.3
Finally, the third line in step 1 invokes Detachment, which instructions (D3) replace
by an instance of (the proof of) theorem 1.15:
` Œ.A/ ) .B/ ) .A/ step 2.1,
` Œ.A/ ) .B/ ) Œ.A/ ) .B/ step 2.2,
` Œ.A/ ) .B/ ) .B/ theorem 1.15.
Hence the proof no longer assumes .A/ ) .B/ as a hypothesis, but it still
assumes A as a hypothesis, thus proving that
Step 3.
Finally, the Deduction Theorem allows for the removal of the first hypothesis, A,
from step 2. Here step 2 consists of steps 2.1, 2.2, and 2.3.
Step 3.1
In step 2.1 the first line consists of this hypothesis, A, whence instructions (D1)
replace A with .A/ ) .A/ by a complete proof of theorem 1.14. In other words,
replace the first line in step 2, ` A, by the following lines:
Proof of theorem 1.14:
` ŒA
˚ ) fŒA
) ŒAg ) ŒA axiom P1,
` ŒA) fŒA ) ŒAg )ŒA axiom P1 : : :
) ŒA ) fŒA ) ŒAg ) fŒA ) ŒAg : : : continued,
` ŒA ) fŒA ) ŒAg ) fŒA ) ŒAg Detachment,
` ŒA ) fŒA ) ŒAg axiom P1,
` ŒA ) ŒA Detachment.
End of proof of theorem 1.14.
Step 3.2
The second line in step 2.1 is an instance of axiom
P1, which instructions (D2)
replace by .A/ ) .A/ ) fŒ.A/ ) .B/ ) .A/g .
Step 3.3
The third line in step 2.1 yields Œ.A/ ) .B/ ) .A/, from Detachment, which
instructions (D3) replace by a complete proof of .A/ ) fŒ.A/ ) .B/ ) .A/g,
as in theorem 1.15. In this case, however, such a proof would be correct but not
www.pdfgrip.com
Step 3.4
The result of step 2.2, Œ.A/ ) .B/ ) Œ.A/ ) .B/, is proved by theorem 1.14.
Hence, instructions (D2) replace it by .A/ ) Œ.A/ ) .B/ ) Œ.A/ ) .B/.
Step 3.5
Fully written out, the remaining lines in step 2.3 would follow the proof of
theorem 1.15. Removing the hypothesis A then amounts to theorem 1.20, which
forms the last line of the final proof:
` .A/)fŒ.A/).B/).A/g axiom P1 (from 3.3, replacing 2.1),
` Œ.A/).B/)Œ.A/).B/ theorem 1.14 (from step 2.2),
` .A/)fŒ.A/).B/)Œ.A/).B/g theorem 1.12 (from 3.4, replacing 2.2),
` .A/)fŒ.A/).B/).B/g Theorem 1.20 (from 3.5, replacing 2.3).
Thus the Deduction Theorem has provided some guidance for the construction
of a proof of the theorem .A/ ) fŒ.A/ ) .B/ ) .B/g. t
u
Proof. The formula to be proved has the form .A/ ) Œ.B/ ) .C/, with A denoting
.Q/ ) .R/, B denoting .P/ ) .Q/, and C denoting .P/ ) .R/:
A B C
‚ …„ ƒ ‚ …„ ƒ ‚ …„ ƒ
Œ.Q/ ) .R/ ) fŒ.P/ ) .Q/ ) Œ.P/ ) .R/g:
Expanding the proof of theorem 1.16 which invokes theorems 1.12 and 1.15,
helps discharging the hypothesis .P/ ) .Q/:
` .Q/ ) .R/ hypothesis A,
` .P/ ) Œ.Q/ ) .R/ theorem 1.12,
` .P/ ) .R/ conclusion C, by theorem 1.15.
The preceding three steps form a derivation A; B ` C of C from A and B,
but without invoking B, which plays a hidden rôle in the proof of theorem 1.15.
Replacing the citation of theorem 1.15 by its proof leads to a derivation of .B/ )
.C/ from A:
` .Q/ ) .R/ hypothesis A,
` .P/ ) Œ.Q/ ) .R/ theorem 1.12,
` Œ.P/ ) .Q/ ) Œ.Q/ ) .R/ .B/ ) .C/ by theorem 1.13.
Replacing the citation of theorem 1.13 by its proof, which uses only axioms and
Detachment, gives a derivation of .B/ ) .C/ from A directly from the axioms:
1 ` .Q/).R/ hypothesis A,
2 ` .P/)Œ.Q/).R/ theorem 1.12,
3 ` f.P/)Œ.Q/).R/g)fŒ.P/).Q/)Œ.P/).R/g axiom P2,
4 `Œ.P/).Q/ ) Œ.P/).R/ 2, 3, Detachment.
Lines 1–4 complete the derivation of Œ.P/ ) .Q/ ) Œ.Q/ ) .R/ from the first
hypothesis .Q/ ) .R/ with axiom P2 and Detachment. Since the second hypothesis,
.P/ ) .Q/, has not been used, it need not be discharged.
Phase 2: discharging A.
An application of the Deduction Theorem (1.22) discharges the first hypothesis and
yields a proof of .A/ ) Œ.B/ ) .C/. The resulting proof can be shortened, or
alternative proofs may result from trial and error. To this end, H, K, L refer to
theorem 1.15:
` Œ.Q/ ) .R/ ) f.P/ ) Œ.Q/ ) .R/g axiom P1,
„ ƒ‚ … „ ƒ‚ …
H K
Proof. With notation as in the proof of theorem 1.27, the formula to be proved has
the form .B/ ) Œ.A/ ) .C/, with A denoting .Q/ ) .R/, B denoting .P/ ) .Q/,
and C denoting .P/ ) .R/:
B A C
‚ …„ ƒ ‚ …„ ƒ ‚ …„ ƒ
Œ.P/ ) .Q/ ) fŒ.Q/ ) .R/ ) Œ.P/ ) .R/g:
Thus designing a proof of .B/ ) Œ.A/ ) .C/ can start with a derivation B; A ` C,
in this case .P/ ) .Q/; .Q/ ) .R/ ` .P/ ) .R/, which is exactly theorem 1.16.
Hence steps as in the proof of theorem 1.27 discharge the hypotheses, but in the
reverse order. The resulting proof might then be shortened or give clues for a shorter
alternative proof. For example, apply axiom P1 with theorems 1.27, 1.13, and 1.16:
U V
‚ …„ ƒ ‚ …„ ƒ
` Œ.P/ ) .Q/ ) fŒ.Q/ ) .R/ ) Œ.P/ ) .Q/g axiom P1,
„ ƒ‚ … „ ƒ‚ … „ ƒ‚ …
K H K
H K H L
‚ …„ ƒ ‚ …„ ƒ ‚ …„ ƒ ‚ …„ ƒ
` fŒ.Q/ ) .R/ ) Œ.P/ ) .Q/g ) fŒ.Q/ ) .R/ ) Œ.P/ ) .R/g 1.13,
„ ƒ‚ … „ ƒ‚ …
V W
U W
‚ …„ ƒ ‚ …„ ƒ
` Œ.P/ ) .Q/ ) fŒ.Q/ ) .R/ ) Œ.P/ ) .R/g 1.16.
t
u
Besides the Deduction Theorem (theorem 1.22), another guide to design proofs
consists of replacing any occurrence of a formula by an equivalent formula, thanks
to theorem 1.29 [18, p. 101, 124, 189], [108, p. 48].
1.29 Theorem (Substitutivity of Equivalence in the Pure Positive Implicational
Propositional Calculus, preliminary version). For all well-formed implicational
logical formulae U and V, if ` .U/ ) .V/ and ` .V/ ) .U/, and if a formula Q
results from substituting any (zero, one, several, or all) occurrence(s) of U by V in
a well-formed formula P, then ` .P/ ) .Q/ and ` .Q/ ) .P/.
Proof (Outline). This proof proceeds by cases and subcases.
In all cases, if Q is P, which results by substituting none of the occurrences of U
by V, then each of .P/ ) .Q/ and .Q/ ) .P/ is .P/ ) .P/, which is theorem 1.14.
www.pdfgrip.com
The following derived rules allow for substitutions within implications subject to
hypotheses, for instance, a substitution within an intermediate hypothesis.
1.31 Theorem (derived rule). For all well-formed formulae H, K, L, M, if
.H/ ) Œ.L/ ) .M/ and
.K/ ) .L/
are theorems, then
.H/ ) Œ.K/ ) .M/
is also a theorem.
Proof. Apply theorems 1.28 and 1.16:
` .K/ ) .L/ hypothesis,
` Œ.K/ ) .L/ ) fŒ.L/ ) .M/ ) Œ.K/ ) .M/g theorem 1.28,
` Œ.L/ ) .M/ ) Œ.K/ ) .M/ Detachment. t
u
` .H/ ) Œ.L/ ) .M/ hypothesis,
` .H/ ) Œ.K/ ) .M/ theorem 1.16.
The second theorem allows for a substitution in the conclusion.
1.32 Theorem (derived rule). For all well-formed formulae H, L, M, N, if
.H/ ) Œ.L/ ) .M/ and
.M/ ) .N/
are theorems, then
.H/ ) Œ.L/ ) .N/
is also a theorem.
Proof. Apply theorems 1.28, 1.16, and 1.17:
` .H/ ) Œ.L/ ) .M/ hypothesis,
„ ƒ‚ …
P
` Œ.L/ ) .M/ ) fŒ.M/ ) .N/ ) Œ.L/ ) .N/g theorem 1.28,
„ ƒ‚ … „ ƒ‚ …
P Q
` .H/ ) fŒ.M/ ) .N/ ) Œ.L/ ) .N/g theorem 1.16,
„ ƒ‚ …
Q
` .M/ ) .N/ hypothesis,
` .H/ ) Œ.L/ ) .N/ theorem 1.17.
t
u
The following three derived rules of inference will simplify subsequent proofs.
1.33 Theorem (derived rule). If .K/ ) .L/ is a theorem, then the following
formula is also a theorem: Œ.H/ ) .K/ ) Œ.H/ ) .L/.
www.pdfgrip.com
` Œ.K/).L/)Œ.H/).L/ Detachment,
„ ƒ‚ … „ ƒ‚ …
P R
` Œ.H/).L/).M/ hypothesis,
„ ƒ‚ …
R
` Œ.K/).L/).M/ 1.16.
„ ƒ‚ …
P
t
u
The following “Law of Commutation” allows for yet another change in the order of
hypotheses:
The Law of Commutation is one of Frege’s axioms [39, p. 146, eq. (8)]. In Frege’s
1879 words, the Law of Commutation states that
If a proposition is the consequence of two propositions, their order is immaterial [39,
p. 147].
1.38 Theorem (law of assertion). The formula .H/ ) fŒ.H/ ) .C/ ) .C/g is
a theorem.
Proof. Apply theorems 1.14 and 1.37 with Detachment:
P P
‚ …„ ƒ ‚ …„ ƒ
` Œ.H/ ) .C/ ) Œ.„ƒ‚…
H / ) .„ƒ‚…
C / theorem 1.14,
„ ƒ‚ …
P Q R
` .„ƒ‚…
H / ) fŒ.H/ ) .C/ ) .„ƒ‚…
C /g commutation (1.37) and Detachment.
„ ƒ‚ …
Q P R
t
u
Also relying on the law of commutation, the following derived rule will shorten
the proof of subsequent results in particular, theorem 1.50.
1.39 Theorem (derived rule). If ` .U/ ) Œ.V/ ) .W/
and ` .H/ ) Œ.W/ ) .R/,
then ` .H/ ) f.U/ ) Œ.V/ ) .R/g.
Proof. Apply the transitivity of implication (theorem 1.27):
` Œ.V/).W/)fŒ.W/).R/)Œ.V/).R/g
theorem 1.27,
` .U/) Œ.V/).W/)fŒ.W/).R/)Œ.V/).R/g theorem 1.12,
` .U/)Œ.V/).W/ hypothesis,
` .U/)fŒ.W/).R/)Œ.V/).R/g theorem 1.15,
` f.U/)Œ.W/).R/g)f.U/)Œ.V/).R/g theorem 1.13,
` .H/)Œ.W/).R/ hypothesis,
` .U/)f.H/)Œ.W/).R/g theorem 1.12,
` .H/)f.U/)Œ.W/).R/g commutation (1.37),
` .H/)f.U/)Œ.V/).R/g transitivity.
t
u
The foregoing theorems involve only implications but no negation, and their
proofs do not involve any negation either. Nevertheless, there exist other theorems
involving only ) but not : for which there does not exist any proof involving only
implications. Examples of such theorems are hidden in the exercises, to be revealed
later. Investigate whether the formulae in the following exercises are theorems, using
any of the axioms P1 and P2, any rules of inference, and any of the theorems just
proved.
1.1 . Œ.H/ ) .L/ ) f.H/ ) Œ.K/ ) .L/g
1.2 . Œ.K/ ) .L/ ) f.H/ ) Œ.K/ ) .L/g
1.3 . Œ.A/ ) .B/ ) fŒ.A/ ) .B/ ) .B/g ) .B/
www.pdfgrip.com
1.4 . Œ.A/ ) .B/ ) fŒ.A/ ) .B/ ) .A/g ) .A/
1.5 . fŒ.P/ ) .P/ ) .P/g ) .P/
1.6 . .P/ ) fŒ.P/ ) .P/ ) .P/g
1.7 . fŒ.P/ ) .Q/ ) .P/g ) .P/ (Peirce’s law.)
1.8 . Œ.P/ ) .R/ ) fŒ.P/ ) .Q/ ) .R/g ) .R/
1.9 . fŒ.R/ ) .Q/ ) .P/g ) fŒ.R/ ) .Q/ ) Œ.S/ ) .P/g
1.10 . Œ.R/ ) .Q/ ) fŒ.R/ ) .Q/ ) .P/g ) Œ.S/ ) .P/
The Classical Implicational Calculus just presented rests on the rules of Detachment
and Substitution with Frege’s axioms P1 and P2:
Frege’s axiom P1 .P/ ) Œ.Q/ ) .P/.
Frege’s axiom P2 f.P/ ) Œ.Q/ ) .R/g ) fŒ.P/ ) .Q/ ) Œ.P/ ) .R/g.
Other selections of axioms exist, for instance, Stephen Cole Kleene’s [72, p. 15],
Kleene’s axiom 1a .A/ ) Œ.B/ ) .A/,
Kleene’s axiom 1b Œ.A/ ) .B/ ) .f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C//,
and Tarski’s [129, p. 147],
Tarski’s axiom I .P/ ) Œ.Q/ ) .P/,
Tarski’s axiom II f.P/ ) Œ.P/ ) .Q/g ) Œ.P/ ) .Q/,
Tarski’s axiom III Œ.P/ ) .Q/ ) fŒ.Q/ ) .R/ ) Œ.P/ ) .R/g.
Frege’s, Kleene’s, and Tarski’s implicational axiom systems are mutually equiva-
lent, in the sense that each system leads to the same Pure Positive Implicational
Propositional Calculus.
Indeed, their first axioms, P1, 1a, and I are mutually identical.
Second, Kleene’s axiom 1b results from applying the law of commutation
(theorem 1.37) to Frege’s axiom P2. Consequently, both of Kleene’s axioms 1a
and 1b are theorems derivable from Frege’s axioms P1 and P2. Thus, prepending
derivations of Kleene’s axioms from Frege’s axioms to any proof of any theorem
from Kleene’s axioms yields a proof of the same theorem from Frege’s axioms.
Therefore, every theorem derivable from Kleene’s axioms is also derivable from
Frege’s axioms.
Similarly, Tarski’s axiom II is theorem 1.24, while Tarski’s axiom III is the-
orem 1.28. Consequently, all three of Tarski’s axioms I, II, and III are theorems
derivable from Frege’s axioms P1 and P2. Therefore, every theorem derivable from
Tarski’s axioms is also derivable from Frege’s axioms.
www.pdfgrip.com
For the following exercises, use only Kleene’s axioms 1a and 1b with Substitution
and Detachment.
1.11 . Prove .P/ ) .P/.
1.12 . Establish the derived rule of inference that if T is a theorem, then .S/ ) .T/
is also a theorem.
1.13 . Prove fŒ.P/ ) Œ.P/ ) .Q/g ) Œ.P/ ) .Q/.
1.14 . Establish the derived rule of inference that if .H/ ) .K/ and .K/ ) .L/ are
theorems, then .H/ ) .L/ is also a theorem.
1.15 . Establish the derived rule of inference that if T is a theorem, then f.A/ )
Œ.T/ ) .C/g ) Œ.A/ ) .C/ is also a theorem.
1.16 . Prove Œ.B/ ) .C/ ) f.A/ ) Œ.B/ ) .C/g.
1.17 . Prove
n o
Œ.B/ ) .C/ ) f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C/
1.18 . Prove
fŒ.B/).C/)Œ.A/).B/g
h n o
) Œ.B/).C/) Œ.A/).B/) f.A/)Œ.B/).C/g)Œ.A/).C/
n oi
) Œ.B/).C/) f.A/)Œ.B/).C/g)Œ.A/).C/ :
www.pdfgrip.com
1.19 . Prove
1.20 . Prove
For the following exercises, use only Tarski’s axioms I, II, and III, with Substitution
and Detachment.
1.23 . Establish the derived rule of inference that if .H/ ) .K/ and .K/ ) .L/ are
theorems, then .H/ ) .L/ is also a theorem.
1.24 . Establish the derived rule of inference that if T is a theorem, then .S/ ) .T/
is also a theorem.
1.25 . Prove .P/ ) fŒ.P/ ) .Q/ ) .P/g.
1.26 . Prove
fŒ.P/ ) .Q/ ) .P/g ) Œ.P/ ) .Q/ ) fŒ.P/ ) .Q/ ) .Q/g :
1.27 . Prove
.P/ ) Œ.P/ ) .Q/ ) fŒ.P/ ) .Q/ ) .Q/g :
1.28 . Prove
Œ.P/ ) .Q/ ) fŒ.P/ ) .Q/ ) .Q/g ) fŒ.P/ ) .Q/ ) .Q/g:
1.30 . Prove
f.P/ ) Œ.Q/ ) .R/g ) fŒ.Q/ ) .R/ ) .R/g ) Œ.P/ ) .R/ :
1.31 . Prove
fŒ.Q/ ) .R/ ) .R/g ) Œ.P/ ) .R/ ) f.Q/ ) Œ.P/ ) .R/g:
1.34 . Prove
f.Q/ ) Œ.P/ ) .R/g ) Œ.P/ ) .Q/ ) f.P/ ) Œ.P/ ) .R/g :
1.35 . Prove
Œ.P/ ) .Q/ ) f.P/ ) Œ.Q/ ) .R/g ) Œ.P/ ) .R/g :
1.36 . Prove
The following proofs demonstrate the use of the converse law of contraposition.
www.pdfgrip.com
1.40 Theorem (law of denial of the antecedent). For all well-formed formulae P
and Q, the following formula is a theorem: Œ:.P/ ) Œ.P/ ) .Q/.
Proof. Apply axioms P1 and P3 with the transitivity of implication (theorem 1.16):
` Œ:.P/ ) fŒ:.Q/ ) Œ:.P/g substitution in axiom P1,
` fŒ:.Q/ ) Œ:.P/g ) Œ.P/ ) .Q/ axiom P3,
` Œ:.P/ ) Œ.P/ ) .Q/ theorem 1.16.
t
u
1.41 Theorem. The formula .P/ ) fŒ:.P/ ) .Q/g is a theorem (schema).
Proof. Apply theorem 1.40 and the law of commutation (theorem 1.37):
` Œ:.P/ ) Œ.P/ ) .Q/ theorem 1.40,
` fŒ:.P/ ) Œ.P/ ) .Q/g ) .P/ ) fŒ:.P/ ) .Q/g commutation (1.37),
` .P/ ) fŒ:.P/ ) .Q/g Detachment.
t
u
The following two theorems establish the complete law of double negation.
1.42 Theorem (law of double negation). The formula Œ::.P/ ) .P/ is a
theorem (schema).
Proof. Apply the transitivity of implication (theorem 1.19) and theorem 1.25:
` f:Œ:.P/g
) fŒ::::.P/ ) Œ::.P/g axiom P1,
` f:Œ:::.P/g ) f:Œ:.P/g ) fŒ:.P/ ) Œ:::.P/g axiom P3,
` Œ:.P/ ) f:Œ::.P/g ) fŒ::.P/ ) .P/g axiom P3,
` Œ::.P/ ) fŒ::.P/ ) .P/g theorem 1.19,
` Œ::.P/ ) .P/ theorem 1.25.
t
u
1.43 Theorem (converse law of double negation). The formula .P/ ) Œ::.P/
is a theorem (schema).
Proof. Apply the law
of double negation (theorem 1.42) and contraposition (P3):
` ˚:f:Œ:.P/g ) Œ:.P/ theorem 1.42,
` :f:Œ:.P/g ) Œ:.P/ ) .P/ ) f:Œ:.P/g axiom P3,
` .P/ ) f:Œ:.P/g Detachment.
t
u
With axiom P3, the following theorem gives the complete law of contraposition.
1.44 Theorem (law of contraposition, principle of transposition). The follow-
ing formula is a theorem (schema): Œ.P/ ) .Q/ ) fŒ:.Q/ ) Œ:.P/g .
Proof. Apply transitivity (theorem 1.16) with A, B, C, D as in theorem 1.35:
` f:Œ:.P/g ) .„ƒ‚… P / theorem 1.42,
„ ƒ‚ …
A B
` . Q / ) f:Œ:.Q/g theorem 1.43,
„ƒ‚… „ ƒ‚ …
C D
www.pdfgrip.com
` Œ.„ƒ‚…
P / ) . Q / ) fŒ::.P/ ) Œ::.Q/g theorem 1.35,
„ƒ‚… „ ƒ‚ … „ ƒ‚ …
B C A D
` fŒ::.P/ ) Œ::.Q/g ) fŒ:.Q/ ) Œ:.P/g axiom P3,
` Œ.P/ ) .Q/ ) fŒ:.Q/ ) Œ:.P/g theorem 1.16.
t
u
Theorem 1.44 is the theoretical basis for reasoning by contraposition. Theo-
rem 1.45 syncopates several steps for later use.
1.45 Theorem. For all well-formed formulae P and Q, the following formula is a
theorem: fŒ:.P/ ) .Q/g ) fŒ:.Q/ ) .P/g.
Proof. Apply the principle of transposition and the law of double negation:
` fŒ:.P/ ) .Q/g ) Œ:.Q/ ) f:Œ:.P/g transposition (1.44) ,
` f:Œ:.P/g ) .P/ double negation (1.42) ,
` fŒ:.P/ ) .Q/g ) fŒ:.Q/ ) .P/g derived rule (1.32).
t
u
The general case follows by several applications of the previous case and the
cases in theorem 1.29, in a way that may be specified more explicitly after the
availability of the Principle of Mathematical Induction in chapter 4. t
u
Within classical logic, a proposition and its negation together form an “absurdity”
that cannot hold. In particular, if a hypothesis implies a conclusion and its
negation — an absurdity — then the hypothesis may be rejected. The following
theorems establish the validity of such a pattern of reasoning, called reduction to
the absurd.
1.47 Theorem (special law of reductio ad absurdum). For each well-formed
formula P, the following formula is a theorem: f.P/ ) Œ:.P/g ) Œ:.P/ .
Proof. Start with theorem
1.44 and the denialof the antecedent (theorem 1.40):
` f.P/ ) Œ:.P/g ) f:Œ:.P/g ) Œ:.P/ (1.44),
` f:Œ:.P/g
) Œ:.P/
) f:Œ.P/ ) .P/g (1.40),
` f:Œ:.P/g
) Œ:.P/ ) f:Œ.P/ ) .P/g
) f:Œ:.P/g ) Œ:.P/
) f:Œ:.P/g ) f:Œ.P/ ) .P/g (P2),
` f:Œ:.P/g ) Œ:.P/ ) f:Œ:.P/g ) f:Œ.P/ ) .P/g (M.P.),
` f:Œ:.P/g ) f:Œ.P/ ) .P/g ) fŒ.P/ ) .P/ ) Œ:.P/g (P3),
` f.P/ ) Œ:.P/g ) fŒ.P/ ) .P/ ) Œ:.P/g (1.19),
` .P/ ) .P/ (1.14),
` f.P/ ) Œ:.P/g ) Œ:.P/ (1.17).
t
u
` fŒ:.P/ ) .R/g ) Œ.Q/ ) .R/ ) fŒ.P/ ) .Q/ ) .R/g theorem 1.39.
„ ƒ‚ … „ ƒ‚ … „ ƒ‚ …
H U V
t
u
www.pdfgrip.com
For the following four exercises, use the rules of inference and only the following
six axioms, due to Frege [39]:
Axiom F2 [39, p. 137, eq. (2)]: f.P/ ) Œ.Q/ ) .R/g ) fŒ.P/ ) .Q/ )
Œ.P/ ) .R/g.
Axiom F3 [39, p. 146, eq. (8)]: f.P/ ) Œ.Q/ ) .R/g ) f.Q/ ) Œ.P/ ) .R/g.
1.43 . Define a “negation” .P/ to be an abbreviation for .P/ ) .F/. Prove that
every theorem of classical logic is a theorem of Church’s logic.
1.44 . Using the theorem :.F/ and axioms P1, P2, P3, prove the theorems fŒ.P/ )
.F/ ) .F/g ) f:Œ:.P/g and f:Œ:.P/g ) fŒ.P/ ) .F/ ) .F/g .
1.45 . Using axioms
P1, P2, P3 and anyof the classical theorems already
proved,
prove Œ:.P/ ) .P/ ) f:Œ.S/ ) .S/g and .P/ ) f:Œ.S/ ) .S/g ) Œ:.P/.
1.46 . In classical logic define a constant f to be an abbreviation for :Œ.S/ ) .S/.
Prove that every theorem of Church’s logic is a theorem of classical logic.
There exist logical connectives other than the negation .:/ and implication .)/,
for example, the conjunction .^/, disjunction ._/, and equivalence .,/. Such other
connectives can be specified by other axioms, for instance, by Tarski’s axioms IV–
VI to specify the equivalence, as in example 1.87 on page 55. Alternatively, other
connectives can be defined as abbreviations of longer formulae in terms of negations
and implications, as presented in this section.
The logical connectives : and ) suffice to define all the other logical connectives,
for instance, the conjunction ^, the disjunction _, and the equivalence ,, as
outlined here.
1.51 Definition (conjunction, disjunction, and equivalence). In the full Classi-
cal Propositional Calculus, the connectives ^, _, and , may be defined as follows:
.P/ ^ .Q/ stands for :f.P/ ) Œ:.Q/g;
.P/ _ .Q/ stands for Œ.P/ ) .Q/ ) .Q/;
.P/ , .Q/ stands for Œ.P/ ) .Q/ ^ Œ.Q/ ) .P/.
The logical conjunction .^/ can also be introduced into a logic by additional
axioms, for instance, as in Hilbert’s Positive Propositional Calculus, Brouwer &
Heyting’s Intuitionistic Logic and Kolmogorov & Johansson’s Minimal Logic [18,
26, p. 140–146]:
www.pdfgrip.com
The following theorems reveal that within the Classical Propositional Calculus,
such axioms also follow from definition 1.51.
The first theorems show that if .P/ ^ .Q/ holds, then P holds and Q holds.
1.52 Theorem (AND.1). The formula Œ.P/ ^ .Q/ ) .Q/ is a theorem (schema).
Proof. Apply transposition, double negation, transitivity, and the definition:
` Œ:.Q/
) f.P/ ) Œ:.Q/g axiom P1,
` :f.P/ ) Œ:.Q/g ) f:Œ:.Q/g contraposition (1.44) of P1, and M.P.,
` Œ.P/ ^ .Q/ ) f:Œ:.Q/g definition 1.51 of ^,
` f:Œ:.Q/g ) .Q/ theorem 1.42,
` Œ.P/ ^ .Q/ ) .Q/ theorem 1.16.
t
u
1.53 Theorem (AND.2). The formula Œ.P/ ^ .Q/ ) .P/ is a theorem.
Proof. Apply theorems 1.40, 1.44, 1.42, and 1.16:
` Œ:.P/
) Œ.P/ ) .Q/
denial of the antecedent (1.40),
` :f.P/ ) Œ:.Q/g ) f:Œ:.P/g contraposition (1.44) of 1.40,
` Œ::.P/ ) .P/ double negation (1.42),
` Œ.P/ ^ .Q/ ) .P/ transitivity (theorem 1.16).
t
u
The next theorem demonstrates a “converse” to the preceding two theorems. so
that if P and Q hold, then their conjunction .P/ ^ .Q/ also holds.
1.54 Theorem (AND.3). If P and Q are both theorems, then .P/^.Q/ is a theorem;
equivalently, for all well-formed formulae P and Q, ` .P/ ) f.Q/ ) Œ.P/ ^ .Q/g.
Proof. Apply the
law of assertion, contraposition,
and transitivity:
` .P/ ) f.P/ ) Œ:.Q/g ) Œ:.Q/ assertion (theorem 1.38),
` f.P/
˚ ) Œ:.Q/g
) Œ:.Q/ contraposition : : :
) .Q/ ) :f.P/ ) Œ:.Q/g : : : continued,
˚
` .P/ ) .Q/ ) :f.P/ ) Œ:.Q/g theorem 1.16,
` .P/ ) f.Q/ ) Œ.P/ ^ .Q/g definition of ^.
Hence, if P and Q are both theorems, then .P/ ^ .Q/ is a theorem:
` .P/ ) f.Q/ ) Œ.P/ ^ .Q/g just proved,
` .P/ hypothesis,
` .Q/ ) Œ.P/ ^ .Q/ Detachment,
www.pdfgrip.com
` .Q/ hypothesis,
` .P/ ^ .Q/ Detachment.
t
u
Theorem 1.54 allows for the following derived rule of inference.
1.55 Theorem (derived rule). For all well-formed formulae H, K, and L, if .H/ )
.K/ and .H/ ) .L/ are theorems, then .H/ ) Œ.K/ ^ .L/ is a theorem:
` Œ.H/ ) .K/ ) Œ.H/ ) .L/ ) f.H/ ) Œ.K/ ^ .L/g I
conversely, if .H/ ) Œ.K/ ^ .L/ is a theorem, then .H/ ) .K/ and .H/ ) .L/ are
theorems.
Proof. Apply theorem 1.54 and transitivity
(theorem 1.32) with M for .K/ ^ .L/:
` .H/ ) f.K/ ) Œ.L/ ) .M/g
) Œ.H/ ) .K/ ) f.H/ ) Œ.L/ ) .M/g axiom P2,
The logical equivalence .,/ can also be introduced into a logic by additional
axioms, for instance, Hilbert’s Positive Propositional Calculus, Brouwer & Heyt-
ing’s Intuitionistic Logic and Kolmogorov & Johansson’s Minimal Logic [18, 26,
p. 140–146], by Tarski’s axioms IV, V, and VI [129, p. 147]:
The following theorems reveal that within the Classical Propositional Calculus,
such axioms also follow from definition 1.51.
Combining theorem 1.54 with the definition of , gives Tarksi’s Axiom VI.
1.60 Theorem (Tarski’s axiom VI). For all well-formed formulae U and V,
Tarski’s axioms VI is a theorem: ` Œ.U/ ) .V/ ) fŒ.V/ ) .U/ ) Œ.U/ ,
.V/g.
Proof. Combine theorem 1.54 with the definition of ,:
.U/,.V/
‚ …„ ƒ
Œ.U/ ) .V/ ) Œ.V/ ) .U/ ) fŒ.U/ ) .V/ ^ Œ.V/ ) .U/g :
„ ƒ‚ … „ ƒ‚ … „ ƒ‚ … „ ƒ‚ …
P Q P Q
t
u
A particular instance of the foregoing theorem allows for any proof of any
equivalence .I/ , .J/ to be split into two separate proofs of .I/ ) .J/ and
.J/ ) .I/.
1.61 Theorem (derived rule). If .I/ ) .J/ and .J/ ) .I/ are theorems, then so
is .I/ , .J/. Conversely, if .I/ , .J/ is a theorem, then so are .I/ ) .J/ and
.J/ ) .I/.
Proof. Apply theorem 1.54 with .I/ ) .J/ for P, and with .J/ ) .I/ for Q, so that
if .I/ ) .J/ and .J/ ) .I/ are theorems, then Œ.I/ ) .J/ ^ Œ.J/ ) .I/ is also a
theorem, which is .I/ , .J/ by definition, and conversely. t
u
Theorem 1.62 discharges the hypothesis .I/ , .J/ in theorem 1.61, which yields
Tarski’s axioms IV and V.
1.62 Theorem (Tarski’s axioms IV and V). For all well-formed formulae I and
J, Tarski’s axioms IV Œ.I/ , .J/ ) Œ.I/ ) .J/ and Tarski’s axioms V Œ.I/ ,
.J/ ) Œ.J/ ) .I/ are theorems.
Proof. By definition 1.51, the formula .I/ , .J/ is an abbreviation for Œ.I/ )
.J/ ^ Œ.J/ ) .I/. The present theorem results from substitutions in theorems 1.52
and 1.53. t
u
Hence the following theorem establishes the reflexivity of equivalence.
1.63 Theorem (reflexivity of ,). For each well-formed formula P, ` .P/ , .P/.
Proof. Apply the definition of , and the reflexivity of ) (theorem 1.14) with
theorem 1.61:
` .P/ ) .P/ theorem 1.14,
` .P/ ) .P/ theorem 1.14,
` .P/ , .P/ theorem 1.61.
t
u
www.pdfgrip.com
The logical disjunction ._/ can also be introduced into a logic by additional
axioms, for instance, as in Hilbert’s Positive Propositional Calculus, Brouwer &
Heyting’s Intuitionistic Logic and Kolmogorov & Johansson’s Minimal Logic [18,
26, p. 140–146]:
The following theorems reveal that within the Classical Propositional Calculus,
the first two axioms also follow from definition 1.51. The third axiom is derived in
theorem 1.78.
1.67 Theorem (OR.1, OR.2). For all well-formed formulae P and Q, the formulae
.Q/ ) Œ.P/ _ .Q/ and .P/ ) Œ.P/ _ .Q/ are theorems.
Proof. For the first formula, apply axiom P1 and definition of _:
` .Q/ ) fŒ.P/ ) .Q/ ) .Q/g axiom P1,
` .Q/ ) Œ.P/ _ .Q/ definition of _.
For the second formula, apply theorem 1.14, the Law of Commutation, and the
definitions:
` Œ.P/ ) .Q/ ) Œ.P/ ) .Q/ theorem 1.14,
` .P/ ) fŒ.P/ ) .Q/ ) .Q/g Law of Commutation (theorem 1.37),
` .P/ ) Œ.P/ _ .Q/ definition 1.51 of _.
t
u
The next theorem shows that the disjunction _ is idempotent:
1.68 Theorem (idempotency of _). For each well-formed formula Q,
Proof. The first formula is a substitution in the proof by cases (theorem 1.49).
The second formula results from the law of denial of the antecedent (theo-
rem 1.40) and a derived rule (theorem 1.33):
` Œ:.P/ ) Œ.P/ ) .Q/ theorem 1.40,
„ƒ‚… „ ƒ‚ …
I H
` fŒ.P/ ) .Q/ ) . Q /g ) fŒ:.P/ ) . Q /g theorem 1.33.
„ ƒ‚ … „ƒ‚… „ƒ‚… „ƒ‚…
H K I K
t
u
The next theorem establishes the law of excluded middle in classical logic.
1.70 Theorem (law of excluded middle). For each well-formed formula B, the
formula .B/ _ Œ:.B/ is a theorem.
Proof. Apply theorems 1.14 and 1.69, and the definition of the disjunction ._/:
` Œ:.B/ ) Œ:.B/ theorem 1.14,
` .B/ _ Œ:.B/ theorem 1.69 and definition of _ with : and ).
t
u
The following theorem shows that the disjunction _ is associative:
1.71 Theorem (associativity of _). For all well-formed formulae P, Q, and R,
Proof. Apply the definition of _ with De Morgan’s laws and theorem 1.55:
Œ.P/ _ .Q/ ^ Œ.P/ _ .R/
m definition of _ by theorem 1.69,
fŒ:.P/ ) .Q/g ^ fŒ:.P/ ) .R/g
m theorem 1.55,
Œ:.P/ ) Œ.Q/ ^ .R/
m definition of _ by theorem 1.69.
.P/ _ Œ.Q/ ^ .R/
t
u
The following theorem shows that conjunctions distribute over disjunctions.
1.75 Theorem (Distributivity of ^ over _). The following formula is a theorem:
For the following exercises, prove that the stated formulae are theorems (schemas),
using the classical propositional calculus and any of the results just proved.
1.47 . fŒ.P/ ) .Q/ ) .P/g ) .P/
1.48 . Œ.P/ ) .R/ ) fŒ.P/ ) .Q/ ) .R/g ) .R/
www.pdfgrip.com
The preceding sections have demonstrated the theoretical foundations for patterns of
deduction in the implicational calculus, for example, the transitivity of implication
and the law of commutation, and with contraposition, for example, the law of
reductio ad absurdum. Similarly, this section presents patterns of deduction with
conjunctions and disjunctions, for instance, proofs by cases or by contradiction.
The first theorem shows yet another form of the transitivity of the logical implica-
tion.
1.76 Theorem. For all well-formed formulae P, Q, and R,
Proof. The first part of the proof assumes the two hypotheses.
` .U/ ) .W/ hypothesis,
` Œ:.W/ ) Œ:.U/ contraposition and Detachment,
` .V/ ) .W/ hypothesis,
` Œ:.W/ ) Œ:.V/ contraposition and Detachment,
` Œ:.W/ ) fŒ:.U/ ^ Œ:.V/g theorem 1.55,
` Œ:.W/ ) f:Œ.U/ _ .V/g De Morgan’s second law,
` Œ.U/ _ .V/ ) .W/ axiom P3 and Detachment.
The second part of the proof dispenses with the two hypotheses.
` fŒ.U/).W/ ^ Œ.V/).W/g)Œ.U/).W/ theorem 1.53,
` Œ.U/).W/)fŒ:.W/)Œ:.U/g transposition,
` fŒ.U/).W/ ^ Œ.V/).W/g)fŒ:.W/)Œ:.U/g theorem 1.16;
` fŒ.U/).W/
^ Œ.V/).W/g
) fŒ:.W/)Œ:.U/g ^ fŒ:.W/)Œ:.V/g theorem 1.55;
` fŒ:.W/)Œ:.U/g
^ fŒ:.W/)Œ:.V/g
) Œ:.W/)fŒ:.U/ ^ Œ:.V/g theorem 1.55;
` Œ:.W/)fŒ:.U/ ^ Œ:.V/g ) fŒ.U/ _ .V/g).W/ contraposition;
` fŒ.U/).W/ ^ Œ.V/).W/g) fŒ.U/ _ .V/g).W/ theorem 1.55.
t
u
Theorem 1.78 shows that an intuitionistic and minimalist axiom for the disjunc-
tion is a theorem in the Classical Propositional Calculus.
1.78 Theorem (OR.3). For all propositional forms P, Q, and R,
` Œ.P/ ) .R/ ) Œ.Q/ ) .R/ ) fŒ.P/ _ .Q/ ) .R/g :
www.pdfgrip.com
Proof. This proof relies on theorems 1.77, 1.53, 1.52, 1.67, 1.31:
` fŒ.U/).W/ ^ Œ.V/).W/g
„ ƒ‚ …
H
)fŒ.U/ _ .V/).„ƒ‚…
W /g theorem 1.77,
„ ƒ‚ …
L M
` Œ.U/ ^ .V/)Œ.U/ _ .V/ 1.53, 1.52, 1.67,
„ ƒ‚ … „ ƒ‚ …
K L
` fŒ.U/).W/ ^ Œ.V/).W/g)fŒ.U/ ^ .V/).„ƒ‚…
W /g theorem 1.31.
„ ƒ‚ … „ ƒ‚ …
H K M
t
u
The third theorem shows that if one hypothesis leads to either of two conclusions,
then the same hypothesis also leads to the disjunction of both conclusions.
1.81 Theorem. For all well-formed formulae P, Q, and S,
` fŒ.P/ ^
.R/ ) .P/g
) Œ.P/ ) .Q/ ) fŒ.P/ ^ .R/ ) .Q/g theorem 1.28,
` Œ.P/ ) .Q/ ) fŒ.P/ ^ .R/ ) .Q/g Detachment;
` fŒ.P/ ^ .R/ ) .Q/g ^ fŒ.P/ ^ .R/ ) .S/g
„ ƒ‚ … „ ƒ‚ …
H H
) fŒ.P/ ^ .R/ ) Œ.Q/ ^ .S/g theorem 1.55;
„ ƒ‚ …
H
The following theorems establish further derived rules of inference. The first
theorem forms a part of the basis for an algorithm — called the Completeness
Theorem — to design proofs within the propositional calculus, as explained in
section 1.9.
1.85 Theorem (proof by cases). If .H/ ) .R/ and Œ:.H/ ) .R/ are theorems,
then R is a theorem. Hence the following formula is also a theorem (schema):
Œ.H/ ) .R/ ^ fŒ:.H/ ) .R/g ) .R/:
Proof. Apply theorem 1.77 and the law of excluded middle (theorem 1.70):
` .H/ ) .R/ hypothesis,
` Œ:.H/ ) .R/ hypothesis,
` f.H/ _ Œ:.H/g ) .R/ theorem 1.77,
` .H/ _ Œ:.H/ theorem 1.70,
`R Detachment;
` Œ.H/ ) .R/ ^ fŒ:.H/ ) .R/g ) f.H/ _ Œ:.H/g ) .R/ theorem 1.77,
` .H/
_ Œ:.H/ theorem 1.70,
` Œ.H/ ) .R/ ^ fŒ:.H/ ) .R/g ) .R/ theorem 1.17.
t
u
Theorem 1.85 is the theoretical basis for proofs by cases, in the sense that if a
conclusion R is a necessary consequence of a case H, and if R is also a necessary
consequence of all the other cases, lumped together into :.H/, then R is a theorem.
The second theorem establishes the classical principle of proof by contradiction.
1.86 Theorem (proof by contradiction). If Œ:.R/ ) .S/ and Œ:.R/ ) Œ:.S/
are theorems, then R is a theorem: for all well-formed formulae R and S,
` fŒ:.R/ ) .S/g ^ fŒ:.R/ ) Œ:.S/g ) .R/:
` Œ:.R/)f.S/ ^ Œ:.S/g) :f.S/ ^ Œ:.S/g )f:Œ:.R/g
contraposition,
` Œ:.R/)f.S/ ^ Œ:.S/g ) fŒ:.S/ _ .S/g).R/ De Morgan,
` Œ:.S/
_ .S/ theorem 1.70 ,
` Œ:.R/)f.S/ ^ Œ:.S/g ).R/ theorem 1.17 ,
` fŒ:.R/).S/g ^ fŒ:.R/)Œ:.S/g ).R/ theorem 1.16.
t
u
Theorem 1.86 is the theoretical basis for proofs by contradiction: if a conclusion
S and its negation :.S/ are both necessary consequences of the negation :.R/ of a
statement R, then R is a theorem.
1.57 . Determine whether the following derived rule holds. “If .V/ ) .W/ and
.R/ ) .S/ are theorems, then Œ.V/ _ .R/ ) Œ.W/ _ .S/ is also a theorem”:
1.58 . Determine whether the following derived rule holds. “If .I/ ) .R/ and
.I/ ) Œ:.S/ are both theorems, then :Œ.R/ ) .S/ is also a theorem”:
Œ.I/ ) .R/ ^ f.I/ ) Œ:.S/g ) f:Œ.R/ ) .S/g:
1.59 . Determine whether the following derived rule holds. “If .U/ ) .W/ or
.V/ ) .W/ is a theorem, then Œ.U/ _ .V/ ) .W/ is also a theorem”:
1.60 . Determine whether the following derived rule holds. “If .P/ ) .Q/ or
.R/ ) .S/ is a theorem, then Œ.P/ _ .R/ ) Œ.Q/ _ .S/ is also a theorem”:
1.61 . Determine whether the following derived rule holds. “If .P/ ) .S/ and
Œ:.Q/ ) Œ:.S/ is a theorem, then .P/ ) .Q/ is also a theorem”:
Œ.P/ ) .S/ ^ fŒ:.Q/ ) Œ:.S/g ) Œ.P/ ) .Q/:
This section addresses the question how to determine whether a formula admits of
a proof from selected axioms, and, if so, how to find such a proof. The main tool to
this end consists of multi-valued logics, also called “fuzzy” logics [102].
Multi-valued fuzzy logics are “models” of propositional logics that assign a “value”
to each propositional variable, and hence also to each formulaic letter and each well-
formed propositional form, by means of a table or a formula.
www.pdfgrip.com
Table 1.94 Church’s logic :.P/ P Q .P/ ) .Q/ .Q/ ) Œ.P/ ) .Q/
with values u, w, and
distinguished value v [18, u v v v v v v
p. 113]. u v w w w v w
u v u u u v u
w w v v v v v
w w w v w v v
w w u u u v u
v u v v v v v
v u w v w v v
v u u v u v v
1.90 Definition. A value in propositional logic may be any symbol that is not
already allowed in well-formed propositional formulae. Thus a value may not be
a connective, formulaic or propositional letter, parenthesis, bracket, or brace.
For the present purposes three values suffice, denoted here by u, v, and w. A logic
with exactly three values is also called a triadic logic. A multi-valued logic also
designates any one of its values as the distinguished value, for example, v (Fig. 1.1).
1.91 Definition. A propositional form holds, or is valid or a tautology, in a multi-
valued logic if and only if it has the distinguished value regardless of the values of
its components. The notation ˆ P indicates that P is valid [72, p. 12 & p. 14].
1.92 Remark. Such software packages as John Harrison’s program [54, 55] and
Stephen Wolfram’s Mathematica [142] provide facilities to calculate and print
multi-valued Truth tables of propositional forms, as explained in [102].
1.93 Example (Church’s triadic logic). Table 1.94 defines (by fiat) the values of
the negation :.P/ and of the implication .P/ ) .Q/ from the values of P and Q in
Church’s triadic logic [18, p. 113, un-numbered table, penultimate column]. The last
column shows how to derive the values of the compound formula .Q/ ) Œ.P/ )
.Q/ from the values of its components. The formula .Q/ ) Œ.P/ ) .Q/ is valid
because it has the distinguished value v regardless of the values of P and Q.
1.95 Example (Łukasiewicz’s triadic logic). Table 1.96 defines (by fiat) the values
of the negation :.P/ and of the implication .P/ ) .Q/ from the values of P and
Q in Łukasiewicz’s triadic logic [18, 79, p. 113, un-numbered table, last column].
The last column of table 1.96 shows how to derive the values of the compound
formula .Q/ ) Œ.P/ ) .Q/ from the values of its components. The formula
.Q/ ) Œ.P/ ) .Q/ is valid because it has the distinguished value v regardless of
the values of P and Q.
1.97 Definition. A multi-valued fuzzy logic is sound if and only if all its theorems
are valid: for every formula L, if ` L (if L is a theorem), then ˆ L (then L is valid).
Every axiom is also a theorem and thus must be valid in a sound logic.
1.98 Example. Table 1.94 in example 1.93 shows that Frege’s axiom P1 (the law
of affirmation of the consequent) is valid in Church’s triadic logic. Similarly,
exercise 1.83 confirms that Frege’s axiom P2 (the law of self-distributivity of
implication) is also valid in Church’s triadic logic. Thus all the axioms of the Pure
Positive Implicational Propositional Calculus are valid in Church’s triadic logic.
If a sound logic allows for inferences by Detachment, then Detachment must be
“sound” or “preserve” valid formulae, in the sense that from valid premisses P and
.P/ ) .Q/ Detachment must produce a valid conclusion Q.
www.pdfgrip.com
1.99 Example. Table 1.94 in example 1.93 contains only one line where P and
.P/ ) .Q/ both have the distinguished value .v/: only in the first line. In that line
Q also has the distinguished value .v/. Thus Detachment preserves valid formulae
in Church’s triadic logic.
1.100 Example. Table 1.96 in example 1.95 contains only one line where P and
.P/ ) .Q/ both have the distinguished value .v/: only in the first line. In that line
Q also has the distinguished value .v/. Thus Detachment preserves valid formulae
in Łukasiewicz’s triadic logic.
Conversely, theorem 1.101 confirms that sound axioms and a sound Detachment
suffice for a propositional logic to be sound.
1.101 Theorem. Suppose that a multi-valued propositional logic is such that its
axioms are valid and Detachment from valid premisses produces a valid conclusion:
for all propositional forms H and C, if H is valid, and if .H/ ) .C/ is valid, then C
must be valid. Then such a logic is sound.
Proof (Outline). This proof shows that each step of a proof is valid. Each step C of
a proof is either an axiom or the result of Detachment. In particular, every formal
proof starts with an axiom.
If C is an axiom, then C is valid by hypothesis.
Suppose that all the steps up to but not including C have already been proved
valid. If C results from Detachment from previous steps H and .H/ ) .C/, then H
and .H/ ) .C/ are valid, and hence C is valid by the hypotheses of the theorem.
Thus every step of a proof is a valid formula. In particular, the last step of a proof
is also a valid formula.
A rigorous proof uses the Principle of Mathematical Induction in chapter 4. u t
1.102 Example. Examples 1.98 and 1.99 show that in Church’s triadic logic
Detachment is sound and all the axioms of the Pure Positive Implicational Proposi-
tional Calculus are valid. Consequently, theorem 1.101 shows that the Pure Positive
Implicational Propositional Calculus with Church’s triadic logic is sound: every
theorem of the Pure Positive Implicational Propositional Calculus is a valid formula
in Church’s triadic logic.
specific formulae, and thereby an explanation of the concept — though not the
proof — of such impossibilities as Arrow’s Impossibility Theorem in voting
presented in chapter 7, ruler-and-compass angle trisection, duplication of the cube,
or a proof of Euclid’s fifth postulate in geometry, and a solution to the decision
problem in first-order logic, or Gödel’s Incompleteness Theorem in logic.
Theorems 1.67, 1.68, and 1.78 show that the formula Œ.P/ ) .Q/ ) .Q/ uses
only the connective ) but is equivalent to .P/ _ .Q/ in the Classical Propositional
Calculus. Similarly, without negations, using only implications, Charles Sanders
Peirce’s Law
With P and Q replaced by .P/ ) .Q/ and P respectively, if the denial .P/ ) .Q/
is False, then Œ.P/ ) .Q/ ) .P/ is True, whence, by Peirce’s Law (1.1) and
Detachment, the consequent P is True. In other words, from the Falsity of the denial
of P follows the Truth of P. Equivalently, if :.P/ is False, then P must be True. In
this sense, using only the implicational connective ) Peirce’s Law (1.1) expresses
the law of the excluded middle .P/ _ Œ:.P/ [104, p. 189–190].
For each of the following formulae, determine whether it is a triadic tautology with
Church’s table 1.94, or Łukasiewicz’s table 1.96, or both, or neither.
1.77 . Œ.P/ ) Œ.Q/ ) fŒ.Q/ ) .R/ ) Œ.P/ ) .R/g.
1.78 . .P/ ) .P/.
1.79 . fŒ.P/ ) .Q/ ) .P/g ) .P/.
1.80 . Œ.P/ ) .Q/ ) f.P/ ) Œ:.Q/g ) Œ:.P/ .
1.81 . fŒ.P/ ) .Q/ ) .Q/g ) Œ.Q/ ) .P/ ) .P/g.
1.82 . Œ:.P/ ) Œ.P/ ) .Q/.
1.83 . f.P/ ) Œ.Q/ ) .R/g ) fŒ.P/ ) .Q/ ) Œ.P/ ) .R/g.
1.84 . Œ.P/ ) .Q/ ) f.P/ ) Œ:.Q/g ) Œ:.P/ .
1.85 . f.P/ ) Œ.Q/ ) .R/g ) f.Q/ ) Œ.P/ ) .R/g.
1.86 . fŒ:.Q/ ) Œ:.P/g ) Œ.P/ ) .Q/.
www.pdfgrip.com
Boolean logic is a dyadic logic: a multi-valued logic with exactly two values. The
distinguished value may be called “True” and denoted by v, whereas the other value
may be called “False” and denoted by u. Table 1.113 defines the Boolean Truth
values of compound propositional formulae with selected connectives.
Boolean logic will lead to an algorithm to design proofs in the Full Propositional
Calculus in section 1.9.
Based on [101], this subsection clarifies the Truth table of the logical implication.
True logical implications with a False hypothesis are convenient in classical
mathematics, but other versions of mathematics do not include them [33]. Moreover,
True logical implications with a False hypothesis rarely occur in practical reasoning:
Actually, the rule that any conditional is true if its antecedent is known to be false has almost
no parallel in natural logic. Examples of the type “if snow is black, then 2 2 D 5,” which
keep cropping up in textbooks, are only capable of confusing the student, since no natural
subsystem in our language has expressions with this semantics [81, p. 36].
Correspondingly, some computers and logical circuits do not include any facility to
test the Truth value of logical implications. Therefore the following four examples
serve solely to demonstrate the difference between Boolean algebraic logic and
practical reasoning. There exist other logics, but they are less used and more
complicated than Boolean logic [18, p. 142, p. 146, 26:11].
The present subsection shows that the Truth table for the Boolean logical
implication is the only Truth table that satisfies certain requirements. Specifically,
the present considerations confirm that the complete law of contraposition
and the nonequivalence of an implication .P/ ) .Q/ with its converse .Q/ ) .P/
hold only with implications defined as in table 1.113. To this end, denote by any
candidate connective for a logical implication. To reflect common experience, as in
Table 1.113 Boolean dyadic logic with values u and distinguished value v.
:.P/ P Q .P/ ) .Q/ .P/ ^ .Q/ .P/ _ .Q/ Œ.P/ NOR .Q/ .P/ , .Q/
u v v v v v u v
u v u u u v u u
v u v v u v u u
v u u v u u v v
www.pdfgrip.com
example 1.1, any concept of logical implication may have to satisfy the following
two requirements.
The first two requirements, (Implication. 1) & (Implication. 2), dictate the first two
lines of Truth table 1.114, where the hypothesis P is True. There remain only
four possibilities for in the last two rows, where the hypothesis is False. For
convenience, denote these four connectives by ), #, , and Õ respectively
(these last three symbols are used in this manner only in the present discussion).
Table 1.115 shows their Truth values. For comparison, the last two requirements,
(Implication. 3) & (Implication. 4), dictate the last two lines of the desired Truth
table.
Verifications based on table 1.113. confirm that the logical implication ) has
the same Truth values as its contraposition has, but not as its converse has.
In contrast, .P/ # .Q/ and its contraposition Œ:.Q/ # Œ:.P/ have different
Truth values, as in table 1.116. Also, for the connective #, neither the law of
contraposition nor its converse holds.
Similarly, .P/ .Q/ and its contraposition Œ:.Q/ Œ:.P/ do not have the
same Truth values, as in table 1.117. Moreover, for the connective , neither the
law of contraposition nor its converse hold.
www.pdfgrip.com
Finally, .P/ Õ .Q/ and its converse .Q/ Õ .P/ have the same Truth values, as
in table 1.118. However, for the connective Õ, both the law of contraposition and
its converse hold.
Thus the Truth table 1.113 specified for .P/ ) .Q/ is the only one that reflects
experience. There exist other concepts of logical implication, but they do not lend
themselves to Truth tables [18, p. 146, #26.12].
Some logical connectives — for instance, NOR — can combine so as to play the rôle
of every logical connective.
1.119 Definition. A logical connective is called primitive, or also universal, if
and only if every propositional form is logically equivalent to a propositional form
containing only that connective.
1.120 Example. The logical connective NOR, defined so that .A/ NOR .B/ stands
for :.A _ B/, is universal; in particular, the following equivalences are tautologies:
1.121 Example. The logical connective NAND, defined so that .A/ NAND .B/ stands
for :.A ^ B/, is universal; in particular, the following equivalences are tautologies:
1.122 Example. For logic and arithmetic, Westinghouse DPS-2402 Computers used
only NAND gates:
Its function can be considered fundamental in that all sequential and combinational logic
functions can be performed entirely by NAND gates [138, Section 4, 4–1(7)(a), p. 4–4].
1.123 Example. The Apollo spacecraft contained two electrically identical Apollo
Guidance Computers (AGC): a Command Module Computer (CMC) in the Com-
mand Module (CM), and a Lunar Module Computer (LMC) in the Lunar Module
(LM) [51, 2.1, p. 23], pictured in figure 1.2.
Each AGC used only one type of logical circuit: a NOR gate with three variables,
such that NOR .A; B; C/ is equivalent to :.A _ B _ C/ [51, 3.2.1, p. 60, fig. 3–1].
Setting C to False .0V/ shows that NOR .A; B; F/ is equivalent to :.A _ B/ and
hence to .A/ NOR .B/. By universality of NOR, every logical, arithmetic, reading,
writing, and copying operation necessary during space flight was implemented by
circuits consisting entirely and exclusively of NOR gates [51, 3.2.1, p. 62]. The
universality of NOR provides a reliability greater than several different connectives
would:
The single logic type simplified packaging, manufacturing, and testing, and gave higher
confidence to the reliability predictions [51, 1.1, p. 10].
Relying on the Deduction Theorem 1.22, the Provability Theorem and the Com-
pleteness Theorem will provide not only guidance but an algorithm to design proofs
within the propositional calculus.
The full Classical Propositional Calculus based on axioms P1, P2, and P3 is
absolutely complete: every tautology is a theorem. Moreover, there are algorithms
to determine for each propositional form whether it is a theorem, and, if it is, to
design a proof of it. The demonstration relies on the following notation.
1.124 Definition. For each proposition P, define a proposition P0 by
P if P is True;
P0 WD
:.P/ if P is False:
P0 ; : : : ; R0 ` S0 :
Proof (Outline). This proof proceeds by cases, removing from S one connective at
each step.
www.pdfgrip.com
Negation
If S is :.V/, then V has one fewer connective than S has. Suppose that P0 ; : : : ; R0 `
V 0 has already been proved, and consider two cases.
S True If S is True, then S0 is S. However, if S is True, then V is False, and V 0 is
:.V/, which is S and hence also S0 . Thus P0 ; : : : ; R0 ` V 0 by the hypothesis on
V, and substituting S0 for V 0 yields P0 ; : : : ; R0 ` S0 .
S False In contrast, if S is False, then S0 is :.S/. However, if S is False, then V is
True, and V 0 is V. Thus P0 ; : : : ; R0 ` V 0 by the hypothesis on V, and substituting
V for V 0 yields P0 ; : : : ; R0 ` V. Hence, appending a proof of the converse law
of double negation (theorem 1.43) produces a proof of P0 ; : : : ; R0 ` f:Œ:.V/g,
and substituting S for :.V/ gives P0 ; : : : ; R0 ` Œ:.S/, whence substituting S0 for
:.S/ yields P0 ; : : : ; R0 ` S0 .
Implication
If S is .V/ ) .W/, then V and W have fewer connectives than S has. Suppose that
P0 ; : : : ; R0 ` V 0 and P0 ; : : : ; R0 ` W 0 have already been proved, and consider two
cases. The first two cases occur if S is True, which occurs if W is True or V is False.
S True, W True If W is True, then W is W 0 and by the hypothesis on W there exists
a proof of P0 ; : : : ; R0 ` W. Again because W is True, it follows from theorem 1.12
that .V/ ) .W/ is also True, and appending a proof of theorem 1.12 after
the proof of P0 ; : : : ; R0 ` W produces a proof of P0 ; : : : ; R0 ` Œ.V/ ) .W/.
However, because .V/ ) .W/ is True and .V/ ) .W/ is S, it also follows that S
is S0 , whence P0 ; : : : ; R0 ` Œ.V/ ) .W/ is P0 ; : : : ; R0 ` S0 .
S True, W False If V is False, then V 0 is :.V/ and True. Thus there exists a proof
of P0 ; : : : ; R0 ` V 0 by the hypothesis on V, which is thus a proof of P0 ; : : : ; R0 `
Œ:.V/. Hence the law of denial of the antecedent (theorem 1.40) gives a proof
of Œ:.V/ ) Œ.V/ ) .W/, which is Œ:.V/ ) .S/, and thence the transitivity
of implications (theorem 1.16) yields a proof of P0 ; : : : ; R0 ` S, which is also a
proof of P0 ; : : : ; R0 ` S0 .
S False The third case occurs if S is False, which occurs if and only if V is True
and W is False. Then S0 is :.S/ and W 0 is :.W/ but V 0 is V. By the hypotheses
on V and W, there exist proofs of P0 ; : : : ; R0 ` V 0 and P0 ; : : : ; R0 ` W 0 , which
are thus proofs of P0 ; : : : ; R0 ` V and P0 ; : : : ; R0 ` Œ:.W/. Appending a proof
of theorem 1.54 then gives a proof of P0 ; : : : ; R0 ` f.V/ ˚ ^ Œ:.W/g, whence the
definition (1.51) of ^ produces a proof of P0 ; : : : ; R0 ` : .V/ ) f:Œ:.W/g .
Thence the converse law of double negation, transitivity applied to
˚
Œ.V/ ) .W/ ) .W/ ) f:Œ:.W/g ) .V/ ) f:Œ:.W/g
The general case follows by several applications of the previous cases, in a way
that may be specified more explicitly after the availability of the Principle of
Mathematical Induction in chapter 4. t
u
The Completeness Theorem shows that within the full classical propositional
calculus every tautology is a theorem, provable from the axioms and the rules of
inference.
1.126 Theorem (Completeness Theorem). Within the full classical propositional
calculus, every tautology is a theorem.
Proof. This proof uses the Deduction Theorem 1.22 and the Provability The-
orem 1.125, removing at each step one propositional variable that occurs in a
tautology.
For every tautology S with propositional variables P; : : : ; Q; R, theorem 1.125
produces a proof of P0 ; : : : ; Q0 ; R0 ` S, because S0 is S. Two cases arise with the last
variable R.
R True If R is True, then R0 is R, whence from the proof of P0 ; : : : ; Q0 ; R ` S, the
Deduction Theorem gives a proof of P0 ; : : : ; Q0 ` Œ.R/ ) .S/.
R False If R is False, then R0 is :.R/, whence from the proof of P0 ; : : : ; Q0 ; R0 ` S,
the Deduction Theorem gives a proof of P0 ; : : : ; Q0 ` fŒ:.R/ ) .S/g.
A proof of P0 ; : : : ; Q0 ` S follows by the principle of proofs by cases (theorem 1.85):
` Œ.R/ ) .S/ ^ fŒ:.R/ ) .S/g ) .S/:
The following considerations demonstrate how to plan the design of a proof by the
Completeness Theorem (theorem 1.126), here with the example of Peirce’s Law:
0
P0 ; Q0 ` fŒ.P/ ) .Q/ ) .P/g ) .P/ :
In all cases, S has the propositional form .V/ ) .W/, where W is P, and where V
is .H/ ) .K/, with .P/ ) .Q/ for H, and P for K:
S
‚ …„ ƒ
H K
‚ …„ ƒ ‚…„ƒ
fŒ.P/ ) .Q/ ) .P/ g ) .P/ :
„ ƒ‚ … „ƒ‚…
V W
Alternatively axiom P1 yields the conclusion directly, but the foregoing deriva-
tion serves to illustrate the use of the Completeness Theorem.
P True, Q False Because P is again True, the preceding reasoning remains valid
because it does not use the Truth value of Q.
P False, Q True If P is False, then so is W. Hence the Provability Theorem calls
for a proof of V 0 . Here V is Œ.P/ ) .Q/ ) .P/, which has the form .H/ ) .K/.
With P False, H is True and K is False, whence V is False. Consequently, V 0 is
:.V/, which has the form :Œ.H/ ) .K/. Therefore, the Provability Theorem
calls for proofs of P0 ; Q0 ` H and P0 ; Q0 ` Œ:.K/.
Here P0 ; Q0 ` Œ:.K/ is Œ:.P/; Q0 ` f:Œ:.P/g, which follows from the
substitution Œ:.P/ ) Œ:.P/ in the proof of theorem 1.14.
Also, P0 ; Q0 ` H is Œ:.P/; Q0 ` Œ.P/ ) .Q/, where P is False. Thus the
Provability Theorem calls for a proof of Œ:.P/; Q0 ` Œ:.P/, which again
follows from the substitution Œ:.P/ ) Œ:.P/ in the proof of theorem 1.14.
Hence Œ:.P/ ) Œ.P/ ) .Q/ by the law of denial of the antecedent
(theorem 1.40).
These proofs of Œ:.P/; Q0 ` H and Œ:.P/; Q0 ` Œ:.K/ complete the proof of
Œ:.P/; Q0 ` f:Œ.H/ ) .K/g, which is Œ:.P/; Q0 ` Œ:.V/. Again the law of
denial of the antecedent gives a proof of Œ:.P/; Q0 ` Œ.V/ ) .W/, which is
www.pdfgrip.com
Œ:.P/; Q0 ` S. The proof just obtained gives the following main steps (the final
proof replaces every theorem cited by a complete proof of that theorem).
` Œ:.P/ ) fŒ.P/ ) .Q/ ) .P/g ) .P/ substitutions.
„ ƒ‚ … „ƒ‚…
V W
P False, Q False Because P is again False, the preceding reasoning remains valid
because it does not use the Truth value of Q.
From the preceding proofs of .P/ ) .S/ and Œ:.P/ ) .S/, the principle of
proofs by cases (theorem 1.85) yields a proof of Peirce’s Law (S). Subsequent
examinations of the proof produced by the Completeness Theorem can yield
simplifications.
` .P/ ) fŒ.P/ ) .Q/ ) .P/g ) .P/ axiom P1,
` .P/ ) .S/ substitution;
1.87 . Assume that Œ:.P/ ) .P/ holds and prove that P holds. In other words,
prove that fŒ:.P/ ) .P/g ` .P/.
1.88 . Assume that .P/ ) Œ:.P/ holds and prove that :.P/ holds. In other words,
prove that f.P/ ) Œ:.P/g ` Œ:.P/.
1.89 . Apply the Deduction Theorem to prove fŒ:.P/ ) .P/g ) .P/ .
1.90 . Apply the Deduction Theorem to prove f.P/ ) Œ:.P/g ) Œ:.P/ .
1.91 . Assume .P/ ) .Q/ and prove f.P/ ) Œ:.Q/g
) Œ:.P/. In other words,
prove that Œ.P/ ) .Q/ ` f.P/ ) Œ:.Q/g ` Œ:.P/ .
1.92 . Apply the Deduction Theorem to prove the law of reductio ad absurdum
Œ.P/ ) .Q/ ) f.P/ ) Œ:.Q/g ) Œ:.P/ :
Chapter 2
First-Order Logic: Proofs with Quantifiers
2.1 Introduction
This chapter introduces quantifiers and first-order logic. The first few sections
demonstrate methods for designing proofs through preliminary versions of the
Deduction Theorem for first-order logic, Substitutivity of Equivalences, and trans-
formations into prenex forms. A final section derives features of predicates for
equality and inequality, either as primitive predicate constants, or predicates defined
from other primitive binary predicate constants. The prerequisite for this chapter
is a working knowledge of the Classical Propositional Logic for instance, as in
chapter 1.
Pure first-order logic includes quantifiers corresponding to phrases such as “for
each object” or “there exists an (at least one) object” with templates for functions
of objects and relations between objects. Applied first-order logic replaces such
templates with functions and relations specific to areas such as algebra, arithmetic,
geometry, or set theory.
In grammar, the noun “predicate” designates the verb or verbal phrase that makes a
statement about the subject of a clause. In logic, similarly, a predicate is a part of an
atomic formula that makes a statement about individual objects in applications.
The logical concept of predicate depends upon the theory under consideration.
2.1 Example (predicates in arithmetic). Some versions of arithmetic have only two
predicates, which state that a number is the sum or product of two numbers:
M D K C L (read “M equals the sum of K and L”),
N DKL (read “N equals the product of K and L”),
or equivalent formulae with a different notation [18, p. 318], [72, p. 202–203].
These predicates are called “ternary” because each involves three variables.
2.2 Example (predicates in geometry). In geometry, a predicate may state that a
point is on a line, or that a point lies between two other points on the same line, or
that two segments, or two angles, are congruent, or that a line lies in a plane:
P2L (read “the point P is on the line L”),
X<Y<Z (read “the point Y is between the points X and Z”),
PQ RS (read “the segment PQ is congruent to the segment RS”),
†ABC †PQR (read “the angle ABC is congruent to the angle PQR”),
LE (read “the line L lies in the plane E”),
or equivalent formulations with a different notation [61, Ch. I].
2.3 Example (predicates in set theory). Some versions of set theory have only one
predicate, which states that a set is an element of a set:
X2Y (read “X is an element of Y”),
¿ 2 Y (read “the empty set is an element of Y”),
X 2 ¿ (read “X is an element of the empty set”),
¿ 2 ¿ (read ‘the empty set is an element of the empty set”).
This predicate is called “binary” because it involves two variables, X and Y.
The formulae in the foregoing examples are called terms or atomic formulae
because they are the simplest formulae in arithmetic, geometry, and set theory. Thus,
X 2 Y is a term, or, in other words, an atomic formula.
In arithmetic, the symbols 0 and 1 are called individual constants, because they
always denote the numbers zero and one, respectively. Similarly, in set theory, the
symbol ¿ is an individual constant, because it always denotes the empty set.
In contrast, the symbols D and 2 are called predicate constants, because they
always denote the relations of equality and set membership, respectively. Logics that
include such constants are called applied predicate calculi; they may also include
other functional constants or relational constants corresponding to other relations
between objects. In contrast, logics that do not include any constants but allow for
variables representing arbitrary individuals, predicates, functions, and relations are
called pure predicate calculi. Thus a pure predicate calculus is a general logic that
may later apply to algebra, arithmetic, geometry, and set theory as well.
In applied logics, if an atomic formula contains a variable, then it may, but need
not, have a Truth value. For example, the formula X 2 Y has no Truth value, because
different substitutions for X and Y can yield different Truth values. However some
formulae may contain variables and yet have a Truth value.
www.pdfgrip.com
2.4 Example. In logics with an “equality” relation, the formula X D X is True for
every X.
2.5 Example. In binary arithmetic the formula 0 X D 0 is True for every X.
2.6 Example. In binary arithmetic 0 X D 1 is False for every X.
2.7 Example. In set theory the formula X 2 ¿ is False for every X.
2.8 Example. In the theory of well-formed sets the formula X 2 X is False for
every X.
Furthermore, only strings of symbols built from letters or variables through appli-
cations of the rules W1–W4 can be well-formed formulae. Equivalent definitions
apply to other connectives and to prefix and postfix notations.
In the logic presented here, only individual variables may appear immediately
after either quantifier, 8 (read “for each”) or 9 (read “there exists”). Because
of this restriction, this logic is a first order logic. Logical systems allowing for
propositional variables to appear immediately after a quantifier are of second or
higher order.
In Boolean logic, if a formula P is True regardless of X, but if P also contains
another variable Z, then substituting Z for X can change the Truth value of P.
2.10 Counterexample. Consider any context with at least two different objects,
for instance, two binary numbers in arithmetic, two points in geometry, or two
sets in set theory. Thus for each object X there exists a different object Z,
whence 8Xf9ZŒ:.X D Z/g is True. Replacing Z by X in 9ZŒ:.X D Z/ gives
9XŒ:.X D X/, which is False, because each object equals itself, by example 2.4.
Thus, replacing Z by X in the True formula 8Xf9ZŒ:.X D Z/g yields the False
formula 8Xf9XŒ:.X D X/g. Similarly, replacing X by Z in the True formula
8Xf9ZŒ:.X D Z/g gives the False formula 8Zf9ZŒ:.Z D Z/g.
One way (not pursued here) to avoid the phenomenon exhibited in counterexam-
ple 2.10 consists of substituting parameters other than variables [117]. Alternatively,
counterexample 2.10 shows that substitutions of a variable by another must obey
certain rules, for instance, with the concepts introduced in definition 2.11.
2.11 Definition (free or bound variables). For each individual variable X and for
each logical formula P, an occurrence of the variable X is bound in the formula P
if and only if in P that occurrence of the variable X immediately follows 8 or 9,
or if it appears in the scope of the quantifier, which is defined to be between either
8X. or 9X. and the corresponding right parenthesis //. An occurrence of the variable
X is free in P if and only if that occurrence of X is not bound in P. A logical formula
is closed if and only if it does not contain any free occurrence of any variable.
A logical sentence is a closed logical formula.
2.12 Example. This example focuses on the formula from counterexample 2.10.
In the formula 9ZŒ:.X D Z/, both occurrences of the variable Z are bound.
In the formula 9ZŒ:.X D Z/, the only occurrence of the variable X is free.
The formula 9ZŒ:.X D Z/ is not closed, because it contains a free occurrence of X.
form a system of three axioms for the Pure Classical Propositional Calculus.
The second axiom (schema) (Q1) corresponds to the notion that if an individual
variable X may occur in a formula P, and if P is True regardless of X, in other words,
if 8X.P/ is True, then P remains True with X replaced by any individual variable or
constant Z. If X and Z are the same variable, then axiom Q1 gives Œ8X.P/ ) .P/.
The third axiom (schema) (Q2) describes the relation between the universal
quantifier (“for each”) and the logical connective of the Pure Positive Implicational
Propositional Calculus (“if : : : then”).
The fourth axiom, for the existential quantifier (Q3), states that a formula P is
False for every X if and only if there does not exist any X for which P is True.
Similarly, the fifth axiom (schema), for the existential quantifier (Q4), states that
there exists some X for which P is False if and only if it is False that P is True
for every X. Axiom Q4 asserts the existence of an object. Consequently, axiom Q4
applies neither to “empty” theories where nothing exists, nor to logics that require
not only existence but also the determination of which objects satisfy a formula.
Besides the axioms, the predicate calculus allows for proofs of theorems through
the following rules of inference.
2.24 Definition (rules of inference). The following rules of inference hold.
2.25 Rule (“Modus Ponens” (abbreviated by M. P.), or “Detachment”).
If P is a theorem, and
if .P/ ) .Q/ is a theorem,
then Q is a theorem.
www.pdfgrip.com
Each of the following ten exercises lists one formula P. Identify a formula that is
logically equivalent to :.P/ among the same ten exercises.
2.1 . 8XŒ9Y.X 2 Y/
2.2 . 8XŒ9Y.Y 2 X/
2.3 . 8XŒ.X 2 A/ ) .X 2 B/
2.4 . 8Xf.X 2 C/ , Œ.X 2 A/ ^ .X 2 B/g
2.5 . 8Xf.X 2 C/ , Œ.X 2 A/ _ .X 2 B/g
2.6 . 9X f.X 2 C/^Œ:.X 2 A/^Œ:.X 2 B/g_fŒ:.X 2 C/^Œ.X 2 A/_.X 2 B/g
2.7 . 9Xf.X 2 A/ ^ Œ:.X 2 B/g
2.8 . 9Xf8YŒ:.Y 2 X/g
2.9 . 9X .X 2 C/ ^ fŒ:.X 2 A/ _ Œ:.X 2 B/g _ fŒ:.X 2 C/ ^ Œ.X 2 A/ ^ .X 2
B/g
2.10 . 9Xf8YŒ:.X 2 Y/g
The examples of theorems and proofs selected for this and the subsequent subsec-
tions gradually build up a tool to design proofs by substituting mutually equivalent
formulae for one another. As a first step, the following derived rules of inference
will simplify proofs by avoiding potentially lengthy instances of axiom Q2.
www.pdfgrip.com
2.33 Example. Let P denote the formula 8Xf9ZŒ:.X D Z/g, and let Q denote the
formula 8Wf9YŒ:.W D Y/g. The variables Y[ and Y[[ occur in neither P nor Q.
In P, let U[ denote the atomic formula X D Z. Then ` Œ9Z.U[ / ,
f9Y[ ŒSubfZY[ .U[ /g by theorem 2.31. Let U[[ denote the resulting formula 9Y[ Œ:.X D
Y[ /. Then theorem 2.31 shows that ` Œ8X.U[[ / , f8Y[[ ŒSubfXY[[ .U[[ /g, which is
8Y[[ f9Y[ Œ:.Y[[ D Y[ /g. The same formula results from the same procedure applied
to Q.
The following selection of theorems also relates the present axioms to other
axiom systems in subsection 2.2.8. Their proofs follow Church’s [18, p. 186–188].
2.34 Theorem. For all P, Q, and X, ` f8XŒ.P/ ) .Q/g ) fŒ8X.P/ ) .Q/g.
Proof. Apply the Implicational Calculus with axiom Q1:
` f8XŒ.P/ ) .Q/g ) Œ.P/ ) .Q/ axiom Q1,
` Œ8X.P/ ) .P/ axiom Q1,
` f8XŒ.P/ ) .Q/g ) fŒ8X.P/ ) .Q/g derived rule (theorem 1.31).
t
u
2.35 Theorem. For all P, Q, and X, ` f8XŒ.P/ ) .Q/g ) fŒ8X.P/ )
Œ8X.Q/g.
Proof. Apply the Implicational Calculus with Generalization, axiom Q2, and
theorems 2.30 and 2.34:
` f8XŒ.P/ ) .Q/g ) fŒ8X.P/ ) .Q/g theorem 2.34,
„ ƒ‚ … „ ƒ‚ …
R S
R 8XfSg
‚ …„ ƒ ‚ …„ ƒ
` f8XŒ.P/ ) .Q/g ) 8XfŒ8X.P/ ) .Q/g theorem 2.30,
` 8XfŒ8X.P/ ) .Q/g ) fŒ8X.P/ ) Œ8X.Q/g axiom Q2,
` f8XŒ.P/ ) .Q/g ) fŒ8X.P/ ) Œ8X.Q/g transitivity (1.16).
t
u
2.36 Counterexample. The converse of theorem 2.35, which would be
is False in contexts with two different objects Y and Z, so that :.Y D Z/ is True:
• 8XŒ.X D Y/ ) .X D Z/ is False, because substituting Y for X gives Œ.Y D
Y/ ) .Y D Z/, which is False, because of the True hypothesis Y D Y and the
False conclusion Y D Z.
• 8X.X D Y/ is False, because substituting Z for X gives .Z D Y/, which is False
by the assumption that :.Y D Z/.
• Œ8X.X D Y/ ) Œ8X.X D Z/ is True, because of its False hypothesis.
• fŒ8X.X D Y/ ) Œ8X.X D Z/g ) f8XŒ.X D Y/ ) .X D Z/g is False,
because of the True hypothesis and the False conclusion.
www.pdfgrip.com
2.37 Theorem (derived rule). For all P, Q, and X, if ` .P/ ) .Q/, then
` Œ8X.P/ ) .Q/ and ` Œ8X.P/ ) Œ8X.Q/ .
Proof. Apply theorems 2.34 and 2.35:
` .P/ ) .Q/ hypothesis,
` 8XŒ.P/ ) .Q/ Generalization,
` fŒ8X.P/ ) .Q/g theorem 2.34 and Detachment,
` Œ8X.P/ ) Œ8X.Q/ theorem 2.35 and Detachment.
t
u
2.38 Theorem. For all P, Q, and X, if P does not contain any free occurrence of X,
then ` f.P/ ) Œ8X.Q/g , f8XŒ.P/ ) .Q/g .
Proof. Axiom Q2 gives ` f8XŒ.P/ ) .Q/g ) f.P/ ) Œ8X.Q/g. For the
converse, use the Pure Positive Implicational Propositional Calculus with axioms Q1
and Generalization:
` Œ8X.Q/).Q/ axiom Q1,
` f.P/)Œ8X.Q/g)f.P/)Œ8X.Q/g theorem 1.14,
` f.P/)Œ8X.Q/g)Œ.P/).Q/ theorem 1.32,
` 8X f.P/)Œ8X.Q/g)Œ.P/).Q/ Generalization,
„ ƒ‚ … „ ƒ‚ …
R S
R 8XfSg
‚ …„ ƒ ‚ …„ ƒ
` f.P/)Œ8X.Q/g) 8XfŒ.P/).Q/g theorem 2.29, no free X in .P/)Œ8X.Q/.
t
u
2.39 Theorem. For all P and X, if P does not contain any free occurrence of X,
then ` Œ8X.P/ ) .P/ and ` Œ8X.P/ ( .P/ .
Proof. Axiom Q1 gives ` Œ8X.P/ ) .P/. For the converse, apply theorems 1.14
and 2.30:
` .P/ ) .P/ theorem 1.14,
` .P/ ) Œ8X.P/ theorem 2.30.
t
u
2.40 Remark. The statement of theorem 2.39 suggests that if P contains a free
occurrence of X, then the implication .P/ ) Œ8X.P/ may differ from the
Generalization rule, from ` P to infer ` 8X.P/, which applied only if P is a
theorem.
2.41 Example. If P denotes the formula X D ¿, then .P/ ) Œ8X.P/ becomes
.X D ¿/ ) Œ8X.X D ¿/, which is not a theorem. Indeed, if .X D ¿/ )
Œ8X.X D ¿/ were a theorem, then substituting ¿ for the free occurrences of X by
specialization and Detachment would yield .¿ D ¿/ ) Œ8X.X D ¿/, which is
not a theorem, because ¿ D ¿ is True while 8X.X D ¿/ is False in set theory.
Theorem 2.42 provides a converse for theorem 2.35 if X is not free in P.
2.42 Theorem. For all P, Q, and X, if X does not occur freely in P, then `
fŒ8X.P/ ) Œ8X.Q/g ) f8XŒ.P/ ) .Q/g.
www.pdfgrip.com
The following theorems invoke the full Classical Propositional Calculus, including
contraposition and its converse for negations, or Tarski’s axioms for equivalences.
2.43 Theorem. For all P, Q, and X,
` f8XŒ.P/ , .Q/g ) fŒ8X.P/ , Œ8X.Q/g.
Proof. Apply theorems 2.37 and 2.35 with the transitivity of implication:
` Œ.P/ , .Q/ ) Œ.P/ ) .Q/ definition of ,,
` f8XŒ.P/ , .Q/g ) f8XŒ.P/ ) .Q/g theorem 2.37,
` f8XŒ.P/ ) .Q/g ) fŒ8X.P/ ) Œ8X.Q/g theorem 2.35,
` f8XŒ.P/ , .Q/g ) fŒ8X.P/ ) Œ8X.Q/g transitivity.
The converse conclusion results from the symmetry of , and swapping P and Q.
The final result then follows from theorem 1.55. t
u
For the records, theorem 2.44 combines theorems 2.35, 2.42 and 2.43.
2.44 Theorem. For all P, Q, and X, if X does not occur freely in P, then `
fŒ8X.P/ ) Œ8X.Q/g , f8XŒ.P/ ) .Q/g.
Proof. Apply theorems 2.35, 2.42 and 2.43. t
u
2.45 Theorem (derived rule). For all P, Q, and X, if ` .P/ , .Q/, then
` Œ8X.P/ , Œ8X.Q/.
Proof. Apply Generalization, theorem 2.43, and Detachment:
` .P/ , .Q/ hypothesis,
` 8XŒ.P/ , .Q/ Generalization,
` f8XŒ.P/ , .Q/g ) fŒ8X.P/ , Œ8X.Q/g theorem 2.43,
` Œ8X.P/ , Œ8X.Q/ Detachment.
t
u
Theorems 2.46 and 2.47 show that 9 could be defined in terms of 8 and
double negation, or vice versa, provided that axiom Q0 includes the full Classical
Propositional Calculus.
www.pdfgrip.com
2.46 Theorem. For all P and X, ` Œ9X.P/ , :fŒ8XŒ:.P/g :
Proof. Apply the full propositional calculus and axiom Q3:
` f:Œ9X.P/g
, f8XŒ:.P/g
axiom Q3,
` :f:Œ9X.P/g , :f8XŒ:.P/g contraposition and its converse, t
u
` Œ9X.P/ , :f8XŒ:.P/g double negation and transitivity.
2.47 Theorem. For all P and X, ` Œ8X.P/ , :fŒ9XŒ:.P/g :
Proof. Apply the full propositional calculus and axiom Q4:
` f:Œ8X.P/g
, f9XŒ:.P/g
axiom Q4,
` :f:Œ8X.P/g , :f9XŒ:.P/g contraposition and its converse, t
u
` Œ8X.P/ , :f9XŒ:.P/g double negation and transitivity.
2.48 Theorem (existential generalization). For all X, Y, and P, ` ŒSubfXY .P/ )
Œ9X.P/. In particular, ` .P/ ) Œ9X.P/.
Proof. Apply the propositional calculus with axioms Q1 and Q3:
` f8XŒ:.P/g ) fSubfXY Œ:.P/g axiom Q1,
` f:Œ9X.P/g ) f8XŒ:.P/g axiom Q3,
` fSubfXY Œ:.P/g ) f:ŒSubfXY .P/g remark 2.20,
` f:Œ9X.P/g ) f:ŒSubfXY .P/g transitivity,
` ŒSubfY .P/ ) Œ9X.P/
X
converse contraposition & Detachment.
t
u
Theorem 2.49 provides a converse to theorem 2.48 if X is not free in P.
2.49 Theorem. For all P and X, if X is not free in P, then ` Œ9X.P/ , .P/.
Proof. Apply the propositional calculus with theorems 2.46 and 2.39:
` Œ:.P/
) f8XŒ:.P/g
theorem 2.39, no free X in P,
` :f8XŒ:.P/g ) f:Œ:.P/g
contraposition,
` Œ9X.P/ ) :f8XŒ:.P/g theorem 2.46,
` Œ9X.P/ ) f:Œ:.P/g transitivity,
` f:Œ:.P/g ) .P/ double negation,
` Œ9X.P/ ) .P/ transitivity.
The converse is theorem 2.48. t
u
The following exercises show that Margaris’s and Rosser’s axioms A4–A6 are
derivable from the rules of inference with axioms Q1–Q4 and the Classical
Propositional Calculus.
2.11 . Prove that the abbreviation 9X.P/ for :f8XŒ:.P/g is derivable from the
rules of inference with axioms Q1–Q4 and the Classical Propositional Calculus.
2.12 . Prove that Margaris’s and Rosser’s axioms A4, A5, and A6 are theorems
derivable from the rules of inference with axioms Q1–Q4 and the Classical
Propositional Calculus.
The following exercises show that axioms Q1–Q4 are derivable from Margaris’s
and Rosser’s axioms A4–A6 and the Classical Propositional Calculus.
2.13 . Prove that axiom Q2 is a theorem derivable from the rules of inference with
Margaris’s and Rosser’s axioms A4–A6 and the Classical Propositional Calculus.
2.14 . Prove that axiom Q1 is a theorem derivable from the rules of inference with
Margaris’s and Rosser’s axioms A4–A6 and the Classical Propositional Calculus.
2.15 . Prove that axiom Q4 is a theorem derivable from the rules of inference with
Margaris’s and Rosser’s axioms A4–A6 and the Classical Propositional Calculus.
2.16 . Prove that axiom Q3 is a theorem derivable from the rules of inference with
Margaris’s and Rosser’s axioms A4–A6 and the Classical Propositional Calculus.
The following exercises show that Kleene’s schema and rules are derivable from
the rules of inference with axioms Q1–Q4 and the Classical Propositional Calculus.
www.pdfgrip.com
2.17 . Prove that Kleene’s 9-rule is derivable from the rules of inference with
axioms Q1–Q4 and the Classical Propositional Calculus.
2.18 . Prove that Kleene’s 8-rule is derivable from the rules of inference with
axioms Q1–Q4 and the Classical Propositional Calculus.
2.19 . Prove that Kleene’s 9-schema is derivable from the rules of inference with
axioms Q1–Q4 and the Classical Propositional Calculus.
2.20 . Prove that Kleene’s 8-schema is derivable from the rules of inference with
axioms Q1–Q4 and the Classical Propositional Calculus.
If other considerations guarantee that a well-formed formula P has a proof but do not
produce any proof of it, then writing down all the proofs of the predicate calculus,
for instance, in increasing order of complexity, eventually yields among all such
proofs a proof of P [18, p. 99–100, footnote 183]. However, if the shortest proof of
P is very long, then this method may take longer than the time available to the user
to arrive at any proof of P. Thus for all practical purposes this method may also fail
to determine whether a formula is a theorem.
The problem of deciding whether a well-formed formula is a theorem, derivable
from specified axioms and inference rules, is called the decision problem. For the
pure predicate calculus, no algorithms can provide a step-by-step recipe applicable
to all well-formed formulae to determine whether any such formula is a theorem, as
proved by Church [16, 17]. Nevertheless, methods exist to help in deciding whether
a well-formed formula is a theorem.
Trial and error is an option [114, p. 31], sometimes working backward from the
particular well-formed formula as a final goal, or forward from the axioms, inference
rules, and previous theorems as starting points or intermediate steps [72, p. 54–55].
The methods presented in this section guide this method of designing proofs.
Proof (Outline). Theorems 1.29 and 1.46 have already established the conclusions
for logical implications and negations.
In all cases, if Q is P, which results by substituting none of the occurrences of U
by V, then .P/ , .Q/ is .P/ , .P/, which is theorem 1.63.
For the universal quantifier, if P is 8X.U/, then Q is either 8X.U/ or 8X.V/.
If Q is 8X.V/, then P with .U/ ) .V/ yield Q, by theorem 2.37, and conversely, Q
with .V/ ) .U/ yield P, by theorems 2.37, or also by theorem 2.45.
For the existential quantifier, theorem 2.46 reduces to the previous cases a
formula P of the form 9X.U/.
The general case follows by several applications of the previous case and the
cases in theorems 1.29 and 1.46, in a way that may be specified more explicitly
after the availability of the Principle of Mathematical Induction in chapter 4. t
u
2.55 Example. Let P denote the formula 8X Œ9Y.Y 2 X/ _ f8ZŒ:.Z 2 X/g , U
the formula 9Y.Y 2 X/, and W the formula :f8YŒ:.Y 2 X/g. Then .U/ , .W/
by theorem 2.46. Moreover, let V denote the formula :f8ZŒ:.Z 2 X/g. Because
Z does not occur in :.Y 2 X/, theorem 2.31 shows that .W/ , .V/. Hence
.U/, .V/ by transitivity. Consequently, .P/ , .Q/ whereQ denotes the formula
8X :f8ZŒ:.Z 2 X/g _ f8ZŒ:.Z 2 X/g , which is 8X :fVg _ fVg . Since
.V/ _ Œ:.V/ is a theorem, by Generalization Q and hence P is also a theorem.
Thus, the selection of axioms P1–P3 and Q1–Q4 leads to the Deduction
Theorem (2.56) more directly than would other selections of otherwise equivalent
axioms [108, p. 47]. In practice, however, a derivation of H ` C of C from H
may already suggest other logical steps that shortcut or bypass the entire procedure
outlined in the proof of the Deduction Theorem 2.56.
To demonstrate such shortcuts, the following theorems provide means for
bringing quantifiers to the “front” of a formula. For example, axioms Q3 and Q4
with theorem 2.54 already allow the replacement of :Œ9X.P/ by 8XŒ:.P/, and of
:Œ8X.P/ by 9XŒ:.P/. In an implication .R/ ) .S/, each of R and S can be of the
form .P/, or 8X.P/, or 9X.P/, starting with 8, or 9, or no quantifiers, which leads
to nine different cases. In the case where neither R nor S begins with a quantifier,
then no quantifiers need to be brought to the front of .R/ ) .S/. The other eight
cases form the object of the following theorems.
Theorem 2.57 handles a case where R is 9X.P/ while S is Q.
2.57 Theorem. If X does not occur freely in Q, then
` f8XŒ.P/ ) .Q/g , fŒ9X.P/ ) .Q/g.
Proof. Let H denote 8XŒ.P/ ) .Q/, and let C denote Œ9X.P/ ) .Q/.
` 8XŒ.P/ ) .Q/ hypothesis,
` f8XŒ.P/ ) .Q/g ) Œ.P/ ) .Q/ specialization (Q1),
` .P/ ) .Q/ Detachment,
` Œ:.Q/ ) Œ:.P/ contraposition,
` 8XfŒ:.Q/ ) Œ:.P/g Generalization,
` Œ:.Q/ ) f8XŒ:.P/g
theorem 2.29, no free X in Q,
` :f8XŒ:.P/g ) f:Œ:.Q/g contraposition,
` Œ9X.P/ ) .Q/ double negation (1.42) and 2.46.
Hence the Deduction Theorem (2.56) leads to a proof of f8XŒ.P/ ) .Q/g )
fŒ9X.P/ ) .Q/g. Yet the foregoing derivation suggests shortcuts:
` f8XŒ.P/).Q/g)Œ.P/).Q/ axiom Q1,
` Œ.P/).Q/)fŒ:.Q/)Œ:.P/g contraposition,
` f8XŒ.P/).Q/g)fŒ:.Q/)Œ:.P/g
transitivity,
` f8XŒ.P/).Q/g) 8XfŒ:.Q/)Œ:.P/g
theorem 2.30,
` 8XfŒ:.Q/)Œ:.P/g)Œ:.Q/)f8XŒ:.P/g theorem 2.29,
` Œ:.Q/)f8XŒ:.P/g
) :f8XŒ:.P/g )f:Œ:.Q/g contraposition,
` :f8XŒ:.P/g )f:Œ:.Q/g)fŒ9X.P/).Q/g 1.42, 2.46,
` f8XŒ.P/).Q/g)fŒ9X.P/).Q/g transitivity.
For the converse, let H denote Œ9X.P/ ) .Q/, and let C denote 8XŒ.P/ ) .Q/.
` Œ9X.P/ ) .Q/ hypothesis,
` .P/ ) Œ9X.P/ theorem 2.48,
` .P/ ) .Q/ transitivity,
` 8XŒ.P/ ) .Q/ Generalization.
www.pdfgrip.com
The following theorems are examples of theorems involving more than one quanti-
fier. The first theorem allows for the deletion of a redundant universal quantifier.
2.70 Theorem. ` Œ8X.Q/ , f8XŒ8X.Q/g .
Proof. Apply theorem 2.39 to 8X.Q/, which has no free X. t
u
The second theorem allows for the swap of two consecutive universal quantifiers.
2.71 Theorem. ` f8XŒ8Y.P/g , f8YŒ8X.P/g:
Proof. Apply axiom Q1, theorem 2.29, and Generalization:
` f8XŒ8Y.P/g ) Œ8Y.P/ axiom Q1,
` Œ8Y.P/ ) .P/ axiom Q1,
` f8XŒ8Y.P/g ) .P/ transitivity (theorem 1.16),
` f8XŒ8Y.P/g ) Œ8X.P/ theorem 2.37,
` f8XŒ8Y.P/g ) f8YŒ8X.P/g theorem 2.29.
t
u
The third theorem allows for the swap of two consecutive existential quantifiers.
www.pdfgrip.com
f8YŒ9X.P/g ) f9XŒ8Y.P/g;
can be False. For instance, in every context with at least two different objects V and
W, consider the logical formula X D Y for P.
` 8YŒ9X.X D Y/ for each Y, choose X WD Y;
9XŒ8Y.X D Y/ is False: no X equals V and W;
f8YŒ9X.X D Y/g ) f9XŒ8Y.X D Y/g is False because .T/ 6) .F/.
The following exercises focus on details of the proof of theorem 2.54, with the
logical equivalence , defined either by Tarski’s axioms IV, V, VI in example 1.87
on page 55, or with ) and ^ in definition 1.51 on page 38.
www.pdfgrip.com
2.21 . Prove that if P denotes .U/ ) .W/, if Q denotes .V/ ) .W/, and if `
.V/ , .U/, then ` .P/ ) .Q/.
2.22 . Prove that if P denotes .W/ ) .U/, if Q denotes .W/ ) .V/, and if `
.V/ , .U/, then ` .P/ ) .Q/.
2.23 . Prove that if P denotes .U/ ) .W/, if Q denotes .V/ ) .W/, and if `
.V/ , .U/, then ` .Q/ ) .P/.
2.24 . Prove that if P denotes .W/ ) .U/, if Q denotes .W/ ) .V/, and if `
.V/ , .U/, then ` .Q/ ) .P/.
2.25 . Prove that if P denotes 8X.U/, if Q denotes 8X.V/, without free occurrences
of X in U and V, and if ` .V/ , .U/, then ` .P/ ) .Q/.
2.26 . Prove that if P denotes 8X.U/, if Q denotes 8X.V/, without free occurrences
of X in U and V, and if ` .V/ , .U/, then ` .Q/ ) .P/.
2.27 . Prove that if P denotes :.U/, if Q denotes :.V/, and if ` .V/ , .U/, then
` .P/ ) .Q/.
2.28 . Prove that if P denotes :.U/, if Q denotes :.V/, and if ` .V/ , .U/, then
` .Q/ ) .P/.
2.29 . Prove that if P denotes either .U/ ) .W/ or .W/ ) .U/, if Q denotes either
.V/ ) .W/ or .W/ ) .V/, respectively, and if ` .V/ , .U/, then ` .P/ , .Q/.
2.30 . Prove that if P denotes :.U/, if Q denotes :.V/, and if ` .V/ , .U/, then
` .P/ , .Q/.
This subsection presents theorems involving the universal quantifier .8/ and a con-
junction .^/ or disjunction ._/, beginning with an equivalence with a conjunction.
2.75 Theorem. ` f8XŒ.P/ ^ .Q/g ) fŒ8X.P/ ^ Œ8X.Q/g.
Proof. Apply theorems 1.53, 2.37, 1.52, 1.55:
` Œ.P/ ^ .Q/ ) .P/ theorem 1.53,
` f8XŒ.P/ ^ .Q/g ) f8X.P/g theorem 2.37,
` Œ.P/ ^ .Q/ ) .Q/ theorem 1.52,
www.pdfgrip.com
may be False. For instance, in every context with exactly two different objects V and
W, consider the formulae X D V for P and X D W for Q:
` 8XŒ.X D V/ _ .X D W/ because either .X D V/ or .X D W/;
8X.X D V/ is False if X WD W;
8X.X D W/ is False if X WD V;
Œ8X.X D V/ _ Œ8X.X D W/ is False by the preceding two lines;
f8XŒ.X D V/ _ .X D W/g
) fŒ8X.X D V/ _ Œ8X.X D W/g is False because .T/ 6) .F/.
However, theorem 2.79 shows a converse of theorem 2.77 in a particular case.
2.79 Theorem. If P has no free X, then ` f8XŒ.P/ _ .Q/g ) f.P/ _ Œ8X.Q/g and
` f8XŒ.P/ _ .Q/g ) fŒ8X.P/ _ Œ8X.Q/g
www.pdfgrip.com
This subsection presents theorems involving the existential quantifier .9/ and a con-
junction .^/ or disjunction ._/, beginning with an equivalence with a disjunction.
2.80 Theorem. ` fŒ9X.P/ _ Œ9X.Q/g , fŒ9XŒ.P/ _ .Q/g:
Proof. Apply
contraposition with
theorems
2.75, 2.76, and 2.45:
` 8XfŒ:.P/ ^ Œ:.Q/g , f8XŒ:.P/g ^ f8XŒ:.Q/g 2.75, 2.76,
m contraposition,
: f8XŒ:.P/g ^ f8XŒ:.Q/g , : 8XfŒ:.P/ ^ Œ:.Q/g
m 2.45,
:f8XŒ:.P/g _ :f8XŒ:.Q/g , : 8Xf:Œ.P/ _ .Q/g
m axiom Q4,
:f:Œ9X.P/g _ :f:Œ9X.Q/g , : :f9XŒ.P/ _ .Q/g
m double negations.
fŒ9X.P/ _ Œ9X.Q/g , f9XŒ.P/ _ .Q/g
t
u
A similar equivalence with a conjunction requires that X be not free in P.
2.81 Theorem. If P has no free X, then ` f9XŒ.P/ ^ .Q/g , f.P/ ^ Œ9X.Q/g.
Proof. Apply the full propositional calculus with theorems 2.38, 2.61, 2.46, and
axiom Q4, and theorem 1.69:
9XŒ.P/ ^ .Q/
m double negation,
9X :f:Œ.P/ ^ .Q/g
m De Morgan’s first law and theorem 2.61,
www.pdfgrip.com
9X :fŒ:.P/ _ Œ:.Q/g
m definition of _,
9X :f.P/ ) Œ:.Q/g
m axiom Q4,
: 8Xf.P/ ) Œ:.Q/g
m theorem 2.38,
: .P/ ) f8XŒ:.Q/g
m definition of _ by theorem 1.69,
: Œ:.P/ _ f8XŒ:.Q/g
m De Morgan’s second law and double negation,
.P/ ^ :f8XŒ:.Q/g
m theorem 2.46.
.P/ ^ Œ9X.Q/
t
u
For the following exercises, prove that the stated formulae are theorem schema.
2.31 . f9XŒ.P/ _ .Q/g , f9XŒ.Q/ _ .P/g.
2.32 . f8XŒ.P/ ^ .P/g , f8X.P/g.
2.33 . f9XŒ.P/ _ .P/g , f9X.P/g.
2.34 . .9XfŒ.P/ _ .Q/ _ .R/g/ , .9Xf.P/ _ Œ.Q/ _ .R/g/.
2.35 . .8XfŒ.P/ ^ .Q/ _ .R/g/ , .f8XŒ.P/ _ .R/g ^ f8XŒ.Q/ _ .R/g/.
2.36 . .9XfŒ.P/ _ .Q/ ^ .R/g/ , .f9XŒ.P/ ^ .R/g _ f9Œ.Q/ ^ .R/g/.
2.37 . Œ9X.Q/ , Œ9X .9X.Q//.
2.38 . If P has no free X, then f.P/ ^ Œ8X.Q/g , f8XŒ.P/ ^ .Q/g.
2.39 . If P has no free X, then f.P/ _ Œ8X.Q/g , f8XŒ.P/ _ .Q/g.
2.40 . If P has no free X, then f.P/ _ Œ9X.Q/g , f9XŒ.P/ _ .Q/g.
2.5 Equality-Predicates
Applications of logic, for instance, algebra, arithmetic, and geometry, may include
concepts of “equality” that allow for substitutions of mutually equal objects in
statements and formulae, which results in mutually equivalent statements and
formulae.
www.pdfgrip.com
Different applications may define equality differently [8, p. 6–7]. For instance, in
some versions of integer arithmetic, the equality a D b means that a and b are
two symbols for one integer [25, p. 44], [76, p. 1]. In contrast, in some versions
of set theory, the equality A D B means that A and B denote sets with identical
set-theoretical features: they have the same elements, and they are elements of the
same sets [8, 35, p. 6–7]; the question whether A and B denote the same set does
not arise in the theory. Nevertheless, such different concepts of equality happen to
conform to a logical predicate, denoted by I to suggest identity, subject to the
following axioms (which might also be called postulates to distinguish them from
logical axioms) [18, 48].
Axiom I 2 (substitutivity of equality) ` ŒI .X; Y/ ) Œ.P/ ) .Q/ for all well-
formed formula P and Q such that Q results from the substitution of any one free
occurrence of X in P by Y, provided that the resulting occurrence of Y is also free,
or, in other words, provided that in P this occurrence of X is not within the scope of
a quantifier .8X; 8Y; 9X; 9Y/ bounding X or Y.
Using only the Pure Positive Implicational Propositional Calculus, theorem 2.84
extends axiom I 2 to a converse implication, so that substituting mutually equal
objects results in mutually equivalent formulae.
2.84 Theorem (substitutivity of equality). ` ŒI .X; Y/ ) Œ.P/ , .Q/ for all
well-formed formula P and Q such that Q results from the substitution of any one
free occurrence of X in P provided that in P this occurrence of X is not within the
scope of a quantifier .8X; 8Y; 9X; 9Y/ bounding X or Y.
Proof. The implication ` ŒI .X; Y/ ) Œ.P/ ) .Q/ is axiom I 2.
For the converse, the hypothesis also states that in Q the resulting occurrence
of Y is not within the scope of a quantifier .8X; 8Y; 9X; 9Y/ bounding Y or X,
which allows swapping X and Y, and swapping P and Q, in axiom I 2, so that
` ŒI .Y; X/ ) Œ.Q/ ) .P/. Hence the conclusion follows from the symmetry
` ŒI .X; Y/ ) ŒI .Y; X/ by theorem 2.82 and the transitivity of implication. ut
Repeated applications of theorem 2.84 and the proof of substitutivity of equiv-
alence then show that substituting mutually equal objects in a formula leads to an
equivalent formula.
Some applications of logic might omit all propositional variables and instead have
only atomic formulae with a few predicates, or perhaps only one predicate, which
might be denoted by some constant E . Such applications are called simple applied
predicate calculi. For instance, a version of set theory has no propositional variables
and only one predicate, for set membership, so that E .X; Y/ stands for X 2 Y.
In such applications, an additional equality predicate I allows for substitutions
of mutually equivalent objects in statements and formulae if and only if I is
reflexive (a condition that replaces axiom I 1), symmetric, transitive, and satisfies
the following two conditions, which replace axiom I 2 [18, p. 283, exercise 48.3]:
In applied logics with other predicates, for instance, predicates for the sum and
products of integers in arithmetic, two similar conditions must be appended for each
predicate to ensure the substitutivity of mutually equal objects. By the postulated
symmetry of the equality predicate I these two conditions are equivalent to
These conditions suffice to ensure that if I .A; B/ holds, then substituting any free
occurrence of A for any free occurrence of B according to the conditions stipulated
by axiom I 2 in any formula P produces an equivalent formula Q, because well-
formed formulae include only atomic formulae of the form E .Z; W/. The proof
follows the pattern of the proof of the substitutivity of equivalence. The resulting
theorem is called the substitutivity of equality.
In particular, if a simple applied predicate calculus has exactly one predicate,
E , which is binary (involving exactly two individual variables), then the same
conditions may serve to define an equality predicate I so that I .A; B/ is merely
an abbreviation for
8XfŒE .X; A/ , ŒE .X; B/g ^ 8YfŒE .A; Y/ , ŒE .B; Y/g : (2.1)
The equality I .A; B/ of sets A and B is then an abbreviation of each of the formulae
8XfŒE .X; A/ , ŒE .X; B/g and 8YfŒE .A; Y/ , ŒE .B; Y/g.
The following theorems confirm that every equality-predicate defined by for-
mula (2.1) is reflexive, symmetric, and transitive.
2.86 Theorem (reflexivity of defined equality-predicates). Every equality-
predicate I .A; B/ defined by formula (2.1) is reflexive:
Proof. One method to design a formal proof transforms the objective, here the
yet unproved formula 8CŒI .C; C/, first into its defining formula (2.1), and then
into logically equivalent formulae, for instance, in prenex form, until one such
equivalent formula appears that is a theorem, thanks to an axiom or to a previously
proven theorem. For instance, substituting C for A and also C for B in the defining
formula (2.1) gives
I .C; C/ yet unproved,
m definition of I
8XfŒE .X; C/ , ŒE .X; C/g ^ 8YfŒE .C; Y/ , ŒE .C; Y/g
m theorem 2.75,
8X8Y fŒE .X; C/ , ŒE .X; C/g ^ fŒE .C; Y/ , ŒE .C; Y/g
www.pdfgrip.com
where each logical formula ŒE .W; Z/ , ŒE .W; Z/ has the pattern of the reflexivity
of the logical implication .P/ , .P/ (theorem 1.63). Thus, a complete proof may
proceed as follows:
` .P/,.P/ theorem 1.63,
` ŒE .X; C/,ŒE .X; C/ substitution in .P/,.P/,
` ŒE .C; Y/,ŒE .C; Y/ substitution in .P/,.P/,
` fŒE .X; C/,ŒE .X; C/g ^ fŒE .C; Y/,ŒE .C; Y/g theorem 1.54,
` I .C; C/ formula (2.1).
Hence ` 8CŒI .C; C/ results by Generalization and theorem 2.75. t
u
2.87 Theorem (symmetry of defined equality-predicates). Every equality-
predicate defined by formula (2.1) is symmetric: if ` I .A; B/, then ` I .B; A/;
moreover.
Proof. One method to design a formal proof transforms the objective, here the
yet unproved formula ` 8A8BfŒI .A; B/ ) ŒI .B; A/g, first into its defining
formula (2.1), and then into logically equivalent formulae, for instance, in prenex
form, until one such equivalent formula appears that is a theorem, thanks to an
axiom or to a previously proven theorem. Here an equivalence will emerge:
ŒI .A; B/,ŒI .B; A/ yet unproved,
m definition of I
8XfŒ E .X; A/,ŒE .X; B/g ^
8YfŒ E .A; Y/,Œ E .B; Y/g
, 8XfŒE .X; B/,ŒE .X; A/g ^ 8YfŒE .B; Y/,ŒE .A; Y/g ;
which suggests invoking the symmetry of the logical equivalence Œ.P/ , .Q/ ,
Œ.Q/ , .P/ (theorem 1.64). Thus, a complete proof may proceed as follows:
` Œ.P/,.Q/,Œ.Q/,.P/ theorem 1.64,
` fŒE .X; A/,ŒE .X; B/g,fŒE .X; B/,ŒE .X; A/g substitution,
` Œ.R/,.S/,Œ.S/,.R/ theorem 1.64,
` fŒE .A; Y/,ŒE .B; Y/g,fŒE .B; Y/,ŒE .A; Y/g substitution,
` fŒ.P/,.Q/ ^ Œ.R/,.S/g,fŒ.Q/,.P/ ^ Œ.S/,.R/g theorem 1.82.
Hence the conclusion results by Generalization and theorem 2.75. t
u
2.88 Theorem (transitivity of defined equality-predicates). Every equality-
predicate defined by formula (2.1) is transitive: if ` I .A; B/ and ` I .B; C/,
then ` I .A; C/; moreover,
` 8A8B8C fŒI .A; B/ ^ ŒI .B; C/g ) ŒI .A; C/ :
Proof. One method to design aformal proof transforms the objective, here the yet
unproved formula ` 8A8B8C fŒI .A; B/ ^ ŒI .B; C/g ) ŒI .A; C/ , first into
its defining formula (2.1), and then into logically equivalent formulae, for instance,
www.pdfgrip.com
in prenex form, until one such equivalent formula appears that is a theorem, thanks
to an axiom or to a previously proven theorem.
fŒI .A; B/ ^ ŒI .B; C/g ) ŒI .A; C/ yet unproved,
m definition of I
8XfŒ E .X; A/,Œ E .X; B/g ^
8YfŒ E .A; Y/,Œ E .B; Y/g
^ 8XfŒE .X; B/,ŒE .X; C/g ^ 8YfŒE .B; Y/,ŒE .C; Y/g
) 8XfŒE .X; A/,ŒE .X; C/g ^ 8YfŒE .A; Y/,ŒE .C; Y/g ;
which suggests invoking the transitivity of the logical equivalence (theorem 1.65):
with the commutativity and associativity of the logical conjunction (theorems 1.57
and 1.66) combined with theorem 1.82:
fŒ.H/ , .K/ ^ Œ.K/ , .L/g ^ fŒ.P/ , .Q/ ^ Œ.Q/ , .R/g
) fŒ.H/ , .L/ ^ Œ.P/ , .R/g:
For every binary predicate variable or predicate constant F , involving only two
individual variables,
˚
` 8X8Y8W8Z ŒI .X; Y/ ) ŒI .W; Z/ ) fŒF .X; W/ ) ŒF .Y; Z/g :
www.pdfgrip.com
For every ternary predicate variable or predicate constant F , involving only three
individual variables,
` 8U8V8X8Y8W8Z
˚
ŒI .U; V/) ŒI .X; Y/) ŒI .W; Z/)fŒF .U; X; W/)ŒF .V; Y; Z/g :
Similar stipulations also hold for predicate variables or constants involving more
than three individual variables.
Formula (2.3) is also denoted by A B (read “A precedes B”) instead of R .A; B/.
The resulting predicate R is reflexive and transitive, but not necessarily symmetric,
as verified in the exercises.
2.41 . Verify that the ranking-predicate R .A; B/ defined by formula (2.3) is reflex-
ive: prove ` 8AŒR .A; A/.
2.42 . Investigate whether the ranking-predicate R .A; B/ defined by formula (2.3)
is symmetric: is 8A8BfŒR .A; B/ ) ŒR .B; A/g is a theorem?
2.43 . Verify that the ranking-predicate
R .A; B/ defined by formula
(2.3) is transi-
tive: prove ` 8A8B8C ŒR .A; B/ ^ ŒR .B; C/g ) ŒR .A; C/ .
Exercises 2.45, 2.44, and 2.46 focus on the alternative ranking predicate A .A; B/
defined in terms of the same binary predicate constant E as an abbreviation of
formula (2.4):
Chapter 3
Set Theory: Proofs by Detachment,
Contraposition, and Contradiction
3.1 Introduction
This chapter introduces set theory from two parallel perspectives: as an intuitive
mathematical theory, and as a simple applied predicate calculus of first order.
Starting from first-order logic and some of the Zermelo-Fraenkel axioms (exten-
sionality, empty set, pairing, power set, separation, and union), where all objects
under consideration are sets, the chapter first derives relations between sets, subsets,
supersets, unions, intersections, and Cartesian products of sets of sets. Subsequent
sections introduce relations, functions, injections, surjections, bijections, composite
functions, and inverse functions. Another section focuses on the duality between
partitions and equivalence relations. The last section deals with pre-orders, partial
orders, linear or total orders, and well-orders. Many proofs begin with an informal
intuitive proof, then demonstrate how to design a more formal proof, and finally
present a detailed outline of such a formal proof in first-order logic. The other
Zermelo-Fraenkel axioms (choice and infinity or substitution) are only mentioned
here, because they form the topic of subsequent chapters. The prerequisites for
this chapter consist of a working knowledge of first-order logic, for instance, as
described in chapters 1 and 2, which contain all the logical theorems cited in this
chapter.
For some practical problems, features that are essential to their solutions can be
specified in terms of sets or collections of objects.
3.1 Example (Binary arithmetic). The binary arithmetic of computers relies on a
set of two symbols, 0 and 1, which will be defined with yet other sets in this chapter.
3.2 Example (Geometries). Geometries can be designed entirely with sets. Points
are sets (of sets of coordinates), while lines, planes, and space are sets of points.
Points, lines, planes, and space are “primitive” objects that may remain undefined,
but relations between them are specified through axioms. For instance, the first
axiom of incidence geometry specifies that through any two distinct points passes
exactly one line [61, p. 3]. Likewise in this chapter, mathematical “sets” are
“primitive” objects that remain undefined, while features of sets and relations
between sets are specified by axioms.
The foregoing examples already demonstrate a major difficulty in using problems
about “real” objects to illustrate logical and mathematical concepts: no exact
answers might be available. For an example as elementary as binary arithmetic,
electronic digital computers internally do not use anything like the symbols 0 and
1: indeed, they use two electrical potentials confined to two mutually exclusive
ranges, neither of which need contain any zero [48, p. 642], with different ranges on
different machines [51, p. 60, fig. 3–1; p. 83], [138, p. 4–5, 1], in the reverse order
on other machines [103, p. 1–4], sometimes reversing the order within the same
machine [23, 5]. A precise answer would involve more advanced engineering,
logic, mathematics, and physics. Therefore, the sets in the present exposition will
not contain “real” objects; instead, all the following sets will contain only abstract
objects defined by precise rules. The judicious use of such abstract mathematical
sets in applied disciplines from astronomy to zoology is precisely the task of such
disciplines.
This section introduces mathematical sets by means of the concept of set member-
ship. The predicate of set membership then allows for the definition of the concepts
of subset, superset, and a derived predicate of equality.
One of the major mathematical achievements around the beginning of the twentieth
century was the realization that most of mathematics and computer science consists
of logical relations between abstract objects called sets [8, p. 3], [83]. There is no
definition of mathematical sets. Indeed, such a definition would have to define sets in
terms of yet more foundational objects, but sets are the most foundational objects.
Henceforth, here and as in other texts [22, p. 50], [141, p. 60], all mathematical
objects are sets, and all quantified variables designate sets:
andere Objekte als Mengen existieren für uns überhaupt nicht
(“for us objects other than sets simply do not exist”) [36, p. 271].
X 2 A;
:.X 2 A/I
A 2 Y;
:.A 2 Y/:
Yet another way to state that two sets have exactly the same characteristics
involves a derived binary relation (derived from 2) denoted by D and called
“equality.” For each set A and each set B, the formula A D B (read “A equals
B”) means that A and B have exactly the same characteristics. By the axiom of
extensionality, the equality of two sets can be stated in two logically equivalent
ways:
˚
` 8A 8B .A D B/ , f8XŒ.X 2 A/ , .X 2 B/g ;
˚
` 8A 8B .A D B/ , f8YŒ.A 2 Y/ , .B 2 Y/g :
The notation A D B is a shorthand to state that the following two formulae hold:
8XŒ.X 2 A/ , .X 2 B/;
8YŒ.A 2 Y/ , .B 2 Y/:
The axiom of extensionality states that these two formulae are logically equivalent.
A variation consists in defining A D B as an abbreviation of the first formula,
8XŒ.X 2 A/ , .X 2 B/, [8, p. 4–5], [36, p. 272–273, Def. 2], and then in adopting
the axiom that if A D B then the second formula holds: 8YŒ.A 2 Y/ , .B 2 Y/
[36, p. 274, Axiom I].
There is another presentation of set theory with two undefined relations, equality
.D/ and membership .2/. Then the axiom of extensionality specifies that two sets
are the “same” set if and only if they contain the “same” elements. In this exposition
the distinction just made does not matter, because equality .D/ serves only as a
shorthand: all operations with sets pertain to elements of those sets. (See also the
discussions by Bernays [8, p. 53] and Fraenkel [8, p. 6–8].) For the negations of
membership and equality, the following abbreviations prove convenient.
3.3 Definition. The symbols … and 6D denote the negations of 2 and D so that
At the elementary stage of set theory, most formal logical proofs of relations
between sets are straightforward, in the sense that they use only axioms and
definitions to establish a sequence of equivalences between the objective of the
proof and a theorem or universally valid formula. Such formal logical proofs are
usually longer than “informal” proofs. To show a first example of a proof within set
theory — a formal version and an informal version — the following theorem states
that each set equals itself.
www.pdfgrip.com
3.4 Remark. In designing a proof we may at any stage start from the conclusion
— but we may not assume it as True — and then search for logically equivalent
formulae that connect the conclusion to other formulae that we know how to prove.
Smullyan’s method of tableaux uses such an approach [117, Ch. II, p. 15–30].
3.5 Theorem. Each set equals itself: the formula 8S .S D S/ is universally valid.
Proof. An informal proof can consist of the following statements.
• Every set X is an element of S if and only if X is an element of S;
• hence S D S by definition of the equality of sets and extensionality (S1).
One method to design a formal proof transforms the objective, here the yet unproved
formula 8S .S D S/, into logically equivalent formulae, until one such equivalent
formula appears that is a theorem thanks to an axiom or to a previously proven
theorem. For instance, substituting S for A and also S for B in the axiom of
extensionality gives
SDS yet unproved,
m definition of D
8XŒ.X 2 S/ , .X 2 S/,
which is in prenex form, and where the logical formula .X 2 S/ , .X 2 S/ has the
pattern of the theorem .P/ , .P/. Thus, a complete proof may proceed as follows:
` .P/ , .P/ reflexivity of equivalence (theorem 1.63),
` .X 2 S/ , .X 2 S/ substitution in the theorem .P/ , .P/,
` 8Sf8XŒ.X 2 S/ , .X 2 S/g generalizations, first on X, then on S,
` 8S.S D S/ definition of D and extensionality (S1).
The proof just presented relies on one of the two formulae for the axiom of
extensionality: 8XŒ.X 2 A/ , .X 2 B/. Another proof could rely on the other
formula: 8YŒ.A 2 Y/ , .B 2 Y/. t
u
More generally, by the properties of a logical equality-predicate derived from a
binary predicate, here 2, as explained in section 2.5, the equality of sets is
• reflexive: ` 8S.S D S/ (every set equals itself), also proved in theorem 3.5,
• symmetric: ` 8A8BŒ.A D B/ ) .B D A/ (if A D B, then B D A),
• transitive: ` 8A8B8CfŒ.A D B/ ^ .B D C/ ) .A D C/g (if A D B and B D C,
then A D C),
and equality also allows substitutions of mutually equal sets in theorems.
The axiom of extensionality merely provides two logically equivalent criteria to
test whether sets have exactly the same characteristics. However, so far in this theory
there is no “set” yet. The “existence” of at least one set — or, more accurately, a
convention about an abstract concept of a specific set — requires a second axiom.
www.pdfgrip.com
The second axiom, called the axiom of the empty set, guarantees the existence of
at least one set, denoted by ¿ or also by f g; this set contains no element.
Set theory could dispense with the constant ¿ and state the axiom of the
empty set in the alternative form ` 9EŒ8X.X … E/. In the present theory, this
alternative form is a consequence of axiom S2. Indeed, with P denoting the formula
8X.X … E/, theorem 2.48 becomes
The alternative form ` 9EŒ8X.X … E/ is more cumbersome, because it does not
provide a name for any such set.
The determination of the Truth value of an equality A D B requires prior
definitions of both sets A and B. In contrast, there exists a different use of the
same concept of equality, denoted by C WD D (read “let C equal D”) to specify
a hitherto undefined set C in terms of an already defined set D [59, p. 8], [121, p. 5].
Alternatively, the notation D DW C (also read “let C equal D”) can also serve to
specify C in terms of D, especially where a derivation leads to a lengthy formula D,
which can thus be abbreviated by a shorter variable or string C [121, p. 271, p. 347].
3.6 Example. The notation E WD ¿ specifies that E stands for ¿.
In some circumstances, only some of the elements of a set prove useful; the
following definition then allows for the grouping of all such elements into a “subset.”
3.7 Definition (Subsets and supersets). For each set A, for each set B, the set A is
a subset of the set B if and only if each element of A is also an element of B. Either
notation A B or A j B indicates that “A is a subset of B”; thus,
˚
` 8A 8B .A B/ , f8XŒ.X 2 A/ ) .X 2 B/g :
formula, for each set X the logical implication .)/ states that if X is an element of
A, then X is also an element of B.
The concept of subset is so different from the concept of element as to warrant
different terminologies, for instance, reading A B as “A is a subset of B” but
reading A 2 B as “A is an element of B.” In contrast, such vague phrases as “A is
in B” or “B contains A” do not have any significance, unless they are supplemented
with “as an element” or “as a subset” [36, p. 272].
There also exist symbols more specific than A B. For instance, A B or A ¨ B
or A ¤ B indicate that A is a subset of B different from B; thus,
˚
` 8A 8B .A ¤ B/ , .A ¨ B/ , .A B/ , Œ.A B/ ^ .A 6D B/ :
m definition of
f8XŒ.X 2 A/ ) .X 2 B/g ^ f8XŒ.X 2 B/ ) .X 2 C/g
) f8XŒ.X 2 A/ ) .X 2 C/g
* theorem 2.35,
˚
8X fŒ.X 2 A/ ) .X 2 B/g
^ fŒ.X 2 B/ ) .X 2 C/g
) Œ.X 2 A/ ) .X 2 C/
which is in prenex normal form, with a matrix of the type
which is another form of the transitivity of the implication (theorem 1.76). Thus the
last line is a theorem. Reversing the order of the steps and inserting 8A and 8B
before each step (generalizing) then completes the proof. t
u
3.10 Theorem (anti-symmetry of ). Two sets are subsets of each other if and
only if they equal each other:
` 8A 8Bf.A D B/ , Œ.A B/ ^ .B A/g :
m definition of subset,
m definition of ,,
8XŒ.X 2 A/ , .X 2 B/
ADB
Inserting 8A and 8B before each step (generalizing) then completes the proof. t
u
The reflexivity, anti-symmetry, and transitivity of the concept of subset also result
more generally from the properties of a logical ranking-predicate, here , derived
from a binary predicate, here 2, as explained in section 2.5.
The next theorems focus on the subsets and supersets of the empty set.
3.11 Theorem. The empty set is a subset of every set: ` 8S.¿ S/.
Proof. An informal proof can use the converse law of contraposition (axiom P3):
• Every set not in S is also not in ¿, because no set belongs to ¿;
• the contraposition then means that every element in ¿ is also in S;
• hence ¿ is a subset of S, by the definition of subsets.
The design of a more formal proof can transform the set-theoretic formula ¿ S
into a logical formula, and verify that it is a logical theorem.
8S.¿ S/ yet unproved,
m definition of
8S f8XŒ.X 2 ¿/ ) .X 2 S/g
m contraposition and theorem 2.45
8S f8XŒ.X … S/ ) .X … ¿/g
which is in the prenex form 8S8XŒ.P/ ) .Q/, where the matrix .P/ ) .Q/ is a
theorem, because so is Q. Thus a complete proof may consist of the following steps.
www.pdfgrip.com
3.1 . Write a logical formula stating that a set S is not the empty set.
3.2 . Write a logical formula stating that a set A is not a subset of B.
3.3 . Write a logical formula stating that a set A is not a superset of B.
3.4 . Write a logical formula stating that a set A is not equal to a set B.
3.5 . Prove that ¿ 2 ¿ is not a theorem.
3.6 . Prove that ¿ ¿ is a theorem.
3.7 . Prove that ¿ D ¿ is a theorem.
3.8 . Use the second formula, 8YŒ.A 2 Y/ , .B 2 Y/, in the axiom of
extensionality to write a proof that S D S for each set S.
3.9 . Prove that ¿ is the only subset of ¿: ` 8SŒ.S ¿/ ) .S D ¿/.
www.pdfgrip.com
.A Z/ , .B Z/:
3.18 . For all sets A and B, prove that A B if and only if for each set Z
.B Z/ ) .A Z/:
3.19 . For all sets C and D, prove that C D if and only if for each set W
.D W/ ) .C W/:
3.20 . For all sets R and S, prove that R S if and only if for each set W
.W 2 S/ ) .W 2 R/:
This section introduces axioms to form sets with one or two elements (pairing), all
subsets of a set (power), or selections of specific elements into a subset (separation).
3.3.1 Pairing
A theory allowing for sets other than the empty set requires additional axioms. For
instance, the axiom of pairing states that for every set H and every set K, there
exists a set L, also denoted by fH; Kg, which contains only the elements H and K.
In the axiom of pairing (S3), the equivalence .,/ states that a set X is an element
of L if and only if X equals H or X equals K. Because the logical “or” is inclusive,
the axiom of pairing thus allows fH; Kg to contain both H and K. Moreover, because
the logical “or” commutes, the order in which H and K appear does not matter.
3.13 Theorem. For each set H and for each set K, ` fH; Kg D fK; Hg.
Proof. An informal proof can compare the elements of fH; Kg and fK; Hg:
• The set fH; Kg contains the elements H and K, but no other element;
• the set fK; Hg contains the elements K and H, but no other element;
• thus fH; Kg and fK; Hg contain exactly the same elements;
• therefore fH; Kg D fK; Hg by the axiom of extensionality (S1).
A formal proof can use ` Œ.P/ _ .Q/ , Œ.Q/ _ .P/ (theorem 1.79) to show that
3.17 Remark. The distinction between ¿ and f¿g is crucial. With˚ different
˚ nota-
tions, to appreciate better the difference between ¿ D f g and ¿ D f g ,
consider ¿ as an empty bag; then f¿g is a bag f: : :g with another empty bag ¿
inside it, also known as a “double bag” in the market place. There, a single empty
bag f g D ¿ might not be sufficiently strong to hold a six-pack of heavy glass
bottles filled with your favorite beverage. (Bottled
˚ water,
˚of course,
what were you
thinking?) That’s why you ask for a double bag ¿ D f g in which to put and
then carry the heavy six-pack. If the six-pack also n˚ comes with
n˚ a wrapping, then
o o
the combined packaging becomes a “triple bag”: ¿ D f g , which is yet
another set.
Theorem 3.18 confirms that the sets ¿ and f¿g have different characteristics.
3.18 Theorem. The sets ¿ and f¿g are two distinct sets: ¿ 6D f¿g.
Proof. An informal proof can utilize substitutions in previous axioms and
theorems:
• By definition of the empty set, ¿ … ¿ (by a substitution in axiom S2);
• moreover, ¿ 2 f¿g (by a substitution in theorem 3.14);
• hence the two sets ¿ and f¿g have different elements: ¿ 2 f¿g but ¿ … ¿;
• consequently, ¿ 6D f¿g, by the axiom of extensionality.
The following formal proof uses contraposition in the theorems
` .P/ ) fŒ.P/ ) .Q/ ) .Q/g law of assertion (theorem 1.38),
` .P/ ) Œ:.Q/ ) f:Œ.P/ ) .Q/g theorem 1.54.
whence if P and :.Q/ are theorems, then so is :Œ.P/ ) .Q/ by Detachment:
For sets with more than two elements, a new axiom becomes necessary. The axiom
of the power set states that for each set A, the collection of all subsets of A forms a
new set, denoted by P or by P.A/ and called the power set of A:
In the axiom of the power set (S4), the equivalence .,/ states that a set S is an
element of the power set P.A/ if and only if S is a subset of the set A.
3.21 Example. The empty set ¿ is the only subset of itself, by theorem 3.12. Hence
its power set has only one element, ¿, so that P.¿/ D f¿g:
A set A and its power set P.A/ might have no elements in common. Theo-
rem 3.22 shows that for every element X of A, the singleton fXg is an element
of P.A/.
3.22 Theorem. For all sets A and X, if X 2 A, then fXg 2 P.A/, and conversely:
˚
8A8X .X 2 A/ , ŒfXg 2 P.A/ :
Proof. An informal proof can rely on the definitions of subsets and power sets.
• X 2 A if and only if fXg A, by definitions of fXg and subsets;
• fXg A if and only if fXg 2 P.A/ by definition of power sets.
A more formal proof carries out similar verifications from the formal definitions.
8A 8Xf.X 2 A/ , ŒfXg 2 P.A/g yet unproved,
m definition of power sets,
8A 8Xf.X 2 A/ , ŒfXg Ag
˚ m definition of subsets,
8A 8X .X 2 A/ , f8ZŒ.Z 2 fXg/ ) .Z 2 A/g
˚ m definition of fXg (pairing),
8A 8X .X 2 A/ , f8ZŒ.Z D X/ ) .Z 2 A/g
˚ m theorems 2.34 and 2.38,
8A 8X8Zf.X 2 A/ , Œ.Z D X/ ) .Z 2 A/g
which holds by extensionality and substitutivity of equality. t
u
Theorem 3.23 shows that the power set of a singleton has exactly two elements.
˚
3.23 Theorem. For each set H, P.fHg/ D ¿; fHg .
Proof. An informal proof can list the subsets of fHg by cases:
www.pdfgrip.com
The axioms of the empty set (S2), of pairing (S3), and of the power set (S4) allow
for increasingly large sets, for instance,
¿;
P.¿/ D f¿g;
˚
P.f¿g/ D n ¿; f¿g ˚;
˚ ˚ o
P ¿; f¿g D ¿; f¿g; f¿g ; ¿; f¿g :
::
:
For each set A and for each logical formula P, the axiom schema of separation
“separates” or “sets aside” from A a subset S consisting of all the elements X of A
for which P is True. The logical formula P may not involve any free occurrence
of the variable S, but it may involve free occurrences of X and of other variables,
as indicated in the following axiom by the ellipsis : : :, which may represent other
free variables. Such a “separation rule” to form subsets differs from other axioms,
because the logic used here allows for the quantification of the elements, with the
symbols 8X, but this logic has no provision for the quantification of formulae: it
does not allow for expressions like 8P with P standing for formulae. Thus, the
“separation rule” provides a schema for infinitely many axioms, in effect one axiom
for each logical formula, whence its name.
Axiom S5 (Axiom schema of separation) For each set A, and for each logical
formula P that does not contain any free occurrence of the variable S, there exists a
subset S A that consists of only those elements X 2 A for which P is True:
` 8A 9S 8X f.X 2 S/ , Œ.X 2 A/ ^ .P/g :
S D fX W .X 2 A/ ^ .P/g;
S D fX 2 A W Pg:
An alternative notation replaces the colon (W) by a vertical bar (j), but in the context
of further mathematics such vertical bars become difficult to recognize against
absolute values, norms, and other similar symbols.
3.26 Example. Consider the set
A WD P.f¿;
˚ f¿gg/
D ¿; f¿g; ff¿gg; f¿; f¿gg ;
The existence of this set S would not have followed from the previous axioms.
www.pdfgrip.com
S D fX W .X 2 A/ ^ .X 2 B/g;
S D fX W .X 2 A/ ^ Œ:.X 2 B/g;
S D fX W .X 2 A/ ^ Œ.X 2 B/ _ .X 2 C/g;
) 9B8Y .Y 2 B/ , f9X.X 2 A/ ^ ŒSubfW Y .Q/g :
This formula can be better described with the concept of “mathematical function”
defined in section 3.6. The axiom of separation follows from the axiom of
replacement by substituting .P/ ^ .W D X/ for Q.
www.pdfgrip.com
3.35 . For theorem 3.13, explain how the word “and” in the informal proof
corresponds to the logical connective _ in the formal proof.
3.36 . For each set S, prove that f¿; Sg P.S/.
3.37 . Prove
that two sets equal each other if and
only if they have the same power
sets: ` 8A 8B f.A D B/ , ŒP.A/ D P.B/g :
˚
3.38 . Identify the set f¿g \ ¿; f¿g .
˚ ˚
3.39 . Identify the set f¿g \ ¿; f¿g .
˚
3.40 . Identify the set f¿g \ f¿g .
˚
3.41 . Identify the set f¿g [ ¿; f¿g and one of its supersets.
˚ ˚
3.42 . Identify the set f¿g [ ¿; f¿g and one of its supersets.
3.43 . For each set S, identify the set S \ ¿.
3.44 . For each set S, identify the set S [ ¿.
3.45 . Prove that ` 8AŒ.A n ¿/ D A.
www.pdfgrip.com
Axiom
S S6 (Axiom of union) For each set F , there exists a set U, also denoted by
F , which consists of all the elements that belong to any element of F :
˚
` 8F 9U 8X .X 2 U/ , f9SŒ.S 2 F / ^ .X 2 S/g :
In the axiom ofSunion (S6), the equivalence .,/ states that a set X is an element
of the union U D F if and only if there exists an element S of F such that X 2 S.
S
3.31 Example. If F D ¿, then . ¿/ D ¿, because in the axiom of union (S6)
the condition
S .S 2 F / is False for everySset S, and then the equivalent condition
X 2 . ¿/ is False for every set X. Thus ¿ contains no element.
For the union of two sets, a special notation proves convenient.
S
3.32 Definition. The notation A [ B stands for fA; Bg.
S
With only two sets in F , the definition of F simplifies considerably.
3.33 Theorem. For all sets A and B, if F D fA; Bg, then
nh [ i o
8X X2 fA; Bg , Œ.X 2 A/ _ .X 2 B/ :
Proof. For F D fA; Bg, the axiom of pairing (S3) shows that
.S 2 F / , Œ.S D A/ _ .S D B/;
whence the axiom of union (S6) gives the following condition for X 2 .A [ B/:
ŒX 2 .A [ B/ , f9SŒ.S
2 fA; Bg/ ^ .X 2 S/g
, 9SfŒ.S D A/ _ .S D B/ ^ .X 2 S/g
, 9SfŒ.S D A/ ^ .X 2 S/ _ Œ.S D B/ ^ .X 2 S/g
, f9SŒ.X 2 A/ _ .X 2 B/g
, Œ.X 2 A/ _ .X 2 B/:
3.35 Example. Abbreviations are common [81, p. 98], [83, p. 453], [128, p. 129]:
www.pdfgrip.com
0 WD ¿;
˚ n˚ o n ˚ o
3 WD 2 [ f2g D ¿; f¿g [ ¿; f¿g D ¿; f¿g; ¿; f¿g ;
˚ n ˚ o
4 WD 3 [ f3g D ¿; f¿g; ¿; f¿g ; ¿; f¿g; ¿; f¿g ;
5 WD 4 [ f4g;
6 WD 5 [ f5g;
7 WD 6 [ f6g;
8 WD 7 [ f7g;
9 WD 8 [ f8g:
The sets 0; 1 are the binary digits; 0; 1; 2; 3; 4; 5; 6; 7; 8; 9 are the decimal digits.
3.36 Example. For all sets A, B, and C, there exists a set V whose only elements are
A, B, and C. Indeed, the axiom of pairing yields two sets fA; Bg and fB; Cg. With
˚
F D fA; Bg; fB; Cg ;
S
the axiom of union produces a set F D fA; Bg [ fB; Cg such that for each set X,
The common notation fA; B; Cg can replace fA; Bg [ fB; Cg, so that
3.37 Example. SIf the set F contains only three sets A, B, and C, so that F D
fA; B; Cg, then fA; B; Cg is also denoted by A [ B [ C, so that
[ [
A[B[C D fA; B; Cg D F
t
u
3.39 Theorem. The union of sets is associative:
` 8A8B 8C fŒ.A [ B/ [ C D ŒA [ .B [ C/g :
Proof. An informal proof can merely point out that a set “A or B, or C” is equivalent
to “A, or B or C.” A formal proof reveals that the associativity of the union [
corresponds to the associativity of the logical connective _ in a translation of
Œ.A [ B/ [ C D ŒA [ .B [ C/ into a theorem with atomic formulae and connectives:
Œ.A [ B/ [ C D ŒA [ .B [ C/ yet unproved,
m axiom of extensionality (S1),
8X .fX 2 Œ.A [ B/ [ Cg
, fX 2 ŒA [ .B [ C/g/
m theorem 3.33 twice,
8X .fŒ.X 2 A/ _ .X 2 B/ _ .X 2 C/g
, f.X 2 A/ _ Œ.X 2 B/ _ .X 2 C/g/
which is in prenex form with a matrix that is a theorem:
Proof. An informal proof can substitute B for S in the axiom of union (S6):
• If B 2 F , then
S for each X 2 B there exists S 2 F Swith X 2 S, namely S D B;
• hence X 2 F for S every X 2 B, by definition of F (axiom S6);
• consequently B F by definition of subsets (definition 3.7).
www.pdfgrip.com
m definitions of ,
h ˚ S i
(4) 8F 8B ŒB 2 F ) 8X f.X 2 B/ ) ŒX 2 . F /g
S
m definitions of ,
h ˚ i
(3) 8F 8B ŒB 2 F ) 8X .X 2 B/ ) f9AŒ.A 2 F / ^ .X 2 A/g
Based on the axiom of union and the axiom of T separation, definition 3.42 specifies
for each nonempty set F a new set, denoted by F and called the intersection of
F , which consists of every element X that is an element of every element of F .
3.42 Definition (intersection
S of sets). For each set F , apply the axiom of union
to define A WD F , and apply the axiom of separation to the set A and to the
formula
8YŒ.Y 2 F / ) .X 2 Y/:
T
Then define the intersection of F , a set denoted by F , through the formula
\ n [ o
F WD X 2 F W 8YŒ.Y 2 F / ) .X 2 Y/ ;
so that
˚ \ [
8X .X 2 F / , .X 2 F / ^ f8YŒ.Y 2 F / ) .X 2 Y/g :
T
The definition
T of the intersection F of a set of sets F states that a set X is an
element of F if and only if 8YŒ.Y 2 F / ) .X 2 Y/ holds, which occurs if and
only if X is an element of every element T Y of F . This definition also holdsS if F
is empty because ofTthe requirement S that F first be a subset of the union F.
(This definition of in terms of conforms to Bernays’s
T [8, p. 14].) If the set
F contains only two elements, then the definition of F simplifies considerably.
3.43 Theorem. For all sets A and B, if F D fA; Bg then
\
fA; Bg D fX W ŒX 2 .A [ B/ ^ Œ.X 2 A/ ^ .X 2 B/g :
Proof. Apply theorem 1.55, Œ.H/ ) .K/ ) Œ.H/ ) .L/ ) f.H/ ) Œ.K/ ^
.L/g :
` 8Yf.Y 2 fA; Bg/ ) .X 2 Y/g
) Œ.A 2 fA; Bg/ ) .X 2 A/ specialization SubfYA ,
` A 2 fA; Bg axiom of pairing (S3),
` X 2 A Detachment;
` 8Yf.Y 2 fA; Bg/ ) .X 2 Y/g
) Œ.B 2 fA; Bg/ ) .X 2 B/ specialization SubfYB ,
` B 2 fA; Bg axiom of pairing (S3),
`X2B Detachment;
www.pdfgrip.com
` 8Yf.Y 2 fA; Bg/ ) .X 2 Y/g
) Œ.X 2 A/ ^ .X 2 B/ theorem 1.55;
Because the definition of the intersection of sets relies upon the conjunction ^,
the intersection has formal features similar to the logical features of the conjunction,
for instance, commutativity and associativity, as demonstrated in the following
theorems.
3.46 Theorem. The intersection of sets commutes:
` 8A f8BŒ.A \ B/ D .B \ A/g :
Proof. An informal proof can state that a set X is an element of A and B if and only
if X is an element of B and A. A formal proof shows that the commutativity of the
intersection \ corresponds to the commutativity of the logical connective ^:
.A \ B/ D .B \ A/ yet unproved,
m axiom S1,
8X fŒX 2 .A \ B/ , ŒX 2 .B \ A/g
m theorem 3.43,
8X fŒ.X 2 A/ ^ .X 2 B/ , Œ.X 2 B/ ^ .X 2 A/g
www.pdfgrip.com
which holds by the commutativity ` Œ.P/ ^ .Q/ , Œ.Q/ ^ .P/ (theorem 1.57). t
u
3.47 Theorem. The intersection of sets is associative:
` 8A 8B 8CfŒ.A \ B/ \ C D ŒA \ .B \ C/g :
Proof. An informal proof can rely on the equivalence of “A and B, and C” with
“A, and B and C.” A formal proof shows that the associativity of the intersection \
corresponds to the associativity of the logical connective ^.
Œ.A \ B/ \ C D ŒA \ .B \ C/ yet unproved,
m axiom of extensionality (S1),
8X fX 2 Œ.A \ B/ \ Cg
, fX 2 ŒŒA \ .B \ C/g
m theorem 3.43 twice,
8X fŒ.X 2 A/ ^ .X 2 B/ ^ .X 2 C/g
, f.X 2 A/ ^ Œ.X 2 B/ ^ .X 2 C/g
which holds thanks to the associativity of ^ (theorem 1.66):
fŒ.P/^.Q/^.R/g , f.P/^Œ.Q/^.R/g: t
u
T
Theorem 3.48 shows that the intersection F of a set of sets F Tis a subset of
every element of F . In other terms, for each set B, if B 2 F , then . F / B.
T
3.48 Theorem. For each (nonempty) set F and for each B 2 F , . F / B:
n h\ io
` 8F 8B .B 2 F / ) F B :
n h ˚
8F 8B 8X .B 2 F / )
io
f8YŒ.Y 2 F / ) .X 2 Y/g ) .X 2 B/
m theorem 1.37,
P
n h ‚ …„ ƒ
8F 8B 8X f8YŒ.Y 2 F / ) .X 2 Y/g
io
) .B 2 F…/ ) .X
„ ƒ‚ 2 B/
„ƒ‚…
Q R
which holds by specialization (SubfYB ) and the law of commutation (theorem 1.37):
` .P/ ) f.Q/ ) .R/g , .Q/ ) f.P/ ) .R/g :
t
u
3.49 Definition (disjoint sets). Two sets A and B are disjoint if and only if
A \ B D ¿. Similarly, a set of sets F is pairwise disjoint if and only if either
A D B or A \ B D ¿ for all elements A and B of F .
For the union of disjoint sets, a special notation proves convenient.
3.50 Definition (disjoint unions). A union A [ B is disjoint if and only if A \S BD
P stands for A [ B. Similarly, a union F
¿; only for disjoint sets, the notation A[B
is pairwise disjoint if and only if either A D B or A \
SB D ¿ for all elements
S A and
B of F ; only for pairwise disjoint sets, the notation P F stands for F .
The indexed notation then denotes the union and the intersection of G as follows:
[ [
.U n S/ WD G;
S2F
\ \
.U n S/ WD G:
S2F
The notation on the left-hand sides avoids the need to write a formula for the set G .
www.pdfgrip.com
Because the definitions of the intersection and union of sets rely on conjunction
and disjunction, the union and intersection have formal features similar to the logical
features of the conjunction and disjunction, for instance, distributivity.
3.52 Theorem (De Morgan’s Laws). For each set U, and for each set F
P.U/ of subsets of U, the complement of the union equals the intersection of the
complements,
[ \
Un F D .U n A/;
A2F
whereas for each set U, and for each nonempty family F P.U/ of subsets of U,
the complement of the intersection equals the union of the complements,
\ [
Un F D .U n A/:
A2F
Proof. For the complement of the intersection, an informal proof can proceed as
follows.
T T
• For each set X, X 2TU n . F / if and only if X 2 U but X … . F /;
… . F / if and only if there exists A 2 F with X … A;
• by definition, X T
• hence X 2 U n . F / T if and only if X 2 U and there exists A 2 F with X … A;
• equivalently, X 2TU n . F / if and only ifSthere exists A 2 F with X 2 .U n A/;
• hence X 2 U n . F / if and only if X 2 A2F .U n A/.
The foregoing informal proof does not justify the permutation of the two
statements “X 2 U” and “there exists A 2 F ” but the following formal proof
justifies such a permutation by the absence of any free occurrence of A in X 2 U.
T S
U n . F / D A2F .U n A/
m (axiom S1),
˚ T S
8X ŒX 2 fU n . F /g , ŒX 2 A2F .U n A/
m .n/,
˚ T S
8X Œ.X 2 U/ ^ f: .X 2 F /g , ŒX 2 A2F .U n A/
S T
m , ,
8X fŒ.X 2 U/ ^ .: f8A Œ.A 2 F / ) .X 2 A/g/
, f9A Œ.A 2 F / ^ f.X 2 U/ ^ Œ:.X 2 A/ggg
:Œ8A.P/;
m
9AŒ:.P/;
8X fŒ.9A f.X 2 U/ ^ f: fŒf:.A 2 F /g _ .X 2 A/ggg/
, Œ9A f.A 2 F / ^ Œ.X 2 U/ ^ Œ:.X 2 A/gg
m De Morgan,
8X fŒ.9A f.X 2 U/ ^ f.A 2 F / ^ Œ:.X 2 A/gg/
, f9A Œ.A 2 F / ^ f.X 2 U/ ^ Œ:.X 2 A/ggg
www.pdfgrip.com
T T
Œ. F / [ . G /
T
D Œ fA [ B W .A 2 F / ^ .B 2 G /g unproved,
m (S1),
T T
8X .fX 2 Œ. F / [ . G /g
T
, fX 2 Œ fA [ B W .A 2 F / ^ .B 2 G /gg/
T
m def. [, ,
T T
8X .fŒX 2 . F / _ ŒX 2 . G /g
, 8A 8BfŒ.A 2 F / ^ .B 2 G / ) ŒX 2 .A [ B/g
T
m definition: ,
Œ8X .f8AŒ.A 2 F / ) .X 2 A/g _ f8BŒ.B 2 G / ) .X 2 B/g/
, 8A 8BfŒ.A 2 F / ^ .B 2 G / ) ŒX 2 .A [ B/g
m no free A, B,
f8X Œ8A .8BfŒ.A 2 F / ) .X 2 A/ _ Œ.B 2 G / ) .X 2 B/g/g
, 8A 8Bf.A 2 F / ^ .B 2 G / ^ ŒX 2 .A [ B/g
m Œ.P/^.Q/).P/,
A 4 B WD .A [ B/ n .A \ B/:
3.87 . For each set F P.U/, and for each subset B U, prove that
[ [
F \BD .A \ B/:
A2F
P n A/.
3.108 . Prove that if A \ B D ¿, then A [ B D .A n B/[.B
P \ B/.
3.109 . Prove that A [ B D .A 4 B/[.A
P n A/.
3.110 . Prove that A 4 B D .A n B/[.B
Beyond logic and sets, much of mathematics consists of connections between types
of sets called Cartesian products, mathematical functions, and mathematical rela-
tions. These types of sets allow for mathematical specifications, analysis, synthesis,
and processing of such concepts as graphs, maps, algorithms, and rankings.
Cartesian products contain certain sets with two elements. Whereas fX; Yg D fY; Xg
for all sets X and Y, however, some situations require a method for listing the
elements of a set in a specific order by means of “ordered pairs” or otherwise, for
example, in geography and in navigation as in figure 3.2. The following definition
(attributed to Wiener and Kuratowski [83, p. 455]) and theorem 3.58 derive such
X,Y
Y,X
ordered pairs from sets, which shows that the concept of ordered pairs does not
require any additional axiom.
3.56 Definition (ordered pairs). For all sets X and Y, the ordered pair .X; Y/ is
the set defined by three applications of the axiom of pairing as
˚
.X; Y/ D fXg; fX; Yg :
X is the first coordinate of .X; Y/, and Y is the second coordinate of .X; Y/.
3.57 Example.
˚
• If X D 0 and Y D 1, then .X; Y/ D ˚f0g; f0; 1g. ˚
• If X D 1 and Y D 0, then .X; Y/ D ˚f1g; f1; 0g D ˚f1g; f0; 1g . ˚
• If X D 0 and Y D 0, then .X; Y/ D f0g; f0; 0g D f0g; f0g D f0g .
The following theorem confirms that, in contrast to sets with two elements,
ordered pairs record the order of their coordinates.
3.58 Theorem. For all sets X and Y, if X 6D Y, then .X; Y/ 6D .Y; X/.
Proof. An informal proof can consist in showing that the two sets .X; Y/ and .Y; X/
contain different elements, whence .X; Y/ 6D .Y; X/.
• For all sets X and Y, X 2 fXg and Y 2 fYg by pairing (S3);
• if X 6D Y, then X … fYg and Y … fXg, by pairing (S3);
• hence if X 6D Y, then fXg 6D fYg by extensionality (S1);
• from X … fYg and X 2 fY; Xg it follows that fXg ˚ 6D fY; Xg by
(S1);
• from fXg 6D˚ fYg , fXg D
6 fY; Xg follows fXg … fYg; fY; Xg D .Y; X/;
• yet fXg 2 fXg; fX; Yg D .X; Y/;
• from fXg 2 .X; Y/ and fXg … .Y; X/ follows .X; Y/ 6D .Y; X/, by (S1).
A formal proof can parallel the same reasoning.
(1) ` 8ZŒ.Z 2 fXg/ , .Z D X/ axiom of pairing (S3),
(2) ` 8ZŒ.Z 6D X/ , .Z … fXg/ contraposition and transposition,
(3) ` .Y 6D X/ , .Y … fXg/ specialization SubfZY ,
(4) ` Y 6D X hypothesis,
(5) ` Y … fXg Detachment;
(6) ` Y 2 fX; Yg axiom of pairing (S3),
(7) ` fXg 6D fX; Yg (5), (6), and extensionality (S1);
(8) ` Y 2 fYg axiom of pairing (S3),
(9) ` fXg 6D fYg (5), (8), and extensionality (S1);
(10) ` fXg … ffYg; fY; Xgg D .Y; X/ (7), (9), and extensionality (S1);
(11) ` fXg 2 ffXg; fX; Ygg D .X; Y/ pairing (S3),
(12) ` .X; Y/ 6D .Y; X/ (10), (11), and extensionality (S1).
t
u
www.pdfgrip.com
The definition of the ordered pair .X; Y/ holds for all sets X and Y, in particular,
for all elements X and Y of two sets A and B. The following theorem shows that all
these ordered pairs .X; Y/ are themselves elements of a set.
3.59 Theorem. For all sets A and B, for each X 2 A and for each Y 2 B, the pair
.X; Y/ belongs to PŒP.A [ B/.
Proof. An informal proof can trace back the definition of the ordered pair .X; Y/:
• From X 2 A follows X 2 A [ B, because A A [ B, whence fXg 2 P.A [ B/;
• from Y 2 B it follows that Y 2 A [ B, because B A [ B;
• from X 2 A [ B and Y 2 A [ B it follows that fX; Yg 2 P.A [ B/;
˚
• from fXg 2 P.A [ B/ and fX; Yg 2 P.A [ B/, it follows that fXg; fX; Yg 2
PŒP.A [ B/.
A formal proof can parallel the foregoing argument.
`X2A hypothesis,
` A .A [ B/ theorem 3.40 ,
` X 2 .A [ B/ definition of subsets and specialization,
` fXg .A [ B/ definitions of subsets and singletons,
` fXg 2 P.A [ B/ definition of power sets.
`Y2B hypothesis,
` Y 2 .A [ B/ as for X 2 .A [ B/,
` fX; Yg 2 P.A [ B/ X 2 .A [ B/ and Y 2 .A [ B/,
` f fXg; fX; Yg g P.A [ B/ fXg 2 P.A [ B/, fX; Yg 2 P.A [ B/,
` f fXg; fX; Yg g 2 PŒP.A [ B/ definition of power set.
t
u
Because every ordered pair .X; Y/ belongs to the set PŒP.A [ B/, the axiom
of separation guarantees that the collection of all such ordered pairs is a set.
3.60 Definition (Cartesian product). For all sets A and B, the Cartesian product
of A and B is the set A B (read “A cross B”), consisting of all ordered pairs .X; Y/
with X 2 A and Y 2 B. Thus,
A B
n o
D C 2 P ŒP.A [ B/ W 9X 9Yf.X 2 A/ ^ .Y 2 B/ ^ ŒC D .X; Y/g
D f.X; Y/ W .X 2 A/ ^ .Y 2 B/g
with 9X 9Yf.X 2 A/ ^ .Y 2 B/ ^ ŒC D .X; Y/g for P in the axiom of separation
(S5). The sets A and B are the factors of the Cartesian product A B.
www.pdfgrip.com
A WD f0; 1; 2g;
B WD f0; 1g;
A B D f.0; 0/; .1; 0/; .2; 0/; .0; 1/; .1; 1/; .2; 1/g;
B A B
0 1 2 A
For practice and for future use, the following theorems establish relations
between Cartesian products and other operations with sets.
3.62 Theorem. For all sets A, B, C, and D,
Proof. An informal proof can establish that Œ.A \ C/ .B \ D/ and Œ.A B/ \
.C D/ have exactly the same elements. These two sets are Cartesian products,
and consequently their elements are ordered pairs.
• An ordered pair .X; Y/ is an element of .A\C/ .B\D/ if and only if X 2 .A\C/
and Y 2 .B \ D/;
• hence .X; Y/ 2 Œ.A \ C/ .B \ D/ if and only if X 2 A and X 2 C, and Y 2 B
and Y 2 D,
• which is equivalent to X 2 A and Y 2 B, and X 2 C and Y 2 D,
• which is equivalent to .X; Y/ 2 .A B/ and .X; Y/ 2 .C D/,
• which is thus equivalent to .X; Y/ 2 Œ.A B/ \ .C D/.
As the preceding informal proof swapped X 2 C and Y 2 B, a formal proof can rely
on the commutativity Œ.P/ ^ .Q/ , Œ.Q/ ^ .P/ of ^ by theorem 1.57:
www.pdfgrip.com
fŒ.P/ ^ .Q/ _ Œ.P/ ^ .S/ _ Œ.R/ ^ .Q/ _ Œ.R/ ^ .S/g , fŒ.P/ _ .R/ ^ Œ.Q/ _ .S/g;
Thus,
.X; Y/ 2 Œ.A B/ [ .C D/ [ .A D/ [ .C
B/
S
m definition of ,
Œ.X; Y/ 2 .A B/ _ Œ.X; Y/ 2 .C D/
_Œ.X; Y/ 2 .A D/ _ Œ.X; Y/ 2 .C B/
m definition of ,
Œ.X 2 A/ ^ .Y 2 B/ _ Œ.X 2 C/ ^ .Y 2 D/
_Œ.X 2 A/ ^ .Y 2 D/ _ Œ.X 2 C/ ^ .Y 2 B/_
m theorem 1.75,
www.pdfgrip.com
S S S
Proof. An informal proof can show that . F / . G / and .A;B/2F G .A B/
have exactly the same elements.
S S S
• A pairS.X; Y/ is an element of . F / . G / if and only if X 2 . F / and
Y 2 . G /,
• which is equivalent to X 2 A for some A 2 F and Y 2 B for some B 2 G ,
• which is equivalent to .X; Y/ 2 .A B/ for some .A; B/ 2 .F G /.
A formal proof can parallel the foregoing reasoning.
S S
.X; Y/ 2 . F / . G /
m definition of ,
S S
ŒX 2 . F / ^ ŒY 2 . G /
S
m definition of ,
f9AŒ.A 2 F / ^ .X 2 A/g ^ f9BŒ.B 2 G / ^ .Y 2 B/g
m no B, no A,
9A 9BfŒ.A 2 F / ^ .X 2 A/ ^ Œ.B 2 G / ^ .Y 2 B/g
m properties of ^,
9A 9BfŒ.A 2 F / ^ .B 2 G / ^ Œ.X 2 A/ ^ .Y 2 B/g
m definition of ,
9A 9BfŒ.A B/ 2 .F G / ^ Œ.X; Y/ 2 .A B/g
S
m definition of .
S
.X; Y/ 2 .AB/2.F G / .A B/
t
u
www.pdfgrip.com
T T T
Proof. An informal proof can show that . F / . G / and .A;B/2F G .A B/
have exactly the same elements.
T T T
• A pairT.X; Y/ is an element of . F / . G / if and only if X 2 . F / and
Y 2 . G /,
• which is equivalent to X 2 A for every A 2 F and Y 2 B for every B 2 G ,
• which is equivalent to .X; Y/ 2 .A B/ for every .A; B/ 2 .F G /.
The foregoing informal proof glosses over the case in which at least one of F or G
is empty, which would require invoking the corresponding unions as supersets of the
intersections, because of the definition of intersections. One remedy could consist in
proving such cases separately. Indeed, F D ¿ or G D ¿ if and only if .F G / D
¿. A formal proof can parallel the foregoing reasoning with theorem 1.82,
to prove that
\ \ \
F G .A B/;
.A;B/2F G
with the hypotheses C 2 F for P and D 2 G for R for the converse, so that
` fŒ.P/ ) .Q/ ^ Œ.R/ ) .S/g ( Œ.P/ ^ .R/ ^ fŒ.P/ ^ .R/ ) Œ.Q/ ^ .S/g :
Thus,
T T
.X; Y/ 2 . F/ . G/
m definition of ,
T T
ŒX 2 . F / ^ ŒY 2 . G /
T S
m H H,
f8AŒ.A 2 F / ) .X 2 A/g ^ f8BŒ.B 2 G / ) .Y 2 B/g
^ f9CŒ.C 2 F / ^ .X 2 C/g ^ f9DŒ.D 2 G / ^ .Y 2 D/g
m no B, no A,
8A 8BfŒ.A 2 F / ) .X 2 A/ ^ Œ.B 2 G / ) .Y 2 B/g
^ 9C 9DfŒ.C 2 F / ^ .X 2 C/ ^ .D 2 G / ^ .Y 2 D/g
m theorems,
www.pdfgrip.com
8A 8BfŒ.A 2 F / ^ .B 2 G / ) Œ.X 2 A/ ^ .Y 2 B/g
^ 9C 9DfŒ.C 2 F / ^ .X 2 C/ ^ .D 2 G / ^ .Y 2 D/g
m definition of ,
8A 8BfŒ.A; B/ 2 .F G / ) Œ.X; Y/ 2 .A B/g
^ 9C 9DfŒ.C; D/ 2 .F G / ^ Œ.X; Y/ 2 .C D/g
T
m definition of .
T
.X; Y/ 2 .AB/2.F G / .A B/g
t
u
Another common notation for .X; Y/ 2 R is XRY, especially if such a special symbol
as , ¤, or D denotes the relation. Relations may also be represented by graphs.
3.67 Definition (directed graph). A directed graph is an ordered pair G WD
.V; E/ with a relation E V V between elements of the same set A D V D B.
Elements of V are called vertices. A pair .X; Y/ 2 E is called an edge from X to Y:
(X,Y)
X Y
A D f.X; X/ 2 A A W X 2 Ag:
www.pdfgrip.com
Thus, the diagonal corresponds to the relation called “equality,” or, equivalently,
“identity”: a pair of elements .X; Y/ lies on the diagonal A if and only if X D Y.
Because X D X, and hence .X; X/ 2 A , for every X 2 A, it follows that the
diagonal A relates every element of A to itself, whence D.R/ D A D R.R/. For
example, if
˚
A WD ¿; f¿g ;
then
.f¿g; f¿g/
A D ;
.¿; ¿/
because the only pairs .X; Y/ with X D Y are the pairs .¿; ¿/ and .f¿g; f¿g/. As a
graph, the diagonal consists of a single loop at each vertex:
(X, X) X
3.69 Example. For all sets H and K, consider their respective power sets A WD
P.H/ and B WD P.K/. The relation R P.H/ P.K/ of inclusion consists
of all, but only those, pairs .V; W/ of subsets V H and W K such that V W:
The domain of the relation of strict inclusion consists of all subsets of A included
in, but not equal to, at least one subset of B; its range consists of all subsets of B that
contain, but do not coincide with, at least one subset of A.
www.pdfgrip.com
3.71 Example. For all sets A and B, the relational constant 2 relates every element
X 2 A that is an element of some Y 2 B. Denote this relation on A B by E.
(The symbol 2 cannot denote a subset E of A B because 2 is neither an individual
variable nor an individual constant.) Thus, E .A B/ is defined by
For example, if
˚
A WD ˚ ¿; f¿g ;
B WD ¿; f¿g ;
then
.¿; f¿g/
ED ;
because in A B the only pair .X; Y/ such that X 2 Y is the pair .¿; f¿g/.
3.72 Example. For the sets
˚
A WD n ¿; f¿g ;
˚ o
B WD ¿; f¿g; ¿; f¿g ;
because in A B the only pairs .X; Y/ such that X 2 Y are those just displayed.
In some contexts a relation R A B may also prove useful with B and A listed
in the reverse order. Then the “inverse” relation Rı1 B A contains similar pairs
as R does but with coordinates listed in the reverse order. Some texts denote the
inverse relation by R1 , which can cause confusion because the same notation also
represents reciprocals in arithmetic and algebra. The notation adopted here for the
inverse relation, Rı1 , conforms to [70].
www.pdfgrip.com
3.77 Example. For each relation R A B between any sets A and B, the
composition R ı Rı1 contains the diagonal 4R.R/ for the range of R. Indeed, by
definition of its range, R relates every Y 2 R.R/ to an element X 2 A so that
.X; Y/ 2 R; then the inverse Rı1 relates Y to X so that .Y; X/ 2 Rı1 . From
.Y; X/ 2 Rı1 and .X; Y/ 2 R it follows that .Y; Y/ 2 .R ı Rı1 / for every
.Y; Y/ 2 4R.R/ . Therefore 4R.R/ .R ı Rı1 /. Similarly, Rı1 ı R contains the
diagonal 4D.R/ for the domain of R. Indeed, by definition of its domain, R relates
every X 2 D.R/ to an element Y 2 B so that .X; Y/ 2 R; then the inverse Rı1
relates Y to X so that .Y; X/ 2 Rı1 . From .X; Y/ 2 R and .Y; X/ 2 Rı1 it follows
that .X; X/ 2 .Rı1 ı R/ for every .X; X/ 2 4D.R/ . Therefore 4D.R/ .Rı1 ı R/.
Because every relation R between sets A and B is a subset R .A B/, operations
with sets apply to all relations.
3.78 Definition (unions and intersections of relations). For all sets A, B, C, E,
and for all relations R A B and T C E, the union of the relations R and T
is the relation R [ T between A [ C and B [ E, so that
3.79 Definition (restricted relation). For each relation R .A B/, and for each
subset S A of A, the restriction of R to S is the relation RjS .S B/ defined by
RjS D R \ .S B/
D f.X; Y/ 2 R W X 2 Sg:
3.128 . For all sets A and B, prove that P.B/ is the range R of the relation on
P.A/ P.B/.
3.129 . Provide examples of sets A and B, such that P.A/ is the domain D of the
relation on P.A/ P.B/.
3.130 . Provide examples of sets A and B, such that P.A/ is not the domain D of
the relation on P.A/ P.B/, so that D ¤ P.A/.
Functions are relations relating exactly one element of their domain to each element
of their range.
In some applications of relations, the domain and the range can contain measure-
ments.
3.80 Example. Results from astronomical observations can consist of a relation
between two coordinates of position, with ordered pairs .X; Y/ where X is the
observed ascension (elevation) and Y is the observed declination (azimuth) of an
asteroid. For example, the following pairs .X; Y/ record the ascension X and the
declination Y of the asteroid Pallas, measured by Baron von Zach about 1800 A.D.
[14, p. 5]:
.0; 408/ .30; 89/ .60; 66/ .90; 10/ .120; 338/ .150; 807/
.180; 1238/ .210; 1511/ .240; 1583/ .270; 1462/ .300; 1183/ .330; 804/
1583
1511
1462
1238
1183
Declination [minutes]
806
408
338
89
10
-66
Fig. 3.3 Positions include at most one declination for each ascension.
The range of F, denoted by R.F/, consists of every Y 2 B such that there exists
some X 2 A such that .X; Y/ 2 F:
˚
R.F/ D Y 2 B W 9XŒ.X 2 A/ ^ f.X; Y/ 2 Fg :
FW A!B
CZ WD f.X; Z/ W X 2 Ag:
In particular, D.CZ / D A for the domain of CZ , and R.CZ / D fZg for its range.
3.86 Example. For each set A, there exists an identity function
IA W A ! A;
X 7! X:
IA WD f.X; X/ W X 2 Ag:
3.87 Example. For all sets A and B, the canonical projection functions from the
Cartesian product A B into its factors A and B are the functions PA and PB with
PA W .A B/ ! A;
.X; Y/ 7! XI
PB W .A B/ ! B;
.X; Y/ 7! Y:
Thus, PA maps .X; Y/ to its first coordinate X in A, whereas PB maps .X; Y/ to its
second coordinate Y in B. The domain of PA and PB is A B, but the range of PA is
A, whereas the range of PB is B.
3.88 Example. For all sets A and B, for any subset V A and any element Z 2 B,
the slice function SV;Z maps each element X 2 V to .X; Z/ 2 .A B/:
SV;Z W V ! .A B/;
X 7! .X; Z/:
S W A ! 2;
1 if X 2 S,
X 7!
0 if X … S.
The following theorem provides a means to compare two functions to each other.
3.90 Theorem. Two functions F W A ! B and G W A ! B are equal, F D G, if and
only if they have the same domain D A and F.X/ D G.X/ for every X 2 D.
Proof. This proof rewrites F.X/ D G.X/ in terms of the definition of functions:
FDG
m extensionality,
8X 8Y fŒ.X; Y/ 2 F , Œ.X; Y/ 2 Gg
m functional notation.
8X 8Y fŒY D F.X/ , ŒY D G.X/g
t
u
Some situations involve only parts of a function, or combinations of several
functions, for instance, as defined by the concepts introduced here.
www.pdfgrip.com
FjS W S ! B;
X 7! F.X/:
Thus, FjS .X/ D F.X/, but FjS maps only the elements of S \ D.F/.
3.92 Example. With A WD f6; 7; 8; 9g and B WD f1; 2; 3; 4g, let
Then the restriction of F to the subset S WD f6; 8g is FjS D f.6; 2/; .8; 2/g.
3.93 Example. For each set B, and for each subset W B, the inclusion function,
denoted by the Greek letter (“iota”), is the restriction of the identity function to W:
W W W ! B;
X 7! X:
P W .A[C/
F [G P ! .B [ E/;
F.X/ if X 2 D.F/ A,
X 7!
G.X/ if X 2 D.G/ C.
F \ G W D ! .B \ E/;
X 7! F.X/ D G.X/;
Some situations involve the images by functions of not only single elements in the
domain, but also subsets of the domain, as defined here.
3.97 Definition (image of a set). For each function F W A ! B and each subset
V A, the image of V by F is the subset F”.V/ consisting of all images of each
X 2 V \ D.F/ by F:
C¿ .f¿g/ D ¿;
C”¿ .f¿g/ D f¿g:
C¿ .f¿g/ D ¿;
C¿ .f¿g/ D f¿g:
3.101 Definition (pre-image). For each function F W A ! B, and for each element
Y 2 B, the pre-image of Y by F is the subset of A denoted by F ı1 ”.fYg/ and
consisting of all elements X 2 A such that F.X/ D Y:
A if Z 2 W,
CZı1 ”.W/ D
¿ if Z … W.
[ [
F ı1 ” G D F ı1 ”.W/
W2G
\ \
F ı1 ” G D F ı1 ”.W/
W2G
[ [
F” F D F”.V/
V2F
\ \
F” F F”.V/:
V2F
S
Proof. Apply the definitions of union and intersection. For F ı1 ” . G /,
S S
F ı1 ” . G / D W2G F ı1 ”.W/ yet unproved,
m extensionality (S1),
˚ ı1
S
8X X 2 F ” . G /
S
, X 2 W2G F ı1 ”.W/
www.pdfgrip.com
S
m definitions of F ı1 ” and ,
S
8X fŒF.X/ 2 .
G /
, 9W f.W 2 G / ^ ŒF.X/ 2 Wg
S
m definition of ,
˚
8X 9W f.W 2 G / ^ ŒF.X/ 2 Wg
, 9W f.W 2 G / ^ ŒF.X/ 2 Wg
T
which holds thanks to theorem 1.63: .P/ , .P/. For F ı1 ” . G /,
T T
F ı1 ” . G / D W2G F ı1 ”.W/ yet unproved,
m extensionality (S1),
˚ ı1
T
8X X 2 F ” . G /
T
, X 2 W2G F ı1 ”.W/
T
m definitions of F ı1 and ,
T
8X fŒF.X/ 2 . G / ,
8W f.W 2 G / ) ŒF.X/ 2 Wg
T
m definitions of ,
˚
8X 8W f.W 2 G / ) ŒF.X/ 2 Wg
, 8W f.W 2 G / ) ŒF.X/ 2 Wg
S
which holds thanks to theorem 1.63: .P/ , .P/. For F” . F /,
S S
F” . F / D V2F F”.V/ yet unproved,
m extensionality (S1),
˚ S S
8Y ŒY 2 F” . F / , Y 2 V2F F”.V/
S
m definitions: F”, ,
˚ S
8Y 9X ŒX 2 . F / ^ ŒY D F.X/
, 9V fŒV 2 F ^ ŒY 2 F”.V/g
S
m definitions: , F”,
8Y 9X f9VŒ.V 2 F / ^ .X 2 V/g ^ ŒY D F.X/
˚
, 9V .V 2 F / ^ 9Xf.X 2 V/ ^ ŒY D F.X/g
m no X in V 2 F ,
8Y 9X f9VŒ.V 2 F / ^ .X 2 V/g ^ ŒY D F.X/
˚
, 9V 9X .V 2 F / ^ f.X 2 V/ ^ ŒY D F.X/g
T
which holds thanks to the associativity of ^ (theorem 1.66). For F” . F /,
www.pdfgrip.com
T T
F” . F/ V2F F”.V/ yet unproved,
m definition 3.7,
˚ T T
8Y ŒY 2 F” . F / ) Y 2 V2F F”.V/
T
m definitions: F”, ,
T
8Y f9X ŒX 2 . F / ^ ŒY D F.X/g
) 8V f.V 2 F / ) ŒY 2 F”.V/g
T
m definitions: , F”,
8Y f9X Œ8V .ŒV 2 F ) ŒX 2 V/ ^ ŒY D F.X/g
˚
) 8V .V 2 F / ) 9Xf.X 2 V/ ^ ŒY D F.X/g
theorem 2.63: m no X in .V 2 F /,
8Y f9X Œ8V .ŒV 2 F ) ŒX 2 V/ ^ ŒY D F.X/g
˚
) 8V9X .V 2 F
„ ƒ‚ …/ ) f.X 2
„ƒ‚…V / ^ ŒY D F.X/g
„ ƒ‚ …
P Q R
3.131 . Determine whether F WD f.0; 1/; .2; 3/; .1; 2/; .0; 4/g is a function.
3.132 . Determine whether G WD f.9; 2/; .7; 3/; .8; 2/; .6; 1/g is a function.
3.133 . Determine whether the following relation is a function:
R WD f.0; 1/; .1; 2/; .2; 4/; .3; 8/; .4; 8/; .5; 4/; .6; 2/; .7; 1/; .8; 0/g:
S WD f.0; 2/; .1; 5/; .2; 7/; .3; 5/; .4; 2/; .5; 5/; .6; 7/; .7; 5/; .8; 2/g:
Z WD f.0; 0/; .1; 0/; .2; 0/; .3; 0/; .4; 0/; .5; 0/; .6; 0/; .7; 0/; .8; 0/; .9; 0/g:
1A W A ! f1g; 1A .a/ D 1:
Recall that the superscript ” indicates images of subsets (rather than of elements).
˚ ˚
(1) ˚¿; f¿g is a subset of A. Find its image: F” ¿; f¿g ˚:
(2) ¿; f¿g is a subset of B. Find its inverse image F ı1 ” ¿; f¿g :
www.pdfgrip.com
˚ ˚
(3) ¿; f¿g is an element of A. Find its image
n F ¿; f¿g
o :
˚ ı1
˚
(4) ¿; f¿g is an element of B. Find F ” ¿; f¿g :
C WD 4 D f0; 1; 2; 3g
˚ n ˚ o
D ¿; f¿g; ¿; f¿g ; ¿; f¿g; ¿; f¿g ;
D WD C;
G WD f.0; 3/; .1; 0/; .2; 3/; .3; 1/g:
Recall that the superscript ” indicates images of subsets (rather than of elements).
(1) 3 is a subset of C. Find its image: G”.3/:
(2) 3 is a subset of D. Find its inverse image Gı1 ”.3/:
(3) 3 is an element of C. Find its image G.3/:
(4) 3 is an element of D. Find Gı1 ”.f3g/:
.G ı F/ .X/ WD G ŒF.X/
Their composition
.G ı F/W A ! C;
has values
so that
Proof. For each X 2 F ı1 ”fGı1 ”ŒD.H/g, apply the definition of ı repeatedly:
t
u
In contrast to its associativity, the composition of functions is not commutative.
3.108 Counterexample. Consider the following functions F and G:
Their composition
.F ı G/W C ! A;
www.pdfgrip.com
has values
so that
In contrast,
3.116 Definition. For all sets A, B, C, D, and for all functions F W A ! B and
G W C ! D, define the function
FGW A C ! B D;
.W; Z/ 7! F.W/; G.Z/ :
Thus .W; Z/; F.W/; G.Z/ 2 F G if and only if W; F.W/ 2 F and
Z; G.Z// 2 G, so that W; F.W/ ; Z; G.Z// 2 F G.
F.W/; G.Z/ D .F G/.W; Z/ D .F G/.R; S/ D F.R/; G.S/ :
Hence F.W/ D F.R/ and G.Z/ D G.S/, by equality of ordered pairs. Consequently,
W D R and Z D S, by injectivity of F and G. t
u
3.118 Definition (Bijectivity). A function F W A ! B is bijective if and only if F
is both injective and surjective, which can be denoted by F W A B or F W A • B.
3.119 Example. With A WD f0; 1; 2; 3g and B WD f1; 2; 4; 8g, the function P W A !
B defined by P WD f.0; 1/; .1; 2/; .2; 4/; .3; 8/g is bijective.
3.120 Example. For each set A the identity function IA W A ! A, X 7! X is bijective.
Indeed, IA is injective, because if IA .W/ D IA .X/ then W D IA .W/ D IA .X/ D X.
Similarly, IA is surjective, because for each Y 2 A there exists X 2 A, in effect
X WD Y, with Y D IA .Y/.
3.121 Example. For each proper subset S ¤ A, the inclusion function W S ! A,
X 7! X, is injective but not surjective. Indeed, is injective, for if W; X 2 S and
W 6D X, then .W/ D W 6D X D .X/, whence .W/ 6D .X/. However, is not
surjective: because S is a proper subset of A, there exists some element Z in A n S; in
particular, Z 6D X for each X 2 S, and, consequently, .X/ D X 6D Z, which means
that is not surjective. Thus, is not bijective either.
3.122 Example. For each nonempty set A and for each set B containing more than
one element, the canonical projection PA W .A B/ ! A is surjective but not
injective. Indeed, B contains at least one element Y 2 B, because B 6D ¿; hence,
X D PA .X; Y/ for each X 2 A. However, B also contains some Z 2 B such that
Y 6D Z. Consequently, .X; Y/ 6D .X; Z/, and yet PA .X; Y/ D X D PA .X; Z/, which
means that PA is not injective. Thus, PA is not bijective either.
3.123 Theorem. For all functions F W A ! B and G W B ! C,
• if F and G are both injective, then G ı F is also injective;
• if F and G are both surjective, then G ı F is also surjective;
• if F and G are both bijective, then G ı F is also bijective;
• if G ı F is injective, then F is injective;
• if G ı F is surjective, then G is surjective;
• if G ı F is bijective, then F is injective and G is surjective.
Proof. Assume that F and G are both injective. For all distinct elements W 6D X in
F ı1 ”ŒD.G/, the injectivity of F ensures that F.W/ 6D F.X/. Hence G.F.W// 6D
G.F.X// by the injectivity of G. Hence, .G ı F/.W/ D G.F.W// 6D G.F.X// D
.G ı F/.X/, whence .G ı F/.W/ 6D .G ı F/.X/, so that G ı F is injective.
Assume that F and G are both surjective. For each Z 2 C, the surjectivity of
G ensures the existence of an element Y 2 B such that G.Y/ D Z. Hence, by the
surjectivity of F, there exists an element X 2 A for which F.X/ D Y. Therefore,
.G ı F/.X/ D G.F.X// D G.Y/ D Z, which means that .G ı F/ is surjective.
In particular, if F and G are both bijective, then G ı F is also bijective.
www.pdfgrip.com
F ı1 ı F D ID.F/ ;
F ı F ı1 D ID.G/ :
is invertible, with inverse F ı1 WD f.1; 0/; .2; 1/; .4; 2/; .8; 3/g.
3.126 Theorem. Each function F W A ! B admits at most one inverse function.
Moreover, if G W B ! A is an inverse function for F, then G consists of all pairs
obtained by swapping the coordinates in each pair of F.
Proof. Assume that G is an inverse function for F, which means that G ı F D ID.F/
and F ı G D ID.G/ , and consider the set
H WD f.Y; X/ W .X; Y/ 2 Fg :
G WD f.Y; X/ W .X; Y/ 2 Fg :
G WD f.1; 0/; .2; 0/; .3; 1/; .4; 0/; .5; 0/; .6; 0/; .7; 0/; .8; 0/; .9; 2/g:
3.130 Example. For A WD f4; 6; 8; 9g and B WD f2; 3g, the function F W A ! B with
F WD f.4; 2/; .6; 3/; .8; 2/; .9; 3/g has a right inverse G WD f.2; 4/; .3; 6/g
3.131 Example. Perspectives to draw a picture of space A on a flat screen B A
can be represented by a function F W A ! B, mapping each point X in space A to its
image F.X/ on the screen B. For such perspectives, each point Y 2 B on the screen
B is its own image, so that F.Y/ D Y. Thus the inclusion function B W B ! A is a
right inverse for F, because .F ı B /.Y/ D FŒB .Y/ D F.Y/ D Y for every Y 2 B,
www.pdfgrip.com
This subsection shows that all the functions from any fixed set into any fixed co-
domain form a set.
Theorem 3.132 shows that all the functions between two fixed sets form a set.
3.132 Theorem. For all sets A and B, all the functions from any subset of A to B are
the elements of a set, of which all such functions defined on all of A form a subset.
Proof. By definition 3.66, the power set P.A B/ is the set of all the relations
between A and B. By definition 3.81, every function F W A ! B defined on any
subset of A is also a relation, so that F 2 P.A B/. By the axiom of separation
(page 124), with a formula stating that a relation F 2 P.A B/ is a function, all
such functions form a subset of P.A B/, denoted here by FA!B :
FA!B
˚
WD F 2 P.A B/ W 8W8X8Y Œ.X; Y/ 2 F ^ Œ.W; Y/ 2 F ) .X D W/ :
For each subset D A, again by the axiom of separation (page 124), with a formula
stating that the domain of F is D, the functions F W D ! B defined on all of D form
a subset of FD!B , denoted here by BD :
˚
BD WD F 2 FD!B W 8X .X 2 D/ ) 9YŒ.X; Y/ 2 F :
zero. This function of zero variable allows for a minimal set of starting functions in
some contexts, for instance, with primitive recursive functions [109, p. 926].
[
3.135 Theorem. For all sets A and B, BD D FA!B .
DA
Proof. For each subset D A, the set BD consists of all the functions F W D ! B
defined on all of D A, whence BD FA!B , or, equivalently, BD 2 P.FA!B /
and hence fBD 2 P.FA!B / W D 2 P.A/g S is also a set. Consequently so is its union
by the axiom of union (page 127): DA BD FA!B . Conversely, each function
F 2 FA!B is defined Son all of its domain, which is a subset D A, so that F 2 BD .
Hence also FA!B DA BD . t
u
3.136 Theorem. For all sets A and B, for each set D P.A/ of subsets of A, all
the D
S sets B Sfor every D 2 D also form a set FD;B P.FA!B /. Hence the union
FD;B D D2D BD is also a set.
Proof. For each set D P.A/ of subsets of A, all the sets BD for every D 2 D
also form a subset of P.FA!B /, denoted here by FD;B :
˚
FD;B WD E 2 PŒP.A B/ W 9DŒ.D 2 D/ ^ .E D BD / ;
H W CAB ! .CB /A ;
F 7! H.F/;
fŒH.F/.X/g.Y/ WD F.X; Y/:
www.pdfgrip.com
L W .CB /A ! CAB ;
G 7! L.G/;
ŒL.G/.X; Y/ WD ŒG.X/.Y/:
Then H ı L D I.CB /A and L ı H D ICAB , whence H and L are inverses of each other:
(X, X) X Y (Y,Y)
X Y Z
(Y, X)
(X, Z) (Y, Z)
R WD f.0; 0/; .1; 1/; .2; 2/; .3; 3/; .0; 2/; .2; 0/; .1; 3/; .3; 1/g
D A [ f.0; 2/; .2; 0/; .1; 3/; .3; 1/g:
www.pdfgrip.com
• The relation R is symmetric: it contains .0; 2/, .2; 0/, as well as .1; 3/, .3; 1/.
• The relation R is transitive because if it contains .X; Y/ and .Y; Z/, then it
contains .X; Z/, for instance, .0; 2/ and .2; 0/ hence .0; 0/; .2; 0/ and .0; 2/ hence
.2; 2/, as well as .1; 3/ and .3; 1/ hence .1; 1/; .3; 1/ and .1; 3/ hence .3; 3/.
3.153 Counterexample. The relation of membership 2 is not an equivalence
relation in general, because it is not symmetric and not transitive.
3.154 Definition (partition). A partition of a set A is a set F P.A/ of subsets
of A with all of the following properties.
• No member of F is empty: 8Vf.V 2 F / ) .V 6D ¿/g: S
• The union of all the members of F “covers” A, which means that A D . F / :
• All pairs of distinct members of F are disjoint:
8V 8WfŒ.V 2 F / ^ .W 2 F / ^ .V 6D W/ ) Œ.V \ W/ D ¿g :
3.155 Example. The empty set ¿ admits only one partition: the empty set F D ¿.
3.156 Example. Each nonempty set S has a partition: the singleton F D fSg.
3.157 Example.
˚ The set AWD f0; 1; 2; 3g admits a partition F into two disjoint
sets, F WD f0; 2g; f1; 3g :
• No member of F is empty: f0; 2g 6D ¿ and f1; 3g 6D ¿.
• The unions of the members of F covers A, so that f0; 1; 2; 3g D f0; 2g [ f1; 3g.
• Distinct members of F are disjoint: f0; 2g \ f1; 3g D ¿.
3.158 Definition (relations from partitions). For each partition F P.A/ of
each set A, define a relation RF on A so that R relates elements X 2 A and Y 2 A if
and only if X and Y belong to the same element of the partition F :
3.159 Theorem (equivalence relations from partitions). For each set A and for
each partition F of A, the relation RF is an equivalence relation.
Proof. The relation RF is reflexive: for each X 2 A, the partition F has an element
B 2 F that contains X, because F covers A. Thus X and X both belong to the same
B 2 F , whence .X; X/ 2 RF . Symbolically,
www.pdfgrip.com
3.162 Example. For the set A WD f0; 1; 2; 3g, the equivalence relation
R WD f.0; 0/; .1; 1/; .2; 2/; .3; 3/; .0; 2/; .2; 0/; .1; 3/; .3; 1/g
D A [ f.0; 2/; .2; 0/; .1; 3/; .3; 1/g
PW A ! A=R;
X 7! ŒXR :
www.pdfgrip.com
3.166 . For the set A WD f0; 1; 2; 3; 4; 5g, verify that the following relation S is an
equivalence relation, and list all its equivalence classes:
8 9
ˆ
ˆ .5; 2/ .5; 5/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 4/ .4; 4/ >
>
ˆ
< >
=
.0; 3/ .3; 3/
S WD :
ˆ
ˆ .2; 2/ .2; 5/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 1/ .4; 1/ >
>
:̂ >
;
.0; 0/ .3; 0/
3.167 . For the set B WD f0; 1; 2; 3; 4; 5; 6; 7g, verify that the following set F of
subsets is a partition, and list the corresponding equivalence relation:
˚
F WD f0; 2; 4; 6g; f1; 3; 5; 7g :
3.168 . For the set B WD f0; 1; 2; 3; 4; 5; 6; 7g, verify that the following set G of
subsets is a partition, and list the corresponding equivalence relation:
˚
G WD f0; 4g; f1; 5g; f2; 6g; f3; 7g :
3.166 Example. The empty relation ¿ is strict, because the conclusion .X; X/ … ¿
is universally valid.
3.167 Example. For each set A, the relation of strict inclusion is strict on P.A/.
Indeed, for all subsets V A and W A, the definition of V W includes the
requirement that V 6D W, so that .V W/ ) .V 6D W/. Contraposition then
confirms that .V D W/ ) .V 6 W/, so that V 6 V.
One method to specify a ranking or direction on a set removes the requirement
of symmetry from the concept of equivalence, which gives the following concept of
preorder.
3.168 Definition (preorder or quasi-order). For each set A, a relation R A A
is a preorder or a quasi-order, if and only if R is reflexive and transitive. It is a
strict preorder if and only if R is irreflexive and transitive.
3.169 Example. Consider the set A WD f0; 1; 2g.
The relation Q WD f.0; 0/; .0; 1/; .1; 0/; .1; 1/; .2; 2/g is a preorder.
The relation R WD f.0; 0/; .0; 1/; .1; 1/; .2; 2/g is a preorder.
The relation S WD f.0; 1/g is a strict preorder.
3.170 Example. For each set A, the diagonal A is a preorder. If A 6D ¿, then A is
not strict, because there exists X 2 A and then .X; X/ 2 A .
3.171 Example. For each set A the relation is a preorder on P.A/. The relation
is not strict because ¿ 2 P.A/ and ¿ ¿.
3.172 Example. For each set A, the relation is a strict preorder on P.A/.
The concept of a preorder R allows for “circular” rankings, with .X; Y/ 2 R and
.Y; X/ 2 R even though X 6D Y. To specify different types of rankings, a different
concept — anti-symmetry — becomes necessary.
www.pdfgrip.com
(Y, X)
As their names might suggest, neither preorders nor partial orders need relate all
elements to one another. Relations that do so are called strongly connected.
3.184 Definition (Strong connectivity). For each set A, a relation R A A
is strongly connected if and only if .X; Y/ 2 R or .Y; X/ 2 R (including both
possibilities) for each X 2 A and for each Y 2 A:
8X 8Y Œ.X 2 A/ ^ .Y 2 A/ ) fŒ.X; Y/ 2 R _ Œ.Y; X/ 2 Rg :
The geometric “direction along a line” corresponds to a total order, also called
linear order, complete order [71, p. 14], or simple order [71, p. 14], [128, p. 69].
3.187 Definition (total order). For each set A, a relation R A A is a total order,
or total ordering, if and only if R is a strongly connected partial order (strongly
connected, reflexive, anti-symmetric, and transitive). It is a strict total order if and
only if R is connected, irreflexive, and transitive.
3.188 Example. The empty relation on the empty set is a strict total order.
3.189 Example. For each set S and the singleton A D fSg, the relation on
P.A/ D f¿; fSgg, is a total order. Indeed, ¿ ¿, ¿ fSg, and fSg fSg.
3.190 Counterexample. The relation is not a total order in general. Thus, if
˚
A WD ¿; f¿g ;
because
˚
˚ f¿g
6 f¿g ;
f¿g 6 f¿g:
˚
Thus
˚ the contains neither the pair f¿g;
relation f¿g nor the pair
f¿g ; f¿g . Instead, for this set A the relation takes the following form:
˚ ˚
f¿g ¿; f¿g
[ [
¿ f¿g
Some relations that are not total orders can restrict to total orders on subsets.
3.191 Definition (chain). For each set A, and for each relation R A A, a subset
B A is a chain if and only if the restriction RjB is a total order on B.
In particular, for each total order R on each set A, the set A is a chain relative to R.
In partially ordered sets, a subset might have a first element, which precedes
every element of that subset, or a last element, which follows every element of that
subset.
3.192 Definition (first or last element). For each set A, for each subset B A,
and for each partial order R A A, an element X 2 B is a minimum, or first, or
smallest, element in B if and only if .X; Y/ 2 R for each Y 2 B. A first element of
B is also denoted by min.B/. Similarly, an element Z 2 B is a maximum, or last,
or largest, element in B if and only if .Y; Z/ 2 R for each Y 2 B. A last element of
B is also denoted by max.B/.
3.193 Example. For each set A and for the relation on P.A/, the element ¿ 2
P.A/ is a first element of P.A/. Indeed, ¿ C for each C 2 P.A/. Also, the
element A 2 P.A/ is a last element of P.A/. Indeed, C A for each C 2 P.A/.
3.194 Definition (well-ordering). For each set A and for each partial order R
A A, the set A is well-ordered by R if and only if each nonempty subset B A has
a first element. A relation R is called a well-ordering (for the lack of a grammatical
and logically equivalent terminology) if and only if A is well-ordered by R.
3.195 Remark. Instead of requiring that a well-ordering be a partial order [30,
p. 31], some texts impose the stronger requirement that a well-ordering be a total
order [71, p. 29]. Yet the insistence on a total order is redundant; indeed, if each
nonempty subset has a first element, then a partial order is automatically a total
order [128, p. 74–76].
www.pdfgrip.com
3.196 Example. The set f¿g D P.¿/ is well-ordered by the relation of inclusion
. Indeed, the only nonempty subset of f¿g D P.¿/ is f¿g, and it has a first
element, namely ¿, because ¿ C for each C 2 f¿g.
3.197 Theorem. Every subset of a well-ordered set is well-ordered.
Proof. Each nonempty subset E B of a subset B A of a set A with a well-order
is also a subset E A and hence has a first element. Thus induces a well-order
on B. t
u
3.198 Definition (upper or lower bound). For each set A, for each subset B A,
and for each partial order R A A, an element Z 2 A is an upper bound for B
if and only if .Y; Z/ 2 R for each Y 2 B. Similarly an element X 2 A is a lower
bound for B if and only if .X; Y/ 2 R for each Y 2 B.
Thus a last element can differ from an upper bound because an upper bound
need not belong to the same subset. Similarly a first element can differ from a lower
bound because a lower bound need not belong to the same subset.
3.199 Definition (maximal element). For each set A, for each subset B A, and
for each partial order R A A, an element Z 2 B is a maximal element of B if
and only if Œ.Z; Y/ 2 R ) Œ.Y; Z/ 2 R for each Y 2 B. In other words, Z 2 B
is a maximal element of B if and only if every element Y 2 B that follows Z also
precedes Z (which allows for the possibility that no Y 2 B follows Z).
Similarly, an element Z 2 B is a minimal element of B if and only if Œ.Y; Z/ 2
R ) Œ.Z; Y/ 2 R for each Y 2 B.
Thus, either a maximal element precedes no element, or it follows every element
that it precedes, but in either case it belongs to the same subset. For instance, a
last element is a maximal element. However, a maximal element need not be a last
element.
3.200 Example. Consider the set A WD f0; 1; 2g and the relation
Q WD f.0; 0/; .0; 1/; .1; 0/; .1; 1/; .2; 2/g:
The concept of well-ordered sets allows for many logically equivalent statements
of the last axiom of Zermelo-Frænkel set theory [30, Ch. 1, 9, p. 23, #9.2(3)].
www.pdfgrip.com
For the following exercises, determine all of the characteristics — among connec-
tivity, strong connectivity, reflexivity, irreflexivity, symmetry, anti-symmetry, asym-
metry, transitivity — of the given relation on the set A WD f0; 1; 2; 3; 4; 5; 6; 7; 8; 9g.
3.171 .
8 9
ˆ
ˆ .0; 9/ .1; 9/ .3; 9/ .6; 9/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 8/ .1; 8/ .2; 8/ .4; 8/ .6; 8/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 7/ .1; 7/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 6/ .1; 6/ .2; 6/ .3; 6/ .4; 6/ >
>
ˆ
< >
=
.0; 5/ .1; 5/
Q WD :
ˆ
ˆ .0; 4/ .1; 4/ .2; 4/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 3/ .1; 3/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 2/ .1; 2/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 1/ >
>
:̂ >
;
3.172 .
8 9
ˆ
ˆ .0; 9/ .1; 9/ .3; 9/ .6; 9/ .9; 9/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 8/ .1; 8/ .2; 8/ .4; 8/ .6; 8/ .8; 8/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 7/ .1; 7/ .7; 7/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 6/ .1; 6/ .2; 6/ .3; 6/ .4; 6/ .6; 6/ >
>
ˆ
< >
=
.0; 5/ .1; 5/ .5; 5/
R WD :
ˆ .0; 4/
ˆ .1; 4/ .2; 4/ .4; 4/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 3/ .1; 3/ .3; 3/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 2/ .1; 2/ .2; 2/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 1/ .1; 1/ >
>
:̂ >
;
.0; 0/
www.pdfgrip.com
3.173 .
8 9
ˆ
ˆ .0; 9/ .1; 9/ .3; 9/ .6; 9/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 8/ .1; 8/ .2; 8/ .4; 8/ .6; 8/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 7/ .1; 7/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .0; 6/ .1; 6/ .2; 6/ .3; 6/ .4; 6/ .8; 6/ .9; 6/ >
>
>
>
< =
.0; 5/ .1; 5/
S WD :
ˆ .0; 4/
ˆ .1; 4/ .2; 4/ .6; 4/ .8; 4/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .0; 3/ .1; 3/ .6; 3/ .9; 3/ >
>
>
>
ˆ
ˆ >
>
ˆ
ˆ .0; 2/ .1; 2/ .4; 2/ .6; 2/ .8; 2/ >
>
ˆ
ˆ >
ˆ .0; 1/ .2; 1/ .3; 1/ .4; 1/ .5; 1/ .6; 1/ .7; 1/ .8; 1/ .9; 1/ >
>
>
:̂ ;
.1; 0/ .2; 0/ .3; 0/ .4; 0/ .5; 0/ .6; 0/ .7; 0/ .8; 0/ .9; 0/
3.174 .
8 9
ˆ
ˆ .0; 9/ .1; 9/ .3; 9/ .6; 9/ .9; 9/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 8/ .1; 8/ .2; 8/ .4; 8/ .6; 8/ .8; 8/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 7/ .1; 7/ .7; 7/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .0; 6/ .1; 6/ .2; 6/ .3; 6/ .4; 6/ .6; 6/ .8; 6/ .9; 6/ >
>
>
>
< =
.0; 5/ .1; 5/ .5; 5/
U WD :
ˆ
ˆ .0; 4/ .1; 4/ .2; 4/ .4; 4/ .6; 4/ .8; 4/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .0; 3/ .1; 3/ .3; 3/ .6; 3/ .9; 3/ >
>
>
>
ˆ
ˆ >
>
ˆ
ˆ .0; 2/ .1; 2/ .2; 2/ .4; 2/ .6; 2/ .8; 2/ >
>
ˆ
ˆ >
ˆ .0; 1/ .1; 1/ .2; 1/ .3; 1/ .4; 1/ .5; 1/ .6; 1/ .7; 1/ .8; 1/ .9; 1/ >
>
>
:̂ ;
.0; 0/ .1; 0/ .2; 0/ .3; 0/ .4; 0/ .5; 0/ .6; 0/ .7; 0/ .8; 0/ .9; 0/
3.175 .
8 9
ˆ
ˆ .1; 9/ .9; 9/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 8/ .8; 8/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 7/ .7; 7/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 6/ .6; 6/ >
>
ˆ
< >
=
.1; 5/ .5; 5/
V WD :
ˆ
ˆ .1; 4/ .4; 4/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .1; 3/ .3; 3/ .9; 3/ >
>
>
>
ˆ
ˆ >
>
ˆ
ˆ .1; 2/ .2; 2/ .4; 2/ .8; 2/ >
>
ˆ
ˆ >
>
ˆ .0; 1/ .1; 1/ >
>
:̂ ;
.1; 0/
www.pdfgrip.com
3.176 .
8 9
ˆ
ˆ .1; 9/ .3; 9/ .9; 9/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 8/ .2; 8/ .8; 8/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 7/ .7; 7/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 6/ .6; 6/ >
>
ˆ
< >
=
.1; 5/ .5; 5/
W WD :
ˆ
ˆ .1; 4/ .2; 4/ .4; 4/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .1; 3/ .3; 3/ .9; 3/ >
>
>
>
ˆ
ˆ >
>
ˆ
ˆ .1; 2/ .2; 2/ .4; 2/ .8; 2/ >
>
ˆ
ˆ >
ˆ .0; 1/ .1; 1/ .2; 1/ .3; 1/ .4; 1/ .5; 1/ .6; 1/ .7; 1/ .8; 1/ .9; 1/ >
>
>
:̂ ;
.1; 0/
3.177 . Prove that for each set A the empty relation ¿ A A is a partial order.
3.178 . Prove that a relation R is symmetric if and only if R D R ı1 .
3.179 . Prove that a relation R on A is irreflexive if and only if R \ A D ¿.
3.180 . Prove that a relation R on a set A is anti-symmetric if and only if R \
R ı1 A .
3.181 . Prove that every strict partial order is asymmetric
3.182 . Prove that if a relation is irreflexive and anti-symmetric, then it is also
asymmetric
3.183 . Prove that every asymmetric relation is also anti-symmetric.
3.184 . Prove that every asymmetric relation is also irreflexive.
3.185 . Prove that a relation is a strict partial order if and only if it is asymmetric
and transitive
3.186 . Provide an example of an anti-symmetric but not asymmetric relation.
3.187 . Provide an example of an asymmetric and strongly connected relation.
3.188 . Exhibit an example of a set A for which the relation on P.A/ is not a
total order: supply subsets B and C such that neither B C nor C B.
3.189 . Exhibit an example of a set A and a subset S P.A/ that is a chain with
respect to the relation on P.A/.
3.190 . Prove that if a function is surjective, then it has a right inverse.
www.pdfgrip.com
Chapter 4
Mathematical Induction: Definitions and Proofs
by Induction
4.1 Introduction
The Principle of Mathematical Induction forms the theoretical basis for such
algorithms as counting and arithmetic, by means of sequences of definitions,
computations, and verifications, where the length of the sequence depends on the
situation. One framework to allow for sequences of yet unspecified lengths consists
in embedding all such sequences into one set N that already allows for sequences
of all lengths. To this end, with a construction attributed to John von Neumann [8,
p. 22], [128, p. 129], [135], for each element X 2 N, the set N also contains a “next”
element X [ fXg.
4.1 Definition. For each set X the successor of X is X [ fXg.
4.2 Example. The successor of the empty set ¿ is the set ¿ [ f¿g D f¿g.
A subsequent theorem will verify that for each set X 2 N the successor X [ fXg
is strictly larger than X, so that X ¤ .X [ fXg/. Within the theory presented so
far, however, nothing guarantees the existence of a set containing the successor of
each of its elements. To this end, a new axiom becomes necessary [8, p. 21]. The
following version is identical to that of axiom S7 already mentioned in chapter 3.
Axiom S7 (Axiom of infinity) There exists a set I, such that ¿ 2 I, and for each
element C 2 I its successor C [ fCg is also an element of I:
` 9I .¿ 2 I/ ^ 8C f.C 2 I/ ) Œ.C [ fCg/ 2 Ig :
Digits or other symbols can abbreviate the set notation for the elements of I.
4.3 Example. The set I described in axiom S7 contains the following elements.
0 WD ¿;
˚ n˚ o n ˚ o
3 WD 2 [ f2g D ¿; f¿g [ ¿; f¿g D ¿; f¿g; ¿; f¿g ;
˚ n ˚ o
4 WD 3 [ f3g D ¿; f¿g; ¿; f¿g ; ¿; f¿g; ¿; f¿g ;
::
:
Such elements are called “natural numbers” (see remark 4.7 about 0).
The axiom of infinity does not restrict I to contain only the elements in
example 4.3, so that the set I might also contain other elements. Therefore, a further
construction with the preceding axioms becomes necessary to form a set containing
only the natural numbers. The construction selected here forms the intersection of
subsets of I, using a method common to several parts of mathematics [30, p. 66],
[34, p. 21], [60, p. 132]. Specifically, denote by P the logical formula defined by
.P/ , .¿ 2 A/ ^ 8X f.X 2 A/ ) Œ.X [ fXg/ 2 Ag :
Thus the formula P asserts that a set A contains among its elements the empty set
and the successor of every one of its elements. The Axiom of Infinity asserts that
there exists a set I for which SubfAI .P/ is True:
` 9I .¿ 2 I/ ^ 8X f.X 2 I/ ) Œ.X [ fXg/ 2 Ig :
so that
G W N ! N;
X 7! X [ fXg:
The following theorem forms the theoretical basis for the methods of proof by
induction and recursive computation. Specifically, the “Principle of Mathematical
Induction” shows that if a subset S N contains ¿, and if S also contains the
successor X [ fXg of each of its elements X, then S contains all the natural numbers:
S D N.
4.15 Theorem (Principle of Mathematical Induction). For each subset S N, if
S contains the
˚ empty set and the successor
one of its elements, then S D N.
of every
Thus, ` 8S .S N/ ^ SubfAS .P/ ) .S D N/ ; or, with SubfAS .P/ spelled out:
` 8S .S N/ ^ .¿ 2 S/ ^ 8X f.X 2 S/ ) Œ.X [ fXg/ 2 Sg ) .S D N/ :
T
Proof. If S N then S N D F I. If moreover SubfAST .P/ is True, then
ST2 F by definition
T of F . From S 2 F , it then follows that TF S. Thus,
F S F , whence (by theorem 3.10) equality holds: S D F D N. u t
4.16 Remark. The proof of theorem 4.15 derives, or, equivalently, deduces the
Principle of Mathematical Induction from the preceding axioms and inference rules
of logic and axioms or postulates of set theory. Therefore, in mathematics, all
inductive proofs are deductive:
In mathematical English, the words “inductive” and “deductive” may be used interchange-
ably for any kind of reasoning, except that the principle listed in [theorem 4.15] is usually
called the Induction Principle [114, p. 195, Remark].
The proof of the following theorem shows the use of the Principle of Mathemat-
ical Induction to verify the pattern of example 4.17 for all natural numbers.
4.18 Theorem. For all L 2 N and N 2 N, if L 2 N then L N.
Proof. This proof proceeds by induction with N. Define a set S N by
S WD fN 2 N W 8LŒ.L 2 N/ ) .L N/g:
Initial step
Inductive hypothesis
Inductive step
From the initial step ¿ 2 S, and from the inductive step, if K 2 S then .K C 1/ 2 S;
it follows from the Principle of Mathematical Induction (theorem 4.15) that S D N.
Thus N 2 S for every N 2 N: for every N 2 N, if L 2 N then L N. t
u
Theorem 4.18 holds for all natural numbers, but it can fail for other sets, as
demonstrated in counterexample 4.19.
˚
4.19 Counterexample. If X D f¿g and Y D f¿g , then X 2 Y but X 6 Y,
because ¿ 2 X but ¿ … Y . Theorem 4.18 does not apply to Y because Y … N.
Similarly, theorem 4.20 shows that every natural number is also a subset of N.
4.20 Theorem. For every set N, if N 2 N, then N N.
www.pdfgrip.com
Initial step
F0 W f0g ! C;
0 7! A:
Then the function F0 just defined satisfies the following two conditions:
(0:0) F0 .0/ D A.
(0:1) F0 .I C 1/ D GŒF0 .I/ for each I 2 S0 n f0g D ¿.
Moreover, there exists only one function of singletons F0 W f0g ! fAg.
Induction hypothesis
Assume that there exists a natural number J 2 N such that the theorem holds for
N WD J, so that for each L 2 SJ D J [ fJg there exists exactly one function FL
satisfying the two conditions .N:0/ and .N:1/ for N WD L.
Induction step
Let S N consist of every N 2 N for which there exists exactly one function FN
satisfying .N:0/ and .N:1/.
There exists exactly one function of singletons HJ W fJ C 1g ! fGŒFJ .J/g,
because there exists exactly one function FJ and hence exactly one value FJ .J/.
Define FJC1 WD FJ [H P J , which is a function because of the disjoint domains
SJ \ fJ C 1g D ¿ (by definition 3.94 in chapter 3).
For I WD J C 1, the definition of FJC1 gives FJC1 .J C 1/ D HJ .J C 1/ D
GŒFJ .J/ D GŒFJC1 .J/. For each L 2 SJ n fJg D J, the definition of FJC1 gives
FJC1 .L C 1/ D FJ .L C 1/ D GŒFJ .L/ D GŒFJC1 .L/ so that the following two
conditions hold:
(J C 1:0) FJC1 .0/ D FJ .0/ D A,
(J C 1:1) FJC1 .L C 1/ D GŒFJC1 .L/ for every L 2 SJC1 n fJ C 1g.
Moreover, there exists exactly one such function FJC1 because there exists exactly
one restriction FJ jSJ , namely FJ , and exactly one value for FJC1 .J C 1/ D GŒFJ .J/.
This subsection defines the addition of natural numbers and establishes some of
its properties, all by induction. Besides explaining the foundations of integer arith-
metic, the following considerations also provide examples of proofs by induction.
4.22 Definition (Addition). For every M 2 N, define
M C 0 WD M; .A0/
M C 1 WD M [ fMg: .A1/
Then for every M 2 N, define an addition function by induction, so that for every
N 2 N,
M C .N C 1/ WD .M C N/ C 1: .A2/
G W N ! N;
N 7! N C 1;
www.pdfgrip.com
Second, assume that there exists some K 2 N such that the theorem holds for R WD
K, so that .P C Q/ C K D P C .Q C K/ for each P and each Q; then
t
u
The following three theorems show that addition commutes. The first theorem
shows that adding 0 commutes.
4.25 Theorem. For each N 2 N, 0 C N D N.
Proof. Proceed by induction with N. First, establish the conclusion for N WD 0:
0 C 0 D 0 by (A0) with M WD 0:
Second, assume that 0 C K D K for some K 2 N, so that the theorem holds for
N WD K; then
0 C .K C 1/ D .0 C K/ C 1 by (A2);
D KC1 induction hypothesis:
t
u
The second theorem shows that adding 1 commutes.
4.26 Theorem. For each P 2 N, 1 C P D P C 1.
www.pdfgrip.com
Second, assume that 1 C K D K C 1 for some K 2 N, so that the theorem holds for
P WD K; then
PC0 D P by (A0);
D 0 C P by theorem 4.25:
Second, assume that there exists K 2 N such that the theorem holds for Q WD K, so
that R C K D K C R for each R 2 N; then for each P 2 N,
t
u
M 0 WD 0; .M0/
M .N C 1/ WD .M N/ C M: .M1/
www.pdfgrip.com
G WD F .M/ W N ! N;
N 7! N C M;
Second, assume that there exists K 2 N such that the theorem holds for R WD K, so
that .P C Q/ K D .P K/ C .Q K/ for each P and each Q; then
.P C Q/ .K C 1/ D Œ.P C Q/ K C .P C Q/ (M1), M WD P C Q;
D Œ.P K/ C .Q K/ C .P C Q/ induction hypothesis;
D .P K/ C Œ.Q K/ C .P C Q/ theorem 4.24;
D .P K/ C Œf.Q K/ C Pg C Q theorem 4.24;
D .P K/ C ŒfP C .Q K/g C Q theorem 4.27;
D .P K/ C ŒP C f.Q K/ C Qg theorem 4.24;
D Œ.P K/ C P C Œ.Q K/ C Q theorem 4.24;
D ŒP .K C 1/ C ŒQ .K C 1/ (M1) twice:
t
u
The following three theorems show that multiplication commutes. The first
theorem shows that multiplication by 0 commutes.
www.pdfgrip.com
0 0 D 0 .M0/:
Second, assume that there exists K 2 N such that the theorem holds for N WD K, so
that 0 K D 0 ; then
0 .K C 1/ D .0 K/ C 0 .M1/;
D 0C0 induction hypothesis;
D0 .A0/:
t
u
The second theorem shows that multiplication by 1 commutes.
4.33 Theorem. For each nonnegative integer N 2 N, 1 N D N.
Proof. Proceed by induction with N. First, establish the conclusion for N WD 0:
1 0 D 0 .M0/:
Second, assume that there exists K 2 N such that the theorem holds for N WD K, so
that 1 K D K ; then
1 .K C 1/ D .1 K/ C 1 .M1/;
D KC1 induction hypothesis:
t
u
Finally, the third theorem shows that multiplication commutes.
4.34 Theorem. For all nonnegative integers P; Q 2 N, P Q D Q P.
Proof. For each P 2 N, proceed by induction with Q. First, for Q WD 0,
P0 D 0 .M0/;
D 0 P theorem 4.32:
Second, assume that there exists K 2 N such that the theorem holds for Q WD K, so
that P K D K P for each P 2 N; then
P .K C 1/ D .P K/ C P .M1/;
D .K P/ C P induction hypothesis;
D .K P/ C .1 P/ theorem 4.33;
D .K C 1/ P theorem 4.31:
t
u
The next theorem shows that multiplication distributes over addition also on the
left-hand side.
www.pdfgrip.com
P .Q C R/ D .Q C R/ P theorem 4.34;
D .Q P/ C .R P/ theorem 4.31;
D .P Q/ C .P R/ theorem 4.34:
t
u
The following theorem shows that multiplication is associative.
4.36 Theorem. For all P; Q; R 2 N, P .Q R/ D .P Q/ R.
Proof. For all P; Q 2 N, proceed by induction with R. For R WD 0,
P .Q 0/ D P 0 .M0/;
D0 .M0/;
D .P Q/ 0 (M0) with M WD .P Q/:
Second, assume that there exists K 2 N such that the theorem holds for R WD K, so
that P .Q K/ D .P Q/ K for all P; Q 2 N ; then
t
u
There are other arithmetic operations with natural numbers, such as the factorial.
4.37 Definition. For each N 2 N, define NŠ (read “N factorial”) recursively by
0Š WD 1;
.N C 1/Š WD .NŠ/ .N C 1/:
The following exercises involve the following sets (also defined in example 3.35):
0 WD ¿;
1 WD f0g;
2 WD f0; 1g;
www.pdfgrip.com
3 WD f0; 1; 2g;
4 WD f0; 1; 2; 3g;
5 WD f0; 1; 2; 3; 4g;
6 WD f0; 1; 2; 3; 4; 5g;
7 WD f0; 1; 2; 3; 4; 5; 6g;
8 WD f0; 1; 2; 3; 4; 5; 6; 7g;
9 WD f0; 1; 2; 3; 4; 5; 6; 7; 8g:
4.45 . Prove or disprove that addition distributes over multiplication on the left.
4.46 . Prove or disprove that addition distributes over multiplication on the right.
The following exercises refer to the factorial specified by definition 4.37.
4.47 . Identify a set C and functions F and G that fit theorem 4.21 to justify
definition 4.37.
4.48 . Calculate 0Š, 1Š, 2Š, 3Š.
4.49 . Prove or disprove that .P C Q/Š D .PŠ/ C .QŠ/ for all P; Q 2 N.
4.50 . Prove or disprove that .P Q/Š D .PŠ/ .QŠ/ for all P; Q 2 N.
The set of natural numbers can model geometric concepts, for instance, a direction
from left to right, and algebraic concepts, for instance, increasing magnitudes:
0 < 1 < 2 < 3 < 4 < 5 < 6 < 7 < 8 < 9 < :::
.M N/ , Œ.M D N/ _ .M 2 N/:
From this well-ordering relation will result the laws of arithmetic cancellations,
which will also allow for the solutions of certain equations. The first results define
a strict order < in terms of the foundational relation 2 of set membership.
4.38 Definition. For all M; N 2 N, define M < N (read “M is less than N”) by
.M < N/ , .M 2 N/:
8L8M8NfŒ.L 2 N/ ^ .M 2 N/ ^ .N 2 N/
Initial step
Induction hypothesis
Assume that there exists some K 2 N such that the theorem holds for N WD K, so
that Œ.L 2 M/ ^ .M 2 K/ ) .L 2 K/ for all L; M 2 N.
Induction step
Initial step
Induction hypothesis
Assume that there exists some K 2 N such that the theorem holds for M WD K, so
that K … K.
Induction step
M 2 N;
M D N;
N 2 M:
Proof. First, observe that at most one of the three formulae may hold. Indeed, M …
M by theorem 4.42, whence .M 2 N/ ^ .M D N/ is False, and .M D N/ ^ .N 2 M/
is False. Similarly, also by theorem 4.42, .M 2 N/ ^ .N 2 M/ is also False. Second,
at least one of the three formulae must hold. This proof uses induction with N.
For N WD 0, and for each M 2 N either M D 0 D N is True, or M 6D 0, and then
0 2 M is True by theorem 4.45. To complete the induction, assume that there exists
K 2 N such that the theorem holds for N WD K, so that .M 2 K/ _ .M D K/ _ .K 2
M/ for each M 2 N, and examine all three formulae.
If M 2 K, then M 2 K K C 1, whence M 2 K C 1.
If M D K, then M D K 2 fKg K [ fKg, whence M 2 K C 1.
If K 2 M, then K C 1 2 M C 1 D M [ fMg, (theorem 4.44); two cases occur.
In the first case, K C1 2 fMg, whence K C1 D M. In the second case, K C1 2 M
already. Thus, .M 2 ŒK C 1/ _ .M D ŒK C 1/ _ .ŒK C 1 2 M/ is True. t
u
The foregoing result completes the proof that < is a strict total order (irreflexive,
asymmetric, and transitive) on the natural numbers. The following theorem shows
that the natural numbers are well-ordered by the relation .
4.47 Theorem (Well-Ordering Principle). Each nonempty subset of the natural
numbers has a smallest element.
Proof. By contraposition, this proof establishes that every subset of the natural
numbers without a smallest element (definition 3.192) is empty. To this end, assume
that a subset E N has no smallest element. Thus every N 2 N is not the smallest
element of E, which means that N … E or that E contains an element M such that
M < N.
www.pdfgrip.com
S WD fN 2 N W .N [ fNg/ .N n E/g:
Initial step
Induction step
S WD fN 2 N W 8IŒ.I 2 E/ ) .I N/g:
fI; : : : ; Ng WD fK 2 N W .I K/ ^ .K N/g;
which thus consists of all the natural numbers from I through N . Similarly,
which thus consists of all the natural numbers larger than or equal to I .
4.52 Example. If E WD f2; : : : ; 7g, then E D f2; 3; 4; 5; 6; 7g. Also, min.E/ D 2
and max.E/ D 7.
4.53 Example. If E WD f2; 3; : : :g, then min.E/ D 2 but E has no maximum.
This subsection establishes rules to cancel terms in equations with additions or mul-
tiplications. The material also provides more practice with proofs by mathematical
induction. The first rule shows how to solve equations of the form M C 1 D L C 1.
4.54 Theorem. For all L; M 2 N, if M C 1 D L C 1, then M D L.
Proof. If M C 1 D L C 1, then M [ fMg D L [ fLg, and the sets on both sides have
the same elements. In particular, L 2 .L [ fLg/ and thus L 2 .M [ fMg/, whence
two cases arise: L 2 fMg, or L 2 M.
In the first case, L 2 fMg, and then L D M indeed.
In the second case, L 2 M; then M … fLg, for otherwise M D L and L 2 L would
contradict theorem 4.42. However, M 2 L [ fLg, whence M 2 L, but that would also
contradict theorem 4.42. Therefore, this second case does not occur. t
u
The following theorem allows for the cancellation of an additive term N common
to both sides of an equation of the type M C N D L C N.
4.55 Theorem. For all M; N; L 2 N, if M C N D L C N, then M D L.
Proof. For all natural numbers L; M 2 N, proceed by induction with N. For N WD 0,
if M C 0 D L C 0, then M D M C 0 D L C 0 D L, by hypothesis and by (A0).
Second, assume that there exists K 2 N such that the theorem holds for N WD K,
so that for all natural numbers L; M 2 N, if M C K D L C K then M D L.
www.pdfgrip.com
L C .K C 1/ D L C .1 C K/
D .L C 1/ C K
< .M C 1/ C K
D M C .1 C K/
D M C .K C 1/:
t
u
4.57 Theorem. For all M; N 2 N, if M 6D 0, then N < N C M.
Proof. Set L WD 0 in theorem 4.56. If M 6D 0 then 0 < M by theorem 4.45, whence
L < M. Hence N D 0 C N < M C N by theorem 4.56. t
u
The following theorem forms the basis for the concept of subtraction of a natural
number from a larger natural number.
4.58 Theorem. For all M; N 2 N, if M < N, then there exists exactly one natural
number L 2 N such that M C L D N.
Proof. This proof establishes the existence and the uniqueness separately.
Existence
Uniqueness
Second, verify the uniqueness of L, which results from the theorem that if M C L D
N D M C K, then L D K (theorem 4.55). t
u
The following definition specifies the concept of subtraction of a natural number
from a larger natural number.
4.59 Definition (Subtraction). For all L; M; N; 2 N, if M < N, then N M WD L
is the natural number L defined in theorem 4.58 such that M C L D N
4.60 Example. 8 5 D 3 because 5 C 3 D 8 by exercise 4.37.
The following theorem shows that multiplication by a nonzero natural number
preserves the ordering.
4.61 Theorem. For all L; M; N 2 N, if M < N and 0 < L, then L M < L N.
Proof. For all nonnegative integers M; N 2 N, use induction with L, the smallest
nonzero value of L being 1. For L WD 1, if M < N, then 1 M D M < N D 1 N.
Assume that there exists K 2 N with 0 < K such that the theorem holds for
L WD K, so that for all natural numbers M; N 2 N, if M < N, then K M < K N.
Thus, if M < N, then apply theorem 4.56 twice:
.K C 1/ M D .K M/ C M < .K M/ C N < .K N/ C N D .K C 1/ N:
t
u
The following theorem allows for the cancellation of a nonzero multiplicative
term on both sides of an equation, which forms the basis for the division of a natural
number by a nonzero natural number, and the solution of equations of the type
L M D L N.
4.62 Theorem. For all L; M; N 2 N, if 0 < L and L M D L N, then M D N.
Proof. Proceed by contraposition.
If M 6D N, then either M < N or N < M.
If M < N, then L M < L N and L M 6D L N.
Similarly, if N < M, then L N < L M and L M 6D L N. t
u
4.63 Definition (division). For all K; M; N 2 N, if 0 < L and K L D N, then
define N=L by N=L WD K, as defined uniquely by theorem 4.62.
www.pdfgrip.com
.M N/ , Œ.M D N/ _ .M 2 N/:
4.5 Integers
Several operations remain undefined with natural numbers. For instance, the
ordering relation does not include elements smaller than the empty set, which
precludes the use of natural numbers to model relations extending in two opposite
directions. Also, the arithmetic of natural numbers does not contain provisions
for the “difference” from a larger natural number to a smaller one. Finally, the
arithmetic of natural numbers does not include any concept of division other than
special cases.
There exist several methods to extend the ordering and the arithmetic of natural
numbers, to allow for elements “smaller” than zero, and for differences of any two
elements. Such methods begin with the specification of a larger set of “integers” Z
[from the German “Zahl(en)” for “number(s)”].
One method of defining a larger set, outlined by Kunen [74, p. 35], is sufficiently
general to produce not only the integers, but also the rational numbers and the real
numbers. Essentially, the method consists in defining the new numbers in terms of
equivalence classes. For the integers, the strategy consists in introducing the concept
of the “difference” between two natural numbers M and N by means of the pair
.M; N/. Then relate .M; N/ to every other pair .P; Q/ with the same “difference”
between P and Q. Because the concept of “difference” has not yet been defined for
all pairs of natural numbers, however, a precise definition uses sums instead.
Two cases arise: either N M, or N > M. In the first case, if N M, then
definition 4.59 specifies their difference J WD M N. If also .M; N/ and .P; Q/
represent the same difference, then P Q D J D M N. An equivalent statement
without subtractions results from adding J to both extremes:
M N D J; M D N C JI
P Q D J; P D Q C J:
N M D J; N D M C JI
Q P D J; Q D P C J:
In either case, the statement that .M; N/ and .P; Q/ represent the same “difference”
can be reworded without subtractions but with sums instead: there exists J 2 N
with
• either M D N C J and P D Q C J (if M > N and P > Q),
www.pdfgrip.com
M N
P Q
N M
Q P
.M; N/ l .P; Q/
m
9J .J 2 N/ ^ fŒ.M D N C J/ ^ .P D Q C J/ _ Œ.N D M C J/ ^ .Q D P C J/g W
.M; N/ l .P; Q/
m
M C Q D N C P:
M C Q D .N C I/ C Q D N C .I C Q/ D N C .Q C I/ D N C P:
M C Q D M C .P C I/ D .M C .I C P/ D .M C I/ C P D N C P:
For the converse, assume that M C Q D N C P. Then two cases can arise.
If M N, then there exists I 2 N such that N D M C I. Hence
www.pdfgrip.com
M C Q D N C P D .M C I/ C P D M C .P C I/
N C P D M C Q D .N C I/ C Q D N C .Q C I/
Reflexivity
Symmetry
Transitivity
K C .M C Q/ D K C .N C P/ D .K C N/ C P D .L C M/ C P D L C .M C P/
Thus the pairs .3; 0/, .4; 1/, .5; 2/, .6; 3/, .7; 4/ are elements of the equivalence class
Œ.3; 0/l .
Each pair .I; J/ 2 N N is equivalent to a pair of the type .K; 0/ or .0; K/.
4.73 Theorem. If I > J, then .I; J/ l .I J; 0/.
If I J, then .I; J/ l .0; J I/.
Proof. If I > J, then I D .I J/ C J and J D 0 C J whence .I; J/ l .I J; 0/.
If I J, then I D 0 C I and J D .J I/ C I whence .I; J/ l .0; J I/. t
u
The following diagram shows a few elements from three equivalence classes:
Œ(0, 2)l , Œ.0; 0/l , and Œ(3, 0)l relative to the relation l on A WD N N:
The following theorem verifies that addition and multiplication of pairs com-
mute.
4.75 Theorem. For all M; N; P; Q 2 N,
.M; N/ C .P; Q/ D .M C P; N C Q/
D .P C M; Q C N/
D .P; Q/ C .M; N/I
t
u
The next theorem checks that arithmetic with equivalent pairs yields equivalent
results: different pairs representing the same difference yield the same sum or
product.
4.76 Theorem. For all I; J; K; L; M; N; P; Q 2 N, if
then
Proof. If .I; J/ l .K; L/, then I C L D J C K. If also .M; N/ l .P; Q/, then
M C Q D N C P. Moreover,
whence
.I C M/ C .L C Q/ D .I C L/ C .M C Q/
D .J C K/ C .N C P/
D .J C N/ C .K C P/;
which means that .I; J/ C .M; N/ l .K; L/ C .P; Q/. For the multiplication,
www.pdfgrip.com
IMCJNCKQCLP D INCJMCKPCLQ
m
I MCI QCJ N CK QCL P D I N CI QCJ MCK PCL Q
m
I .M C Q/ C J N C K Q C L P D I N C J M C K P C .I C L/ Q
m
I .N C P/ C J N C K Q C L P D I N C J M C K P C .J C K/ Q
m
IPCJNCLP D JMCKPCJQ
m
IPCJNCJPCLP D JMCKPCJPCJQ
m
I P C J .N C P/ C L P D J M C .K C J/ P C J Q
m
I P C J .M C Q/ C L P D J M C .I C L/ P C J Q
m
0 D 0;
multiplication commutes,
multiplication is associative,
multiplication distributes over addition, and
Œ.1; 0/l is a multiplicative unit: Œ.M; N/l Œ.1; 0/l D Œ.M; N/l .
Proof. The proof forms the object of exercises. t
u
The following definition specifies a subtraction.
4.79 Definition (Integer subtraction). Define a subtraction of integers by
Œ.M; N/l Œ.P; Q/l WD Œ.M; N/l C Œ.Q; P/l ;
Z WD Z n fŒ.0; 0/l g;
Z WD fŒ.M;
˚ N/l 2 Z W M Ng;
WD Œ.M; N/l 2 Z W 9If.I 2 N/ ^ Œ.M; N/ l .0; I/g ;
ZC WD fŒ.M;
˚ N/l 2 Z W M Ng;
WD Œ.M; N/l 2 Z W 9If.I 2 N/ ^ Œ.M; N/ l .I; 0/g ;
Z WD fŒ.M;
˚ N/l 2 Z W M < Ng;
WD Œ.M; N/l 2 Z W 9If.I 2 N / ^ Œ.M; N/ l .0; I/g ;
ZC WD fŒ.M;
˚ N/l 2 Z W M > Ng
WD Œ.M; N/l 2 Z W 9If.I 2 N / ^ Œ.M; N/ l .I; 0/g :
and
The following theorem verifies that the ordering does not depend on the choice
of the pairs representing the equivalence classes.
www.pdfgrip.com
then
Proof. If .I; J/ l .K; L/, then I C L D J C K. If also .M; N/ l .P; Q/, then also
M C Q D N C P. Consequently, adding L C P to each side of the inequalities yields
t
u
A definition of an order on Z can thus use any pair from an equivalence class.
4.85 Definition. Define an order < on Z by
Theorem 4.86 shows that < is a strict total order (definition 3.187).
www.pdfgrip.com
and
then
Proof. This proof verifies that < is irreflexive, connected, and transitive.
Irreflexivity
(The relation does not relate any element to itself.) For each pair .M; N/ 2 N N,
M C N – N C M, whence .Œ.M; N/l / – .Œ.M; N/l /.
Connectedness
Transitivity
K C N < L C MI
M C Q < N C P:
.K C N/ C .M C Q/ < .L C M/ C .N C P/
KCQ<LCP
then
Proof. The proof uses the equivalence .I; J/ l .I J; 0/. First, from Œ.M; N/l <
Œ.P; Q/l it follows that M C Q < N C P by definition of < on the pairs. Hence a
multiplication throughout by I J > 0 yields
.I J/ .M C Q/ < .I J/ .N C P/
whence
Swapping the roles of .I; J/ with .K; L/, and .M; N/ with .P; Q/, then yields
.I; J/ .P; Q/ D .P; Q/ .I; J/ < .P; Q/ .K; L/ D .K; L/ .P; Q/:
Consequently,
t
u
www.pdfgrip.com
The following theorem shows that the square of every integer is nonnegative.
4.88 Theorem. For every X 2 Z, X X 0.
Proof. For every X 2 Z there exists K 2 N such that X D Œ.K; 0/l or X D
Œ.0; K/l , by theorem 4.73. However,
.K; 0/ .K; 0/ D .K K C 0 0; K 0 C 0 K/
D .K K; 0/
D .0 0 C K K; 0 K C K 0/
D .0; K/ .0; K/
Z WD fK
˚ 2 P.N/ W 9N Œ.N 2 N/ ^ .N 6D 0/ ^ .K D fNg/g
D fNg W N 2 N
D f: : : ; ff¿; f¿gg; ff¿ggg
D f: : : ; f3g; f2g; f1gg :
. M/ C . N/ WD .M C N/;
. N/ C M WD M C . N/ WD .N M/ if M < N;
. N/ C M WD M C . N/ WD M N if N < M:
Similarly,
. N/ . M/ WD M N;
. N/ M WD M . N/ WD .M N/:
The addition and multiplication thus defined for integers remain associative and
commutative, multiplication distributes over addition, N C0 D N and 1N D N for
each integer N 2 Z. Proofs proceed by cases, depending on the sign of the operands,
and are straightforward but lengthy. See [76, Ch. IV] or the exercises.
4.91 Remark. Common usage abbreviates each nonnegative integer Œ.I; 0/l by I.
Also, any variable can denote an integer, for instance, M D Œ.P; Q/l .
M 0 WD 1:
Then for each integer M and for each nonnegative integer J define
M JC1 WD .M J / M:
G WD H .M/ W N ! N;
L 7! M L;
4.94 Example. Here are the first four nonnegative powers of the integer 2:
20 D 1I
21 D .20 / 2 D 1 2 D 2I
22 D .21 / 2 D 2 2 D 4I
23 D .22 / 2 D 4 2 D 8:
www.pdfgrip.com
.M N/J D .M J / .N J /;
N ICJ D .N I / .N J /:
Proof. This proof proceeds by induction with the exponent J. The first equation,
.M N/J D .M J / .N J /, holds for all integers M and N, and for J 2 f0; 1g:
.M N/0 D 1 D 1 1 D .M 0 / .N 0 /I
.M N/1 D M N D .M 1 / .N 1 /:
˚ WD
S
J 2 N W 8M8NfŒ.M 2 N/ ^ .N 2 N/ ) Œ.M N/J D .M J / .N J /g :
Thus, assume that there exists K 2 S, or, equivalently, K 2 N such that the theorem
holds for J WD K; then
whence K C 1 2 S, and hence, S D N, which means that the first equation holds.
The second equation, N ICJ D .N I / .N J /, holds for each integer N, for each
nonnegative integer I, and for J 2 f0; 1g:
N IC0 D N I D .N I / 1 D .N I / .N 0 /;
N IC1 D .N I / N D .N I / .N 1 /:
Next, let
and assume that there exists K 2 S, or, equivalently, K 2 N such that the theorem
holds for J WD K, so that N ICK D .N I / .N K / for all I 2 N and N 2 N; then
www.pdfgrip.com
.P; Q/ .M; N/
m
PN DQM
Reflexivity
Symmetry
Transitivity
If
then
I L D J K;
K N D L M;
whence multiplying the left-hand sides and the right-hand sides together gives
.I L/ .K N/ D .J K/ .L M/;
.I N/ .L K/ D .J M/ .K L/;
IN DJM
Proof. This proof applies the Well-Ordering Principle to the set of all nonnegative
numerators of a rational number. To this end, for each P=Q 2 Q define the set
www.pdfgrip.com
The comparison of two rational numbers P=Q and M=N can proceed through
equivalent fractions with a common denominator, for instance, .P N/=.Q N/ and
.Q M/=.Q N/. Common denominators also lead to an arithmetic with fractions
(ordered pairs) and then with rational numbers (equivalence classes).
4.100 Definition. For all pairs .I; J/ and .K; L/ in Z Z , define functions C and
on .Z Z / .Z Z / by their counterparts C and already defined on Z Z :
.I; J/ .K; L/ WD .I K; J L/ :
The symbols C and on the left-hand sides are the functions being defined, whereas
the symbols C and on the right-hand sides are the addition and multiplication of
integers. Yet common usage employs C and for both.
The following theorem shows that equivalent fractions lead to equivalent results.
4.101 Theorem. If
then
where
.I N C J M/ .L Q/ .J N/ .K Q C L P/
D ŒI N L Q C J M L Q ŒJ N K Q C J N L P
D Œ.I L/ .N Q/ C .J L/ .M Q/
Œ.J K/ .N Q/ C .J L/ .N P/
D Œ.I L/ .N Q/ C .J L/ .M Q/
Œ.I L/ .N Q/ C .J L/ .M Q/
D 0:
where, by hypotheses,
.I M/ .L Q/ .J N/ .K P/
D .I L/ .M Q/ .J K/ .N P/
D .I L/ .M Q/ .I L/ .M Q/
D 0:
I K ILCJK
C WD ;
J L JL
I K IK
WD :
J L JL
The following theorems establish algebraic characteristics of rational arithmetic.
4.103 Theorem. The addition of rational numbers commutes.
Proof. For all I=J and K=L in Q,
I K ILCJK LICKJ K I
C D D D C :
J L JL LJ L J
t
u
www.pdfgrip.com
I K IK KI K I
D D D :
J L JL LJ L J
t
u
4.105 Theorem. The addition of rational numbers is associative.
Proof. For all I=J, K=L, and M=N in Q,
I K M ILCJK M
C C D C
J L N JL N
Œ.I L C J K/ N C Œ.J L/ M
D
.J L/ N
ŒI .L N/ C J .K N/ C ŒJ .L M/
D
J .L N/
ŒI .L N/ C ŒJ .K N C L M/
D
J .L N/
I KNCLM
D C
J LN
I K M
D C C :
J L N
t
u
4.106 Theorem. The multiplication of rational numbers is associative.
Proof. For all I=J, K=L, and M=N in Q,
I K M IK M
D
J L N JL N
.I K/ M
D
.J L/ N
I .K M/
D
J .L N/
www.pdfgrip.com
I KM
D
J LN
I K M
D :
J L N
t
u
4.107 Theorem. For each N 2 N and each I=J 2 Q, multiplications of the
numerator and denominator by a nonzero common factor yields the same rational
number: I=J D .I N/=.J N/.
Proof. Verify the criterion for equivalent fractions: I .J N/ D J .I N/. t
u
4.108 Theorem. The multiplication of rational numbers distributes over addition.
Proof. For all I=J, K=L, and M=N in Q,
I K M ILCJK M
C D
J L N JL N
.I L C J K/ M
D
.J L/ N
.I L/ M C .J K/ M/
D
J .L N/
.I M/ L C J .K M/
D
J .L N/
.I M/ .L N/ C .J N/ .K M/
D
.J N/ .L N/
IM KM
D C
JN LN
I M K M
D C :
J N L N
t
u
The following theorem shows that adding 0=1 does not produce any change.
4.109 Theorem. For each K=L 2 Q, .K=L/ C .0=1/ D .K=L/.
Proof. .K=L/ C .0=1/ D .ŒK 1 C L 0=L 1/ D .K=L/. t
u
Thus, the rational number 0=1 plays the same role as the integer 0 in additions.
The following theorem shows that each rational number has an additive inverse.
www.pdfgrip.com
4.110 Theorem. Each I=J 2 Q has an additive inverse: .I=J/ C .Œ I=J/ D .0=1/.
Proof.
I I I J C J . I/
C D
J J JJ
J I C J . I/
D
JJ
J ŒI C . I/
D
JJ
J0
D
JJ
0
D
JJ
0 .J J/
D
1 .J J/
0
D :
1
t
u
The following theorem shows that multiplying by 1=1 changes nothing.
4.111 Theorem. For each K=L 2 Q, .K=L/ .1=1/ D .K=L/.
Proof. .K=L/ .1=1/ D .ŒK 1=ŒL 1/ D .K=L/. t
u
Thus, the rational number 1=1 plays the same role as the integer 1 in multiplica-
tions. Similarly, each nonzero rational number has a multiplicative inverse.
4.112 Theorem. Each I=J 2 Q such that I 6D 0 has a multiplicative inverse: .I=J/
.J=I/ D .1=1/.
Proof.
I J IJ IJ 1 .I J/ 1
D D D D :
J I JI IJ 1 .I J/ 1
t
u
Rational arithmetic thus satisfies the algebraic properties in table 4.115.
4.113 Definition (Field). A field (of numbers) is a set F with at least two different
elements 0 and 1, so that 0 6D 1, and binary operations C and , satisfying the
algebraic properties in table 4.115.
www.pdfgrip.com
Table 4.115 These properties hold for all I=J; K=L; P=Q 2 Q.
(1) Associativity of C Œ.I=J/ C .K=L/ C .P=Q/ D .I=J/ C Œ.K=L/ C .P=Q/
(2) Commutativity of C .I=J/ C .K=L/ D .K=L/ C .I=J/
(3) Additive identity .K=L/ C .0=1/ D .K=L/ D .0=1/ C .K=L/
(4) Additive inverse .K=L/ C .ŒK=L/ D .0=1/
(5) Associativity of Œ.I=J/.K=L/.P=Q/ D .I=J/Œ.K=L/.P=Q/
(6) Commutativity of .I=J/.K=L/ D .K=L/.I=J/
(7) Multiplicative identity .K=L/.1=1/ D .K=L/ D .1=1/.K=L/
(8) Multiplicative inverse If .K=L/ 6D 0,
then .K=L/.L=K/ D .1=1/
(9) Distributivity .I=J/Œ.K=L/ C .P=Q/ D Œ.I=J/.K=L/ C Œ.I=J/.P=Q/
Technically, the pair .C; / suffices to identify the set F and the elements 0 and
1, but for emphasis a field can be defined as the quintuple .F; C; 0; ; 1/.
4.114 Example. The quintuple .Q; C; 0; 1/ is a field (of numbers).
The next theorem forms the basis for a concept of division of rational numbers.
4.116 Theorem. For each K=L 2 Q, and for each I=J 2 Q such that I 6D 0,
Œ.K=L/ .J=I/ .I=J/ D .K=L/.
Proof.
K J I KJ I .K J/ I K .J I/ K .I J/ K
D D D D D :
L I J LI J .L I/ J L .I J/ L .I J/ L
t
u
4.117 Definition. For each K=L 2 Q, and for each I=J 2 Q such that I 6D 0, define
K I K J
WD
L J L I
JC1 J
K K K
WD :
L L L
The notation introduced here proves convenient to define and investigate sums and
products of finite sequences of numbers.
4.119 Definition. A finite sequence of rational numbers is a function S W N ! Q
defined on some N 2 N. The value S.K/ is also denoted by SK ; then the function S
is also denoted by .SK /.
4.120 Example. The function S W 9 ! Q defined by SK WD . 2=3/K is a finite
sequence of numbers:
S0 D . 2=3/0 D 1;
S1 D . 2=3/1 D 2=3;
S2 D . 2=3/2 D 4=9;
S3 D . 2=3/3 D 8=27;
S4 D . 2=3/4 D 16=81;
S5 D . 2=3/5 D 32=243;
S3 D . 2=3/6 D 64=729;
S4 D . 2=3/7 D 128=2187;
S5 D . 2=3/8 D 256=6561:
The next definition gives a notation for the product of a finite sequence of
numbers.
4.121 Definition (Product notation). For each finite sequence of numbers
S W N ! Q, define the “empty product” to be 1:
Y
SK WD 1:
K<0
Then define the product of the first value to be the first value:
0
Y
SK WD S0 :
KD0
Hence for each L 2 N, such that 0 < L < N, define the product of the first L values
of the sequence S “inductively” [77, p. 5] or “recursively” [49, p. 133] by
!
Y
L Y
L1
SK WD SK SL :
KD0 KD0
www.pdfgrip.com
Q1 Q
0
KD0 SK D KD0 SK S1
D .1/ 2=3;
Q2 Q
1
KD0 SK D KD0 SK S2
D .1 2=3/ 4=9;
Q3 Q
2
KD0 SK D KD0 SK S3
D .1 2=3 4=9/ 8=27;
Q4 Q
3
KD0 SK D KD0 SK S4
D .1 2=3 4=9 8=27/ 16=81;
::
:
Q8 Q7
S
KD0 K D KD0 SK S8
256
81 243 729 2187 /
D .1 23 49 278 16 32 64 128
6561
:
The next definition gives a notation for the sum of a finite sequence of numbers.
4.123 Definition (Sum notation). For each finite sequence of numbers S W N ! Q,
define the “empty sum” to be 0:
X
SK WD 0:
K<0
Then define the sum of the first value to be the first value:
0
X
SK WD S0 :
KD0
Hence for each L 2 N, such that 0 < L < N, define the sum of the first L values of
the sequence S inductively by
!
X
L X
L1
SK WD SK C SL :
KD0 KD0
www.pdfgrip.com
P1 P
0
KD0 SK D KD0 SK C S1
D .1/ C 2=3;
P2 P
1
KD0 SK D KD0 SK C S2
D .1 C 2=3/ C 4=9;
P3 P
2
KD0 SK D KD0 SK C S3
D .1 C 2=3 C 4=9/ C 8=27;
P4 P
3
KD0 SK D KD0 SK C S4
D .1 C 2=3 C 4=9 C 8=27/ C 16=81;
::
:
P8 P7
S
KD0 K D KD0 SK C S8
256
81 C 243 C 729 C 2187 / C
D .1 C 23 C 49 C 278 C 16 32 64 128
6561
:
X
N1
1 XN
XK D :
KD0
1 X
Proof. This proof uses induction with N. For N WD 1, and for every X 2 Q n f1g,
11
X 0
X 1 X1
XK D XK D X0 D 1 D :
KD0 KD0
1 X
Assume that there exists I 2 N such that the theorem holds for N WD I and for
every X 2 Q n f1g, so that
www.pdfgrip.com
X
I1
1 XI
XK D :
KD0
1 X
Then
.IC1/1
!
X X
I1
X DK
X K
C XI
KD0 KD0
1 XI
D C XI
1 X
1 XI .1 X/ X I
D C
1 X 1 X
.1 X I / C .X I X IC1 /
D
1 X
1 X IC1
D :
1 X
t
u
An alternative proof of the same formula proceeds along the following outline:
PN1
KD0 XK D1C X C C X N2 C X N1
PN1
X KD0 XK D X C X2 C C X N1 C X N
PN1
.1 X/ KD0 XK D 1 C 0 C C 0 XN
X
N1
1 XN
XK D :
KD0
1 X
Yet such a proof also requires induction to rearrange the terms in the subtraction.
www.pdfgrip.com
1. 2=3/9
D
1 2=3
1 512=19,683
D 1=3
19,171=19683
D 1=3
D 3=1 19171=19683
D 19171=6561:
The determination of whether two rational numbers P=Q and M=N coincide can
utilize equivalent fractions with a common denominator, for instance, .P N/=.Q
N/ and .Q M/=.Q N/, and then with the comparison of the numerators P N and
Q M. The same comparison leads to a concept of order on the rational numbers.
4.127 Definition. Define a relation < on Q as follows. First, 0 < .P=Q/ if and only
if 0 < P Q, so that either P and Q are both positive, or P and Q are both negative:
The following theorem shows that the square of any nonzero rational number,
and the sum and product of positive rational numbers, are positive rational numbers.
4.128 Theorem. If .M=N/ > 0 and .I=J/ > 0, then .M=N/ C .I=J/ > 0 and
.M=N/ .I=J/ > 0. Moreover, .K=L/ .K=L/ > 0 for every .K=L/ 6D .0=1/.
Proof. For the square, let P=Q WD .K=L/2 D .K 2 /=.L2 /. Then P Q D .K 2 /
.L2 / D .K L/2 > 0 (theorem 4.88). Thus .K=L/2 D P=Q > 0 (definition 4.127).
For the product, let P=Q WD .M=N/.I=J/ D .M I/=.N J/. By the hypotheses,
M N > 0 and I J > 0. Hence P Q D .M I/ .N J/ D .M N/ .I J/ > 0,
whence .M=N/ .I=J/ D P=Q > 0 (definition 4.127). For the sum, let
www.pdfgrip.com
P M I .M J/ C .N I/
WD C D :
Q N J NJ
Then
Indeed, J 2 > 0 and N 2 > 0 (theorem 4.88), .M N/.J 2 / > 0 and .N 2 /.I J/ > 0
by hypothesis, whence P Q D .M N/ .J 2 / C .N 2 / .I J/ > 0. t
u
The next theorem shows that < is a strict total order (definition 3.187) on Q.
4.129 Theorem. The relation < is a strict total order on the rational numbers.
Proof. This proof verifies that < is connected, irreflexive, and transitive.
Irreflexivity
Connectedness
Transitivity
If .I=J/ < .K=L/ and .K=L/ < .M=N/, then .K=L/ .I=J/ > 0 and .M=N/
.K=L/ > 0, whence .M=N/ .I=J/ D Œ.M=N/ .K=L/ C Œ.K=L/ .I=J/ > 0
(theorem 4.128) and then .M=N/ > .I=J/ (definition 4.127). t
u
4.130 Definition. The sets of nonzero rationals .Q /, negative rationals .Q /,
positive rationals .QC /, non-positive rationals .Q /, and nonnegative rationals
.QC / are
Q WD Q n f0=1g;
Q WD fP=Q 2 Q W P=Q 0=1g;
QC WD fP=Q 2 Q W P=Q 0=1g;
Q WD fP=Q 2 Q W P=Q < 0=1g;
QC WD fP=Q 2 Q W P=Q > 0=1g:
www.pdfgrip.com
4.131 Definition. Define the absolute value jP=Qj of a rational number P=Q by
P=Q if P=Q 0;
jP=Qj WD
.P=Q/ if P=Q < 0:
A Ð B;
ANN D B;
NN
jAj D jBj;
#.A/ D #.B/;
card .A/ D card .B/:
Definition 4.134 does not yet define the concept of cardinality; it only defines
the concept of same cardinality. Because such a definition leaves the notation #.A/
www.pdfgrip.com
˚
Fig. 4.1 Same cardinality: Earth and Moon, or Pluto and Charon, or ¿; f¿g .
yet undefined, this exposition adopts the notation A Ð B, which merely means that
there exists a bijection from A to B.
4.135 Example. All empty sets have the same cardinality. Indeed, by the axiom of
extensionality there exists only one empty set, namely ¿, and the empty function
¿ W ¿ ! ¿ is a bijection, whence ¿ Ð ¿.
4.136 Example. All singletons have the same cardinality. Indeed, for all sets X and
Y and all singletons fXg and fYg, the function F W fXg ! fYg defined by F WD
f.X; Y/g is a bijection. Thus fXg Ð fYg.
4.137 Example. The sets A WD f4; 9g and B WD f2; 3g have the same cardinality,
thanks to the bijection F W A ! B with F WD f.4; 2/; .9; 3/g. The other bijection,
G WD f.4; 3/; .9; 2/g, could also serve to prove that A and B have the same
cardinality.
The following theorem forms the basis for the relation between the addition of
natural numbers and the union of disjoint sets.
4.138 Theorem. For all sets A, B, C, and D, if
A Ð C;
B Ð D;
A\B D ¿;
C\D D ¿;
then
P Ð .C[D/:
.A[B/ P
www.pdfgrip.com
F W A ! C;
G W B ! D:
P ! .C[D/
H W .A[B/ P
defined by
P
H WD .F [G/ P
.A[B/ P
.C[D/
so that
8
< Y D F.X/ if X 2 A;
Œ.X; Y/ 2 H ,
:
Y D G.X/ if X 2 B:
The relation H just defined is a function, because for each X 2 .A[B/ P the relation
H contains only one pair .X; Y/. Indeed, thanks to A \ B D ¿, either X 2 A and
then Y D F.X/, or X 2 B and then Y D G.X/, but not both (definition 3.94).
To verify the injectivity of H, assume that W 2 .A[B/ P and X 2 .A[B/ P have the
same image H.W/ D Y D H.X/ in C [ D. Because of C \ D D ¿, either both
images lie in C or both lie in D. In the first case, if both images lie in C, then W 2 A
and X 2 A, and then
Z D F.W/ D H.W/:
Z D G.X/ D H.X/:
P ! .C[D/
Therefore, H W .A[B/ P is a bijection. t
u
www.pdfgrip.com
The following theorem forms the basis for the relation between multiplication of
natural numbers and Cartesian products of sets.
4.139 Theorem. For all sets A, B, C, and D, if
A Ð C;
B Ð D;
then
.A B/ Ð .C D/:
F W A ! C;
G W B ! D:
H W .A B/ ! .C D/;
.W; X/ 7! F.W/; G.X/ :
The relation H is a function, because F and G are functions, so that for each W 2 A
and each X 2 B there exists at most one Y 2 C and at most one Z 2 D with
.W;
Y/ 2 F and .X; Z/ 2 G, so that there exists at most one .Y; Z/ 2 C D with
.W; X/; .Y; Z/ 2 H. To verify the injectivity of H, assume that .W; X/ 2 .A B/
and .U; V/ 2 .A B/ have the same image H.W; X/ D H.U; V/ in C D:
To verify the surjectivity of H, assume that .R; S/ 2 .C D/. Then the surjectivity
of F guarantees the existence of an element W 2 A such that
R D F.W/;
www.pdfgrip.com
S D G.X/:
Consequently,
Therefore, H W .A B/ ! .C D/ is a bijection. t
u
The following definition establishes the concept of cardinality for finite sets.
4.140 Definition. For each set S, the set S is finite if and only if there exists N 2 N
and a bijection F W N ! S. Such a natural number N is then called the number of
elements in S, or the cardinality of S, which is denoted by #.S/, jSj, or S. NN
4.141 Example. For each natural number N 2 N the set N is finite and has N
elements, because the identity function IN W N ! N; K 7! K is a bijection.
4.142 Remark. Because every bijection has an inverse function, a set S is finite if
and only if there exist a natural number N 2 N and a bijection G W S ! N, for
instance, the inverse function G WD F ı1 for any bijection F W N ! S.
The following theorem shows that the insertion of a new element into a set
corresponds to the arithmetic addition of 1 to its cardinality.
P
4.143 Theorem. The equality #.A[fZg/ D Œ#.A/ C 1 holds for each finite set A,
and for each set Z … A.
Proof. For each finite set A there is a natural number N and a bijection F W N ! A.
Moreover, for each set Z there exists a bijection of singletons G W fNg ! fZg.
Consequently, because A \ fZg D ¿ by hypothesis, and because N \ fNg D ¿
P
by theorem 4.42, it follows that theorem 4.138 gives a bijection N C 1 ! A[fZg:
P
H D .F [G/ P
W N C 1 D .N [fNg/ P
! .A[fZg/:
t
u
The following theorem shows that the cardinality of the union of two disjoint
finite sets equals the arithmetic sum of their two cardinalities.
P D Œ#.A/CŒ#.B/ holds for all disjoint finite
4.144 Theorem. The equality #.A[B/
sets A and B.
www.pdfgrip.com
Proof. This proof proceeds by induction with the cardinality of the second set.
If #.B/ D 0, then B D ¿ by definition, whence for each finite set A,
P D #.A[¿/
#.A[B/ P D #.A/ D Œ#.A/ C 0 D Œ#.A/ C Œ#.B/:
Hence, assume that there exists a natural number N 2 N for which the theorem
P D Œ#.A/ C Œ#.B/ holds for all disjoint finite sets
holds, so that the equality #.A[B/
A and B with #.B/ D N. For each set C with N C 1 elements, there exists a bijection
F W N C 1 ! C. Consequently, the subset B WD F".N/ D F".f0; : : : ; N 1g/ has
N elements, because the restriction FjN W N ! B is a bijection. Hence, with the
P
element Z WD F.N/, it follows that C D B[fZg with Z … B, whence
P
#.A[C/
D P
because C D B[fZg;
P [fZg/
#.A[ŒB P
D associativity of [;
P [fZg
#Œ.A[B/ P
D theorem 4.143;
P C #.fZg/
#.A[B/
D induction hypothesis;
Œ#.A/ C #.B/ C #.fZg/
D associativity of C;
#.A/ C Œ#.B/ C #.fZg/
D theorem 4.143;
#.A/ C #.C/:
t
u
The following two theorems confirm that every subset of a finite set is also finite.
4.145 Theorem. For each N 2 N, every subset S N is also a finite set, with at
most N elements. Moreover, each proper subset S N has fewer than N elements.
Proof. This proof proceeds by induction with N.
If N WD 0, then N D ¿, and the only subset S N is S D ¿, which is finite.
As an induction hypothesis, assume that there exists a natural number K 2 N
such that the theorem holds for N WD K, so that each subset S K is finite with
at most K elements, and that each proper subset S K has fewer than K elements.
Hence, consider a subset R K C 1. Two cases arise: either K … R or K 2 R.
If K … R, then R Œ.K C 1/ n fKg D K whence R is finite with at most K
elements by induction hypothesis.
If K 2 R, then the set C WD R n fKg is a subset of Œ.K C 1/ n fKg D K, whence
C is finite with at most K elements by induction hypothesis. Thus, there exists a
www.pdfgrip.com
B \ .A n B/ D ¿;
B [ .A n B/ D A;
Œ#.B/ C Œ#.A n B/ D #.A/;
4.149 Theorem. For all sets A and B with the same finite cardinality, and for each
function F W A ! B with domain D.F/ D A, the following conditions are mutually
equivalent:
.P/ F is injective,
.Q/ F is surjective,
.R/ F is bijective.
Proof. If F is bijective, then F is also injective and surjective, because .R/ , Œ.P/^
.Q/ by definition of bijectivity (definition 3.118); therefore both .R/ ) .P/ and
.R/ ) .Q/ hold, by theorems 1.53 and 1.54.
For the reverse implications, because A and B have the same finite cardinality,
there exist a natural number N 2 N and a bijection G W N ! A.
If F is injective, then F".A/ B is a finite set with cardinality L N, so that
there exists a bijection H W F".A/ ! L. Then the composition
G F H
H ı F ı G W N ! A ! F".A/ ! L
J F I
I ı F ı J W .N 1/ ! S ! F".S/ ! N
cannot be a bijection, because by theorem 4.148 its inverse could not be an injection
from N to N 1. Hence F cannot be surjective, which proves .Q/ ) .P/. t
u
The following theorem shows that the number of elements in a Cartesian product
equals the product of the numbers of elements in its factors.
4.150 Theorem. For all K; L 2 N, the Cartesian product K L has K L elements.
Proof. By induction with L, if L D 0, then L D ¿, whence for every K 2 N
K 0 D K ¿ D ¿;
#.K 0/ D #.¿/ D 0 D K 0:
#.K M/ D K M
www.pdfgrip.com
holds for every K 2 N. Hence, from the disjoint union M C 1 D M [fMg,P and from
the distributivity of Cartesian products over unions (theorem 3.63), it follows that
K P
.M C 1/ D K .M [fMg/
P
D .K M/[.K fMg/;
#ŒK P
.M C 1/ D #Œ.K M/[.K fMg/
D Œ#.K M/ C Œ#.K fMg/
D .K M/ C K
D K .M C 1/;
4.117 . Prove or disprove that all ordered pairs have the same cardinality.
4.118 . For all finite sets A and B, prove that .A B/ Ð .B A/.
4.119 . For all finite sets A, B, C, prove that Œ.A B/ C Ð ŒA .B C/.
4.120 . For all finite sets A and B, prove Œ#.AB/ D Œ#.A [ B/ Œ#.A \ B/.
Some sets are not finite, for they do not admit any bijection onto any natural number.
4.151 Definition (infinite sets). A set Z is infinite if and only if Z is not finite,
which means that there are no bijections from Z onto any natural number.
For instance, the set N is infinite.
www.pdfgrip.com
1 if K D N;
ŒJ.N/.K/ WD
0 if K 6D N:
In other words, the function J.N/ is the characteristic function fNg of the singleton
fNg (from example 3.89). Consequently, J is injective; indeed if M 6D N, then
ŒJ.M/.M/ D 1 but ŒJ.N/.M/ D 0, whence J.M/ 6D J.N/.
To prove the absence of any surjection X 2X , for each function J W X ! 2X ,
this proof demonstrates a method known as Cantor’s diagonalization to show that
J is not surjective. Each such function J W X ! 2X maps each element N 2 X to a
function J.N/ 2 2X , so that J.N/ W X ! 2. In particular, J.N/ maps N to an element
ŒJ.N/.N/ 2 2 D f0; 1g. Thus, define a function F W X ! 2 by
www.pdfgrip.com
0 if ŒJ.N/.N/ D 1;
F.N/ WD
1 if ŒJ.N/.N/ D 0:
Thus, G 2 R but G.N/ 6D ŒJ.N/.N/ for every N 2 N, whence G 6D ŒJ.N/ for every
N 2 N. Consequently G … J".N/, and therefore J is not surjective. Hence there does
not exist any bijection J W N ! R.
Example 4.157 reveals that the set R of all real numbers is “more infinite” than
the set N of all natural numbers. Moreover, applying theorem 4.156 to X WD R
R
shows that 2R is also “more infinite” than R. Then 2.2 / is also “more infinite” than
2R . And so forth, thus there exists an “infinite” variety of “infinite” sets.
There exist several infinite sets that have the same cardinality as N has. For instance,
using only addition and multiplication from integer arithmetic, this subsection
presents a proof that the set of all nonnegative integers N and the set of all integers
Z have the same cardinality; similarly, N and the Cartesian product N N have the
same cardinality. The following terminology conforms to [8, p. 152], [30, p. 47], and
[128, p. 151].
4.158 Definition. A set is denumerable — or has cardinality @0 (read “aleph
zero”) — if and only if it has the same cardinality as the set N of all natural numbers.
A set is countable if and only if it is either finite or denumerable.
The following theorem shows that the set Z of all the integers has the same
cardinality as the set N of all the nonnegative integers.
www.pdfgrip.com
T.0/ WD 0;
T.N C 1/ WD T.N/ C .N C 1/:
T.0/ WD 0;
T.1/ WD T.0/ C .0 C 1/ D 0 C .0 C 1/ D 1;
T.2/ WD T.1/ C .1 C 1/ D 1 C .1 C 1/ D 3;
T.3/ WD T.2/ C .2 C 1/ D 3 C .2 C 1/ D 6;
::
:
The values of T are called triangular numbers because they correspond to the
number of elements in the following patterns:
www.pdfgrip.com
The following theorem provides a different formula to compute the same function
T.
4.162 Theorem (arithmetic series). For each natural number N 2 N,
N .N C 1/
0C1C2C C .N 1/ C N D T.N/ D :
2
T.N/ D 0 C 1 C C .N 1/ C N;
T.N/ D N C .N 1/ C C 1 C 0;
T.N/ C T.N/ D .N C 1/ C .N C 1/ C C .N C 1/ C .N C 1/;
whence 2 T.N/ D N .N C 1/. However, a proof along this outline also requires
induction to rearrange the terms of the sum with associativity and commutativity.
The following definition provides a formula for Cantor’s diagonal enumeration
of N N, and the subsequent theorems will verify that indeed it enumerates N N.
4.163 Definition. Define a function T W .N N/ ! N by
T .M; N/ WD M C T.M C N/
.M C N/ .M C N C 1/
D MC :
2
www.pdfgrip.com
The value T .M; N/ corresponds to the sum of the number of elements in the
triangular pattern counted by the “triangular number” T.M C N/ and a last partial
row with M elements (instead of a complete last row of M C N elements for the next
triangular number). For example, with M WD 1 and N WD 2,
T .1; 2/
T.M C N/
M
.M C N/ .M C N C 1/
IDMC :
2
Proof. This proof proceeds by induction with I.
First, if I WD 0, then I D 0 C T.0 C 0/ with M WD 0 and N WD 0.
Second, assume that there exists K 2 N such that the theorem holds for I WD K,
so that there exist M; N 2 N with K D M C T.M C N/.
If N > 0, then K C 1 D 1 C ŒM C T.M C N/ D .M C 1/ C T.ŒM C 1 C ŒN 1/.
If N D 0, then K C 1 D 1 C ŒM C T.M C 0/ D .M C 1/ C T.M C 0/ D
0 C T.0 C ŒM C 1/. t
u
The following theorem shows that the function T W .N N/ ! N is injective.
4.165 Theorem. For all K; L; M; N 2 N, if K C T.K C L/ D M C T.M C N/, then
both K D M and L D N.
Proof. If
If K > M, then K M > 0, so that the left-hand side is positive, but then the right-
hand side must also be positive: T.M C N/ T.K C L/ > 0. Hence M C N > K C L,
but then
by definition of T. Thus
K M D T.M C N/ T.K C L/ M C N
M C N > K C L Œ.2 M/ C N C L
M L;
The following theorem guarantees the existence of a bijection between two sets,
provided that there exist injections from one set to the other and vice versa. Accord-
ing to Suppes [128, p. 95], Cantor conjectured the theorem and then Bernstein and
Schröder proved it independently of each other in the 1890s. Fraenkel [37, p. 102]
credits the following proof to J. M. Whitaker.
4.167 Theorem (Bernstein–Cantor–Schröder). For all sets A and B, if there are
injections F W A ! B and G W B ! A, then there is a bijection H W A ! B.
Proof. The strategy of this proof consists in producing a subset E A such that
G"ŒB n F".E/ D .A n E/
Because these inclusions hold for every element V 2 D, it follows that they also
hold for their union:
[ n h [ io
D A n G" B n F" D :
Let
h [ i
E WD A n G" B n F" D :
S
From D E it follows that
h [ i
A n G" B n F" D fA n G"ŒB n F".E/g;
so that
E fA n G"ŒB n F".E/g
S S
whence E 2 D. Consequently, E D, whence E D D, but then the definition
of E gives
t
u
The next theorem shows a use of the Bernstein–Cantor–Schröder Theorem.
4.168 Theorem. There exists a bijection H W Z ! Q.
Proof. First, there exists an injection F W Z ! Q with F.N/ WD N=1.
Second, there exists an injection I W Q ! .Z Z/ such that I.P=Q/ WD .P; Q/
with P 2 N minimum (theorem 4.99).
www.pdfgrip.com
This subsection shows that the set of all finite sequences of natural numbers is
denumerable.
Theorem 4.171 shows that allowing the number of variables to change from one
function to another still produces a set of functions.
www.pdfgrip.com
4.171 Theorem. For all sets E and C, all the functions F W EK ! C of any finite
number of variables from the domain E and with values in the co-domain C form a
set
n K o
C.E / W K 2 N :
Also let D WD FN;E and B WD C, the set C in the hypothesis of the theorem. Thus
C.E / is the set of all functions F W EK ! C defined on all of EK , with EK 2 D for
K
each K 2 N, and
n K o
FD;C WD C.E / W K 2 N
[
4.173 Theorem. The set NK of all finite sequences of natural numbers is
K2N
denumerable.
Proof. For each K 2 N , denote by TK W NK ! N any enumeration of all natural
sequences of length K, as in theorem 4.172. Again following [86, p. 799], this proof
[ all the bijections TK of N by means of K and T2 . Specifically,
K
pieces together
define S W N ! N by
K
K2N
The present text has defined a set to be finite if and only if there exists a bijection
onto a natural number, and infinite if and only if there does not exist any such
bijection. There exists a different definition of infinite sets, called “Dedekind-
infinite” [128, p. 107], corresponding to Dedekind’s definition [25, V, #64, p. 63].
4.174 Definition. A set Z is Dedekind-infinite if and only if there exists a proper
subset Y ¤ Z and a bijection F W Z ! Y, or, equivalently, F ı1 W Y ! Z.
Thus, a set is Dedekind-infinite if and only if it has the same cardinality as that
of one of its proper subsets.
4.175 Example. The set N is Dedekind-infinite because there is a proper subset
N N and a bijection defined by the successor function:
G W N ! N;
X 7! X [ fXg:
G F Gı1
H WD Gı1 ı F ı G W N ! Z ! Y ! N
would be a bijection from N onto a proper subset Gı1 ".Y/ ¤ N, which would
contradict theorem 4.148. t
u
The proof of the converse requires the Axiom of Choice (exercises 4.139, 4.140).
4.177 Theorem. Every infinite set contains a denumerable subset.
Proof. Apply recursion (theorem 4.21). If Z is infinite, then Z 6D ¿, whence there
exists some X 2 Z. Define F0 W f0g ! Z by 0 7! X. Assume that there exists
an injection FN W N ! Z. Then Z 6D FN ".N/ because Z is infinite, whence Z n
FN ".N/ 6D ¿ and there exists X 2S Z n FN ".N/. Let HN W fNg ! fXg and let
FNC1 WD FN [HP N . Finally, let F WD N2N FN . Then F W N ! Z is an injection. u t
Theorem 4.177 explains the subscript 0 in the notation @0 for the cardinality of N.
Because in the Zermelo-Frænkel set theory with the Axiom of Choice every infinite
set Z contains a denumerable subset, there exists an injection F W N ! Z, so that
the cardinality of N cannot exceed that of Z. If there is also an injection G W Z ! N,
then the Bernstein–Cantor–Schröder Theorem (theorem 4.167) guarantees that there
is also a bijection H W N ! Z; thus, the cardinality of Z cannot be strictly smaller
than that of N. Thus @0 represents the “smallest” infinite cardinality.
4.178 Theorem. Every infinite set is also Dedekind-infinite.
Proof. If W is infinite, then W contains a denumerable subset Z W, by
F W Z ! Y onto a proper
theorem 4.177. By example 4.175, there exists a bijection
subset Y ¤ Z. Extend F to all of W by setting H WD F [P IW jWnZ . t
u
4.179 Theorem. Every denumerable union of disjoint denumerable sets is denu-
merable.
Proof. If F is denumerable, then there is a bijection A W N ! F with I 7! A.I/. If
each A.I/ is denumerable, then by the Axiom of Choice there is a bijection
S FI W N !
A.I/ with J 7! FI .J/ 2 A.I/. Hence the function G W N N ! F defined by
G.I; J/ WD FI .J/ is a bijection. t
u
Bolzano [10] and Dedekind [25, V, Theorem 66, p. 64] argued for the “existence”
of an infinite set from practical considerations. Yet the “existence” of an infinite set
does not follow from the other axioms, but requires an axiom of infinity [8, p. 21].
www.pdfgrip.com
Zermelo [145] introduced such an axiom with a set such that if X is an element, then
fXg is also an element. The variant adopted here, with fXg replaced by X [ fXg, is
attributed to John von Neumann [135] and has proved more convenient [8, p. 22].
There also exist “infinitely” many other nonequivalent axioms of infinity [18, 57,
p. 342–346].
Axiom A2 For each natural number N there exists exactly one natural number
denoted by N 0 and called the successor of N.
Axiom A4 If K 0 D L0 , then K D L.
From Peano’s axioms, and two additional axioms for recursive definitions of
addition and multiplication [128, p. 136], the same proofs as in this chapter verify
all the algebraic properties of arithmetic and ordering with natural numbers [76,
Ch. 1, p. 1–18]. However, in situations that involve other topics, for examples,
rational numbers or cardinality of sets, the use of Peano’s axioms would require
some theoretical link between Peano’s natural numbers and other sets, in other
words, some means of including Peano’s arithmetic and applications within the
same framework. This theoretical link can be the development of arithmetic from
within set theory, as done here.
The progression from N to Z and then to Q allows for subtractions with integers
without requiring rational numbers, and then allows for divisions and subtractions
with rational numbers without requiring the development of “real” numbers.
www.pdfgrip.com
I J
N ,! S ,! 2N
I J
X ,! S ,! 2X
P is denumerable.
4.123 . With A denumerable, B finite, disjoint, prove that A[B
4.124 . Prove that if A is denumerable and B finite, then A [ B is denumerable.
4.125 . Prove that if A and B are denumerable, then A [ B is denumerable.
4.126 . Prove that there exists a bijection from 2N to P.N/.
4.127 . Prove that P.N/ is not countable.
4.128 . Prove: if A is denumerable and B is finite, then A B is countable.
4.129 . For each nonempty set X prove that there is no injection 2X ,! X.
4.130 . Prove that if Œ#.X/ D Œ#.Y/, then Œ#.2X / D Œ#.2Y /.
4.131 . Prove: if Œ#.X/ < Œ#.Y/ are both finite, then Œ#.2X / < Œ#.2Y /.
4.132 . Prove that if X is a finite set, then 2X is also finite.
4.133 . Prove that if X is a finite set, then Œ#.X/ < Œ#.2X /.
4.134 . Prove that every infinite subset S N is denumerable.
4.135 . For each denumerable set A prove that every subset S A is countable.
4.136 . Prove: if A is uncountable and A B, then B is uncountable.
4.137 . Prove that if A, B, and C are denumerable, then so is .A B/ C.
4.138 . Prove that if A B is denumerable, then A and B are countable.
4.139 . Show where the proof of theorem 4.179 invokes the Axiom of Choice.
4.140 . Show where the proof of theorem 4.177 invokes the Axiom of Choice.
www.pdfgrip.com
Chapter 5
Well-Formed Sets: Proofs by Transfinite
Induction with Already Well-Ordered Sets
5.1 Introduction
This chapter focuses on “well-formed” sets, which are defined by means restricted
to the axioms of Zermelo-Fraenkel set theory from chapters 1, 2, 3, and 4. The
main result states that no two well-formed sets are members of each other, and
consequently that every well-formed set is not an element of itself.
This chapter shows the dependence of one axiom on the others for sets — called
well-formed sets — defined in specific ways solely through the axioms of set theory.
Such well-formed sets suffice for most of logic, mathematics, computer science, and
their applications to the sciences and engineering. Among other features, the result
shows that no well-formed set is an element of itself, and that this result is provable
from the other axioms of set theory [74]. This chapter also provides a way to revisit
induction (chapter 4) on a different level.
268 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
WC WD fB 2 W W .B C/ ^ .B 6D C/g;
Proof. If V D W, then the formula is a tautology: Œ.P/ ^ .Q/ ) .P/. For the
converse, let U WD W n V. If U 6D ¿, then U has a first element A 2 U. Thus, if
B A but B 6D A, then B … U, whence B 2 W n U D V. Hence the initial interval
WA V, but then A 2 V by hypothesis on V, which contradicts A 2 U D W nV. u t
Well-ordered sets also lend themselves to a method of definition known as
transfinite construction, which relies on the concept of “ideal” in a well-ordered
set.
5.4 Definition (ideal). For each set W well-ordered by a relation a subset V
W is an ideal of W if and only if V contains every element preceding any of its
elements. Thus V is an ideal if and only if WC V for every C 2 V:
The following theorems show that the set of all ideals is well-ordered by
inclusion.
5.8 Theorem. For each set W well-ordered
T by a relation and for each nonempty
set F of ideals of W, the intersection F is also an ideal of W.
T
Proof. If C 2 F , then C 2 V for each ideal V 2 F . Hence, if B 2 W and B C,
B 2 V because
then T T V is an ideal. This conclusion holds for each V 2 F , whence
B 2 F . Thus F is an ideal of W. t
u
5.9 Theorem. For each set W well-ordered by a relation
T and for each nonempty
set F of ideals of W, the smallest element of F is F . Therefore the set I .W/
is well-ordered by inclusion ./.
Proof. For each nonempty setT F of ideals of W, there exists at least one ideal
U 2 F , and the intersection F is also an ideal, by theorem 5.8. Let
n [ \ o
Z WD C 2 W W C 2 F n F :
S T
If Z D ¿, then F D F , whence F contains only one ideal, in effect U D
T
F. T
If Z 6D ¿ then it has a smallestTelement A 2 Z W. BecauseT F is an ideal
by theorem 5.8, and because
T A … F , it follows that WA D F by theorem 5.7.
Also because A … F , there exists an ideal V 2 F with A … V. Hence V WA
by theorem 5.6. T T T
From V TWA and WA D F it follows that V F , but F V.
Consequently F D V 2 F .
T Therefore every nonempty set F I .W/ has a smallest element, in effect
F , so that I .W/ is well-ordered by set inclusions. t
u
270 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
denote the set of all functions with domain equal to an initial interval WB and with
range in E (definition 4.153). For each Z 2 E, and for each function P W Y ! E,
there exists exactly one function F W W ! E such that F.A/ D Z and F.B/ D
P .FjWB /.
Proof. This proof establishes the uniqueness and existence separately.
Uniqueness
S WD fC 2 W W F.C/ 6D G.C/g:
Existence
Let F denote the set of all ideals V W for which there exists a function
FV W V ! E such that FV .A/ D Z and FV .B/ D P .FV jWB /.
For each ideal U 2 F , applying the uniqueness just proved to the well-ordered
set U \ V instead of W shows that the functions FU and FV coincide on the well-
ordered subset U \ V in W. S
Hence, define a function FF W F ! E by settingS FF .B/ WD FU .B/ for
any ideal U 2 F with B 2 U. In other terms, FF D U2F FU . The preceding
argument confirms that this definition does not depend on which ideal U contains
B, because if B 2 U and B 2 V, then FU .B/ D FV .B/.
Next, if an ideal U is an initial interval, U D WB for some B 2 W, and if WB 2 F ,
then WB [ fBg 2 F . Indeed, a function FU W U ! E extends to WB [ fBg by the
definition FWB [fBg .B/ WD P.FU /.
Suppose that W … F . Then let G denote the set of Tall the ideals of W that are not
elements of F . In particular, W 2 G . Define V WD G , which is then the smallest
ideal of W in G . If V had a last element D, then V D W TD [ fDg by definition of
WD and of an ideal; however, WD 2 F , otherwise V 6D G , but from WD 2 F it
follows that WD [ fDg 2 F . Thus V cannot have a last element.
www.pdfgrip.com
S
If V does not have a last element, then V D B2V WB by definition
S of an ideal.
S
Again, it follows that WB 2 F for each B 2 V, whence V D B2V WB F
and then V 2 F because of the existence of FF , contradicting the definition of V.
Therefore, W 2 F , which means that F extends to all of W. t
u
5.1 . Prove that in each well-ordered set each initial interval is an ideal.
5.2 . Provide an example of an ideal that is not an initial interval in a well-
ordered set.
5.3 . Prove that if Z is well-ordered by 4, then 4 differs from .
5.4 . Prove that if Q is well-ordered by 4, then 4 differs from .
5.5 . Provide an example of a well-order on a set of modular integers ZM D
fŒ0M ; : : : ; ŒM 1M g and modular integers ŒIM , ŒKM , ŒLM , such that ŒKM ŒLM
but ŒIM C ŒKM 6 ŒIM C ŒLM .
5.6 . Provide an example of a well-order on a set of modular integers ZM D
fŒ0M ; : : : ; ŒM 1M g and modular integers ŒIM , ŒKM , ŒLM , such that Œ0M ŒIM
and ŒKM ŒLM but ŒIM ŒKM 6 ŒIM ŒLM .
5.7 . Prove that every subset of a well-ordered set is also well-ordered.
S
5.8 . Determine whether for each set F of ideals in W the union F is also an
ideal in W.
S
5.9 . Determine whether for each set G of initial intervals in W the union G is
also an initial interval in W.
5.10 . Determine whether for each set G of initial intervals in W the intersection
T
G is also an initial interval in W.
Sets defined exclusively through the axioms of set theory adopted here are called
well-formed sets. They have the advantage of avoiding certain contradictions that
would arise from defining sets by means not so strict. The definition of well-formed
sets involves the concept of sets that are “transitive” relative to the relation 2.
www.pdfgrip.com
272 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
5.11 Definition (Transitive Sets). A set A is transitive if and only if every element
of A is also a subset of A, so that 8XŒ.X 2 A/ ) .X A/, or, equivalently,
¿;
f¿g;
˚
n¿; f¿g ;˚ o
¿; f¿g; ¿; f¿g ;
n ˚ ˚ o
¿; f¿g; f¿g ; ¿; f¿g :
˚
5.13 Counterexample. The set A WD f¿g is not transitive, because it contains
an element X WD f¿g that is not a subset of A: X 6 A, because ¿ 2 X but ¿ … A.
Power sets, unions, and intersections of transitive sets are also transitive.
5.14 Theorem. If a set A is transitive, then P.A/ is also transitive.
Proof. If S 2 P.A/, then S A. Thus if X 2 S, then X 2 A, and X A by
transitivity of A. Hence X 2 P.A/ for each X 2 S, whence S P.A/. t
u
S
5.15 Theorem. If a set F is transitive, then F is also transitive.
S
Proof. If S 2 F , then there exists A 2 F with S 2 A. YetS A F by transitivity
of F . From S 2 A and A F follows S 2 F , whence S F . t
u
S T
5.16 Theorem. If F is a nonempty set of transitive sets, then F and F are
also transitive.
S
Proof. If S 2 F , then S S 2 A for some
S A 2 F , whence S A by transitivity of
A, and hence
T S A F , so that F is transitive.
If S 2 F T , then S 2 A for
T every A 2 F , whence S A by transitivity of A,
and hence S F , so that F is transitive. t
u
5.3.2 Ordinals
Well-formed sets will rely on the concept of ordinals (also called “ordinal numbers”
[30, p. 42]). The following definition conforms to Kunen’s [74, p. 16].
5.17 Definition (ordinals). A set A is an ordinal if and only if it is a transitive set,
and the relation 2 is an irreflexive well-ordering on the set A.
www.pdfgrip.com
¿;
f¿g;
˚
n¿; f¿g ;˚ o
¿; f¿g; ¿; f¿g :
X WD ¿;
˚
Y WD f¿g ;
then X 2 A and Y 2 A, so that fX; Yg A, but fX; Yg does not have any smallest
element relative to the relation 2, because X … Y and Y … X.
In particular the subset fX; Yg is not an ordinal.
5.20 Theorem. The empty set is an element of every nonempty ordinal.
Proof. By definition, every ordinal A is transitive, so that if X 2 A then X A.
Consequently, if Y 2 X and X 2 A, then Y 2 A. Therefore, if X 2 A and X 6D ¿,
then X is not the smallest element of A. Yet every nonempty ordinal A has a smallest
element. Hence contraposition shows that the smallest element must be ¿. t
u
5.21 Theorem. If A is an ordinal, then A … A. Moreover, A … X for each X 2 A. In
particular, if A and X are ordinals, then A … X or X … A.
Proof. If A is an ordinal, then 2 is connected (definitions 3.185, 3.194) and
irreflexive (definition 5.17): exactly one of X 2 Y, X D Y, or Y 2 X holds for
all X; Y 2 A. Because A D A, it follows that A … A. Moreover, if A is an ordinal
and X 2 A, then X A, whence if Y 2 X then Y 2 A. With Y WD A, it follows by
contraposition that A … X. t
u
5.22 Theorem. Every element of an ordinal is an ordinal.
Proof. If A is an ordinal and X 2 A, then X A because A is a transitive set. Hence
2 well-orders X, because 2 well-orders A. If moreover Z 2 Y and Y 2 X, then
Z 2 X because A is a transitive set, whence Y X. Thus X is also a transitive set.
Furthermore, the relation 2 remains irreflexive on the subset X A. u
t
5.23 Theorem. If A isSan ordinal, then either A contains a last element D and
A D D [ fDg, or A D A is the union of all its elements.
www.pdfgrip.com
274 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
S
an ordinal and B 2 A, then B A; consequently, A A.
Proof. If A is S
Let D WD A. If D 6D A, then there exists X 2 A n D. Then X D because
X 2 A;
S hence X is an ordinal by theorem 5.22, whence X 6D A by theorem 5.21, and
X A D D.
Conversely, still with X 2 A nS D, for each B 2 A it follows
S thatSX … B, whence
B 2 X, and hence B X, so that B2A B X. Thus, D D A D B2A B X.
Therefore, if D 6D A, then D X D, whence X D D is the only element of
A n D, and hence A D D [ fDg. t
u
5.24 Theorem. If B is an ordinal, then B [ fBg is also an ordinal.
Proof. First, B [ fBgis transitive. Indeed, if X 2 B [ fBg, then either X 2 B, whence
X B B [ fBg, or X 2 fBg, whence X D B B [ fBg.
Second, 2 well-orders B [ fBg. Indeed, for each nonempty subset S B [ fBg,
two situations can occur: S B or S \ fBg 6D ¿. If S B then S has a smallest
element, because B is an ordinal. If S \ fBg 6D ¿, then either S D fBg has the
smallest element B, or S \ B 6D ¿ and then S \ B is a subset of B and hence has a
smallest element, which is then also a smallest element of S D .S \ B/ [ fBg.
Moreover, the relation 2 remains irreflexive and transitive on B [ fBg. Indeed,
B … B by theorem 5.21. Furthermore, if X 2 Y and Y 2 Z in B [ fBg, then Y 2 B
from either Z 2 B or Z 2 fBg. Also, Y 6D B by theorem 5.21, which forbids B 2 Z
and Z 2 B [ fBg, and hence X 6D B also by theorem 5.21, which forbids B 2 Y
Y 2 Z, and Z 2 B [ fBg. Consequently only two cases can occur: Z 2 B or Z D B.
If Z D B, then X 6D B and Y 6D B, whence X 2 Z.
If Z 2 B, then X, Y, and Z all three lie in B, whence X 2 Z because 2 is transitive
on the well-ordered set B.
Finally, 2 is strict on B [ fBg. Indeed, because 2 is strict on B, it follows that if
X 2 B [ fBg and Y 2 B [ fBg, then two different cases can arise.
If X 2 B and Y 2 B, then X 2 Y and Y 2 X cannot both hold, for transitivity
would yield X 2 X which cannot hold by strictness of 2 on B.
If X 2 B and Y 2 fBg, then X 2 Y and Y 2 X cannot both hold. Otherwise Y D B
and then X 2 B and B 2 X. However, X is also an ordinal by theorem 5.22, whence
2 is also transitive on X, so that B 2 X and X 2 B yield B 2 B, which cannot hold
by strictness of 2 on X. t
u
The following theorems show that every set of ordinals is well-ordered. First is
strongly connected (definition 3.184) on every set of ordinals.
5.25 Theorem. For all ordinals A and B, either A B, or A D B, or B A.
Proof. If B 6 A, then there exists X 2 B n A, and hence there exists a smallest such
element: X 2 B n A, so that X 2 Y for every Y 2 B n A with Y 6D X. Also, X 6D ¿,
www.pdfgrip.com
E WD fB 2 G W B 2 Cg D C \ G :
The union and the intersection of every nonempty set of ordinals is an ordinal.
S
5.28 Theorem. For each set F of ordinals, F is also an ordinal.
S
Proof. The union F is transitive, by theorem 5.15.
www.pdfgrip.com
276 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
S S
The union F is well-ordered. Indeed, for each nonempty subset S F,
let
E WD fA 2 F W A \ S 6D ¿g:
then either W D ¿ or V 6D ¿.
www.pdfgrip.com
The following definition establishes sets that contain all the well-formed sets.
5.30 Definition. For each ordinal A, define a set R.A/ by transfinite construction:
• R.¿/ WD ¿;
• R.A [ fAg/
S WD PŒR.A/;
• R.A/ WD B2A R.B/ if there does not exist any ordinal B such that A D B [ fBg,
but if R.B/ has been defined for every B 2 A.
5.31 Definition. A set X is well-formed if and only if there exists an ordinal A such
that X 2 R.A/.
5.32 Remark. The transfinite construction proceeds as follows. For each ordinal W,
let
E WD PŒP.W/;
S
Y WD A2W EWA ;
PW Y ! E;
S
P.RjWA / WD B2A R.B/:
278 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
5.33 Example. The first few sets of the form R.A/ are also ordinals:
R.¿/ D ¿;
R.f¿g/ D R.¿ [ f¿g/ D PŒR.¿/ D PŒ¿ D f¿g;
˚ ˚ ˚
R ¿; f¿g D R f¿g [ f¿g D PŒR.f¿g/ D PŒf¿g D ¿; f¿g :
which is of the form R.A/, but it is not well-ordered by 2 and hence not an ordinal,
by counterexample 5.19. The set R.A/ is also different from the ordinal
˚ n ˚ o
¿; f¿g; ¿; f¿g ; ¿; f¿g; ¿; f¿g :
WC D fB 2 W W .B 2 C/ ^ .B 6D C/g
5.4.2 Regularity
The following theorems confirm that every element, subset, pairing, power set,
union, intersection, and Cartesian product of well-formed sets is again a well-
formed set.
5.35 Theorem. For each well-formed set X there is a smallest ordinal A with
X 2 R.A/.
Proof. If X is a well-formed set, then there exists an ordinal C such that X 2 R.C/ 2
PŒR.C/ D R.C [ fCg/. Hence X 2 R.C [ fCg/ by transitivity (theorem 5.34). By
theorem 5.27, there exists a smallest ordinal A 2 C [ fCg such that X 2 R.A/.
For every ordinal D, either D D A, or D 2 A 2 C [ fCg and then X … R.D/, or
A 2 D and then D is not the smallest such ordinal. u
t
5.36 Theorem. If X is well-formed, then every Y 2 X is well-formed.
Proof. If X is well-formed, then there exists an ordinal A such that X 2 R.A/.
If A D ¿, then X 2 R.A/ D R.¿/ D ¿, whence Y 2 X is vacuously well-
formed.
If there exists an ordinal B such that A D B [ fBg, then X 2 R.A/ D PŒR.B/,
whence X PŒR.B/ by transitivity of PŒR.B/, Consequently,
S Y 2 R.A/ D
PŒR.B/ is well-formed for each Y 2 X. If R.A/ D B2A R.B/ and the theorem
holds for each Z 2 R.B/ for each B 2 A, then for each X 2 R.A/ there exists B 2 A
such that X 2 R.B/ whence every Y 2 X is also well-formed. t
u
S
5.37 Theorem. If X and Y are T well-formed sets, then so are fX; Yg, P.X/, X,
X Y, every subset of X, and X provided X 6D ¿.
Proof. If X and Y are well-formed sets, then there exist ordinals A and B such that
X 2 R.A/ and Y 2 R.B/. Either A D B (whence R.A/ D R.B/), or A 2 B (whence
R.A/ R.B/), or B 2 A (whence R.B/ R.A/). For instance, assume that R.A/
R.B/. Thus X 2 R.B/ and Y 2 R.B/, so that fX; Yg 2 PŒR.B/ D R.B [ fBg/,
whence fX; Yg is well-formed, because B [ fBg is an ordinal.
Similarly, if X 2 R.B/ is a well-formed set, then X R.B/ by transitivity, whence
P.X/ PŒR.B/ D R.B [ fBg/ and hence P.X/ 2 PŒR.B [ fBg/ D R.A [ fAg/
with A WD B [ fBg. Thus, P.X/ is well-formed.
Consequently, from P.X/ 2 R.A [ fAg/ it follows that P.X/ PŒR.A [ fAg/,
whence if S X,Tthen S 2 PŒR.A [ fAg/ is T well-formed.
S S
In particular, X is well-formed because X X and X is well-formed.
Also, if X 2 R.B/ is a well-formed set, then X R.B/. Hence, S if Z 2 Y and
Y 2 X, then Y S X whence Z 2 X and Z 2 R.B/. This shows that X R.B/.
Consequently, X 2 PŒR.B/ D R.B [ fBg/ is well-formed. S
In particular, because fX; Yg is well-formed, it follows that X [ Y D fX; Yg is
well-formed, whence P.X [ Y/, PŒP.X [ Y/, and PfPŒP.X [ Y/g are also
well-formed. Therefore, X Y 2 PfPŒP.X [ Y/g is also well-formed. t
u
www.pdfgrip.com
280 5 Well-Formed Sets: Proofs by Transfinite Induction with Already Well-Ordered Sets
attributed independently to Zermelo and von Neumann [128, p. 53]. The axiom of
regularity has the disadvantage of asserting a condition about sets already defined by
previous axioms. Yet within the theory of well-formed sets, exercise 5.31 confirms
that the axiom of regularity is not independent but is a theorem derivable from
the other axioms. Because well-formed sets suffice for most of logic, mathematics,
computer science, and their applications, the foundations of these fields can restrict
themselves to well-formed sets [74]. In contrast to the derivability of the axiom
of regularity from the other axioms of the theory of well-formed sets, neither the
generalized continuum hypothesis nor its negation are derivable from the other
axioms of the theory of well-formed sets [19, 20, 44, 45]. Thus, the “axiom of
regularity” is an example of an axiom that is “dependent” on the other axioms,
whereas the generalized continuum hypothesis is an example of an axiom that is
“independent” from the other axioms.
www.pdfgrip.com
5.27 . Prove that the set N of all natural numbers is a well-formed set.
5.28 . Prove that every natural number N 2 N is a well-formed set.
5.29 . Prove that the set Z of all integers is a well-formed set.
5.30 . Prove that the set Q of all rational numbers is a well-formed set.
5.31 . Prove that for each well-formed set X there exists Y 2 X such that Y \X D ¿.
5.32 . Prove that if X is a well-formed set, then so is fXg.
5.33 . For each well-formed set X, prove that if A is the smallest ordinal such that
X 2 R.A/, then there exists an ordinal B such that A D B [ fBg.
5.34 . Prove that every finite set of well-formed sets is a well-formed set.
5.35 . Determine whether every countable set of well-formed sets is a well-formed
set.
5.36 . Determine whether every set of well-formed sets is a well-formed set.
www.pdfgrip.com
Chapter 6
The Axiom of Choice: Proofs
by Transfinite Induction
6.1 Introduction
This chapter presents several statements, which are called “principles” because
they are well-formed formulae but not propositions, in the sense that neither of
them nor their negations are theorems, in the Zermelo-Frænkel set theory. The
first sections show how Zorn’s Maximal-Element Principle implies Zermelo’s Well-
Ordering Principle, which in turn implies the Choice Principle. Thus any extension
of the Zermelo-Frænkel set theory that includes Zorn’s Maximal-Element Principle
as an axiom also includes the other two principles as theorems. From the Choice
Principle, subsequent sections demonstrate the converse implications, known as
Zorn’s Lemma and Zermelo’s Theorem, so that all three principles are logically
equivalent within the Zermelo-Frænkel set theory. Hence all three principles are
theorems in the Zermelo-Frænkel-Choice set theory, which includes the Choice
Principle as the Axiom of Choice. The material also introduces yet other principles
that are logically equivalent to the Axiom of Choice, for example, the principle of
the distributivity of intersections over unions of families of sets. Any theory that
requires any such equivalent principle thus also requires the Axiom of Choice.
Other consequences of the Axiom of Choice include the existence of extrema for
continuous functions on closed and bounded sets in Euclidean spaces.
This section presents several mutually equivalent forms of the Choice Principle.
One version of the Choice Principle relies on the concept of a “choice function”:
6.1 Definition (choice function).S For each set F of nonempty sets, a choice
function is a function C W F ! F such that C.A/ 2 A for every set A 2 F .
S S
S If F D ¿, then F D ¿ D ¿: there exists exactly one function
6.2 Example.
C W F ! F , namely ¿ W ¿ ! ¿, which is “vacuously” a choice function.
S S
6.3 Example. With 1 D f¿g, if F DSf1g, then F D f1g D 1 D f¿g: there
exists exactly one function C W F ! F , namely C W f1g ! f¿g with C.1/ D ¿,
which is a choice function because C.1/ 2 1.
6.4 Example. For eachSnonempty set A, if F D fAg, then there exists X 2 A,
whence .A; X/ 2 F F D fAg A, and C WD f.A; X/g is a choice function.
S
6.5 Example. With 1 D f¿g and 2 D f¿; 1g, if F D f1; 2g, then
S F D
f1; 2g D 1[2 D f¿; 1g D 2. There are two choice functions C W f1; 2g ! f¿; 1g:
The requirement that C.1/ 2 1 D f¿g imposes that C.1/ D ¿.
The requirement that C.2/ 2 2 D f¿; 1g allows for the two possibilities:
either C.2/ D ¿ 2 f¿; 1g, or C.2/ D 1 2 f¿;S1g.
Thus the two
S choice functions are F W F ! F with F.1/ D 0 and F.2/ D 0,
or G W F ! F with G.1/ D 0 and G.2/ D 1.
More generally, each finite set of nonempty sets has a choice function.
6.6 Theorem. Within the Zermelo-Frænkel set theory, each finite set of nonempty
sets has a choice function.
Proof. This proof proceeds by induction on the number N 2 N of sets.
For N D 0, example 6.2 shows that the empty set has a choice function.
Example 6.4 proves the theorem for N D 1.
As an induction hypothesis, assume that the theorem holds for some N D M 2 N
and every finite set F with exactly M elements, all nonempty. For each set G with
exactly M C 1 elements, all nonempty, there exists B 2 G , so that F WD G n fBg has
exactly M elements, all
S nonempty. By the induction hypothesis, there exists a choice
function F SW F ! F . Example 6.4 shows that there exists a choice function
G W fBgS! fBg D B.SBecause F \ fBg D ¿, the union C WD F [ G W F [ fBg D
G ! . F / [ B D G is a function. Also, C.A/ D F.A/ 2 A for each A 2 F
and C.B/ D G.B/ 2 B for each B 2 fBg, whence C is a choice function for G . u t
Still, theorem 6.6 applies only to finite sets.
6.7 Example. No choice functions are known for G D P ŒP.N/ n f¿g, which
is the set of all nonempty sets of subsets of the set N of natural numbers, which
corresponds to the set of all nonempty sets of real numbers between 0 and 1.
The statement of the existence of choice functions is one version of the Choice
Principle, called the Choice-Function Principle to distinguish it from other versions.
www.pdfgrip.com
Neither the Choice-Function Principle 6.8 nor its negation are propositions in the
Zermelo-Frænkel set theory: there it is merely formula (6.1).
The concept of a “family” of sets provides a convenient way to specify more than
one operation on a set, for instance, to choose more than one element from a set.
6.9 Definition (family of sets). A family of sets F D fAi W i 2 I g is a set (of sets)
with a function I W I ! F , i 7! Ai , from a set I , called the indexing set or set of
indices, to F .
6.10 Example (Self-Indexed Family of Sets). Each set (of sets) F is a family of
sets: the same set I WD F serves as an indexing set, and the identity function
W F ! F , E 7! AE WD E shows that F D fE W E 2 F g D fAE W E 2 I g.
Still other versions of the Choice Principle rely on families of sets, as in
definitions 6.11 and 6.12.
Another version of the Choice Principle relies on choice functions only for
pairwise disjoint sets, as specified by definition 6.14.
6.14 Definition (Choice-Function Principle for Pairwise Disjoint Sets). Each
set of pairwise disjoint nonempty sets has a choice function.
Yet another version of the Choice Principle relies on functions and relations, as
specified by definition 6.15.
6.15 Definition (Choice-Relation Principle). For each relation R there exists a
function F R such that F and R have the same domain [128, p. 243, AC3 ].
In the Zermelo-Frænkel set theory, a relation is a subset of a Cartesian product:
R A B. There is a different principle, called the Relational Axiom of Choice,
where R may be a relation defined for all sets [88, p. 22].
This subsection provides different statements of the Choice Principle that are
logically mutually equivalent within the Zermelo-Frænkel set theory. One version
of the Choice Principle relies on “choice sets” rather than choice functions.
Definition (choice set). For each set F of nonempty sets, a choice set is a set
6.16 S
S F such that for each A 2 F , S \ A contains exactly one element.
As with the Choice-Function Principle 6.8, neither the Choice-Set Principle nor
its negation are propositions in the Zermelo-Frænkel set theory: it is merely a
formula, stated in words in definition 6.17.
6.17 Definition (Choice-Set Principle). Each set of nonempty sets has a
choice set.
S S
Example. If F D ¿, then F D ¿ D ¿: there exists exactly one subset
6.18 S
S F D ¿, namely S D ¿, which is “vacuously” a choice set, because S \ A
“vacuously” contains exactly one element for each A 2 F D ¿.
S S
6.19 Example. With 1 D f¿g, if F D f1g, then S F D f1g D 1 D f¿g,
whence there exists exactly one choice set S F D 1, namely S D 1. Indeed,
the only element in F D f1g is 1, and S \ 1 DS1 \ 1 D f¿g, which contains exactly
one element, namely ¿ 2 1. In contrast ¿ F D 1 is not a choice set.
Theorem 6.20 shows that choice sets are equivalent to choice functions.
6.20 Theorem. Within the Zermelo-Frænkel set theory, the Choice-Function Prin-
ciple 6.8 is logically equivalent to the Choice-Set Principle 6.17.
www.pdfgrip.com
S
Proof. Each choice set S F corresponds to the choice function FS defined by
n [ o
FS WD .A; X/ 2 F F W X 2A\S :
S
Conversely, each choice function is a subset of a Cartesian product: F F F.
Thus F consists of pairs .A; X/ with exactly one X 2 A for each A 2 F . Hence
n [ o
SF WD X 2 F W 9Af.A 2 F / ^ Œ.A; X/ 2 Fg
6.23 Definition (Axiom of Choice). The Axiom of Choice is any of the principles
6.8, 6.14, 6.15, 6.17, 6.21, included as an axiom in a set theory.
6.1 . Find the number of choice functions for each finite set F D fA0 ; : : : ; AN1 g
of nonempty finite sets where each element Aj has exactly Nj elements.
6.2 . Find the number of choice sets for each finite set F D fA0 ; : : : ; AN1 g of
nonempty finite sets where each element Aj has exactly Nj elements.
6.3 . Prove that each finite set of nonempty finite sets has a choice set.
6.4 . Prove that the Choice-Function Principle 6.8 is logically equivalent to the
“Pairwise Disjoint” Choice-Function Principle 6.14. within the Zermelo-Frænkel
set theory.
6.5 . Translate the Choice-Set Principle 6.17 into a logical formula, similar to
formula (6.1) for the Choice-Function Principle 6.8.
6.6 . Translate the “Pairwise Disjoint” Choice-Set Principle 6.21. into a logical
formula, similar to formula (6.1) for the Choice-Function Principle 6.8.
6.7 . Prove that the Choice-Relation Principle 6.15 is logically equivalent to the
Choice-Function Principle 6.8 in the Zermelo-Frænkel set theory.
6.8 . Translate the “Pairwise Disjoint” Choice-Function Principle 6.14 into a
logical formula, similar to formula (6.1) for the Choice-Function Principle 6.8.
This section introduces two principles and shows that each of them implies the
Choice Principle.
6.25 Example. The order is a well-order on the set N of all natural numbers.
6.26 Example. Every finite set has a well-order. Indeed, by definition of “finite” for
each finite set E there exists a natural number N 2 N and a bijection F W E ! N.
The subset N N is well-ordered by example 6.25 and theorem 3.197. Hence the
relation defined on E by X Y if and only if F.X/ F.Y/ well-orders E.
Theorem 6.27 shows that the existence of a choice function on a set follows from
the existence of a well-ordering on its union.
S
6.27 Theorem. For each set F of S nonempty sets with a well-ordered union F ,
there exists a function C W F ! F such that C.A/ 2 A for every set A 2 F .
S
Proof. Let C.A/ be the first element of A relative to the well-order on F . t
u
6.28 Example. If F D P.N/ nSf¿g, which is the set of all nonempty subsets of the
set N of natural numbers, then F D N, where is a well-order. By the proof of
theorem 6.27, the function C W F ! N with C.A/ D min.A/ chooses the smallest
element of A and thus is a choice function for F .
6.29 Example. No well-orders are known for R D P.N/, which is the set of all
subsets of the set N of natural numbers, which
Sis also isomorphic to the set of all real
numbers from 0 through 1. Because R D G with G D P ŒP.N/ n f¿g from
example 6.7, any specific well-order on R would yield a specific choice function on
G , by theorem 6.27.
Theorem 6.30 shows a logical relation between the foregoing principles.
6.30 Theorem. Zermelo’s Well-Ordering Principle 6.24 implies the Choice-
Function Principle 6.8.
Proof. Theorem
S 6.27 shows that if every set admits a well-ordering relation, in
particular, F , then every set F of nonempty sets has a choice function. t
u
As sets (of pairs), relations are partially ordered by inclusion. Theorem 6.31
shows that every total order is maximal relative to inclusion among partial orders.
6.31 Theorem. Each reflexive total order on a set is maximal among all partial
orders on that set.
Proof. If R is a reflexive total order on a set A, and if T is a relation on A such
that R ¨ T, then there are X; Y 2 A such that .X; Y/ 2 T but .X; Y/ … R, whence
.Y; X/ 2 R by totality of R. Hence X 6D Y by reflexivity of R. Consequently, .Y; X/ 2
R T and .X; Y/ 2 T with X 6D Y. Thus T is not anti-symmetric and therefore not
a partial order. Therefore, R is not properly contained in any partial order on A. ut
The concept of maximality in a partially ordered set leads to Zorn’s Maximal-
Element Principle.
www.pdfgrip.com
6.9 . Prove that the hypothesis of Zorn’s Maximal-Element Principle 6.32 implies
that the partially ordered set is not empty.
6.10 . Prove that the conclusion of Zorn’s Maximal-Set Principle 6.34 requires the
hypothesis that the set be not empty.
6.11 . Translate Zorn’s Maximal-Set Principle 6.34 into a logical formula, similar
to formula (6.1) for the Choice-Function Principle 6.8.
6.12 . Translate Zermelo’s Well-Ordering Principle 6.24 into a logical formula,
similar to formula (6.1) for the Choice-Function Principle 6.8.
6.13 . For each partial order on a set E, let C P.E/ be the set of all chains
relative to in E. Thus each element of C is a subset of E on which is a linear
order. Partially order C by inclusion. Prove that the union of each chain in C relative
to inclusion is an element of C .
6.14 . Prove that Zorn’s Maximal-Set Principle 6.34 is logically equivalent to
Zorn’s Maximal-Element Principle 6.32 within the Zermelo-Frænkel set theory.
This section shows that the Choice Principle is equivalent to the distributivity of
intersections over unions of sets.
6.37 Theorem. Within the Zermelo-Frænkel set theory, the Family Choice-
Function Principle 6.12 is logically equivalent to the Multiplicative Principle 6.36.
Proof. Definitions 6.11 and 6.35 show that the Cartesian product consists of all the
family choice-functions as its elements. Thus the Cartesian product is nonempty if
and only if there exists a family choice-function. t
u
This subsection shows that the Choice Principle is equivalent to the distributivity
of intersections over unions of families of sets. The following development follows
Dugundji’s [30, Ch. I, § 9.7–9.8, p. 25].
6.38 Definition (Distributive Principle). For each family of sets fAi W i 2 I g and
each partition fI` W ` Q2 L g of the indexing set I , which sets up equivalence
classes in I , let K WD `2L I` ; then
! 0 1
[ \ \ [
Ak.`/ D @ Ai A : (6.2)
k2K `2L `2L i2I`
6.39 Example. For the self-indexed family F WD fA; B; C; Dg and the self-indexed
partition G WD ffA; Bg; fC; Dgg, the right-hand side of equation (6.2) becomes
[
E D A [ B;
E2fA;Bg
[
E D C [ D;
E2fC;Dg
0 1
\ [
@ Ai A D .A [ B/ \ .C [ D/:
`2L i2I`
The Distributive Principle 6.38 and its negation are not propositions in the
Zermelo-Frænkel set theory; there they are mere formulae. Nevertheless, theo-
rem 6.37 shows that one of the inclusions, but only one, implied by equation (6.2)
is a theorem in the Zermelo-Frænkel set theory.
www.pdfgrip.com
6.40 Theorem. Within the Zermelo-Frænkel set theory, for each family of sets
fAi W i 2 I g and each partition fI` W ` Q 2 L g of the indexing set I , which
sets up equivalence classes in I , let K WD `2L I` ; then
! 0 1
[ \ \ [
Ak.`/ @ Ai A : (6.3)
k2K `2L `2L i2I`
Proof. This proof unravels the definitions of unions, intersections, and Cartesian
products: T
S
X 2 k2K `2L Ak.`/ S
m definition of
T
9k .k 2 K / ^ X 2 `2L Ak.`/
T
m definition of
˚
9k .k 2 K / ^ 8` .` 2 L / ) X 2 Ak.`/
.k 2 K / )
+
.8`f.` 2 L / ) Œk.`/ 2 I` g/
8` Œ.` 2 L / ) f9i Œ.i 2 I` / ^ .X 2 Ai /g
S
m h i definition of
S
8` .` 2 L / ) X 2 i2I` Ai
T
m S definition of
T
X2 `2L i2I` Ai
t
u
Theorem 6.41 shows that in the Zermelo-Frænkel set theory, the Distributive
Principle 6.38 is equivalent to the foregoing choice principles.
6.41 Theorem. Within the Zermelo-Frænkel set theory, the Multiplicative
Principle 6.36 is logically equivalent to the Distributive Principle 6.38.
Proof. Reversing the last two equivalences in the proof of theorem 6.40 gives
2 0 13
\ [
4X 2 @ Ai A5 , f8` Œ.` 2 L / ) f9i Œ.i 2 I` / ^ .X 2 Ai /gg ;
`2L i2I`
fi 2 I` W X 2 Ai g 6D ¿:
If the Distributivity of Intersection over Union Principle 6.38 holds, then equa-
tion (6.2) gives
!
[ \
Ak.`/ D f¿g;
k2K `2L
Q
which is not empty, whence neither is K D `2L I` , so that the Multiplicative
Principle 6.36 holds. t
u
6.15 . Translate the Multiplicative Principle 6.36 into a logical formula, similar to
formula (6.1) for the Choice-Function Principle 6.8.
6.16 . Translate the Distributive Principle 6.38 into a logical formula, similar to
formula (6.1) for the Choice-Function Principle 6.8.
Based on Dugundji’s proof [30, Ch. II, § 2], this section proves that the Choice-
Function Principle 6.8 implies Zorn’s Maximal-Element Principle 6.32 by means of
the concept of a “tower” of sets relative to a function.
www.pdfgrip.com
S
In either case R 2 S .
(T.C) For each A 2 S , either M [ fF.M/g A or A M, by definition of S .
In the first case, if M [ fF.M/g A, then M [ fF.M/g A A [ fF.A/g.
In the second case, if A M and A 2 S M , then A [ fF.A/g 2 M because
M is a tower, whence either M [ fF.M/g A [ fF.A/g or A [ fF.A/g M
because M is medial in M . Thus A [ fF.A/g 2 S .
Thus S is a sub-tower of the smallest sub-tower M . Hence S D M , so that
A M or M [ fF.M/g A for each medial element M 2 M and each A 2 M . u
t
6.46 Theorem. Every element of the smallest sub-tower of any tower is medial in
the smallest sub-tower. The smallest sub-tower is linearly ordered by inclusion.
Proof. This proof consists of verifying that the set V of all medial element of M is
a sub-tower of M .
(T.A) ¿ 2 V because ¿ is medial in M by example 6.43.
(T.B) If U V is a chain in V and B 2 U V , then B is medial in M , so
that B [ fF.B/g A or A B for each A 2 M , by theorem 6.45.
there exists any A 2 U such that B [ fF.B/g A, then
In the first case, if S
B [ fF.B/g A U .
In the second case, if there does not exist any A 2 U such that B
S[ fF.B/g A,
then A B forSevery A 2 U V , by definition of V whence U B.
In either case U 2 V .
(T.C) Every A 2 V is medial in M by definition of V . By theorem 6.45 for
every B 2 M either B A, in which case B A A[fF.A/g, or A[fF.A/g
B. In either case A [ fF.A/g is medial in M , so that A [ fF.A/g 2 V .
Thus V is a sub-tower of the minimal sub-tower M . Hence V D M , so that every
element of M is medial in M . Hence M is linearly ordered by inclusion. t
u
6.47 Theorem. For every tower T P.X/ relative to a function F W T ! X
there exists A 2 T such that F.A/ 2 A.
S
Proof. Let M be the smallest sub-tower, and let A WD M . Then A 2 M by
definition 6.42 of a tower and because M M and M is linearly ordered, by
theorem 6.46. Hence
S also A [ fF.A/g 2 M by definition 6.42 of a tower, whence
A [ fF.A/g M D A, and hence F.A/ 2 A. t
u
This subsection shows that Zorn’s Maximality Principle follows from the Choice
Principle.
6.48 Theorem. The Choice-Function Principle 6.8 implies Zorn’s Maximal-
Element Principle 6.32.
www.pdfgrip.com
Proof. For each set X with a reflexive partial order let C be the set of all chains
relative to in X.
For each nonempty chain C 2 C , let UC be the set of all upper bounds for C
relative to in X. The hypotheses of Zorn’s Maximal-Element Principle 6.32 imply
that UC 6D ¿.
If the Choice-Function Principle 6.8 holds, then there exists a function
F W fUC W C 2 C g ! X such that F.UC / 2 UC , which chooses one upper bound
F.UC / for each nonempty chain C 2 C . Hence define DC WD fK 2 X W ŒF.UC /
K ^ f:ŒK F.UC /gg, which is the set of all the elements of X that strictly follow
the upper bound F.UC /. If there exists C 2 C such that DC D ¿, then F.UC / is a
maximal element of X.
Alternatively, if DC 6D ¿ for every C 2 C , then, again by the Axiom of
Choice 6.8, there is a function G W fDC W C 2 C g ! X such that G.DC / 2 DC
for each C 2 C . The hypotheses of Zorn’s Maximal-Element Principle 6.32 also
imply X 6D ¿, so that there exists Z 2 X. Define H.¿/ WD Z and H.C/ WD G.DC /.
The remainder of the proof verifies that T WD C [ f¿g is a tower relative to H.
(T.A) ¿ 2 T by the definition T WD C [ f¿g. S
(T.B) If S is a chain in T relative
S to inclusion, then S is a chain relative to
in X. Indeed, for all K; L 2 S there exist A; B 2 S such that K 2 A and
L 2 B. Also, S is linearly ordered by inclusion, so that A B or B A. Hence
K; L 2 B or K; L 2 A. Moreover, A; B 2 S T are chains relative to in X,
S A and B are linearly ordered, and hence K L or L K, in B or A.
whence
Thus S 2 T .
(T.C) If C 2 T , then H.C/ is an upper bound for C relative to in X, whence
C [ fH.C/g is also a chain relative to in X. Hence C [ fH.C/g 2 T .
By theorem 6.47 there is a chain C 2 T such that H.C/ 2 C. Hence H.C/
F.UC / because F.UC / is an upper bound for C. Yet H.C/ 2 DC is a strict upper
bound with :ŒH.C/ F.UC /. This contradiction completes the proof: there is a
chain C 2 C for which DC D ¿, so that F.DC / is maximal relative to in X. t
u
Zermelo’s Theorem [144] consists of the logical implication that the Well-
Ordering Principle follows from the Axiom of Choice within the Zermelo-Frænkel-
Choice set theory.
Zorn’s Lemma consists of the logical implication that Zorn’s Maximal-Element
Principle follows from the Axiom of Choice within the Zermelo-Frænkel-Choice
set theory.
6.49 Definition (Interval). In a set E pre-ordered by a relation a subset S E
is an interval if and only if for all U; W 2 S and every V 2 E, if U V W, then
V 2 S.
6.50 Example. The empty subset ¿ and the whole set E are intervals.
www.pdfgrip.com
This section states principles that are related to the Axiom of Choice, in particular,
in the sense that they are not propositions within the Zermelo-Frænkel set theory.
This subsection states principles that are equivalent to the Choice Principle in the
Zermelo-Frænkel set theory.
6.51 Definition (Hausdorff’s Particular Maximal-Chain Principle). In a set of
sets partially ordered by inclusion, each chain is a subset of a maximal chain [128,
p. 248, H1 ].
6.55 Definition (Counting Principle). For each set E there is an ordinal O and a
bijective function F W O ! E with domain O [88, p. 117], [128, p. 241].
www.pdfgrip.com
6.57 Definition (Trichonomy Principle). For all sets A and B, there exists an
injection F with F W A ,! B with domain A, or F W B ,! A with domain B [88,
p. 118], [128, p. 241].
6.58 Definition (Mapping Principle). For all nonempty sets A and B, there exists
a surjection F with F W A B, or F W B A [88, p. 118].
This subsection states principles that follow from the Choice Principle in the
Zermelo-Frænkel set theory.
6.61 Definition (Finite Sets and Infinite Sets). A set E is finite if and only if there
is N 2 N and a bijection F W N ! E; it is infinite if and only if it is not finite.
6.62 Theorem. In the Zermelo-Frænkel-Choice set theory, every infinite set con-
tains a denumerable subset.
Proof. For each set E, let E be the set of all injections from any subset of N into E.
Thus each element F 2 E is a subset of N E , and E P.N E/ is partially
ordered by inclusion. If E is infinite, then for each N 2 N there exists an injection
F W N ,! E, which is also a subset of N E.S In particular E 6D ¿.
S Each chain F E defines an injection F 2 E . Indeed, for each .X; Y/ 2
F , there exists F 2 F such that .X; Y/ 2 F. If also .X; Z/ 2 G 2 F , then F G
S G F, because F is a chain, whence Y D Z. Moreover, if .U; Y/; .X; Y/ 2
or
F , then there exist F; G 2 F such that .X; Y/ 2 F and S .U; Y/ 2 G, with F G
or G SF, whence U D X by injectivity of F or G. Thus F is an upper bound for
F and F 2 E . By Zorn’s Maximal-Element Principle 6.32 the set E contains a
maximal element F.
If the domain of F was finite, then a reparametrization of its domain would give
an injection G W N ,! N, which would not be injective, by definition of infinite.
Hence the domain of F is an infinite subset of N.
www.pdfgrip.com
Chapter 7
Applications: Nobel-Prize Winning Applications
of Sets, Functions, and Relations
7.1 Introduction
This chapter shows what mathematicians recognize in the argument: links between
applications and mathematical sets, functions, and relations. Expositions considered
concise and elegant by mathematicians are listed in the references.
The mathematical analysis of games, called game theory, explains why and
predicts how people will rationally make decisions against their interest. Examples
include decisions whether to arm one’s self. For their contributions to game theory,
John Harsanyi, Reinhard Selten, and John Forbes Nash, Jr., received the Nobel
Prize in Economics in 1994 [93–95]. This introduction to game theory relies on
mathematical functions and ordering relations.
7.2.1 Introduction
Fig. 7.1 Early days of the nuclear arm race. Left: a Soviet modified SS-6 (Sapwood) with
Sputnik 2 on 3 November 1957, photograph courtesy NASA. Right: an Atlas rocket lifts off,
photograph courtesy U.S. Air Force. (http://www.history.nasa.gov/SP-4408pt1.pdf) (http://www.
nationalmuseum.af.mil/shared/media/photodb/photos/050406-F-1234P-014.jpg)
www.pdfgrip.com
7.1 Example (The Prisoner’s Dilemma [131, 132]). The police charge Al and Bo
with a crime and hold them in separate cells, so that Al and Bo cannot communicate
with each other. Al and Bo believe that they have at least a first strategy:
(PD.1) If Al and Bo do not confess to the crime, then they both go free.
However, the police gives each of them another strategy, to confess:
(PD.2) If both Al and Bo confess to the crime, then each gets sentenced to prison.
(PD.3) If one confesses but the other does not confess, then whoever confesses
gets a reward and goes frees, while whoever does not confess gets a death
sentence.
Al and Bo’s dilemma is whether to confess or not to confess.
The Prisoner’s Dilemma also applies to two countries, or their rulers, or any two
individuals, Al and Bo, deciding whether to cooperate not to arm themselves, or to
defect the agreement of cooperation and arm themselves. The Prisoner’s Dilemma
predicts that they will both arm themselves, as shown in figure 7.1, and explains
why [13, 126], as demonstrated in subsection 7.2.2. Game theory also explains
animal behavior, for example musth in male African elephants [87]. To get a sense
of Nash’s work, the reader may at this stage try to design a method to analyze
mathematical games such as The Prisoner’s Dilemma but with any numbers of
players and strategies.
This subsection describes ways to design a mathematical model and analyze games
such as The Prisoner’s Dilemma.
7.2 Example (The Prisoner’s Dilemma, continued [131, 132]). What Al and Bo
eventually get (reward, freedom, prison, or death) is called their payoff. The first
step in designing a mathematical model of the game consists in modeling the
players’ preferences with ordering relations on each player’s set of payoffs. In this
example, it seems reasonable to assume that Al and Bo rank all four payoffs in the
same order of preference, with > meaning “better than”:
The analysis of the game then examines every combination of strategies in the
table. For instance, Al may examine each of Al’s own strategies and seek to avoid
the worst payoff.
Al confesses. If Al confesses, then meanwhile two cases can occur: Bo can either
confess or not confess.
Bo confesses. If Bo also confesses, then Al faces a prison term,
Bo does not confess If Bo does not confess, then Al goes free with a reward.
Thus, if Al confesses, then the worst that can happen to Al is a prison term.
Al does not confess. Similarly, if Al does not confess, then two cases can occur:
Bo can either confess or not confess.
Bo confesses. If Bo confesses, then Al faces death.
Bo does not confess If Bo does not confess either, then Al goes free.
Thus, if Al does not confess, then the worst that can happen to Al is death.
Consequently, to avoid death with certainty but without knowing Bo’s action, Al
must confess.
Table 7.3 is symmetric in the sense that swapping Al’s and Bo’s rôles does not
change their payoffs. Thus Al’s reasoning also applies to Bo, who must also confess
to avoid death with certainty.
Therefore, to avoid death with certainty, Al and Bo both confess, even though
they would both be better off by not confessing.
7.4 Example (Arm race [13, 126]). If two individuals or countries cooperate not to
arm themselves, then they remain free and safe. If both defect and arm themselves,
then they remain safe but incur a penalty with the cost of armament. If either
one defects and arms itself while the other one does not arm itself, then whoever
arms itself stays free and gets a reward by conquering and looting the other, who
suffers from death, destruction, and loosing the property to the conqueror. The
ordering of the outcomes just described are the same as those in table 7.3 for
The Prisoner’s Dilemma. Therefore, the analysis of The Prisoner’s Dilemma in
example 7.2 predicts and explains why they will both arm themselves.
Example 7.5 shows another way to analyze The Prisoner’s Dilemma.
Some of the modeling and analysis of The Prisoner’s Dilemma carry over to other
games. To this end, this subsection shows that a special type of strategy contains an
equilibrium, from which no players has any incentive to switch to another strategy.
For a game with exactly two players, Al and Bo, assume that Al has a nonempty set
of strategies SAl and a nonempty set of payoffs PAl , while Bo has a nonempty set of
strategies SBo and a nonempty set of payoffs PBo . Assume also that Al and Bo have a
linear order “” on their sets of payoffs. If Al plays a strategy A 2 SAl and Bo plays
a strategy B 2 SBo , then Al gets a payoff pAlA ;B
while Bo gets a payoff pBo
A ;B
. Thus for
each Ci 2 fAl; Bog, p is a function p W SAl SBo ! PCi . Table 7.6 shows the payoff
Ci Ci
another strategy. A strategy is strongly dominant for a player (Al) if and only if,
for each fixed combination of the other players’ strategies, that player (Al) always
gets a lower payoff by switching to another strategy.
7.8 Example. In The Prisoner’s Dilemma (examples 7.1 and 7.5), Al’s strongly
dominant strategy is to confess: If Bo confesses, then Al gets a higher payoff by
confessing (prison) than by not confessing (death). Similarly, if Bo does not confess,
then Al also gets a higher payoff by confessing (freedom and a reward) than by
not confessing (freedom). By symmetry of The Prisoner’s Dilemma, Bo’s strongly
dominant strategy is also to confess.
7.9 Definition (dominant strategy equilibrium). A position is an array of strate-
gies, with one strategy for each player. A weakly (respectively strongly) dominant
strategy equilibrium is a position where each player plays a weakly (respectively
strongly) dominant strategy.
7.10 Example. In The Prisoner’s Dilemma, the position where Al and Bo both
confess is a dominant strategy equilibrium, where their dominant strategies meet.
Neither Al nor Bo has any incentive to switch to another strategy (not to confess),
because whoever decides not to confess while the other still confesses faces death.
7.11 Definition (Nash equilibrium). A Nash equilibrium is a position from
which no players can get a higher payoff by changing only their own strategy, while
all the other players’ strategies remain fixed [93, p. 287].
7.12 Example. In The Prisoner’s Dilemma, the position where Al and Bo both
confess is a Nash equilibrium: If Al decides not to confess but Bo still confesses,
then Al’s payoff drops. Similarly, if Bo decides not to confess but Al still confesses,
then Bo’s payoff drops. In contrast, the position where neither Al nor Bo confess is
not a Nash equilibrium, because whoever decides to confess gets a higher payoff if
the other one still does not confess.
7.13 Theorem. Every weakly dominant equilibrium is a Nash equilibrium.
Proof. Table 7.14 shows the conditions for either of Bo’s strategic to be weakly
dominant in a game with only one other player (Al).
If R is Bo’s single weakly dominant strategy (Bo’s R column in table 7.14), then
Bo plays R. Indeed, if Al plays T , then R gets Bo a larger payoff, because pBo T ;L
pBo
T;R
;
Bo Bo
if Al plays B , then R also gets Bo a larger payoff, because pB;L pB;R .
Since Al knows the game, Al knows that R is Bo’s single weakly dominant
strategy. Thus, Al knows that Bo will play R. Consequently, Al has only two payoffs
available: pAl
T;R
or pAl
B;R
. Therefore, Al chooses a strategy that yields the larger available
payoff. For instance, if pAlT ;R
pAl
B;R
, then Al plays B.
Hence the game ends at (B,R) with payoffs pAl B ;R
to Al and pBo B ;R
to Bo. From (B,R)
Bo cannot get a higher payoff by switching to L because pBo B ;L
pBo B ;R
. From (B,R) Al
cannot get a higher payoff by switching to T because pAl T;R
pAlB ;R
. Thus (B,R) is a
Nash equilibrium.
If R and L are both weakly dominant strategies for Bo, then Bo’s payoffs are
identical in R and L. If pAl B ;R
is the largest of Al’s four payoffs in Bo’s weakly
dominant strategies, then (B,R) is still a Nash equilibrium. Yet Al no longer knows
which of R or L Bo chooses: the game might not end at a Nash equilibrium. t
u
Some games or situations might not occur more than once in the players’ life time,
for instance, The Prisoner’s Dilemma. Yet other games or situations may occur more
than once in the players’ life time. Each player may then adopt a mixed strategy,
defined by choosing one strategy some of the time and another strategy at other
times. Each player’s payoff may then be a weighted average of the payoffs from
the different strategies in the mixed strategy. For emphasis, strategies that are not
mixed, for instance, as in The Prisoner’s Dilemma, are called pure strategies.
7.15 Example (The Battle of the Sexes [92]). Al and Bo would like to go to a
show together. Al prefers a Jazz concert while Bo prefers a play at the theater.
Nevertheless, they would prefer to go to the same show together, rather than to
different shows separately. After some negotiation they agree on a show. On the
day of the show, however, they each forget which show they had agreed on, but
cannot communicate with each other. So they each must decide to which show to
go, without knowing in advance where the other will go. Table 7.16 shows their
payoffs. In hope of finding Bo, Al goes to the Jazz concert one half of the time, and
Table 7.17 Payoffs for the Battle of the Sexes over four days.
BO GOES TO
JAZZ (J) PLAY (P)
PAYOFFS TO PAYOFFS TO
AL BO AL BO
JAZZ (J) 3 (DAY 1) 2 1 (DAY 3) 1
AL GOES TO
PLAY (P) 0 (DAY 2) 0 2 (DAY 4) 3
to the play at the theater the other half of the time. In hope of finding Al, Bo does
the same. Table 7.17 shows the four combinations of strategies; for this example,
assume that they occur equally frequently. Thus Al’s and Bo’s average payoffs over
four days are
3C0C1C2 3
pAl .1=2; 1=2/ D D I
4 2
2C0C1C3 3
pBo .1=2; 1=2/ D D :
4 2
Al now decides to go to the Jazz concert every day, while Bo still goes to Jazz every
other day and to the play every other day. There are then only two kinds of days,
and their average payoffs become
3C1
pAl .1; 0/ DD 2I
2
2C1 3
pBo .1=2; 1=2/ D D :
2 2
Hence arises the question whether other mixed strategies can increase both Al’s
and Bo’s payoffs.
This section shows that if neither player has any weakly dominant pure strategy, then
the game still has a Nash equilibrium for some combination of mixed strategies.
7.18 Definition (mixed strategy). The same players might play multiple rounds
of the same game and independently choose any available strategy for each round.
Over M rounds, Al may play T exactly T times and B exactly B times, with T C
B D M. Thus Al plays T with a frequency t D T=M and B with a frequency
b D B=M D 1 t. The ordered pair of frequencies .t; 1 t/ is called a mixed
www.pdfgrip.com
strategy for Al. Meanwhile, Bo may play L exactly L times and R exactly R times,
with LCR D M. Thus Bo plays L with a frequency ` D L=M and R with a frequency
r D R=M D 1 `. Thus .`; 1 `/ is a mixed strategy for Bo. Mixed strategies are
also subject to the condition that Al and Bo play (T,L) with frequency t ` and hence
also receive payoffs pAl
T ;L
and pBo
T ;L
with frequency t `. Similar products of frequencies
apply to the other plays (T,R), (B,L), and (B,R). The sum of such payoffs to Al and
Bo are
7.19 Theorem. Every game for two players with two pure strategies has a Nash
equilibrium.
Proof. For Bo’s payoff, collecting similar powers of t and ` leads after some algebra
to the equivalent formula
t Œ.pBo
T ;L
pBo
T ;R
/ C .pBo
B;R
pBo
B ;L
/ .pBo
B;R
pBo
B ;L
/ D 0;
which does not depend on ` and thus strips away from Bo all controls over Bo’s
own payoff. In particular, Bo cannot get a higher payoff by switching to a different
frequency `. Yet if Bo can choose a frequency ` so that Al cannot control Al’s own
payoff, then they are at a Nash equilibrium .t ; ` /. The following considerations
show that such a Nash equilibrium exists provided that neither Al nor Bo has any
weakly dominant strategy. Table 7.20 shows the conditions for Bo not to have any
weakly dominant strategy, obtained from the logical negation of Table 7.14. Two
cases emerge: either pBo B ;L
< pBo
B ;R
and pBo
T ;L
> pBo
T ;R
, or pBo
T ;L
< pBo
T ;R
and pBo
B ;L
> pBo
B ;R
.
In the first case, pB;L < pB;R and pT;L > pT;R . Hence pB;R
Bo Bo Bo Bo Bo
pBo
B;L
> 0 and
pBo
T ;L
pBo
T;R
> 0. Consequently,
.pBo
T;L
pBo
T ;R
/ C .pBo
B ;R
pBo
B ;L
/ > .pBo
T;L
pBo
T ;R
/ > 0:
Thus, Al can choose a frequency t between 0 and 1 where Bo has no controls over
Bo’s payoff:
www.pdfgrip.com
.pBo
B;R
pBo
B ;L
/
1 > t D > 0:
.pBo
T;L
pT;R / C .pB;R
Bo Bo
pBo
B ;L
/
.pAl
B;R
pAl
T ;R
/
1 > ` D > 0:
.pAl
B;L
pT;L / C .pB;R
Al Al
pAl
T ;R
/
Neither Al nor Bo can change their own payoff by switching to another frequency
while the other player’s frequency is fixed. Thus .t ; ` / is a Nash equilibrium. The
second case, where pBo
T;L
< pBo
T ;R
and pBo
B ;L
> pBo
B ;R
, is entirely similar.
If a player has a dominant strategy, then such a strategy contains a Nash
equilibrium, by theorem 7.19. t
u
For Nash’s equilibria with two players and two strategies, see also
[106, p. 138–139], [107], [119, p. 155–168].
7.21 Example (The Professors’ Problem). Al and Bo are on the tenured faculty
at King Game’s College. Each may either teach students or sit on committees.
Teaching does not hamper any one’s work, but committees hamper other faculty
members’ work, to the extent summarized by Al’s and Bo’s end-of-the-year bonus
payoffs in table 7.22. The sum of Al’s and Bo’s payoffs reflects the value of their
work to the College. Al and Bo know table 7.22 but work in different offices and
thus do not cooperate with each other; hence they must choose a strategy without
knowing in advance what the other is doing.
Bo has exactly one weakly dominant strategy, which is to sit on committees;
consequently, Bo declines to teach and decides to sit. Al knows table 7.22 and thus
www.pdfgrip.com
knows that Bo will sit on committees; therefore, to get the higher payoff available
to Al under Bo’s decision to sit, Al also declines to teach and decides to sit on
committees.
Not only do Al and Bo choose the Nash equilibrium where they are both worse
off than in the other Nash equilibrium, but the sum of the values of their work to the
College is the worst of all possibilities.
The administration’s challenge is to entice the faculty to teach by modifying the
payoff table.
Table 7.23 shows the new game on campus after the administration capped
payoffs from committees to 5 units. If Bo sits on committees, then Al may also
sit for a payoff of 4 rather than teach for 6: Al now values teaching only up to 4.
To get a deeper sense of one of Nash’s many contributions, the reader may
attempt proving the existence of a Nash equilibrium in games for any number of
players with any number of strategies.
7.24 Example (Blue against Red [26]). In Melvin Dresher’s initial context, the
Blue and Red commanders led opposing armed forces [26, p. 4], but they might also
lead sports teams [12]. Table 7.25 shows only the payoffs to the Blue commander.
7.1 . Identify all the weakly dominant pure strategies in The Battle of the Sexes
(table 7.16 in example 7.15).
www.pdfgrip.com
7.2 . Identify all weakly dominant pure strategies in the Administration’s Solution
to the Professors’ Problem (table 7.23 in example 7.21).
7.3 . Identify all Nash equilibria with pure strategies in The Battle of the Sexes
(table 7.16 in example 7.15).
7.4 . Identify all Nash equilibria with mixed strategies in the Administration’s
Solution to the Professors’ Problem (table 7.23 in example 7.21).
7.5 . Prove or disprove that every Nash equilibrium is a dominant strategy
equilibrium.
7.6 . Prove or disprove that every two-player game restricted to pure strategies has
a Nash equilibrium.
7.7 . For each function f W A B ! C such that C is linearly ordered and for each
nonempty subset E A B the image f 00 .E/ has a first element minŒf 00 .E/ and a
last element maxŒf 00 .E/, prove that
7.8 . Design a function f W A B ! C such that C is linearly ordered and for each
nonempty subset E A B the image f 00 .E/ has a first element minŒf 00 .E/ and a
last element maxŒf 00 .E/, with a strict inequality
Denote by “Al’s strategies” and “Bo’s strategies” the set of all strategies available
to Al and Bo respectively.
7.9 . Prove that if each player plays so as to avoid the worst payoff available, then
they get the payoffs
payoff to Al D maxfminfpAl
SAl ;SBo W SBo 2 Bo’s strategiesg W SAl 2 Al’s strategiesg;
payoff to Bo D maxfminfpBo
SAl ;SBo W SAl 2 Al’s strategiesg W SBo 2 Bo’s strategiesg:
www.pdfgrip.com
7.10 . Assume that in a game for two players with two strategies pAl SAl ;SBo pSAl
Al
0 0
;SBo
if and only if pSAl ;SBo pS0 ;S0 for all positions .SAl ; SBo / and .SAl0 ; SBo
Bo Bo 0
/. Prove that if
Al Bo
each player plays so as to avoid the worst payoff available, then they get the payoffs
payoff to Al D minfmaxfpAl
SAl ;SBo W SBo 2 Bo’s strategiesg W SAl 2 Al’s strategiesg;
payoff to Bo D minfmaxfpBo
SAl ;SBo W SAl 2 Al’s strategiesg W SBo 2 Bo’s strategiesg:
7.11 . For noncooperative games with any number of players and any number of
pure strategies, prove or disprove that if at least one player has at least one weakly
dominant strategy, then such a strategy contains a Nash equilibrium.
7.12 . For noncooperative games with two players but any number of pure strate-
gies, prove or disprove that if at least one player has at least one weakly dominant
strategy, then such a strategy contains a Nash equilibrium.
7.13 . Analyze the game of Blue against Red defined by table 7.25 in example 7.24.
7.14 . Prove or disprove that the formulae from exercise 7.10 still hold for more
than two strategies.
Match making pairs up items or individuals from two groups. For example, the
National Resident Matching Program (NRMP) uses an algorithm developed by
David Gale, Alvin E. Roth, and Lloyd S. Shapley to match medical doctors with
internships in hospitals [68, 96]. For their work on such problems, Alvin E. Roth
and Lloyd S. Shapley received the Nobel Prize in Economics in 2012. The precise
statements and proofs of their algorithms involve the mathematical concepts of sets,
functions, relations, and induction.
7.3.1 Introduction
The problems considered here have been documented for millennia since Plato’s
time (Figure 7.2): how to arrange for proper marriages [97, p. 27], and how to admit
students to schools [127, p. 44, footnote 11]. The problems consist in matching in
some “optimal” way individuals from two groups, for example, boys and girls for
marriage, students and schools for education, doctors and hospitals for internships,
or, more generally, beggars and choosers (table 7.26). “Optimality” may mean, for
instance, that there are no beggar from one couple and chooser from another couple
that prefer each other to their current mate. A similar optimality applies to schools
www.pdfgrip.com
admitting several students. To get a sense of Gale and Shapley’s work, the reader
may at this stage try to design such optimal match-making procedures.
To specify preferences for certain beggars over others, each chooser c 2 C ranks
all beggars by a strict well-order c on B. Thus relative to c each nonempty subset
V B has a unique first-ranked element, denoted by firstc .V/.
The two disjoint sets B and C, with B well-ordered by each chooser and C well-
ordered by each beggar, complete the mathematical model of the situation. The
second step consists in designing a model of a successful relation.
For each relation M B C, denote its domain by Domain.M/ and its range by
Range.M/. Also, call “single” each beggar b0 2 B0M WD B n Domain.M/ and each
0
chooser c0 2 CM WD C n Range.M/: neither b0 nor c0 are related to anyone by M.
Definition 7.27 specifies the notion of a successful relation by a concept of stability.
7.27 Definition. A relation M B C is unstable if and only if at least one of the
following conditions holds:
(US.1) A beggar and a chooser prefer each other to their current mates: there
are different couples .b1 ; c1 /; .b2 ; c2 / 2 M for whom b2 c1 b1 and c1 b2 c2 , a
condition denoted by .b1 ; c1 / ‰ .b2 ; c2 /.
(US.2) There is a couple .b; c/ 2 M and a “single” beggar b0 2 B0M D
B n Domain.M/ for whom b0 c b and c b0 c0 for every “single” chooser
0
c0 2 CM D C n Range.M/, a condition denoted by .b; c/ ‰ b0 .
0
(US.3) There is a couple .b; c/ 2 M and a “single” chooser c0 2 CM D Cn
Range.M/ for whom c b c and b c0 b for every “single” beggar b 2 B0M D
0 0 0
This subsection defines and demonstrates an algorithm for a match maker (person or
machine) to find a stable relation. First, a match maker with access to all well-orders
from all beggars and choosers can extend any stable relation.
www.pdfgrip.com
Also, for each first-ranked single chooser c1 2 C1 , the set Bc1 B0M of single
0
beggars who rank c1 first in CM ,
is not empty by definition and hence has a unique first element bc1 WD firstc1 .Bc1 /.
Moreover, if c1 6D c2 2 C1 , then bc1 6D bc2 , because bc1 ranks c1 first, ahead of c2 .
Thus the function C ,! B0M , c1 ! bc1 is injective. Define
M1 WD f.bc1 ; c1 / 2 B0M 0
CM W .c1 2 C1 / ^ Œbc1 D firstc1 .Bc1 /gI
O WD M [ M1
M M:
7.32 Example. Consider Gale and Shapley’s unlabeled example [40, 41, p. 389],
adapted in table 7.33. At the intersection of the row for bi and the column for cj ,
the ordered pair .mi ; nj / states that bi ranks cj in position mi , whereas cj ranks bi in
position nj . For instance, at the intersection of the row for b3 and the column for c2 ,
the ordered pair .1; 4/ states that b3 ranks c2 in position 1, whereas c2 ranks b3 in
position 4.
Start from the stable empty relation ¿ B C. Thus B0¿ D B and C¿ 0
D C. The
set of first-ranked choosers is C1 D fc1 ; c2 ; c3 g. They are ranked first by
Among those beggars who ranked them first, their first-ranked beggars are
O to
where all beggars have their first choice. Theorem 7.30 then extends M
where b2 and c4 have their worst choice, but then c4 was ranked last by every beggar.
This subsection describes Gale & Shapley’s algorithm to find a stable relation
without any match maker [40, 41]. In the context of marriages, boys and girls carry
out the algorithm themselves through rounds of proposals and rejections.
www.pdfgrip.com
7.15 . For disjoints sets B and C with the same finite number N of elements, prove
that if in some round any boy b 2 B receives his .N 1/-th rejection, then Gale &
Shapley’s algorithm terminates at the next round.
7.16 . For disjoints sets B and C with the same finite number N of elements, prove
that Gale & Shapley’s algorithm terminates after at most N 2 C 2 2N rounds.
7.17 . Prove or disprove that for all disjoint sets B and C with the same infinite
cardinality there must exist a total stable relation.
7.18 . Carry out Gale & Shapley’s algorithm with Gale & Shapley’s example 7.32.
7.19 . Extend algorithm 7.31 (with a match maker) to stable relations that may
relate each chooser to more than one beggar, with a quota qc of beggars for chooser
c (polygamy, polyandry, schools admitting more than one student, etc.).
7.20 . Extend algorithm 7.34 (without a match maker) to stable relations that may
relate each chooser to more than one beggar, with a quota qc of beggars for chooser
c (polygamy, polyandry, schools admitting more than one student, etc.).
7.21 . Denote by S the subset of the power set P .B C/ consisting of all stable
relations,Spartially ordered by inclusion. Prove that for each chain T S , its union
UT WD T 2 S is an upper bound for T in S .
www.pdfgrip.com
7.3.6 Projects
7.35 Project. Develop concepts, theorems, and algorithms for match making where
beggars might order any subset, not necessarily the whole set, of choosers, and
choosers might order any subset, not necessarily the whole set, of beggars. For
example, the National Resident Matching Program (NRMP) evidently uses an
algorithm to this effect [96].
7.36 Project. Develop concepts, theorems, and algorithms for tri-partite match
making. (Tri-partite reproduction occurs in Isaac Asimov’s novel The Gods Them-
selves [5]. See also the “three-parent” therapy [2].)
7.4.1 Introduction
7.37 Example. In the United States presidential election of 1824, the popular and
electoral votes had ranked Andrew Jackson (Democrat-Republican Party) ahead of
John Quincy Adams (Coalition Party), William H. Crawford, and Henry Clay (Whig
Party); nevertheless the House of Representatives elected John Quincy Adams ahead
of Andrew Jackson (http://www.archives.gov/federal-register/electoral-college/
scores.html).
In the scientific arena, the decision criteria may be measurements from sensors,
which may act as voters, while the alternatives, which play the rôles of candidates,
may be decisions on how to proceed with a mission. In some missions, a simple
majority of votes may fail to select the best alternative (Figure 7.3).
7.38 Example. The Mars Climate Orbiter spacecraft was lost on 23 September
1999, due to small thrusters controlled by faulty unit conversions that compounded
during the year-long flight, which had been tracked by three telemetric systems:
Doppler only, range only, and Doppler and range combined. Two out of the three
systems submitted the same vote and won. Range only, and Doppler and range
combined, both showed a flight path clearing the planet, allowing the mission to
proceed without corrections. Both systems were wrong and the probe crashed.
The minority vote was right: “The Doppler-only solutions consistently indicated
a flight path insertion closer to the planet. These discrepancies were not resolved”
[78, p. 13]. In some situations, discrepancies must be investigated, not voted away.
Among other procedures, the voting procedures considered here are based on
voters ranking all the candidates.
7.39 Example. Table 7.40 shows Judges’ (voters’) ranking of three skaters (candi-
dates) at the 1994 Winter Olympic Games in Lillehammer, Norway [112, p. 22].
The judges’ plurality voting procedure selects the skater rated first by the largest
number of judges, here Baiul ahead of Kerrigan.
www.pdfgrip.com
Many features have been deemed desirable from a voting procedure, for instance,
the following features.
(AIT.1) Unrestricted Domain. Each voter may submit any ranking of the candi-
dates: no rankings are forbidden.
(AIT.2) Unanimity. If every voter prefers candidate Al to Bo, then the voting
procedure must rank Al ahead of Bo.
(AIT.3) Independence of Irrelevant Alternatives. If every voter prefers candidate
Al to Bo in two different elections, then the voting procedure must rank Al
ahead of Bo in both elections, independently (regardless) of changes in the
ranking of any candidate Ci other than Al and Bo between the two elections.
(AIT.4) Absence of Dictators. The ranking from the voting procedure does not
coincide with the ranking of any fixed voter: no voters can dictate the outcome
of all elections.
(AIT.5) Anonymity. The ranking from the voting procedure does not depend
on who cast what vote. (Anonymity may be desirable for public votes, but
undesirable for the votes of elected representatives or scientific sensors.)
To get a sense of Arrow’s work, the reader may at this stage try to design a voting
procedure with all the features just listed to elect one among three candidates: either
find such a voting procedure, or prove that there are none.
Arrow’s Impossibility Theorem states that the first four features, (AIT.1),
(AIT.2), (AIT.3), and (AIT.4) are already mutually incompatible: no such voting
procedures are possible. As just stated, the four features are somewhat vague. For
instance, a voting procedure might need only to accommodate finitely many voters.
A precision sufficient for rigorous reasoning may have to be inserted and culminate
in a mathematical model of the descriptions of the features. To get a deeper sense
of Arrow’s work, the reader may at this stage try to formulate those four features
mathematically.
In general, a set V of voters must rank a set C of mutually exclusive alternatives,
for instance, candidates, decisions, laws, policies, etc. To this end, each voter
submits one ranking of the set C of candidates. From the rankings submitted by
the voters a “social welfare function” F produces a final “aggregate” ranking of the
set C of candidates (Table 7.44).
7.45 Definition. A weak ranking allowing for ties on a set C is a transitive and
strongly connected relation R C C .
To allow for ties, weak rankings need not be anti-symmetric; thus weak rankings
are not partial orders.
The notation .X; Y/ 2 R, also abbreviated by XRY, means that R ranks X before
or tied with Y.
The transitivity of R is defined by 8X8Y8Z fŒ.XRY/ ^ .YRZ/ ) .XRZ/g.
www.pdfgrip.com
Not all social welfare functions are considered here: only those that satisfy the
following four conditions.
(AIT.1) Unrestricted Domain. The domain of each social welfare function F is
all of R .C /V . This means that each voter may submit any ranking of the
candidates for the social welfare function to produce an aggregate ranking
(Table 7.47).
(AIT.2) Unanimity. For all candidates X and Y in C , if in a voters’ profile r each
voter V in V prefers X to Y, then so does the aggregate ranking from the social
welfare function F :
˚
8X8Y8r Œr 2 R .C /V ^ 8VŒ.X; Y/ 2 Pr.V/ ) Œ.X; Y/ 2 PF .r/ :
(AIT.4) Absence of Dictators. The social welfare function does not coincide with
the ranking of any voter:
˚ ˚
: 9V 8r Œr 2 R .C /V ) ŒF .r/ D r.V/ :
For all elements Bo; Ci 2 C , the transposition Bo;Ci is the permutation Bo;Ci 2
SV that swaps Bo and Ci but fixes every other element Al 2 C n fBo; Cig:
f.X; Y/ 2 C C W Bo;Ci .X/ Bo;Ci .Y/g:
V
Arrow’s Impossibility Theorem applies to every set V of voters that is, or can be,
strictly well-ordered by a relation so that each nonempty subset E V has a first
element first.E / and V has a last element last.V /. The strict well-order remains
fixed for the entire proof. For instance, some sets of voters may be in alphabetical
order.
7.49 Theorem (Arrow’s Impossibility Theorem). No social welfare functions
satisfy all four conditions (AIT.1)–(AIT.4).
Proof. This proof expands on [4, 6, 42, 143], showing that every function satisfying
(AIT.1), (AIT.2), (AIT.3) violates (AIT.4). For all distinct candidates Al; Bo 2 C ,
at one extreme focus on any profile r where every voter V 2 V prefers Al to Bo, so
that .Al; Bo/ 2 Pr.V/ for every V 2 V :
By the rule of unanimity so does the aggregate ranking: .Al; Bo/ 2 PF .r/ .
At the other extreme, focus on any profile s where every voter V 2 V prefers Bo
to Al, so that .Al; Bo/ 2 Ps.V/ for every V 2 V :
By the rule of unanimity so does the aggregate ranking: .Bo; Al/ 2 PF .s/ .
Between extremes, for each voter V 2 V define a profile rV 2 R .C / such that
each voter U V prefers Bo to Al, so that Bo U Al for each U V, while each
voter W V prefers Al to Bo, so that Al W Bo for each W V:
Let EBo=Al be the subset of voters for whom F .rV / ranks Bo ahead of Al:
Thus last.V / 2 EBo=Al 6D ¿ and EBo=Al has a first elements VBo=Al WD first.EBo=Al /. This
first element VBo=Al is called a pivotal voter for Bo over Al.
By the rule of Independence of Irrelevant Alternatives VBo=Al is a pivotal voter
for Bo over Al starting from every profile r0 where every voter ranks Al before Bo.
Indeed, for each voter V 2 V , in r0 and in r, also in s0 and in s, as well as in r0V and
in rV , each voter U V, V, and W V ranks Al and Bo relative to each other in
the same way.
The following considerations show that VBo=Al can also dictate the ranking of Bo
relative to any third candidate Ci 2 C n fAl; Bog. To this end, consider the voters’
profile qV defined for V WD VBo=Al by
8
<U V W Bo U
Ci U
Al;
qV VW Al Bo Ci;
: V V
V W W Al W
Bo W
Ci:
Since V D VBo=Al is pivotal for Bo over Al, the aggregate ranking F .qV / still ranks
Al Bo. Also, by unanimity F .qV / ranks Bo Ci. By transitivity the aggregate
ranking F .qV / ranks Al Bo Ci. If V D VBo=Al swaps Al and Bo, then the voting
profile changes from qV to rV , defined by
8
<U V W Bo U
Ci U
Al;
rV VW Bo Al Ci;
: V V
V W W Al W
Bo W
Ci:
Since V D VBo=Al is pivotal for Bo over Al, the aggregate ranking F .rV / now ranks
Bo Al. By Independence of Irrelevant Alternatives F .rV / still ranks Al Ci,
because each voter ranks Al and Ci in the same order in both profiles qV and rV . By
transitivity, the aggregate ranking F .rV / ranks Bo Al Ci. If any, some, or all
voters other than V D VBo=Al swap Bo and Ci, then the voting profile changes from
rV to sV , defined by
8
<U V W Ci U
Bo U
Al or Bo U
Ci U
Al;
sV VW Bo Al Ci;
: V V
V W W Al W
Ci W
Bo or Al W
Bo W
Ci:
Then Al and Bo are ranked in the same order by every voter in rV and sV , whence the
aggregate ranking F .sV / still ranks Bo Al. Similarly, Al and Ci are ranked in the
same order by every voter in rV and sV , whence the aggregate ranking F .sV / still
ranks Al Ci. By transitivity the aggregate ranking F .sV / ranks Bo Al Ci,
even though every voter other than V ranks Ci Bo.
www.pdfgrip.com
Candidate Al plays only an auxiliary rôle in the proof that VBo=Ci is a dictator for
Bo over Ci. If any, some, or all voters swap Al and any candidate Ig other than Bo
and Ci, then the voting profile changes from sV to pV , defined by
8
<U V W Ci U Bo U Al or U
Ci U Bo U ;
pV VW Bo V Al V Ci or Bo Ci ;
: V V V
V W W Al W Ci W Bo or W
Ci W
Bo W
:
Then Bo and Ci are ranked in the same order by every voter in sV and pV , whence
the aggregate ranking F .pV / still ranks Bo Ci, even though every voter other
than V ranks Ci Bo.
For the uniqueness of the dictator, in the strict well-ordering of the voters V ,
a pivotal voter VBo=Ci for Bo over Ci cannot appear later than such a dictator for Bo
over Ci as VBo=Al , because such an earlier dictator VBo=Al would determine the ranking
of Bo over Ci before a pivotal voter VBo=Ci does; consequently, VBo=Ci VBo=Al .
Similarly, a pivotal voter VCi=Bo for Ci over Bo cannot come earlier than such a
dictator for Bo over Ci as VBo=Al , for otherwise such a later dictator could reverse a
pivotal vote from the pivotal voter; therefore VBo=Al VCi=Bo .
Combining VBo=Ci VBo=Al with VBo=Al VCi=Bo gives VBo=Ci VBo=Al VCi=Bo by
transitivity.
Reversing the rôles of Bo and Ci gives the reverse ranks VCi=Bo VCi=Al VBo=Ci .
Consequently, VBo=Ci D VBo=Al D VCi=Bo .
Therefore there is exactly one pivotal voter, who is the pivotal voter and the
dictator for every pair of candidates.
Thus every social welfare function satisfying (AIT.1), (AIT.2), (AIT.3) violates
(AIT.4). t
u
Another rule that might be imposed on a social welfare function pertains to the
anonymity of voters. One way to model the concept of anonymity of voters consists
in requiring that a social welfare function remains invariant under all permutations
of the voters.
7.50 Definition. A social welfare function F W R .C /V ! R .C / is invariant
under permutations if and only if F .r ı / D F .r/ for every profile r W V !
R .C / and every permutation 2 SV .
(AIT.5) Anonymity. A social welfare function respecting voters’ anonymity is
invariant under permutations.
Yet another condition attempts to model the condition that no voters have a right
of veto.
(AIT.6) No Veto. Except perhaps for voter Val, if none of the other voters prefer
any other candidate to Al, in other words, if all the other voters prefer Al to all
the other candidates, or are indifferent between Al and all the other candidates,
then the social welfare function does not rank any other candidate ahead of Al,
regardless of Val’s rankings [115, p. 386].
(AIT.7) Weak Unanimity. Except perhaps for voter Val, if all the other voters do
not prefer Bo to Al, so that they might prefer Al to Bo or be indifferent between
Al and Bo, but Val prefers Al to Bo, then the social welfare function prefers Al
to Bo:
n o
8X8Y8r Œr 2 R .C /V ^ Œ.X; Y/ 2 Pr.Val/ ^ 8VŒ.Y; X/ … Pr.V/
) Œ.X; Y/ 2 PF .r/ :
1.1 Use axioms P1 and P2, theorem 1.12, and Modus Ponens:
` .L/ ) Œ.K/ ) .L/ axiom P1,
` .H/ ) f.L/ ) Œ.K/ ) .L/g theorem 1.12,
` .H/ ) f.L/ ) Œ.K/ ) .L/g ) Œ.H/ ) .L/ ) f.H/ ) Œ.K/ ) .L/g axiom P2,
` Œ.H/ ) .L/ ) f.H/ ) Œ.K/ ) .L/g Modus Ponens.
1.7 The formula fŒ.P/ ) .Q/ ) .P/g ) .P/ cannot be proved using only
implications.
1.9 Use the tautology Œ.H/ ) .L/ ) f.H/ ) Œ.K/ ) .L/g:
` .P/
) Œ.P/ ) .P/ substitution in axiom 1a,
` .P/ ) fŒ.P/ ) .P/ ) .P/g ) Œ.P/ ) .P/ Detachment.
` Œ.P/ ) .P/ Detachment.
1.13 Apply exercise 1.11
and Detachment:
` Œ.P/ ) .P/ ) fŒ.P/ ) Œ.P/ ) .Q/g substitution in 1b,
) Œ.P/ ) .Q/
` fŒ.P/ ) Œ.P/ ) .Q/g ) Œ.P/ ) .Q/ Detachment.
1.15 Apply exercise 1.12 and Detachment:
`T hypothesis,
` .A/ ) .T/ exercise 1.12,
` Œ.A/ ) .T/ ) f.A/ ) Œ.T/ ) .C/g substitution in 1b,
) Œ.A/ ) .C/
` f.A/ ) Œ.T/ ) .C/g ) Œ.A/ ) .C/ Detachment.
1.17
` Œ.B/
) .C/ ) f.A/ ) Œ.B/ ) .C/g 1a,
` Œ.B/ ) .C/ ) f.A/ ) Œ.B/ ) .C/g )
˚
Œ.B/ ) .C/ ) f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C/
) fŒ.B/ ) .C/ ) Œ.A/ ) .C/g 1b,
˚
` Œ.B/ ) .C/ ) f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C/
) fŒ.B/ ) .C/ ) Œ.A/ ) .C/g Detachment.
1.19 Apply exercises 1.15,
1.18, and Detachment:
` Œ.A/ ) .B/ ) f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C/ 1b,
` Œ.B/˚) .C/
) Œ.A/ ) .B/ ) f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C/ 1.12,
` fŒ.B/
n ) .C/ ) Œ.A/) .B/g o
) Œ.B/ ) .C/ ) f.A/ ) Œ.B/ ) .C/g ) Œ.A/ ) .C/
Œ.P/ ) .Q/ ) f.P/ ) Œ.Q/ ) .R/g ) Œ.P/ ) .R/g :
www.pdfgrip.com
1.41
fŒ.B/ ) .F/ ) Œ.A/ ) .F/g ) fŒ.B/ ) .F/ ) .A/g ) fŒ.B/ ) .F/ ) .F/g axiom C2;
fŒ.B/ ) .F/ ) .F/g ) .B/ axiom C3;
fŒ.B/ ) .F/ ) Œ.A/ ) .F/g ) fŒ.B/ ) .F/ ) .A/g ) .B/ transitivity;
.A/ ) fŒ.B/ ) .F/ ) .A/g C1;
fŒ.B/ ) .F/ ) Œ.A/ ) .F/g ) Œ.A/ ) .B/ transitivity:
1.43 Substitutions of :.B/ for .B/ ) .F/ and :.A/ for .A/ ) .F/ into the solution
of exercise 1.41 transform Church’s third axiom into fŒ:.Q/ ) Œ:.P/g )
Œ.P/ ) .Q/. Consequently, all three axioms of classical logic remain valid in
Church’s logic, and, therefore, Church’s logic allows for proofs of all the theorems
of classical logic.
1.45 For the first theorem,
1.49
fŒ.P/ ) .Q/ ) .R/g ) fŒ.R/ ) .P/ ) .P/g
m definition of _,
: :fŒ:.P/ _ .Q/g _ .R/g _ :fŒ:.R/ _ .P/g _ .P/
m de Morgan’s second law,
˚
: :fŒ:.P/ _ .Q/g ^ Œ:.R/ _ f:Œ:.R/g ^ Œ:.P/ _ .P/
m double negations,
fŒ:.P/ _ .Q/g ^ Œ:.R/ _ f.R/ ^ Œ:.P/g _ .P/
m distributivity,
fŒ:.P/ ^ Œ:.R/g _ f.Q/ ^ Œ:.R/g _ Œ.R/ _ .P/ ^ fŒ:.P/ _ .P/g
m excluded middle,
˚
fŒ:.P/ ^ Œ:.R/g _ f.Q/ ^ Œ:.R/g _ Œ.R/ _ .P/ ^ .T/
m identity,
fŒ:.P/ ^ Œ:.R/g _ f.Q/ ^ Œ:.R/g _ Œ.R/ _ .P/
m de Morgan’s second law,
f:Œ.R/ _ .P/g _ f.Q/ ^ Œ:.R/g _ Œ.R/ _ .P/
m commutativity,
associativity,
f:Œ.R/ _ .P/g _ Œ.R/ _ .P/ _ f.Q/ ^ Œ:.R/g
m excluded middle,
.T/ _ f.Q/ ^ Œ:.R/g
m identity,
.T/
www.pdfgrip.com
1.51
1.53 One implication is axiom P1 and the other implication is Pierce’s law.
1.55 Œ.P/ ) .Q/ , :f.P/ ^ Œ:.Q/g :
.P/ ) .Q/
m double negations;
.P/ ) f:Œ:.Q/g
m double negations;
: : .P/ ) f:Œ:.Q/g
m
definition of ^;
:f.P/ ^ Œ:.Q/g
1.57
1.59 No, the suggested rule fails if U and W are False but V is True.
1.61
` Œ.P/ )
.S/ ^ fŒ:.Q/ ) Œ:.S/g
) f.P/ ^ Œ:.Q/g ) f.S/ ^ Œ:.S/g theorem,
` f.P/
^ Œ:.Q/g ) f.S/
^ Œ:.S/g )
:f.S/ ^ Œ:.S/g ) :f.P/ ^ Œ:.Q/g contraposition,
www.pdfgrip.com
` :f.S/
^ Œ:.S/g ) :f.P/ ^ Œ:.Q/g
) fŒ:.S/ _ .S/g ) Œ.P/ ) .Q/ de Morgan, definition of ^,
1.65 Use fŒ:.R/ ) .F/g ) fŒ:.F/ ) .R/g and Œ.T/ ) .R/ ) .R/.
1.67 Definition 1.51 of .A/ , .B/ as Œ.A/ ) .B/^Œ.B/ ) .A/ with theorems 1.57
Œ.P/ ^ .Q/ ) Œ.Q/ ^ .P/ and 1.52 Œ.P/ ^ .Q/ ) .Q/ yield Tarski’s axiom IV.
1.69 Theorem 1.61 gives a derivation of .I/ , .J/. from .I/ ) .J/ and .J/ ) .I/.
The Deduction Theorem (1.22) then yields a proof of Œ.I/ ) .J/ ) fŒ.J/ )
.I/ ) Œ.I/ , .J/g.
1.71 Subsections 1.3.10, 1.3.11, and 1.3.12 show that axioms P1 and P2 are
derivable from Tarski’s axioms I–III. Moreover, axiom P3 coincides with Tarski’s
axioms VII. Consequently, axioms P1, P2, and P3 are derivable from Tarski’s
axioms I–VII. Therefore, every theorem of the Classical Propositional Calculus is
also derivable from Tarski’s axioms I–VII.
1.73 Substitute R for S in theorem 1.82, which gives fŒ.P/ ) .Q/ ^ Œ.R/ )
.R/g ) fŒ.P/ ^ .R/ ) Œ.Q/ ^ .R/g. Then apply the reflexivity of the
logical implication (theorem 1.14), the law of contraposition (theorem 1.44), and
transitivity, to derive Rosser’s axiom R3 from axioms P1, P2, and P3.
1.75 Kleene’s axiom 7 is the law of reductio ad absurdum (theorem 1.48).
1.77 fŒ.P/ ) .Q/ ) .P/g ) .P/ is a triadic tautology in Łukasiewicz’s triadic
logic.
1.79 fŒ.P/ ) .Q/ ) .P/g ) .P/ is not a triadic tautology in Church’s system or
in Łukasiewicz’s.
1.81 fŒ.P/ ) .Q/ ) .Q/g ) Œ.Q/ ) .P/ ) .P/g is a triadic tautology in
Łukasiewicz’s system, but not in Church’s.
1.83 f.P/ ) Œ.Q/ ) .R/g ) fŒ.P/ ) .Q/ ) Œ.P/ ) .R/g is a triadic
tautology in Church’s system, but not in Łukasiewicz’s.
www.pdfgrip.com
1.85 f.P/ ) Œ.Q/ ) .R/g ) f.Q/ ) Œ.P/ ) .R/g is a triadic tautology in both
Church’s system, and in Łukasiewicz’s.
1.87
1.89 In the proof of fŒ:.P/ ) .P/g ` .P/, the first step lists the hypothesis H, here
` fŒ:.P/ ) .P/g. The Deduction Theorem replaces this step, ` H, by a complete
proof of .H/ ) .H/, from theorem 1.14, with Œ:.P/ ) .P/ substituted for H
everywhere. The second step invokes theorem 1.14, which the Deduction Theorem
replaces by a complete proof of theorem 1.14. The third step uses the deduction rule
which the Deduction Theorem replaces by a complete proof of this deduction rule.
m
.Q/ ) fŒ.P/ ) .Q/ ) Œ.Q/ ) .R/g ) .R/ :
„ƒ‚… „ ƒ‚ … „ƒ‚…
K H L
To prove either formula with the Deduction Theorem, this proof starts by assuming
that the hypotheses H and K, here Q and Œ.P/ ) .Q/ ) Œ.Q/ ) .R/, are True.
` Œ.P/ ) .Q/ ) Œ.Q/ ) .R/
` .Q/ ) fŒ.P/ ) .Q/ ) Œ.Q/ ) .R/g ) f.Q/ ) Œ.P/ ) .Q/g ) f.Q/ ) Œ.Q/ ) .R/g
` f.Q/ ) Œ.P/ ) .Q/g ) f.Q/ ) Œ.Q/ ) .R/g
` .Q/ ) Œ.P/ ) .Q/
` .Q/ ) Œ.Q/ ) .R/
`Q
` .Q/ ) .R/
`Q
`R
The result then follows from the Deduction Theorem.
1.97 The formula S defined by fŒ:.P/ ) .P/g ) .P/ has only one propositional
variable, P. Moreover, S has the form .V/ ) .W/, with Œ:.P/ ) .P/ for V,
www.pdfgrip.com
and P for W. Thus the first step consists in applying the Provability Theorem
(theorem 1.125) to prove P0 ` S0 .
has the form .V/ ) .W/, with .P/ ) .Q/ for V, and f.P/ ) Œ:.Q/g ) Œ:.P/
for W.
P True, Q True If P is True and Q is True, then W is True. However, W has the
form .H/ ) .K/, with .P/ ) Œ:.Q/ for H, which is False, and :.P/ for K,
which is also False. Hence the Provability Theorem calls for a proof of P0 ; Q0 `
H 0 , here P; Q ` :.H/. Because H has the form .P/ ) Œ:.Q/, and hence :.H/
has the form .P/ ^ f:Œ:.Q/g, the Provability Theorem calls for proofs of P; Q `
P and P; Q ` Q, which follow from substitutions in the proof of theorem 1.14:
From the proofs of .P/ ) Œ.Q/ ) .S/ and .P/ ) fŒ:.Q/ ) .S/g follows a proof
of .P/ ) .S/, and thence from the proof of Œ:.P/ ) .S/ follows a proof of S.
1.101 The propositional form fŒ.P/ ) .Q/ ) .R/g ) fŒ.R/ ) .P/ ) .P/g has
the form .V/ ) .W/.
P True If P is True, then axiom P1 gives a proof of .P/ ) .W/, and hence a proof
of .P/ ) .S/.
P False, R False If P is False, then .P/ ) .Q/ holds, by the law of denial of the
antecedent (theorem 1.40):
If R is also False, then Œ:.R/ ) Œ:.R/, whence Œ.P/ ) .Q/ ^ Œ:.R/ holds,
whence also :fŒ.P/ ) .Q/ ) .R/g, which is :.V/.
P False, R True If P is False and R is True, then Œ.R/ ) .P/ is False, but so is P,
whence Œ.R/ ) .P/ ) .P/, which is W, is True.
1.103 The propositional form fŒ.P/ ) .R/ ) .Q/g ) .Q/ ) fŒ.Q/ ) .R/ )
Œ.P/ ) .R/g has the form .V/ ) .W/.
P True If P is True, then axiom P1 gives a proof of .P/ ) Œ.S/ ) .P/, whence
a proof of .P/ ) .W/, and hence a proof of .P/ ) .U/.
P False, S False If S is False, then the law of denial of the antecedent, Œ:.S/ )
Œ.S/ ) .P/ gives a proof of Œ:.S/ ) .W/ and hence of Œ:.S/ ) .U/.
P False, S True, R False With P False, S True, R False, the law of denial of the
antecedent gives Œ:.R/ ) Œ.R/ ) .Q/, while .S/ ) .P/ is False, whence V is
False, and then the law of denial of the antecedent gives .V/ ) .W/, which is U.
P False, S True, R True With P False, S True, R True, .R/ ) .P/ is False, whence
the law of denial of the antecedent gives Œ.R/ ) .P/ ) Œ.S/ ) .P/, which is
W; hence the law of denial of the antecedent gives .V/ ) .W/, which is U.
P False, S True The foregoing two cases give a proof of Œ:.P/ ) Œ.S/ ) .U/.
P False The proofs of Œ:.P/ ) Œ.S/ ) .U/ and Œ:.P/ ) fŒ:.S/ ) .U/g
then combine into a proof of Œ:.P/ ) .U/.
Finally, the proofs of .P/ ) .U/ and Œ:.P/ ) .U/ combine into a proof of U.
2.15 Axiom Q4 follows from the abbreviation :f8XŒ:.P/g, double negation and
theorem 2.45:
` Œ8X.P/ , f8XŒ::.P/g
theorem 2.45,
` f:Œ8X.P/g , :f8XŒ::.P/g
contraposition and its converse,
` f9XŒ:.P/g , : 8Xf:Œ:.P/g abbreviation, .R/ , .R/,
` f9XŒ:.P/g , f:Œ8X.P/g transitivity.
2.17 Kleene’s 9-rule is derivable from the rules of inference with axioms Q1– Q4
and the propositional calculus.
` .P/ ) .Q/ hypothesis,
` Œ.P/ ) .Q/ ) fŒ:.Q/ ) Œ:.P/g contraposition,
` Œ:.Q/ ) Œ:.P/ Detachment,
` 8XfŒ:.Q/ ) Œ:.P/g Generalization,
` Œ:.Q/ ) f8XŒ:.P/g theorem 2.29,
` f8XŒ:.P/g ) f:Œ9X.P/g axiom Q3,
` Œ:.Q/ ) f:Œ9X.P/g transitivity,
` Œ9X.P/ ) .Q/ converse contraposition
and Detachment.
2.19 Kleene’s 9-schema is derivable from the rules of inference with axioms
Q1– Q4 and the propositional calculus.
` f8XŒ:.P/g ) fSubfXY Œ:.P/g axiom Q1,
` fSubfXY Œ:.P/g ) f:ŒSubfXY .P/g remark 2.20,
` f8XŒ:.P/g ) f:ŒSubfXY .P/g Detachment,
` ŒSubf
Y .P/ ) :f8XŒ:.P/g
X
contraposition and double negation,
` :fŒ8XŒ:.P/g ) Œ9X.P/ theorem 2.46,
` ŒSubfXY .P/ ) Œ9X.P/ transitivity.
2.21
` .V/ , .U/ hypothesis,
` Œ.V/ , .U/ ) Œ.V/ ) .U/ theorem 1.62,
` .V/ ) .U/ Detachment,
` Œ.V/ ) .U/ ) fŒ.U/ ) .W/ ) Œ.V/ ) .W/g transitivity
(theorem 1.27),
` Œ.U/ ) .W/ ) Œ.V/ ) .W/ Detachment.
www.pdfgrip.com
2.23
` .V/ , .U/ hypothesis,
` Œ.V/ , .U/ ) Œ.U/ ) .V/ theorem 1.62,
` .U/ ) .V/ Detachment,
` Œ.U/ ) .V/ ) fŒ.V/ ) .W/ ) Œ.U/ ) .W/g transitivity
(theorem 1.28),
` Œ.V/ ) .W/ ) Œ.U/ ) .W/ Detachment.
2.25 Theorem 2.45 shows that if ` .V/ , .U/, then ` .P/ ) .Q/.
2.27 If ` .V/ , .U/, then ` Œ:.V/ , Œ:.U/, by contraposition and
transposition.
2.31
2.33
2.35
8X fŒ.P/ ^ .Q/ _ .R/g ,
8X fŒ.P/ _ .R/ ^ Œ.Q/ _ .R/g ,
f8XŒ.P/ _ .R/g ^ f8XŒ.Q/ _ .R/g
2.37
Œ9X.Q/ , 9Xf:Œ:.Q/g
, :f8XŒ:.Q/g
, :8X f8XŒ:.Q/g
, :8X f: Œ9X.Q/g
, : : f9X Œ9X.Q/g
, f9XŒ9X.Q/g
2.41 Invoke the reflexivity of the logical implication (theorem 1.63): by definition
of R,
ŒR.A; B/ , 8XfŒE .X; A/ ) ŒE .X; A/g ;
which is in prenex form, and its matrix is an instance of theorem 1.63: .P/ ) .P/.
2.43 Invoke the transitivity of the logical implication (theorem 1.65):
which is in prenex form, and its matrix is an instance of theorem 1.63: .P/ ) .P/.
2.47 The equality predicate I defined as in example 2.85 for set theory is
reflexive, as proved by the solutions to exercises 2.41 and 2.45, so that I satisfies
axioms J 1. Also, the set theory described in example 2.85 has only one predicate,
E , and formula (2.2) shows that I satisfies axiom J 2.
2.49 Axiom J 1 from subsection 2.5.3 coincides with axiom I 1 from subsec-
tion 2.5.1. Axiom J 2 from subsection 2.5.3 is a special case of axiom I 2 from
subsection 2.5.1, which allows for atomic formulae as particular cases of P and Q.
3.5 By definition of the empty set (axiom S2), the formula :.X 2 ¿/ is universally
valid. By specialization, with ¿ substituted for X, it follows that :.¿ 2 ¿/ is True,
whence ¿ 2 ¿ is False, by definition of False.
3.7 Use substitutions in the axiom of extensionality and the definition of equality.
3.9 For each set S, ¿ S, by theorem 3.11. Consequently, ¿ is a subset of ¿.
Conversely, if S is a subset of ¿, so that S ¿, then S D ¿ by theorem 3.10:
`S¿ hypothesis on S,
`¿S theorem 3.11,
`SD¿ theorem 3.10.
3.11 This proof proceeds by contraposition, showing that if S 6D ¿, then S has a
subset different from S. Because ¿ S by theorem 3.11, it follows that if S 6D ¿,
then S has a subset, ¿, different from S.
3.13 If S is a subset of every set, then S is a subset of the empty set: S ¿.
Moreover, ¿ S by theorem 3.11. Consequently, S D ¿, by theorem 3.10.
3.15 If A ¨ B and B ¨ C, then A B and B C, whence A C, by theorem 3.9.
However, because A ¨ B, there exists some Z 2 B such that Z … A. Consequently,
Z 2 C but Z … A, so that A 6D C, whence A ¨ C.
3.17 This proof establishes each implication () and () independently.
For one implication, assume ` 8YŒ.A Y/ , .B Y/.
` 8YŒ.A Y/ , .B Y/ hypothesis,
+ SubfYA ,
` .A A/ , .B A/
+ ),
` .A A/ ) .B A/
`AA theorem 3.8,
`BA Modus Ponens;
`AY hypothesis,
` 8XŒ.X 2 A/ ) .X 2 Y/ definition of subsets,
`BA hypothesis,
` 8XŒ.X 2 B/ ) .X 2 Y/ transitivity of ),
`BY definition of subsets.
Swapping A and B yields the converse. Hence ` 8YŒ.A Y/ , .B Y/.
3.19 If C D and D W, then D C and W D, whence W C. For the
converse, let W WD D, whence D D, the hypothesis .D D/ ) .C D/, and
Modus Ponens yield C D.
˚
3.21 Let X WD ¿, Y WD f¿g, Z WD f¿g .
˚
3.23 Let X WD ¿ and Y WD f¿g .
˚
3.25 Let X WD f¿g and A WD f¿g .
3.27 .X 2 f¿g/ , .X D ¿/ whereas .X 2 f f¿g g/ , .X D f¿g/. Yet ¿ 6D f¿g.
Consequently, f¿g and f f¿g g have different elements. Therefore f¿g 6D f f¿g g.
3.29 f¿g 2 f ¿; f¿g g. Yet .X 2 f¿g/ , .X D ¿/ and ¿ 6D f¿g, whence
f¿g … f¿g. Hence f¿g 6D f ¿; f¿g g.
˚ ˚ ˚
3.31 From 8S.S S/ specialization with S WD f¿g gives f¿g f¿g .
˚
3.33
˚ The set f¿g ˚ has only
˚one element f¿g, which is also an element of the set
¿; f¿g . Hence f¿g ¿; f¿g .
3.35 For theorem 3.13, the word “and” in the informal proof corresponds to
the logical connective _ in the formal proof through the universal quantifier,
specializations with H first and then K, and the axioms of extensionality and pairing,
along the following outline:
` .X D H/ ) Œ.X D H/ _ .X D K/ .P/ ) Œ.P/ _ .Q/,
` Œ.X D H/ _ .X D K/ ) .X 2 L/ pairing,
` .X D H/ ) .X 2 L/ transitivity,
` .X D H/ ) Œ.X 2 L/ ) .H 2 L/ extensionality,
` .X D H/ ) .H 2 L/ transitivity,
` .H D H/ ) .H 2 L/ specialization SubfXH ,
`HDH extensionality,
`H2L Modus Ponens;
`K2L as for H;
` .H 2 L/ ^ .K 2 L/ .P/ ) f.Q/ ) Œ.P/ ^ .Q/g.
3.37 If A D B, and if S A, then each element of S is also an element of B, whence
S B. Thus P.A/ P.B/, and conversely with the rôles of A and B switched.
If P.A/ D P.B/, and if X 2 A, then fXg 2 P.A/, whence fXg 2 P.B/, so that
fXg B, and hence X 2 B. Thus A B, and conversely B A with the rôles of A
and B switched. Therefore A D B.
www.pdfgrip.com
˚ ˚ ˚
3.39 f¿g \ ¿; f¿g D f¿g .
˚ ˚ ˚ ˚
3.41 f¿g [ ¿; f¿g D ¿; f¿g in the superset ¿; f¿g
3.43 8SŒ.S \ ¿/ D ¿.
A WD ¿;
B WD f¿g;
A n B D ¿ n f¿g
D ¿
6D f¿g
D f¿g n ¿
D B n A:
X WD f¿g;
˚
A WD f¿g I
˚
Y WD f¿g ;
n˚ o
B WD f¿g I
˚
X [ Y D ¿; f¿g ;
n ˚ o
A [ B D f¿g; f¿g :
www.pdfgrip.com
3.65
[
X2 fAg , f9YŒ.Y 2 fAg/ ^ .X 2 Y/g
, f9YŒ.Y D A/ ^ .X 2 Y/g
, Œ9Y.X 2 A/
, X 2 A:
3.67 Use the tautology fŒ.P/ _ .Q/ ^ .R/g , fŒ.P/ ^ .R/ _ Œ.Q/ ^ .R/g.
3.69 Use the tautology Œ.P/ _ .F/ , .P/:
3.75 Use the tautology .B/ _ Œ:.B/ and the contradiction .B/ ^ Œ:.B/, with X 2 U
for B:
fX 2 Œ.A n B/ n Ug
, fŒX 2 .A n B/ ^ Œ:.X 2 U/g
˚
, .X 2 A/ ^ f:Œ.X 2 B/g ^ Œ:.X 2 U/
, f.X 2 A/ ^ Œ:.X 2 U/g ^ Œ:.X 2 B/
˚
, ŒX 2 .A n U/ ^ Œ:.X 2 B/ _ f.X 2 U/ ^ Œ:.X 2 U/g
, f.X 2 A/ ^ Œ:.X 2 U/g ^ fŒ:.X 2 B/ _ .X 2 U/g ^ fŒ:.X 2 B/ _ Œ:.X 2 U/g
, f.X 2 A/ ^ Œ:.X 2 U/g ^ fŒ:.X 2 B/g ^ fŒ:.X 2 B/ _ Œ:.X 2 U/g
, f.X 2 A/ ^ Œ:.X 2 U/g ^ fŒ:.X 2 B/ _ Œ:.X 2 U/g
, f.X 2 A/ ^ Œ:.X 2 U/g ^ f:f.X 2 B/ ^ Œ:.X 2 U/g
, ŒX 2 .A n U/ ^ f:ŒX 2 .B n U/g
, fX 2 Œ.A n U/ n .B n U/g:
www.pdfgrip.com
3.77
3.79
fX 2 ŒU \ .A n B/g
, f.X 2 U/ ^ ŒX 2 .A n B/g
, Œ.X 2 U/ ^ f.X 2 A/ ^ Œ:.X 2 B/g
, fŒ.X 2 U/ ^ Œ:.X 2 B/g ^ .X 2 A/
˚
, fŒ.X 2 U/ ^ Œ:.X 2 B/g ^ .X 2 A/ _ fŒ.X 2 U/ ^ Œ:.X 2 B/g ^ Œ:.X 2 U/
, fŒ.X 2 U/ ^ Œ:.X 2 B/g ^ f.X 2 A/ ^ Œ:.X 2 U/g
, fŒ.X 2 U/ ^ Œ:.X 2 B/g ^ :fŒ:.X 2 A/ _ .X 2 U/g
, fX 2 Œ.U n B/ n .U n A/g:
A WD f¿g;
˚
B WD f¿g ;
˚
A [ B WD ¿; f¿g ;
C WD A [ B;
C 2 P.A [ B/;
C … P.A/;
C … P.B/;
C … P.A/ [ P.B/:
3.87
S S
. F / \ B D A2F .A \ B/
m ˚
S S
8X ŒX 2 . F / \ B , X 2 A2F .A \ B/
m
S
8X .fŒX 2 . F / ^ .X 2 B/g , Œ9A f.A 2 F / ^ ŒX 2 .A \ B/g/
m
8X .fŒ9A.A 2 F / ^ .X 2 A/ ^ .X 2 B/g
, Œ9A f.A 2 F / ^ Œ.X 2 A/ ^ .X 2 B/g/
3.99 No, Œ.A [ C/.B [ C/ 6 Œ.AB/ [ C, because the left-hand side does not
contain .A [ C/ \ .B [ C/ while the right-hand side contains all of C.
3.101 Yes, Œ.A \ C/.B \ C/ Œ.AB/ \ C.
3.103 Yes, Œ.A n C/.B n C/ Œ.AB/ n C.
3.105 No, Œ.C n A/.C n B/ 6 ŒC n .AB/, because the left-hand side does not
contain C n .A [ B/ while the right-hand side contains all of C n .A [ B/.
3.107 No, ŒP.A/P.B/ 6 ŒP.AB/, because the left-hand side does not
contain ¿ 2 P.A/ \ P.B/ while the right-hand side contains ¿ 2 P.AB/.
3.109
3.111 As defined here the Cartesian product of sets is not associative (but a slightly
different version of the Cartesian product is associative). For example, .A B/ C 6D
A .B C/ for the sets A D f0g, B D f1g, and C D f2g, because
which reveals that f0g 2 .0; .1; 2// 2 A .B C/ whereas f0g … ..0; 1/; 2/ 2
.A B/ C. Consequently, .0; .1; 2// 6D ..0; 1/; 2/ by extensionality, and then
.A B/ C 6D A .B C/ again by extensionality. (A slightly different definition of
the Cartesian product through functions of integers will later provide an associative
Cartesian product.)
3.113 (The last step still requires further symbolic substeps.)
.A D ¿/ _ .B D ¿/
m extensionality,
f:Œ9X.X 2 A/g _ f:Œ9Y.Y 2 B/g
m de Morgan’s Law,
:fŒ9X.X 2 A/ ^ Œ9Y.Y 2 B/g
m definition of A B,
:f9X9YŒ.X; Y/ 2 .A B/g
m uniqueness of ¿.
.A B/ D ¿
3.115 The Cartesian product distributes over unions: for all sets A, B, and C,
An informal proof can establish that Œ.A B/ [ .C B/ and Œ.A [ C/ .B/ have
exactly the same elements. These two sets are Cartesian products, and, consequently
their elements are ordered pairs.
• An ordered pair .X; Y/ is an element of .A B/ [ .C B/ if and only if .X; Y/ 2
.A B/ or .X; Y/ 2 .C B/;
• hence .X; Y/ 2 Œ.A B/ [ .C B/ if and only if X 2 A and Y 2 B, or X 2 C and
Y 2 B,
• which is equivalent to X 2 A or X 2 C, and Y 2 B;
• thus .X; Y/ 2 Œ.A B/ [ .C B/ if and only if X 2 .A [ C/ and Y 2 B, which
is equivalent to .X; Y/ 2 Œ.A [ C/ B.
Just as the preceding informal proof concatenated the two occurrences of Y 2 B into
one such occurrence, a formal proof can rely on the distributivity of ^ over _ by the
tautology
.X; Y/ 2 Œ.A B/ [ .C
B/
m definition of union,
Œ.X; Y/ 2 .A B/ _ Œ.X; Y/ 2 .C B/
m definition of Cartesian products,
Œ.X 2 A/ ^ .Y 2 B/ _ Œ.X 2 C/ ^ .Y 2 B/
m distributivity of ^ over _,
Œ.X 2 A/ _ .X 2 C/ ^ .Y 2 B/
m definition of union,
ŒX 2 .A [ C/ ^ Œ.Y 2 B/
m definition of Cartesian products.
.X; Y/ 2 Œ.A [ C/ B
3.117
3.119 No, Œ.A n C/ .B n D/ 6 Œ.A B/ n .C D/, because the right-hand side
contains all of .A n C/ B.
3.121 No, Œ.AC/ .BD/ 6 Œ.A B/.C D/. For instance, if B D D, then
BD D ¿, whence .AC/ .BD/ D ¿ on the left-hand side. Yet on the right-
hand side, still with B D D, if C D ¿, then C D D ¿, whence .A B/.C D/ D
.A B/.
3.123 No, .ŒP.A/ ŒP.B// 6 P.A B/, because the left-hand side consists of
pairs of subsets of A and B, whereas the right-hand side consists of subsets of A B.
For instance, if A D ¿ D B, then A B D ¿, whence P.A B/ D P.¿/ D f¿g,
whereas ŒP.A/ ŒP.B/ D ŒP.¿/ ŒP.¿/ D f¿g f¿g D f.¿; ¿/g.
3.125 If S denotes the relation of strict inclusion on A WD P.H/, then
R WD f¿g;
˚
S WD f¿g ;
P
A WD R[S;
B WD f¿g;
F W A ! B;
˚
F WD .¿; ¿/; .f¿g; ¿/ :
www.pdfgrip.com
Then
R \ S D ¿;
F".R \ S/ D F".¿/ D ¿;
F".R/ D F".f¿g/ D fF.¿/g D f¿g D B;
˚ ˚
F".S/ D F" f¿g D F.f¿g/ D f¿g D B;
F".R/ \ F".S/ D B \ B D B 6D F".R \ S/:
3.145 For all sets and for every function, ŒF".K/ n ŒF".H/ F".K n H/.
3.147 For each X 2 A, from fXg A it follows that F".fXg/ B, with F".fXg/ D
fF.X/g by definition of F". Consequently, X 2 A, the pair fXg; F".fXg/ 2
for each
P.A/ P.B/ corresponds to the pair X; F.X/ 2 F A B.
˚ ˚
3.149 F" ˚¿; f¿g D F.¿/; F.f¿g/
˚ D fF.0/; F.1/g D f2g.
Fı˚ 1" ¿; f¿g D Fı 1" 0; 1 D f2g.
F ¿; f¿g
n˚ D F.2/ D 0 D ¿.
o
Fı 1" ¿; f¿g D Fı 1".f2g/ D f0; 1g.
A WD f0; 1g;
B WD f0; 1; 2g;
F WD f.0; 0/; .1; 1/g;
www.pdfgrip.com
Then G ı F D IA , and H ı F D IA .
3.161 The empty relation is vacuously reflexive, symmetric, and transitive because
.X; Y/ 2 ¿ is False>
Reflexivity:
m definition of ¤,
m definitions of and D,
m: Œ9Y.P/ , f:8YŒ:.P/g
m de Morgan’s Law,
m substitutions,
False
Consequently, with .V ¤ W/ ^ .W ¤ V/ False, the implication
Œ.V ¤ W/ ^ .W ¤ V/ ) .V D W/
is True.
3.165 Here are the equivalence classes:
8 9
ˆ
ˆ .1; 7/ .3; 7/ .5; 7/ .7; 7/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 6/ .2; 6/ .4; 6/ .6; 6/ >
>
ˆ
ˆ >
ˆ
ˆ
ˆ .1; 5/ .3; 5/ .5; 5/ .7; 5/ >
>
>
>
< =
.0; 4/ .2; 4/ .4; 4/ .6; 4/
RF D :
ˆ
ˆ .1; 3/ .3; 3/ .5; 3/ .7; 3/ >
>
ˆ
ˆ >
>
ˆ
ˆ .0; 2/ .2; 2/ .4; 2/ .6; 2/ >
>
ˆ
ˆ >
>
ˆ
ˆ .1; 1/ .3; 1/ .5; 1/ .7; 1/ >
>
:̂ >
;
.0; 0/ .2; 0/ .4; 0/ .6; 0/
3.169 Outline: For each equivalence relation R, the partition FR consists of all the
equivalence classes corresponding to R. The relation RFR then consists of all the
pairs from all the equivalence classes of R, which is then R again.
3.171 The relation Q is
antisymmetric, vacuously, because it does not contain any two pairs .X; Y/ and
.Y; X/,
asymmetric, because it does not contain any two pairs .X; Y/ and .Y; X/,
not connected, because .2; 9/ … Q and .9; 2/ … Q,
irreflexive, because it does not contain any element on the diagonal,
not reflexive, because .0; 0/ … Q,
not strongly connected, because .2; 9/ … Q and .9; 2/ … Q,
not symmetric, because .0; 1/ 2 Q but .1; 0/ … Q,
not transitive, because .2; 6/ 2 Q and .6; 9/ 2 Q but .2; 9/ … Q.
3.173 The relation S is
not antisymmetric, because .0; 1/ 2 S and .1; 0/ 2 S but 0 6D 1,
not asymmetric, because it does not contain any two pairs .0; 1/ 2 S and
.1; 0/ 2 S ,
not connected, because .8; 9/ … S and .9; 8/ … S ,
irreflexive, because it does not contain any element on the diagonal,
not reflexive, because .0; 0/ … S ,
not strongly connected, because .8; 9/ … S and .9; 8/ … S ,
symmetric, because if .X; Y/ 2 S then .Y; X/ … S ,
not transitive, because .9; 6/ 2 S and .6; 8/ 2 S but .9; 8/ … S .
3.175 The relation V is
not antisymmetric, .0; 1/ 2 V and .1; 0/ 2 V but 0 6D 1,
not asymmetric, because it does not contain any two pairs .0; 1/ … V and
.1; 0/ … V ,
not connected, because .0; 2/ … V and .2; 0/ … V ,
not irreflexive, because it contains at least one element on the diagonal,
not reflexive, because .0; 0/ … V ,
not strongly connected, because .0; 2/ … V and .2; 0/ … V ,
not symmetric, because .1; 2/ 2 V but .2; 1/ … V ,
not transitive, because .0; 1/ 2 V and .1; 2/ 2 V but .0; 2/ … V .
www.pdfgrip.com
4.1
( )
˚ n ˚ o ˚ n ˚ o
5D ¿; f¿g; ¿;f¿g ; ¿;f¿g; ¿;f¿g ; ¿; f¿g; ¿;f¿g ; ¿;f¿g; ¿;f¿g
˚
but ¿ … f¿g; f¿g whence
˚
f¿g 6 f¿g; f¿g ;
˚ ˚
.A [ fAg/ \ .B [ fBg/ D .¿ [ f¿g/ \ f¿g [ f¿g D ¿;
˚ ˚
.A \ B/ [ fA \ Bg D ¿ \ f¿g [ ¿ \ f¿g D f¿g;
Initial Step
Induction hypothesis
Assume that there exists N 2 N such that the theorem holds for M WD N, so that for
all K; L; 2 N, if K 2 N and L 2 N, then K [ L 2 N.
Induction step
For all K; L; 2 N, if K 2 .N [ fNg/ and L 2 .N [ fNg/, then only four cases can
arise.
If K 2 N and L 2 N, then .K[L/ 2 N by induction hypothesis; from N N[fNg
it then follows that .K [ L/ 2 .N [ fNg/.
www.pdfgrip.com
m definition of F ,
ŒH 2 P.H/
^ .¿ 2 H/ ^ 8Xf.X 2 H/ ) Œ.X [ fXg/ 2 Hg
which is True by the definition of H and the Axiom of Infinity.
4.13
8Xf.X 2 N/ ) Œ.X [ fXg/ 2 Ng yet unproved,
m definition of N,
T T
8Xf.X 2 F / ) Œ.X [ fXg/ 2 F g
T
m definition of ,
8X f8BŒ.B 2 F / ) .X 2 B/g
) 8Bf.B 2 F / ) Œ.X [ fXg/ 2 Bg
*
8X 8B fŒ.B 2 F / ) .X 2 B/g
) f.B 2 F / ) Œ.X [ fXg/ 2 Bg
˚ * axiom P2,
8X 8B .B 2 F/
) f.X 2 B/ ) Œ.X [ fXg/ 2 Bg
which is True by the definition of F . Moreover, ¿ 2 N by theorem 4.8.
www.pdfgrip.com
4.15
`I2F theorem 4.4,
T
` .I 2 F / ) . F I/ theorem 3.48,
T
F I Modus Ponens;
T
` .S N/ , .S F/ definition of F ,
SI transitivity of ;
` .S N/ ^ .¿ 2 S/ ^ 8Xf.X 2 S/ ) Œ.X [ fXg/ 2 Sg hypothesis on S,
` .¿ 2 S/ ^ 8Xf.X 2 S/ ) Œ.X [ fXg/ 2 Sg Œ.P/ ^ .Q/ ) .Q/ and M.P.,
` .¿ 2 S/ ^ 8Xf.X 2 S/ ) Œ.X [ fXg/ 2 Sg ) .S 2 F / definition of F ,
`S2F M.P.
T
` .S 2 F / ) . F S/ theorem 3.48,
T
F S Modus Ponens;
T T T
F DS F S and S F .
`SDN definition of N.
T T
4.17 Because ¿ 2 N it follows that N ¿ whence N D ¿.
˚
P
4.19
˚ S WD ˚V [ ¿; f¿g
Let . This proof shows that S D N, whence V D S n
¿; f¿g D N n ¿; f¿g . ˚ ˚
For each˚ X 2 S, either X 2 ¿; f¿g or X … ¿; f¿g .
If X 2 ¿; f¿g , then either X D ¿ or X D f¿g.
In the first case (if X D ¿), then ¿ 2 S, by pairing and union. Moreover, ¿ [
f¿g D f¿g 2 V by hypothesis, whence ¿ [ f¿g ˚ 2 S. ˚
In the second case (if ˚X D f¿g),
then f¿g[ f¿g 2 V whence f¿g[ f¿g 2 S.
Otherwise (if X 2 Sn ¿; f¿g ), then X 2 V, whence X [fXg 2 V by hypothesis,
and hence X [ fXg 2 S by union.
Thus, X [ fXg 2 S for each X 2 S, and ¿ 2 S. Consequently, S D N by the
Principle of Mathematical Induction.
Therefore, V D S n f¿g D N n f¿g.
4.21 1 C 1 D 1 [ f1g by definition of M C 1 D M [ fMg, and 1 [ f1g D 2 by
definition of 2.
4.23 3 C 1 D 3 [ f3g by definition of M C 1 D M [ fMg, and 3 [ f3g D 4 by
definition of 4.
4.25 5 C 1 D 5 [ f5g by definition of M C 1 D M [ fMg, and 5 [ f5g D 6 by
definition of 6.
4.27 7 C 1 D 7 [ f7g by definition of M C 1 D M [ fMg, and 7 [ f7g D 8 by
definition of 8.
4.29 2C2 D 2C.1C1/ D .2C1/C1 by definition of M C.N C1/ D .M CN/C1,
and .2 C 1/ C 1 D 3 C 1 D 4 by the preceding exercises.
www.pdfgrip.com
( )
˚ n ˚ o ˚ n ˚ o
¿; f¿g; ¿;f¿g ; ¿;f¿g; ¿;f¿g ; ¿; f¿g; ¿;f¿g ; ¿;f¿g; ¿;f¿g
www.pdfgrip.com
.Œ.K; L/l C Œ.M; N/l / C Œ.P; Q/l D Œ.K C M; L C N/l C Œ.P; Q/l
D Œ..K C M/ C P; .L C N/ C Q/l
D Œ.K C .M C P/; L C .N C Q//l
WD Œ.K; L/l C Œ.N C M; Q C P/l
WD Œ.K; L/l C .Œ.P; Q/l C Œ.M; N/l / :
D Œ.K .M P C N Q/ C L .M Q C N P/;
K .M Q C N P/ C L .M P C N Q//l
D Œ..K P C L Q/ M C .K Q C L P/ N;
.K P C L Q/ N C .K Q C L P/ M/l
D Œ.K P C L Q C K M C L N; K Q C L P C K N C L Ml
1 .1 C 1/ D 1 2 D 1;
.1 1/ C .1 1/ D 0 C 0 D 0:
.I Q C J P/ .J Q/ D .I J/ .Q Q/ C .J J/ .P Q/ > 0
Alternatively, # P PfPŒP.¿/g D 16 because
P PfPŒP.¿/g
D P hn
PfPŒf¿gg
˚ ˚ oi
D P ¿; f¿g; f¿g ; ¿; f¿g
D ¿;
˚ n˚ o n˚ o
f¿g; f¿g ; f¿g ; ¿; f¿g ;
˚ n ˚ o n ˚ o
¿; f¿g ; ¿; f¿g ; ¿; ¿; f¿g ;
n ˚ o n ˚ o n˚ ˚ o
f¿g; f¿g ; f¿g; ¿; f¿g ; f¿g ; ¿; f¿g ;
n ˚ o n ˚ o
¿; f¿g; f¿g ; ¿; f¿g; ¿; f¿g ;
n ˚ ˚ o n ˚ ˚ o
¿; f¿g ; ¿; f¿g ; f¿g; f¿g ; ¿; f¿g ;
f0; 1; 2; 3; 4; 5; 6; 7; 8; 9; 10; 11; 12; 13; 14; 15g:
:f9Z8YŒ.Y Z/ _ .Y D Z/g
whence
In other words, for each Z 2 W there exists Y 2 W such that :.Y Z/ and Y 6D Z.
In particular, Y … WZ . In yet other words, there does not exist any Z 2 W such that
W D WZ . Because totally orders W, however, it follows
S that Z Y must hold.
Let G consists of all the initial intervals in W. Then SG D W, but there does not
exist any Z 2 W such that W D WZ . Consequently, G is not an initial interval
of W.
5.15 There is a transitive set on which 2 is not a transitive relation. For example,
the set
n ˚ o
A WD ¿; f¿g; f¿g
6.3 Theorem 6.6 shows that each finite set of nonempty sets has a choice function.
The proof of theorem 6.20 shows that each choice function corresponds to a choice
set.
6.5 The Choice-Set Principle 6.17 translates into formula (1):
S
6.7 For each set F of nonempty sets, define a relation R F F by R WD
f.A; X/ W .A 2 F / ^ .X 2 A/g. If the Choice-Relation Principle 6.15 holds, then
there exists a function F R with the same domain as that of R. Hence F is a choice
function. Thus the Choice-Function Principle 6.8 holds.
Conversely, for each relation R A B, and for each set X in the domain of R,
define the vertical section of R at X by RX WD fY 2 B W .X; Y/ 2 Rg. In particular,
RX 6D ¿ because X is in the domain of R. If the
SChoice-Function Principle 6.8 holds,
then there exists a choice function C W F ! F for F WD fRX W 9YŒ.X; Y/ 2 Rg.
Thus C.RX / 2 RX for every X. Hence parametrize the vertical sections of R by
S W A ! P.B/ with S.X/ WD RX , and set F WD C ı S. Thus F is a function,
with F.X/ D CŒS.X/ D C.RX / 2 RX , whence F R. Thus the Choice-Relation
Principle 6.15 holds.
6.9 In every partially ordered set A the empty subset ¿ A is a chain. If the
hypothesis of Zorn’s Maximal-Element Principle 6.32 holds, then A contains an
upper bound for ¿. Hence A 6D ¿ [88, p. 118].
6.11 Zorn’s Maximal-Set Principle 6.34 translates into formula (3):
˚ ˚
8F .F 6D ¿/ ^ 8G ŒG 2 P.F / ^ 8A8B fŒ.A 2 G / ^ .B 2 G /
[
) Œ.A B/ _ .B A/ ) G 2F (2)
!
˚
) 9S .S 2 F / ^ f8A Œ.A 2 F / ) .S 6 A/g :
www.pdfgrip.com
6.17
(T.A) ¿ 2 T by example 6.50; S
(T.B) if A T is linearly ordered by inclusion, and if U; W 2 A , then there
exist intervals H; K 2 A such that U 2 H and W 2 K, with H K or K H.
case, if H SK, then U; W 2 K. If also V 2 N and U V W, then
In the firstS
V 2 K A ; Thus A 2 T ; the second case is similar;
(T.C) if A 2 T , then either A D ¿ and A [ fF.A/g D f0g 2 T , or A 6D ¿ and
A [ fF .A/g D A 2 T , because F.A/ D min.A/ 2 A.
6.19
(T.A) ¿ 2 P.E/; S S
(T.B) if A P.E/ then A E whence A 2 P.E/;
(T.C) if A 2 P.E/, then A [ fF.A/g 2 P.E/, because F.A/ 2 E.
6.21 For each denumerable element D` of a denumerable family F D fDk k 2 Ng
there exists a bijection F W N ! D` . Thus the set B` of all such bijections is not
empty. Let B WD fB` ` 2 Ng. In the Zermelo-Frænkel-Choice
S set theory, there
exists a family choice-function F W N ! B such that F` WD F.`/ 2 B` for S each
` 2 N; thus each F` W N ! D` is a bijection. Hence the function N N ! B
with .k; `/ 7! F` .k/ is a bijection.
7.1 There are no dominant pure strategies in The Battle of the Sexes.
7.3 The two positions where they both go to the same show are Nash equilibria.
7.5 There are two-person games with a Nash equilibrium but without any dominant
strategy and hence without dominant strategy equilibrium, for instance, The Battle
of the Sexes (exercises 7.1 and 7.3).
7.7 For all x 2 A and y 2 B,
Consequently,
where the left-hand side depends only on x while the right-hand side depends only
on y. Therefore,
7.9 For each strategy SAl available to Al, Al identifies the worst payoff pAl
SAl ;SBo under
every strategy SBo available to Bo:
To avoid the worst of the worst, Al plays the strategy SAl that returns the best among
the worst payoffs:
7.11 If a third player Ci has a single weakly dominant strategy, then Ci will play
that strategy, leaving Al and Bo with a game for two players and any number of pure
strategies, which need not have any Nash equilibrium.
7.13 To avoid failure, the Blue commander must play Middle. Thus the Red
commander knows that Blue plays Middle; to avoid failure, Red must then play
either Left or Right. In either case, (M,L) or (M,R), both commanders get a Draw.
7.15 Each of the N 1 girls who rejected him did so because she received and holds
a better proposal. In particular, the last girl who rejected him received at least two
proposals. Consequently, there is still at least one girl c who has not yet received
any proposal. In the next round, b must then propose to c, which terminates the
algorithm, because all the other girls have received at least one proposal (from b)
and therefore hold one.
7.17 Depending on the beggars’ and choosers’ rankings, infinite sets need not admit
of any total stable relations. Let the beggars and choosers consist of all negative and
positive integers, indexed by their values:
Because beggar bm begs for m, every chooser c ranks all beggars with the same
well-order bk c b` if and only if k < `. In contrast, beggar bm ranks choosers
by swapping 1; : : : ; 2m with 2mC1; : : : ; 4m, and otherwise leaving the natural order
on ZC unchanged:
2m C 1; : : : ; 4m; 1; : : : ; 2m; 4m C 1; 4m C 2; : : :
www.pdfgrip.com
With such rankings, there are no total stable relations. The proof proceeds by
contraposition, showing that every total relation is unstable. Indeed, if a relation
M B C is total, then there exists k 2 ZC with .bk ; c1 / 2 M, preceded by
k 1 0 couples
Yet bk prefers 2k > k 1 choosers c2kC1 ; : : : ; c4k to c1 . Hence there exists n > k
such that .bn ; c`n / 2 M with 2k C 1 `n 4k, whence .bn ; c`n / ‰ .bk ; c1 /, so
that M is unstable.
7.19 Extend algorithm 7.31 (with a match maker) to stable relations that may relate
each chooser to more than one beggar, with a quota qc of beggars for chooser c
(polygamy, polyandry, schools admitting more than one student, etc.). See Gale &
Shapley’s references [40, 41].
7.21 For all .b1 ; c1 /; .b2 ; c2 / 2 UT there exist stable relations M1 ; M2 2 T with
.b1 ; c1 / 2 M1 and .b2 ; c2 / 2 M2 . Yet M1 M2 or M2 M1 because T is a chain.
Thus .b1 ; c1 /; .b2 ; c2 / 2 M WD maxfM1 ; M2 g 2 T , but M is stable, so that none of
the three conditions in definition 7.27 can hold. Therefore UT is stable.
7.25 From .X; X/ 2 R ı1 if and only if .X; X/ 2 R follows .X; X/ … R n R ı1 .
7.27 If .X; Y/; .Y; Z/ 2 PR D R n R ı1 R, then .X; Z/ 2 R, by transitivity
of R.
If also .X; Z/ 2 R ı1 , then .Z; X/ 2 R, but .X; Y/ 2 R by hypothesis, whence
.Z; Y/ 2 R by transitivity of R, so that .Y; Z/ 2 R ı1 , contradicting the hypothesis
that .Y; Z/ … R ı1 .
7.29 Combine the preceding solutions.
7.31 With at least two voters Val and Vic and at least two alternatives Al and Bo,
there is a voters’ profile r such that Val prefers Al to Bo while Vic prefers Bo to Al:
If Val is a dictator, so that F .s/ D s.Val/ for every voters’ profile s, then F .r/ D
r.Val/ D f.Al; Bo/g while F .r ı / D .r ı /.Val/ D f.Bo; Al/g. Thus F .r/ 6D
F .r ı /, so that F is not invariant under permutations.
7.33 Weak unanimity implies unanimity.
www.pdfgrip.com
References
1. L.V. Ahlfors, Complex Analysis, 3rd edn. (McGraw-Hill, New York, 1979). ISBN 0-07-
000657-1. QA331.A45 1979. LCCC No. 78-17078. 515’.93; MR0510197 (80c:30001)
2. Anonymous (ed.), ‘Three-parent’ therapy approved. Science 347(6222), 592 (2015). http://
scim.ag/_mtvote
3. V.I. Arnold, From Hilbert’s superposition problem to dynamical systems. Am. Math. Mon.
111(7), 608–624 (2004). ISSN 0002-9890
4. K.J. Arrow, A difficulty in the concept of social welfare. J. Polit. Econ. 58(4), 328–346 (1950)
5. I. Asimov, The Gods Themselves (Spectra, New York, 1990). ISBN-10: 0553288105. ISBN-
13: 978-0553288100
6. S. Barberá, Pivotal voters: a new proof of Arrow’s theorem. Econ. Lett. 6(1), 13–16 (1980)
7. S. Becattini1, D. Latorre1, F. Mele1, M. Foglierini1, C. De Gregorio1, A. Cassotta1, B.
Fernandez1, S. Kelderman, T.N. Schumacher, D. Corti1, A. Lanzavecchia1, F. Sallusto1,
Functional heterogeneity of human memory cd4+ t cell clones primed by pathogens or
vaccines. Science 347(6220), 400–406 (2015). doi:10.1126/science.1260668. http://www.
sciencemag.org/content/347/6220/400.long
8. P. Bernays, Axiomatic Set Theory with a Historical Introduction by Abraham A. Fraenkel
(Dover, Mineola, 1991). ISBN 0-486-66637-9. QA248.B47 1991. LCCC No. 90-25812.
511.3’22–dc20
9. G.D.W. Berry, Peirce’s contributions to the logic of statements and quantifiers, in Studies
in the Philosophy of Charles Sanders Peirce: First Series, ed. by P.P. Wiener, F.H. Young
(Harvard University Press, Cambridge, 1952), pp. 153–165
10. B. Bolzano, Paradoxien des Unendlichen, in Essays on the Theory of Numbers, ed. by R.
Dedekind (Dover, Leipzig, 1851)
11. R.T. Boute, How to calculate proofs: bridging the cultural divide. Not. Am. Math. Soc. 60(2),
173–191 (2013)
12. S.J. Brams, P.D. Straffin Jr., Prisoner’s dilemma and professional sports drafts. Am. Math.
Mon. 86(2), 80–88 (1979)
13. S.J. Brams, M.D. Davis, P.D. Straffin Jr., Module 311: Geometry of the Arm Race (COMAP,
Lexington, 1978)
14. W.L. Briggs, V.E. Henson, The DFT: An Owner’s Manual for the Discrete Fourier Transform
(Society for Industrial and Applied Mathematics, Philadelphia, 1995). ISBN 0-89871-342-0;
QA403.5.B75 1995; 95-3232; 515’.723–dc20
15. D.R. Chalice, How to teach a class by the modified Moore method. Am. Math. Mon. 102(4),
317–321 (1995)
374 References
16. A. Church, Correction to a note on the Entscheidungsproblem. J. Symb. Log. 1(3), 101–102
(1936)
17. A. Church, A note on the Entscheidungsproblem. J. Symb. Log. 1(1), 40–41 (1936)
18. A. Church, Introduction to Mathematical Logic. Princeton Landmarks in Mathematics and
Physics (Princeton University Press, Princeton, 1996). Tenth printing. ISBN 0-691-02906-7
19. P.J. Cohen, The independence of the continuum hypothesis. Proc. Natl. Acad. Sci. U. S. A.
50(6), 1143–1148 (1963)
20. P.J. Cohen, The independence of the continuum hypothesis, II. Proc. Natl. Acad. Sci. U. S. A.
51(1), 105–110 (1964)
21. P.J. Cohen, Set Theory and the Continuum Hypothesis. (W. A. Benjamin, New York, 1966)
QA9.C69
22. P.J. Cohen, Set Theory and the Continuum Hypothesis (Dover, Mineola, 2008). ISBN-10 0-
486-46921-2; QA248.C614 2008; LCCC No. 2008042847 511.3’22—dc22
23. L. Dadda, Some schemes for parallel multipliers. Alta Frequenza 34(2), 349–356 (1965)
24. C. Davis, The role of the untrue in mathematics. Math. Intell. 31(3), 4–8 (2009). Text of a talk
presented at the Joint Mathematics Meeting in Toronto on Monday 5 January 2009
25. R. Dedekind, Essays on the Theory of Numbers (Dover, New York, 1963). ISBN 0-486-21010-
3; 63-3681. (Also Open Court Publishing, Chicago, 1901; QA248.D3)
26. M. Dresher, Games of Strategy: Theory and Applications. Prentice-Hall Applied Mathematics
Series (Prentice-Hall Inc., Englewood Cliffs, 1961)
27. J.L.E. Dreyer, A History of Astronomy from Thales to Kepler, 2nd edn. (Dover, New York,
1953). SBN 486-60079-3; QB15.D77 1953; 53-12387
28. U. Dudley, Mathematical Cranks. MAA Spectrum (Mathematical Association of America,
Washington, DC, 1992)
29. U. Dudley, The Trisectors. MAA Spectrum, revised edition (Mathematical Association of
America, Washington, DC, 1994)
30. J. Dugundji, Topology (Allyn and Bacon, Boston, 1966). ISBN 0-205-00271-4. LCCC No.
66-10940
31. H.B. Enderton, A Mathematical Introduction to Logic (Academic Press, New York, 1972).
LCCC No. 78-182659
32. J. Evans, The History and Practice of Ancient Astronomy (Oxford University Press,
New York, 1998). ISBN 0-19-509539-1; QB16.E93 1998; LCCC No. 97-16539; 520’.
938–dc21
33. S. Feferman, Does mathematics need new axioms? Am. Math. Mon. 106(2), 99–111 (1999)
34. G.B. Folland, Real Analysis: Modern Techniques and Their Applications (Wiley, New York,
1984)
35. A. Fraenkel, Axiomatische theorie der geordneten mengen. (untersuchungen über die grund-
lagen der mengenlehre. ii.). J. Reine Angew. Math. (Crelle’s Journal) 1926(155), 129–158
(1926)
36. A. Fraenkel, Einleitung in die Mengenlehre, volume IX of Die Grundlehren der math-
ematischen Wissenschaften in Einzerldarstellungen mit besonderer Berücksichtigung der
Anwendungsgebiete (Dover, New York, 1946). First edition published by Julius (Springer,
Berlin, 1919)
37. A. Fraenkel, Abstract Set Theory (North-Holland, Amsterdam, 1953)
38. G. Frege, Begriffsschrift, eine der arithmetischen nachgebildete Formelsprache des reinen
Denkens (Nebert, Halle, 1879)
39. G. Frege, Conceptual Notation, and Related Articles; Translated [from the German] and
Edited with a Biography and Introduction by Terrell Ward Bynum (Clarendon Press, Oxford,
1972). ISBN 0198243596; B3245.F22 B94
40. D. Gale, L.S. Shapley, College admissions and the stability of marriage. Am. Math. Mon.
69(1), 9–15 (1962)
41. D. Gale, L.S. Shapley, College admissions and the stability of marriage. Am. Math. Mon.
120(5), 386–391 (2013). Reprint of MR1531503
www.pdfgrip.com
References 375
42. J. Geanakoplos, Three brief proofs of Arrow’s impossibility theorem. Econ. Theory 26(1),
211–215 (2005)
43. K. Gödel, Über formal unentscheidbare Sätze der Principia Mathematica und verwandter
Systeme I. Mon. Math. Phys. 38, 173–198 (1931) Leipzig
44. K. Gödel, The consistency of the axiom of choice and of the generalized continuum
hypothesis. Proc. Natl. Acad. Sci. U. S. A. 24, 556–557 (1938)
45. K. Gödel, The Consistency of the Axiom of Choice and of the Generalized Continuum
Hypothesis with the Axioms of Set Theory. Annals of Mathematics Studies, vol. 3 (Princeton
University Press, Princeton, 1940). QA9.G54
46. K. Gödel, On formally Undecidable Propositions of Principia Mathematica and Related
Systems (Dover, Mineola, 1992). ISBN 0-486-66980-7. QA248.G573 1992. LCCC No.
91-45947. 511.3–dc20
47. G. Gonthier, Formal proof—the four-color theorem. Not. Am. Math. Soc. 55(11), 1382–1393
(2008)
48. C. Gram, O. Hestvik, H. Isaksson, P.T. Jacobsen, J. Jensen, P. Naur, B.S. Petersen, B. Svej-
gaard, GIER—a Danish computer of medium size. IEEE Trans. Electron. Comput. EC-12(5),
629–650 (1963)
49. W. Grassmann, J.-P. Tremblay, Logic and Discrete Mathematics: A Computer Science Per-
spective. (Prentice Hall, Upper Saddle River, 1996). ISBN 0-13-501296-6; QA76.9.M35G725
1996; LCCC No. 95-38351; 005.1’01’5113–dc20
50. T.C. Hales, Formal proof. Not. Am. Math. Soc. 55(11), 1370–1380 (2008)
51. E.C. Hall, MIT’s role in project Apollo: final report on contracts NAS 9-163 and NAS 94065.
Technical report R-700 (Charles Stark Draper Laboratory, MIT, Cambridge) (1972)
52. P. Hamburger, R.E. Pippert, Venn said it couldn’t be done. Math. Mag. 73(2), 105–110 (2000)
53. P. Hamburger, R.E. Pippert, A symmetrical beauty: a non-simple 7-Venn diagram with a
minimum vertex set. Ars Combinatoria 66, 129–137 (2003)
54. J. Harrison, Formal proof—theory and practice. Not. Am. Math. Soc. 55(11), 1395–1406
(2008)
55. J.R. Harrison, Handbook of Practical Logic and Automated Reasoning (Cambridge University
Press, London, 2009). ISBN-13: 9780521899574
56. F. Hausdorff, Bemerkung §ber den Inhalt von Punktmengen. Math. Ann. 75(3), 428–433
(1914)
57. F. Hausdorff, Set Theory, 2nd edn. (Chelsea, New York, 1962). Translated from the German
by John R. Aumann, et al. QA248.H353 1962; LCCC No. 62-19176
58. M. Henle, Which Numbers Are Real? Classroom Resource Materials Series (Mathematical
Association of America, Washington, DC, 2012)
59. P. Henrici, Essentials of Numerical Analysis with Pocket Calculator Demonstrations (Wiley,
New York, 1982). ISBN 0-471-05904-8; QA297.H42; 81-10468; 519.4
60. E. Hewitt, K. Stromberg, Real and Abstract Analysis (Springer, New York, 1965). LCCC No.
65-26609
61. D. Hilbert, Foundations of Geometry, 10th edn. (Open Court, La Salle, 1971). Eighth printing,
1996. Translated by Leo Unger and revised by Paul Bernays. ISBN 0-87548-164-7; LCCC
No. 73-110344
62. D. Hilbert, W. Ackermann, Principles of Mathematical Logic (Chelsea Publishing Company,
New York, 1950). BC135.H514. (Translation of David Hilbert & Wilhelm Ackermann,
Grundzüge der theoretischen Logik, Julius Springer, Berlin, Germany, 1928 & 1938)
63. T.W. Hungerford, Algebra (Holt, Rinehart and Winston, New York, 1974). ISBN 0-03-
086078-4; QA155.H83; LCCC No. 73-15693; 512
64. International Business Machines Corporation, 590 Madison Avenue, New York 22, NY. IBM
604 Electronic Calculating Punch Manual of Operation, 1954. http://www.bitsavers.org/pdf/
ibm/604
65. International Business Machines Corporation, 590 Madison Avenue, New York 22, NY. IBM
650 Magnetic-Drum Data Processing Machine Manual of Operation, revised edition, June
1955. http://www.bitsavers.org/pdf/ibm/650
www.pdfgrip.com
376 References
66. T.J. Jech, The Axiom of Choice, II. Studies in Logic and the Foundations of Mathematics,
vol. 75 (North Holland, Amsterdam, 1973). ISBN 0-444-10484-4; QA248.J4 1973; LCCC
No. 73-75535
67. M. Kaku, Venus envy. Wall Street J. CCXLVIII(40), A8 (2006). Western Edition
68. E. Kalai, Foreword: the high priest of game theory. Am. Math. Mon. 120(5), 384–385 (2013)
69. H. Karttunen, P. Kröger, H. Oja, M. Poutanen, K.J. Donner (eds.), Fundamental Astronomy
(Springer, New York, 1994). ISBN 0-387-57203; QB43.2.T2613 1993; 93-31098; 520–dc20
(second enlarged edition)
70. L. Keen, Julia sets, in Chaos and Fractals: The Mathematics Behind the Computer Graphics
(American Mathematical Society, Providence, 1989), pp. 57–74
71. C.T. Kelley, in Iterative Methods for Linear and Nonlinear Equations. Frontiers in Applied
Mathematics, vol. 16 (Society for Industrial and Applied Mathematics, Philadelphia, 1995).
0-89871-352-8; QA297.8.K45 1995; 95-32249; 519.4–dc20
72. S.C. Kleene, Mathematical Logic (Dover, Mineola, 2002). ISBN 0-486-42533-9; QA9.A1
K54 2002; LCCC No. 2002034823 (originally published by Wiley in 1967)
73. S. Kortenkamp, Why Isn’t Pluto a Planet?: A Book about Planets (Capstone Press, Mankato,
2007). ISBN 0-7368-6753-8; QB701.K57 2007; LCCC No. 2006025648; 523.4–dc22
74. K. Kunen, Set Theory: An Introduction to Independence Proofs (North Holland, Amsterdam,
1980). ISBN 0-444-85401-0. QA248.K75 1980. LCCC No. 80-20375. 510.3’22
75. E. Lakdawalla, Pluto and the Kuiper belt. Sky Telesc. 127(2), 18–25 (2014)
76. E. Landau, Foundations of Analysis (The Tan Chiang Book Co., 1951)
77. S. Lang, Algebra (Addison-Wesley, Reading, 1965). LCCC No. 65-23677
78. L.S. LaPiana, F.H. Bauer, Mars Climate Orbiter mishap investigation board phase I report.
Technical report, National Aeronautics and Space Administration, 10 Nov 1999. http://sse.
jpl.nasa.gov.news
79. J. Łukasiewicz, Zur vollen dreiwertigen Aussagenlogik. Erkenntniss 5, 176 (1935)
80. J. Łukasiewicz, Zur Geschichte der Aussagenlogik. Erkenntniss 5, 111–131 (1935)
81. Y.I. Manin, A Course in Mathematical Logic (Springer, New York, 1977). ISBN 0-387-90243-
0; QA9.M296; LCCC No. 77-1838; 511’.3
82. M.F. Mann, R.R. Rathbone, J.B. Bennett, Whirlwind I Operation Logic, vol. 29 (Massachus-
setts Institute of Technology, Digital Computer Laboratory, Cambridge, 1954). http://www.
bitsavers.org/pdf/mit/whirlwind
83. V.W. Marek, J. Mycielski, Foundations of mathematics in the twentieth century. Am. Math.
Mon. 108(5), 449–468 (2001). ISSN 0002-9890
84. A. Margaris, First Order Mathematical Logic (Dover Publications Inc., New York, 1990).
Corrected reprint of the 1967 edition
85. B. Mates, Elementary Logic, 2nd edn. (Oxford University Press, New York, 1972).
BC135.M37 1972. LCCC No. 74-166004
86. R. Miller, Computable fields and Galois theory. Not. Math. Am. Soc. 55(7), 798–807 (2008).
http://www.ams.org/notices
87. K. Mitchell, J. Ryan, Game theory and models of animal behavior, in UMAP/ILAP Modules
2002–03: Tools for Teaching, ed. by P.J. Campbell (COMAP, Bedford, 2003), pp. 1–48.
Reprinted as Module 783, COMAP, Bedford, MA, 2003
88. J.D. Monk, Introduction to Set Theory (McGraw-Hill, New York, 1969). LCCC No. 68-20056
89. J.D. Monk, Mathematical Logic (Springer, New York, 1976). ISBN 0-387-90170-1.
QA9.M68
90. M.R. Mugnier, G.A.M. Cross, F.N. Papavasiliou, The in vivo dynamics of
antigenic variation in Trypanosoma brucei. Science 347(6229), 1470–1473 (2015).
doi:10.1126/science.aaa4502. http://www.sciencemag.org/content/347/6229/1470.full
91. J.R. Munkres, Topology, 2nd edn. (Prentice-Hall, Englewood Cliffs, 2000). ISBN 0-13-
181629-2; QA611.M82 2000; 514–dc21; LCCC No. 99-052942
92. R.B. Myerson, An Introduction to Game Theory. Studies in Mathematical Economics. MAA
Studies in Mathematics, vol. 25 (Mathematical Association of America, Washington, DC,
1986), pp. 1–61
www.pdfgrip.com
References 377
93. J. Nash, Non-cooperative games. Ann. Math. Second Ser. 54(2), 286–295 (1951)
94. J.F. Nash Jr., The bargaining problem. Econometrica 18(2), 155–162 (1950)
95. J.F. Nash Jr., Equilibrium points in n-person games. Proc. Natl. Acad. Sci. U. S. A. 36(1),
48–49 (1950)
96. National Resident Matching Program, http://www.nrmp.org/match-process/match-algorithm/
97. O. Neugebauer, The Exact Sciences in Antiquity, 2nd edn. (Dover, New York, 1969).
SBN 486-22332-0; LCCC No. 69-20421. Reprint of the 1957 second edition from Brown
University Press
98. O. Neugebauer, A History of Ancient Mathematical Astronomy. Part I. Studies in the History
of Mathematics and Physical Sciences, vol. 1 (Springer, New York, 1975)
99. O. Neugebauer, A History of Ancient Mathematical Astronomy. Part II. Studies in the History
of Mathematics and Physical Sciences, vol. 1 (Springer, New York, 1975)
100. O. Neugebauer, A History of Ancient Mathematical Astronomy. Part III. Studies in the History
of Mathematics and Physical Sciences, vol. 1 (Springer, New York, 1975)
101. Y. Nievergelt, The truth table of the logical implication. Math. Gaz. 94(531), 509–513 (2010).
ISSN 0025-5572
102. Y. Nievergelt, H. Sullivan, Undecidability in fuzzy logic. UMAP J. 31(4), 323–359 (2010).
Also reprinted as UMAP Unit 804
103. Packard Bell Electronics, Packard Bell Computer, 1905 Armacost Avenue, Los
Angeles 25, California. pb 250 Technical Manual, Volume 1, 15 July 1961.
http://www.bitsavers.org/pdf/packardBell
104. C.S. Peirce, On the algebra of logic: a contribution to the philosophy of notation. Am. J. Math.
7(2), 180–196 (1885)
105. F. Quinn, A revolution in mathematics? What really happened a century ago and why it
matters today. Not. Am. Math. Soc. 59(1), 31–37 (2012)
106. A. Rapoport, Two-Person Game Theory: The Essential Ideas (The University of Michigan
Press, Ann Arbor, 1966)
107. A. Rapoport, Two-Person Game Theory (Dover Publications Inc., Mineola, 1999). Reprint of
the 1966 original [Univ. Michigan Press, Ann Arbor, MI, 1966; MR0210463 (35 #1356)]
108. J.W. Robbin, Mathematical Logic: A First Course (Dover, Mineola, 2006). SBN 486-61272-
4; LCCC No. 65-12253 (originally published by W. A. Benjamin, New York, 1969)
109. R.M. Robinson, Primitive recursive functions. Bull. Am. Math. Soc. 53, 925–942 (1947)
110. J.B. Rosser, Logic For Mathematicians (McGraw-Hill, New York, 1953). BC135.R58; LCCC
No. 51-12640
111. H. Rubin, J.E. Rubin, in Equivalents of the Axiom of Choice, II. Studies in Logic and the
Foundations of Mathematics, vol. 116 (North Holland, Amsterdam, 1980). ISBN 0-444-
87708-8; QA248.R8 1985; LCCC No. 84-28692; 511.3’22
112. D.G. Saari, Chaotic Elections: A Mathematician Looks at Voting (American Mathematical
Society, Providence, 2001). ISBN 0-8218-2847-9
113. D.G. Saari, Mathematics and voting. Not. Am. Math. Soc. 55(4), 448–455 (2008). http://
www.ams.org/notices
114. E. Schechter, Classical and Nonclassical Logics (Princeton University Press, Princeton 2005).
ISBN 0-691-12279-2; QA9.3.S39 2005; 2004066030; 160–dc22
115. R. Serrano, The theory of implementation of social choice rules. SIAM Rev. 46(3), 377–414
(2004) (electronic)
116. T.A. Skolem, Einige Bemerkungen zur axiomatischen Begrundung der Mengenlehre, in
Matematikerkongressen i Helsingfors den 4–7 Juli 1922, Den femte skandinaviska matem-
atikercongressen, Redogšrelse (Helsinki, SF, 1923), pp. 217–237. Akademiska Bokhandeln.
Cited in [36] and [128]
117. R.M. Smullyan, First-Order Logic (Dover, Mineola, 1995). ISBN 0-486-68370-2. QA9.S57
1994. LCCC No. 94-39736. 511.3–dc20
118. M. Spivak, Calculus, 2nd edn. (Publish or Perish, Wilmington, DE, 1980). ISBN 0-914098-
77-2. LCCC No. 80-82517
www.pdfgrip.com
378 References
119. S. Stahl, A Gentle Introduction to Game Theory. Mathematical World, vol. 13 (American
Mathematical Society, Providence, 1999)
120. J. Stillwell, The Real Numbers. Undergraduate Texts in Mathematics. An Introduction to Set
Theory and Analysis (Springer, Cham, 2013)
121. J. Stoer, R. Bulirsch, Introduction to Numerical Analysis, 2nd edn. (Springer, New York,
1993). ISBN 0-387-97878-X; QA297.S8213 1992; 92-20536; 519.4–dc20
122. R.R. Stoll, Sets, Logic, and Axiomatic Theories (W. H. Freeman, San Francisco, 1961). LCCC
No. 61-6784
123. A.A. Stolyar, Introduction to Elementary Mathematical Logic (Dover, Mineola, 1983). ISBN
0-486-64561-4. BC135.S7613 1983. LCCC No. 83-5223. 511.3
124. P.D. Straffin Jr., The prisoner’s dilemma. UMAP J. 1(1), 102–103 (1980)
125. P.D. Straffin Jr., Changing the way we think about the social world. Two Year Coll. Math. J.
14(3), 229–232 (1983)
126. P.D. Straffin Jr., Game theory and nuclear deterrence. UMAP J. 1(1), 87–92 (1989)
127. D.J. Struik, A Concise History of Mathematics, 4th edn. (Dover, New York, 1987). ISBN
0-486-60255-9; QA21.S87 1987; LCCC No. 86-8855; 510’.09 (revised edition)
128. P. Suppes, Axiomatic Set Theory (Dover, Mineola, 1972). ISBN 0-486-61630-4. LCCC No.
72-86226
129. A. Tarski, Introduction to Logic and to the Methodology of Deductive Sciences (Oxford
University Press, Oxford, 1941)
130. A.E. Taylor, Introduction to Functional Analysis (Wiley, New York, 1958) 58-12704
131. A.W. Tucker, The prisoner’s dilemma: a two-person dilemma. UMAP J. 1, 101 (1980). Dated
“Stanford, May 1950”
132. A.W. Tucker, A two-person dilemma: the prisoner’s dilemma. Two Year Coll. Math. J. 14(3),
228 (1983). Dated “Stanford, May 1950”
133. G. van Brummelen, The Mathematics of the Heavens and the Earth: The Early History
of Trigonometry (Princeton University Press, Princeton, 2009). ISBN 978-0-691-12973-0;
QA24.V36 2009; 516.2409–dc22; LCCC No. 2008032521
134. B.L. van der Waerden, Geometry and Algebra in Ancient Civilizations (Springer, Berlin,
1983). ISBN 3-540-12159-5. QA151.W34 1983. LCCC No. 83-501. 512’.009
135. J. von Neumann, Zur einführung der transfiniten Zahlen. Acta Szeged 1, 199–208 (1923)
Cited in [128, p. 129]
136. R.J. Walker, Algebraic Curves (Springer, New York, 1978). ISBN 0-387-90361-5.
QA564.W35 1978. LCCC No. 78-11956. 516’.352. (First published by Princeton University
Press, Princeton, 1950)
137. D.A. Weintraub, Is Pluto a Planet?: A Historical Journey through the Solar System (Princeton
University Press, Princeton, 2007). ISBN 0-691-12348-9; QB602.9.W45 2007; LCCC No.
2006929630
138. Westinghouse Electric Corporation, Surface Division, P.O. Box 1897, Baltimore, MD
21203. Technical Manual for DPS-2402 Computer, Volume I, 1 February 1966.
http://www.bitsavers.org/pdf/westinghouse
139. F. Wiedijk, Formal proof—getting started. Not. Am. Math. Soc. 55(11), 1408–1417 (2008)
140. R.L. Wilder, Introduction to the Foundations of Mathematics, 2nd edn. (Dover, Mineola,
2012)
141. R.S. Wolf, A Tour Through Mathematical Logic. The Carus Mathematical Monographs, vol.
30. (Mathematical Association of America, Washington, DC, 2005). ISBN 0-88385-036-2;
LCCC No. 2004113540
142. S. Wolfram, The Mathematica Book, 3rd edn. (Wolfram Media, Champaign, 1996). ISBN
0-9650532-1-0; QA76.95.W65 1996; LCCC No. 96-7218; 510’.285’53–dc20
143. N.N. Yu, A one-shot proof of Arrow’s impossibility theorem. Econ. Theory 50(2), 523–525
(2012)
144. E. Zermelo, Beweis, daß jede Menge wohlgeordnet werden kann. Math. Ann. 59(4), 514–516,
(1904)
www.pdfgrip.com
References 379
145. E. Zermelo, Untersuchungen über die Grundlagen der Mengenlehre, I. Math. Ann. 65(2),
261–281 (1908). Cited in [8, p. 21]
146. D.E. Zitarelli, The origin and early impact of the Moore method. Am. Math. Mon. 111(6),
465–486 (2004)
147. M. Zorn, A remark on method in transfinite algebra. Bull. Am. Math. Soc. 41(10), 667–670
(1935)
148. G. Zukav, The Dancing Wu Li Masters: An Overview of the New Physics (Morrow, New York,
1979). ISBN: 0-688-03402-0; QC173.98.Z84 1979; 530.1’2; LCCC No. 78-25827
www.pdfgrip.com
Index
382 Index
Index 383
384 Index
Index 385
386 Index
Index 387
M N
major Nash equilibrium, 308
premiss, 7 Nash, John Forbes, Jr., 304
map National Resident Matching Program, 315
canonical, 178 natural numbers, 192
quotient, 178 negative integer, 221
mapping notation
principle, 300 product, 237
match maker’s algorithm, 318 sum, 238
match making, 315 number
matrix field, 235
of a prenex normal form, 95 of elements, 248
maximal rational, 230
element, 184 real, 254
Maximal-Chain Principle triangular, 255
Hausdorff, 299 numerator, 230
Maximal-Element Principle
Zorn, 290
Maximal-Order Principle O
Kuratowski, 299 operator
Maximal-Set Principle overloading, 158
Zorn, 291 order
Maximal-Subset Principle of natural numbers
Hausdorff, 299 definition, 205
maximum, 183, 210 on Z, 221, 222
member of a set, 111 ordered pairs, 143
membership ordinal, 272
of sets, 111 overloading operators, 158
Mercury, 3
Minimal
Logic, 7, 38, 41, 44 P
minimal pairing, axiom (S3), 119
element, 184 pairwise disjoint
propositional calculus, 7, 38, set, 135
41, 44 union, 135
minimum, 183, 209 partial order, 181
www.pdfgrip.com
388 Index
Index 389
390 Index
Index 391
V well-order, 183
valid induced on subsets, 184
propositional form, 57 Well-Ordering
value Principle
distinguished, 57 Zermelo, 288
propositional, 57 Westinghouse DPS-2402 computer
value of a function, 155 NAND, 66
variable Whirlwind I computer
bound, 78 division, 213
free, 78 Whitaker, J. M., 258
individual, 77
propositional, 5, 77
Venn diagram, 127 Z
vertices, 149 Zermelo
veto, 329 Theorem, 185
von Neumann, John, 1, 303 Well-Ordering Principle, 288
voters’ profile, 325 Zorn
Maximal-Element Principle, 290
Maximal-Set Principle, 291
W Zorn’s Lemma, 185
well-formed formula, 5, 77 Zorn, M., 185