You are on page 1of 437

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the
text directly from the original or copy submitted. Thus, some thesis and
dissertation copies are in typewriter face, while others may be from any type of
computer printer.

The quality of this reproduction is dependent upon the quality of the copy
submitted. Broken or indistinct print colored or poor quality illustrations and
photographs, print bleedthrough, substandard margins, and improper alignment
can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and
there are missing pages, these will be noted. Also, if unauthorized copyright
material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning


the original, beginning at the upper left-hand comer and continuing from left to
right in equal sections with small overlaps.

Photographs included in the original manuscript have been reproduced


xerographically in this copy. Higher quality 6” x 9” black and white photographic
prints are available for any photographs or illustrations appearing in this copy for
an additional charge. Contact UMI directly to order.

Bell & Howell Information and Learning


300 North Zeeb Road, Ann Arbor, Ml 48106-1346 USA
®

IJM I 800-521-0600

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
STRICT CONSTRUCTIVISM AND THE PHILOSOPHY
OF MATHEMATICS
Volume I

i
s

| FENG YE
i
1

i
A DISSERTATION

PRESENTED TO THE FACULTY

OF PRINCETON UNIVERSITY

IN CANDIDACY FOR THE DEGREE

OF DOCTOR O F PHILOSOPHY

RECOMMENDED FO R ACCEPTANCE

BY THE DEPARTMENT OF

PHILOSOPHY

Ja n u a ry 2000

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
UMI N um ber 9957375

UMI*
UMI Microform9957375
Copyright 2000 by Bell & Howell Information and Learning Company.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.

Bell & Howell Information and Learning Company


300 North Zeeb Road
P.O. Box 1346
Ann Arbor, Ml 48106-1346

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
© Copyright by Feng Ye, 2000. All rights reserved.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
iii

Abstract
The dissertation studies the mathematical strength of strict con­
structivism, a finitistic fragment of Bishop’s constructivism, and ex­
plores its implications in the philosophy of mathematics.
It consists of two chapters and four appendixes. Chapter 1 presents
strict constructivism, shows that it is within the spirit of finitism, and
explains how to represent sets, functions and elementary calculus in
strict constructivism. Appendix A proves that the essentials of Bishop
and Bridges’ book Constructive Analysis can be developed within strict
constructivism. Appendix B further develops, within strict construc­
tivism, the essentials of the functional analysis applied in quantum
mechanics, including the spectral theorem, Stone’s theorem, and the
self-adjointness of some common quantum mechanical operators. Some
comparisons with other related work, in particular, a comparison with
S. Simpson’s partial realization of Hilbert’s program, and a discussion
of the relevance of M. B. Pour-El and J. I. Richards’ negative results
in recursive analysis are given in Appendix C.
Chapter 2 explores the possible philosophical implications of these
technical results. It first suggests a fictionalistic account for the ontol­
ogy of pure mathematics. This leaves a puzzle about how truths about
fictional mathematical entities are applicable to science. The chapter
then explains that for those applications of mathematics that can be
reduced to applications of strict constructivism, fictional entities can
be eliminated in the applications and the puzzle of applicability can
be resolved. Therefore, if strict constructivism were essentially suffi­
cient for all scientific applications, the applicability of mathematics of

R ep ro d u ced with p erm ission of the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
iv

mathematics in science would be accountable. The chapter then ar­


gues that the reduction of mathematics to strict constructivism also
reduces the epistemological question about mathematics to that about
elementary arithmetic. The dissertation ends with a suggestion that
a proper epistemological basis for arithmetic is perhaps a mixture of
Mill’s empiricism and the Kantian views.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
V

Acknowledgement
First of all, I want to express my deep gratitude to Princeton Uni­
versity and the Department of Philosophy for the generous award of a
graduate fellowship, without which this dissertation would have been
impossible.
I also want to express my deep gratitude to my advisors Professor
John P. Burgess and Professor Paul Benacerraf for the great help they
gave me during my five years of study at Princeton. This disserta­
tion is under their supervation. Their academic influence is obvious.
The conceptual framework under which the work in this dissertation
is conducted is shaped by their (among others’) researches, and what
I learned from them in various occations, in seminars and discussions,
helped to shape the main ideas and arguments in this dissertation (al­
though I must mention that they may not agree with all the philosoph­

ical oppinions I expressed here). Here I want to especially mention


that their help constantly goes beyond the academic aspects, which
I am particularly grateful to. For example, Professor John Burgess
patiently corrected numerous errors in my English.
I am deeply indebted to Professor Edward Nellson for what I learn
from his courses in mathematical logic, for the stimulating discussions
about this dissertation, and for a very valuable, detailed comment on
the dissertation.
I am indebted to Professor Solomon Feferman for a long commen­
tary that helps to improve this dissertation greatly. Thanks are also
due to Professor Douglas Bridges and Hajime Ishihara for giving helpful
comments and sending me their papers.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
VI

Many professors and fellow students gave helpful comments on sev­


eral occasions. Thanks are due to Professor John Cooper and Professor
Gilbert Harman for the comments in my dissertation talks at Prince­
ton; to Professor Gideon Rosen for the comments in a discussion; to
Professors Robert Adams, Theodore Brennen, Karsten Harries, Shelly

Kegan and Philip Kremer for some comments in a talk at Yale; to


Cian Dorr, Christopher Erlenkamp, Benjamin Friedman, Seahwa Kim,
Brian Lee and Cei Maslen for the comments in some seminars.
The work in this dissertation was partially inspired by Geoffrey
Heilman’s papers on the sufficiency of constructive mathematics for
developing quantum mechanics. I am deeply indebted to him although

my study gradually goes to the opposite direction.


Finally, the influence of Errett Bishop’s and Douglas Bridges’s work
is clear. Other academic influences and credits will be cited in due
course in the dissertation.

'H u s ^
, t| "T- no a f , p < la i.
\ OL* v * /v-1 /~ \

r u -Cd 1
>\/<2_ ~th(2.0f
Tnc.Of u oi

W ooJ o\

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
Contents

1. Introduction 1
1.1. The Background 1
1.2. Motivations 3
1.3. Summary of Contents 9
1.4. Further Research Topics 12

Chapter 1. A Logical Framework for Strict Constructivism 14


1. Introduction 14
2. The formal system P R O 25
3. The system SC 38
4. Inductions and Inductive Constructions in SC 49

5. Sets and Functions in SC 65


5.1. Introduction 65
5.2. Sets 67
5.3. Functions 76
5.4. Partial Functions 79
5.5. Complemented Sets 81
6. Calculus and the Real Numbers in SC 83
6.1. The Real Number System 83
6.2. Sequences and Series of Real Numbers 85
6.3. Continuous Functions 89
6.4. Differentiation 93

vii

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
Contents viii

6.5. Integration 95
6.6. Certain Important Functions 96
6.7. A Final Remark 98

Chapter 2. Toward a Philosophy of Mathematics 100


1. Fictionalism 100
1.1. Introduction 100
1.2. Geometric Figures and Sets 110
1.3. The Case of Numbers 115
1.4. The Indeterminacy of Fictional Entities 123
1.5. Fictionalism and the Indispensability Argument 129
1.6. Truth and Consistency in Fictionalism 137
1.7. Fictionalism and Other Positions in Philosophy of
Mathematics 146
2. On Applications of StrictConstructivism 157
2.1. The Problem of the Applicability of Mathematics 157
2.2. On Applications of Strict Constructivism 161
3. Remarks on the Epistemology of Arithmetic 172
3.1. Introduction 172
3.2. On Empiricism in Elementary Arithmetic 175
3.3. On Reducing Arithmetic to Logic 185
3.4. More Remarks on Kant and Mill 194

Appendix A. Constructive Analysis in SC 202


1. Introduction 202
2. Metric Spaces 204
2.1. Basic Definitions and Associated Structures 204
2.2. Completeness 206
2.3. Total Boundedness and Compactness 209

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
Contents ix

2.4. Spaces of Functions 211


2.5. Locally Compact Spaces 214
3. Complex Analysis 218
3.1. The Complex Plane 218
3.2. Derivatives 218

3.3. Integration 218


3.4. The Winding Number 221
3.5. Estimates of Size, and Location of Zeros 222
3.6. Singularities and Picard’s Theorem 250
4. Integration 251
4.1. Integration Spaces 251

4.2. Complete Extension of an Integral 255


4.3. Integrable Sets 258
4.4. Profiles 259
4.5. Positive Measures on R 261
4.6. Approximation by Compact Sets 262
4.7. Measurable Functions 267
4.8. Convergence of Functions and Integrals 269
4.9. Product Integrals 271
4.10. Measure Spaces 271
5. Normed Linear Spaces 272
5.1. Definitions and Examples 272
5.2. Finite-Dimensional Spaces 273
5.3. The Lp Spaces and the Radon-Nikodym Theorem 277
5.4. The Extension of Linear Functionals 278
5.5. Quasinormal Linear Spaces; the Space L * 281
5.6. Dual Spaces 281
5.7. Extreme Points 281

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
Contents x

5.8. Hilbert Space and the Spectral Theorem 287

Appendix B. Constructive Functioned Analysis for Quantum Mechanics


in SC 301
1. Introduction 301
2. Spectral Decompositions of Unitary Operators 308
3. Unbounded Operators 325
4. The Spectral Theorem 332
5. Stone’s Theorem 342
6. Fourier Transformations and Sobolev Spaces 348
7. Self-adjointness of Quantum MechanicalOperators 372
7.1. Position operators 372
7.2. Differential operators with constant coefficients 374
7.3. Hamiltonians 376
7.4. Angular momentum operators 387

Appendix C. Comparison with Related Work 389

Appendix D. Normal Forms and Conservativeness 401


1. The Existence of Normal Forms 401
2. The Church-Rosser Property 403
3. Conservativeness of P R O over P R A 405

Bibliography 415

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 1

1. Introduction

1.1. T h e B ack gro u n d . A large part of contemporary philosophy


of mathematics is shaped by a dilemma. On the one hand, there is the
apparent indispensability of classical mathematics in scientific applica­
tions, which leads to realism in the so-called Quine-Putnam indispens­
ability argument1. On the other hand, there are some epistemological
and-semantic difficulties with the naive picture of realism, as discussed
by Paul Benacerraf in two well-known papers2.
There are two major strategies that tackle the dilemma by reject­
ing the indispensability argument. The first strategy, used by Charles
Chihara, Hartry Field and Geoffrey Heilman among others3, tries to un­
dermine the argument by showing that applications of mathematics can
be achieved within some subsystems of classical mathematics that are
susceptible to non-realistic interpretations. This will be called Reduc­
tion Strategy below. Their projects are usually called nominalization
programs4. The second strategy appears in some recent publications
by Jody Azzouni, Mark Balaguer, Penelope Maddy and Elliott Sober
among others5. It argues that the success of mathematical applica­
tions does not confirm the objective existence of mathematical entities
or structures, even if it is agreed that the success of scientific theo­
ries does confirm the existence of unobservable physical entities such
as electrons. This strategy tries to accept all classical mathematics for
scientific applications and reject the full-blooded realism in mathemat­
ics at the same tim e. It will be called Non-confirmation Strategy below.

1See, for instance, Putnam [81].


2Benacerraf [9], [10].
3See, Chihara [38], Field [44], and Heilman [52].
4See John Burgess and Gideon Rosen’s recent book A Subject With No Object
[36] for an extensive evaluation of this strategy.
sSee, for example, Assouni [3], Balaguer [5], Maddy [66], and Sober [96].

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. INTRODUCTION 2

These two strategies lead to non-realistic views about mathematics and


they are mostly pursued by philosophers.
On the other hand, there are some new developments in several
schools of foundations of mathematics that first emerged in the first
quarter of the 20th century, pursued mostly by mathematicians. For
example, Solomon Feferman develops predicativism6, Stephen Simpson
tries a partial realization of Hilbert’s program7, and there are also new
developments in various schools of constructivism8. These approaches
to foundations of mathematics also accept only some subsystems of
classical mathematics that are supposed to be philosophically more ac­
countable. So they are also using the Reduction Strategy. They are also
among the alternatives to realism, although they are chiefly motivated
by Poincare’s criticism of impredicative definitions and Brouwer’s crit­
icism of classical logic, which are different from Benacerraf’s problems
about abstract entities.
This Reduction Strategy must answer three general questions: (1)
Is the accepted mathematics sufficient for scientific applications? (2)
Is the accepted mathematics really philosophically innocent? (3) Can
we reject the argument that as long as classical mathematics is a part
of our best scientific world view, a naturalistic attitude toward episte-
mology and ontology already demands the acceptance of the existence
of mathematical entities and the truth of mathematics?
There is a trade-off between (1) and (2): the more mathematics
one accepts, the easier it is to verify (1), but the harder it is to answer
(2). Approaches that follow the Reduction Strategy m ostly start with

6See, for example, Feferman [42], [43].


7Simpson [95]
’See, for example, Beeson [8], Bridges and Richman [30], and Troelstra and
van Dalen [103].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 3

some philosophical ideals that can sanction at least some substantial


mathematical concepts, and then develop mathematics in conformity
with the ideals and hope for a positive answer to (1). (3) is not very
much discussed on those approaches.
In contrast, the Non-confirmation Strategy aims at rejecting the
argument mentioned in (3) directly, without considering whether or
not (1) and (2) can be successfully pursued. If it is successful, it will
show that one can actually accept all classical mathematics for scien­
tific applications without committing oneself to realism, and therefore
nominalization programs and other approaches such as predicativism
and constructivism will be redundant in some sense. Naturally, there
are some doubts about this kind of strategy. It sounds cheap. In
particular, even if the strategy does succeed in discrediting the indis­
pensability argument, it still leaves a puzzle about why non-existent
mathematical entities and non-true mathematical theorems would be
so useful in science.

1.2. M o tiv a tio n s. This dissertation follows the Non-confirmation


Strategy, but it takes resolving the puzzle as one of its major tasks.
Some preliminary analyses will show that realism actually faces the
same kind of puzzle, and that the essence of the puzzle is about the
role of infinity in mathematical applications, which was also one of the
motivations for Hilbert’s program as Hilbert explained in his paper
‘On the infinite’ [57]. The puzzle is this: The part of the universe
to which we can have direct or indirect access is always finite, and
the universe could be totally discrete and finite. On the other hand,
infinity is everywhere in mathematics, and it seems that we will still
use infinitary mathematics in ordinary physics, in engineering, biology,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 4

social sciences and so on, even if physicists someday agree on a dis­


crete and finite model of the universe. Scientists freely apply infinitary
mathematics without any qualm about the radical difference between
the mathematical world and the physical world. Put in this way, the
puzzle becomes a general puzzle about applications of mathematics.
It becomes a problem facing all philosophies of mathematics that take
the task of accounting for the applicability of mathematics seriously.
In* this dissertation, the puzzle is simply understood as a puzzle.
Unlike the common interpretations of Poincare’s and Brouwer’s criti­
cism of classical mathematics and Benacerraf’s problems for realism, it
is not taken as a criticism of mathematics. It is not supposed to have
any normative consequences for mathematical practice, at least not
without further arguments. It is a puzzle about the well accepted and
very successful norm of mathematical practice. Instead of suggesting a
new norm, the puzzle invites some descriptive and analytic studies of
mathematical applications in science. The puzzle is to be resolved by
some detailed logical analyses of mathematical applications.
This motivates a different approach to philosophy of mathemat­
ics. It is not to criticize the foundations of classical mathematics and
to suggest new norms for mathematical practice, nor is it simply to
defend the accepted norm or to deduce realism from it. It is to ana­
lyze mathematical applications in science in order to resolve the puzzle
about the applicability of mathematics. Of course, it is also expected
that resolving the puzzle will reveal the nature of mathematics and the
function of mathematics in science.
This is perhaps similar to Benacerraf’s original intention in his 1973
paper on mathematical truth [10]. In other words, the epistemological

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 5

difficulty about abstract mathematical entities raised there was origi­


nally understood as a puzzle about mathematical language and mathe­
matical knowledge. It is to be resolved by a more careful analysis of the
semantics of the language and a more careful analysis of mathematical
knowledge. It was not meant to be an objection to the legitimacy of
the accepted norm of mathematical practice.
For the puzzle presented here, the infinity of mathematical entities,
rather than their abstractness, is emphasized, and the puzzling point is
more about the logical structure of mathematical applications, due to
the structural dissimilarity between mathematical entities and physi­
cal objects, than about the epistemological inaccessibility of abstract
entities.
This way of posing philosophical problems about mathematics may
appear to devalue the philosophy of mathematics, but this seems in­
evitable. The situation as regards classical mathematics today is dif­

ferent from the situation at the beginning of the 20th century. Clas­
sical mathematics has been successfully practiced for a century. The
question today is not which kind of mathematics we should practice,
though that question indeed bothered Hilbert eighty years ago, because
of Brouwer’s criticism and Weyl’s apparent conversion. A descriptive
and analytic approach to philosophy of mathematics is perhaps the
most appropriate one today.
On the other hand, if we take analytic and descriptive studies seri­
ously, we can actually raise some new interesting questions in philoso­
phy of mathematics. These are just the questions this dissertation will
study.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 6

The first is the question, about the scope of finitism. Here ‘finitism’
means Hilbert’s finitism9. Some analyses will show that the puzzle of
applicability does not arise for applications of finitistic mathematics.
Finitism is more restrictive than the systems of mathematics accepted
by predicativism, constructivism, or various nominalization programs.
They all accept at least potential infinity, and therefore accept some­
thing that could be essentially alien to the real physical world. Only
applications of finitistic mathematics are absolutely transparent. On
the other hand, some analyses will show that infinity is usually intro­
duced to build simplified models in mathematical applications and it
is used to approximate the finite but far more complicated physical
reality. For that reason, infinity is probably not strictly indispensable
in applications, and finitism is perhaps already essentially sufficient for

scientific applications.
This is merely an intuitive conjecture. This dissertation will try to
study this conjecture by doing some substantial technical work. It will
develop a logical framework for finitism and then examine how much of
applied mathematics can be developed within finitism. The work here
is not complete yet, for this is a very big topic, but it has been proved
that a significant part of applied mathematics can indeed be developed
within finitism. Therefore, this seems to be an interesting topic and it
deserves more attention from philosophers and logicians.
This is the first step in the attempt at resolving the puzzle of ap­
plicability. If indeed ordinary mathematical applications could all be
reduced to applications of finitistic mathematics, then we would be
able to claim that infinity only provides some efficient tools, and the

9See Hilbert [57] and Tait [98].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 7

puzzle of applicability will disappear if infinity is eliminated in math­


ematical applications. This is only the first step, for to resolve the
puzzle completely, we must further study how exactly classical mathe­
matics provides more efficient tools, and whether or not scientists im­
plicitly follow some intuitive constraints in using mathematics so that
the mathematical inferences they make are essentially finitistic. A full
study of these problems is beyond the scope of this dissertation.
Secondly, a closer analysis of applications of mathematics will moti­
vate fictionalism about mathematics. Fictionalism is the view that the
ontological status of mathematical entities is like that of fictional char­
acters in stories. A closer analysis will show that there is a borderline
between things in the mathematical world and those in the real physical
world in mathematical applications. It will show that scientists treat
mathematical entities and truths differently from physical objects and
truths. If infinity is further dispensable in mathematical applications,
then fictionalism seems to be the most natural response to the onto­

logical question about mathematical entities. On the other hand, the


analysis also shows that it is actually very casual to claim, as some re­
alists seem to do, that the mathematical entities and physical objects
that scientific theories referred to are on a par, and therefore both are
mind-independent entities accepted by scientific theories. Fictionalism
is consistent with the basic ideas of the Non-confirmation Strategy, and
some philosophers following that strategy actually support fictionalism.
Finally, the analysis of mathematical applications and the study of
fictionalism and finitism suggest a different way of posing the episte­
mological question about mathematics. According to fictionalism, pure
mathematical truths are truths within a story, and therefore there is no

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. INTRODUCTION 8

genuinely interesting epistemological question about pure mathemat­


ics. On the other hand, when mathematics is applied to reality, the
truths obtained are truths about reality. Then there is the question
about the nature of such truths. How do we know that the assertions
about real things obtained from applications of mathematics are true?

What justifies the belief that they are true?


If applications of classical mathematics can be reduced to applica­
tions of finitistic mathematics, the epistemological question will also
be reduced to the question about applications of finitistic mathemat­
ics. Some analyses will show that applications of finitistic mathematics
are actually very similar to applications of elementary arithmetic. So
the epistemological question about mathematics will be reduced to the
epistemological question about elementary arithmetic.
This implies that traditional philosophies of arithmetic, in particu­
lar those of Kant and Mill, are still of genuine interest today. In other
words, modem developments of mathematics and logic do not really

make them obsolete. The reduction of classical mathematics to finitis­


tic mathematics in applications will push the questions that traditional
philosophies are concerned with to the center of contemporary philos­
ophy of mathematics. This is another interesting research topic that a
more careful analysis of mathematical applications will bring up.
So the attempt at resolving the puzzle of applicability suggests a
two-level description of mathematics as it is actually practiced: Fiction­
alism accounts for the ontology and epistemology of pure mathemat­
ics, while the studies in the scope of finitism and the possible further
logical analyses of mathematical applications explain the applicability
of mathematics in science. The former is a phenomenal description
of pure mathematics, while the latter gives the logical mechanism of

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 9

mathematical applications. The study also reduces the epistemology


of mathematical applications to that of applications of finitistic math­
ematics, for which we can go back to Mill and Kant to look for inspi­
rations.

1 .3. S u m m a ry o f C o n te n ts. The dissertation will start with a

study of the scope of finitism. First, Chapter 1 presents a logical frame­


work called strict constructivism10. It is actually within the spirit of
finitism and is weaker than intuitionism, Bishop’s constructivism, pred-
icativism, or the systems of mathematics accepted by various nominal-
ization programs, but its language can still express ordinary mathemat­
ical propositions. It is not exactly primitive recursive arithmetic. It is a

non-essential extension of primitive recursive arithmetic, though it will


be argued that it is still finitistic for the same reason that primitive
recursive arithmetic is. In particular, every statement in strict con­
structivism has real finitistic content (in Hilbert’s sense), so strict con­
structivism is different from those conservative extensions of primitive
recursive arithmetic that contain ideal elements, namely statements
with no finitistic content in Hilbert’s sense. The name ‘strict construc­
tivism ’ is chosen to indicate the fact that it is actually a restriction of
Bishop’s constructivism in conformity with the spirit of finitism. Chap­
ter 1 will also explain how to represent sets and partial functions and
how to develop elementary calculus in strict constructivism. These will

show that when we do mathematics in strict constructivism informally,


it does not look much different from classical mathematics.
It will be proved in the dissertation that the essentials of Bishop
and Bridges’ book Constructive Analysis can be developed within strict

10The name is suggested by my advisor Professor John Burgess.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 10

constructivism, and moreover the essentials of functional analysis ap­


plied in quantum mechanics can also be developed within strict con­
structivism. These mathematical details will be put in Appendix A
and Appendix B to the dissertation.
These invite us to consider the following Conjecture of Finitism:

Nothing beyond strict constructivism, or actually finitism, is logically

indispensable in formulating current scientific theories and deducing

ordinary scientific beliefs.

Chapter 1 and the two appendices show that the conjecture is plau­
sible, though it is still a genuine open question.
There will also be a brief discussion on the connection between
strict constructivism and other approaches by Feferman, Simpson and
other nominalization programs. The relevance of Pour-El and Richards’
negative results in recursive analysis will also be discussed. These are
put in Appendix C to avoid distractions.
The formal system for representing strict constructivism had been
studied long ago by Kleene and Parsons among others, and the con­
servativeness of it over primitive recursive arithmetic has been well
known11. But the following are new in this dissertation: (1) A recon­
struction of the system, following Bishop’s interpretation of construc­
tive mathematics, can show that it is finitistic by itself; (2) A significant

“ See Kleene [65], Parsons [72], [73], and see Troelstra [101] and Avigad and
Feferman [2] for more details. The formal system for strict constructivism is essen-
#
tially the system HA% in [2]. I want to thank Professor Solomon Feferman for
informing me about some of the literature cited here.

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 11

part of Bishop’s constructive analysis can be developed in this finitis-


tic system; (3) The theory of unbounded linear operators on Hilbert
spaces can be developed within this finitistic system.
Chapter 2 takes up the philosophical issues. The chapter starts
with an exposition of fictionalism. It is the view that (1) the ontologi­

cal status of mathematical entities and structures is like that of fictional


characters in stories and a branch of mathematical theory is a story
about those fictional entities and structures expressed in a rigorous lan­
guage, and (2) fictional mathematical entities and structures are used
by scientists to build models to simulate aspects of nature in math­
ematical applications. Some of the basic ideas of the exposition are

from the literature following the Non-confirmation Strategy mentioned


above. The exposition in this dissertation tries to show that fiction­
alism is less assertive and more acceptable than realism. It squares
better with the way we actually invent, learn, and use mathematics.
And further it can incorporate good insights from other schools of phi­

losophy of mathematics, such as modalism, structuralism, formalism,


deductivism and so on. The indeterminacy of fictional entities will be
emphasized in the exposition. Possible objections to fictionalism will
be discussed, including a discussion of the objection in connection with
consistency and an analysis of the indispensability argument. The ex­
position is expected to make a convincing case for fictionalism. These
constitute Section 1 of Chapter 2.
Then in Section 2, the puzzle about the applicability of mathematics
will be characterized more carefully. It will then be explained how
mathematics within strict constructivism can be applied and why its
applications are logically clearer. The basic idea is that when we apply
finitistic mathematics, mathematical theorems about fictional entities

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. INTRODUCTION 12

can be directly translated into truths about real things, and therefore
mathematical truths can be seen as generalizations of truths about
real things (about some combinatorial features of macroscopic physical
objects). This explains the applicability of mathematics.
Finally, Section 3 of the chapter discusses the epistemological ba­
sis of elementary arithmetic. Comments on Kant’s, Mill’s and Frege’s
views on arithmetic will be given. Some reasons supporting Mill’s em­
piricism in arithmetic will be discussed and some reservation about
empiricism in arithmetic will also be considered. The reservations are
due to some Kantian considerations. It will also be argued that the
Fregean reduction of arithmetic to logic does not really help in clarify­
ing the epistemological basis of applications of elementary arithmetic.
Finally, it is suggested that the proper epistemological basis for arith­
metic is perhaps some sort of mixture of Mill’s empiricism and the
Kantian view that arithmetic is a priori because it is determined by
our sensibility. This is merely a suggestion. There will no attempt at

giving a complete theory. It is merely an indication about where the


studies in this dissertation will naturally lead.

1.4. F u rth er R esea rch T opics. The studies in this dissertation


suggest some topics for further research.
One is the status of the conjecture of finitism. It has been shown
to have some plausibility, but it is still open. Aside from its possible

philosophical implications, it seems to be an interesting topic by itself.


Another is the further logical analysis of mathematical applications
in science. For instance, can we characterize the restrictions on math­
ematical reasoning that scientists implicitly accept in order to avoid
deducing unreliable predictions from general laws due to the fact that
the laws (for example, physical laws) are only approximately true of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 13

reality? It is unclear if there could be a uniform logical characteriza­


tion, but one suggestion is that they should be essentially finitistic, or
even more restrictive.
Set theory and mathematical logic have provided good logical anal­
yses for pure mathematics, but the logical structures of science-cum-
mathematics in mathematical applications are still very unclear. Closer
logical analyses of these must be very helpful for our understanding of
the function of mathematics in science.
Finally, there is the question about the epistemological basis of
applications of finitistic mathematics, or elementary arithmetic. Some
suggestions are already given in the dissertation, but a coherent and
complete theory is still to be explored.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
CHAPTER 1

A Logical Framework for Strict Constructivism

1. In tro d u ctio n

This chapter presents a formal system for strict constructivism. It


will be shown that the system is within the spirit of finitism. Induc­
tions in strict constructivism will be discussed. Then the basics of
sets, functions and elementary calculus will be developed within strict
constructivism. In the appendices, it will be shown that the essentials
of Bishop and Bridges’ book Constructive Analysis and the basics of
the theory of unbounded linear operators on Hilbert spaces needed in
quantum mechanics can be developed within strict constructivism.
Strict constructivism is finitistic in the following sense: (i) It is
quantifier-free, though free variables of various types are allowed to
express schematic assertions; (ii) The objects it treats are numerals
and terms (and therefore concretely representable), and the functions
it accepts are all primitive recursive functions (on numerals and terms);
(iii) No abstract concepts such as constructive proofs, computable (or
primitive recursive) functionals of finite types, or finitely performable
procedures are assumed as primitive notions; (iv) It is a conservative
extension of primitive recursive arithmetic P R A , but not a proper
extension by adding ideal elements in Hilbert’s sense; it is merely a
more schematic reformulation of P R A .1

1The exact meaning of this characterisation will be dear at the end of Section 2
and the beginning of Section 3. Here I want to thank Professor Solomon Feferman
14

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
1. INTRODUCTION 15

The idea of strict constructivism is inspired by Errett Bishop’s pa­


per ‘Mathematics as a numerical language’ [14]. Strict constructivism
is the part of Bishop’s constructivism that conforms to the spirit of
finitism. Bishop did not link constructivism with finitism. The formal
system he used to formalize his constructive analysis is an extension of
Heyting’s arithmetic H A . Other studies in formalizations of Bishop’s
constructive analysis, for instance Goodman and Myhill [50], Myhill
[70], Friedman [46], and Feferman [41], all attempt to extend H A to
ease the formalization. Strict constructivism originated from an explo­
ration in the opposite direction, finding the minimum logical system
that can accommodate the essentials of Bishop’s constructive analysis.

The rest of this section is an informal introduction to strict con­


structivism.
The formal system Bishop considered in [14] is Spector’s system £2
in [97] with the axiom (F) dropped. The system W E -H A " described
in Troelstra’s introductory notes [102] to Godel’s paper is equivalent
to Bishop’s system. A related system is To in Troelstra’s notes [102],
which is obtained from W E -H A " by dropping quantifiers and the ex-
tensionality rule. (See Troelstra [101] for proof theoretical results on
these systems.) In Section 2 and Section 3 below, two systems P R O
and SC will be introduced to represent strict constructivism, where
SC is a non-essential extension of P R O .
The base system, called P R O for Prim itive Recursive Operations,
is a fragment of To, obtained by restricting the recursion operator to
numerical functions. This restriction has been studied by Kleene [65]

for the comments that bring up the question of the difference between PR O and
other conservative extensions of P R A , such as P R A +£i-induction , or WKLo in
Simpson [95].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 16

and Paxsons [72]. It is essentially the system called To in Avigad and


Feferman [2]. So the language of P R O is the quantifier-free language
of primitive recursive arithmetic P R A , extended by adding the sym­
bols of typed lambda calculus and a recursion operator on numerical
functions. The language is many sorted, just as the language of finite
type theory, but it is quantifier-free. Axioms of P R O consist of ax­
ioms of typed lambda calculus, axioms of primitive recursive arithmetic
P R A , and axioms for the recursion operator. It has a quantifier-free
induction rule as P R A does. It will be argued that P R O is finitistic
in the sense stated above. It is already known that P R O is conser­
vative over P R A , and its terms representing numerical functions can
represent only primitive recursive functions. See Kleene [65], Parsons

[72] and Avigad and Feferman [2].


P R O is essentially weaker than ordinary intuitionistic systems, for
instance, Heyting’s arithmetic H A and its extensions, or Bishop’s sys­
tem W E -H A " and its quantifier-free fragment To- Intuitionistic sys­
tems employ the primitive concept of arbitrary constructive proofs to
interpret quantifiers and implications. Similarly, the concept of com­
putable functionals o f finite types is needed to interpret To (See Godel
[48] and Troelstra [102]). Non of these systems are conservative over
P R A . It is known that the language of T 0 can represent all < eo-
recursive functions, which are strictly more extensive than the class
of primitive recursive functions (see [6] p. 568). Similarly, provably
recursive functions of H A are just <«0-recursive functions. While the
language of P R O is very close to that of To, its interpretation does not
need the notion of computable functionals of finite types. There is a
way of interpreting the language that accepts only primitive recursive
functions on numerals or terms (as strings of symbols).

of the copyright ow ner. Further reproduction prohibited w ithout p erm ission .


1. INTRODUCTION 17

Moreover, P R O is not just a conservative extension of P R A in the


sense that P R A plus E°-induction (formulated in first-order language)
is conservative over P R A , or that the system W KLo in Simpson [95]
(formulated in second-order language) is conservative over P R A . Actu­
ally, any formula of P R O with no higher type free variables is provably

(in P R O ) equivalent to a formula in P R A , and the reduced formula


in P R A can be constructed by a fixed primitive recursive function. So
P R O extends P R A only in that it allows one to prove theorems of
P R A in some more schematic forms. More concretely, it only provides
theorems with free variables of higher types, for which one can plug
in higher type closed terms to obtain theorems of P R A . That is why
P R O is finitistic in itself.
Strict constructivism conforms to Bishop’s informal interpretation
of constructive mathematics given in [14], pp 57-58, which can be
summarized as follows: (i) The objects of constructive mathematics
are finitely performable programs (with numerals as special cases of
programs); (ii) Any constructive theorem can be stated in the form
‘there exist programs y i,...,y n such that for all programs x \,...,x m,
A ( x i,...,x m, y i ,...,y n) ’, where A is a decidable predicate; (iii) The
constructive proof of the theorem must contain the constructions of
y i ,...,y n) though maybe only implicitly; (iv) The real statement to be
proved as true is the quantifier-free schema A (xl t ..., xm, tl t ..., tn) with
i i , ...,xn as free parameters, where t i , ...,tn are the constructed pro­
grams y i , ..., yn. In a word, constructive mathematicians’ work consists
of two parts: (1) constructing some programs, for instance, t \ ,..., fn
above, and (2) proving some schemata of quantifier-free, decidable as­
sertions about what are constructed, for instance, A ( x i , ..., xm, i i , ..., tn)

accordingly.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. INTRODUCTION 18

Here we give some elaboration of this interpretation. Informally


speaking, these are what constructive mathematicians are doing: (1)
constructing some programs; and (2) convincing people that the pro­
grams will do some sort of jobs. When they state a theorem as above,
they are saying that they can construct the programs t i , ..., t„ that will
do such and such jobs. Their proof of the theorem contains the pro­
grams, and they convince people of the function of the programs by
proving a decidable, quantifier-free, and schematic assertion about the

programs.
The assertion is to describe the function of the programs. It is
schematic, that is, it contains some free variables, because a program
is supposed to be something that can work in different situations. The
values of the free variables can represent specific situations, for in­
stance, specific inputs to the programs. The assertion with specific
values assigned to the free variables describes the outcomes of the
programs in a specific situation. Then, if you are convinced of the

schematic assertion, and if you want to use the programs for your spe­
cific inputs, you can plug in concrete items for the free variables and
you are convinced of what will be the real outcomes.
The assertion must be decidable, because after the free variables
are filled in, the assertion describes the outcomes of the programs in
a definite circumstance. Then the description must be either true or

false of the real outcomes, for that is a matter of fact. Notice that it is
implicitly assumed here that the termination of the programs should
never be a question.
Finally, the assertion about the programs must be quantifier-free,
because the process and the outcome of a program at a definite sit­
uation is something plainly finite and observable, and its description

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 19

should not involve quantifications over infinite domains. Another rea­


son for this is that we want to avoid both the classical and the intu­
itionistic interpretations of quantification. The classical interpretation
assumes a finished infinite domain of entities. As for the standard
intuitionistic interpretation, namely Heyting’s proof interpretation, it
resorts to the primitive concept of ‘constructive proof*. Our objective
is to investigate if these are logically indispensable in applied math­
ematics. Bishop believes that in realistic constructive mathematics
we never need such abstract concepts. We construct numerals and
programs that operate on other numerals and programs. We never re­
ally construct procedures that operate on arbitrary constructive proofs,
where it is assumed in intuitionism that they are really arbitrary, that

is, not limited to any formal system.


The phrase ‘there exist programs y i,...,y n such that for all programs
*ii —i ira’ that precedes the statement of a constructive theorem is the
only context in which we use the phrases ‘there exist’ and ‘for all’ in
stating propositions in constructive mathematics, and they should not
be confused with the classical or intuitionistic quantifiers. When con­
structive mathematicians state the theorem, they are only saying that
they can construct some programs that will do such and such
jobs. So ‘there exist’ reads simply as ‘we can’ as in ordinary language,
not ‘there objectively exists’ or ‘an ideal human agent can’. It is a
promise that they can do something or an announcement that they
have done something. And ‘for all’ is merely to indicate that x i , ..., z m
are free parameters in the assertion that follows. Since this is the
only context in which the terms ‘there exist’ and ‘for all’ are used, in
other words, since these terms are never entangled with other logical
constants ‘and’, ‘or’, ‘if....then....’ and ‘not’, the interpretation given

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 20

here is already complete. Constructive mathematicians use these terms


only to inform people about what programs they have constructed and
what quantifier-free schematic assertions about the programs they have
proved. Constructive mathematicians’ work is similar to that of engi­
neers, architects, or computer programers. They ail design something
and convince people that their products will do some jobs. They are
not interested in proving that some programs or machines (or designs
of machines) objectively exist, without giving the actual designs.
When constructive mathematicians convince people of the schema,
they present a schema of arguments so that when the free variables in
the schema of arguments are filled in, it becomes an argument for the
truth of an instance of the statement schema. This is exactly what
we do when we prove a formula in a quantifier-free system like P R O .
Infinity, even potential infinity, is not needed in order to reach the truth
of a specific instance of the schema.
Finally, notice that for Bishop the programs are arbitrary finitely
performable procedures. This is still an abstract notion, and to show
that a particular procedure expressed in the language of To is really
finitely performable, one will need inductions on quantified formulas or
transfinite inductions up to the ordinal e0. On the other hand, all pro­
cedures considered in strict constructivism are primitive recursive pro­
cedures. The termination of each procedure is recognizable within the
spirit of finitism. Then, combined with the informal description above,
we can see that strict constructivism is within the spirit of finitism.
This is the difference between strict constructivism and Bishop’s con­
structivism. A formalization of Bishop’s informal development of con­
structive analysis in strict constructivism actually translates Bishop’s

p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
1. INTRODUCTION 21

constructive analysis into primitive recursive analysis, which is in con­


trast with the recursive analysis studied by some mathematicians, for
example, Aberth [1] and Pour-El and Richards [78].
This is an informal description of strict constructivism, following
Bishop’s ideas. Now we explain how the formal systems represent this
interpretation. Return to the constructive theorem ‘there exist pro­
grams yi, ..., 2/n such that for all programs x : , ...,x m, A ( x 1} ...,x m,yx, ...,y„)’.
In the formalizations, the programs constructed, for instance, t x, ..., tn
here, will be terms in the typed language of the base system PRO,

and A (x i, ...,x TO,t i , . . . , i n) will be a formula of PRO . PRO has its


axioms and rules, just as the traditional formal systems. The proof of
A ( x i,...,x m, t i , . . . , t n) is as usual a sequence of formulas of PRO. So
PRO represents only the second part of constructive mathematicians’
work, namely proving some schematic assertions. The system SC (for
Strict Constructivism) is designed to represent both parts. All sen­
tences of SC are of the form 3y1...3y„Vx1...VxmA ( x 1, ..., xTO, yx, ..., y„),

where A ( i i , . . . , x m, y i,...,y n) isaform ulaof PRO (and hence quantifier-


free). A proof of such a sentence in SC consists of some terms t i , ..., tn
and a proof of A (x x ,...,x m,tx ,...,£ n) in PRO . In particular, the con­
structed programs tl f ..., tn constitute a part of the proof and the other

part is actually a proof in PRO. Clearly, SC represents the informal


interpretation given above faithfully.
This construction of SC from the quantifier-free system PRO is
somewhat special. In the paper [14], Bishop’s own formulation of his
formal system for formalizing constructive analysis follows the ordi­
nary approach, that is, starting with a quantified, typed language, the
language of H A ". It is not very clear how his formal system fits his
informal description of constructive mathematics.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 22

Notice that the proper logical system that represents the proposi­
tions literally proved by constructive mathematicians is the base system
P R O , while SC functions merely as a book-keeping tool, for construc­
tive mathematicians generally do more than just proving truths, they
construct programs. Therefore SC is still finitistic in spirit.
It will also be shown that Bishop’s numerical interpretation of impli­
cation (see [14] pp 58-60) can be naturally introduced into finitism as
a defined notion. Actually, some defined symbols “'*> V*, A*, —♦*, V*, 3*
will be introduced into the language of SC , and then it will be shown
that they ‘happen’ to follow the laws of intuitionistic predicate logic,
including the axiom of choice. This explains why in the informal proofs
in constructive mathematics we can employ quantifiers quite arbitrar­
ily, as long as we follow the laws of intuitionistic logic.
To state this more accurately, consider a sequence of defined for­
mulas of SC , using only the defined symbols -i*, V*, A*, —►*, V*, 3* as
logical connectives. Suppose that the sequence is a proof in intuitionis­
tic predicate logic (for the typed language and with the axiom of choice)
if these defined symbols are seen as intuitionistic logical constants. Sup­
pose that the last formula of the sequence is 3yi...3ynVxl ...VxmA ( x i , ..., xm, y i , ..., y„)
when the defined symbols are eliminated in terms of the original sym­
bols. There is a master program that operates on any such sequence of
defined formulas in SC and outputs some terms and a proof
of A ( x i ,..., x ro, t i , ..., fn) in P R O , which constitute a proof of the last
formula in SC . The master program is itself primitive recursive. These
imply that we can treat SC as if it were an intuitionistic formal system
with the defined symbols -i*, V*, A*, —►*, V*, 3* as logical connectives.
The definitions of the symbols ->*, V*, A*, -»*, V*, 3* are just Godel’s
Dialectica interpretations. They reduce any sentence in the quantified

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 23

typed language to a sentence of the form 3yi...3ynVxi...VxmA ( x i , x m, y x, ..., yn) ,


exactly what we need. In particular, the definition of —►
* includes

(3 x V y ^ [®>y] —** [u, u])

= # (3U 3YVxV v (tp [x, Y (x, u)] ( x ) ,« ])),

where <p [x, y ] , ip [u, t>] are formulas of P R O . This is just Bishop’s nu­
merical implication. Since —►
* is a defined symbol, not a primitive
logical constant, its coherence is not in doubt, as long as the primitive
logical constants in the finitistic system P R O are coherent.
Numerical implication is different from intuitionistic implication.
Some logical laws that are not valid in intuitionistic logic become
valid with the numerical interpretation of implication. These include
Markov’s principle, the independence-of-premiss principle, and the equiv­
alence of the both sides of Godel’s Dialectica interpretations. At the

end of the paper [14] Bishop considered adding Godel’s Dialectica in­
terpretation schemata to W E -H A " as new axioms. Since numerical
implication is different from the intuitionistic implication, there are
some doubts as to whether that move is coherent. See, for example,
Goodman and Myhill [50] p. 93. The presentation here is formally
different from Bishop’s. Numerical implication is introduced here by
definition. This avoids the charge of incoherence.
___
Formally speaking, the system SC is essentially the system H A "
in Avigad and Feferman [2]. Its axiomatic characterization was also
given in Troelstra [101]. However, the interpretation of the formal
system here is different. According to this interpretation, it is finitistic
in itself, not just a conservative extension of the quantifier-free P R A .

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. INTRODUCTION 24

Since the axioms and rules of intuitionistic predicate logic are avail­
able in SC , the difference between SC and other intuitionistic exten­
sions of H A lies only in the available inductions. SC does not have
the full induction schema, since the recursion operator in the language
of SC is restricted to numerical functions. However, in Section 4 it
will be shown that SC has a derived induction rule of this format:

if SC b x fa] - » <P[0, u ] ,

SC h x fa] (<Pfa, u} - n p [Sn, u ]),

then SC I- x fa] —* <Pfa, u ] ,

where x [U1 *s ^ arbitrary formula of SC with free variables u, and


tp fa, u] is a formula of the form

3m i ... 3m fcVni.. .Vn^' [m i,..., m k, n i,..., m, n, u]

where <p' is quantifier-free and m i, ...,mfc,ni, ...,n j,n are type o vari­
ables (variables ranging over natural numbers). This is a E§-induction
with assumption. In Appendix A it will actually be proved that the
inductions used in Chapter 2 through Chapter 7 of the book Construc­
tive Analysis, with only a few exceptions, can all be reduced to this
induction rule.
Section 5 explains how to represent sets and partial functions in
strict constructivism. Sets will be defined by formulas. In Bishop’s
constructive integration theory, there are some apparent quantifica­
tions over all subsets of a set. It will be shown how to avoid such
quantifications.
Then Section 6 develops elementary calculus in SC . This will show
that the informal language that fits SC is almost exactly the same as

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
2. THE FORMAL SYSTEM PR O 25

the informal language of classical mathematics. Mathematics devel­


oped in SC looks almost exactly like the ordinary mathematics. Strict
constructivism is not an exotic type of mathematics. It is a reinterpre­
tation of ordinary logical constants that assigns primitive recursively
computational contents to ordinary mathematical propositions.
The developments of other more advanced parts of Bishop’s con­
structive analysis and the developments of the theory of unbounded
linear operators on Hilbert spaces will be put in Appendix A and Ap­
pendix B to this dissertation respectively.

2. T h e form al s y ste m PRO

This section presents the base system PRO, which is essentially


the system T 0 in Avigad and Feferman [2], namely, the system To de­
fined in Troelstra [102] with recursion operators restricted to numerical
functions. It was first studied by Kleene [65]. The language of PRO

is the language of typed A-calculus, plus constants for the number 0


and all primitive recursive functions, and plus operators for recursion
and definition by case. They are summarized as follows:
typ es: o is a type, and if <rx, ..., <rn, a are types, (<Xx,..., an -* a ) is
a type.
variables: for each type <j , there are variables x f , x j, ... .
con stan ts: 0, S , and for each n > 0 and each definition of an
n-place primitive recursive function there is a constant /•*, i = 0 ,1 ,....
Term s: (1) 0 is a term of the type o, 5 is a term of the type
(o —» o), each x f is a term of the type cr, and each /•* is a term of the
type (o ,..., o —►o) (with n o’s before the arrow);
(2) if f i , ..., tn, t are terms of the types <rx,..., an, and (o x ,..., an —* a )
respectively, A p (t, tx, ..., £n) is a term of the type a ;

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. THE FORMAL SYSTEM PR O 26

(3) if t is a term of the type a, ...x £ .t is a term of the type

(o-!, ...,£Tn —» o');


(4) if t [n,m], r, s are terms of the type o, are variables
of the type o, and i , j are not free in t,r , s, then R eij (j ,r ,£ [t,j]) is a
term of the type o;
(5) if t is a term of the type o, fi, t 2 are terms of the type a, then

J (i, ti, t 2) is a term of the type a.


Form ulas: (1) if t, s are terms of the type o, t = s is a primitive
formula;
(2) if <p, ip are formulas, (tp V %p), (<p A ip), (ip —* ip) are formulas;
(3) ~«p is defined as <p —» 50 = 0.
o is called the base type and terms of the type o are called numerical
terms. For each type a its height |<x\ is defined as follows:

M = o,
1( 0-!, -♦ o')) = max flax I, |cr„|, |<r|} + 1.

and J (t, t\, t 2) are called a A-term and a /-term respec­


tively. A p (t, s i , ..., 3n) is sometimes denoted as <(si, ...,s n). Bold face
letters such as x , t , a denote sequences of variables, terms, and types.
Letters m, n, i , j and so on are reserved for variables of the base type o.
They are called numerical variables. The notion of free/bound variables
and changes of bound variables are as usual, in particular, x ^ }
in (3) and i , j in (4) of the definition of terms above are bound in
the terms. There is the convention that two terms (and hence formu­
las) differing by changes of bound variables only are considered as the
same term (and formula). We will use = as a meta-language symbol
to denote syntactical equalities. So, for example, s = t means s and

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. THE FORMAL SYSTEM P R O 27

t denote the same term ignoring the variances, t [s/x] denotes the re­
sult of substitution of terms s for all the free occurrences of x in t,
and variances for bound variables in t are presumed so that no free
variables in s would become bound after the substitutions, t [s] means
t [s/x] for some x, so when this notation is used, it is assumed that free

variables of s in the indicated occurrences of s in t [s] axe not bound by


other means in t [s]. The notion of subterms of a term is obvious and
notice that we do not count the variables right after a A or an Re as
subterms. We use t {3 } to indicate an occurrence of a subterm s of t ,
and when this notation is used in some context, t {s'} will accordingly
denote the term obtained by replacing that indicated occurrence of 3

in t by a1. In this case, free variables of 3 or s' may become bound in


the terms f { 3 } or t {s'}. Terms 0, SO, SSO, ... will be called numerals
and denoted by 0, 1) 2 and so on. On the other hand, if n is used as
a variable for natural numbers in the meta-language, we let n denote
the term SS...S0 with n occurrences of S.
Some points about the language of P R O must be especially noted.
R et; (a,r, t [t,;]) is to mean recursion. By definition, s ,r ,t here must

be numerical terms, so recursion is restricted to numerical functions.


The term J ( t ,< i,t 2) is to mean a definition by selection. See the se­
lection axioms below. Here t \ , t 2 can be of any type. This is needed,
because without recursions on the higher type terms, we cannot oth­

erwise construct definitions by cases for higher type terms, t = 3 is


a formula only if t, s are numerical terms. There are no equalities be­
tween higher type terms in the language of P R O . Also notice that
there are no quantifiers in the language of P R O .
A x io m s: (1) axiom schemes of classical propositional logic;

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
2. THE FORMAL SYSTEM PRO 28

(2) Identity Axioms: for terms i,a ,r of the type o,

t = t, t = r A s = r —>t = s,

t = 3 —* r [t/n ] = r [s/rx];

(3) Arithmetic Axioms: S t = Sr —►t = r, and the defining


axioms for every constant /";
(4) Reduction Axioms:

a { A p ( J ( t , M 2) , s ) } = s { J ( t ,A p ( t x ,s ) ,A p ( t 2,s ) ) } ,

a (A p(A x.t, r)} = a { t [r /x ]} ;

(5) Recursion Axioms:

Rex; (0 ,r ,<[»,;]) = r,

Rex; (5 a ,r, t [x',jJ) = t [a, Rex; (a,r, < [*,;])];

(6) Selection Axioms:

a { J (0 ,< i,< a)} = a { t i } ,

a { J ( 5 t , t l ,t 2)} = 4{<a};

(7) A-Axioms: when x = ( z x , z n) are not free in t

s { X x .A p ( t,x ) } = a { t } ,

a {Ax .J (t, t u t2)} = a { J (t , Ax.fx, Ax.f2) } ;

Rules: (1) (p - » ij>, (p = > ip ;


(2) Induction Rule, <p [0], <p [rx] —» ip [5rx] = > <p [t].
Notice that, according to our convention, if t' is a variant of t, then
t = t' is considered as an instance of identity axioms, for t and t' are
considered as the same term. Similar remarks apply to all other axioms.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
2. THE FORMAL SYSTEM PR O 29

Axioms (1), (2), (3) plus the rules constitute the axioms and rules of
primitive recursive arithmetic P R A . So P R O is an extension of P R A .
It is known that P R O is conservative over P R A and any closed term
of the type (o, ...,o —» o) represents a primitive recursive function. See
Kleene [65], Avigad and Feferman [2]. For our purpose of claiming
P R O as finitistic, we will need some standard results in the theory of
typed lambda calculus: the existence of normal forms relative to some
notion of reductions and the Church-Rosser property of the reduction.
These will also lead to the conservativeness of P R O and the conclusion
about functions representable in P R O . Below we will present a proof

along this line. Readers not interested in technical details can skip
these and proceed to the main theorem, Theorem 3.7 below, directly.
The proofs of the existence of normal forms and the Church-Rosser
property for ordinary lambda calculus and typed lambda calculus can
be found from [6]. Here we only verify that the proofs can be adapted to
our case with the operator J added, and that the proofs are formalizable

in P R A , that is, only primitive recursive functions and inductions on


quantifier free statements are used in the proofs.
A one-step reduction is a pair of terms of one of the following forms:

(s {A p (J (t, ti, t2) , s ) } , s { J (t , Ap (ti, s ) , Ap (<a, s))}>,

<a{Ap(Ax.t, r )> , a {* [r /x ]}).

Recall that by our convention about variances and syntactical equalities


between terms, if (t , s) is a one-step reduction and t 1, s' are variances
of t, s respectively, then (i', s') is also a one-step reduction. Terms of
the forms A p (J (t, ti, t2) , s) or Ap(Ax.£, s) are called redices and the
indicated redices above are the reduced redices in each reduction. We
will call Ap ( J (t, ti, <a), s) a J-redex and A p (Ax.t, s) a A-redex. A term

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission .
2. THE FORMAL SYSTEM PR O 30

is normal if it contains no redices. A reduction sequence is a sequence


of terms such, that each pair of adjacent terms is a one-step reduction.
We say that t is reducible to t' if we can construct a reduction sequence
t = to, ..., tn = t 1. If t' is further normal, t' is called a normal form of
t.
To prove the existence of normal forms we must construct a prim­
itive recursive function cr, called a complete reduction function, and
show (by methods formalizable in P R A ) that for an arbitrary term t,
cr (t) is a reduction sequence from t to a normal term. Here we assume
some coding for terms and sequences of terms, and then we can talk
about primitive recursive functions on terms and sequences of terms.
The construction below is a generalization of Turing’s construction,
which is originally for typed lambda calculus without /-term s. Tur­
ing’s proof was first reported by Gandy in [47]. While Gandy appealed
to a n ° induction in the proof, it can be shown that only quantifier free
inductions are needed if we proceed more carefully. To avoid techni­
calities, the details are put in Appendis D to this dissertation.
Next we need the Church-Rosser property of the reduction and the
uniqueness of normal forms. The proofs of these results for ordinary
lambda calculus can be found in [6], Section 11.1, pp 275-279. Simi­
larly, it can be shown that the proofs can be adapted to our case with
/-term s added and they are formalizable in P R A . Details are given in

Appendix D.
Then it follows that the following theorem is provable within P R A :

THEOREM 2 .1 . (Normal Form Theorem) There is a primitive re­

cursive function cr such that fo r arbitrary term t, c r ( t ) is a sequence


of one-step reductions from t to a normal term, and fo r two arbitrary

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. THE FORMAL SYSTEM PRO 31

sequences of reductions from t to normal terms, the ending terms of


the two sequences are equal.

So, every term t has a unique normal form, which we will denote
as t*. Normal terms in P R O have some special properties which the
normal terms of the stronger system To do not shaxe. As a consequence,
the definable numerical functions in P R O are only primitive recursive
functions.
We define a proper subtype of a type (<Xi,..., <rn —» <r) to be one of
(T\ , ..., crn, a or one of their proper subtypes.

LEMMA 2 .2 . I f t [ x \ ,.. . , x ^ i s a normal term of the type <r, i i , . . . , x n

are all free variables in t and their types are a \ , ..., an respectively, then
any subterm s o f t must satisfy one o f the following:
(i) s is one o f x i, ...,x n> or one o f the constants S, /■“ , or
(ii) the type of s is o, or a , or a proper subtype of one of<T\ , ..., crn, a.

PROOF. Suppose that there exist subterms of t which do not satisfy


(i) or (ii). Let s be one such with the highest type, s is not a o type
term, or a constant, or a free variable of t , or t itself, s cannot be a
bound variable of t , for in that case t has a subterm, a A-term that
binds s, which has a higher type and does not satisfy (i) or (ii) either,
a contradiction. So, 5 cannot be a variable, s cannot be a term o f the
form A p (si, r), for s i would have a higher type and could not satisfy

(i) or (ii) either. So, s must be a A-term or a J -term.


Suppose that s is J(so, *i, ■sj). We may assume that s does not occur
is a context like J (s', s, s") or J (s', s", s ) , for this larger J-term would
not satisfy (i) or (ii) either. Since s is a proper subterm of t, it has
an immediate context in t. Since s is not of the type o and a context
like J ( s ' ,s ,s u) is already ruled out, the possible cases remained to

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. THE FORMAL SYSTEM PR O 32

consider axe (1) Ax.s: ruled out, for this would be a higher type term
not satisfying (i) or (ii) either; (2) A p ( tu ..., s , ...): ruled out, for t\
would be of higher type and not satisfy (i) or (ii); (3) A p (s,r): ruled
out, for a would be a redex, contradicting the normality of t. So, in
the end, s cannot be a J-tenn.
Now, suppose that s is Ay.si and consider similarly the immediate
context of the occurrence of s. By a similar argument to that above,
we can rule out all possibilities except a context of the form J (so, a, s 2)
(or a switch of s and s2). Then the J-term J ( sq, s , s2) has the same
type as s and hence does not satisfy (i) or (ii) either, contradicting the

above conclusion.
The proof is formalizable in P R A , for all the predicates involved are
primitive recursive and the apparent uses of quantifiers are all bounded
quantifiers. |

COROLLARY 2.3. / / t [ m i , . . . , m j ] is a numerical term whose free

variables are all among the numerical variables m 1,...,m j , then the
normal form of t [m i, ...,m (] contains no occurrences o f X and all its
subterms of a type other than o are among the constants S, /" , ... .
Further we can find, primitive recursively from t, a constant f- such
that

P R O H t [mi, ...,TO|] = //( m i,...,m i) .

PROOF. The first part is obvious by the lemma. Then the normal

form t [m i,...,mj]* is a term built from m i , ...,m i,0 ,5 , /•*,... with the
operators J, Re, and Ap only. It is easy to see that such a term can
be interpreted as a definition of a primitive recursive function with
arguments m i, ...,m j, and we can find, primitive recursively from t, a
constant f[ corresponding to that definition. Since normal forms can

o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


2. THE FORMAL SYSTEM P R O 33

be computed primitive recursively and P R O has the reduction axioms,


we can find a proof of t [mi, ...,mj] = t [ m i , m / ] * in P R O primitive
recursively from t. Clearly, a proof of t [mi, ...,m/]* = f- (m x, ...,m j) in
P R O can also be found primitive recursively from £*, by induction on
the structure of t*. |

From the corollary we see that for any closed term t of the type o,
there is a numeral n such that P R O b t = n, though this cannot be
proved within P R A , for the proof will involve a diagonalization over
all definitions of primitive recursive functions. If t is a closed term of
the type ( o ,..., o —►o ) , for a sequence of numerical variables m , t (m )
will satisfy the conditions and hence we can find, primitive recursively
from t, a constant /" and prove

P R O h t(m ) = /•* ( m ) .

So t represents a primitive recursive function.


More generally, any closed term s of the type (<7i,..., an —> a) can
be interpreted as a program, which, with closed normal terms <1,..., tn
as inputs, outputs the normal form of A p(s, <1 ,..., tn). By the theorem
above, such programs are primitive recursive as functions of terms.
Since it is known that definable numerical functions in the language
of To are exactly the < e 0-recursive functions (cf. [6] Appendix A.2),
the language of P R O is strictly less powerful.
Now consider the conservativeness of P R O over P R A . This can
be proved as follows. Consider a proof of a formula of P R A in P R O .
We first instantiate all free variables of types higher than o by some
closed terms. Then we reduce all the terms in the proof to normal
forms. This will result in a proof containing no A-terms. Then we can
eliminate Re-terms and J-terms by assigning a constant /* to each

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. THE FORMAL SYSTEM PRO 34

numerical term t[7ii, ...,7i*] with free variables all in 7ii,...,7i* so that
fi (n i i n * ) = t is a defining axiom and then replacing the term t by
f t (ni> ..., n*). This will result in a proof in P R A . Then we have

THEOREM 2 .4 . P R O is conservative over P R A . More specifi­

cally, there is a prim itive recursive function that transforms a proof


of a formula of P R A in P R O to a proof in P R A .

PROOF. See Appendix D for details. |

Before closing this section, we give some final comments about


P R O . The standard interpretation of the language of To is Godel’s
computable functional interpretation. If the language of P R O is in­
terpreted in the same way, it will be obviously beyond the scope of
finitism. To understand P R O in the spirit of finitism, we must take
a syntactic point of view. It means that we do not consider terms of
P R O as names denoting other objects such as functions or function­
als, for in that case they have to be some abstract objects (no matter
whether they are intensional or extensional). Terms themselves should
be the objects of our concern.
First, a closed normal term of the type o can be translated into an
ordinary computer program (with no inputs needed) rather straightfor­
wardly, for such a term contains only constants for primitive recursive
functions and operators J, Re and Ap. In particular, it contains no A-
terms. (A A-term has no immediate counterparts in ordinary computer
programing languages.) We will call these programs basic programs, or
executable programs, for each can be executed to output a numeral.
By the theorems we proved, there is a master program that is itself
a primitive recursive function on terms as strings of symbols and that
produces the normal form of any term of any type. We can take an

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. THE FORMAL SYSTEM PRO 35

arbitrary closed term of the type o as a proto-program. It cannot be


executed directly, but it can be transformed into an executable pro­
gram by the master program. Any formula of P R O is equivalent to an
equality t = a, where t, 3 are of the type o. So a sentence (closed for­
mula) of P R O can be transformed into an equality between two closed
normal terms of the type o, which asserts that two basic programs out­
put the same numeral when executed. After we prove a sentence t = s
in P R O , we can assert that t and s are equivalent proto-programs.
Then an arbitrary closed term t can be seen as a program trans­
lator. For any closed terms s of some appropriate types, t (s) will
also be a closed term, which is a proto-program or a program transla­
tor. For closed terms Si, ...,s n of some appropriate types, t ( s i ) ... (sn)
will eventually become a closed term of the type o, namely, a proto-
program. When we prove an open formula t [x] = s [x] in P R O , we
prove that two program translators Ax .t and Ax.3 generate equivalent
proto-programs for the same inputs (which can be numerals, proto-
programs, or other program translators). In that sense, we can assert
that the two translators are equivalent.
It is easy to see that these can all be realized on computers. Then
the master program will be something like a compiler and it is primitive
recursive in itself. All programs subjected to direct executions are also
primitive recursive.
Program translators are needed to express more general construc­
tions in constructive mathematics. In applications, a rational number
is coded as a numeral; a real number is represented as sequence of nu­
merals, namely, a translator that translates a numeral (as input) into a
proto-program that outputs a numeral interpreted as a rational number
approximating to the real number; and a real function is represented by

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. THE FORMAL SYSTEM PRO 36

a translator that, with translators representing real numbers as inputs,


will output translators representing real numbers. So, for instance, the
function sin is represented by a term that is a program translator.
Notice that the language of To can also be interpreted in a similar
way, but in that case, either the master program is not < eo-recursive,
or the executable programs should include all < eo-recursive programs.
Note that the definition of normal form of terms of P R O here is slightly
different from the ordinary definition. A normalization process does
not expand recursions. Recursion operators can appear in the normal

form. That is why the master program is primitive recursive. When we


consider To, the normalization can only be obtained by a function that
majorizes all < eo-recursive functions. On the other hand, if we take
non-normalized terms as executable programs, executable programs
should essentially contain all < eo-recursive programs.
To prove the termination of < eo-recursive programs, one must re­
sort to either transfinite inductions up to the ordinal e0> or inductions
on quantified formulas. Infinity or the intuitionistic interpretation of
quantifiers is implicitly assumed. That is how To differs from P R O
and why it is beyond the scope of finitism.
Finally, let’s recapitulate the claim that P R O is finitistic. It has
no quantifiers. Generality is obtained by using free variables and prov­
ing schematic assertions. In the intended interpretation, entities are
ordinary computer programs. They are finite and physically realizable.
These programs are all primitive recursive. We can convince ourselves
of their terminations by tracking the constructions of the programs.
Counting the number of iterations of loops and sub-procedures, we can
also have an estimation of the length of computation. There is no need

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
2. THE FORMAL SYSTEM PR O 37

for a proof of termination by principles alien to finitism, for exam­


ple, inductions on formulas with quantifiers, no matter whether they
are classical quantifiers or intuitionistic quantifiers. No vague notions
such as ‘constructive proof’, ‘computable functionals of finite types’,
or ‘finitely performable procedures’ are assumed.
P R O is conservative over primitive recursive arithmetic P R A , but
it is more than that. All sentences in P R O are meaningful. Each
sentence is actually equivalent to a sentence of P R A obtained by ap­
plying the master program, and it eventually asserts that two primitive
recursive (executable) programs will output the same numeral when ex­
ecuted. That is the only kind of assertion with definite meanings we
can make in the language of P R O . P R O is an extension of P R A only
in the sense that in P R O one can prove theorems of P R A in some
more general schematic forms, namely, theorems with free variables
for higher type terms (not just numerals) such that when all higher
type variables are instantiated by arbitrary closed terms of appropri­
ate types, we actually obtain theorems of P R A . (This follows from
Theorem 3.7.) So P R O is not a conservative extension of P R A by
adding some ideal, non-meaningful elements in Hilbert’s sense. It is
merely a more general, more schematic way of formulating P R A .
Here are some more concrete examples. There are no variables for

functions in P R A , but there is no harm in stating theorems of P R A


in a schematic form such as

/(m , n) + g(m , n) = g (m , n) + /( m , n),

leaving f , g as variables. When they are substituted by constants for


primitive recursive functions in P R A , a theorem o f P R A is obtained.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM SC 38

Then we can consider similarly

F (/> 9) (« 0 + G ( / , g ) (m ) = G ( / , g) (m) + F ( / , g) (m ) ,

where F, G are also free variables to be substituted by A-terms of ap­


propriate types. After the substitutions, A-symbols can be eliminated
and we still obtain a theorem of P R A .
It is in this sense that P R O only provides a more schematic way
of proving theorems of P R A . Hilbert’s finitism bans quantifiers but
allows schematic assertions. In this sense, we claim that P R O is itself
finitistic, and it is essentially different from other conservative exten­
sions of P R A such as PRA(form ulated in a quantified language)-}-£°-
induction , or the system W K Lo in Simpson [95].

3. T h e s y ste m SC

Now we introduce the system SC . This is a system of notations we


use to represent what is constructed and what is proved in constructive
mathematics. The basic ingredients of the notational system SC are
its symbols, terms, formulas, and proofs. The symbols of SC are those
of P R O plus 3 and V. Terms of SC are just those of P R O . Formulas
of SC are

3x i ... 3z„Vyi... ,

where tp is a formula of P R O , m > 0 ,n > 0, ® i,..., xn, ..., ym are


different variables. We also write 3®i...3xn as 3x and write Vyi...Vym
as Vy. ip may contain free variables other than ®i, •••, xn, yi, ym-
They are the free variables o f the formula 3xVy<p, and if 3xVy<p has
no free variables we call it a sentence (closed formula), x, y are bound
variables in BxVy^. As before, we treat variances of a formula as the
same formula, x, y can be empty. If both are empty, the formula is just

p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. THE SYSTEM SC 39

a formula of P R O . A proof of a sentence 3xVyy> in SC consists of a


sequence of closed terms t with the same types as x and a sequence $
of formulas of P R O such that $ is a proof of tp [t/x] in P R O . Notice
that terms t are parts of the proof and tp [t/x] generally contains free
variables y . We call sentence 3xV ytp a theorem of SC , denoted by
SCI- 3xVyy>, if we can construct such a proof.
Clearly, SC is not a formal system in the traditional sense. SC
represents constructive mathematics in the following sense. First, the
theorems in constructive mathematics, for example, the fundamental
theorem of algebra, spectral theorems and so on, can be translated
into sentences of SC . A sentence 3xVy<p of SC is what Bishop called
an ‘incomplete proposition’ in [14]. It states that some programs, the
terms t, can be constructed and the complete proposition tp [t/x] can
be proved. Second, the informal proofs of the theorems in constructive
mathematics can be translated into proofs in SC . This implies that the
programs which the incomplete proposition claims to be constructive
are always implicitly contained in the informal proofs and the formal­
izations in SC make them explicit. The two points together represent
Bishop’s idea that constructive mathematics consists of constructing
programs and proving deddable assertions about the programs.
Moreover, this realization of Bishop’s interpretation of constructive
mathematics is finitistic. The programs constructed are closed terms
of P R O and the assertions about the programs are formulas of P R O
and they are proved in P R O . So, whenever we prove a theorem in SC ,
we actually prove a stronger formula in P R O . It has been indicated
that P R O is finitistic. Therefore, SC is also finitistic.
It needs to be emphasized that for the sentences of SC , we never put
the combination 3xVy in a context other than 3xV y^. The informal

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM SC 40

interpretation of 3xV y<p is ‘terms t can be constructed and tp [t/x]


can be proved in P R O ’. The modal word ‘can’ here is used in the
most innocent way. It is never entangled with quantifiers, implications,
or other modal words, ‘can be’ really means ‘can be’, for a proof
must explicitly contain the constructions. It is not merely a proof

that something objectively exists or can be constructed by some ideal


human agent.
Now we introduce the defined symbols -i*, V*, A*, —>*, 3*, and
V*. Suppose that tp = HxVytpi [x, y], i/> = 3uV v0i[u, v], are formulas
of SC, where tpi, ip\ are formulas of PRO, x,y., u ,v axe different
variables. Let (we will also use X , Y , U for variables):
(1) (ip A* ip) = 4 3xuVyv(y>1 A ipx)\
(2) (tp V* ip) =df (<pi V ip\) if x , y , u, v are all empty, otherwise,

(<p V* ip) = # 3n xu V yv((n = 0 A < p ) V ( n / 0 A ip) ) ;

(3) (3*z<p) 3zxVy<pi if z does not occur in x ,y , otherwise

(3 'zip) =df <p\


(4) (V*z<p) = $ 3XVzyv?i [ X ( z ) , y] if z does not occur in x , y , oth­

erwise (V*z<p) =df tp\


(5) (,tp —• ii) = 4/ 3UYVxv(v>1 [x, Y (x, v)] lii (U ( x ) , v]);
(6) ( - » = 4 , (*> - • SO = 0) = 3Y V xH >i [x, Y (y)]).
Literally, ->*, V*, A*, —►*, 3*, and V* are symbols in the meta­
language. We only use them to facilitate our references to formulas of
SC . However, as we will use them so heavily, sometimes we talk as if
they were logical constants in the language of SC . For example, we will
call <pA*ip a conjunction and call <p,ip its components, though, literally,
tp A* ip is the formula 3xuVyv(<px A ip\) according to the definition.
No real ambiguities will arise. Notice that when p , ip are formulas of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM SC 41

P R O , (<p v* tp) = ((p v t p) , {tp a* tp) = {tp a VO, (<p i>) = (<p -* i>) ■
Similarly, if tp, tp are formulas of SC and 3ztp, 'iztp axe still formulas
of SC , we have 3“ztp = 3ztp and V'zip = Vzifr. So we can drop all the
stars without generating real ambiguities. Further we define

{tp <-►tp) = jf {tp -» tp) A {tp - » <p) •

Definition (5) is the definition of numerical implication. See [14].


We want to translate the assertion ‘If tp, then tp1 in informal construc­
tive mathematics into tp —»* tp. To justify this, one has to show that
in the informal language we do mean numerical implications when we
use ‘if... then...’. One way to see this is to examine how we prove the
conditional statement ‘if tp, then tp' in informal contexts, where tp,tp
are as in the above definitions. Bishop gave some explanations in [14].
Here we present Bishop’s main points and add some more discussion.
In the informed proof we usually take tp as an assumption and de­
rive tp. Consider the typical way of proving 3uVvV>i[u, v] from the
assumption 3xVy<pi [x, y]. One must take an arbitrary x and derive
3uVvi/’i [u, v] from Vy<px [x, y ] . That means one must construct u and
derive Vv^i (u, v] from Vy<px[x,y]. Now, by the intuitionistic inter­
pretation of implication, u can depend on any hypothetical proof of
Vytpi [x, y ] , besides x. However, in realistic constructive mathemat­
ics we never construct operations that operate on proofs of our as­
sumptions. So Bishop conjectured that we always simply construct

an operation U that operates on x and derive V v^i [U ( x ) , v] from


'iytpi [x,y]. This is the first major difference between the two interpre­
tations. Bishop’s judgment certainly comes from his experience with
constructive mathematics. We can agree that it is reliable to some

extent.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM SC 42

Next, for an arbitrary v , one has to derive fa [U ( x ) , v] from V y ^ [x, y].


According to the intuitionistic interpretation, a proof of Vyi^ [x, y] —*
fa [U ( x ) , v] can operate on any hypothetical proof of Vyipx [x, y] to
generate a proof of fa [U ( x ) , v j . Similarly, in realistic mathematics
we never seem to construct such a general operation. In particular, it
is totally unclear how to operate on a proof if the proof is not already
formalized. Then here is another point of Bishop’s in justifying the nu­
merical interpretation of implication. Bishop conjectured that the only
way to use the universal statement Vyipi [x, y] as an assumption is to
construct some sequences of terms fy [x, v ] ,..., tn [x, v], which may con­
tain x , v as free variables, and derive fa [U ( i ) , v] from <p\ [x, tj [x, v]],

..., Vi [x, tn [x, v]J, that is, prove

(1) <px [x, t : [x, v]] A ... A [x, t n [x, v]] -♦ fa [U ( x ) , v]

in some quantifier-free logic, which we take as P R O . For each tpx [x, t, [x, v]],
we can construct its representing term a* [x, v] = t Vl[x,t.j, that is, a term
such that PROH <p\ [x, tt] *-* a,- [x, v] = 0. Let

t [x, v] = J [ai, J [a2, ... J [an, u, t n] ,..., t 2] , t x] .

Then it is easy to show that

P R O I- [x, tjA .-.A ^ i [x, t ,] A ^ i [x, t»+i] {<Pi [x, t,-+1] tpx [x, t]) .

This implies that we can prove in P R O the formula <px [x ,t [x, v]] —*■

ipi [U ( x ) , v]. So, we finally have

(2) BUYVxv (<pi [x, Y (x, v)] -♦ fa [U ( x ) , v ] ) .

Moreover, we can actually allow a format more general than (1).


Suppose that we have terms t [n, x , v] and another term a [x, v] and we

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM S C 43

can prove in P R O

Vn < s [x, v] {<px [x, t [n, x, v]]) -* V»i [U ( x ) , v ] ,

where the bounded quantifier is defined as in the primitive recursive


arithmetic. For any quantifier-free p [n,x], let

p[m ,x] = 1 - R e ij (m, 1 - tv[0iX], **p+i,x])) .

where t v is the representing term of p \ then Vn < rrup [n, x] is defined as


p [m, x] = 0. Then, using Re we can construct a term d [x, v] such that
it is intuitively the minimum m < s [x, v] for which ~'<pi [x, t [m, x, v]]
if it exists, otherwise it is just s [x, v]. Then it is easy to show that

P R O b tpx [x, t [d [x, v ] , x, vjj -» V»i [U ( x ) , v ] ,

and hence we still have (2).


So the conjecture is that the only way to make use of a univer­
sal statement as an assumption in constructive proofs is to use a fi­
nite number of instances, and the number of instances is bounded
by a different term we can construct. Then, when we want to prove
Vy<pi [x, y] —> ipi [U ( x ) , v ] , we can always do it by constructing an
instance p i [x, t [x, v]] of Vyp x [x, y] and deriving ipi [U ( x ) , v] from
<Pi [x, t [x, v ]]. This is the second major difference between the two in­
terpretations of implication. We can also agree that it is justified to
some extent by Bishop’s experience.
From the explanations it seems that for a sentence like <p —>
the numerical interpretation conforms with the way we actually prove
such a statement better than the intuitionistic interpretation. Suppose
that we put the numerical interpretation of —►in tp —» if) in a way
similar to the proof interpretation in intuitionism, that is, a proof of
(p —» $ is a construct that includes operations U , Y and a proof of

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. THE SYSTEM SC 44

tpi [x, Y (x, v)] —» ip\ [U ( x ) , v ] . Then any formula will be equivalent to
a formula of the form 3xVy<pi [x, y] when this interpretation is applied
thoroughly, and hence the interpretation of —> in contexts like tp —* ip
will already be sufficient for determining the meaning of —►
. So it
seems that the numerical interpretation is at least as legitimate as the

intuitionistic interpretation.
Certainly, such arguments cannot be very conclusive. The prob­
lem of how meanings of logical constants can be coherently fixed is too
complicated for us to discuss here. In our presentation, —»* is merely
a defined symbol. Its legitimacy is no problem. If there a question, it
must be something like this. In translating informal statements into
formulas of SC, we simply translate ‘if..., then...’ into the defined nu­
merical implication —►
* . One may suspect that by such translations
we weaken the propositions as they are originally understood in intu-
itionism. If that is the case, the theorem proved in SC may only be a
weaker version of the original theorem understood in intuitionism, and
we canno t claim that we have formalized constructive mathematics in

SC .
Notice that, from the intuitionistic point of view, the right hand
side of the definition of —>* is stronger than the left hand side. So,
if a theorem takes the form tpi A* ... A* tpm —►
* ip, where <pi, ..., tpm
are of the structure as tp, then this is not a problem, for actually the
interpretation in SC is stronger than the original when both are un­
derstood intuitionisticaly. The problem may arise when the theorem
is of a structure like (<pi —♦* <pj) ip. The numerical interpretation
interprets (tp\ —** as a formula that is stronger when understood
intuitionistically, so the interpretation of (<pi —♦* tp2) —»* ip might be­
come weaker. However, if we do always prove a theorem like tp\ —»* tp2

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM S C 45

in the way described above, that is, in conformity with the numerical
interpretation of —►*, then this is still not a real problem. For exam­
ple, if (tfx —** —** fp is applied to derive tp from <pi —>* tpi, then,
as the proof of tpx —►
* <p2 actually proves its numerical interpretation,
the numerical interpretation of (tpx —** <p3) —** tp is just what is really
needed.
On the other hand, this is not a problem peculiar to the numerical
interpretation. There exists the very same problem when we translate
classical theorems into theorems in intuitionism. Since in interpreting
Bishop’s constructive analysis in SC our real concern is the sufficiency
of finitism for mathematical applications in science, we do not worry if
SC is a faithful representation of some preconceived constructivism or
intuitionism. As for the sufficiency question, we will focus on presenting
mathematics in SC.
There is indeed the other side of the question: can those applica­
tions of classical mathematics in various circumstances also be trans­
lated into applications of mathematics in SC? In other words, if appli­
cations themselves demand the classical (or intuitionistic) interpreta­
tions of the theorems applied, then even if we can prove in SC some
theorems that look similar, we still cannot claim the SC is sufficient
for the applications. Generally, we believe that as long as the math­
ematical theorems axe provable in S C , this will not be a problem by
itself. More discussion on this point is given at the end of Appendix C.
Now we can show that symbols -i*, V*, A*, —>*, 3*, and V* behave as
if they were logical constants in intuitionistic logic. First, for each for­
mula <p of P R O we can find a term t v of P R O , with the same free vari­
ables, such that P R O b tp *-* t v = 0; cf. [101], p. 45. If p [xi, ...,x m]
is a formula of SC where X i,..., xm are all its free variables, a proof of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM SC 46

tp [ i i , x m] in SC is a proof of tixe sentence V*x1...V*xm<p[xi,...,xm]

in SC , where the latter has been defined above since it is closed. Simi­
larly, we call tp [ x i,..., xTO] a theorem of SC , denoted SCI- tp [x ^ ..., xm] ,
if we can construct such a proof. Then it can be shown that Godel’s
system of axioms and rules for intuitionistic predicate logic hold for the
symbols V *, A *, 3*, and V* (we keep the stars to emphasize
their real characters)

THEOREM 3 .1 . For any formulas tp, tp, x> and term. t of S C ,

(1) the following formulas are theorems of SC :

0 = SO —►
* tp, t p M * tp —>* tp, tp —>* tp A * <p,

tp —►* tp V * iff, tp A * ip —►* t p , tp V * tp —»* “tp V * tp,

tp A * tp —►* tp A * tp, V*xtp [x] —»* tp [t], tp [t] —►* 3*x^ [x] j

(2) S C r tp, S C r tp —>* tp =>• S C r tp,


S C r tp —►
* tp, S C r tp —>* x => S C r tp —►
* x>
S C r tp A * tp —¥* x ^ S O - tp —♦* (tp -»* x)i

S C - tp (tp —** x ) =► SC ~ tp A * tp —>* X i


S O - tp —»* tp =s>S C r x V* tp - * * x V* tp,
S O - tp —** tp => SC ~ tp — V*xtp, x not free in tp,
S O - tp —** tp =>■SO~ 3*xtp —>* tp , x not free in tp.
Further, in (1), the proofs of the formulas in S C are constructed by
some prim itive recursive functions o f tp ,tp ,x ,t and so on, and in ( 2 ),
the proofs o f the formulas at the right side o f => are also constructed
by prim itive recursive functions from the proofs of the formulas at the
left side of=>, as well as the formulas tp,tp themselves.

PROOF. The proof is exactly the same as the proof of the soundness

of Godel’s D ialectica interpretation in [101], pp 234-236 (disregarding

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. THE SYSTEM SC 47

the proof for induction). Notice that definition by cases is obtained


by the operator J of the language of P R O and the existence of terms
representing quantifier free formulas was remarked above. |

Now, given a sequence of formulas of SC such that each one is


either one of the formulas in (1) of the theorem, or obtained from the
previous formulas by one of the rules in (2), we can certainly construct,
primitive recursively, a proof in SC of the last formula in the sequence.
So, to construct proofs of formulas of SC , we can proceed as if SC
were an intuitionistic logical system and the symbols V*, A*,
3*, and V* were logical constants of the intuitionistic logic; that is, we
can construct a sequence of formulas which is an intuitionistic proof in
the intuitionistic formal system, and then a master program, which is
primitive recursive, will automatically transform the proof into a proof

in SC .
Clearly, the fact that —►
* is defined as numerical implication is cru­
cial for Theorem 3.1 to hold. On the other hand, the classical axiom
3 * ip [x] V* [x] does not have a general proof in SC even if p [x]
is quantifier free and x is numerical, for by definition a proof would
contain closed terms t , s of the type o and a proof o f {tp [s] A t = 0) V
( - 'p [x] A t > 0 ). Then by deciding if t = 0 or t > 0 we would be able
to prove either p [s] or ->p [x] in P R O , contradicting Godel’s incom­
pleteness theorem.
Suppose that $ is a collection of some finitely many formulas (or
schemes of formulas) of SC . We use the notation S C ,$ hj p to mean
that we have a sequence of derivations of ^ from the theorems of SC
and the formulas in $ (or instances of the schemes in $ ) with the
intuitionistic logical rules as in Theorem 3.1, and no generalizations
are applied to free variables of $ . Then we have

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. THE SYSTEM SC 48

COROLLARY 3 .2. If S C ,$ b / tp, then for some tpi,...,tpn in

S C b (pi - ♦ * - V tpn -»* tp.

Further, the proof in S C can be constructed primitive recursively from


the derivations and the proofs (in S C ) of the theorems o f S C used.

This corollary represents the general style of informal constructive

proofs. That is, to prove a theorem of the form tpx —»*.....—►


* <pn —♦* tp,
we take tpx,...,tp n and other theorems already proved as assumptions
and use the rules of intuitionistic logic to derive tp. Clearly, in deriving
tp, we can further use temporary assumptions to derive conditional for­
mulas. That is, the deduction theorem generally holds for derivations
with intuitionistic rules. But notice that, as SC is actually not a logical
system in the traditional sense, the notation ‘SC,<p b tp’ does not have
a natural meaning. That is why we use b / .
Some other intuitionistic laws and some features of intuitionistic
systems are also shared by SC . For example, the axiom o f choice is
trivially provable:

(AC) SC b Vx3y<p [x, y] -> lY V x tp [x, Y (x ) ],

for, as can be checked, the antecedent and the consequent denote the
same formula in SC . The existential property is also trivial by the
definition of the theorems of SC . The disjunction property directly
follows from the definition of the symbol V* and the theorems of SC
and the assertion that for any closed term t of the type o, we have
either P R O b t = 0 or P R O b t / 0. This is obvious though a general
proof involves the evaluation of all closed terms of type o, which is not
formalizable in P R O .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 49

On the other hand some formula schemes not generally valid in


the intuitionistic logic become provable in S C , because the meaning
of —» in SC is actually different from that in intuitionistic logic. For
instance, similar to the proof of Lemma 3.5.7 and Theorem 3.5.10 of
[101], pp 238-240, we can prove that

(M) SC h Vx(y? V -iyj) A -'-’Bxip —» 3x<(0,

(IP)
SC H Vx (v? V ~>tp) A (Vxtp -> 3yV>) -» 3y (Vxv? - » t p).

M is the generalized M arkov’s principle. IP is called independence-


of-premiss schemata. Further, if tp is a notation for a formula of SC
using the symbols V*, A* and so on, and <p denotes the formula tp of SC,
then the formula tp «-»* tp is provable in SC , for it is actually tp *-** tp.
However, if both <p and tp are read as formulas of intuitionistic logic by
themselves this equivalence generally will not hold.
_
This implies that SC is formally equivalent to the system H A " in
Avigad and Feferman [2] when the symbols V*, A* and so on are taken
as primitive symbols.

4. In d u ctio n s and In d u ctiv e C o n stru ctio n s in SC

Besides the intuitionistic logical axioms and rules discussed above,


in constructive analysis we still use inductive constructions to construct
sequences and use induction rules to prove assertions. Since recursions
are restricted to numerical terms in the language of P R O and P R O
has only quantifier-free induction rules, we cannot construct sequences
of higher type entities by inductions directly, and not all induction
rules are available in SC . However, we can reduce some higher type

p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 50

inductive constructions to primitive recursions and we can prove some


induction rules for quantified formulas in SC.
First let’s see some numerical recursions reducible to simple primi­
tive recursions. First we have recursions with shifted secondary numer­
ical arguments: Given terms r[m ], t [i,^ , m] of the type o and
some sequences of terms Si [n, m ] ,..., s*. [n, mj, each, of the same type
as m , we can construct a term q [n, m] such that

P R O b q [0, m] = r [m ],

P R O I- q (Sn, m] = t [n, q [n, Si [n, m ]], ..., q [n, sfc [n, m ]], m ].

This is obvious. Notice that q [n, m] may contain free variables other
than n, m , if r or t contains other free variables.
Simultaneous recursions are also reducible: Given terms [m],
h = 1 we can construct terms ^ [ n ,m ] such
that, for h — 1 ,..., /,

P R O b qh [0, m] = r h [m ],

P R O b qh [Sn, m] = t/» [n, [n, m ],..., qt [n, m ] , m]

We can also combine these two: Given o type terms tf, [i, j, mfc],
r/ifma], h = 1 where j = ( ii, and sequences of terms
Ph,k [nih] , h ,k = 1 ,..., I, each being of the same type as m*, we can
construct terms [n, m^] such that, for h = 1,..., /,

P R O b qh [0, in/,] = r h [m *],

P R O b qh [Sn, m*] = t h [n, qx [n, pA,i [mfc]], ..., qi [n, p h,i [mh]], m k] .

Further, there can be several occurrences of terms like q&[n ,...... ]


in the last equation and their occurrences can be iterated as in this
example: Given o type terms r[m ], and sequences of terms

with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 51

s [71,111], p [n ,j], each with the type of m, we can construct a term


q [n, m] such that

g[0,m ] = r [ m ] ,

q [ S n t m] = t[n, q[n, p [n, q [n, s [n, m]]]]].

To see why, let’s consider a simplified format. We can construct a term


q [n, m] such that

9 [0,m] = r [m],

q[ Sn, m] = q[n, q[n, m] ],

for, intuitively, q [72,771] is just r [r [ r [m].....]] with 2n iterations. This


can be constructed by simple primitive recursions. Clearly, there can
be more than two iterations of q's in the recursion equation. The
important thing is that q is iterated a fixed number of times. So, for
example, the recursion pattern

q [Sn,m] = Rezj (m ,r, q [n,;])

is not reducible, for it intuitively implies

q [Sn, 771] = q [ n , q [ n , q[n, r } ]]

with 771 iterations of q [77, ...]’s. Then, when we try to calculate the
right hand side, we find that we have to iterate 9 [n —1, ...] as many
as q [ n , q [n, r] .....] times (with tti —1 iterations of q [n, ...]’s). This
is beyond primitive recursion. (This is the recursion pattern for acker-
mann’s function.).
Notice that secondary arguments m, m* above must all be numer­
ical variables, for otherwise spelling out the recursion equations would
result in indefinite iterations of higher type terms, which can be ob­
tained only by higher type recursions.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 52

In formalizing Bishop’s constructive analysis, recursions are mostly


used to construct sequences. A sequence of type a objects is a term of
the type (o —* a). We also call a term T [n] of the type a a sequence,
by which we mean An. T [n]. Constructions of sequences are frequently
followed by inductions, proving that the sequences have some desired
properties. Applications of induction rules are also frequently justified
by constructing the relevant sequences.
To apply the available numerical recursion patterns to construct
sequences of higher type objects, we generally proceed this way. For a
term T of any type, T ( x i ) ... (x*) becomes a numerical term for some I
and some variables X i,..., xj of appropriate types. Instead of construct­
ing a term T [n] to satisfy some recursion equations directly, we con­
struct a numerical term q [n, x x, ..., x/j, which is to be T [n] (x x) ... (x j),
to satisfy some appropriate recursion equations, and then let T [n] be
Axx Ax t.q[n, x x,...,X i ] .
For this strategy to work, we usually need some so-called extension-
ality conditions. Recall that equalities between two higher type terms
are not defined in the language of P R O or SC . However, we can define
various kinds of equality, depending on the needs. The simplest one
is extensional equality, denoted by ~ . For two terms T, R of the same
type, T (x x) ... (xj) becomes a type o term for any variables x x, ...,X ; of
appropriate type and we denote

(T ~ R) = Vxx...xt (T ( x x)...( x i) = R (x x) ...( x i) ) ,

where x x, ..., Xj are not free in T , R . A related notion is preserving


extensional equality: We say that a term t [x] preserves extensional
equality for z if we can prove the extensionality rule for t [z] with respect

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN S C 53

to x :

SC h Vx, y (x ~ y —►t [x] ~£ [y]).

Similarly, we say that a formula tp [x] preserves extensional equality for


x if we can prove

SC h V x ,y (x ~ y (<p [x] y>[j/])).

We say that a sequence of terms t [x], or a formula tp [x], preserves


extensional equality in the sequence of variables x, if each term in
t [x], or tp [x], preserves extensions! equality for each of x.
As an example of the use of these extensionality notions and of

reducing of higher inductive constructions we will discuss the coding


of finite sequences. First, we can code a sequence of terms £1,..., U of
the type o into a single term (£x,..., ti) of the same type by a primitive
recursive function and the decoding function is also primitive recursive.
For terms t\, ...,ti of some higher type a, the coding is available only
with respect to extensional equality. Let

(*1, ..., tl) = Axi AXn.(fx (Xi) ... (Xn) ,..., tt (x X) ... (Xn)),

where X i,..., x* are new variables of some uniquely determined types


such that £1 (xi)...(X n ) is of the type o. In reverse, we define, for
numerical variable k,

(t)k = Axi Ax„. (t (x x )... (x„))fe,

where (,..)fc is the corresponding decoding for the type o. Concatenation

is defined by

t * s = Axx Ax«. (£ (x x) ... (xn) * s (Xx) ... (Xn)) .

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 54

Then we have

t * ( t l+1) s (tu ...,th ti+i),

(t)fc ~ ifc, for k = 1,

(t * s) * r ~ t * (a * r ) , and so on.

For t = ( i i , ti), (f (x x )... (x„))0 = Z, the length of the sequence, and

(t)0 ~ A xi Axn.r

As an abuse of the notation, we will simply write Axi Ax„.7 as Z, so


( t ) 0 ~ Z. This means, not every object of the type a is a code of a finite
sequence, for (t)0 may not be constant. So, we introduce the definition:

Seq( t ) = 3l ( ( t ) 0 ~ l ) ,

which reads‘t is a finite sequence’. We will also simply write (t ( x i ) ... (Xn))0

as (t)0.
Given a term s of the type tr, there is a standard way of transforming
it into a term of the type (ox,...,<7n —» a), that is, A zi...xn.a. So we
can also code a sequence of terms of the types a or (o x ,..., an —>a)
into a term of the type ( 01 ,..., crn —> a). We will use such conventions
in coding terms of different types.
For a variable x of the type ( 0 —►a ), we can construct a term x of
the same type, containing x as the only free variable, such that

x (0 ) ~ <x(0)),

x ( n + 1) ~ x (n ) * (x (n + 1 ) ).

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 55

That is, intuitively, z ( n ) = (z (0 ) , ...,z (n ) ). This is obtained by first


constructing a numerical term q [n, z , X i,..., x*] such that

q [0, z, x : , ..., x,] = (z (0) ( x i ) ... (x,)),

q[n + l,z ,X i, Xf] = q[ n, x, xi , . . . , xi ] * (z (n + 1) (x x) ... (xj)),

and then letting z = An.Axx Axj.g[n, z,X i,...,x < ]. So, given a se­
quence, we can always construct the sequence of the codes of its finite
initial segments. However, we must remember that z (n ) shows the
properties for the code of the finite sequence only in contexts which
preserve extensional equality, for only in those contexts can we replace
z (n + 1) by z (n ) * (z (n + 1)). In exactly the same way, given z of
the type (o —►a), we can construct a term z such that V n 5 eq (z(n ))
implies

z (0 ) ~ z ( 0 ) ,

i(n + l)~ i(n )* z (n + l),

that is, z (n) is the concatenation of z (0 ),..., z (n).


Now we give a general treatment of constructing sequences of ob­
jects of the types (o —►o) or o, and with that we will prove some useful
induction rules in SC . In formalizing Bishop’s constructive analysis,
real numbers are represented by sequences of rational numbers, which
in turn are coded by natural numbers. So a real number is just an
object of the type (o —» o). A general discussion of the constructions
of even higher type sequences would be too complicated. In the actual
formalization work below, the reduction of such constructions will be
treated in different ways depending on the context.
As a convenient notation, we understand s ( m ) as just s, if j is
actually of the type o. Let z and t [z] each be of the type (o —►o) or

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 56

o. We call t [x] extensional in x, if (i) x is of the type o, or else (ii)


we can show that, for a new numerical variable m, t [x] (m) is provably
equal to some term t' such that all free occurrences of x in t' are in
a context of the form x (p ), and such contexts can be iterated but
no free occurrences of variables in the term p become bound in the
whole term t1, that is, x (p) is substitution-free in t'■ Intuitively this
means, according to the syntax of f', that to evaluate t [x] (m ), the
m -th’member of the sequence t [x] (or to evaluate t [x], in case it is
already numerical), the only things needed about x are some values of
x, that is, x (p). In other words, x itself as a program is not needed.
For we can evaluate t [x] ( m ) , beginning with the innermost contexts
of the form x (p ), and using the value of x (p) to calculate the value
of an outer context, say x (x (p)) and so on. In this way, we may need
to apply the function x iteratively, but the fact that all contexts of
the form x(p ) are substitution-free guarantees that we need only to
iterate x a fixed number of times. This corresponds to the pattern of
numerical recursions given above. In this case, we can construct the
sequence x ( m ), t [x] (m ) , t [t [x]] (m ), by primitive recursion.
Notice that y (x) is not extensional in x, where y is a variable of the
type ((o —►o) —* o) or ((o —» o) —» (o —♦ o)). On the other hand, if no
free variables of t [x] (m ) are of types higher than (o —» o), then, in the
normal form of t [x] (m ), there are no A terms and all occurrences of x

are in a context like x (p), and hence, the only things that can violate
extensionality in x are contexts like R e ij ( s , r , t [x (?[?]), i, j\) . j is
a bound variable, so x ( p [ j ]) is not substitution-free in that context.
In this case, to evaluate t [x] (m ), we need to iterate x up to s times. If s
contains m, we cannot construct the sequence x (m ), t [x] ( m ) , t [t [x]] (m ),
by primitive recursions.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 57

For x a sequence of variables and t [x] a sequence of terms, we call


t [x] extensional in x, if each term in t [x] is extensional in each variable
in x . For s a term of the type ( o i , on —►cr0) , where each of Cq, ..., crn
is (o —♦ o) or o, we call s extensional if s (x\, ...,z n) is extensional in
each of X i , x n. We call a primitive formula t [x] = s [x] extensional
in x , if both t [x] and s [x] are extensional in x . For a formula <p [x] of
SC , we call tp [x] extensional in x if all primitive subformulas of tp [x]
are extensional in x.
The informal discussion above shows that extensionalities imply
preserving extensional equalities. Notice that for tp [x] to preserve ex­
tensional equality for x, x can appear in tp [x] in a context x (p) where
some free variables in p are bound by quantifiers of tp [x], because we
have SC b t ~ s —» Vm (t (m ) = s (m )).
Then we can summarize the numerical recursive patterns given
above this way.

LEMMA 4 .1 . Let x = (x i,...,x j) be a sequence o f variables each


being of the type o or (o —►o), th [ti, x ] , rh be term s o f the same type as
Xh, h = 1 ,..., I. Suppose that th [n,x] is extensional in x . Then we can
construct type o term s qh [n,m] such that, for h = 1 ,..., I, the following

are provable (in P R O ):

qh [0, m] = rh (m ) ,

qh [Sn, m] = th [n, Am.gx [n, m ], ..., Am.qi [n, m]] (m ).

PROOF. By the definition of extensionality in x, we can rewrite the


recursion equations above into a pattern like

qh [Sn,m] = gfc[n,p[m]] ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 58

where there may be several iterated occurrences of terms like q* [n, p [m]],
for instance, a context like qy [n, p' [qy> [n, p"[ m]], m]], but all such
contexts are substitution-free. This can be reduced to simple primitive
recursions as discussed at the beginning of this section |

From the lemma, we have the following theorem about inductive


construction:

THEOREM 4 .2 . I f x , y are of the type (o —►o) or o, t [ x , n ] is ex­


tensional in x, and r is a term of the same type as x, then we can find

some term T [n] of the same type such that

S C I- T [0] ~ r A T [Sn] ~ t [ T[ n] , n] .

PROOF. By the lemma, we can construct a term q [n, m] such that

?[0, m] = r ( m ) ,

q [5n, m] = t [Am.q [n, m ], n] (m ).

Let T [n] be Am. q[n, m]. Then

T [0] ~ r

T [5n] ~ t [T [n], n ].

In analysis, when we construct a sequence (zn) of real numbers


by iterating the operation z => t [z, n], the extensionality conditions
here are always satisfied. In calculating z n+i (m) = t [zn, n] (m ), we
may need some values like xn (m'), or even z n (z n (m')). It seems very
unlikely that we need z n (z n (...z „ (m /) ...)), where the number of iter­

ations depends on m.
Now we can prove some induction rules. First we have

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 59

THEOREM 4 .3 . Let x , y be of the type (o —» o) or o, and ip [x,y,n]

be a formula of P R O . Suppose that

s c i- 3 x v y 0 [;c)y)O]

SC 1- 3xV y0[x >yj7i] —►3xVyV'[I iy>‘S'n]

with terms r, X , Y of appropriate types such that

P R O b ^ r .y .O ] ,

P R O b i p [ x , Y (n) (x, v ) , n) —►ip [X (n) ( x ) , u, 5 n ] .

Suppose that iff [x, y , n] preserves extensional equalities f o r x , y, X (n) (x)


is extensional in x, and Y (n )(x ,u ) is extensional in v . Then

SC b BxVyifi [x, y , n ].

PROOF. First, by the theorem above, we can construct a term T [n]

such that

T [0] ~ r,

T [ 5 n ] ~ X (n ) (T [n ] ).

By the assumptions we have

* [T [n], Y (n) (T [n], v ) , n] V [X (n) (T [n ]), ®, S n ].

Since ip [x,y,n] preserves extensional equality for x, we have

ip [T [n], Y (n) ( T [n ], v ) , n \-> ip [T [5 n ], v, S n ].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 60

Then similarly we can construct a term U [m], which may contain free
variables n and v, such that

U [0] ^ v,

U [ S m ] ~ Y ( n ) [T[n], U[m]},

ip [T [n], U [Sm ], n] —* ip[T [Sri], U[m] , S n ].

From .these we have

(; < n -» ip [T [j] , U [ n - j ], j])

- > ( S j < n -*• ip [T [ S j ] , U[n — S ; ] , S ; ] ) .

By the assumption and T [0] ~ r we already have ip [T [0], U [n], 0 ].


A quantifier free induction in P R O gives

j < n - u p [T\j\, U [ n - j ] , j ] .

Substituting n for j , we obtain ip [T [n], u, n], which is SC h 3x iy ip [i, y, n]. |

Using the coding techniques, we can see that the theorem can be
generalized.

COROLLARY 4 .4 . The theorem still holds if x ,y are replaced by


x, y , sequences o f variables of the types o or (o —* o), and r, X , Y are
replaced by r, X , Y , some sequences o f terms o f appropriate types, and
the various extensionality conditions still hold.

The conditions for the induction rule in Theorem 4.3 refer to terms
X (n) and Y (n), which means, to apply the induction rule, we must
spell out the proof of

3x iy ip [z, y, n] —» 3x'iyip [z, y, Sn]

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 61

and check the witness terms. As a matter of fact, in applications we


rarely mention this theorem, because the real difficulties are in the con­
structions of witness sequences. After that, the inductions are trivial.
On the other hand, in some special cases, the witness sequences are au­
tomatically obtained. Notice that any term or formula is extensional
in its numerical variables. Therefore, if x , y in Corollary 4.4 are all nu­
merical variables, then all the extensionality conditions automatically
hold. In a word, we always have E° inductions in SC .
We will use the common conventions about the notations E°, 11°,
Eij, that is, for tp quantifier free, any formula known to be equivalent
to 3 m \..3 m ntp, or Vm^.-Vm*^, or 3m !...3m„Vk\...'iknp will be called
E°, or 11°, or E° respectively. However notice that the combination
3n < N m cannot be transformed into Vm3n < I in general, though all
bounded quantifiers immediately applied to quantifier free formulas can
be eliminated and Vn < 13m can be transformed into 3mVn < I. Recall
that, primitive recursive arithmetic contains only E° inductions. What

we have here is apparently stronger. But notice that we don’t have


general 11° inductions. A quantifier combination 'imSntp [m, n] must
be interpreted as 3xVmy> [m, x (m)], which involves a non-numerical
quantification. As a matter of fact, IIj induction would enable us to
evaluate all primitive recursive functions in SC and thus would make
SC not conservative over the primitive recursive arithmetic.
What are more useful in applications are inductions on formulas
with parameters ranging over a domain, or inductions with assump­
tions, also called relativized inductions. We first consider a simple

case.

LEMMA 4 .5 . Suppose that tp [n, x] is a quantifier-free formula whose


free variables are all in n, x , and x [x] is any formula o f SC whose free

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 62

variables are all in x , and suppose that

SC h X [x] —> v [0, x]

SC I- X [x] -* O [n, x] -* <p [Sn, x ] ) .

Then

SC h x [x ] -»< p[n ,x].

P r o o f . Suppose that x [ x ] = 3yV zxi [ y ,z ,x ] . By the assump­


tions, we have closed terms Zo, z such that

P R O h x i [y, zo (y , x ) , x] - » <p [0, x ] ,

P R O h x i [y, z (n, y , x ) , x ] - » (<p [n, x] - » tp [5n, x ] ) .

Let z' = A n yx.J(n , z0 (y , x ) , z (n - 1, y , x)). Then

P R O I- Xi [y, z' (0, y , x ) , x] «-►Xi [y, *0 (y , x ) , x ] ,

P R O I- x i [y, z ' ( 5 n , y , x ) , x ] « Xi [y» * (n >y>x ) . x l •

Let [n] = Vi < n x i [y, z' (i, y , x ) , x] —» ^ [n, x ] . Then we will have

P R O I- tf) [0] A (if> [n] -» ^ [<Sn]).

By an induction in P R O we obtain

P R O 1- Vi < n x i [y, z' (i, y , x ) , x] -*• if [n, x ] .

By the discussion on page 42, we can construct terms z" such that

P R O h xx [y, z* (n, y , x ) , x] - » <p [n, x ] ,

which is SC h x [x ] —* <P[n >x l- ■

Then we have the general case.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 63

THEOREM 4 .6. Suppose that y , z are variables o f the type (o —►o)


or o, <p [n, x] = ByVz^i [y, z, n, x] is any formula whose free variables
are all in n , x , and x [x] is any formula o f S C whose free variables are
all in x . Suppose that

SC b x[x] -► ¥>[0,x]

SC b x [x] -► (tp [n, x] - » <p [Sn, x ] ) ,

and further we can find terms Y q, Y , Z such that

SC b x [x] -♦ <px [Y0 ( x ) , z, 0, x ] ,

SC b X [x] -* {<P\ [y, z ( y , v , n, x ) , n, x] -* <px [Y (y , n, x ) , v , Sn, x ] ) .

Suppose further that <p\ [y, z, n, x] preserves extensional equality for


y ,z , Z ( y ,v ,n ,x ) is extensional in v, and Y (y ,n , x) is extensional

in y . Then
SC b x [x ] -ȴ> [n,x].

PROOF. The proof is almost exactly the same as the proof of The­
orem 4.3, with the application of the induction in P R O replaced by
an application of Lemma 4.5. |

Similarly, in case tp is a E , formula, we have a simpler format:

THEOREM 4 .7 . Suppose that <p [n, x] is a Hi formula whose free


variables are all among n, x, and x [x] w any formula o f S C whose
free variables are all in x . Suppose further that

SCI-xW —¥>[0,*1
s c I- x W — 0 1 ", x] ->v[Sn,x\).

Then
SC b x [x] —► [n, x ] .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INDUCTIONS AND INDUCTIVE CONSTRUCTIONS IN SC 64

PROOF. Suppose that x [x] = 3yV zxi [y, z , x] and <p [n, x] = 3mVk<pi [m , k , n, x].

From the assumption SCI- x [x ] —►p [0,x], we have closed terms Yo


such that

SC h Vzxi [y, a, x] - » p [Yo (y, x ) , k, 0, x ] .

Now p [n, x] —►p [5n, xj is interpreted as

3M , KVm, k (p i [m, K (m , k ) , n, x] -♦ p x [M ( m ) , k, S n , x ] ) .

So, from the assumption SCI- x l x ] (v? [n, x] —» ^[SVi, x]), we have
closed terms M*, K* such that

SC f- V z x i[y ,z ,x ] -♦

(<pi [m , K* ( y , n, x ) ( m , k ) , n, x] -*■ tpx [M* (y, n, x ) ( m ) , k , Sn, x ] ) .

Since m , k are numerical variables, the extensionality conditions of

Theorem 4.6 all hold (with x [x] replaced by x! [y,x] = Vzxi [y, z, x]).
The conclusion then follows easily. |

So, we always have inductions with assumptions. This will be


the most frequently used induction rule. Further, notice that double
inductions on 11° formulas are instances of this induction with assump­
tion. That is, if x M is Vmcp [n, m] where tp is a 11° formula, then if we
can prove

'im p [0, m ],

i m p [n, m] —►p [n + 1 , 0] A {p [n + 1, m] - » p [n + 1, m + 1 ] ) ,

then i n i m p [n, m] will follow. Here, an induction with assumption is


used to prove i m p [n, m] —» i m p [n + 1 , m ].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 65

5. S e ts an d F u n ction s in SC

5.1. In tro d u ctio n . We will start developing mathematics in SC .


This section explains how to express basic notions, such as sets, func­
tions, partial functions, complemented sets and real numbers in SC.
The original constructive formulations of these mathematical concepts

are given in Chapter 3 of Bishop and Bridges’ book Constructive Anal­


ysis. The task here is to show that they can be formalized in SC . The
basic ideas of the formalization come from Bishop’s paper [14], though
our presentations are different. The next section will show that the ele­
mentary calculus developed in Chapter 2 of the book can be developed
in SC . Appendix A and Appendix B will contain the developments of
more advanced mathematics in SC .
Obviously, it is impossible to present the details of the actual for­
malizations. It is too lengthy and tedious. It is also unnecessary, for our
objective is to show the formalizability, not the actual formalizations.
We want to explore the minimum logical system that can represent or­

dinary constructive mathematics. Given this philosophical objective,


the formalizability is the real issue. This is different from the purpose
of other studies concerning the appropriate systems for formalizing
Bishop’s constructive mathematics. For example, Myhill considered
extending Bishop’s system to include set variables because he wanted
‘to make the process of formalization completely trivial’ ([70], p. 347).
So we will only give informal arguments to show the translatability of
informal proofs into proofs in SC , and the simplicity of formalization
will not concern us very much.
However, after the basic notions have been represented in the lan­
guage of SC , the formalization is actually routine. Whenever there are
some obstacles, they are m ostly due to the limitedness of the inductions

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 66

and recursions available in SC , which we impose to make PR O in con­


formity with finitism. If we do not want this finitism and we can use
the system To instead of P R O , the formalization will be quite simple
and most of the contents of the following sections will be redundant.
On the other hand, from a computational point of view, the lan­
guage of P R O is more natural than the more powerful language of
To or other intuitionistic systems. Terms of P R O representing nu­
merical functions will naturally correspond to definitions of primitive
recursive functions. Definitions of primitive recursive functions can be

translated into ordinary computer algorithms using loops and callings


of sub-procedures very straightforwardly. In contrast, the terms of To
may contain recursions on higher type functionals which have no direct
counterparts in ordinary computer algorithms and whose realizations
must essentially use constructs such as recursive calling procedures.
Put in another way, if we want the word ‘constructive’ to have some
more realistic meaning, for example, to be closer to ‘computable on
ordinary computers’, the mathematics in SC will square better with
what is constructive.
Mathematics in SC can very well be called primitive recursive
mathematics. Of course, being primitive recursive is still very far from
being feasible. Given the fact that the exponential function is ubiq­
uitous in applied mathematics, a feasible mathematics must be quite

different from the ordinary mathematics. Generally, subtantial work


is needed in finding feasible algorithms. What we will show here is
that much of ordinary mathematics can be translated into primitive
recursive mathematics rather straightforwardly.
Since all intuitionistic logical laws (except inductions) are available
in the system SC , to translate the constructive proofs in Bishop and

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 67

Bridges’ book Constructive Analysis into proofs in SC , it will suffice to


show that various inductions used can be replaced by ones available in
SC . So we will usually skip the proofs or steps of the proofs which do
not involve any inductions. These axe understood as trivially formaliz-
able. On the other hand, every inductive construction of sequences and
every application of induction rules will be examined carefully. More­
over, some presentations in the book Constructive Analysis apparently
use quantifications over sets, for instance, in the theory of integration.
We will also find some way to avoid that.
We will frequently refer to the book Constructive Analysis by Bishop
and Bridges [15] and hereafter we will refer to it as B&B. When we
make references to theorems in B&B, we will use phrases like ‘The­
orem (5.11) on p.157 of B&B’, or simply ‘Theorem (5.11) on p .157’.
In particular, when ‘p. x x x ’ is used without explicitly mentioning
which paper or book is referred, we always mean B&B. Furthermore,
when we claim that something can be proved, we always mean that it

is provable in SC.

5 .2. S e ts. First we represent sets in SC . A set will be a pair of


SC formulas, defining respectively the membership condition and the
equality relation between two members. For A = (<p [a], i/> [a, b]), we
will call A a set form o f the type a , if a is the only free variable in <p}
a and b are the only free variables in rj), a and b are of the type a. In
that case, ‘A is a set of the type a ’ is translated as the conjunction of
the following formulas (i) and (ii) in SC , where recall that a ~ b is the
extensional equality introduced in Section 4:
. ( V [a] A <p [5] A <p [c] -+
(i) Vo, b, c
^ xl>[a,a]A(il)[a,b]-*4>[b,a\)/\(rl)[a,b]Ail>[b,c] -» ^ [ a ,c ] )
(ii) Va, b(a ~ b A <p [a] - » <p [6] A [a, 5]) •

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN S C 68

We usually write <p[a\ as a 6 A , and write ip [a,b] as a =& b


(or simply a = b when no real ambiguities will occur), and call them
respectively the defining formula and equality formula of A. (i) is called
equivalence condition and (ii) is the extensionality condition, (ii) says
that <p [a] preserves extensional equality for a and the equality relation
for a set must be weaker than the extensional equality for the members
of the set. The extensionality condition was not explicitly stated in
B&B,'but it conforms with the idea that we treat only extensional
concepts in Bishop’s constructivism. Further, it is not a real restriction
on sets, for the concrete sets we will see in analysis all satisfy this
condition. We will need this condition in order to code an arbitrary
finite sequence of elements of a set A into a single entity of the same
type, and thus to form the set A <oa of all finite sequences from A.

If (a 6 A ) d is BxVy^o [a, x, y] and x are of the types px, ..., pm, we


call (<x, p i , ..., pm) the signature of the set A and call p \ , ..., pm the wit­
ness types of A. And we use a 6 X A. to denote the formula Vyt^o [a, x, y],
read as ‘a belongs to A with x as the witnesses’.
The need for witnesses is the major difference between the way we
talk of sets in constructive analysis and in the classical analysis. As
we will define them below, the constructive functions defined on a set
operate on the witnesses for an element’s being in the set, as well as
the element itself. For instance, the inverse function x -1 defined on
non-zero real numbers should operate on a real number z , which is
a sequence (x (ti))^Lx of rational numbers such that |x (n) — x (m )| <
n~l + m -1 for all m , n, together with a witness n for x ^ 0, that is,
an n such that |x (n )| > n _1; cf. B&B, p.24. When real numbers
are defined as such sequences, it is impossible to construct the inverse
(x -1 (n))^Lj as a sequence solely from the sequence (x (n ))n , so that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 69

whenever x is a non-zero real number, x -1 is also a real number and


x _1x = 1. To see this, notice that if (x _1 (n))n is a real number, then
|x -1 | is bounded by |x -1 (1)| + 2 from above. Choose x a real number
such that presently we don’t know if x = 0 or |x| > 0 and construct
x _1 (1). Then, |x| > 0 would imply that (x -1 ( n ) ) ^ is a real number

and |x_ l| < |x_1 (1)| -I- 2, and hence |x| > (|x _1 (1)| + 2)- 1 . By the
laws of intuitionistic logic, we have

(3) |x| < (|x_1 (1)| + 2) 1 - » x = 0.

Let N > |x-1 ( l ) | + 2 be a integer. Then |x(4JV)| > j ( | x _1 (1)| + 2)-1


implies |x| > 0, and |x(4JV)| < | (|x -1 (1)| + 2)-1 implies |x| < (|x -1 (1)| + 2 )-1
and hence x = 0 by (3). So we would be able to decide if x = 0 or
|x| > 0, a contradiction. Therefore, the witness for |x| > 0 must be
needed in constructing (x -1 (n ))n .
As usual, we use Vx G A to mean Vx (x 6 A —►...) and use 3x 6 A
to mean 3x (x 6 A A ...). When sets A and B are of the same type, lA
is a subset of B \ or A C B , is the formula

Vx € A (x 6 B ) A Vx, y G A (x = * y «-♦ x = b y) A

Vx, y 6 B (x G A A x —b y —* y 6 A ) .

The last component says that the membership condition for A is ex­
tensional relative to the equality relation of B . This is called the ex­
tensionality condition for subsets. We will make this convention: when
■we mention a subset of a set, unless otherwise stated, we always as­
sume that its equality formula is ju st that o f the set. So, a subset is
determined when its defining formula is given. Further, A — B is
( A C B ) a ( S C A). We also need stronger notions of containment and

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 70

equality between sets. For A, B of the same signature, A X B is

A C B A Va, x (a 6 x A —> a € x B ),

and A = B is A ■< B A B ~< A.


The intersection and union of two sets A and B are defined only in
case they have equivalent equality formulas. In that case, the equality
formulas of the intersection and union are that of A , while the defining
formulas are respectively

x e A n B = x e A a x e B,

x 6 AU B = x € A V x 6 B.

There are some generalizations of sets. First, sets of n-tuples, or


n-place multipit sets. A set form for an n-place multiple set of the type
(<Ti,..., an) will be apair A = (<p [a], ip [a, b]), wherethe single variables
a and b arereplaced by sequences of n variables a and b of the type
(a-! , ..., <rn). Then lA is a set’ is translated similarly. The signature
of the multiple set will be (o-1,...,a n,p i, ...,pm). Other notations, for
instance, a € A, a = a b, a 6X A, A C B and so on, are defined
similarly. Notice that n here is not a symbol of the language of SC . It
is a symbol in the meta-language. We cannot define sets with varying
tuples as members.
Second, parametrized sets. A set form A [w] for a parametrized set
with the parameters w is a pair of formulas (ip [a, w ], ip [a, b, w]) such
that <p, ip may contain free variables w other than a, b. Then lA [w]
is a set’ is translated as the conjunction of (i) and (ii) above, with
[a], -0 [a, b] replaced by <p [a, w j, ip [a, 6, wj. Similarly, notations like
a £ A [w], A [w] C B and so on are clear. We frequently write A [w]
as Aw. A form fo r a fam ily o f sets is a pair (A [w], x [w]) consisting

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 71

of a set form A [w] for a parameterized set and a formula x [w ] with w


as the only free variables. Then ‘(A [w ], x (w l) is a family of sets’ is to
mean ‘For all w , if x [w], then A [w] is a set’. We also say ‘{A W • X[w]}
is a family of sets’. It is more useful to consider a family {A, : i € / } of
sets indexed by a set I. In using this notation we always assume that

the following extensionality condition relative to the indices holds:

Vii, ia £ I (ii = / is —►At, = A t,).

We will also use the notation {A [}l€/.


Clearly, we can also combine these generalizations. So, we have
parameterized multiple sets, families of parameterized sets or multiple

sets and so on.


A sequence (A n)n of sets will be a parameterized set with n as (one

of) the parameter(s). In case the equality formula does not depend
upon the parameter n, especially when An is a subset of a set B for all
n, the union and intersection of the sequence are sets defined as usual:

x € Un_oAn = 3n (x £ An) ;

x e n ^ o A n E E V n fc e A n ).

Similarly, for any family {Ai}t6/ of subsets of a set B , we can define


the set Ute /A | and Djg/Aj. Clearly, these are still subsets of B.
Let A j ,..., An be n sets of the types <7i,..., <rn respectively. The
product A i x ... x An is the multiple set of the type (o i, ...,<7n) with

the defining and equality formulas:

((* 1,..., x„) £ Ai x ... x An) = (x i £ Ai A ... A xn £ A n ),

((x i, •••) i n) = A ix ...x A n y*0) = (* i Ai Vi a ... a xn yn) •

p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
5. SETS AND FUNCTIONS IN SC 72

Similarly, if each Ai is a multiple set of the type <7^) for i =


1, then the product will be a multiple set of the type ( < r j , akl, ....
Clearly A^’s can also be parametrized sets. Also notice that n here is
not a variable in the language of SC . We cannot construct a product
of an indefinite number of sets this way.
However, when an indefinite number of sets are given as instances
of a parametrized set, we can construct their products in a different
way.' Suppose that An is a set form for a parameterized set with the
parameter n. Then for any term N we can construct the defining and
equality formulas for the product fl^L o which is a set of the type
(o -4 tr):

f x 6 JJ Ah') = V n < N ( x n € A n ) ,
\ n=0 /

( * = rin=0 ** y) - Vn - N ( Zn =A~ Vn) •

Then it is easy to prove the statement ‘if An is a set for all n < N,
then rin=o An is a set’. We use the same format of definition to define
n “ o An- So, its members are sequences (xn) such that xn 6 An for all
n. Clearly, An can also be multiple sets or sets with other parameters.
Recall that we can code a finite sequence of objects of a type into
a single object of the same type. So, given a set A, we can define A"
as a parameterized set with n as a parameter:

x € An = Seq (z ) A (x )0 = n A V f c > 0 ( & < n - » (x)k 6 A ) ,

x = An y = ' i k > 0 ( k < n —> (x)fc = A (y )k) .

Notice that the extensionality condition for A is necessary for An to


behave like a product set, for instance x 6 AnA y G Am —* x*y G An+m,
for we have only this extensional equality (x * y)k ~ (x)fc for 1 < k < n.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 73

Further we can also define the set A <0° of all finite sequences of elements
of A:

x G A <0° = Seq (x) A VA: > 0 (k < (x)0 —» (x )fc G A ) ,

x =x<» y = (x ) q = (y)0 A Vfc > 0 (k < (x)0 -» (x)4 = a (y )fc) •

To simplify the notations we make a convention: from now on,


Wien, we mention a set A, unless otherwise explicitly stated, we always
assume that it can aho be a multiple set, a parameterized multiple set
and so on, although we will simply use the plain letters as variables
and write x G A rather than x G A, and we usually suppress mention
of parameters unless they are needed.
Notice that we have no variables for sets and we cannot quantify
over sets in the language of SC . Assertions about arbitrary sets are
translated into schematic formulas containing meta-linguistic variables
for arbitrary formulas. When we state a theorem, we may say ‘Given a
set A , ’, or ‘For any set A , but that should mean: for any set
form A, the schematic formula ‘If A is a set, t h e n ’ is provable in
SC . Similarly, when we say that (A„)n is a sequence of sets, we mean
actually that An is a parametrized set with n as one of the parameters.
So we can concede only a sequence of sets of the same type. Whenever
there is an apparent quantification over sets, we will try to replace it
with quantifications over parameters of some parameterized sets. For
instance, ‘for all sets in the sequence (A«)n...... 1 is no other than ‘for all
n This is the main trick we will use to eliminate quantifications
over sets. The formalization of the integration theory will use this trick

extensively.
When we say “‘A is a set” is translated a s ’, we directly give the
literal translation of an informal sentence into a formula in the language

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 74

of SC , but when we say ‘Let A i , A n be n sets of the types ..., crn.


Then we can define the product A \ x ... x An as the multiple set of the
type ( o i , < r n) with the defining and equality formulas ’ and so on,
we are not literally translating informal sentences into formulas in the
language of SC . We are only explaining the strategies of formalization
and conventions about notations. To express it more accurately and
completely, we should say ‘Let A i,...,A n be n set forms (that is, n
pairs'of formulas). We agree that Ai x ... x An is to denote a form for
a multiple set consisting of the defining formula and the equality
formula Further, the schematic formula (where ‘A,- is a set* should
further be translated)

(Ai is a set) A ... A (An is a set) —►(Ai x ... x An is a set)

is always provable in S C ’. So here we make a convention about the


symbol ‘Ai x ... x An’ at the meta-language level and also claim that
a schematic formula is provable in SC . In particular, one should not
take it to mean that we define Ai x ... x A„ only if all A i , ..., An are
sets, or that the set A \ x ... x An exists if all A \ , ..., An are sets, for all
these conditions may not be decidable. Conventions we make at the
meta-language level should be universal, that is, should not depend on
the provability of any formulas, although we may say loosely things like
the above or ‘If for any n, A„ is a set, then we can construct the set

n « 0 An’. Here, similarly, the convention about the notation n^Lo4»


is made for any set form An for a parameterized set, while the im­
plicit assertion is that (the translation of) ‘If An are sets for all n, then
n “ 0 An is a set’ is provable in SC . It would be much too tedious to
give such details always. Recall that our objective is to show formal­
izability, not the formalizations. We are not interested in presenting

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 75

literal translations of all theorems and proofs in B&B. So we will keep


using the informal and loose explanations, as long as we axe sure that it
is obvious what are the conventions about meta-linguistic symbols im­
plicitly made, what should be translated into formulas of SC directly,
or what formulas of SC are implicitly claimed to be provable in SC.
Here are some examples of sets. We assume a fixed coding of ra­
tional numbers into natural numbers, so we have the set of rationed
numbers, Q, as a set of type o, whose defining and equality formulas
are some quantifier-free formulas. We will use the common notations
for the common functions or predicates of rational numbers, considered
as primitive recursive functions or predicates, for instance, + , —, / , pq,

Ip Ij P_1 » P ^ 9> P > 9 so on- (A default value for 0-1 can be set.)
We will write Vm (m > 0 —►...) as Vm > 0..., and similarly for 3m > 0
and so on. The set of real numbers, R, is of the type (o —►o) and sig­
nature (o —» o) and has the following defining and equality formulas:

x 6 R = V m ,n > 0 (x ( n ) 6 Q A |x (m ) —x (n )| < 1/m + 1 / n ) ,

x = R y = Vn > 0 ( |x ( n ) - y ( n ) | < 2 /n ) .

The equivalence and extensionality conditions are easily verified.


Here are some common subsets and parameterized subsets of R.
Their equality formulas are all the same as that of R, so we give only
defining formulas. The relevant extensionality conditions are all trivial.
Positive real numbers R +, signature: ((o —» o ) , o),

x 6 R + = x € R A 3 n > 0 ( x (n) > 1 / n ) ;

R~ is defined similarly.
Non-negative real numbers R +0, signature: (o —►o),

x 6 R +0 = x e R A V n > 0 ( x (n) > —1 / n ) ;

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 76

Intervals [a, b], parameters: a, b of the type (o —►o); signature:

(o - o),
x € [a, 6] = x € R A a < i A z < 6,

where x < y is Vn > 0 (x (n) < y (n) + 2/n);


Intervals (a, b), parameters: a, b of the type (o —►o) ;signature: ((o —►o ) , o, o):

i 6 (o,6) = i 6 R A o < i A i < 4 ,

where x < y is 3n > 0 (z (n) < y (n) —2/n).


Intervals [a, 6), (a, 6], (oo, 6], (a, oo] and so on are defined similarly.

[a, 6] is called a compact interval.


Sometimes we consider sets with inequality relations, which means
a set A together with a formula defining a relation ^ . Then is an
inequality relation on A ’ is translated as:

Vx, y , 2 e A ( i ^ j - + - i i = y A y / i A ( i / z V y ^ z ) ) .

Usually there is a natural definition for inequality. For example, we


always assume
(x £ y) = (x > y V x < y)

for the real numbers.

5.3. F u n ctio n s. Suppose that A and B are sets of signatures

(ffo,<7i,...,orn ) and {p o,P \,—,Pm) respectively, and / is a term of the

type (cto, 0 i, On -♦ Po)- ‘/ is a function from A to B \ or / : A —* B,


is the formula

VxoX! (x0 6 Xl A -* f { x o, Xi) 6 B ) A

VxoXiy0y i (so 6*t A A y0 € yi A A x0 = a yo f (so, Xi) = b / (yo, y i)) •

From / : A -» 5 it follows that x0 A A x0 € XJ A —* f (x0, Xi) = b


/ (xq, X2). So we frequently simply write / (xo) instead of / (x 0, Xi), as

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 77

equal members of a set can be treated as the same in most contexts.


Similarly, sometimes we use notations like Vxo 6 A ( . . . . / (xo) —•)> while
literally it should be VxoXx(x0 € Xl A —►. . . / ( x0, x W e will use
such simplified notations frequently. They are more readable and the
context can always determine how to complete them, as long as we
always keep in mind that a function operates on the witnesses as well
as the elements.
It is clear that we don’t take a function as an entity carrying with
it a domain and a range. Syntactically / is a term of some appropriate
type and hence it signifies a primitive recursive operation. As such
it operates on all entities of its argument types. The assertion */ is
a function from A to 5 ’ is an assertion involving / , A, and B. For
instance, consider the inverse function x -1 on the set R +. It operates
on an arbitrary sequence (xn)n of natural numbers and an arbitrary
natural number m and generates a sequence (x ~ x)n of natural numbers.
We can prove in SC:

VxVm (x € m R + —►x -1 € R ) ,

in other words, if x is a real number and m is the witness for x > 0,


then the generated sequence (x^l )n is also a real number. That is the
assertion x -1 : R+ —* R. Nothing is said about the case when x is not

a real number or m is not the witness.


The set of functions from A to 5 , F ( A , B ), is defined as follows:

f £ F (A, B ) = f : A -* B,

f = f(a j) 9 = VxoXx (x0 GXl A -» / (x0, Xx) = B g (x0l Xx)) .

This equality formula, which can be expressed as Vx 6 A ( / (x) = g (x))


in a simplified manner, is called extensional equality fo r Junctions. We

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission .
5. SETS AND FUNCTIONS IN SC 78

agree that when defining other sets of functions, for instance, the set
of continuous functions on R and so on, the extensional equality for
functions is always assumed unless otherwise stated.
Notice that when A is an n-place multiple set, Oq, x0t and yo above
are actually sequences of types and variables <rQ, Xo, yo- To emphasize
this fact, we will call / an n-place function. Similarly, when B is a
multiple set, po is a sequence po of types and / is a sequence f of terms
of the type (<r0, o i ,. . ., a n —►po). To emphasize that, we will call f a
multiple function.
Notions like */ : A —» B is onto’, ‘/ : A —+ B is an inverse of
g : B —* A ’, lf : A —i> B is one-one’ and so on can be formalized
straightforwardly. ‘A is countable’ is translated into

3 / ( ( / : N —►A) A ( / is o n to )),

and ‘A is subfinite’ is translated as

3x3m (Vn < m (x (n) G A) A Vz G A3n < to (x (n) = * z ) ) .

The extensionality condition enables us to code a subfinite set into a


single object of the same type. Suppose that x, m witness A ’s being a
subfinite set and let o = x (m) = (x ( 0 ) ,..., x (m — 1)). Then

Vn < m ( ( a )n+1 € A) A Vz G A3n < m ( ( a ) n+1 '

So a codes A. Notice that the extensionality condition is needed be­


cause we have only (a)n+1 ~ x (n) and we need extensionality to infer

(« W i« A-
Here are some more notions and notations about functions. If A,
B are sets with inequality relations, */ : A —» B respects inequalities’
is to mean ‘if / (x) ^ / (y) then x ^ y for all x , y G A ’. If / : A \ —* B
and Ax = A i then / : Aj —►B . The stronger equality is necessary for

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission .
S. SETS AND FUNCTIONS IN SC 79

this assertion to hold, for otherwise A \ and A2 may even have different
signatures. On the other hand, if in a context we have f : A - * B and
C C A , then there is a natural way to construct a function from C
to B as the restriction of / to C. First, C C A implies, by the axiom
of choice, that there exists W such that for any a, u, a € u C implies
a €w(a,u) A. W are the witnesses for C C A . Then let f be defined
by / ' (a, u) = / (a, W (a, u)) . Clearly, / ' : C —* B. We will denote f
by f \ c , °r simply / , while noticing that it actually depends upon the
witnesses for C C A . We will frequently use such simplified notations
without stating so explicitly. For example, after defining a function on
R, we always use the same notation for its restrictions to subsets of R.
Further, if in a context we have f \ : Ai —» B and / 2 : A 2 —> B , and

C C A\ A C C Aj, then ‘/ i = /a on C” is t^ mean

VxuiUj (x € C A x Ai A x 6 UJ A2 -» /1 (x, Ui) = f 2 (x, u2) ) ,

or V i £ C ( f i (x) = / 2 (x)) in simplified notations. Notice that the


assertion actually involves A\, A2 which do not appear in the simplified

expression. Also notice that it is equivalent to / i | c = f ( c ,b ) /a le under


the given assumptions.

5 .4. P a r tia l F u n ctio n s. In the theory of integration we must


treat partial functions from a set X to the set of real numbers. The

domain of a partial function on X is supposed to be a subset of X .


Subsets of X may have different witness types and hence they cannot
all be put into a single family of subsets. So, we cannot quantify over all
such partial functions and cannot define the set of all partial functions.
However, given a family of parameterized subsets V = {D \ : i € / } of
X indexed by a set I, we can define the set T D, Y ) of partial functions

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. SETS AND FUNCTIONS IN SC 80

from X to a set Y with domains in the family V. This is an (n + 1)-


place multiple set (or (n + m)-place multiple set if Y is an m-place
multiple set):

( i J ) e f ( V , Y ) = ( i z l A f : D { -+Y),

(i i,/ .) = ( i a , / i ) s ( A , = A , A Vx S A , ( / . (x) = / , (x) )).

(Notice that the latter involves some simplified notations.) The equiv­
alence and extensionality conditions are easily verified. We will say
that D\ is the domain of (i, / ) . So the domain (actually the parameter

of the domain) is uniquely determined for a partial function. ‘( i i , / i )

is a restriction of (ia, f i ) ’ is to mean

D h C D \7 A (fi = f 2 on D \x) .

And ‘the restriction of (ia> /b) to D ^ ’ is to mean (ix, /a)- Notice that
if (ii,/i) is a restriction of ( i a , /a)j then (ii,/x ) = (ii,/a), so the two
conventions just introduced are coherent. Further, we can quantify over
all partial functions in T (V, Y ), by which we mean a quantification like

v iv /((i,/)e m n - )•

When no real ambiguities will occur, we simply call / a partial function


and call D\ the domain of / , denoted by D m n ( f ). So for example
(i, / ) € F ( V , Y ) wiU be simplified as / 6 Y).
We say that the family V is closed under finite intersections if for
any finite sequence ix ,..., i„ of indices there exists j such that =
D J w h e r e the sequence ii,...,in can be coded by a single i, or
understood as an initial segment of an infinite sequence. We say that
V is closed under countable intersections if for any sequence (i„) of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
5. SETS AND FUNCTIONS IN SC 81

indices there exists j such that Dj C DjJLjAn* Notice that here we


need only inclusion instead of equality.

5.5. C o m p lem en ted S e ts. Let X be a set with an inequality re­


lation A complemented set in X is a pair A = (A1, A0) of subsets
of X such that x ^ y for x 6 A1, y 6 A0, cf. B&B, pp.72-75. Asser­
tions about arbitrary complemented sets in X should be understood
as schematic assertions, for they involve arbitrary formulas that are
the defining formulas for A1, A0. The union V“ xAi of a sequence of
complemented sets Ai = (A*, A°) is

( ( u £ , 4 ) n (n g , ( 4 u A?)), n g .x ? ) .

The intersection A“ 1At- is

(n g .A } , (u g iA ? ) n ( n g , ( ^ | u 4 ° ) ) ) .

The union and intersection of a finite sequence of complemented sets


are defined similarly. The complement —A of A is (A0, A 1) and A — B

is A A ( - B ) .
We will not define the characteristic function for an arbitrary com­
plemented set A, for, intuitively, it has to be a partial function with the
domain A1 U A0 and we want to consider only partial functions with
domains in a given family of subsets. Let V = { A : i 6 1 } be a family
of subsets of X indexed by a set I. ‘x is a characteristic function for

A in V ’ is to mean that

x e ^ ( T > ,{ o ,i } ) A

Vx 6 D m n (x ) ( ( x (*) = 1 -» * € A 1) A ( x ( z ) = 0 —» x € A0) ) .

Notice that we have the containment D m n (x ) Q A1 U A0 rather than


the equality.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
5. SETS AND FUNCTIONS IN SC 82

As an example, we give a detailed construction of a characteristic


function for V%=0Ak in V from the characteristic functions for each
Ak, k = 0, ...,n , in V, in case V is closed under finite intersections.
So, suppose that for k < n, Ak is a complemented set in X and Xfc
is a characteristic function for Ak in A and suppose that V is closed
under finite intersections. Recall that according to our convention,

Xik = (ifc.Xfc) € F ( V , { 0 , 1}) with D m n (x k ) = The assumption


about V is

ViVn(Vfc < n ( ifc 6 / ) -» 3j 6 I ( D } = n£=0A J ) •

By the axiom of choice and the numerical interpretation of implication,


this implies that there exists an operation C such that for all i,n , v,

Vfc < n (it € v(fc) / ) -» Dc(i,n,v) = n Jt=oAfc-

This further implies that there exists W such that for all i, n, v, x, u, k

(4)
' i k < n ( U € v (ife) / ) -* > ( i 6 U A :(i,n ,v ) —♦ ® G w (i,n ,v 1*,u1ik) A fc) •

C, W axe the complete witnesses for the fact that V is closed under
finite intersections. Since Vfc < n (i* € I ) , there exists V such that
Vfc < n (it €v(fc) i ) , that is, V are the witnesses for that assertion.
We know that for t [A] a term of the type o, we can construct a term
max£_0 (t [A:]) such that intuitively it is the maximum of t [0],..., t [n].
Let j = C (i, n, V ) and

X = Axu.max(xfc (*, W (i, n, V , x, u, k ) ) ) .

From the assumptions and (4) it follows that x : A ~ * { 0 ,1 } . We


denote ( j ,x ) by V£_0Xfc- Then, the assumptions plus (4) imply that
Vfc=0X* is a characteristic function for V£_0Ajfc in V .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 83

This is a rather detailed construction. The result (j, %) is a sequence


of terms containing witness terms C, W , V . Later on, when doing sim­
ilar things, we will not give such complete details any more. For ex­
ample, in a case like this, we will simply explain it this way: It follows
from the assumption that there exists j such that D] = fl%=0D m n (xk) •

Let x be such that D m n ( x ) — Ai and x ( x ) ~ max£_0 Xk(x ) for


x 6 D m n (x ) • Then x is a characteristic function for V]J_0j4fc in V.
In this way, we did not mention the witnesses C, W , V . Such simpli­
fied explanations are more readable and after some practice it becomes
routine to supply the details. However, we always remember that an
expression like V£_0x* may actually involve some parameters which are
the witnesses needed in constructing the term but do not appear in the

simplified notation.
Finally, notice that a characteristic function A”=1Xi for A"=1i4i in V
can be constructed similarly.

6. C alculus and th e R ea l N u m b ers in SC

This section corresponds to Chapter 2 of B&B. It will be shown


that everything in that chapter can be developed in SC . Since the
laws of intuitionistic logic are available in SC , the focus will be on the
formulations of the concepts and on the proofs that involve inductions
or inductive constructions.

6 .1 . T h e R e a l N u m b er S y s te m . The set R of real numbers is


already defined. Notice that the set of rational numbers Q is a set of the
type o that is different from the type of R. However, clearly x 6 Q —♦
Am.x 6 R. For convenience, we frequently ignore the type differences
between rational numbers and real numbers and simply write 0 € R,
x 6 Q —» x € R and so on.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 84

Common real functions on Rn, such as + , —, x , |- |, max, min and


so on, are constructed as closed terms of appropriate types, cf. B&B
p. 19. For instance,

+ is Xxy.Xn (x (2n) + y (2 n )),

x is Xxy.Xn (x (2Af [x, y] n) y (2M [x, y] n ) ) ,

where M [ x ,y \ is max (J |x (1)| + 2^, ^ |y (l)| + 2^ and V means the


least integer greater than t (as a rational number). We will use common
notations x + y, xy and so on for functions. All these functions are

extensional, that is, the term x + y, xy and so on are extensional in


x and y (cf. the definitions in Section 4 above). It is easy to verify
that these are all real functions, that is, + , x £ F (R2, R). Common
relations < , < , > , > , ^ and so on are defined by SC formulas. See the
last section. Notice that x < y, x > y, x ^ y are S j formulas, while
3 < y , x = y are 11° formulas. The basic properties of these

functions and predicates, those listed on pp. 19— 23, B&B, are easily
proved (in SC ). Ignoring the type differences we will also use notations
like 2 -f x, x > 1 and so on when x is a variable for real numbers.
The inverse function --1 is defined on R + U R “ , whose signature
is ((o —» o) ,o, o, o) because the defining formula is x 6 R + V x 6 R~,
and x € R + and x 6 R~ each contains one existential quantifier and

the disjunction V generates one more existential quantifier. So there


are three witnesses i,n ,m for x € R + U R “. t = 0 or 1, indicating
the element’s belonging to R + or R"; n and m witness the element’s
belonging to R+ or R" respectively, that is, x (n ) > 1 /n or x (m ) <
—1/m . Then •~1 operates on ( x ,i,n ,m ) , actually (cf. B&B, p.24)

•-1 = Xxinm.Xj.x (m a x (j, m ax(n, m ))m ax(n , m )2) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 85

We can verify that this is a function from R +UR~ to R and hence we can
simply write -_1 (x , i, n, m) as x _1 when it is verified that x € R + U R _ .
Propositions on pages 24-26 of B&B are proved easily. Here we
check the proof of Theorem (2.19) on p.27, the constructive version of
Cantor’s theorem. It uses some inductive constructions. It suffices to
show that for any rational numbers xo < yo and a sequence (an) of
real numbers, we can construct a real number x such that %0 < x <
yo and x ^ a„, for all n > 0. The idea is to construct a sequence
(xn, yn, kn), n = 1 ,2 ,..., where xn, y n are rational numbers and kn is
a natural number such that for any n > 0, (i) xn < yn, and (ii) kn
witnesses x„ > On or yn < a„, and (iii) xn < xn+1 and yn+i < yn, and
(iv) yn —xn < 1/ti. Then we can let x = (xn) . xo, yo are given and ko
can be arbitrary. Suppose that (xn, yn, kn) has been constructed. Then

(®n+i,2/n+i, &n+i) is constructed as follows: set any default values for


the case yn < x„. Otherwise, decide if On+1 > xn or an+i < yn, and
in the former case let xn+1 = xn and choose yn+i between xn and
m injon+i, xn + (n + l ) -1 }, and in the latter let y n+i = y n and choose
xn between max ja n+i , yn - ( n + l ) -1 } and yn . The witness kn+i is
obtained accordingly. The construction is obviously primitive recursive
and hence available in SC . Then a quantifier free induction shows that
(i) — (iv) hold.

6 .2. S eq u en ces and Series o f R ea l N u m b er s. A sequence of


real numbers is a function from N to R. As usual we use (an) to
denote a sequence and then On will mean the term (an) (n). The set of
sequences in R, .F(N, R ), will also be denoted as R N. The statement

‘(a,,) converges to y \ or lim n_ 0o a* = y is

Vfc > 03nVm > ndom —y\ < 1 /k ) ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 86

or equivalently,

3NVk > OVm > N (k) (loro - y \ < 1/Jfe).

N is the witness for the convergence, also called a modulus of con­


vergence, that is, ‘N is a modulus of convergence for limn-.® &n — y ’
is translated as the formula following 3 N above. Similarly, ‘(an) is a
Cauchy sequence’ is

VA: > 03nVi,y > n(|a* - ay| < 1/k ),

or equivalently,

3NVk > OVi,; > N ( k ) ( \a i - oy| < 1/As),

and N is a modulus of Cauchyness for (an) .


The proofs of Theorem (3.3) on p. 29 and Proposition (3.4) on
p. 30 of B&B are formalizable in SC directly, for no inductions are
used. From the proof we can see that, if N is a modulus of Cauchyness
for (a * ), the limit of (a„) can be constructed as a term containing a
and N . Actually, it is Am.a^ (2 m ), where M = max {3m, N (2 m )} .
Denote this term by limn-.oo an we can show that if N is a modulus
of Cauchyness for (a„), then (a,,) converges to limn_ 00an. However,
we should remember that limn_ O0 a„ actually involves a parameter N
which is supposed to be a modulus of Cauchyness for (dn), and which
does not appear in the notation, and which actually cannot be construc­
tively determined from the sequence (an) alone. We must be careful in
using such notations.
Suppose that (a„) is a sequence of real numbers. We want to
construct the partial sum £"_0 <x». Clearly, to approximate this sum
up to ± l / m degree of precision, it suffices to use the approximations

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 87

ai ((n + 1) m). So, we let

n n
£ a,- = Am. £ a,- ((n + 1 ) m ) ,
i= 0 i= 0

where for t [x] any term of the type o, £ "_ 0 1 [x] can be constructed as
a term which is intuitively the sum of t[ 0 ], t [n] as rational num­
bers. Notice that £"_0 Oi is a term with free variables a, n. It is easy
to show that if a,- G R for x < rx, then € R. We denote the
sequence An. £ ?=0 a* by a»- The latter will also denote the limit

limn_oo £?=o of the sequence ambiguously. This is the series con­

structed from a sequence. Notations a«> S S m a»> !C?=i t [*] ond so


on are defined similarly.
By definition it easily follows that if a* € R for x < n + 1, then

£?= 0 a. + Qfi+i =R Oi- Then, using the inductions available in SC


we can show that basic properties of partial sums hold. For instance,

(dn) 6 R N A (6„) £ R n implies

Vn (o„ —r bn) -» Vn ai = R £ M ,
\i= o 1=0 /

n m n
(m < n ) ~ * £ a , = R £ o , + 2 °*»
i= 0 i= 0 i= m + l

(Oi + &»)= R $ 3 <*i + $ 3


isO i= 0 i= 0

(Vx < n ( a i > bi)) -» o, > 6,.


i= 0 i= 0

The first conclusion is obtained by an induction with assumptions on


n, with a»- = r £"=0 48 inductive formula. It is a 11° formula.
The other conclusions are similar.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 88

The convergence of a series can be characterized as usual, that is,

VJfe > 03/V m ,n > I £


t= tn + l
a« < l/k

lL is a modulus of convergence for a,-’ is to mean Vfc > 0Vm,n >

L {k) (IZILto+i0* — 1/^) • Similar to the notation limn_ 00 a*, we let


denote (ambiguously) a term containing a and a parameter L

such that if L is a modulus of convergence for £ So ai> then X^L0 aj


converges to a^.
The divergence of sequences and series are defined straightforwardly.
Then the divergence of i~l and the tests of convergence and diver­
gence on pp 32— 35 of B&B can be verified: The proofs of Propositions
(3.5)— (3.7) do not involve inductions. For the proof of Lemma (3.8)
on p.34 one needs a relativized induction on n to show that

(yn) € R n A c € R A Vn (yn > 0 A n (y„y“+! - l ) > c)

“ » VlfVn1 > 1 + C ^ fc "1.


k -N

Finally, the proof of Proposition (3.9) does not involve extra inductions.
Basic results for series can be generalized to double series. The
partial sum 5Z"J=0 a«,j can be constructed as a term and we can show
that if Hij=a l°ijl converges (as a sequence) then £ “ o converges for
each z and £ ” o E " o a«j ^ so converges to £ “ =0 Oij, where £ “ =0 <h j
is a series obtained by fixing a (primitive recursive) arrangement of
( i , j ) , i , j = 0 ,1 ,..., into a sequence.
Finally, the constructions of constants e and t as limits of series on
p.35, B&B, involve no difficulties.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 89
6 .3 . C on tin u ou s F u n ction s. Continuity can be defined straight­
forwardly. We call an interval [a, 6] a compact interval. Here are some
(parametrized) sets of continuous functions:

/e C 7 ( [ a ,6 ] ,R ) = ( / : [ a ,6 ] - » R ) A
3wVx,y € [a, 6] Vn > 0 (|x —y| < u > (n )-+ \ f (x) - f (y)\ < 1 / n ) ,
f € C ((a, b ) , R) = ( / : (a, b) —* R)AVc, d £ (a ,b )(c < d —* f € C ([c, d], R )),
/ € C (R , R) = ( / : R -♦ R)AVc, d e R ( c < < £—► /€ ( ? ([c, d ], R)).
Notice that a modulus of continuity u here is of the type (o —> o ) ,
which is different from the definition on B&B p. 38. A member of set
C ([a, b], R ) has u as a witness among others. Further C ([a, 6], R ) and
others are all subsets of F ([a, b], R).
For A a set of real numbers, ‘A is bounded above’, ‘6 is an upper
bound of A ’, *b is a supremum of A’ and so on are defined naturally.
We use b > A to denote ‘6 is an upper bound of A ’ and use b = sup (A )
to denote ‘6 is a supremum of A ’.
Now we consider Proposition (4.3) on p.37 of B&B, which says, for a
non-empty set of real numbers A that is bounded above, the supremum
of A exists if and only if for any real numbers x < y, either y > A or
for some a € A, a > x. The necessity part can be proved directly, but
the sufficiency part involves some inductions. We prove it below.

THEOREM 6.1. We can prove in S C that if A is a set (of the


type (o —» o)), then (i) A C R, and (ii) 3 x (x € A), and (Hi) 3x €

R (x > A), and (iv)

Vx, y 6 R (x < y —> y > A V 3 z 6 A (z > x))

imply (v) 3x € R (x = sup (A )).

PROOF. The informal idea of the proof is to construct two se­


quences of rational numbers (an), (6n) such that (1) a*. < On+i < bn+ 1 <

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 90

bn and |6n+1 - an+1| < 11bn - an|, and (2) bn > A A 3x € A ( x > a^).
Then it is easy to see that we can construct the limit of (a„) and show
that the limit is just a supremum of A. Here we explain how to con­
struct the sequences and show that the assertions (1) and (2) follow
from the assumptions (i) to (iv).
For this we must spell out the existential assertions implied in the
assumptions (i) to (iv). (ii) plus (i) implies that for some xo, xo €
R A*io 6 A. (iii) means for some y0l yo € R A yo > A. (iv) implies

Va, 6 6 Q (a < 6 —> b > A V 3z £ A ( z > a ) ) ,

which in turn implies

( a < b —* \
3sV a,6 6 Q
^ (s (a, 6) = 0 A 6 > A) V (s (a, b) > 0 A 3z 6 A (z > a)) J

So, instead of proving (i)A(ii)A(iii)A(iv)—*(v) in SC , it suffices to prove


that if (i') A C R, and (ii;) x0 € A, and (iii') yo > A , and (iv')

Va, 6 6 Q(a < 6 —* (s (a, 6) = 0 A 6 > A)V(s (a, 6) > 0 A 3z 6 A [z > a ) ) ,

then (v), where xo,yo, s are free variables.


The sequences (a*) and (6n) are constructed by recursion: Let ao =

x o (l) — 2 and 60 = y o (l) + 2. Suppose that On and 6n have been


constructed. Denote a'n = On + \ ( b n —On) and 6J, = On -(- |( 6 n —On).
By the assumptions we can decide if 6^ > A or 3z € A (z > a'n ) by
deciding if s (a'n, b'n) = 0 or > 0. In the former case we let a„+i = a„
and 6n+x = 6^, and in the latter we let On+x = a'n and 6n+1 = 6n.
Clearly all this can be accomplished by a primitive recursion. So the
sequences (a n ), (6n) can be constructed in SC . From the construction
we have

On £ an+l < 6n+x < 6n,

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 91

i&n+l ~ an+l| < - |6n —a„| .


4

Then it remains to prove that (i')— (iv') imply


(1) 3 z E A { z > On), and

(2) in > A.
Here we prove (2). The proof of (1) is similar. Suppose that

z 6 A = BuVvx (u, v, z)

with x quantifier free. Notice that 6o > yo by our definition. Then (iii')
implies

3V 0VuVzVm > 0 (x (u, V 0 (u, z } m ) , z) -* i 0 > z (m ) — 1 /m ) .

Similarly, (i')— (iv7) imply

a 6 Q A 6 6 Q A o < i A s ( a , 4 ) = 0-+

( VuVzVm > 0 (x (u, V (a, 6) (u, z, m ) , z) -> b > z (m) — 1/m ) ;


.

So it suffices to show that (2) follows from (i')— (iv') together with
(iii") VuVzVm > 0 ( x ( u , V 0 (u, z ,m ), z) -* b0 > z (m ) — 1 /m ), and
a € Q A i £ Q A a < i A s ( a , b) = 0 —*

(
VuVzVm > 0 (x (u, V (a, 6) (u, z , m ) , z) —* b > z (m) — 1/m )
V 0 and V (a, b) above are the witnesses for bQ > A and b > A re­
spectively. To use inductions to prove (2) we must construct a sequence
T [n] such that T [n] witnesses bn > A for all n. First, construct a term

D [n] such that

D [ 0] = 0 ,

D [n + 1] = J (a « , O , n + 1, D [n]),

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 92

that is, D [n] is the minimum m such that bm = 6m+i = ... = bn. Let

T [n] s J (D [n], V 0, V 6'0W_ ,) ) .

It suffices to show that (i')— (iv'), (iii"), (iv"), and m > 0 imply

X(u, T [n] ( u ,z ,m ) , z) -» bn > z (m) - 1/m ,

that is, T [n] witnesses bn > A. For this, by the quantifier free induction,
it suffices to show that the assumptions imply

X (u, T [0] ( u ,z ,m ) , z) -* b0 > z ( m ) - 1 / m ,

(5) (X (“ . T [n] (u, z, m ) , z) -» bn > z (m) - 1/m )

— (X (u, T [n + 1] (u, z , m ) , z) -» bn + 1 > z ( m ) - l / m ) .

The first follows from (iii") and the definition of bo and T. As for the
second, we distinguish between s (a'n, b'n) > 0 and s (a'n, b'n) = 0. In the
former case, D [n + 1] = D [n], 6n+i = 6„, and (5) becomes a formula of

the form <p —►(p after substituting D [n] for D [n + 1] and bn for 6n+i •
In the latter, D [n + 1] = n + 1 > 0 , and i = Ki = bn+ i• Then,
taking a = a'n, b = b'n in (iv"), it follows directly that

X (u, T [n + 1] (u, z, m ) , z) A m > 0 —> 6n+i > * (m ) - 1/m ,

and hence (5) follows from (i')— (iv7), (iii"), (iv") and m > 0. This
completes the proof of (2). |

Corollary (4.4) to Theorem (4.8) on pages 38— 41, B&B, can be


proved without further application of induction. The definitions of
f + S> f g, | / | , max { f , g } , g~l and so on on p. 39 are straightforward.
Notice that the term representing g~l should contain a parameter that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 93

is supposed to be a witness for Vx 6 A( g ( x ) / 0 ). The remarks fol­


lowing Theorem (4.8) on p. 40 are also valid, for the conclusion follows
from Proposition (4.3) without further application of induction.
For I an interval, a sequence of functions in C ( / , R) is a function
from N to C ( / , R). We will use the notation ( / n). Notions like ‘conver­
gence (uniformly) on I to g ’ and ‘Cauchy on / ’ are defined naturally
and the basic properties (Theorem (4.11)) about them are trivial.
Given a sequence of functions (/„ ) in C ( / , R), the corresponding
series is defined as the sequence
n
(gn) = An.Ax. £ fi ( x ) .
«=o

We will use U to denote the series. We must prove that gn =


Ax. £"_0 fi (x) 6 C ( / , R). The assumption Vn (/„ € C ( / , R)) implies
that for some u>, u (n) is a modulus of continuity of f n. By primitive
recursions we can construct a term r such that

r [n ](m ) = m in{u/(0) ((n + l ) m ) , ..., u ;(n )((n + l ) m ) } .

Then it is easy to verify that r [n] is a modulus of continuity of gn.


The comparison tests and ratio tests for convergence are easily
proved to hold. Finally, power series are defined naturally and the
basic properties (Proposition (4.12) on page 43) are trivial.

6 .4 . D ifferen tia tio n . For I a compact interval, ‘g is a derivative


of / on P is translated as

^ € C (J, R ) A / € C ( / , R ) A

3S ix , y € /V n > 0

j ( n) > 0 A | ' I—
\ f ( y ) - f ( x ) - g ( x ) ( y - x ) | < |y - x |/n

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 94

We use the notations g = f , g = D f , and g (x) = as well. Then ‘/


is differentiable1 is 3g (g = / ' ) and l6 is a modulus of differentiability
for / ' = g' is the formula with ‘3S’ dropped. The basic properties
of differentiation, Theorem (5.2) to Theorem (5.6) on pp 44-48 are
straightforward.
For n a variable, ‘g is the n-th derivative of / ’ is translated into

- geC(I,R)AfeC{I,R)/\

3F ( F ( 0) = / A Vi < n ( F ( i + 1) = F (i)') A g = F { n j ) ,

where F is intuitively the sequence / = ..., . We also use


the notation g = f^nK Notice that F is a witness for 3g (g = / ^ ) , that
is, ‘/ ’s n-th derivative exists’, and it actually contains all derivatives of
/ up to the n-th. So, we should also be careful in using the notation
/ ( n) by itself, not in the context g = /("). As a term it is just F (n) and
then we can show that if F witnesses 3g (g = , then F (n
where the latter is translated as above. It is more natural to see ^
as a function from C n ( / , R) to C ( / , R ) , where Cn (I, R) is the set of
functions on I whose n-th derivative exists. So, / ’s n-th derivative is
among the witnesses for / 6 C n ( / , R) and the function just extracts
the witness as the value. Of course, when / is given as an elementary
function, for instance, a polynomial, sin, cos, e* and so on, / ’s n-th
derivative can be constructed directly from / and n. Then we can take
/W simply as a term with n as the only free variable.
For the proof of Taylor’s theorem we need

Vi (/,- is differentiable) -» f e / i ) = 5 1 ?**


\»=o / i= 0

which can be proved directly as in the proof of the continuity of £"=o fn-

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 95

Notice that Vn ( / is n times differentiable on 7), or / € C°° (I , R ) ,


implies 3GVn (Gn = So, under that assumption, the Taylor se­
ries ( i —a)n can. be constructed, using the witness G, and
its convergence to f ( x ) can be proved under the ordinary assumptions
together the assertion that G is the witness. Similarly, in case G can be
constructed from / directly, which is the case for all concrete functions
we treat in analysis, the notation for Taylor series can be used freely
as in the informal language.

6.5. Integration. A finite sequence of real numbers P = (a0, ..., On)


is a partition of an interval I = [a, i] if a = ao < ax < ... < an = 6.
We can represent the Riemann sum £ ”r0x / (a,-) (a,+x - a,) by a term
5 ( / , P) and show that for / 6 C (7, R), P a partition of [a, 6], we have

S ( f , P ) e R.
To define the integration, we choose a sequence of partitions (Pn),
Pn = (a, ..., a + ij-^ , ..., 6). It can be proved that a < b and / 6
C (7, R) imply limn_ O0 S ( / , P„) exists. Then we can construct a term
T[/,u>, a, b] such that if a < 6, / : I —►R, and a> is a modulus of
continuity of / on [a, 6], then T[ f , w, a, 6] = limn-,,* S (f , P n). Notice
that the construction of the limit of the sequence ( 5 ( / , Pn)) needs the
witness for its being a Cauchy sequence, and that in turn depends
on the modulus of continuity of / . So u must appear in the term.
However, it is clear that if w' is also a modulus of continuity of / on
[a, b] then T [/, u/, a, b] = r T [/, w', a, 6]. So we simply use the notation
f t f ( x ) d x . On the other hand, we can take integration as a function
from the set C (7, R) to R, for a function operates on the members
of C (I, R ) together with their witnesses, which include the moduli of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 96

continuity. For arbitrary o, b 6 R, we define

J f(x)dx=J f(x)dx — j f(x)dx,

where c = m in(a, 6). When a < b , this is equivalent to the above case.
The basic theorems about integration, Theorem (6.3) to Theorem
(6.10) on pp 51— 55 are easily proved, no novel inductions are needed.
But notice that the formulation of some theorems needs some caution.
In one way they may involve the relevant moduli of continuity as free
parameters. For instance, the fundamental theorem of calculus can be
stated this way: I f / : / —►R, is a modulus of continuity for / on I,
and a £ I, then the term A x ./“ f ( x ) d x , with a, f , u as free variables,
is a continuous function from I to R and / is its derivative. We can also
eliminate w as a free variable by using extra existential quantifiers. For
example, the theorem can also be stated as: If / 6 C (I, R ), then for
some uj, u is a modulus of continuity of / a n d There will be similar
problems whenever notations implicitly involving some witnesses are
used in stating a theorem, and they can be solved in a similar way. We
will not repeat this comment afterwards.

6.6. Certain Important Functions. Functions e*, ln x , sinx,


cos x are defined as limits of series or integrations; cf. pages 55— 58 of
B&B. The basic properties of these functions are easily proved, except
Theorem (7.15) on page 59, which constructs the first positive zero of
cosx. Here we need the fact that cosx can be approximated up to
± l / m degree of precision by an expression like (cos (x (t [m]))) (s [m ]),
where t [m ], s [m] are some terms. This is generally true for all contin­
uous functions. Suppose that u is a modulus of continuity for / on I.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission .
6. CALCULUS AND THE REAL NUMBERS IN SC 97

Then, clearly, for x € I and m > 0,

1/ (x ) - / (* (w (2m ))) (2m)| < 1/m .

To construct the first positive zero of cosx, we first construct a


sequence of real numbers rn such that

i"o = 1,

rn+i = r„ + cos r„.

This is obtained by the following recursion which is reducible to prim­

itive recursions:

rn+1 (m) = rn (2m) + cos(rn (2u (4m ))) ( 4 m ),

where a; is a modulus of continuity for cos on R such that ui (m ) > 0 for


all m. A induction on n gives rn € R, and then rn+i = rn + cos rn

follows.
Then we use an induction on n to show that

(6) cos rn > 0 A r , i > 0 A V g € Q ( 0 < g A g < r n —►cos q > 0) .

Notice that this is a formula, so the induction rule is available in


SC . The initial step n = 0 can be verified directly. For the inductive
step, supposing (6), it follows that rn+1 > rn > 0. For any x > rn we

have

cos x = cos Tn — f
Jrn.
sin tdt.

From cosrn > 0 we have |sinrn| < 1 and then the continuity of sin
implies that / r* sin tdt < x — rn. In particular, for x such that rn <
x < rn+1 we have

cos x > cos rn - (rn+1 - rn) = 0.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 98

So (6) holds with n replaced by n -f 1. Therefore (6) holds. By the


continuity of cos, it follows that for i € R, 0 < i < rn implies cos x > 0 .
Since sin' i = cos i , cos' x = —sin x, by Rolle’s theorem, it follows that
for x 6 R, 0 < x < r« implies sin s > sinO = 0. Applying Rolle’s
theorem again we have cosx > cosrn > 0 for 0 < x < rn. Rolle’s
theorem also implies that

rn+i - r„ = rn - rn_i + cosrn - cos rn_x

is arbitrarily close to (r„ —rn_ i) (1 —s in i) for some x € [rn_ i,r n]. For
n > 1, we have x > 1 and hence sin s > sin 1. So,

rn+i - r„ < (1 - sin 1) (rn - rn. t ) .

Then some inductions show that (rn) is a Cauchy sequence. The rel­
evant inductive formulas are all 11° formulas, as the formula above.
Finally, it is trivial to show that the limit is the first positive zero of
cos. It is denoted by ir/2. Thus we have proved Theorem (7.15) on
page 59 of B&B.
The above proof also shows that sin is strictly increasing on ( —7r/2, it/ 2),
so arcsin can be defined on ( —ir/2,7r/2). The relevant properties are

easily proved.

6 .7 . A F inal R em ark . The last proof is bit more complicated


than the original proof in B&B, because we have to show that the
inductions are reducible to those available in SC . On the other hand,
the informal language for stating the theorem and the proof is exactly
like that of classical mathematics. A proof in SC largely corresponds to
a proof in classical mathematics in which all existential assertions are
proved by providing relevant prim itive recursive algorithms. A proof
of a theorem in SC certainly provides more information.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
6. CALCULUS AND THE REAL NUMBERS IN SC 99

The development of mathematics in S C will continue in the appen­


dices, where it is proved in Appendix A that, with a few exceptions,
Chapter 4 through Chapter 7 of B&B can be developed within SC.
Then it is proved in Appendix B that the basic theory of unbounded
linear operators on Hilbert spaces can be developed in SC.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
CHAPTER 2

Toward a Philosophy of Mathematics

This chapter explores a philosophical account of mathematics that is


inspired by the technical results presented in Chapter 1 and the appen­
dices. Note that the t e c h n ica l results are not final. At this point, it is
known only that some significant part of analysis and some basic math­
ematical tools used in quantum mechanics can be developed within
strict constructivism. However, this already suggests some philosoph­
ical points. I will explore these in this chapter. The basic idea of this
chapter has been introduced in the Introduction to this dissertation, so
it will not be repeated here. Below, Section 1 contains an exposition
of fictionalism. Section 2 explains how mathematics in strict construc­
tivism can be applied to reality, and how references to fictional things
can be eliminated and the applicability can be explained in such appli­
cations. Finally, Section 3 contains some remarks on the epistemology
of arithmetic. In particular, it suggests that the proper epistemological
basis of arithmetic is perhaps a compromise between Mill’s empiricism
and some Kantian views.

1. Fictionalism

1.1. Introduction. The kind of fictionalism defended here may


have a long history. I will not try to pursue that in this dissertation,
100

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 101

but below are some publications that express views close to fiction­
alism and that inspire the exposition given here. Azzouni [3], Bala-
guer [5], Maddy [66] and Sober [96] share a common idea in rejecting
the Quine-Putnam indispensability argument for realism. They hold
that the references to mathematical entities in scientific theories do
not commit scientists to the objective, mind-independent existence of
those entities. This is one of the basic ideas behind the type of fiction­
alism presented here. Papineau [71] also defends fictionalism. Leslie
Tharp’s conceptualism [99] [100] is also close to the kind of fiction­
alism exposed here. Mathematicians Mycielski [69], Renyi [86] and
Robinson [91] express similar views1. There are also connections with
fictionalism about possible worlds. See Gideon Rosen [92] [93].
The name ‘fictionalism’ was originated by Hartry Field for his po­
sition expressed in [44]. There are some differences between Field’s
views and the position presented here, which will be explained near
the end of this section. The name ‘fictionalism’ is retained, because it
accurately describes the position presented here. That is, the ontolog­
ical status of mathematical entities is like that of fictional characters
in stories and a mathematical theory is a story about those fictional
entities.
This name may cause some misunderstandings. It hints that math­
ematics is an arbitrary, whimsical creation, which certainly induces
suspicions. How could such a fiction be so useful and become the para­
digm of knowledge for thousands of years? But Renyi’s dialogues have
answered this sort of question: Mathematical entities provide general

1Renyi’s dialogues [86] are the best informal description of the kind of fictional­
ism I want to defend here. This section can be seen as an exposition of Renyi’s ideas
with the contemporary philosophers of mathematics as the intended audience. I
did not know Renyi’s interesting and eloquent dialogues until I read Mycielski [68].

o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.


1. FICTIONALISM 102

and simplified models of nature. Presumably, some fictions are quite


arbitrary, but some fictions are created with specific purposes, and as
such they conform to some special constraints. Mathematics should
belong to the latter. Because of the constraints, it is true that in
some sense mathematical entities are not as ‘non-real’ as the ordinary
fictional characters, but fictionalism emphasizes that notwithstanding
those constraints, mathematical theories are still fictions, not collec­
tions of truths about some mind-independent entities. Because of the
constraints, the difference between fictionalism and realism, in particu­
lar, some sort of more sophisticated (not so full-blooded, not too naive)
realism, becomes subtle in some respects. But the basic ontological
point should be clear: Mathematical entities are not mind-independent
entities. They are unlike stars or electrons when we come to the onto­
logical and epistemological questions about them. The purpose of this
section is just to illustrate the subtle differences and to explain why
fictionalism is perhaps in a better position than realism.
I will give a positive and straightforward exposition of fictionalism.

The exposition will emphasize that fictionalism is very close to common


sense. It is less assertive and it should be more acceptable to ordinary
scientists than realism is. It is not an exotic view that only philoso­
phers will like to entertain. In particular, I will emphasize that it is
consistent with commonsense realism about the outer world and sci­
entific theories. Moreover, it will be argued that fictionalism makes it
possible to integrate good insights from other positions in philosophy of
mathematics, such as formalism, structuralism, modalism, if-thenism
and so on.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 103

Views close to fictionalism have been expounded by many philoso­


phers and some of the ideas are probably just common sense. There­
fore it is impossible to cite the source of every idea presented in this
section. The references to the published literature in the following ex­
position aim more at indicating differences than at crediting sources.
This exposition is better to be judged as an attempt at a fairly com­
plete exposition of fictionalism, rather than an exposition of an original
philosophical position. Though, the points cited in the last paragraph
which I want to emphasize are not seen to have been emphasized in
the literature. I did not know other detailed discussions of the in­
determinacy of fictional entities and that of truth and consistency in
fictionalism, which will be given in Subsection 1.4 and 1.6 below. More­
over, the discussion on the semantic function of numerals in Subsection
1.3 and the analysis of (the false appealingness of) the indispensability
argument in Subsection 1.5 are perhaps new. But it is possible that
none of these is really original. They may have been implicit in some

literature I cited above, or in some other literature.


Finally, in my exposition, I will try to avoid unclear philosophical
jargon and assertions that might obscure the simplicity and naivete
of a view like fictionalism. Moreover, some questions not closely re­
lated to the main thesis will not be discussed. In particular, I will
not discuss what exactly is the ontological status of fictional charac­
ters, for the focus is on the question whether mathematical entities are
mind-independent entities.
A summary of the basic points of fictionalism will be given at the
end of this subsection. Then Subsection 1.2 focuses on geometric figures
and sets, explaining how they are fictional. There will be a discussion
on the concern that mathematical entities are timeless and therefore

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 104

cannot be creations of minds, as well as a discussion on the concern


that mathematics is the most certain knowledge we have and cannot
just be fictional. Subsection 1.3 treats the case of natural numbers,
which is a little more subtle. Subsection 1.4 discusses a special feature
of fictional things, their indeterminacy, which is in some sense an in­
trinsic distinction between fictional things and real things. Then I will
discuss two other objections to fictionalism. The first is the indispens­
ability argument in Subsection 1.5. The basic ideas in the discussion are
inspired by the publications of Azzouni, Balaguer, Maddy and Sober
cited above, but I will give an analysis of the indispensability argu­
ment to explain why the apparent appealingness of it is an illusion.
Next, Subsection 1.6 discusses the objection from consistency raised
by Resnik [87]. Finally, Subsection 1.7 contains some comparisons be­
tween fictionalism and other positions. The objective is to show that
fictionalism can share many insights of other philosophical positions.
In the rest of this subsection, I want to clarify some points. Then
a summary of the basic points of fictionalism will be given.
(1) First, the focus here is on the presentation of fictionalism as a
coherent and common-sense view. It is not to refute realism in math­

ematics. As for realism, it seems that we still need a positive and co­
herent formulation of the position that can answer the common doubts
about it. That is worth trying. Realism is an ideal that dates back to
ancient time, and the debates between realism and its opponents may
last forever. The only point that I do believe to be wrong is the mod­
ern Quinean claim that modem science has already justified realism in
mathematics.
(2) Second, I want to make it explicit that the debate between
realists and their opponents is a genuine debate. Sometimes, arguments

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 105

in philosophy of mathematics seem to assume that most scientists and


mathematicians accept realism in mathematics, and it is only a few
philosophers who are trying to reject it. If that were indeed true, any
anti-realistic views would be in a bad position from the start. Anti-
realists would be like skeptics, arguing for something which no one

would take seriously.


Certainly we can take a poll to see what ordinary scientists believe,
but we must be careful in putting the questions. We cannot simply
ask ‘Are there numbers?’ This is too ambiguous. An answer ‘Yes’
might just mean that there are numerals, or that there are numerical
properties, or that we can talk about numbers as if we are talking about
objects. One basic idea of the type of fictionalism I want to support is
just that we do talk about fictional things constantly in everyday life
and science, and there is no need to avoid that. Realism is the view that
numbers are abstract, mind-independent objects, or at least that the u i
sequence is a mind-independent structure, and moreover meaningful
statements about them are either true or false, and the truth or falsity

is a matter of fact. The trouble is that ordinary people do not seem to


have a clear notion of abstract, mind-independent objects or structures.
We must carefully put our questions in such a way that an answer ‘Yes’
will be definitely a realistic answer.
The following may be some such questions: (1) ‘Supposing that
the universe is closed and finite, so that Euclidean straight lines or
planes cannot be embedded into the real space-time, do you still believe
that Euclidean lines, planes, and the 4-dimensional Euclidean space
which we use to represent space-time in Newtonian mechanics exist
independent of mind, independent of our thinking of them, and that
they would have existed even if human beings had never invented (or

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 106

discovered?) Euclidean geometry?’ (2) ‘Conceding the possibility that


the universe may be finite and the space-time may be discrete, so that
the structure of w-sequence could not be instantiated in the universe,
do you still believe that the structure of u/-sequence is objective, mind-
independent?’. For mathematicians we can certainly ask this question:
‘Do you believe that the continuum hypothesis must be either true or
false, and that its truth or falsity is a matter of fact for us to discover?’
In these questions, we avoid the general notion of mind-independent
abstract object, but we also carefully distinguish the supposed mind-
independent abstract mathematical objects from the concrete things
in the universe. This excludes the possible misunderstanding that we

are asking about structures of some aspects of the real world. Realism
demands the answer ‘Yes’ for these questions.
I did not try taking the poll, but here I will assume that there will be
no unanimous answer from scientists. Probably, most scientists will de­
cline to give answers, because the questions are typically philosophical
questions. They are not the type of questions that scientists constantly
ask themselves and try to answer by themselves. If philosophers ask
physicists why they believe that there are atoms, molecules, electrons
and so on, physicists will have a lot to explain, but the questions given
above do not seem to concern physicists. Even if some of them happen
to be interested by the questions, it seems that bewilderment will be
a more com m on reaction than a prompt realistic answer. Moreover,
physicists tend to call mathematics a formalism, or a language, while
they generally have a solid sense of reality for things in the universe. So
I assume that there may be more physicists who are willing to answer
negatively than those who tend to answer positively.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
1. FICTIONALISM 107

The Benacerrafic difficulty for the naive picture of realism is well


known. Realists certainly want to reject the picture, that is, the pic­
ture that abstract objects dwell in a different universe cut off from us.
However, in the questions above, the mathematical entities and struc­
tures are infinite and the universe is assumed to be finite. Then it is
very explicit that the mathematical structures cannot be manifested
or be innate in the universe in any way. Therefore, it becomes hard
to avoid the naive picture if one still claims that those structures are
mind-independent. That is the reason why the questions given here
may be less likely to get positive responses.
As for mathematicians, there is a folklore that ordinary mathemati­
cians are realists in their everyday work, but they become formalists
when pressed. It seems that they are more conscious about making
correct statements when pressed than not. So, if the folklore is indeed
true, probably most mathematicians should not be classified as realists.
So, in this section, I will take it that there is a genuine open ques­
tion here. Realism in mathematics is not part of common sense and
rejecting it is not like skepticism. Then, I will argue that if scientists
are bewildered by the questions given above, the answer fictionalists
provide will make them more comfortable. It is fictionalism in math­
ematics, rather than realism, that is closer to our scientific common
sense, although realism can perhaps still stand self-consistently as a

philosophical position.
(3) Thirdly, I want to emphasize that the discussions in this section
always assume the common-sense realistic attitude toward the outer
world and scientific theories. The assumptions about the finitude of
the universe, the existence of atoms, electrons and so on are all under­
stood literally. Since philosophers of mathematics, in particular, those

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 108

realistically inclined philosophers, generally tend to hold that realism


in mathematics is more akin to scientific realism than the opponents,
the attitude here can be taken as a concession for the sake of argument.
Finally, the basic points of fictionalism are summarized as follows:
Mathematical entities and structures such as numbers, sets, functions,
groups, topological spaces, points, lines and so on are fictional entities
or structures. Their ontological status is like that of fictional characters
in stories. A branch of mathematical theory is a story about those
fictional entities or structures.
Because mathematical entities are simple, stories about mathemat­
ical entities can be made rigorous. The most extreme form of such
rigorization is the axiomatic approach. Mathematicians lay down the
basic assumptions about those fictional things, the axioms, and then
deduce consequences from the axioms. In deducing the consequences,
mathematicians follow the logical laws we ordinarily follow when we
reason about real things. But usually the intuitive imagination of the
fictional entities precedes the axiomatization, and the axioms can be re­
vised or extended according to our possibly very vague intuition about
the coherence, simplicity and possible applications of the story.
A scientific theory, in particular, a physical theory, is frequently pre­
sented as a story about mathematical objects and structures combined
with other fictional objects or structures that are abstractions or ide­
alizations of physical objects or structures in the real world. Examples
of the latter are Newtonian particles, quantum mechanical particles,
ideal gases, rigid bodies and so on. This is a way of stating the theory
in a general format. When such a theory is applied to the real world,
some of the fictional entities or structures are interpreted as real phys­
ical objects, structures, or properties, and some assertions about the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 109

fictional entities in the theory are translated into assertions about real
things. In particular, numbers are usually combined with physical units
to represent physical quantities.
Fictional things are used to build models to simulate aspects of na­
ture. The success of the application of a model depends on whether
or not the model provides a sufficiently accurate description of nature.
The question if the model itself is a mind-independent object is irrel­
evant. Fictional models can equally provide accurate descriptions of
nature. Moreover, there is no need to avoid conceiving of and using
fictional objects and structures in representing nature. As a mater of
fact this is done ubiquitously and it might even be practically indis­
pensable, because everywhere we have to make idealizations, to ignore
trifling details. The epistemological standards assumed by scientists
never forbid using fictional things such as fnctionless planes and ideal

gases to model nature.


When we reflect upon how we first think of things such as an u
sequence or an infinitely long line, we realize that, as a matter of fact,
we first use our imagination to imagine something, for we can only
observe finite things, and it is only after that do we start to ask if
they are not just products of our imagination, and if they are indeed
something actually independent of our minds out there and discovered
by us. Conceiving infinite mathematical entities comes first and the
leap to realism is secondary3. Now if there are some difficulties in
realism and there is no need for it in accounting for the applicability
aWe cannot really imagine an infinite structure. All we can do is to feel that a
sequence can be extended and extended. At some point our imagination becomes
vaguer and vaguer, but we cannot imagine a lim it for the extension either. We just
use words such as ‘so on and so forth', ‘repeatedly forever’ to express our feelings
and call it potential infinity. Sometimes we also talk as if such a potentially infinit
process could be finished. We talk, feel, and imagine first. Only after that do we
start to ask if there are some really infinite structures out there for us to discover.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 110

of mathematics in science, then, compared with fictionalism, realism


seems to be making some not fully warranted but redundant and more
assertive assertions. That is how fictionalism is perhaps less assertive
and more acceptable to ordinary scientists.
This is a su m m ary. The rest of this section contains more detailed
exposition.

1.2. G e o m e tr ic F igu res and S e ts . In this subsection I will dis­


cuss how geometric figures and sets are fictional. I will also discuss the
concern that mathematical entities are timeless and therefore cannot
be creations of minds, as well as the concern that geometric knowledge
is the paradigm of certain knowledge and therefore cannot be fictional.
First consider geometric figures. Fictional mathematical entities
and structures are sometimes idealizations of and abstractions from
real things. Geometric figures are typical examples. We observe the
shape of a piece of paper and then we conceive of a rectangle, although
we know that nothing in reality could be exactly like a rectangle. Real
things always have fuzzy boundaries. Moreover, real space is curved.
A rectangle is much simpler than the real shape of the piece of paper.
As a matter of fact, there is no such thing as the real shape of the
piece of paper. However, taking the rectangle as the shape of the
piece of paper makes the estimation of the area of the paper much
easier. We conceive of the perfect rectangle by imagining that the
boundaries of it are absolutely clear cut, that is, they are lines with
zero width, and imagining that its angles are strictly rectangular and
so on so forth. Fictionalists claim that the rectangle itself is a fictional
entity, a creation of minds. We create the figure just as novelists create
fictional characters with the intent to reveal some aspects of human
personalities.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. FICTIONALISM 111

Note that the real piece of paper with its fuzzy shape is real and
mind-independent. Fictionalists only claim that the perfect rectangle
is our fiction. This is an assertion about the ontological status of the
geometrical figures. It is the opposite of Platonism, which claims that
geometrical figures axe mind-independent entities. One can press the
point that the perfect rectangle might not have instances in the real
world, and then ask how could it be that it exists independent of mind,
even if no one ever thinks of it. Of course, realists do not mean that
geometric figures must be instantiated in the real world. However, the
Platonic notion of an abstract, mind-independent rectangle seems to
be less clear than fictionalists’ explanation that we imagine the perfect
rectangle. Realists might be right in claiming the mind-independent
status of the perfect rectangle, but fictionalists’ account is less assertive
and more easily acceptable.
Sometimes generalizations, extensions and conscious exercises of
our imaginative power are involved in conceiving of fictional entities
or structures. We observe some physical lines. Then we conceive of
a geometrical line segment as an entity with no breadth. Abstraction
is the way by which we create the fictional line segment here. Then
we imagine that the line segment can be arbitrarily extended, up to
infinitely long. Here we consciously use our imagination. Now we know
that Euclidean infinitely long lines might have no instances in the real
physical space. However, we can (and we do) still conceive of Euclidean
straight lines. Fictionalists take the description here seriously: we
conceive of the infinitely long lines. They reject the characterization
that we are perceiving something (which presumably should be infinite)
independent of the mind and they reject the leap to the assertion that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 112
there are some ‘true’, mind-independent lines, not just the lines we
conceive of.
Similarly, here we must distinguish the real structure of space-time
from geometric figures. Fictionalists only claim that the geometric
figures are fictional. We conceive of them in order to construct a model
to represent space-time. Our geometry may or may not represent space­
time faithfully, which is a question for physicists to answer, but either
way'the geometric objects are fictional.
No matter if we are realists or not, it seems that we can always con­
ceive of geometric figures, or we can always talk as if there were such
things. This is enough for our purpose of describing space-time. Real­
ists want to claim that there are mind-independent geometric figures
and it is not just that we are imagining something. This is a stronger
assertion. They must provide reasons for the claim, or at least show
that the stronger claim is inevitable in some sense. Therefore, if there
are some difficulties in the realistic picture of geometric figures as mind-
independent entities and the indispensability argument can.be rejected,
fictionalism will be more acceptable.
Now consider sets. We observe some physical objects. Then in our
mind we abstract the objects from their actual context and imagine the
collection of them. That is a set. We further conceive of the sets thus
obtained as individual objects that can be collected into other sets. In
this way, beginning with a single physical object, we can imagine the
set of the object, the set of the set of the object and so on. Then we
imagine that this process can be carried out forever and we can still
collect those infinitely many sets into a single set. Here we reach the set
belonging to K,+1 — Vu in the cumulative hierarchy. Then we imagine
that we can go on further and further. Fictionalism holds that all these

R e p ro du ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
1. FICTIONALISM 113

sets are fictional. We literally conceive of them, and moreover, in doing


this, we consciously exercise our imagination.
Cantor once defined a set as a collection of objects collected by
minds, and fictionalists would add that sets are thus fictional. It seems
that it is only after we have been so used to talking about sets that we
unconsciously slide from the recognition that we can talk about sets
to the belief that sets are always totally independent of the mind, are
something for us to discover.
Conceiving fictional things is very common in everyday life and
scientific research. When scientists present a scientific theory, they
conceive of fictional things such as frictionless surfaces, ideal gases,
mass points, absolutely rigid bodies, absolutely elastic balls, perfect
fluids and so on. Geometric figures are clearly very much like these

fictional and idealized things. We can conceive of a cube. In geometry,


a cube is not supposed to be solid, but certainly we can conceive of
a solid cube with some positive mass in mechanics. The color and
the material may be irrelevant at this point, but certainly in some
situations we do talk about a red solid cube or an ice cube in a perfectly
cubic shape . If the latter are fictional, the geometric figures should
be the same. Geometric figures are much simpler, more rigorous and
more universal. That may be the reason why we tend to take them as
mind-independent.
Newtonian mechanics is presented as a theory about systems of
mass points. When we apply the theory to describe the motion of the
earth around the sun, we take the sun and the earth as mass points,
while when we apply the theory to a sample of gas, we take molecules
as mass points. Mentioning fictional mass points is a way to present
the theory in its most general format. When the theory is applied to

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 114

reality, fictional things will be replaced by real entities, structures, or


properties.
Fictional entities are not created arbitrarily. Geometric figures are
to simulate the shapes of real things. Mathematical entities and struc­
tures are intentionally conceived to represent some aspects of nature.
The word ‘fictional' is to express the ontological status of those entities.
It should not hint that they are whimsical or they have no connection
with reality.
There is a concern that geometric figures or sets are supposed to
be timeless and therefore cannot be creations of minds at a particular
time. Here we must distinguish two points. Within the story about
fictional mathematical entities, we certainly speak as if those entities
are timeless. We never imagine that those entities start to exist at
some point or end existence at some point. The same is true for other
fictional entities, such as frictionless surfaces, ideal gases, perfect ice
cubes and so on. But on the other hand, there is no problem in saying
that people started to imagine perfect circles, squares, or cubes and
so on about twenty-six hundred years ago. If indeed those geometric
figures are mind-independent abstract objects, of course they are really
timeless, not just timeless in the story. However, this argument clearly
begs the question and fictionalists need not worry about it.
To see another example, consider the supposed abstract entity, the
letter ‘a’. Was it created at some stage of history? The answer seems to
be positive. It sounds counter-intuitive to say that the letter ‘a’ already
existed five thousand years ago, or that its existence is timeless. We
started to take the mark ‘a’ as referring to a letter at some particular
point in our history. We created the letter. Similarly, there seems to be
no harm in saying that we create geometric figures. The only difference

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. FICTIONALISM 115

is that the choice of ‘a’ is quite arbitrary, while the rectangle is in some
sense a very natural abstraction of the real shapes.
Finally, there is the concern that geometric knowledge, or generally
mathematical knowledge, is the most certain, most rigorous knowledge
we have, and therefore it is inconceivable that such knowledge would
be fictional. Mill had given the answer to this question in Chapter V,
Book II of A System of Logic. The same answer is also given in the first
dialogue of Renyi’s [86]. Basically, it is this: The best explanation for
the certainty of geometric knowledge is just that it is about fictional
things. To use modern logical language, we ourselves lay down the
axioms to characterize the fictional things and then deduce the conclu­
sions. As long as we proceed carefully, there are no other sources for
the errors. On the other hand, any assertions about real things in the
real world cannot be so certain. Today we know that when we try to
speak of things very far from our experience, for instance the funda­
mental particles, even the simple logical laws in classical propositional
logic may become uncertain. Absolute certainty can be obtained only
for the world we intentionally conceive of.

1.3. T h e C ase o f N u m b er s. I have tried to explain how geo­


metric figures and sets are conceived. The case of natural numbers is
a little more subtle. Here I will first give an analysis of the semantic
function of numerals that differs from the Fregean analysis. This will
help to make clear how numerals are really used in mathematical ap­
plications. Then I will describe how we come to talk about numbers
as if we are talking about objects.
According to the Fregean analysis, numerical properties are second
order properties, or put slightly differently, properties of sets. ‘There
are three apples on the table’ means that the concept ‘apples on the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 116

table' falls under the second order concept 3, or that the set of apples
on the table has the property 3. This analysis has a problem in dealing
with sentences such as ‘The apples on the table weigh 3 pounds'. Here
it is unclear of which set 3 is to be a property, or which concept is
supposed to fall under the second order concept 3.
It seems that ‘3 pounds’ in this example simply signifies a property
of the apples on the table, not the set of the apples, but the physical
stuff, or the mereological sum of the apples. ‘3* by itself does not
signify a property of the physical stuff, but ‘3 pounds’ does. Similarly,
‘3 inches’, {3 square feet’ and so on all signify properties, ‘pound’,
‘inch’, ‘square foot’ and so on are called units. This suggests another
way of analyzing the semantic function of numerals: Numerals plus
units signify properties. This analysis seems to be consistent with what
appears in physical applications. In physics, physical quantities are
represented by numbers together with units. Units are essential in
expressing physical quantities.
Now this analysis can also cover the cases such as ‘There are three
apples on the table’. We know that the set of the apples is not the
same as the mereological sum of the apples. The mereological sum
can be cut up in many different ways, and the set is determined by
one special way of cutting the mereological sum. The Fregean analysis
takes the set as an object and claims that ‘3’ signifies a property of
that object. The alternative would be this: The mereological sum,
or the physical stuff, is the only object involved here, and this object
has the property expressed by ‘3* together with the way of cutting up
the stuff into parts, that is, cutting it into separate apples. This is
similar to ‘3 pounds’, ‘3 inches’ and so on. The word ‘apple’ functions
just as the units such as ‘pound’, ‘inch’. The only difference is that

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 117

the way of cutting the mereological sum signified by ‘apple’ is unique.


This gives a uniform explanation of the semantic function of numerals
that encompasses both types of usage exemplified by ‘3 pounds’ and ‘3
apples’.
The following are some other examples that favor the analysis here:
‘There are 31 apples on the table’, ‘3 | pounds’, ‘3 j pairs of shoes’. A
Fregean analysis of rational numbers is much too complicated. How­
ever, these phrases seem to have not just the same syntactic structures
as former examples, but the same logical structures as well. 3 j pounds
are just 56 ounces. The analysis here provides a uniform solution. ‘3 j ’
has the same semantic function as ‘3’.
In this analysis ‘3’ does not signify an object, nor a property, or a
property of a property. The function of ‘3’ is just that ‘3’ plus a unit
signifies a property. A unit provides a scale, or a framework of cali­
bration. Then numerals combined with units signify some properties
relative to the framework of calibration.
We count fictional things as well as real ones. The function of nu­
merals in counting fictional things can be analyzed in the same way.
As a matter of fact, the ways of dividing fictional things into individ­
uals can be more subtle. For example, suppose one said ‘I propose 3
arguments in this paper’ and the other retorted ‘Actually, there are
only 2, for the second and the third are just the same argument’. The
disagreement is about which is the proper way of dividing the argu­
ments. ‘3’ combined with one way of dividing the arguments signifies
a property of the arguments, but ‘2’ combined with another way also
signifies a property of the arguments.
Given the semantic function of ‘3’ as described above, ‘the number
of pounds o f ...’ is much like ‘the degree of hardness o f ...’. The former

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 118
can be 3, 2 and so on, while the latter can be ‘extreme’, ‘moderate*
and so on. The word ‘pound’ fixes a unit, then ‘3 pounds’ expresses a
property. Note that to be extremely hard is to be so relative to what
is to be hard. So when we describe the level of hardness as ‘extreme’,
some sort of unit, the standard of hardness, is also implicitly assumed.
Then ‘extreme’ or ‘moderate’, together with the standard, signifies a
definite property. We see that the semantic function of numerals is
much like the semantic function of adverbs.
When mathematics is applied in physics, numbers are combined
with units to represent physical quantities. There is a worry about
unit-free physical constants, such as the ratio of the mass of a neutron
to that of an electron. Putnam [81] once took this as evidence that
numbers cannot be eliminated in physics. However, if the ratio is the
rational number r, it only means that if we take the mass of an electron
as the unit, then the mass of a neutron will be r units. Here, the se­
mantic function of r is still to signify a scale point relative to a scaling
system. In other words, r plus a unit signifies a property. Nowhere in
physics is it implied that constants should refer to objects. Moreover,
common physical constants such as the gravitational constant in New­
tonian mechanics, Planck’s constant, the speed of light and so on, can
be eliminated (reduced to 1) by choosing units appropriately. Finally,
we constantly say something like ‘This is 3 times as long as that’. Here
‘3 times as long as’ signifies a relation. The semantic function of ‘3’ in
this phrase is similar to that in ‘3 pounds’. The ratio of the length of
A to that of B in inches is just like the number of inches of the length

of A.
To know the color of a piece of paper one just needs to observe
the piece of paper. The Benacerrafic epistemic difficulty does not arise

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 119

here. Similarly, to know the number of apples on the table, we just


observe the apples and see how they can be divided into individual
apples. There is no Benacerrafic difficulty here, either. To know that
the color of this piece of paper is the same as the color of that piece of
paper, we just need to observe the two pieces of papers and compare our
impressions. Similarly, to know that the number of apples on the table
is equal to the number of oranges on the table, we just need to count
the apples and the oranges. When the semantic function of numerals
is construed this way, the related epistemological question will be easy
to answer.
Given this analysis of the semantic function of numerals, it is easy
to see that numerals are primarily used in arithmetic assertions as in
the following examples: ‘3 apples plus 2 apples are 5 apples’, (3 pounds
plus 2 pounds are 5 pounds’, ‘3 inches plus 2 inches are 5 inches’. Each
of these sentences expresses some physical phenomenon we constantly
observe. Here I literally mean something like (3 pounds of sugar and 2
pounds of salt together weigh 5 pounds’. So these sentences state some
observable physical phenomena, rather than the so-called arithmetic
facts about pure concepts or abstract objects. (3 pounds’, '2 pounds’,
‘5 pounds’ all signify properties. The sentence ‘3 pounds plus 2 pounds
are 5 pounds’ is like ‘red mixed with yellow gives orange’. It expresses
something we constantly observe.
There is a temptation to account for such empirical truths about
the real world by resorting to the assumption that ‘3 + 2 = 5 ’ by itself
expresses some arithmetic truth (about some abstract entities or con­
cepts). The explanation will go like this: '3 apples plus 2 apples are
5 apples’ because ‘3 + 2 = 5 ’. Fictionalists would rather reverse the ex­
planation process: We observe empirical facts of the same pattern:

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 120
‘3 apples plus 2 apples are 5 apples’, ‘3 pounds plus 2 pounds axe 5
pounds’, ‘3 indies plus 2 inches are 5 inches’ and so on. Then ‘3 + 2 = 5 ’
is a notation that summarizes these empirical facts. After that, by
analogy, one can apply ‘3 + 2 = 5 ’ to new cases. This is basically John
Stuart Mill’s explanation. (See [67] Chapter VI.) It seems to be what
we actually did when we learned arithmetic in schools. We started with
counting concrete physical objects. We gather two groups of physical
objects and count them together. Then we memorize ‘3 + 2 = 5 ’ and ap­
ply it in new counting tasks. If this analysis is correct, then ‘3 + 2 = 5 ’
needs not express a statement about any objects. It should primarily
be taken as a schematic sentence, summarizing sentences about real
things and their quantities of the same pattern.
Now, when we develop the theory of numbers or other branches
of pure mathematics, we indeed seem to treat numbers as objects.
According to fictionalism, here we just start to talk as if ‘3’, ‘2’, ‘5’ and
so on were names of objects and ‘3 + 2 = 5 ’ expresses an assertion about
those objects. Sometimes we constiously conceive of numbers. The
way we conceive of objects here is much like the way we conceive of the
letter ‘a’ as a representative object assodated with similar marks. The
reason for doing this is simply that it provides a more effident m anner

of speech. We can ignore apples, pounds, inches and so on and focus on


‘3 + 2 = 5 ’ itself. Counting these numbers will become representatives of

counting other objects, and we can talk as if ‘3 + 2 = 5 ’ is itself a sentence


with definite meaning, not just a schema.
Here we should distinguish two things. ‘3 pounds’ can be a real
property. If 3 pounds of apples put on a scale cause the scale to tilt,
it is this real property that is relevant here. What is fictional is the
number 3 as an object. Ignoring this distinction may induce us to

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 121

think that the number 3 as an object plays some essential role in the
mind-independent physical event of causing the scale to tilt, and then
to conclude that the number 3 must exist independently of mind. Sim­
ilarly, consider the claim that there is a number between 3 and 5. We
have properties such as ‘3 pounds’, ‘5 pounds’, ‘3 inches’, ‘5 inches’

and so on. It is a matter of fact that some physical objects weigh 4


pounds and that is between 3 pounds and 5 pounds. There are similar
cases for ‘inches’, or other units, or apples, chairs and so on. These
are concrete facts. Now when we intentionally take numbers as some
sort of objects, fictionalists claim that they are fictional. In that case,
‘There is a number between 3 and 5’ is a statement in the story about
the fictional objects and it is true in the story.
Certainly, we do not conceive of small numbers only. We tend to
make generalization and therefore imagine that there are arbitrarily
large numbers. So we conceive of the natural number sequence.
Here one must be cautious about another point. Properties can
also be fictional. There is a natural sense in which we can say that a
property is real, that is, it is a property of some existent things. Now
in this sense, for a numeral N, whether or not ‘N pounds’, ‘N inches’, or
‘N miles per hour’ are real properties is contingent on the real world.
For large numerals, these may not be properties of any real things.
Certainly, we can still conceive of physical objects with the properties,
and in that case we have to admit that the properties are fictional,
because they can only be properties of fictional things. For example,
in Newtonian mechanics we can conceive of some object moving faster
than the speed of light waves. This distinction between properties of
real things and fictional properties is also frequently ignored, because
in most cases properties of fictional things are also properties of real

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. FICTIONALISM 122

things. Ignoring the distinction may induce us to think that all prop­
erties such as ‘one pound’, ‘2 pounds’, ‘3 pounds’, ... are real and then
to conclude that at least the structure of the natural number sequence
is mind-independent. However, we do not really know if all these are
properties of some real things, given that the universe could be finite.
Possibly the universe is intrinsically infinite in some aspects and it
can instantiate the u/-sequence, but philosophy of mathematics should
not depend on such a physical hypothesis which we don’t even know to
true. Certainly, realists do not mean that the abstract w-sequence must
be exemplified in the real world. However, the fact is that we can ob­
serve only finite sequences. If we reflect on the origin of our conception
of k/-sequence, we see that, as a matter of fact, we start by imagining
that a finite sequence can be extended forever and ever. We have to use
words such as ‘so on and so forth’ to describe the infinitely many repe­
titions in the imaginary extension and presume that these words have
definite meanings. The question if our imagination does accurately re­
flect some aspects of nature is a secondary question for physicists to
answer. Realists need not mean that the physical question must have
a positive answer, but they make the further move when they claim
that there really exists the abstract mind-independent o;-sequence, not
just the sequences we conceive of. In contrast, fictionalists just try to
be modest.
After conceiving of the natural number sequence, we conceive of
further sets of natural numbers, real numbers, functions and so on. In
connection with the real number system there are similar remarks to be
made. If indeed a geometric line in physical space is isomorphic to the
real number system, we can admit that the structure of the real number
system is real, mind-independent, for it is exemplified by something

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 123

real. However, we don’t really know that. Similarly, the philosophical


account of mathematics should not resort to the physical hypothesis
that space-time is indeed continuous and complete in the mathematical
sense. Fictionalists claim that we conceive of the continuum in order
to build a simple model for the real space-time. To use the words in
ordinary textbooks, we define or construct the continuum. Simplicity
seems to be main reason for modeling space-time by a continuous and
complete model. Now realists claim that even if space-time is not
continuous, even if any aspect of the universe is finite and discrete,
there still exists the mind-independent abstract structure of continuum
and it is not just that we imagine such a continuum. It seems that
fictionalists’ more cautious claim is more understandable to ordinary

scientists.

1 .4. T h e In d e te r m in a c y o f F ictio n a l E n titie s. The indeter­


minacy of fictional entities is a definitive feature of fictional things.
Here I will explain the indeterminacy and then give a characterization
of the distinction between fictionalism and realism based on this feature
of fictional things. This will also help to show that fictionalism is closer
to our common sense and to answer some objections to fictionalism.
Fictionalists claim that mathematical objects are fictional. Their
ontological status is like that of fictional characters in stories. We
have some mental pictures of the objects and structures we conceive
of. Mental pictures are fuzzy, non-rigorous. To communicate them and
to reason about them, we also carefully develop a rigorous language for
talking about them. After this is achieved, we can just talk as if we
are talking about objective objects and structures, not just each one’s
mental pictures. The most extreme form of such a rigorous language
is a formal system. We lay down axioms as the basic assumptions for

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 124

the entities or structures we conceive of, and then we deduce conse­


quences. We want to conceive of mathematical objects as something
definite, existing by themselves. So in particular we accept the law of
excluded middle in our stories. We want to do this, because in most
cases the fictional structures are supposed to simulate the real struc­
tures of some aspects of nature, and our realistic attitude toward the
real world implies that the law of excluded middle holds for the real
world.
However, stories are mostly incomplete. Some statements about
the fictional things may be undecided. We cannot always specify every
characteristic of fictional characters. So, for instance, the cardinality
of the continuum in set theory may be like the length of H a m l e t’s nose
in Shakespeare’s drama. There is no a fact of the matter about them
so long as they are logically independent of the statements accepted in
the story. On the other hand, we can agree that in the story the length
of Hamlet’s nose is either shorter than 3 inches, or longer than or equal
to 3 inches, even though there is not a matter of fact as to which one
is true, so long as Shakespeare did not say it. Similarly, we claim in
the theory that the continuum hypothesis is either true or false, even
though there is not a matter of fact as to which one is true. This is
the most important feature of fictional things: Some statements about
fictional things are supposed to be meaningful and either true or false
in the story, while at the same time there is not a matter of fact as
to whether they are indeed true. Fictional things are indeterminate in
this sense.
Although fictional things are indeterminate in some aspects and to
some extent, they can still be used to build models to simulate reality.
We can compare fictional things with real things meaningfully. Given

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 125

our story about natural numbers, it is true that there are as many
planets in the solar system as there natural numbers less them 9 and
there are more natural numbers than stars in the Galaxy. I will not
discuss the nature of such truths here. They are indeed special, in
that fictional things and real things are mixed. But we constantly say
something like ‘Of course, Hercules is stronger than me’ and we agree
that sentences like this make sense. Given the story about Hercules,
there is a fact of the matter about the truth value of the sentence.
Moreover, we can claim that there are more real numbers than stars in
the universe, although there is some indeterminacy in the cardinality of
real numbers and further the comparison between cardinalities is itself
something we imagine.
One comparison is the structural similarity between some aspects
of the real world and some fictional structures we use to model nature.
Such similarity is mostly vague and intuitive. For example, in studying
thermodynamics we model air by mass points, but we use continuous
models to simulate the motion of air in aerodynamics. Both models
are approximations of air appropriate to their specific objectives. This
notion of approximation may not be definable rigorously, but it is as a
matter fact the basis of scientific studies.
Because of indeterminacy, stories about fictional things can be ex­
tended by adding new axioms about the fictional things. Fictionalists
concede the possibility that set theory plus the continuum hypothesis
will be useful for some aspects of science, while set theory plus the de­
nial of the continuum hypothesis will be useful for some other aspects.
This is like the case of geometry, where different kinds of geometry are
suitable for different applications. Of course, fictionalism also concedes
the possibility that we will subsume different extensions of set theory

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 126
under a more general theory about something more general than sets.
Then we will be able to reduce sets to those newly introduced entities,
and probably we will have sets that satisfy the continuum hypothesis
and sets that do not. This is what we have done to geometries by
reducing them to set theory.
On the other hand, fictionalism respects the common-sense realistic
attitude toward the outer world, which, I assume, is also the attitude
accepted by most scientists. So it is a matter of fact whether or not
there are aliens, even if both our scientific theories and our observa­
tions will never decide it. This is the basic distinction between our
ontological attitudes toward fictional things and the real outer world.
Then we can give a characterization of realism in this context. Re­
alists believe that there is a ‘true’ sequence of natural numbers and a
‘tru e' real continuum which are mind-independent and fully determi­

nate. All meaningful sentences about them are definitely either true or
false, and the truth or falsity is a matter of fact, independent of the
way we think of them. Therefore, these ‘true’ natural and real numbers
must differ from the numbers we conceive of.
Consider this example: A novelist wrote a story about a character
called ‘Joe’. It might happen that the novelist had a prototype for

the character who was also called ‘Joe’, and everything said about
Joe in the story was true of the real Joe. As long as the story is
considered as a fiction and not non-fiction, we can still distinguish
the Joe as the character in the story from the real person Joe. The
story might continue and the character might diverge from the real
Joe. As a character, the Joe in the story must be indeterminate. Some
statements are undecided in the story. I will not go into the question
of what exactly is the ontological status of the character Joe. Maybe

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 127

he is some sort of objective but not mind-independent entity, or there


is no such entity and we just talk as if there were. It suffices to assume
that we can talk as if there were such a character Joe and as if we were
referring to the same character Joe.
The relation between the numbers we conceive of and the supposed
‘true’ numbers is similar to the relation between the Joe as a character
and the real Joe. However, there is an important difference. The real
Joe exists independently, not because the novelist happened to write a
story that happened to be true of him, although the causal connection
might be the other way, that is, the novelist observed the true Joe
first and than created the character Joe. For the case of numbers, we
cannot get in touch with the ‘true’ numbers directly and independently.
It seems that we first conceived of the numbers, or we first just talked
as if we were talking about some objects. It was only after that realists
came to believe that there are indeed those mind-independent ‘true’
numbers.
Realists have to justify the claim and explain how we know that
the axioms we lay down for our fictional numbers are also true of those
‘true’ numbers. That is where the well known Benacerrafic problem
arises. To avoid the difficulties realists might try to avoid the naive
picture of mathematical objects as something just like physical objects
but dwelling in a different universe. However, simply rejecting the pic­
ture without supplying other explications leaves the following question
unanswered: what exactly are those objects? why call them ‘objects’ ?
and what is it for them to be mind-independent? Our prototype of an
object is a physical object such as a chair, a pebble and so on. For such
physical objects, the claim that they are mind-independent is also un­
derstandable. If realists want to call numbers ‘objects’ and claim that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 128

they are also mind-independent, while at the same time they want to
reject drawing any analogy between numbers and pebbles, what ex­
actly realists have said is just unclear. When calling numbers ‘objects’,
the word ‘object’ is stripped of its original meaning, but no explication
of the supposed new meaning is provided. It is fictionalists who can
take the naive picture of numbers as objects, because conceiving of the
sequence of natural numbers is similar to conceiving of a sequence of
infinitely many pebbles. The word ‘object’ is still used in its ordinary
sense. Because of the admission that the objects are fictional, we avoid
the obscure notion of mind-independent abstract object, avoid the Be­
nacerrafic difficulty and avoid any commitment to the unique truth
value of the continuum hypothesis. It seems that fictionalism should
be more acceptable to ordinary people.
Moreover, realists should hold that the numbers we conceive of and
the ‘true’ numbers are related in this way: the ‘true’ real number con­
tinuum is in some sense a completion of the fictional real continuum,
because realists certainly hold that the axioms we lay down for the
fictional real continuum are true of the ‘true’ real continuum. The fic­
tional continuum is indeterminate in some aspects, but the ‘true’ con­
tinuum is supposed to be fully determinate. So the ‘true’ continuum
must be one fully determinate completion of the fictional continuum.
Since there are many ways of extending the story about the continuum
to make the fictional continuum more determinate, we see another dif­
ficulty for realists here: how could it be possible to identify which way
will lead to the complete true story about the ‘true’ continuum?
We cannot uniquely identify the real Joe from the story about Joe,
for there could be many persons about whom all statements in the
story are true. If Joe exists in the real world, we refer to him by

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 129

pointing to him, independently of any story. For the supposed ‘true’


continuum, we don’t have any independent means of access to it. It
seems that realists try to take a leap. Their strong convictions might
be self-consistent, and it is certainly very hard or even impossible to
refute their position. However, they clearly assert more than fiction­
alists do. So the really interesting question is: Is there anything that
would support their more assertive conclusion, and can their position
provide a better account for the nature of mathematics, in particular,
the applicability of mathematics in science? If the answer is ‘N o’, it
seems that fictionalism is more acceptable than realism.
The description above fits the full-blooded realism such as Godel’s.
Realism in Quine’s style is a little ambiguous at some points. In the re­
cent book [84] (p. 57) Quine seems to agree that for independent state­
ments, such as the continuum hypothesis, or the existence of some large
cardinals, their truth and falsity may be indifferent both to our theory
and observation, and therefore may be unknown to us forever. Earlier
([83] p. 94-95), Quine remarked that they can be decided by consider­
ing the simplicity, economy and naturalness of the theory. From that
consideration, Quine suggested accepting the axiom of constructibility
in set theory, which decides the continuum hypothesis.
It is unclear if mathematical reality is really fully determinate for
Quine. However, as long as Quine’s realism is not fictionalism, the
above description of realism seems to be still correct; that is, Quine’s
realism still claims that there are some ‘true’ numbers besides the num­
bers we conceive of. Then the position is still more assertive than
fictionalism.

1.5. Fictionalism and the Indispensability Argument. The


previous subsections have focused on the point that fictionalism looks

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 130

less assertive and more acceptable. Now, if there are reasons for be­
lieving that the existence of mathematical entities as mind-independent
entities is justified by science, then our assessment of the case will be
different. In this subsection, I will argue that the indispensable ap­
plication of mathematics does not demand that mathematical entities
should be mind-independent. They can very well be fictional. The
basic ideas of the argument already appeared in Azzouni [3], Balaguer
[5], Maddy [66] and Sober [96]. Here I want to emphasize what I see
to be the essence of the problem and hope this can make the argument
simpler and more convincing.
The main point of the counter-argument against the indispensabil­
ity argument, as I understand it, is the following: When mathematics
and science are casually put together, it perhaps gives the impression
that mathematical entities are on a pax with physical objects such
as molecules or electrons, and that we must commit ourselves to the
objective existence of mathematical entities just as we do commit to
molecules and electrons. However, a closer analysis of mathematical
applications in science will reveal that physical objects and mathemat­
ical entities play different roles in our scientific explanations of nature.
Physical objects are parts of nature. They exist in the universe, mind-
independently, and they are parts of the causal nexus that explains the
observable phenomena. On the other hand, mathematical entities are
something we use to build models to simulate nature, to represent the
causal nexus, in our scientific studies, and as such they can just be
fictional.
Here is a more concrete example. In modeling the Brownian motion
of a pollen grain floating on a cup of water, we use real numbers to
set up space coordinates, we represent atoms of water by mass points,

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 131

and represent the pollen grain by a spherical rigid body3. We assume


that the atoms move in random directions and with speeds randomly
distributed in some way. Then we use the techniques of probability
theory to estimate the pattern of the motion of the pollen grain. The
model truthfully replicates the observed motions of the pollen grain.
In the model, the atoms collide with the pollen grain and cause it
to move in a specific way macroscopically. If we modeled water as
some continuously distributed physical medium, we would not correctly
replicate the Brownian motion. Physicists believe that this is strong
evidence for the reality of atoms.
Now philosophers of science holding a different view might question
this assertion. This is the type of the question in the debate between
scientific realism and anti-realism. However, the question whether the
real numbers used to build the model are mind-independent entities is
a different question. Physicists will not be moved if they are questioned
about the reality of real numbers, or questioned about how they can be
sure that the mathematical theorems about real numbers they accept
are indeed true. To use real numbers to build the model, it suffices
that we can conceive of real numbers rigorously and coherently. For
instance, suppose we imagine that there were some ‘true’ real numbers
and some of the mathematical beliefs we presently hold happened to
be false of those ‘true’ real numbers. This certainly would not bother
physicists, for they just need the real numbers they conceive o f to build
the model. What are those ‘true’ real numbers and what are true for
them are just irrelevant, because we do not use them.

3This is a simplified model of Brownian motion. The actual experiments that


confirm the particle nature of air or liquid are much more complicated.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 132

It is misleading to say that mathematical entities enter into the


causal explanation of Brownian motion. Atoms enter into the causal
explanation, because it is supposed that atoms collide with the pollen
grain and cause it to move. Fictional atoms cannot do that. Physical
properties of the atoms, the mass, velocity and so on, are also relevant

in the explanations. However, to say that real numbers enter into the
causal explanation is like saying that computers enter into the causal
explanation of the formation of a hurricane when we use computers
to model the formation of a hurricane. Computers do not cause the
hurricane. The relation between mathematical entities and read things
in the universe is the relation between a model and what is to be
modeled. This is the way models enter into the scientific explanations.
Models can be real things by themselves, such as computers, but they
can also be fictional. In the model here, mass points and rigid bodies
obviously do not exist in the real world. Fictionalism claims that these
idealizations and mathematical entities are all fictional.
Here I want to emphasize that physical quantities are represented by
numbers plus units. Physical quantities are indeed relevant in causal
explanations, but numbers as objects are what we use to represent
physical quantities. It is incorrect to say that numbers enter into causal
explanations.
The success of modeling reality by mathematical models depends
on whether or not the models replicate reality in some sufficiently ac­
curate way. As for the mathematical assertions about the models, they
are trivially true as long as they logically follow from the axioms about
the model, because models are conceived by us. When we give di­
rections to others, we try to com m unicate our mental route map to
others. The question whether the sentences we say are indeed true of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 133

our mental route map is trivial. The only possible errors are those due
to our carelessness. The real question is whether the mental route map
faithfully reflects the real roads.
The question whether models faithfully replicate reality and the
question whether the models are themselves mind-independent entities
are clearly two different types of questions. To investigate the latter,
the issue of realism in mathematics, we should first explicate what
are abstract entities, what it is to be meant by saying that they exist
and what could be the evidence for their existence. The success of
using the models in modeling aspects of reality does not hint anything
relevant to the answers to these questions. If we can explicate the
notion of abstract entities, and with the explication we can show that
mathematical entities are indeed mind-independent abstract entities,
the conclusion should be the same even if we never fully succeed in
modeling the physical reality by mathematical models. As a matter of
fact, we know that none of our physics theories provides a literally true
model of reality.
Some people may doubt that building models to simulate reality
is all there is to mathematical applications. This can be checked by
examining various concrete examples, though ‘building models’ should
be understood broadly.
When we count or measure things, or do arithmetic in counting or
measuring, we use numbers together with units to represent physical
quantities, for instance, ‘3 apples’, ‘4 inches’, ‘5 pounds’ and so on.
These quantities can be quantities of real things, but the numbers 3,
4, 5 as objects can just be fictional. There is no need to take these

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 134

numbers as mind-independent objects4. In a little more sophisticated


applications, we do not just count or measure. We consider statistical
features of the data we gathered. In other words, we report means,
standard deviations and so on. Suppose we obtain that the mean height
of American male adults is 180.123cm. This may not be the height of
any real person at the moment, but we are still talking about a possible
property of real things. There is still no need to take 180.123 as a mind-
independent object.
As a further step, we may want to describe the distribution of

heights. Here we use a distribution function # to represent the distri­


bution. This is a simple example of modeling reality by a mathematical
model. The function can just be fictional. It is easy to see how to trans­
late assertions about the function into assertions about real things. If in
the mathematical story we have # (1 8 5 ) —# (1 7 5 ) = 0.6, that means 60
percent of American male adults are between 185cm and 175cm. Even
more complicated applications are similar to the example of Brownian
motion cited above.

The essential point is this: in textbooks and classrooms, or in do­


ing calculations, we deal with fictional things, but when we apply these
to reality, we meet real physical stuff with various properties. No one
really confuses the two sides in real life. When we go into our labo­
ratories, we look for atoms or electrons in this real world and look for
physical properties such as ‘3 cm’, ‘5 g1, ‘2 cm /sec’ and so on. We do
not look for mass points or numbers. Numbers are combined with units
to represent physical quantities, and physical quantities can indeed be
real. No matter how complicated the applications are, we can always

4Actually, there is no need to take ‘3’, ‘4’, ‘5’ as names of objects here. See the
analysis of the semantic function of numerals given in Subsection 1.3.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 135

distinguish the two sides: for one, the mathematical entities and the
idealized physical objects such as mass points we refer to in doing the
calculations, and for the other, the physical stuff and properties in the
real world5. Physical phenomena in the real world are accounted for by
referring to other physical objects and physical quantities. Numbers,

functions and so on as objects are used to represent those objects and


quantities. The relation between the two sides is the relation between
a model and what is to be modeled.
The conclusion of this analysis is that mathematical realism is a
purely philosophical issue. The methodology adopted by ordinary sci­
entists and the so-called naturalized epistemology say nothing about
the issue. Fictionalism is consistent with the scientific methodology
and epistemology. In particular, it is consistent with the realistic un­

derstanding of science.
Now, what exactly is wrong with the indispensability argument?
The argument sounds appealing, at least on a first appearence.
The argument is based on an analogy between mathematical enti­
ties and unobservable physical objects such as atoms, electrons and so
on. One thing that the argument simply ignores is the fact that we
do refer to fictional things and real things simultaneously in everyday
speech. For instance, sometimes we say ‘Hercules is stronger than any
real man’. Here we even consciously use the predicate ‘real’ to dis­
tinguish recil things from fictional things. If we also consciously add
this predicate ‘real’ in scientific discourse whenever we intend to re­
fer to physical objects and properties in the real universe, the analogy
between mathematical entities and physical objects will break down.

sAssouni [3] emphasised this distinction between the physical and the
mathematical.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 136

When we refer to electrons, we mean real things in this real universe,


but when we introduce mathematical entities and use them to repre­
sent and describe real things and their properties, there is no need to
apply the predicate ‘real’ to those mathematical entities. We can use
them in the same way as we use Hercules. Quine had this well known
slogan: ‘To be is to be the value of a variable’, but it seems that Quine
never considered the fact that we constantly use this predicate ‘real’ in
our languages. When ‘real’ appears among the predicates of a theory
couched in a first-order language, the theory does not really commit to
all the entities that its variables range over.
The indispensability argument does demand that the entities a the­
ory really commits to are entities indispensable to the theory. But no
one ever argued that fictional things must be dispensable, or that any­
thing indispensable must be real and mind-independent. The need for
such an argument is due to the fact that we do intentionally refer to
fictional things (for instance, Hercules) to make assertions about real
things. The analogy between mathematical entities and unobservable
physical objects hides this point, for physical objects are of course
mind-independent. So, there is a gap in the indispensability argument,
and what is supposed to fill the gap is actually the most important
point. Missing this point, the argument is almost empty. The ap­
parent appealingness of the argument is due to the fact it implicitly
presumes the main point: What we indispensably refer to in order to
express scientific theories must be real and mind-independent.
Now fictionalists actually contend that there is a close connection
between what is fictional and what is practically indispensable for sum­
marizing our knowledge of real things. Real things are always too fuzzy

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 13T

and too complicated to be fully graspable. So we have to imagine non­


existent things to build simplified models to summarize our knowledge
about real things. Doctors may feel that they have to use fictional
cases to summarize their clinical experiences. Physicists refer to mass
points, rigid bodies and so on to build simple models to make the cal­

culations manageable. Such idealizations do seem to be indispensable.


They are useful just because they are not real. Similarly, it is much
simpler to say ‘2 + 3 = 5 ’ as if we are talking about some objects than to
say something like ‘2 objects + 3 objects = 5 objects’.
Fictionalists claim that these idealizations, as well as mathematical
entities, are fictional, while realists claim that they axe real, mind-
independent. Probably, neither side has conclusive arguments, but
given the way we actually first come to think of these idealized entities,
fictionalism should be less assertive and more acceptable.

1.6. T ruth and C o n siste n c y in F ictio n a lism . Here I want to


analyze two other possible objections to fictionalism in connection with
the notion of truth and consistency. I will refer to the characterization
of realism from the perspective of fictionalism given in Subsection 1.4.
First I want to consider Resnik’s [87] objection that as long as we
use mathematics in science, we have to agree that mathematics is true,
and then some sort of realism is inescapable. The answer is that our
stories are indeed true of those fictional things, but we do not need
to assume that those fictional things are real, mind-independent. For
example, suppose that the story about Joe was told by a psychiatrist
in a book of psychiatry. The author used the story as a hypothetical
clinical case. Now an intern tried to apply the case to a real patient
Tom. Whether or not the application will lead to correct conclusions
depends on whether or not the descriptions of Joe in the book really fit

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. FICTIONALISM 138

the real person Tom. On the other hand, it seems irrelevant whether or
not there was a real person called Joe who is the prototype of the Joe
in the story, or whether or not there is a mind-independent abstract
entity Joe for whom the story is accurately true. The author might use
the single character Joe in the story to summarize a huge amount of
clinical experience. If we ask the ontological status of this Joe per se,
we have to agree that he is fictional.
Now suppose we agree that no matter whether there are those ‘true’
numbers claimed by realists, we can always conceive of our fictional
numbers, or in other words, we can talk consistently and rigorously as
if there were numbers. Then, similarly, for the fictional real number
continuum to be applicable, for instance in modeling the space-time
structure, the relevant question is whether our descriptions of the fic­
tional continuum are approximately true of some relevant aspects of
real space-time. It is irrelevant whether there is a mind-independent
‘true’ continuum for which our axioms are true.
Now consider another possible objection also raised by Resnik [87]
in a footnote. It comes from the claim that a mathematical story
should be consistent. Before examining what exactly the objection is,
I will first explain how fictionalists understand the consistency of a

mathematical theory.
Fictionalists would distinguish two kinds of consistency statements.
First there are statements about deriving (or the future possibility of
deriving) contradictions in the real world. These are statements about
the real world and they have realistic meanings for fictionalists. They
are of the same nature as scientific predictions put in a conditional
format, such as: human beings will not derive a contradiction from
the axioms, or if a computer generates proofs of the formal system,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONAL ISM 139

there will be no proofs of contradictions generated. Such statements


are never absolutely rigorous because they refer to concrete physical
things such as human beings or computers. But they are indeed like
ordinary scientific predictions.
On the other hand, we can imagine that sentences and proofs can be
inde“finitely long. In other words, we can conceive of a formal system
just as we conceive of fictional mathematical objects such as numbers.
We conceive of symbols, strings of symbols, sequences of strings of
symbols and so on. We define the symbols of a formal language and
define sentences, proofs and so on. Then the consistency statement will
be a statement about the sentences and proofs we conceive of, and thus
involves a quantification over all proofs. Here we are creating another
story, a story about a fictional formal system. Then the consistency
statement becomes a statement within the story.
Let me call the two types of consistency assertions ‘material con­
sistency assertion’ and ‘theoretical consistency assertion’ separately.
Sometimes people use ‘could be’ to express something beyond what
will actually happen. Modem philosophers are very much interested in
speculating about various kinds of possibility and necessity. From the
point of view of fictionalism, when we talk about what could be, we
already start imagining something. When we say that a contradiction
could never be derived, we are imagining that an ideal agent can derive
theorems forever while following the some rules. Fictionalists do not
deny that we can imagine that way. Even if we strongly believe that
human beings will all be wiped out in the next year, we can still imagine
that way. Similarly, even if scientists all believe in a model of the
universe with an ending in time, we can still imagine that way. Or we
can simply imagine arbitrarily long proofs as timeless entities.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
1. FICTIONALISM 140

Moreover, fictionalists do not deny that we follow some constraints


in imagining the fictional situations. To say what could be is to imagine
what could happen in various fictional situations while following the
constraints of imagining. Different types of constraints correspond to
different types of possibility. Given the constraints, sometimes we can
decide what could be and what could not be. When we imagine formal
systems, we want them to simulate some aspects of actual proofs in
the real world, ignoring some other aspects that we take as irrelevant,
such as the concrete physical realizations, the limits on the lengths of

proofs and so on.


For that reason, the two kinds of consistency statements are related.
Every concrete derivation in the real world can be seen to correspond
to a fictional proof. If we derive a contradiction in the real world and
we want the fictional proofs to replicate real derivations in telling the
story, clearly we will claim that the formal system is inconsistent in the
story. On the other hand, if in our story about the fictional proofs we
deduce that no proof is a proof of a contradiction, then, by analogy,
we can convince ourselves that no reed derivations in the real world
will deduce a contradiction. This is just the typical way of applying
fictional models to reality by analogy as we constantly do in physics.
The formal system is a fictional model to simulate our real theorem
proving activities in the real world.
However, fictionalists reject the assertion that those fictional situ­
ations are real, mind-independent. They also reject the assertion that
there is always a fact of the matter as to what could be and what could
not be. Consider the consistency statement about a formal system
that is an extension of set theory. Suppose that right now we cannot
imagine an extension of set theory that would decide the statement. So

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 141

right now we cannot decide if a contradiction could, be derived. Within


the story about the formal system we still agree that either there is
a proof of contradiction or there is no such proof. But fictionalists
do not believe that there is always a fact of the matter as to whether
we could imagine a situation in which a contradiction is derived, and
therefore they do not believe that there must be a fact of the matter as
to whether a contradiction could be derived. It is one thing to follow
some objective constraints in imagining what could be and what could
not, so that there is some objectivity in claiming what could be and
what could not be, but it is a quite different thing to leap to the con­
clusion that there is a mind-independent fact of the matter as to what
could be and what could not. The latter is actually a very big move.
Similarly, it is one thing to follow classical logic in conceiving the fic­
tional situation, so that we can agree within the story that anything
either happens or does not happen, and it is a quite different thing to
assert that there is a fact of the matter as to which really could be the
case.
Generally, fictionalists reject the leap from some limited objectiv­
ity, due to some constraints on our imagining activity, on the fictional
things, to the full-blooded realistic assertion that there is a mind-
independent domain of entities, or that there is a mind-independent,
context-independent, absolute possibility. The rejection is of course
partially due to the Benacerrafic style difficulties and partially due to
the fact that science does not commit to such a move and the move is
not really indispensable to account for science, either.
Now what exactly is the objection from consistency? Here I will
try several ways of fleshing out the objection. One way is to say that
for a story to be applicable, it must be consistent, and it must be truly

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 142

consistent, not merely that we can tell another story about the former
story and claim the old story to be consistent in the new story. The
answer is that what really matters is material consistency, which has
a realistic meaning for fictionalists. In other words, indeed consistency
in a story by itself is not sufficient, but there is no need to assume the
so-called ‘true’ consistency, either. For example we can conceive of such
a scenario: There is a contradiction in set theory, but all derivations
of contradictions are so long that human beings could never discover
them. Would that be a problem for using set theory? The answer
seems to be ‘No’. Conceive this scenario: We can actually prove a
theorem VnA (n) with A (n) a decidable predicate in set theory, but
there is a proof of ->A (N ) for some extremely large numeral N , though
it is so long that no human beings could possibly discover it. This
would not cause any problem in our applications of VnA(n), because
all instances of the form A (n) we can actually draw happen to be true
(when interpreted as assertions about real calculations).
Moreover, when a mathematical theory is combined with physical
assumptions and applied to reality, statements in the theory are trans­
lated into statements about real things in some complicated way. A
theory with hidden contradictions can still help us to get truths. This
is just a matter of fact. Eighteenth century calculus was contradictory.
Some modern physical theories such as quantum field theory are al­
leged to contain literal contradictions, not to mention psychological or
sociological theories. Finally, we must recognize that set theory might
contain hidden contradictions. Moreover, if one demands a logically
absolutely clear account for mathematical applications, then my sug­
gestion, as has been indicated in the Introduction to this dissertation
and as will be discussed in the next section, is that we should eliminate

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 143

infinity totally and reduce the applications to applications of finitary


mathematics.
Another attempt to flesh out the objection might be this: fiction­
alists cannot prove the consistency of our theory by resorting to the
existence of models of the theory. The answer is that, as a matter fact,
we resort to fictional models, rather than the alleged ‘true’ models, to
convince ourselves of the consistency of a theory. The consideration
of how to convince ourselves of consistency actually favors fictionalism
rather than realism.
For instance, consider Peano Arithmetic P A . We conceive of a se­
quence of natural numbers and assume the basic axioms of our story
about those natural numbers. To convince ourselves that our story is
consistent, we imagine that the story is formalized as a formal system.
That is, we extend our story about natural numbers to include sen­
tences and proofs in the formal system as objects of the story. We lay
down some basic combinatorial assumptions about the symbols and
strings of symbols of the formal system, which we recognize as true
about reed physical realizations of sentences. Then we define the in­
terpretation relation between the expressions of the formal system and
the natural numbers and deduce (within the story) that the theorems
of the formal system are satisfied by the natural numbers. From that
we conclude that 0 = 1 is not a theorem. As we intentionally create the

story of the formal system to simulate our theorem proving activities


in the real world, we apply the conclusion about ‘0 = 1’ in the story
to reality and conclude that no inconsistency will be found in the real
world.
This is a phenomenal account of how we actually convince ourselves
of consistency, and there is no reference to the alleged ‘true’ model of

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 144

natural numbers. From the perspective of fictionalism, the phenome­


nal account fits what both realists and fictionalists are actually doing.
Realists just want to draw an extra conclusion that there is a mind-
independent, fully determinate ‘true’ sequence of natural numbers, not
just the vague and indeterminate natural numbers we conceive of. Tak­
ing this extra conclusion as a reason for believing in consistency clearly
begs the question.
Sometimes in philosophical arguments people seem to assume that
we know for sure that set theory is consistent and only realism can
explain that. That is clearly not true. For set theory we similarly first
of all conceive of the cumulative hierarchy of sets. We also practice
reasoning about sets conceived in this way. Then, after a while, we
gradually convince ourselves that no contradictions will be found. So
the same is true here: we imagine something first, and then we come
to believe in the consistency of the imaginations, and only after that
do some people try to make a further move to claim the reality of those
things.

Another way to put the objection might be this. We need some


minimum mathematical knowledge in order to prove consistency, and
that minimum knowledge should be ‘true’ knowledge, or it should be
understood realistically.
The assertion is true to some extent. Consider again the descrip­
tion of how we convince ourselves of the consistency of the Peano
Arithmetic. To say that we intentionally conceive of the fictional for­
mal system to simulate the real theorem proving activities in the real
world, we actually presume some combinatorial knowledge about sym­
bols, expressions and operations on them. As I will explain in the
next section (and have already indicated in the Introduction to( this

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 145

dissertation), fictionalists agree that there is an elementary part of


mathematics that can be interpreted realistically. That corresponds to
what Hilbert called the part of mathematics with content. In particu­
lar, strict constructivism is within this elementary part of mathematics
and it does include some elementary combinatorial knowledge. That is
what we need when we try to conceive of a formal system in order to
simulate the real deductions and claim that the real deductions can be
translated into the fictional deductions in the fictional formal system.
However, this does not imply that infinitary mathematics must be true
in some realistic sense.
There might be another attempt: one might claim that we need

‘true’ consistency to account for material consistency. Material consis­


tency is an assertion about what will happen in the real world, while
the alleged ‘true’ consistency refers to an infinite domain of sentences
and proofs. The claim has the consequence that even if the real uni­
verse happens to be totally finite, we still need to refer to arbitrarily
long sentences and proofs to account for some events in the finite real
world. This is a typical Platonic claim, but we don’t know if there are
good reasons for claiming the necessity of such accounts.
In particular, we never need such an account in science. For causal
explanations, such as the explanation of the Brownian motion of a
pollen grain as the effect of the collisions of atoms, the causes and
effects are both real concrete events and only physical objects and
properties are involved in the causal nexus. As for the explanation of
physical reality by building a model, the model can be fictional.
Some realists might have the conviction that to account for the
phenomena in this concrete, vague, finite, but extremely complicated
real world, one has to resort to the abstract, idealized, infinite, but

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
1. FICTIONALISM 146

relatively simpler world, and moreover one has to assume that the
latter is also mind-independent. Fictionalists accept the first assertion,
the pragmatic need to refer to some idealized world, because that is
the most efficient way to describe the real world, but fictionalists reject
the assertion that the idealized world is mind-independent. This may
simply be due to caution, because there is no real need for that extra
assertion, while there are difficulties in it.

1.7. F ictio n a lism and O th er P o sitio n s in P h ilo so p h y o f M a th


e m a tic s. Here I want to compare fictionalism with some other posi­
tions in philosophy of mathematics. The objective is to show that
fictionalism makes it possible to integrate different good insights about
mathematics and its applications. The discussions are all very brief. I
will not attempt to give detailed criticism of the various non-fictionalist
philosophical positions.
As I have mentioned, fictionalism is not supposed to be an exotic
or extremist view. It is supposed to provide a phenomenal account
of mathematics as it actually is. It seems that various positions in
philosophy of mathematics try to emphasize some special aspects of

mathematics, and then try to prescribe what mathematics ought to


be, or to reveal what mathematics actually is upon the right inter­
pretation. I take it that a phenomenal account just tries to describe
mathematics as it apparently is, but for the description to be complete
and informative, it should also be able to explain what other accounts
were emphasizing and what they missed.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 147

First, consider formalism6. As has been noted, to present the fic­


tional objects or structures rigorously, mathematicians develop a rig­
orous mathematical language, and the most extreme form of it is a
formal language. After laying down the basic axioms about the fic­
tional objects, mathematicians’ work mostly consists in deriving the­
orems from the axioms. So formalism does capture some important
feature of mathematics, that is, the essence of mathematical activities
is deriving theorems in a formal system.
Now clearly mathematicians are not just deriving theorems in a
formal system or proving meta-theorems about a formal system. They
do seem to talk some objects: numbers, sets, functions and so on. They
do seem to assign some meanings to the mathematical expressions and
sentences. Obviously they are not just manipulating symbols. It is
almost impossible to prove ordinary mathematical theorems by taking
the axioms of a formal system as strings of symbols and the inference
rules as formal transformation rules on strings of symbols, without any
intuitive understanding of the meanings of the axioms and the rules.
So conceiving of some system of mathematical objects and taking them
as the intended interpretation of the formal system is as a matter fact
a part of mathematicians' work, and it is indispensable.
Moreover, formalists sometimes talk about the consistency of the
formal systems.7 Now, how do we come to believe in consistency?
As it is explained above, we conceive of models of the formal system.
Actually, the actual process is mostly the reverse. We first conceive of
some structures and then develop the formal system. We can never be

6Here I mean the view expressed by Curry [51]. Robinson’s view in [91] is
actually very close to fictionalism.
7Curry did emphasise that there is no absolute necessity in proving the consis­
tency, but the remak here should also apply.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 148

absolutely certain about consistency. That is not a problem, because


we have an intuitive, though vague, picture of the fictional structure.
If we randomly pick 10 sentences from a first order language as the
axioms, we can have no idea about the consistency of the system.
Similarly, formalists hold that the applicability of a formal system
is a pragmatic question. However, obviously, it is not true that mathe­
maticians study all possible formal systems and then scientists choose
the'ones that are applicable. Actually we first have some structures
in mind, and only after that do we try to develop a formal system to
capture our vague intention of conceiving of some sort of structures.
The structures we conceive of are intuitively similar to some aspects
of nature. That is why we expect that they can be used in science to
model reality. It seems that formalism misses this point.
This remark applies to the so-called ‘if-thenism’ as well. It is not
true that mathematicians are interested in drawing conclusions from
any arbitrarily chosen axioms. They are especially interested in the
axioms that intuitively characterize some fictional structures.
Moreover, when mathematics is combined with physics and applied
to reality, the axioms, or the premises, are almost never literally true.
We use real numbers to represent space-time points, but in most cases
physicists do not really care if real space-time structure is literally
isomorphic to the mathematical structure. What accounts for the ap­
plicability of mathematics is not simply that the axioms become true
under appropriate interpretations in applications and then the conse­
quences we deduce are also true. In most cases, reality only provides
an approximate model for the mathematical axioms, and this intuition
of ‘approximation’ is what makes scientific work non-trivial.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 149

Mathematical models are mostly much simpler than the reality to


be modeled. For example, in studying fluid dynamics, we use continu­
ous models to model water. The introduction of infinity and continuity
is to simplify the models. The usefulness of mathematics is not merely
due to the fact that mathematics provides some strict logical conse­
quence relations, some theorems following logically from some axioms.
Mathematics is useful because the mathematical models, though not
exactly true of reality, are much simpler than reality and can still pro­
vide a sufficiently accurate description of reality.
Structuralism8 certainly captures something rightly as well. As a
fictional entity in a story does not subsist beyond the story, it is fully
characterized by its properties in the story and its relations with other
fictional entities in the story. If someone repeats the plot of the story
of Romeo and Juliet, but replaces the names ‘Romeo’ and ‘Juliet’ by
other names, we will claim that it is the same story. What is essential
about a story is the structure, the properties and mutual relations of the
characters in the story. Moreover, when we define natural numbers as
some sets, we are embedding the story of natural numbers into a more
comprehensive story, the story of sets. As the entities are fictional, it
does not matter if we embed the story in one way or another, as long
as the relevant properties and relations are preserved.
However, fictionalists emphasize that mathematical structures are
fictional. Here it is necessary to distinguish two cu e s. Some structures
we treat in mathematics may have instances in the real world, for
instance, some finite groups. For this type of mathematical structures
there could be some ambiguity when mathematicians talk about them.
For example, they may intentionally mean a fictional group, or they

s See Reanik [88].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 150

may indefinitely mean things in the real world with the kind of mutual
relations that make them to form a group.
On the other hand, for most structures treated in mathematics, in­
finity or continuity is intrinsic; then we don’t really know if the struc­
tures tire exemplified in reality. The first time we think of such struc­
tures is typically something like this: We draw a straight line on a
blackboard and then imagine that the line can be extended infinitely
long while kept straight, or we observe some objects arranged in a
sequence and then imagine that the sequence can be infinitely long.
The structure of a straight line or a w-sequence is thus obtained. Fic­
tionalists take this process of conceptual formation very seriously, and
they are modest and cautious in ontology, claiming that first of all we
conceive of the structures.
The need for fictionalism also arises in the following consideration.
According to Resnik’s account [88], first-order structuralists do not
commit themselves to structures as objects. Variables range over only
positions in a structure. However, when we study branches of mathe­
matics, we still need some reasons to believe that we are talking about
something, a structure, even though we agree that a structure is not
to be an object. Now the question is: Under what conditions can we
claim that there is such and such a structure? How do we know (or how
do we convince ourselves) that, for instance, there is a structure with
sets as positions? Could we have always been talking about nothing?
In the account of first-order structuralism, a structure is uniquely
determined by a first order theory. Clearly, the first order theory must
be consistent9. So at least we should have reasons to believe that the

9There might be some doubts as to whether consistency is sufficient, but I will


not discuss the details here.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. FICTIONALISM 151

first order theory is consistent. Now how do we come to believe that


set theory is consistent? I have explained this from the fictionalist
point of view in the last subsection, that is, we come to believe in
consistency by conceiving of the cumulative universe of sets. Rejecting
this, we will be trapped in circularity: We believe in the consistency of
set theory because we believe there is a structure of sets characterized
by the theory, and we believe in the structure because we believe in the
consistency of the theory. Generally speaking, we need a place where
we can get off the ground, and that means we must take our activity
of conceiving of structures seriously.
Modal structuralists try to say that mathematicians study logically
possible structures, but the remark above still applies. That is, how
do we know that the w-sequence is logically possible? In the 18th
century, some mathematicians believed that the notion of real infinity
involves contradictions. It is after Cantor that we gradually come to
believe that we can talk about the infinite consistently. The case of
o/-sequence is indeed simpler than that of sets. Real infinity is not
so obviously involved. However, the fact is still that we can observe
only finitely many things. There should be a different ground for be­
lieving in the possibility of an infinite sequence. Fictionalists have a
simple explanation: It is logically possible because it is conceivable.
So modal structuralists’ reduction of mathematics to modal assertions
about possible structures may really reveal something interesting about
mathematics, but fictionalists’ claim seems indispensable.
Similarly, Charles Chihara [38] gave a reinterpretation of math­
ematical language that reduces references to mathematical entities to
references to possible constructions of open sentences. From the fiction­
alist point of view, this is indeed like a hermeneutic study of literature.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 152

It reveals that the apparent assertions about the fictional characters


can be interpreted as assertions about something else. This is inter­
esting by itself and it may serve some philosophical purpose. On the
other hand, fictionalism is just a phenomenal account of mathematics
as it apparently is.
Maddy’s [66] discussion about the methodology of set theory can
be naturally fitted into the framework of fictionalism. If we want a
grand story that can provide sufficient fictional structures to model
every aspect of nature, Maddy’s maxims such as maximizing and uni­
fication are indeed what we will expect of the story. Maddy wants to
suspend ontological judgement about mathematical objects. So fiction­
alism claims a bit more.

On the other hand, the discussion of methodology would be puzzling


if it were combined with realism. For natural science, methodology
is about what is our best theory of nature. However, our sense of
physical reality is that the best theory could be wrong as a matter of
fact. Methodological maxims only instruct us as to what is our best
guess. They do not determine reality. We do not think that to be real
is just to satisfy the maxims. In the case of set theory, the discussion
of methodology does seem to make reality depend on our methodology.
The discussion does not leave room to see set theory as our best guess
at an unknown mathematical reality, which means that mathematical
reality consists of sets, rather than numbers, or categories, or some
suspected unknown things that might be discovered by mathematicians
two centuries later. The paradoxes will all disappear if we grant that
we are talking about a fictional world.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 153

Hartry Field’s position expressed in [44] and [45] is also called ‘fic­
tionalism’. Cleaxly the fictionalism presented above is close to Field’s
views. Here I want explain the differences.
Field’s emphasis was on the elimination of fictional things in sci­
ence. However, mentioning fictional things such as mass points is so
common in science that there seems to be no reason why they should
be eliminated. Scientists use fictional mathematical entities and struc­
tures to model aspects of reality, because they find that it is the most
efficient way for describing nature, and it does give us a more and
more accurate picture of nature. If we hold a naturalistic attitude to­
ward epistemology, that is, if we agree that science provides the norm

for knowledge, then we should accept ordinary scientists’ methodol­


ogy. That means there is no need for eliminating fictional entities in
scientific theory, unless the reason for doing so is a pertinent scientific
reason, not just the concern that fictional things (or abstract objects)
do not exist.

Moreover, it is possible that referring to fictional things is prag­


matically indispensable in stating scientific theories. Just consider
Newtonian mechanics. How can one state the general theory with­
out mentioning something like mass points? It must be very awkward.
Similarly, there seems to be no way to avoid mentioning frictionless sur­
faces, ideal gases and so on, because we have to resort to idealizations
and simplifications in doing calculations.
On the other hand, it is an interesting logical question if mathe­
matical entities are as a matter of fact logically indispensable in the
cases of applications. If some fictional things are strictly indispensable,
it will be logically puzzling. How can truths about fictional things be

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. FICTIONALISM 154

essential to truths about real things? My opinion is that nominaliza-


tion programs such as Field’s are still insufficient for answering that
question, because many fictional entities are still assumed there. Most
importantly, infinity, at least potential infinity, is accepted in all such
programs, while in most mathematical applications infinity is not ob­
tained or not known to be obtained. In the next section I will discuss
this question and suggest that strict constructivism developed in this
dissertation can provide a way of answering the question.
Fictionalism shares some common ideas with Tharp’s conceptual­
ism [99] [100]10. Here I want to explain the differences. A major
difference lies in a possible ambiguity in Tharp’s assertion that mathe­
matical truths are conceptual truths. Tharp seems to claim that some
relatively simple concepts in mathematics, for instance, the concepts
in arithmetic, or the concept ‘tallies’ in [99], are sufficiently definite
(or ‘sharp’, to use Tharp’s description) that any sentence about the
concepts has a definite truth value. This is not explicitly asserted in
Tharp’s papers, though some discussions in the papers seem to imply
this point11 and Chihara [37] explicitly reports it as Tharp’s view. If
that is the case, it is different from the view presented in this section,
as it is from the discussion of consistency in Subsection 1.6 above.
The type of fictionalism expounded in this section holds a more
consistent view on this point. Concepts in arithmetic are treated in
the same way as concepts in geometry. The really essential distinction
lies between real things in this real universe and their properties on the

10Here I want to thank Professor Solomon Feferman for bringing to my atten­


tion Tharp’s papers and the issue of their relation to fictionalism.
u See, for example, the second paragraph on pp. 182 of [99], where it hints
that it makes sense to refer to ‘all truths of of arithmetic’. But Tharp did claim
that the concepts in Euclidean geometry and some concepts in set theory are not
absolutely sharp.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. FICTIONALISM 155

one side and other things about which we just talk as if they exist on the
other side. At least we should leave open the question whether concepts
in arithmetic are ‘absolutely sharp’ so as to endow truth values to all
sentences. Tharp also held that mathematical concepts are creations
of minds [100]. However, it is unclear how such a mental creation,
for instance, the concept of w-sequence or Tharp’s tallies, can endow
all sentences about it with truth values. The fact is that we know
no way to give a complete axiomatic characterization of the concept.
Fictionalism takes this stance: Within the story involving the concept
we talk as if there were an objective w-sequence and any meaningful
sentence about it were either true or false. Namely, we accept the law
of excluded middle in the story. But without the story, we agree at
most that the story can be extended in some natural way in order to
capture our vague intuition of a concept better. There is no a priori
guarantee that every sentence about the concept has a truth value and
no guarantee that indeterminacy in the concepts can be eliminated.
Tharp emphasized that his interpretation of mathematics is object-
free, while in fictionalism we simply say that those objects are fictional.
This is perhaps not an essential difference. Our grasping of a mathe­
matical concept such as the concept of u>-sequence and our imagining
some objects, a particular sequence, seem to go hand in hand. It does
not seem to be the case that we grasp the concept before we conceive of
a particular ut-sequence and have some practice of reasoning about the
sequence. Claiming that mathematical truths are conceptual truths
does not say much in itself if we agree that the concepts are also fic­
tional.
Finally, another point is the following. Tharp’s interpretation of
mathematics is not committed to infinitely many objects, but it is

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
1. FICTIONALISM 156

not finitism in the ordinary sense. Tharp clearly accepted infinitary


mathematics. Then there is the question about the applicability of
mathematics. Mathematical concepts, or the example ‘tallies’, are dif­
ferent from physical concepts. The former necessarily ‘permit’ infinity,
as Tharp claimed, and (iterated) quantifiers ranging over infinite do­
mains are used to express truths about the concepts. Then it is unclear
how such concepts will be applicable to the physical reality, in partic­
ular, in case the physical reality is finite.
In the next section, a suggestion will be made to resolve this puzzle
by reducing applications of infinitary mathematics to applications of
strict constructivism.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 157

2. O n A p p lica tio n s o f S tric t C o n stru ctiv ism

2 .1. T h e P ro b lem o f th e A p p lica b ility o f M a th e m a tics. The


exposition of fictionalism in the last section leaves two puzzles: (1) How
are truths about fictional mathematical entities related to truths about
real things in mathematical applications? or how can truths about
fictional entities help to deliver truths about real things? (2) What is
the epistemological nature of the truths thus obtained? Fictionalism
is about the ontology of mathematics. The questions raised here are
logical and epistemological questions. This short section will discuss
the first puzzle. The epistemological question will be discussed in the
next section. In this subsection, I will first characterize the logical
problem.
The intuitive answer to the logical question should be clear from
the last section: Fictional things are used to build models to simulate
aspects of nature. In applications, truths about models are translated
into truths about aspects of nature. That is how truths about fictional

things are related to truths about reality.


This explanation is far from complete. In most applications, the
models are not true representations of reality. Consider a more specific
example13. In fluid dynamics we use continuous models to simulate
fluids. We assume that the mass of a sample of fluid is distributed
continuously over a space region. The model is not really true of real
fluid, for real fluid is composed of discrete tiny particles, but the model
works very well as long as only mechanical (non-thermodynamical)
properties of fluids are concerned. The puzzle here is why the model
works, since it is false.

iaSimilar examples have been discussed by several philosophers, in particular,


Maddy [66].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 158

Suppose we deduce some predictions about the motion of an air­


plane in a continuous model of air. On first appearence, the predictions
logically follow from Newtonian mechanical laws and the assumption
of continuous mass distribution of air. Do the predictions also logically
follow from Newtonian mechanical laws about particles and the true
assumption that the air is composed of particles? Certainly we expect
the answer to be ‘Yes’. However, clearly not every conclusion one can
draw about the continuous model is also true of the particle model.
In the extreme case, it follows from the continuous model that there
are no atoms, for mass is distributed continuously there. So the puz­
zle is: what kind of mathematical reasonings can guarantee that the
conclusions deduced from the continuity assumption are also at least
approximately true of the particle model of air?
The puzzle is actually about the logical structure of mathematical
applications. Axiomatic set theory and first-order logic have provided
a logical analysis of pure mathematics. On the other hand, the logical
mechanism of mathematical applications is still very unclear. It is
unclear what exactly follows from what in applications. What are the
real true premises and what are the true conclusions? Besides general
laws and observational statements, are truths about fictional things
(or abstract mathematical entities, if you like) also really among the
premises from which we deduce scientific predictions about real things?

Given all the defense of fictionalism in the last section, the situation is
still very puzzling if the answer to this last question is ‘Yes’.
This example of fluid dynamics is not special. The situation is the
same for almost all mathematical applications in science. The essential
point is that the part of the universe that human beings can have direct
or indirect access is essentially finite. Given the present state of science,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 159

we have no idea about what are below the scale of 10-40cm or 10~43 sec.
What we do know is that things become too much for our imagination
when they are more and more remote from us (both microscopically
and cosmologically). Any assertions about things beyond some finite
scope must be very speculative. This has nothing to do with any anti-

realistic interpretation of science. One can accept everything physicists


confidently assert, including the assertions in fundamental physics. But
we know that direct generalizations of the physical laws that we know
to be approximately valid in some finite scope can go wrong wildly.
Moreover, the universe could just be totally discrete and finite.
On the other hand, in mathematics we assume infinity from the very
beginning. We apply mathematics not just to cosmology or fundamen­
tal physics. The majority of applications of mathematics are in the
cases where infinity definitely cannot be found (for instance, in applied
physics, engineering, biology, social sciences and so on.). We certainly
do not want a philosophical or logical explanation of the applicability
of mathematics to depend on some unknown contingent facts such as
the infinity of the universe. So there is a general puzzle here: how can
those entities so unlike things in the real world help to deliver truths
about things in the real world?
Sometimes it is claimed that realism in mathematics is the best
explanation for the applicability of mathematics. That is not correct
for the puzzle here. Realists face the same puzzle. It does not help to
claim that the continuous model of fluid is itself a mind-independent
abstract entity and our mathematical theorems about it are literally
true of it. The real puzzle is due to the fact that the model is so
radically different from the real fluid and our mathematical theorems

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 160

about the model could not be translated into truths about real fluid
directly.
If the fluid were indeed structurally isomorphic to the model, that is,
if it were indeed a kind of continuous stuff, then we could try to dispense
with the model and talk about the fluid directly. This is Hartry Field’s
strategy in his nominalization of physics. In other words, assertions
about the model would be translated into true assertions about the
fluid. In that case, the applicability of the model would be accountable
without assuming realism. Actually even realists would need the same
explanation for its applicability, for what is to be explained is how
truths about mathematical things are related to truths about physical
stuff. Realists would also need the isomorphism to explain the relation.
The mind-independent status of the continuum itself does not seem to
be relevant here.
One can see that the puzzle is logical in nature and the solution
should be mostly technical rather them philosophical. We need a more
detailed logical analysis of how mathematics helps to deliver truths
about real things in applications. Moreover, it is a puzzle for logicians
and philosophers. Scientists are mostly guided by intuitions, experi­
ences and pragmatic considerations. They do not worry about the lack
of strict logical clarity in their work. Physicists use non-rigorous math­
ematical tools freely. We should distinguish what is a good methodol­
ogy for working scientists from what logicians and philosophers want to
know in studying how exactly the methodology works. Logicians and
philosophers can demand a clearer and logically more detailed expla­
nation of the applicability of mathematics, even though the failure of
logicians’ and philosophers’ search does not imply that scientists should
abandon their methodology. This is the true spirit of analytic study of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 161

mathematics and its applications in science, where ‘analytic study’ is


understood seriously. It is different from the casual claim that mathe­
matical entities are on a par with electrons because scientific theories
refer to them. Substantial logical analyses are needed to resolve the
puzzle here.13

2.2. On Applications of Strict Constructivism. One way to


approach the question is to study whether infinity can be eliminated in
mathematical applications. Infinity accounts for the essential difference
between the mathematical world and the known physical world. It
seems that in applications infinity almost never obtains in reality, and
infinity is mostly introduced to simplify our models (as in the case of
fluid dynamics). So it is natural to think that the logical structure
of applications may become clearer if we eliminate infinity and restate
the mathematical inferences in applications in the language of finitary
mathematics.

This is the strategy already explained in the Introduction at the


beginning of this dissertation. That is, mathematics can be divided
into two parts: the finitary part and the infinitary part. Strict con­
structivism developed in Chapter 1 belongs to the finitary part whose
applications can be accounted for in the sense that fictional things
can be eliminated directly and it becomes explicit how the conclusions
about real things logically follow from the true assumptions about real
things.

I3Physiciat Eugene Wigner’s paper ‘The unreasonable effectiveness of math­


ematics in the natural sciences’ [105] is constantly cited by philosophers in dis­
cussions of the applicability of mathematics. It seems that it is logicians’ and
philosophers’ duty to look for a logical analysis of mathematical applications that
may help to answer Wigner’s question.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 162

Now, Chapter 1 and Appendices A, B prove that a significant part


of classical applied mathematics can be translated into mathematics
within strict constructivism. These include the ordinary (classical)
mathematical tools needed for developing the basics of quantum me­
chanics. Therefore the applications of that part of classical infinitary
mathematics can be accounted for by reducing to applications of strict
constructivism. In other words, from the methodological point of view,
classical mathematics is more efficient, but when one demands an exact
logical account of its applications, we can reduce the applications to
applications of strict constructivism.

In this subsection I want to explain how finitary mathematics is


applied and why its applications are easy to account for.
First, it has been explained in the last section how truths about
natural numbers as fictional entities are related to truths about physical
objects in counting and measuring. The structure of the sequence of
natural numbers from 1 to n is isomorphic to that of a sequence of
n objects. Actually, we first observe some combinatorial properties
of ordinary physical objects, and then conceive of numbers as objects
sharing the same combinatorial properties as physical objects. As a
result, the truths such as ‘2 + 3 = 5 ’ will become representatives of truths
about real things of the same format, namely, ‘2 apples plus 3 apples
are 5 apples’, ‘2 inches plus 3 inches are 5 inches’ and so on.

Here, the applicability of fictional things is clear. Fictional things


are intentionally created to simulate real things, and as far as com­
binatorial properties of real things are concerned, fictional things are

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 163

isomorphic to real things in applications .14 Therefore truths about


fictional things summarize truths about real things of some patterns.
The applicability of algebraic laws such a s m + n = n + m can be
explained in the same way. Such general laws first summarize truths
about numbers of some patterns, namely, 2 + 3 = 3 + 2 , 3 + 5 = 5 + 3 and
so on. Then they further summarize truths about real things of some
similar patterns.
Consider mathematics developed in the system P R O . The language
of P R O is quantifier-free. Its theorems are actually very similar to
elementary arithmetic laws such a s m + n = n + m, though programs
more complicated than addition are involved. To make things clearer,
we must distinguish two stages of applications.
The objects of strict constructivism are numerals and programs.
We must distinguish concrete realizations of programs from fictional
programs. We can conceive of fictional numerals and programs as rep­
resentatives of concrete realizations. These are ordinarily called nu­

meral or program types. We can conceive of arbitrarily large numerals


and arbitrarily complicated programs. Strict constructivism as a the­
ory can be seen as a story about these fictional numerals and programs.
On the other hand, there are concrete realizations of numerals and pro­
grams. They can be marks on papers or memory states of computers.
It is possible that not all fictional numerals are concretely realizable,

for the real world may be finite.


In the first stage, we consider applications of theorems in P R O to
concrete realizations of numerals and programs. A theorem of P R O

u The sequence of natural numbers may not be isomorphic to any aspects of


reality, for there may only be finitely many physical objects. But in any concrete
applications of numerical identities of the form m +n=k, we have a one-one corre­
spondence of some finite segment of natural numbers and the real things counted.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 164

can be reduced to an equality t = s for some terms t, 3 of P R O of the


type o. A proof of this theorem is a sequence of formulas of P R O . To
apply t = s in a concrete situation, we substitute for the free variables
in t , s some closed terms, normalize and then obtain an equality be­
tween two executable programs t* = s*. See the discussion at the end

of Chapter 1, Section 2. When these programs are realized as com­


puter programs, the equality is interpreted as an assertion about the
outcomes of two concrete computer programs15. This is how theorems
of P R O are related to truths about real things.
In Chapter 1 it is proved that from the proof of t = 3 in P R O we
can obtain, primitive recursively, a proof of t* — 3 * in P R A . Because

t* and 3 * are closed, inductions in P R A can also be eliminated. So we


actually have a derivation of t * = s* from some concrete instances of
the definition axioms of primitive recursive functions. These instances
of axioms of P R A can all be translated into true assertions about
concrete realizations of the terms as computer programs. Then the

proof in P R A will be translated into a derivation of a truth about


real things from some other truths about real things. The references
to fictional things are thus eliminated, and the proof about fictional
things in the theory is translated literally into a proof about real things
in which no more fictional things are referred to. That is how proofs
in P R O can help to derive truths about real things.
The fact that the language of P R O is quantifier free plays a crucial
role here. It makes it possible that P R O is applicable to reality even
if the real world is essentially finite. Without using quantifiers, one

lsIn the realistic situation, one must consider questions such as memory over­
flows, running time limits and so on. So the assertion about real programs should
be stated as something like ‘If the physical conditions are normal, the outcomes
will be ...’

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 165

cannot even claim that the successor function S is a total function.


To express that we must use quantifiers, namely ‘For all numerals n,
Sn is also a numeral.’ Indeed in P R O we have the open formula
m + Sn = S (m + n) as an axiom, but it is understood as a schema. It
means that when it is applied to reality, it can be ontologically neutral.
It does not demand the existence of real things by itself. If we can
realize some numerals m ,n and the programs + , S , the schema only
says that if the programs with the numerals as inputs as indicated by
fn + Sn and S (m + n) can be successfully executed (say, there are no
memory overflows), then the output will be the same. In other words,
even if there are only finitely real numerals, we can still interpret some
instances of the schema as true sentences about the real numerals. The
applicability of the schema to reality in a specific case does not demand
that all fictional numerals be instantiated in reality. Actually it does
not depend on what exactly are in the domain of real things. The
application has this feature of locality: One can just take whatever one
knows to be real and instantiate the open schema by those real things.
The case is different for quantified sentences such as Vm3n (n = S m ) .
If we want to interpret it into a statement about real things, we must
determine the domain of the interpretation of quantifiers Vm and 3n.
The interpretation of a single sentence must refer to all entities in the
domain. There is no locality as above. Moreover, for the intended
interpretation of the sentence to be true, the domain must be infinite.
Therefore, if the sentence is used as an axiom in a proof, the proof
cannot be literally translated into a derivation of a truth about real
things from truths about real things, for there is no way to know that
the literal interpretation of the axiom is true.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 166

So far only applications to concrete realizations of programs are con­


sidered. At the second stage, we will consider applications to science.
The basic idea is that in such applications we use concrete programs
to model real things. Generally, in applying mathematics within strict
constructivism to physics, we use programs to express physical quan­
tities, for instance, positions and velocities of physical objects. These
quantities may depend on some parameters, for instance, the time pa­
rameter, which will be inputs to the programs. So programs summarize
information about physical systems. Physical laws are then expressed
as conditions on these programs. Resolving a physical problem corre­
sponds to constructing the programs and verifying the conditions.
Here we consider a more concrete example. Recall that rational
numbers, real numbers, functions of real numbers are represented by
programs. Suppose we want to describe the motion of a physical object.
For simplicity, suppose that the object is constrained to a straight
line. We associate the numeral 0 with a chosen point on the line as
the origin of a coordinate system and then associate other programs
representing real numbers with points on the line. We do not have
to make the universal assertion that any real number corresponds to
a point. Actually, it is undecidable whether a program represents a
real number, or whether two programs represent the same point. For
practical uses, it suffices to use rational numbers to represent points.
On the other hand, we can agree that any physically determinable point
can in principle be represented by a concrete program approximately.
Similarly, we choose an instant as the origin of the temporal coordinate,
then physically determinable instants are represented by programs.
The position of the object is represented by a real function x (t),
which is also a program. With a program t representing an instant as

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 167

input, it outputs the coordinate of the object. Suppose the force acting
on the object depends only on time and position; then it is represented
by a function / (x, t), which is also a program. The mass of the object
is represented by a real number m. Further suppose that the object can
be treated as a particle. Then the third law of Newtonian mechanics
is a condition about the program x (t ) representing the position of the
object:
cPx

If we can prove in strict constructivism (the system SC ) that the


solution to this differential equation (for t in some finite interval, with
some initial conditions) exists, it means that we have constructed some
programs x ( t ) , v (i) and the related moduli of differentiability, and
proved in P R O some (quantifier-free) formulas expressing that — = v
and TOg|r = / (x, t) (with the related moduli of differentiability). (See
Chapter 1 , Section 6 for the relevant definitions.) More specifically,

= f ( x >0 becomes (for t X)t2 in the interval)

(7) ^ < |ia - i , | ( M ) < 1 1


M 1‘ ' u (N ) M

\ k )
E j — Ei 771

^ 1 1
“ N + K'
where N , M , K , t i , t 2 are free variables, and u is the modulus of differ­
entiability.
Now consider how such an application makes the applicability of
mathematics clearer.
First, it shows that fictional things are eliminable. To eliminate
references to fictional things (namely, fictional programs), we stipulate
that we use concrete programs to model the motion of the object. Then

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 168

physical laws state what concrete programs correspond to the position


of the object in various contexts (determined by the nature of the
force and initial conditions). They characterize the programs by some
conditions about the programs such as the condition ( 8 ). Mathematical
theorems state that those programs can be actually constructed. A
proof in P R O of a closed instance of condition such as ( 8 ) can be
translated into a derivation of a truth about the real programs from
ofher truths about real programs.
So both physics and mathematics deal with real things here. Mathe­
matics deals with things that we can directly manipulate, for instance,
data and programs on computers16. In doing physics, we use things
that we can directly manipulate to model things out of our direct con­
trol. Physical laws state how the former model the latter. This is an
indirect way of describing the course of nature, but it is more efficient.
Second, approximations can be made clearer. Real physical quanti­
ties are always fuzzy and can only be measured approximately. When
programs are used to represent points, instants and other quantities,
degrees of approximations correspond to limits on the magnitude and
precision of inputs and outputs. If x is a program representing a real
number, x (n) gives approximate positions of a point given a coordi­
nate system. Physical laws are also valid only approximately. This is
expressed by the stipulation actually that not all instances of the condi­
tions on programs such as ( 8 ) need to be true for real physical quantities
represented by the programs. For example, for some numerals N, M , K
sufficiently large, the condition ( 8 ) for rational numbers t\ < satis­
fying the antecedent of the condition says that the acceleration of the

16Of course, computers are referred to as examples. In the old days we did
calculations on paper. Programs should also be understood as instructions in some
general sense.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 169

object calculated by measuring the velocity increment v (£x) — v (t2)


in the short temporal interval [ti,ta] is approximately ^ / ( i (tx) , £x).
This says exactly that Newton’s third law is approximately true. As the
numerals N , M , K increase, the degree of approximation also increases.
On the other hand, if a condition involves universal and existential
quantifiers ranging over all real numbers or even all real functions, it is
unclear in what sense the condition can hold only approximately of real
physical quantities. The simple explanation above will not be available.
Of course, the real point should be the other way around: Given that
physical laws are only approximately true of reality, we suspect that
the laws can eventually be expressed in a quantifier-free language.
Third, it becomes clearer how mathematical proofs preserve approx­
imate truths. Since general physical laws are only approximately true
of reality, the mathematical reasonings we use to deduce predictions
must be able to preserve approximate truths. In other words, what
we demand of the mathematical proofs in applications is approximate

truth preserving, not merely truth preserving.


Now suppose we deduce that x ( l ) = 1 in the above example, with
the initial conditions x (0) = 0, v (0) = 0. That means we have a proof
in SC of x ( l ) = 1 from some assumptions, for instance,

(8) x (0) = 0, v (0) = 0, Vtx, t u N, K , M

and a similar formula for the condition ^ = v. A quantifier free


instance of the conclusion x ( 1 ) = 1 will be |x ( 1 ) (L ) — 1 | < 1fL , stat­
ing that x ( l ) is approximately 1 . From the discussion in Chapter
1, Section 3 we know that the proof actually constructs some terms
t\, t\, ..., t\ , Nk,K k ,M k and deduces | x ( l ) ( L ) - 1| <
l / L from the instances of the condition ( 8 ) obtained by substituting

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 170

these terms for the free variables t i , t 2, N , K , M there (together with


other assumptions). So the proof actually proves that for the predic­
tion x ( 1 ) = 1 to be approximately true to some degree (determined by
the magnitude of L), it suffices that the condition ( 8 ) is approximately
satisfied at some instants t\, ..., t*, t \ and to some degree (deter­

mined by K \, M \ , N k , Kfe, Mk). The proof explicitly constructs


these instants and the required degree of approximations (determined
by' N{, K{, Mi). Therefore, by examining the proof, we can see that
if Newton’s third law is approximately valid to some degree at some
specific instants, the prediction will also be approximately true to some
degree. This is how the proof preserves approximate truths.
This analysis relies on the fact that when we prove a sentence of
the form Vn<p(n) —» Vmi/> (m) in strict constructivism, we actually
construct terms si [m], ..., Sk [m] and proves

<p ( s i [rrc]) A ... A tp (ak M ) - * 4 > ( m ) .

That can be interpreted as saying that the consequent is true up to


m as long as the antecedent is true up to some s [m]. In the example
above, this becomes: the prediction is approximately true to some de­
gree as long as the laws are approximately valid to some related degree.
Moreover, recall that terms [m] all represent primitive recursive func­
tions. Their growth rates are relatively clear if we obtain the definition
of the functions. The case will be different if we use the language of To.
Of course, the real point is again the other way around. If we expect
that mathematical inferences in applications can preserve approximate
truths in some realistic sense, the growth rates of the terms st- [m] must
be limited. So we expect that the terms st- [m] can be constructed from
some simple common functions.

R ep ro d u ced with p erm ission of the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. ON APPLICATIONS OF STRICT CONSTRUCTIVISM 171

Generally, because the language for expressing the physical laws is


quantifier-free and the mathematical proofs are within a finitistic and
constructive system, the logical structure of the applications becomes
clearer. References to fictional entities are not essential. The way by
which physical laws approximate reality is clearer, and the reason why
logical inferences preserve approximate truths can be explained.
The example treated here is relatively simple. A full study of the
logical structure of mathematical applications should be a very big
topic. But I hope this already indicates that the puzzle of applicability
can be resolved for applications of strict constructivism and therefore
it can also be resolved for applications of classical mathematics that
are reducible to applications of strict constructivism.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 172

3. R em ark s on th e E p iste m o lo g y o f A r ith m etic

3.1 . In tr o d u ctio n . Now, consider the epistemological basis of ap­


plications of mathematics. Here I emphasize applications of math­
ematics, for according to fictionalism, pure mathematical truths are
truths within stories, and therefore there is no interesting epistemo­
logical question about pure mathematics. On the other hand, when
mathematics is applied to reality, the truths obtained are truths about
reality. Then epistemological questions will arise.
First I want to indicate that focusing on applications can help to
clear away some confusions in epistemology of mathematics. For exam­
ple, the epistemology of arithmetic is plagued by the lack of consensus
on the meanings of pure arithmetic statements such as ‘2 + 3 = 5 ’. Are
they about some mind-independent entities? or should they be con­
strued in a different way? On the other hand, applying ‘2 + 3 = 5 ’ we will
obtain ‘2 apples plus 3 apples are 5 apples’, ‘2 pounds and 3 pounds are
5 pounds’ and so on. These are assertions about real things and their
meanings are very clear. So we can ask what kind of truths they are.
Are they logical or analytical truths? Or otherwise, are they empirical
truths just as other scientific truths?

After these questions are answered, there seems to be no more in­


teresting epistemological question remaining for ‘2 + 3 = 5 ’ itself. When
some philosophers try to hold to realism in arithmetic, it seems that
they just create some extra and more troublesome epistemological prob­
lems for themselves, while the original question about ‘2 apples plus 3
apples are 5 apples’ and so on is still to be answered. Fictionalism plus

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 173

the shift of focus to applications seems to put the real epistemological


question on the right track.
When applications of infinitary mathematics are reduced to those
of strict constructivism, epistemological questions about applications
are also reduced to those about applications of strict constructivism.
In the last section we saw that applications of strict constructivism are
not essentially different from applications of elementary arithmetic in
counting. As is indicated there, when applied in science, mathematics
in strict constructivism deals with concrete realizations of programs
and numerals. So actually, it deals with combinatorial properties of
macroscopic physical objects. The assertion that the operation
applied to numerals ‘2’ and ‘3’ will result in the numeral ‘5’ is much
like the assertion that counting 2 apples, counting 3 apples and counting
them together again will result in 5 apples.
This shows that philosophy of elementary arithmetic should be of
great significance in philosophy of mathematics. It also shows that
traditional philosophies of arithmetic, in particular, Kant’s and Mill’s
philosophies, are still of genuine interests in the contemporary context.
The development of modem mathematics and modern logic does not
really make them obsolete.
Clearly, the reduction of applications of infinitary mathematics
helps to pose the epistemological problem in this way. The analysis

given in the last section helps to separate applications of mathemat­


ical theorems from applications of physical laws. The former deals
with combinatorial properties of things we can manipulate directly,
while the latter deals with how to use these to model other things,
namely, natural processes. Therefore, epistemological questions can
also be separated. The validity of physical laws is clearly an empirical

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 174

matter, while the epistemological basis for combinatorial truths about


macroscopic physical objects is perhaps more subtle. For example, the
Kantian view may need to be considered. On the other hand, when
infinitary mathematics is used in building models to simulate nature,
what are mathematical and what are physical are mixed. Then either
one just considers the nature of truths in pure mathematics and nat­
urally accepts realism or formalism (or similarly, conventionalism), or
when applications of mathematics in science are considered, one simply
ignores the role of mathematics and considers only the epistemological
basis of empirical science. The specific question of the epistemological
basis of arithmetic is hidden.
Frege’s logicism, Mill’s empiricism and Kant’s view that arithmetic
is synthetic a priori are representatives of traditional views on arith­
metic. In this section I will give some comments on these views on
arithmetic. I am unable to present a complete theory of the episte­
mology of arithmetic. This section will contain only some preliminary
remarks. The objective is to indicate what an epistemological basis
of arithmetic should look like and to suggest directions for further re­
searches.
This section will begin with a discussion of some reasons supporting
Mill’s empiricism. The main point is that when we apply numerical
equations in counting physical objects, the applicability depends on

some features of the physical objects, for instance, the features that
they are determinate and identifiable17. These may be in doubt if
we try to apply numerical equations to strange things such as funda­
mental particles. But I will also express some reservation about the

17Theae attributes are intended to characterise physical objects treatable in


classical mechanics, but I did not attempt to give any complete characterisation of
objects to which arithmetic is applicable, neither here nor elsewhere in this section.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 175

radical empirical view of arithmetic, owing to some Kantian consid­


erations: Given our human faculty of cognition, the objects we can
directly perceive or conceive of may necessarily conform to arithmetic
laws. Arithmetic can still be necessary or a priori in this sense. These
discussions of empiricism will constitute Subsection 3.2 below.

Then the Fregean reduction of simple numerical identities to logic


will be discussed in Subsection 3.3. I want to argue that while the
reduction may have some theoretical interest, it does not help in clar­
ifying the epistemological basis of our knowledge of applications of
arithmetic. For one thing, when we consider applications to things
remote from us (say, things in the quantum world), even the applica­
bility of logic is contingent on the outer world, and for another, when we
consider applications to things closest to us (say, the ordinary physical
objects immediately perceived by us), the reduction failed to show that
arithmetic truths can be known by analyzing the concepts alone, for
intuition and arithmetic knowledge are already assumed in the alleged
analysis of concepts.

In Subsection 3.4, I will come back to give more remarks on Mill’s


and Kant’s views on arithmetic. I will discuss the possibility of a
compromise between them, which seems to be the only choice for a
proper philosophy of arithmetic. The subsection will give only some
suggestions. No attempt will be made to reach a complete theory.

3.2. On Empiricism in Elementary Arithmetic. B y ‘empiri­


cism in arithmetic’, I mean the view that truths obtained from the
applications of arithmetic in counting ordinary physical objects, such
as ‘2 apples plus 3 apples are 5 apples’, are not essentially different from
other empirical truths. They are contingent on the way the world is.
This is very close M ill’s empiricism in arithmetic, though Mill will add

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 176

that those truths can be subject to empirical tests and the apparent
difference between them and other empirical truths is due to the fact
that they have been verified numerous times. See, in particular, Mill’s
A System of Logic, Book II, Chapter V and Chapter VI [67].
In this subsection I will first discuss reasons that may support em­

piricism in arithmetic, and even empiricism in logic, from the modem


perspective. One of the main reasons comes from the modem discovery
that'the logic for describing things in the quantum world may have to
be different from classical logic18. I will argue that these reasons are co­
gent as far as they go, and actually they entitle us to claim empiricism
more confidently than Mill did. But I will also indicate that Kant’s
view on the a priori status of arithmetic is still of interest, and it is not
necessarily contradictory to empiricism when properly formulated.
As has been explained, the focus will be on applications of arith­
metic. For simplicity, I will consider applications of the simple example
‘2 + 3 = 5 ’. It seems that the applications cam be largely classified into
three types. This is not supposed to mean a strictly exhaustive and
clear-cut classification. It suffices that there are clear distinctions in
typical cases.
The first type of applications fits Mill’s empiricism best. This
includes the instances of measuring physical quantities and counting
‘strange’ physical objects such as fundamental particles. Examples of
the former are: ‘If A drives 2 miles per hour and B derives 3 miles per
hour relative to A, then B derives 5 miles per hour’, ‘If 2 cups of water
and 3 cups of water are poured into a container, it results in 5 cups of
water’, *2 pounds of apples and 3 pounds of oranges weigh 5 pounds’

iaThis is the so-called quantum logic. It is fiist presented by von Neuman in


1930’s. Putnam is perhaps the first who emphasised the philosophical implication
of quantum logic, the implication that logic is empirical. See Putnam [80].

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 177

and so on. These are true statements about measuring physical quan­
tities, but the truthfulness is contingent on some special features of the
physical objects and quantities. Moreover, they are only approximate
truths, and they are exactly true only in the hypothetical sense.
The example about speed is obvious. After the discovery of rela­

tivity, it is clear that whether or not speeds can be added in such a


simple way is contingent on some special features of the outer world.
It is contingent on if the outer world is Newtonian or relativistic, and
it is contingent on if we use an inertial frame to measure the speeds,
which can only be determined empirically.
Of course, the assertion is necessarily true for a hypothetical and
strictly Newtonian universe. However, epistemology is primarily epis­
temology of our knowledge of this universe we live in. So, whenever
we consider applications of arithmetic, we mean the truths about this
universe obtained from the applications. In other words, we mean the
applications of arithmetic in concrete contexts of the real world. Then,
empiricism claims that the truths thus obtained are contingent on the

real world.
The second example above is contingent on the fact that when two
samples of the same kind of liquid are mixed, the volume is stable. The
statement will not be true if it is about mixing water and alcohol. This
example may look trivial or even absurd to some people. Certainly
we do not apply ,2-i-3=5, to mixing water and alcohol in counting vol­
umes. But the point is just that it is an empirical question if arithmetic
is applicable in a specific situation. Experimental tests are needed to
justify the applicability. There is no guarantee that applications of
arithmetic in whatever way always lead to truths. Of course, empiri­
cists are not objecting that whenever arithmetic is applicable it leads

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 178

to truths, where ‘being applicable’ can only mean ‘leading to truths


when applied’. This is merely a tautology.
The example about mass is similar. When the apples are considered
in isolation, we can assign (rest) mass to them, and the same is true for
the oranges. Now when we consider the apples and oranges together in
the real world, we should consider the interaction between them. The
interaction must exist (for instance, the gravitational interaction must
exist), though it is negligible. Interaction means (binding) energy and
energy means mass. So strictly speaking, the (rest) mass of the system
of apples and oranges is not the sum of the (rest) mass of the apples
and that of the oranges. Surely the difference does not really make
sense for apples and oranges, but it is very significant for neutrons and
protons, and it is the well known basis for atomic bombs. So the truth
is also contingent on some special features of the outer world and it
should be subject to empirical test, just as special relativity and the
equivalence of mass and energy should be subject to test. One might

contend that although the rest mass of the system is not the sum of
the rest mass of the subsystems, we can define a different notion of
mass so that the rest mass of the system is still the sum of the newly
defined mass of the subsystems. But the point of empiricism about
the applicability of arithmetic is just that experiences are needed in
judging which applications lead to truths and which do not.
The above examples are all about measuring physical quantities.
One might contend that the primary applications of arithmetic are
in counting objects, or in other words, they are about combinatorial
properties of things, not other physical quantities. However, when we
try to express how many fundamental particles there are in the com­
bination of two subsystems of particles, the same problem as in the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 179

case of mass may occur. In fundamental physics, there is the so-called


random creation and annihilation of particles. The particle-number
property of a physical system is treated in the same way as other phys­
ical quantities, such as energy, mass, charge and so on, and moreover
the particle-number property is related to other physical quantities. It
might happen that the number of particles in a combined system is not
simply the sum of those of subsystems.
I am unable to give a very concrete example about this here. So I
will turn to the applicability of classical logic in quantum mechanics.
If the applicability of logic is already in question, the applicability of
arithmetic certainly should be the same. Now, it is well known that
the meanings of logical constants such as conjunction, disjunction can
become unclear when we try to apply them to describe the quantum
world. Due to the incompatibility of properties, the conjunction of two
properties may become meaningless. Similarly, due to indeterminacy
in the quantum world, the disjunction of two properties may obtains
without either of the disjuncts being obtained. When we attribute
properties to particles and try doing the logical calculus of properties,
some laws of classical logic (specifically, the distribution laws) may fail.
Notice that we do not need to assume that modem physics is lit­
erally true. To reject the claim that applications of logical laws and
numerical identities always lead to necessary truths, it suffices to real­
ize that the picture m odem physics gives us is possible. In other words,
things far from us could be beyond our imagination, so that our famil­
iar logical and numerical concepts might not be suitable for describing
them. Only the possibility of the strange, ineffable quantum world is
needed for rejecting the universal necessity of logic and arithmetic. Put
in another way, what m odem physics has done is to open our eyes and

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 180

make us realize the possibility of a world radically different from our


familiar world, so that some forms of beliefs, logical laws, of which we
feel so certain may lose their meanings for that different world. There
is no absolute guarantee for any forms of beliefs.
Certainly, modern physics is not unique on this point. There are
many similar cases in the history of science, from the ancient belief
that the earth is necessarily flat to the 17th century belief that distant
action is impossible. Mill had a long discussion on the thesis that geom­
etry and arithmetic are necessary because the contrary is inconceivable
([67] pp 236-251), citing many such examples to show that there is no
a priori knowledge based on the inconceivability of the contrary. To­
day, with the development of modern cosmology and quantum theory,
we can have examples that apply to logic, arithmetic and geometry
directly. We can express Mill’s view on the contingency of arithmetic
more confidently than Mill did.
This concludes the discussion on the first type of applications and
the morals we can draw from it. In the second type of applications,
we apply numerical identities in counting determinate and identifiable
macroscopic physical objects, such as pebbles, apples and so on. Here
we seem to always have exact and certain truths. However, by com­
paring them with the previous examples, it seems still reasonable to
say that the truths obtained here are still contingent on the fact that
the objects counted are determinate and identifiable, and are therefore
different from fundamental particles. Moreover, there is no clear-cut
distinction between these two types of objects.
Consider the discussion on the logic of the quantum world again.
One may object that we actually still use classical logic and arithmetic
in describing the quantum world. This doubt can be explained away by

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 181

considering the third type of applications. Here we apply arithmetic


to fictional entities, including fictional mathematical entities such as
numbers, functions, sets and so on. Our imagination is constrained
by our direct perceptual experience with the determinate and macro­
scopic part of the outer world. So we have to conceive of fictional things
as determinate and identifiable. That means arithmetic and classical
logic are automatically true of fictional entities just as they are true
of macroscopic physical objects. Arithmetic and logical truths do have
necessity here, because we decide them to be thus. Arithmetic and
logical truths are also conventional in this type of applications in the
following sense: In telling the story about the fictional things, we al­
ready implicitly assume that those fictional entities are determinate,
just as ordinary physical objects, and therefore arithmetic and logic
are true of them. On the other hand, the adoption of this convention
is contingent on the fact that our imaginative ability is constrained by
the outer world and fictional things have to be similar to macroscopic

physical objects.
Then we can see how we use classical logic and mathematics in
describing the quantum world. First, in classical mechanics, our models
of particles directly correspond to the particles. The position of a
particle p corresponds to a point x in 3-dimensional Euclidean space.
The physical statement ‘Particle p is in the space region R * about the
particle corresponds to a mathematical statement ‘x £ D \ where D is
the set of points in 3-dimensional Euclidean space corresponding to the
region R in real space. Then the model directly corresponds to reality
in the following sense: The physical statements *p is in the space region
R i and/or p is in the space region R 2 are translated as ‘x E D \ and/or
x € D 2\ where D \ t D 2 represent R i, R 2 respectively. It means that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 182

the logical constants ‘and’, ‘or’ have the same meaning when they are
used in physical statements about the particle as when they are used
in mathematical statements about the model of the particle.
When we describe the positions of a particle in quantum mechanics,
we do it differently. If p a quantum mechanical particle, the position
property of p will be represented by a vector i in a Hilbert space. The
physical statement ‘ p is in the space region R' about the particle will
correspond to a mathematical statement ‘x € D \ where D is a sub­
space of the Hilbert space corresponding to the space region R. Then
the physical statement ‘ p is in the space region Ri or p is in the space
region R 2 will not be translated into ‘x 6 D\ or x 6 D 2 . Instead it
will be translated into ‘x 6 D \ U D 2\ where D\ U D 2 is the subspace
spanned by D \ and D 2. Therefore, the logical constant ‘or’ used in
statements about the particle and the ‘or’ used in statements about
mathematical entities have different meanings. When we talk about
mathematical entities such as vectors, subspaces, numbers and so on,
we still use classical logic. That must be the case, because we can
only conceive of entities that conform to classical logic. But when we
translate physical statements about quantum mechanical particles into
statements about their mathematical representations, we do it indi­
rectly. Physicists cleverly define some mathematical operations and let
them represent the logical operations on physical statements, for these
logical operations mean something different from the logical constants
we use in mathematics.
So what we do here is try to use conceivable things to build a
model to describe things we cannot directly conceive of. We may not
be completely successful in doing this, and actually that is just the
source of conceptual puzzles about quantum mechanics, but that is the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 183

only thing we can do given the limitation of our imagination. Indeed,


we can imagine that if there were intelligent beings who could perceive
the quantum world directly, their mathematics might be different from
ours, in that they might study some mathematical structures we cannot
imagine, and their theory of quantum mechanics might also be different.
This seems to show that there is no contradiction in claiming that
the logic for the quantum world could be different from our familiar
world, while granting that we still use classical logic in quantum me­
chanics. It also seems to corroborate the claim that at least things
remote from us could be radically different from the familiar things, so
that our ordinary logical concepts are not applicable to them. Recall
that only the possibility of this is needed for rejecting the universality
of classical logic. Therefore, we seem to have strong support for the
claim that classical logic does not have unlimited universality19.
This finishes the discussion of supports for Mill’s empiricism. On
the other hand, the discussion also hints that classical logic and arith­
metic may have necessity in one sense: It is probably true that the only
things we human beings can perceive or conceive o f directly and clearly
must obey classical logical and arithmetic laws. Recall that even when
we want to study other kinds of logic, we still have to use classical logic
and arithmetic. All we can do is to represent other kinds of logic in

our mathematics in some indirect way.


This seems plausible. However, it should be distinguished from the
metaphysical picture of logical truths as necessary truths, which is the
picture that there is a mind-independent domain of logical concepts and

19Also notice that the discussion fits the framework of fictionalism and the
emphasis on applications nicely. If logical and arithmetic truths are understood
truths about abstract, mind-independent entities or concepts, there is indeed little
sense in empiricism about arithmetic and logical truths.

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 184

logical truths, and they are invariably applicable to all possible things.
This metaphysical picture is rejected in the empirical view espoused
above, for it was argued that the scope of universality of classical logic
does not cover all possible things. Classical logic and arithmetic can
only be necessary relative to human faculty of cognition, and for things
we can directly perceive or conceive of.
This means that Kant’s view is still of some interest and it is not
contradicted by empiricism. Even though the universality of arithmetic
is restricted to things we human beings can directly perceive or conceive
of, it is still of some interest to explain this limited universality by
resorting to the idea that we impose our form of sensation on the senses.
If we stop at empiricism, we seem to commit ourselves to the picture
that our minds are originally blank and they passively but objectively
receive whatever is impressed on them, and the picture of nature our
minds have is just the picture from God’s eye. One difficulty with
this picture is that it makes it hard to explain the limitation of the
universality of arithmetic. In other words, since we already grant that
logic and arithmetic are not universal, there must be an explanation
about why we never perceive otherwise. If one refers to the limitation

of human capacity of cognition, it amounts to conceding that our minds


are not really blank so that they can receive anything objectively. They
are actually selective, and thus they do have some innate structure.
The picture given by Kant seems to be more reasonable. So, con­
ceding all the reasons supporting empiricism, we may still have to look
for a compromise between Mill and Kant for a proper account of arith­
metic from the human perspective. What we need is a more careful
formulation of the situation we are in. Logic and arithmetic are not
a priori, if being a priori means being universal to all possible things,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 185

but they are still a priori in the sense that they are necessary, relative
to us, for things we can perceive or conceive of. Then we will see that
there is probably no contradiction here. I will come back to this point
afterwards.

3 .3 . O n R e d u c in g A r ith m e tic to L ogic. Here I will discuss the


Fregean thesis that arithmetic truths are logical truths. I will argue
that although the reduction of some arithmetic truths to logical truths
may have some theoretical interest, it does not really help in clarifying
the epistemological basis of arithmetic. Similar views have been lightly
mentioned by some authors20, but I want to show that the framework of
fictionalism, the focus on the epistemology of arithmetic applications
instead of pure arithmetic truths, and empiricism of the last section
can make the point clearer and more convincing.
Many contemporary philosophers seem to agree that at least simple
numerical identities such as ‘2 + 3 = 5 ’ are reducible to logic, though it
is unfortunate that Frege’s program of reducing all arithmetic does not
work. To avoid difficulties that may arise with second-order logic or
Frege’s axiom (V), it is frequently asserted that ‘2 + 3 = 5 ’ can already
be reduced to a first-order logical truth, namely,

(9) 3 j x F x A 3$xGx A - ’3 s (F x A G x) -* 35s ( F x V G x ) ,

where 32 s (...), 33x(...) and so on are the standard first-order expres­


sions for numerical properties (using quantifiers and identities). The
sentence is certainly a logical truth in first-order logic. I will focus on
this reduction of ‘2 + 3 = 5 ’ to logic below.
From the epistemological point of view, what can we really achieve
by this reduction? Originally, the reduction is supposed to prove at

J0See for example Parsons [74], Young [106].

p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 186

least the following: ( 1 ) Arithmetic truths, just as logical truths, can be


stated as universal and topic neutral. They are valid for any domain
of things, not just the numbers21. More concretely, F, G in the formula
can be any predicates of any objects. Therefore the formula shows how
‘2 + 3 = 5 ’ is applicable in counting any objects; (2) Arithmetic truths
are conceptual, or analytical truths. They can be known by analyzing
the concepts involved alone, and no intuition is necessarily involved; (3)
Arithmetic truths are true by virtue of their logical forms, and therefore
they have nothing to do with the speciality of human sensibility. As
we distinguish the applications of arithmetic from the pure arithmetic
statements and focus on applications, by ‘the arithmetic truth 2 + 3 = 5 ’
we always mean truths of the same form as ‘2 apples plus 3 apples
are 5 apples’. So, the question becomes if the reduction can help to
reveal whether or not (1) They are necessarily and universally true; (2)
They can be known by analyzing the relevant concepts alone, without
resorting to intuition; (3) They are true by virtue of uni versed logical
concepts, rather than human sensibility.
First, from the empirical point of view, it is easy to understand that
the reduction does not achieve anything significant in revealing the ne­
cessity and universality of arithmetic, for even logic is not universal.
The reduction indeed shows that whenever logical concepts in first or­
der logic are applicable, arithmetic is also applicable. Informally, this
can be explained as follows: The logical concepts in first-order logic
are supposed to include identities, inequalities and quantifications, and
these can be combined and iterated any finite number of times in con­
structing formulas. If all such formulas can be interpreted meaningfully

J1Frege held that numbers are logical entities, so arithmetic is still actually
universal, although in appearence arithmetic is about some special entities called
numbers.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 187

in some domain of entities and all classical logical laws hold, those en­
tities must be determinate and identifiable, just like ordinary macro­
scopic physical objects, and therefore arithmetic must is true of them.
The reduction gives a formal proof of this point. This seems to the
only merit the reduction has in connection with the universality and
necessity of arithmetic.
As for the second question, the answer is clearly negative. To make
sure if logic and arithmetic are applicable to some domain of things
directly, for instance, to the borderline cases between the classical and
the quantum world, one must resort to experiments. There is no way
to know the truths by analyzing the concepts alone. The reduction
is indiscriminate between the applications of numerical identities to
ordinary physical objects and those to things in the quantum world. It
does not help in revealing where arithmetic is directly applicable and
where it may be problematic.
Even for applications to ordinary physical objects, the reduction
failed to show that arithmetic truths can be known by analyzing con­
cepts alone. The point is that we already assume some arithmetic
knowledge in the so-called ‘analyzing the concepts’. Therefore intu­
ition is still necessarily involved in knowing the logical truths resulted
from reduction.
For example, the so-called knowing ‘2 + 3 = 5 ’ by analyzing concepts
is supposed to be deducing the sentence (9) in first-order logic. Notice
that Bs is expressed by 2 existential quantifiers in first-order language,
and H3, B5 are similar. At some stage of the deduction, we must con­
catenate 2 existential quantifiers and 3 existential quantifiers to form
a sequence of 5 quantifiers. To manipulate formal sentences this way
and to be sure that there are indeed 5 existential quantifiers in the

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 188

result, we must implicitly have the simple combinatorial knowledge


that 2 symbols concatenated with 3 symbols result in 5 symbols. In
other words, ‘2 + 3 = 5 ’ is already resorted to in the so-called ‘analysis of
concepts’.
This will be clearer when we consider other more extreme cases.
Suppose that VnP (n) is a sentence in primitive recursive arithmetic
P R A (formulated in the language with quantifiers) expressing the con­
sistency of an axiomatic system T that is either Z FC or some extension
of it, and suppose that <p is the conjunction of all the (closed) primi­
tive recursive defining axioms of the primitive recursive functions and
predicates involved in constructing P. For any numeral constant mo,
we know that

( 10 ) <p-*P(m5)

is provable in first-order logic if P (mo) is true. Suppose for some reason


we believe that T is consistent. In other words, we believe that P (m)
is true for any numeral m. Let mo be a very large number so that
we have no other way to come to the belief in the truth of P ( m o ).
Then we believe that (10) is a logical truth based on our belief in the
consistency of T. On the other hand, let ip be the conjunction of the
axioms of T and a stronger axiom (for instance, a large cardinal axiom)
that proves the consistency of T. Then we also know that

( 11 ) ^ (y, _* p (m^))

is a logical truth in first-order logic. In this case, even if we do not


believe in the consistency of ZFC, we will still believe that we can
derive ( 11 ) in first order logic.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. REMARKS ON THE EPISTEMOLOGY OP ARITHMETIC 189

Obviously, no one would simply claim that we can actually know


the logical truth ( 10) by analyzing concepts alone, for the existence
of this ‘analysis’ is based on the consistency of T, about which we
may not be so sure, and our belief in it may even be based on a huge
amount of past experience. A proper question to ask here is ‘What
is the basis of our belief in a particular sentence?’ For the sentence
( 10 ), it seems clear that the basis does not consist of first-order logical
axioms alone. It should include the belief in the consistency of T or
whatever justifies the belief, for instance the axiom ip. Comparing (10)
with ( 11) we see that in ( 10 ) we just relegate some of our knowledge
basis from the objective language level to the meta-language level. As
long as intuition or anything else is involved in knowing the basis, we
are not entitled to claim that the sentence can be known by analyzing

concepts alone.
The case of simple numerical identities such as ‘2 + 3 = 5 ’ is similar
though not so extreme. The knowledge of ‘2 + 3 = 5 ’ is already assumed
in the alleged analysis of concepts, namely, the derivation of (9) in first-
order logic. Here, the reduction involves a circularity when we consider
the whole basis on which we come to know ‘2 + 3 = 5 ’. The reduction
does not reduce knowing ‘2 + 3 = 5 ’ to knowing something else.
From Kant’s account of how we know ‘7+ 5= 12’ it seems obvious
that we know arithmetic truths about ordinary physical objects by
intuition, that is, by the immediate awareness of some particular in­
stances, rather than by analyzing concepts. When they are reduced to
logical formulas such as (9), it appears that they become logical truths
and are therefore true by virtue of concepts, and then it appears that
no intuition is needed in knowing them. However, if one really starts
examining how we deduce a sentence like (9), one will realize that we

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 190

still resort to intuition. Instead of five fingers in Kant’s account of


‘7+ 5= 12’, we take quantifier symbols 3s (which is a sequence of 5 ex­
istential quantifiers) as objects of intuition22. In the logical deduction,
we must concatenate 5 quantifiers to 7 quantifiers and then, based on
the perceptual awareness of this singular presentation of a sequence of
symbols, judge that they are 12 quantifier symbols. Clearly, the reduc­
tion just replace our intuition of fingers by our intuition of marks of
symbols. Actually the reduction makes the situation more complicated,
for there must be more than 12 symbols involved.

Now consider the third question, that is, whether or not the re­
duction shows that arithmetic truths are true by virtue of their logical
forms and therefore have nothing to do with the special features of
human sensibility. We seem to have an apparently paradoxical situa­
tion. On the one side, the reduction of arithmetic truths to logic indeed
seems to show that arithmetic truths are true by virtue of their logi­
cal forms alone, while on the other side we conclude that intuition is
indispensable in knowing arithmetic truths. The resolution lies in rec­
ognizing that the so-called ‘true by virtue of logical forms’ is a fictional
notion, and it is constrained by human imagination, which is in turn
constrained by our capacity of senses.
First consider the notion of logical validity. From the point of view
of fictionalism, validity as it is ordinarily defined is a notion introduced
in a story, for we refer to arbitrary sets, functions and relations in the
definition. We just imagine that there is a domain of all sentences
and there are entities such as sets, functions and relations, and then

aaIt doesn’t matter if these are concrete marks on papers or the so-called ab­
stract symbol types, for I hold the fictional interpretation for abstract entities.
When we imagine a sequence of symbols as a special entity different from all con­
crete marks, our imagination is still based on our perceptual experience of concrete
marks of sequences of symbols.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 191

we define the notion of validity in the story about these sentences and
entities. The definition of validity refers to all possible interpretations
of a sentence. In the formal definition, that means all possible math­
ematical interpretations. It is commonly assumed that these should
have essentially covered all possible physical interpretations. Moreover,
considering the fact the physical universe could be finite, these may be
essentially more extensive than physical interpretations. However, no­
tice that they are conceivable by us human beings, and that in modem
physics we seem to believe that there are things not clearly conceivable
by us. Therefore, on the other side, the definition of validity in classical
logic may also be too restrictive. Sentences valid in classical logic may
become invalid when we consider quantum interpretations.
The classical notion of validity resulted from our pretending to have
a God’s eye view of reality. On the one hand, we ignore the possibility
that there may be things we cannot clearly imagine, such as things in
the quantum world. We pretend that we already have God’s eyes and
everything that could exist is in principle conceivable by us and can
be modeled in mathematics. On the other hand we ignore our human
limitation in knowing truths and making logical inferences. We ignore
the basis we need in proving the validity of a particular sentence. This is
an attempt at an epistemology for God. Sentences classified as logical
truths have the characteristic that they are true for all conceivable
interpretations. Therefore, it is concluded that the classification is
determined by the form of a sentence, not the content of a particular
interpretation of the sentence. Now God knows the truth or falsity of
all possible interpretations of sentences. The classification divides the
truths into two kinds (for God): those that are true by their forms

alone and those that are true by their particular contents.

of the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 192

Of course, we do not mention ‘pretending God’s eye view’ in our


ordinary explication of validity. Instead, we say that logic truths are
true by virtue of their logical forms alone. But the fact is still that
the picture is devoid of the special faeature of the human perspective.
Some sentences are such that all sentences of the same forms are true,
and therefore they are true by virtue of their forms alone, and some
others are true by virtue of their specific contents. The former are
logical and analytical truths and the latter must be empirical. Then, if
arithmetic truths belong to the former, they are true by virtue of the
concepts, and therefore have nothing to do with human sensibility.
This notion of validity may have some theoretical interest. By pre­
tending a God’s eye view, we can have a description of our beliefs by
classifying them according to some hypothetical standards. This may
help us to get a clearer view of the structure of our beliefs. However,
we must realize that the picture is something we conceive and we are
constrained by our sensibility. The so-called ‘true by virtue of logical
forms’ should actually be ‘true by virtue of logical concepts that fit
all our possible intuitions’. If something is true by virtue of its logi­
cal form, it only means that statements of the same form about any
possible human intuitions are all true, and then the explanation is just
that they are true by virtue of the human capacity of intuition. On the
other hand, if there were other intelligent beings with sensibility differ­
ent from ours, they would have a different notion of ‘true by virtue of
logical forms’, and our arithmetic might not be true by virtue of logical
forms for them. Therefore, granting something to be true by virtue of
logical forms does not exclude the relevance of human sensibility, for
our logical truths are indeed true by virtue of our sensibility.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 193

In conclusion, the reduction of numerical identities to first order


logical truths does not help in clarifying the epistemological basis of
applications of arithmetic, though it may have some other theoretical
merits. It clearly does not invalidate Kant’s view that arithmetic truths
are true by virtue of our sensibility, although we may have to abandon
the analytic/synthetic distinction and treat logic in the same way as
arithmetic.
It is constantly claimed that Kant’s philosophy of arithmetic suf­
fered seriously from the limitations of the logic of his time. The discus­
sion above shows that if the objective is to examine what is the basis
of arithmetic truths fo r us, that is, whether it is some particular expe­
rience of ours or the form of our senses, this limitation is perhaps not
a serious defect. This should not be surprising. Arithmetic knowledge
is the simplest form of knowledge we have. A reduction of arithmetic
truths to logical truths by some sophisticated logical constructions can
only complicate the issue.
One standard strategy widely used by philosophers in studying phi­
losophy of mathematics is extending logic. The general idea in the dis­
cussion above also implies that this kind of maneuver does not really
help in clarifying the epistemological basis of mathematical applica­
tions. What we must do is perhaps the reverse. Instead of extending
logic in order to claim more mathematical truths as logical truths and

to endow them with some special status, we must grant that even log­
ical truths are constrained by our sensibility. I will not go into details
about this point, although the choosing of finitism as the logical ba­
sis in the technical work in this dissertation is clearly related to that
concern.

o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 194

3 .4 . M ore R em ark s on K a n t and M ill. So far I have argued


that (1) empiricism seems to be forced upon us when we consider arith­
metic applications to the quantum world, but a Kantian account is
more attractive when we consider applications to ordinary physical ob­
jects directly perceived by us; (2) reducing arithmetic to logic does
not help in clarifying the epistemological basis of arithmetic. If these
observations are correct, then we are forced to look for a compromise
between the empirical and Kantian accounts of arithmetic.
I am unable to give a complete theory. I want to touch on this point
here, because I want to indicate where the studies in this dissertation
may lead us. This section contains only some preliminary remarks.
Clearly, many things Kant said have to be revised for the view to be
tenable. Moreover, there are some commonly held serious objections
to the Kantian views, widely believed to have made the views obsolete.
The attack on the a priori /a posteriori and the analytic/synthetic
distinctions are among these objections. The criticism on Kant’s views
on the a priori status of geometry is also among them. Here, I want
to give some remarks on what must obviously be revised in the basic
Kantian notions, and I want to indicate that the objections are not
really decisive.
First of all, it is clear that one cannot expect a clear-cut distinction
between what is a priori and what is a posteriori. We expect that the
applicability of arithmetic to ordinary physical objects directly percep­
tible or conceivable by us is a priori, in the sense that the form of our
senses determines that those things must conform to arithmetic. Now
there is not a clear-cut borderline between physical objects perceptible
or conceivable by us and strange things in the quantum world. So we

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 195

have to abandon any commitment to a clear-cut borderline between


what is a priori and what is a posteriori.
This is perhaps not a disadvantage. First notice that the distinc­
tion between conceivable physical objects and things in the quantum
world is not the Kantian distinction between appearances and things-
in-themselves. Things we deal with in quantum mechanics should be
among appearances, although they are too small and cannot appear
to our unaided senses directly. We extend our senses first by instru­
ments, for instance, microscopes, and then by theoretical models. With
microscopes we can see many things we originally did not know. For
even smaller things, we try to build models to describe them. When
we claim that water is actually composed of tiny particles although it
appears to be a kind of continuous stuff, we present a model of a huge
number of particles and then claim that if a tiny part of a cup of water
is magnified about a hundred million times23, it will be approximately
like the model. In this way, unobservable phenomena can ‘appear’ to
our theory. So we need a division of phenomena. Some phenomena are
directly (or indirectly) perceptible, some others can only ‘appear’ to us
in our scientific theories. Things in the quantum world are among the
things that ‘appear’ only to our theories.
Then we can say that Kant was not fundamentally wrong. He just
did not realize that appearances could be too small to be perceptible
directly and could only ‘appear’ to our theories. More importantly, he
did not realize that things that ‘appear’ to our theory could ‘appear’
in a way radically different from the way by which things appear to
our unaided senses. In other words, our familiar logical concepts may
become inapplicable to them.

33The radius of an atom is about 10~*cm.

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 196

Our form of sensibility, if there is such a thing, only determines


the way we perceive perceptible things. It does not cover things that
are not perceptible by us and that can only ‘appear1 to us in theories.
Perceptible things may necessarily obey logical and arithmetic laws.
In other words, some form s of truths about perceptible things may be
true by virtue of our sensibility. It is in this sense that these truths
are a priori (for us), for they cannot be otherwise, as long as we do
perceive something. But this certainly does not (and does not have to)
imply that the same form s are true of things that can only ‘appear’ to
us in theories. When things become more and more remote from us,
they may differ more and more from familiar things. When we try to
use models to describe things that are far beyond our perception and
that can ‘appear’ to us only in theories, we may realize that arithmetic
and classical logical laws are not valid for them.
If that is the only point that Kant missed, it is perfectly under­
standable, given the status of science of his time. Kant just made

some too hasty generalizations which are not really necessary for the
fundamental point of his philosophy.
Of course we must also abandon any commitment to a clear-cut dis­
tinction between what is fully determined by our sensibility and what
is not. Appearances are vague and very complicated. This should be
clear if we reflect on our real perceptual experiences. The boundaries
of the macroscopic perceptible things are fuzzy. Any physical object
we perceive is also fuzzy to some extent. Similarly, we can expect that
sensibility, if there is such a thing, must also be vague and compli­
cated. Abandoning clear-cut distinctions should not be considered as
a disadvantage in our philosophical account of appearances and our

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 197

knowledge of them. It only means that we abandon some wishful ex­


pectations for some clean and simple generalizations and that we treat
our real perceptual experiences more faithfully.
There are similar remarks to be made about geometry. It is a
common view today that Kant’s theory of geometry is out-of-date.
That means geometry is obviously empirical. Sometimes it is held that
Kant had better reasons to believe that arithmetic is a priori and it is
only suspicious if it is synthetic, but Kant’s views on the a priori status
of geometry is fundamentally wrong. Another constantly cited point is
that Kant did not make the distinction between a formal, mathematical
geometry and a physical geometry. While the former may be a priori
in some sense, the latter must be empirical.
Since epistemology is primarily epistemology of our knowledge about
real things, the primary concern is certainly physical geometry, that is,
our knowledge of space as it appears to us. As we agree that there
is no clear-cut distinction between the a priori and the a posteriori,
the proper question we should ask is ‘Is there a basic core of our ge­
ometrical knowledge of space that is a priori, namely, determined by
our sensibility?’ The answer must very hard to reach, but it is not
obviously ‘No’. Presumably, it is not obvious that every piece of our
geometrical knowledge can be subject to empirical test. As early as
infants stretch out their hands to grasp things, they already have some
implicit geometric knowledge of space. When we try to interpret a
four dimensional manifold as a model of the universe, a lot of primi­
tive geometrical knowledge is presumed. We can test if the model fits
astronomical observations, but we do not seem to be testing some very

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 198

primitive knowledge in doing this, for instance, the primitive knowl­


edge we implicitly have in order to set up our telescopes and direct
them to some appropriate directions.
Kant certainly made a mistake in claiming that Euclidean geometry
is among the core that is a priori. However, perhaps he (and geome­
tricians before his time) just made some too hasty generalizations in
claiming this. We do know from our perceptual experience that as far
as the part of space within the scope our perception is concerned, it is
approximately Euclidean. However, we cannot directly perceive things
very far from us or things very small. So our form of intuition can­
not really determine a strictly Euclidean space, with continuous and
infinite lines. We must distinguish what may be determined by our
form of intuition and what we imagine by consciously exercising our
imagination ability. The latter can be supposedly beyond the former.
Given the vagueness of appearances and the limitedness of our percep­
tion, our concepts of points, lines and so on must also be vague. When
we conceive of points without extension, lines without breadth and
infinitely long straight lines, we are clearly doing the latter, namely
imagining by conscious by exercising our imagination ability. We are
using thought and imagination, rather than intuition, to characterize
things that could not be immediately presented to us as objects of in­
tuition. We should not take such idealized points and lines as objects
of the possible a priori core of geometry.
Of course I do not have a definite answer as to whether there is
such a core. However, there are some very primitive assertions which
do seem to have necessity for us as in the case of arithmetic. To look
for examples we can consider the axioms that are common to all kinds
of geometries proposed by mathematicians. Here may be a concrete

o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 199

one. Recall that in axiomatic geometries, a primitive concept is ‘be­


tween’. Suppose that we have concepts of points and straight lines,
notwithstanding their vagueness. Consider a simple axiom such as ‘If
point b is between points a and c and point c is between points b and d
on a straight line, then point b is between points a and d.' Suppose this
is about points and lines we can directly perceive. It is not obvious in
what way we could perceive otherwise. This geometric assertion may
be necessary relative to our faculty of space perception.
Although no positive conclusion is reached yet, it seems we can still
claim that Kant’s view of geometry, while certainly wrong in technical
details, is not fundamentally wrong. In particular, while empiricism in
geometry is clearly correct when we mean the geometry of large scale or
microscopic scale space, it misses something epistemologically impor­
tant. That is, it is still a possible and attractive view that some very
primitive geometrical truths are determined by our form of intuition.
And we have to consider this if we want a faithful account for our
geometrical knowledge. Moreover, there is no contradiction between
empiricism and apriorism in geometry if we abandon the demand for a
clear-cut distinction between the two24.
It is also conceivable that an absolute distinction between the syn­
thetic and the analytic should not be expected. Kant’s simple charac­
terization of the analytic (for instance, ‘the concept that is the pred­
icate is contained in the subject concept’, or ‘knowable by analyzing
concepts alone’) is certainly based on the assumption that the concepts
involved are well defined. Now we know that even the familiar logical

a4There are some other attempts at saving the a priori status of Euclidean
geometry, for example, the attempts by referring to visual space or the attempts
by referring to conventionalism in geometry as in Barker [7]. I am not convinced
by those but I will not discuss them here.

of the copyright ow ner. Further reproduction prohibited w ithout p erm ission .


3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 200

concepts may loose their sense for strange things in the quantum world
and it becomes an empirical matter what are the appropriate logical
concepts there. Moreover, we know that other non-logical knowledge
may be assumed in analyzing the concepts. These are the two points
that force us the reconsider the notion of analyticity.
I do not have a definite opinion about whether or not the ana­
lytic/synthetic distinction is still of some interest, but I want to suggest
that this does not annihilate the Kantian question of the possibility of
synthetic a priori knowledge. Two points above only push more truths
to the side of synthetic truths in Kant’s sense. This is contrary to
what is expected by logicists or empirical positivists. On the other
hand, as was explained in the previous sections, arithmetic can still be
considered as a priori if we restrict to applications to ordinary physical
objects. Therefore the Kantian question is still interesting.25
The conclusion of this section is as follows. Fictionalism and the
reduction of applications of classical mathematics to applications of
strict constructivism will reduce the epistemology of mathematics to
that of applications of finite mathematics, which is close to elementary
arithmetic. Therefore, philosophy of arithmetic is of central impor­
tance for philosophy of mathematics. Empiricism in arithmetic forces
itself upon us for some applications of arithmetic and Kant’s view is
more attractive for some other applications, while logicism is off the
point for the pursuing of the epistemological basis of arithmetic. So
a proper philosophy of arithmetic is probably a mixture of empiricism
2SSome conclusions in this and the preceeding subsection are somewhat similar
to Quine’s rejection of a priori/a posteriori and analytic/synthetic distinctions, but
the overall picture is clearly different. There is no assumption of holism and the
centrality of pure arithmetic (as truths about abstract entities) in the web of beliefs,
while there is a distinction between things we consciously imagine in our heads,
namely mathematical entities, and things given to us in experience. But I cannot
discuss the details here.

of the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


3. REMARKS ON THE EPISTEMOLOGY OF ARITHMETIC 201

and Kant’s position. Finally, a mixture is possible, if we discard the


demand for a clear-cut distinction between the a ■priori and the a pos­
teriori, and grant that things can ‘appear’ only to our scientific theories
and thus out of the determination of our sensibility, and that in that
case they can ‘appear’ in a way different from the way they appear
to our senses directly, and then, as a consequence, while arithmetic
truths about perceptible things are a priori for us, the applicability of
arithmetic to other appearances becomes empirical.26

J6Some recent literature on Kant’s philosophy of mathematics focuses on Kant’s


notion of intuition and its role in mathematical reasoning. See Parsons’ and Thomp­
son’s papers in the collection [75]. I am unable to give comments on these debates,
but it suffices to mention that these are not supposed to be serious objections to
Kant’s fundamental points. There is indeed an open question raised by Parsons
(see [74] pp. 75): why an arithmetic truth known by intuition of a singular case
(say, counting fingers) ‘can be applied to all experience a priori and with certainty’.
Kitcher’s paper ‘Kant and the foundations of mathematics’, also in [75], objects to
the possibility of an explanation, which is basically an objection to the posibility of
a priori intuition, for instance, an intuition of a triangle that involves 'only the act
whereby we construct the concept, and abstract from the many determinations (for
instance, the magnitude of the sides and of the angles), which are quite indiffer­
ent, as not altering the concept ‘triangle’.’ (Kant, Critiqe of Pure Reason, B 742)
But the Kantian transcendental strategy is presumably not to prove the a priori
certainty of arithmetic truths. It is to provide an explanation, or to argue, for the
case of a priori intuition of a triangle here, as in the transcendental deduction of
categories and principles, that if any knowledge is after all possible, such a priori
intuition should be possible. Of course I do not know if this is workable, but the
objection is perhaps not conclusive. (For the same reason, Young’s [106] attempt
at answering the question does not seem to be complete. The objective there does
seem to be proving the a priori certainty, and therefore the circularity problem
raised by Kitcher is still lurking there.)

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission .
i
i
i
i
i

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
STRICT CONSTRUCTIVISM AND THE PHILOSOPHY
OF MATHEMATICS
Volume II

FENG YE

A DISSERTATION

PRESENTED TO THE FACULTY

O F PRINCETON UNIVERSITY

IN CANDIDACY FOR THE DEGREE

OF DOCTOR O F PHILOSOPHY

RECOMMENDED FO R ACCEPTANCE

BY THE DEPARTMENT OF

PHILOSOPHY

Ja n u a ry 2000

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX A

Constructive Analysis in SC

1. Introduction

In this appendix we continue to show that the essentials of Bishop’s


constructive analysis are fonnalizable in SC . We will examine Chapter
4 through Chapter 7 (except Section 7 of Chapter 5) of Bishop and
Bridges’ book Constructive Analysis. Chapter 2 and Chapter 3 of the
book have already been treated in Chapter 1 of this dissertation. Each
section below will deal with one of the other chapters.
The following are the parts of the book Constructive Analysis that
we have not examined and the exceptions we have found. We have to
leave open Chapter 5 Section 7 of the book tentatively. This section
is on the Riemann mapping theorem in complex analysis. It uses the
Jordan curve theorem whose proof is complicated and not included
in the book. Theorem (2.12) in Chapter 7 of the book is left open
as well. This is a theorem in constructive approximation theory. We
can not formalize the proof given in the book and we do not know if
there are other proofs that are formalizable within our more limited
system. We are not clear if this means a serious omission. It seems
that this is better to be examined in the more general background of
constructive approximation theory, and we noticed a remark by Bridges
([18] p. 43) that this theorem is generally not needed in constructing
best approximations in realistic cases. The examination of constructive

202

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 203

approximation theory has to be left for the future. The section and the
theorem omitted are never mentioned elsewhere in the book.
Besides, there are two other theorems that can not be fully formal­
ized in our finitistic system. The first is the Baire category theorem.
We have to assume that the relevant metric space is separable. This
may not be a serious compromise, for the basic metric spaces considered
in constructive analysis are all separable (ignoring the non-located sub­
spaces). The second is Theorem (6.7) in Chapter 6 of the book. It says
that any integrable set with a positive measure can be approximated
from within by the so-called strongly integrable sets. We can only for­
malize two weaker versions: One says that any integrable set with a
positive measure can be approximated by strongly integrable sets, not
necessarily from within, but in the sense that the symmetric differ­
ence between them can have arbitrarily small measure. This is close
to Lemma (6.5) in the same chapter. The other weaker version says
that any integrable set with a positive measure can be approximated
from within by compact integrable sets. (Note: Strong integrability
implies compactness.) The original theorem is never used elsewhere in
the book. We believe that this weaker version is probably sufficient for
applications.
All other sections and theorems in Chapter 2 through Chapter 7 of
the book Constructive Analysis will be carefully examined and shown
to be formalizable in our finitistic system. On the other hand, we
have not examined the last two chapters of Bishop and Bridges’ book,
which are on locally compact Abelian groups and commutative Banach
algebras, and furthermore there are many other aspects of constructive
mathematics not covered in the book. It is still not very clear if all

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. METRIC SPACES 204

the essentials of constructive mathematics can be formalized in this


finitistic system.
We will keep the notations of Chapter 1 of this dissertation.

2. M etric Sp aces

This section deals with Chapter 4 of B&B.

2.1. Basic Definitions and Associated Structures. Suppose


that X is a set, fp is a metric on X ’, or ‘(X, p) is a metric space’,
is translated as : ‘p : X x X —►R 0* and for all x , y , z 6 X , (i)
p ( x , y ) = 0 <-►x = y, and (ii) p { x , y ) = p( y , x) , and (iii) p ( x , y ) <
p ( x , z ) + p ( z , y ) \ Recall that according to our convention X can also
be a multiple set and/or a parameterized set. If p is a metric on X, and
Y is a subset of X , p is also a metric on Y, called the induced metric
on Y. The following are easily formalized: ‘(X, p) is bounded’, ‘c is
a bound of ( X, p ) \ lY is a bounded subset of X ’, ‘the sequence (xn)
of elements of X converges to an element y of X \ 1f is a uniformly
continuous function from the metric space (X , p) to the metric space
(X ',/>')’, ‘the sequence (/„ ) of functions from the set S to the metric
space (X , p) converges uniformly to a function / : 5 —» X ’, ‘the subset
A of X is located in X ’, ‘(X , p) and (X ', p') are metrically equivalent’
and so on; cf. B&B p. 82-89.
The open sphere S (x ,r ) is a (parameterized) subset of X:

(y € S ( x , r ) ) = p ( y , x ) < r,

and the closed sphere S c ( x , r ) is defined similarly. Then notions like


open subsets, closed subsets, dense subsets, one subsets’ being an­
other’s interior or closure and so on are naturally defined by referring
to these open spheres as neighborhoods.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 205

The witnesses for ‘A is a located subset of X ’ include an operation


L such that

Vx 6 X ( L (x) = inf { p ( x , y ) : y € A } ) .

We sometimes use the notation p (x, A) for L (x), with the understand­
ing that it is a witness for A’s being located, implied by the context.
Notice that, for instance, p( x, A) > 0 is equivalent to

3r € R (r > 0 A (r = inf { p (x, y ) : y 6 A }))

and hence does not actually depend on the witness. p (x ,A ) = r,


p (x, A) < r and so on are similar. For the occurrences of p (x , A) in
other contexts we need some caution. The metric complement X — A
of a subset A is a subset of X defined by

( x € X —A ) = x € X A

3r 6 R (r > 0 A (r = inf { p (x, y ) : y € A }))

and hence if p (., A) is the witness for the locatedness of A, then

(x € X - A) is equivalent to x 6 X A p (x, A) > 0.

cf. p. 85— 88 of B&B.

Propositions and remarks on p. 83-84 are simple. For instance, if


( X i t pi), ..., (X n, pn) are metric spaces, let
n
p = Ax y . J 2 P i ( * » V i ) -
i=i

Then p is a metric on rn=i^«- Similarly, suppose that ((X m /J n ))^


is a sequence of metric spaces, which means X n is a parameterized set
and (pn) is a sequence such that for all n > 0, pn is a metric on X n,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 206

and suppose that each (X n, pn) is bounded by 1, and let


OO

P = A x y . ^ 2 ' np„(x„,yn) .
n=l

Then p is a metric on X n. Propositions (1.8) and (1.9) on p. 85


are also simple; so is the assertion that any metric is equivalent to a
bounded metric.

The Cauchy-Schwarz inequality and Minkowski inequality for real


numbers and integrals axe easily provable, with no novel inductions
needed. Hence we can show that d defined by

is a metric on Rn, and d defined by

is a metric on C ([a, 6], R).

2.2. Completeness. Given a metric space ( X , p), ‘(xn) is a Cauchy


sequence of X ’ and ' X is complete’ are easy to formalize. Given (X, p),
the completion of it, ( X, p J , can be constructed: X is the set of se­
quences (xn) such that

Vn, m > 0 (xn 6 X A p (xn, xm) < rT 1 + m -1) ,

and the equality on X is defined by

(® = x v ) = Vn > 0 (p ( * n , yn) < 2 n 'x) .

For x , y 6 X it can easily be verified that (p (x n,y n)) is a Cauchy


sequence of real numbers:

Vfc > OVn, m > 4k (|p (x n,y„) - p (x m,ym)| < 1/fc).

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 207

So we can construct the term limn_ O0 p ( x n, yn) (with N = XkAk as


the witness for ( p( xn, yn)y* being a Cauchy sequence) and define

P = >&y- Hm p ( x n, yn).

Then we can prove that p is a metric on X and (X , pj is complete.


Further, the inclusion map i : X —* X can be defined and it is easy to
see that when X is complete, i is isomorphic.
It is also straightforward to prove that the product of a sequence

of complete metric spaces is still complete and a uniformly continuous


function from a dense subset of a metric space to a complete metric
space can be extended into a uniformly continuous function on the
whole space (Theorem (3.6) and Lemma (3.7) on page 91).
Lemma (3.8) on page 92 needs some attention for its proof involves
an inductive construction of a sequence. The lemma says, for A a
complete, non-void, located subset of X , and x a point of X , there
exists a point a £ A such that p ( x , a ) > 0 entails p ( x , A ) > 0. To
prove this, we construct a Cauchy sequence (an) of elements of A and
take the limit as a. First, the assumption of locatedness implies that
we can use the notation p (., A) and

(12) V u £ X ( p (u, A) = inf { p (u, y ) : y 6 A } ) ,

which in turn implies that for n > 0, if p ( x , A ) < 1/n, there exists
y 6 A such that p (z, y) < 1/n. Then it follows that there exists 6 such
that

(13)
Vm, n > 0 (p (x, A) (m ) < 1 /n — l / m —> S (m, n) € A A p (x, 6 (m, n)) < 1 / n ) .

Also, the non-voidness of A implies that for some xo,

(14) xq 6 A.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 208

Now to prove the conclusion it suffices to show the existence of a under


the original assumptions plus (12), (13), and (14).
Let k [n] be the representing term for the quantifier free formula
p(x, A) (4 (n + 1)) < 3/4 (n + 1). Then k [n] = 0 implies6 (4 (n + 1) ,n + 1) 6
A and p(x, £ ( 4 ( n + 1) , n + 1)) < l / ( n + l). We can construct a
term h [n] by primitive recursion such that when k [0] = 1, h [n] = 0
for all n, and when k [0] = 0, h [n] is the largest m < n such that
k [0] *= ... = k [m] = 0. Finally let

On = J (fc [0], 6 (4 (h [n] + 1), h [n] + 1 ), x0) .

Then its straightforward to verify that (an) is a Cauchy sequence and


the limit is the required a. cf. the proof on p. 92 of RfeB.
Now consider the Baire category theorem on p. 93. Unfortunately,
we cannot formalize the full theorem in SC , for the construction of the
sequence (xn) given there is not generally available in SC . However,
when the metric space is separable, we can get over the difficulties.
A metric space (X , p) is separable, if there exists a sequence (c„) of
elements of X such that its range is dense in X , that is, for any x 6 X
and m > 0, there exists n such that p(x, Cn) < 1/m . Then we have,

THEOREM 2.1. If (Un) is a sequence of dense open subsets of a


separable complete metric space { X, p) , then U = n^_017n is also dense
inX.

PROOF. Suppose that (c*) is a sequence that witnesses the sepa­


rability of (X , p) as mentioned above. It suffices to prove that for any

l0 > 0 and &o, there exists x 6 U n S (c^ , /o). The assumptions imply
that for any number n, k and any I > 0, there exist k', I' such that

S c (c k,} 1 / 0 C Un n S ( c fe, 1//) A I' > n.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 209

So, by the axiom of choice, we can prove the conclusion under the extra
assumption

Vn,kVl > 0 ( 5 c (<:*(„,*,/), 1 /L ( n ,k,I)) C U n C[ S(cfc, l / l ) A L ( n ,k,I) > n ) ,

with K , L as new free variables. Since i f (n, A, /) and L (n, fc, /) are
numerical terms, we can construct sequences (k„) and (/„) by primitive
recursion such that kn+j = K (n, kn, l n) and /„+i = L( n, kn, l n). Then
it remains to prove that (c ^ ) is a Cauchy sequence and its limit is the
required x. For this we need an induction on m to show that

m > n —> Sc^Ckn, 1 /lm) Q S c ^ , 1/ Q ■

Since S c ( c ^ , l / / m) C S c ^ , 1//„ ) is equivalent to p ^ , CfcJ +


1/lm < 1/ U, which is a 11° formula, the induction is available in SC . |

In constructive analysis, the concrete metric spaces we treat are all


separable, so we expect that this theorem is sufficient for applications.
This is the first compromise we made in the formalization.

2.3. T otal B o u n d ed n ess and C o m p a c tn ess. Recall that we


can code a finite sequence of elements of X into a single object of the
type of X , and given the extensionality condition in the definition of
sets, this coding technique does function properly in our contexts. So,
‘(X, p) is totally bounded’ can be formalized as

VI > 03x (Seq (x ) A Vy € X 3 i < (x)0 (p ( ( x ^ , y) < l / l ) ) .

Then it is trivial that total boundedness implies separability. (Notice


that this assertion strictly relies on the extensionality condition, for
to obtain the infinite sequence that witnesses the separability we must
concatenate all the finite sequences which witness the total bound­
edness. This is done in Section 4; cf. the construction of x there.)

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. METRIC SPACES 210

The basic propositions about total boundedness, Proposition (4.2) to


Proposition (4.6) on p. 94—95 are trivial. Also, when X is totally
bounded, the diameter

diam (X ) = sup {p (x, y ) : x ,y e X } ,

of X exists, though notice that the notation diam (X ) as a term con­


tains a parameter that is supposed to be a witness for X ’s being totally
bounded. Similarly, the notation p (x, A) will also be used for a totally
bounded subset A of a metric space X .
Compactness is easily defined. We consider the proof of Theorem
(4.8) on p. 96-98. The theorem says, given a compact space X and a
positive number e, there exist finitely many compact subsets X \ , ..., X n
of X , each with diameter at most e , whose union is X . In the proof, X{
are constructed as parameterized subsets of X . Supposethat I > 2/e.
From the assumption we have a function 8 such that for t > 0, 8 (i) is
a finite sequence of elements of X that constitutes a 3- *-2 /-1 approx­
imation to X . Let £(0) = (xx,...,x„). In the informal proof (cf. p.
96-98, B&B), one constructs a sequence of finite sets X° , X} , . . . , for
each j = 1, ...,n, such that X° = {xy} and X j +1 is obtained as fol­
lows: beginning with AT’+1 = the empty set, for each y in the sequence
£(z + 1), decide if p (y,X jJ < 3- '-1 /-1 or p (j/i-Xy) > and
add y to X j +l in the former case. Then X j is the closure of the union

u 2 o * }-
Since elements of X j are selected from the sequence 6 (i), to for­
malize the construction in SC, we replace X jX by the indices of its
members in the sequence 6 (i). So instead of constructing sequences of
finite subsets of X , we construct sequences of finite sequences of natural
numbers, Zj, i = 0,1,..., such that Z° = (j), and Zj +1 is the sequence

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 211

° f k, 1 < k < ( 8 (i -f 1))0, such that for some h > 1 and h < we
have

P ( ( * « )( * } ) , ■ (<(*+ !))») (8 '3i+1l) < 3 — Ir I - 8 - ' 3 - i- ‘i - 1.

So Zy is the sequence of the indices of the elements of Xy in the sequence


8 (i). Let Y j be the set such that

y € Yj = 3* < (Z5)o ( k > 1 A y = ( f (i))w ) J .

Let Fy = and let Xy be the closure of Yj. By an induction on


m ' we can show that

m' > m V, € Yj"' (p (», >7") < E ).


\ h=m /

This formula is equivalent to a £° formula given the assumptions, be­


cause we can replace y by ( 8 35 shown by the definition of
Yj*' and then the universal quantification Vy 6 Yj71' becomes a bounded
numerical quantification V/b < (z™ '^ and p (y, Vy") < UhLm 3-/l-1/-1
is £ °. So the induction rule is available in SC . Then it follows that
Ujlpfry is a 3~ml~l approximation to Yj and Xy is totally bounded.
Suppose that p(y,xy) < 3 By another induction on m we can
show that p (y , Yj*') < 3-m-a/_1. Notice that this formula is

3k < { Z f) a (* > 1 A , ( ,, (f(m )){ J r)i) <

which is £ °. Then it follows that y € Xy and X = Uy=1Xy.


Theorem (4.9) and Proposition (4.11) on p. 98-100 are derived
from Theorem (4.8) with no more inductive constructions.

2 .4 . S p a ces o f F u n ction s. For a compact space X and any met­


ric space Y , the set C (X , Y ) of (uniformly) continuous functions from

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. METRIC SPACES 212
X to Y , the norm ||/|| for / 6 C ( X fY) , and the notion of equi-
continuity are easily defined; cf. B&B p. 100. Notice that ||/|| as a
term actually contains the modulus of continuity for / . Theorem (5.2)
to Lemma (5.5) on p. 101— 102 axe trivial. We examine Theorem (5.6)
whose proof involves some inductive constructions.
Here, given a sequence (x,) of elements of a compact space X , a
real number K > 0, and a finite sequence (ux, ...,it„) of real numbers
such'that

(15) |«i| < K, |u< - Uj\ < S ( p (xi} x j ) )

for i , j = l , . . . , n , where 8 : R +0 —►R +0 is strictly increasing, £(0) =


0, and £(u + u) < 8 (u) + 8 (v) for all u , v G R +0, we must extend
(it\ ,..., un) into an infinite sequence (ui) such that (15) holds for all
1, j = 1 , 2 , ...... For this, we first construct a term q [i,m] such that

q [t, m] = U{ ( m ) , for i = 1,..., n,

q[i + l,m] = m a x ( - K ( m ) , m i n ( K ( m ), min{g[A,2m ] + 6 ,(p (xfc,x ,+1))(2m) |A

for i > n. Clearly, this can be reduced to primitive recursion. Let


Vi = Am. q [i, to]. By an induction on k we can show that

Vi < k (u,- € R A |vj| < i ) ,

noticing that this is a 11° formula. Similarly, an induction on k gives

V*,y < f c d u i - u j l < ^ ( ^ ( x i . x , ) ) ) .

cf. the proofs of Lemma(5.3) (5.4) on pplOl— 102. So, (15) holds for
(Vi). With this construction, the rest of the proof can be formalized in
SC straightforwardly.
A polynomial p of degree m in n variables can be represented by
a sequence of (m + l ) n + 2 real numbers, where the first two numbers

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 213

are m and n and the others are the coefficients. For a sequence x of n
real numbers, the value of p at x can be represented by a term p( x) .
Similarly, for a sequence h = ( hi , ..., h*) of n functions in C ( X ) , the
composition p o h can be constructed and is also a function in C ( X) .
Then, for G a subset of C (X), the algebra A ( G ) of functions generated
from G is the subset of C ( X ) such that / £ A ( G ) if and only if for
some polynomial p and some sequence h of functions from G, f = poh.
Lemma (5.9), (5.11), and (5.12) on p. 104— 106 are straightforward.
The notion of separating subset (of C (X ) ) is also easily defined. Then
we examine the proof of Stone-Weierstrass theorem on p. 106— 108.
The point that needs some special attention is the following claim:
(*) if H is the closure of A ( G ) in C ( X) and for all i < n, ft 6 H,

then max { /,|t < n } £ H, £ ,= 0 /,• £ H, and n?=o f* £ H -


We check the last conclusion here. The others are similar. Since
for each i < n there exists M such that ||/»|| < M, it follows that there
exists M > 1 such that for all i < n, ||/j|| < M. Since./< £ H for all
i < n , there exists g such that for all i < n and all k > 0, g (i )k £ A (G)
and for all i < n, ||</ (z)fc —/.|| —> 0 as —►oo. Then we can show that

||nr=o5 (i)fc — niLo AH -* 0 as k -* oo. For this, we need the following


assertion: if M > 1, and for all i < n a,- £ R, 6, £ R, |a,| < M, |6t| <

M, and |at - < r, then |n?=0 Oi - FIt=o ^1 < (2Af)n r. This can be
proved by an induction on k for a 11° formula

k <n < ( 2 M ) k r.
i=0 i=0

Now it remains to prove that ni=o 9 (0* S A ( G ) . This follows from


the observation that the product of a finite sequence of polynomials (in
different variables) is still a polynomial, which can be proved directly.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. METRIC SPACES 214
So we have the claim (*). Then, the rest of the proof on p. 106— 108
can be formalized easily. Corollary (5.15) to (5.17) on p. 108-109 axe
also straightforward. No more inductive constructions are needed.

2.5. L ocally C o m p a c t Sp aces. The definition of local compact­


ness needs some elaboration, for it quantifies over ail bounded sub­
sets and all compact subsets of the metric space. Quantifications over
bounded subsets of a metric space X can usually be replaced quantifi­
cations over parameterized subsets {x 6 X : p (x, y) < r}, where y 6 X
and r > 0 are the parameters. Now a compact subset Y of X is deter­
mined by a sequence 8 (i) such that for i > 0, 8 (i) is a finite sequence
of elements of Y that is a 1 /i approximation to Y, for in that case Y
is the closure of

(16) {x 6 X : 3i > 03k < (S( i ) ) 0 (k > 1 A x = (8 (i))fc) } .

We introduce a term a generator of a compact subset: ‘8 is a generator


of compact subset of X ’ is translated as ‘for all i > 0 8 (i ) is a finite

sequence of elements of X , and for all i, j > 0 and any member x of


8 (j ), there exists a member y of 8 (i) such that p (x, y) < 1 / i ’. For Y a
compact subset, the 8 such that 8 (i) gives a 1 /i approximation to Y is
clearly a generator of a compact subset, and reversely, for 8 a generator
of compact subset, the closure of (16) is a compact subset, the compact
subset generated by 6 . Then, quantifications over compact subsets can

be replaced by quantifications over generators of compact subsets. So,


lX is locally compact’ is translated as ‘for any x 6 X and any r > 0,
there exists a generator of compact subset 8 such that, for all y € X ,
if p [ y , x ) < r then y is in the compact subset generated by 8 \ For X
a locally compact metric space, / : X —* Y is continuous when for any
x € X and r > 0, / is uniformly continuous on { y 6 X : p (y , x) < r}.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 215

It is trivial to show that locally compact spaces are complete and


separable. Proposition (6.2) to (6.4) on p. 110-111 are also straightfor­
ward. The notion of homeomorphism and one-point compactification
are defined naturally, where we quantify over the generators of compact
subsets in the latter. The proofs of Proposition (6.7), Theorem ( 6 . 8 ),
and Proposition (6.9) on p. 113-116, though quite long, do not involve
novel inductive constructions and can be formalized in SC straight­
forwardly. The notion of vanishing at infinity and the set (X ) of
continuous functions vanishing at infinity are also easily defined and
Lemma (6.11), Corollary (6.12), and the assertion that C * (X ) is com­
plete and separable for locally compact A- are all straightforward. The
definition of test functions and Proposition (6.14) and (6.15) are also
straightforward.
Now we consider the proof of Theorem (6.16), Tietze extension the­
orem, on p. 120-121. The original proof uses an inductive construction
to construct a sequence of functions from X to R. We will eliminate
the inductive construction and replace it by a uniform construction
of a function fin> with n as a parameter, such that limn_ Q0 fin is the
desired extension. So, suppose that Y is a locally compact subset of
X such that there are two points x°, x 1 6 X — Y and p ( x ° i x l ) > 0.
Let f : Y —» [0 , 1] be continuous Then we will show that there exists
a function fi : X —►[0, 1] such that fi is uniformly continuous on any
bounded subset of X and fi (y) = / (y) for y 6 Y.
Let Z = Y U {x°, x 1}. Then Z is also locally compact. There
exists a function g : Z —» [0 , 1] such that g (y) = f ( y ) for y € Y,
g( x°) = 0, g (x 1) = 1, and g is also continuous. Since Z is locally
compact, by Theorem (4.9) on p. 98, there exists a sequence (/3k)
of positive real numbers such that /3* —> oc as fi —►cx> and Dk =

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. METRIC SPACES 216

{ z £ Z : p (z , x°) < Pk} are all compact. Then, by Theorem (4.9) again
(notice that countable unions of countable sets are still countable),

are all compact for i = 0,..., 2n — 1 and k, n = 0 , 1 , 2 ,... . So,

are locally compact. Since x° 6 Ai>n and x 1 6 Bi<n, they are instanti­
ated and hence p (x, i4j>n) + p (x, Bi>n) exists for any x € X . It is easy
to see that, for any bounded subset 5 of X and n, there exists 8 > 0
such that p (x, A i tn ) + p (x, B i <n) > 6 for all x € S and i = 0,..., 2n —1.
(This is because \g (z) —g (z')\ > for z 6 i4»> and z' 6 Bi%n and g is
uniformly continuous on any bounded subset of Z\ cf. p. 120) Let

Then h* is uniformly continuous on any bounded subset of X . (Here


we use the notations p(x, >!»>)> p ( z , Biltl) freely, but notice that as
terms of the language of SC they should contain other free variables as
parameters which are supposed to be the witnesses for At-,n and s
being locadly compact, and which have been proved to exist.)
We show that h n (z ) —» g { z ) for z 6 Z. For each n > 0, i =
0 ,..., 2n — 1 , let

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. METRIC SPACES 217

For each z, n, and i, g (z ) € Ai>n U Ci<n U Bi<n. It is easy to see that,


for each z and n, there exists k, 0 < k < 2 n — 1 such that g (z ) 6 Bi>n
for i = 0 , A; — 1, and g { z ) 6 A{<n for i = k + 1, 2n — 1. Then

|M * ) - f | - h ^ So> "+ ffW “ B -♦


oo.
It remains to show that lim n_>0o Ki (s ) exists uniformly, for then it
follows that the limit function is uniformly continuous on any bounded
‘subset. Notice that

^n+l(x )
2B—1
__ 1 y ' ^1 / ________rP( x , •A2i,n+l)
r*2t,n+i/_________ |______ p( x, Aji+i.n+i)
2n t=0 2 \p (x, i42,> +i) + p (x, 5 jj>n+x) p (x, Aai+li„+i) + p (x, B 2i+i,n+i),

Since

•^2 » ,n + l Q ^ i . n Q ■ ^ i + l . n + l Q ■ ^ 2 (» + l),n + l >

® 2 i,n + l 2 f t.n 3 ^ 2 t + l , n + l 5 -® 2 (» + l),n + l j

w e have p(*.Au,n+i)------ > _____ p(x.At[nJ____ > _______ p(x.Aa,^ttn±tJ----


p(x,Ajitn+l)+P(^<^7i,n+l) — p(x,Ai>n)+p(x,BjlFV) — p(Xi-Ajt+L,n+t )+p(l,Bjj+i, n+l)
> -7—-— p(x"Aa<*'+I>'B+0--------- Then it is easy to see that
p^r,A j(i+ l)lB+ i J+P^*."j(«+I).n+1)

l^ n + 1 (x) — /i n (x )|

2B—1 p ( x ,A 2itn+1)
^ y ___ _____________P_________________________
( * , A 3( i + i ) , n + l )

2n i=o \ P ( x ) ^•2»,n+l ) + P (®j # 2 i,n + l) p ^X, A jfj+ x ^ n + x ) + P ( x i ^ ( i + x j . n + l )

< —.
- 2n

So, h„, (x) converges uniformly for x 6 X .


This completes the proof. All the steps of the proof are straightfor­
wardly formalizable in SC and no novel inductions are needed.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 218

3. C o m p lex A n alysis

3 .1 . T h e C o m p lex P lan e. A complex number x + yi will be


represented as a pair (x, y) of real numbers. So a complex number
is an object of the type (o —> o ) . Other notions, such as norm, con­
jugation, sum, product and so on, are easily defined and the basic
properties are also simple. For convenience we will make this con­
vention: When in a context we treat z = (x, y) as a complex num­
ber, we let z (m ) = (x(2m ) , y ( 2m )), instead of (x ( m ),y (m )). Fur­
ther, for rational numbers a, b, we also call (a, b) a complex number,
by which we mean (Am.o, Am.b). So, z ( m ) is a complex number and
|z —z ( m) \ < 1/m . That is, z ( m ) is a 1/m approximation to z in the
norm of complex numbers. We will frequently use the complex num­
bers with rational real and imaginary parts to approximate an arbitrary
complex number. We will call them rational complex numbers. Notice
that quantifications over rational complex numbers are quantifications
over objects of the base type o.

3 .2 . D e riv a tiv es. This section, p. 130-134, is simple. The con­


cepts and results there are similar to the corresponding ones for the

real analysis. We will skip the details.

3.3. Integration. ‘7 is a path with the parameter interval [a, 6]’


is translated as ‘7 : [a, b] —►<C, and there exists a finite sequence

(to, •••,*»») ° f real numbers such that <0 = a < t\ < ... < tn = b
and 7 is differentiable on each [f»,ft+i], i = 0 ,..., n — 1.’ Sometimes
we ignore the parameter interval and simply say ‘7 is a path’. This
should be understood as a simplified manner of speech. Other notions,
the equivalence of two paths, the sum of two paths and so on, are
formalized easily.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 219

A form u> = f d x + gdy on a compact subset K of C is represented


by a pair (f , g ) of continuous functions on K . The integration of a
form along a path is constructed as a term I of an appropriate type
such that if a; is a form on K and 7 is a path with parameter interval
[a, b] in K , then

j u = /( w ,7 ,( a ,i»
•nr

is the integration, that is, the limit of some sequence of partial sums.
Notice that I should operate on the witnesses for u/’s being a form
and 7 ’s being a path with the parameter interval [a, 6], though we do
not indicate that in the notation. The basic properties of integration

mentioned on p. 135-136 are simple.


A polygonal path 7 = poly ( z i , ..., z„) is a path with the parameter
interval [0 ,n — 1] such that for some sequence (z i,...,z „ ) of complex
numbers, 7 is linear on [z, i + 1] and 7 (1) = z *+1 for z = 0 , ...,rz — 1 .
We can also take poly as a function from the set of finite sequences of
complex numbers to the set of polygonal paths.
The definition of a convex subset 5 of C needs some caution, for
x € 5 may be a complicated formula and we cannot use induction
to generalize the convex combinations of two elements to the convex
combinations of arbitrarily many finite elements. So, we will define
convexity this way: S is convex if for any finite sequence (z1, ..., z„)
of complex numbers in 5 and any finite sequence (ai,...,O n) of real

numbers such that a* 6 [0 , 1] for each z and £?=i a< = 1 , we have


O iZ i 6 S. Then the convex set spanned by a finite sequence of
complex numbers can be defined and proved to be convex.
The definition of analytic functions on p.137 is simple. Consider
the proof of Proposition (3.6) on p. 137-138. It uses an inductive
construction to construct a sequence of triangular paths ( 7 „). Clearly,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 220

to prove the proposition, it suffices to prove that for any triangular


paths 71 = poly (zi,Z 2 ,Z3 , z i ) determined by rational complex num­
bers Zj, 2 2 , z 2 whose span is well contained in U, we have J* / (z ) dz =
0. Then we can construct the sequence (7 „ ) by constructing a se­
quence ((zri'i, z „ i2 , Z n i3) ) of triples of rational complex. Given any
e > 0, the sequence is constructed informally as follows. Suppose
that ( z „ , i , z n , 2 , Zn,3 ) and 7„ = poly (zntl, z ^ , z „ i3, z „ , i ) have been con­
structed. Write

7 n ,l = poly (zn. 1 , £ (Z n ,l + 2 n ,2 ) , £ (Zn,l + Zn ,3 ) , Z * i) ,

7 n ,2 = P0ly ( * n ,2 , £ (^ n ,2 + ^ n .s ) , £ (Z n .l + * n , 2 ) , 2 n ,2) ,

7 n .3 = poly ( z nj , ^ (Z n ,i + Z „ i3 ) , £ ( z „ ,2 + Zn,3 ) , Zn ,3 ) ,

rn,4 = poly Q (Zn,l + 2n,2) , £ (2n,2 + *n,3) , £ ( * n ,l + * m ) , £ (*n,l + V * ) )

Compute / fni f ( z ) d z , i = 1 ,2 ,3 ,4 , to a sufficient degree of precision


and choose i such that

1/ f ( z ) d z > -l ( £ \ f S( z ) d. - 2-n4~ne
r'fcM 4 \i= i r 'w

. ( Jdz - 2_n4~ne.
*il l /w *

Then let Zn+x,x — ^n,l, ^n+1,2 — j (^n.l ^ 1,2), ^n,3 — j (^n.l d" 2n,3 )
when i = 1 and so on. Notice that, the term /7b i f (z) dz for example
should involve the witnesses for 7 being a path, but they can be
constructed directly from the points z^x, | (z„(x + Zn,a), 5 (z,,^
There is no need for constructing a sequence of witnesses separately.
Clearly, the construction of ((z n ,i, 2n,2, 2^ 3)) is obtained by a primitive
recursion. Then the assertions (3.6.1) and (3.6.2) on p. 138 follow from

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 221

the construction directly and (3.6.3) can also be obtained by induction,


for it is a formula. After that, the rest of the proof is straightforward.
The notions and theorems in the rest of the section pose no further
difficulties.

3.4. The Winding Number. The concepts and proofs from Propo­
sition (4.1) on p. 142 to Lem m a (4.9) on p . 148 are straightforward. The
proof of Theorem (4.10) on p. 148 uses an induction to show the exis­
tence of all finite order derivatives of / from the differentiability of
/ ' . This can be eliminated as follows. Notice that ‘/ has any order
of derivatives on U ’ is translated as ‘for any n > 0, there exists a se­

quence ( / o , . . . , / n) of functions on U such that /o = / and / i +1 = / '


for i = 0, ...,n — 1.’ Now, suppose that / is differentiable on U. For
any z 6 U, there exists r > 0 such that S c ( z , r ) C U. r is among the
witnesses for z 6 U. So we can define a function f n on U by

fn (*) s ( 2trz)~x n! I / ( ( ) ( ( - z)~n~l d£,


•'70

where 70 is the circular path of radius r about z. /„ is a function


on U for the right hand side is independent of the choice of r. From
Theorem (4.7) on p.147 it follows that /o = / . We show that = / n+1
on U. From the proof of Lemma (4.9) we see that it suffices to show
that f'n = / n+1 on any closed sphere well contained in U. Let 5 c (z 0>t)
be such a sphere and choose t' > t such that Sc(zQ,t‘) is still well
contained in U. Let 7 ' be the circular path of radius t 1 about z0. Then
we have

/„ (zo) = ( 2 ™ )-1 n! / / ( 0 (e - *o)'n - 1 4 .


jy
For each z 6 Sc(zo,£), choose r sufficiently small such that S c ( z , r ) C
Sc ( z o, t ,)\ then 7 ' is homotopic in U with the circular path of radius r

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 222

about z. Therefore

fn(2) =(2’ri)”1n-J'nf /(OK-*)-*-1


= ( 2w i y ' n l f
Jyi

Similarly,

/ n+1 (z) = (27tz)-1 (n + 1)! / / ( 0 (£ - z ) ' n" 2 #


Jyi

on S c ( z 0 ,t). Then by Proposition (4.8) on p.147 we have f n = / n+1 on


Sc ( z 0 ,t). This completes the proof that / has derivatives of any order

an d f n = / (n)-

Now we prove the formula (4.10.1) on p.148. It suffices to prove


the conclusion for any rational complex number z in U — car ( 7 ) . Let
r > 0 be a rational number such that S c ( z , r ) is well contained in
U — car (7). The statement ‘for any rational complex number u such
that |u — z| < r,

j ( 7 .« ) f n (« ) = (2 tti)-1 n\ f f ( 0 (£ - u p " 1 d£’


J-i
is n ° and is equivalent to the statement ‘for any u 6 S c ( z , r ) the same
equation holds.’ Then, by Theorem (4.7) and Proposition (4.8) and
an induction, it follows that the equation holds on S c ( z , r ) for any
n. (4.10.1) is obtained by taking u = z. This completes the proof of
Theorem (4.10).
The rest of the section poses no more difficulties.

3.5. Estimates of Size, and Location of Zeros. There are no


difficulties for Definition (5.1) through Lemma (5.4). The proof of
Lemma (5.5) on p. 154 uses a generalization of Corollary (4.14) that if
a differentiable function f on U vanishes at points z l t ..., zn in U such
that |Zj — Zfc| > d > 0 (1 < j < k < n), then the function h (£) =

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 223

(g-nM l-tn) 011 ^ ~ { zii —i zn} extends to a differentiable function on


U. This can be proved without further inductions, for the proof of
Lemma (4.6) on p. 145-146 works for any finite number of points
z i , zn in place of a single point, as long as they satisfy \zj — z*! > d
(1 < j < k < n) for some d > 0 .

Lemma (5.6) and (5.7) are straightforward. Lemma (5.8) is ob­


tained by applying Lemma (5.7) recursively to construct a sequence
(Sn) of closed spheres converging to a zero point of the given function
/ . Examining the proof of Lemma (5.7) we see that if z, £ are rational
complex numbers and r > 0 , e > 0 , t > 0 are also rational such that
( € S (z, r) and m ( / , T (z , r)) > \ f (£)| + e, then, with some minor re­
visions, we can construct rational complex numbers w, and positive
rational numbers t 0 < t and £q such that ( ' 6 5 (to, t0) C S (z, r) and

m ( / , T (w , tQ)) > | / (£')|+£o- S o , to construct the sequence (Sn) we can


instead construct sequences of rational complex numbers (zn) , (£„) and
sequences of rational numbers (rn), (en) satisfying the conditions above
and further rn < n"1. Such sequences can be obtained by primitive
recursions available in SC .

Now consider the proof of Theorem (5.10), the fundamental theo­


rem of algebra, on p. 156-157. The proof applies Lemma (5.8) recur­
sively to construct zeros z i,z a ,... of the given polynomial p ( z ) . Since
each Zi in turn is obtained by constructing a sequence (z< (m ))“ =1 ap­
proximating to Zi, this involves an iterated recursion. However, it can
be reduced to primitive recursions, because we will show that to com­
pute Zi+i (m + 1), we will need z,+i (m ) and z, (m') for some m', where
m! may further depend on z,- (m") for some m", but we will not need

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 224
Zi+i (m). That is, the recursion pattern is largely of the format

z ( i + l,m + l) = / ( z ( i + l , m ) , z (i, g ( z ( i , h ( i , m ) ) , i,m ) ) , i, m ) .

This can be reduced to primitive recursions. It differs from the re­


cursion pattern for Ackermann’s function just in that z (z -f 1, m) does
not appear in any context of the form z (x,......). The details are given
below. Regrettably, they are a bit lengthy because of the complexity
of the construction.
The following is the recursion pattern we actually need.

LEMMA 3.1. Given primitive recursive functions p, q, r, s, t, / , 5 , A,


we can construct primitive recursive functions b, z such that

b( 0 , m ) = q ( m ) , z ( i -I-1, 0 ) = p ( i ) , z ( 0 ,m ) = t(m ),

(17)
z ( i + l,m + 1) = r (z (z + l , m ) , b ( i , g { m , S ) ) , 5, m ),

(18) b(i + l,m ) = s ( z ( i + 1, $ ) , $ , m, i),

where S is b ( i , f ( b (i, h (b (t, 1 ))))), and $ , 'J are expressions that may
contain several iterations of the contexts of the form b ( i , ) but con­
tain no z or any context of the form b(i + 1 ,....).

PROOF. First we use the course of value functions:

Z ( i , m ) = ( z ( i , 0 ) ,...,z ( i,m ) ) ,

B ( i , m ) = (6 ( t,0 ),...,6 (i,m )).

Obviously, i?(0, m ), Z (i -(-1,0 ), and Z ( 0 ,m ) can be computed di­


rectly. Notice that b (i, N ) — B (x, M ) N for any M > N . Suppose that

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
3. COMPLEX ANALYSIS 225
the coding of sequences guarantees that B (i, N ) > N. Then we can
see that the recursion equation (17) can be reduced to

(19) Z ( i + l,r a + 1) =

f l ( Z ( i + l ,m ) , 1 ))))))) , m, i)

for some primitive recursive functions R, F, G, H.


Let Q (m) denote B ( i , G ( m , B (t, F (B (i, H ( B (i, 1 ))))))). If we
apply the equation (19) to itself m + 1 times we should obtain

Z ( i + l , m + l) = R ( R ( p ( i + 1 ))Q (0 ) ,0 , i ) Q ( m ) , m , i ) .

We can choose G such that G ( k , x ) > G ( j , x ) > m ax{y, x } for k > j,


then we can see that for some P ,

Q(k) = P ( Q ( m ) , k )

for k < m . So, we see that the equation can further be reduced to

(20) Z ( i + l,m + 1 ) = iZ '(Q (m ), m, z)

for some R!.


By a similar argument, the equation (18) can first be reduced to

b (i + 1, m) = S' ( Z (i -I-1, V ) , m, i ) ,

where expressions contain some iterations of the contexts of the


form B ( i , ..... ). Write and 9 ' as (m ) and (m) respectively. By
the same argument as above, we have Pi, P 3 such that

&(k) = Pi(& (m ),k ),

Z ( i + 1, ¥ ' (*)) = P 2 (Z (i + 1, (m ) ) , k) ,

for k < m . There exists S such that

S (x, y , m , i ) = (S' (P 2 (x, 0 ), Px ( y , 0 ) , 0 ,» ) ,..., S' (P 2 (x, m ) , Px ( y , m ) , m, i)) .

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. COMPLEX ANALYSIS 226
So the equation (18) is finally reduced to

(21) B (i + 1, m) = S {Z (i + 1 , (m )), ( m ) , m, i ) .

Substitute (20) into this (letting m = (m ) — 1 ) we obtain a re­


cursion equation for B (i + I, m) such that the right hand side contains
some iterations of the contexts B (i ,...) but no contexts like B (i + 1,...)
and no occurrences of Z. This is one of the recursion patterns we dis­
cussed in Section 4. It can be reduced to the plain primitive recursions.
Construct B to satisfy this recursion equation and define Z according
to ( 20 ); then we can recover z and b to satisfy the original equations. |

To formalize the proof of the fundamental theorem of algebra we


need some more subtle formulations of Lemma (5.6) and (5.7) on p.
155-156. They are given below. The main point is to show that to
construct the zeros of a polynomial function to the precision ± m -1 we
need only the approximate values of the coefficients to a precision deter­
mined by some primitive recursive functions of m. For a = (ao,..., a*) a
sequence of complex numbers, we use pu to denote the polynomial func­
tion aozn-f...+ a n . Seep. 153 for the notations r(u>, a ) , m ( f , r ( w , s ) ) .

LEMMA 3 .2 . There are primitive recursive functions F , G , E such


that if r, 6 , e , 0 are positive rational numbers such that r > 38, w is

a rational complex number, a = (a0 )..., a*) is a sequence of complex


numbers, and S = S c ( w , r ) , T = S c ( w , r —38), then for

s = G (r, £,e, 0 , w, a (.F f y .^ e ,# , 10, 0 ( 1 )))),

c = E ( r , 8 , e , 0 , w , a ( 1))

if ||Pa|lx > 0 , then 8 —e < s < 8 and m (pa, T (u>, s )) > c. Farther, F
can be replaced by any primitive recursive F' such that F' > F fo r any
values of the arguments.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 227
PROOF. The existence of s , c is an instance of Lemma (5.6) on p.
155. We will follow the proof of that lemma on p. 155, with pa in place
of / , to verify that the construction of s, c uses only approximations to
a of the form a (F (r, 6 , e, fi, w, a (1 ))), a (1) for some F. First, a bound
M for |a0| ,..., |On| can be computed from a (mo) for an arbitrary mo.

Then a bound M ' for jjp,,^ can t>e estimated from M and r,u;. So,
N, ri,...,rj\r, d, and a in the proof on p. 155 can all be found with

fi in place of ||p„||x in place of ||pa||s- Notice that a is a


primitive recursive function of r, 6 , e, fi, w , a (1 ). Next, we must decide
if m (p„, T (w, r*)) < a or > f a for each k, choose one that makes the
latter true, and let s = r*. and c = |a . c is a primitive recursive function
of r, 6 , e , 0 , w , a ( l ) . We must estimate m (p a, T (w , r*)) to some degree
of precision determined by a. From a , the bound M , the center w,
and the radius r we can find, primitive recursively, a number m such
that for any m' > m , pa (u) —p„(TO') (u)| < ^ for all u 6 S c ( w , r ) . Let
mo = 1 in estimating the bound M. Then m = F(r,S,e,./3, w , a ( 1)) for
some F and F can be replaced by any F' as stated in the lemma. Also,

the modulus of continuity ui for p„ on Sc ( w, r ) can be determined from


the bound Af, and w , r . Further, we can find an u; approximation
{ z u ...,zi} to

j u : rk - w < |u - w\ < rk + u ^ j |

consisting of rational complex numbers. Then it suffices to compute

Pa(m) (z»)|) i = 1, •••, I, and decide if each is less than £ct or greater than
|a . If all are greater than fat then m (pa,r(u/,rfc)) > f a , otherwise
m (p 0 ,r ( t i 7,rfc)) < a . We can let m 0 = .F ( r ,f ,e ,/? ,a ( l) ) when it is
used other than computing m above, so only the approximation to a of
the form a ( F (r, 6 , e, fi, a (1))) is ultimately used in determining r*. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 228
LEMMA 3 . 3 . There exist -primitive recursive functions F , E , T such
that if R, r, e, t are positive rational numbers, and z , ( are rational com­
plex numbers such that 0 < t < r, \z\ + r < R, £ € S ( z , r), and
m (p„, T (z, r)) > |pa (£)| + e, then there exist positive rational numbers
to, £o, rational complex numbers w, 6 such that to < t, |u; —z 14-to < r,

6 € S (to,to), \pa ( 6 ) 1 < f , £o < f , arid

m (p0, T (w , t0)) > |pa (6)1 + £o-

Further, w, tQ, 6 , £o are primitive recursive functions of R, r,e, t, z, (,


and the approximation a ( F ( R , t, t, a ( l) ) ) to a. And we have £<j >
E ( R , t , e , a ( 1)) > 0 and t 0 > r ( / 2 Jf ,e ,o ( l ) ) > 0. In particular, the
approximation to a used in determining to, to, 6 ?£o does not dependent
on z , £ , r and z, (, r do not appear in the arguments o f E ,T either.

PROOF. The existence of to, to, 6 ) £o is an instance of Lemma (5.7)


on p .155-156. To verify that the constructions depend only on the
approximation to a of the form a ( F ( R , t, e, a ( l ) ) ) , we examine the
steps of the proof on p.156, letting K = S c ( z , r ) , B = T (z ,r ). For
any a, 0 < a < t < r, K' = S c ( z , r —a) is compact. A bound M
of the coefficients of pa can be computed from a (mo) for an arbitrary
mo and then the modulus of continuity ui of pa on Sc (R) (and hence
on K since K C S c ( R ) ) can be determined from M and R. Let
a = | m in {a; , t j . Then a is primitive recursive on R, t, e, M, and

m (po, K fl B 3a) = m ( p a, { u : r - 2a < \ u —z \ < r } )

> m ( p a, B ) ~ | > |.

Notice that a does not depend on z, 6 r.


Choose a approximation { * i , ..., to S c ( R ) consisting of ra­
tional complex numbers. This depends on a and R only. Notice that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 229

this is different from the approximation to K ' used in the original


proof on p .156. For each j = let 5 = S c ( z j , 2 R + a ), T =
S c ( z j , 2 R ) . We have S c ( R) C T, and hence IIpoIIt > m (Pa, K) > t.
Applying Lemma 3.2 with S , T as given here, 6 = f , e = f — | , and
0 = £ here, we can compute a positive rational number t j ,primitive
recursively on R, cl, £, Zj and an approximation a ( F (R, a , £, Zj, a (1 ))),
and a rational number Cj, primitive recursively on R , a , £ , j , such that
tj 6 f ) and m (pa, F (zj, tj)) > Cj. Replacing F by a function ob­
tained by taking maximum for j = 1 ,..., N, we see that tj is a function

of R, at , j and the approximation a ( F ( R, t , e , a ( l ) ) ) for some F. Let


7 = j m in { f , ci( ...,c ^ | and 7 ' = m in {a,a; ( j ) } - We see that 7 , 7 '
are functions of R, t, e, a (1).
Choose a 7 ' approximation {£ 1, ...,& } to S c { R ) consisting of ra­
tional complex numbers. So, & is a primitive recursive function of
R, t,e , i and a ( 1) . From 7 , the radius R, and the bound M for the
coefficients of p0 we can compute m such that for any m' > m and
u 6 S c ( R) , Pa (u) - pa(m>) (u)| < J. So, m = F (R, t, £, a (1)) for some
primitive recursive function F. For each i = 1 ,..., I, we can decide

if IP«(m) ( 6 ) I > i or < J. If |p0(m) (fi)| > \ for aU f,- € K (which is


decidable since K = S c ( z , r ) and z , r are rational), then |pa (f»)| > J
for all f, 6 K . For any u 6 K , u 6 K D B 2a or u 6 K' . It is easy
to verify that in both cases |pa (u)| > J. This contradicts with the
condition m ( p a, B ) > |p„ ( f ) | + e because of Corollary (5.3) on p .153.
Choose fi € K such that p „ (m )(f i)| < \ and choose Zj such that
fi 6 Sc (Zj, | ) . We have |pa ( f i ) | < 7 < f . Since m (pa, K n B 2a) > f i

we have |f i — z\ < r — | o , \zj —z\ < r - f a , and S c ( z j , t j ) C K . So,


we finally let z 0 = Zj, to = tj, f0 = fi, and £q = 7.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 230

Clearly, zq, tQ, £0) £o are primitive recursive functions of R, t, e, z, r, (


and the approximation a (m ) = a ( F (R, t , e, a (1))). We can replace F
by ma x{ F, G } we see that they are functions of R , t , e , z , r , ( and the
approximations of the form a ( F ( R , t , e, a ( l ) ) ) . £o = 7 and to > a /4
while 7 and a both are functions of R, t, e, a ( 1 ). |

The following lemma elaborates a step in the original proof on p. 157.

LEMMA 3 .4 . There exists a primitive recursive function H such

that if a = (a o ,...,a n) is a sequence of complex numbers, 0 < j < n,


and e > 0 is a rational number such that \ | > c, then there exists a
positive rational number r, depending prim itive recursively on j , £ and
the approximation a ( H (e,j, a (I))), such that

m (Pa, T (r)) > |P o ( 0 )| + ? .

Further, H can be replaced by any function H' > H.

PROOF. First we can find M , depending o n a ( l ) such th at

w > E M + 1* . (o)| + i.
»=0

Construct a sequence (e n - j - m ) m such that e n - j = f, £ n - j - m - i =


2$ * $ ' ^et N be any integer such that JV > ^ + 1. N is a function
of e ,j , and a ( l ) . Denote a 1 = a ( N ) and a[ = at (N). For m = j , ...,n ,

®n—m S n -m < Since |a„_y| > e, we have ja'n_j > f = en- j • If

Gn —m > £n
> C n—
-jj —
-1 for som em = j + 1 , ...,n , then > en-y-x > en - m -
Repeating the process in less than n —j times we must obtain some

jo, j < jo < n, such that a'n_K > and < _ m < £n- » - i for
m = jo + 1, ...,n . Clearly, jo is a primitive function of s , j and a ( N ) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 231
Let r = Some direct calculations will show that
o

|an-iolr '° > £ |On-m|rm + £ |<ln-m|rm + \jpa ( 0 )| +


Tn~0 m=jo+l
The conclusion then follows. |

Now we can begin formalizing the proof of Theorem (5.10) on p.156-


157. So, suppose that a = (ao,..., On) is a sequence of complex numbers,
0 < A < n, A < j < n, and t is a positive rational number such that
|on_j| > e. We will construct a sequence z i , ..., Zk of complex numbers,
a sequence bQt..., 6* with b0 = a and &,• = (b*Q,..., 6J,), z = 1,..., A, such
that

Pa ( z ) = (z ~ *0 (2 ~ z <)Pbi (2)

for z = 0 ,..., A. Then the theorem follows.


*(We can assume that for any A, <z(A) > A if taken as a natu­
ral number itself, for otherwise we can replace a by another sequence
a' = (oq, a'n) satisfying the condition and such that for z = 0 , ...,ra,
a, = a[ as complex numbers.)* To apply the lemmas repeatedly to con­
struct sequences (zj (m ))“ _0 and ( 6, (m))®_0, we must simultaneously

construct sequences ( f t ( m ))m > (& (m ))m positive ratio­


ned numbers and rational complex numbers, and sequences ( j i ( m) ) m,
(e* (m ))m of integers and positive rational numbers which are supposed

to satisfy

( 22 ) & (m ) 6 S c ( z i( m ) ,t i( m ) ) , 0 i(m )> 0

(23) m a x ( & ( m ) ) |, f t (m )} < \&i (m — 1),

(24)
m (pbi-L , r (Zi (m ) ,ti (m ))) > ( £ (m ))| + ft ( m ) ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 232

(25) j i (m ) = j i ( 1 ) , d ( m ) = £i (1) ,

(26) |* U ( i ) | > f t ( l ) , j i ( l ) > j . - - i ( l ) - l

for all m and i = 1 , k and thus provide the necessary conditions for
the applications of the lemmas at stage m -f 1 . So we will construct 2
double sequences

2 (i, m) = {zi ( m ) , U (m ) , f t (m ), & (m )),

b (i , m) = (bi ( m ) , ji ( m ) , a (m)).

b( 0 ,m ) can be determined from the given a ,j,£ directly. z ( 0 ,m )


and z ( i + 1,0) will be irrelevant. To compute z ( i + l,m + 1) we dis­
tinguish two cases. First suppose that m = 0. By Lemma 3.4 above,
we have a primitive recursive function H such that if &,■ = (blQ, ..., bln),
ji, and e, were already constructed and satisfy the conditions (25) and
(26), then we can use the approximation b ( i , H ( b ( i , 1))) to compute
an r such that

™ (P6,-.r(r)) > |pb.-(0 )| + J.

Then apply Lemma 3.3 with R = r, r = r, z = 0, £ = 0, e =


and t = m i n | j , r | and obtain tu,£o,£o,*o- So, we let z ,+ i(l) = w,
U+i (1) = *0) ft+ i (1) = So, and &+i (1) = ^0. The recursion equation
for z (» + 1 , 1 ) takes the form

(27) z ( t + l , l ) = r 1 (fe(i, F { b ( i , f f ( b ( i , l ) ) ) ) ) ) .

Now suppose that m > 0. Let E, T be the functions in Lemma 3.3,


where we may assume that T (R, t, e, a ( 1)) < | given the assumptions

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. COMPLEX ANALYSIS 233

of that lemma. We can construct E ^ T ^ , such that

£• = 0 M (1 ), t ; = t j+, ( 1 ) < i

(1)X.-Si. Ml)),
r ; +i =r(ii+1(i)X,XMi))-
So, E xm and are primitive recursive functions of m and b (z, F (b (z, H (6 (z, 1)))))
by (27). In particular, we don’t need z (z + 1, m) in computing T^, E xm.
Further 0 < < 2-m . Suppose that the construction up to stage
m has guaranteed that tj+i (rn) > and 0 i+1 (m ) > E xm - Apply
Lemma 3.3 above with Zi+i (m ) as z, t i+1 (m ) as r, ti+ i (m ) as t , ti+i ( 1 )
as R, E xm as t, as t, and the supposed bi as a. Then we have
w, to, £o, e0, depending primitive recursively on z (z + 1, m ), z (i + 1, 1),

m, b (z, F (b (z, H (b (z, 1 ))))), and

6 (z, G ( z(z + 1, 1) , m , b ( i , F ( 6 (z, H ( 6 (z, 1))))), 6 ( i , l ) ) ) .

Notice that z ( i + l,m ) does not appear in the context b ( i , ...) because
we use E^, instead of U+i ( m ) , /3,+i (m). We will let Z{+i (m + 1) =
w, ti+1 (m + 1) = to, &+i ( m + 1) = to, and ft+ i ( m + 1) = e0. Then
we have <i+i (m + 1) > T U i and /?t+i (m + 1) > E^+i- Use the equa­
tion (27) for z (z -f 1 , 1 ) and notice whenever b(i, 1 ) is used in the com­
putation we can use instead b ( i , H ( b ( i , 1 ) ) ), we can reduce the above

expression into

6(t,G(m,4(i,F(6(i,ff(4(i,l))))))).
So we see that the recursion equation for z (z + 1, m + 1 ) takes the form

(28)
z ( z - t - l,m + l) = r ( z ( z + l , m ) , m, b ( i , G ( m , Q ) ) , Q ) .

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. COMPLEX ANALYSIS 234

where Q is b(i, F 1 ))))). We can subsume (27) under this


pattern. So this is the equation for any m > 0.
Prom the construction we see that *»+i (m + 1) < T ' m < 2"mand
for rn! > m, 2,-+1 {w!) 6 S c ( z i+i (m ), ti+x (m)). So, (z ,-+1(m ))“ =1 is
supposed to a complex number. Now we can give the recursion equation
for b(i + l,m ) . The supposed connection between = (b'0, ..., bln) and

bi+i = (&o+1i —>^n+1) is given by the equation

E hn-kzk = (z ~ *i+i) E *«-***•


fc=0 fc=0

The construction should have guaranteed that 6J, = ... = = 0. So,

we set 6J,+1 = ••• = K+1 = 0. Then we have the relations

= 4S« + = » !+ , +

^1+3 = b'i+2 + ^!+2Zt+l = K + 2 + ^+1++1 + K Zi + l ,

To compute 6{+1 (m ), I = i + l ,...,n , which are supposed to be the


approximations to b\ +1 to the degree of precision within ± ^ , we need
b\ (N) and Zj+i (N), where N depends on m and the bounds for |6{|
and |zi+i|. So, there are primitive recursive functions C, &i such that

(29)
bi+1 (m)

= 5! ( 2 ( i + l , C ( 6 ( i , l ) , z ( i + l , l ) , m ) ) , 6 ( i , C ( 6 ( t , l ) , *( * + 1 , 1 ) , m ) ) , t ) .

If = n —i, we set y,-+i = n —i — 1 and e,+i = et. If ji < n — t, we


have

e u - .i - e w . =

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 235

Use some approximations 6,+1 ( N 1), z,-+i (IV7), depending on £< and
bounds for |z,-+i| and &J.+1 , we can choose s = or ji — 1 such that

2 (|zi+, ( 1)1 + 2 )

We set (1) = s, e ,+ i(l) = and j.+ i (m ) = j 1+ I(l) ,


£i+1 (m) = £j+i ( 1) for m > 1. Notice that N ‘ depends on

bi+i (1 ), and b(i, 1 ). So, j<+i (to) and £,+i (m) are primitive recursive
functions of

bi+i ( A ( z ( i + 1 , 1 ) , bi+i ( l ) , b(i, 1 ))),

* + i ( A ( z ( i + l , l ) , bi+1 ( 1 ), 6 (i, 1))),

6 (* , 1) , z (i + l , l ) , i , m.

Using the equation (27) for z (i + 1,1) and (29) for bi+ 1 (...), and com­
bining with (29) we see that the recursion equation for b (i + 1, m) takes
the form

(30) b(i + l,m ) = s ( z ( i + 1, tf), $ , m, i ) ,

where $ , $ are expressions which contain several iterations of the con­


texts of the form b (i , ..... ) but contain no contexts of the form b (i + 1, ...... ).
By Lemma 3.1 above, we can construct primitive recursive functions
z, b that satisfy equation (28) and (30) and the equations for the initial
values. We can use induction to prove simultaneously that for i < k, (i)
6} are complex numbers for I = 0 , ...,n; (ii) b\_j. > £<; (iii) j i > k —i\

(iii) (z,- (m ) ) " =1 is a complex number; (iv) pb{_l (z,) = 0 ; (v)

£ bn-lzl = (z ~ 2») £ b ^ Z 1.
1=0 1=0

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
3. COMPLEX ANALYSIS 236
These are ail II? statements where the last equation should be un­
derstood as some equations between the coefficients of the polynomi­
als. Another induction with assumption is needed for proving the in­
ductive step. That is, assuming that (i) to (v) hold for i, we can
prove by an induction on m such that equations (22), (23), (24),
Sc (z,- (m -I-1 ), t{ (m + 1)) C S (z* (m ), U (m )), and U (m + 1) < 2-m
together with the initial condition for the case m = 1 all hold with i
replaced by i + 1 . Then it follows that (iii) and (iv) above hold for z + 1
and then others of (i) to (v) for z + 1 follow easily. The conclusion of
the fundamental theorem of algebra follows for z = k.
Now we consider the proof of Theorem (5.11) on p.157-159. It needs
a similar construction. First we introduce some notations. Let Z =
(z\ ,..., Zk) be a sequence of complex numbers and d > 0 be rational. For
each z = l , ..., k, by computing \z\ — Zj|, ..., |z* — Zj| to some sufficient
degree of precision we can choose a rational number r 6 (3d, Ad) such
that \\zj — Zi\ —r\ > ^ for j i. So, there exists R such that for
Z , d , i as above,

11Zj - Zi\ - R ( i , Z , d ) \ > for j ^ i.

Let 7 (i, Z, d) denote the circular path T ( z i , R ( i , Z , d ) ) with radius


R (i, Z, d) and center z,-. When Z, d are fixed in the context, we will
simply denote it by 7 ». We have p (z j, 7 ,) > ^ for all j = 1,..., k.
Throughout the discussions of Theorem (5.11) on p .157 below, we
make the following assumptions: ( 1 ) / is a differentiable function on a
compact set K C C with border B such that m ( / , B ) > m ( / , K ) = 0;

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. COMPLEX ANALYSIS 237

(2) d < | is a positive real number such that K fl B m ,

K ' = { u e K : p ( u , B ) > I d} , and

K" = {u e K : p (u, B) > 2d}

are all compact and m ( f , K C\ B m ) > ^ m (/, B)\ (3) a; is a modulus

of continuity for / on K \ (4) M is a positive numbersuch that K C


5 and ||/ ||x < M; (5) {«i, ...,vj} is a d approximation to K ' and
{iyi,...,tui»} is a d approximation to K", and

V = {u 6 C : \u —i/{| < d for some i = 1, / } ,

V' = { u £ C : \u —Wi\ < d for some i = 1, / ' } .

and (6 ) u0 6 V and 0 O > 0 are such that m ( f , B ) > | / ( u 0)| +/?o-


Notice that since B is the border,

K' C V C {u 6 K : p(u,B) > 6 d} C

K" c V' C {u G K : p { u , B ) > d } C K .

Further, z £ V and z € V' are formula and when z is a rational


complex number, they are actually £?. When we constructfunctions
or sequences, we will use d, M, and the witnesses for the compactness
of K, K' , K" and so on without explicit mentioning them.
We can construct a term G such that if Z = ( z i , ..., Zk) is a sequence
of complex numbers with p (zj, B ) > 4d for t = 1 ,..., k and ( E K , then

(i)

^ \t - * i ( m ) > i d f° r * = • • • > “ d (“ )

Q( % ^ f _________ f iU)_________
2 ici Jyj (u - z i ) ... (u - zk) (u - 0

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 238

if j is the least number such that |f — Zj| ( [ | l ) < \d . Notice that


G actually applies to the witnesses for 2* 6 V, i = 1 , k, and for
f 6 K , though these do not appear in the notation. Further notice
that the conditions in the cases (i) and (ii) imply |f — Zi\ > d for
i = 1, k and |f — zy| < 2d respectively. The construction guarantees
that G ( Z , f) is a complex number if Z , { satisfy the condition, but it
does not guarantee that G ( Z , £ ) is a function of f for given Z, for which
we need the condition that f = £ f' implies G ( Z , ( ) = c G ( Z , ('). We
have the following lemma that will be needed in the induction below.

LEMMA 3 . 5 . Under the assumption ‘Z = (21, ...,2*.) and € V for

i = 1 ,..., k the statement ‘\£ .G (Z,£) is differentiable function on K


that extends the function on K —{ 2 1 , ..., Z k } ’ is equivalent
to a n ° formula. Further, there is a primitive recursive function Q
of k, e such that when the statement holds, Ae.fi (k, e) is a modulus of
continuity o f \ ( .G (Z , £) on K , in particular, the modulus of continuity
does not depend on Z.

PROOF. Consider the conjunction of the following three sentences:


(i) For any rational complex number f and j = 1,..., k, if £ 6 V
|f — 2j| > 0 for i = l , ..., k , and |f — Zj\ < | d, then

m = j_ [ /(* ) iu
(£ 2i) ... (f 2fc) 2 r iJ rj( v - z l) ...( u - g k ) ( v - ( )

(ii) For any rational complex numbers f and j, j ' = 1, ...k, if f 6 V',
- 2y| < |d and |f - Zj>\ < | d, then

/ ?—
(u - 21)r...r (uM- - r (un - rf) * - / (W
Zk) ?—- r (ur r (u -- '1 Z i ) ... - Zk) 0

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. COMPLEX ANALYSIS 239

(iii) For any rational complex numbers £1 , ( 2 ,£ 3 , if 6., £2, £3 6 V'

and 16 ~ 61 < d, |6 - 61 < d, and |6 - 61 < d, then

L G ( Z , u ) d u = 0.

Three formulas are all 11°, for the antecedents in the ‘if’ clauses,

£ € V, |£ - Zi\ > 0 and so on, are all Ej since £ is of the type of a


rational complex number, and the consequents are equalities, which are
n °. We will show that the conjunction of (i), (ii), and (iii) is equivalent
to the statement under the assumption.
First assume that (i), (ii), and (iii) hold. Suppose that 6 > 6 £ K
and 6 = c 6 - We must show that G( Z , 6 ) = c G( Z, 6 ) - First assume
that p ( ( i , B ) = p ( 6 ) 5 ) > 3d. It suffices to consider two cases: (1)

|6 — Zi\ > d for i = 1,..., k and |6 - Zj\ < 2d for some j and

G(Z, 6 ) =
/(6)
(6 - z i ) - ( 6 -* * )’

G ( Z , 6 ) = 5^ / -------- , / (U)' w ------


2 tti J y j (tx - zx) ... ( u - Z k) (u - 6 )
In this case, for any rational complex number £ sufficiently close to

6 = 6 we must kave |f - Z{\ > 0 for t = 1,..., k and |£ — Zj\ < fd, and

further ( € K" C V'. So, by (i),

/(0
/(O . 1 f ____r /(«) -du .
(£ - zx)
zi ) ..... . {(
( £ -- Zk) 2iri Jyj (u - Z \ ) ... (u - zfe) (u - 0
2iri Jy,
Let £ -* 6 = 6 - We have G ( Z , 6 ) = G (Z , 6 )> (2) For some j , j ' ,

16 — < 2d, | 6 — Zj>| < 2 d and

For this case, it follows similarly from (ii) that G ( Z , 6 ) = G ( Z t 6 )-


Now, for any 6 = 6 £ p ( 6 >-®) is either > 3d or < 4d. In the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 240

latter case, | 6 — Zi\ > 2d for i = 1, k since Zi 6 V, and then

G ( Z , 6 ) = - --------- -------------------- r = - r = G (Z ,fc).


(6 - *0 (6
- - zk) (6 - *\) - (6 - zk)

So, G ( Z y£) is a function of £ on K .


Now we show that G ( Z , ( ) extends • So suppose that
£ 6 K and |£ — zt-| > 0 for i = 1,..., k. If p ( 6 5 ) < 4d, then G ( Z , ( ) =
by definition. So suppose that p { £ , B ) > 3d. If G ( Z , ( ) =
by definition, it is done, otherwise for some j , |£ — Zj\ <
2d andG ( Z , £ ) = — f ----- 1 f t ) . ——du. We can choose rational
complex number (' arbitrarily close to ( such that |6 — Zi\ > 0 for
i = l,...,jfe and |£' - Zj\ < |d . Let -► 6 It follows from (i) that

Before we prove the differentiability of G ( Z , ( ) , we must prove

its continuity. Suppose that 6 ) 6 6 K and |6 —£2 1 < d / 8 . In case


16 —Zi\ > d /4 for all z = 1,...,&, we have | 6 — Zi\ > d/8 for all i

and hence |G (Z ,{ ,) - G ( Z ,6)1 < c if 1 6 - 6 1 < u ( ( l ) ‘ £) • Now


suppose that |6 — z,| < d/2 for some j . Then | 6 — Zj\ < d. By the
definition and the proof of the functionality of G ( Z , ( ) above, we must
have

G ( Z ,W = t o l l , ft )* * ’

G(Z,6) = S i I , ( ,u - z ,).J u - z k) ( u - ( ,) dU-


Notice that |tz — 61 > d, |zx —61 > d, and |u - z,| > i = 1,..., k, for

u in the path 7 So, |G ( 2 , 6 ) - <3(2,6)1 < « if 16 - 61 <


So, the modulus of continuity for G ( Z , ( ) is given by

"'(£) =min{ i ' “ ( © £) ' S W £}'

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 241

To show that G { Z , £)is differentiable on K it suffices to show that


G (Z, £) is differentiable on the open set V' and a compact set KCiB# for
some d! € (3d, 4d). For the former it suffices to prove that G (Z, £) is an­
alytic on V' and this in turn will follow if for any 6 i 6 >6 €E V' such that

16 - 61 < d / 2 , | 6 - 61 < d / 2 , | 6 - 61 < d / 2 , and sp a n ( 6 , 6 . 6 ) is


well contained in V', we have fpoiy((ll(a,u,(l) G (Z, u) du = 0. For this we
can approximate 6 >6 > 6 by rational complex numbers in V' and use
(iii) and the continuity of G ( Z , ( ) on K . On the other hand, for any
£ 6 KCiBd>, we have |£ - z,| > dfor all i = 1 ,..., k, and hence G (Z, f ) =
‘ ^ben clearly, G ( Z , ( ) is differentiable on K fl B
Now assume that \ ( . G ( Z , £ ) is differentiable function on K that
extends the function on K —{ z X)..., zk} . Since for u in the
range of the path 7 ,, |u — Zi\ > 0 for i — 1,..., k , we have

_L f // W
W du
2wi J7i (u - z x) ... (u - z k) (u - 6
1 f G{Z,
a ^ \iu
2x i Jfj (ii —0

= G(Z, 0

when ( belongs to the open disc with 7 j as the border. Then (i) and
(ii) are obvious, (iii) is also trivial. So we have the equivalence. |

LEMMA 3 .6 . There exists a prim itive recursive function F such

that Z = {zX)...,zk) is a sequence o f points in V and \£ . G ( Z , £ ) is


a differentiable function on K that extends the function
then fo r N > F { k , m ) , \ G ( Z , { ) - G { Z ( N ) ,()\ < l / m for all £ € K

and \ G ( Z , 0 - G { Z , 6)1 < 1/m for all 6 6 6 K with |£ - 61 < 1/ N.

PROOF. The second inequality is easy to obtain since the modulus


of continuity of G ( Z , 6 on K is already given in the proof of the last

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
3. COMPLEX ANALYSIS 242

lemma. We consider the first. Clearly, there exists a primitive recursive


function F such that for any £ 6 K , and Z\, z [ , ..., zk, z'k € K , we have

/(0 /(fl < 1/m

if for some N > F ( k , m ) and all z = 1,..., k, |£ - Z{\ > d/2, |£ - z[| >
d / 2 , |Zi — z[\ < 1/N ; and

JL/JfeLf 1____________ 1____ - U


27TI J-j (u - £) \ ( u - Zi) ... (u - z k) (u - z [ ) ... (u - z'k)J

< 1/m

if for some N > F (k , m ) and all z = 1 ,..., k, |u —Zi\ > \u —z[\ >

lw ~ £ l ^ <d/2, for all u € 7 , and |z,-—zt'| < 1 /N, where 7 is


a circular path centered at some z'-. Further we may assume that
F(k,m) > Then for N > F ( k , m ) , it suffices to consider two
cases: (i) |£ — *i(lV)| > d for all z = 1, . . . , fc and G ( Z ( N ) , £ ) =

case we ^ave - 2il > ^ for all z = 1,..., k


and, by the assumption on G( Z , £ ) , G ( Z , £ ) = • So, the
conclusion follows, (ii) |£ — Zj (IV) | < 2d and

/(«)
G ^ ^ A(i.^(N)J)
2zrz hj, z{N)J) (u - Zi ( N ))... (u - zk (N
( N)))) ((uu -- (£Y) du'

In this case, £ is also in the disc bordered by 7 (j, Z (N) .d) and hence
by the assumptions on G [Z, £),

Q f Z £) = — / du.
’ 2zrz J^j,z(N).d) ( u - Z i ) ... (u - z fc) ( u - 0

It can easily be checked that |u —Zi\ > ^ and |u — (\ > d/2, for all
u 6 7 (j, Z (IV) .d). So, the conclusion follows. |

Lemma 3 .7 . I f Z = {zi, ...,zk), Z' = (zx, ...,zk, z k+l) are sequences


of points in V, A£. G( Z, £) is a differentiable function on K that extends

i
R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 243
the function , and G (Z , Z k+ 1) = 0, then A£.G ( Z \ £) is a dif­

ferentiable Junction on K that extends the function ^_Zl) (£-^)(7 ^xIM.l ) -

PROOF. From Lemma (4.6) on p .145 it easily follows that


extends to a differentiable function h o n K and h extends 77— .
({-*1 )•••(€-**)U~-
So, it suffices to show that h = A£.G (£ ',£ ). If |£ — Z{\ > 0 for

i = 1 , A; + 1 and G { Z ' , ( ) = (<- n )..^/!S)(g- , t+l), we already have


. h ( ( ) = G( Z' , £) . So, suppose that

G (Z\i) = — [ du
2 %i J-,(j,z\d) (u - z x) ...(u - 2fc+1) (u - 0

for some j•> = 1,...,


’ ’ k + 1. We have 7(u-*l
----- -----------
}...(«-**+1r) = h l\ u ) ’, since / z;»
for u E 7 (j, Z', d). So, G{ Z' , £ ) = h ( ( ) as h is differentiable and (
must be in the sphere with 7 (j, Z ' , d ) as the border. |

The following lemma is needed in deciding if till zeros of / in K


have been found.

LEMMA 3.8. F o r a n y Z = ( z x, ..., z ^ th e r e e x i s t j, r a t i o n a l c o m p le x


n u m b e r u , a n d r a t i o n a l n u m b e r (3 s u c h t h a t i f Z i € V f o r i = 1,..., k , a n d
\ ( . G ( Z , £ ) i s a d i f f e r e n ti a b le f u n c t i o n o n K t h a t e x te n d s fc"*i& -;*)>
t h e n j = 0 o r j = 1, j = 0 i f m ( \ £ . G { Z , Q , K ) > 0 , a n d j = 1 i f
m ( \ £ .G ( Z ,( ) , K) = 0. I n th e l a t t e r c a s e , u E K", j3 > 0, and

m (\(.G (Z,(),B)>\G (Z,u)\+f).

Further, j, u, / 3 are primitive recursive functions of Z { H { k ) ) and k for


some primitive recursive H.

PROOF. Notice that for ( E K C\ B u , \ G ( Z , t ) \ > > 0. De­


note 131 = and let No be a integer > 8//9'. Let H (k) — F (k, No),
where F is the primitive recursive function in the last lemma. Choose
a l / H (k) approximation {ui, ...,Up} to K". For j = 1, ...,p, decide if

R ep ro d u ced with p erm ission of the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 244

\ G ( Z ( H ( k ) ) , uj ) \ < \ t 3 ' o T > \ P ' . l i \ G { Z { H { k ) ) , uy)| > \ 0 ' for all


j = 1, p, then by the last lemma |G ( Z , uy)| > \ 0 ' for all j and there­
fore \ G( Z, f ) \ > \ 0 ' for all £ 6 K". So, m ( \ £ . G ( Z , £ ) , K ) > 0 and it is
enough to set j = 0. Otherwise, for some j , \ G( Z (H (A:)), Uj)\ < \0 '.
Then \G(Z,Uj)\ < |/3' and it is enough to set j = 1, u = uy, and

0 = ;/?'• ■

The following three lemmas play the roles of Lemma 3.2 and 3.3 as
in the proof of the fundamental theorem of algebra.

LEMMA 3 .9 . There is a primitive recursive function F such that


if r, S, e , 0 are positive rational numbers such that r > 26, w is a ra­
tional complex number, Z = ( z i,...,Zk) is a sequence, then there ex­
ists a positive rational number s, depending prim itive recursively on
r, 6 , e, 0 , w, k, and Z (F (r, 6 , e, 0 , w, k)), and a positive rational number
c,depending prim itive recursively on r, 6 , £, 0, w, k, such that if Zi € V,
i = 1,..., k, \ ( . G ( Z , £ ) is a differentiable function on K that extends

U -J )% -z h)> S ~ S c ( w , r ) C K , T = S c { w , r - 26), and ||A £.G (Z ,0H T ^


0, then 6 —e < s < 6 and m ( \ ( . G ( Z , £ ) , T ( w , s ) ) > c. Further, F
can be replaced by any prim itive recursive F' such that F' > F for any
values of the arguments.

PROOF. Similar to that of Lemma 3.2, where recall that an upper


bound of ||A £.G (£,£)|| can be estimated independent of Z (except k )
and the modulus of continuity of A£. G( Z, £) on K can be determined
from k. |

LEMMA 3 .1 0 . For any sequence Z = ( z i,..., z*), rational complex


number z, and rational number 0 > 0 , there exist rational complex
number z', and rational number 0', t \ such that if Zi € V , i = 1,..., k,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 245

A£ . G ( Z , ( ) is a differentiable function on K that extends ^_Zi^ ^ _ Zky


z E K, a n d m { \ ( . G { Z , ( ) , B ) > \ G( Z, z ) \ + 0, then z' G K', t' < d,
('eS{z',t'),0 ' < 0 , and

m ( \ t . G ( Z t(),r(z',t')) > \G(Z,t')\+0'.

Further, z', £',0', t‘ are prim itive recursive functions of £, 0, and Z ( F (k)),
for some prim itive recursive F , and F can be replaced by any F'vrith
F' > F.

P r o o f . (i)W e h a v e m (A {.G (Z ,{),.K T iB M) > s& .


Let {ui, ...,ua/-} be a d / 2 approximation to K ' . For each j = 1 , N,
Apply the last lemma with r = p ( u j , B ) , 8 = d, e = d/2, 0 = 0 O,
w = Uj, we obtain tj G {d/2, d ) , depending on k, j , Z { F{ k) ) primitive
recursively for some primitive recursive F, and c3 > 0, depending on
k , j primitive recursively, such that m ( \( .G ( Z, £) , T (uj, , tj)) > Cj.
Let

7 = i m in {/So, e x , . . . , cat},

7 is a primitive recursive function of k. Choose a u' ( 7 / 4 ) approx­


imation to K', where u>' is the modulus of continuity of
A£ . G ( Z , £ ) on K . For each i = 1, I decide if \G{Z, Vi)\ > 7 /2 or < 7 .
For this, by Lemma 3.6, we need only Z ( H (k, 7 )) for some primitive

recursive H. Replace F by max{jET, F1}, we can use Z { F (k )) instead.


If \G{Z, Vi)\ > 7 /2 for all i, it would follow that \ G ( Z , ( ) \ > 7 / 4 on K,
which contradicts the assumptions on A{.(7(Z, { ) . So, we can choose
Vi such that |G { Z , Uj)| < 7 . vt- G S (tty, tj) for some j = 1 ,..., N. Then

let z' = Uj, t' = tj, = V{, 0' = Cj/2. Clearly, these are primitive
functions of k, Z (F { k )) for some primitive recursive F. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 246

LEMMA 3 .1 1 . There exist primitive recursive functions E ,T such


that fo r any sequence Z = (z i, ..., z*), rational complex number z,£,
and positive rational number 0 , r, t, there exist rational complex num­
ber z',£', and rational number 0 ',r', such that if Z{ G V, i = 1,..., k,
A( . G ( Z , £ ) is a differentiable function on K that extends
0 < t < r, £ G S (z, r) C K", 0 > 0, and

m ( H - G ( Z , ( ) , r ( z , r ) ) > \G(Z, £)\ + /?,

thenO < r ' < t / 2, 0 < 0 ' < 0 /4 , (' G S ( z '.r ') , Sc(z' , r' ) C S ( z , r ) , |G(Z,£')I <
0 /4 , and

0' > E (k, 0, t ) > 0, r ' > T { k, 0, t ) > 0.

m ( A ^ ( ^ 0 , r(z',r')) > | G ( Z , O I + ^ .

Further, z 1, £',0', r' are primitive recursive functions of z , £ , 0 , r , t , k ,


and Z (F ( k , 0 , t ) ) , for some primitive recursive F, and F can be re­
placed by any F' with F' > F.

PROOF. From the modulus of continuity of A£ . G( Z, £) on K , 0,


and t we can choose a < min { t / 2 , d} such that

m ( \ ( . G ( Z , Z ) , {u 6 C : r - 2a < |u - z| < r}) > 0 / 2 .

Let {u'l } ..., u'N} be a a /8 approximation to K " . Notice that for any u G
K", S c ( u , 2 d ) C K . So, for each j = 1, ...,n we can choose a rational
complex number Uj G K such that Uj —u' < a / 8 . Then
is an a /4 approximation to points in K " consisting o f rational complex
numbers in K* = {tx G K : p (u, B ) > d}. For each j = 1,..., N, we can
apply Lemma 3.9 with r — p (u,-, B) > d,S = a / 3 , e = a / 3 —a /4 , 0 =
0o, where 0o is as in the proof of the last lemma, and w = Uj to obtain

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 247

tj £ (a /4 , a /3 ) and Cj > 0 such that m ( \ £ . G ( Z , £ ) , T (uj , t j )) > c,-.


Let

7 = im in { /3 /2 ,c 1,...,c Ar}.

Then 7 is a function of k , 0 , t , a n d tj is a function of k , 0 , t , Z { F { k , 0 , t ))


for some primitive recursive F. In the same way we can choose an
a / (7 / 4 ) approximation to all points in K" consisting of
rational complex numbers in K*. For each Vi £ S c ( z , r ) decide if
\G{Z,vi)\ > 7 /2 or < 7 . For this we need the approximation Z ( H (k, 7 ))
for some H, which is Z ( H (k, 0, t , )) for some H. Replace F by max {F, H} ,
we can use Z ( F ( k , 0 , t ) ) instead. Similarly, there must be one v, £
S c ( z , r ) such that \G{Z,Vi)\ < 7 and then we must have Vi 6 S ( z , r — | a ) .
Choose Uj such that V{ 6 S (Uj, a / 4 ) , then Uj 6 S ( z , r — a ). So we
let z‘ = Uj, = Vi,r' = tj, and 0' = m in {cj/2 , >3/4}. r' > a /4 and

0' ^ 7 /4 . Both a and 7 are primitive recursive functions of k , 0 , t. So,


E , T can be found. |

Finally, we can construct the proof of Theorem (5.11) on p .157. We


will construct, for each k > 0, a sequence Zk = ( z \ , ..., Zk) of complex
numbers and a number J (k) = 0 or 1 such that if J (k) = 1 then
Zj £ V, i = 1, . . . , k and A£.G (£ * ,£ ) is a differentiable function on K
that extends )j and further J ( k + 1) = 1 if and only if

m (At . G ( Zk, t ) , B ) > m (A( . G (Zk, ( ) , K ) = 0.

Each Zk will be the limit of a sequence Zk (m) constructed by repeat­


edly applying Lemma 3.11, and for that we must construct sequences
0 (m ), £ (to) , and r (to) , simultaneously. So, we will construct a dou­
ble sequence Z (k, to) = (zl (to) ,..., zk ( m ), 0 k (to) , & (to) , rk (m)) to­
gether with J (k ). Here we explain only how to compute J (k + 1) and
Z {k + 1, to + 1 ), using j ( k ), Z (k + l , m ) , Z ( k , m + 1 ), and Z (k, Q),

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 248

where Q is an expression which may contain Z (k, Q') for some Q'
again but contains no contexts of the form Z (k + 1 ,...). Then, as in
the proof of the fundamental theorem of algebra, we can see the recur­
sion can be reduced to primitive recursion. We will denote Zk (m) =
{z\ (m ),..., zk (m)).
J ( k + 1) is computed according to Lemma 3.8. When J ( k + 1) =
0, Z ( k + l , m ) will be irrelevant for ail to. When J (k + 1) = 1,
given Z ( k , m + 1 ), to compute Z(fc + l,m + l) it suffices to com­
pute zk+i (m -I-1 ), (k+i (m + 1), 0k+i (to + 1 ), and r k+ 1 (to + 1). For
m — 0, we apply Lemma 3.8 and 3.10 to obtain <' there as
Zk+i (1 ), 6 +1 (1 ), Pk+i (1 ), and r k+ 1 (1), where an approximation of
the form Zk { H (fc)) is used. For m > 0, we use the same trick as we
used in constructing zeros of polynomials, that is, construct T ^ ,E ^
such that

r ‘ = rt + l( l ) , S f = A + i ( 1 ),

T ‘ +1 = r ( * , f £ , : r ‘ ) ,

£ £ +I = £ ( M ‘ , r ‘; ) ,

where E , T are the functions in Lemma 3.11. T f ,E* are functions of


k and Zk { H{ k ) ) , so E ^ ,T ^ are functions of fc,m, and Zk ( H ( k ) ) .

Further, T* = r 4+ 1 ( l ) < d < J, £ { < T * +1 < J I J ,


and by Lemma 3.8 and 3.11, so Tj^ < 2 ~m. The
construction up to stage m should have guaranteed that /3k+i (m ) >
E £ and r k+i (m ) > T£. We apply Lemma 3.11 with E ^, as (3,t
there and zk+i (rn) , 6 + 1 (m ), and rt +1 (m ) as z , ( , r there to obtain
Zk+i (m + 1 ), 6 +1 (m, -h 1) , 0 k+l (m + 1 ) , and rfc+i (m + 1). Then, be­
sides Z(ifc + l , m ) , which provides z , £ , r there, only an approxima­
tion of the form Zk (F (fc, m, Zk (H (A)))) is used. So, Z (fc + l ,m + 1)

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. COMPLEX ANALYSIS 249

can be computed in the way stated above. Further, by the definition,


r*+i (m + 1 ) > T *+1 and 0 k+l (m + 1) > £ * +1.
After Z, J have been constructed, we can use an induction on k to
prove the formula ‘If J (k) = 1 then (i) Z{ 6 V for i = 1,..., k, and (ii)
\ ( .G (Zk, £) is a differentiable function on K that extends
The component (i) in the formula is £° and by Lemma 3.5, under the
assumption (i), (ii) is equivalent to a 11° formula. So, the formula is
equivalent to a £ ° formula. As before, another induction with assump­
tion is needed to prove the inductive step: Assuming (i) and (ii), by
the construction of J ( k + 1) and Z ( k + l,m ) and the lemmas above,
we can prove by induction on m that if J ( k + 1) = 1 then r *.+1 ( 1 ) < | ,

ffc+1 (m + 1 ) < < 2 "m, and

Sc(z*+i (m + 1) ,r*+i (m + 1)) C S (zk+i (m ) ,r k+l ( m ) ) .

So (zk+i (m ))“ =1 is a complex number. Similarly, we have E^+i <

4 and \ G( Zk,Zk+i (m + 1))| < £?*. So, G ( Z k)zk+1) = 0 .


And then it follows that zk+i 6 K ' and (i) holds for fc + 1. Finally by
Lemma 3.7, (ii) holds for fc + 1 .
Following the argument in the second paragraph on p.158 we have
that J (N ) = 1 leads to a contradiction for some N . So, there exists
n such that J (n) = 1 and J (n + 1) = 0. Then \£ .G ( Z n, £) is a differen­
tiable function on K that extends and m (A(.G (Zn, ( ) , K ) >
0. The conclusion of Theorem (5.11) then follows.
The proof of Lemma (5.12) on p .159 uses an induction on k to prove

the statement ‘if k < n, then ll/llsefro*),*) > O’- This is available, since
the statement is £ °. Theorem (5.13) to Corollary (5.19) on p.159-162
pose no more difficulties.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. COMPLEX ANALYSIS 250

3 .6 . S in g u la rities and P ica rd ’s T h eo rem . This section, though


a bit long, poses no real difficulties. Lemma (6.1) to Lemma (6.4) on
p. 163-166 are straightforward. The proof of Lemma (6.5) on p. 166—
167 implicitly uses an induction on k to show ‘if k < v, then either
a_v+t = 0 for all z < k, or o_v+i ^ 0 for some i < k \ which is a £?
statement. Lemma ( 6 . 6 ) to Lemma (6.13) on p. 167-172 pose no real
difficulties, while a II? induction is used in the proof of Lemma (6.14)
fo prove an inequality. Lemma (6.16) to Theorem (6.19) on p. 173-177
do not involve any inductions.
In the proof of Corollary (6.20) on p. 177-179, an inductive con­
struction is used to construct a sequence Sn (t), for an arbitrary £, such
that £n+i (t ) = $ (£n ( t ) , J) , where $ (a , is continuous for a 6 R +.
This is available in SC for it is clear that to compute £„+i (t) (m ),
we need only Sn (t) ( N ) for some N depending on m and the modulus
of continuity of $ (a , on a bounded interval in R + that contains
Sn ( t) , for example [0, |6n (£ )(l)| + 2 ] , and N is actually a primitive
recursive function of m and Sn (t) (1 ). Such a recursion pattern can be
reduced to primitive recursions. The proof still uses another induc­
tion on rz to prove the inequality (6.20.4) on p .178. The statement,
that is, ‘|/ ( z ) | < £n+i (jn ( / .T ( j ) ) ) for all z € Sc (z*, involves a
quantifications over complex numbers, but it is equivalent to c|/ ( * ) | <
8 n+i (m ( / , T ( | ) ) ) for all rational complex number z 6 s ( z n , i i ) , b y
the continuity of / . The latter statement is 11° for ‘z 6 5 (zn, ^ ’, that

is ‘|z - Zn| < ^ ’ is £5 and ‘|/ ( * ) | < 8 n+l (m ( / ,T ( j ) ) ) ’ is II?. So the


induction is available in SC .
Finally, Theorem (6.21) and (6.22) on p. 179-180 pose no more
difficulties.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 251

4. Integration

4.1. Integration Spaces. Through, this section we consider only


sets with inequality relations and functions that respect the inequality
relations. The definition of integration spaces needs some elaboration,
for the definition on p.217 of B&B apparently refers to the set of all
partial functions from a set X to R.. As we have noted in Section 5
above, this cannot be defined in SC. We will define integration in a
slightly different way. If V = {Dj : j € J } is a family of parameterized
subsets of X that is closed under finite intersections and if for k < n,

fk 6 F ( V , R ) , then £fc=o /* , U L o f k , V?=0/< 6 F ( V , R ) are partial


functions with the domain n£_0drrm (/*) and are defined in the obvious
way.

DEFINITION 4 .1 . An integration apace ( X , U , L , I ) consists of a


set X , a family V = {Dj : j 6 J } o f parameterized subsets o f X that
is closed under finite intersections, a subset L of the set T (V , R) of

partial functions with domains in V , and a function I : L —» R such


that (where f l\g means m in { /,p } ; and we assume that f A n has the
same domain as f )
(1) If / . 9 i e L f o r i < n , and a 6 R, then a f , Ti?=o9 i, V"=0&, |/ |,
and / A 1 all belong to L, and I ( a f ) = a l ( / ) , I (£ ? =0 &) = £"=o I (p,);
(2) If f € L and (f n) is a sequence of non-negative functions in

L such that (/n ) converges and E£Lo I (fn) < 1 ( f ) , then there

exists x 6 d m n ( f ) D (D£L0 d m n ( f n)) such that ££Lo f n (x) converges

and E ^ lo/nO c) < / ( * ) ;


(3) There exists p € L such that I (p) = 1;
(4) For each f 6 L, limn_ 00 1 ( f A n ) = I ( / ) a n dlim n_ O0 I ( | / | A n~l ) =

0.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 252

The appearance of | / | in (1) is redundant for | / | = ( / V 0/ ) V


( ( —! ) / V Of). There are two new conditions in this definition. First,

the domains of the partial functions are restricted to a family of param­


eterized subsets. This is necessary because some notions, for instance
measurability, are defined by apparently quantifying over all integrable
functions in L. Without such a restriction, these are not definable in
SC . On the other hand, it will be clear that the restriction does not
damage the development of the integration theory. In particular, the
complete extension of a integration space can still be constructed. Sec­
ond, it is not sufficient to assume that f , g 6 L implies / + g € L
and / V g & L , for there is no available induction to infer £ ”=o gi 6 L
and V"_0<fc from gi € L. So these has to be added as assumptions on
integration spaces in (1). We will see that these new conditions are
always easily verified in the concrete examples of integration spaces.
The basic properties of an integration space and Lemma (1.2) to
Proposition (1.6) on p. 217-218 are simple. The definition of a pos­
itive measure p also needs the stronger condition that m(£"=0 fi) =

£?=o M(/<)• Lemma (1.8) and (1.9) on p.219 are straightforward.


We consider Theorem (1.10), which says that for a locally compact
metric space X , and a positive measure p on C ( X ) , (X , C ( X ) , p)
is an integration space. Notice that all functions in C (X ) are total
functions on X , so we can omit the family V. The proof on p. 220-
221 uses inductive constructions to construct a sequence of functions
in C ( X ) in order to prove the following claim:

(Claim) If / , f n are test functions in C (X ), / n are non­


negative, I f n d p exists, and Z%L0 f f„d(i < f fd p ,
then there exists a sequence (gi) of test functions such
that for / > 0 the diameter of a compact support of gi is

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 253

less than l / l and

(3 1 ) 2 J f n 9 i - 9 i d f i < J fg i...g id fi.


71=0

Here we will show how to replace it by an inductive construction of


a sequence of natural numbers. Suppose that K is a compact support
of / . By Proposition (6.15) on p .119, for each / > 0 there exists non­
negative functions glQ, ..., glN^ , each with a compact support of diameter

less than l / l , such that Yli^o 9i < 1 and for x E K , Yli'Jo 9i ix ) — 1-


We will construct a sequence k (I) of natural numbers such that k (I) <
N (/) and

5 2 J ^ 9k{\)—9k{f,)dy- < J /S t(i)—9k{i)dfi,

and then let gi be 9 k(iy First notice that the formula

n n i
Vz < n (a* E R A bi E R) A m > 0 A 5 3 a« ^ ------
t= 0 i= o m

—* 3z < rz3m' > 0 ^a, < bi — ^

is provable in SC. So there are terms operations H, M such that

n n j
Vz < n (a,- E R A b{ E R) A m > 0 A 5 3 a* ^ ------
<=o t=0 m

/ H ( n , a , b , m ) < n A M [n.a.b.m) > OA >

^ Aa/f(n,a,fc,m) < ~ 6,m) >

Let G ( m ) = n<Ti° 9 (m).> where we take g'k = glN^ for k > N (z). So,
G((z‘i , ...,z"j)) is supposed to be g^.-.g1^. By the assumptions, there ex­
ists mo > 0 such that / f ndfi < f fdfi — Construct sequences

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 254

k(l), k(l ) (supposed to be ( k ( 1 ) , k(l))), and m ( l ) such, that

m ( l) = M ( n ( 1 ) , fn9 i f a *»• J m 0^ ,

k ( 1 ) = H ^ N ( 1 ), Ax'. 5 3 J fn 9 }dfi, XL J f g\ dp , m 0j ,

* ( 1) = (*(!)>,

m ( l + 1) = M ( n (I + 1 ), Ax'. 5 3 / fnG (k ( l )) gl+'dfi, XL f f G (£ ( /)) g[+1 dfi, m ( l )


V n —o J l

k ( l + l ) = H ^ N ( l + l ) , X L ± J f nG ( k ( l ) ) g l + ' d , , XL J f G ( k ( l j ) gf ' df i , m ( /)j

k ( l + 1) = k ( l ) * ( k ( l + 1)).

Then we have

£ ' J f . a (*w) to < J f a (k (0) to - ^ -


£ j fnG (* (i + 1)) d p < j fG (E(i + 1)) ip. - ^ T T y
n=0

for from the antecedent it follows that


N(I+1) *
E E / /•<*(*(0) s!+1^ = E / /-o (*(0) to
«=1 n=0 n=0

W('+l) /• ,_ N 1
= e J fG (m )^to -W )
(cf. the proof of Lemma (1.8) on p.219). Then by the definition of
jb,m, the consequent follows. An induction on a 11° formula then gives

£ f f nG ( k ( l ) ) t o < [ f G { m ) t o-
n=0 J

This proves the claim.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 255
The rest of proof on p. 220-221 axe straightforward given the claim.
In particular, the new conditions on finite sums £?=o g, and finite max­
imum VJL0# are trivial. The discussions following Theorem (1.10) on
p. 221-222, including the definition of Lebesgue measure on R, are also
simple.

4 .2. C o m p lete E x te n sio n o f an In tegral. Given an integration


space ( X, V, L, / ) , we will construct the complete extension of it. First
define an index set T:

(/„ ) 6 r = V n (/„ 6 L) A 3a [ a = £ / ( | / „ | ) ) .
V n=0 /

Then the domains of integrable functions will be in the family V* =


{ p \ f nj : (fn) 6 r } of subsets of X indexed by T:

1 € D Un) ~ x e n “=orfmn (In) A 3a (a = f ; |f n (x )|) .


V n=0 /

Finally, we can define the set Lj of integrable functions with respect


to (X ,£ > ,Z ,/):

« ( /.) , / ) e £ i )

= (C/n)i / ) € F {V ,R)A

3 (Sn) e r ( d 'm C D-i m A Vx 6 /)(•„, ( / ( * ) = (* )) ) ■

So, integrable functions are members of the set R) of partial


functions with domains in the family V*. We willsimply call/ an
integrable function and D (/n) the domain of / and denote as d m n ( f ) .
Among the witnesses for / € L\ isa sequence (gn) 6 T. This will
be called a representation sequence of / . This definition isslightly
different from the definition on p.222 in that domains of integrable
functions are restricted to subsets in the family V *. Also notice that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 256
any (/„ ) 6 r is a representation of an / G L\ with the domain

where / (x) = fn (x ) f°r x € £(/„)• The appearance of (y„) in the


definition is to make the definition more general, for we still want to
consider ((f n) , / ) as an integrable function if for some x G D'fn) we
have / (x) ^ /„ ( x ) . This is to reflect the classical idea that for
an integrable function its values on a set of measure zero are irrelevant.
For / integrable and (gn) a representation of / , 1 (gn) con­
verges, so there exists a function I* : L\ —* R such that / * ( / ) =
/(</„). Notice that this depends on the fact that if (g'n) is a rep­
resentation of / ' in L \ such that / = £ l that is,

/ ( i ) = / ' ( * ) for * 6 D'w , then £ £ * / ( * ) = ' « ) ■ The proof


on p.223 actually proves this. For / G L, let (/„ ) be the sequence
such that / 0 = / , /„ = 0 • / for n > 0. Then for some f* we have
( ( /n ) ./* ) 6 d mn( f * ) = d m n ( f ) , and /* (x) = / ( x ) for
x G d mn( f * ) . Then /* € L\ with (/„) as a representation of /* , and
further I ( / ) = /* ( / ) . That means there is a natural way of transform­

ing a function / 6 I into an /* G Li. So, we will simply write /* and


/* as / and / . /* is called the complete extension of I.
It can be shown that V* is closed under finite intersections and L\ is
closed under finite linear combinations and finite maximum operations.
Further / G L\ implies / A 1 G Lx. (cf. p.224) Here we check only
the maximum operation. So suppose that /; G L\ and is a

representation of ( /,) for i = 1, ...,n . Let jo = /i° V ... V and

m
( m \ / m —X
*7m E / f v- v E £ - E/f
k=0 k= 0 / \k = 0

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 257

for m > 0. We have gm e L, \gm\ < | / ” l + - + |/™l on d mn ( g m).


Then it is easy to see that the sequence

is a representation of f i V ... V /„.


A subset of X is a full set if it contains a domain of an integrable
function. Assertions about an arbitrary full set are schematic. How­
ever, we can make the convention that when we quantify over full sets
we always mean a quantification over domains of integrable functions,
that is, sets in the family V*. For example, ‘/ = g on a full set’ is to
mean ‘there exists h 6 L\ such that / ( x ) = <7 ( 1 ) for x 6 d m n (h )’.
Then Proposition (2.5) to Lemma (2.10) are all straightforward, where
in Proposition (2.9), the condition / 6 T (AT) should be replaced by
f 6 f { W , R).
Now we define a new equality relation between members of L\.
From now on, / = g will mean ‘there exists (/„ ) 6 T such that D * ^ C
dmn ( / ) (1 dmn( g ) and f ( x ) = g (x) for x 6 that is, / and g are
equal on a full set. This is weaker than the old equality relation =&t
of the set L \. Linear combinations and the function I* all respect this
new equality relation. From now on, we will consider L\ as a set with
this new equality relation. That means, we expect that the functions
and concepts defined on L\ will respect this new equality relation. For
example, if A is any one of the relations = , < , > , < , or > , then f& g
is to mean / (x) A g (x) for all x in a full set. Similarly, ‘/ is bounded’
is to mean | / | < c on a full set for some c > 0. The norm of / 6 L\
is ll/d x = / ( | / | ) . Further, for F a full set, sometimes we construct
a function / by defining / ( x ) for all x £ F and claim that / with
the domain F is integrable, by which we actually mean that for some

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. INTEGRATION 258

Dlfn) Q ■f’j ((fn) i / ) £ L\. The choice of ( / n) does not matter in view
of the new equality for Lx.
Proposition (2.12) is trivial. The proofs of Lemma (2.14) and The­
orem (2.15) on p. 228-229 are also straightforward. Corollaries (2.16)
and (2.17) on p.230 are also simple, so L is dense in Lx and Lx is
complete for the norm ||-||r Finally, the proof of Theorem (2.18) is for-
malizable straightforwardly. Given the results we already have, it uses
only some properties of absolute convergent series of real numbers. So
( X , V * , L XiI*) is an integration space which cannot be extended fur­
ther. It is called the completely extended integration space.

4 .3 . In te g r a b le S e ts. Let (X , V * , Lx, I*) be a completely extended


integration space. An integrable set A = (A 1, A0) is a complemented
set in X such that it has a characteristic function \ a in 7?* that is inte­

grable, that is, x a € Lx and for z £ dmn ( x a )> X (x ) = 1 or X (z ) = 0>


and z £ A1 if x (x) = 1 and z £ A0 if x ( x ) = 0. Clearly x a is unique
up to the equality for Lx. So we call xa the integrable characteristic
function of A and denote fi (A) = I (xa)- /*(A) is called the measure

of A. Any function i i such that x ( I ) = 1 ° r x ( z ) = 0 for all


x G dmn (x x ) is the integrable characteristic function for the integrable
set

({ z G d m n ( x ) : x(*) = 1}» {* £ d ™ n ( x ) ' x ( * ) = 0 } ).

So we make the convention that when we use a quantification over all


integrable sets we mean a quantification over all integrable character­
istic functions. T he basic properties of integrable sets, Lemma (3.2) to
Proposition (3.7) on p. 232-233 are then trivial.
Notice that Proposition (3.5) can be generalized to finite sequences
of integrable sets, that is, V"_0An and AJLqA^ are integrable if A,,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 259

i < n, are integrable. For a infinite sequence (An) of integrable sets,


A = VJJL0A„ is integrable if and only if there exists an integrable char­
acteristic function xa such that

d m n (xx) Q A 1 U A 0

= (n~ o (4 , u < ) n u“ 0^i) u n“ X


and for x 6 d m n (xx), if x ( x ) = 1 tben * € A1, and if x ( x ) = 0 then
i € A0. So, VJJL0An is integrable if and only if for some integrable

function Xa , Xa = VJJL0xx« on some full set. The similar is true for


intersections. Then Proposition (3.8) to Proposition (3.11) on p. 234-
235 are straightforward.

4.4. P rofiles. The notion of profile, Definition (4.1) on p.236, is


simple. Notice that it involves a set, so lemmas about profiles are
schematic. The profile ( / , / ) associated with an integration space
(X, T)*,Li, I) and a function / 6 L\ is also straightforward. The notion
‘[a, 6] has a profile lower them e relative to the profile (£, A)’, which is
denoted by [a, 6] ■< e, and the notion ‘[a, 6] has arbitrarily low profile
relative to (£, A)’ are also easy to formalize.
The proofs of Lemma (4.3) to Lemma (4.6) on p. 237-239 are easily
formalizable. We consider the proof of Lemma (4.7) on p. 239-240
which contains some inductive constructions. We can assume that a, b
in the assumptions are rational numbers, for [a, 6] is in the interior of
the interval J and we can enlarge it to rational endpoints. We can still
assume that the end points of J are rational for we can replace it with
a smaller interval. Similarly we assume that t is rational. Since (£, A)
is a profile for the interval J , there exists H such that for Xo, z i 6 J,
xQ < xx we have H ( x 0,x i) e £ , f f ( x 0,x i) ( x ) = 0 for x 6 J and
x < xq, and H (xq ,xi) (x) = 1 for x € J and x > x\ . Then, for any

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 260

[x, y] in the interior of J and 6 > 0, [x, y] C 8 is equivalent to

n / i q < q' < x f \ r > r' > yA '


3 q , q , r , r 6<Q>
\ A { H { q ,j ) ) - K { H { T ',r ) ) < 6 )

which is a formula. When x , y , 6 are rational, we will use [x,y] < p


6 to mean that p = (q,q',r,r') is the sequence of the witnesses for
[x,y] <C S as in the above formula. This is a quantifier free formula.
From the proof of Lemma (4.4) on p. 237-238 we see that we can
choose x and £i, £i there to be rationed numbers, and then from Lemma
(4.6)on p.238 it follows that there exists X , P i , P 2 , A , B such that for
rational numbers X i,x 2, integer M > 0, and a sequence number p, if
®i < x2, [xi, x2] is in the interior of J and [xx, x 2j < p (Af + 1) e, then

for X = X ( x i,x 2,p, Af), Pi = P i ( x u x 2 , p , M ) , P2 = P 2 ( i i , x 2,p, M),


A = i4 (x x,x 2,p, M) , B = 5 ( x i , x 2,p, M) , X is a rational number,
Pi, P2 are sequences of rational numbers, and A , B are integers such
that X e ( x i,x 2), m a x {X - x x, x 2 - X } < \ { x 2 - x i ) , [xi,x]
( A + l ) e , [x ,x 2] (B + l ) s , and A + B = Af.
With these, we now construct the sequence required in the proof.
By the assumption of the lemma there exist M, P0 such that [a, b] <Cp,

(M + 1) e. For any sequence e of 0 and 1 we construct rational numbers


ae, be, integer Me, and a sequence number Pe such that

oe = a , b e = b, Pe = PQ, Me = M , for e = ( ),

®e«(0) = &e, &e*(0) = X ( t t e > P t, Mg) ,

P s«(o) — Pi (®e> bg, P s1Mg ) , Mg,(p} = A (ae, be, P ., M g),

®e*(l) = X {o>e, be, P 1, Mg) , = be,

P s»(i) = Pi (®ei be, P M g ) , Mg,(1 ) — B (ae, bg,Pe, Mg) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 261

Then a simple induction, shows that the conditions (i) — (vi) on p.239
hold when / e = [ae,6e] and with (iii) replaced by Ie <Cp, (Me + l) e .
(vii) also follows by an induction, for it is equivalent to the 11° formula
‘for any rational number x £ [o, b] and r > 0, there exists e, |e| =
n, such that x £ [ae, 6e]’, where the second quantifier is a bounded
quantifier.
The construction of the sequence s* = ( i nifc)^L0, 1 < fc < Af, on
p.240 is simple, for they are sequences of rational numbers and the
conditions are quantifier free. Then the rest of the proof is straightfor­
ward.

The proof of Theorem (4.8) on p.241 uses no further inductive con­


structions. The notion of smoothness and the proof of Theorem (4.10)
on p. 241-242 are all straightforward.
For / € L\, ( / > £) denotes the complemented set

({x 6 dmn ( / ) : / (x) > *}, {x 6 dmn ( / ) : / (x) < *}).

Notations ( / > t), ( / < t), ( / < t ) are similar. Then the proofs of
Theorem (4.11) to Corollary (4.17) on p. 242-245 pose no difficulties,
where notice that the construction of sequences in the proof of Propo­
sition (4.13) does not use inductive constructions. They are defined
uniformly for an arbitrary n.

4 .5 . P o sitiv e M ea su res on R . Suppose that A is a dense subset

of [a, 5], a < b, a, b £ A, a is an increasing function on A, and /


is a continuous function on [a, 6]. To construct the Riemann-Stieltjes
integration of / it suffices to construct a sequence of partitions (Pn)>

Pn = (ao» —>xk») I z o = °> xln. = x? £ A for i = 0 ,..., such that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 262

meah{ Pn) —►0 and

s ( / , p») = E f W ) ( “ ( * ? « ) - “ ( * ? ) )
«=o
converges. Then it will follow from the continuity of / that the limit
is independent of the choice of the partitions. Here we construct a
sequence of partitions (Pn), Pn = (x", ...,x£n) such that x" = xjij*1 6 A
for all n ,m > 0 and i = 0 ,...,2n and meah{Pn) —» 0. Then the proofs
for the existence and uniqueness in the case of Riemann integrations
are still valid. This construction needs some illustration, for we need
that x" 6 A and there may be no inductions applicable to the formula
i? 6 A.
Here is a substitute. Since A is dense in [a, 6], it follows that there
exists a sequence (a*) of elements of A that is dense in [a, 6]. So,
there exists H such that for any x ,y ,m , if x ,y 6 [a, 6] and y — x >

( aT (6 “ a )> then

aff(*,Vlm) 6 ( x + i (y - x ) , x + ^ (y - x )) .

With H, we can construct Pn by constructing a finite sequence of in­


dices of (a*) We may assume that do = a, ai = b. Then p° = (0,1).

Suppose that pn = (po>—iPa") has been constructed. Then p n+ 1 =


(Po+1.-.P ? « + \) is “ follows: p t f 1 = p?, p j£ \ = H (ap?, Op?+i, n ). By
some inductions it is easy to show that p" = pjm” , dp? < ®p?+li and
Op«+i —Op« < . So, Pn = (op*,..., aj,nn) is the required partition.
This completes the construction of Riemann-Stieltjes integrations.
The proofs of Lemma (5.2) to Theorem (5.5) on p. 247-251 are then
straightforward. No novel inductive constructions are needed.

4 .6 . A p p ro x im a tio n b y C o m p a ct S e ts. The definitions and


Propositions (6.2), (6.3) on p. 252-254 are simple. The proof of Lemma

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
II
! 4. INTEGRATION 263

(6.5) on p. 254-256, though a bit long, is actually straightforward. No


i

I inductive constructions are sued. Lemma (6.6) on p. 257 is also simple.


We don’t know if Theorem (6.7) is formalizable in SC . Its proof
on p. 257-258 uses Lemma (6.5) repeatedly to construct a sequence of
integrable sets. We don’t know if that can be unraveled. But we have

two weaker versions of the theorem. One follows from Lemma (6.5)
directly:

THEOREM 4 .1 . If p is a positive measure on a locally compact

\ space X , and A is an integrable set vritk fi(A) > 0, then fo r each


e > 0 there exists a strongly integrable set K such that p ( K —A) 4-
| p(A -K )<£.

] This is weaker in that it does not guarantee that K < A. On the

other hand, the following shows that we can obtain K < A if we want
i
only that K is compact and integrable. Remember the convention that
when A is a located subset of a locally compact space X , we simply
denote the complemented set (A , X — A) by A.

THEOREM 4 .2 . If p. is a positive measure on a locally compact


{ space X , and A is an integrable set with p. (A) > 0, then fo r each
1 t > 0 there exists a compact and integrable set K such that K < A and

\ n(,A -K )<e.
1

| PROOF. First, by the proof of Lemma (6.5) we know that there


i

exist a closed integrable set B = ( B l , B°) and a strongly integrable set


B\ such that B < A, p ( A ) — p ( B ) < e/2 , and B 1 C B\ . By Lemma
(6.3), for any x 6 X and any real numbers a, b with 0 < a < 6, there
exists r 6 (a, b) such that the complemented set 5 c ( i , r ) is strongly
integrable. So there exists Xi such that for x 6 X , and a,b, 0 <

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 264

a < b, X i { x t a>b) is an integrable characteristic function for a strongly


integrable set S c ( x , r ) for some r E (a, b) . For u a finite sequence of
points in X we denote

p (x, u) = min {p (x, («),-) : 1 < i < («)„} .

Then similarly there exists X 2 such that for u a finite sequence of points
in X and a,b, 0 < a < b, X i ( u t a >b) is an integrable characteristic
function for a strongly integrable set

S c (u ,r ) = {x 6 X : p (x ,u ) < r }

for some r € (a, b ).


Since B\ is compact, there exist N, U such that for I > 0, U (I)
is a sequence of N (I) points that constitute a /-1 approximation to
B\. Choose lo > e-1 and construct sequences (En) , (A/n) of natural
numbers such that

E\ = 4/0| M 1 = N( AEl ),

Ek+i = 4A/fc£?fc, Mk+i = N (AEk+i) .

If u is a sequence of points in X and H is a. sequence of positive integers,


let (u)H denote the subsequence of u obtained by selecting the entries
whose indices are in H . Denote Un = U (4En) . For each n we construct
a finite sequence Hn of indices of Un as follows. H q = () the empty
sequence. To compute Hk+i, we begin with the empty sequence. Then
for each i = 1,..., (Uk+1 )0 denote

x ' = Xl ( ( t / t + o . . (4 £ * +1) - 1 , ( 2 £ » „ ) - ■ ) A x s a

A‘= , X2 ((tfy )H j , (2JS,-)-1 , S f l ) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 265

x 'is integrable. Decide if /z (x') > ( i M k+i E k+i)~l or < (2Mfc+i £ ,Jfe+1)"1
and add i to Hk+\ in the former case. Clearly this can be reduced to a
primitive recursion.
Denote

= W O * ,.

Si = Xl ( W ) » , , (2E i) - ' , £ -■ ) , for j > 0,

9n = X B A A % J j , for n > 0.

Then gn+i = yn A / n+i- Recall that /,• is tin integrable characteris­


tic function for some strongly integrable set S c ( h j , a 3) for some a , €
((22?,-)- 1 , E ~l ^j . Let be the indices of Uk that are not in Hk. Since
Un+i is a ( 4 £ n+1) _1 approximation to B \ and X i ( W + x ) , , ( 4 £ „ + i ) * ' , (2 £ n+i) " ‘)
is a characteristic function for a strongly integrable set Sc ( ( Un+ i ) i , r)
for some r € ((4£?n+i ) - 1 , (2i?Tl+i ) -1j , and since gn < x b and B 1 C B\,
we have

ft. = ft. A v £ r ll’ X. ( W + i ) . , { 4 K +i ) _ 1 . (2 £ „ « )-■ )

< (ft. A /„ +i ) V ( j „ A V i£H ; 4 iX i ((K .+ i)j , ( 4 £ n + l) ‘ l , (2£„+1) _1) .

For i 6 Hn
' +l, by definition

/• (ft. A x i ((£f«+i)j, (4£„+I) - ‘ , (2 £ n+l) - 1) ) < (2 M ,+ lf:„+I) - 1 .

So,

- S « + l) </*(»« A VieHi+lXl ((C n+llj, (4 6 .+ 1 )'1 , (2£„+,) ~ l )

< 2 (2M„+1£ „ +1) ‘ 1

< (2£„+1) ' 1 < 2-Jft-3e.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. INTEGRATION 266

So lirnn_ 00 fi (gn) exists and x = Xs A Aylx/y is integrable and

V- ( x ) = fl
J im
100 fi (gn)

= H ( x b ) - i n ( x b ) - (i (gi)) - ( n ( g i ) - f i (g 2)) - -

^ M( x b ) — 2-3e —2 ~ s e —...

> m (x s ) - c/2.

So, m ( x a ) -A i( x ) < £•
Recall that /y is a characteristic function for a strongly integrable
set Sc(hj, ay) for some ay 6 ((22?y)- 1 , E j l J . Let

K 1 = B 1 D n j l 1 S c ( h j , a J) ,

= B 1 n n '.jS c C A j.a ,).

So x is a characteristic function for the complemented set A = ( A 1, A10)


for some A 0. Since A 1 C B 1 C A1, x is also an integrable character­
istic function for A A A . Replace A- by A A A we m ay assume that
A < A. It remains to show that A 1 is compact. A 1 is closed. Recall
that gn is a characteristic function for An = (A*, A °) for some A °. Let
n > 0. For each x 6 An+i by definition x i (x, (4An+1)- 1 , (2An+1)-1 ) is
a characteristic function for a strongly integrable set S c ( x , r ) for some
r £ ((4i?„+i) 1 ,(2A n+1) and

f i ( S c ( x , r ) A An) > (4Afn+1An+i ) _ l .

N oticeth atr < (2An+i ) _1 < atn+i and therefore S c (x ,r ) C Sc ( hn+ i , a n+ i ) .


Since gn + 1 = gn A XSc(fcn+llan+l), we also have

l i ( S c { x , r ) A A n+i) > (4A/n+1£ n+1)~l = E ~ l2.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 267

For any m > 0 we have

TO—1
( ^ i+ l ^n+l+m) (^n+l+i f^n+l+t+l)
«=0

< £ (2£„t l+ i) - < \ e £ , .


1=0 6

So fi (Sc (x, r) A K n+i+m) > for all m > 0. Therefore p (Sc (x, r) A K ) >
0. So there exists y E K l such that p ( x , y ) < r < On+1. Let Y be the

sequence of all such y for x € /in+i- For any z E K 1 we have z E Kn+i


and hence p ( z , /in+i) < otn+j, that is, p (z, x) < atn+i for some x E hn+1.
By the definition of Y, p (x, Y ) < On+i and therefore p (z, V) < 2an+1.
So Y is a 2 ctn+i approximation to K 1. Since a„+i -+ 0 as n —►oo, K 1

is totally bounded and hence compact. |

4 .7 . M easu rab le F u n ction s. The definition of measurability in­


volves quantifications over all integrable sets, so according to our con­

vention, these are translated into quantifications over integrable char­


acteristic functions. Given the completely extended integration space
‘/ is measurable’ is translated as ' / E .F(1?*,R) and
for any integrable characteristic function Xi and each e > 0, there

exists integrable characteristic function x i and <j € Li such that for

x E rfmn(xa) and X i { x ) = 1 we have x € d m n (x i) and XiC*) = 1>

I ((Xi - Xa)+) < £, and \ f - y| < e on {x E dm n(xa) : Xi (*) = !}•’


Then a measurable set is a complemented set with a characteristic func­
tion in V* that is measurable. It is easy to see that if two functions in
T (V*, R) are equal on a full set and one is measurable then the other
is also measurable. So, if a complemented set is measurable, its any
characteristic function in V* is also measurable.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 268

Lemma (7.2) is then simple. Notice that ‘there exists an integrable


set B ’ literally should be ‘there exists an integrable characteristic func­

tion x b 1 and B 1 should be {x 6 d m n (x fl): Xfl (x ) = 1}> though A


can still be an arbitrary complemented set and then the Lemma is a
schematic theorem in SC . Similar interpretations hold for all the the­
orems below. We will not repeat this comment, for the contexts can
always determine what should be interpreted as quantifications over
idtegrable characteristic functions and what can be arbitrary comple­

mented sets.
A simple function Y%=i akXk can be represented by a finite sequence
(Xi) —iXn) of integrable characteristic functions and a finite sequence
(aij.-.jOn) of real numbers. The proofs of Lemma (7.4) to Proposi­
tion (7.7) on p. 259-261 are simple. Further, Proposition (7.7) can
be generalized to finite sequences of measurable functions, that is, if

fi is measurable for i < n, then fi, n?=o fi, ^i=ofi, V?=o/i are
also measurable. Because, for any e > 0, there exist c,- > 1, Bi
with B} C A 1 and /z(A — B,) < e/n, and simple functions x* such
that |xi| < <k, < (ncx-.Cn)-1 e on B\. Then, for example,
inr=o /« - nr=o x.l < e o n B 1, where B = Bx A ... A Bn. We have
f i ( A — B) < t and it is also easy to show that rK*=oXi is integrable.
Consider Lemma (7.8) on p..261. We code the finite set { l , . . . , n }
by a sequence number (1, ...,n ). Then the detachable subsets of it are
the subsequences (i‘i , ..., z*), 1 < k < n, and z/ < z'j+x for 1 = 1,..., k —1 .
Two such subsets are distinct if they are distinct sequences. Then the
proof of the lemma is easily formalized. Proposition (7.9) and Corollary
(7.10) are also simple. The sequences in the proof of Theorem (7.11) on
p.263 are all defined uniformly. No inductive constructions are used.
Finally, Corollary (7.12) and Proposition (7.13) are also easy.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 269

4 .8 . C o n v erg en ce o f F unctions and In teg ra ls. The definitions


of ‘(/„ ) converges to / in measure (almost everywhere, almost uni­
formly)’ are formalizable directly, where the quantifications over all
integrable sets should be understood as quantifications over all inte­
grable characteristic functions. We have to assume that / 6 f (P*, R).
The proofs of Proposition (8.2) to Theorem (8.8) on p. 265-268 are
straightforward. No inductive constructions are used.
The definitions of finite and a-finite spaces are also simple as well
as Proposition (8.10) to Lemma (8.12) on p. 269-270. Notice that the

/-basis (En) is determined by a sequence of integrable characteristic


functions.

The various notions of Cauchyness, Definition (8.13) on p.270, are


straightforward, as well as Lemma (8.14) and (8.15) on p.271. Let’s
consider the proof of Theorem (8.16) on p.272. Given a sequence (/„ )
that is Cauchy in measure, in the second part of the proof, we must con­

struct a subsequence (//c(n)) and a sequence (i4n) of integrable sets such

that C fi{En —A*) < 1/n , and (/jr(n)) is uniformly Cauchy on


j4^, where (En) is the /-basis. Here we show how to replace the original
inductive construction by one available in SC . Since ( / n) is Cauchy in
measure, it follows that there exist H, B such that for all A;, all n > 0
and i , j > H ( k , n ) , B { i tj t k, n) is an integrable set (an integrable
characteristic function), B (z, j, A:, n)1 C E\ , n ( E n — B ( i , j , k , n ) ) <

2“*n_ l, and |/,• —fj\ < 2~*n _1 on B (z, 7', A:,n.)1. We may further as­
sume that H (k + 1, n) > H (fc, n) for n > 0 and hence H (k, n) > k for
n > 0. Construct F so that

F(0,fc) = fc,

F ( n + l,Jfe) = F(n,fr(Jfe,n + l ) ) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 270

Then we have F ( n , k + 1) > F ( n , k ) , F ( n + l,fc) > F( n , k ) , and


F (n, k) > H (k , n ). Clearly we can construct G such that F ( m + n , l ) =
F (m, G (m , n, I)) and we have G (m , n, I) > I. We have

n ( A Z = i B ( F ( n , k ) , F ( n , k + 1), k,n)) - y. B (F (n, k ) , F ( n , k + 1 ), k , n ))

So. An = A jix5 (F ( n , k ) , F (n, k + 1 ), k, n ) is integrable and

(En - A ,) < f ; H ((E„ - B ( F (n, * ), F (n, k + 1 ), k, n)))


k-1

< 1/n.

Clearly ( /F ( n ,* ) ) fc_ 0 is uniformly Cauchy on A ^ . Let K ( n ) = F ( n , n ) .


For m > n,

K (m) = F (n, G (n, m —n, m))

where G ( n , m —n , m ) > m. Then it is easy to show that (//r(n)) is


uniformly Cauchy on This finishes the construction. The rest of
the proof is straightforward.
Lemma (8.17) is simple. For B an integrable set, the definition of
the integration I s relative to B as well as Lemma (8.18) and (8.19)
pose no difficulties. The proof of Theorem (8.20) is directly formaliz-
able. Actually the first part of the proof can be simplified. That is,
assuming that / is a finite integration and |/ | < c and hence integrable,
by Theorem (4.11) on p. 242, for all but countable a > 0 ( / > a ) is
integrable and hence ( / < a ) is integrable, and also for all but count­
able a > 0 ( —/ > a) is integrable and hence ( / < —a ) is integrable.
The conclusion then follows easily. The second part of the proof on p.
276-277 is simple.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. INTEGRATION 271

4 .9 . P r o d u ct In teg ra ls. For two sets X, Y with inequality rela­


tions and complemented sets A, B in X, Y respectively, the product
A x B is the (multiple) complemented set

(A 1 x B 1, ( A 1 x B°) U (A0 x B 1) U (A 0 x B 0) )

in X x Y. If V 1 = : Ai € A i}, V 2 = : ^2 6 A? j are families


of subsets of X, Y respectively, then V 1 x V 2 is the family

{D Jl x ^ , : ( A l l A , ) e A , x A J}

of subsets of X x Y . Clearly, if x a is a characteristic function of A in V 1

and xb is a characteristic function of B in D2, then Xa * X b : (x, V) •-*


Xa ( 2 ) X b (V) with the domain dmn (x a ) x dmn (x b ) is a characteristic
function of A x B in V 1 x I?2.

Let (X, V X, L ( I ) , I) and (Y , V 2, L (J ) , J ) be two integration spaces.


Let L be the subset of F ( V 1 x V 2 , R ) such that / 6 L iff for some n >
0, a finite sequence ( c i ,..., cn) of real numbers, a sequence ( x j , x i )
integrable characteristic functions in ^ ( V 1 , L ( I ) , I ) , and a sequence

(Xi> —iXn) integrable characteristic functions in T ( V 2, L ( J ) , J ),


dmn ( / ) = (n"=1dmn (x,-)) x (n^=1d m n (x 2)) and / = E?=iC,Xi x x?
on dmn (x). Lemma (9.3) and Lemma (9.4) are simple. So

(Z x J )(/) = f v ( * ? ) . / (x?)
1= 1

defines a function on L. Theorem (9.6) is then straightforward, that is,


( X x Y, L, I x J) is an integration space and the definition of product
integrals and Fubini’s theorem on p. 281-282 pose no difficulties either.

4 .1 0 . M easu re S p a ces. We will skip this section because it gives


only an alternative but equivalent approach to the theory of integration.
All applications of integration can be based on the previous sections.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 272

5. N o rm ed Linear Spaces

5.1. D e fin itio n s and E x a m p les. The definition of linear space


on p.300 needs some revisions, for, without higher type recursions, we
cannot generally construct the sum of an arbitrary finite sequence of
elements from the operation + as a function on two elements only.
Throughout this section let IF be R or C. For X a set, let X <a0 be the
set of finite sequences of elements of X . If X , Y are sets and f : X —*Y,
let /* be the function from X <co to Y <0° defined by

/* = Ati.AxX....AX n . { f ((tt)j) (x X) ... ( X n ) , ........, / ( ( u ) ^ ) (x X) ... (x*)).

A permutation is a numerical function that maps a finite sequence


of positive integers into a permutation of the sequence. A grouping
function is a numerical function that maps a finite sequence of positive
integers into a sequence of finite sequences of positive integers obtained
by grouping the original sequence. If r is a permutation and u 6
X <ao we let 7r (it) denote the permutation of u obtained in the way
7r permutes the sequence (1,2, ...,(u )0). 7r(u) can be constructed as a
term containing t and u. Similarly, if 7 is a grouping function, then
let 7 (u) 6 ( X <00)<oa be the sequence obtained by grouping u as 7
groups (1 , 2 ,..., (tt)0). Then the definition of linear space can be stated
as follows:

DEFINITION 5 .1 . A linear space consists of a set X , a function


( i , y ) ■-» x + y from X x X to X , a function (o, x) *-* ax from W x X to
X , and a function £ from X <eo to X , such that the ordinary conditions
on the operations x + y and a z hold, and further,

£ ( ( x ) ) = x, £ ( u * (x)) = £ ( n ) + z,

£ (x (u )) = £u, £ ( £ * (7 (u))) = £ ( « ) ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 273

fo r x 6 X , u 6 X <oa, it a permutation, and 7 a grouping function.

These extra conditions are always easily proved for concrete ex­
amples of linear spaces, for instances, spaces of real or complex value
functions on some sets. For easy reading we will write S (x) as X{
where n = Ith (x) and x, = (x){. Notice that the extensionality condi­
tion in the definition of sets is extensively utilized here.
' The notions of semi norm, norm, normed space, and the metric
associated with a norm are defined easily. For the definition of a linear
mapping / from a linear space X to another linear space Y we also
need one more condition:

/(E (x )) = E ( /- ( x ) ) ,

for x 6 X <0°. Other notions, such as kernel, hyperplane, nonliabil­


ity, Banach space and so on, pose no difficulties. The lemmas and
propositions in the section are also simple.

5 .2 . F in ite -D im e n sio n a l S p a ces. The definitions of finite-dimensional


spaces and a subspace of a normed space on p. 307 are direct. The
basic properties and Proposition (2.3) and Corollary (2.4) are also sim­
ple. We consider the proof of Lemma (2.5) on p. 308 which involves
an inductive construction of a sequence of sets X \ , ..., X n.
First we need a notion. For a normed space X , we call a non-empty
finite sequence y = (yi, •••, yn) of elements of X linearly independent,
if the exists c > 0 such that for any sequence a = (ax, ..., On) of real (or

complex) numbers, ||£)iLx a»y*|| — c l®*l f°r eac^ * = ti. For any
sequence y = (yx, ..., yn), the linear subset [y] generated by y is defined
by: x e [ y \ i S x e X and for some sequence a = (ax, ..., On) of real (or
complex) numbers, x = £ ”=1 a,yt. Then it is easy to show that y is

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
5. NORMED LINEAR SPACES 274

linearly independent iff [y] is an n dimensional subspace with the basis

y-
Second we show that if y = (yx, ...,yn) is linearly independent with
the witness c > 0 and p (x , [y]) = e > 0, then ||x|| > e and y * ( i)
is linearly independent with the witness d = min je , | c | . From
p { x, [y]) = e it follows that ||aox + EILi ^ k o |s. Further, for
each i = l , ...,n ,

a Qx + £ a im - l<*ol I N
» = x i 1=1

Then, for any 8 > 0, by distinguishing the cases (i) c|ai| < 8 , (ii)
c|a i| > 0 and |a| ||x|| < jc |a i|, and (iii) c|aj| > 0 and |a| ||x|| > jc |a i|,
we see that ||a0x + a,-y,-|| + 8 > d |ai| for i = 0, ...,n . Since 8 is
arbitrary, the conclusion follows.
Now for the proof of Theorem (2.5) it suffices to prove the following
claim

(Claim) If x = (xx, ...,x n) is a finite sequence of elements


in X and e > 0, then there exists a subsequence x' =
(xfcj,..., 2 ^ ) , 0 < m < n, such that x' is linearly indepen­
dent and p(x{, [x7]) < e for all i = l,...,n . (For conve­
nience we use the empty sequence and take [( )] to be the
empty space.)

We may assume that e is rational. The sequence (ki) can be con­


structed along with another sequence (c,) of rational numbers, where c,
is supposed to be the witness for the linear independence of (x ^ ,..., x^.).
For each i = 1,..., n we can decide if ||xt-[| > 0 or ||x,|| < e. If there

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 275

exists the first i such that ||xt|| > 0, let &i = i and cx be a posi­

tive rational number less than ||x»||, otherwise let k\ = 0 and C\ = 0.


Suppose that k i , . . . , k i , C i,...,q have been constructed. If ki = 0, let
ki+i = 0, Ci+i = 0. Otherwise, denote x‘ = (x ^ , ...,xj.,), and for
each i, ki + I < i < n, decide if p(xi,[x']) > e /2 or < e. If there
exists the first i such that p(x,-, [x/]) > e/2, then let = % and
Q+i = m i n ||, otherwise, let ki+i = 0. The construction
uses only numerical recursions. Notice that the term p(x*, [x;]) should
contain the witness for the locatedness of [x;], which in turn depends
on the witness c/ for the linear independence of x'. That is why we need
the sequence (q ). We can show that if ki+i > 0 and q is the witness
for the linear independence of (x ^ ,..., i*,), then q+i is the witness for
the linear independence of (x ^ , . . . , X f c , , ifc,+l). The statement ‘c is a
witness for the linear independence of y ’ is equivalent to ‘for any se­
quence a of rational numbers, ||£"=1 a,y,|| > c |a»| for z = 1,..., n, where
n = l t h ( y ) \ This is a 11° formula, so an induction shows that if ki > 0
then ci is a witness for the linear independence of (x ^ ,..., x^). Then
the claim follows easily.
The proof of Theorem (2.6) on p.308-309 is now easy. Lemma
(2.7) is simple. We consider the proof of Lemma (2.8) on p. 310—
311. The proof uses inductive constructions of some sequences of real
numbers and we will replace them by constructions of sequences of
rational numbers.
First, horn the assumptions of the lemma it follows that there exists
R such that for any rational number 6 > 0, R ( 6 ) is a rational number
and \ \ x - R ( 6 ) e|| < p (x, Re) + S. Notice that

II* ~ (P*i + (1 “ P) h ) e|| < m ax{||x - t^H , ||x - tae ||} ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 276

for p, 0< p < 1. It follows that if r £ (tx, t 2) and ||x —re|| < ||x —txe||
and ||x —re|| < ||x —t2e||, then ||x —txe|| < ||x —te|| for t < t x and
||z —t2e|| < ||x — te\\ for t > t 2. This implies that if for some r 6
(tx, t2) and rational number 8 > 0, ||x —fxe|| > ||x —re||, ||x — t2e|| >
j|x —re||, and also ||x — txe|| > p (x ,R e )+ 6 , and ||x —txe|| > p (x , R e)+
8 , then R ( 8 ) 6 (tx, t2). From the assumptions it also follows that there
exists D such that for rational numbers tlt t2, if t x < t2, then D (tx, t 2)
is *a positive rational number such that

m a x {||x - *xe||, ||x - t2e||} > d + D (t x, t 2) .

Further, there exists H such that for rational numbers t x, t 2 , 8 , if 8 > 0


and m ax{||x — txe||, ||x —f2e||} > d -F 8 , then H { t \ , t 2 , 8 ) is one of tx
or t2 such that

llx-fffe.MWI ><<+ !«.


We may assume that p ( x , R e ) < d + j . Then we can construct se­
quences of rational numbers An.an.in.Cn.r,, as follows. Choose ax, b i , r x
such that rx € (ai,i»i)> ||x - oxe|| > d + 1, |[x - &ie|| > d + 1, and
||x —r*xe|| < So, for any t outside (ax, bx) we must have ||x —te|| >
d + 1. Let Ax = 0 and cx = 1. Suppose that A„, cn, On, bn, r„, have
been constructed. If An = 1, let

(32)

An+1 = An , Cn+l = Cn, fln+1 = On, ^n+ l = bn , r n+ l = Tn .

If An = 0, denote a = + \ ( b n - a^) and b = + § (6n - an)


and decide if p (x , Re) > d or < d + J m in {|c „ , |D ( a ,6 ) } . In the
former case, assign values as in (32). In the latter, let An+1 = 0,
Cn+x = m in ||c n , §I>(a,&)}, r„+1 = R(±Cn+i). Then, denote h =
H ( a , b ,D (a ,b )) , when h < rn+i, let On+i = h and 6n+i = bn, and

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 277

when h > rn+x, let a„+i = an and 6„+x = h. Clearly the con­
structions use only numerical recursions. Some inductions show that
if Xn = 0, the Cn < 0 < (bn - On) < (§)" 1 (6X- ax), rn 6

[«n,^»] C [ fl n- XA _ X], P (x, Re) < rf + ^C*, ||x - r«e|| < d + \cn,
||x —ane|| > d + c*, and ||x — 6ne|| > d + c„. Notice that these formu­
las are all II? or £?. Further, if A„ = 1, then dn = dn_x, an = On_x,
bn = bn- x, rn = r n- i , and p(x, Re) > d. Then, following the rest of the
proof on p. 310-311 we can easily show that (rn) converges to a limit
£ such that ||x — £e|| > d entails p (x, Re) > d.
The definition of quotient space and Lemma (2.10) axe easy. The
proof of Theorem (2.12) on p. 312-313 involves an induction on the
dimension of the space. We don’t know if that can be unraveled. This
is one of the theorems in B&B that we have to leave open. On the other
hand, notice that for a fixed numeral n and an n-dimensional subspace
F, the proof can be formalized, for we can relegate the induction to
the meta-language level. What cannot be formalized is the assertion
for arbitrary n-dimensional subspace F with n as a free variable.

5.3. The Lp Spaces and the Radon-Nikodym Theorem. We


assume a completely extended a-finite integration space (X , V, L \ , I ).
Lemma (3.1) on p.314 is simple. Corollary (3.2) needs some inductions.
We can first use an induction to prove the II? statement: for any finite
sequence a x, ..., ctn of rational numbers of length n with a x+ . . . + a n = 1,

a?l ...a“'1 < atxa x + ... + ctnO n-

Then the corollary follows by approximation. Lemma (3.3) and The­


orem (3.4) are easy. The definition of Lp and its norm ||*|| are di­
rect, as well as Proposition (3.6) and Theorem (3.7). Notice that
Minkowski’s inequality in Theorem (3.7) can be generalized to sums

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 278

of finitely many functions by an induction, because an inequality is a


II? formula. Lemma (3.8) to Theorem (3.25) on p. 316-324 pose no
difficulties to the formalization. No inductions are ever used and all
the inferences are based on the previous results and the intuitionistic
logical axioms and rules. The notion of absolute continuity is also di­
rect. The equality (3.34.1) in the proof of Theorem (3.34) on p.329 can
be obtained by an induction on a 11° formula. Otherwise, the proofs of
Lemma (3.27) to Theorem (3.40) on p.326-334 are straightforwardly
formalizable. No inductive constructions are involved.

5.4. T h e E x te n sio n o f Linear F u nctionals. The proof of Lemma


(4.1) on p.335 is simple. For the formalizations below, we need a gen­
eralization of the lemma:

LEMMA 5 .1. Let K be a bounded located subset of a normed linear


space X , n > 0, a = ( a i , ..., On) be a sequence of elements in X . Then
the set

5 = | ^ U a i + t0x : x e K, 0 < U < 1, t = 0 ,...,n, £ U = l }


L=i t=o J

is located.

The proof is similar, where the total boundedness of the set

j ( * o , * i , tn) 6 R n + 1 : 0 < t{ < 1, for z = 0 ,..., 1, and = 1j

is used instead of that of (0, a ). For N > 0, let

** - { (£ 1- - V * ) : * " ° ’ 1....... 2* p ’ s 2" 1 •

Then it is easy to see that A s gives arbitrary close finite approxima­


tions to the set as IV —» oo.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 279

For the definition of a convex set, we need the arbitrary finite convex
combinations, that is, K is convex if £ ”=1 t{X{ £ K whenever n >
0, Xi £ K , 0 < U < 1 for i = and £"=1 £, = 1. A similar
generalization is needed for the definitions of a cone. The definition of
the cone c ( K ) generated by a convex set K needs no changes. When
K is convex and open then S in the lemma above is also convex and
open. We will call it the open convex hull of K and a, and denoted by
Hull ( K , a).
Lemma (4.2) is simple. Consider the proof of Theorem (4.3) on
p. 336-340. An inductive construction is used to prove the following
claim (Note: The original max in (3) below should be a misprint.)

(Claim) If i f is a bounded, located, open, convex subset of


a separable normed linear space X such that p (0, K ) > 0,
x = ( i i ) “ 0 is a sequence dense in X , and -X q £ K , then
there exists a sequence K = Kq Q K \ C ... of bounded,
located, open, convex sets such that
(1) p ( x o , c ( K n)) > (1 - 2 - n) p ( i 0, c ( i f n_ 1)),
( 2 ) p ( 0 , K n) > 0 ,
(3) m in {p (x n, c ( i f n)), p ( - i n, c ( i f n))} < n"1.

Informally, the sequence ( K n) is constructed this way: At each


stage i = 1 ,2 ,..., we either choose x,-, or choose —x,-, or choose noth­
ing. Then K n is the open convex hull of K and the chosen elements
at the stages i = 1, ...,n . So we can instead construct by recursion a
sequence (fc,) of 1, or —1, or 0, indicating the choices above. Then the
corresponding sequence y = (y,) of chosen elements and the sets K n
can be constructed uniformly. First, there exists s such that for any
finite sequence r = (rx, ...,r m) of 1, or —1, or 0, s (r) is the sequence
(yu ..., Vm1) of elements among X i,..., xm, —X i,.., —xm chosen according

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 280

to r. Then, for any n > 0 and r as above, p (x n, H u l l ( K , s [ r ) ) ) ex­


ists by the lemma above and similarly p ( i 0, c(Hu ll ( K , a (r)))) exists.
We have to construct the witnesses (/i„) for (2) above along with the
sequence (&„.). So, let &o = —1 (because —io 6 K ) and ho be a posi­
tive rational number less than p ( 0 t K ) . Suppose that kn-i, h^-i have
been constructed. To construct kn, hn, let r = (ki, ..., kn-i), compute
p ( x n, H u l l ( K , 3 (r))) and p ( —xn, H u l l ( K , s ( r ) ) ) to some sufficient
degree of precision and decide if they are less then 1 /n or greater than
1/ (2n). If one of them is less than 1/n, then let kn = 0 and hn = hn-i-
Otherwise, Let
t hn—1_____
4n(/in_i + ||x „ ||)‘

L e tr + = r * { 1) a n d r~ = r * ( —1). Compute^(xo, c ( H u l l ( K , a (r + ))))


and p ( x o, c ( H u l l ( K , s ( r ~ ) ) ) ) to some sufficient degree of precision
and decide if they are (i) greater than

( l — 2-n ) p(xo, c(H ul l ( K , 3 (t)))) ,

or (ii) less than

( l - 2 - " ) / . ( x „ , c ( f f u H ( i i r , J ( r ) ) ) ) + 2 - " - I An. I .

If (i) is true for the case of r+ , let kn = 1, and let kn = - 1 in all other
cases. Notice that the term p ( x n, f f u l l ( K , s ( r ) ) ) should contain the
witnesses for the locatedness of Hull (K , a (r)), but the locatedness is
proved for an arbitrary r uniformly, so there is no need to construct the

witnesses inductively along with kn. Then let K n = Hull ( K , a (kn) ) ,


where kn = (k j,..., kn). Then the condition (3) can be verified directly.
A n ° induction shows that p (0 ,K n) > hn (cf. the proof on p.337).
Since —x0 6 K n for all n, following the argument at the beginning of

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 281

the last paragraph on p.336 we see that p(xo, c ( K n)) > A*. Hence

( l - 2-") p(x„, c ( K n- 1)) + 2 - " - ' K - i < ( l - 2 -”"1) p (x 0, c t if n - O ) .

Then, following the argument in the last paragraph on p.337 we see


that if (ii) above is true for both the cases of r+ and r~, there will be
a contradiction. Hence the condition (1) above holds directly.
The rest of the proof on p. 338-340 are formalizable straightfor­
wardly. No further inductive constructions are needed. Finally, Corol­
lary (4.4) to (4.6) on p. 341-342 are simple.

5.5. Quasinormal Linear Spaces; the Space L«,. The defini­


tion of quasinorm on p.343 is straightforward: a quasinorm is a func­
tion from a set I to the set of functions from X to R+0 such that for

i e l , ||*||,- is a seminorm on X . Other notions in the section are also


straightforward and the proofs pose no difficulties either. In particular,
no inductive constructions are used. We will skip the details.

5.6. Dual Spaces. The notion of double norm and Proposition


(6.2) to Theorem (6.8) on p. 350-356 are straightforward. The in­
duction in the proof of Corollary (6.9) on p.357 can be dismissed:
First by Theorem (6.8), for each n > 0 there exists yn in X such
that |v>(u) —u (y n)| < 2“n for all u € S *. Then the original proof
works after we set X\ = yi and xn = yn —yn- i for n > 2.

5.7. Extreme Points. Definition (7.1) and Lemma (7.2) on p.


357-359 are simple. Notice the notation

M (u, K ) = sup {u ( x ) : x 6 K } ,

S (u, a, K ) = {x 6 K : u (x) > a }

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 282

for a compact K and bounded functional u. For the formalization of


Lemma (7.4) below we need a more informative form of Lemma (7.2)
as follows.

LEMMA 5 .2. Suppose that K is a compact convex subset of a normed


space X over R, u, v are linear functionals on X boundedby p, q re­
spectively, and e > 0 is a rational number. Then there exist rational
numbers a, b, r such that for w = u + rv,

M (w , K ) —e < a < b < M (w, K ),

w —u is bounded by e, \w (x) —u (z)| < e for all x £ K , and if c € (a, b)


makes S (w, c, K ) compact, then the variations ofu and v on S (in, c, K )
are less than t .

PROOF. From the proof of Lemma (7.2) on p. 358-359 we see


that there exist d , a , 0, and 5 > 0 such that 5 < , a < f3,
S~l (/3 — a) < 8 , and for w ‘ = u + 8 ~l (0 —a) v, M (w‘, K ) —e < d <

M { w \ K ) , w' — u is bounded by e, S ( w ' , d , K ) is compact, and the


variations of u ,v on S {w', d , K ) are less than e. Choose a rational
number d such that

( -i t o s M (w ',K )-d \
0 < d < min < £ ( 0 —a ) ,

and choose a rational r such that £_1 ( 0 —a) — d < r < £-1 ( 0 —a),
and choose rational numbers a, b such that

d + ^ { M K , K ) - d ) < a < b < d + ? ( M (wf, K ) - d ) .

Let w = u + rv. Clearly, w is also bounded by e and for x € K,


|u> (z) —u (z)| < e. Further, for z 6 K ,

|u /(z) — u /( z ) | < d\v (z)| < ^ (M ( w ', K ) —c ') .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
S. NORMED LINEAR SPACES 283

So,

\M (ti>, K ) - M ( 10', K ) | < ^ ( M (w ‘, K ) - c ').

It follows that M ( w , K ) — e < a < b < M ( w , K ) . For c 6 (a, 6) we


have S (w , c , K ) C S (w', c7, K ), so the variations of u, v on S (w , c, i f )
are less than e. |

Lemma (7.3) on p. 359-360 is simple. We consider the proof of


Proposition (7.4) on p. 360-362. We must unravel the construction
of the sequence (un) of elements of X* and the sequence (o^) of real
numbers satisfying the conditions (7.4.1) to (7.4.6) on p ..360. Notice
that, for (un)“ 2 a sequence that is dense in 5* relative to the dou­
ble norm, the original construction gives un = u + £"_2 r»ti» for some
r2,...,r n. So, instead of constructing (un), (an) directly, we construct
four sequences (rn), (<x„), (6„), (dn) of rational numbers such that for

Vn = U + E "=2 TiUi

(1’) M (vn, K ) — 1/n < cin < bn < M {vn, K ),


(2’) | (M (wn, K ) - bn) < dn < l ( M ( v n, K ) - bn),

(3’) S ( v n, a n , K ) C S K i . L i + C O ,
(4’) for any a 6 (on,bn) such that S (vn, a , K ) is compact, the
variance of Un on S (vn, a , i f ) is at most 1/n,
(5’) |v „ (z) —i/n_i (x)| < 2"n-1 m in{d2 :1 < j < n — 1} for all x 6

(6’) vn —vn- i is bounded by 2~nS.

Then, for each n > 0 choose any atn 6 («„, bn) such that S (v„, a*, K )
is compact, we can easily verify that (7.4.1) to (7.4.6) on p.360 hold.
Here we show how to construct the sequences of rational numbers.
From the lemma above it follows that there exist operations iZ, A,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 284

B, D such that for functionals u, v on X bounded by rational num­


bers p , q , and any rational numbers e > 0, r = R ( u , v , p , q , e ) , a =
A ( u ,v ,p ,q ,e ) , b = B ( u , v , p , q , £ ) satisfy the conditions stated in the
lemma and further d = D ( u , v, p, q, e ) satisfies | (M (w , K ) - b) < d <
| (M (w, K ) —b). With these, we can construct the sequences: There
exists a rational number p\ that bounds u and we may assume that
6 is rational. Let r x = 1, uj = u, a\ ,b i, d i be any rational numbers
such that (1’) and (2*) hold for n = 1. Suppose that Tj, a.j, by, dj
have been constructed for j = 1, n — 1. Let u' = u 4- E y lj rytty,

P = P i + E"=2 rj> v> = «n» and

z' = min j n _ 1 , 2 ^ - 1 , 2 ~nS, 2~n~1 min {dj : 1 < j < n - 1}

Then we let rn = R (uf, v ', p, 1, s'), On = A (u', v', p, 1, e'), bn = B (u', v', p,
and dn = D ( u \ u',p, 1, s'). Clearly, this is a primitive recursion. Let

n
Vn = U + ^ 2 Tj Uj .
1= 1

Then vn = u„_i 4- rnu„. (1’), (2’), (4’), (5’), and (6’) follow from
the definition and the lemma above directly. By the lemma we also
have |un (x) - v„_x(x)| < for all x 6 K . So, M ( v n, K ) >
M { v n. l t K ) - and hence

K ) - ^ .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 285

So, if un (z) > On, then

> M ^

> M , K ) - -2 ( I ( M (v „ .„ K ) - K - i ))

>6.-i + j(M0v-i,jr)-4„-i)
> ^n—1 + ^n-1-

So, (3’) also holds.


Then the rest of the proof on p. 361-362 can be formalized straight­
forwardly. No more inductive constructions are needed.
The original proof of Theorem (7.5) on p.363 contains repeated
applications of Lemma (7.4), which means another induction. However,
this can be eliminated as follows. Let’s assume the hypotheses of the
theorem. Let e be an arbitrary positive number. Let y = (yi, •••, yw)
be a ^ approximation to K . For any subsequence (yjL, ...,y Jm) of the
sequence y, let

be the closed convex hull. Notice that [(yy,,..., yJm)] is totally bounded.
For each j = 1,..., N and each subsequence ( j \ , ..., j m) of (1 ,..., N), we
can decide if p (yy, [(yyu ...,yym)]) is greater than | or less than *. In
the former case, by Theorem (4.3) on p.336, there exists a functional
u of norm 1 such that

u (y y ) > s u p { ti ( y ) : y 6 [ ( y * , . . . , yym ) ] } + | .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 286

Then, by Proposition (7.4) on p.360, there exists an extreme point

> sup {u (y) : y € [(y*, ...,y Jm)]} + - .

So,

P [(^ii j •••! Vim)])

> inf { |i t ( i jyii...jm) - u(y)| : y 6 [(yh , ...,yJm)]}

> u (xj d i ) - sup {u (y) : y 6 {(yh ,..., yim)]}


t
> i
Collect all such extreme points z Jai into a sequence z, it suffices to
show that p (z , [z]) < e. We prove instead that p (z, [z]) > | leads to
a contradiction. So, suppose that p( x, [z]) > | . For each j = 1 ,..., N
decide if p ( y y, [z]) > £ or < f . Let j 1} be all those j = 1 ,..., N
such that p (y y ,[x ]) < §. Since ( y i,...,y jv ) is a ^ approximation to
K , there exists y, such that p ( x , y j ) < Then p(yy, [x]) > |§. For
each yjk there exists xk 6 [z] such that p (yjk, z fc) < | . So, for any
a x, ...,ctm € [0,1] such that SJLx a t = 1 , we have

- II
~ k=l
E a*zlII ~ k=i
E (**-
7e e
> 16 “ 8
5e
" 16'

So, p (y j, [(y*«—!%*>]) > f “ d we have the corresponding extreme


point xJaii...am such that p ( z jai im, [(yiu ....y y j]) > f. There exists
yj> such that p (z , v t ) < -h- Since z is an element in the
sequence z, j ' is a m o n g ;!,...,; m. This contradicts p (xia W m , [ ( y * , yym)

| . This completes the proof. No inductive constructions are used here.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
5. NORMED LINEAR SPACES 287

Theorem (4.3) and Proposition (7.4) are uniformly applied to a finite


sequence of points to get the extreme points.

5 .8 . H ilb er t S p a ce and th e S p ectra l T h eo rem . The basic def­


initions and results up to Definition (8.8) on p.368 are simple. Notice
that the linearity of inner products for linear combinations of finitely
many elements follows from the condition for two elements by a 11?
induction, because an equality between two complex numbers is 11°.
We consider the Gram-Schmidt orthogonalization process. To avoid
confusions with sequences we use (x , y ) to denote inner products. To

unravel the inductive construction involved, we need a lemma.

LEMMA 5 .3 . Suppose that the operation F is such that for any


natural number n > 0, (i) if r is a sequence of n real numbers, then
F (r, n) 6 R ; (ii) for any M > 0 there exists P > 0 such that if r
is a sequence of n real numbers and |(r)J < M for i = 1, ...,n , then
Ii7’ (r, n) | < P ; (Hi) for any M > 0, e > 0 there exists 8 > 0 such that
for any two sequences r, r' o f n real numbers, t/Kr)^ < M , ((r')^ < M,
|(r )j — (r'JJ < 8 for all i = l,...,n , then \ F ( r , n ) — F ( r /,7i)| < e.
Suppose that the same holds for G in place of F. Then, given any real
number a and a non-decreasing function h from positive integers to
positive integers, there exists a sequence (kn) of 0 and 1 and a sequence
(rn) of real numbers such that ( 1 ) if k \ = 0 then T\ = 0 and |a| < h ( 1),
and if k\ = 1 then |a| > j and ri = |a|-1 ; (2 ) for n > 1, i f kn = 0
then G ((r i, ...,r n_ i),n — 1) < h (n )-1 and rn = 0, and if kn = 1 then
G { ( r u ...,r n_ i),n - 1) > ( 2 k { n ) ) "

PROOF. From the assumptions it follows that there exist 8 and P


such that for any n > 0, and rational numbers M > 0 and e > 0,
P = P (n, M ) and 6 = 8 (n, M, e) are rational numbers such that the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 288

conditions in (ii) and (iii) above hold for both F and G. We show how
to compute kn and rn (m) by recursion on n. The case n = 1 is trivial.
Notice that to compute rn (m) we can use (ri ( m' ) , rn_x (m')) if m' is
given by a primitive recursive function of n, m, and (ri ( 1 ) , rn_x (1)).
Let M = E fe 11 (|ri (l) | + 2 )l P = P (n , Af), let m0 be the smallest
natural number such that l/m 0 is less than \8 (n, M, and
and let m' be a natural number greater than mo and such that 1/m ' is
lo tt lu m I f (n ,M , 2^ / {p„ y J and Notice that mo i>
independent of m. If
3
C?({ri (m0) ) ...,rn_l (m0)), n - l) ( m 0) < ^ ^ ,

let kn = 0 and rn (m) = 0, otherwise let kn = 1 and


, = F ((rx (m '), ...,r n_i (m')), n - 1) (m')
G ({r1(m ')) ...,r n_ i(m '))) n - l)(m ')'

We must show that for all t = 1 , n, r,- 6 R. This is a 11° formula,


so we can use an induction on n. Suppose that t\ 6 R for i = 1 , n —1.
Recall that the value of kn is independent of m. If kn==0 then rn =
0 € R. Otherwise, keep the notations above and suppose that m i > m.
It suffices to show that |r„ (m i) —rn (m )| < 1/m . We have
3
G ((ri (m o),..., rn- i (m0)), n — 1) (m0) >

Since ..., r n- i are real numbers and m' > mo, we have |r,- (m')| < M,
(mo)| < M , and

lr»(m') - Ti (mo) | < 2/m 0 < 6 ^n, M,

for i = 1 , ..., n - 1. So by the assumptions on G ,

|G ((r! (m '), ...,rn_: (m')), n - 1) - G((rx (m o), ...,r n_x (mo)), n - 1)|
1
< 1 2 h{n)'

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 289

Since m ' > mo > 12h (n), we have

|G ((ri {m' ) , ...,r n_i (m ')), n - 1) - G ( ( r x (m'))> n - l)(m ')|

< -_ L -
“ I 2 h{n) '

l^ ( ( r i (™o) r - i V i W , n - 1) - G ( ( ti (m0) , ...,r n_x (m0)), n - l) ( m 0)|

— 1 2 h(n)
So,'

1
(33) £? ((ri(m /) , . . . , r n_i(m ')), n - l ) ( m ' ) >
/ft (n )
The same holds with m 1 replaced by m'lt the corresponding number
used in computing rB(m i). Further,

|r< (m i) - n (m')| < l / m [ + 1/m' < 8 ( n , Af, --2 - — — ]


\ 24ft (n) ( P - t - l ) m /
for i = 1, ...,n — 1. So,

l^ ( ( r i (m i) n - 1) - F’ ((r1(m ;) , (m ')), n - 1)|

< ‘ .
~ 24A(n) ( P + I) m

|F ((rx (m;) , (m')), n - 1) - F « r x(m '), ...,rn_i (m#)), n- 1) (m')|

< ______ J ________ .


~ 24h(n) (P + l ) m

The same holds for m[ in place of ml. So, we have

(34)

l^ ( ( r i (mi ) . -.T*n-x (™i))> « - 1) (™i) ~F((ri (m '), ...,rn_i (m')), n - 1) (m')|

< _______ I_______


~ %h (n)2( P + l ) m

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
5. NORMED LINEAR SPACES 290

The same holds for G in place of F. From (1) the inequality (33)
and the same inequality for m[, (2) the inequality (34) and the same
inequality for G, and finally (3)

\G ((rx {rn! ) ,..., rn_x (m')), n - 1) (m')| < P + 1/m ' < P + 1

and the same inequality for m [ t it directly follows that jrn (m x) —rn (m)|
1/m . This completes the proof of rn 6 R. The rest is simple. |

Now consider the orthogonalization process. Let (yn) be a sequence


dense in the Hilbert space X . We may assume that for each y 6 X
and any e > 0 there exists arbitrarily large n such that |y - y„| < £•
Informally, the orthonormal basis (en) is constructed as follows: If

llyill > 1/2 let ex = IlyiU-1 yx, if ||yx|| < 1, let ex = 0. Suppose that
ex, ...,en_x have been constructed. Decide if

n —X n —1
Vn ~ £ (2 /n , 6i) e,- > — or Vn - £ {y»> ei ) e«
t=l 2n i=1

In the former case, let

n1 —1 -I
en = V n ~ J 2 (yn, e.) ei Vn ~ £ fr"' *»') e* >
i=l i= l

and in the latter, let en = 0. We can express en as D ”=1 r”y;. So we can


construct instead the sequence rj, rx, r f , .... For m > 0, let n = N (m),
j = H (m ) be the unique numbers such that the m-th term of the
sequence rj, rx, r j , ... is r". Denote 6tJ = (y*, yy); then we should have

n —1 n —1
£E fan, ei) e. = £ I yn, £ r*-,yy/ £ rjyy
i= l i=i i '= i J
n - l ( n -1 *
= £ £ £ ) yy.
i=i \t=j j'=i

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 291

(Notice that the new conditions in Definition 5.1 are needed here.) So
let F, G be the terms such that
/ N (m )-1 i \
F (r, m - 1) = J N' (m) - H (m ), 1, - £ £ &N(m).i,r}'rH(m) .
\ i=H{m) j ' = l /

N(m)-1 /N(m)-1 i
G ( r , m — 1) = I(N(m) - E ( E Z ] y,
j=i \ i=j j ’=i

for r = { r \ , r l , r l , r^^ Z ij). Applying the lemma with N in place


of the function h, we have the sequence r | , .... and (Am). No­
tice that N (m) = n for m = sfSpH + 1,...., nfcpll -f. n . Let ln =

k n(n-i) | x = ... = k *(n-i) [ n. From the construction we see that for any
n, if ln = 0, then G ( r , m — 1) < 1/n for any m with N (m ) = n and
r = and hencer" = 0, fo r; = l ,...,n . Similarly,
if ln = 1, then G ( r , m — 1) > 1/2n for any m with N (m) = n and
n —1 i
r" = —G { r , m — 1) 1 £ £ bnJ'rj‘r) ’ for ■? = *> ••*»« “ 1.
»=i j'=i
-x
r” = G ( r , m — 1)

Let ^ r"yj. Then we have if /n = 0 then en = 0, and


if /„ = 1 then ||e„|| = 1 and (em',^m) = 0 for all m ,m ' = l ,...,n
and m ^ w! (by an induction on n). Let M n-1 denote the finite
dimensional space expanned by e i , ..., en_i. For any y 6 X and e > 0,
choose n > 2 /e such that ||y —y„|| < e/2 . If /„ = 0, then
n—i /n—l » n —1
Vn-E E E h nj 'r ^ r) ] yj V n - J 2 {yn, e.) e, < - < e/2 ,
n
1 \i = j j'=l »=1

and hence p { y , M n- i ) < e. If /„ = 1 then yn E Mn, and hence


p (y, Mn) < e. Then following the proof on p.369 it is easy to infer
that (en) is a orthonormal basis. This completes the Gram-Schmidt
orthogonalization process.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 292

The definitions and lemmas from Lemma (8.10) on p.370 to Lemma


(8.15) on p.372 axe simple. Consider the proof of Lemma (8.16) on
p.373. Here, given a Hermitian operator A with bound 1 and a unit
vector Ui, we must construct a sequence (un) of unit vectors satisfying
this recursion formula

(35)

«n+l = ^ A ti* + Un - ^ ( A t X n .U n ) ^ ! (^ A v ^ + V ^ - ^ (A lin , U * ) U n ^j ,

where a direct calculation gives

1 1 ^ 1
-A lin + U n - “ (AtinjUnJtin = | K ||2 + “ || Alin “ ( A ^ , Un) U„|

To construct the sequence, we need an orthonormal basis (eJ) ° l1 and


construct the coordinates of tin in the basis. Denote a{j = (Ae*, ey).
We have a i j = oj j and for any i > 1, Ey^i |aitJ|2 = ||Ae^||2 < 1. So,
there exists N& such that for any i > 1 and e > 0,

£ k 'j l2 < e.
i=Nx(«,«)

If un = E “ i Viei for all n > 1, then Atin = EyLi E “ i the


recursion formula becomes

(36) c 1= 1 /2 *

(e* li Ei +K - )
We will construct, by recursion on n, the sequence 6JJ (m) of rational
(complex) numbers along with the sequence N (n, m ) of natural num­
bers which is supposed to be the modulus of convergence for E ^ i l&?|2>
that is, E f i n,m) |6?|2 - 1 < 1/m . To compute 5J+1 (m), we can dis­
card the terms in the sum E * in the denominator for k sufficiently large
and for other k we use approximations (m ;) for some m'. b\ (m ) and

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
5. NORMED LINEAR SPACES 293

N ( I , m) are all given. Let

N (n + 1, m) = N (n, 8m )+m ax (i, 2N(n>8m)+3N (n, 8m) m ) : 1 < i < N (n, 8m )}

Suppose that (i) 6" € R (or C) and |fi?| < 1 for all * and (ii) |£ £ S ,,m) \b? \ 2 - l| <
1/m for all m. Let un = E “ x 6"et-. Then

= | (i4tin , u „ ) | < 1.

Denote N = N (n + l,m ) and / = N (n, 8m). We should have

< l/8 m ,

E Wl’ ^ l/8m.
fc=N

and further

E
*=W
X > xI,*

< 2 E
fc=AT i=i
E ‘X.<
fc=l »=l+i

1=1 k=N «=l+i

< 2 /+1l ( 2 ,+2lm )" + 2 2 |£ |


i=i+i
< 1/2m.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 294

This justifies the formula for N ( n + l,m ), because a direct computa­


tion shows that for the denominator of (36)
2

£
k ~ t i,j

= I M 2 + ~ ||Aun - (Aun, u„) un||2

> 1.

Similarly, we can find primitive recursive functions P, Q such that sup­


posing (i) and (ii) about (6") and N (n ,m ) above and denoting

I = N (n + l,Q (n ,m )),

r = P (n ,m ,l)

I£ » + bk ~ ^ f c E . j $
b=
/ 2 \ 1/ 2 ’
(£ * + is - £ .„ )

I E L 6" ( 0 q q (?) + *2 (/') - & (f) E U (Q ai,j (Q 6? (/')


6' =

(e U || EL 4?((') <n*GO+ 4?(10 - |4E((') E U s?('0 «MGO4?G0|)


then |6 — b'\ < 1/m . So, we let 6J+1 (m) = 5'. Then the argument
above actually shows that if
N (n ,m )
V i > 0 (6 " € R ) A V m > 0 £ |4?|a - l < 1/m ,
i=l

then the same holds for n + 1 in place of n. Since this is a 11° formula,
an induction concludes that (tin) is a sequence of unit vectors. Then
the recursion equation (35) follows from the identity (36) by some com­
putations. With these available, the rest of the proof of Lemma (8.16)
on p.373 is simple.
Now consider the proof of Lemma (8.17) on p. 373-374. The original
proof is obtained by an induction on n. We will replace it by a direct

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout perm ission.
5. NORMED LINEAR SPACES 295

construction of the required vector y there. We assume an orthonormal


basis (en). Notice that for any sequence r i , ..., rn_x of complex numbers
with rational real and imaginary parts such that lri|2 < 1 and
rn = (l — ^ , u = £"=1 r,et- is a unit vector, and any unit
vector can be approximated by such unit vectors. We call such unit
vectors rational unit vectors and we make the convention that when
we quantify over rational unit vectors we mean a quantification over
the sequence (r*i, . . . , r n_x) and so it is a type o quantification. Clearly
it suffices to prove Lemma (8.17) for any rational vector x and rational
e > 0. Given the assumptions of the lemma, from Lemma (8.16) it
follows that for any z, 1 < z < n, any rational 8 > 0, and any rational
unit vector u, there exists a rational unit vector y such that ( A # j , y ) >
(A{U, u) —8 / 2, ||Ajy — (.A»-y,y)y|| < 8 , and for any Hermitian operator
B commuting with >4,-, ||£ y || < 2^324-^ -1 ||flu||, where [a] means the
least integer greater than a. So, there exists Y such that for any z, e,
and u as above, y = Y (z, u, 8 ) is a rational unit vector that satisfies
the condition.
Now we construct the sequence needed for the proof. First construct
a sequence (£,•) such that eo = e/2, £,+i = 2"f32*» ^£,-. Then construct
a sequence (it*) of rational vectors such that U\ = x, and

^ t+ l = ¥ (* j ^ i i ^ n - t ) •

A simple induction shows that u,+i is a rational unit vector, for z =


1, ...,n . The construction guarantees that

||AtU»+l (•Aj'Ut+X) Wt+lJ'Ut+lH ^ £n-it

(Aiiii+i,Vi+i) > (AjtijjUj) —£n-i,

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
5. NORMED LINEAR SPACES 296

and for any Hermitian operator B commuting with A{,

(37) ll-Bm+ill < 2 ^ ' - ' \\BmW .

Then an induction on j shows that for i + 1 < j < n + 1,

The inductive step is obtained by letting B = A\ —(AiUi+i, ■u.+i) I and

i ~ j in (37). Let y = u ^ . Then ||A<y - (i4i«i+i,t*i+i)y|| < e /2 . A


direct computation gives (cf. p.374)

\\A& ~ (<A.-y,y)y|| < ll^y - (A^.n,u*+i)y|| < e

for all i = 1, ...,n . Further,

(A iy , y ) > (Aiuaitia)(y,y)- IKAjy-^tia.uaJy, y)||

= ( j4 ix , x ) —e.

This completes the proof.

Lemma (8.18) on p. 374-374 is simple. The product space X =

I E * [~1)1] ^ the set of sequences of real numbers in [—1,1] with


the metric p ( (x „ ), (y„)) = 2 ~n |xn - y j - (*■«) is a sequence of
functions and then the real algebra P generated by (xn) consists of
functions / 6 C { X ) such that
N

for some integer N > 0 and / > 0 and some finite sequence (e,-!,„„*,) of
real numbers. So P is a subset of C (X ) with these as witnesses. For

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 297

a sequence A = (An) of commuting Hermitian operators, to construct


the algebra V (A) generated by (A„) we must construct the product

^(n = A ,,...Atl

for an arbitrary sequence {i\, of positive integers. First we prove


a more general lemma:

‘ LEMMA 5 .4 . For any sequence (An) of bounded linear operators,


there exists a sequence (Bn) of bounded operators such that £ n+i =

An+i-Sn f 0T n-

In order to prove the lemma, we will develop some notations so


that we can construct elements of H as we construct real or complex
numbers. We fix an orthonormal basis (e^) of H. For any x £ H,

llz l|2 = E “=i |(a;, e*)|2 - So, there exists Q such that for any x 6 H ,
fi(x ) is a modulus of convergence for |(x, e,)|2 , that is, for any

m>0, £“n(x)(m
)+i|(*»e»)|2<1/™. Let
x (m ) = ((x, d ) (N ) , . . . , (x , eL) ( N )),

where N = [(2m2 (fl (x) (2m 2)))* ], L = Q (x ) (2m 2) . Notice that x (m )


is a finite sequence of rational complex numbers. Clearly, x (m ) is
supposed to represent a 1/m approximation to x. For any sequence
r = (ri, ...,rj) of rational complex numbers let

i
E (r ) = £ r.-e» 6 H.
i=i

Then it is easy to verify that

(38) ||x — E (x (m ))|| < 1/m , for all m > 0, x 6 H.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
S. NORMED LINEAR SPACES 298

So, instead of constructing x, we can construct the sequence (x (m ))m of


rational complex numbers and then let x = limm_ 00 E ( x ( m ) ) , where
the modulus of convergence is fixed. Now we prove the lemma.
Proof of the lemma. There exists a sequence (M„) of integers such
that An is bounded by Mn. Let x € H be arbitrary. We can construct
a term q [n, x , m] such that

g [l,x ,m ] = (A ix K m ),

q [n + 1, x, m] = (A»+1 (E (q [n, x, 2Afn+1m]))) (2m) .

Then, by an induction on n we can prove the 11? formula

||£ ( g [n ,x,m ]) — E ( q [n,x,m '])|| < 1/m + 1/m ', for all m ,m ' > 0.

The case n = 1 follows from (38) directly. Suppose that it holds for n.
Then we have

||E (q [n , x, 2M n+lm\) - E ( q [ n , x, 2Mn+1m'])|| < ^m '

and therefore

\\An+i (.E { q [ n , x, 2 Af„+im ])) - An+i (E ( q [ n , x, 2 M„+im/]))||

<- L + -L.
2m 2m'

From (38) it follows that

||E (q [n + 1 , x, m]) - An+1 ( £ (g [n, x, 2 JW„+iTn]))|| <

II# (? [n + 1, x, m 1]) - An+i ( £ (g [n, x, 2JWn+im ,]))|| < ^ 7.

So, the formula holds for n + 1 . Let Bnx = limm_ 0OE (q [n.x.m]). Let
m —» 00 in the above inequality. Then we see that Bn+ix = An+xBnx. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. NORMED LINEAR SPACES 299

Then given the sequence A = (An) of commuting Hermitian op­


erators, for any sequence (ii,...,ij) of positive integers one can easily
construct A ^ such that

.•(■•i+i) — ... »i)

for any positive integers ii ,...,i i +i. If ir is a permutation of finite se­


quence of integers, the equality

can be proved by some inductions on some 11° formulas of the form

xi y) = y) •
Then the mapping / / (A) from V to V (A) can be constructed.
The proofs of Lemma (8.19) and (8,20) on p. 375-377 are straight­
forward. Notice that in the latter we need some inductions to prove
the inequalities, which are 11° formulas. The definition of the positive
measure \i on C { X ) on p.377 is direct. Theorem (8.22), the spectral
theorem, and Corollary (8.23) on p. 378-380 pose no difficulties either.
The consequences of the spectral theorem mentioned on p.380 are also
direct.
The definition of algebra seminorm ||•||, on C (X, F) is simple. Sup­
posing that X is compact and ||*||/ is an algebra seminorm on C (X , F),
the proof of Proposition on p. 381-382 uses an inductive construc­
tion to construct a sequence (/„ ) of functions in C (X , F) and a se­
quence of compact sets ( K n) from any given / 0 and compact set Kq
that supports /o, such that for n > 1, ||/n|| > 0, K n supports / n,
and d i a m ( K n) < 1/n. Here we explain how this can be unraveled. Let
9 i >•••> be the functions in C (X , F) such that each g" has a compact
support X" of diameter less than 1 /n and g" = 1. According to

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
5. NORMED LINEAR SPACES 300

the idea of the construction, f n = fog^ —<7*1, for some f c i , kn- It suf­
fices to construct instead the sequence (kn) such that IJ/o^ > 0
and then let f n = fog^ —gk„. K n = KJ^. The construction is almost
exactly the same as the one used in proving Theorem (1.10) of Chapter
6 on p.220 of B&B (see page 252 above), with the inequality (31) there
replaced by [[/o ^ —<7£,|| > 0- Notice that we don’t have K q D K \ D ...
here as in the original proof. For each n > 0, since ||/n|| > ||/n|| > 0,
we can choose xn such that |/ n (xn)| > 0. So, xn 6 Ki for all i = 0,..., n.
Therefore, (x„) converges to a point x0 6 K q. Then the rest of the proof
can be formalized straightforwardly. The strict condition K q D K i D ...
is never used.
Finally, the rest of the section, Proposition (8.25) to Corollary
(8.30) on p. 382-389 pose no more difficulties.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
APPENDIX B

Constructive Functional Analysis for Quantum


Mechanics in SC

1 . In tro d u c tio n

In this appendix, we will prove that some basics of the theory of


unbounded self-adjoint operators on Hilbert spaces can be developed
within strict constructivism. We will first present the definitions, the­
orems, and proofs in the syle of Bishop’s constructive analysis. Such
a style is perhaps more familiar to readers. For most contents of this
chapter, the translations into strict constructivism are straightforward.
In case this is not obvious, in particular, in case inductions are used
in a proof, we will add notes to illustrate how the translations can be
carried out. These notes will be printed in italics.
We will focus on the topics that are essential to the applications
in quantum mechanics. We will prove the Spectral Theorem for un­
bounded self-adjoint operators, Stone’s Theorem, and the self-adjointness
of some quantum mechanical operators. The quantum mechanical
operators considered will include the position and momentum oper­
ators on Rn, angular momentum operators on R3, Hamiltonians with
some general forms of potentials, including the Hamiltonians of an
N-electron atom in an external constant magnetic field (the Zeeman
effect case), Hamiltonians of an iV-electron atom in a constant electric

301

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
j 1. INTRODUCTION 302
1
field (the Stark effect case), and the Hamiltonians of harmonic and an-
i harmonic oscillators. (These include all but the quantum field theory
operators summarized in [85] p.350.) With these mathematical tools,
the basic framework of non-relativistic quantum mechanics can be de­
veloped. The constructive proof of self-adjointness also implies that
the dynamical equations of the relevant physical systems can be solved
constructively.
This work is a response to Geoffrey Heilman's remarks that con­
structive mathematics may not be sufficient for the applications in
quantum mechanics (cf. [55], [56]). The results proved here are among
those challenged by Heilman. Heilman’s question has not been fully
answered yet, but the work has already shown that what construc­
tive mathematics (actually the finitistic strict constructivism) can do
is quite impressive.
The non-constructive status of the closed graph theorem was once
thought to be the major obstacle in constructivizing the theory of un­
bounded operators (cf. [23]). We found that we can bypass the closed
graph theorem and some other obstacles by defining self-adjointness
this way: an operator A is self-adjoint if it is symmetric and the ranges
of A + i and A — i are the whole space. This means that the re­
solvents (A ± i) -1 exist constructively (as bounded operators defined
on the whole space). It is well known that choosing the appropriate
constructive versions of classical concepts is frequently crucial in the
constructivization of classical results. This definition seems to give the
real constructive contents of self-adjointness, while the common condi­
tion in the classical definition of self-adjointness, A — A*, is perhaps
constructively too weak. W ith this definition, on the one side, we can
use Cayley transformations to prove the Spectral Theorem, while on the

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. INTRODUCTION 303

other side, we still have the Kato-Rellich Theorem on self-adjointness,


and then there follows the self-adjointness of common quantum me­
chanical operators.
On the other hand, it is still unclear if constructive mathematics is
sufficient for all applications in quantum mechanics. There are some
powerful classical techniques whose constructive substitutes have not
been found. Two major obstacles are: (i) Riesz Representation Theo­
rem is not constructively provable; (ii) the decomposition H = M @ M L
is not constructively available for all closed linear subset M of a Hilbert
space H. As a consequence, we have not yet found constructive ver­
sions of some well known classical theorems: Friedrich’s extensions
of positive operators, von Neumann’s theorem on deficiency indices
and self-adjoint extensions, Weyl’s theory on Sturm-Liouville opera­
tors (limit circle/limit point alternatives), some strengthenings of the
Kato-Rellich Theorem, for instance, the W ust’s Theorem and KLMN
Theorem, and their corollaries on self-adjointness. We are still not
clear if these can be constructivized and if they would cause essential
obstacles in physics applications in case some are not constructivizable.
Some applications of these classical results can be constructivized in
a case by case manner, using the classical results as hints. Further,
we have not yet considered some other essential mathematical tools
in quantum mechanics, for instance, spectral analysis and scattering
theory. The question if constructive mathematics is sufficient for de­
veloping the full quantum mechanics is still open. Though, we believe
we have shown that it is sufficient for some basics of quantum me­
chanics and the answer to the whole question is not trivially negative.
Further studies in this area are worthwhile, both philosophically and
mathematically.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. INTRODUCTION 304

Throughout this chapter we work within the framework of Bishop’s


constructive mathematics. We will frequently refer to the definitions
and results in the book Constructive Analysis by Bishop and Bridges
[15], and we will refer to this book as B&B. We also try to keep till the
notations of B&B. B&B is a revision and expansion of Bishop’s earlier
book Foundations of Constructive Analysis [13]. [30] contains some
topics not covered in B&B. The work here is a continuation of B&B,
especially Chapter 7 Section 8 of it. There are some new developments
in constructive analysis after the publication of those books. Here are
some examples. [19] gives some general forms of constructive interme­
diate value theorems. Some improvements of the spectral theorem for
bounded self-adjoint operators are given in [28]. [29] discusses con­
structive versions of the open mapping theorem. Some properties of
compact operators are discussed in [59], [61]. See also [31], [64], [89]
for some other constructive versions of well known classical results. [25]
and [26] contain some results concerning some notions which are trivial
in classical mathematics but of great importance in constructive anal­
ysis, for instance, locatedness. [27] discusses a similar notion: sharp
intersection. [20], [62], [63] study the connections between various
continuity properties. Some more results of meta-mathematical nature
can be found in [24] and [60]. [22] contains a constructive theory of
the representation of preference relations by utility functions.
However, we did not see results concerning unbounded operators on
Hilbert spaces given here published in the literature. In particular, it
seems that attention has not been focused on constructivizing more and
more applied mathematics, especially the mathematical tools that are
essential for modern physics. ([17] is an exception; cf. also [23]). From

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 305

the philosophical point of view this is of greater interest. This is obvi­


ously connected to the much discussed indispensability argument. The
question can be put this way: Do our ordinary scientific beliefs about
this physical world logically and indispensably rely on our willingness to
conceive an infinite domain o f objectively existing, m ind independent,
and mostly infinite mathematical entities or structures, while some of
those structures m ay not even remotely resemble anything in this phys­
ical world? This appendix belongs to the efforts for answering that
question. More about the related philosophical issues are discussed in
Chapter 2 of this dissertation.
Before closing this introduction, let us give some remarks on the sta­
tus of an important theorem in the foundations of quantum mechanics,
Gleason’s theorem. Whether or not it is constructively provable is still
open. Since the ordinary text book expositions of quantum mechan­
ics do not rely on Gleason’s theorem, this does not affect the claim
that constructive mathematics is sufficient for the basics of quantum

mechanics.
Heilman [54] attempted to give a weak counter example to Glea­
son’s Theorem, but the proposition he considered contains redundant
non-constructive ingredients, and hence is not a natural constructive
version of Gleason’s theorem. (We refer to [54] for the definitions of
the relevant notions below.) The proposition Heilman considered is

(i) If / is a frame function on R3, then there exist an orthonormal


basis {"p*, ~q, T*} of R 3 and 3 real numbers m , s , M , m < s < M, such
that

f ( x * ) = m x \ + s x \ + M x 3 , for "5* = x\~p + x2~q + x 3T \

with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. INTRODUCTION 306

The proposed counter example is actually based on the fact that


inequalities among reeil numbers are generally undecidable. So, for
example, even if we do have

f ( x * ) = ax\ + bxj + cx 3, for ~x = Xie? + x{e% + x^e^, | "x*| = 1 ,

for some orthonormal basis { e?, Tf, e?}, as we cannot decide the order
of o, 6, c, we still cannot decide which of { e?, e j , e?} should be ~p,or
~q~,OT~?. But we believe this only shows that (i) is not an appropriate
constructive version of Gleason’s Theorem. It implicitly contains the
decidability of the inequalities between the coefficients as part of the
assertion.
Similarly, if / is given as a quadratic form we may not be able to

find an orthonormal basis { e?, e^, e?} such that / takes a diagonal
form /(af*) = a x \ + &Xj + cx\ on the basis. Because, doing this is
no other than finding normalized eigenvectors of Hermitian matrices.
This cannot be done constructively in general for there is a Brouwerian
counter example as follows. Pick two recursively inseparable r.e. sets,
for instance, K q = {n : (n } (n ) ],= 0}, K \ = { n : (n } ( n ) 1 = 1 }. For
each n, we can construct two real numbers otn^n, otn < 1 , /3n < 1
such that atn > 0 «-» n 6 K q and /3n > 0 «-> n € K \. Clearly,
On > 0 —♦ 0n = 0 and /?„ > 0 —» a„ = 0. Suppose that the existence of
normalized eigenvectors of any 3 x 3 symmetric is constructively prov-
1 0 0
able. Consider the matrix 0 0 Oln . When 0 n > 0 and hence

0 Qn j3»
\
1 f 0 ' f ° \
On = 0 , it has normalized eigenvectors 0 > 1 and 0

0 / , 0 ; . 1 ;

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
1. INTRODUCTION 307

/ \

When On > 0 and hence /3n = 0, it has normalized eigenvectors 0


0

0 '
&
f *5° ^
and So, constructing the eigenvectors of the
2 2
&
2 / \ 2 )
matrix and comparing them with the vectors above we should be
able to decide ->a„ > 0 V -'/3n > 0, that is, a,, = 0 V /3n = 0.
From this, assuming Church’s Thesis, there should be a recursive set,
A = {n : if we decide atn = 0 before 0 n = 0 } , and this set would sep­
arate Kq and K \ , a contradiction.
So, It is not enough just to drop the condition m < s < M in the
claim (i) . We suggest that the constructive Gleason’s Theorem on R 3
must look like:

(ii) Any frame function / on R 3 is a quadratic form, or in


other words, is represented by a symmetric matrix A on
R: f ( ~ x ) = (! ? ,A ~ x ) .

Clearly, this is just Gleason’s theorem in its original form, that


is, any measure on the subspaces of a Hilbert space is represented by
a statistical operator (a trace-one Hermitian operator). Presently we
don’t know if there are counter examples to (ii). It is because we cannot
find a way to construct frame functions which are not already known to
be quadratic forms. Remember that frame functions primarily available
are all quadratic forms. Further it can easily be proved constructively
that if a function on S 2 — {"5* : I"?! = 1, "5* € R 3} can be uniformly
approximated by quadratic forms, it must itself be a quadratic form.
So, the limit process does not help in constructing frame functions
which are not known to be quadratic forms. Unfortunately, we cannot

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 308

prove (ii) either. So the question is still open. (Note: Billinge [12]
presents another version of a constructive Gleason’s theorem. It can
be proved that (ii) here is equivalent to Billinge’s.)

2. S p e c tr a l D e co m p o sitio n s o f U n ita ry O p era to rs

In this section, we construct spectral decompositions for unitary


operators, following the classical proof in [90] p.281. This will be used
in "Section 3 to construct spectral decompositions for unbounded self-
adjoint operators.
An operator U defined on the whole Hilbert space H is called uni­
tary if its adjoint U* (see B&B p.371 for definitions of adjoint and
self-adjointness) exists and UU* = U*U = I . Throughout this section
we assume that U is a unitary operator on a Hilbert space H . Let
V denote the algebra of complex formal polynomials, CkZk, in
variables z and z -1 . We will construct functional calculus for U. For
this we will first define a mapping from V into H o m (H ),the set of
bounded operators defined on the whole space H . Then we will ex­
tend it into a mapping from C ( 5 1), the set of continuous functions
on S l = {z : |z| = 1, z € C}, into H o m ( H ) . After that, one can
easily follow the proof of the spectral theorem for bounded self-adjoint
operators on pp 378-379 of B&B.
We say that £ * =_n c*z* is n- degree. So n-degree polynomials are

also m-degree for m > n . Suppose that p(z) = cfc2* € V and


p(z) = 0 for all z G S 1. Then we have 2xic_n = f s i p (z )z n~1d z =
0, 2xtc_(n_i) = fsi p(z)zn-2dz = 0, (see B&B, Chapter 5). So
p = 0. Therefore, p i(z) = Pa(z) on S x implies pi = pa. For p(z) =
£ £ _ _ n CfcZ* G V , p i—►p(U) = £ CkUk defines a linear, multiplicative
mapping from V into H om (H ). Obviously, p(U)* = P*(U), where

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 309

P*(z ) = H ckz ~k- If p is real on S 1, then p = p* and hence p(U) is


self-adjoint. Recall that an operator A on H is called positive (denoted
as A > 0), if (A x , x ) > 0 for all x € H. We want to prove that
the mapping p h-+ p(U) preserves positivity and bounds, but first we
need a lemma, a constructive version of the lemma on [90] p .118. For

3 = Er=-m &***, define

Coef ( s ) = max{|6_m| , ..., |6m|}.

LEMMA 2.1. If p is m-degree and p(z) > e > 0 on 5 X, then fo r


any 6 > 0 , there exist m-degree q and s such that p = q*q + s and
Coef (s) < S.

PRO O F. Let p(z) = E J tL -m ckZ k ■ As p is real on S l , c*k = c_fc.


We assume that 8 < For each k , |cfe| > 0or |ct| < 6 . We may
assume that there is n > 0 such that |c„| > 0 and |cfc| < 8 for k =
± (n + l) ,...,± m . Let p'(z) = £ ,k=- n ckZk- Then for z € S 1, p \ z ) is
also real and p'{z) > e /2 . Now we show that p' can be expressed as
p' = q*q.
The case n = 0 It is trivial, so suppose that n > 0. Let r(z) =

z np'(z). Then |r(e)| > | on S l , and hence r( z) = 0 implies ||z| — 1| =


d(z, S l ) > |u ;(|) > 0 , where uj is the modulus of continuity for r(z)
. This means, zeros of r are either inside or outside the unit circle.
Similarly, since |r(0)| = |c_„| > 0 , r(z) = 0 implies |z| > |u/(|c_n|).
So the zeros of r are away from 0. Since |c„[ > 0, by the fundamental
theorem of algebra (Theorem(5.10) on p.156, B&B), there exits z\ such
that r(zi) = 0. Since ck = c_*,

(39) r(z) = ^ r ( L ) ' for |2 | > 0,


z

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 310

and therefore r(^-) = 0 as we have \zx\ > 0. Since either |zi| > 1 or
|zx| < 1, we have |p-| < 1 or — > 1 correspondingly. So zx ^ j? are
different zeros of r. Two successive polynomial divisions give
1
(40) r(z) = (z - z*)(z - — )rt (z)

for some r x( z ) (cf. the proof of Theorem(5.10) on p .156, 157, B&B).


Substitute this into (39),

z,r(*) = „ ^ > ( i - * ) • ( ! -
z

= -
‘>(* - *»)(* -

Compare this with (40) we see that ziri(z) = z2^n_1^(ziri(p-))*. This is


similar to (39) and the coefficient of its highest degree term is z xCn ^ 0.
So we can repeat the factorization and finally obtain

r ( z ) = Cn(z - Z l)(z - - ^ ) ...( z - Zn)(z - ^ ) .

For z £ S 1,

p'(z) = z “nr(z)

= b(z - Z i)...(z - z„)(z* - z{)...{z* - z*n),

for some constant b. Since p'(z) > | , we must have b > 0. So, p' = qq*
for q = V b(z - z x)...(z - zn). |

Note. To see that the proof is foamalizable in S C we must unravel


the induction. We need some preliminaries. Recall that a polynomial
p = £?=0 c\zx of n-degree can be coded by a sequence p = (c q , ..., c„) of
complex numbers. For any polynomial p o f n-degree and any complex
number (, there exist a polynomial q of (n — 1)-degree and a complex
number s such that p = (z — q + a, where the equalities between
polynomials are understood as equalities between all coefficients. So,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 311

there exist Q, S such that fo r any polynomial p and complex number f ,

Q(p> £) *■* a polynomial and S( p , £ ) is a complex number, and

p = ( z - £ ) Q ( p , ( ) + s (p , £)-

Further, it is easy to see that, for any polynomial q,

p ( 0 = o — p = (* - 0 Q (p, 0 »

p = ( z - £ ) q -> Q{ p , £ ) = q-

A closer examination will reveal that to compute the coefficients of


Q( p, £) to a degree of precision within ± l / m , it suffices to use p ( N ) ,
£ (N ) for some N depending on m and p ( l) ,£ (1 ), that is,

Q( p, £) ( ™) = G ( m , p ( F ( m , p ( l ) , £ ( l ) ) ) , £ ( F ( m , p ( l ) , £ ( 1 ))))

for some prim itive recursive functions F, G. Then we can see that the
construction is repeatable. That is, there exists D iv such that for any
n-degree polynomial p and £ = £m) a sequence o f complex num­
bers, m < n,

D iv (0,p,£) = p ,

D iv (i + 1, p, 0 = Q (Di v (*, p, £), &+1),

fo r i < m . Intuitively, D iv (i,p, £) is the polynomial quotient of p


by (z —£ i ) ... (z —£ i ) . ( We will not introduce the notation for Ike

residue.)
The following lemma characterizes D iv.

Lemma 2.2. (1) If for all i < m, D i v ( i , p , £ ) ( £ i+i) = 0, then for


any I,

/ < m -> p = (z - £ i ) ... (z - £t) Di v (l , p , £);

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 312

(2) If p = ( z —£ i ) ... (z - ( m)q fo r some polynomial q, then for all


i < m, D i v ( i , p , ( ) ( £ i +i) = 0, and D i v ( m , p , ( ) = q.

PROOF. (1 ) Use an induction on I. Notice that the formula is 11°.

(2) It suffices to show by an induction on I that

I < m —* D i v ( l , p , £ ) = (z - £ i + 1 ) . . . ( z - ( m) q ,

which is 11°. |

Now we can translate the proof o f 2.1. Let r = znp '(z) be as in the
proof. Since r = CnZ2n + ... with |c„| > 0 , by the fundamental theorem

of algebra,

r = Cn(z - 6 ) ...(z - 6 n )

for some complex numbers zl t z2n- We know that for each Zi, |z»| > 0
and |z*| ^ 1. So, we may assume that z \ , . . . , z m are those Zi such that
|z,| < 1 . We m ay assume that m > n, fo r otherwise we can reason
with zm+h ■■■■>z 2n &nd the arguments below still work. For I < m , let

„ . 1 1.
Zi = ( z i , — , . . . , z t, — ).
z\ zi
From the original proof we know that

( 41 ) r ( z ) = z 2nr ( ^ J

for z ^ 0 and therefore r (jr) = 0 fo r i = 1 We want to show

that

I < n -» r = JJ (z —Zi) ^z — j D iv (21, r, Z{ ) .

Then it will follow that

r= (* -* » )(* “

and the conclusion follows directly.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 313

This is a 11° formula, so we can use an induction on I. The case


I = 0 is trivial. So, suppose that I + 1 < n < m and

(42) r = n (* ~ * i) ^ ~ D iv (21, r, Zt) .

Let Z[ = (Z\,..., zi). Then by Lemma 2.2 we have

( ‘ " I ; ) ......

= D iv (I, r, Z{)

= (z - Zl+1 ) ...... (z - z2n) ■

So (z/+1 - .... (zt+i - ^ D iv (2l,r, Zi) (zi+ i) = 0. Since|zf+1| <

1 and |p-| > 1 fo r i = 1 we must have D iv (21, r, Zt) (zj+i) = 0.


From (41) we have for z 0,

- *»*n (j; - « .) ' (^ )‘

= 6 H (z - z{) ^ z ^ - ^ D i v (21, r, Zt) ,

Z* Z*
where b = 1 " 1.
Since this holds fo r all z 0, it must hold as an equality between
two polynomials. Then by Lemma 2.2, it follows from (42) and this
equality that

D iv (21, r, Zt) = bz^n~l)D iv (21, r, Zt) ( ^ j *.

Since b ^ 0, we have D iv (21, r, Zt) = 0- Since zj+1 ^ jr~ ,

D iv (21, r, Zt) = ( z - z l+i ) [ z - ) r'


\ zi+iJ
fo r some polynomial r'. Then by Lemma 2.2 again we see that (42)
holds fo r I + 1. This completes the proof.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 314

For convenience, we use ||j4|| < M to mean that M is a bound of


A , without considering if A is normable or not.

LEMMA 2 .3 . (a) I f p( z ) > 0 on S 1, then p(U) > 0;

P>) If\p(*)\ < M on S 1, then ||p(Cf)|| < M .

PROOF, (a) By the previous lemma, for any t > 0 and 8 > 0
we can find q, s such that p + e = qq* + s and C o e f ( s ) < 8 .So
we have |(s (U) x, ®)| < 8(2n + l ) | | i | | Z, assuming that p is n-degree.
Therefore, ((p { U) + e)x, x) > —8(2n + 1) ||x||2 . 8 and e are arbitrary,
so (p(U)x,x) > 0.
(b) Since M 2 —p*p > 0, this follows from (a). |

Next we extend the mapping V —> Hom( H) to C ( S 1). Let

1ZV = {p 6 V : p(z) is real on 5 1} C C ( S l ).

1ZV is a real algebra and p(U) is self-adjoint for p in TZV. It is easy


to see that TZV is separating in C ( S 1) (cf. B&B p .106). By Stone-
Weierstrass Theorem (B&B p .106), 1ZV is dense in C'(51). Then the
extension is straightforward because of Lemma 2.3.

LEMMA 2 .4 . The mapping TZV —> Hom( H), p i—►p(U) can be


extended into a linear, multiplicative mapping C ( S 1) —►Ho m( H) , f •-»
f { U) , such that f { U ) is self-adjoint and (i) f ( U ) > 0 whenever f ( z ) >

0 on S l ; (ii) ||/(C 0|| — M whenever \f(z)\ < M on S l .

To extend the functional calculus further, we follow the idea in


B&B pp 377-379. So choose an orthonormal basis {e*} of H. Define

p.: C{Sl ) - R by

tif) =
k -1

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 315

Since / 6 C (5 1) is bounded, f { U ) is bounded by Lemma 2.4, and hence


fi is well defined, ? is clearly linear and It is positive by Lemma 2.4,
so ? is a positive measure on 5 1.(cf. B&B p.219) Then we can extend
the functional calculus to L00( 5 1, /i), the space of essentially bounded
measurable functions (cf. B&B p.346, 239). Notice that LO0( 5 1,/x) C

L1( 5 1, ?) as fi is a finite measure. This is summarized in the theorem.

THEOREM 2 .5 . For each ? defined above, there exists a linear and

multiplicative mapping f •—►f ( U) , from L ^ S 1, p.) to the algebra of self-


adjoint operators on H , which extends the mapping TZV —►Ho m{ H)
and is such that
f i ) \ \ m \ \ < \ \ f \ L f o r f e i t s 1, ?) ;

fi) if fn, f € L00( 5 1,^ ) such that \\fn - f\h -> 0, then f n(U)
converge to f ( U ) strongly, where ||/„ —f\\^ = / |/ n —f\dp. is the norm
of h i S 1,?).

The proof is exactly the same as the proof of Theorem (8.22) on


pages 378-379 of B&B, with X replaced by 5 1and V replaced by C (S 1),
so we omit it here.
To construct spectral decompositions for U, we must locate inte­
g r a te arcs on S 1. We first introduce some notations. For z , w € S 1,
\z — w\ > 0, let { z , w ) denote the open arc from z to w along the
positive (counterclockwise) direction on S 1. This description is geo­
metrical, but it can easily be translated into an analytical definition.
We define closed, semi-closed arcs [z,tu], [z,w], [z,w) similarly. It is
easy to verify that for z ^ w, the complement S 1 ~ (z, w) = [to, z\ and
S 1 ~ [z,to] = (u/,z) (cf. B&B, pp 195-196, for the definition of ~ ) .
Other familiar relations also hold: for instance, when (z,u>) C { z ' , w‘),

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 316

Z z', W w\

(z 1, w ‘) ~ (z, w) = ( z \ z] U [w, w')

((z ',w ') ~ (z;, z)) ~ (w ,w ') = [z, to] and so on

For convenience, we also let (z, w) denote the complemented set (cf.

B&B, p.73, p.232) ((z,u>), S 1 ~ (z,™ )) and let X{*,w) denote the corre­
sponding characteristic function. Notations (z,u/], [z, to), and [z , u j ] are
similar. The contexts will make clear if we mean sets or complemented
sets.

A sufficient condition for boundaries of integrable arcs is the smooth­


ness defined as follows:

DEFINITION 2 .1. A point z E 5 1 is called smooth (relative to fi),


if there exists a sequence of arcs (un,wn) such that z E [u„,u)n] C
(Un-i.u/n-x) fo r all n, |u„ - z| -» 0, |u/n - z| -» 0, and /i(u„,u;n) -» 0
as n —►oo. Let pt (U) denote the set o f all smooth points.

First we prove a lemma.

LEMMA 2.6. For any z ,w E S l t \z - w\ > 0, and e > 0, there


exist z', w' E S 1, such that \z' — z| < e, \w' —w\ < t , and ( z ',w ') ,
[z;, to'] are integrable. We can choose z', w' such that [z, u;] C (z', w')
or [z',u>'] C (*,u;) •

PROOF. Choose z i,w x € Cz,uO, such that zx ^ u;X) |zi - z| < e,


|u/x —to| < £, and [zi,ti>i] C (z, u;). We can construct a continuous

function / on S 1 such that / ( z ) = 1 on [zi,u/x], f ( z ) — 0 on [u j , z ],

and / is monotonic and invertible on each of [z, zx] and [u/i,u/]. By


Theorem (4.11) on p.242, B&B, there exists a t E (0,1) such that
( / > t) and ( / > t) are integrable. By the assumption on / , we can

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 317

find z' £ w 1 £ (u>i,w) so that f ( z ' ) = f{ vi ' ) = t. Clearly,


(z', w 1) = ( / > t), [z', w 1} = ( / > i). So both are integrable. |

The following shows that smooth points have the nice properties
we need to construct the spectral families on S 1.

LEMMA 2 .7 . Suppose that z ,w £ pt (U), z f ^ w . Then,

{ z }) is integrable and p. ({ z }) = 0;
(b) (z , w ), [z,w\, (z,w], [z,w) are all integrable and have the same
measure, and furthermore i f w n £ p, ( U ) andw n —►w , then /s((z,u/n|) —»

fcj i / u 6 (z, to), u 6 p,(C/), then p( { z, w\ ) = fi((z,u]) + /a((u,u/j);


(d) if Zn £ pt (U) and z„ -* z, then (i) if Zn+i € (*»»» z) for all n,
we have p((z, z„]) —> S x); (ii) if zn+i € (z, .Zn) /o r a// n, uie Aave

K(*»*i*]) -* °-

PROOF, (a) This follows directly from the definition and Proposi­
tion (3.8) on p.234, B&B.

(b) Choose, by Lemma 2.6, a sequence of arcs (zn, w n) such that


[*.H c [a»,wn] C (zn-ijU /n-i), Zn ►z, wn * w, and (Zn,wn) and
[znjU/n] are all integrable. As m ,n —►00 and m < n, [zTO, Zn) and
(wn, tom] each is contained in any given open arcs containing z and w
respectively. By the definition of smoothness we have

If* ([*m> U>m]) - fi ([ z ^ , tO „])| = ^ ( [ z ,* , Z n )) + p((v>n, tOm ]) - > 0

as m < n and m , n —» oo. So the sequence {/s ([zn,to„])} converges


and hence by Proposition (3.8) on p.234, B&B, [z,w] = An [zn,to„]
is integrable and /i([z„,tun]) —* p( [ z, w\ ) . The rest can be proved
similarly.

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 318

(c) This follows from the fact that the characteristic functions
and X(*,u] + X(u,m] 316 equal on the set (z, u] U (it, u;] U (w , z], which is a
full set because it contains the intersection of the full sets (z, u;] U (w, z]
and (z, u] U (u, z\. (cf. B&B, p.226, p.242)
(d) Similar to (b). |

The following lemma guarantees that there are enough smooth


points.

LEMMA 2 . 8 . There exists a countable set E C 5 1 such that S 1 ~

E C Pi (U ) and hence pt ( U ) is dense in S 1.

PROOF. Notice that the complement of any countable set in S 1

is dense in S 1 (cf. the proof of Theorem (2.19) on p.27, B&B). To


construct E, we will pick two countable sets from the arcs (1, —i) and
(—l , i ) separately and take the union as E .
So, first construct a continuous function / on S 1 such that / (1) = 0,
/ ( —0 = 1) / maps each of the arcs [1, —i] and [—i, 1] onto [0,1], and
/ is monotonic and invertible on each arc. By Theorem (4.11) on
p.242, B&B, there exists a countable set Ti = { ii,ia , of positive
real numbers such that for each t 6 R + ~ Ti, ( / > t) and ( / > t) are
integrable, and for any e > 0, there exists 5 (e ) > 0 such that whenever
t' € R + ~ Tx and |i —1 '\ < 6 (e) we have \y. ( / > t) — p ( / > t')l < £-
Take a positive number r small enough, for instance, r < 1 —/ (c***) •
For each U in Tx we can decide if £» < 1 or ^ > 1 —r. Collect all the
ti determined as < 1 into a set I^. This is still a countable set and is
contained in (0,1). Let

Ex s {*6 (1 ,-*):/(*)€ rtf.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 319

Since / is invertible and monotonic on the arc [1, —i], we can see that
Ei is also countable. Furthermore, we can find u 6 (1, —i) such that /
maps ( l,u ) onto (0,1 —r).
We want to prove that points in ( l,u ) ~ Ei are smooth, that is, for
each z 6 ( l,u ) ~ E j, t > 0, there exists z it z 2 such that z 6 ( z i , z 2),

( 21, 22) integrable, and f i ( ( z i , z 2)) < s. Let t = f ( z ) 6 (0,1 —r) ~ V[.
Since t < 1 —r, we must have t 6 R + ~ IV Pick t 2 € (0,1 —r) ~
such that t x < t < t 2 and t 2 - t \ < S We have similarly t i , t 2 6

R + ~ I \ and hence

M (/ > *i) ~ t i ( f ^ h )

= ! / * ( / > ti) - > *) I + I/* ( / > 0 - /* ( / > <a) I

< £.

Let 21,23 € ( l,u ) — E i be such that f ( z \ ) = £1, f ( z 2) = <2- Then


2 € («i, za) and (21, z 2) < ( ( / > ti) — ( / > *2)) (as complemented sets).
So /i( ( 2i , 22)) < £.
Doing the same for the arc (—1, i), we obtain a countable set Ea C
(—l,ti/) for some w 6 ( 1,*) and very close to i such that points in
( —l,u;) ~ E3 are smooth. Two arcs ( 1, it) and ( —l,u /) cover 5 1. So
points in S l ~ (E i U E2) are smooth. |

Finally, we show that smoothness is independent of the choice of (i.


We actually give a characterization of smoothness independent of fi.

LEMMA 2 .9. z is smooth (relative to n ) if and only if there exists

a sequence of arcs (zn,w n), z € [*»,«*»] C (z»_i,ti/n- i ) fo r all n, a


sequence o f Junctions f n € Cr( 5 1), /„ = 1 on [zn, u/„], f n = 0 on
[tt/n_ i , 2n_x], and 0 < /„ < 1 on [2„ _ i,z n] and [u/n,u/„-i], such that

fn (U) -* 0 strongly.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 320

PROOF. When z g p, (£7), we can take the sequence of arcs in


the definition as (z*, wn). Since p([ zn, wn]) —» 0, we have \\fn\\i —* 0
and hence f n (U) —» 0. On the other hand, suppose the condition is
satisfied. We have ||/n (U) efc|[ —» 0 for each k, and hence p ( f t ) —♦ 0
by the definition of p. By Lemma 2.6, for each n, we can find z'n, w'n

such that z'n € (z ^ z * ^ ), w'n 6 (wn+i,w n), and (a£ , « 0 is integrable.

Since f t = 1 on [z„,u;n], p (C C u £ )) ^ /*(/n) -* So z is smooth


relative to p.. |

For z u z2 6 Pm{U), define £ ( Xllxa] = X(,VliJ] (U). E(tut7] is a projec­


tion, because X(xv,xj] = X(xltxa] B&B, p.371). By Lemma 2.7, 2.8,
and Theorem 2.5, {i?(*llZj] : 21,22 6 P*(U) >2i ^ *2} has the common
properties of a spectral family: For z, u,w,Zn 6 p, (U), 2 / u ,
(i) £ (x ,x » ] —» 0 if 2n+1 6 (2, Zn) for all n and Zn —* 2; £ (l)in] —►/ if
2n g (z , 2„+i) for all n and z„ —> z;

(ii) £(x,x«] -» E m \izn-*u;


(iii) £(u,xi = I 5(x,ui, £(*,«] = £(.,«]£(.,■,], and E^ZiUj = f^(x,u<] d*
£J(w>u], if in 6 (2, 11).
In (ii) we have continuity rather than semi-continuity, because u is
a smooth point.
If zq, z i , ..., zn+1 = zQ are smooth points on S 1, chosen along the
positive direction, then P = (zo, z \ , ..., 2„) gives a partition of S 1. Let

M e s h ( P ) = max{|zfc - z/t+il : k = 0, ...,n } .

Another partition P' is a refinement of P if P' D P.

For g g C ( 5 l ), define

n
9P = E j ( ^ ) X ( x kllH t]'
fc=i

1
R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 321

Suppose that Pm = (zom\ z[m\ ...) is a sequence of partitions such that


Pm+i is a refinement of Pm and Me s h (Pm) —» 0. The domain of gpm,
Dorn (gpm) = Ufc(z£m\ z[+|], is a full set. By Proposition (2.5) on p.224,
B&B, flmDo m( g pm) is a full set. On this full set gpm —> g uniformly,

so II9Pm ~ slli -► 0. By Theorem 2.5, gPm (U ) -♦ g ( U) strongly. Notice


that

m = E s ( 4 m))

We express this fact as

S (U) = Js i g ( . z ) d E„

where the integration is understood as the strong limit above.


For g = <7i + igj, we define g (U) = g\ (U) + ig 2 (U). Then the same
expression holds as well. So we have

THEOREM 2.10. I f U is a unitary operator on H , g 6 C ( 5 1, C),


{£?(XliiJ] : Z \ , zj 6 Pm( U ) , Z \ ^ z 2} is the spectral fam ily associated viith

U , then

9 (V)= f g { z ) d E t.

Especially,
U= / zdE z .
Js1
Furthermore, if A £ H o m ( H ) , then A U = U A iff A£?(XliXj] = J5(XliX,]A
fo r all z\ / Z2 in p, (17).

PROOF. As in the classical case. |

The uniqueness of the decomposition holds in the following sense.

THEOREM 2 .1 1 . Given the notations in Theorem 2.10, suppose


that r is a dense subset of S l , { £ (XliJJ] : zi, z% € T, zx ^ z2} is another
fam ily of projections such that fo r z ,u , u/,z„ 6 T, z ^ u,

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 322

(i) £?('ZZn] -> 0 if zn+l G (z , Zn) fo r all n and zn -» z ; £ ('ZZn] -» / t /

zn G (z, Zn+i) /o r all n and z^ —* z;

(i{) EUm] “ * E l*M u '<


(™) E[u,z} =1- E[z,u)>E[z,w] = E[z,u\E[z,w]i andEU»1= Ei*M+
EUu]>{f W€(Z>U)>
and finally suppose U = fs i zdE': . Then T C p , (U) and =

/ or 2x>22 6 T; Z i £ z2.

PROOF. W e first prove th at / 5i g ( z ) d.Ez exists for all g G C ( S \ C)

and ||/5l g (z)dE'g\\ < \\g\\. For each partition o f S 1, P = (zo, —, z n),

each / G C (S 1, C ), and each x , w e have

(43)

< ll/ll1 11/11’ 11*11*-


i=0 t= 0

From this It is easy to see that if Pn = ( z ^ , . . . ) is a sequence of


more and more refined partitions and Me s h ( P n) —* 0, then for each
X 6 H, E ^ W n)J ^ ( n ) (»),* is a Cauchy sequence in H and hence
\ ' J >*t+i] J n
converges. Therefore ( g fz-n^) E' (n) (n) ) has a strong limit. From
\ ' ' (*. <z i+\.iJ n
(43) ||^ || is clearly its bound.
Next we want to prove that p { U ) = / Si p { z ) dE't for p G TVP. For
this, it suffices to prove
(1) u - = ; s . z'd E '„

(2) }p g ( z ) d E ' , ■j sl f ( o dE ‘, = /s . j ( o / ( 0 if;;.


(1) can be verified directly. As for (2), notice that for any partition

P = ( z o ,Z n )i

E 9 ( * ) £ ( , , , +l]E f ( * ) EU ^ l x = ! > ( * ) / (*.)


>=o »=o i= 0
Let M e s h ( P ) —* 0 and take the limit we obtain (2).

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 323

From (43) we see that / 5i pn (2) dE't converge to fs i g (z ) dE'z strongly


if pn converge to g uniformly. So we have g (U ) = / 5i g (z ) dE'z for all
geC (S').
Now we prove T C pg (U). Suppose that z 6 I \ Choose zn,w n € I \
z G [zn, u>n] c and zn -> z, wn -> z. Let f n be any
function in C ( S 1) such that /„ = 1 on [zn, u/Bj, f n = 0 on [u;„_x, zn_xj,
and 0 < / n < 1 on [zn_x,2n] By the conditions (i) and
(iii) on Z3]}, E ^ w j —♦ 0 as n —►oo. So, for each x,

fn(z)dE'zx
- I L[* n —1 iM/n—l ]
2
< 11/. 2 IIEl( z n - i . u ' r v - i ] 1

- 0

a s n - ^ o o . So / n (U ) -> 0 strongly. By Lemma 2.9, z 6 p, (U ).


Finally, suppose that z\ ± z2, z x%z 2 6 V C p. ( U) t we prove

E U ,n] = Choose un,w n e T such that [zu z2] C [«»,«>»] C


(U n -x^ n -x), Un -► zx, and wn -+ z2. Choose functions /„ 6 C (S 1)
such that / n = 1 on [*x,*a]. /n = 0 on [lUn.z*], and 0 < /„ < 1 on

[«n,*i] and [z2, u/n]. Clearly |/ » - XCn^lIj So> by Theorem 2.5,


f n {U) -» £ (jl|jj] strongly. Since f n (U) = / Si f n (z)dE'tl we need only
to prove f Si f n (z) dEz —* E[XliXjj- From the condition on E' we can see

that

Uz)det + ( U z)iz + J f.{*)ie.+ ( u


- u

■a ,/.( * ) * ? .+ /
A]
. / . w < « ; ) * + £ ; ,.« ! * •
/

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
2. SPECTRAL DECOMPOSITIONS OF UNITARY OPERATORS 324

By the conditions on E \ E 1^ ^ —►0 and -E(ZjiUJn] —♦ 0- So,

-» 0 .

The same is true for f n (z) dE'zx . Therefore, fsl f n (z) dE'z -*

EU * l 1

N o te . Here, an induction is needed to prove that

n—1

>=0

for Zq, ..., z,i 6 S 1 such that Zi 6 (z 0 ,Zn) , Zi S( 2, - 1, z»+i)> for i =


— 1. Notice that an equality x = y between two elements of
a Hilbert space H is equivalent to a 11° formula ||z —y|| = 0, so the
induction is available in SC .
Finally we prove a corollary which will be used later.

COROLLARY 2 .1 2 . If zq € pt (U), then (U —Zq)x = 0 implies

x = 0.

P R O O F . Suppose that (U —z q ) x — 0. Then ((U* — z q ) (U — zo) x, x)


0, that is, Jgi |z - zo\2 d\\Et x \\2 = 0. Let u , w 6 p , ( U ) , z 0 € (u,tz>).
Then we should have Jjwl4j |z — z 0 |J d ||f?*a;||2 = 0. Suppose that |u — z0| >
s and |w — z0\ > e for some e > 0. Then

So E ( WiUj x = 0. Let u —* z q , w —* z q . Then ■£(«,«] —> I - So x = 0. I

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. UNBOUNDED OPERATORS 325

3. U n b o u n d ed O p erators

In this section, we present basic notions and facts about unbounded


operators. We will consider linear operators defined on linear subsets
of a Hilbert space H. The definition is as follows:

DEFINITION 3 .1. A linear operator A on a Hilbert space H is a

pair consisting of a specification of a domain D (A), a linear subset of


H, and a linear mapping from D (A) into H .

The range of A , R { A ) = { A x : x 6 D (A )} , the null space of A,


N ( A ) = { x € D ( A ) : Ax = 0}. A is called injective, if Ax = Ay im­
plies x = y for any x , y £ D ( A ) . If A is injective, we can define
A -1 : D { A ~ l ) = i2(A ). For x 6 R { A ) , there exists y such that
Ay = x. We define A- l x = y. The sum, product of two operators and
the inclusion relation between two operators are defined just as in the
classical case. Some familiar relations also hold, for instance,

( B + C ) A = B A + CA,

A ( B + C ) 2 A B + AC.

(Cf. [90], p.299)


When D (A) is dense in H, we can define A*:

D (A*) = {y : for some y*, (Ax, y) = (x, y*) for all x 6 D ( A ) } ,

A*y = y *, for y 6 D ( A * ) .

y* is uniquely determined by the fact that D (A) is dense in H. Notice


that D (A) = H does not imply D (A*) = H. However, some familiar

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. UNBOUNDED OPERATORS 326

relations still hold:

(ciy = s i ,

(A + B Y d A' + B',

(ABY D B *A \

( A B Y = B'A* if D ( A ) = D (A*) = H,

(A + B Y = A* + B ' if D (A) = D (A*) = H,

B* D A* i£ A d B.

N o te . To express these in the language of SC, we can proceed, as


follows. Given a Hilbert space H , a form A for a linear operator on
H is a pair consisting of a set form D (A) fo r a subset of H and
a term A of the type of a function from D (A) to H. Then ‘A is a
linear operator on H ’ is translated as ‘D (A) C H , D (A) is closed
under linear combinations, A : D (A) -* H, and A preserves linear
combinations’, where the components are translated obviously. With
this convention, assertions about arbitrary linear operators should be
translated into schematic formulas o f SC and we cannot quantify over
linear operators. This is no problem for our purposes in this chapter,
though some tricks as we used in formalizing the theory of integration
might be needed fo r the further development o f functional analysis.

Notice that fo r two linear operators A , B , A = B i s t o mean D (A) =


D ( B ) and A (x ) = B { x ) for all x € D ( A ) . Sometimes the domains
of operators are not explicitly given, but they can always be properly
determined from the context. There are no special difficulties in trans­
lating the contents o f this section into SC . In particular, no inductions

are involved.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
3. UNBOUNDED OPERATORS 327

A is called closed if whenever z n € D (A), xn —►i , Axn —» y, we


have x £ D (A) and Ax = y. Suppose that whenever xn, x'n £ D (A),
xn —* x, x'n —* x, A xn —» y, and A xn —> y', we have y = y'. Then we
say that A is closable and its closure A is defined as

D ( A) - {x : for some xn £ D ( A ) , and some y, xn —» x, and A in —►y }

Ax = y for x £ D ( a ) .

It is easy to verify that A* is always closed. If B is closed and A C B,


then A must be closable and A c B . If A is closed, then A = A.
A is called symmetric if D (A) is dense and A C A*. It is easy to
see that when A is symmetric, A is closable, and A is also symmetric.
The definition of self-adjointness is not exactly the same as the
common classical definition.

DEFINITION 3 .2 . A is self-adjoint if A is symmetric and R ( A ± i) =

H. A is essentially self-adjoint, if It is closable and A is self-adjoint.

Before we proceed, note a useful equation: if A is symmetric, then


for x £ D (A),

(44) \ \ ( A ± i ) x \\3 = \\Ax \\2 + \\x\\2 .

LEMMA 3 .1 . Suppose that D { A ) is dense.

(a)R(A)x = N(A-);
( b ) R ( A ± i f = { 0} * /A* = A;
(c) If A is self-adjoint, then A* = A
(d) If A C B , A is self-adjoint, and B is symmetric, then A = B .
(e) If A is sym m etric and R ( A ± i ) are dense in H , then A is
essentially self-adjoint.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. UNBOUNDED OPERATORS 328

PROOF, (a) is trivial b y definition, (b) follows from (a) and the

equation (44).

(c) By definition R ( A ± . i ) = H. We must prove D ( A *) C D (A).


Let x £ D(A*). Choose y £ D ( A ) such that (A —i ) y = (A* —i ) x.
So (A* —z) (x —y) = 0. Therefore, by (a),

x - y £ W(A* - z ) = R ( A + i f = {0 }.

Hence x = y £ D (A).
(d) follows from (c) directly.
(e) A is closable. By the assumption, for any y £ H there ex­
ists xn £ H such that (A + z)x n -» y. By the equation (44)wesee
that xn —» x and Axn —►y' for some x and y'. So, x £ D ( a ) and
( A + z) x = y. Hence R ( a + ij = H. Similarly, R ^ A —ij = H. |

(c) justifies the term ‘self-adjoint’ in Definition 3.2. In the classical


analysis, for A closed and symmetric, R (A ± z)x = {0} is equivalent to
R (A ± z) = H, but we don’t have a constructive proof. So R (A ± i) =
H is perhaps stronger than A* = A.
We still need to show that for bounded operators defined on the
wliole space H, Definition 3.2 is equivalent to the original definition
of self-adjointness. We need a lemma whose proof is the same as the
classical one.

LEMMA 3 .2 . If D ( A ) = H and ||A|| < 1 — e fo r some e > 0, then


(1 - A )-1 exists, is bounded by e~l , D ((1 - A )_x) = R (1 — A) = H,
and
(1 — A) 1 = 1 + A + A2 + ...,

•where the infinite sum is understood as the strong limit.

Then we can prove the lemma which guarantees the equivalence.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. UNBOUNDED OPERATORS 329

LEMMA 3 .3. If D (A ) = H , A is bounded, and A* = A, then


R ( A ± i ) = H.

PROOF. Suppose that ||A|| < M - e for some e > 0. We assume

that M > 1. By Lemma 3.2, (A + A /i)-1 = (ilfi)-1 ((A /i)-1 A + l)


exists and R (A + A/i) = D ((A + A /i)-1) = H. For any x, because of
A* = A,
||(A + A /i) xj|2 = j|Ax||2 + A/2 ||x||2 .

Therefore (A + A /i)-1 || < A /-1 . Note that

A + i = (l - (A/ —l ) i ( A + A/i)-1) (A + A/i).

Now ||(A /— l ) i ( A + A /i) - 1 1 < (Af — 1) A/-1 < 1. By Lemma 3.2


again, R ( l — ( M — 1) i (A + A /i)-1 ) = H. So R (A + i) = H. Simi­
larly, R ( A —i) = H. |

DEFINITION 3 .3. TAc resolvent set o f an operator A is

p (A) = ( z 6 C : z - A t s injective, (z — A )-1 is bounded, and D ((z —A )-1) = H j ,

and ike spectrum o f A is

o-(A) = C ~ p { A ) .

We need the boundedness of (z — A)-1 to be explicitly specified in


the definition of resolvent set, because we don’t have the closed graph

theorem.

Lemma 3 .4 . If Zo 6 p ( A) , ||(z0 - A ) -1 < M, and \z\ < A/- 1 ,

then zq + z € p { A) .

P r o o f . Since

Zq + z —A = ( j + z (zQ - A )-1) (zo - A) ,

this follows from Lemma 3.2. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. UNBOUNDED OPERATORS 330

We have a characterization of self-adjointness similar to that in the


classical analysis, (cf. [104] p .108, Theorem 5.21)

LEMMA 3.5. Suppose that A is sym m etric. Then


(a) A is self-adjoint iff p ( A ) D C ~ IR.
(b) if there are two points z0, z x with Imzo > 0 and Im zi < 0 such
that R (A + zo) = R (A + z i) = H , then A is self-adjoint.

PROOF. These follow from Lemma 3.4 as in the classical case. |

Let A, B be two linear operators on H. B is said to be A-bounded


if D ( B ) D D (A ) and there exist a, b > 0 such that

||£ x || < a ||Ax|| + b ||x|| for all x 6 D (A ).

a is called an A-bound of B .

LEMMA 3 .6 . Suppose that A and B are operators on H and B is


A-bounded with an A-bound a < 1. Then
(a) if A is closable, then A + B is also closable with D (A + fl) =

D & ).
(b) if A is closed, then A + B is also closed.

PROOF, (a) By the assumption, there exists b > 0 such that ||B x|| <
a ||Ax|| + b ||x|| for all x 6 D (A). So,

||(A + 5 ) x|| < (a + 1) ||Ax|| + b ||x ||.

On the other hand, ||(A + B )x || > ||Ax|| — ||B x|| > (1 —a )||A x|| —
b ||x ||. So

|| Ax|| < (1 - a ) ' 1 ||(A + B ) x || + 6(1 - a )’ 1 ||x ||.

Suppose that A is closable and xn, x^ 6 D (A + B ) = D { A ) such


that x„ -* x, x'n -+ x, (A + B ) x n -*■ y , (A + B) x'n -*■ y'. Then by

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
3. UNBOUNDED OPERATORS 331

the second, inequality above, A xn and Ax'n both converge, and then,
since A is closable, A (x„ —x'n) —> 0. By the first inequality, we have
(A + B ) (x„ — x'n) —* 0, that is, y = y'. So, ( A + B ) is closable.

D (A + B^J = D (a ) follows similarly.


(b) follows from (a) directly. |

THEOREM 3 . 7 . (Kato-Rellich Theorem) Suppose that A and B are

operators on H and B is sym m etric and is A-bounded with an A-bound


a < 1. Then
(a) if A is self-adjoint, then A + B is also self-adjoint;
(b) if A is essentially self-adjoint, then A + B is also essentially
self-adjoint with A + B = A + B and D (A + B ) = D ( a ) .

PROOF, (a) By the assumption, there exists b > 0 such that ||Bx|j <
a || Ax|| + b ||x|| for all x 6 D (A). We may assume that a > 0 and b > 0.
We will prove R ^ A + B ± = H . Now,

(*4 )-
,-i
By Lemma 3.2, we need only to prove that B (A ± < a, which

is a direct calculation.
(b) By the assumptions it can be proved that B is A-bounded with
an A-bound a. So, A + B is self-adjoint and hence closed. Then, by
Lemma 3.6 one can easily verify that A + B = A + B is self-adjoint. |

Two operators A, B are unitarily equivalent if there is a unitary

operator U such that B = U~l AU.

LEMMA 3 .8 . If A and B = U~XA U with U unitary are unitarily

equivalent, then
(a) A is closed if and only if B is;

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 332

(b) A is closable in and only if B is closable and in that case B =


U~l AU;
(c) A is symm etric if and only if B is;
(d) A is self-adjoint if and only if B is.

PROOF. The proof is straightforward. |

4. T h e S p e c tr a l T h eo rem

We prove the Spectral Theorem for unbounded self-adjoint opera­


tors in this section. We will use the Cayley transformation method.

LEMMA 4 .1 . If A is self-adjoint, then the Cayley transformation


of A,

U = ( A - i ) { A + i)~1 ,

is a unitary operator on H . 1 —U is injective, D (A) = i2 (l — U), and

A = i(l + U )(l-U )~ 1.

PROOF. This is the place where we need the stronger definition of


self-adjointness. By that definition, U, which maps each (A + i ) x to

(A —i ) x, is defined on the whole space H. The rest of the lemma can


be proved just as in the classical case (cf. [90], p.321-322). |

LEMMA 4 .2 . If U is the Cayley transformation of a self-adjoint


operator A, then 1 6 P i ( U ) .

PROOF. Choose u/n,Un 6 p, (U), wn -+ 1, ix« -♦ 1, 1 € [u;n iUn] C


(u/n—i,Un—i)- Take f n 6 C ( 5 1), f n = 1 on [tn„,Un], / B = 0 on
[tin_i,ti/n_i], and 0 < f n < 1 on 5 1. We will prove that f n(U) —» 0
strongly. If x € R (1 — U ), then x = (1 — U ) y for some y. fn (U ) x =
f n (U) (1 —U ) y . It is obvious that / n (z) (1 — z) —* 0 uniformly on

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. THE SPECTRAL THEOREM 333

S 1. So, f n ( U ) ( l — U) —* 0 strongly. Hence f n ( U ) x —►0. Since


i2 (l — U) = D (A) is dense in H and f n (U ) is uniformly bounded, we
must have f n (U) —> 0 strongly. |

tp i ►elv is a continuous and invertible mapping from (0, 2 t) onto


S l —{1}, and A i—►2 arccot (—A) is a continuous and invertible mapping
from ( - o o , + 0 0 ) onto (0 ,2x), so A r (A) = ea*rccot(- *)* is a continu­
ous and invertible mapping from ( —oo, -t-oo) onto 5 1 — {1} such that
when A increases, r (A) moves along the positive direction on S 1 —{1}.
Let

Pm(A) = {A € R : r (A) € pt (C^)} •

We also call points in p, (A) the smooth points of A. Note that p, (A)

is dense in R. Then we can construct the spectral family on p, (A)


corresponding to the spectral family on p , ( U ):

Ex = E(i tr(A)], for A € p, ( A ) .

By Lemma 4.2 and the properties of ^(*ll4,j we can easily verify that
E \ have the common properties for a spectral family:

(i) Ex —<►0 as A —* —oo, Ex -* / as A —» +oo;


(ii) -* Ex as A' —» A;
(iii) Ex'Ex = E x Ey = Ex when A < A'.
We first construct the decomposition of A restricted to R (Ex2 — Ext )•

Lemma 4 .3 . For Ax, Aa 6 p, (A), Ax < A3;

(a) R{ Ex 2 - E x , ) C D (A), A{ Ex 2 — Ex, ) D (Ex, - EX, ) A ;


(b) A( Ex a — Ex,) is bounded, self-adjoint, D ( A (Ex, - Ex,)) = H,

where the integration is understood as the strong limit.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 334

PROOF. Consider z i , z 2 € pt ( U), {zy,z2} C S 1 - { !} , f (z) =


We have

(1 - U) (1 - U - ) f ( U ) = (1 - U * ) f ( U ) ( l - U ) = E(zuZi]

by Theorem2.5. So, R C R( 1 — U) — D (A), and (1 - U)~l E(ZliZ7] D


E(zl)Zj] (1 — U)~l . And therefore AE(ZliZ, j D E(Zlt.3]A. (a) follows when
t(Ax) = zu t ( Aa) = 22-

Take g { z ) = € L00( 5 1). Then g( U) is bounded, self-


adjoint, defined on the whole space H, and (1 — U) g (U ) = i (1 + U) E^Zl il3j.
On the other hand,

(1 - U) A B f , ^ = (1 - U ) i (1 + U) (1 - U )-'

= i ( l + U) E ( n ,„l-

Since (1 — U) is injective, we have AE(ZuZJi = g(U). So AE(ZliZi\ is


bounded, self-adjoint, and defined on the whole space H. Following
the proof of Theorem 2.11, it can be seen that JjZl 2jj i^ ^ d E z exists as
a strong limit and (1 - U) Jj^^, t jtjd E , = i (1 + U) £ („ ,„]. So, we
have g (U) = /[ZliZJ] i\f^ d E z . By the familiar trigonometrical equations
(cf. Proposition(7.13) on p.59, B&B), = A. Hence (b) follows
when r(Ax) = z i} t (A2) = z2. |

To generalize the decomposition we must define Ad E \. For the


latter use, we need a more general definition

DEFINITION 4 .1 . For g € C ( R , Q , the set of continuous Junc­

tions from R to C, we define

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 335

F o T x e D ( j £ g ( X ) d E k) , it can be verified that lim*,— .«> f x( g (A) dE \


Aa-*+oo
exists and we define g (A) d E \x as the limit.

Then the following corollary can be proved straightforwardly.

C o r o l l a r y 4 .4 . ( a ) x e D ( /+ “ g ( X ) AS*) |s (A)|2< f p Ai ||3


exists, and in that case,

II r ° ° g (A) dExx f = / +°° Ig (A)|2 d \\Exx \\2 ,


\\J —’OO I J —00

where the right hand side o f the equation is understood as the limit
of Riemann-Stieltjes integrations, linul_ - 0o | < 7(A)|2 d||.£*x||2. (cf.
Aj—»+oo
B&B, p.246)

(b) I f g is bounded and ||^|| < M , then D (j*™ g (A) dEx) = H and
S -» 5 (A) dEx^ < M , and if h £ C (R ,C ) is also bounded,then

g(X)dE^ = J*“ g(\)h(\)dEi.

Then we have the Spectral Theorem.

THEOREM 4 .5 . Suppose that A is self-adjoint, Ex the spectral fam ­


ily on p, (A) constructed above. Then
f+oo
+00
/ *00
MEx,

and for any B £ Horn (H ), B A C A B i f and only if B E X = EXB fo r


all Ex.

PROOF. Choose Am £ p , (A), m — 0, ± 1 , ± 2 , ..., such that Am < An


for m < n, Am —> +oo as m —* +oo, and Am —* —oo as m -♦ —oo.
For any x, let xm = [Exn - E x_m) x, m = 1 ,2 ,.... Then xm -* x.
By Lemma 4.3, Axm = A (Exm — E x-n) x = f£™n XdExx. Now, if
x £ D XdExx), limm_oo Axm exists. Since A is closed, x £ D (A)
and Ax = limm_ O0 ~ XdExx = XdExx. Thus A D XdExx.

R ep ro d u ced with p erm ission of th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
4. THE SPECTRAL THEOREM 336

On the other hand, if x 6 D ( A ) , by Lemma 4.3, f£™m \d E \x =


A ( £ Am - Ex. m) x = (Exm - Ex_m) Ax. So, l i m ^ * XdExx ex­
ists. That means x E D AdExx}- So, A = / * “ XdEx-
Now suppose that B A C AB. For each x there exists y, x =
(A + i ) y . Let U be the Cayley transformation of A. Then Ux =
(A —i ) y and B U x = (A - i ) B y = UBx. So U B = BU. By Theo­
rem 2.10, BE( i it] = E(iit]B for all z 6 p, (U). That is BEx = ExB.
Conversely, suppose that BEx = ExB for all E\. We can also infer
by Theorem 2.10 that B U = UB. Then for x 6 D (A), x = (1 - U) y
for some y and Ax = i (1 + U ) y . Then A B x = i ( l + U) B y = B Ax .
That is B A C AJ3. I

As a corollary we have decompositions of bounded operators.

COROLLARY 4 .6 . I f A is self-adjoint, Ex is the spectral fam ily for


A, and ||A|| < M, then ( —0 0 , - M ) U (M , + 0 0 ) C p, (A), Ex = 0 for
X < —M , Ex = I fo r X > M, and A = f!?M„ XdEx for any M', M" >
M.

P R O O F . We need only to prove that E x i — E x l = 0 for M < X i < Aj

or Ai < A2 < —M. Consider the first case. Let x 6 R ( E x t —E x v ) -


Then x 6 D { A ) , Ax = f £ XdExx, and ||Ax||a = / £ X2 d\\Exx \\2 >
AJ ||x||2. So, M 2 ||x||2 > AJ ||x||2. Since M < Ai, we must have x = 0.

That is Exj — Exx — 0. |

Uniqueness holds in the following sense.

THEOREM 4 .7 . Suppose that h i s a dense subset of R, E'x, A 6 A,


is a fam ily o f projections such that
(i) E'x-* 0 as X —►—00, E'x —♦ I as X —» + 0 0 ;
(ii) E'x, - E'x as X' - A;

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 337

(Hi) E'XE'X, = E'X,E'X = E'x when A < A',


and suppose that A = / * “ AdEx. Then A C p, (A) and E \ = Ex
for A € A.

PROOF. Let r = T (A) C S1 ~ {1}. For zi, z 2 € I \ z\ ^ z2, we can


choose At, A2 £ A such that r (Ax) = zi, r (A2) = z2, and decide Ai < A2
or At > A2. We define E[ZuZi] = (E'Xj - E'Xl) if Ax < A2 and E[Zi ia] =
-/ — (E'Xj — J otherwise. Then It is easy to check that the family
has the properties mentioned in Theorem 2.11. By the proof

of the theorem, U' = Jsi zdE'z exists. It is easy to verify that U' is a
unitary operator. Notice that E[ZUZ7] -* 0 as z\, z2 —» 1 and 1 € (zi, z2).
From the proof of 2.12 we can see that 1 —U' is injective. Define A' =
i (1 + U') (1 — U')~l . Follow the proof of Theorem 4.5 we obtain A 1 =
J*-<x M E x = A. Then we can recover U' = (A! — i ) ( A' + i ) -1 = U. So
we have two decompositions of U. The conclusion then follows from
Theorem 2.11, the uniqueness theorem for U. I

As an application of the Spectral theorem, we discuss the spectrum


of an operator. As in the classical case, we can characterize p { A ) by
the spectral family E \ , A 6 pt (A) (cf. [104] p.200, Theorem 7.22).

LEMMA 4 .8 . Suppose that A is self-adjoint and E \ is the spectral


fam ily of A . Then fo r each A £ R., A € p ( A ) if and only if there exist
Ai, A2 £ p, (A) such that Ai < A < A2 and E \x = E \7.

PROOF. First suppose that Ao £ p { A ) and ||(Ao — A ) -1 < M.


Then ||(Ao — A )x || > jjf||x|| for any x £ D ( A ) . Choose Ax,A2 £
p, (A) such that Ai < Ao < A2 and A2 — Ai < We need to prove
R ( E \ , - ^ A t ) = {0}. Let x £ R { E \ 7 - E \t ). Then x £ D { A ) and

||(A „ -X )x ||* = < ( j j y ) 2 11*11 '■ So we hove

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 338

||i|| < 2^- ||as||.Hence x = 0. Conversely, suppose that Ax, A2 €

p, (A), Ax < A0 < A2, and E\ , = E \x. Let e = min{A2 — A0, A0 - Ax}
and z = A0 + ei. By Lemma 3.5, z 6 p (A). For A < Ax or A >
A,, |z - A|! > 2£j . We have ||(* - A ) * ||J = \z - A|J i ||£ i * ||J >

2£! ||x||! , since Ex, = Ex,- So, j|(z - yt)_1|| < Now |ei| < i/2£, by
Lemma 3.4, we have A0 = z —ei 6 p (a4). I

. As an example, we compute the spectrum of a self-adjoint operators


on a finite-dimensional space H. Note that any linear operator defined
on the whole space H is bounded. Finite-dimensional spaces are locally
compact. By Proposition (6.30) on p.110, B&B, and Theorem (2.6) on
p.308, B&B, all subspaces of H are also finite-dimensional. Let A be
a self-adjoint operator on H. Suppose that H is n-dimensional and
choose an orthonormal basis {e x ,..., en} of H. A is represented by an
n x n matrix on this basis, also denoted by A. Let |B | denote the
determinant of an arbitrary matrix B. Then, as in the classical case,
we have

T h eo rem 4.9. A € a (A) i f and only if |A/ - A\ = 0.


PROOF. First, by a pure algebraic computation, we see that when
|A / —A\ ^ 0 , the matrix (AI — A) has an inverse, and hence the op­
erator (AI —A ) has a bounded inverse defined on the whole space H,
that is, A 6 p ( A ) . On the other hand, when A € p( A) , the opera­
tor (AI — A )-1 is represented by a matrix B. It is easy to verify that
products of operators axe represented by the products of correspond­
ing matrices, so we have (AI — A ) B = I. Finally, |(AI — A) B\ =
|(A / —A)| \B\ can be proved by pure algebraic computations. So we
must have |(A7 — A )| / 0. Hence we have proved that A € p ( A)

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 339

iff | (AI — A) |^ 0. The conclusion, follows from the continuity of


|(A / — A)| as afunction of A. |

Since a (A) C R, zeros of the equation |(AI —A)| = 0 are all real.
By the Fundamental Theorem of Algebra (Theorem (5.10) on p .156,
B&B), we can find n real numbers A1(..., An such that

|(A/ — A)| = (A — Ai) • • • (A — An) .

Then we have another characterization of a (A):

THEOREM 4 .1 0 . A 6 O’(A) if and only if fo r any e > 0, there

exists a A* such that |A — A*| < e ; or equivalently,

* ( A ) = n ?=1 u ?=1 1 1
‘ k' ifcj

PROOF. The sufficiency is clear by the continuity of |( A /— A)|.

Conversely, when A 6 a ( A ) , for each t, we can decide if |A — A»| < e or


> | . If |A - Aj| > | for all i = 1, ...,n , we would have |(AI — A)| ^ 0,
a contradiction. So we can find one Ai with |A — A»| < e. I

Certainly { A i,..., A„} C o '(A ). However, generally we cannot claim


that <r(A) = {Ax,..., An}. Similarly, we cannot generally claim that
the eigen-space of the eigen-value A< exists. Indeed, for any e > 0,
i = l ,...,n , we can find A, A' 6 A < A, < A', A' — A < e,
and an orthonormal basis for the subspace R ( E X> —Ex), which is the
eigen-space of the eigen-values in (A, A') in the classical sense. But
we cannot guarantee that E x> — E \ converges as t —» 0. Note that
N (A,- — A) is a linear subset of H, but we may not be able to find a
basis for it, for it may not be located and hence fail to be a subspace.
We will see that this is due to our inability in deciding At- = Ay V A* ^
A,-, i , j = 1, ...,n , for, in case the A,’s appeared in the eigen-equation

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
4. THE SPECTRAL THEOREM 340

|(A / —A)| = (A — Ai) • • • (A — An) = 0 are mutually distinguishable,


everything will be the same as in the classical case (Theorem 4.12
below). First we need a lemma. For any self-adjoint operator B on
any (not only finite-dimensional) space, a point A 6 a (B ) is called
isolated if there exists t > 0 such that (A —e, A + e) ~ {A} C p ( B ).

Lemma 4 .1 1 . If B is self-adjoint, E \ is the spectral fam ily of B,


and A0 6 o { B ) is isolated, then N (A0 - B ) = R ( E \ 0 +e, E \ 0- c) for
sufficiently small e > 0.

P r o o f . Note that for e sufficiently small, Ao ± e 6 p, (B ) and

| | ( A „ - B ) i | |: = [ ^ ( X o - X f d W E x z W 3
J —ao

= e 2 \ \ ( I - E Xo+' + EXo- e)x\\2 .

So, if x 6 N ( A0 — B ), then (I — Ex0+t + E \0..e) x = 0, that is, x 6


R ( E \ 0+e, Exo-e)- Conversely, suppose that x 6 R (E \0+e, E \0. e) for all
sufficiently small e. Then
. 0+* . .
ll(Ao - 5 ) x || = j (Ao - A) (£||£as||
JAo—c

Since e can be arbitrary small, we have x € IV (Ao — B ). |

Back to the finite-dimensional case, we have

THEOREM 4 .1 2 . If A is a self-adjoint operator on a finite-dimensional


space H , and

lAZ-Al^fA-AxJ-.^A-A*),
then

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout perm ission.
4. THE SPECTRAL THEOREM 341

(a) if fo r each j = either Xi = Ay or A; ^ Ay, then


N (Xi — A) is a finite-dimensional subspace of H with dimension > 0;
(b) if fo r all i , j = 1, ...,n , either A, = Xj or Xi Ay, then fo r any
A, A 6 o'{A) iff X = X{ fo r some i = and if A ^ , A ,-m are the
mutually unequal representatives from A i,...,A n, then

H = N (Ait - A) ® • • • © TV(A,m — A ) ,

A = A^ P i -I— -Ai m P m ,

where Pj is the projection onto TV (A(j. —A j .

P r o o f , (a) If the condition holds, we can find e > 0 such that


Ay = Ai or |Ay — Ai| > e for all j . Then for A 6 (Ai —e, Ai + e) ~
{Ai}, we have |Ay - A| > 0 for all j , and hence | A / - A\ ^ 0. So

(Ai —e, A, + e ) ~ {Ai} C p ( A) , that is, At- is isolated. By Lemma


4.11 TV(Ai — A) is a subspace of H and consequently It is also finite­
dimensional. If the dimension of TV(Ai — A) is 0, we would have E \i+e =
E xi-t by Lemma 4.11, and then Ai 6 p ( A ) by Lemma 4.8, contradicting
to Theorem 4.9. So N (Ai — A) has a positive dimension.
(b) If the condition holds, we can find £o > 0 such that Ay = Ai

or |Ay - A,| > e0 for all i , j . Now, suppose that A € o { A ) . Find, by


Theorem 4.10, an i such that |A — Ai| < Then for any positive
e < by Theorem 4.10, there exists a Ay such that |A - Ay| < e.
However, we must have Ay = Ai. Since e can be arbitrarily small,
A = A,. So o-(A) = {A i,...,A n}. The rest follows from (a), Lemma
4.11, and the Spectral Theorem. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. STONE'S THEOREM 342

The representation in (b) above is exactly the same as the clas­


sical case. It seems that indistinguishable V s arise only in artifi­
cial constructions. In other word, we expect that operators on finite­
dimensional spaces appear in natural physical contexts ail satisfy the
conditions in Theorem 4.12(b). More generally, in quantum mechan­

ics, the spectra of an operator A are supposed to be the values of the


observable corresponding to A. Any realistic observation can be per­
formed only up to a finite precision. So, the first thing an observation
can determine is that the value of the observable is in a interval (Ai, A2)
of positive length. Since p, (A) is dense in R, there is no harm in as­
suming that Ax,A2 E Pm(A) . After the observation, the state of the
physical system is determined by the projection E \ 2 — E \l , which ex­
ists constructively. We can assign an exact value Ao to the observable
only if the values are discrete. We expect that, in that case, we can
prove that A0 is an isolated spectral point. Then, by Lemma 4.11, we
can still construct the eigen-space of A0.

N ote. There are no novel applications o f inductions in this sec­


tion. All reasonings are based on the previous results and the laws of
intuitionistic logic.

5. S to n e ’s T heorem

We constructivize the proof of Stone’s Theorem, following the clas­


sical presentations in [104] p.220-223 and [30] p. ?.
Recall that a family of unitary operators on H , {U ( t ) : t E R } ,
is a (one-parameter) group if (0) — I and U (a) U (t) = U (s + 1 ),
for s , t E R. The group is strongly continuous if for each x E H,
t 1—►U (t) 1 is a continuous function from R into H. It is obvious
that { U (t) : t E R } is strongly continuous if, for each x, t t-» U (t) x

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. STONE'S THEOREM 343

is continuous at t = 0, that is, for any e > 0, there exists S > 0 such
that ||(17 (i) — / ) x || < e whenever |t| < S. Furthermore, in that case,
t 1-> U {t) x is uniformly continuous on R. Note that

11(0- («) - / ) x|| = ( ( / - £ /(()) x, x) + ( ( / - U (-< )) x, x).

So, { U (t) : t E R } is strongly continuous if 1 i-» (U (t ) x, x) is continu­


ous at t = 0 .
The infinitesimal generator of {17 (i) : t 6 R }, limt_o 7 (U (t ) - I),
is defined as

•0 ( l i m j ( ^ ( t ) - - f ) ) = { x 6 » i l i m i f U ' t i ) - / ) ! exists J ,

(itjg \ (V (t) - / ) ) x = lim ( i (U (t) - /):x ) for x € D (lim | (V (t) - / ) ) .

The following theorem says that every self-adjoint operator is an infin­


itesimal generator of a strongly continuous unitary group.

THEOREM 5 .1. Suppose that A is a self-adjoint operator on H,


E \, X € p, {A), is the spectral fam ily o f A . For t € R define

U( t ) = e't A = r ^ d E x .
J —ao

Then {U ( t ) : t G R } is a strongly continuous unitary group with the


infinitesimal generator iA and U ( t ) x E D (A) f or x 6 D (A).

PROOF. By Corollary 4.4, U (f) is bounded, defined on the whole


space, and { U (t) : t 6 R } is a group. It is also easy to verify that U (t)
is unitary.

We prove that { U ( t ) : t 6 R } is strongly continuous. Let x E H.


For any e > 0, first choose Ai,Aa € Pi ( A) , Ai < Aa, such that
||( / — E x 3) x \ \ 2 -I- ||£ Alx ||2 < y , and then choose 6 > 0 such that

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. STONE’S THEOREM 344

|e,tA — 1 < 2^||x^|3+1^ whenever |f| < S and Ax < A < A2. Then, as

1*1 < 6,

||(£/(f) - 7 )x ||’ = + £ " ) |e*» - 1 <f||E»x||’ + P ’ \eia - l \ ' d W E n f


fXi
JXl
.2
< 4 ( | | ( / - £ A , ) * f + ||B»1*||, ) + 11(25*,-15Al) *1

< £ J.

So, { U (t) : t (= R } is strongly continuous.

Now we prove i A = lime_o 7 (U (t ) — I). Let 1 6 D (A). For any


e > 0, choose Ai, A2 6 p, (A), Ax < A2, such that

xll’ < |
{ £ . +r h ,dt*
Note that

costA —1 ,s in tA -f A \,
+ 1 ------- ------ I A
tx tx
So, we can choose 6 > 0 such that |j- (e“A - 1) - i X 2 <

whenever t < 6 , t ^ 0, and A 6 [AX,A2]. (cf. Definition (2.8) on p.133


and equation (7.11) on p.58) Furthermore |j- (e‘tA — l j | < |A| for all
A 6 R and t ^ 0. Therefore, as t ^ 0 and |f| < 5,
1 2
-{U {t)-I)x-iA x
y

= ( £ +jT) I?(eiU■^ ■*Af * +jC Ii (e'“ *x) ■“f ■


i||£ii||:
s 4 ( £ > C ) + 2 ( i i ^ F + r y l l (£” - * “ > • “*

< £2

So, lime_o 7 {U (t) — I) x = iAx. It remains to prove D (lime_o 7 ( U (t) — I)) C


D (A ). Suppose that lim«_o 7 {U (*) — I) x exists, which means, for any

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. STONE'S THEOREM 345

e > 0, there exists 6 > 0 such that for any Ai, A2 € p, (A), Ax < A2,

II . i - 2
JIIW
/Ai

d\\Exx\

whenever |t| < S, |t'| < S. Let t' —» 0. Since p (eli'x — l) —♦ iA


uniformly for A € [Ai, A2], we have

/ * ' It ( ' i,A _ 0 - a f ^ j-

Fix a t such that |t| < 6 . j (e‘‘A — l) is bounded for A 6 R, so we can


find M > 0 such that whenever A2 > Ax > M or Ai < A2 < —M ,

d\\Exx\\2 < T

So, for any e > 0, we have found M > 0 such that whenever Ax, A2 6
pt (A) and A2 > Ax > M or Ax < A2 < —M, we have

fAj
f 3 |A|2 d\\Exx \\2
JIX
Xii
/ .1. I1 2 «ln I1 |2 >

<£'

That means x 6 D (A).


Finally, for any Ax, A2 € pt (A), Ax < A2,

/" ’ lAl'rfllfWTCf)*!!1 = /** |A|a < J ||t f ( i) f \* ||J = / * ’ |A|* d l l W -


J Ai «/Ai J ai

So, x 6 D (A) if and only if U (t) x 6 D (A). |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout perm ission.
5. STONE’S THEOREM 346

Stone’s Theorem asserts that any strongly continuous unitary group


can be represented as in Theorem 5.1.

THEOREM 5.2. If { U ( t ) : t £ R } is a strongly continuous unitary

group on H , then there exists a unique self-adjoint operator A on H


such that

U ( t ) = eltA = f +°° eitXdE \,


J —ao
where E \, A 6 p, ( A ) , is the spectral fam ily of A.

PROOF. The uniqueness follows from Theorem 5.1. We prove the

existence. Define

A = -iH m i(C /(i)-/),

we want to prove that A is self-adjoint. For each n = 1 ,2 ,..., x 6 H,


define

Tnx = [ e_™[/* ( s ) xd s , Snx = f e '^ U ( - s ) x d s ,


Jo Jo

where

r e~ntU { ± s ) x d s = Um e~ntU ( ± s ) x d s ,
Jq Af—
*00 Jo

and the integrations on [0, M] are understood as the limits of partial


Riemann sums. All the limits clearly exist because \\U (s) x|| is bounded
by ||x||. Tn and Sn are linear operators defined on the whole space H
and bounded by n -1 . Note that

- r e - ,M(tf(a -M )-C /-(s ))x (fc


t Jo

= - r * - « - « U {s)xds-1
- re-~U(s)xds
t Jt t Jo
_ 1 too 1 rt
= jf e ' ^ U (a) xds ~ ^ j q e~ntU (a) xds.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
5. STONE’S THEOREM 347

As t -» 0, ^ - n, / “ e - ^ U i a ) xds - Tnx, J J* e~ntU (a) xds -> x.


So, we have

ATnx = —inTnx + ix.

Similarly,

ASnx — in S nx —ix.

As a consequence we have (A + i) 7 \x = ix and (A —i) 5 :x = - i x . So,


R ( A ± i ) = H.

Next we prove that D (A) is dense. Since D (A) D {Tnx : x E H, n = 1 ,2 ,...},


we only need to prove that Tnnx —» x as n —♦ oo for every x E H. As
{17 (t) : t E R } is strongly continuous, for any e > 0, there exists 8 > 0
such that ||(17 (s) - I) x|| < f when |s| < 8 . Note that / “ n e-n*ds = 1.
So,

\\Tnnx - x11 = / Tie-™ (U (s) - I) xds


\\Jo

< | + 2 ||x ||e —*.

Clearly, ||Tnnx —x|| < e as n sufficiently large. So Tnnx —> x.


It remains to prove A C A*. For x ,y E D (A),

Let t —* 0 we have (A x , y ) = (x, Ay). So A C A*. Put these together,


we conclude that A is self-adjoint.
Finally, it remains to prove U ( t ) = e,tA. Let V (t) = ettXd E \,
where Ex is the spectral family of A. Fix an x E D (A) and let

f (t ) = || V ( t ) z - V (t) *||* = 2 Hxll1 - 2 Re(£f (() x, V (t) x).

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 348

We want to prove that f ' [ t ) = 0 for t € R. We have

/(* )-/(* )
3 —t

= - 2 Re ( (U (t) U (3 ~ t ) - I X} v x) + (t ) z> V V (s ~ t h I x)
^ 3 — 1 3—t

Let s —* t. By the definition of A, —* iAx. Since U (t ) is


uniformly bounded for t € R, we see that U (t) —» i U (t) Ax
uniformly for t 6 R. Similarly, by Theorem 5.1, V ( t ) x —»
iV (<) A i uniformly for t 6 R. Then, because of the continuity of the
inner products, we can see that

ISfl S- l S l ) _> _ 2 Re ((iU (t) A x, V (t ) x) + (U (t ) x, iV (t ) Ax))


3 —t

uniformly for t 6 R. By the definition of A and Theorem 5.1 we can


easily verify that U ( t ) A = A U (t) and V (<) A = A V (t ). Then, since
A is self-adjoint,

(iU (t) Ax, V (t ) x) + (U (t ) x, iF (t) Ax) = 0.

So, we have —►0 uniformly for t € R. That is / ' (t) = 0.


(cf. Definition (5.1) on p.44, B&B) Since / ( 0 ) = 0, we must have
f ( t ) = 0 for all t 6 R. (cf. Theorem (5.6) on p.48, B&B) Therefore
U ( t ) = V (t ) = eltA on D (A). D (A) is dense, so U ( t ) = V (<) = eltA

N o te . This section poses no problems fo r the formalization in S C


either.

6. F ourier T ran sform ations a n d S o b o le v S p aces

This section contains tools needed for proving the self-adjointness of


quantum mechanical operators in the next section. We will follow the

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 349

classical presentations in [104] Chapter 9. First we need some lemmas


about the Lebesgue integration on R m.
For a,b G Rm, a< <bi,i = 1, we denote

[a, b] = {z G Rm : cti <Xi < bi, i = 1 , m }.

(a, b], [a, 6), (a, b) are defined similarly. These are called intervals. For
a real number M > 0, we denote

[ - M , M] m = { x G Rm : - M < x, < M, i = 1, .

For r > 0, we denote

B m (x, r ) = { y G R m : \ y - x \ < r } ,

S™ (x, r ) = { y G R m : \y - x\ = r } ,

where |x| = (xj H 1- x^,)5 is the standard norm in Rm. We usually


omit the superscript m, and write B (0 ,r), 5 (0 ,r) as B (r), S (r).
The theory of integration developed in the Chapter 6 of B&B is
presented for real functions. It can be extended to complex functions
straightforwardly: / = /x + t/a is /-integrable when both f i and / j are
integrable, and we define 1 ( f ) = I ( / i ) + i l (fi). Other notions, for
instance, measurability, convergence in measure, convergence almost
everywhere, and convergence almost uniformly and so on, are extended

similarly.
For a, b G Rm, if / : [a, 6] —» C is continuous, the Riemann inte­
gration J]a 6j / (x) dx can be defined as in the case of one-variable real
functions and It is easy to show that

f f(x)dx= I ■( f ( x ) dx\ • • • dxm.


J[a,b] Am Jav

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 350

For f E Co (R m, C ), the space of continuous complex functions on Rm


with compact support, we define

vm (/) = f f d x = [ f (x) dx,


J •/[o16]

where [a, b] is some interval supporting / . This is a positive measure


on Rm (B&B p.219). Let fin be the complete extension of v m (B&B
C h.l, Sec.2). Hm is the Lebesgue integration on R m. We also write

Pm ( / ) as //«**• Forp > 1,

Lp (R m) = { / : | / | p is Lebesgue integrable}

is a normed linear space (of complex functions) defined as in Sec.3,


Ch.7 of B&B. It is easy to verify that Lj (Rm) is a Hilbert space with
the inner product (f , g ) = J f*gdx.
The intervals [a, 6], (a, b), (a, 6], [a, b) are all Lebesgue integrable
and have the measure n™i (&» — <*»)• When / is continuous on Rm, we
have / X[a,b\fdx = Jja6j / (x) dx, where the latter can be understood as a
Riemann integration. More generally, we also denote J XAfdx = fA f d x
for any Lebesgue measurable set A.
Hm is <r-finite and [—n ,n ], n = 1,2, • • *, constitute a pm-basis of
Rm (p.269, B&B). So, if / is Lebesgue measurable, then / is Lebesgue
integrable, if and only if |/ | X[-«,n] is Lebesgue integrable for all n and
limn_ 00 /[_nnj f d x exists. Therefore, for continuous / , / is Lebesgue
integrable just in case / is absolutely integrable in the sense of improper
Riemann integration.

For f E Co (R m, C ), by translating into an iterated integration, It is


easy to see that / f d x = \r\m f f ( tx + a) dx for r E R, r ^ 0, a E Rm.
Now suppose that / E L i(R m) and ( / n) is a representation of / (p.222,
B&B). Then ( / n (rx + a)) is a representation of / (rx + a). So we also

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 351

have

J fdx = |r|mJ f (rx + a) dx.


We can cover the unit sphere 5 (1 ) with intervals whose total volume
is arbitrarily small. Prom this It is easy to show that B (1) is integrable.

For s > 0, xb(.) (*) = XB(i) (? )• So, (B (s)) = sm/xm (B (1)). Note
that {13 (n )} , n = 1,2,* * *, is also a ^TO-basis of Rm. As in classical
.calculus, we have

LEMMA 6 .1 . I f r > m , then

J -^ d x = J 1 . X~ ~ dx
[II j |x|

exists.

PROOF. ~~| ^ ^ Xr(N) -* monotonically as N -» oo. By the


monotone convergence theorem, it suffices to prove that f
are arbitrarily small as N < M are arbitrarily large. It is easy to see
that

f XB(M) - XB(ff) = Um y '1 / X B ( N H i + l ) & p ) - XB(N+i ! ^ )


J i^ jr |x |r

However,

n_1 f Xfl(N-Kt-H)^^g) ~ X B ( N + i * f Z )
£/
i= 0 - I*

2 1 + (U 1) ^ + i^ ) "

As n —►oo, the sum approaches to Riemann integration fjjf dt,


which tends to 0 as IV —►oo, since r > m. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 352

We define Lebesgue integration on R m directly. On the other hand,


given Lebesgue integration pm on Rm and pn on Rn, we also have the
product /Xtn x fin on Rm+n (see B&B Ch.6, Sec.9). We can prove that

this is just /W n -

LEMMA 6 .2 . Suppose that f is pm-integrable andg is pn-integrable.

Then f - g is (im+n-integrable and f f (x) g ( y ) d ( x , y ) = (J f d x ) ( / gdy).

PROOF. Suppose that (A) is a representation of / and (gi) is a


representation of g. For each k, it can easily be shown that (A<7j) is a
representation of f kg, Pm+n (As) = Mm(A) Pn (g), and

M m +n ( l A t f l ) = ,l i m M m +n
(—♦00

Then, by Theorem(2.15) on p.229 of B&B, the conclusion follows. |

THEOREM 6 .3 . Pm+n is the same as pm x fin.

PROOF. Suppose that / 6 Co (R m+n, C). Then / can be uniformly

approximated by step functions, that is, functions of the form CiXin


where are intervals. Step functions are both pm x pn integrable
and Pm+n integrable and have the same integration. So, / is pm x Pn
integrable and (pm x pn) ( / ) = Pm+n (/) • Since pm x pn is complete,
This must be true for all / 6 Li (R m).
It remains to prove that all pm x pn integrable functions are Pm+n
integrable. By the definition of product integrations, we only need to
prove that when A and B are pm and pn integrable respectively, A x B
is pm+n integrable. But this is a special case of Lemma 5.1. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 353

C orollary 6 .4 . If f 6 L\ (R m+n), then fo r a full set A c Rm,


f (x, •) 6 L i (R n) for all x E A and x *-* f f ( x , y ) d y is /zm integrable
and

PROOF. This follows from Lemma 5.2 and Fubini’s Theorem (The-

orem(9.7), p.281, B&B). |

Now we prove some lemmas on differentiations. We denote Sff =


(Su , • • where is the Kronecker notation. Suppose that / is
a real or complex function on Rm. Then the partial derivative
j = 1, if exists, is a continuous function such that for any e > 0
and M > 0, there exists 8 > 0 such that

f ( x + rSf) - / ( * ) - ^ < c |r |

whenever x, x + rS’p 6 [—M, M\. and |r| < 6 . C “ (R m) denotes the


space of functions on Rm with compact supports and all iterated partial
derivatives. A sequence of non-negative integers a = ( a i , . . . , a m) is
called an index and we denote |a| = c*i -I ctm, xa = x“l • • • x“m, and

j)* = ^ -------- .
9x “l • • • 5 x “m

Let Ma denote the operator Maf = x“/ .


The Schwartz space of Rm is

/ : / is a complex function on Rm and for any indices a , /3,


5 (R m) = • there exists a constant cafi such that D af exists and
x^Z?a/ ( x ) | < Cafi for all x 6 Rm.

Clearly, / 6 5 (Rm) if and only if for any a and p > 0, there exists
Cd'p such that ( l + |x|2) P \Daf ( x ) [ < Caj, for all x € Rm. We have, by

Lemma 6.1, C0“ (Rm) C S (Rm) C Lv (R m) for aU p > 1.

with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 354

N o te . Recall that fo r a = ( a t i , q ^ ) and

Da = ^|a|
d x * 1 ...dx%*'
the witnesses for the existence of the higher order derivative D a f should
include all the derivatives one obtains in calculating D a f , that is, a
sequence of all functions f & \ for all 0 = (f t, ...,/3m) , f t < on, i =
1, such that fo r all 0 and i = 1, if f t < on, then
9 fW
1’ = /M ,
dx{

where f t = ( f t , f t _ i , f t + l,f t + i, ...,ftn) and the equality - i ^ =


/(^ ‘) should be read as • / M is a derivative of f W with respect to
x, ’ which is translated according to the definition o f differentiation.
Similarly, the witnesses fo r f ’s being in the Schwartz space S (Rm)
should involve an infinite sequence of all derivatives D ^ f o f f , that

is, a sequence i.f^ ')S such that —i^ g ~ = fo r all f3 and all
i = We take D a as a function from S (R m) to itself. So, D a
applies to a member f £ S (R m) together with its witnesses and merely
extracts the a-derivative f ^ form the sequence o f all derivatives.
The Fourier transformation on 5 (Rm) is the linear mapping Fq on
5 (Rm) defined by

(F „ /) (x) = ( 2 x ) - ? J e - ^ f ( y ) i y , for / S S ( R " ) .

To prove the properties of the Fourier transformation, we need some


lemma on the exchangeability of differentiations, limits, and integra­
tions.

LEMMA 6 .5 . Suppose that f : Rm+n -♦ C, f ( x , •) 6 f t (R n), §£:

exists, and fo r any M > 0, there exist 6 > 0 and g 6 f t (R n) such that

|f ( x + r 8 ? , y ) - f { x , y )| < g ( y ) \r\, for a l l y £ Rn,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 355

whenever x, x + rS™ E [—M , M] and |r| < S. Then

d r . . v. r df{x,y)

PROOF. Given M > 0, e > 0, we first find 8X> 0 and g E Lx (R n)

such that

j/ (x + r S f , y ) - f (x, y)| < g (y) |r|

for y E Rn, x, x + rSj 1 E [ - M , M ] , and |r| < 8 X. From this we see that
< g (y), and hence is integrable, for x E [—M, M], Take N

sufficiently large such that

df(x,y)
[
J R ”- [ - N , N )
f (x + r S ? , y ) - /( s , y ) --------■—
v 1 J OXj
dy < 2 lr l
for x , x + rS j1, E [ - M , M ] , and |r| < 8 X. Since exists, there exists

S < 8 X such that

whenever x ,x + r 6 ™ E [—Af, M ] , y E [—N, N] , and |r| < S. Then, in


this case, we have

d f ( x , y)
d y < - |r|

So,

|J f (z + r 6? ty ) d y - j f{x,y)dy- (j < t lrl

The conclusion follows by definition. |

COROLLARY 6 .6 . The conditions of the lemma are satisfied when

m = n, a n d /o r some h E S (Rm), / ( x , y ) = e~i*vh (y ).

We also need integration by parts.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 356

LEMMA 6 .7 . Suppose that f , g are continuous on RTO, ez-

ist, € L\ (R m), and fo r any fixed Xj, f g € L\ (R m_1) and

Mm-i (l/^ l) -* 0 as Xj —> ± 00 . Then


.dg, fdf
f f— —dx = - [ — gdx.
J Xj J Xj*
PROOF. By the assumptions, is continuous and integrable, so
f Q i s l d x can be expressed as the limit

to
M -*oa J - M J-M Xj

of Riemann integrations and the fundamental theorem of calculus (The-


orem(6.8), p.54, B&B) applies. The conclusion follows easily. |

COROLLARY 6.8. The assumptions o f Lemma 6.7 hold when f €


S (R m), g is continuous on Rm, exists, and fo r some constants
p > 0, c > 0, we have
dg(x)
|g (x)| < c ( l + |x |J) F, < c ( l + I*!1) ” , fo r all x S R m.

LEMMA 6 .9 . Suppose that f : R —>R is differentiable on R, f (x )+

z / ( z ) = 0 fo r all x 6 R, and f (0) = 0. Then f ( x ) = 0 for all x £ R.

PROOF. ([58] p. 15-16) Take any c > 0 and let F (z) = c f £ \f (<)| dt.
Since /0* ( / ' (t) + t f (t)) dt = 0 and / (0) = 0, we have (cf. Theo-
rem(6.8), p54, B&B)

1/ (®)| < I / t f (t) dt < F (z) for z 6 [0, c ].


\jQ

Since F' { x ) = c | / ( x ) | , we have ^ ( e ~ exF { x ) ) < 0 on [0,e]. So,


e-crF ( x ) is decreasing on [0, c] (cf. p.48, B&B). As / ’ (O) = 0 and
F ( x ) > 0 on [0, c] by definition, F ( x ) = 0 on [0,c]. Hence, / ( z ) = 0
on [0, c]. c is arbitrary, so / (z) = 0 for z > 0. The case of z < 0 can

be proved similarly. |

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 357

Lemma 6.10. J e-^dx = y/2ir.


P r o o f . By Corollary 6.4,

Given integer M > 0, for each integer N > 0, let Si = B (jj 'j—B ( ^ ) >
i = 1,..., N M . e ' K w ) Xs. converges to e“ Kl2 +v*) uniformly on

B ^ M) as IV —►oo. On the other hand,

j E e"’ ( ^ XSid(x,y) = £ e'K jf) * j

= 27rEe -L
> fW
i f F* 1 -H
F ^ e ■(»i )) 2 1

Let N —* oo. We see that


rM
fg M e ^ +^ d ( x , y ) = 2ir J r e ~ ^ 2d r = 2ir ( l - e *M *'j .

So, ( / = 2tt. |

With these preparations, we can derive the basic properties of

Fourier transformations. ( [104] Theorem 10.1 to Theorem 10.6)

T h eo re m 6 .1 1 . F0 S ( Rm) c S ( R m). For f € S ( R m) and index


at,

0 a f ,o / = ( - i ) M f y < j 1

M .F e f = F J T f .

PROOF. B y Corollary 6.6,

D ’ Faf = (-i)W (2x)-» /

So, for any index 0 ,

z * ir F o f = (_i)W+W (2x)-? J D> («-**) (if* / (,)) <ty.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 358

Applying Corollary 6.8 several times, we have

xs D “ f 0/ = ( - . ) W ( 2 t ) - ? J
So, x*D a F0 f \ = (2 * ) ~ 7 f e~t!tvD^ ( y af ( y ) ) \ d y = Ca# a constant.
Hence Fqf € S (Rm). Let a = 0 and /3 = 0 separately. We obtain the
equations in the theorem. |

Note. The first equation of the proof should be understood this


way. Denote

= ( _ i ) H ( 2 , ) - ? J y ° e- * >f ( y ) d y ,

then by Corollary 6 . 6 , w among the witnesses for F0f 6 S (R m)


which gives the infinite differentiability of Fof. After that we can see
that D a F of is given by the expression g^a\ The last equation o f the
proof is also obtained by an induction, but this time the inductive for­
mula is an equation between two terms for complex numbers with x as
a free variable, so it is a II? induction.

THEOREM 6 .1 2 . Fq is isometric in the norm of L 2 (R m) with

{f q 1 s ) = (2 * )’ * J e ^ g (y) dy.

PROOF. ([104] p.291) Let h ( x ) = A direct computation

gives {F 0 h) ( x ) = ri/Li ^0 («-***)• Set k{ t ) = F0 (e~5,a) - e"**2, The­


orem 6.11 gives k' (t ) + tk (t) = 0 and Lemma 6.10 gives k (0) = 0. So,
by Lemma 6.9, k (t) = 0 and we have F0h = h.
For any £ > 0, / 6 S (Rm), using Corollary 6.4, we can directly

compute that

J h (ey) e**v (F0f ) (y) dy = j f ( e z + x ) h (z) dz.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 359

When e —►0, h ( e y ) e ixv(Fof) (y) —►elxv(F0 f ) ( y ) almost uniformly,


and f (ez + x) h( z ) —►/ (x) h( z ) almost uniformly. Lebesgue’s dom­
inated convergence theorem applies and we have / eixv (F0 f ) (y) dy =
f ( x ) f h (z) dz = (2tt)t / (x). That is

(2*)-> J e * '( F 0f)(y)dy = f(x).


Let y' = —y in it. We will get

(2t)‘ » _ /V “» ((2x)-» J dy = /( * ).

That means Fq maps S(Rm) one-one onto 5 (Rm) and F0_1 is given as

in the theorem.
For f,g £ S (Rm), by Corollary 6.4, we can directly compute that
(.F Qf , F Qg ) = (f,g), that is, F0 is an isometry. I

Next, we extend Fq to Lj (Rm). We need some more lemmas. First,


recall that we can patch two continuous (one variable) functions on two
contiguous intervals [a, 6] and [6, c] to form a continuous function on
[a, c] if the two functions are equal at b. The following shows that we
can also get a smooth function if the derivatives of the two function

are equal at b.

LEMMA 6 .1 3 . For f d e fin e d o n (0 ,+ o o ), th e r e e x is ts a fu n c tio n g

on ( —oo,-|-oo) su c h th a t i f f W e x is ts o n ( 0 , + 00), and f ^ (x) —» 0


os x —» 0+, fo r k = 0 , 1 , ...,n, th e n g W e x is ts , g W ( x ) = f W (x), on

(0, + 00), and gW (x) = 0 on (—00, 0], fo r k = 0,..., n.

PROOF, g is uniquely determined by the condition that g(x) = 0


on (-o o ,0 ] and g(x) = f( x ) on (0,+ oo). Let h be the continuous
function such that h (x) = 0 for x < 0 and h (x) = f (x) for x > 0. We

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 360

prove that g' — h. Given M > 0 and e > 0, we must find 8 > 0 such
that

(45) \g (y)-9 (x )-h {x ){y -*)| < e \y - x \

whenever x ,y 6 [—M , M ] and |y — x| < 8. Clearly, we only need this


for x ^ y. First, choose 8 \ > 0 such that |/'( x ) | < | for x 6 (0, £i). /
is differentiable on M , so we can find 8 j such that

min (<$1, 62)) we prove


(45). Suppose that x , y 6 [—M , M ], x ^ y, and \y —1 | < 8. We can
distinguish 3 cases.
(1) x < 0, y < 0. (45) is trivial.
(2) x , y 6 [^ , M . It follows from (46) and the definition of g, h.
(3) x , y € (—£i,£i). By the choice of £i and the continuity of h,

IM*)I - fi so we nee£l °nly to prove that |g ( y ) —y (x )| < § |y — x|.


This will follow if we can prove

(47) \ g { y ) - g { z ) \ < | | y - * l + e'

for any e' > 0. Given e‘ > 0. Since y ^ x and g (0) = 0, we can find
8' > 0, 8 ’ < min (|y - ®| >£ ) » such that \g(z)\ < j whenever \z\ < S'.
We can further distinguish 3 cases.
(4) x,y < 0. (47) is trivial.
(5) x,y > 0. Since x ^ y, we may assume that x < y. Then,
by Theorem(5.6) on p.48, B&B, there exists z € ( i , y ) such that
\ g ( y ) - g ( x ) - 9 ' ( z ) ( y - x ) \ < e ' . As |, | < * , \g'{z)\ = \ f'(z)\ < §

So, (47) holds.


( 6 ) One of 1 , y, say 1 , is in { —j , j ) • Since 8 ’ < \y —x|, we have y ^
0. If y < 0 , (47) holds because \g (x)| < e'. If y > 0, then y > j and,

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 361

by the same reason as in (5), we have g ( y ) —g ( y ) | < | y - is + f


So,

Is ( y ) - $ ( * ) ! < +

<-
s' 3e'
y —
~ 2 + T

< - |y - 1 | + e1.

So, we have proved that g1 exists and g'(x) = 0 for x < 0 and
g' (x) = / ' (x) for x > 0. The theorem is obtained by repeatedly
applying this argument. |

Note. T h e a p p a r e n t u s e o f i n d u c t io n h e r e c a n be e lim in a te d . For

k — 0, l e t g it b e t h e u n i q u e c o n t i n u o u s f u n c t i o n s u c h t h a t g k ( x ) =
0 fo r x < 0 a n d gk (x) = / ^ (x) fo r x > 0. T h e n b y th e s a m e a r g u m e n t

w e h a v e g'k = g k + i, k = 0 , ..., n —1. T h e c o n c lu s io n th e n fo llo w s .

LEMMA 6.14. C7o°(Rm) is d e n s e in Lp(Rm), p > 1.

PROOF. By Lemma(3.13) on p.318, B&B, C7o(Rm,C ) is dense in

Lp (R m). Each function in Cq (Rm, C) is bounded, has a compact sup­


port, and hence can be almost uniformly approximated by step func­

tions £ y cyX[«i .&,-]> so it suffices to prove that any characteristic function


of an interval, X[a,6j can be approximated by functions of (Rm) in
Lp(R m). Let

for xy 6 n > 0, j = 1, ...,m. Then, for any k > 0, gf£J exists


on (o y, bj) and gj^J ( xj ) —» 0 as xy —►o,*+ or xy —» bj—. So, By Lemma

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 362

6.13, there exist f nj 6 (Rm) such that

f n j (xj) = 0 for i j < aj or xj > bj,

= 9nj (xj) for Xj e (aj, bj) .

Let f n = nr=i fnj- Then f n 6 C0°° (Rm) and / n -» X M almost uni­


formly as n - > oo. Lebesgue’s dominated convergence theorem implies

fn -* X[a,b] in I p ( R m). I

The following theorem follows from Lemma 6.14 and Theorem 6.12

directly.

THEOREM 6 .1 5 . fll04] Theorem 10.4) Fo can be extended into a


unitary operator F on L 2 (Rm) with F~l = F*, the extension of F q 1.

We move to another topic, the Sobolev spaces. We follow the clas­


sical presentations in [104] Section 10.2. For s > 0, denote k, (z) =
( l + |x|2) J/2. In the following, the inner product and norm of L2 (R m)
will be simply denoted as (•, •) and ||-j|. The Sobolev space of order s

is

Wu ( l m) = { / 6 L2 (R m) : kt F f 6 L 2 (Rm) } ,

with the inner product

( f , g ) l = (kl F f , k tFg)

for f , g e W3<t (Rm). We denote ||/(|, = ( /, / ) „ the norm of W2,t (R m).


/ *-* kt F f is an isomorphism of W2l, (Rm) onto L2 (R m), so W2i, (Rm)
is a Hilbert space with the inner product (•, •),. Obviously, S (Rm) C
W2,t (R m) for all a > 0, and W2t. (R TO) C W2>r (R m) for 3 > r > 0.
Define the operator D a = F ~l MaF . When / € W2i, (R m) and
a is an index such that |a| < s, then Ma F f € L2 (Rm) , and hence

of the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 363

D ( D a ) D W 2 i t ( Rm) and

D a f = F ~ lMaF f , for / € W2>l (R m) .

By Theorem 6.11, this is an extension of the operator D a defined on

5 (Rm).

LEMMA 6 .1 6 . ([104] Theorem 10.6 ) Suppose that r > s > 0. Then

fo r any rj > 0, there exists c > 0 , such that

\ \ f \ \ 2t < V \ \ f \ \ r2 + c \ \ f \ \ 2 f o r f e W 2,r.

PROOF. This is obvious from the inequality

< (l + AT1) ’ | | F / f + (l + JV1) " ' " ' 1 If/I’

< ( l + AT1) ' ll/U 1 + ( l + J V I ) ' (" ‘, | | / | | r2 .

LEMMA 6.1 7 . ([104] Theorem 10.8) There are constants c i , c 2 > 0

such that

«1 E n ^ / l l 1 < ll/ll. < < * E l |0 “ / f /<” • / € W i ,.


|a |< r |o |< r

PROOF. There are constants c i,c 2 such that

ci £ |x*|2 < ( l + |* |2) r < c2 X ) I1*!2 for all i 6 Rm.


|a |< r |a |< r

The inequalities follows from the definition of ||-||r, D a , and the uni-

tarity of F . |

LEMMA 6 .1 8 . ([104] Theorem 10.10) C™ ( Rm) is dense in W 2 ) t ( Rm)

in the norm ||-||4 .

p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 364

PROOF. ([1 0 4 ] p.299) First we prove that S (Rm) is dense in

Let / 6 W2<t and e > 0. Then kt F f 6 L 2 ( Rm) and by Lemma 6.14


there exists h € C7“ (R m) such that \\kgF f — A|| < e. k j xh 6 S (R m).
Let g = F~ l ( k ^ h ) 6 S (R m). Then | | / —g ||4 < e. So, S (Rm) is dense
in W2<m.
Then we prove that C * (Rm) is dense in S (Rm) in the norm ||-||4.

Let g ( t ) = exp (2- t)a e ) • Then, as t —> 1+, g { t ) —» 1 and


g(i)-(t) —►0 for j > 0; and as t —* 2—, gW (t) —►0 for all j > 0. So, by
Lemma 6.13, we can find a function gi on R such that gx (t ) = g (f) on
( 1 , 2 ), gi (t ) = 1 for t < 1, gi (t ) = 0 for t > 2 , and g[^ exists for all
j . Repeat the construction, we have a function h 6 C “ (R) such that
h( t ) = 0 for |t| > 2 and h( t ) = 1 for |f| < 1. Now given / € 5 (R m)
and e > 0, we need to find g € C ” (Rm) such that | | / — <7^ < e. By
Lemma 6.17, we need only that ||D af — Dag\\ <e for |a| < s. Let
g„(x) = h ( ! g ) f ( x ) e C ? ( R” ).

(* ) = h o * /(* )+ £
V” / |,|>0.,+«=» \n J

Each D^h is bounded by a constant independent of n, and in


case I7 I > 0, D^h ( ^ * ) = 0 for |z| < n. Then It is clear that
|jD af — D agn || < £ as n is sufficiently large. |

Finally, we want to prove an inequality ([104] Theorem 10.17)


which will be used in the next section. First we need some notations
and lemmas ( [104] Section 10.3) For p £ R, p < m or p > m , we
define a space of measurable functions Mp (Rm) as follows:
1) When p < m , q € Mp (Rm) if and only if q is measurable and

there exists a constant M qiP > 0 such that jg (z)| 2 XB(x,i) (*) |* — z\p~m

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 365

is integrable as a function of x and

/ |? (x )|2 \x - z \p- m dx < Mq,p for all x € R m.


J B [ z , 1)

2) When p > m , q € M p (R m) if and only if q is measurable and

there exists a constant MqiP > 0 such that |q (x )|2 XB(2,i) (2 ) inte'
grable as a function of x and

/ |q (x)l2 dx < Mqp for all z 6 R m.


J B (t, i) v

Mq>p is called a p-bound of q. If p < p ' , then Mp (R m) C Mp>(R m)

and MqtP is also a p'-bound of q.

LEMMA 6 .1 9 . Suppose that f , g , h 6 L i(R m) are non-negative,


k ( x ) < c fo r some constant c, and fo r some full set F C R 2”*, (x, y) 6
F implies y — x € Dam (g). Then f (x) g (y — x )h (y ) 6 L\ (R Jm).

PROOF. Let gn — g A n lot positive integer n. Then f gn (y) dy ->


fg(y)dy. By Lemma 6.2, / ( x ) gn ( y ) , / { x ) h ( y ) 6 Lx (R 2m). Since
|f { x ) g n (y — x) h( y ) \ < n \ f (x) h(y)\ on the full set F fl D o m ( f h ) ,

f (*) 9 n ( y - x ) h (y ) 6 L\ (R 2”1). ( / (a;) gn (y - x ) h (y)) is an increas­


ing sequence converging pointwise to / (x) g ( y —x ) h (y) on the full set
F fl Do m( f h ) . By Theorem(8.6) on p.267, B&B, it suffices to prove
that ( / f ( x ) g n (,y - x ) h ( y ) d ( x , y ) ) converges. But for m > n ,

| / f{x)gm { y - x ) h { y ) d { x , y ) - j f (x)g n ( y - x ) h(y)d(x,y)

= J f{x) J (gm ( y - x ) - g n ( y - x ) ) h ( y ) d y d x

< c j f { x ) j (gm ( y - x ) - gn (y - x)) dydx

= c j f ( x ) d x j ( g m (y) - gn (y)) dy.

As ( f g n ( y) dy) converges, we see that ( / f ( x) gn (y - x) h ( y ) d ( x ,y))


converges. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 366

LEMMA 6.20. Suppose that r > 0, s > 0. Then the Junction


(*) belongs to L \ (R m) and fo r some constant c indepen­
dent o f s,

PROOF. As in the proof of Lemma 6.1, we have

for some constant c and some any q > 8 > 0. The conclusion follows
immediately. |

THEOREM 6.21. ([104] Theorem 10.17) Suppose that r > 0 is an


integer, p < 2r. Then
(a) there exists c > 0 such that fo r any q G Mp (R m) and MqiP a

p-bound of q, we have, fo r all f G W2,r (Rm), q f G La (Rm) and

(b) fo r each q G M„ (Rm) and tj > 0, there exists Cv > 0 such that
f o r a l l f e W 2 iT( Rm),

ik /ii2 < v m l + c , ii/ii2 .

P r o o f . We follow the proof in [104]: first prove (a) and (b) for
/ G (7“ (Rm) and then extend to W2>r (Rm) by the density of (Rm).
Some revisions are needed in this constructive version.
Let / G C7“ (Rm). Choose a function 8 (x) 6 C " (Rm) such that
8 (0) = 1 an d S (x) = Ofor [x[ > 1. For any w G S' (1) = {x G Rm : |x| = 1}
and s G (0,1], “ *n Cq°(R ) as a function of t, and

o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission .


6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 367

^ (5 ( y ) |t=J = 0 for all n > 0. Integrating by parts re­

peatedly, we have

For any a , 5 “ 5 is bounded and |u;“ | < 1, so

I-
[eft

for some constant ci independent of s , w, t . Let

k(z)=-£
M<»*
We have, for some constant C2 independent of s , w, t ,

(48) |/ ( 0 ) | < c , [ ' h ( w t ) t r~ldt.


Jo

The next step in the classical proof is to integrate the right hand
side of this inequality over the unit sphere 5 (1 ) and then transform
it into an integration over 5 ( 1 ) , the unit ball. To avoid developing a
full constructive theory of surface integration, we can instead construct
Riemann sums over 5 ( 1 ) directly. Here we sketch how to do that.
First we must divide 5 ( 1 ) into appropriate regions. Consider the
mapping r from the interval

m —I
I = [0,7r] x • • • x [0, it] x [0,2%] C R1

into S (1) C R m: for 0 = ( 0 i,..., 8 m- i )

( COS 01 ^

sin 0i cos 8 2

sin 0i • • • sin 0m_2 cos 0m_i


^ sin 0i • • • sin 0m-a sin 0m_i J

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 368

For each N > 0, divide each side of I into N equal sub-intervals, then I
is divided into JVm-1 sub-intervals, and their images under r are N m~l
subsets of 5 (1 ). For each K > 0, divide the interval
(0, s] into a sequence of sub-intervals (s ( l — j^ 3 , a ( l — j ^ 3], j —
0 ,1 ,2 ,.... For each i = 1,..., Wm-1, j = 0 ,1 ,2 ,..., let

^ = jp x : x 6 Si, 1 - ^ < p < 1J ,

Each V * , V * . j = 0 ,1 ,2 ...... is a section of the cone generated by


Si. It can be proved that all V * , are integrable. (We can cover
their boundaries by intervals of R m with arbitrary small total volume
and hence approximate (almost uniformly) each region by intervals.)
Furthermore, V * , i = 1 constitute a partition of B ( l ) —

B (1 ~ f ) *n t^ie sense ^ a t Xb(i) ~ = H i X v f on a set-


Similarly, V^j, i = 1, ...,JVm_1, j = 0 ,1 ,2 ,..., constitute a partition of

5 00 -

Note that n ( K j ) = ( l — jf)*™ fi> ( v f ) . Foreach i = 1,..., N m~l


pick one it;,- 6 5,-. For any e > 0, we have,

J h ( w i t ) t T l dt

= Um | > ( « * ( l - ■ !)’) ( l - i ) J) (s ( l - ^ ) ' - * ( l -

= & 7c % k {w“ ( l - w ) ’) (J( l - w ) ’)

- f a ~ w (s ~ w +e

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 369

for all i = 1, whe n K is sufficiently large. Since (48) is true


of all Wi, we have, for sufficiently large K ,

(/ o* a (w ) t ' - 1* ) M(v;K)

E 2 T £,"o ft (<m (i - if) (* (i - £ )’) ' m(v;K)


— c2 ------------------------------- 7------------------ 7--------TT\--------------------------- ^ c 2e
Kn(B(l)- B ( l - i ) )

£ .= r s ” o ft («*» (i - iy) (* (i - * ) ') " “ /. (k j)


jr„ (* (i)-u (i-* )) +CJ£-
As K , N —* oo, the numerator converges to JB^ h ( x ) \ x \ r~m dx and
the denominator

x> ( b (i ) - . b ( i - I ) ) = #. ( s ( i ) ) jc ( i - ( i - ^ r ) “ )

—ft m f i ( B ( 1 ) ) .

Since e is arbitrary, we obtain

|/( 0 ) |< c 3 / h( x) \ x\ r- m dx

for some constant C3.


Now for any e > 0,

l/(o)|2< 4f/,
< £ [ |«|— dx [ \ h( x ) \ 2 \x\2r-'~ m dx
J\X\<M J |*|<«
<C43* [ h 2 {x)\ x\ 2r— m dx
J\z\<l
FI*I<1

for some constant C4 independent of s. Replace / ( x ) by g (x) =


/ ( y + x), we have

(49) | / (iOl2 < W t h 2 (x) |x - y |2r~*_m fa,


J |* -v |< l

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 370

We will prove that q f 6 £2 (Rm) and for some 8 > 0,

(50) ||g / | | 2 < cAM qiPss J h2 dx.

Distinguish two cases: 2r > m and 2r < m. First consider the case
2r > m. We may assume that p > m. Let s = 2r —m in (49). Then
| / ( y )|2 < c*3 2r~m f x s (» ,i)(i —y ) h 2 {x)dx. Suppose that the interval
2 2
[—N, N] supports / . It is easy to see that X[-n,n] kl *s integrable. I/I
is bounded, so q f = qX[-N,N]f £ ^2 (Rm) and

lk / ||2 < C4s 2r_m J J \ q {y)|2 x[-n.n] (y) Xfl(i) (* - y ) ^2 (*) dxdy

= cA3 2r~m J ( J k (y)|2 xi-tr,N] (y) X B ( 1) (* - y) dy) h2 (z) dz

< c^s2 r~mM^p J h2 (z) dz.

So, (50) holds. When 2r < m, we have p < m . Let e = 2r —p, in (49).

Then

l/(*r)l* < < * * ' ' [ , ^ (x jk -v r " * .

We have similarly q f 6 £2 (Rm) and

Ik /ll2 ^ c4*2t~pJ J k ( y ) l2 x [ - n ,n ) (y ) x b (i) (* - y ) I* - y lp_ro ^ 2 (* ) dxdy-

Note that { ( x , y ) : x = y, z ,y 6 Rra} C R 2m is an integrable set with


measure 0. So, {(® ,y) : 2 y , z ,y € Rm} is a full set. By Lemma
6.19 the order of integration can be changed, and (50) follows with
8 = 2 r —p.
We have f h 2 ( x) dx < c5 E |a|<r s 2(1“l_r) ||D a / | | 2 for some constant
C5. Let 3 = 1 in (50). We obtain the conclusion (a). To prove (b), note

o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.


6. FOURIER TRANSFORMATIONS AND SOBOLEV SPACES 371

that, by Lemma 6.16 and 6.17, we have

I h' ( x) & < * ( • £ W I T f f + s - 1' £ III T f t


\|o |= r |a |< r

< * (ll/ll.1 + ^ lr ll/ll.2- , )


< * ( ii/ ii; + ii/ ii;+« tw ii/ ii1)
< =»ii/ h;+<^(^)ii/n:'
for some constants eg, eg (a), where eg is independent of a. Then (b) is
obtained by taking s sufficiently small in (50).
To prove (a) and (b) for / € (R m) we will find a sequence of

/„ € C “ (Rm) SUCk tilat ll/n - /II -* °i II/n - /IIr “ » “ d II?fn ~ ? /|| “ »


0. By Lemma 6.18 and 6.17, we can choose f n 6 Cq0 (R to) such that
ll/n ~ /II —♦ 0 and ||/n - / | | r -* 0. Since ||/n - / | | -> 0, by a remark
near the end of p.267, B&B, |/ n —/ | 2 —►0 in measure. So, f n -» /
in measure. We want to prove that q fn —> q f in measure. So, given
an integrable set A and e > 0, we first find M > 0 and an integrable
set Bisuch that B \ C A 1, n {A —B \) < | , and B \ C B ( M ) . Now

Sbi l?|2 dx exists, so we can find an integrable set B 2 and M\ > 0


such that B\ C Bf, f i ( B\ — Bf) < | , and |? (x )| < M i on B \. Now, as
f n —» / in measure, there exists N > 0 such that for each n > N, we can
find an integrable set f?3 such that B ] C B \ and /x(B 2 — B 3 ) < | and
1/ — f n\ < on B \. Then we have ft ( A — B 3) < £ and |q f —qfn| < s
on B j. That is, qfn —* q f in measure. Since by what we have proved,

||?/n - ? /« || = II? (/n - / . ) | | < c M q,p ||f n - f m\\r ,

and since ||/„ — / | | r -* 0, we see that q fn -* h. for some h e L 2 (R m).


It follows that qfn —* h in measure. B y Lemma (8.12) on p.270, B&B,
we have h = q f . So, q fn -* q f in L 2 (R m). This completes the proof. |

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited without perm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 372

N o te . The proof of this theorem, though very long, poses no real


difficulties for translating into S C . A 11° induction is needed at the
beginning to reach that equality obtained by integrating by parts repeat­
edly. The expansion of the derivative gp (j> / (tutfj into partial
derivatives of 8 and f can be constructed uniformly fo r r, given that 8

and f are infinitely differentiable. So, inequality (48) is obtained. The


sequences used in estimating the integration over 5 ( 1 ) are all con­
structed uniformly and inequality (49) is thus obtained. After that the
proofs are straightforward. No more inductions are involved.

7. S elf-ad join tn ess o f Q u a n tu m M ech an ical O p erators

In this section we prove the self-adjointness of some most common


quantum mechanical operators. We begin with the position operators.

7.1. P o sitio n o p era to rs. We consider the general case. Suppose


that q : RTO—» C is Lebesgue measurable. The operator Af, on L 2 (Rm),
called the maximum multiplication operator induced by q, is defined

as:

D (Af,) = { / € L 2 (R m) : q f e L 2 (R m) } ,

Mq = q f for / € D (A f,).

T heorem 7 .1. ([104] p .51 and 91)


(a) D (Af,) is dense;
(b) I f \q\ is essentially bounded, then D (Af,) = L 2 (Rm); if\q\ > c
on a full set fo r some c > 0, then R ( M q) = L% (Rm);
(c) If q is real on a full set, then Af, is self-adjoint;
(d) I f \q\ is essentially bounded on any bounded integrable set, then

C * (Rm) is a core o f Af,.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 373

PROOF, (a) By Theorem(8.20) on p.274 of B&B, we can find a

sequence of positive numbers Cx,C2,..., Cn —> + 00 , and a sequence of


measurable sets A i, A2, s u c h that |g| < Cn on A^ and |g| > Cn on A£.

Let / E L2 { Rm). Each x a , / € D (Af,). We show that x a „ / -» / in


L 2 (R m). By the dominated convergence theorem, we can prove instead
that \xAnf ~ S I* —* 0 in measure. Given t > 0 and an integrable set A,
since \q\ is measurable, we can find an integrable set B with B 1 C A1
and an integrable function k such that f i ( A — B ) < | and ||q| — k\ < e
on B 1. Since \k\ is integrable, we can find Af > 0 and an integrable
set C such that C 1 C B 1, f i ( B — C) < | , and \k\ < Af on C 1. So
|?| < Af + e on C 1. Take n > M + e. Then C l D A° = 0, which means
C A An = C. So,we have p ( A — C A A„) < e and |x a * / - / | 2 = 0 on

C l H A^. That is, \xAnf — / | 2 —►0 in measure.


(b) and (c) are obvious.
(d)Given / 6 D (Af,) and e > 0, we must find g 6 C * ( ^ m) suc^
that H / - 5 II < £ and \\qf — qg\\ < £• Since f , q f 6 £r2 (Rm), there

exists N > 0 such that ||x b ( N ) / - / [ < | and ||xR (N )g/- ? /|| < f -
Prom the proof of the density of C“ (Rm) in L 2 (Rm) (Lemma 6.14),
we see that there is a sequence of gj 6 (R m), with some common
support B (Af), such that gj -* XB(N)f L 2 (Rm). By assumption,
q is essentially bounded on B (Af), so qgj —* qXB{N)f■ Therefore,

there exists gj such that gj - Xb(N)/|| < f and qgj - q X fl(N )/|| < f
simultaneously. So we have | | / —gj\\ < £ and \\qf — || < e. I

Position operators on L 2 (Rm) are special cases of Af, with q = Xj ,


a real function bounded on any bounded set, so they are self-adjoint

with <7* (Rm) as a core.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 374

7 .2 . D ifferential o p e ra to rs w ith constant coefficients. Now


we consider differential operators with constant coefficients. These will
contain momentum operators on L 2 (Rm). We follow the classical pre­
sentations in section 10.2 of [1 04]. Let P ( x ) = £|a|<r caxa be a poly­
nomial. Call P r degree if |ca | ^ 0 for some |a| = r. We denote
P (D ) = £ | a|<r CaDa . We define the minimum operator Tp associated
with P (D ) as:

D (r £ ) = (R m) ,

T p f = P (D ) f for / e D ( 2 ? ) .

And we define Tp — F~l Mp F. Notice that when P = i “ , this is


the same as the operator D a we defined in the last section. Clearly,
Tp is the restriction of Tp to C” (R m) and Tp is self-adjoint when the
coefficients of P axe real, since It is unitarily equivalent to Mp (Lemma

3.8).

THEOREM 7.2. Suppose that P is r degree and the coefficients of

P are real. Then


(a) D (TP) D W2,r ( R m ), and if c ( l + \P (x)\) > ( l + | x | 2) 5 for
some c > 0, then D (Tp) = W2,r ( R m ) /
(b) Tp = Tp and hence £7“ (R m) is a core of Tp.

PROOF, (a) follows from the definitions directly. For (b), consider
the extension Tp of Tp:

D f o ) = S ( R m) ,

Tpf = P(D) f for / 6 £ (R m).

By Theorem 6.11, Tp = F~l Mp F, where Mp is the restriction of


Mp to S ( R m). By Theorem 7.1, Mp = Mp. So, Lemma 3.8 gives

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 375

Tp = Tp. As Tp C Tp, Tp exists and Tp C Tp. It remains to prove


D (Tp'j D D (Tp) = S (Rm). Let / 6 5 (R m) and e > 0. We must find
g 6 <?0°°(R m) such that | | / - $ | | < £ and \\T },f-T ^ g \\ < e. This is
obvious, for, in the proof of Lemma 6.18, It is shown that there exists
g 6 Co°(Rro) such that \\Daf — D ag\\ axe arbitrary small for all a,
|at| < r. |

We can represent momentum operators as Fourier transformations


of position operators. Then the self-adjointness is obvious. On the

other hand, the following theorem shows that D a are differential op­
erators on W2f For j = l,...,m , e ^ 0, / € Z j(R m), let Sj =
•••) <^mj) 6 Rm, index ay = (£Xy ,..., £my), and /y,e be the function
defined by /* .( * ) = / (x + eSj) for x 6 Rm.

THEOREM 7.3. ([104] Theorem 10.8) Suppose that r > 1, / 6


W2,r (Rm)» and 1 < j < m . Then

(a) lim*_0 ^ (fj,t - / ) = D * if in L 2 (Rm);


(b) If exists, then f belongs to W2<r (Rm) and 7 §■£■ = D * 1 f in
L 2 (R m). And this can be iterated so that D af is the iterated partial
derivatives for |a| < r.

P r o o f , (a) It suffices to prove that ( / Jit —f ) —* MSjF f in


La(Rm). Choose / n € S (R m) such that /„ —►/ in L 2 (R m), then

(A)* “ > /* .. - F f, F i M j , - Ffu- Now F ( / .) * ,( * ) =


f t ( /.) * (» ) = So, (/■ /* .)(* ) = =“ ** (J ’/ ) (x). Hence,

(51)

\{F h {fi- ~ ! ) ~ M “ F f ) (l)f = \i _ 1- “ * ) ( f ( l ) f ■

— l —ixj-e) < 2\ x j \ < ( l + |x |2) s . So the right hand side of

(51) is integrable and bounded by the integrable function ( l -f |x |2 ) 3 (F f ) (x)

R ep ro d u ced with p erm ission of the copyright ow ner. Further reproduction prohibited w ithout p erm ission .
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 376

for all e 7^ 0 . 1 (eiXje — 1 —iXj-e)| —►0 uniformly on any bounded set.


Since \ F f \2 is integrable, It is easy to show that the right hand side of
(51) converges to 0 in measure. The dominated convergence theorem
implies that F ± ( / J>e — / ) —» M XjF f in L 2 (R"1).
(b) By the assumption, £ (/,-,* —/ ) —► uniformly on any bounded
set and hence also in measure.' On the other hand, by (a) and the re­
mark at the bottom of p.267, B&B, £ ( / j ,e ~ f ) D ai f in measure.
Then Theorem(8.12) on p.270 of B&B gives = D a’ f on a full set,
and hence in L 2 (Rm). The iterated case holds as well because, from

the definition, D a f 6 W^.r-ial (R m) when |a| < r. |

7.3. H a m ilto n ia n s. We will prove the self-adjointness of some


of the most common Hamiltonians appearing in quantum mechanics.
These, together with Stone’s Theorem, imply that the corresponding
dynamic equations can be solved constructively. We will first follow
[104] to prove a (constructive version of) a general theorem, Theorem
10.23 of [104]; then we apply it to the special cases.
In the following we let

^ 2,/oc (Rm) = { / : / is measurable and X[-fr,N]f € L 2 (Rm) for all N > oj

Notice that M p (Rm) C La ioc (R m) for all p. Consider operators Ho of


this general form:

D ( H 0) = C ? ( R m),

Ho f = ' £ ( D j - b j )2 f + q f
y=i

= - A o/ - 2 £ ; b j Dj f + (b2 + iD ivb + q) /,
j=i

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 377

where D j = — and bj is a differentiable real function on Rm for

j = 1, ...,771, b2 = Zj Vj , 9 € I-2joc(Rm), and A 0 = E , - i s the


operator with domain C “ (R m). By Theorem 7.2, Ao is essentially
self-adjoint. Let A = Ao- Then D { A ) = W i t2 (R m).

THEOREM 7 .4. ([104] Theorem 10.22) Let Ho be as above. Sup­


pose that bi, ..., bm are bounded with bounded derivatives and q € M p (R m)
for -some p < 4. Then Ho is essentially self-adjoint and D (# o ) =

^ 2l2(Rm).

P r o o f . Let,

Vo = —2 bjDj + 4" idivb + q^j

m
= 'Z .tjD j + qa-

Since bj, b2, and div (b) are bounded and q £ Mp (R m) C L2,/oc (R m)i we
have D (Vo) 3 Co° (Rm)> So, Ho = - Ao + VQ. Vo is clearly symmetric
on (R m) (for Ho and —Ao are), so,by Lemma 3.6, it suffices to
prove that Vo is Ao-bounded with a bound less than 1. From Lemma
6.20 we know that each qj 6 M„> (R m) for some p' < 2. So, for each j ,
Theorem 6.21 implies that for any rj > 0 there exists > 0 such that

h i D j f f < V p , / | * + 0, W D j f t for aU / € C ? ( R - ) .

By definitions,

< C l||A o /||2 + cx ii/ ii2

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 378

for some constant Ci > 0. By Lemma 6.16 and 6.17, for any tj > 0,
there exists a constant dr, such that

WDjfW2 < V ll/ll, + dr, ll/ll2 for all / 6 C0~ ( R » ) .

Since |( l 4- |x |2) F f < ||/ ||2 -I- ||A 0/ | | 2, we have finally, for
any rj > 0,

\\qjDjfW2 < r, || A o /||2 + e , || / ||2 for all / € C0~ (Rm)

for some constant ev > 0. Similarly, as div (6), b are bounded, and
q € Mp (Rm), we have

qo = ( b2 -(- id iv (b) + 9) € Mp (R m) ,

and for any rj > 0,

Iko/H2 < V || A o /ll2 + e' ll/d2 for all / 6 C0°° (Rm)

for some constant e' > 0. So, VQ is Ao-bounded with arbitrary small
bounds. |

Next we will prove a generalization of this theorem. It is Theorem


10.23 of [104] and we will follow the proofs there closely. However we
have to strengthen the assumption of the theorem a little in order to
constructivize the proofs: in the original theorem, 6 below is just 2.
Let

M„ m (R ” ) = { / £ (R” ) : * „ ( „ ,/ £ M , (R ” ) for any N > o} .

THEOREM 7 .5. Let Ho be as above. Suppose that bi,...,bm have

continuous first order derivatives and q = qi + qi such that


?i € Mp (Rm) fo r some p < 4;
q2 € Mpjoc (Rm) fo r some p < 4 and ? j(x ) > —c|x|* fo r some
c > 0, 6 < 2 , and all x.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 379

Then Hq is essentially self-adjoint.

Following [104] we first prove some auxiliary lemmas. The proofs


in [104] can be adapted directly.

LEMMA 7 .6 . Let bj, qi be as in the theorem. Then, fo r any rj > 0,

there exists c„ > 0 such that for all f 6 C " (Rm),

(toll /./)< 1EIKA-W /II1+ <*ll/ll2-


J=1
PROOF. Given / , choose <p € C ® (R m) such that <p(x) = 1 for x

in a compact support of / and let / , (x) = ( | / ( x ) | 2 + e) * for e > 0.


Notice that

(52) (toil/. /) = ||toil*/f < Itoil’/.v f.

(53) ll/rf-f = ll/ll2+ e|M|J. II/.DjV\? = C\\DM\2 ■


We first prove that |^i | * 6 A f^ R ”1) for some a < 2. By the
assumption qx 6 Afp (Rm) for some p < 4. We may assume that p > 3.
Let cr be any number such that p/ 2 < a < 2. When m = 1, we have
m < a < p, and for all z 6 Rm,

f \qi(x)\dx<[[ dx f |gi|2 dx) < Ci


JB(z, i) 1) J B (t, 1) J

for some constant cx. So |q i|J 6 Af„ (Rm). When m > 2, we have
a < m and for all 2 6 Rm,

I M x ) | | x - z r m dx
Jb (»,i)
= / |X - ) " T |g i ( x )| |x _ 2 | f - T d x
JB{jt, 1)

- L, ^ “ z\(2tr~p)~m dx f |gx (x )|2 |x - , r » dx


JB{t, 1) ■'*(*,1)

< c2

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 380

for some constant C2 because of the assumption q\ 6 Afp(Rm) and


Lemma 6.20. So l^ l5 6 Af, (R m).
Then, by Theorem 6.21 and (53), there exists a constant C3 (de­
pending on tj) such that for all / and e,

| tail’ /«¥>[

< V WfMll + c3 H / e r f

= , | | ( l + | x | 1) ^ ( / , ¥.) ||! + C 3 | | / , r f

= 1 f l I x) l-F( M \ 2 dx + (c3+ l) WfMI2


j=iJ

- 4 / I-0J ( M l 2 dx + (c3 + 1) \ \ M 2
^ 7 E / V l-Dj (/7)|2dx + £ ||/,U jV ||J + (c3 + 1) Wf M?
* J=1 J=1

^ 1E / 1/ c-Oi/ - W - r (Di/ - 6,/)i2*>+££ hb^ii2+ (c3+ 1) v M f


i=i i=i
m
<1 E 11(0,' - « /II2+(X3+1)ll/ll2+£I s |\DM ? +(x3+i)Ml11■
Let e —►0, together with (52). We have the conclusion. |

For n > 1, let

Rn = { x G R m : n < |z| < 2n} , Qn = j z 6 Rm : j < |z | < 3 n j .

LEMMA 7 . 7 . Let Hq be as in the theorem. Then, there exists a


constant ci > 0 such that fo r all f € C * (Rm) and n,

£ 4 v .°i - w /i2 $ ( 4 iff»/i2 * + n<4 i/i2 *) •


PROOF. Choose r) € C“ (R m) such that 0 < r/(z) < 1 for all

z 6 Rm, Tj (z ) = 1 for all z with 1 < |z| < 2, and 77 (z ) = 0 for all z

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
T. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 381

with |x| < | or |x| > 3. Let 7/n (x) = 7/ For any / G C * (R m) and
n > 1,

(54) £ J \{Dj - bi) /|2 < E \\V n ( D j - bi) / ||2 ,


i= iJRn j=i

(55) EKDirjn) f \\2 < c2 f | / | 2 dx,


j=i

(56) ||T7„/||2 < [ \f\ 2 dx, ||77nJ7o/ | | 2 < / \H0 f \ 2 dx,


JQn JQn

where c2 is a constant such that \ ( D j r j ) (x)| < for all x, j . By Lemma


7.6 and the above inequalities,

(57)

l( ? .W , l . f ) \ < i E ll(0 j - + =3IIWII2


* J=1

= 7 E IIW n .) / + ( 0 ; - W /II2 + ‘3 M l 1
* J=1

< l JE i i ' ? . ( 0 , ' - 4 ; ) / i i I + ^ JQn


/ 1 i/ i2 * .

where c2 and c4 are some constants independent of / and n. B y the


assumption on 92,

~ { 9 2 Vnf, Tinf) < c(3n)* ||W I |2 < C5n* f | / | 2 dx


JQn

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 382

for some constant c5 > 0 . Now,

f>n(0y-*y)/||2
y=i
y=i
= t ( V n V i - 6;)2 / . / ) + 2 £ > n ( DjVn) (Dj - bj) f, f )
y=i y=i
= Be{r}nH of, rjnf ) - {qmnf , T}nf ) - {q2 r)nf , 77^ /) + 2 £ ] Re(r?n (D , - 6,-) / , (Dy77„)* / )
y=i
< ^ (llW II* + llln-ffo/ll’ ) + < 9 1 ^ /, 1?„/)| - (qiVnf, y * l ) +

+ £ ( J 111. W - bi) / f + 4 l l P i - l . ) / I I 2) •

Substitute (55), (56), and (57) into it we obtain

y£=i Ifrn (-Dy - bj) / | | 2 < j


4y=i \\rin (D j - bj) f \\2+ce ( f
V/Qn\H f 0 \2 dx + n6 f
«/<?„\ f \2 d z)
/
for some constant c% independent of / and n. The conclusion follows
from this and (54). |

Proof of Theorem 7.5. Choose <p 6 C " (Rm) such that 0 < <p (z) <
1, <p(z) = 1 for |z| < 1, and <p(x) = 0 for |z| > 2. For n > 1,
j = 1,..., m, and I = 1,2, let <pn (z ) = ( * ) , bjiTl = <p3 nbj, qiitl = <p3ng,,
and Hn be H 0 with all bj and qi replaced by 6y,n and qi<n. By Theorem
7.4, each Hn is essentially self-adjoint and (Rm) is a core of
Fix an arbitrary g 6 L j (Rm). There exists /„ 6 C£° (Rm) such that

(58) IIS- ( « . + ■ ) / . » < - •


71
Since |p ( z ) | < 1, we have

(59) \\(png - <Pn (Hn + ») /«|| <


71

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 383

Some direct calculations give

(60) - H n {v „ fn )\\

m
- 2 £ (D i V n ) ( D j - bj%„) /„ + /„Aov>„
J=1

<2 Z ( D j V„ ) ( D j - b j , n) f „ 1 + ^ l l / n l l ,
TV
i=i
where ci is a constant such that |Ao¥>(x)| < cx for all x € Rm. Since
(Z?yy)n)(x ) = 0 for all x with |x| < n or |x| > 2n, and &y,n(x) = 6y(x)
for all x with |x| < 3n, we have

(61)
2

4
Z ( D , V n )(D j - h - ) U
j=i
where cj is a constant such that \Dj<p (x)| < ci for all x and j . Noticing
that Hofn (x) = Hnf n (x) for x 6 Qn, we have, by Lemma 7.7, for some
constant C3 independent of n,

(62) E /„ \{Dj - bj) /.I* <b < c3 (||/r „ /„ ||2 + n‘ ||/„ ||2) .
j= 1

By (58), H fU .ll2 + ||/„ ||2 = ||(fT. + i ) / 4 ‘ < ( ||,|| + i ) 2, so ||fr„/.ll


and ||/n|| are bounded. Then, from (60), (61), (62), and the condition

6 < 2, we see that || <pnHnf n - # n (^ n /n )|| -* 0 as n —» 0 0 . Recalling


that Hn (<Pnfn) = Ha (<Pnfn), and obviously \\<png - y|| -» 0, we have,
by (59), ||^ — (H 0 + i) (tpnfn)\\ —►0 as n —►00 . So, R( Ho + i) is dense
in Xrj(Rm). Similarly, R( Hq —i ) is dense. So, Hq is essentially self-
adjoint (Lemma 3.1 (e)). |

N o te . Notice that the sequences appearing in the proofs o f the


lemmas and the theorem are all constructed uniformly. No inductions
are involved.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 384

To apply Theorem 7.4 and 7.5, we must verify that the relevant
potentials are in Mp (R n) for some p < 4. We say that a function q
belongs to (Lp + L ^ ) (Rn), if q = qi + q2 for some q\ 6 Lp (Rn) and
q2 G I * (R n). Clearly, L * (Rw) C Mp(Rn).

Lemma 7.8. If q € (Lp + La,) (Rn) with (i) n < 3 and p = 2; or


(ii) n > 4 and p > y, then q € Mp (Rn) fo r some p < 4.

PROOF. It suffices to show that Lv (R n) C Mp(Rn). The case of

n < 3 is obvious. Suppose that n > 4. Then p > 2 and 4 > 2a. Choose
p such that 4 > p > y , then < n. By Holder’s inequality
(Theorem(3.4) on p.314, B&B),

2 2 / \ 8=2

/x B ( ^ ) j ^ “ U < [JxB (z,D k r ) ” •

Then it follows that q € Mp (Rn). I

Take R3n as the phase space of a system of n particles. Write x €


R3n as x = (x ? ,...,x ^ ), where x? = (x3,_ 2, x3j_i, x3i) is the position
coordinates of the z-th particle. Denote A , = + ff* t +
g— , and let |x?|, |x? — xy| be the norm in R 3.

LEMMA 7 .9. For i 7^ j and any constant a > 0, jg»"gVj 6


Mp (R 3n) fo r some p < 4.

PROOF. The case n = 1 is obvious. Suppose that n > 2. Wecan

assume that p > 3. For z € R3", let

z x = (...,5 ili,ii+ x ,...) G R3(n_1),

z lJ = (..., z f lt , ..., z ^ t , ...) 6 R3(n-2>.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 385

Note that

XB(x.i) ( s ) < X B (i) (s ? ) XB(x>,i ) ( a s * )


|i ? |2 |S - Si3""' - J i^ l2 ' 11* - *

By Lemma 6.20 and Lemma 6.2, we see that j=*y E Mp (R 3n).


Denote ~y = a x j . We have

XB(x.l) ( g ) < XB(x?,l) ( x ? ) XB(a*?,a) ( Y ) XB(«»J,1)


|s ? - a x } \ 2 \x - z\3n~p ~ \ x i - y*\2 | i ? - 2 ? |3 _ £ fi I3*"” 2) - * ?

The second factor is integrable with an integration bounded for x‘,J.


The first factor is integrable on R6 (as a function of x?, 1/*) by Lemma
6.19. To see that the integration is bounded for ~zl, note that

| ~zt - a z j | - a - 1 < | i i ~ V l < \ X ~ a'zj I + a + 1

for Xi E B ( z i , 1) and ~y* E B (a z y , a ) . Consider the cases | ~Zi — a~Zj |—

a — 1 < 1 and |z? — a z j \ — a — 1 > 0 separately, we see that the


integration is bounded by

XB(*?,1) ( Xi ) Xfl(a*?,a) ( V ) j ^ \ XB(*?,1) ( X* )


;/(/ -----=T2------------d y ~ -----~ 3 e = i d X i

_ 1 f XB(2a+3) + (xB(r+,+l) ~ XB(r— 1)) f XB(x?,1) (* ? )


~ _ I I—>.2 y / _. _, , o-3 dxj
aJ ly l J - ~zZ\

for some r > a + 1. By the proof of Lemma 6.20,

f T ^ n id~y < ((»■ + « + 1 ) - ( r - a - 1 )) e = 2 ( a + 1) c


J B (r+ « + l)-B (r -a -l) | y*|2 " K '

for some constant c. So the integration is bounded by a constant and

hence isisrI 6 ^p(R3")- ■


As corollaries we have the following self-adjointness results.

THEOREM 7 .1 0 . (cf. [16] Theorem 14.1.2 and I 4 .I . 4 )

1
R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 306

(a) If V € (Lp + Loo) (Rn) with (i) n < 3 and p = 2; or (ii) n > 4
and p > y, then —A 0 + V is essentially self-adjoint on (R n) and
D(-A 0 + v ) = W 2 ,2 ( Rn).
(b) (Kato Theorem) If V (x ) = £" = i K ( x?) on R3*1, and V- G
(L 2 + L00)( R 3) /or each i, then —Ao + V is essentially self-adjoint

on <?0“ (R 3n) and D ( - A o + K) = Wa.j (R 3*1).

PROOF, (a) is clear. The proof of 6 Mp (R3n) above also applies


to Vi ( x t ). |

T h e o re m 7.11. Write x e R3n+3 as X = ( *o, x t, ...,xH) and de­


note A ,• = g g j, di = £ t f o r i = 0 ,...,n.
faj ^[16] p.446 Example 14-1.5) The Hamiltonian o f a neutral atom
with N electrons,

ii- 1 A 1 V a £ N e* e2
2^ ° 2 m fe ‘ £ I * - iJI + 2 £ |i? - i? |

is essentially self-adjoint on (R3n+3).


W ft®5 ] P-f^O Example 4) Let ~a ( f x) = |~x x B0, where B 0* 6 R3
is constant. The Hamiltonian o f an atom with N electrons in a constant
external magnetic field,

H = - m {d° ~ i N e ^ { n ) ! c ) 1 ~ ' L ' i=i


£ ‘ { d i+ , v f ( * ) /<=)’

A N e2 1^ e2

is essentially self-adjoint on C * (R 3n+3).

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 387

(c) ({85] p.200 Example 5) The Hamiltonian o f an atom in a con­


stant electric field E q,

1_ 1 ^ N e2 ,1 ^ e2
2M ° 2m f r [ ‘ \xi - i £ | 2 g j | i t - Xj |

+eS • ~ Y ,^ J >

ts essentially self-adjoint on C “ (R3n+3).

PROOF. These follow from Theorem 7.4, 7.5, and Lemma 7.9. Con­

stants ^ and so on can be eliminated by changing variables, and


we clearly have, for any p > 0, eE0 • ( x £ — 6 M Ptioc( R 3n+3)

and eEo • ( xq — ®i) > ~ c lx l f°r some constant c. |

THEOREM 7 .1 2 . ({85] p .174) Harmonic and anharmonic oscilla­


tor Hamiltonians —gjy + ui2 x 2 + b2 x2m (m = 2 ,3 ,...) are essentially
self-adjoint on (R ).

PROOF. This follows from Theorem 7.5 since u/2 x 2 +b 2 x 2m € MPtioc(R)

for any p > 1 and is non-negative. |

7 .4 . A n gu lar m o m en tu m o p era to rs. Consider the angular mo­


mentum operator L° = xi pXl — x2pXl = MXlpXl — M X3 pXl defined on

Cq° (R 3), where pXi = - i ^ .

THEOREM 7 .1 3 . L \ is essentially self-adjoint on (R 3) and for


f e S ( R 3)

L \ f = (xxp„ - x 2p ,J / .

PROOF. We follow the proof on [79] p.369, using Stone’s Theorem.


For t 6 R, define U ( t ) : L%(R 3) - L 2 (R3)

U( t ) ( / ) (x) = / ( x i cost —x2 sint, i i sin t + x 2 cosf, X3 ) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
7. SELF-ADJOINTNESS OF QUANTUM MECHANICAL OPERATORS 388

For / E (R 3), it can be shown that \\U (t) / | | = ||/||. Since C“ (R 3)


is dense in L 2 (R 3) , U (t) is unitary. { U (t) : t E R } is clearly a group.
For / 6 (R 3), It is obvious that ||(Z7 (t ) — 1) f\\ —» 0 as t —» 0. Since
U (t) is uniformly bounded for t E R, it follows that { U (t) : t E R } is
strongly continuous. By Stone’s Theorem (Theorem 5.2),

= i l i m i ( « 7 ( 0 — 1)
1 t— »o t
is self-adjoint. Suppose that / E C " (R 3), from the proof of Stone’s
Theorem, {L\ + i ) T xf = i f , where Txf = f e ~ 'U (s) fda. Let g be
defined by g ( x ) = / e~*U (s) ( / ) (x) ds. It is easy to see that T \ f = g
and g € C ? (R 3). So (L x + i) (C 0« (R 3)) D (R 3) is dense in L 2 (R 3).
Similarly, (Lx - i ) (C “ (R 3)) is dense. So C " (R 3) is a core of L \. It
remains to prove that when f E S (R 3), L \ f = (xipXj - x 2pXl) / , for it
implies L \ = Li|(7q°(R3) and hence Lx = L°. By the assumptions it
can be shown that, as t —►0,

J (U ( 0 ~ !) ( / ) (*) (*1 P*i ~ * 8 * 0 / ( * )

uniformly on any bounded subset of R 3. So i (U (t) - 1) / - » (xxpSi - x 2pXl) /


in measure. For any x E R3 and t E R, t ^ 0, by the mean value theo­
rem (B&B, p.48), there exists t' E [0,£] (or [t,0]) such that

1*1
— 1 |2
* |‘ + i
So,

5 ^ (*■ )(/)(*) +TTiT


i*r+i
Since / E S (R 3), (£') ( / ) (x)| < ~ ^c~i for some constant c. So

| \ { U ( 0 — 1) f \ is dominated by an integrable function. Hence j ( U (t) — 1) /


(*1Px7 - xiPxt) f in Li (R 3) . |

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
APPENDIX C

Comparison with Related Work

This appendix will briefly compare strict constructivism with some


other approaches to the foundations of mathematics. The relevance of
negative results in recursive analysis will also be discussed.
First of all, there is a fundamental difference between the objective
of strict constructivism and that of some other approaches to the foun­
dations of mathematics. Strict constructivism is not mean to provide a
foundation for mathematics. It is a logical tool primarily for studying
this problem: Is infinity really indispensable in mathematical applica­
tions in science? It is also expected to be useful in analyzing the logical
structure of mathematical applications. Adherence to finitism is cru­
cial for this purpose. This has been explained in Section 2, Chapter
2. Various nominalization programs and predicativism all accept real
infinity, and therefore cannot serve the purpose here.
The case of intuitionism and Bishop’s constructivism is a little
more subtle. When Bishop’s constructive analysis is formalized in a
quantifier-free language such as the language of To, in appearence, it
does not commit to infinity, for the language is quantifier-free. How­
ever, it has been indicated at the end of Section 2, Chapter 1 that
infinity or quantification is implicitly assumed, for they are needed in
proving that terms of To terminate. It is also indicated at the end of
Section 2, Chapter 2 that the language of P R O , whose terms represent

389

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 390

primitive recursive functions only, can help better in explaining why


mathematical inferences preserve approximate truth.
Since strict constructivism is weaker than predicativism and the sys­
tems accepted by various nominalization programs, anything developed
in strict constructivism is also available to other systems. However, the
nature of the question whether strict constructivism is sufficient for sci­
entific applications is different from that of the same question for other
systems. Those approaches generally expect a wholesale answer to the
question. For example, Feferman [42], [43] conjectured that predicative
systems he developed are sufficient for all scientific applications. This
is based on the general observation that there are some fundamental
axioms for developing modern analysis, namely, inductions (including
transfinite inductions), comprehension axioms and so on. As long as
these axioms are available, ordinary mathematical proofs in analysis
can all be obtained. Therefore, there is no need for examining each
particular theorem and its proof.
On the other hand strict constructivism is too restrictive. This
type of general approach does not work for strict constructivism. One
has to study the actual applications of mathematics in each branch of
science, in particular, in each branch of physics, to see if real infinity is
indispensable and if strict constructivism is sufficient there. This is far
more laborious, but it also makes possible a more accurate description
of mathematical applications and a better understanding of the role
of mathematics in science. It makes it possible to answer questions
such as ‘Is infinity really indispensable in the scientific description of

nature?’

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 391

Friedman., Simpson and some coworkers have shown that a signif­


icant part of modem analysis can be developed in two weak subsys­
tems W KLo and W K L q of second-order arithmetic Zj. These two
systems are II, conservative over primitive recursive arithmetic P R A .
This work is interpreted as a partied realization of Hilbert’s program by
Simpson [95]. The differences between strict constructivism and these
systems are described as follows.
It has been explained in Section 2, Chapter 1 that every theorem
of P R O is a schema of some theorems of P R A , in the sense that
when free variables of higher types are instantiated by closed terms,
the theorems of P R O are simply reduced to theorems of P R A . In this
sense, P R O does not contain ideal elements in Hilbert’s sense. It be­
longs to the part of mathematics that has real content. The system SC
just provides some convenient notations (some schematic notations) for
expressing proofs in P R O . So developing mathematics in SC means
developing mathematics with real content. More specifically, mathe­
matical theorems in strict constructivism have computational content.
They are about (primitive recursive) programs.
Hilbert’s idea is that the inffnitary part of classical mathematics
should be considered as consisting of ideal elements. Mathematical
propositions in that part of mathematics, for example, propositions
expressed by sentences with quantifiers ranging over infinite domains,
should be considered as meaningless in themselves. They are useful
tools that can help to streamline mathematical proofs and derive the­
orems with real content. Hilbert’s program aims at using mathemat­
ics with real content to justify this instrumental function of infinitary
mathematics.

p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 392

Therefore, the question whether strict constructivism is sufficient


for mathematical applications is just the question whether ideal ele­
ments are really needed in applications and whether Hilbert’s program
is needed as far as applications of mathematics are concerned. It is dif­
ferent from Simpson’s question about to what extent Hilbert’s program

can still be realized.


The systems W K Lo and WKLq certainly contain ideal elements.
They are formulated in second-order language, with second-order quan­
tifiers. Theorems of these systems generally do not have computational
content unless they are sentences. For instance, suppose K (m, n, p)
is the Kleene predicate. Then the sentence

\/m 3 p iq ( K ( m ,m ,p ) V ->K (m ,m ,g ) )

is provable in classical first-order logic, and therefore in the systems


W KLo and W K L q . But there is no recursive function / such that

VmVg(K (m, m, / (m )) V -'K (m ,m , q ) ) .

So proving theorems in these systems and proving theorems in strict


constructivism have different meanings.
When physical laws are formulated as mathematical propositions in
classical mathematics, the propositions are sometimes ideal elements in
Hilbert’s sense. Then these laws should be considered as meaningless
according to Hilbert’s approach. This means that the instrumentalist
philosophy must be extentded to physics. On the other hand, applying
strict constructivism is applying mathematics with real content.
As for concrete mathematical theorems, we can expect that the
systems W K Lo and W K L q can prove more theorems than strict con­
structivism can do, for the logic and axioms of W KLo and W K L J

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPEN DIX C. COMPARISON WITH RELATED WORK 393

are more powerful. For example, the Heine-Borel covering theorem,


the uniform continuity theorem, the Cauchy-Peano theorem for ordi­
nary differential equations and the Hahn-Banach theorem for separable
Banach spaces are provable in W K Lo. Moreover, the open mapping
theorem and the closed graph theorem for separable Banach spaces
are provable in W K L q . See Simpson [95], [94], Brown and Simpson
[32], [33]. These theorems are not constructive in their original forms.
Given the open mapping theorem and closed graph theorem, the theory
of unbounded linear operators on Hilbert spaces, including the materi­
als developed in Appendix B above, is expected to be formalizable in
W KLq.

When some theorem is proved to be equivalent to the main axiom


of W K Lo, the weak Konig lemma, it is generally non-constructive.
However, proving classical theorems in strict constructivism means ex­
tracting computational contents of the theorems. In many cases, we
can still prove some constructively weaker versions of the original the­
orem by stating the assumptions of the theorem in some constructively
more informative format, or by adding some classically trivial but con­
structively informative assumptions, or by weakening the conclusion
into approximations in place of an exact assertion. These weaker ver­
sions are supposed to reveal the computational content of the original
theorems. For example, a weaker version of the Hahn-Banach theorem
is proved in B&B. In Appendix B, it is shown that we can bypass the
closed graph theorem and the open mapping theorem in developing
the constructive theory of unbounded operators. This is supposed to
reveal the real computational content implicit in the spectral theorem,
Stone’s theorem and self-adjointness and so on.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 394

Sometimes, it even happens that classical theorems that are in­


dependent of the systems W K L o and W K L q have some construc­
tive versions that are provable in strict constructivism. For exam­
ple, some versions of Lebesgue’s dominated convergence theorem and
Radon-Nikodym theorem are shown to be equivalent to arithmetic com­

prehension, and are therefore independent of the systems W K L o and


W K L q • (See Yu [108], [107].) However, some constructive versions

of these theorems have been proved in B&B and are formalizable in


SC . So the question how much of mathematics can be developed in
strict constructivism is not directly related to the same question for
the systems W K L o and W K L J •
Pour-El, Richards and coworkers obtained some results of non­
recursiveness in analysis. See, for example, Pour-El and Richards [76],
[77], [78].1 Here is a discussion of the relevance of these negative results
to the sufficiency of strict constructivism for applications.
From the constructivism’s point of view, some of these results axe
Brouwerian weak counter examples to some classical theorems. A
prominent example is the result in [76]. It actually gives a Brouwerian
weak counter example to the following classical theorem: If the function
/ ( x , y ) is uniformly continuous on the rectangle { —1 < x < 1, —1 < y < 1},
then there exists a differentiable function ip (x) such that ip (0) = 0
and ip' (x) = f ( x , <p(x)) for x in some interval [—a, a].2 In such
cases, a non-recursiveness result indeed shows that the theorem is non­
constructive.

1See also Baes [4] for some positive results on the recursiveness of Hamiltonians
in quantum mechanics. The work in this dissertation improves that work although
the background framework is different.
3When / (*, y) satisfies further the Lipschits condition, the existence of the
solution can be proved constructively.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 395

However, some other examples actually show that the framework


of constructive analysis is more natural for proving theorems formally
similar to their classical counterparts than the framework of recur­
sive analysis. For example, the First Main Theorem in Pour-El and
Richards [78] says that an unbounded linear operator on a Banach

space maps some computable vectors in its domain to non-computable


vectors in its range. From the constructivism’s point of view, all vec­
tors' are constructive and if the operator cannot map a constructive
vector to another constructive vector, the former vector cannot belong
to the domain of the operator. So the theorem becomes meaningless
when translated into constructive analysis literally. The theorem ac­
tually shows that when we consider the same unbounded operator in
constructive and classical settings, the domains may be different. Some
vectors provably belong to the domain in the classical setting may not
belong to the domain in the constructive setting, even if the vectors
themselves are constructive.3
A more concrete example is this: There are constructive continuous
functions that are classically differentiable but not constructively dif­
ferentiable. The non-recursive solution to wave equation constructed
in Pour-El and Richards [77] is based on this example. In constructive
analysis, the solution to the wave equation

d 2u d 2u d 2u d 2u _
d^+ d^+ d^~d^~ ’

3It is trivial that some vectors in the classical setting are not recognisable in
the constructive setting. For example, some functions are not constructive. This
First Main Theorem is certainly deeper than that.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 396

with the initial conditions

u ( x ,y ,z , 0) = f(x,y,z),
du
— ( x ,y ,z , 0) = 0,

given by Kirchhoff’s formula

u(x,t) = f f [f (x + tn) + 1 (grad / ) (x + tn) • nj do (n)


J J unit cphere
(cf. [77]) obviously exists if / is indeed constructively differentiable, for
the formula already constructs the solution from / and partial deriva­
tives of f . 4 Pour-El and Richards’ example uses a function / that is
classically differentiable but not constructively differentiable.
When we do mathematics in the classical setting, we recognize both
recursive and non-recursive objects, and we can treat unbounded oper­
ators, have results such as the spectral theorem and have solutions to
wave equations. But recursive objects may be mapped to non-recursive
objects and solutions may be non-recursive for recursive initial condi­
tions. Therefore, some mathematical theorems will be lost if we decide
to accept recursive objects only.
On the other hand, when we do mathematics in the constructive
setting, we recognize only constructive objects, and therefore we actu­
ally discard some unwanted objects. We have to expell those objects
from the domain of our discussion if we cannot map them into objects
we can recognize. As a result, we can still treat unbounded operators,
have results such as the spectral theorem and have solutions to wave
equations. Things will go smoothly again. That is, the theorem is still

4The theory of distributions is still to be constructivised. Here we are only


concerned with the fact that the solution given by the formula is constructive.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 397

formally valid, although the contents will be different from the classical
point of view.
Moreover, there are actually two types of constructive settings:
Bishop’s constructivism and strict constructivism. In the frame work of
strict constructivism, we accept only primitive recursive objects. The

work in this dissertation shows that things will go equally smoothly if


we restrict everything to the primitive recursive setting, which means,
for example, that only primitive recursive real numbers are recognized.
Of course, not everything will go smoothly when we restrict to
the constructive setting. In case there are genuine Brouwerian weak
counter examples, some classical theorems will have to be abandoned.
Pour-El and Richards’ example about ordinary differential equations
cited above belongs to these cases. The initial condition is construc­
tively uniformly continuous, but the existence of the solution cannot
be proved constructively. These cases should be distinguished from the
cases like the First Main Theorem.
Sometimes a classical existential theorem can be proved to have a
recursive solution, but the theorem is not constructive. An interesting
example of this type is the Intermediate Value Theorem proved on pp.
41 of Pour-El and Richards [78]. Brouwerian weak counter examples
to this classical theorem are well known. The proof of the recursiveness
of the intermediate value given in [78] is non-constructive. It is based
on a distinction between two cases: (1) / ( c ) = 0 for some rational
number, and (2) / ( c ) ^ 0 for all rational number. These are generally
undeddable. The proof does not provide an algorithm for finding the
intermediate value, although it proves that, ‘platonically’, there exists

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited without p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 398

such an algorithm. This seems to show that if finding the computa­


tional content of mathematics is the recil concern, constructivism is a
better framework than recursive analysis.
Finally, we consider another problem: Even if we have some math­
ematical theorems in strict constructivism that are formally like the
classical theorems used in mathematical applications, can it happen
that restricting to constructive or primitive recursive setting will still
compromise mathematical applications because of the limitation of the
real numbers and functions accepted? In the papers [55], [56] published
several years ago Heilman suspected that this is the case. However, we
believe that this doubt can be put aside, at least as far as applications
in quantum mechanics are concerned.
Generally, there are perhaps two possible ways by which the restric­
tion to strict constructivism may compromise applications. The first is
the one cited above: If some physical quantities or events can only be
represented by non-primitive recursive real numbers or functions and so
on, strict constructivism will not be able to 'represent those quantities
or events.
There seems to be no problem of this type for strict constructivism,
at least as far as applications in quantum mechanics are concerned.
Recall that physical theories can only give approximate descriptions of
nature. We use real numbers to represent space-time points in most
applications, although we are still not sure if space-time is continu­
ous. Moreover, the continuity of space-time does not mean the same
thing as the continuity of the real number system. Continuity in the
physical sense only means density in the mathematical sense. In other
words, there are no minimum gaps in space-time structure. Rational

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX C. COMPARISON WITH RELATED WORK 399

numbers are already dense. Further, there are various kinds of non­
standard models of the theory of real numbers, including models with
infinitesimal numbers. Clearly, in physics there is no way to decide if
real space-time is literally isomorphic to the classical continuum. So
it seems we can confidently claim that real numbers in strict construc­

tivism are sufficient for representing space-time points.


Similarly, any continuous function on R can be approximated on
compact interval by polynomial functions with rational coefficients, or
by step functions on rational intervals with rational coefficients. The

same is true for Lebesgue integrable functions. In physics, if the abso­


lute difference between the values of two functions on any point (ex­
cept for a zero-measure set) is insignificant when compared with some
physical constants, such as the Planck constant or the minimum space
distance presently reachable (say, 2-43cm), those two functions make
no difference in expressing physical states. So we can always replace
a function in classical mathematics by a function acceptable to strict

constructivism. In Appendix A we see that starting with continuous


functions in strict constructivism, we can construct various Lp spaces
of functions and prove their completeness. Therefore functions recog­
nizable in strict constructivism seem to be sufficient for representing
physical states, at least in quantum mechanics.

In general, presently we do not know any particular reason for


doubting the sufficiency of mathematical entities in strict construc­
tivism for representing physical quantities, states or events. Certainly,
the question is still open and further studies are interesting.
The second possible way is this: If some crucial mathematical the­
orems related to a physical theory are not available in strict construc­
tivism, and as a result a satisfactory framework of the physical theory

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
APPENDIX C. COMPARISON WITH RELATED WORK 400

cannot be set up (unless one develops a new physical theory to replace


to old one), then there will be serious doubts for the sufficiency of strict
constructivism, even if some physical results can still be obtained in
some ad hoc manner. Fortunately, in Appendix B it is already shown
that the basic mathematical theorems for the framework of quantum
mechanics are available in strict constructivism. So we can state the
formalism of quantum mechanics in the same way as we do when we
use classical mathematics. There is an underlying difference: In strict
constructivism, all objects, namely, real numbers, square integrable
functions and so on, are primitive recursive in some proper sense. But
formally everything goes as smoothly as in the classical case: We still
have Stone’s theorem that guarantees the existence of a Hamiltonian
for any physical system; we still have the spectral theorem based on
which we can state the rules for probability predictions; and so on and
so forth.

It is still unknown what will happen if we consider the more ad­


vanced developments of quantum mechanics or quantum field theory.
This seems to be an interesting topic to pursue.
Finally, it may be necessary to emphasize that the above discussion
has nothing to do with the anti-realistic interpretation of science. One
can concede that electrons, quarks and so on all exist and that physical
theories give approximately true descriptions about them. The discus­
sion only rejects that entities in classical mathematics provide exact
models of them. Moreover, we do not want to make any dogmatic
or normative assertion here. The study started in this dissertation is
analytic in nature. The outcome about the sufficiency of strict con­
structivism for mathematical applications could be either way. We
expect that either outcome will be of interest.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
APPENDIX D

Normal Forms and Conservativeness

1. T h e E x iste n c e o f N orm al Form s

Here we show that the existence of normal forms of terms of P R O


can be proved finitistically.
First we need some notions. The height of a redex Ap (J (t, ti, t 2) , s)

or A p ( \x .t, r) is the height of the type of J (t,ti,ta ) or ^x -*- A. one-


step strict reduction is a one-step reduction of a term t such that the
reduced redex is of the maximum height among the redices in t, and
either it is the rightmost occurrence of a J-redex of that height in t if
there is such a J-redex, or it is the rightmost occurrence of a A-redex
of that height if there is no 7 -redex of that height. (We agree that if sj

occurs in sj and s 2 occurs in t, then the corresponding occurrence of si


in t should be counted as an occurrence at the right of the occurrence

of s 2 in t.)
So if t is not normal, there is a unique ti such that (t , ti) is a strict
reduction. Clearly, t\ = sr (t) for some primitive recursive function sr,
where we agree that s r (t) = t in case t is already normal. Beginning

with any term t we can then construct a primitive recursive sequence of


strict reductions t = sr° ( t ) , sr ( t ) , s r 2 ( t ) ,.... We must find a primitive
recursive function i such that ar'W (t) is a normal term for any t. Then
the complete reduction sequence cr (t) can easily be obtained.
For each term t, let m (t) be the maximum height of the redices in
t, which is 0 if t is normal, and let n (t) be the number of occurrences of
401

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
1. THE EXISTENCE OF NORMAL FORMS 402

J-redices of the height m (t), and let I (i) be the number of occurrences
of A-redices of that height. Now suppose that t is an arbitrary term
and m ( t ) > 0. If n(£) > 0, in the one-step strict reduction from t,
the reduced redex is a J-redex A p ( J (t0, t i , t 2) ,s ) and it is reduced to
J(<o, A.p{tu s) , A p { t 2, s ) ) . In case ti or t 2 or both are still J-terms,
one or two more J-redices of the same height will be created in this
one-step strict reduction, but no new redices of any higher height will
be created. If f2 is a J-term, Ap (i2, s) must be the reduced redex in the
next one-step strict reduction. Continue strict reductions we will reach
a term in the position of t 2 which is not a J-term and then for the next
strict reduction we will go back to the corresponding A p ( t i , s ) and see
if ti is still a J-term. Continuing the procedure we will finally reduce
all J-redices of the same height created by the original reduction and
decrease the number of J-redices in t of that height by one. This is
similar to a right-first binary tree search. We can define a primitive
recursive function / such that / (t ) is the number of strict reductions
in the process. That means n (sr W (t)J = n ( i) —1 if n( t ) > 0. To see
this, it suffices to define f such that

/'(*> A p ( J ( t 0 }t u t3) ,s)) = f {t, A p ( t u s)) + f ' ( t , A p ( t 2, s)) + l,

if J(fo,*i»* 2) is o{ the height m ( t ), while f ' ( t , t') = 0 if t' is of other


forms. Then / (t ) = f (t, t) and by quantifier-free inductions alone we
can prove that s r ^ (t) contains no J-redices of the same height as t.
By iterating the function t —* ar*® (t) up to n (<) times we have g such
that n (tfj = 0, where g ( t ) = 0 if we already have n ( t ) = 0.
Next, for any t', if n (<') = 0, the reduced redex in the strict reduc­
tion from t is a A-redex of the height m (t '). It is routine to check that af­
ter the one-step strict reduction no new J-redices of the same height will

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. THE CHURCH-ROSSER PROPERTY 403

be created and the number of A-redices of that height decreases by one.


That means, after / (t') strict reductions, we must reach a term with a
strictly less maximum height of redices, that is, m (sr1® (t'fj < m (t1) .
Then, for an arbitrary t, let h ( t ) = g ( t ) + I (sra® (t)) . We have
m (« •* « (*)) < m ( t ) . By iterating the function t —♦ s r h® (i) up to

m ( t ) times we can finally define i such that m (sr*® (t)j = 0. That is,
sr*® (t ) is normal.
Assertions here, namely n ( s r 9® (Z)) = 0, m ( s r h® (t)j < m ( t )
and m (<)) = 0, are all proved by inductions on the defining
equations of primitive recursive functions. So the proofs axe formal-
izable in PRA. This completes the proof of the existence of normal
forms.

2. The Church-Rosser Property

Here we show that the Church-Rosser property can be proved fini-


tistically.
First we introduce two new symbols to the language: A* and J *,
called marked A and J, and we add the following to the formation
rules for terms: A p ( \ * x . t , s ) and Ap(J* (t,ti,ta ) ,s ) are also terms if
t, ti, t2, and s are terms (of appropriate types). These are called marked
redices, and marked symbols occur only in such contexts. For the
reductions we add the following: Ap(X*x.t, s) is replaced by t [s/x] and
Ap{J* ( t , t i , t 2) ,s ) is replaced by J ( t , A p ( t i , s ) , A p ( t 2 ,s)). Clearly,
the result of a reduction is still a term in the extended language. The
notion of normality does not change. Notice that normal terms cannot
contain marked symbols. For any term t of the extended language, jt|
denotes the term obtained from t by erasing all stars.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
2. THE CHURCH-ROSSER PROPERTY 404

The operation, f that reduces all marked redices simultaneously is


primitive recursive, for it can be defined by recursion:

f ( A p ( J ' ((, („ h ) ,«)) = J ( / ( i ) , / (Xp ( « , , . ) ) , / ( A t (<,, s ) ) ) ,

f ( A p ( \ " x . t , .) ) = / ( 4 ) [ / ( s ) / x ] ,

and / preserves all other ways of constructing terms. An induction (on


the construction of t) will show that

/ (t [■ /* ]) = / ( * ) [ / (■) / X ] .

We can define a primitive recursive function g and prove, by induction


on the construction of i, that if (f, a) is any one-step reduction, then
g ( ( t , s ) ) is a sequence of reductions from f (t) to / ( a ) , where in case
f ( t ) = f (a) we take ( / (t)) as g ( ( t ,s ) ) . Here are a few cases in the
definition of g: (1) if t is J (<0, t x, t 2) and a is J (t0, t'u t 2) , where (ti,t[)
is a one-step reduction, then g ( ( h , t [ ) ) = ( / ( t i ) , ...,/( * ( ) ) is available
by the inductive hypothesis and

s((M » = M ) J ( f (to) , / ( t ( ) , f ((,))) i

(2) if t is Ap(J* ( t o , t i , t 2) ,r ) and a is J(to, A p { t u r ) , A p ( t 2 , r ) ) , then


f ( t ) = / ( a ) ; (3) if t is Ap(Xx. t 0 , r ) and a is t0 [r /x ], then /(< ) is

Ap (A x ./ (t0) , / (r)) and / (a) is / (<0) [ / (r) / x ] , and ( / (<), / (a)) is a


one-step reduction. Then by an induction on the length of a sequence
we can further define a primitive recursive function h such that if a is a
sequence of reductions from t to a, then h (a) is a sequence of reductions
from f ( t ) to / (a).
Now suppose that t is a term in the original language (with no
marked symbols), a is a sequence of reductions from t to a, and (t, t')
is another one-step reduction. We mark the occurrence of the symbol

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER P R A 405

A or J in t that is reduced in the one-step reduction (t, t') and obtain


a term t* such that |t*| = t and /(< * ) = t'. Clearly, corresponding to
the sequence a of reductions from t to s, we can construct a sequence
a* of reductions from t * to some term a* such that when all the stars
in a * are erased we will obtain exactly a. Notice that we must have
|a*| = s. Then h(a*) is a sequence of reductions from / (t*) = t' to
In case s is normal, we must have s* = 3 and / (a*) = a. So
we have a primitive recursive function p such that if (t , t‘) is a one-step
reduction, a is a sequence of reductions from t to a normal term a, then
p ( ( t , t ' ) , a) is a sequence of reductions from t' to a.
Finally, by induction on the length of the sequence 6, we can de­
fine a primitive recursive function q and prove that if b and a are two
sequences of reductions from a term t to normal terms a and s' respec­
tively, then 5 ( 6, a) is a sequence of reductions from a to a '. Since a, s'
are normal, we must have a = a'.
All the constructions involved are primitive recursive and all induc­
tions are on quantifier-free formulas. Therefore, the proof is finitistic.

3. C o n serv a tiv en ess o f P R O over P R A

Now we prove the conservativeness of P R O over P R A . The idea


of the proof is as follows. Consider a proof of a formula of P R A in
P R O . We first instantiate all free variables of types higher than o by
some closed terms. Then we reduce all the terms in the proof to normal
forms. This will result in a proof containing no A-terms. Then we can
eliminate Re-terms and J-terms by assigning a constant /* to each
numerical term t[ni, ...,n*] with free variables all in n i, ..., 71* so that
f t (n i, ...,rifc) = t and then replacing the term t by /* (ni, ...,n*.).

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 406

We first give the details about eliminating A-terms. For every for­
mula if, we define its normal form if* as follows: (t = a)* is t* = s*,
(<f V ip)* is if* V ip*, (ip A ip)* is f* A ip*, (if —» ip)* is if* —►ip*, tf is
called normal if all terms in it are normal. Then we have

LEMMA 3 .1 . (1) F o r r a term of the type o, t [r/n]* = t*\r*/n\,


V5 [r/n]* = <p* [r*/ n] ;
(2) The normal form s o f the axioms o f P R O except the selection
axioms and X-axioms are still axioms of P R O ]
(S) The operation * preserves inference rules o f P R O .

PROOF. (1) Consider such a sequence of reductions beginning with


t [r/n]: first reduce all occurrences of r obtained from the substitution
to the normal form r* and obtain t [r*/n], and then reduce other redices
in t [r*/n]. Since r* is of the type o and no free variables of r* become
bound in t [r*/n ], it is routine to check all the possible cases to see that
if {t, t') is a one-step reduction, then (f [r*/n],£'[r*/n]) is also a one-
step reduction. So t [r*/n] is finally reduced to t* [r*/nj, which must
be normal as can be verified easily. By the uniqueness of the normal
form, this must be t [r/n]*. The rest of (1) follows by induction.
(2) and (3) follow from (1) easily. |

Now we consider the selection axioms. We need a lemma.

LEMMA 3 .2 . (1) There is a primitive recursive function f such


that if s\ = 3j is a selection axiom and {sx,s[) is a one-step reduction,
then / ( a i , 3a , 3 i) is a sequence of terms = r x ,...,r n = s '2 such that
r» = r\+i is a selection axiom fo r i = 1,..., n — 1, and either (sj, s'2) is
a one-step reduction or 53 = s2.
(2) There are prim itive recursive functions g\,g? such that i f s x = s 2
is a selection axiom and a is a sequence o f reductions from Si to s[,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 40T

then gi (si, 3 %, a) is a sequence of terms s^ = r i,...,r n = s '2 such that


ri = r,+i is a selection axiom fo r i = 1 , n — 1, and gj (s i, sj, a) is a
sequence of reductions from S2 to s^.

PROOF. (1) We consider the case Sx = s { J ( 5 r /,,r /,r )} and S2 =

s { r } . The case of J (0, r', r) is similar. / is defined according to the re­


duction (sx, si). If the reduced redex does not overlap with J (Sr", r', r)
or it is contained in one of r/,,r ,Jr, this is obvious. So suppose that
J (Sr",r ',r ) occurs in the reduced redex. First consider the case when
it is a A-redex, that is, s x = s {i4p (A x.t,t')}, s[ = s { i [ t ' / x ] } , and
J (Sr", / ,r) appears in t or t'. In the former case, clearly we can write
s[ as s i { J (S r" [ t '/x ], r' [t '/x ], r [t'/x ])} , supposing that appropriate
variances have been chosen. Then we can simply let s'2 = s i {r [ t ' / x l ) .
si = s i is a selection axiom and (^2,-si) is a one-step reduction that
corresponds to (sx ,si). In the latter case, J ( S r " , r ' , r ) appears in t'.
Then si takes the form

s i { J ( 5 r w, r ' , r ) , . . . , J ( 5 r " , r /,r)>,

where the occurrences of J (Sr" ,r ' ,r ) may be multiplied because of


the substitutions. Clearly, a corresponding reduction reduces s2 to

s i{ r , ...,r } and this can also be obtained from si by several replace­


ments of J ( S r " , r l, r ) with r. Hence the sequence r i,...,r n is easily
obtained. The case when the reduced redex is a J-redex is similar.
(2) By an induction on the length of a. |

Now suppose that s i = Sj is a selection axiom and a = c r (s i) is


the sequence of strict reductions from s 2 to its normal form sj. By
the lemma we can find a sequence o f reductions from s3 to some term
s2 and s i is obtained from sj by some replacements as in a selection

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 408

axiom. So we have a proof of aj = s '2 using only the selection axioms


and the identity axioms. It is easy to verify that if s { J (S i,r ',r )} is
normal, a {r } must also be. So we see that the formulas in the proof
of a* = aj axe all normal and also s 2' = a£. That is, we can construct a
proof of (ai = a2)* consisting of normal formulas only. We can have a
similar construction for the A-axioms, for we can prove a lemma similar
to Lemma 3.2. So we have

LEMMA 3 .3 . I f tp is a selection axiom or X-axiom, then <p* has


a proof in P R O consisting o f normal formulas only and the proof is
constructed by a prim itive recursive function o f <p.

Now we give some details about eliminating Re and J-terms. The


idea mentioned above should be straightforward, though the detailed
constructions are a bit lengthy.
Recall that for each definition of a primitive recursive function,
there is a constant symbol in the language of PRA. For convenience,
we will fix some notations for these symbols. We will let Zn be a sym­
bol for (a definition of) an n-place zero function, t" be a symbol for
a projection function, and j be a symbol for the 3-place function h
defined by j (n, m, I) = m if n = 0, and I otherwise. Given constant

symbols /• * ,/” , . . . , / £ , let C om p (.ft*,/™, . . . , / £ ) be the symbol cor­


responding to the definition by composition from /J1, / ” , —1/ £ , and
similarly let Re (/•*, /J*+2) be the symbol corresponding to the defini­
tion by recursion. Consider normal terms of type o with no higher
type free variables. Recall that any subterm of such a term must either
be of the type o or be one of the constants 5 , /" ,.... For each such
term £, and each sequence of n numerical variables x = ( i i , ..., xn) that
contains all free variables of t and is arranged in the ascending order of

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
3. CONSERVATIVENESS OF PRO OVER PRA 409

the indices, we assign a constant symbol f t<x in the language of P R A


as follows:
(i) /o lX is Zn, / IiiX is x f;
(ii) if t is A p ( f , t u ...,ti) then f t,x is C o m p ( /,/tliX, . . . , / t(iX);

(iii) if t is J { t o , t i , t 2) then f t<x is Co mp { j , f tuX, f ti,x)\


(iv) if t is R ei j (a, r, t-i let y, z be the two numerical variables
following the last one in x in the ascending order of the indices, and
let g be Re ( / r,x, /{l[v,i],x„*); then f t,x is Comp(g, / , lX,x ? ,..., *£).
Let x, u be two sequences of numerical variables of lengths n, m

respectively, each arranged in the ascending order of the indices. Let


t be an order-preserving mapping from members of x to those of u.
(Hence we must have m > n.) r (x) denotes the sequence of images of
r. And for a sequence of terms s = ( s i , ..., s„), let r* (s) be the sequence
with length m and with Si at the position of r ( x i ) in u, i = 1, ...,n ,
and with any new variable z at other positions. Then we have the
following lemma, which basically says that superfluous variables in x
do not really affect f t<x.

LEMMA 3.4. If all free variables o f t are in x , and r, u are as


above, then

P R A 1“ f t)X (s) = /t[r(x)/x],u (t* (s)) .

Further, the proof in P R A is a prim itive recursive function o f the


given terms and variables.

PROOF. This is proved by induction on the construction of t. We

discuss the case t — Re i j (s ,r , tx [i, j ]) only. Denote t' = t [ r ( x ) / x ] ,


t[ = [t (x ) /x ], r' = r [r (x) /x], s ' = s [ r (x) /x]. Let g be Re ( / r,x , /t l[v,I],xvx) ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 410

and g' be Re (/r'|U, /*»[«,wj.uvw)* Then, in P R A , we can prove

/t.x (s) = s ( / . , * ( s ) , s ) ,

ft'.u (T* (s)) = g' (/,',ii (t* (s)) , r* ( s ) ) .

By the inductive assumptions, P R A h / <iX(s) = /,/,u (r* (s)), so it suf­


fices to prove g (f t>x ( s ) , s) = g' ( / f)X ( s ) , r* (s)) in P R A . By induction
again, it suffices to prove

ff(0,s) = g' ( 0, T9 (s )),

g (x , s) = g' (x, r* (s)) -► g ( S x , s) = g' ( S x , r* ( s ) ) .

The former follows from the induction hypothesis about r. As for the
latter, notice that

<7 (S x ,s ) = /t l[y*],xy* (s, x , g (x, s ) ) ,

g ' ( S x , T * ( s ) ) = / t-(Hiuwu(r* ( s ) , x , g ' ( x , T 9 (s ))).

Then it is clear that it also follows from the induction hypothesis about

h-

Finally, a careful examination shows that the proof of the equation


in P R A is obtained by a primitive recursion on the construction of the
term t. I

C o r o l l a r y 3 .5 . I f x c y , then P R A \- / e.x ( x ) = f t,y (y ).

For any normal term t of the type o whose free variables are all
numerical and are exactly x = ( x i , ..., xn) in the ascending order (take
n = l,x x = xj in case t has no free variables), let be the term
ft,x (x). So ri contains no Re or J, and is in the language of P R A , and is
supposed to be equal to t. Define tp* by (t = s)* = (t* = s* ), (cp V =

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
3. CONSERVATIVENESS OF P R O OVER P R A 411

(tpi V i j ) ( t p A ip y = (tp^ A ^ *), and —►VO* = (v**- ^V’*)- Then


we can prove

LEMMA 3 .6 . Suppose that t, r, <p are normal and all their free vari­

ables are numerical.


(1) PRAY- i[r/x]* = /x ;
(2) If p is an axiom of P R O , then PRAY- p*;
(S) If PRAY- <p[0]*, and PRAY- p [n]t —» p [ S n \ \ then PRAY-
p[tf.
Farther, in (1) and (2) the proofs are primitive recursive functions
of t , r and p respectively, and in (S) the proof is a prim itive recursive

function o f the supposed proofs in the assumptions.

PROOF. (1) By some routine inductions. We discuss the case t =

R et; ( s , s i , t i [t,;]) only. Denote t' = t [ r / x ] , t[ = ti[r /x ], 3[ =


s i[r /x ], s' = s \ r j x \ . Let y be all the free variables in t [ r / x ] . We
can assume that x does occur in t but does not occur in r. So y x
contains all free variables in t. (t')* is f t>,y (y). By definition, PRAh

/*'.y ( y ) = $ '(/•',y ( y ) . y ) > where g' is Re ( f t>


uy,ft[,yuv) . Similarly, by
the corollary above, PRA I- = f t<yx (y, x ) , and by definition, PRAh
ft,yx (y , x) = 9 (/.,yx ( y , x ) , y , x ) , where g is R e ( /f,y i, ft#**,) • By the
induction hypothesis, /,<iy (y ) = f t<yx (y , , so it remains to prove

5/ (n i y ) = 9 (n, y , r f) . B y induction and the definitions of g,g', this is


reduced to

f*[,y ( y ) = / . . y * ( y . ^ ) .

9 ' (n, y ) = g (n, y , r f) -► g' (S n, y ) = g (S n , y , r f) .

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 412

The former is the induction hypothesis about s. As for the latter,


notice that

g' ( S n , y ) = / t. tjru„ ( y , n, g' (n , y )),

9 ( S n , y , r f) = f tiYXUV ( y , r \ n , g (n .y .r * ) ) .

Then we see that the latter also follows from the induction hypothesis.
(2) Propositioned axioms are simple. The case for the identity ax­
ioms follows from Corollary 3.5. The arithmetic axioms and recursion
axioms are guaranteed by the definitions of the translation <p*. No­
tice that (p cannot be a reduction axiom or a A-axiom as it is nor­
mal. So it remains to consider the selection axioms. We consider
s{J ( 0 , 3 i , 3 2 ) } = > * { s i } only. The other case is similar. We need an
induction on 3. Notice that . / ( 0 , 3 i , 3 2 ) must be a normal term of the
type o. So the base case is J (0 , 3i, 32) = 3i, which is simple. By (1), if
J (0, Si, s 2) is substitution-free in the context s { J (0 , 31, 32)} >then the
conclusion follows from . / ( 0 , 31, 32)* = 3J. So, for the inductive steps it
suffices to consider the case

(63)
Reij ( s , r , t [ i , j } { J ( 0 , 31, 32)})* = R eij ( s , r , t [ i , j ] {3X})f .

Let x, x 7 be all free variables in the two terms above respectively. We


can assume that i , j are not among them. Then by definition, the
following are provable in P R A :

R eij ( 3 , r, t [i, j] { J (0, s l} 32)})* = Re ( / r,x , / tiX„ ) ( / , iX( x ) , x ) ,

R eij ( 3 , r, t [i,j] {3X} )f = R e ( /r,x., / t<tX<o) ( x ' ) , x ;) ,

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 413

where t' = t [i,j] {sx} is obtained from t by the replacement. By Corol­


lary 3.5, f SiX (x) = / , iX' ( x ' ) . So it suffices to prove

Re (/r,x> f t , xi j ) ( n ) x ) = Re (f r , x ‘J f t' .x' ij ) (n i x ) •

We use the induction rule of P R A . The base case n = 0 is trivial. We


have

Re(/r,X) f t , x i j ) {Sn, x) = f t ,x i j (x, 71, Re(/r,x, f t . x i j ) ( n , x )) ,

Re (/r,x'j f t ' , x ' i j ) {Sn, x ) = f t ' , x ' i j (x i n i Re (/r,x'> f t ' , x ' i j ) (^i x ))

as axioms of P R A and

(64) PRA h (x , n, p ) = f t>iX>ij (x', n, p)

by the inductive hypothesis about t and Corollary 3.5, where p =

Re (/r,xi f t , x i j ) { n, x ) . Then it is clear that (63) is provable in P R A .


Further, a careful examination will show that a proof of <pt can be
obtained by a primitive recursive function of <p. For example, in the
case just discussed, the inductive assumption plus Corollary 3.5 gives
a proof of fttxij ( x , i , i ) = /t'.x'.i ( x ' , z , ; ) , which is = (t')f . Simply
substituting z, j by n ,p we will obtain a proof of (64) and then obtain

a proof of (p* in that case. This is a primitive recursion.


(3) follows from (1) easily. |

Finally we have

THEOREM 3 .7 . P R O is conservative over P R A . More specifi­


cally, there is a prim itive recursive function that transforms a proof
of a formula of P R A in P R O to a proof in P R A .

PROOF. Consider a proof of a formula <p of P R A in P R O . First

instantiate all free variables of higher type in the proof by closed terms

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
3. CONSERVATIVENESS OF PRO OVER PRA 414

of appropriate types. It is easy to check that all axioms and applications


of rules are preserved. So we obtain a proof of <p with no free higher
type variables. Then, by Lemma 3.1, we can construct a proof of
<p consisting of normal formulas with no higher type variables. The
conclusion finally follows from Lemma 3.6. |

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
Bibliography
[1] 0 . Aberth, Computable Analysis, New York: McGraw-Hill 1980.

[2] J. Avigad and S. Feferman, ‘Godel’s functional ( “dialetica”) interpretation’,


in S. R. Buss (ed.), Handbook of Proof Theory, Elsevier Science B.V., 1998,
pp. 337-405.

[3] J. Assouni, ‘Applied mathematics, existential commitment, and the Quine-


Putnam indispensability thesis’, Philosophia Mathemaiica (3), 5(1997), pp.
193-209.

[4] J. C. Baes, ‘Recursivity in quantum mechanics’, Trans. Amer. Math. Soc.,


280(1983), pp. 339-350.

[5] M. Balaguer, ‘A fictionalist account of the indispensable application of math­


ematics’, Philosophical Studies, 83(1997), pp. 291-314.

[6] H. P. Barendregt, The Lambda Calculus, Its Syntax and Semantics, North-
Holland, Amsterdam, 1981.

[7] S. Barker, ‘Kant’s view of geometry: a partial defense’, in [75] pp. 221-244.

[8] M. J. Beeson, Foundations of Constructive Mathematics, Springer-Verlag,


1985.

[9] P. Benacerraf, ‘What numbers could not be’, Philosophical Review, 74(1965),
47-73. Reprinted in [11].

[10] P. Benacerraf, ‘Mathematical truth’, Journal of Philosophy, 70(1973), 661-


679. Reprinted in [11].

[11] P. Benacerraf and H. Putnam, Philosophy of Mathematics: Selected Read­


ings, 2nd ed., Cambridge University Press, Cambridge, 1983.

[12] H. Billinge, ‘A constructive formulation of Gleason’s theorem’, J. Phil. Logic,


26(1997) 661-670.

[13] E. Bishop, foundations of Constructive Analysis, McGraw-Hill, New York,


1967.

415

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout p erm ission .
BIBLIOGRAPHY 416

[14] E. Bishop, 'Mathematics as a numerical language’, in A. Kino, J. Myhill, and


R.E. Vesley (eds.), Intuitionism and Proof Theory, pp.53-71, North-Holland,
Amsterdam, 1970.

[15] E. Bishop and D. S. Bridges, Constructive Analysis, Springer-Verlag, 1985.

[16] J. Blank, P. Exner, and M. Havlicek, Hilbert Space Operators in Quantum


Mechanics, American Institute of Physics, 1994.

[17] D. S. Bridges, ‘Towards a constructive foundation for quantum mechanics’, in


F. Richman, ed., Constructive Mathematics, Lecture Notes in Mathematics,
No. 873, pp 260-73, Springer, 1981.

[18] D. S. Bridges, 'Recent progress in constructive approximation theory’, in The


L. E. J. Brouwer Centenary Symposium, pp. 41-50, North-Holland 1982.

[19] D. S. Bridges, ‘A general constructive intermediate value theorem’, Zeitschr.


Math. Logik Grundlagen Math. 35(1989), 433-435.

[20] D. S. Bridges, ‘Sequential, pointwise, and uniform continuity: a constructive


note’, Math. Log. Quart. 39(1993), 55-61.

[21] 0 . S. Bridges, 'Constructive notions of strict convexity’, Math. Log. Quart.


39(1993), 295-300.

[22] D. S. Bridges, ‘The constructive theory of preference relations on a locally


compact space II’, Math. Social Science 27(1994), 1-9.

[23] D. S. Bridges, 'Constructive mathematics and unbounded operators — a re­


ply to Heilman’, J. Phil. Logic, 24(1995), 549-561.

[24] D. S. Bridges and O. Demuth, ‘On the Lebesgue measurability of continuous


functions in constructive analysis’, Bull. Amer. Math. Soc. (NS) 24(1991),
no.2, 259-.

[25] D. S. Bridges and H. Ishihara, 'Linear mappings are fairly well-behaved’,


Aech. Math. 54(1990), 558-562.

[26] D. S. Bridges and H. Ishihara, ‘Locating the range of an operator on a Hilbert


space’, Bull. London Math. Soc. 24(1992), 599-605.

[27] D. S. Bridges and H. Ishihara, ‘Complements of intersections in constructive


mathematics’, Math. Log. Quart. 40(1994), 35-43.

[28] D. S. Bridges and H. Ishihara, ‘Absolute continuity and the uniqueness of the
constructive functional calculus’, Math. Log. Quart. 40(1994), 519-527.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
BIBLIOGRAPHY 417

[29] D. S. Bridges, W. Julian, and R. Mines, ‘A constructive treatment of open


and unopen mapping theorems’, Zeitachr. Math. Logik Grundlagen Math.
35(1993), 29-43.

[30] D. Bridges and F. Richman, Varieties of Constructive Mathematics, Lond.


Math. Soc. Lecture Notes, no.97, Cambridge University Press, 1987.

[31] D. S. Bridges, F. Richman, W. H. Julian, and R. Mines, ‘Extensions and


fixed points of constructive maps in R.n’, J. Math. Analys. Appl. 165(1992),
438-456.

[32] D. K. Brown and S. G. Simpson, ‘Which set existence axioms are needed
' to prove the separable Hahn-Banach theorem?’, Annals of Pure and Applied
Logic, 31(1986), pp. 123-144.

[33] D. K. Brown and S. G. Simpson, ‘The Baire category theorem in weak subsys­
tems of second-order arithmetic’, The Journal of Symbolic Logic, 58(1993),
pp. 557-578.

[34] J. P. Burgess, ‘Epistemology and nominalism’, in A. D. Itvine (ed.), Physi-


calism in Mathematics, Dordrecht: Kluwer Academic Publishers, pp. 1-15.

[35] J. P. Burgess, ‘How foundational work in mathematics can be relevant to


philosophy of science’, PSA 1992, Volume Two, p.433-441.

[36] J. P. Burgess and G. Rosen, A Subject with No Object, Oxford University


Press, Oxford, 1998.

[37] C. Chihara, ‘Tharp’s "Myth and mathematics’” , Synthese 81(1989), pp. 153-
165.

[38] C. Chihara, Constructibility and Mathematical Existence, Clarendon Press,


Oxford, 1990.

[39] J. Conway, A Course in Functional Analysis, Springer-Verlag,

[40] H. B. Curry, ‘Remarks on the definition and nature of mathematics’, in [11].

[41] S. Feferman, ‘Constructive theories of functions and classes’. In: Logic Col­
loquium *78, pp 159-224, Amsterdam, North-Holland, 1978.

[42] S. Feferman, ‘Weyl vindicated: “Das Kontinuum” 70 years later’, in Termi


e prospettive della logica e della filosofia della scicnza contemporanee, vol. I,
pp.59-93, CLUEB, Bologna, 1988.

[43] S. Feferman, ‘Why a little bit goes a long way: logical foundations of scien­
tifically applicable mathematics’, PSA 1992, Volume 2, pp. 442-445.

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
BIBLIOGRAPHY 418

[44] H. Field, Science without Numbers, Princeton University Press, Princeton,


1980.

[45] H. Field, Realism, Mathematics and Modality, Oxford: Basil Blackwell. 1989.

[46] H. Friedman, ‘Set theoretic foundations for constructive analysis’, Ann. of


Math. (2) 105 (1977), 1-28.

[47] R. 0 . Gandy, ‘An early proof of normalisation by A. M. Turing’, in: J.


R. Hindley and J. P. Seldin (eds.), To H. B. Curry: Essays on Combinatory
Logic, Lambda Calculus and Formalism, Academic Press, New York and Lon­
don, 1980.

[48] K. Godel, 'On a hitherto unutilized extension of the Unitary standpoint’, in


[49].
[49] K. Godel, Collected Works, Volume II, edited by S. Feferman, et aI, Oxford
University Press, 1990.

[50] N. Goodman and J. My hill, ‘The formalization of Bishop’s constructive math­


ematics’. In: Toposes, Algebraic Geometry, and Logic, pp 83-96, Lecture Notes
in Mathematics 274, Springer-Verlag, 1972.

[51] H. B. ‘Remarks on the definition and nature of mathematics’, in [11].

[52] G. Heilman, Mathematics without Numbers, Oxford University Press, Oxford,


1989.

[53] G. Heilman, 'On the scope and force of indispensability arguments’, PSA
1992, Volume Two, p.456-464.

[54] G. Heilman, ‘Gleason’s Theorem is not constructively provable’, J. Phil.


Logic, 22(1993), 193-203.

[55] G. Heilman, ‘Constructive mathematics and quantum mechanics: Unbounded


operators and the Spectral Theorem’, J. P h i. Logic, 22(1993), 221-248.

[56] G. Heilman, ‘Quantum mechanical unbounded operators and constructive


mathematics a rejoinder to Bridges’, J. Phil. Logic, 26(1997), 121-127.

[57] D. Hilbert, ‘On the infinite’, reprinted in [11], 183-201.

[58] E. Hille, Lectures on Ordinary Differential Equations, Addison-Wesley, 1969.

[59] H. Ishihara, ‘Constructive compact linear mappings’, Bull. London. Math.


Soc. 21(1989), 577-584.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
BIBLIOGRAPHY 419

[60] H. Ishihara, ‘An omniscience principle, the Konig lemma and the Hahn-
Banach theorem’, ZtiUehr. Math. Logik Grundlagen Math. 36(1990), 237-240.

[61] H. Ishihara, ‘Constructive compact operators on a Hilbert space’, Ann. Pure


Appl. Logie 52(1991), 31-37.

[62] H. Ishihara, ‘Continuity and nondiscontinuity in constructive mathematics’,


J. Symbolic Logic 56(1991), no. 4, 1349-1354.

[63] H. Ishihara, ‘Continuity properties in constructive mathematics’, J. Symbolic


Logic 57(1992), no. 2, 557-565.

[64] H. Ishihara, ‘Constructive existence of Minkowski functionals’, Proc. Amer.


Math. Soc. 116(1992), no. 1, 79-84.

[65] S. C. Kleene, ‘Recursive functionals and quantifiers of finite types, I’, Trans­
actions of American Mathematics Society, 91(1959), pp. 1-52.

[66] P. Maddy, Naturalism in Mathematics, Clarendon Press, Oxford, 1997.


[67] J. S. Mill, A System of Logic, Books I-III, Collected Works of John Stuart
Mill, Volume VII, University of Toronto Press and Routledge Sc Kegan Paul,
1973.

[68] J. Mycielski, ‘Analysis without actual infinity’, The Journal of Symbolic Logic,
46(1981), pp. 625-633.

[69] J. Mycielski, 'The meaning of pure mathematics’, Journal of Philosophical


Logic, 18(1989), pp. 315-320.

[70] J. Myhill, ‘Constructive set theory’, J. Symbolic Logic 40(1975), no. 3, 347-
382.

[71] D. Papineau, ‘Mathematical fictionalism’, International Studies in the Phi­


losophy of Science, 2(1988), pp. 151-174.

[72] C. Parsons, 'On a number-theoretic choice scheme and its relation to induc­
tion’, in A. Kino, J. Myhill, and R.E. Vesley (eds.), Intuitionism and Proof
Theory, pp. 459-473, North-Holland, Amsterdam, 1970.

[73] C. Parsons, ‘On n-quantifier induction’, Journal of Symbolic Logic, 37(1972),


pp. 466-482.

[74] C. Parsons, ‘Kant’s philosophy of arithmetic’, reprinted in [75], pp43-8Q.

[75] C. Posy (ed.), Kant’s Philosophy of Mathematics, Kluwer Academic Publish­


ers, 1992.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
BIBLIOGRAPHY 420

[76] M. B. Pour-El and I. Richards, ‘A computable ordinary differential equa­


tion which possesses no computable solution’, Annals of Mathematical Logic,
17(1979), pp.61-90.

[77] M. B. Pour-Bl and I. Richards, 'The wave equation with computable initial
data such that its unique solution is not computable’, Advances in Mathe­
matics, 39(1981), pp. 215-239.

[78] M. B. Pour-El and I. Richards, Computability in Analysis and Physics,


Springer-Verlag, Berlin-Heidelberg-New York, 1989.

[79] E. Prugovecki, Quantum Mechanics in Hilbert Space, 2nd ed., Academic


- Press, 1981.

[80] H. Putnam, ‘Is logic empirical?’, Boston Studies in the Philosophy of Science,
5(1969), pp. 216-241.

[81] H. Putnam, 'Philosophy of logic’, in H. Putnam, Mathematics, Matter and


Method: Philosophical Papers, vol. i, 2nd ed., Cambridge University Press,
Cambridge, 1979.

[82] W. V. Quine, Ontological Relativity and Other Essays, Columbia university


Press, New York and London, 1969.

[83] W. V. Quine, Per suit of Truth (Revised Edition), Harvard University Press,
1992.

[84] W. V. Quine, From Stimulus to Science, Harvard University Press, 1995.

[85] M. Reed and B. Simon, Methods of Modem Mathematical Physics, II: Fourier
analysis, Self-adjointness, Academic Press Inc., 1975.

[86] A. Renyi, Dialogues on Mathematics, Holden-Day, San Fransisco, 1967.

[87] M. Resnik, Philosophia Mathematica (3), 3(1995), pp. 166-74.

[88] M. Resnik, ‘Structural relativity’, Philosophia Mathematica (3), 4(1996), pp.


83-99.

[89] F. Richman, 'Transcendental operators on a Banach space’, Rocky Mountain


J. Math. 22(1992), no. 2, 697-704.

[90] F. Riess and B. Ss.-Nagy, Functional Analysts, Frederick Ungar Publishing


Co., 1955.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.
BIBLIOGRAPHY 421

[91] A. Robinson, ‘Formalism 64’, Logic, Methodology and Philosophy of Science,


Proceedings of the 1964 Congress in Jerusalem, (Y. Bar Hille, Editor), North-
Holland, Amsterdam, 1965, pp. 228-246.

[92] G. Rosen, ‘Modal fictionalism’, Mind, 99(1990) pp. 327-354.

[93] G. Rosen, ‘Modal fictionalism fixed’, Analysis 55(1995), pp. 67-73

[94] S. G. Simpson, ‘Which set existence axioms are needed to prove the
Cauchy/Peano theorem for ordinary differential equations?’, The Joulnal of
Symbolic Logic, 49(1984), pp783-802.

[95f S. G. Simpson, ‘Partial realizations of Hilbert’s program’, The Journal of


Symbolic Logic, 53(1988), pp. 347-363.

[96] E. Sober, ‘Mathematics and indispensability’, Philosophical Review,


102(1993), pp. 35-57.

[97] C. Spector, ‘Provably recursive functionals of analysis’, Proc. Symp. on Pure


Mathematics, Amer. Math. Soc., Providence 5 (1962), 1-27.

[98] W. Tait, ‘Finitism’, Journal of Philosophy, 78 (1981), pp. 524-546.

[99] L. Tharp, ‘Myth and Mathematics: a conceptualistic philosophy of mathe­


matics I', Synthese 81(1989), pp. 167-201.

[100] L. Tharp, ‘Myth k math, part II (priliminary draft)’, Synthese, 88(1991),


pp. 179-199.

[101] A. S. Troelstra, Meiamathematical Investigation of Intuitionistic Arithmetic


and Analysis, Lecture Notes in Mathematics, no. 344, Springer-Verlag, Berlin,
1973.

[102] A. S. Troelstra, ‘Introductory note to 1958 and 1972’, in [49].

[103] A. S. Troelstra and D. van Dalen, Constructivism in Mathematics: An Intro­


duction, vol. 1, North-Holland, Amsterdam, 1988.

[104] J. Wieidmann, Linear Operators in Hilbert Space, Springer-Verlag, 1980.

[105] E. Wigner, ‘The unreasonable effectiveness of mathematics in the natural


sciences’, reprinted in Symmetries and reflections, Indiana University Press,
Bloomington, pp. 222-237.

[106] J. M. Young, ‘Construction, schematism and imagination’, in ?? pp. 159-176.


[107] X . Yu, ‘Radon-Nilcodym theorem is equivalent to arithmetical comprehen­
sion’, in W. Sieg (ed.), Logic and Computation, Contemporary Mathematics,

R ep ro d u ced with p erm ission o f th e copyright ow ner. Further reproduction prohibited w ithout perm ission.
BIBLIOGRAPHY 422

106, American Mathematical Society, 1990, pp. 289-297.

[108] X. Yu, ‘Lebesgue Convergence Theorem and Reverse Mathematics’, Math.


Log. Quart. 40(1994), pp. 1-13.

R ep ro d u ced with p erm ission o f the copyright ow ner. Further reproduction prohibited w ithout p erm ission.

You might also like