Anda di halaman 1dari 125

OXFORD UNIVERSITY

Part C Mathematics, Mathematics and Philosophy, Mathematics and Computation


M.Sc. in Mathematics and Foundations of Computer Science
B.Phil. and M.St. in Philosophy

16 lectures on
Gödel’s Incompleteness Theorems
Michaelmas Term 2010

Daniel Isaacson
Faculty of Philosophy
Oxford University

1st December 2010

Copyright c 2010 by Daniel Isaacson


All rights reserved. No part of this publication may be reproduced
without prior permission by anyone other than for their own use in
studying this subject. Enquiries and corrections to
daniel.isaacson@philosophy.ox.ac.uk
Contents

0 Background: first-order logic and formal systems 2


0.1 First-order formal languages with identity . . . . . . . . . . . . . . . 2
0.2 Interpretations of first-order formal languages with identity; truth of
a first-order formula in an interpretation; logical validity and logical
consequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
0.3 A system of natural deduction for first-order logic with identity . . . 4
0.4 Prenex normal form . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
0.4.1 Model theory and proof theory . . . . . . . . . . . . . . . . . 11
0.5 Completeness of a system of natural deduction with respect to first-
order logical consequence . . . . . . . . . . . . . . . . . . . . . . . . . 11
0.5.1 Lindenbaum’s Lemma . . . . . . . . . . . . . . . . . . . . . . 11
0.5.2 The Completeness Theorem . . . . . . . . . . . . . . . . . . . 11
0.5.3 The Compactness Theorem for first-order logic . . . . . . . . . 11
0.5.4 The existence of non-standard models of the truths of arithmetic 11
0.5.5 Completeness of other systems of derivation with respect to
first-order logical consequence . . . . . . . . . . . . . . . . . . 11

1 Introduction: a weak form of Gödel’s First Incompleteness Theo-


rem; the symbols and expressions of a language for arithmetic LE ;
Gödel numbering of the expressions of LE 12
1.1 Introduction: a weak form of Gödel’s First Incompleteness Theorem 12
1.2 The symbols and expressions of a language for arithmetic LE . . . . . 16
1.3 Gödel numbering of the expressions of LE . . . . . . . . . . . . . . . 17

2 Terms and formulas of the language LE ; expressibility of diagonal


substitution in the language LE 20
2.1 Terms and formulas of the language LE . . . . . . . . . . . . . . . . . 20
2.1.1 Terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.2 Formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.3 Free and bound variables; open formulas and sentences . . . . 22

1
CONTENTS 2

2.2 Designation by terms in LE , truth of sentences of LE , and express-


ibility of sets and relations of natural numbers by formulas of LE . . 23
2.2.1 Designation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2.2 Truth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2.3 Expressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3 Concatenation of numbers in a given base notation is Arithmetical. . 24
2.4 Substitution and diagonal functions, and their arithmetization . . . . 25

3 The Diagonal Lemma; expressibility of properties of sequence num-


bers 28
3.1 The Diagonal Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2 Expressibility of properties of sequence numbers in the language LE 29
3.2.1 Properties of sequences of digits . . . . . . . . . . . . . . . . . 29
3.2.2 Sequence numbers . . . . . . . . . . . . . . . . . . . . . . . . 31
3.3 Coding of finite sequences of Gödel numbers . . . . . . . . . . . . . . 31

4 A formal system PAE for arithmetic; an Arithmetical proof predi-


cate for PAE ; a weak version of Gödel’s first incompleteness theo-
rem for PAE 34
4.1 A formal system PAE for arithmetic . . . . . . . . . . . . . . . . . . . 34
4.2 An Arithmetical proof predicate for PAE . . . . . . . . . . . . . . . . 37
4.3 An inefficient and a weak version of Gödel’s First Incompleteness
Theorem for PAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

5 The system PA with zero, successor, addition, multiplication, and


≤ as primitive; Σ0 - and Σ1 -formulas; a Σ0 -coding of finite sets of
ordered pairs; the relation xy = z is ∆1 -expressible in the language
of PA 42
5.1 The system PA with zero, successor, addition, multiplication, and ≤
as primitive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.2 Σ0 -formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.3 Σ1 - and Π1 -formulas; Σ1 - Π1 - and ∆1 -relations . . . . . . . . . . . . . 44
5.4 Arithmetization of syntax in the language of PA . . . . . . . . . . . . 46
5.5 A Σ0 -coding of finite sets of ordered pairs of numbers . . . . . . . . . 47
5.6 The relation xy = z is ∆1 -expressible in the language of PA . . . . . . 49

6 Every Σ-formula is provably equivalent to a Σ1 -formula; the arith-


metized proof predicate for PA is Σ1 ; the arithmetical hierarchy 50
6.1 Σ-formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.2 The arithmetized proof predicate for PA is Σ1 . . . . . . . . . . . . . 52
6.3 The arithmetical hierarchy . . . . . . . . . . . . . . . . . . . . . . . . 53
CONTENTS 3

7 Σ0 -completeness and Σ1 -completeness; weak systems of arithmetic


Q and R (without induction); Σ0 -completeness of systems R, Q, and
PA; Σ0 -soundness and Σ1 -soundness 55
7.1 Σ0 -completeness and Σ1 -completeness . . . . . . . . . . . . . . . . . . 55
7.2 Weak systems of arithmetic Q and R (without induction) . . . . . . . 57
7.3 Σ0 -completeness of systems R, Q, and PA . . . . . . . . . . . . . . . 59
7.4 Σ0 -soundness and Σ1 -soundness . . . . . . . . . . . . . . . . . . . . . 62

8 The notions of consistency, ω-consistency and 1-consistency; incom-


pleteness from the assumption of 1-consistency; truth of the Gödel
sentence; ω-incompleteness. 63
8.1 The notions of consistency, ω-consistency and 1-consistency. . . . . . 63
8.2 Incompleteness of PA from the assumption of 1-consistency . . . . . . 67
8.3 Truth of the Gödel sentence . . . . . . . . . . . . . . . . . . . . . . . 68
8.4 PA is ω-incomplete . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

9 Enumerability and the Separation Lemma; incompleteness of PA


from the assumption of consistency (Rosser’s Theorem); weak and
strong definability of a function in a system; formal provability of
the Diagonal Lemma 71
9.1 Enumerability and the Separation Lemma . . . . . . . . . . . . . . . 72
9.2 Incompleteness of PA from the assumption of consistency (Rosser’s
Theorem) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
9.3 Weak and strong definability of a function in a system . . . . . . . . 74
9.4 Formal provability of the Diagonal Lemma . . . . . . . . . . . . . . . 76

10 Arithemization of consistency; provability predicates; Gödel’s Sec-


ond Incompleteness Theorem; Löb’s Theorem; analyzing and strength-
ening the First Incompleteness Theorem 79
10.1 Arithmetization of the statement that a system S is consistent . . . . 79
10.2 Provability predicates. . . . . . . . . . . . . . . . . . . . . . . . . . . 80
10.3 Gödel’s Second Incompleteness Theorem . . . . . . . . . . . . . . . . 82
10.4 Löb’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
10.5 Analyzing and strengthening the First Incompleteness Theorem . . . 84
10.5.1 S 0∼ GS cannot be proved from the consistency of S . . . . . 85
10.5.2 Strengthened second half of the First Incompleteness Theorem 85
10.5.3 Consistency of S∪{ConS} is strictly weaker than 1-consistency
of S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

11 Provable Σ1 -completeness 88
CONTENTS 4

12 The ω-rule and uniform reflection; PA proves that PA proves every


instance of the Gödel sentence; Π1 -uniform reflection and consis-
tency; PA is Π1 -conservative over PAΠ2 ∪ {ConP A } 98
12.1 The ω-rule and uniform reflection . . . . . . . . . . . . . . . . . . . . 98
12.2 PA proves that PA proves every instance of the Gödel sentence . . . . 100
12.3 Equivalence of Π1 -Uniform Reflection and consistency . . . . . . . . . 101
12.4 PA is Π1 -conservative over PAΠ2 ∪ {ConP A } . . . . . . . . . . . . . . 102

13 Provability logic: the system GL 104


13.1 The language of GL . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
13.2 The axioms and inference rules of GL . . . . . . . . . . . . . . . . . . 105
13.3 Some derivations in GL . . . . . . . . . . . . . . . . . . . . . . . . . . 107
13.4 Closure of GL under substitution by provably equivalent formulas . . 108
13.5 Closure of GL under substitution of provably equivalent formulas is
provable in GL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
13.6 Strengthened proof that the closure of GL under substitution of prov-
ably equivalent formulas is provable in GL . . . . . . . . . . . . . . . 111

14 The fixed-point theorem for GL 114


14.1 The notion of a sentence letter modalized in a sentence, and arithme-
tized substitution for modalized sentences . . . . . . . . . . . . . . . 114
14.2 The fixed point theorem for GL . . . . . . . . . . . . . . . . . . . . . 117
Lecture 0

Background: first-order logic and


formal systems

This unspoken lecture reviews essential results on first-order logic and formal systems
as background to the development of Gödel’s incompleteness theorems.
The notion of logical consequence is the starting point. A rough characterization
of this notion is the following. A sentence φ is a logical consequence of a set of
sentences Γ, symbolized as Γ  φ, if and only if φ is true in every interpretation of
the language of Γ ∪ {φ} in which all the sentences of Γ are true. If Γ is empty, Γ  φ
reduces to the condition that φ is true in every interpretation of the language of φ,
which is to say, φ is logically valid, symbolized by  φ.
Accordingly, we need to make precise the notion of a formal language and the
notion of interpretation of a formal langauge.

0.1 First-order formal languages with identity


The notion of a formal language begins from the specification of a finite alphabet
of symbols which are strung together (concatenated) to produce the expressions of
the language. The symbols of the alphabet of a formal language have no intrinsic
meaning (they are purely formal symbols) but are chosen with the intention that
they should be interpretable in certain ways (which in general does not rule out
their being interpreted in other ways). Some but not all of the expressions of the
language constitute well-formed expressions. Which expressions are well-formed is
a matter of stipulation, according to our intended use of the formal language. The
symbols of a first-order language are of three sorts, logical, non-logical, and syntactic.
The logical symbols include two sorts: propositional, and quantificational, and may
include a third, identity. The non-logical are of two sorts, function symbols and
predicate or relation symbols. Functions and relations are of a particular ’arity’

5
LECTURE 0 6

(unary, binary, ternary, and so on). A zero-ary function symbol is a constant term,
a zero-ary relation symbol represents a sentence. The shape of a symbol is completely
arbitrary, though there are some conventional choices, e.g. the universal quantifier
is usually written “∀”, but used often to be written “( )” V (the brackets enclosing
the variable of quantification), and is sometimes written “ ” (called a California
quantifier). But the symbols can be anything and in particular, as we shall exploit,
they can be digits also used to generate numerals.
We shall in this course always have full first-order classical (as opposed to intu-
itionistic) predicate logic with identity as our background logic. Quite apart from
the shapes of the symbols, there are choices to be made as to which are our prim-
itive propositional functions and quantifiers. In our background logic we shall take
as our primitives of propositional logic negation, conjunction, disjunction, and im-
plication, for which the symbols will be ∼, ∧, ∨, and ⊃, and both the universal
and existential quantifiers as primitive, written as ∀ and ∃. As you will know from
a previous logic course, we could take as primitive just negation with any one of
conjunction, disjunction, and implication and either the universal or the existential
quantifier, and in our official formal language we shall take as primitive ∼, ⊃, and ∀.
We shall operate with the langauge as if it also contained symbols for conjunction,
disjunction, and the existential quantifiers, but strictly (A ∨ B) will be an abbrevi-
ation for (∼ A ⊃ B), etc. There is also the connective “if and only if”, which we
will symbolize as ≡. Even in our background logic we will take (A ≡ B) to be an
abbreviation for ((A ⊃ B) ∧ (B ⊃ A)).
Stipulating that our background logic is first-order classical logic with identity,
means that we have a primitive symbol for identity, which we will take as the
.
common symbol for identity = (though others are sometimes used, e.g. ≈, ≏, =),
but having a symbol and first-order axioms for identity is not sufficient to make a
logical system first-order logic with identity. Identity is a two place relation which
every object bears to itself and to no other object. The first property is easily
expressed: ∀v1 v1 = v1 , but the second property is not easily expressed, in particular
not expressed by ∀v1 ∀v2 (v1 6= v2 ⊃ v1 6= v2 ) (if v1 is a different object from v2 then
v1 6= v2 ).
terms in the language
formulas in the language
LECTURE 0 7

0.2 Interpretations of first-order formal languages


with identity; truth of a first-order formula in
an interpretation; logical validity and logical
consequence
0.3 A system of natural deduction for first-order
logic with identity
This system is based on natural deduction as developed by Gerhard Gentzen. In
Gentzen-style systems of natural deduction, deductions consist of branching trees.
For ease of formulation on the page, I give here a variant form of natural deduction in
which deductions are linear1 . I shall call this system LN D, standing either for Linear
Natural Deduction (but in that case be aware that this usage has nothing to do
with Girard’s usage in what he calls Linear Logic), or Lemmon Natural Deduction.
[Strictly I should call this system something like CLN D, for Classical, as opposed
to Intuitionistic, Linear Natural Deduction, but since in this course our background
logic is always classical and not intuitionistic, I won’t mark the distinction.]
Deductions are generated by using the Rule of Assumption and fourteen Rules of
Inference. There are no axioms. The rules of inference come in pairs, an Introduction
and an Elimination rule for each one of the seven logical constants ∧, ∨, ⊃, ∼, ∀, ∃,
and =. We take (F ≡ G) to be defined as ((F ⊃ G) ∧ (G ⊃ F )). The first formula
of a Deduction is always an Assumption; later formulas can also be Assumptions.
An Introduction Rule for a given logical constant results in a formula that has that
logical constant as its main logical constant and is deduced from formulas that that
do not have that logical constant as their main constant (though that constant may
occur in a subformula). An Elimination Rule for a given logical constant deduces a
formula that does not have that logical constant as its main logical constant from a
formula that has that logical constant as its main logical constant.
A deduction consists of a numbered sequence of formulas, with each of which
is associated on the left a finite set of numbers (possibly empty) which are the
numbers of the formulas that constitute the assumptions on which the formula in
that line depends, and on the right notation indicating how and from what other
formulas that formula was derived, if it’s not an Assumption. Viewed vertically
rather than horizontally, a deduction consists of four columns. The middle two
columns are a numbered sequence of formulas; the third column is a sequence of
formulas and the second column is an enumeration of those formulas. To the left of
1
A linear formulation of natural deduction was given by E.J. Lemmon in Beginning Logic,
Nelson, London, 1965; the system here is, in essentials, as in that book.
LECTURE 0 8

each numbered formula are the numbers of the Assumptions, if any, on which that
formula depends. The entry in the right-hand column gives the basis on which the
formula in the third column at that line is introduced into the deduction, i.e. either
as an Assumption or by one of the Rules of Inference. An Assumption depends on
itself, so in an application of the rule of Assumption there is one number in the first
column which is the same number as in the second column. If the formula in the
third column of a given line is introduced by a Rule of Inference, the entry in the
fourth column for that line says what Rule of Inference has been used and numbers
of the formulas to which that Rule of Inference has been applied. The Assumptions
on which those formulas depend, as given by the numbers in the first column at
the lines for those formulas, are gathered together as the Assumptions on which
the formula that results from application of that Rule of Inference depends. Four
of the Rules of Inference, ⊃-Introduction, ∼-Introduction, ∨-Elimination, and ∃-
Elimination, discharge an Assumption, so it is possible to arrive at a formula which
depends on no assumptions. Such formulas are, by the Soundness of the Rules of
Inference, logically valid.

∧-Introduction

assumptions numbering formulas justifications


A (a) F [whatever]
B (b) G [whatever]
A∪B (c) (F ∧ G) (a) (b) ∧-Introduction

∧-Elimination1

assumptions numbering formulas justifications


A (a) (F ∧ G) [whatever]
A (b) F (a) ∧-Elimination1

∧-Elimination2

assumptions numbering formulas justifications


A (a) (F ∧ G) [whatever]
A (b) G (a) ∧-Elimination1

∨-Introduction1
LECTURE 0 9

assumptions numbering formulas justifications


A (a) F [whatever]
A (b) (F ∨ G) (a) ∨-Introduction1

∨-Introduction2
assumptions numbering formulas justifications
A (a) G [whatever]
A (b) (F ∨ G) (a) ∨-Introduction1

∨-Elimination
assumptions numbering formulas justifications
A (a) (F ∨ G) [whatever]
{b} (b) F Assumption
B ∪ {b} (c) H [whatever]
{d} (d) G Assumption
C ∪ {d} (e) H [whatever]
A∪B∪C (f) H (a)(c)(e) ∨-Elimination

⊃-Introduction

assumptions numbering formulas justifications


{a} (a) F Assumption
A ∪ {a} (b) G [whatever]
A (c) (F ⊃ G) (a)(b) ⊃-Introduction

⊃-Elimination
assumptions numbering formulas justifications
A (a) (F ⊃ G) [whatever]
B (b) F [whatever]
A∪B (c) G (a)(b) ⊃-Elimination

The Introduction and Elimination rules for negation in this system of rules are
not as natural as those for the other logical connectives. A more natural formulation
of logic in natural deduction is to take negation to be defined in terms of a primitive
false sentence, sometimes symbolized as ⊥, by the equivalence ∼ ψ ≡ ψ ⊃⊥. How-
ever, in terms of formalizing informal maths talk it is more natural to take negation
LECTURE 0 10

as a primitive symbol, which we do here. It is subject to the following Introduction


and Elimination rules.

∼-Introduction

assumptions numbering formulas justifications


{a } (a) F Assumption
A ∪ {a} (b) (G∧ ∼ G) [whatever]
A (c) ∼F (a)(b) ∼-Introduction

∼-Elimination

assumptions numbering formulas justifications


A (a) ∼∼ F [whatever]
A (b) F (a) ∼-Elimination

∀-Introduction

assumptions numbering formulas justifications


A (a) F (vi ) [whatever]
A (b) ∀vi F (vi ) (a) ∀-Introduction, if vi does not occur free
in any formula enumerated by A

∀-Elimination

assumptions numbering formulas justifications


A (a) ∀vi F (vi ) [whatever]
A (b) F (t) (a) ∀-Elimination, for t any term free for vi in F (vi )

∃-Introduction

assumptions numbering formulas justifications


A (a) F (t) [whatever] for t any term free for vi in F (vi )
A (b) ∃vi F (vi ) (a) ∃-Introduction

∃-Elimination
LECTURE 0 11

assumptions numbering formulas justifications


A (a) ∃vi F (vi ) [whatever]
{b} (b) F (vi ) Assumption
B ∪ {b} (c) G [whatever]
A∪B (d) G (a)(b)(c) ∃-Elimination, if vi does not occur free
in any formula enumerated by B

General restriction on terms t in ∀-elimination and ∃-introduction: The term t


must not contain any free variable which in ψ(x) is quantified by a quantifier whose
scope in ψ(x) includes an occurrence of the variable x, i.e. the substitution of t
in ψ(x) must not result in ψ(t) having a different quantifier structure from that of
ψ(x).

=-Introduction

assumptions numbering formulas justifications


(a) t=t =-Introduction, for t any term

=-Elimination

assumptions numbering formulas justifications


A (a) t1 = t2 [whatever] for t1 , t2 any terms
B (b) F (t1 ) [whatever]
A∪B (c) F (t2 ) =-Elimination

Definition 1 For F a formula and Γ a set of formulas, we say that F is (logically)


derivable from Γ, notated Γ ⊢LDN F , or just Γ ⊢ F , if there is deduction from the
Rule of Assumption and the 14 Rules of Inferences of LDN in which the last line
has the numbered formula F and the assumptions on which F in that line depends
on are exactly the formulas of Γ.

⊢LDN (F ∨ ∼ F )

(1) (1) ∼ (F ∨ ∼ F ) Assumption


(2) (2) F Assumption
(2) (3) (F ∨ ∼ F ) (2) ∨-Introduction
(1)(2) (4) ((F ∨ ∼ F )∧ ∼ (F ∨ ∼ F )) (1)(3) ∧-Introduction
(1) (5) ∼F (2)(4) ∼-Introduction
LECTURE 0 12

(1) (6) (F ∨ ∼ F ) (5) ∨-Introduction


(1) (7) ((F ∨ ∼ F )∧ ∼ (F ∨ ∼ F )) (1)(6) ∧-Introduction
(8) ∼∼ (F ∨ ∼ F ) (1)(7) ∼-Introduction
(9) (F ∨ ∼ F ) (8) ∼-Elimination

A modification to make LND more readily usable: Any truth functional


tautology can be written down as a line of a derivation, resting on no Assumption,
with Tautology as the Justification, e.g.

(1) (F ∨ ∼ F ) Tautology

0.4 Prenex normal form


Lemma 1 (change of quantified variable) For any formula in a first-order lan-
gauge, any variable of quantification can be changed to any one of infinitely many
other variables while preserving the free variables of the formula, if there are any, in
a way that results in a formula that is logically equivalent to the original formula.

Remark. There are restrictions on logically equivalent changes of bound vari-


able, e.g. ∃v1 ∀v2 v1 = v2 is not logically equivalent to ∃v2 ∀v2 v2 = v2 (the latter is
logically valid while the former is not).

Lemma 2 (prenex equivalences) The following formulas are logically valid.


(ia) (∼ ∀vi F (vi ) ≡ ∃vi ∼ F (vi ))
(ib) (∼ ∃vi F (vi ) ≡ ∀vi ∼ F (vi ))
In the following the variable vi does not occur free in the formula G.
(iia) ((∀vi F (vi ) ∨ G) ≡ ∀vi (F (vi ) ∨ G))
(iia′ ) ((G ∨ ∀vi F (vi )) ≡ ∀vi (G ∨ F (vi )))
(iib) ((∃vi F (vi ) ∨ G) ≡ ∃vi (F (vi ) ∨ G))
(iib′ ) ((G ∨ ∃vi F (vi )) ≡ ∃vi (G ∨ F (vi )))
(iiia) ((∀vi F (vi ) ∧ G) ≡ ∀vi (F (vi ) ∧ G))
(iiia′ ) ((G ∧ ∀vi F (vi )) ≡ ∀vi (G ∧ F (vi )))
(iiib) ((∃vi F (vi ) ∧ G) ≡ ∃vi (F (vi ) ∧ G))
(iiib′ ) ((G ∧ ∃vi F (vi )) ≡ ∃vi (G ∧ F (vi )))
(iva) ((∀vi F (vi ) ⊃ G) ≡ ∃vi (F (vi ) ⊃ G))
(iva′ ) ((G ⊃ ∀vi F (vi )) ≡ ∀vi (G ⊃ F (vi )))
(ivb) ((∃vi F (vi ) ⊃ G) ≡ ∀vi (F (vi ) ⊃ G))
(ivb′ ) ((G ⊃ ∃vi F (vi )) ≡ ∃vi (G ⊃ F (vi )))
LECTURE 0 13

Proof.

Theorem 3 (Prenex Normal Form) Each formula F in a first-order language


with ∀ and ∃ is logically equivalent to a formula F ∗ whose quantifiers all occur as a
string at the beginning of the formula, and such that F ∗ contains exactly the same
free variables as F does.

Proof The proof is by induction on the number of quantifiers in F .


Base case: F is atomic. Then F has no quantifiers so, vacuously, all its quanti-
fiers, namely none, occur in a string at the beginning of the formula.
Induction steps: (i) F is of the form ∼ G. By induction hypothesis G has a
prenex normal form G∗ . If G∗ has no quantifiers, we are done, as in the base case. If
G∗ has quantifiers, they occur in a string at the beginning of the formula. Then by
logical equivalences (ia) and (ib) from Lemma 2, the negation sign can be pushed
past all the quantifiers, changing each to the other quantifier in the process.
(ii) F is of the form (G ⊃ H). By Induction Hypothesis G and H have logically
equivalent prenex normal forms G∗ and H ∗ . Then F is logically equivalent to (G∗ ⊃
H ∗ ). By Lemma 1 and (iva), (ivb), and (iva′ ), (iv′ ) of Lemma 2, the quantifiers of
G∗ and H ∗ can be pulled out into prenx normal form, those from G∗ changing to
the other quantifier.
(iii) F is of the form (G ∧ H), (iv) F is of the form (G ∨ H). These two cases
are as (ii) except simpler, with no changes of quantifiers as the quantifiers are put
into prenex form. (v) F is of the form ∀vi G. Then by induction hypothesis G has
prenex normal form G∗ which has the same free variables as G has, particular vi
(in the case when the quantification ∀vi of F is not vacuous. Then F is logically
equivalent to ∀vi G∗ , which is in prenex normal form. (vi) F is of the form ∃vi G.
Same argument as for (v). N

Note that the Prenex Normal Form Theorem holds only on the assumption that
all domains of interpretation are non-empty. Otherwise we have, for example, that
for v1 not free in G, and G true in the empty domain (e.g. G = ∀v1 v1 = v1 , or
equally G = ∀v1 ∼ v1 = v1 ),
(∀v1 F (v1 ) ⊃ G) is true in the empty domain, since G is true, but ∃v1 (F (v1 ) ⊃ G))
is false, since every existentially quantified statement is false in the empty domain.

Note that prenex normal forms are not in general unique, e.g.
(∀v1 F (v1 ) ⊃ ∀v2 G(v2 ) has as prenex normal form both ∃v1 ∀v2 (F (v1 ) ⊃ G(v2 )) and
∀v2 ∃v1 (F (v1 ) ⊃ G(v2 )).
LECTURE 0 14

0.4.1 Model theory and proof theory

0.5 Completeness of a system of natural deduc-


tion with respect to first-order logical conse-
quence
0.5.1 Lindenbaum’s Lemma
0.5.2 The Completeness Theorem
The completeness theorem for a given formal system of first-order logic means that
a model theoretic argument for a logical consequence (as above) establishes the
existence of a formal derivation of that logical consequence in the system whose
completeness has been proved without actually finding a derivation.
While we talk about the completeness theorem for first order logic, there are ac-
tually many completeness theorems, one for each different complete formal system
for first-order logical consequence. On the other hand there is an intrinsic com-
pleteness theorem, namely completeness of a system which consists of exactly those
axioms and rules of inference needed for the proof of the completeness theorem.
Comparison between the completeness theorem for first-order logical consequence
and the incompleteness theorem for truth in arithmetic. Both these results are
due to Gödel, in 1930 and 1931. (The first was his Ph.D. thesis, the second his
Habilitation thesis. They were written under the nominal supervision of the very
notable mathematician Hans Hahn, but Gödel was essentially working on his own.)

0.5.3 The Compactness Theorem for first-order logic


0.5.4 The existence of non-standard models of the truths of
arithmetic
0.5.5 Completeness of other systems of derivation with re-
spect to first-order logical consequence
Definition 2 Let S1 and S2 be formal systems such that the language L1 of S1 is a
sub-language of the language L2 of S2 . We say that S1 is a subsystem of S2 if for
every formula φ in L1 , if S1 ⊢ φ, then S2 ⊢ φ

Lemma 4 If S1 is a complete system for first-order logical consequence and S1 is a


subsystem of S2 , then S2 is complete for first-order logical consequence.
Lecture 1

Introduction: a weak form of


Gödel’s First Incompleteness
Theorem; the symbols and
expressions of a language for
arithmetic LE ; Gödel numbering
of the expressions of LE

(Tuesday, 12 October 2010)

1.1 Introduction: a weak form of Gödel’s First


Incompleteness Theorem
The context of Gödel’s discovery of the phenomenon of formal incompleteness is
David Hilbert’s programme for giving mathematics a secure foundation by estab-
lishing the consistency of systems formalizing it. In 1918 he declared that

we must make the concept of specifically mathematical proof itself into


an object of investigation (Hilbert [2]).

Hilbert formulated the distinction between finitary and infintary mathematics.


The paradigm of finitary mathematics is arithmetical calculation. Finitary mathe-
matics is mathematical bedrock, corresponding to observation statements in science.
A calculation such as 27 = 128 is finitary, but the claim that exponentiation to the

15
LECTURE 1 16

power 2 always yields a value, i.e. ∀x∃y(2x = y) is infinitary, and more generally,
quantification over the infinite domain of natural numbers is infinitary. However,
quantification over a bounded, i.e. initial segment of the natural numbers, which is
finite, belongs to finitary mathematics.
Hilbert’s deep insight was to recognize that the formal manipulation of all sym-
bols, not just the symbols for numbers, i.e. numerals and terms built up from
numerals and symbols for arithmetical operations, belongs to finitary mathematics.
In particular,

a formalized proof, like a numeral, is a concrete and surveyable object.


([3], p. 383 and also in [4], p. 471.)

Hilbert recognized two sorts of finitary statements, general and particular (though
he did not introduce terminology for this distinction). Particular finitary statements
are decided by computations, e.g. 7 × 5 = 35, and 210 = 1024 and truth functional
combinations of them (the truth values of such combinations being computable from
the truth values of the component statements). General finitary statements contain
free variables, and can be thought of as a template for particular finitary state-
ments that result by substitution of numerals for the free variables, for example
x + y = y + x, and (n > 2 ⊃ xn + y n 6= z n ). On the other hand, ∀x∀y x + y = y + x
and ∀n∀x∀y∀z(n > 2 ⊃ xn + y n 6= z n ) are infinitary.
For F (v1 ) a general finitary statement with free variable v1 , bounded quantifica-
tion on the variable v1 , which is finitary, is expressible using (apparently) unbounded
quantification by, in the case of universal quantification, ∀v1 (v1 ≤ t ⊃ F (v1 )), for
t a term in the language of arithmetic, which we abbreviate as (∀v1 ≤ t)F (v1 ),
and in the case of existential quantification, ∃v1 (v1 ≤ t ∧ F (v1 )), which we abbre-
viate (∃v1 ≤ v2 )F (v1 ). For t a numerical term (a numeral or a composition of
arithmetical functions applied to numerals), (∃v1 ≤ t)F (v1 ) and (∀v1 ≤ t)F (v1 ) are
particular finitary statements if v1 is the only free variable in F (v1 ). If t is a free
variable or a composition of arithmetical functions applied to one or more variables,
(∃v1 ≤ t)F (v1 ) and (∀v1 ≤ t)F (v1 ) are general finitary statements.
Hilbert noted that general finitary statements are not closed under negation, i.e.
the negation of a general finitary statement cannot be expressed as a general finitary
statement. For example, Fermat’s Last Theorem is expressible as a general finitary
statement, (n > 2 ⊃ xn + y n 6= z n ), but to say (falsely) that Fermat’s Last Theorem
is false requires existential quantification, ∃n∃x∃y∃z(n > 2 ⊃ xn + y n 6= z n ). On
the other hand, the statement that a specific quadruple of numbers a, b, c, d is a
counterexample, i.e. (a > 2 ∧ an + bn = cn ), is a particular finitary statement.
Statements about particular formal proofs are, as Hilbert recognized, finitary
statements, e.g. that a particular formal proof is or is not a proof of a particular
statement.
LECTURE 1 17

Hilbert missed something about his insight which Gödel realized, namely that
formal proofs can be literally identified with natural numbers, i.e. they could be
taken to be numerical expressions, rather than merely like them. As Gödel put this
point in (1931),

Of course, for metamathematical considerations it does not matter what


objects are chosen as primitive signs, and we shall assign natural numbers
to this use, that is, we map the primitive signs one-to-one onto some
natural numbers.

Numbers assigned to formulas of a formal language in this way are called Gödel
numbers.

Definition 3 We denote the Gödel number of a formula F by pF q.

To carry out the arithmetization of syntax, the system must be able to ”talk”
about numbers, i.e. there must be for each natural number a formal numeral in the
language of the system that denotes that numbers.

Definition 4 For formal languages that have a numeral for each natural number,
we denote by n the numeral for the natural number n.

Definition 5 Whenever in these notes I speak of a sentence in the specified language


of arithmetic as being true, I mean true in the usual (intended) structure consisting
of the domain of natural numbers with the usual arithmetical functions and relations
on the natural numbers (also known as the standard model).

What Gödel showed is that the property of being (the Gödel number of) a
provable formula is expressible within any system which can express basic arithmetic,
i.e. there is a formula P r(v1 ) with one free variable, in the language of a formal
system for arithmetic, S, such that for every formula X in the language of S, S ⊢ X
if and only if P r(pXq) is true. We shall establish the existence of such a formula for
a particular formal system of arithmetic in Lecture 4. Gödel also showed that for any
formula with one free variable (in particular a formula that expresses the property
of being the Gödel number of an unprovable formula) there is a diagonal sentence,
i.e for formula F (v1 ) there is a sentence D such that the equivalence (D ≡ F (pDq))
is true. We shall establish this result in Lecture 3.
From these results and on the assumption that everything provable in a given
system S is true (about the natural numbers) (a strong assumption, much stronger
than is needed to establish incompleteness, but it is illuminating to consider this
simple case), it is easy to see that for G such that (G ≡∼ P r(pGq)), S 0 G, and also
thereby that G is true, which implies, from the assumption that everything provable
in a system S is true, that S 0∼ G.
LECTURE 1 18

Theorem 5 (weak form of Gödel’s first incompleteness theorem) Let S be


a theory such that for each natural number n there is a numeral n in the language
of S, and assume that
(i) there is a sentence G such that the sentence (G ≡∼ P r(pGq)) is true;
(ii) there is an assignment of numbers to the formulas of the language of S and
a formula P r(v1 ) in the language of S such that for each formula X, S ⊢ X if
and only if the sentence P r(pXq) is true, so in particular S ⊢ G if and only if the
sentence P r(pGq) is true.
(iii) every theorem of S is true.
Then S 0 G, G is true, and S 0∼ G.

Proof. (1) Suppose that S ⊢ G. (2) Then by (ii), P r(pGq) is true. (3) From
(2) and (i), G is false. (4) From (3) and (iii), S 0 G. (5) Since (4) contradicts (1),
we have by reductio ad absurdum that S 0 G (from assumptions (i), (ii), and (iii)).
(6) From (5) and (ii), P r(pGq) is false. (7) From (6) and (i), G is true.
(8) From (7) and (iii), S 0∼ G. N

Remarks about this result:


(1) This is a weak version of the first Gödel incompleteness theorem since, while
assumptions (i) and (ii) are provable for weak systems of arithmetic, assumption (iii),
soundness of the system with respect to truth in arithmetic, is a strong (highly non-
finitistic) assumption, and unprovability of the Gödel sentence holds from the much
weaker assumption that S is consistent. Gödel sketches the proof of this weak form
of the First Incompleteness Theorem in section 1 of his 1931 paper, and notes that
“The purpose of carrying out the above proof with full precision in what follows is,
among other things, to replace the second of the assumptions just mentioned [every
provable formula is true in the interpretation considered] by a purely formal and
much weaker one.” (van Heijenoort Sourcebook p. 599).
(2) In his introductory section Gödel notes that this argument is “closely re-
lated” to the argument for the “Liar” paradox. The argument does not lead to a
contradiction since it starts from the assumption that G is provable, and so by RAA
establishes that G is not provable. Use of the Liar paradox also shows, as we shall
see, that unlike provability in a formal system, truth in a language of arithmetic
cannot be expressed in the language.
(3) We shall establish how much basic arithmetic is required for arithmetization
of syntax, i.e. to prove assumptions (i) and (ii), to be made precise by the notion
of Σ0 -arithmetic, essentially just computations with addition and multiplication. In
particular we don’t need exponentiation. This shows that arithmetized syntax is a
proper sub-part of what Hilbert meant by finitist mathematics. Hilbert never gave
a precise characterization of finitist mathematics, but it is clear that it includes all
primitive recursive functions, so plus and times, but then also exponentiation and
LECTURE 1 19

so on. On the other had, both addition and multiplication are needed for incom-
pleteness: there is a complete theory of zero, successor, and addition (Presburger
Arithmetic).
(4) The independence of the Gödel sentence from formal arithmetic was unprece-
dented. In the previous hundred years the independence of Euclid’s fifth postulate
from the other postulates of geometry had been established. Gödel’s result differs
from this earlier one in two crucial ways. One was the technique used. The result
concerning the fifth postulate was established by the construction of models. The
Gödel result is purely syntactic (exploiting Hilbert’s insight). The other difference
is even more fundamental. The fifth postulate is neither true nor false, per se. It
is true in Euclidean geometry and false in non-Euclidean geometries. The Gödel
sentence is demonstrably true, though not demonstrable in the system for which it
is constructed.

Proposition 6 For any system S and sentence X, a proof that S 0 X from the
assumption of consistency is best possible.

Proof. If S is inconsistent, it proves everything, so in particular S ⊢ X.

Remark. That the Gödel sentence G for a system S is not refutable, i.e. ∼ G
is not provable, requires a stronger condition on S than consistency but a condition
much weaker than the soundness of S is sufficient.

1.2 The symbols and expressions of a language


for arithmetic LE
A formal language is generated by combining symbols from a specified finite alpha-
bet. For our formal language for arithmetic, as in [7], this consists of the following
13 symbols:
0 ′ ( ) f ′ v ∼ ⊃ ∀ = ≤ ♯
These formal symbols will be used with the following intended meanings:
The symbol 0 denotes the natural number zero1 .
The symbol ′ denotes the successor function
The symbols ( and ) are left and right brackets
The symbols f and v are for functions and variables, to which numerical sub-
scripts in tally notation, i.e. iterations of the subscript ′ are attached. The strings of
1
Note that in this sentence I am being casual about the distinction between use and mention.
That distinction is easily but cumbersomely dealt with by using quotation marks, in which case
this given sentence would read: “The symbol ’0’ denotes the natural number zero”, which is fine,
though fussy, but the next sentence becomes just about unreadable if use vs mention is spelled out
in this way.
LECTURE 1 20

symbols f′ , f′′ , f′′′ will denote the functions addition, multiplication, and exponenti-
ation, respectively, which we will write informally as +, ·, and exp or xy in the usual
notation. There are an infinity of variables v′ , v′′ , v′′′ , . . ., which we will usually write
as v1 , v2 , v3 , . . .. If we want to signify a variable without specifying which variable it
is, we will write vi , vj etc or use informal variable letters x, y, z, u, v, w.
The symbol for the propositional connectives negation and implication are ∼
and ⊃. The symbol for the universal quantifier is ∀.
The symbols = and ≤ are for the two-place relations of equality and less than
or equals.
The symbol ♯ will be used to mark breaks between strings of symbols that are
terms and formulas of the language (to be defined in the next lecture) in sequences
of terms and formulas.
An expression in the language is (almost) any finite string of these symbols.
The set of expressions for the language LE is specified by the following recursive
definition.

Definition 6 (expressions) basis: Each one of the symbols 0 ′ ( ) f ′ v ∼ ⊃


∀ = ≤ ♯ is by itself an expression.
recursion: If Ei and Ej are expressions, and Ei 6= ′ , then the result of writing Ei
directly followed by Ej , which we call the concatenation of Ei and Ej and symbolize
as EibEj , is an expression.

Remark: It’s for a technical reason (to do with our choice of Gödel numbering)
that we excluded from the class of expressions as here specified strings of more than
one symbol that begin with the symbol ′ .

1.3 Gödel numbering of the expressions of LE


.
We assign Gödel numbers to the expressions of LE . This can be done in infinitely
many ways. The way we shall do it, following Smullyan following Quine, makes the
link between Gödel numbering of expressions as strings of formal symbols particu-
larly transparent. Gödel’s original method involved coding by exponents of prime
factors. On our method each number is the Gödel number of an expression, while
on Gödel’s method not every number is a Gödel number. Having every number be
a Gödel number makes the formulation of some results a little simpler but is not
essential.

Definition 7 (notation for expressions and Gödel numbers) En =df the ex-
pression with Gödel number n.
pEq =df the Gödel number of expression E.
LECTURE 1 21

Corollary 7 (of Definition 7) pEn q = n.

We are used to the idea that numbers are denoted by numerals and that numerals
are not the same thing as numbers. The Roman numerals for the first five non-zero
natural numbers are I, II, III, IV, V, while the Arabic numerals are 1, 2, 3, 4 , 5. The
crucial property of the Arabic numerals is that they are constructed on a place-value
system with a base of 10. That the system of numerals in common use is base 10 is
presumably down to the contingent fact (it could have been otherwise) that human
beings have 10 fingers. Any other number greater or equal to 2 gives a perfectly
good numeral system with that base. Base 2 is used in machine code for computers,
with 0 and 1 represented by current off and current on. The number we write as 15
in base 10 we write as 1111 in base 2 and as 13 in base 12. In the formal language for
arithmetic we shall be using the numerals for numbers use a tally notation, rather
n
z }| {
than place values. The formal numeral for the number n is the expression 0′ . . . ′ ,
i.e. concatenation of the symbol 0 with n-many concatenations of the symbol ′ .
The following function plays a key role in our chosen system of Gödel numbering.

Definition 8 (concatenation of base b numerals) For natural numbers m and


n, we denote by m ∗b n the number designated by the base b numeral that results from
concatenating the base b numeral for m with the base b numeral for n.

Note that ∗b is a function mapping pairs of natural numbers to natural numbers,


and that natural numbers are not intrinsically in base b or any other base notation.
The role of b in this function is to specify the method of calculating this function.
Examples: For m = 673, n = 32 (written in base 10), m ∗10 n = 67332 and
n ∗10 m = 32673. For m = 59, n = 0, m ∗10 n = 590 and n ∗10 m = 059 = 59.
Remark : As illustrated by these examples, ∗b is not commutative. It is also not
associative, e.g. (17∗b 0)∗b 59 = 17059 6= 1759 = 17∗b (0∗b 59). Non-associativity only
arises when the middle value is 0, but since we will include 0 as a Gödel number we
cannot suppress parentheses in multiple computations with ∗b except by adopting a
convention for reinstating them; we adopt the common convention of association to
the left, i.e. x ∗b y ∗b z = (x ∗b y) ∗b z.
We assign Gödel numbers to expressions by first stipulating the Gödel numbers
of the symbols. We assign to these thirteen symbols the numbers denoted by the
thirteen digits of base 13 notation, where the digits for 10, 11, and 12 (as we write
them in base 10) are taken to be η, ǫ, and δ, respectively.

Definition 9 (assignment of Gödel numbers to expressions) By recursion over


the recursive definition of expressions.
Base case: The assignment of numbers to symbols is specified by
LECTURE 1 22


0 ( ) f ′ v ∼ ⊃ ∀ = ≤ ♯
1 0 2 3 4 5 6 7 8 9 η ǫ δ

Recursion: For expressions X and Y , pXbY q = pXq ∗13 pY q

By these stipulations, p′ q = 0, p′ . . .′ q = 0, and p′ ∀q = 09 = 9 = p∀q. It is in


order for each expression to have a unique Gödel number that we stipulated above
that the class of expressions does not contain strings of more than one symbol that
begin with the prime symbol ′ .
There is a technical advantage in taking the base b to be a prime number but
it is by no means essential. We can use base 10 and the operation ∗10 even with
thirteen symbols by, for example, assigning the thirteen symbols respectively the
following numbers (written in base 10):

0 ( ) f ′ v ∼ ⊃ ∀ = ≤ ♯
1 0 2 3 4 5 6 7 89 899 8999 89999 899999

Of course on this assignment not every number is a Gödel number. But we can
effectively tell the ones that are, i.e. we know that if an 8 or a 9 occurs its base 10
notation it must occur within a string of the form 89, 899, 8999, 89999, 899999, and
we know which symbol is coded by counting the number of 9s in that string.
Lecture 2

Terms and formulas of the


language LE ; expressibility of
diagonal substitution in the
language LE

(Wednesday, 13 October 2010)

2.1 Terms and formulas of the language LE


2.1.1 Terms
Terms in the formal language LE are expressions that, on the intended interpretation
of LE as a language for arithmetic, denote a number if the term does not contain a
free variable, and if it contains one or more free variables, then the term that results
from substituting a numerals for each variable denotes a number.

Definition 10 (Variables) v′ is a variable, and if the expression E is a variable


then the expression E′ , i.e. the concatenation of E and the subscript symbol ‘′ ’, is
a variable.

Remark: Formal variables are expressions of the form v′ , v′′ , v′′′ , . . .. We will
abbreviate the string of symbols consisting of the formal variable symbol v followed
by n subscripts as vn .
Further remark: Smullyan’s formal variables are of the form (v′ ), (v′′ ), (v′′′ ), . . .,
i.e. formal variables as we have defined them enclosed in brackets (p. 15). Unique
readability for terms within expressions does not require enclosing variables within

23
LECTURE 2 24

brackets in this way. The motivation for this unnecessary use of brackets might
be an artefact of what’s simple to write in LaTeX. Evidently Smullyan produces
e.g. the substring v′′ of his variable (v′′ ) by the LaTeX code v_{’’}. If this was
v2 , then the seemingly natural way to write LaTeX code for the concatenation of
v2 with ′ would be $v_{’’}^{\prime}$, but this produces v′′′ , in which the ′ sym-
bol occurs interposed over the string v′′ rather than concatenated at the end of
that string. On the other hand, $(v_{’’})^{\prime}$ produces (v′′ )′ , in which ′
is correctly concatenated at the end. Nonetheless, more complicated LaTeX code
(also with compounded subscript command to lower the apostrophe further so it
looks more convincing as a subscript) produces the required concatenation with vari-
ables not enclosed in brackets, namely $v_{_{’’}}$\hspace{-.12ex}$^{\prime}$,
which compiles to the required concatenation of symbols v′′ ′ .

Definition 11 (Numerals) The symbol 0 is a numeral (which denotes the number


zero). If the expression E is a numeral then the expression Eb′ is a numeral (which
denotes the successor of the number denoted by E, so the expressions 0, 0′ , 0′′ , 0′′′ , . . .
are formal names of the natural numbers 0, 1, 2, 3, . . ., respectively).

Notation: For natural number n we write n for the numeral that denotes the
number n, e.g. 7 = 0′′′′′′′ .

Corollary 8 (of Definition 11 and Notation) For any natural number n, n + 1


is n′ , i.e. the numeral for the number n + 1 is the concatenation of the numeral for
the number n and the symbol ′ .

Definition 12 (Terms) Among expressions of LE , the class of terms is specified


by the following recursive definition:
Base clause: Each variable and each numeral is a term.
Induction clauses: If t is a term, then t′ is a term. If t1 and t2 are terms,
then (t1 f′ t2 ), (t1 f′′ t2 ), and (t1 f′′′ t2 ) are terms. [Recall that the notations in LE for
addition, multiplication, and exponentiation are, respectively, f′ , f′′ , and f′′′ .]

Definition 13 (expression E1 occurs in expression E2 ) (i) Every expression oc-


curs in itself.
(ii) Expression E1 occurs in expression E2 if there exist expressions E3 , E4 such
that E2 = E3bE1bE4 or E2 = E3bE1 or E2 = E1bE4

Definition 14 (constant terms) A term in which no variable occurs is called a


constant term, or a closed term.
LECTURE 2 25

2.1.2 Formulas
Definition 15 (Atomic formulas) An atomic formula is any expression of the
form t1 = t2 or of the form t1 ≤ t2 , where t1 and t2 are terms.

Definition 16 (Formulas) The class of formulas is specified by the following re-


cursive definition:
Base clause: Every atomic formula is a formula.
Induction clauses: If F and G are formulas, then ∼ F and (F ⊃ G) are formulas,
and for every variable vi , the expression ∀vi F is a formula. [Note that the formula
(F ⊃ G) is enclosed in brackets, but that the other two formation rules do not
introduce new brackets.]

We will use logical equivalences between conjunction, disjunction, and existential


quantification and expressions in terms of ∼, ⊃, and ∀ as abbreviations, i.e.

Definition 17 For formulas A and B,


(A ∧ B) =df ∼ (A ⊃∼ B);
(A ∨ B) =df (∼ A ⊃ B);
∃vi A =df ∼ ∀vi ∼ A.

2.1.3 Free and bound variables; open formulas and sen-


tences
Definition 18 (a variable occurs free in a formula) (i) If F is an atomic for-
mula, all occurrences of variables in F are free.
(ii) For F and G formulas, if variable vi occurs free in F , then vi occurs free in
∼ F and in (F ⊃ G) and (G ⊃ F ).
(iii) For any formula F , if vi occurs free in F , vi is free in ∀vj F iff j 6= i.

Definition 19 (bound variables) A variable occurs bound in a formula F iff it


occurs in F and does not occur free in F .

Definition 20 (open formulas) A formula with one or more free variables is an


open formula

Notation: We write F (vi ) to signify a formula in which the variable vi occurs


free. Other unspecified variables may occur free as well unless we stipulate that
vi is the only variable free in F (vi ). In the latter case we may say that F (vi ) is a
one-place formula. Similarly we write F (vi1 , . . . vik ) for a formula in which variables
vi1 , . . . vik occur free, possibly with other free variables unless we stipulate that these
are the only ones.
LECTURE 2 26

[Note that this convention on possible occurrence of free variables other than
those explicitly shown is different from Smullyan’s: “We write F (vi1 , . . . vik ) for any
formula in which vi1 , . . . vik are the only free variables.” (p. 16). Since there are
situations, e.g. in stating the Induction axioms, in which we need to allow for the
possibility of other free variables than those explicitly shown, Smullyan’s convention
has in those situations to be violated, e.g. “F (v1 ) is to be any formula at all (it
may contain free variables other than v1 )” (Smullyan, p. 29). It seems to me more
coherent for the convention to allow other variables, which also then allows us under
the same convention to stipulate in a given situation that there are no other free
variables.]

Definition 21 (closed formulas a.k.a. sentences) A formula with no free vari-


ables is a closed formula, also called a sentence.

Notation: Substitution of numerals for free variables in formulas: We write


F (n) to signify the result of substituting the numeral n for every free occurrence
of vi in the open formula F (vi ). If vi is the only variable free in F (vi ), then F (n)
is a sentence. For numbers n1 , . . . , nk , F (n1 , . . . , nk ) signifies the result of substi-
tuting the numerals n1 , . . . , nk for all free occurrences in F (vi1 , . . . vik ) of vi1 , . . . vik ,
respectively. If vi1 , . . . vik are the only variables that occur free in F (vi1 , . . . vik ), then
F (n1 , . . . , nk ) is a sentence.

Definition 22 (regular open formulas) An open formula is said to be regular if


its k-many free variables are the first k variables.

For F (v1 , . . . , vk ) a regular open formula, the expression F (n1 , . . . , nk ) is unam-


biguous; we don’t need to stipulate which number is substituted for which variable.
Note that for F (v1 ) a formula with one free variable, F (n) is a sentence while
F (n) is not. The latter can be construed as shorthand for the statement that the
open formula F (v1 ) is satisfied by the number n, and that statement will be true if
and only the sentence F (n) is true.

2.2 Designation by terms in LE , truth of sentences


of LE , and expressibility of sets and relations
of natural numbers by formulas of LE
2.2.1 Designation
On the intended interpretation of the formal language LE each constant term des-
ignates a particular natural number, by the following recursive specification:
LECTURE 2 27

(i) the numeral n designates the number n.


(ii) If constant term c designates n, then constant term c′ designates the next
natural number after n; if constant terms c1 and c2 designate n1 and n2 respectively,
then (c1 f′ c2 ) designates the sum of n1 and n2 , (c1 f′′ c2 ) designates the product of n1
and n2 , and (c1 f′′′ c2 ) designates n1 raised to the power n2 .

2.2.2 Truth
Truth for a sentence of LE in the structure of the natural numbers (the intended
interpretation) can be defined by recursion over the recursive generation of the
sentence in the usual way. For none of the results in this course do we require a
formal definition of truth, and I will take it as known informally what it means for
a formula in the language of arithmetic to be true in the structure of the natural
numbers.

2.2.3 Expressibility
Definition 23 (expressibility of relations) A formula F (v1 , . . . , vn ) in LE is said
to express a relation R ⊆ Nn iff for every n-tuple < k1 , . . . , kn > of natural numbers
the sentence F (k 1 , . . . , k n ) is true iff < k1 , . . . , kn >∈ R. In such case the relation
R is said to be expressible in LE .

Definition 24 (expressibility of functions) A function f (v1 , . . . , vn ) : Nn → N


is expressible in LE iff the relation f (v1 , . . . , vn ) = vn+1 is expressible in LE .

Definition 25 (Arithmetical) A relation or a function is Arithmetical if it is


expressible by a formula in LE .

Definition 26 (arithmetical) A relation or a function is arithmetical [with a


lower-case “a”] iff it is expressible by a formula in LE in which the expression f′′′
(for exponentiation) does not occur.

2.3 Concatenation of numbers in a given base no-


tation is Arithmetical.
Lemma 9 For a fixed number b ≥ 2, the condition that v1 is a power of b, which
expression we abbreviate as P owb (v1 ), is Arithmetical.

Proof. P owb (v1 ) iff ∃v2 (v1 = bv2 ), or more formally


P owb (v′ ) iff ∼ ∀v′′ ∼ v′ = (bf′′′ v′′ ). N
LECTURE 2 28

Lemma 10 For ℓb (n) the length of the base b notation for n, i.e. the number of
digits in the base b notation of n, the two-place relation bℓb (v1 ) = v2 is Arithmetical
Proof. This relation is expressed by the following condition on v1 and v2 .

((v1 = 0∧v2 = b)∨(v1 6= 0∧P owb (v2 )∧v1 < v2 ∧∀v3 ((P owb (v3 )∧v1 < v3 ) ⊃ v2 ≤ v3 ))).
This equivalence is seen as follows: If v1 = 0, ℓ(v1 ) = 1 and b1 = b. The first
disjunct takes care of this case. If v1 6= 0, then the length of v1 (in base b notation) is
the least power of b > v1 , e.g. ℓ10 (935) = 3, and 103 = 1000 is the least power of 10
> 935. [Note that we need to treat these two cases separately, since ℓ(0) = 1, but—
writing µv3 F (v3 ) for the least v3 s.t. F (v3 )—µv3 (10v3 > 0) = 0 since 100 = 1 > 0.]
That v2 is the least power of b greater than v1 is expressed by the two conditions
that bv2 > v1 , and for any v3 such that bv3 > v1 , v2 ≤ v3 .
The above condition is expressible in LE by Lemma 9 and the fact that v1 < v2
is equivalent to (v1 ≤ v2 ∧ ∼ v1 = v2 ). N
Theorem 11 For any number b ≥ 2, the relation v1 ∗b v2 = v3 is Arithmetical.
Proof. The relation v1 ∗b v2 = v3 is expressed by the condition that
v1 · bℓb (v2 ) + v2 = v3 .
For example, 1570 ∗10 365 = 1570365 = 1570000 + 365 = 1570 · 103 + 365 =
1570 · 10ℓ10 (365) + 365.
This condition is equivalent to
∃v4 (bℓb (v2 ) = v4 ∧ ((v1 · v4 ) + v2 ) = v3 ).
By Lemma 10, this relation is Arithmetical. N

2.4 Substitution and diagonal functions, and their


arithmetization
The operation of substituting a numeral for a free variable lies at the heart of
the construction of a ‘self referential’ arithmetical sentence (‘This sentence is not
provable in the given formal system’). To describe the formula F (n) obtained by
substituting the numeral n for the free variable vi in the formula F (vi ) is complicated
(requiring recursion over the logical complexity of formulas). We follow Smullyan
in utilizing a trick due to Tarski by which we construct a formula F [n] which is not
the same formula as F (n) but is logically equivalent to it for which there is a simple
general description that does not depend on the logical complexity of the formula
F (vi ).
LECTURE 2 29

Definition 27 (quasi-substitution) For F (v1 ) any formula of LE with one free


variable and n any numeral, F [n] =df ∀v1 (v1 = n ⊃ F (v1 )).
Quasi-substitution is logically equivalent to substitution, i.e.
Lemma 12 (∀v1 (v1 = n ⊃ F (v1 )) ≡ F (n)) is logically valid.
Proof. (i) Suppose ∀v1 (v1 = n ⊃ F (v1 )). Then by universal instantiation,
(n = n ⊃ F (n)). Since n = n is logically valid, by modus ponens, F (n).
(ii) Suppose F (n). Then by substitutivity of identity, (v1 = n ⊃ F (v1 )). By
universal generalization, ∀v1 (v1 = n ⊃ F (v1 )). N

Note We could also have defined F [n] as ∃v1 (v1 = n ∧ F (v1 )) since:
Lemma 13 (∀v1 (v1 = n ⊃ F (v1 )) ≡ ∃v1 (v1 = n ∧ F (v1 ))) is logically valid.
Proof. Exercise. N

Our first step in the arithmetization of syntax, i.e. showing that syntactic oper-
ations on expressions of LE can be reflected into arithmetically definable operations
on their Gödel numbers, is to show that the function
s(v1 , v2 ) = p∀v1 (v1 = v2 ⊃ Ev1 )q
is Arithmetical.
The value of the function s(v1 , v2 ) is the Gödel number of a formula logically
equivalent to the substitution of the numeral of the number v2 into the expression
Ev1 when Ev1 is a formula in which v1 occurs free. Note that for some values of v1 ,
Ev1 is a formula in which the variable v1 occurs free, and for other values it is not.
Indeed for some values of v1 , the expression whose Gödel numbers if v1 , i.e. Ev1 ,
will not be a formula, in which case s(v1 , v2 ) is the Gödel number of an expression
which is not a formula—a ‘don’t care’ case. If Ev1 is a formula in which v1 does not
occur free then s(v1 , v2 ) is the Gödel number of a formula that has nothing to do
with substitution of the numeral for v2 into it—another ‘don’t care’ case.
Before we can establish that the two-place function (three-place relation) s(v1 , v2 ) =
v3 is Arithmetical, we must calculate the Gödel numbers of the numerals.
Lemma 14 The Gödel number of the numeral v 2 is 13v2 (in our base-13 assignment
of Gödel numbers ).
v
z }|2 {
Proof. The numeral v2 is the expression 0′ . . . ′ . The symbol 0 is assigned Gödel
number 1, the symbol ′ is assigned the number 0, so that whole expression is, by
v
z }|2 {
concatenation, assigned the number written in base-13 notation as 1 0 . . . 0, which
is 13v2 . N
LECTURE 2 30

Theorem 15 The function s(v1 , v2 ) = p∀v1 (v1 = v2 ⊃ Ev1 )q is Arithmetical.

Proof. Let k = p∀v1 (v1 =q, a particular number whose base 13 notation—
given our assignment of base 13 digits to the symbols of our language—is 965265η
(or if we use the base 10 Gödel numbering also given in Lecture 1, whose base
10 notation is 899652658999). Given that pv2 q = 13v2 , p⊃q = 8, pEv1 q = v1 ,
p)q = 3, s(v1 , v2 ) = v3 iff ∃v4 (v4 = 13v2 ∧ v3 = k ∗ v4 ∗ 8 ∗ v1 ∗ 3), which by left-
hand association of ∗13 and repeated use of Theorem 11 is Arithmetical, i.e. there
is a formula S(v1 , v2 , v3 ) such that for all numbers m, n, k, S(m, n, k) is true iff
s(m, n) = k. Expressing this argument from Theorem 11 more strictly, the formula
needs to be the following:

∃v4 ∃v5 ∃v6 ∃v7 (v4 = 13v2 ∧ k ∗ v4 = v5 ∧ v5 ∗ 8 = v6 ∧ v6 ∗ v1 = v7 ∧ v7 ∗ 3 = v3 ) N

Definition 28 (diagonal substitution) The diagonal substitution function d(v1 )


is s(v1 , v1 ), i.e. d(v1 ) =df p∀v1 (v1 = v1 ⊃ Ev1 )q.
Remark. Since by Definition 27, ∀v1 (v1 = x ⊃ Ev1 ) =df Ev1 [v1 ], Definition 28
means that d(v1) = pEv1 [v1 ]q.

Corollary 16 (corollary to the proof of Theorem 15) The relation d(v1 ) = v2 is


Arithmetical.

Proof. Let D(v1 , v2 ) be the formula that results by substituting v1 for v2 and
v2 for v3 in S(v1 , v2 , v3 ). For all numbers m, n, S(m, m, n) is true iff s(m, m) = n.
By the definition of d(v1 ), s(m, m) = d(m). So D(m, n) is true iff d(m) = n. N
Lecture 3

The Diagonal Lemma;


expressibility of properties of
sequence numbers

(Tuesday, 19 October 2010)

3.1 The Diagonal Lemma


Theorem 17 (The Diagonal Lemma) For each one-place formula F (v1 ) in LE ,
there exists a sentence C in LE such that the equivalence (C ≡ F (pCq)) is a true
sentence in LE .
Proof. (1) Let D(v1 , v2 ) be the formula in LE from the proof of Corollary 16
that expresses the relation d(v1 ) = v2 .
(2) Let k =df p∀v2 (D(v1 , v2 ) ⊃ F (v2 ))q.
(3) Let C =df ∀v1 (v1 = k ⊃ ∀v2 (D(v1 , v2 ) ⊃ F (v2 ))).
(4) By Lemma 12, (C ≡ ∀v2 (D(k, v2 ) ⊃ F (v2 )).
(5) By (1), (∀v2 (D(k, v2 ) ⊃ F (v2 )) ≡ ∀v2 (d(k) = v2 ⊃ F (v2 ))).
(6) By Lemma 12, (∀v2 (d(k) = v2 ⊃ F ((v2 ))) ≡ F (d(k))).
(7) By the chain of equivalences (4), (5), (6), (C ≡ F (d(k))).
(8) By (2) and (3) and Definition 28, pCq = d(k) (fuller explanation below).
(9) From (7) and (8) by substitutivity of identity, (C ≡ F (pCq)). N
Explanation of step (8): k is the Gödel number of a formula with one free variable,
C is the formula that results from the diagonal substitution of quasi-substituting
into that formula its own Gödel number, and the one-place function d(v1 ) generates
from the Gödel number of a formula with one free variable the Gödel number of the
formula that results from diagonal substitution into that formula.

31
LECTURE 3 32

The proof of the Diagonal Lemma is by a kind of double substitution into the one-
place formula for which a diagonal sentence is being established, first the substitution
of the Arithmetical expression of the diagonal function, and then the substitution
of the numeral for the Gödel number of the formula that results from that first
substitution. Both of these substitutions are quasi-substitutions in this construction.
It might make the idea of the proof more perspicuous if we consider how it goes
with actual substitutions if we have a term s(v1 ) in our language such that s(v1 ) =
pEv1 (v1 )q. (We could have such a language, at the cost of taking more functions as
primitive.)
Theorem 18 (variant diagonal lemma) Given a formula F (v1 ) with one free
variable in a language for arithmetic L that has a term s(v1 ) such that for each
number n, s(n) = pEn (n)q, there is a sentence C in L such that (C ≡ F (pCq)) is
true.
Proof. Consider the formula F (s(v1 )) formed by substituting the term s(v1 ) for
the free occurrences of v1 in F (v1 ). (1) Let k = pF (s(v1 ))q. (2) Let C =df F (s(k)).
(3) Then s(k) = pCq. (4) The numeral s(k) designates the same number as is
designated by the term s(k), i.e. the equation s(k) = s(k) is true, so by substitutivity
of identity, (F (s(k)) ≡ F (s(k))). (5) Hence by (2) and (3), (C ≡ F (pCq)). N

3.2 Expressibility of properties of sequence num-


bers in the language LE
3.2.1 Properties of sequences of digits
The first tool we need in order to code sequences of numbers in LE is to show that
we can express in LE the relations that the base b notation of a number m begins,
or ends, or is part of the base b notation of a number n.
Definition 29 (x begins y) x begins y in base b notation iff the base b notation
of x is a (not necessarily proper) initial segment of the base b notation of y. We
write this as xBb y.
Examples. In base 10, 2 begins 20, but note that in base 13, 2 (as it is written
in base 10, and in base 13) does not begin 20, since 20 base 10 is written 17 base
13. Other examples (base 10): The numbers which written in base 10 are 7, 76, 760,
7600, 76007, 760074, and 7600748 all begin 7600748 in base 10. The last of these
examples points up the fact that every number begins itself, i.e. an initial segment
need not be a proper initial segment. Note that the number 0 does not begin any
number except itself, i.e. we don’t say that 0 begins 760748, even though 0760748
= 760748.
LECTURE 3 33

Definition 30 (x ends y) x ends y in base b notation, which we write as xEb y,


if the base b notation of x is an end segment (not necessarily proper) of the base b
notation of y.
Examples. In base 10, the following numbers all end 7600748: 7600748, 600748,
748, 48, 8.
Given the notion of one number beginning another in a given base representation
and the notion of one number ending another in a given base, we can define the
notion of a number being part of another in a given base in terms of these two
notions:
Definition 31 (x is part of y) x is part of y, in base b notation, which we write
as xPb y, if x ends some number that begins y.
Remark. Every number is a base b part of itself. Given a base b notation for
x, every proper sub-segment of the base b notation that does not begin with a 0 is
the base b notation of a number y that is a base b part of x. In base 10, the parts
of 2600748 are all the numbers that begin or that end it, and 60074, and all the
numbers that begin or end it, and 7.
Theorem 19 For any b ≥ 2 the following relations are Arithmetical: (1) xBb y, (2)
xEb y, (3) xPb y and, for any natural number n ≥ 2, (4) x1 ∗b . . . ∗b xn Pb y
Proof. We will prove a stronger result, which we need later, that these relations
not only are expressible in LE , but also that this can be done using only bounded
quantifiers, i.e. that these are finitary properties of numbers.
1. If 0 does not occur in the base b numeral for y, then x begins y just in case
there exists z such that x ∗b z = y. However, if a zero or a string of zeros occurs
in the base b numeral for y and the base b numeral for x is an initial segment of
the base b numeral of y which ends just before the 0 or string of 0s in the base b
numeral for y or includes some but not all of those 0s, then the numeral of x has
to be extended by the remaining 0s before it can be concatenated with a numeral
to result in the numeral for y. The extension of the base b numeral for x by the
required number of 0s is accomplished by multiplying x by b raised to the power of
how many 0s need to be appended. This condition can be expressed in terms of the
previously expressed notions P owb (w) and x ∗b z = y, as follows:
xBb y iff (x = y ∨ (x 6= 0 ∧ (∃z ≤ y)(∃w ≤ y)(P owb (w) ∧ (x · w) ∗b z = y)))
The bounds on the quantifiers hold from the fact that if z is part of y, then
z ≤ y, and any number of the form 10...0 in base b with a string of 0s of a string of
0s in y is ≤ y.
2. xEb y iff (x = y ∨ (∃z ≤ y)(z ∗b x = y)
For this case we don’t have any complications from the occurrence of zeros.
3. xPb y iff (∃z ≤ y)(xEb z ∧ zBb y)
4. x1 ∗b . . . ∗b xn Pb y iff (∃z ≤ y)(x1 ∗b . . . ∗b xn = z ∧ zPb y). N
LECTURE 3 34

3.2.2 Sequence numbers


Treating sequences of expressions as expressions with Gödel numbers is very conve-
nient. It requires making sequences of expressions into single expressions, which is
done by expanding the langauge of arithmetic by introducing the symbol ♯ to mark
out the beginning and end of a sequence of expressions and the boundary between
two successive expressions. Thus among the primitive symbols of LE (Lecture 1) is
♯, which did not enter into the rules for the formation of terms and formulas of LE .
The use of this symbol is to allow us to concatenate a finite sequence of formulas of
LE into a single expression of the language, by serving as a marker between different
formulas in a sequence of formulas. When that is done, a sequence of formulas, as
an expression, will have a Gödel number.

Definition 32 (sequence number) x is a sequence number if it is the Gödel


number of an expression of the form ♯Ei1 ♯Ei2 ♯ . . . Eik ♯ in LE where each expression
Eij does not contain the symbol ♯,

3.3 Coding of finite sequences of Gödel numbers


Recall the assignment of the first 12 digits of base 13 representation of natural
numbers to the 12 symbols that enter into formulas of the language LE :

0 ( ) f ′ v ∼ ⊃ ∀ = ≤
1 0 2 3 4 5 6 7 8 9 η ǫ

Thus if a number is the Gödel number of a formula in LA on the particular


Gödel numbering we have adopted, then the 13th digit, δ, will not occur in its base
13 representation. Call the class of such numbers Nδ .
A formal proof is a finite sequence of formulas, so to code a proof by a number
it suffices to find a way of coding finite sequences of numbers in Nδ . We code such a
sequence (a1 , . . . , an ) by the number δ ∗13 a1 ∗13 δ ∗13 a2 ∗13 δ ∗13 . . . ∗13 δ ∗13 an ∗13 δ. In
future I shall mostly suppress the explicit notation of base 13 (or more generally base
b for any b ≥ 2) concatenation and write v1 = v2 v3 for v1 = v2 ∗b v3 , i.e. symbolize
the concatenation relation by concatenation itself.
There are several points about the concatenation relation that need to be borne
in mind. (1) It is a three place relation and not a two-place function. (2) It is
a relation between numbers, and numbers are expressible in base b notation for all
b ≥ 2 but are not in base b notation. The situation is similar to what it is in number
theory generally. When we compute with natural numbers we do so using their base
10 notation (or in the case of computers, base 2 notation). But when we prove
something about numbers what we prove is proved using properties of numbers that
LECTURE 3 35

are not specific to decimal notation, even if what is being proved specifically refers
to decimal notation, as in “if the digits of a number in its base 10 notation add up
to 9 then the number is divisible by 9”, which is proved from general properties of
the congruence relation and the fact that 10 ≡ 1 (mod 9).

Proposition 20 (sequence numbers) A natural number n is a sequence number


iff n = δa1 δa2 δ . . . δan δ for ai ∈ Nδ .

Proof. Immediate from the definition of sequence number above.

Proposition 21 The property of being a sequence number, Seq(v1 ), is Arithmetical,


i.e. it is expressible in LE .

Proof. The property of being a sequence number is expressible by the following


formula:

(δBv1 ∧ δEv1 ∧ δ 6= v1 ∧ ∼ δδP v1 ∧ (∀v2 ≤ v1 )(δ0v2 P v1 ⊃ δBv2 ))


The first four conjuncts characterize the required occurrences of the digit δ in the
base 13 representation of v1 . The last conjunct rules out the occurrence of a string
of zeros of length greater than one. The reason for this requirement is that sequence
numbers code sequences of numbers in Nδ . Since 00 = 0, if, e.g., δ00δ were allowed
as a sequence number it would code the sequence (0) but which is also coded by
δ0δ. Since δ00δ 6= δ0δ, if we allowed, e.g. δ00δ as a sequence number, any sequence
which includes the number zero would have more than one (indeed infinitely many)
sequence numbers. N

Definition 33 For v1 a sequence number, we say that v2 is in v1 , symbolized as


v2 ∈ v1 , iff v2 is one of the numbers coded by v1

Proposition 22 v2 ∈ v1 is Arithmetical.

Proof. v2 ∈ v1 iff (Seq(v1 ) ∧ δv2 δP v1 ∧ ∼ δP v2 ). It is a necessary condition for


v2 ∈ v1 that δv2 δP v1 but not sufficient since numbers of the form δa1 δa2 δ satisfy it;
the condition ∼ δP v2 rules out those cases. N
In expressing the condition that a number is the Gödel number of a proof in
a formal system we need to be able to express the condition that one part of a
sequence number occurs earlier in the sequence than another. We do this as follows.

Definition 34 v2 ≺ v3 iff v1 is the sequence number of a sequence in which v2


v1
and v3 occur and the first occurrence of v2 in the sequence is earlier that the first
occurrence of v3 in the sequence.
LECTURE 3 36

Proposition 23 The three-place relation v2 ≺ v3 is Arithmetical.


v1

Proof. v2 ≺ v3 iff (v2 ∈ v1 ∧ v3 ∈ v1 ∧ (∃v4 ≤ v1 )(v4 Bv1 ∧ v2 ∈ v4 ∧ ∼ v3 ∈ v4 ))


v1
Note that the formulas v2 ∈ v1 and v3 ∈ v1 each contains the condition Seq(v1 ),
so we don’t need a separate conjunct Seq(v1 ). N
Lecture 4

A formal system PAE for


arithmetic; an Arithmetical proof
predicate for PAE ; a weak version
of Gödel’s first incompleteness
theorem for PAE

(Wednesday, 20 October 2010)

4.1 A formal system PAE for arithmetic


We now begin the investigation of formal first-order axiomatic systems of arithmetic.
A theory is first-order if its formal language is first-order. A formal language is first-
order if its quantifiers range only over the objects in its domain of interpretation
and not over collections (pluralities) of those objects. A second-order language has
quantifiers that range over collections (pluralities) of objects (possibly also over
relations between objects). Formal systems of first-order logic are complete (which
Gödel proved in 1930, in his doctoral thesis–the incompleteness theorem was his
Habilitation thesis). There is no complete axiomatization of full second-order logic.
Accordingly, where we are interested in properties of what can, or cannot, be proved
in formal systems, we shall be concerned only with first-order systems. In this course
we will investigate properties of a number of different formal systems for arithmetic.
All the formal systems we investigate in this course
will explicitly or by assumption contain a com-

37
LECTURE 4 38

plete axiomatization of classical first-order logic


with identity.
We now set out a formal axiomatic system for arithmetic, PAE in the language
LE . ‘PA’ stands for Peano Arithmetic, a standard misnomer1 The subscript E
signifies that in this system exponentiation is taken as primitive, i.e. it is governed
by its own axioms. In Lecture 5 we shall see that exponentiation need not be taken
as a primitive and that via coding of ordered pairs of natural numbers the relation
xy = z can be expressed in terms of zero, successor, addition, and multiplication.
We shall follow Smullyan in the specification of PAE . The axioms are in four
groups, the first two of which, with two rules of inference, is a formal system of
first-order predicate logic, and the third and fourth groups are axioms specific to
arithmetic.
Group I are axiom schemata for propositional logic: These are the standard
axioms for propositional logic with ∼ and ⊃ as primitive. L1 and L2 are exactly the
axioms for ⊃ required to establish the Deduction Theorem (taken as proved in a
previous course). L3 establishes the classical logic of ∼. These axioms are complete
for truth-functional validity.
Group II are axiom schemata for first-order predicate logic. The axiomatization
of first-order predicate logic by the Group II schemata is highly unnatural in terms
of establishing formulas as logically valid. Its virtue for us is that it is very easy to
arithmetize since it involves no substitution of terms for free variables, which more
natural axiomatizations of predicate logic with identity do. Note that L6 strictly
should be written ∼ ∀vi ∼ vi = t. To prove the following valid formulas from these
schemata is non-trivial: vi = vi , (vi = vj ⊃ vj = vi ), (∀vi F (vi ) ⊃ F (t)) for t any
term of LE not containing a variable that is quantified in F (vi ) within the scope
of which vi occurs. For proofs of these formulas from these schemata see Donald
Kalish and Richard Montague, “On Tarski’s formalization of predicate logic with
identity”, Archiv für mathematische Logik und Grundlagenforschung 7 (1965), pp.
81-101, Lemmas 2, 3, and 8 on pp. 85-87.
Group III are axioms specific to each of the primitive non-logical notions of the
language
Group IV is all instances of (a version of) the axiom schema of induction. A
usual formulation of the induction schema is

(F (0) ⊃ (∀v1 (F (v1 ) ⊃ F (v1′ )) ⊃ ∀v1 F (v1 )))

.
1
It was Dedekind who established the first axiomatization of arithmetic, in 1888, which Peano
took over in his publication a year later. Peano cites Dedekind 1888 as the source of his axioms.
It seems to have been Russell who introduced the misnomer Peano Arithmetic.
LECTURE 4 39

However, two formulas within this schema are generated by substitution, namely
F (0) and F (v1′ ), and for ease of arithmetization we want to use quasi-substitution
instead of substitution. We can’t use quasi-substitution directly on F (v1 ) to express
F (v1′ ), since F [v1′ ] would be ∀v1 (v1 = v1′ ⊃ F (v1 )), in which no variable occurs
free, so not equivalent to F (v1′ ). We could change the variable in the auxiliary
quantification, say to v2 if v2 does not occur free in F , i.e. ∀v2 (v2 = v1′ ⊃ F (v2 )),
which is logically equivalent to F (v1′ ). But this involves substitution of the variable
v2 for all free occurrences of v1 in F , which would defeat the purpose of avoiding
substitution. However, we can use a quasi-substitution to obtain from F (v1 ) a
formula logically equivalent to F (vi ), namely ∀v1 (v1 = vi ⊃ F (v1 )) where vi is any
variable that does not occur in F (v1 ). Then we can use quasi-substitution to obtain
a formula without any substitutions that is logically equivalent to F (v1′ ), namely
∀vi (vi = v1′ ⊃ ∀v1 (v1 = vi ⊃ F (v1 )). We abbreviate this formula as F [[v1′ ]]. It’s
easily seen that F [[v1′ ]] is logically equivalent to F (v1′ ). This logical equivalence only
requires that vi does not occur free in F (v1 ), but the sufficient condition that it does
not occur at all in F (v1 ) is easier to express in arithmetized syntax, and that’s the
condition we take.

Definition 35 (the system PAE ) The axioms and rules of inference of PAE are
the following:
A. Logical axioms and rules of inference
Group I propositional logic. All instances of the following schemata:
L1 (F ⊃ (G ⊃ F ))
L2 (F ⊃ (G ⊃ H)) ⊃ ((F ⊃ G) ⊃ (F ⊃ H))
L3 ((∼ F ⊃∼ G) ⊃ (G ⊃ F ))
Group II predicate logic. All instances of the following schemata:
L4 (∀vi (F ⊃ G) ⊃ (∀vi F ⊃ ∀vi G))
L5 (F ⊃ ∀vi F ), provided vi does not occur in F .
L6 ∃vi (vi = t), provided vi does not occur in t.
L7 (vi = t ⊃ (X1 vi X2 ⊃ X1 tX2 )), where X1 and X2 are any expressions such
that X1 vi X2 ) is an atomic formula and t is any term of LE .
Rules of inference
R1 From F and (F ⊃ G), infer G. [Modus Ponens]
R2 From F , infer ∀vi F . [Generalization]
B. Non-logical axioms
Group III axioms specific to each of the primitive non-logical notions of the
language
N1 (v1′ = v2′ ⊃ v1 = v2 )
N2 ∼ 0 = v1′
N3 (v1 + 0) = v1
N4 (v1 + v2′ ) = (v1 + v2 )′
LECTURE 4 40

N5 (v1 · 0) = 0
N6 (v1 · v2′ ) = ((v1 · v2 ) + v1 )
N7 (v1 ≤ 0 ≡ v1 = 0)
N8 (v1 ≤ v2′ ≡ (v1 ≤ v2 ∨ v1 = v2′ ))
N9 (v1 ≤ v2 ∨ v2 ≤ v1 )
N10 v10 = 0′
v′
N11 v12 = (v1v2 · v1 )
Group IV axiom schema of mathematical induction
N12 (F [0] ⊃ (∀v1 (F (v1 ) ⊃ F [[v1′ ]]) ⊃ ∀v1 F (v1 ))),
where, for vi any chosen variable that does not occur in F (v1 ), F [[v1′ ]] is
∀vi (vi = v1′ ⊃ ∀v1 (v1 = vi ⊃ F (v1 ))). Recall that when we write a schematic formula
F (vi ), unless we stipulate otherwise, variables other than vi may occur free in it.
These other free variables are referred to as parameters.

Definition 36 (proof ) A proof in PAE is a sequence of formulas each one of which


is either an axiom of PAE or follows from an earlier formula in the sequence by the
rule of Generalization or follows from two earlier formulas in the sequence by Modus
Ponens.

Definition 37 (provable) A formula F of LE is provable in PAE , symbolized as


P AE ⊢ F , if there exists a proof in PAE of which F is a member.

4.2 An Arithmetical proof predicate for PAE


Each numbered paragraph in this section is both a definition of a property or relation
of numbers, and a proposition that that property or relation is Arithmetical, the
proof of which is established by the formula that follows. Because it will be needed
for later results we prove, in all but cases (4) and (6), a stronger result than is needed
for the present theorem, namely that the expressing formulas from LE require only
bounded quantifiers. The quantifications in (4) and (6) can also be bounded, but
these cases are considerably more complicated. The property of being the Gödel
number of a provable formula requires an unbounded existential quantifier.
Both for ease of reading and of typesetting in the following I will, after the first
case, abbreviate the abbreviation ∗13 for base 13 concatenation by using concatena-
tion itself, i.e. for vi ∗13 vj I will write vi vj , and for 4 ∗13 5, which is an abbreviation
for 0′′′′ ∗13 0′′′′′ and could also be written pf q ∗13 p′ q, I will write 45.
(1) V ar(v1 ): Ev1 is a variable, an expression of the form v′′ ···′ , i.e. the variable
symbol followed by a finite string of subscript symbols. Recall that the Gödel
number of the subscript symbol, on our chosen Gödel numbering, is 5, and that of
the variable symbol is 6.
LECTURE 4 41

(∃v2 ≤ v1 )((∀v3 ≤ v1 )(v3 P13 v2 ⊃ 0′′′′′ P13 v3 ) ∧ v1 = 0′′′′′′ ∗13 v2 )


In this formula I write out the formal numerals 0′′′′′ and 0′′′′′′ rather than ab-
breviating them as 5 and 6, respectively, to bring out the fact that the relation
0′′′′′ P13 v3 ) is between numbers, in this case between the number 5 (as we write it
in base 10, and also, as it happens, in base 13 notation) and some other number
and not a relationship between numerals, though the relation between numbers is
determined to hold or not by going from the number to its, in this case, base 13
representation. The number required to exist by the quantification over the variable
v2 is the Gödel number of a string of subscripts, by the condition that the subscript
symbol is a part of every part of that expression. All of which is to way that when
the formula V ar(v1 ) is written in the primitive notation of LE , the numbers in it
will be expressed by numerals, i.e. 0′′′′′ for 5 and 0′′′′′′ for 6, and similarly in the rest
of the formulas expressing arithmetized syntax.
(2) N um(v1 ): Ev1 is a numeral, i.e. an expression of the form 0′...′
P ow13 (v1 )
(3) Seqt(v1 ): Ev1 is a formation sequence for terms, i.e. a sequence of expressions
each one of which is either a variable or a numeral or the result of applying one of
the four functions of successor, addition, multiplication, or exponentiation to an
expression or expressions occurring earlier in the sequence, i.e. of the form t′ or
(t1 f′ t2 ) or (t1 f′′ t2 ) or (t1 f′′′ t2 ).
(Seq(v1 ) ∧ (∀v2 ≤ v1 )(v2 ∈ v1 ⊃ (V ar(v2 ) ∨ N um(v2 ) ∨ (∃v3 ≤ v1 )(v3 ≺ v2 ∧ v2 =
v1
v3 0)∨(∃v3 ≤ v1 )(∃v4 ≤ v1 )(v3 ≺ v2 ∧v4 ≺ v2 ∧(v2 = 2v3 45v4 3∨v2 = 2v3 455v4 3∨v2 =
v1 v1
2v3 4555v4 3)))))
(4) T m(v1 ): Ev1 is a term.
∃v2 (Seqt(v2 ) ∧ v1 ∈ v2 )
Note: The formula Seqt(v2 ) in (4) is obtained from the formula Seqt(v1 ) in (3)
by changing the free variable from v1 to v2 . In changing the free variable in this
way corresponding changes of bound variables in Seqt(v1 ) must be made so that v1
is free for v2 in a logically equivalent transform of Seqt(v1 ), e.g.
(Seq(v1 ) ∧ (∀v5 ≤ v1 )(v5 ∈ v1 ⊃ (V ar(v5 ) ∨ N um(v5 ) ∨ (∃v3 ≤ v1 )(v3 ≺ v5 ∧ v5 =
v1
v3 0)∨(∃v3 ≤ v1 )(∃v4 ≤ v1 )(v3 ≺ v5 ∧v4 ≺ v5 ∧(v5 = 2v3 45v4 3∨v5 = 2v3 455v4 3∨v5 =
v1 v1
2v3 4555v4 3)))))
If we had given the formula in (4) as the logically equivalent formula ∃v5 (Seqt(v5 )∧
v1 ∈ v2 ), the only change needed to obtain Seqt(v5 ) from Seqt(v1 ) is to replace all
occurrences of v1 by v5 .
Note: The formula in (4) above contains an initial unbounded existential quan-
tifier. This quantifier can be bounded by the correlate in arithmetized syntax of
Problem 2 on Problem sheet 1, i.e. decidability of whether or not an expression is
LECTURE 4 42

a term, but it’s a delicate question in which languages, i.e. with what primitives,
that bound can or cannot be expressed.
(5) AF (v1 ): Ev1 is an atomic formula, i.e. of the form t1 = t2 or t1 ≤ t2 for t1 , t2
terms.
(∃v2 ≤ v1 )(∃v3 ≤ v1 )(T m(v2 ) ∧ T m(v3 ) ∧ (v1 = v2 ηv3 ∨ v1 = v2 ǫv3 ))
(6) Seqf (v1 ): Ev1 is a formation sequence for formulas, i.e. a finite sequence of
expressions each one of which is either an atomic formula or of the form ∼ E for E
occurring earlier in the sequence or of the form (Ei ⊃ Ej ) for Ei and Ej occurring
earlier in the sequence or of the form ∀vi E for vi any variable and E occurring earlier
in the sequence.
(Seq(v1 )∧(∀v2 ≤ v1 )(v2 ∈ v1 ⊃ (AF (v2 )∨(∃v3 ≤ v1 )(v3 ≺ v2 ∧v2 = 7v3 )∨(∃v3 ≤
v1
v1 )(∃v4 ≤ v1 )(v3 ≺ v2 ∧ v4 ≺ v2 ∧ v2 = 2v3 8v4 3) ∨ (∃v3 ≤ v1 )(∃v4 ≤ v1 )(v3 ≺
v1 v1 v1
v2 ∧ V ar(v4 ) ∧ v2 = 9v4 v3 ))))
(7) F m(v1 ): Ev1 is a formula.
∃v2 (Seqf (v2 ∧ v1 ∈ v2 ))
Note: The remark as at (4) above applies here also. We know by Problem 2
on Problem sheet 1, that we can determine by a finite search whether an expression
is a formula, but it’s a delicate matter to determine in exactly what language of
arithmetic, i.e. with what primitives, this numerical quantifier can be bounded by
a term of the language.
(8) Ax(v1 ): Ev1 is an axiom of PAE . There are seven schemata of logical axioms
L1 − L7 and eleven axioms of arithmetic N1 − N11 plus one axiom schemata of
arithmetic N12 (Induction). We need formulas Li (v1 ) such that Li (v1 ) iff (Ev1 is an
axiom of form Li ) and Ni (v1 ) such that Ni (v1 ) iff (Ev1 is an axiom of form Ni ). The
property of numbers Ax(v1 ) is expressed by (L1 (v1 ) ∨ . . . ∨ L7 (v1 ) ∨ N1 (v1 ) ∨ . . . ∨
N12 (v1 )). Finding N12 (v1 ) is problem 5 on Exercise sheet 2. I will treat a couple of
cases from each of the other groups of axioms.
Logical axioms:
Group I
L1 (v1 ): (∃v2 ≤ v1 )(∃v3 ≤ v1 )(F m(v2 ) ∧ F m(v3 ) ∧ v1 = 2v2 82v3 8v2 33)
L3 (v1 ): (∃v2 ≤ v1 )(∃v3 ≤ v1 )(F m(v2 ) ∧ F m(v3 ) ∧ x = 227v2 87v3 382v3 8v2 33)
Group II
L4 (v1 ): (∃v2 ≤ v1 )(∃v3 ≤ v1 )(∃v4 ≤ v1 )(F m(v2 ) ∧ F m(v3 ) ∧ V ar(v4 ) ∧ v1 =
29v4 2v2 8v3 3829v4 v2 89v4 v3 33)
L7 (v1 ): (∃v2 ≤ v1 )(∃v3 ≤ v1 )(∃v4 ≤ v1 )(∃v5 ≤ v1 )(∃v6 ≤ v1 )(∃v7 ≤ v1 )(V ar(v2 )∧
T m(v3 ) ∧ v6 = v4 v2 v5 ∧ AF (v6 ) ∧ v7 = v4 v3 v5 ∧ v1 = 2v2 ηv3 82v6 8v7 33)
Group III
N1 (v1 ): Ev1 is the axiom N1 , which in primitive notation is
((v′ )′ = (v′′ )′ ⊃ (v′ ) = (v′′ )).
LECTURE 4 43

v1 = 226530η26553082653η265533.
N7 (v1 ): Ev1 is the axiom N7 . To compute the Gödel number of N7 we must
write it in primitive notation. This requires expressing ≡ in terms of ∼ and ⊃,
given by the truth-functional equivalences ((A ≡ B) ≡ (A ⊃ B) ∧ (B ⊃ B)) and
((C ≡ D) ≡∼ (C ⊃∼ D)), which yields ((A ≡ B) ≡∼ ((A ⊃ B) ⊃∼ (B ⊃ B))).
So N7 =∼ (((v′ ≤ 0 ⊃ (v′ ) = 0) ⊃∼ ((v′ = 0 ⊃ (v′ ≤ 0))
v1 = 7222653ǫ182653η138722653η182653ǫ133
(9) P rfP AE (v1 ): Ev1 is a proof in PAE , i.e. a sequence of formulas each one of
which is either an axiom of PAE , or is the result of applying R1 [Modus Ponens]
to two formulas occurring earlier in the sequence, or is the result of applying R2
[Generalization] to a formula occurring earlier in the sequence.
(Seq(v1 ) ∧ (∀v2 ≤ v1 )(v2 ∈ v1 ⊃ (Ax(v2 ) ∨ (∃v3 ≤ v1 )(∃v4 ≤ v1 )(v3 ≺ v2 ∧ v4 ≺
v1 v1
v2 ∧ v4 = 2v3 8v2 3) ∨ (∃v3 ≤ v1 )(∃v4 ≤ v1 )(V ar(v3 ) ∧ v4 ≺ v2 ∧ v2 = 9v3 v4 )))))
v1
(10) P rovP AE (v1 , v2 ): Ev1 is proved by Ev2 .
P rovP AE (v1 , v2 ) ≡ (P rfP AE (v2 ) ∧ v1 ∈ v2 ))
(11) P rP AE (v1 ): Ev1 is provable
∃v2 P rovP AE (v1 , v2 ).

Theorem 24 (Arithmetical proof relation) The two place relation between num-
bers m and n given by the condition that n is the Gödel number of a proof in PAE
of the formula whose Gödel number is m is expressible in LE .

Proof. The formula P rovP AE (v1 , v2 ) in (10) expresses “Ev2 is a proof of Ev1 ”.
This is evident from this formula and (1) - (9). N

Corollary 25 (Arithmetical proof predicate) The property of a number that it


is the Gödel number of a formula provable in PAE is expressible in LE ,

Proof. PAE ⊢ En if and only if ∃v2 P rovP AE (n, v2 ) is true. N

Remark: As the construction of the proof predicate for PAE shows, arithmeti-
zation by assignment of digits to symbols and of concatenation of corresponding se-
quences of digits (numbers in the given base notation) to concatenation of sequences
of symbols (expressions) makes the correspondence between formal expressions and
numbers, which Hilbert recognized (see quotation in the first lecture) completely
direct. What Gödel achieved, going beyond Hilbert’s insight, was to show that the
formal syntax of strings of symbols by which a formal system of proof is established
corresponds exactly with arithmetically definable properties of the corresponding
numbers.
LECTURE 4 44

4.3 An inefficient and a weak version of Gödel’s


First Incompleteness Theorem for PAE
Theorem 26 (an inefficient proof of Gödel’s First Incompleteness Theorem)
If every sentence provable in P AE is true, there exists a sentence which is true but
unprovable in P AE and whose negation is unprovable.

Proof. Having shown that the property of being the Gödel number of a theorem
of PAE is Arithmetical (expressible in LE ), the existence of a true sentence in the
language of arithmetic not provable in PAE follows immediately from Tarski’s theo-
rem, that the set of Gödel numbers of true sentences in the language of arithmetic is
not arithmetical (Exercise sheet 1 problem 4) and the hypothesis that every sentence
provable in PAE is true: By the fact that the Gödel numbers of provable formulas
is arithmetical, we have that this set is not identical with the set of Gödel numbers
of true sentences. If every sentence provable in PAE is true and the set of provable
sentences does not coincide with the set of true sentences, there must exist a true
sentence which is not provable. N

This argument via Tarski’s theorem is highly inefficient since it fails to generate
a particular true sentence that is unprovable in the given system while at the same
time it requires the construction of an arithmetized proof predicate, from which we
can by the Diagonal Lemma, itself used in the proof of Tarski’s theorem, obtain a
particular sentence, namely a diagonal sentence for ∼ P rP AE (v1 ) which we can show,
on the hypothesis that every sentence provable in PAE is true, is unprovable and
(thereby) true, and so whose negation is false and so by the hypothesis unprovable.
We have thus

Theorem 27 (weak version of Gödel’s First Incompleteness Theorem for PAE )


There is a sentence G in LE such that, if every sentence provable in PAE is true,
PAE 0 G, G is true, and PAE 0∼ G.

Proof. By Corollary 25, there is a formula in LE , P rP AE (v1 ), that expresses the


property of being the Gödel number of a formulas derivable in PAE . By the Diagonal
Lemma (Theorem 17) there is a sentence G such that (G ≡∼ P rP AE (pGq))). This
establishes conditions (1) and (2) of Theorem 5. Hence on the assumption that
every sentence provable in PAE is true (condition (3)), Theorem 5 establishes that
PAE 0 G, G is true, and PAE 0∼ G. N
Lecture 5

The system PA with zero,


successor, addition, multiplication,
and ≤ as primitive; Σ0- and
Σ1-formulas; a Σ0-coding of finite
sets of ordered pairs; the relation
xy = z is ∆1-expressible in the
language of PA

.
(Tuesday, 26 October 2010)

5.1 The system PA with zero, successor, addition,


multiplication, and ≤ as primitive
We now drop exponentiation as a primitive of the language of arithmetic and drop
the axioms governing exponentiation from our first-order system of arithmetic PAE .
The resulting system is (with variant formulations) the standard first-order axiom
system for arithmetic, known as Peano Arithmetic, labeled PA. The mathematical
fact that justifies the move from PAE to PA is that the three-place relation xy = z
is expressible in the language of PA. We shall establish not only that exponentia-
tion is expressible in the language of PA, but that it is expressible in a way that
characterizes the total general recursive functions.

45
LECTURE 5 46

Definition 38 (the system PA) The language L for PA is obtained from the lan-
guage LE for PAE by dropping the condition in the definition of terms for LE , Def-
inition 12, that if t1 and t2 are terms, then so is (t1 f′′′ t2 ), which correspondingly
removes from formulas of the language LE any expressions that contain the expres-
sion f′′′ (without having to make any change to Definition 16). The axioms of PA
are obtained from those of PAE by dropping axioms N10 and N11 (which are not
formulas in the language L of PA).
By simple modification of the construction of an Arithmetical proof predicate
for PAE in Section 4.2, we obtain an Arithmetical proof predicate for PA.
Theorem 28 (proof predicate for PA in LE ) There are formulas P rfPEA (v1 ),
P rovPEA (v1 , v2 ), and P rPEA (v1 ) in the language LE that express the property of being
the Gödel number of a proof sequence for PA, the relation of being the Gödel number
of a formula that occurs in the proof sequence coded by a given number, and the
property of being the Gödel number of a theorem of PA.
Proof. We do this by by modifying the constructions for PAE in Theorem 24 and
Corollary 25. We drop the disjunct corresponding to term formation by the function
expression f′′′ , i.e. v2 = 2v3 4555v4 3, so that Seqt(v1 ) is (Seq(v1 ) ∧ (∀v2 ≤ v1 )(v2 ∈
v1 ⊃ (V ar(v2 ) ∨ N um(v2 ) ∨ (∃v3 ≤ v1 )(v3 ≺ v2 ∧ v2 = v3 0) ∨ (∃v3 ≤ v1 )(∃v4 ≤
v1
v1 )(v3 ≺ v2 ∧ v4 ≺ v2 ∧ (v2 = 2v3 45v4 3 ∨ v2 = 2v3 455v4 3))))). In the formula Ax(v1 )
v1 v1
that expresses “Ev1 is an axiom of PAE ” (p. 39) the disjuncts N10 (v1 ) and N11 (v1 ) are
dropped. With these modifications, the formulas for P rfP AE (v1 ), P rovP AE (v1 , v2 ),
and P rP RE (v1 ) on p. 40 are transformed to corresponding Arithmetical formulas
P rfPEA (v1 ), P rovPEA (v1 , v2 ), and P rPEA (v1 ). N

5.2 Σ0-formulas
Definition 39 (bounded quantifiers) For vi any variable, and n any numeral,
quantification in either of the forms ∀vi (vi ≤ n ⊃ F ) or ∃vi (vi ≤ n ∧ F ) is called
bounded quantification, abbreviated as (∀vi ≤ n)F and (∃vi ≤ n)F , respectively.
Also, for vi any variable, vj any variable such that i 6= j, and F any formula,
quantification in either of the forms ∀vi (vi ≤ vj ⊃ F ) and ∃vi (vi ≤ vj ∧ F ) are called
bounded quantifiers, abbreviated as (∀vi ≤ vj )F and (∃vi ≤ vj )F , respectively.
Remark. The restriction that the variable vj be distinct from the variable vj
when the bound on the quantification is a variable is essential, since ∀vi (vi ≤ vi ⊃ F )
is logically equivalent to ∀vi F , which is unbounded quantification. Also, note that
bounded existential quantifications are, in primitive notation, formulas of the form
∼ ∀vi ∼ (vi ≤ n ∧ F ) and ∼ ∀vi ∼ (vi ≤ vj ∧ F ).
LECTURE 5 47

Definition 40 (Σ0 -formulas) .


(a) Every atomic formula of the language L of PA is Σ0 , i.e. for any terms t1
and t2 , t1 = t2 and t1 ≤ t2 are Σ0 .
(b) If F is Σ0 , then ∼ F is Σ0 .
If F and G are Σ0 then (F ⊃ G) is Σ0 .
(c) If F is Σ0 , vi is any variable, and t is either a variable distinct from vi or a
numeral, then ∀vi (vi ≤ t ⊃ F ) is Σ0 .

Corollary 29 For F any Σ0 -formula, vi any variable, and t either a variable dif-
ferent from vi or a numeral, ∃vi (vi ≤ t ∧ F ) is Σ0 .

Proof. By (b), ∼ F (vi ) is Σ0 . Then by (c), ∀vi (vi ≤ t ⊃∼ F ) is Σ0 . Hence by


(b) again, ∼ ∀vi (vi ≤ t ⊃∼ F ) is Σ0 . ∼ ∀vi (vi ≤ t ⊃∼ F ) is logically equivalent to
∼ ∀vi ∼ (vi ≤ t ∧ F ), which is abbreviated as ∃vi (vi ≤ t ∧ F ), which is to say that
∃vi (vi ≤ t ∧ F ) is Σ0 . N

Proposition 30 (Decidability of Σ0 -sentences) We can effectively decide (com-


pute) the truth or falsity of each Σ0 -sentence, i.e. closed Σ0 -formula.

Proof. By induction over the recursive definition of Σ0 -formulas, corresponding


to the clauses (a), (b), and (c) of the definition:
(a) A closed term is computable to a numeral, and an equation between a term
and a numeral is decidable.
(b) Given the truth value of F and G, we can compute the truth value of ∼ F
and of (F ⊃ G).
(c) Suppose vi is not free in F . Suppose F is true. Then (H ⊃ F ) is true for any
H, so ∀vi (vi ≤ n ⊃ F ) is true. Suppose F is false. Since 0 ≤ n is true, (0 ≤ n ⊃ F )
is false, so ∀vi (vi ≤ n ⊃ F ) is false.
Suppose vi occurs free in F . Then ∀vi (vi ≤ n ⊃ F (vi )) is equivalent to (F (0) ∧
. . . ∧ F (n)). By induction hypothesis each conjunct is decidable, so the conjunction
is decidable. N

For reasons that will be apparent shortly, we also label Σ0 -formulas as Π0 and
as ∆0 .

5.3 Σ1- and Π1-formulas; Σ1- Π1- and ∆1-relations


Definition 41 (Σ1 -formula) A Σ1 formula is any formula of the form ∃vi F where
F is a Σ0 -formula.
LECTURE 5 48

Definition 42 (Σ1 -relation) A relation R ⊆ Nk is Σ1 iff there is a Σ1 -formula


G(v1 , . . . , vk ) that expresses R, i.e. such that for each k-tuple (n1 , . . . , nk ), G(n1 , . . . nk )
is true iff (n1 , . . . , nk ) ∈ R.
Note that a Σ1 -formula begins with one unbounded existential quantifier. (It
may contain other quantifiers so long as they are bounded.)
Proposition 31 Every Σ0 -formula is logically, and hence provably, equivalent to a
Σ1 -formula.
Proof. For vi not free in F , (F ≡ ∃vi F ) is logically valid, and hence provable
in every system complete with respect to first-order logical validity. N
Definition 43 (Π1 -formula) A Π1 formula is any formula of the form ∀vi F where
F is a Σ0 -formula.
Definition 44 (Π1 -relation) A relation R ⊆ Nk is Π1 iff there is a Π1 -formula
G(v1 , . . . , vk ) that expresses R, i.e. such that for each k-tuple (n1 , . . . , nk ), G(n1 , . . . nk )
is true iff (n1 , . . . , nk ) ∈ R.
Lemma 32 (Σ1 and Π1 are dual to each other) The negation of a Σ1 -formula
is Π1 , and the negation of Π1 -formula is Σ1 .
Proof. By logic and the fact that the negation of a Σ0 -formula is Σ0 . N
Definition 45 (∆1 relations) A relation is ∆1 if and only if it is both Σ1 and Π1 .
Remark. Strictly, there is no such thing as a ∆1 -formula, i.e. ∆1 is not a syn-
tactic form. However, we shall sometimes label as ∆1 a formula which is equivalent
both to a Σ1 -formula and to a Π1 -formula.
Corollary 33 A relation is ∆1 iff it is Σ1 and its complement is also Σ1 .
Proof. Immediate from Definition 45 and Lemma 32. N

Remark. The definitions of Σ1 and Π1 -formulas, and by the Remark above


also ∆1 -formulas, include the case of formulas with no free variables, i.e. sentences.
The definitions of Σ1 , Π1 , and ∆1 -relations include the case of 0-ary relations, i.e.
propositions, and 1-ary relations, i.e. sets.
Example of a ∆1 -proposition: The proposition expressed by the equivalent
sentences ∃v1 (v1 = n ∧ F (v1 )) and ∀vi (vi = n ⊃ F (vi )), where v1 is the only free
variable in a Σ0 -formula F (v1 ). However, as we have seen, these sentences are
logically equivalent to F (n), which is Σ0 for F (v1 ) Σ0 . It is an important fact that
there are ∆1 relations, including 0-ary relations, that are ∆1 and not Σ0 , but this is
not a fact we can prove on the basis of results so far obtained.
LECTURE 5 49

Definition 46 (a function is Σ1 , Π1 , ∆1 ) For f an n-ary function from Nn to


N, f is Σ1 iff the n + 1-ary relation f (v1 , . . . , vn ) = vn+1 is Σ1 . Similarly for Πi and
∆1 .

Lemma 34 If a total function f : Nn → N is Σ1 , then it is ∆1 .

Proof. We show that the relation f (v1 , . . . , vn ) 6= vn+1 is also Σ1 , from which
it follows that the relation f (v1 , . . . , vn ) = vn+1 is ∆1 . By the hypothesis that f is Σ1 ,
there is a Σ0 -formula F (v1 , . . . , vn , vn+1 , vn+2 ) such that ∃vn+2 F (v1 , . . . , vn , vn+1 , vn+2 )
expresses the relation f (v1 , . . . , vn ) = vn+1 . Then, since f is total so that, for every
v1 , . . . , vn , there is vn+1 such that f (v1 , . . . , vn ) = vn+1 , the relation f (v1 , . . . , vn ) 6=
vn+1 is expressed by the condition ∃vn+3 ∃vn+2 (F (v1 , . . . , vn , vn+3 , vn+2 )∧ ∼ vn+3 =
vn+1 ). This condition is equivalently expressed by

∃vn+4 (∃vn+3 ≤ vn+4 )(∃vn+2 ≤ vn+4 )(F (v1 , . . . , vn , vn+3 , vn+2 )∧ ∼ vn+3 = vn+1 )

By Definition 40,

(∃vn+3 ≤ vn+4 )(∃vn+2 ≤ vn+4 )(F (v1 , . . . , vn , vn+3 , vn+2 )∧ ∼ vn+3 = vn+1 )

is a Σ0 -formula. Hence the preceding formula is Σ1 . N

5.4 Arithmetization of syntax in the language of


PA
We have already seen that for any base b ≥ 2, concatenation to base b, x ∗b y = z, is
expressible in LE . We now show that for base p for p a prime number, concatenation
to base p is expressible in L, and indeed that it is Σ0 -expressible in this language.
This result is based on an observation by John Myhill (see Smullyan, p. 43) that
the property of a number x that it is a power of a given prime p can be expressed
without using the exponentiation function, since x is a power of the given prime p
if and only if every proper divisor of x is divisible by p.

Lemma 35 For every prime number p the following conditions are Σ0 .


1. x | y —x divides y.
2. P owp (x) —x is a power of p.
3. pℓp (x) = y —y is the smallest positive power of p greater than x.

Proof.
1. ((∃z ≤ y)(x · z = y)∧ ∼ x = 0). Note that by this condition every non-zero
number divides 0 (which reflect the usual practice in number theory that 0 is a
LECTURE 5 50

member of every principal ideal on the natural numbers) and 0 does not divide 0
by the second conjunct, nor any non-zero number by the first conjunct (cf problem
3(a) on Problem sheet 1).
2. (∀z ≤ x)((z | x ∧ z 6= 1) ⊃ p | z).
3. (P owp (y) ∧ y > x ∧ y > 1 ∧ (∀z < y) ∼ (P owp (z) ∧ z > x ∧ z > 1). N

Lemma 36 (base p concatenation is Σ0 ) For any prime p, the relation x ∗p y =


z is Σ0 .

Proof.
x ∗p y = z is Σ0 iff pℓp (y) + y = z iff (∃v1 ≤ z)(v1 = pℓp (y) ∧ ((x · v1 ) + y) = z).
The result follows by part 3 of Lemma 35. N

Lemma 37 The Arithmetical relation xPb y (‘x is part of y’) for base b a prime
number is Σ0 .

Proof. By Theorem 19, Lemma 35 and Lemma 36. N

5.5 A Σ0-coding of finite sets of ordered pairs of


numbers
We now establish a result that exponentiation can be expressed in the language of
PA, which does take exponentiation as primitive. The key to this results is a Σ0 -
coding of finite set of ordered pairs of numbers. This results was proved by Gödel in
his original paper, for his Gödel numbering using the Chinese Remainder Theorem.
Given our different Gödel numbering result, we don’t need the Chinese Remainder
Theorem for this result.

Theorem 38 (Σ0 -coding of finite sets of ordered pairs) There is a Σ0 -formula


K(v1 , v2 , v3 ) such that
1. For any finite set of ordered pairs of natural numbers (a1 , b1 ), (a2 , b2 ), . . . , (ar , br ),
there is a number k such that for any numbers m and n, K(m, n, k) holds if and
only if (m, n) is one of the pairs (a1 , b1 ), (a2 , b2 ), . . . , (ar , br ).
2. For any numbers v1 , v2 , v3 , if K(v1 , v2 , v3 ) holds, then v1 ≤ v3 and v2 ≤ v3 .

Proof.
We need to describe two things (dependently related to each other): a process
whereby a set of ordered pairs of numbers (a1 , b1 ), (a2 , b2 ), . . . , (ar , br ) is coded by a
number k, and a formula K(v1 , v2 , v3 ) whereby, given a code number k, the set of
ordered pairs is decoded.
LECTURE 5 51

(1) Coding. By a frame number we shall mean a number whose numeral in a


specified base, in our case 13, is of the form 2t2 where t is a string of 1s (this idea goes
back to W.V. Quine, “Concatenation as a basis for arithmetic”, Journal of Symbolic
Logic 11 (1946), pp. 105-114; see Smullyan p. 45). The condition that the numeral
for x in base b representation is a string of 1s, which we shall abbreviate as 1b (x),
is expressible by the condition, (∀y ≤ x)(yPb x ⊃ 1Pb y). Examples: 110 (11110 ), but
∼ 110 (11113 ) since 11113 = 18310 , which also means that 113 (18310 ).
Let θ be a finite sequence of ordered pairs, (a1 , b1 ), (a2 , b2 ), . . . , (ar , br ), and let
f be any frame number that has a longer string of 1s in it than the longest string
of 1s that occurs in any of the numbers a1 , b1 , . . . , ar , br . The code k of the set of
ordered pairs is the number f f a1 f b1 f f a2 f b2 f f . . . f f ar f br f f .
(2) Decoding We call x a maximal frame in y if x is a frame number, x is part
of y, and x is as long as any frame that is part of y. This relation is expressed by
the following formula, which we will label M F (x, y):
(xP y ∧ (∃z ≤ y)(1(z) ∧ x = 2z2∧ ∼ (∃w ≤ y)(1(w) ∧ 2zw2P y))).
By Lemmas 36 and 37, M F (x, y) is Σ0 .
Note that if y has a frame number in it, it has a maximal frame number, since the
length of frame numbers in y is bounded by the length of y. Since frame numbers
whose numerals in the given base are of the same length are equal, any number
which contains a frame number contains a unique maximal frame number.
Having expressed the notion of a maximal frame, we are then able to define a
formula K(v1 , v2 , v3 ) with which to decode the ordered pairs coded by the process
specified in (1).
K(v1 , v2 , v3 ) =df
(∃v4 ≤ v3 )(M F (v4 , v3 ) ∧ v4 v4 v1 v4 v2 v4 v4 P v3 ∧ ∼ v4 P v1 ∧ ∼ v4 P v2 ).
By Lemmas 36 and 37, the formula K(v1 , v2 , v3 ) is Σ0 .
Let us suppose that the sequence of ordered pairs (a1 , b1 ), (a2 , b2 ), . . . , (ar , br )
has been coded by the number k using the above procedure. The frame number f
chosen for the coding occurs in k will be a maximal frame number for k since the
string of 1s in f is longer than any strings of 1s from the ai and bi . As noted above,
it is unique. Having recovered f from k, we can then decode each of the pairs of
numbers coded in k. N
Note that there is an error in Smullyan’s proof (p. 45), at the point when he
says, let f be any frame number that is longer than any frame which is part of
any of the numbers a1 , b1 , . . . , ar , br . The problem is if one or more of the ai or bi
is a string of 1s that is longer than any string of 1s occurring in a frame number
in one of the numbers being coded. Let c be the longest such number and let f
be the frame number specified by Smullyan. In this case the maximal frame in
f f a1 f b1 f f a2 f b2 f f . . . f f ar f br f f is 2c2 and not f , and the decoding fails.
LECTURE 5 52

5.6 The relation xy = z is ∆1-expressible in the


language of PA
Theorem 39 (exponentiation is Σ1 ) The relation xy = z is Σ1 .

Proof. The relation xy = z holds if and only if there is a set of ordered pairs
{(0, 1), (1, x), (2, x2 ), . . . , (y, xy )} and (y, z) is a member of that set. Given the coding
and Σ0 -decoding of finite sets of ordered pairs of numbers by Theorem 38, we can
express this by the formula ∃w(K(y, z, w) ∧ (∀u ≤ w)(∀v ≤ w)(K(u, v, w) ⊃ ((u =
0 ∧ v = 1) ∨ (∃r ≤ w)(∃s ≤ w)(K(r, s, w) ∧ r′ = u ∧ v = x · s))). Since K(v1 , v2 , v3 ) is
Σ0 , this whole formula is Σ1 . Notice that for this result we need the second clause of
Theorem 38, i.e. that for any numbers v1 , v2 , v3 , if K(v1 , v2 , v3 ) holds, then v1 ≤ v3
and v2 ≤ v3 in order to bound the universal quantifiers inside the formula, so that
it is, indeed, Σ1 . N

Corollary 40 (exponentiation is ∆1 ) The relation xy = z is ∆1 .

Proof. Immediate from Theorem 39 by Lemma 34. N

A stronger result than Corollary 40 is true, namely:

Theorem 41 (exponentiation is ∆0 ) The relation xy = z is ∆0 .

Proof. The proof is too complicated to give here. See Petr Hájek and Pavel
Pudlák, Metamathematics of First-Order Arithmetic, Springer, Berlin, 1993, pp.
299-303. N

Theorem 41 and its proof are specific to exponentiation and do not generalize
to other functions. On the other hand, the result in Theorem 39 with Corollary 40
generalizes to the following important result:

Theorem 42 A total function is general recursive if and only if it is ∆1 .

Proof. Later. N
Lecture 6

Every Σ-formula is provably


equivalent to a Σ1-formula; the
arithmetized proof predicate for
PA is Σ1; the arithmetical
hierarchy

(Wednesday, 27 October 2010)

6.1 Σ-formulas
We now define a class of formulas each of which is provably equivalent in PA to a
Σ1 -formula, but which are more flexibly expressed than is required in order to be a
Σ1 -formula.
Definition 47 (Σ-formula) . Base: Every Σ0 -formula is a Σ-formula.
Recursion: 1. If F is a Σ-formula, then for any variable vi , the formula ∃vi F is
a Σ-formula.
2. For any Σ-formulas F and G, (F ∨ G) and (F ∧ G) are Σ-formulas.
3. For any Σ0 -formula F and Σ-formula G, (F ⊃ G) is a Σ-formula.
4. If F is a Σ-formula, then for any distinct variables vi and vj , (∃vi ≤ vj )F and
(∀vi ≤ vj )F are Σ-formulas, and for any number n, (∃vi ≤ n)F and (∀vi ≤ n)F are
Σ-formulas.
To show that every Σ-formula is provably equivalent in PA to a Σ1 -formula we
prove the following five Lemmas, corresponding to the five clauses in the definition
of a Σ-formula.

53
LECTURE 6 54

Lemma 43 Every Σ0 -formula is provably equivalent in PA to a Σ1 -formula.

Proof. Any formula F (vi1 , . . . , vik ) is logically equivalent to ∃vik+1 F (vi1 , . . . , vik ),
where vik+1 does not occur free in F (vi1 , . . . , vik ), so by the completeness of first-order
logic in PA, PA ⊢ (F (vi1 , . . . , vik ) ≡ ∃vik+1 F (vi1 , . . . , vik )) N

Lemma 44 If ∃vi F (vi ) is a Σ1 -formula, then ∃vj ∃vi F (vi ) is provably equivalent in
PA to a Σ1 -formula

Proof. Exercise (Problem sheet 3 problem 2(a)) N

Lemma 45 Let ∃vi F (vi ) and ∃vj G(vj ) be Σ1 -formulas. Then (∃vi F (vi ) ∨ ∃vj G(vj ))
and (∃vi F (vi ) ∧ ∃vj G(vj )) are each provably equivalent in PA to a Σ1 -formula

Proof.
(a) ((∃vi F (vi ) ∨ ∃vj G(vj )) ≡ ∃vk (F (vk ) ∨ G(vk ))) is logically valid (for vk any
variable substitutable into F and G) (Cf. Problem sheet 1 problem 5(a)), so provable
in PA.
(b) ((∃vi F (vi ) ∧ ∃vj G(vj )) ≡ ∃vi ∃vj (F (vi ) ∧ G(vj ))) is logically valid (on the
assumption, without loss of generality, that vi does not occur free in G and vj does
not occur free in F ). Hence this equivalence is provable in PA. By Lemma 44,
⊢ (∃v1 ∃v2 (F (vi ) ∧ G(vj )) is provably equivalent in PA to a Σ1 formula. N

Lemma 46 For F a Σ0 -formula and ∃vi H(vi ) a Σ1 -formula, (F ⊃ ∃vi H(vi )) is


provably equivalent in PA to a Σ1 -formula.

Proof. (F ⊃ ∃vi H(vi )) ≡ ∃vi (F ⊃ H(vi )) is logically equivalent, so provably in


PA. We must assume, which we can do without loss of generality, that vi does not
occur free in F . The formula on the right is Σ1 . N

Lemma 47 For ∃vi R(vi , vj ) a Σ1 -formula, the following formulas are provably equiv-
alent in PA to Σ1 -formulas.
(a) (∃vj ≤ vk )∃vi R(vi , vj )
(b) (∀vj ≤ vk )∃vi R(vi , vj )
and similarly for the formulas in (a) and (b) with a numeral in place of the
variable vk .

Proof.
(a) The formula vj ≤ vk is Σ0 by the definition of Σ0 . Hence by Lemma 43 it is
provably equivalent to a Σ1 formula. Hence by Lemma 45 (vj ≤ vk ∧ ∃vi R(vi , vj )) is
provably equivalent to a Σ1 -formula. Hence by Lemma 44, ∃vj (vj ≤ vk ∧∃vi R(vi , vj )),
which we abbreviate as (∃vj ≤ vk )∃vi R(vi , vj ), is provably equivalent to a Σ1 -
formula. This argument holds also for a numeral n in place of the variable vk .
(b) Exercise (Problem sheet 3 Problem 3(b)). N
LECTURE 6 55

Theorem 48 (Σ equivalent to Σ1 ) Every Σ-formula is provably equivalent in PA


to a Σ1 -formula.

Proof. By induction on the recursive definition of Σ-formulas, the base case and
each induction step established by one of Lemmas 43 - 47. N

6.2 The arithmetized proof predicate for PA is Σ1


Theorem 49 The arithmetized proof predicate for PA under our arithmetization of
syntax of PA is arithmetical, i.e. expressed in the language of PA.

Proof. By Theorem 28 and Lemmas 36 and 37.

A much stronger result than Theorem 49 is true and is needed for what is to
come, namely that the arithmetized proof predicate for PA is Σ1 . The first step is
to show that the two-place formula (P rfP A (v2 ) ∧ v1 ∈ v2 ) is Σ1 . With considerable
effort we can actually establish that it’s Σ0 . We know from Problem 2 on Problem
sheet 1 that whether a string of symbols is a term or a formula is decidable, so we
know that there is a bound on those quantifiers, but giving an explicit formulation
of that bound in the language of PA is hard work which we can avoid. However, for
the overall result that the proof predicate for PA is Σ1 it’s sufficient to show that
(P rfP A (v2 ) ∧ v1 ∈ v2 ) is Σ1 .

Lemma 50 The formula (P rfP A (v2 )∧v1 ∈ v2 ), i.e. (Seq(v2 )∧(∀v3 ≤ v2 )(v3 ∈ v2 ⊃
(Ax(v3 ) ∨ (∃v4 ≤ v2 )(∃v5 ≤ v2 )(v4 ≺ v3 ∧ v5 ≺ v3 ∧ v5 = 2v4 8v2 3) ∨ (∃v4 ≤ v2 )(∃v5 ≤
v2 v2
v2 )(V ar(v4 ) ∧ v5 ≺ v3 ∧ v3 = 9v4 v5 ))) ∧ v1 ∈ v2 ) is equivalent to a Σ1 -formula.
v2

Proof. The key point is that the only place in the construction of P rfP A (v2 ) in
which we used an unbounded quantifier was in T m(v1 ) (Ev1 is a term) as ∃v2 (Seqt(v2 )∧
v1 ∈ v2 ) and F m(v1 ) (Ev1 is a formula) as ∃v2 (Seqf (v2 ∧ v1 ∈ v2 )). These occur
the formula Ax(v1 ) in L1 − L7 and L12 (Induction axioms). In none of these oc-
currences is it in the antecedent of a conditional, and the occurrence of Ax(v1 ) in
(P rfP A (v2 ) ∧ v1 ∈ v2 ) is also not in the antecedent of a conditional. Hence in a
prenex normal form of (P rfP A (v2 ) ∧ v1 ∈ v2 ) these several existential quantifiers
come out as prenex existential quantifiers. By Lemma 44, this formula is Σ1 . N

Theorem 51 (proof predicate for PA is Σ1 ) The formula P rP A (v1 ),


i.e. ∃v2 (P rfP A (v2 ) ∧ v1 ∈ v2 ), is equivalent to a Σ1 -formula.

Proof. By Lemma 50 and Lemma 44. N


LECTURE 6 56

Corollary 52 {n : P A ⊢ En } is Σ1 .

Proof. By inspection of the formula ∃v2 (P rfP A (v2 ) ∧ v1 ∈ v2 ). N

Theorem 53 {n : P A ⊢ En }, which by Corollary 52 is Σ1 , is not ∆1 , if every


sentence provable in PA is true.

Proof. We require the result, which we will prove in Lecture 7, that PA is Σ1 -


complete. Suppose the complement of {n : P A ⊢ En } is expressed by a Σ1 -formula,
call it N P rP A (v1 ). Then the Gödel sentence G, such that (G ≡∼ P r(pGq)) is
true, is equivalent to the Σ1 -sentence N P RP A (pGq). By a simple modification of
Theorem 27 so that it applies to PA rather than PAE , G is true and, if every
sentence provable in PA is true, not provable in PA. But by the Σ1 -completeness of
PA (to be proved in the next lecture), if G is true and equivalent to a Σ1 -sentence,
then PA⊢ G, which contradicts the unprovability of G on the hypothesis that every
sentence provable in PA is true. (In Lecture 8 we shall prove this result on a much
weaker hypothesis.) N

Proposition 54 {n : PA ⊢ En [n]} is Σ1 .

Proof. Exercise (Problem sheet 3 problem 1) N

6.3 The arithmetical hierarchy


The kind of classifications we introduced in Lecture 5 with the notions of Σ0 , Π0 ,
∆0 and Σ1 , Π1 , ∆1 can be extended, by Theorem 3 (Prenex Normal Forms) and
a generalization of Theorem 48, to a hierarchy of all formulas in the language of
PA. This correspondingly defines a hierarchy of relations on natural numbers (the
arithmetical hierarchy).
Remark. A simple cardinality argument tells us that most relations on the
natural numbers, in particular the 1-ary relations, i.e. the sets of natural numbers,
are not in this hierarchy (there are uncountably many sets of natural numbers and
there are countably many formulas in the language of arithmetic).

Definition 48 (arithmetical hierarchy of formulas) (a) Σ0 -formulas (=df Π0 -


formulas) are as given by Definition 40.
(b) If F is Σn , then ∀vi F is Πn+1 .
(c) If F is Πn , then ∃vi F is Σn+1 .

There is a corresponding arithmetical hierarchy of sets and relations.


LECTURE 6 57

Definition 49 (arithmetical hierarchy of sets and relations) A relation on nat-


ural numbers is Σn (or Πn ) if and only if it is expressible by a Σn -formula (respec-
tively a Πn -formula) in L.

Definition 50 If a relation is both Σn and Πn , it is said to be ∆n .

In order to generalize Theorem 48, and for many other purposes, we require a
Σ0 -pairing function.

Lemma 55 (Σ0 pairing function) The function p(m, n) = 21 (m + n + 1)(m + n) +


m is a bijection between the natural numbers and pairs of natural numbers which is
strictly increasing in both arguments, and it is Σ0 .

Proof. Exercise (Problem sheet 3 problem 3). N

Theorem 56 (1) for n > 0, formulas provably equivalent in PA to Σn -formulas


are closed under existential quantification and formulas provably equivalent to
Πn -formulas are closed under universal quantification.

(2) formulas provably equivalent to Σn -formulas, and formulas provably equivalent


to Πn -formulas, are both closed under conjunction and disjunction.
(b) Show that formulas provably ∆n are closed under conjunction, disjunction, and
negation.

Proof. By a single induction on n in the conjunction of the three statements.


(Problem sheet 3 problem 4). N

Corollary 57 Every formula in L is equivalent to a Σn or Πn formula for some n.

Proof. For a given formula, find a prenex normal form for it. By Theorem 56
(3), adjacent like quantifiers can be collapsed to a single quantifier. N
Lecture 7

Σ0-completeness and
Σ1-completeness; weak systems of
arithmetic Q and R (without
induction); Σ0-completeness of
systems R, Q, and PA;
Σ0-soundness and Σ1-soundness

(Tuesday, 2 November 2010)

7.1 Σ0-completeness and Σ1-completeness


Definition 51 (Σ0 -completeness) A system S is Σ0 -complete iff for each true
Σ0 -sentence X, S ⊢ X.

Definition 52 (Σ1 -completeness) A system S is Σ1 complete iff for each true


Σ1 -sentence X, S ⊢ X.

Proposition 58 A system is Σ1 -complete iff it is Σ0 -complete.

Proof. Assume S is Σ0 -complete and let X be a true Σ1 -sentence, i.e. a sentence


of the form ∃vi F (vi ) where F (vi ) is a Σ0 -formula. Since X is true, for some number
k, F (k) is a true Σ0 -sentence. By Σ0 -completeness of S, S ⊢ F (k). Then by
predicate logic in S, S ⊢ ∃vi F (vi ).

58
LECTURE 7 59

Assume S is Σ1 -complete and let X be a true Σ0 -sentence. Then ∃v1 (X ∧v1 = v1 )


is a true Σ1 -sentence, so S ⊢ ∃v1 (X ∧ v1 = v1 ). Since X is a sentence, v1 does not
occur free in X, so (∃v1 (X ∧ v1 = v1 ) ⊃ (X ∧ ∃v1 v1 = v1 )) is logically valid and
hence provable in S. So by propositional logic in S, S ⊢ X. N
In this lecture we shall see that PA and much weaker subsystems of PA are
Σ0 -complete.

Definition 53 (A system S correctly decides a sentence X) A system S cor-


rectly decides a sentence X iff either X is true and S ⊢ X, or X is false and S ⊢∼ X.

Lemma 59 A system S is Σ0 -complete iff S correctly decides every Σ0 -sentence.

Proof. Half of the condition that S correctly decides every Σ0 -sentence is that
if X is a true Σ0 -sentence then S ⊢ X, i.e. Σ0 -completeness of S.
Conversely, suppose S is Σ0 -complete, and X is any false Σ0 -sentence. Then
∼ X is a true Σ0 -sentence, so by Σ0 -completeness, S ⊢∼ X, as required. N

Proposition 60 The following two conditions on a system S together imply that S


correctly decides every Σ0 -sentence.
C1 . S correctly decides every atomic sentence.
C2 . For any Σ0 -formula F (vi ) with vi the only free variable and for every number
n, if S ⊢ F (0), . . . , S ⊢ F (n), then S ⊢ (∀vi ≤ n)F (vi ).

Proof. By induction over the recursive definition of Σ0 -formulas.


1. By C1 S correctly decides all atomic Σ0 -sentences.
2. By propositional logic in S, if X and Y are correctly decided by S, then ∼ X
and X ⊃ Y are correctly decided by S. [Exercise]
3. Any Σ0 -sentence Z that is neither atomic nor of the form ∼ X or X ⊃ Y
must be of the form (∀vi ≤ n)F (vi ) where F (vi ) is a Σ0 -formula of lower degree
than Z and contains vi as its only free variable.
Suppose (∀vi ≤ n)F (vi ) is true. That means that each of the sentences F (0), . . . , F (n)
is true. Then by induction hypothesis, S ⊢ F (0), . . . , S ⊢ F (n). Then by condition
C2 , S ⊢ (∀vi ≤ n)F (vi ).
Suppose (∀vi ≤ n)F (vi ) is false. Then for some m ≤ n, F (m) is false. Then
by induction hypothesis we have that S ⊢∼ F (m). Since m ≤ n is a true atomic
Σ0 -formula, by C1 , S ⊢ m ≤ n. Then by 2. S ⊢∼ (m ≤ n ⊃ F (m)).
The following formula is logically valid:

(∀vi (vi ≤ n ⊃ F (vi )) ⊃ (m ≤ n ⊃ F (m)))

and hence provable in S, so by propositional logic in S,


S ⊢∼ ∀vi (vi ≤ n ⊃ F (vi )), which is to say S ⊢∼ (∀vi ≤ n)F (vi ). N
LECTURE 7 60

Proposition 61 The following three conditions on a system S jointly imply that S


is Σ0 -complete.
D1 . All true atomic sentences are provable in S.
D2 . For any distinct numbers m and n, S ⊢∼ m = n.
D3 . For any variable vi and any number n, S ⊢ (vi ≤ n ⊃ (vi = 0∨. . .∨vi = n)).

Proof. We show that conditions D1 , D2 , D3 imply conditions C1 , C2 of the


previous Proposition, which establishes that S correctly decides every Σ0 -sentence,
and is thereby Σ0 -complete.
Specifically, D1 , D2 , D3 together imply C1 , and D3 implies C2 .
1. To establish C1 , i.e. that S correctly decides all atomic sentences.
We are given by D1 that all true atomic sentences are provable in S. So it
remains to show that all false atomic sentences are refutable in S.
(i) If the false atomic sentence is of the form t1 = t2 , there are true atomic
sentences of the form t1 = m and t2 = n, where m 6= n. Then by D1 , S ⊢ t1 = m,
and S ⊢ t2 = n, and by D2 S ⊢∼ m = n. Then by propositional logic in S,
S ⊢∼ t1 = t2 .
(ii) If the false atomic sentence is of the form t1 ≤ t2 , there are true atomic
sentences of the form t1 = m and t2 = n, where ∼ m ≤ n. From D1 and the fact
that t1 = m and t2 = n are true, S ⊢ t1 = m and S ⊢ t2 = n. Since ∼ m ≤ n, all the
sentences m = 0, . . . , m = n are false. Hence by D2 , S ⊢∼ m = 0, . . . , S ⊢∼ m = n.
Then by propositional logic in S, S ⊢∼ (m = 0 ∨ . . . ∨ m = n).
From D3 , by Generalization and Instantiation,
S ⊢ (m ≤ n ⊃ (m = 0 ∨ . . . ∨ m = n)). Hence by propositional logic in S,
S ⊢∼ m ≤ n. Then by substitutivity of identity, ⊢∼ t1 ≤ t2 .
2. To show that D3 implies C2 .
Suppose F (vi ) is a Σ0 formula with vi its only free variable, and that n is a
number such that S ⊢ F (0), . . . , S ⊢ F (n). Then by pure logic (with identity) in
S, S ⊢ (vi = 0 ⊃ F (vi )), . . . , S ⊢ (vi = n ⊃ F (vi )). Then by propositional logic in
S, S ⊢ ((vi = 0 ∨ . . . ∨ vi = n) ⊃ F (vi )). Then by D3 and propositional logic in S
(transitivity of ⊃), S ⊢ (vi ≤ n ⊃ F (vi )). Then by pure logic (Generalization) in S,
S ⊢ ∀vi (vi ≤ n ⊃ F (vi )), i.e. S ⊢ (∀vi ≤ n)F (vi ). N

7.2 Weak systems of arithmetic Q and R (without


induction)
Definition 54 (system Q) The system Q is obtained from the system PA by drop-
ping the axiom schema for induction, N12 . Thus Q has only finitely many (nine)
nonlogical axioms:
N1 (v1′ = v2′ ⊃ v1 = v2 )
LECTURE 7 61

N2 ∼ v1′ = 0
N3 v1 + 0 = v1
N4 v1 + v2′ = (v1 + v2 )′ .
N5 v1 · 0 = 0
N6 v1 · v2′ = (v1 · v2 ) + v1
N7 (v1 ≤ 0 ≡ v1 = 0)
N8 (v1 ≤ v2′ ≡ (v1 ≤ v2 ∨ v1 = v2′ )).
N9 (v1 ≤ v2 ∨ v2 ≤ v1 ).

The system Q is a variant of one due to Raphael Robinson. We will show that
all true Σ0 -sentences are provable Q and so provable in PA since all the axioms of
Q are axioms of PA. We prove this by proving a yet stronger result, namely that an
even weaker system R, also due to Raphael Robinson, is Σ0 -complete. Instead of
the (finitely) many recursion axioms of PA and Q, it has as axioms infinitely many
instances of computations of addition, multiplication, and inequality, in three axiom
schemata, and two axiom schemata expressing properties of ≤.

Definition 55 (system R) The axioms of R are all sentences sentences and for-
mulas of L generated from natural numbers m and n by the following axiom schemata:
Ω1 m + n = m + n.
Ω2 m · n = m · n.
Ω3 ∼ m = n where m 6= n.
Ω4 (v1 ≤ n ⊃ (v1 = 0 ∨ . . . ∨ v1 = n)).
Ω5 (v1 ≤ n ∨ n ≤ v1 ).

Note that by our notational conventions the symbol + in the formulation of Ω1 is


used to abbreviate expressions in the formal language L and does not actually occur
in sentences of the form Ω1 , and also that + abbreviates different formal expressions
in its two occurrences in the formulation of Ω1 . The instances of Ω1 are generated
from pairs of natural numbers, m and n, by writing an equation between the term
(mf′ n) on the left and the term 0 with m + n many occurrences of the symbol ′
suffixed to it on the right. A corresponding remark holds concerning occurrences of
the dot symbol for multiplication in the above formulation of Ω2 , e.g. an instance
of Ω2 is (0′′ f′′ 0′′′ ) = 0′′′′′′ .

We now show that the system R proves the converse of Ω4 .

Lemma 62 For each natural number n, Ω5 ⊢ n ≤ n.

Proof. By ∀-Intro and ∀-Elim on the instance of Ω5 for n, Ω5 ⊢ (n ≤ n ∨ n ≤ n),


so by propositional logic, Ω5 ⊢ n ≤ n. N
LECTURE 7 62

Lemma 63 For each number n, (R) ⊢ (v1 = n ⊃ v1 ≤ n).

Proof.

(1) v1 = n ⊃ (n ≤ n ⊃ v1 ≤ n) substitutivity of identity


(2) (2) v1 = n Assumption
(2) (3) (n ≤ n ⊃ v1 ≤ n) (1) (2) ⊃-elimination
(4) n ≤ n Lemma 62
(2) (5) v1 ≤ n (2)(4) ⊃-elimination
(6) v1 = n ⊃ v1 ≤ n (2)(5) ⊃-introduction N


Lemma 64 For each number k, (R) ⊢ ((v1 = 0 ∨ . . . ∨ v1 = k) ⊃ v1 ≤ k )

Proof. Let m be any number ≤ k


′ ′
(1) k 6= 0, . . . , k 6= m Ω3
′ ′
(2) (k 6= 0 ∧ . . . ∧ k 6= m) (2) ∧-introduction
′ ′
(3) ∼ (k = 0 ∨ . . . ∨ k = m) (3) DeMorgan law
′ ′ ′
(4) (k ≤ m ⊃ (k = 0 ∨ . . . ∨ k = m)) Ω4 ∀-introduction and elimination

(5) ∼k ≤m (3), (4) prop. logic

(6) (v1 = m ⊃∼ k ≤ v1 ) (5) substitutivity of identity

(7) ((v1 = 0 ∨ . . . ∨ v1 = k) ⊃∼ k ≤ v1 ) (6), for each m ≤ k, by ∨-elimin
′ ′
(8) (v1 ≤ k ∨ k ≤ v1 ) instance of Ω5

(9) ((v1 = 0 ∨ . . . ∨ v1 = k) ⊃ v1 ≤ k ) (7), (8) prop. logic N

Theorem 65 R ⊢ ((v1 = 0 ∨ . . . ∨ v1 = n) ⊃ v1 ≤ n).

Proof. The results follows by ∨-elimination from (n = 0 ∨ n 6= 0). The case for
n = 0 is an instance of Lemma 63, i.e. R ⊢ (v1 = 0 ⊃ v1 ≤ 0). If n 6= 0 then there
is some number k such that n = k + 1. This case follows by ∨-elimination from
′ ′
the disjunction ((v1 = 0 ∨ . . . ∨ v1 = k) ∨ v1 = k ). We have v1 ≤ k from the first

disjunct by Lemma 64 and v1 ≤ k from the second disjunct by Lemma 63. So by
′ ′
⊃-Intro, R ⊢ (((v1 = 0 ∨ . . . ∨ v1 = k) ∨ v1 = k ) ⊃ v1 ≤ k ). N

7.3 Σ0-completeness of systems R, Q, and PA


We will now show the R is Σ0 -complete. To do this we need the following lemma.
LECTURE 7 63

Lemma 66 (evaluation of closed terms by R) For each closed term t in L,


there is a unique numeral n such that R ⊢ t = n.

Proof. If t is a closed term it is either a numeral, or there is a closed term t1


such that t is the term t′1 , or there are closed terms t1 and t2 such that t is (t1 f′ t2 ),
i.e. t1 + t2 , or t is (t1 f′′ t2 ), i.e. t1 · t2 . If t is a numeral n then R ⊢ t = n by reflexivity
of identity. If t is of the form t′1 for some term t1 , then by induction hypothesis there
is a numeral n such that R ⊢ t1 = n. Then by the logic of identity, R ⊢ t′1 = n′ .
But n′ is n + 1, which is to say that R ⊢ t = n + 1. If t is of the form t1 + t2 ,
then by induction hypothesis there are numbers m1 and m2 such that R ⊢ t1 = m1
and R ⊢ t2 = m2 . Let m1 + m2 = k. Then by Ω1 , R ⊢ m1 + m2 = k. Then by
substitutivity of identity,R ⊢ t1 + t2 = k. The argument is the same for t of the
form t1 · t2 .
The uniqueness of n for a given term t follows by transitivity of identity. N

Proposition 67 The system R is Σ0 -complete.

Proof. We establish this result by showing that R satisfies the conditions D1


D2 D3 of Proposition 61.
D1 . (i) Suppose t1 and t2 are terms such that t1 = t2 is a true sentence. Since
t1 = t2 is a sentence, the terms t1 and t2 contain no variables. Hence by Lemma 66
there are numbers m1 and m2 such that R ⊢ t1 = m1 and R ⊢ t2 = m2 . Since t1 = t2
is true, m1 = m2 , so m1 and m2 are the same numeral, so by reflexivity of identity
in R, R ⊢ m1 = m2 . Hence by transitivity of identity in R, R ⊢ t1 = t2 .
(ii) Suppose t1 and t2 are terms such that t1 ≤ t2 is a true sentence. Then t1
and t2 have no free variables, so by Lemma 66 there are natural numbers m1 and
m2 such that t1 = m1 and t2 = m2 are true sentences, and so by (i) provable in S.
Since t1 ≤ t2 is true, m1 ≤ m2 . By the logic of identity, R ⊢ m1 = m1 and so by
propositional logic, R ⊢ (m1 = 0 ∨ . . . ∨ m1 = m1 ∨ . . . ∨ m1 = m2 ). By Theorem 65,
with ∀-introduction and ∀-elimination, and Modus ponens, R ⊢ m1 ≤ m2 . Then by
substitutivity of identity, R ⊢ t1 ≤ t2 .
D2 . This is the schema Ω3 .
D3 . This is a direct consequence of Ω4 by ∀-Intro and ∀-Elim. N

Proposition 68 R is a subsystem of Q.

Proof. We must show that each axiom of R, i.e. every instance of Ω1 , Ω2 , Ω3 ,


Ω4 , is provable in Q.
Ω1 : X is of the form m + n = m + n for natural numbers m and n. We argue
by induction on n in the statement Q ⊢ m + n = m + n.
Note that while Q does not contain induction, use of induction for this argument
is valid since we are arguing about Q and not in Q.
LECTURE 7 64

n = 0. By N3 , ∀-I and ∀-E, Q ⊢ m + 0 = m. Since m = m + 0, m and m + 0 are


the same term (formal numeral), so Q ⊢ m + 0 = m + 0.
Induction step.

(1) Q⊢m+n=m+n Induction hypothesis


(2) Q ⊢ m + n′ = (m + n)′ N4 , ∀-I, ∀-E
(3) Q ⊢ m + n′ = (m + n)′ (1), (2), substitutivity of = in Q
(4) Q ⊢ m + n + 1 = ((m + n) + 1) (3), Corollary 8
(5) ((m + n) + 1) = (m + (n + 1)) truth of arithmetic
(6) (m + n) + 1) and (m + (n + 1)) are the same term (5)
(7) Q ⊢ m + n + 1 = m + (n + 1) (4), (6)

Ω2 : We show by induction on n that Q ⊢ m · n = m · n.


n = 0 By N5 , ∀-Intro, ∀-Elim, Q ⊢ m · 0 = 0. Since 0 = m · 0, 0 and m · 0 are
the same term. So Q ⊢ m · 0 = m · 0.
Induction step.

(1) Q⊢m·n=m·n Induction hypothesis


(2) Q ⊢ m · n′ = m · n + m N6 , ∀-I, ∀-E
(3) Q ⊢ m · n′ = m · n + m (1), (2), substitutivity of = in Q
(4) Q⊢m·n+1=m·n+m (3), Corollary 8
(5) Q⊢m·n+1=m·n+m (4), previous case for Ω1
(6) m · n + m = m · (n + 1) truth of arithmetic
(7) m · n + m and m · (n + 1) are the same term (6)
(8) Q ⊢ m · n + 1 = m · (n + 1) (5), (7)

Ω3 : To show that for every m and n such that m 6= n, Q ⊢∼ m = n. Suppose


that m 6= n. Without loss of generality, we may suppose that m > n, since by logic
of identity in Q, if Q ⊢∼ m = n, then Q ⊢∼ n = m. We argue by cases.
n = 0. Then since m 6= 0, there is a number k such that k + 1 = m. By
′ ′
Corollary 8, m is k . By N2 , ∀-Intro, ∀-Elim, Q ⊢∼ k = 0, i.e. Q ⊢∼ m = 0.
n 6= 0. Since m > n, there is a non-zero number d such that m = d + n. By
taking d in place of m in the argument for the previous case, Q ⊢∼ d = 0. By

contraposition of N1 , ∀-Intro, ∀-Elim, Q ⊢ (∼ d = 0 ⊃∼ d = 0′ ), so by Modus

ponens, Q ⊢∼ d = 0′ . By n many applications of this argument,
n
z }| { n
z }| {
′...′
Q ⊢∼ d = 0′ . . . ′ , i.e. Q ⊢∼ m = n.
Ω4 : We show by induction on n that for each n, Q ⊢ (v1 ≤ n ⊃ (v1 = 0∨. . .∨v1 =
n)).
LECTURE 7 65

n = 0. By ∧-Elim from N7 , Q ⊢ (v1 ≤ 0 ⊃ v1 = 0).


Assume, as induction hyposthesis, that Q ⊢ (v1 ≤ n ⊃ (v1 = 0 ∨ . . . ∨ v1 = n)).
By ∧-Elim, ∀-Intro, ∀-Elim from N8 , Q ⊢ (v1 ≤ n′ ⊃ (v1 ≤ n ∨ v1 = n′ )). Then
by ∨-Elim, Q ⊢ (v1 ≤ n′ ⊃ ((v1 = 0 ∨ . . . ∨ v1 = n) ∨ v1 = n′ )), i.e. by Corollary 8,
Q ⊢ (v1 ≤ n + 1 ⊃ (v1 = 0 ∨ . . . ∨ v1 = n ∨ v1 = n + 1)).
Ω5 : N9 ⊢ Ω5 by ∀-Intro and ∀-Elim. N

Proposition 69 Q is Σ0 -complete.

Proof. By Propositions 67 and 68. N

Theorem 70 PA is Σ0 -complete.

Proof. By Proposition 69 and the fact that PA is an extension of Q. N

7.4 Σ0-soundness and Σ1-soundness


Definition 56 (Σ0 -soundness) A system S is Σ0 -sound if and only if for every
Σ0 -sentence X, if S ⊢ X, then X is true (in the structure of the natural numbers).

Definition 57 (Σ1 -soundness) A system S is Σ1 -sound if and only if for every


Σ1 -sentence X, if S ⊢ X, then X is true (in the structure of the natural numbers).

Proposition 71 If a consistent system is Σ0 -complete, it is Σ0 -sound.

Proof. Let S be a Σ0 -complete system and let X be a false Σ0 -sentence such


that S ⊢ X. Since X is Σ0 and false, ∼ X is Σ0 and true. Hence by Σ0 -completeness
of S, S ⊢∼ X. But this means that S is inconsistent, contrary to hypothesis. N
Remark. Though a system is Σ0 -complete if and only if it is Σ1 -complete
(Proposition 58), there is no result corresponding to Proposition 71 that holds for
Σ1 -completeness. As we shall see in Lecture 8, there are Σ1 -complete systems that
are not Σ1 -sound. In any case it is clear that the proof of Proposition 71 does not
extend to the case of Σ1 -completeness since in general the negation of a Σ1 -sentence
is not Σ1 .
Lecture 8

The notions of consistency,


ω-consistency and 1-consistency;
incompleteness from the
assumption of 1-consistency; truth
of the Gödel sentence;
ω-incompleteness.

(Wednesday, 3 November 2010)

8.1 The notions of consistency, ω-consistency and


1-consistency.
Definition 58 (consistency) A system S is consistent if there is no formula X
in the language of S such that S ⊢ X and S ⊢∼ X.

Proposition 72 A system S containing propositional logic is consistent if and only


if there is formula Y in the language of S such that S 0 Y .

Proof. (i) Left to right: We prove the contrapositive. Suppose for every X,
S ⊢ X. Then in particular for any formula Y , S ⊢ Y and S ⊢∼ Y .
(ii) Right to left: We prove the contrapositive. Suppose S is inconsistent, i.e.
there is a formula Y such that S ⊢ Y and X ⊢∼ Y . Then by ∧-introduction (as a
derived rule if not a primitive rule of S), S ⊢ (Y ∧ ∼ Y ). Propositional logic proves

66
LECTURE 8 67

((Y ∧ ∼ Y ) ⊃ Z) for every formula Z. So by Modus Ponens in S, S ⊢ Z for every


formula Z. N
Remark. Proposition 72 shows that we could equivalently have defined consis-
tency by:

Definition 59 (alternative definition of consistency) S is consistent if there


is a formula X such that S 0 X.

Definition 60 (ω-consistency) A system S in a language that contains a closed


term n, i.e. a numeral, for each natural number, is said to be ω-consistent if and
only if there is no formula F (vi ) with one free variable in L such that S ⊢ ∃vi F (vi )
and for each natural number n, S ⊢∼ F (n).

Proposition 73 If a system is ω-consistent, then it is consistent.

Proof. The contrapositive is immediate by ex falso quodlibet: if a system S


is inconsistent S proves every formula in the language of S; in particular for any
formula F (w) with one free variable, S ⊢ ∃wF (w), and for each n S ⊢∼ F (n), i.e.
S is ω-inconsistent. N
The converse of Proposition 73 does not hold, i.e.

Proposition 74 There are consistent systems that are ω-inconsistent.

Proof. Exercise (problem 3(a) of Problem sheet 4). N

Proposition 75 If a system is sound with respect to truth in arithmetic, then it is


ω-consistent.

Proof. Let S be a system whose language contains numerals for the natural
numbers and which is sound with respect to truth in arithmetic. Suppose S ⊢
∃wF (w). Then ∃wF (w) is true, i.e. there is a natural number n such that F (n) is
true, which is to say that ∼ F (n) is false. So S 0∼ F (n), which is to say that S is
ω-consistent. N

The converse holds only to a strictly limited extent, as detailed by the following
two propositions, and in general ω-consistency does not imply truth.

Proposition 76 If a system S is Σ0 -complete and ω-consistent, it is Σ2 -sound, i.e.


if sentence X is Σ2 and S ⊢ X, then X is true.

Proof. Exercise (problem 2(b) of Problem sheet 4). N


Remark This result is best possible, i.e.
LECTURE 8 68

Proposition 77 There is an ω-consistent system that proves a false Σ3 -sentence.

Proof. Exercise (problem 4(c) of Problem sheet 4). N


Gödel introduced the notion of ω-consistency in order to prove the second half
of his First Incompleteness Theorem. It is a very much weaker hypothesis than that
the system is sound, i.e. that all theorems are true (which we saw already in Lec-
ture 1 is sufficient for the result), established by Proposition 77, which shows that
ω-consistency only implies a very limited amount of truth. Even so, ω-consistency
is a considerably stronger hypothesis than is necessary to established formal in-
completeness. From the fact that a proof predicate for a formal deductive system
with arithmetized syntax is Σ1 , the First Incompleteness Theorem can be proved,
as we shall see, with just the assumption that there is no ω-inconsistency with a
Σ1 -formula. Kreisel in 1957 [5] noted that the minimum case of ω-consistency, which
he labeled 1-consistency, is sufficient for the second half of Gödel First Incomplete-
ness Theorem. This special case of ω-consistency perhaps strictly should be labeled
something like Σ1 ω-consistency, but the label 1-consistency introduced by Kreisel
has the virtue of brevity and is standard in the literature.
While 1-consistency is both weaker and more natural than ω-consistency as a
hypothesis for proof of the second half of the First Incompleteness Theorem, we
shall see later (but only after we have proved the Second Incompleteness Theorem)
that 1-consistency is also stronger than necessary for this result.

Definition 61 (1-consistency) A system S in a language that contains a closed


term n, i.e. a numeral, for each natural number, is said to be 1-consistent if and
only if there is no Σ1 -formula ∃vi F (vi ) with one free variable in the language such
that S ⊢ ∃vi F (vi ) and for each natural number n, S ⊢∼ F (n).

Theorem 78 For a Σ0 -complete system, 1-consistency is equivalent to Σ1 -soundness.

Proof. Let S be a Σ0 -complete system.


(i) Suppose S is 1-consistent, let ∃vi F (vi ) be a Σ1 -sentence such that S ⊢
∃vi F (vi ), and suppose ∃vi F (vi ) is false. Then for each number n, ∼ F (n) is true.
Since ∃vi F (vi ) is a Σ1 -sentence, F (vi ) is Σ0 and since Σ0 -formulas are closed under
negation, ∼ F (vi ) is a Σ0 -formula. Hence by Σ0 -completeness of S, for each nat-
ural number n, S ⊢∼ F (n). This violates the hypothesized 1-consistency of S, so
∃vi F (vi ) is true, i.e. S is Σ1 -sound.
(ii) Suppose S is Σ1 -sound, and suppose S ⊢ ∃vi F (vi ). Then by Σ1 -soundness of
S, there is a natural number k such that F (k) is a true Σ0 -sentence, so ∼ F (k) is a
false Σ0 -sentence which, by Lemma 43, is logically equivalent to a false Σ1 -sentence.
Then by Σ1 -soundness of S, S 0∼ F (k). Hence S cannot prove a 1-inconsistency,
i.e. S is 1-consistent. N
LECTURE 8 69

Note that in (ii) of the proof of Theorem 78 the hypothesis that S is Σ0 -complete
is not needed. This part of the proof of Theorem 78 is a sharpening of the argument
for Proposition 75.

Corollary 79 Every 1-consistent system is consistent.

Proof. By the first half of Theorem 78 and the fact that a Σ1 -sound system is
consistent (since there are sentences it doesn’t prove, namely false Σ1 -sentences).
Or we can prove the contrapositive, as in the proof of Proposition 73, i.e. an
inconsistent theory proves everything so in particular a 1-inconsistency. N

The proof of incompleteness from 1-consistency is strictly a stronger result than


proving incompleteness from ω-consistency, in that

Theorem 80 1-consistency is strictly weaker than ω-consistency.

Proof. Exercise problem 3(b) on Problem sheet 4. N

However, Gödel’s original proof made no use of the extra strength of ω-consistency
over 1-consistency, so the proof from 1-consistency is really an improvement in clar-
ity rather than in strength.

Definition 62 (2-consistency) A system S in a language that contains a closed


term n, i.e. a numeral, for each natural number, is said to be 2-consistent if and
only if there is no Σ2 -formula ∃vi F (vi ) with one free variable in the language such
that S ⊢ ∃vi F (vi ) and for each natural number n, S ⊢∼ F (n).

Lemma 81 If a system is Σ2 -sound, then it is Σ1 -sound

Proof. The proof is by vacuous quantification. Let S ⊢ ∃v1 F (v1 ) for F (v1 ) a
Σ0 -sentence (i.e. no free variables). Then S ⊢ ∃v1 ∀v2 F (v1 ), and so by Σ2 -soundness,
∃v1 ∀v2 F (v1 ) is true. Then ∃v1 F (v1 ) is true. N

Theorem 82 For a Σ0 -complete system, 2-consistency is equivalent to Σ2 -soundness.

Proof. (i) Suppose S is 2-consistent and suppose S ⊢ ∃v1 ∀v2 F (v1 , v2 ), where
F (v1 , v2 ) is a Σ0 -formula, and ∃v1 ∀v2 F (v1 , v2 ) is false, which is to say that for
each natural number n, ∃v2 ∼ F (n, v2 ) is a true Σ1 -sentence. Then by the Σ1 -
completeness of every Σ0 -complete theory and predicate logic, for each natural num-
ber n, S ⊢∼ ∀v2 F (n, v2 ). But then S is 2-inconsistent. So by RAA, ∃v1 ∀v2 F (v1 , v2 )
is true.
(ii) Suppose S is Σ2 -sound and suppose S ⊢ ∃v1 ∀v2 F (v1 , v2 ). Then ∃v1 ∀v2 F (v1 , v2 )
is true, so for some number k, ∀v2 F (k, v2 ) is true. Suppose S ⊢∼ ∀v2 F (k, v2 ).
LECTURE 8 70

Then S ⊢ ∃v2 ∼ F (k, v2 ). Since S is Σ2 -sound, by Lemma 81 it is Σ1 -sound, so


∃v2 ∼ F (k, v2 ) is true. But this contradicts the truth of ∀v2 F (k, v2 ), so by RAA,
S 0∼ ∀v2 F (k, v2 ). This means that S is 2-consistent. N

Remark. We could define 3-consistency and n-consistency for larger n exactly


as for 1- and 2-consistency, but these notions are not natural in the way 1- and
2-consistency are since there is no equivalence with truth corresponding to Theo-
rems 78 and 82. Since ω-consistency implies n-consistency for each n and so in
particular 3-consistency, by Theorem ?? and Proposition ??, there is a 3-consistent
system that is not Σ3 -sound.

8.2 Incompleteness of PA from the assumption of


1-consistency
We are now in a position to prove Gödel’s First Incompleteness Theorem. This
theorem is significantly stronger than the version of incompleteness we have already
proved, Theorem 27, in that it proves this result from the hypothesis that PA is
1-consistent, which is strictly weaker than the hypothesis that PA is sound, i.e. that
everything PA proves is true (in the structure of the natural numbers). It is also a
purely syntactic (finitary) property as opposed to a semantic (infinitary) property.
(We replace Gödel’s hypothesis of ω-consistency by the strictly weaker hypothesis
of 1-consistency, so the theorem is stronger, but the argument used is exactly the
same as for Gödel’s result.)
Since we are no longer working with the notion of truth (in the structure of
the natural numbers), we cannot use the Diagonal Lemma, Theorem 17, which
establishes the truth of the diagonal equivalence. However, we use the construction
of the diagonal sentence for the one-place formula ∼ P rP A (v1 ) from the proof of
the Diagonal Lemma to obtain the same Gödel sentence for this theorem as for the
weaker one.

Theorem 83 (Gödel’s First Incompleteness Theorem for PA) There is a Π1 -


sentence G such that
1. If PA is consistent, PA 0 G, and
2. If PA is 1-consistent, PA 0∼ G.

Proof. By Proposition 54, {n : PA ⊢ En [n]} is Σ1 . Let ∃v2 A(v1 , v2 ) be a Σ1 -


formula that expresses {n : PA ⊢ En [n]}. Let a =df p∀v2 ∼ A(v1 , v2 )q, and let
G =df ∀v2 ∼ A(a, v2 ).
1. Suppose PA ⊢ G. By Lemma 12 and the completeness of first-order logic of
PA, PA ⊢ (∀v2 ∼ A(a, v2 ) ≡ ∀v2 ∼ A([a], v2 )), so then PA ⊢ ∀v2 ∼ A([a], v2 ). Then
LECTURE 8 71

a ∈ {n : PA ⊢ En [n]}. Since ∃v2 A(v1 , v2 ) expresses {n : PA ⊢ En [n]}, ∃v2 A(a, v2 )


is true. Since PA is Σ0 -complete and hence Σ1 -complete (Proposition 58), PA ⊢
∃v2 A(a, v2 ). This contradicts the assumption that PA is consistent. So PA 0 G.
2. Suppose PA ⊢∼ G, i.e. PA ⊢ ∃v2 A(a, v2 ). From the assumed 1-consistency
of PA, PA is Σ1 -sound, by Theorem 78, so ∃v2 A(a, v2 ) is true. Since ∃v2 A(v1 , v2 )
expresses {n : PA ⊢ En [n]}, PA ⊢ Ea [a], i.e. PA ⊢ G, which with the assumption
of this argument means that PA is inconsistent. But by the condition that PA is
1-consistent, PA is consistent (Corollary 79), so PA 0∼ G. N

8.3 Truth of the Gödel sentence


By the bivalence of truth, one or other of G and∼ G is true in the structure of
the natural numbers. Though PA cannot decide G, i.e. does not prove G and does
not prove ∼ G, we have good reason to hold that G is true (in the structure of the
natural numbers) by the following considerations.

Lemma 84 G is true if and only if P A 0 G.

Proof. By Proposition 54, there is a Σ1 -formula, ∃v2 A(v1 , v2 ), that expresses


{n : PA ⊢ En [n]}, i.e. PA ⊢ En [n] if and only if ∃v2 A(n, v2 ) is true. Hence for
a =df p∀v2 ∼ A(vi , v2 )q, and G =df ∀v2 ∼ A(a, v2 ), PA ⊢ G if and only if ∃v2 A(a, v2 )
is true, which is to say that PA ⊢ G if and only if ∼ G is true. So by contraposition,
PA 0 G if and only if G is true. N

Theorem 85 If PA is consistent, G is true.

Proof. This is equivalent to the first half of Theorem 83, by Lemma 84. N

Corollary 86 Theorem 85 gives a variant proof of the second half of Theorem 83,
the First Incompleteness Theorem.

Proof. Assume PA is 1-consistent. Then PA is consistent, so by Theorem 85,


G is true. On the assumption that PA is 1-consistent, PA is Σ1 -sound. Then, since
∼ G is a false Σ1 -sentence, PA 0∼ G. N
Remarks (1) By Theorem 85, we are justified in holding that the Gödel sen-
tence for PA is true insofar as we are justified in our conviction (universal among all
mathematicians except a few with very quirky views) that PA is consistent. Identi-
fying the basis of our conviction that PA is consistent lies outside the scope of these
lectures.
(2) Formalizing the proof of Theorem 85 in PA shows that if PA is consistent
it cannot prove its own consistency, since otherwise it would prove G, which we
LECTURE 8 72

have just shown, by Theorem 83 it cannot if it is consistent. This is Gödel’s Second


Incompleteness Theorem. Formalizing the proof of Theorem 85 in PA requires some
hard work, the hardest of which is formalizing the proof that PA is Σ1 -complete
(provable Σ1 -completeness), which I will do in Lecture 11.
(3) Having shown that G for PA can be seen to be true on the basis of the
accepted consistency of PA, it is important to realize that there is no weaker basis
on which to hold that G is true than that PA is consistent, i.e. G for PA cannot be
established as true on the basis of any considerations that do not also establish the
consistency of PA, by the converse of Theorem 85, which also holds.

Theorem 87 The truth of G implies the consistency of PA.

Proof. Suppose G, i.e. ∀v2 ∼ A(a, v2 ), is true. Then every number is not the
code of a proof of Ea [a], which is to say that PA 0 G. But if there is any sentence
that a system doesn’t prove then the system is consistent, which is to say that PA
is consistent. N

There is another argument to show that G is true which is weaker than the ar-
gument for Theorem 85 because it requires a stronger hypothesis, but the argument
itself is of independent interest.

Theorem 88 If a system is Σ0 -complete and does not prove ∃vi F (vi ) for F (vi ) a
Σ0 -formula, then ∃vi F (vi ) is false.

Proof. This theorem is the contrapositive of the implication that every Σ0 -


complete system is Σ1 -complete (Proposition 58), i.e. if ∃vi F (vi ) is true, then for
some number k, F (k) is true and hence provable in any Σ0 -complete system, which
then by predicate logic also proves ∃vi F (vi ). N
Theorem 88 tells us that G is true from the fact that PA 0∼ G. However, we
established that PA 0∼ G on the assumption that PA is 1-consistent, and as we
have seen, the truth of G follows from the fact that PA 0 G, which requires only
the assumption that PA is consistent. So this argument is inefficient as a means to
establishing that G is true.

Proposition 89 For S any Σ0 -complete, consistent, Σ1 -axiomatized theory, the


Gödel sentence for S is true.

Proof. Let S be any Σ0 -complete, consistent, arithmetized theory that has a


Σ1 -predicate that expresses {n : S ⊢ En }. All of the results of the previous two
sections generalize from PA to S. In particular, for any such theory S, the Gödel
sentence G for S is a true Π1 -sentence and ∼ G is a false Σ1 -sentence. N
LECTURE 8 73

8.4 PA is ω-incomplete
Definition 63 (ω-completeness) A system S in a language containing a numeral
n for each natural number n is ω-complete if for every formula F (v1 ) in the language
of S such that for each natural number n S ⊢ F (n), S ⊢ ∀vi F (vi ) If S is not ω-
complete we say that S is ω-incomplete.

Theorem 90 If PA is consistent, PA is ω-incomplete.

Proof. We have shown that if PA is consistent, PA 0 G, i.e. PA 0 ∀v2 ∼ A(a, v2 ).


We have also seen that if PA is consistent then G is true, i.e. for each natural number
n, ∼ A(a, n) is true. The sentences ∼ A(a, n) are Σ0 . Hence by the Σ0 -completeness
of PA, for each n, PA ⊢∼ A(a, n). Hence PA is ω-incomplete. N
Lecture 9

Enumerability and the Separation


Lemma; incompleteness of PA
from the assumption of
consistency (Rosser’s Theorem);
weak and strong definability of a
function in a system; formal
provability of the Diagonal Lemma

(Tuesday, 9 November 2010)

We have noted that since an inconsistent system proves everything, consistency


of a system S is a necessary condition for S 0 X, for any sentence X and in particular
for G the Gödel sentence for S. We have also seen that consistency is a sufficient
condition for S 0 G, where S is any system that can arithmetize its own syntax. The
condition of ω-consistency is, as we have seen, a stronger condition than consistency
(and a weaker condition than that every sentence provable in the system is true). We
saw that 1-consistency arises in a natural way as a condition sufficient to establish
that S 0∼ G for G the Gödel sentence for S, while consistency of S is sufficient to
establish that S 0 G. There turns out to be a form of incompleteness, discovered by
J. Barkley Rosser (1936), that is symmetric with respect to negation. This result
is of definite interest, though it does not supersede Gödel’s First Incompleteness
Theorem, i.e. it does not show that the Gödel sentence for S is undecidable in S
on the assumption that S is simply consistent, and indeed to show that S 0∼ G
requires a stronger hypothesis than just that S is consistent. Rosser constructed

74
LECTURE 9 75

a different sentence from the Gödel sentence which is provably undecidable just on
the hypothesis of consistency of the system.
The Rosser Incompleteness Theorem can be proved from a separation property,
itself of independent interest and which we use also in proving that the diagonal
equivalence in the diagonal lemma is not only true, as we have seen, but also formally
provable.

9.1 Enumerability and the Separation Lemma


Definition 64 (enumeration of a relation by a formula in a theory) A k-place
relation R ⊆ Nk is enumerated by a formula F (v1 , . . . , vk , vk+1 ) in a system S if and
only if:
1. If hn1 , . . . , nk i ∈ R, then there exists a number m such that S ⊢ F (n1 , . . . , nk , m).
(We say that m is a witness to the fact that hn1 , . . . , nk i ∈ R.)
/ R, then for every number m, S ⊢∼ F (n1 , . . . , nk , m).
2. If hn1 , . . . , nk i ∈

Definition 65 A formula F (v1 , . . . , vk ) separates a k-ary relation A from a k-ary


relation B in a system S if and only if for all (n1 , . . . nk ) ∈ A, S ⊢ F (n1 , . . . , nk ),
and for all (n1 , . . . nk ) ∈ B, S ⊢∼ F (n1 , . . . , nk ).

Lemma 91 (1) If F (v1 , . . . , vk ) separates A from B in S, then ∼ F (v1 , . . . , vk )


separates B from A in S. (2) If F (v1 , . . . , vk ) separates A from B in S and S is
consistent, then F (v1 , . . . , vk ) does not separate B from A in S. (3) If F (v1 , . . . , vk )
separates A from B in S and S is consistent, then A and B are disjoint. (4) If S is
inconsistent, then for any formula F (v1 , . . . , vk ) and any k-ary relations A and B,
F (v1 , . . . , vk ) separates A from B in S.

Proof. Exercise. N

Theorem 92 (Separation Lemma) Let S be a system in which Ω4 and Ω5 hold,


and let A and B be disjoint k-ary relations enumerated in S by formulas F (v1 , . . . , vk , x)
and G(v1 , . . . , vk , x), respectively. Then the formula

∃x(F (v1 , . . . , vk , x) ∧ (∀y ≤ x) ∼ G(v1 , . . . , vk , y))

separates A from B in S.

Proof. In order to shorten formulas on the page I shall take A and B to be unary
relations. The proof for A and B as k-ary relations is just a notational variant of
this proof.
i) To show: if n ∈ A, then S ⊢ ∃x(F (n, x) ∧ (∀y ≤ x) ∼ G(n, y)).
LECTURE 9 76

(1) n ∈ A Assumption
(2) there exists k such that S ⊢ F (n, k) (1) and enumeration of A by F (v1 , v2 ) in S
(3) n ∈
/B (1) and the hypothesis that A and B as disjoint
(4) for every m, S ⊢∼ G(n, m) (3) and enumeration of B by G(v1 , v2 ) in S
(5) S ⊢ ((y = 0 ∨ . . . ∨ y = k) ⊃∼ G(n, y)) (4) and substitutivity of =, ∨-elim, ⊃-intro
(6) S ⊢ (y ≤ k ⊃∼ G(n, y)) (5), instance of Ω4 and prop. logic
(7) S ⊢ ∀y(y ≤ k ⊃∼ G(n, y)) (6) by ∀-Intro
(8) S ⊢ (F (n, k) ∧ ∀y(y ≤ k ⊃∼ G(n, y))) (2) (7) ∧-Intro
(9) S ⊢ ∃x(F (n, k) ∧ ∀y(y ≤ k ⊃∼ G(n, y))) (8) ∃-intro
(10) S ⊢ ∃x(F (n, x) ∧ (∀y ≤ x) ∼ G(n, y)) (9) definition of (∀y ≤ x)

(ii) To show: if n ∈ B then S ⊢∼ ∃x(F (n, x) ∧ (∀y ≤ x) ∼ G(n, y)), which is


logically equivalent to S ⊢ ∀x(F (n, x) ⊃ (∃y ≤ x)G(n, y))

(1) n ∈ B Assumption
(2) there exists k such that S ⊢ G(n, k) (1) and enumeration of B by G(v1 , v2 ) in S
(3) n ∈
/A (1) and the hypothesis that A and B as disjoint
(4) for every m, S ⊢∼ F (n, m) (3) and enumeration of A by F (v1 , v2 ) in S
(5) S ⊢ ((y = 0 ∨ . . . ∨ y = k) ⊃∼ F (n, y)) (4) and substitutivity of =, ∨-elim, ⊃-intro
(6) S ⊢ (y ≤ k ⊃∼ F (n, y)) (5), instance of Ω4 and prop. logic
(7) S ⊢ (F (n, y) ⊃∼ y ≤ k) (6) by prop logic (contraposition)
(8) S ⊢ (F (n, y) ⊃ k ≤ y) (7) by Ω5 and prop logic
(9) S ⊢ (F (n, y) ⊃ (k ≤ y ∧ G(n, k))) (2) (8) propositional logic
(10) S ⊢ ∃y(F (n, x) ⊃ (y ≤ x ∧ G(n, y))) (9) ∃-Intro
(11) S ⊢ (F (n, x) ⊃ ∃y(y ≤ x ∧ G(n, y))) (10) predicate logic (anti-prenexing)
(12) S ⊢ ∀x(F (n, x) ⊃ ∃y(y ≤ x ∧ G(n, y))) (11) ∀-Intro
(13) S ⊢ ∀x(F (n, x) ⊃ (∃y ≤ x)G(n, y))) (12) definition of (∃y ≤ x) N

Note that in the proof of the Separation Lemma, the argument for (ii) uses both
Ω4 and Ω5 while the argument for (i) uses just Ω4 .

9.2 Incompleteness of PA from the assumption of


consistency (Rosser’s Theorem)
Theorem 93 If a formula H(v1 ) separates {n : S ⊢∼ En [n]} from {n : S ⊢ En [n]}
in a consistent axiomatizable system S and S, then for h = pH(v1 )q, S 0 H(h) and
S 0∼ H(h).
LECTURE 9 77

Proof. (i) Suppose S ⊢ H(h). Then since S ⊢ (H(h) ≡ H[h]),


h ∈ {n : S ⊢ En [n]}. Then by the separation property of H(v1 ), S ⊢∼ H(h). This
contradicts the assumption S is consistent. So S 0 H(h).
(ii) Suppose S ⊢∼ H(h). Then h ∈ {n : S ⊢∼ En [n]}. Then by the separation
property of H(v1 ), S ⊢ H(h). This contradicts the assumption that S is consistent.
So S 0∼ H(h). N

Lemma 94 If S be a consistent axiomatizable extension of R in which the formula


P d(v1 , v2 ) enumerates {n : S ⊢ En [n]} and the formula Rd(v1 , v2 ) enumerates
{n : S ⊢∼ En [n]}, then the formula

∀v1 (P d(v3 , v1 ) ⊃ (∃v2 ≤ v1 )Rd(v3 , v2 ))

separates {n : S ⊢∼ En [n]} from {n : S ⊢ En [n]}.

Proof. By Theorem 92 and Lemma 91 (1). N

Theorem 95 (Rosser’s Theorem) There is a sentence R such that if PA is con-


sistent, PA 0 R and PA 0∼ R.

Proof. By Theorem 93, Lemma 94, and the fact that the sets {n : PA ⊢ En [n]}
and {n : PA ⊢∼ En [n]} are enumerable in PA.

9.3 Weak and strong definability of a function in


a system
Definition 66 (weak definability of a function in a system) A function f :
Nn → N is weakly definable in a system S iff there is a formula F (v1 , . . . , vn , vn+1 )
such that.
(1) If f (a1 , . . . , an ) = b, then S ⊢ F (a1 , . . . , an , b).
(2) If f (a1 , . . . , an ) 6= b, then S ⊢∼ F (a1 , . . . , an , b).

This notion is also sometimes called expressibility (e.g. Elliott Mendelson, In-
troduction to Mathematical Logic, 4th edn, Chapman and Hall, 1997, p. 170), or
numeralwise expressibility (e.g. Stephen Cole Kleene, Introduction to Metamathe-
matics, D. Van Nostrand, 1950, p. 195).

Definition 67 (strong definability of a function in a system) A function f :


Nn → N is strongly definable in a system S iff there is a formula G(v1 , . . . , vn , vn+1 )
such that if f (a1 , . . . , an ) = b, then

S ⊢ (G(a1 , . . . , an , b) ∧ ∀vn+1 (G(a1 , . . . , an , vn+1 ) ⊃ vn+1 = b)).


LECTURE 9 78

It’s easy to show that strong definability implies weak definability (Proposi-
tion 96). The converse also holds but is considerably more complicated to prove
(Theorem 97).

Proposition 96 Let S be a system in which Ω3 holds. If a function is strongly


definable in S then it is weakly definable in S.

Proof. If f : Nn → N is strongly defined in S by the formula F (v1 , . . . , vn , vn+1 ),


then it is weakly defined in S by the same formula. Assume that f (a1 , . . . , an ) = b.
Then by the first conjunct of the condition for strong definability, S ⊢ F (a1 , . . . , an , b),
which is condition (1) of weak definability. Suppose c 6= b, so f (a1 , . . . , an ) 6= c. By
∀-elimination from the second conjunction of the condition for strong definability
in S, S ⊢ F (a1 , . . . , an , c) ⊃ c = b. If c 6= b, then by Ω3 , S ⊢∼ c = b. So by
propositional logic in S, S ⊢∼ F (a1 , . . . , an , b), i.e. clause (2) of the definition of
weak definability of a function.
We will now show, just with the use of Ω4 and Ω5 , that any function weakly
definable in a system S is strongly definable in S. Weak definability of a function
f (x) = y in S is the condition that there is a formula F (v1 , v2 ) that separates in S
the graph of f (x) = y from its complement. The first conjunct of the condition for
strong definability is the same as the first clause for weak definability. The second
conjunct expresses the functionality condition for a given argument of the function,
i.e. that only number that bears the defining relation to the given argument is
the value of the function for that argument. The way we do this is to define a
new formula from the formula that weakly defines the function with the additional
condition that the relationship between the argument and a number holds just in
case it’s the least number for which the weak definability relation holds.

Theorem 97 If S is an extension of {Ω4 , Ω5 }, then any function weakly definable


in S is strongly definable in S.

Proof. To reduce clutter I give the proof for the case of a unary function, which
is also the case of immediate interest since the diagonal function is unary. The
argument for the general case is a notational variant.
Let F (x, y) be a formula that weakly defines f (x) in S, i.e.
(1) If f (a) = b, then S ⊢ F (a, b).
(2) If f (a) 6= b, then S ⊢∼ F (a, b).
Let G(x, y) be the formula (F (x, y)∧∀z(F (x, z) ⊃ y ≤ z)). We show that G(x, y)
strongly defines f (x) in S, i.e. if f (a) = b then S ⊢ (G(a, b) ∧ ∀y(G(a, y) ⊃ y = b)).
We show that S proves both conjuncts, on the assumption that f (a) = b.
(i) To establish the first conjunct, i.e. S ⊢ G(a, b), i.e.
S ⊢ (F (a, b) ∧ ∀z(F (a, z) ⊃ b ≤ z)):
LECTURE 9 79

(1) Since f (x) in weakly definable in S by F (v1 , v2 ), S ⊢ F (a, b).


(2) To prove the second conjunct we establish (F (a, v1 ) ⊃ b ≤ v1 ) by ∨-
elimination from the instance of Ω5 for b, i.e. (v1 ≤ b ∨ b ≤ v1 ).
(3) We have S ⊢ (b ≤ v1 ⊃ (F (a, v1 ) ⊃ b ≤ v1 )), as an instance of L3 . So it
remains to show: S ⊢ (v1 ≤ b ⊃ (F (a, v1 ) ⊃ b ≤ v1 ))).
(4) Suppose k < b. Then k 6= b, so by clause (2) of weak definability, S ⊢∼
F (a, k). Then by propositional logic S ⊢ (F (a, k) ⊃ b ≤ k). Then by substitutivity
of identity, S ⊢ (v1 = k ⊃ (F (a, v1 ) ⊃ b ≤ v1 ))
(5) Suppose k = b. We know by Proposition 62 that Ω5 ⊢ b ≤ b. So by L3
and Modus ponens, S ⊢ (F (a, b) ⊃ b ≤ b). Then by substitutivity of identity,
S ⊢ (v1 = b ⊃ (F (a, v1 ) ⊃ b ≤ v1 ).
(6) By ∨-elimination from the cases v1 = 0, . . . , v1 = b established in (4) and (5),
S ⊢ ((v1 = 0 ∨ . . . ∨ v1 = b) ⊃ (F (a, v1 ) ⊃ b ≤ v1 )).
(7) From (6) by Ω4 and propositional logic, S ⊢ (v1 ≤ b ⊃ (F (a, v1 ) ⊃ b ≤ v1 )).
(8) By ∨-elimination from Ω5 with (3) and (7), S ⊢ (F (a, v1 ) ⊃ b ≤ v1 ).
(9) by ∀-Intro from (8), S ⊢ ∀y(F (a, y) ⊃ b ≤ y).
(ii) To establish the second conjunct, i.e. S ⊢ ∀y(G(a, y) ⊃ y = b):
(1) By ∧-elimination S ⊢ (G(a, y) ⊃ (∀z(F (a, z) ⊃ y ≤ z)).
(2) Since (∀z(F (a, z) ⊃ y ≤ z) ⊃ (F (a, b) ⊃ y ≤ b)) is logically valid, from (1) it
follows that S ⊢ (G(a, y) ⊃ (F (a, b) ⊃ y ≤ b)).
(3) By condition (1) of weak definability, S ⊢ F (a, b), so from (2), S ⊢ (G(a, y) ⊃
y ≤ b).
(4) We aim to show that S ⊢ (y ≤ b ⊃ (G(a, y) ⊃ y = b)), which with (3) implies
(ii).
(5) If k < b, then k 6= b, so f (a) 6= k, so by weak definability of f (x) by F (v1 , v2 ),
S ⊢∼ F (a, k), so S ⊢∼ G(a, k). So by propositional logic S ⊢ (G(a, k) ⊃ k = b).
(6) From (5) by substitutivity of identity, for k < b, S ⊢ (y = k ⊃ (G(a, y) ⊃
y = b)
(7) By L3 , S ⊢ (y = b ⊃ (G(a, y) ⊃ y = b)).
(8) By ∨-elimination from (6) and (7) and ⊃-introduction, S ⊢ ((y = 0∨. . .∨y =
b) ⊃ (G(a, y) ⊃ y = b)).
(9) From (8) by Ω4 and propositional logic, S ⊢ (y ≤ b ⊃ (G(a, y) ⊃ y = b)).
(10) From (3) and (9) by propositional logic, S ⊢ (G(a, y) ⊃ y = b).
(11) From (10) by ∀-introduction, S ⊢ ∀y(G(a, y) ⊃ y = b). N

9.4 Formal provability of the Diagonal Lemma


The Second Incompleteness Theorem is proved by proving within the system S the
first half of the First Incompleteness Theorem, “If S is consistent, then S 0 G”,
where G is the Gödel sentence for S.
LECTURE 9 80

We began our proof of the First Incompleteness Theorem by establishing the


Diagonal Lemma. In our original proof of the Diagonal Lemma we proved the
truth of the equivalence (C ≡ F (pCq)). In arithmetizing the First Incompleteness
Theorem for S in S we must show that S ⊢ (C ≡ F (pCq)).
We first show the following.

Lemma 98 For S any extensions of R, the diagonal function is strongly definable


in S.

Proof. (1) If the graph of a total function is Σ1 , the complement of its graph
is also Σ1 ) (Lemma 34 in Lecture 5). (2) The diagonal function is total and its
graph is expressed by a Σ1 -formula, ∃v3 A(v1 , v2 , v3 ), i.e. d(a) = b iff ∃v3 A(a, b, v3 ) is
true. (3) By (1) and (2), the complement of its graph is expressed by a Σ1 -formula
∃v3 B(v1 , v2 , v3 ), i.e. d(a) 6= b iff ∃v3 B(a, b, v3 ) is true. (4) Suppose d(a) = b; then
by (2) ∃v3 A(a, b, v3 ) is true, so there is a number m such that A(a, b, m) is true. (5)
By Σ0 -completeness of S, since S is an extension of R, S ⊢ A(a, b, m). (6) Suppose
d(a) 6= b; then by (2), ∃v3 A(a, b, v3 ) is false, so for every m, ∼ A(a, b, m) is true. (6)
By Σ0 -completeness of S, since S is an extension of R, for each m, S ⊢∼ A(a, b, m).
(7) Hence the d(v1 = v2 is enumerated in S by A(v1 , v2 , v3 ). (8) Similarly, d(v1 6= v2 is
enumerated in S by B(v1 , v2 , v3 ). (9) Hence by the Separation Lemma (Theorem 92,
the formula ∃x(A(v1 , v2 , x) ∧ (∀y ≤ x) ∼ B(v1 , v2 , y)) separates d(v1 ) = v2 in S, i.e.
if d(a) = b, then S ⊢ ∃x(A(a, b, x) ∧ (∀y ≤ x) ∼ B(a, b, y)) and if d(a) 6= b, then
S ⊢∼ ∃x(A(a, b, x)∧(∀y ≤ x) ∼ B(a, b, y)), which is to say that d(v1 ) = v2 is weakly
defined by ∃x(A(v1 , v2 , x)∧(∀y ≤ x) ∼ B(v1 , v2 , y)) in S. (10) Then by Theorem 97,
d(v1 ) = v2 is strongly definable in S. N

Theorem 99 (provable substitution) If a total function f (x) is strongly defin-


able in a system S, then for each formula G(v1 ), there is a formula H(v1 ) such that
for each number n, S ⊢ (H(n) ≡ G(f (n))).

Proof. Let F (v1 , v2 ) be the formula that strongly defines f (x) in S. For given
formula G(v1 ) let H(v1 ) =df ∃v2 (F (v1 , v2 )∧G(v2 )). Let n and m be such that f (n) =
m. We establish the provability of the two halves of the required biconditional as
follows:
(i) To show that S ⊢ (G(f (n)) ⊃ H(n)): Since F (v1 , v2 ) strongly defines f (x), by
clause (1) S ⊢ F (n, m). Hence by propositional logic in S, S ⊢ (G(m ⊃ (F (n, m) ∧
G(m)). Then by ∃-introduction in S (note that in inferring ∃v1 A(v1 ) from A(t), v1 is
substituted for some but not necessarily all occurrences of t in A(t)), S ⊢ (G(m) ⊃
∃v2 (F (n, v2 ) ∧ G(v2 ))).
(ii) To show that S ⊢ (H(n) ⊃ G(f (n))): By ∀-elimination from the second
conjunct of the condition for strong definability, S ⊢ (F (n, v2 ) ⊃ v2 = m). Therefore
LECTURE 9 81

by propositional logic in S,
S ⊢ ((F (n, v2 ) ∧ G(v2 )) ⊃ (v2 = m ∧ G(v2 ))). Since ((v2 = m ∧ G(v2 )) ⊃ G(m)) is
logically valid (by substitutivity of identity), S ⊢ ((v2 = m ∧ G(v2 )) ⊃ G(m)), so
by propositional logic in S, ⊢ (F (n, v2 ) ∧ G(v2 )) ⊃ G(m). Hence by ∀-introduction,
S ⊢ ∀v2 (F (n, v2 ) ∧ G(v2 ) ⊃ G(m)), and so by anti-prenexing,
S ⊢ (∃v2 (F (n, v2 ) ∧ G(v2 )) ⊃ G(m)) N

Theorem 100 For S any extension of R, for each formula F (v1 ) in the language of
S, with v1 its only free variable, there is a sentence C such that S ⊢ (C ≡ F (pCq)).

Proof. By Theorem 99, given F (v1 ), there is a formula H(v1 ) such that for
each n, S ⊢ (H(n) ≡ F (d(n))). In particular, for h = pH(v1 )q, S ⊢ (H(h) ≡
F (d(h))). By the construction of d(x), d(h) = pH[h]q. We have noted before that
for any formula F (v1 ), (F (n) ≡ F [n]) is logically valid, so S ⊢ (H(h) ≡ H[h]). So
S ⊢ (H[h] ≡ F (d(h))). Thus taking C =df H[h], we have S ⊢ (C ≡ F (pCq)), as
required. N
Lecture 10

Arithemization of consistency;
provability predicates; Gödel’s
Second Incompleteness Theorem;
Löb’s Theorem; analyzing and
strengthening the First
Incompleteness Theorem

(Wednesday, 10 November 2010)

10.1 Arithmetization of the statement that a sys-


tem S is consistent
Gödel’s Second Incompleteness Theorem for a system S satisfying certain conditions
is the inference that if S is consistent, S cannot prove the consistency of S. Clearly
the condition that S is consistent is necessary, since if S is inconsistent it proves ev-
erything, including any sentence in the language of S that expresses the consistency
of S.
The first condition for Gödel’s Second Incompleteness Theorem to hold for a
system is that the consistency of the system be expressible by a sentence in the
language of the system.

Definition 68 (definition of ConS ) For P r(v1 ) a formula that expresses {n : S ⊢

82
LECTURE 10 83

En } and X any sentence in the language of S such that S 0 X, we let ConS , the
formal expression in the language of S of the consistency of S be ∼ P r(pXq).

The notation ConS does not notate relativity to the sentence X, nor relativity
to the chosen Gödel numbering and arithmetization of syntax. I will say something
later about provable invariance of the Gödel sentence and of ConS with respect to
Gödel numbering and arithmetization of syntax. As to relativity of ConS to the
unprovable sentence X, we have the following

Lemma 101 (justifying the definition of ConS ) A system S is consistent if and


only if ∼ P r(pXq) is true, for P r(v1 ) any formula in the language of S that ex-
presses {n : S ⊢ En }, and X any sentence in the langauge of S such that S 0 X.

Proof. Immediate from Proposition 72 (in Lecture 8). N

Remark. The relativity to X is sometimes dealt with by fixing on a particular X


such that S 0 X if and only if S is consistent. This could, with a lot of work—
which we have now done—be the Gödel sentence itself. But with no work, except
as required to show S ⊢∼ X (which can be very little, as in ∼ 0 = 0′ , immediate
from axiom N2 ), we can take ConS to be ∼ P r(pXq) for some particular such X,
and it is nearly standard to take ConS as ∼ P r(p0 = 0′ ).

The Second Incompleteness Theorem is established by formalizing in S the proof


of the first half of the First Incompleteness Theorem, i.e. if S is consistent, then
S 0 G. Hence if S proves ConS , then S proves that S 0 G. But since S ⊢ (G ≡∼
P r(pGq), in that case S ⊢ G, which, if S is consistent, it doesn’t.
The process of formalization in S is intricate. By work of Paul Bernays and
Martin Löb, the requirements for formalization are reduced to three conditions on
the proof predicate for S. We shall establish the Second Incompleteness Theorem
from the assumption of these three conditions. The first of these three conditions
we have already established (for PA), the second is a problem on Problem sheet 5,
and the third we will establish in Lecture 11.

10.2 Provability predicates.


Definition 69 (provability predicate) A formula P (v1 ) is called a provability
predicate for a system S if for all sentences X and Y in L(S) the following three
conditions hold:
P1 : If S ⊢ X, then S ⊢ P (pXq).
P2 : S ⊢ (P (p(X ⊃ Y )q) ⊃ (P (pXq) ⊃ P (pY q)))
P3 : S ⊢ (P (pXq) ⊃ P (pP (pXq)q))
LECTURE 10 84

Note that P1 is a one-way implication and not a biconditional. The converse


implication is an instance of Σ1 -soundness. The effect of this is that these conditions
on being a provability predicate do not require that a provability predicate expresses
{n : S ⊢ En }. Rather, it can express a superset of that set, in particular, the formula
x = x is a provability predicate.
Being a provability predicate is not extensional, i.e. there are pairs of formulas,
even in the same class of the arithmetical hierarchy, that have the same extension
and one of them is a provability predicate and the other is not.
We established property P1 (and its converse) for the arithmetized proof predi-
cate for PA by Theorem 28.
For Property P2 we have

Theorem 102 For P r(v1 ) the arithmetical proof predicate for PA constructed for
Theorem 28, and X and Y any sentences in the language of PA,
PA⊢ (P r(pX ⊃ Y q) ⊃ (P r(pXq) ⊃ P r(pY q))).

Proof. Exercise (Problem sheet 5) N

Property P3 is the arithmetization of P1 , and follows from provable Σ1 -completeness


of PA, which I shall prove in Lecture 11.

The following three properties of a provability predicate are immediately provable


from the three that define what it is to be a provability predicate.

Lemma 103 For P (v1 ) a provability predicate for a system S,


P4 If S ⊢ (X ⊃ Y ), then S ⊢ (P (pXq) ⊃ P (pY q)).
P5 If S ⊢ (X ⊃ (Y ⊃ Z)), then S ⊢ (P (pXq) ⊃ (P (pY q) ⊃ P (pZq))).
P6 If S ⊢ (X ⊃ (P (pXq) ⊃ Y )), then S ⊢ (P (pXq) ⊃ P (pY q)).

Proof.
P4 : If S ⊢ (X ⊃ Y ), then by P1 , S ⊢ P (p(X ⊃ Y )). Then by P2 and Modus
ponens, S ⊢ (P (pXq) ⊃ P (pY q)).
P5 : By P4 , S ⊢ (P (pXq) ⊃ P (p(Y ⊃ Z)q)). By P2 , S ⊢ (P (p(Y ⊃ Z)q) ⊃
(P (pY q) ⊃ P (pZq))). Then by propositional logic, S ⊢ (P (pXq) ⊃ (P (pY q) ⊃
P (pZq)).
[Notice that there is no corresponding result for S ⊢ ((X ⊃ Y ) ⊃ Z).]
P6 : If S ⊢ (X ⊃ (P (pXq) ⊃ Y )), then by P5
S ⊢ (P (pXq) ⊃ (P (pP (pXq)q) ⊃ P (pY q))). Then by L2 and Modus ponens,
S ⊢ ((P (pXq) ⊃ P (pP (pXq)q)) ⊃ (P (pXq) ⊃ P (pY q))) Then by P3 and Modus
ponens, S ⊢ (P (pXq) ⊃ P (pY q)). N
LECTURE 10 85

10.3 Gödel’s Second Incompleteness Theorem


Theorem 104 (Gödel’s Second Incompleteness Theorem) Let P (v1 ) be a prov-
ability predicate for a system S, and let G be a sentence in the language of S such
that S ⊢ (G ≡∼ P (pGq)). For X any sentence in the language of S, if S is consis-
tent, S 0∼ P (pXq).

Proof.

(1) S ⊢ (G ≡∼ P (pGq)) Provable diagonal equivalence


(2) S ⊢ (G ⊃∼ P (pGq)) (1) ∧-Elim
(3) S ⊢ (G ⊃ (P (pGq) ⊃ X)) (2) Prop Logic
(4) S ⊢ (P (pGq) ⊃ P (pXq)) (3) P6
(5) S ⊢ (∼ P (pXq) ⊃∼ P (pGq)) (4) contraposition
(6) S ⊢ (∼ P (pGq) ⊃ G) (1) ∧-Elim
(7) S ⊢ (∼ P (pXq) ⊃ G) (5) (6) Prop Logic
(8) S ⊢∼ P (pXq) Assumption
(9) S ⊢ G (7) (8) ⊃-Elim
(10) S ⊢ P (pGq) (9) P1
(11) S ⊢∼ G (10)(2) Prop Logic
(12) S is inconsistent (9)(11)
(13) S 0∼ P (pXq) (8)(12) Consistency of S RAA N

Remark 1: Line (5) of this proof of the Second Incompleteness Theorem gives
a formal proof in S of the first half of the First Incompleteness Theorem for S, i.e.
if S is consistent, S 0 G.

Remark 2: Gödel’s Second Incompleteness Theorem is a generalization of the


first half of the First Incompleteness Theorem. The first half of the First Incom-
pleteness Theorem establishes, for G the Gödel sentence S, that if S is consistent,
S 0 G. By the diagonal equivalence, this is tantamount to S 0∼ P (pGq). The Sec-
ond Incompleteness Theorem establishes that for every sentence X in the language
of S, if S is consistent, S 0∼ P (pXq)). The heart of the matter, however, is that
the Gödel sentence for S is provably in S equivalent to the consistency of S. One di-
rection of this equivalence is established at line (7) of the above proof of the Second
Incompleteness Theorem: If S is consistent, i.e. some sentence is unprovable, then
G holds. The converse of (7) for arbitrary X cannot be proved because it doesn’t
hold, if S is 1-consistent: If S ⊢ (G ⊃∼ P (pXq)) for X such that S ⊢ X, then since
by P1 , S ⊢ P (pXq), S ⊢∼ G. However, if S ⊢∼ X, then indeed the implication
holds, i.e.
LECTURE 10 86

Proposition 105 For X any sentence such that S ⊢∼ X, S ⊢ (G ⊃∼ P (pXq)).

Proof. If S ⊢∼ X, by propositional logic S ⊢ (X ⊃ G). Then by P4 , S ⊢


(P (pXq) ⊃ P (pGq)). By contraposition, S ⊢ (∼ P (pGq) ⊃∼ P (pXq)). Then by
the diagonal equivalence for G and transitivity of implication,
S ⊢ (G ⊃∼ P (pXq)). N

Corollary 106 For X any sentence such that S ⊢∼ X, S ⊢ (G ≡∼ P (pXq)).

Proof. By Proposition 105 and line (7) of the proof of Theorem 104. N

10.4 Löb’s Theorem


Löb’s Theorem is a deep result which characterizes the abstract properties of prov-
ability, i.e. it can be used as the fundamental axiom for a theory of provability,
as we shall see in Lectures 13 and 14. It is also a generalization of the Second
Incompleteness Theorem, though this was not immediately realized,. It arose in
response to an almost jokey question in the 1950s by Leon Henkin: Is the sen-
tence that asserts its own provability (there is such a sentence, by diagonalization,
i.e. S ⊢ (P r(pHq) ≡ H)) provable (in which case it is true ) or unprovable (in
which case it is false)? What Martin Löb showed was that from just half of that
diagonal equivalence, i.e. S ⊢ (P (pHq) ⊃ H), it follows that S ⊢ H (so H is prov-
able and true), because for any sentence X and any provability predicate P (v1 ), if
S ⊢ (P (pXq) ⊃ X), then S ⊢ X. (The converse holds by propositional logic, from
L1 and Modus ponens.)

Theorem 107 (Löb’s theorem) Let S be a system in which the Diagonal Lemma
is provable, and let P (v1 ) be a provability predicate for S. For X any formula in the
language of S, if S ⊢ (P (pXq) ⊃ X), then S ⊢ X.

Proof. (1) Assume that S ⊢ (P (pXq) ⊃ X).


(2) Let L be a provably diagonal sentence for the predicate (P (v1 ) ⊃ X), i.e.
S ⊢ (L ≡ (P (pLq) ⊃ X).
(3) S ⊢ (L ⊃ ((P (pLq) ⊃ X), by ∧-Elimination from the diagonal equivalence
in (2).
(4) By P6 from (3), S ⊢ (P (pLq) ⊃ P (pXq)).
(5) From (1) and (4) we have S ⊢ (P (pLq) ⊃ X)
(6) S ⊢ ((P (pLq) ⊃ X) ⊃ L), by ∧-Elimination from the diagonal equivalence
in (2).
(7) By Modus ponens from (5) and (6) S ⊢ L.
(8) Then from (7) by P1 , S ⊢ P (pLq).
LECTURE 10 87

(9) S ⊢ X by (8) and (5). N

Remark. A sentence of the form (P (pXq) ⊃ X) expresses the soundness with


respect to provability of X of the system for which P (v1 ) expresses provability, i.e.
it says that if X is provable, then X, i.e. X is true. Löb’s Theorem says that the
only such statements that can be proved in a system are the ones that hold trivially
by propositional logic, i.e. for which S ⊢ X.

Theorem 108 Löb’s Theorem is a generalization of Gödel’s Second Incompleteness


Theorem.

Proof. The Second Incompleteness Theorem proves Löb’s Theorem in those


cases where S ⊢∼ X, as follows: If S ⊢ ((P (pXq) ⊃ X)) and S ⊢∼ X, then
by propositional logic, S ⊢∼ P (pXq). Then by the contrapositive of the Second
Incompleteness Theorem, S is inconsistent, which is to say that S proves everything,
so in particular, S ⊢ X.
Löb’s Theorem goes beyond what is established by the Second Incomplete-
ness Theorem and propositional logic by establishing the implication from S ⊢
((P (pXq) ⊃ X)) to S ⊢ X for the case of X such that S 0 X and S 0∼ X. N

Despite Löb’s Theorem being a generalization of the Second Incompleteness The-


orem, Löb’s Theorem can be proved from the Second Incompleteness Theorem,
though not uniformly but on a sentence by sentence basis. The situation is the
following:

Theorem 109 The Second Incompleteness Theorem for S ∪ {∼ X} implies that if


S ⊢ (P (pXq) ⊃ X), S ⊢ X.

Proof. Problem 4 on Problem sheet 5. N

10.5 Analyzing and strengthening the First In-


completeness Theorem
The Second Incompleteness Theorem greatly increases our understanding of the
First Incompleteness Theorem. In discussing the truth value of the Gödel sentence
for a system S, we noted that the truth of the Gödel sentence for S is implied by and
implies the consistency of S. The Second Incompleteness Theorem establishes that
the equivalence of G and Con S is provable in S, so that with respect to S itself, we
see that the meaning of G is “S is consistent”. The Second Incompleteness Theorem
enables us to show that the second half of the First Incompleteness Theorem, that
S 0∼ G, cannot be established just from the hypothesis that S is consistent, by
LECTURE 10 88

allowing us to establish the existence of a consistent theory S such that S ⊢∼


GS . It enables us also to find a non-trivial formulation of the minimum sufficient
condition for unprovability of the negation of the Gödel sentence for system S,
namely consistency of S ∪{ConS}, and to show that this condition is strictly weaker
than the 1-consistency of S.

10.5.1 S 0∼ GS cannot be proved from the consistency of S


This result can be established by showing that any of the most usual examples
of consistent, 1-inconsistent theories, S ∪ {∼ GS } for consistent Σ0 -complete Σ1 -
axiomatizable theory S, proves the negation of its Gödel sentence (not the same as
the Gödel sentence for S)—but also (as we shall see below) not every 1-inconsistent
theory proves the negation of its Gödel sentence.

Theorem 110 For P (v1 ) a provability predicate for a consistent theory S, let G be
a sentence in the language of S such that S ⊢ (G ≡∼ P (pGq)), let S ∗ be the system
S ∪{∼ G}, which is consistent, given the consistency of S, let P ∗ (v1 ) be a provability
predicate for S ∗ . Let G∗ be a sentence such that S ∗ ⊢ (G∗ ≡∼ P ∗ (pG∗ q)). Then
S ∗ ⊢∼ G∗ .

Proof. (1) Let X be such that S ⊢∼ X. (2) Then also (by thinning), S ∗ ⊢∼ X.
(3) S ⊢ (X ⊃ (∼ G ⊃ X)), by propositional logic, for any sentence X. Then (4)
S ⊢ (P (pXq) ⊃ P (p(∼ G ⊃ X)q, by property of a provability predicate, and so
by thinning, (5) S ∗ ⊢ (P (pXq) ⊃ P (p(∼ G ⊃ X)q. Then (6) S ∗ ⊢ (P (pXq) ⊃
P ∗ (pXq)) since for all Y , S ∗ ⊢ (P (p(∼ G ⊃ Y )q) ≡ P ∗ (pY q)). By the proof of the
Second Incompleteness Theorem, G is provably equivalent in S to ConS, and so also
by thinning in S ∗ , i.e. (7) S ∗ ⊢ (G ≡∼ P (pXq)), and similarly (8) S ∗ ⊢ (G∗ ≡∼
P ∗ (pXq)). Therefore (9) S ∗ ⊢ (∼ G ⊃∼ G∗ ). Hence (10) S ∗ ⊢∼ G∗ . N
Remark. We know that P A 0∼ GP A . Theorem 110 shows is that this fact can-
not be proved without appealing to a property of PA stronger than it’s consistency.

10.5.2 Strengthened second half of the First Incompleteness


Theorem
Theorem 111 (strengthened First Incompleteness Theorem) Let P (v1 ) be
a provability predicate for a system S, and let G be a sentence in the language
of S such that S ⊢ (G ≡∼ P (pGq)). Let X be any sentence such that S ⊢∼ X. Let
ConS stand for ∼ P (pXq). Then if S ∪ {ConS} is consistent, S 0∼ G.

Proof. Suppose S ⊢∼ G. Then by Lemma ??, S ⊢∼ ConS. This contradicts


the hypothesized consistency of S ∪ {ConS}. Therefore S 0∼ G. N
LECTURE 10 89

10.5.3 Consistency of S ∪ {ConS} is strictly weaker than 1-


consistency of S
We establish this result by constructing a theory S such that if PA∪{ConP A} is
consistent, S is 1-inconsistent and S ∪ {ConS} is consistent.

Theorem 112 Let P A+ =df P A ∪ {ConP A} and S =df P A ∪ {∼ ConP A+ }. If


P A+ is consistent, then S is 1-inconsistent and S ∪ {ConS} is consistent.

Proof. (i) If P A+ is consistent, then ConP A+ is a true Π1 -sentence, so


∼ ConP A+ is a false Σ1 -sentence. Then by Theorem 78, S is 1-inconsistent.
(ii) (1) Suppose S ∪ {ConS} is inconsistent.
(2) Then by propositional logic, S ⊢∼ ConS.
(3) Similarly, S is inconsistent if and only if P A ⊢ ConP A+ , so
(4) S proves the inconsistency of S iff S ⊢ P rP A (pConP A+ q), i.e. P A ∪ {∼
ConP A+ } ⊢ P rP A (pConP A+ q).
(5) Then by the Deduction Theorem, P A ⊢ (∼ ConP A+ ⊃ P rP A (pConP A+ q)),
and so by the contrapositive of (5),
(6) P A ⊢ (∼ P rP A (pConP A+ q) ⊃ ConP A+ ).
(7) By arithmetized Second Incompleteness Theorem, P A ⊢ (ConP A ⊃∼ P rP A (pConP Aq)).
(8) By arithmetization of the logical fact that if PA (or any other theory) does not
prove the consistency of a theory then a fortiori it does not prove the consistency of
an extension of that theory, P A ⊢ (∼ P rP A (pConP Aq)) ⊃∼ P rP A (pCon(P A ∪ {ConP A})q)).
(9) By (7) and (8), P A ⊢ (ConP A ⊃∼ P rP A (pCon(P A ∪ {ConP A})q)).
(10) By (6) and (9), P A ⊢ (ConP A ⊃ ConP A+ ).
(11) Then by Modus Ponens, P A ∪ {ConP A} ⊢ ConP A+ , i.e. P A+ ⊢ ConP A+ .
(12) From the assumption that P A ∪ {ConP A} =df P A+ is consistent, P A+ 0
ConP A+ , i.e. Second Incompleteness Theorem for P A+ .
(13) Then by (1), (11), (12) and RAA, S ∪ {ConS} is consistent. N

Proposition 113 Theorem 111 is stronger than Theorem 83, and is best possible.

Proof. By Theorem 112, the hypothesis of Theorem 111 is strictly weaker


than the hypothesis of the second half of Theorem 83. It is best possible since if
S ∪ {ConS} is inconsistent, S ⊢∼ ConS. By Theorem ??, S ⊢ (∼ ConS ⊃∼ G),
so then S ⊢∼ G, i.e. if the hypothesis doesn’t hold, the result to be proved doesn’t
hold. N
Remark. The hypothesis S ∪ {ConS } for the second half of the First Incom-
pleteness Theorem not only makes the theorem best possible but is the natural
hypothesis for this result. The two parts of the First Incompleteness Theorem for a
Σ0 -complete theory S that is Σ1 -axiomatizable are of different character. To show
LECTURE 10 90

that S 0 G requires the purely formal hypothesis that S is consistent. The consis-
tency of S implies the truth of G, as noted in Theorem 85. This result holds for
every system for which a Gödel sentence can be constructed. Consequently, whether
or not S 0∼ G is equivalent to whether or not S is sound to the extent that it does
not prove this false sentence. We have seen that the first half of the First Incom-
pleteness Theorem establishes the consistency and unsoundness of S ∪ {∼ G}, if S
is consistent, and we have also seen that S ∪ {∼ GS } ⊢∼ GS∪{∼GS } . Given that S
is consistent, ConS is true. Exactly the amount of soundness of S required for the
second half of the First Incompleteness Theorem is that S does not prove the false
Σ1 -sentence ConS, i.e. S ∪ {ConS } is consistent. We have seen that a theory can
have this much soundness and still not be Σ1 -sound (Theorem 112).
Lecture 11

Provable Σ1-completeness

(Tuesday, 16 November 2010)

Proposition 114 For P r(v1 ) a formula in the language of PA that expresses {n :


P A ⊢ En } and X any Σ1 -sentence in the language of PA, the sentence (X ⊃
P r(pXq)) is true.

Proof. (i) If X is true, then by Σ1 -completeness of PA, PA ⊢ X, and since


P r(v1 ) expresses {n : PA ⊢ En }, P r(pXq) is true. So (X ⊃ P r(pXq)) is true.
(ii) If X is false, then (X ⊃ P r(pXq)) is true. N
In this lecture we will establish that all these true sentences are provable in PA
(provable Σ1 -completeness). Since for Y any sentence in the language of PA, the
sentence P r(pY q) is Σ1 , these sentences include all sentences of the form (P r(pY q) ⊃
P r(pP r(pY q)q)), i.e. P3 , the third condition on a provability predicate.
Note that for X a Σ1 -sentence, (X ⊃ P r(pXq)) is ∆2 . As we shall see, PA is not
∆2 -complete. So the provability of these sentences in PA is specific to provability
properties of Σ1 -sentences and of arithmetization of provability in PA.
While provable Σ1 -completeness is a deep theorem whose proof is complicated,
provable completeness for Σ0 -sentences is very easy to show.

Proposition 115 For P r(v1 ) a formula in the language of PA that expresses {n :


P A ⊢ En } and X any Σ0 -sentence in the language of PA, P A ⊢ (X ⊃ P r(pXq)).

Proof. Argument (1): (i) If X is true, then by Σ0 -completeness of PA, PA ⊢ X,


and since P r(v1 ) is Σ1 , and expresses {n : PA ⊢ En }, P r(pXq) is a true Σ1 -sentence.
Then by Σ1 -completeness of PA, P A ⊢ P r(pXq), so by propositional logic in PA,
P A ⊢ (X ⊃ P r(pXq)).

91
LECTURE 11 92

(ii) If X is false, then ∼ X is a true Σ0 -sentence, so by Σ0 -completeness, P A ⊢∼


X, so by propositional logic in PA, P A ⊢ (X ⊃ P r(pXq)).
Argument (2): If X is Σ0 , then (X ⊃ P r(pXq)) is Σ1 , and hence provable in PA
by Σ1 -completeness. N
Remark. Neither of the two arguments for Proposition 115 can be extended to
the case of X a Σ1 -sentence. For Argument (1), (i) holds for X a Σ1 -sentence, but
(ii) fails since the negation of a Σ1 -sentence is not, in general, a Σ1 -sentence. For
Argument (2), we noted above that for X a Σ1 -sentence, (X ⊃ P r(pXq)) is ∆2 ,
and as we shall see later, PA is not ∆2 -complete.

Remark about the strategy for the proof. The definition of what it is to be a
Σ1 -formula, Definition 41, is explicit (rather than recursive), i.e. a Σ1 formula is
any formula of the form ∃vi F where F is a Σ0 -formula. —— So as in our proof
that R and thereby Q and PA are Σ1 -complete (Propositions 58, 67, and 69, and
Theorem 70), the proof of provable Σ1 -completeness has to go via a proof of provable
Σ0 -completeness.
The definition of Σ0 -formula, Definition 40, is recursive, so the proof of prov-
able Σ0 -completeness must proceed by induction over the recursive definition of
Σ0 -formulas. The sequence of formulas by which a Σ0 -sentence is generated by this
recursion will in general contain free variables. Thus to have a strong enough induc-
tion hypothesis for the inductive argument, we must prove provable Σ0 -completeness
for formulas that may contain free variables.
How to formulate this result for Σ0 -formulas with free variables requires some
thought. It is not that P A ⊢ (X ⊃ P r(pXq)) for all Σ1 -formulas, including ones
with free variables. For example,

Proposition 116 If PA is 1-consistent, PA 0 (v1 + v2 = v3 ⊃ P r(pv1 + v2 = v3 q)).

Proof. IF PA ⊢ (v1 + v2 = v3 ⊃ P r(pv1 + v2 = v3 q)), then by instantiation,


PA ⊢ (1 + 2 = 3 ⊃ P r(pv1 + v2 = v3 q)). PA ⊢ 1 + 2 = 3, so by Modus ponens,
PA ⊢ P r(pv1 + v2 = v3 q). By arithmetization of the rule R1 , i.e. Generalization,
PA ⊢ (P r(pXq) ⊃ P r(p∀vi Xq)). Hence PA ⊢ P r(p∀v1 ∀v2 ∀v3 v1 + v2 = v3 q). By
logic of identity in PA, PA ⊢ (∀v1 ∀v2 ∀v3 v1 + v2 = v3 ⊃ 1 + 0 = 0). Hence PA
⊢ (P r(p∀v1 ∀v2 ∀v3 v1 + v2 = v3 q) ⊃ P r(p1 + 0 = 0q)). Then PA ⊢ P r(p1 + 0 = 0q),
which contradicts the 1-consistency of PA. N
The result we need to establish is that PA proves that each true substitution
instance of a Σ0 -formula is provable in PA. So we need to find a way to express the
condition that a substitution instance of a formula is provable.
To do this we need first to modify the definition we gave in Lecture 2 of quasi-
substitution (Definition 27), which was defined just for substitution on the free
LECTURE 11 93

variable v1 , i.e. s(x, y) = p∀v1 (v1 = y ⊃ Ex )q, to allow substitution on any specified
variable, i.e. s(x, y, z) = p∀vz (vz = y ⊃ Ex )q
Proposition 117 For the function s(x, y, z) = p∀vz (vz = y ⊃ Ex )q there is a
Σ1 -formula S(x, y, z, w) in the language of PA such that for all natural numbers
n1 , n2 , n3 , n4 , S(n1 , n2 , n3 , n4 ) is true if and only if s(n1 , n2 , n3 ) = n4 .
Proof.
p∀vz (vz = y ⊃ Ex )q =
. . 5} η ∗ 13y ∗ 8x3.
. . 5} 26 |5 .{z
p∀v′ . . .′ (v′ . . .′ = y ⊃ Ex )q = 96 5| .{z
|{z} |{z} z z
z z
The function f (z) = 5| .{z
. . 5} is generated by the following primitive recursion:
z
f (0) = 5
f (n + 1) = 5 · 13n+1 + f (n), so by generalization of Theorem 39, f (x) = y is Σ1 ,
so the relation s(x, y, z) = z is Σ1 . N
Definition 70 (arithmetized proof predicate with free variables) For P r(v1 )
a formula in the language of PA that expresses {n : P A ⊢ En } and F (vk1 , . . . , vkm )
any formula in the language of PA with exactly the free variables shown, and k =
max{k1 , . . . , km }, P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ) =df
∃vk+1 . . . ∃vk+m ((S(pF (vk1 , . . . , vkm )q, vk1 , k1 , vk+1 ) ∧ S(vk+1 , vk2 , k2 , vk+2 ) ∧ . . .
∧ S(vk+m−1 , vkm , km , vk+m )) ∧ P r(vk+m )).
Note that F (vk1 , . . . , vkm ) and P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ) have the same
free variables.

Remark: Where a formula S(v1 , v2 ) represents a total function s(v1 ) = v2 in


a theory T , we can express the substitution F (s(v1 )) either by ∀v2 (S(v1 , v2 ) ⊃ F (v2 ))
or by ∃v2 (S(v1 , v2 )∧F (v2 )), since (∀v1 ∃v2 ∀v3 (S(v1 , v3 ) ⊃ v3 = v2 ) ⊃ (∀v2 (S(v1 , v2 ) ⊃
F (v2 )) ≡ ∃v2 (S(v1 , v2 ) ∧ F (v2 )))) is logically valid. So strictly we could have defined
P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ) as
∀vk+1 . . . ∀vk+m ((S(pF (vk1 , . . . , vkm )q, vk1 , k1 , vk+1 ) ∧ S(vk+1 , vk2 , k2 , vk+2 ) ∧ . . .
∧ S(vk+m−1 , vkm , km , vk+m )) ⊃ P r(vk+m )). We do not use this latter formula as the
definition since, given that P r(v1 ) is Σ1 , this formula is Π2 , whereas on the given
definition, P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ) is, like P r(v1 ), Σ1 .
Further remark: The notation P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ) stresses the
important point that it is the Gödel number of the formula F (vk1 , . . . , vkm ), and
not the formula itself, that occurs in P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ). Occa-
sionally, to avoid clutter, we will abbreviate this formula as P r[F (vk1 , . . . , vkm )],
which must be read bearing in mind that F (vk1 , . . . , vkm ) is not a sub-formula of
P r[F (vk1 , . . . , vkm )]. Note that this abbreviation cannot be used if we need to show
the result of making a substitution for a free variable of P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ).
LECTURE 11 94

Proposition 118 For a formula F with no free variables,


PA ⊢ (P r[pF q] ≡ P r(pF q)).

Proof. For F with no free variables, consider P r[pF q] with one vacuous quan-
tifier, e.g. ∀v1 (S(pF q, v1 , 0′ , v2 ) ⊃ P r(v2 )). etc. N

Proposition 119 For any natural numbers a1 , . . . am , P r[pF (vk1 , . . . , vkm )q](a1 , . . . , am )
is true if and only if P r(pF (a1 , . . . , am )q) is true.

Proof. By the definition of P r[pF q], and provable equivalence of F (a) and F [a].
N

We need to generalize P1 and P2 to allow for occurrence of free variables. The


generalization of P1 expresses that if a formula with free variables is provable in PA,
then for each sentence that results from substituting numerals for the free variables
of that formula, PA proves the proof predicate for PA applied to the Gödel number
of that sentence.

Theorem 120 (P1∗ = P1 generalized to allow free variables) For any formula
F (vk1 , . . . , vkm ) in the language of PA, if PA ⊢ F (vk1 , . . . , vkm ), then
PA ⊢ P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ).

Proof. The proof is by induction on the number of free variables in F (vk1 , . . . , vkm ).
It consists of arithmetizing the following argument, which for simplicity I formu-
late for the case in which F has just one free variable.
If PA⊢ F (vk ), then by logic in PA, PA⊢ ∀vk F (vk ). Then by P1 , PA ⊢ P r(p∀vk F (vk )q).
For each natural number n, (∀vk F (vk ) ⊃ F [n]) is logically valid, so PA⊢ (∀vk F (vk ) ⊃
F [n]). Then by P1 , PA ⊢ P r(p(∀vk F (vk ) ⊃ F [n])q). Then by P2 , PA ⊢ (P r(p∀vk F (vk )q) ⊃
P r(pF [n]q)). Then by Modus pones, PA ⊢ P r(pF [n]q). We need to formalize this
argument in PA. N

Theorem 121 (P2∗ = P2 generalized to allow free variables) For any formu-
las F (vk1 , . . . , vkm ) and G(vr1 , . . . , vrs ) in the language of PA,
PA⊢ (P [(F ⊃ G)](vk1 , . . . , vkm , vr1 , . . . vrs ) ⊃ (P [F ](vk1 , . . . , vkm ) ⊃ P [G](vr1 , . . . vrs ))).

Proof. Exercise. N

Lemma 122 For all formulas F (vj1 , . . . , vjm ) and G(vk1 , . . . , vkn ), P A ⊢ (P [pF q](vj1 , . . . , vjm ) ⊃
(P [pGq](vk1 , . . . , vkn ) ⊃ P [p(F ∧ G)q])(vj1 , . . . , vjm , vk1 , . . . , vkn ))

Proof Exercise. N
LECTURE 11 95

Definition 71 A term t is free for variable vi in formula F (vi ) if vi in F (vi ) does


not occur within the scope of a quantifier whose variable of quantification is a free
variable in t.

Definition 72 For F (vi ) a formula with free variable vi and t any term, F (vi /t)
is the result of substituting the term t for all occurrences of the variable vi in the
formula F (vi ).

Lemma 123 Let F (vi ) be a formula with free-variable vi . Let vj be a variable and
t a term both free for vi in F (vi ). Then the following equivalence is logically valid:

(P r[pF (vi )](t) ≡ P r[pF (vj )q](t)),


and correspondingly for F with any number of variables.

Proof. From the definition of P r[pF (vi )](vi ), substitutivity of =, and logical
equivalence of F (vi ) and F [vi ]. N

Lemma 124 Let F (vi ) be a formula with free-variable vi and let t(vj ) be a term
free for vi in F (vi with variable vj distinct from vi . Then the following equivalence
is logically valid:

(P r[pF (vi /t(vj ))q](vj ) ≡ P r[pF (vi )q](t(vj ))),

and correspondingly for F with any number of variables.

Proof. From the definition of P r[pF (vi )](vi ), substitutivity of =, and logical
equivalence of F (vi ) and F [vi ]. N

Our proof in Lecture 7 that PA is Σ0 -complete went by way of proving the very
strong result that the extremely weak system R is Σo -complete and then showing
that R is a subsystem of PA. Proving Σ0 -completeness requires heavy use of math-
ematical induction, so that proof cannot be formalized in R or Q. It could be
formalized in PA, but that would be a very roundabout way of proving the provable
Σ0 -completeness of PA in PA. Rather what we want to do is formalize in PA a di-
rect proof of the Σ0 -completeness of PA. I will illustrate an informal direct proof of
the Σ0 -completeness of PA that we will be formalizing in PA, by going through an
informal proof for the case of an atomic Σ0 -formula of the form v1 + v2 = v3 .

Lemma 125 If a + b = c, then PA ⊢ a + b = c.


LECTURE 11 96

Proof. We argue by (informal) induction on the free variable b in the statement,


for all c, if a + b = c, then PA ⊢ a + b = c. (The universal quantifier on the variable c
is to strengthen the induction hypothesis.) The result then follows by ∀-Elimination
on the quantifier ‘for all c’.
b = 0: If a + 0 = c, then a = c. From Axiom N3 by ∀-Intro and ∀-Elim, PA
⊢ a + 0 = a. If a = c, then PA ⊢ a = c. By the logic of identity in PA, PA
⊢ ((a = c ∧ a + 0 = a) ⊃ a + b = c). Hence if a + 0 = c, PA ⊢ a + b = c. This is
proved outright, in particular with no assumption in which the variable c is free, so
for all c, if a + 0 = c, PA ⊢ a + b = c.
Induction step. Suppose that for all c, if a + b = c, then PA ⊢ a + b = c, and
suppose a + b′ = c. Then by the informal version of N4 , (a + b)′ = c, so there is a
number d such that d′ = c. By logic of identity, (a + b)′ = d′ . Then by the informal
version of N1 , a + b = d. Instantiating the quantifier ‘forall c’ in the induction
hypothesis with d, we have that PA ⊢ a + b = d. Then by logic of identity in PA,
′ ′
PA ⊢ (a + b)′ = d . From the fact that d′ = c, PA ⊢ d′ = c. We know that d and
d′ are the same expression, so by logic of identity in PA, PA ⊢ (a + b)′ = c. Then

by N4 and logic of identity, PA ⊢ a + b = c. So from the Induction Hypothesis we

have proved that if a + b′ = c, PA ⊢ a + b = c. Since c is not free in the Induction

Hypothesis, it follows that for all c, if a + b′ = c, PA ⊢ a + b = c. N

We now turn to proof of the main theorem.

Theorem 126 (provable Σ0 -completeness with free variables) For each Σ0 -


formula F (vk1 , . . . , vkm ), PA ⊢ (F (vk1 , . . . , vkm ) ⊃ P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm )).

Proof. By induction over the inductive definition of Σ0 -formulas.


Base case:
F is an atomic formula, i.e. a formula of the form t1 = t2 or t1 ≤ t2 for t1 , t2
terms. The proof of this case is by a double induction over the recursive definition
of terms. We will prove one of these cases. The others are similar.
PA ⊢ (v1 + v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2 , v3 )), i.e. PA ⊢ (v1 + v2 = v3 ⊃
∃v4 ∃v5 ∃v6 (S(pv1 + v2 = v3 q, v1 , 0′ , v4 ) ∧ S(v4 , v2 , 0′′ , v5 ) ∧ S(v5 , v3 , 0′′′ , v6 ) ∧ P r(v6 ))
The following is an informal description of a formal proof within PA. The proof
is by induction on the variable v2 , but rather than argue by induction on v2 in the
formula (v1 + v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2 , v3 )), we argue by induction on v2
in the formula ∀v3 (v1 + v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v3 , v2 )), in order to have a
stronger Induction Hypothesis.
Base case: v2 = 0. We need to show that PA ⊢ ∀v3 (v1 +0 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , 0, v3 )).

(1) (1) v1 + 0 = v3 Assumption


(2) v1 + 0 = v1 N3
LECTURE 11 97

(1) (3) v1 = v3 (1) (2) substitutivity of =


(4) P r[pv1 + 0 = v1 q](v1 ) (2) P1∗
(5) (P r[pv1 + 0 = v1 q](v1 ) ⊃ P r[pv1 + v2 = v3 q](v1 , 0, v1 )) Lemma 124
(6) P r[pv1 + v2 = v3 q](v1 , 0, v1 ) (4) (5) ⊃-elimination
(1) (7) P r[pv1 + v2 = v3 q](v1 , 0, v3 ) (6) (3) substitutivity of =
(8) (v1 + 0 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , 0, v3 )) (1)(7) ⊃-intro
(9) ∀v3 (v1 + 0 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , 0, v3 )) (8) ∀-intro

Induction step:
We have to show, within PA, that there exists a derivation of
(∀v3 (v1 +v2 = v3 ⊃ P r[v 1 + v2 = v3 q](v1 , v2 , v3 )) ⊃ ∀v3 (v1 +v2′ = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2′ , v3 ))).

(1) (1) ∀v3 (v1 + v2 = v3 ⊃ P r[v 1 + v2 = v3 q](v1 , v2 , v3 ) Assumption I.H.


(2) (2) v1 + v2′ = v3 Assumption
′ ′
(3) (v1 + v2 ) = (v1 + v2 ) N4

(2) (4) (v1 + v2 ) = v3 (2)(3) logic of =
(2) (5) ∃v4 (v4′ = v3 ) (4) ∃-Intro
(6) (6) v4′ = v3 Assumption
(6)(2) (7) (v1 + v2 )′ = v4′ (4)(6) subst of =
(6)(2) (8) v1 + v2 = v4 (7) N1 , logic
(1) (9) (v1 + v2 = v4 ⊃ P r[v 1 + v2 = v3 q](v1 , v2 , v4 )) (1) ∀-elimin
(1)(2)(6) (10) P r[pv1 + v2 = v3 q](v1 , v2 , v4 ) (8)(9) ⊃-elim
(11) ((v1 + v2 = v4 ∧ v1 + v2′ = (v1 + v2 )′ ) ⊃ v1 + v2′ = v4′ )
substitutivity of =
(12) P r[p((v1 + v2 = v4 ∧ v1 + v2 = (v1 + v2 ) ) ⊃ v1 + v2 = v4′ )q](v1 , v2 , v4 )
′ ′ ′

(11) P1∗
(13) ((P r[pv1 + v2 = v4 q](v1 , v2 , v4 ) ∧ P r[pv1 + v2′ = (v1 + v2 )′ q](v1 , v2 )) ⊃
P r[pv1 + v2′ = v4′ q](v1 , v2 , v4 )) (12) P2∗
(14) (P r[pv1 + v2 = v4 q](v1 , v2 , v4 ) ≡ P r[pv1 + v2 = v3 q](v1 , v2 , v4 ))
Lemma 123
(15) (P r[pv1 + v2 = v4 q](v1 , v2 , v4 ) ≡ P r[pv1 + v2 = v3 q](v1 , v2′ , v4′ ))
′ ′

Lemma 124

(16) ((P r[pv1 + v2 = v3 q](v1 , v2 , v4 ) ∧ P r[pv1 + v2 = (v1 + v2 )′ q](v1 , v2 )) ⊃
P r[pv1 + v2 = v3 q](v1 , v2′ , v4′ )) (13)(14)(15) propositional logic

(17) P r[pv1 + v2 = (v1 + v2 )′ q](v1 , v2 ) (3) P1∗
(1)(2)(6) (18) P r[pv1 + v2 = v3 q](v1 , v2′ , v4′ ) (16)(10)(17) ⊃-elim

(1)(2)(6) (19) P r[pv1 + v2 = v3 q](v1 , v2 , v3 ) (18)(6) subst =

(1)(2) (20) P r[pv1 + v2 = v3 q](v1 , v2 , v3 ) (6)(19) ∃-elimination
(1) (21) (v1 + v2′ = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2′ , v3 )) (2)(20) ⊃-intro
LECTURE 11 98

(1) (22) ∀v3 (v1 + v2′ = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2′ , v3 )) (21) ∀-intro (v3 not free in (1))
(23) (∀v3 (v1 + v2 = v3 ⊃ P r[v 1 + v2 = v3 q](v1 , v2 , v3 )) ⊃
∀v3 (v1 + v2′ = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2′ , v3 ))) (1)(22) ⊃-intro

By the instance of N12 for induction on v2 in the formula ∀v3 (v1 + v2 = v3 ⊃


P r[pv1 + v2 = v3 q](v1 , v2 , v3 )), we have proved in PA, ∀v3 (v1 +v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2 , v3 )).
Then by one step of ∀-elimination, we have (v1 +v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2 , v3 )),
which was to be proved.

We now turn to the induction steps of the proof.


(i) F is ∼ G where G is Σ0 .
Exercise.
(ii) F is (G ∧ H), for G and H both Σ0 -formulas.

(1) (G ⊃ P r[G]) Induction hypothesis


(2) (H ⊃ P r[H]) Induction hypothesis
(3) (G ⊃ (H ⊃ (G ∧ H))) propositional logic
(4) P r[(G ⊃ (H ⊃ (G ∧ H)))] (3) Theorem 120
(5)(P r[(G ⊃ (H ⊃ (G ∧ H)))] ⊃ (P r[G] ⊃ P r[(H ⊃ (G ∧ H))])) Theorem 121
(6) (P r[(H ⊃ (G ∧ H))] ⊃ (P r[H] ⊃ P r[(G ∧ H)])) Theorem 121
(7) (P r[G] ⊃ (P r[H] ⊃ P r[(G ∧ H)])) (4) (5) (6) propositional logic
(8) ((G ∧ H) ⊃ P r[(G ∧ H)]) (1) (2) (7) propositional logic

(iii) F is (∀v1 ≤ v2 )G(v1 ), for G(v1 ) a Σ0 -formula. This means that v2 is free in
F . To simplify notation we shall take it that no other variables are free in F , i.e. v1
is the only free variable in G(v1 ). We need to give a proof in PA of (F ⊃ P r[F ]),
i.e.
((∀v1 ≤ v2 )G(v1 ) ⊃ ∀v3 (S(p(∀v1 ≤ v2 )G(v1 )q, v2 , 0′′ , v3 ) ⊃ P (v3 )).
The proof is by induction on the variable v2 occurring free in the formula (F ⊃
P r[F ]).
Base case: We need to prove that ((∀v1 ≤ 0)G(v1 ) ⊃ ∀v3 (S(p(∀v1 ≤ 0)G(v1 )q, 0, 0′′ , v3 ) ⊃
P (v3 ))
To simplify notation I shall write v1 as x and v2 as y.

(1) (G(x) ⊃ P r[pG(x)q](x)) Induction hypothesis for the main induction


(2) (G(x) ⊃ P r[G(x)](x))(0) (1) ∀-Intro, ∀-Elim
(3) (G(0) ⊃ P r[G(x)(x)](0)) (2) defn of subst
LECTURE 11 99

(4) (P r[G(x)](0) ≡ P r[G(0)]) Lemma


(5) (G(0) ⊃ P r[G(0)]) (3) (4) propositional logic
(6) (∀x ≤ 0)G(x) ≡ G(0)) provable in PA
(7) ((∀x ≤ 0)G(x) ⊃ P r[(∀x ≤ 0)G(x)]) (5) (6) logical equivalences

Induction step: On the assumption ((∀x ≤ y)G(x) ⊃ P r[p(∀x ≤ y)G(x)q](y)),


which is the induction hypothesis for this sub-induction, we need to establish
((∀x ≤ y ′ )G(x) ⊃ P r[p(∀x ≤ y ′ )G(x)q](y ′ )).

(1) (1) ((∀x ≤ y)G(x) ⊃ P r[p(∀x ≤ y)G(x)q](y)) Assumption IH for sub-induction


(2) (2) (∀x ≤ y ′ )G(x) Assumption
(2) (3) ∀x(x ≤ y ′ ⊃ G(x)) (2) defn of (∀x ≤ y ′ )
(2) (4) ∀x((x ≤ y ∨ x = y ′ ) ⊃ G(x)) (3) N6
(2) (5) ∀x((x ≤ y ⊃ G(x)) ∧ (x = y ′ ⊃ G(x)) (4) propositional logic
(2) (6) (∀x(x ≤ y ⊃ G(x)) ∧ ∀x(x = y ′ ⊃ G(x))) (5) predicate logic
(2) (7) ((∀x ≤ y)G(x) ∧ G(y ′ )) (6) defn of (∀x ≤ y), logic
(8) (8) (G(x) ⊃ P r[pG(x)q](x)) Assumption IH for main induction
(8) (9) (G(y ′ ) ⊃ P r[pG(x)q](y ′ )) (8) ∀-Intro, ∀-Elim
(1)(2)(8) (10) (P r[p(∀x ≤ y)G(x)q](y) ∧ P r[pG(x)q](y ′ )) (7)(1)(9) propositional logic
(1)(2)(8) (11) P r[p((∀x ≤ y)G(x) ∧ [G(y)])q](x, y ′ ) (10) Lemma 122
(1)(2)(8) (12) P r[(∀x ≤ y ′ )G(x)] (11) N8 and logic
(1)(8) (13) ((∀x ≤ y ′ )G(x) ⊃ P r[(∀x ≤ y ′ )G(x)]) (2) (12) ⊃-Intro

Step (9) in the above derivation calls for comment. In deriving (9) from (8)
∀-Intro is applied to (8) and the status of (8), as an induction hypothesis, is that of
assumption. If the variable in (8) were the variable of induction this ∀-Intro would
be illegitimate. However, the induction of which (8) is an induction hypothesis is
over formulas, so the assumption is about the formula G(x), and not about x, i.e.
x is a free variable in (8) to which ∀-Intro may be applied. N

Corollary 127 (provable Σ1 -completeness with free variables) For each Σ1 -


formula F (vk1 , . . . , vkm ), PA ⊢ (F (vk1 , . . . , vkm ) ⊃ P r[pF (vk1 , . . . , vkm )q](vk1 , . . . , vkm ).

Proof. We abbreviate vk1 , . . . , vkm as v, to avoid clutter. That F (v) is a Σ1 -


formula means there is a Σ0 -formula G(v, vi ) such that F (v) = ∃vi G(v, vi ). The
following describes a derivation in PA.

(1) (G(v, vi ) ⊃ P r[pG(v, vi )q](v, vi ) Theorem 126


(2) (G(v, vi ) ⊃ ∃vi G(v, vi )) predicate logic
LECTURE 11 100

(3) (P r[pG(v, vi )q](v, vi ) ⊃ P r[p∃vi G(v, vi )q](v)) ( 2) P1∗ , P2∗


(4) (G(v, vi ) ⊃ P r[p∃vi G(v, vi )q](v) (1) (3) propositional logic
(5) ∀vi (G(v, vi ) ⊃ P r[p∃vi G(v, vi )q](v) (4) ∀-Intro
(6) (∃vi G(v, vi ) ⊃ P r[p∃vi G(v, vi )q](v) (5) anti-prenexing N
Lecture 12

The ω-rule and uniform reflection;


PA proves that PA proves every
instance of the Gödel sentence;
Π1-uniform reflection and
consistency; PA is Π1-conservative
over PAΠ2 ∪ {ConP A}

(Wednesday, 17 November 2010)

12.1 The ω-rule and uniform reflection


We noted in Section 8.4 that the first half of the Gödel Incompleteness Theorem for
a Σ0 -complete system S establishes that if S is consistent then it is ω-incomplete,
i.e. we have, for each n, S ⊢∼ P rov(pGq, n), but S 0 ∀v2 ∼ P rov(pGq, v2 ). If each
numerical instance of a formula F (vi ) with one free variable is true, then ∀vi F (vi )
is true. Hence the following inference, called the ω-rule, is sound with respect to
truth in arithmetic:

F (0), F (1), . . . , F (n) . . .

∀yF (y)

Definition 73 PAω =df PA + ω-rule.

101
LECTURE 12 102

Definition 74 PAω ⊢ X if and only if there is a tree of finite height with formulas
in the language of PA at each node and with X at the bottom node, such that for
each formula at a node without predecessor is a theorem of PA, and such that each
node is either a theorem of PA or there is one node directly above the node at which
X occurs such that for Y the formula at the predecessor node, PA ⊢ (Y ⊃ X), or
X is the result of an application of the ω-rule. By a tree consisting of a single node
(height 0), if PA ⊢ X, then PAω ⊢ X.

For G the Gödel sentence for S, PAω ⊢ G. A derivation using the ω-rule has
infinitely many premisses and hence is an infinite object, unlike a formal proof, or
our usual idea of a informal proof, so derivation of G by the ω-rule cannot be said
to constitute a proof of G. Indeed derivability from the axioms of PA by the ω-rule
is tantamount to truth.

Proposition 128 If a sentence X in the language of arithemtic is true, PAω ⊢ X.

Proof. We argue by induction over the arithmetical hierarchy. We know by


Lecture 7 that if X is a true Σ0 (= Π0 ) or Σ1 -sentence, PA ⊢ X, and hence PAω ⊢ X.
Suppose X is a true Π1 -sentence, i.e. of the form ∀v1 F (v1 ) for F (v1 ) a Σ0 -
formula. Then for each natural number n, PA ⊢ F (n), so by one application of the
ω-rule, PAω ⊢ ∀v1 F (v1 ). Assume for Induction Hypothesis that the result holds
for Σn and Πn -sentences. (i) Let X be a true Σn+1 -sentence ∃vi F (vi ), where F (vi )
is a Πn formula. Then for some natural number m, F (m) is a true Πn -sentence.
Then by Induction Hypothesis PAω ⊢ F (m). Since PA ⊢ (F (m) ⊃ ∃vi F (vi )),
PAω ⊢ ∃vi F (vi ). (ii) Let X be a true Πn+1 -sentence ∀vi F (vi ). Then for each
number n, F (n) is a true Σn -sentence. Then by Induction Hypothesis, for each n,
PAω ⊢ F (n). Then by one application of the ω-rule, PAω ⊢ ∀vi F (vi ). N

The completeness of PAω with respect to truth is entirely to do with the ω-rule,
and essentially nothing to do with the axioms for arithmetic of PA, as shown by the
fact that the corresponding system for R, i.e. Rω =df R + ω-rule, is also complete
with respect to truth.

Proposition 129 If a sentence X in the language of arithmetic is true, Rω ⊢ X.

Proof. The only facts about PA used in the proof of Proposition 128 are that
PA is Σ0 -complete and that PA ⊢ (F (m) ⊃ ∃vi F (vi )). Both these properties are
also facts about R. N

There are constructive versions of the ω-rule, based on the fact that we can
state in a single sentence that all numerical instances of a given formula F (v1 ) are
provable, and by the arithmetization of syntax such single sentences can be expressed
LECTURE 12 103

in the language of arithmetic, namely as ∀v1 P r[pF (v1 )q](v1 ), where P r[pF (v1 )q](v1 )
is defined by Definition 70 in Lecture 11. We can then give finite expression to an
ω-rule by the sentence:

(∀v1 P r[pF (v1 )q](v1 ) ⊃ ∀v1 F (v1 )).


Such sentences are highly sensitive to the axiomatic strength of the system to
which they are applied. They are called Reflection Principles (see Smorynski [6], p.
845). We use the following terminology:

Definition 75 For F (v1 ) a formula in the language of PA with one free variable, a
Uniform Reflection Principle is any sentence of the form
(∀v1 P r[pF (v1 )q](v1 ) ⊃ ∀v1 F (v1 )).

PA extended by Uniform Reflection Principles for all one-place formulas is strictly


weaker than PA extended by the infinitary ω-rule. This is because PA + ω-rule =
true arithmetic, while PA + all Uniform Reflection Principles is axiomatic and hence
incomplete.

12.2 PA proves that PA proves every instance of


the Gödel sentence
Theorem 130 PA⊢ ∀v1 P r[p∼ P rov(pGq, v1 )q](v1 )

Proof. The following derivation shows the existence of a formal proof in PA.

(1) (1) ∼ P r[p∼ P rov(pGq, v1 )q](v1 ) Assumption


(2) (0 = 0′ ⊃∼ P rov(pGq, v1 )) PA ⊢∼ 0 = 0′
(3) (P r[0 = 0′ ] ⊃ P r[∼ P rov(pGq, v1 )]) (2) P1∗ , P2∗ , MP
(4) (∼ P r[∼ P rov(pGq, v1 )] ⊃∼ P r[0 = 0′ ]) (3) contraposition
(1) (5) ∼ P r[0 = 0′ ] (1) (4) MP
(6) (∼ P r[0 = 0′ ] ⊃∼ ∃v1 P rov(pGq, v1 )) line (5) of proof of Theorem 104
and Proposition 118
(1) (7) ∼ ∃v1 P rov(pGq, v1 )) (5) (6) MP
(1) (8) ∀v1 ∼ P rov(pGq, v1 )) (7) pred logic
(1) (9) ∼ P rov(pGq, v1 )) (8) ∀-elim
(10) (∼ P rov(pGq, v1 ) ⊃ P r[∼ P rov(pGq, v1 )]) Theorem 126
(1) (11) P r[∼ P rov(pGq, v1 )] (9) (10) MP
(12) P r[∼ P rov(pGq, v1 )] (1)(11) reductio ad absurdum
(13) ∀v1 P r[∼ P rov(pGq, v1 )] (12) ∀-Intro N
LECTURE 12 104

Corollary 131 PA ∪{(∀v1 P r[p∼ P rov(pGq, v1 )q](v1 ) ⊃ ∀v1 ∼ P rov(pGq, v1 ))}


⊢G

Proof. By the Theorem, Modus ponens, and the fact that


PA ⊢ (G ≡ ∀v1 ∼ P rov(pGq, v1 )). N

12.3 Equivalence of Π1-Uniform Reflection and con-


sistency
Π1 -Uniform Reflection for PA, i.e. the Uniform Reflection Principle restricted to Π1 -
formulas, is provably equivalent in PA to ConP A , the formal consistency statement
for PA, ∼ P rP A (p0 = 0′ q).

Theorem 132 PA + Π1 -uniform reflection ⊢∼ P rP A (p0 = 0′ q).

Proof. Take the instance of Π1 -uniform reflection for the Π1 -formula


∀v1 (0 = 0′ ∧ v1 = v1 ). This is provably equivalent to (P r(p0 = 0′ q) ⊃ 0 = 0′ ).
PA ∪{(P r(p0 = 0′ q) ⊃ 0 = 0′ )} ⊢∼ P r(p0 = 0′ q). N

Theorem 133 For F (v1 ) any Π1 -formula with one free variable,
PA ∪{ConP A } ⊢ (∀v1 P r[pF (v1 )q](v1 ) ⊃ ∀v1 F (v1 )).

Proof.

(1) (1) ∀v1 P r[pF (v1 )q](v1 ) Assumption


(2) (2) ∼ F (v1 ) Assumption
(3) (∼ F (v1 ) ⊃ P r[p∼ F (v1 )q](v1 ) Corollary 127 since
∼ F (v1 ) is Σ1
(2) (4) P r[p∼ F (v1 )q](v1 ) (2)(3) MP
(1) (5) P r[pF (v1 )q](v1 ) (1) ∀-elim
(1)(2) (6) P r[p(F (v1 )∧ ∼ F (v1 ))q](v1 ) Lemma 122
(7) ((F (v1 )∧ ∼ F (v1 )) ⊃ 0 = 0′ ) propositional logic
(8) P r[p(F (v1 )∧ ∼ F (v1 )) ⊃ 0 = 0′ q](v1 ) (7) P1∗
(9) P r[p(F (v1 )∧ ∼ F (v1 ))q](v1 ) ⊃ P r[p0 = 0′ q] (8) P2∗
(10) ∼ P r[p0 = 0′ q] ⊃∼ P r[(p(F (v1 )∧ ∼ F (v1 ))q](v1 ) (9) contraposition
(11) ∼ P r(p0 = 0′ q) given
(12) ∼ P r[p0 = 0′ q] (11) Proposition 118
(13) ∼ P r[(p(F (v1 )∧ ∼ F (v1 ))q](v1 ) (12) (10) ⊃-elimination
(1)(2) (14) ((6) ∧ (13)) (6) (13) ∧-intro
(1) (15) F (v1 ) (2) (6) RAA
LECTURE 12 105

(1) (16) ∀v1 F (v1 ) (15) ∀-intro (v1 not free in (1)
(17) (∀v1 P r[pF (v1 )q](v1 ) ⊃ ∀vi F (v1 ) (1) (16) ⊃-intro

This proof covers the case where F (v1 ) is Σ0 rather than Π0 , in which case the
justification at line (3) is Theorem 126 rather than Corollary 127. N

12.4 PA is Π1-conservative over PAΠ2 ∪ {ConP A}


In the 1920s David Hilbert adumbrated a research programme which had at its
heart the project of giving proofs of the consistency of formal systems of infini-
tary mathematics in finitary mathematics. The motivation for this programme was
foundational and philosophical but Hilbert came to see that it also promised math-
ematical application in terms of establishing “conservative extension” results.
A theory S2 is an extension of S1 if S2 proves everything that S1 proves, and
S2 is a conservative extension of S1 if for every formula X in the language of S1 , if
S2 ⊢ X, then S1 ⊢ X, i.e. S2 proves nothing in the language of S1 that S1 doesn’t
already prove. More precisely and more generally,

Definition 76 For theories S1 and S2 formulated in the same language (or such that
the language of S2 is an extension of the language of S1 or such that the language
of S1 can interpreted in the language of S2 , but we shall not be concerned with these
more general cases), S2 is an extension of S1 if for each formula X in the language
of S1 , if S1 ⊢ X, then S2 ⊢ X (or S2 ⊢ X ∗ where X ∗ is the translation of X into
the language of S2 ).

Definition 77 An extension S2 of S1 is conservative over S1 with respect to a class


of formulas Γ if for each formula X in Γ, if S2 ⊢ X, then S1 ⊢ X.

A fundamental insight of Hilbert’s that lies at the heart of his programme of


proof theory is that if finitary mathematics can prove the consistency of infinitary
mathematics, then infinitary mathematics is a conservative extension of finitary
mathematics with respect to finitary mathematics. Hilbert sketches an argument
for this claim in [4], p. 474.
Gödel’s Second Incompleteness Theorem shows that, insofar as infinitary math-
ematics is an extension of finitary mathematics, the consistency of infinitary math-
ematics cannot be proved within finitary mathematics. Nonetheless Hilbert’s ar-
gument adumbrates a correct mathematical theorem the main content of which is
the proof of Theorem 133 that consistency implies the arithmetized ω-rule (a.k.a.
the uniform reflection principle) for Π1 -sentences. Hilbert formulates his argument
in terms of a particular Π1 -sentence, Fermat’s last theorem: “Let us suppose, for
LECTURE 12 106

example, that we had found, for Fermat’s great theorem, a proof in which the [in-
finitary] logical function ǫ was used. We could then make a finitary proof out of it
in the following way.”
Leaving aside the question, in what minimal system can the consistency of PA
be proved, which is beyond the scope of this course (the answer is, very roughly,
constructive principles of abstract mathematics, rather than finitary principles of
concrete mathematics), a precise working out of the argument Hilbert sketched
requires that the proof of Theorem 133 be carried out in finitary mathematics.
Hilbert never formulated clearly what he meant by finitary mathematics, i.e. he
never gave a formal system of finitary mathematics, and I won’t enter here the
debate over what formal system should be taken to capture the intended notion of
finitary arithmetic. Rather, I will address the question, in how weak a subsystem of
PA can the argument for Theorem 133 be carried out?
The key point is that such a system must be strong enough to prove provable Σ0 -
completeness of PA. Hilbert did not explicitly formulate provable Σ0 -completeness,
but it is implicit in his argument, and implicitly he takes it to be a fact of finitary
mathematics. “Let us assume that numerals p, a, b, c (p > 2) satisfying Fermat’s
equation ap +bp = cp are given; then we could also obtain this equation as a provable
formula by giving the form of a proof to the procedure by which we ascertain that
the numbers ap + bp and cp coincide.”
Proving provable Σ0 -completeness requires mathematical induction, so we may
take the question to be, how much induction, measured by complexity in the arith-
metical hierarchy of formulas to which induction must be applied, is needed for this
proof?
The minimum is Σ1 -induction, i.e. axioms N12 for Σ1 -formulas, as in the step of
the proof in Lecture 11 in which we proved that
((∀v1 ≤ v2 )G ⊃ ∀v3 (S(p(∀v1 ≤ v2 )Gq, v2 , 0′′ , v3 ) ⊃ P (v3 )) by induction on the free
variable v2 . However, we also used Π2 -induction, in our proof in PA that (v1 +
v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2 , v3 )), since we proved this by induction on the
formula ∀v3 (v1 + v2 = v3 ⊃ P r[pv1 + v2 = v3 q](v1 , v2 , v3 )). It might be that there
is some clever way to reconstruct that proof so that the universal quantification of
the induction formula is not needed. In any case, from what has been established,
we have the following theorem corresponding to Hilbert’s argument claiming that
infinitary mathematics is conservative over finitary mathematics with respect to
Π1 -theorems.

Theorem 134 For X any Π1 -sentence in the language of PA, if P A ⊢ X, then


P AΠ2 ∪ {ConP A } ⊢ X.

Proof. By Theorem 133 and analysis of the proof of Theorem 126. N


Lecture 13

Provability logic: the system GL

(Tuesday, 23th November 2010)

A proof predicate P r(v1 ) for a system S can be thought of as an operator on


sentences in the language of S, i.e. it generates a sentence from a sentence. To signify
this viewpoint we write 2A for P r(pAq). In this notation the arithmetization of
Löb’s theorem (problem 1 on Problem sheet 7) is expressed as (2(2A ⊃ A) ⊃ 2A).
As we shall see, this formula provides the fundamental axiom by which to axiomatize
the logic of provability.
When provability logic first began to be developed, in the 1970s, there existed
already, for more than fifty years, systems of logic for a sentence operator 2A with
the intended meaning, “A is necessarily true”, and in the 1930s, in a brief note, Gödel
had obtained results by interpreting the 2 operator of modal logic as provability.
Such systems are called modal logic since necessity concerns not only the truth of
sentences but also the kind or mode of their truth. Modal logic provided a framework
for setting up systems for provability logic (and also a semantics of possible worlds
with an accessibility relation between worlds by which to study properties of such
systems, which has been exploited in the study of provability logic, but we will
establish the results that concern us here purely syntactically, i.e. not using these
semantic techniques). It is important to realize that provability logic is not an
extension of the logic of necessity, for which (2X ⊃ X) is valid, in contrast to which
Löb’s theorem shows that if S 0 X, then S 0 (P r(pXq) ⊃ X).
The system of provability logic was named GL by George Boolos, after Gödel
and Löb. It consists of propositional logic plus the provability operator.

13.1 The language of GL


The primitive symbols of GL:

107
LECTURE 13 108

A sentence ⊥ (standing for a generic false sentence, e.g. 0 = 0′ in the language


of arithmetic).
Infinitely many sentence letters pi , generated from the symbol ‘p’ and iteration
of the subscript symbol ‘′ ’.
The sentential connective ⊃.
The sentential operator 2.

Definition 78 (sentences of GL) By recursion:


base: ⊥ and all pi are sentences.
recursion: If X and Y are sentences, (X ⊃ Y ) is a sentence.
If X is a sentence, 2X is a sentence.

We shall write sentences in the language of GL using the following abbrevia-


tions which, on the intended meaning for ⊃ and ⊥, express negation, conjunction,
disjunction, and equivalence:

Definition 79 ∼ X =df (X ⊃⊥),


(X ∧ Y ) =df ((X ⊃ (Y ⊃⊥)) ⊃⊥),
(X ∨ Y ) =df ((X ⊃⊥) ⊃ Y )
(X ≡ Y ) =df (((X ⊃ Y ) ⊃ ((Y ⊃ X) ⊃⊥)) ⊃⊥)

13.2 The axioms and inference rules of GL


Definition 80 (axioms of GL) A1. (Tautologies) Every sentence in the language
of GL that is a truth functional tautology when ⊥ is assigned the truth value F
(falsity) and ⊃ is interpreted as the truth function ‘if . . . then . . . ’ is an axiom.
A2. (Distribution) For X and Y any sentences in the language of GL, (2(X ⊃
Y ) ⊃ (2X ⊃ 2Y )) is an axiom. (Corresponds to property P2 of provability predi-
cates.)
A3. (Arithmetized Löb’s Theorem) For each sentence X in the language of GL,
(2(2X ⊃ X) ⊃ 2X) is an axiom.

Definition 81 (rules of inference of GL) R1. From sentences X and (X ⊃ Y ),


infer Y . (Modus ponens)
R2. From sentence X infer 2X. (Corresponds to property P1 of provability
predicates. In modal logic this rule is known as Necessitation.)

There is no axiom schema corresponding to property P3 for provability predicates


since, as we shall see by Theorem 147, (2X ⊃ 22X) is derivable from the axioms
and rules of inference specified for GL.
LECTURE 13 109

The axioms and inference rules of GL arise by abstraction from the arithmetized
proof predicate for PA. Conversely, the axioms and inference rules, and hence all
theorems of GL, translate into theorems of PA. This result means that GL is sound
with respect to interpretation in PA.

Definition 82 An interpretation ∗ of GL in PA is given by translating the language


of GL into the language of PA by the following inductive definition:
Base:
(i) ⊥∗ = X for X a specified sentence in the language of PA such that P A ⊢∼ X;
(ii) p∗i = si for i 7→ si a specified enumeration of the sentences in the language
of PA.
Induction:
(iii) (X ⊃ Y )∗ = (X ∗ ⊃ Y ∗ );
(iv) (2X)∗ = P r(pX ∗ q), for P r(v1 ) an arithmetized proof predicate for PA.

Note that an interpretation ∗ in the above specification is determined by the


choice of a sentence X such P A ⊢∼ X and a specific enumeration of the sentences
in the language of PA. Also note that we can reformulate the language and logic
of PA by including the symbol ⊥ in the language of PA with corresponding axioms
for propositional logic so that ⊥ is treated as falsity, in which case an interpretation
of GL in PA is determined by the choice of enumeration of the sentences in the
language of PA.

Theorem 135 (soundness of GL with respect to interpretation in PA) For


every interpretation ∗ from GL into PA as in Definition 82, if GL ⊢ X, then
P A ⊢ X ∗.

Proof. By induction over the recursive definition of theorems of GL.


Basis step, for X an axiom of GL:
(i) If X is a truth functional tautology, then X ∗ is a truth functional tautology,
and hence provable in PA.
(ii) If X is a Distribution Axioms, then X ∗ is an instance of P2 and hence provable
in PA.
(iii) If X is a Löb’s Theorem axiom, then X ∗ is an instance of arithmetized Löb’s
Theorem, which by problem 1 on Problem sheet 7, is provable in PA.
Induction step:
If P A ⊢ X ∗ and P A ⊢ (X ⊃ Y )∗ , then since (X ⊃ Y )∗ = (X ∗ ⊃ Y ∗ ), P A ⊢ Y ∗ .
If P A ⊢ X ∗ , then by P1 from arithemetization of the provability predicate for
PA, P A ⊢ P r(pX ∗ q), i.e. P A ⊢ (2X)∗ . N

There is also a completeness theorem, due to Robert Solovay, for GL with respect
to interpretation in PA.
LECTURE 13 110

Theorem 136 (Solovay completeness theorem for GL) For X any sentence
in the language of GL, if for every interpretation ∗ (as in Definition 82), P A ⊢ X ∗ ,
then GL ⊢ X; or equivalently, for X any sentence in the language of GL, if GL 0 X,
then there is an interpretation ∗ such that P A 0 X ∗ .

Proof. Beyond the scope of this course.

13.3 Some derivations in GL


Unarithmetized Löb’s Theorem holds for GL, i.e.

Lemma 137 If GL ⊢ (2X ⊃ X), then GL ⊢ X.

Proof. Suppose GL ⊢ (2X ⊃ X). Then by R1, GL ⊢ 2(2X ⊃ X). By A3,


GL ⊢ (2X(2 ⊃ X) ⊃ 2X), so by R1, GL ⊢ 2X. Then by the initial supposition
and R1, GL ⊢ X. N

Lemma 138 If GL ⊢ (A ⊃ B), then GL ⊢ (2A ⊃ 2B).

Proof. This is a notational variant of P4 and the proof of P4 in Lemma 103


proves this result. N
Next we note that GL is closed under truth functional consequence, i.e.

Theorem 139 If Y is a truth functional consequence of finitely many formulas


which are each provable in GL, then GL ⊢ Y .

Proof. If Y follows truth functionally from the finitely many formulas X1 , . . . , Xr ,


then (X1 ⊃ (X2 ⊃ (. . . ⊃ (Xr ⊃ Y ) . . .))) is a tautology and hence an axiom of GL,
and GL ⊢ Y by r-many applications of Modus ponens starting with this axiom. N

Remarks. In proofs that depend on this theorem I will say “by propositional
logic in GL” or just “by propositional logic”, rather than citing Theorem 139. By
the compactness theorem for truth functional logic, the above result holds for Y a
truth functional consequence of any set of provable formulas, not just finite sets of
formulas, but we have not need for this generalization.

Theorem 140 GL ⊢ (2(X ∧ Y ) ≡ (2X ∧ 2Y ))


LECTURE 13 111

Proof. (i) Both ((X ∧ Y ) ⊃ X) and ((X ∧ Y ) ⊃ Y ) are tautologies and hence
axioms of GL. By Lemma 138, GL ⊢ (2(X ∧Y ) ⊃ 2X) and GL ⊢ (2(X ∧Y ) ⊃ 2Y ).
Then by propositional logic in GL, GL ⊢ (2(X ∧ Y ) ⊃ (2X ∧ 2Y ))
(ii) Since (X ⊃ (Y ⊃ (X ∧ Y ))) is a tautology, GL ⊢ (X ⊃ (Y ⊃ (X ∧ Y ))).
Hence by Lemma 138, GL ⊢ (2X ⊃ 2(Y ⊃ (X ∧ Y ))). By A2, GL ⊢ (2(Y ⊃
(X ∧ Y ) ⊃ (2Y ⊃ 2(X ∧ Y )))). Then by propositional logic in GL, GL ⊢ (2X ⊃
(2Y ⊃ 2(X ∧ Y ))). The formulas ((2X ∧ 2Y ) ⊃ 2X) and ((2X ∧ 2Y ) ⊃ 2Y )
are tautologies and hence axioms of GL, so by propositional logic in GL, GL ⊢
((2X ∧ 2Y ) ⊃ 2(X ∧ Y )). N

Proposition 141 GL ⊢ (2(X ≡ Y ) ⊃ (2X ≡ 2Y ))

Proof. By Definition 79, (X ≡ Y ) =df ((X ⊃ Y ) ∧ (Y ⊃ X)), so by Theo-


rem 140, GL ⊢ (2(X ≡ Y ) ≡ (2(X ⊃ Y ) ∧ 2(Y ⊃ X))). By A2 and Theorem 139,
GL ⊢ ((2(X ⊃ Y ) ∧ 2(Y ⊃ X)) ⊃ ((2X ⊃ 2Y ) ∧ (2Y ⊃ 2X))). Then by propo-
sitional logic in GL, GL ⊢ (2(X ≡ Y ) ⊃ ((2X ⊃ 2Y ) ∧ (2Y ⊃ 2X))), which by
Definition 79 is GL ⊢ (2(X ≡ Y ) ⊃ (2X ≡ 2Y )). N

The converse of Proposition 141 does not hold.

Proposition 142 If PA is 1-consistent, GL 0 ((2X ≡ 2Y ) ⊃ 2(X ≡ Y )).

Proof. P A 0 ((P r(pGq) ≡ P r(p0 = 0′ q)) ⊃ (P r(pG ≡ 0 = 0′ q))), since P A ⊢


(P r(pGq) ≡ P r(p0 = 0′ q)), by Theorem ??, and P A 0 (P r(pG ≡ 0 = 0′ q)) if PA is
1-consistent. Hence by Theorem 135, GL 0 ((2X ≡ 2Y ) ⊃ 2(X ≡ Y )). N

13.4 Closure of GL under substitution by prov-


ably equivalent formulas
The result of substituting a sentence A for sentential variable p in formula F , Fp (A),
is defined by the following recursion.

Definition 83 (Fp (A)) 1. If F = p, then Fp (A) = A.


2. If F = q where q 6= p, then Fp (A) = q.
3. If F =⊥, then Fp (A) =⊥.
4. If F = (G ⊃ H), then Fp (A) = (Gp (A) ⊃ Hp (A)).
5. If F = 2G, then Fp (A) = 2(Gp (A)).

Proposition 143 (closure of GL under substitution) For X any formula in


the language of GL, and F (p) any formula in the language of GL in which the
sentence letter p occurs, if GL ⊢ F (p), then GL ⊢ Fp (X).
LECTURE 13 112

Proof. By induction on the length of a proof of F (p) in GL.


Base case: The proof is of length 1, i.e. F (p) is an axiom, in which case it is of
the form A1 (tautology), A2 (distribution), or A3 (arithmetized Löb’s Theorem).
Then Fp (X) is an axiom, of the same form as F (p) is, so GL ⊢ Fp (X).
Induction steps: (i) The proof of F (p) ends with an application of R1, i.e. GL ⊢
Y and GL ⊢ (Y ⊃ F (p)). By Induction Hypothesis, GL ⊢ Yp (X) and GL ⊢ (Y ⊃
F (p))p (X). Then by clause 4 of Definition 83, GL ⊢ (Yp (X) ⊃ Fp (X)), so by R1,
GL ⊢ Fp (X).
(ii) The proof of F (p) ends with an application of R2, i.e. F (p) is of the form
2G(p) and GL ⊢ G(p). By Induction Hypothesis, GL ⊢ Gp (X). By R2, GL ⊢
2(Gp (X)). Then by clause 5 of Definition 83, GL ⊢ Fp (X). N

Theorem 144 (provable equivalence of substitution of provable equivalents)


For all formulas A, B, and F and any sentence letter p, if GL ⊢ (A ≡ B), then GL
⊢ (Fp (A) ≡ Fp (B)).

Proof. We argue by induction over the inductive definition of formulas F .


If F = p, what is to be proved is that if GL ⊢ (A ≡ B), then GL ⊢ (A ≡ B),
which holds trivially.
If F = q for p 6= q, what is to be proved is GL ⊢ (A ≡ B), then GL ⊢ (q ≡ q).
The consequent is provable outright, so the implication holds.
If F =⊥, what is to be proved is GL ⊢ (A ≡ B), then GL ⊢ (⊥≡⊥), for which
again the consequent is provable outright.
If F = (G ⊃ H), then we have as induction hypotheses, if GL ⊢ (A ≡ B), then
GL ⊢ (Gp (A) ≡ Gp (B)), and if GL ⊢ (A ≡ B), then GL ⊢ (Hp (A) ≡ Hp (B)). By
Definition 83, what is to be proved is if GL ⊢ (A ≡ B), then GL ⊢ ((Gp (A) ⊃
Hp (A)) ≡ (Gp (B) ⊃ Hp (B))), which follows by propositional logic from the induc-
tion hypotheses.
If F = 2G, then we have by induction hypothesis that if GL ⊢ (A ≡ B), then
GL ⊢ (Gp (A) ≡ Gp (B)). Then by Lemma 138 and propositional logic in GL, GL
⊢ (2(Gp (A)) ≡ 2(Gp (B))), so by Definition 83, GL ⊢ ((2G)p (A) ≡ (2G)p (B)),
which is to say, GL ⊢ (Fp (A) ≡ Fp (B)). N
The proof of Theorem 144 generalizes to establish provable equivalence of sub-
stitution on more than one sentence letter.

Theorem 145 (substitution on more than one sentence letter) For F any
formula with sentence letters pk1 , . . . , pkn , and for pairs of formulas Ai , Bi , i =
1, . . . , n, such that GL ⊢ (Ai ≡ Bi ), GL ⊢ (F (A1 , . . . , An ) ≡ F (B1 , . . . , Bn )), where
F (A1 , . . . , An ) is the result of substituting Ai for pki in F , and F (B1 , . . . , Bn ) is the
result of substituting Bi for pki in F .
LECTURE 13 113

Proof. Exactly the same proof structure as for Theorem 144, with just reformu-
lation of the induction hypothesis so that it’s for multiple substitutions, establishes
this result. N
Theorem 144 immediately establishes that GL is closed under substitution of
provable equivalents.

Corollary 146 If GL ⊢ Fp (A) and GL ⊢ (A ≡ B), then GL ⊢ Fp (B).

Proof. From GL ⊢ (A ≡ B) and Theorem 144, we have by ∧-elimination, GL ⊢


(Fp (A) ⊃ Fp (B)), so by Modus ponens from GL ⊢ Fp (A), we have GL ⊢ Fp (B). N

We are now able to show that GL proves P3 .

Theorem 147 GL ⊢ (2X ⊃ 22X)

Proof. The formula (X ⊃ ((22X ∧ 2X) ⊃ (2X ∧ X)) is a tautology and


hence GL ⊢ (X ⊃ ((22X ∧ 2X) ⊃ (2X ∧ X)). Since, by Theorem 140, GL
⊢ ((22X ∧ 2X) ≡ 2(2X ∧ X))), by Corollary 146, taking Fp as (X ⊃ (p ⊃
(2X ∧ X)), GL ⊢ (X ⊃ (2(2X ∧ X) ⊃ (2X ∧ X)). Then by Lemma 138 (P4 ),
GL ⊢ (2X ⊃ 2(2(2X ∧ X) ⊃ (2X ∧ X)). The formula (2(2(2X ∧ X) ⊃
(2X ∧ X)) ⊃ 2(2X ∧ X)) is an A3 axiom of GL. Then by propositional logic in
GL, GL ⊢ (2X ⊃ 2(2X ∧ X)). Then by Theorems 140 and propositional logic in
GL, GL ⊢ (2X ⊃ (22X ∧ 2X). Since ((22X ∧ 2X) ⊃ 22X) is a tautology, by
propositional logic in GL, GL ⊢ (2X ⊃ 22X). N

13.5 Closure of GL under substitution of provably


equivalent formulas is provable in GL
We now show that the closure of GL under substitution of provably equivalent
formulas, Theorem 144, can be formalized in GL.

Theorem 148 (arithmetized substitution theorem) For all formulas A, B,


and F and propositional variable p, GL ⊢ (2(A ≡ B) ⊃ 2(Fp (A) ≡ Fp (B))).

Proof. The proof is by induction over the recursion that generates the formula
F.
If F = p, what is to be proved is GL ⊢ (2(A ≡ B) ⊃ 2(A ≡ B)), which is a
tautology and hence provable in GL.
If F = q for p 6= q, what is to be proved is GL ⊢ (2(A ≡ B) ⊃ 2(q ≡ q)). Since
(q ≡ q) is a tautology, GL ⊢ (q ≡ q), so by R2 , GL ⊢ 2(q ≡ q). The result follows
by propositional logic in GL.
LECTURE 13 114

If F =⊥, what is to be proved is GL ⊢ (2(A ≡ B) ⊃ 2(⊥≡⊥)). The argument


is as for the preceding case.
If F = (G ⊃ H), we have as induction hypotheses, GL ⊢ (2(A ≡ B) ⊃
2(Gp (A) ≡ Gp (B))), and GL ⊢ (2(A ≡ B) ⊃ 2(Hp (A) ≡ Hp (B))). By propo-
sitional logic in GL, GL ⊢ (2(A ≡ B) ⊃ (2(Gp (A) ≡ Gp (B)) ∧ 2(Hp (A) ≡
Hp (B)))). Then by Theorem 140 and propositional logic in GL, GL ⊢ (2(A ≡
B) ⊃ 2((Gp (A) ≡ Gp (B)) ∧ (Hp (A) ≡ Hp (B)))). The following formula is a tau-
tology and so provable in GL (as an axiom): (((Gp (A) ≡ Gp (B)) ∧ (Hp (A) ≡
Hp (B))) ⊃ ((Gp (A) ⊃ Hp (A)) ≡ (Gp (B) ⊃ Hp (B)))). Then by Lemma 138
(P4 ), GL ⊢ (2((Gp (A) ≡ Gp (B)) ∧ (Hp (A) ≡ Hp (B))) ⊃ 2((Gp (A) ⊃ Hp (A)) ≡
(Gp (B) ⊃ Hp (B)))). From this result and the two steps earlier we have, by proposi-
tional logic in GL, GL ⊢ (2(A ≡ B) ⊃ 2((Gp (A) ⊃ Hp (A)) ≡ (Gp (B) ⊃ Hp (B)))),
which by Definition 83(4), is GL ⊢ (2(A ≡ B) ⊃ 2(Fp (A) ≡ Fp (B))).
If F = 2G, we have as induction hypothesis that GL ⊢ (2(A ≡ B) ⊃ 2(Gp (A) ≡
Gp (B))). Then by Proposition 141 and propositional logic in GL, GL ⊢ (2(A ≡
B) ⊃ (2(Gp (A)) ≡ 2(Gp (B)))). Then by Definition 83 (5.), GL ⊢ (2(A ≡ B) ⊃
(Fp (A) ≡ Fp (B))). Then by Lemma 138 (P4 ), GL ⊢ (22(A ≡ B) ⊃ 2(Fp (A) ≡
Fp (B))). By Theorem 147, GL ⊢ (2(A ≡ B) ⊃ 22(A ≡ B)), so by propositional
logic in GL, GL ⊢ (2(A ≡ B) ⊃ 2(Fp (A) ≡ Fp (B))). N

13.6 Strengthened proof that the closure of GL


under substitution of provably equivalent for-
mulas is provable in GL
This strengthened proof makes use of the following technical definition.

Definition 84 ⊡X =df (2X ∧ X)

Lemma 149 GL ⊢ (⊡X ⊃ X)

Proof. ⊢ (⊡X ⊃ X) is a tautology. N


Remark. Lemma 149 is a triviality but draws attention to a key property of ⊡
that holds for all formulas and which, by Löb’s Theorem for GL (Lemma 137), holds
for 2 only for formulas provable in GL. The constraint of Löb’s Theorem makes it
very difficult to derive an unboxed conclusion from a boxed premiss, i.e. of the form
2X. The situation is much more flexible if we are able to strengthen the premiss to
⊡X.

Lemma 150 For each formula X in the language of GL, the formulas 2X, 2 ⊡ X,
and ⊡2X are provably equivalent in GL.
LECTURE 13 115

Proof. The provable equivalence of 2 ⊡ X and ⊡2X is, by Definition 84, an


instance of Theorem 140, namely GL ⊢ (2(2X ∧ X) ≡ (22X ∧ 2X)).
By Theorem 147 (P3 ) and propositional logic in GL, GL ⊢ (2X ≡ (22X ∧2X)),
which by Definition 84 is GL ⊢ (2X ≡ ⊡2X). N

Corollary 151 GL ⊢ (⊡X ⊃ 2 ⊡ X)

Proof. By Lemma 149. N


Remark. The converse implication is not provable in GL, since
GL 0 (2X ⊃ X) unless GL ⊢ X (Lemma 137).

Lemma 152 If GL ⊢ (⊡X ⊃ Y ), then GL ⊢ (2X ⊃ 2Y ).

Proof. Assume GL ⊢ (⊡X ⊃ Y ). Then by Lemma 138, GL ⊢ (2 ⊡ X ⊃ 2Y ).


Then by Lemma 150 and propositional logic in GL, GL ⊢ (2X ⊃ 2Y ). N

Lemma 153 GL ⊢ (⊡X ⊃ Y ) if and only if GL ⊢ (⊡X ⊃ ⊡Y ).

Proof. (i) From Definition 84 by propositional logic in GL, GL ⊢ (⊡X ⊃ 2X).


Assume GL ⊢ (⊡X ⊃ Y ). Then by Lemma 152, GL ⊢ (2X ⊃ 2Y ). Then by
Theorem 139, GL ⊢ (⊡X ⊃ 2Y ). From the assumption and this last result by
propositional logic in GL, GL ⊢ (⊡X ⊃ (2Y ∧ Y )), i.e. GL ⊢ (⊡X ⊃ ⊡Y ).
(ii) Assume GL ⊢ (⊡X ⊃ ⊡Y ). By Lemma 149, GL ⊢ (⊡Y ⊃ Y ) and hence by
propositional logic in GL, GL ⊢ (⊡X ⊃ Y ). N

Theorem 154 (strengthened arithmetized substitution theorem) For all for-


mulas A, B, and F and propositional variable p, GL ⊢ (⊡(A ≡ B) ⊃ (Fp (A) ≡
Fp (B))).

Proof. The proof is by induction over the recursion that generates the formula
F.
If F = p, what is to be proved is GL ⊢ (⊡(A ≡ B) ⊃ (A ≡ B)), which holds by
Definition 84 and propositional logic in GL.
If F = q for p 6= q, what is to be proved is GL ⊢ (⊡(A ≡ B) ⊃ (q ≡ q)). For any
formula H, (H ⊃ (q ⊃ q)) is a tautology and hence an axiom of GL, so in particular
GL ⊢ (⊡(A ≡ B) ⊃ (q ≡ q)).
If F =⊥, what is to be proved is GL ⊢ (⊡(A ≡ B) ⊃ (⊥≡⊥)). The argument is
as for the preceding case.
If F = (G ⊃ H), we have as induction hypotheses, GL ⊢ (⊡(A ≡ B) ⊃ (Gp (A) ≡
Gp (B))), and GL ⊢ (⊡(A ≡ B) ⊃ (Hp (A) ≡ Hp (B))). By Definition 83(4), what
is to be proved is GL ⊢ (⊡(A ≡ B) ⊃ ((Gp (A) ⊃ Hp (A)) ≡ (Gp (B) ⊃ Hp (B)))),
which follows from the induction hypotheses by propositional logic in GL.
LECTURE 13 116

If F = 2G, we have as induction hypothesis that GL ⊢ (⊡(A ≡ B) ⊃ (Gp (A) ≡


Gp (B))). Then by Lemma 138, GL ⊢ (2⊡(A ≡ B) ⊃ 2(Gp (A) ≡ Gp (B))). By The-
orem 140, Axioms A2, and propositional logic in GL, GL ⊢ (2(Gp (A) ≡ Gp (B)) ⊃
(2(Gp (A)) ≡ 2(Gp (B)))). From these two last results, by propositional logic in
GL and Defintion 83(5), GL ⊢ (2 ⊡ (A ≡ B) ⊃ ((2Gp )(A) ≡ (2Gp )(B))). Then
by Corollary 151 and propositional logic in GL, GL ⊢ (⊡(A ≡ B) ⊃ ((2Gp )(A) ≡
(2Gp )(B))). N

Corollary 155 (variant proof of Theorem 148) For all formulas A, B, and F
and propositional variable p, GL ⊢ (2(A ≡ B) ⊃ 2(Fp (A) ≡ Fp (B))).

Proof. By Theorem 154 and Lemma 152.

Theorem 156 (Theorem 154 generalized to multiple substitutions) Let F (pi1 , . . . , pim )
be a formula with sentence letters pi1 , . . . , pim . Then
GL ⊢ (⊡(A1 ≡ B1 ) ∧ . . . ∧ ⊡(Am ≡ Bm )) ⊃ (F (A1 , . . . , Am ) ≡ F (B1 , . . . , Bm ))).

Proof. Exercise. N
Lecture 14

The fixed-point theorem for GL

(Wednesday, 24 November 2010)

14.1 The notion of a sentence letter modalized


in a sentence, and arithmetized substitution
for modalized sentences
Definition 85 Y is a subsentence of X is defined recursively by
Base case: X is a substence of X
Recursion clauses: If the sentence (Z ⊃ W ) is a subsentence of X, then Z is a
subsentence of X and W is a subsentence of X.
If the sentence 2Z is a subsentence of X, then Z is a subsentence of X.

Definition 86 A sentence letter p is modalized in a sentence X iff every occurrence


of p in X is a subsentence of a subsentence of X of the form 2Y .

Examples of sentences in which the sentence letter p is modalized: (1) 2p, (2)
∼ 2p, (3) 2 ∼ p, (4) ∼ 2 ∼ p, (5) (2p ⊃ q) (6) 2(2p ⊃ p), (7) (2(2p ⊃ p) ⊃ 2p),
(8) 2(p ≡ (2p ⊃ q)), (9) (2(2p ⊃ q)∧ ∼ 2p), (10) ⊥, (11) (⊥⊃⊥), (12) 2 ⊥, (13)
q (where q 6= p), (14) 2q (in these latter cases p is modalized in the sentence since p
does not occur in the sentence and so every occurrence of p in the sentence is within
the scope of 2).
Examples of sentences not modalized in p: (1) p, (2) (p ⊃⊥), (3) (2p ⊃ p), (4)
(p ≡ (2p ⊃ q).

Definition 87 For X a sentence in which the sentence letter p occurs modalized,


a sentence D(pk1 , . . . , pkn ) with sentence letters pk1 , . . . , pkn that do not occur in X

117
LECTURE 14 118

and sentences 2C1 (p), . . . , 2Cn (p) such that X(p) = D(2C1 (p), . . . , 2Cn (p)) (i.e.
the result of substituting each sentence 2Ci (p) for all occurrences of the variable pi
in the sentence D(pk1 , . . . , pkn )) is called a decomposition of X with respect to p;
D(pk1 , . . . , pkn ) is called a decomposition sentence for X, and 2C1 (p), . . . , 2Cn (p)
are called components of X.

Lemma 157 If X is modalized in p, then there is a decomposition of X with respect


of p.

Proof. We give a constructive proof of this result, i.e. a method, or in fact


two different methods, which generate a decomposition of X for X any sentence
modalized in p.
Method 1 (top down): The whole sentence X cannot consist just of a sentence
letter pi or ⊥ since no such sentence is modalized in p. So there are two cases, either
X is of the form 2Y for some sentence Y , or X is of the form (Y ⊃ Z) for sentences
Y and Z.
(i) X is of the form 2Y . Then we are done, with the decomposition sentence for
X any sentence letter pi distinct from p and component 2Y .
(ii) X is of the form (Y ⊃ Z). Then p occurs in Y or in Z, or both, and
wherever it occurs it occurs modalized. If the sentence or sentences in which p
occurs modalized is/are of the form 2W then by (i) we are done. If not then it is
of the form (U ⊃ V ) and we repeat the argument. Since X is generated in finitely
many steps, this process comes to an end, and since p occurs modalized, it ends in
components, i.e. sentences of the form 2C(p) in which p occurs modalized.
Method 2 (bottom up): For each occurrence of p in X, find the innermost oc-
currence of 2 in whose scope that occurrence of p occurs. For each such occurrence
of 2, let Ci (p) be the sentence to which that occurrence of 2 is prefixed. From the
resulting set of sentences 2Ci (p), discard those sentences that are a subsentence of
any of the others in which that occurrence of p occurs. The remaining sentences
2Ci (p) will be the components of the decomposition of X. In X replace each of
these component sentences by a distinct sentence letter not occurring in X, at each
occurrence in X of that component. This gives the decomposition sentence of X for
those components. N
Examples. (2(2(2p ⊃ p) ⊃ p) ⊃ 2p)
Method 1: The sentence is an implication between boxed formulas, so the com-
ponents are 2(2(2p ⊃ p) and 2p and the decomposition formula is (pi ⊃ pj ).
Method 2: When applied to the three occurrences of p in the antecedent sentence
of this implication, the procedure results in sentence 2p for the first occurrence of p
(going from the left), 2(2p ⊃ p)) for the second occurrence of p, and 2(2(2p ⊃ p) ⊃
p) for the third occurrence of p, and for the single occurrence of p in the consequent
sentence, it results in 2p. From the sentences resulting from the occurrences of p in
LECTURE 14 119

the antecedent sentence, we discard 2p because 2p with that occurrence of p occurs


in 2(2p ⊃ p)) (and also in 2(2(2p ⊃ p) ⊃ p)), and we discard 2(2p ⊃ p)) since it
is a subsentence of 2(2(2p ⊃ p) ⊃ p) and the two occurrences of p in 2(2p ⊃ p))
occur in 2(2(2p ⊃ p) ⊃ p). So the components are 2(2(2p ⊃ p) ⊃ p) and 2p,
and the decomposition sentence for these components is (pi ⊃ pj ).
Note that for some sentences modalized in a sentence letter the two methods
results in different decompositions, e.g.
(2(2p ⊃ q)∧ ∼ 2p)
By Method 1:
D2 (p1 , p2 ) = (p1 ∧ ∼ p2 ), 2C1 (p) = 2(2p ⊃ q), and 2C2 (p) = 2p.
By Method 2:
D1 (p1 , q) = (2(p1 ⊃ q)∧ ∼ p1 ), and 2C(p) = 2p.

Some modalized sentences have more than two decompositions, e.g.


222p has three decompositions:
(i) D1 = p1 and 2C1 (p) = 222p.
(ii) D2 = 2p1 and 2C1 (p) = 22p.
(iii) D3 = 22p1 and 2C1 (p) = 2p.
The decomposition of 222p that results from the proof of Lemma 157 is (iii).
The second example generalizes to show that

Proposition 158 For each n, there is a sentence, modalized in a sentence letter,


that has n-many decompositions.

Proof. For each n, |2 .{z


. . 2} p has decompositions
n
Di = |2 .{z
. . 2} p1 with component |2 .{z
. . 2} p, for each i such that 1 ≤ i ≤ n. N
n−i i

For sentences in a which a sentence letter occurs modalized there is an arith-


metized substitution theorem which yields the conclusion of Theorem 154 on the
hypothesis of Theorem 148.

Theorem 159 (arithmetized substitution in modalized sentences) Let F (p)


be a sentence in which sentence letter p occurs modalized. Then
GL ⊢ (2(A ≡ B) ⊃ (Fp (A) ≡ Fp (B))).

Proof. By Lemma 157, there is a decomposition D(p1 , . . . , pm ), with sentence


letters p1 , . . . , pm not in F (p) and components 2C1 (p), . . . , 2Cm (p). The following
is (a recipe for) a proof in GL.
LECTURE 14 120

(1) (2(A ≡ B) ⊃ 2(Cip (A) ≡ Cip (B))) Theorem 148


(2) (2(A ≡ B) ⊃ (2Ci (A) ≡ 2Ci (B))) (1) Proposition 141 prop logic
(3) (22(A ≡ B) ⊃ 2(2Ci (A) ≡ 2Ci (B))) (2) Lemma 138
(4) (2(A ≡ B) ⊃ 2(2Ci (A) ≡ 2Ci (B))) (3) Theorem 147 prop logic
(5) (2(A ≡ B) ⊃ ⊡(2Ci (A) ≡ 2Ci (B))) (2)(4)
(6) (((⊡(2C1 (A) ≡ 2C1 (B))) ∧ . . . ∧ ⊡(2Cm (A) ≡ 2Cm (B))) ⊃
(D(2C1 (A), . . . , 2Cm (A)) ≡ D(2C1 (B), . . . , 2Cm (B))))
Theorem 156
(7) (2(A ≡ B) ⊃ (Fp (A) ≡ Fq (B))) (5)(6) prop logic N

14.2 The fixed point theorem for GL


Definition 88 (fixed point) For X a sentence in the language of GL that contains
the sentence letter pi , a sentence F that contains only sentence letters contained in
X and does not contain pi is a fixed point for X with respect of pi if and only if
GL ⊢ (F ≡ Xpi (F )).

We shall see that every sentence modalized in pi has a fixed point with respect
to pi . We establish this result first for the simplest case of a sentence modalized in
pi , i.e. of the form 2Y (pi ).

Lemma 160 For pi any propositional variable and Y (pi ) any sentence in the lan-
guage of GL in which pi occurs, the result of substituting (⊥⊃⊥) for each occurrence
of pi in 2Y (pi ) is a fixed point for 2Y (pi ).

Proof. In this proof I abbreviate (⊥⊃⊥) as ⊤, and write 2Y (⊤) for the result
of substituting (⊥⊃⊥) for each occurrence of pi in 2Y (pi ). The following derivation
is a proof in GL.

(1) (2Y (⊤) ⊃ (⊤ ≡ 2Y (⊤)) tautology.


(2) 2(2Y (⊤) ⊃ (⊤ ≡ 2Y (⊤)) (1) P1 .
(3) (22Y (⊤) ⊃ 2(⊤ ≡ 2Y (⊤))) (2) P2 Modus ponens
(4) (2Y (⊤) ⊃ 2(⊤ ≡ 2Y (⊤))) (3) P3 propositional logic.
(5) (2(⊤ ≡ 2Y (⊤)) ⊃ 2(Y (⊤) ≡ Y (2Y (⊤)))) Theorem 148.
(6) (2(Y (⊤) ≡ Y (2Y (⊤))) ⊃ (2Y (⊤) ≡ 2Y (2Y (⊤)))) Proposition 141.
(7) (2Y (⊤) ⊃ (2Y (⊤) ≡ 2Y (2Y (⊤)))) (4) (5) (6) transitivity of ⊃.
(8) (2Y (⊤) ⊃ 2Y (2Y (⊤))) (7) prop logic.
LECTURE 14 121

(9) (2Y (⊤) ⊃ ⊡(⊤ ≡ 2Y (⊤))) (1) (4) prop logic


(10) (2Y (⊤) ⊃ (Y (⊤) ≡ Y (2Y (⊤)))) (9), Theorem 154, prop logic
(11) (Y (2Y (⊤)) ⊃ (2Y (⊤) ⊃ Y (⊤))) (10) prop logic.
(12) (2Y (2Y (⊤)) ⊃ 2(2Y (⊤) ⊃ Y (⊤))) (11) Lemma 138 (P4 ).
(13) (2Y (2Y (⊤)) ⊃ 2Y (⊤)) (12) A3 prop logic.

(14) (2Y (⊤) ≡ 2Y (2Y (⊤))) (8)(13)∧-Introduction. N

Before proving the fixed point theorem itself we need another lemma and for
that lemma we need a definition:

Definition 89 For 1 ≤ i ≤ m, let Ai (pk1 , . . . , pkm ), be m-many sentences each of


which includes among its sentence letters the m-many sentence letters pk1 , . . . , pkm .
A system of simultaneous equivalences of the form

(pki ≡ Ai (pk1 , . . . , pkm ))

is solvable (in GL) iff there are sentences F1 , . . . , Fm not containing pk1 , . . . , pkm
such that for 1 ≤ i ≤ m,

GL ⊢ (Fi ≡ Ai (F1 , . . . , Fm )).

Lemma 161 (solving systems of simultaneous equivalences) Every system of


m-many simultaneous equivalences of the form

(pki ≡ 2Ci (pk1 , . . . , pkm ))

is solvable.

Proof. The proof is by induction on m.


m = 1. This case is Lemma 160.
Suppose the Lemma holds for m. Let 2Ci (pk1 , . . . , pkm , pkm+1 ) be a set of (m +
1)-many sentences with m + 1-many sentence letters pk1 , . . . , pkm , pkm+1 . By the
Induction Hypothesis, there are sentences Fi (pkm+1 ), 1 ≤ i ≤ m, each of which does
not contain the sentence letters pk1 , . . . , pkm and does contain the sentence letter
pkm+1 , such that for 1 ≤ i ≤ m,

(1) GL ⊢ (Fi (pkm+1 ) ≡ 2Ci (F1 (pkm+1 ), . . . , Fm (pkm+1 ), pkm+1 )).


LECTURE 14 122

Now by Lemma 160 applied to the sentence that results from 2Cm+1 (pk1 , . . . , pkm , pkm+1 )
by substituting Fi (pkm+1 ) for pki for each i such that 1 ≤ i ≤ m, i.e.
2Cm+1 (F1 (pkm+1 ), . . . , Fm (pkm+1 ), pkm+1 ), there is a sentence Fm+1 such that
(2) GL ⊢ Fm+1 ≡ 2Cm+1 (F1 (Fm+1 ), . . . , Fm (Fm+1 ), Fm+1 )).
By substitution of Fm+1 into the provable equivalences (1), the sentences
F1 (Fm+1 ), . . . , Fm (Fm+1 ), Fm+1
solve the set of equivalences. N

Theorem 162 (Fixed Point Theorem) For every sentence X in the language of
GL modalized in pr , there is a sentence F containing only sentence letters that occur
in X and not containing pr such that GL ⊢ (F ≡ Xpr (F )).

Proof. Let X be any sentence modalized in pr . Then by Lemma 157, X(pr )


has a decomposition D(pk1 , . . . , pkn ), with sentence letters pk1 , . . . , pkn distinct from
pr , and components 2C1 (pr ), . . . , 2Cn (pr ), i.e. X(pr ) = D(2C1 (pr ), . . . , 2Cn (pr )),
where 2Ci (pr ) is substituted for pki in D(pk1 , . . . , pkn ).
By Lemma 161, the set of equivalences (pki ≡ 2Ci (D(pk1 , . . . , pkn ))) indexed by
i = 1, . . . , n, where the sentence D(pk1 , . . . , pkn ) is substituted for pr in 2Ci (pr ), is
solvable. Let F1 , . . . , Fn be a solution, i.e. for each i = 1, . . . , n,
GL ⊢ (Fi ≡ 2Ci (D(F1 , . . . , Fn ))). Then by Theorem 145,
GL ⊢ (D(F1 , . . . , Fn ) ≡ D(2C1 (D(F1 , . . . , Fn )), . . . , 2Cn (D(F1 , . . . , Fn )))),
which shows that D(F1 , . . . , Fn ) is a fixed point for X(pr ) with respect to pr . N
Remark. The sentence F gives the explicit, i.e. non self-referential, content of
a self-referential sentence p such that (p ≡ X(p)), for X modalized in p.
There are a number of different proofs of this theorem, most of which use Kripke
models for the modal operator (three such proofs are expounded by George Boolos in
his book The Logic of Provability, Cambridge University Press, 1993, pp. 104-123).
The purely syntactic proof given here follows Per Lindstrom, “Provability logic—
a short introduction”, Theoria 62 (1996), pp. 31-35; see also Craig Smorynski,
Self-Reference and Modal Logic, Springer, 1985, pp. 78-82.
The fixed points proved to exist by the Fixed Point Theorem are unique to within
provable equivalence.

Theorem 163 (provable equivalence of fixed points) Let X(p) be a sentence


in the language of GL in which the sentence letter p occurs modalized, and let q
be a sentence letter that does not occur in X(p). Abbreviating Xp (q) as X(q), GL
⊢ ((2(p ≡ X(p)) ∧ 2(q ≡ X(q))) ⊃ 2(p ≡ q)).
LECTURE 14 123

Proof. By Theorem 159 (arithmetized substitution into modalized sentences),


taking A as p and B as q,
(2(p ≡ q) ⊃ (X(p) ≡ X(q))) is provable in GL. Then by propositional logic,
(((p ≡ X(p)) ∧ (q ≡ X(q))) ⊃ (2(p ≡ q) ⊃ (p ≡ q))). By Lemma 138 and
Theorem 140,
((2(p ≡ X(p)) ∧ 2(q ≡ X(q))) ⊃ 2(2(p ≡ q) ⊃ (p ≡ q))). Then by Axiom A3
and propositional logic,
((2(p ≡ X(p)) ∧ 2(q ≡ X(q))) ⊃ 2(p ≡ q)). N
The uniqueness of the fixed point to within provable equivalence turns essentially
on the sentence letter of the fixed point equivalence occurring modalized in the
sentence for which the existence of a fixed point follows. We can have fixed points for
non-modalized sentences that are not unique with respect to provable equivalence.
For example, take X(p) as ((r ⊃ r) ⊃ ((q ⊃ q) ⊃ p)). By Definition 88, a fixed point
for X(p) is a sentence X containing only sentence letters that occur in X and not
containing p such that GL ⊢ (F ≡ Xp (F )). By this criterion, since GL ⊢ (q ≡ ((r ⊃
r) ⊃ ((q ⊃ q) ⊃ q))) and GL ⊢ (r ≡ ((r ⊃ r) ⊃ ((q ⊃ q) ⊃ r))), q and r are both
fixed points of ((r ⊃ r) ⊃ ((q ⊃ q) ⊃ p)). By R2 , GL ⊢ 2(q ≡ ((r ⊃ r) ⊃ ((q ⊃ q) ⊃
q))) and GL ⊢ 2(r ≡ ((r ⊃ r) ⊃ ((q ⊃ q) ⊃ r))). Suppose GL ⊢ 2(q ≡ r). Then
by Proposition 143, GL ⊢ 2(⊥≡ (⊥⊃⊥)), in which case GL ⊢ 2 ⊥. But assuming
the PA is Σ1 -sound, by Theorem 135, GL 0 2 ⊥. In which case GL 0 2(q ≡ r).
The previous example can be tweaked to give an example of sentence in which
p occurs not modalized that has a fixed point with respect to p which is unique
to within provable equivalence, e.g. ((q ⊃ q) ⊃ p). It is also worth noticing how
modalizing our original example to non-uniqueness to within provable equivalence
of a fixed point for a non-modalized sentence yield provable equivalence, e.g. ((r ⊃
r) ⊃ ((q ⊃ q) ⊃ 2p)). This sentence is provably equivalent to 2p, so all fixed points
for it are provably equivalent to the Henkin sentence.
Theorem 164 (Strengthened Fixed Point Theorem for GL) For every sen-
tence X in the language of GL modalized in p, there is a sentence F containing only
sentence letters that occur in X and not containing p such that
GL ⊢ (⊡(p ≡ X(p)) ≡ ⊡(p ≡ F )).
Proof. Theorems 162 and 163 establish this result. N
Theorem 165 Theorem 164 implies Theorems 162 and 163.
Proof. Substituting F for p in Theorem 164 results in GL ⊢ (⊡(F ≡ X(F )) ≡
⊡(F ≡ F )) Since (F ≡ F ) is a tautology, GL ⊢ (F ≡ F ), so GL ⊢ 2(F ≡ F ), so by
∧-Introduction, GL ⊢ ⊡(F ≡ F ). Hence GL ⊢ ⊡(F ≡ X(F )), so by ∧-Elimination,
GL ⊢ (F ≡ X(F )). N
Example. To find a fixed point for (∼ 2p ∧ ∼ 2 ∼ p).
Bibliography

[1] George Boolos, The Logic of Provability, Cambridge University Press, 1993.

[2] David Hilbert, “Axiomatische Denken”, Mathematische Annalen 78 (1918), pp.


405-415; English translation in William B. Ewald (ed),

[3] David Hilbert, “On the infinite” (1926); English translation in Jean van Hei-
jenoort (ed), (1927), p. 471.

[4] David Hilbert, “Die Grundlagen der Mathematik”, Abhandlungen aus dem
mathematischen Seminar der Hamburgeshcen Universität 6 (1928), English
translation by Stephan Bauer-Mengelberg, ‘The foundations of mathematics”
in Jean van Heijenoort (ed.) From Frege to Gödel; A Source Book in Mathe-
matical Logic 1879-1931, Harvard University Press, 1967.

[5] Georg Kreisel, “A refinement of ω-consistency” (Abstract), The Journal of Sym-


bolic Logic 22 (1957), pp. 108-109.

[6] Craig Smorynski, “The incompleteness theorems”, Jon Barwise (ed.), Handbook
of Mathematical Logic, Horth-Holland Publishing Company, 1977, pp. 821-865.

[7] Raymond M. Smullyan, Gödel’s Incompleteness Theorems, Oxford University


Press, 1992.

124

Anda mungkin juga menyukai