MATHEMATICS
MADE
DIFFICULT

A Handbook
for the
Perplexed

Carl E.
Linderholm


There was a time when every bookshop
displayed a selection of soporific tracts
expounding the principle that mathe-
matics could be made easy. All that is
now past, and the idea that mathematics
really is difficult has regained its
freedom. Dr. Linderholm's pioneering opus
Mathematics Made Difficult is the hand-
book of the new liberation. As you read
the book the ability to count, let us say,
begins to haze out. You are gradually
coaxed deeper into the eerie landscape
of central mathematics. Forgetful func-
tors stare back at you between wild
trees; arrows fly in all directions; your
feet become entangled in an under-
growth of morphisms; your eyes behold
the universally repelling object. Not since
Alice in Wonderland has mathematical
wizardry been used to such lunatic effect.

Dr. Carl Linderholm was born in Baton
Rouge, Louisiana. He studied at the Uni-
versity of Chicago, is married, and has
four children. Before lecturing at Roose-
velt University and at the universities of
Michigan and Illinois, he was briefly a
waiter but was fired because he could
not remember on which side of the plate
to put the knife, fork, and spoon. His
mathematical interests include abstract
languages and ergodic theory.

He is nearsighted in the left eye but not
in the right and astigmatic in the right
eye but not in the left. He can wiggle
his ears only simultaneously and can
cross the right (left) foretoe over the
right (left) great toe and vice versa, no
other ordered pairs of toes on the same
foot besides these four being crossable.
A small patch of dark skin on his right
side contains hundreds of moles. During
lectures, he removes and replaces his
jacket frequently.

As to likes and dislikes, his favorite wild
insect is the tarantula; favorite tame in-
sect, the cheese mite; favorite instru-
ment, the Acadian accordion; favorite
book, The Anatomy of Melancholy; most
disliked tame plant, privet; most disliked
wild plant, scrub oak; most disliked part
of speech, the adverb; most disliked
country, Mordor.

Besides children and cheese mites, Dr.
Linderholm's house contains a red Per-
sian cat. Guinea pigs live in the garden.


Jacket design by Seymour Schlattner/
Milton Charles

Published by the World Publishing Company
Published simultaneously in Canada
by Nelson, Foster & Scott Ltd.
First American Edition
First printing--1972
Copyright (C) 1971 by Carl E. Linderholm
All rights reserved
ISBN 0-529-04552-4
Library of Congress catalog card number: 76-183082
Printed in the United States of America

World Publishing
Times Mirror


To
CLEMENT V. DURELL, M.A,
WITHOUT WHOM THIS BOOK
WOULD NOT HAVE BEEN NECESSARY
(With apologies to St. Thomas Aquinas)


CONTENTS

INTRODUCTION
1. The structure of the book 10
2. How to read this book 11

1 ARITHMETIC
1. So you think you can count? 13
2. Add astra per aspirin 23
3. Subtraction 36
4. The negative-positive diathesis 43
5. Multiplication 49
6. Division 54
7. Casting out nines 66

2 FACTORS AND FRACTIONS
1. Prime numbers 71
2. Finding prime factors of a number 78
3. The greatest common divisor 80
4. Guess the next number 89
5. Fractions 99
6. Calculations with fractions 108

3 ALGEBRA
1. The wonderful quadratic formula 117
2. The incommensurable of the incommensurable 126
3. The inconstructible of the inconstructible 129
4. Pain de maison 137
5. The Rule of Three 143
6. Polynomials 145
7. What are brackets? 155

4 TOPOLOGY AND GEOMETRY
1. Into the interior 163
2. The insides 171
3. Geometry 186

EXERCISES 199

BIBLIOGRAPHY 205


INTRODUCTION


One of the great Zen masters had an eager disciple who never lost an
opportunity to catch whatever pearls of wisdom might drop from
the master's lips, and who followed him about constantly. One day,
deferentially opening an iron gate for the old man, the disciple asked,
'How may I attain enlightenment?' The ancient sage, though
withered and feeble, could be quick, and he deftly caused the heavy
gate to shut on the pupil's leg, breaking it.

When the reader has understood this little story, then he will
understand the purpose of this book. It would seem to the unen-
lightened as though the master, far from teaching his disciple, had left
him more perplexed than ever by his cruel trick. To the enlightened,
the anecdote expresses a deep truth. It is impossible to spell out for
the reader what the truth is; he can only be referred to the anecdote.

Simplicity is relative. To the great majority of mankind--
mathematical ignoramuses--it is a simple fact, for instance, that
17 X 17 = 289, and a complicated one that in a principal ideal ring
a finite subset of a set E suffices to generate the ideal generated by E.
For the reader and for others among a select few, the reverse is the
case. One needs to be reminded of this fact especially as it applies to
mathematics. Thus, the title of this book might equally well have
been Mathematics Made Simple; whereas most books with that title

9

might equally well have been called Mathematics Made Complicated.
The simplicity or difficulty depends on who is reading the book.

There is no doubt that an absolute ignoramus (not a mere qualified
ignoramus, like the author) may become slightly confused on reading
this book. Is this bad? On the contrary, it is highly desirable.
Mathematicians always strive to confuse their audiences; where there
is no confusion there is no prestige. Mathematics is prestidigitation.
Confusion itself may be taken as the guiding principle in what is done
here--if there is a principle. Just as the fractured leg confused the
Zen disciple, it is hoped that this book may help to confuse some
uninitiated reader and so put him on the road to enlightenment,
limping along to mathematical satori. If confusion is the first prin-
ciple here, beside it and ancillary to it is a second: pain. For too long,
educators have followed blindly the pleasure principle. This over-
simplified approach is rejected here. Pleasure, we take it, is for the
initiated; for the ignoramus, if not precisely pain, then at least a kind
of generalized Schmerz.


1. The structure of the book


The following account of the structure of the book makes no pretence
to be complete; nor is it in every respect completely accurate. It is
a naive account, along taxonomic lines, which it is hoped will be of
use to the reader. The book is divided into chapters, which are divided
into sections, which are divided into paragraphs, which are divided
into sentences, which are divided into words, which are divided into
letters. For reference purposes, each letter in the book has a number
(actually a finite sequence of positive integers) attached to it. The
finite sequence has exactly six (the first perfect number) terms, so
that the book, considered as a finite sequence of letters, is mapped
to N<6>[0]. It has been found very convenient both for author and
reader to do this abstractly, rather than concretely.

Walking home along a deserted street late at night, the reader may
imagine himself to feel in the small of his back a cold, hard object;
and to hear the words spoken behind him, 'Easy now. This is a stick-
up. Hand over your money.' What does the reader do? He attempts

10

to generate the utterance. He says to himself, now if I were standing
behind someone holding a cold, hard object against his back, what
would make me say that? What would I mean by it? The reader is
advised that he can only arrive at the deep structure of this book, and
through the deep structure arrive at the semantics, if he attempts to
generate the book for himself. The author wishes him luck.


2. How to read this book


Reading books, like mathematics, is an art to which many pretend,
to which some aspire, and to which few attain. Nothing outside the
realm of mathematics is more beautiful to behold than a beautifully
read book. The art of reading books beautifully, like mathematics,
cannot be imparted in a few easy lessons. It is difficult to give hints
without incurring the danger of becoming fatally lucid. The following
advice, it is hoped, avoids this deadly lucidity as effectively as it is
avoided in the book itself. Here as in so many things, it is really the
man who is totally at sea who has got both feet firmly on the ground.

It is the author's general impression that, at first reading at least,
it is unwise to read the book on too many levels at once. For most
readers it will in any case be impossible to penetrate beyond the
lowest level or two; these should be quite satisfactory for a beginning.

Consider a category (small, if you wish) in which there is at most
one arrow between two objects, and in which every isomorphism is
an identity. Products and coproducts exist; moreover product dis-
tributes over coproduct in the sense that the dotted arrow exists:

(A (pi) B) (mu) (A (pi) C)
/|\ /:\ /|\
| :... |
| : |
A (pi) B --> A <-------- A (pi) C
| /|\ : |
\|/ | : \|/
B A (pi) (B (mu) C) C
| | |
| \|/ |
\---> B (mu) C <---------/

11

Further, assume initial and final objects 0, 1 and a contravariant
functor from the category to itself .' that is alternating on objects and
preserves no arrow except 0 -> 1. Then it is possible to think, be-
cause the category is a boolean algebra. Certain obvious refinements
will bring the reader face-to-face with logic. A little logic of this
naive sort may occasionally be found helpful in reading the book.

It is finally suggested that the reader begin reading at some definite
point; say the first chapter.

















12

1

ARITHMETIC


1. So you think you can count?


QUESTION 1. Whether anybody really counts?

Objection 1. In order really to count, the counter must know the
names of the numbers. But no-one really knows the true names of the
numbers, since Englishmen call them one thing, Chinamen another;
therefore, no-one really counts.

Reply to Objection 1. It is not necessary to know the real names of
the numbers in order to count. You may call the numbers anything
you like: moreover you need not use what are usually called num-
bers at all in counting--or to put it otherwise, almost anything can
serve as a number if you so desire. (Simply transfer the structure of
the counting system to another set by means of a bijection.) Mathe-
maticians never claim to know about the elements of a system,
only about relations between them and structures imposed on them.

Objection 2. Furthermore, although in a certain language a man
might be said to count as far as a certain number, after that he must
stop. For instance, in English [1] no-one can count higher than nine
hundred and ninety-nine thousand nine hundred and ninety-nine
decillion nine hundred and ninety-nine thousand nine hundred and
ninety-nine nonillion ... nine hundred and ninety-nine thousand
nine hundred and ninety-nine. In Melanesian Pidgin English many
speakers cannot count higher than one-fellow man. But in counting,
if the counter stops at a certain point he is said not really to count.

13

Reply to Objection 2. It is true that most languages have only a few
names for numbers. Whether there are any exceptions to this rule is a
question I leave to linguists. [2] The remark about Melanesian Pidgin
may be misleading, since [3] one-fellow man means twenty, just as
two-fellow hand one-fellow foot means fifteen. As far as it concerns
the English, and most civilised, 'counting systems' the objection is a
valid one. The trouble is that the rule for naming the successor of a
number after you have named the number is not everywhere defined;
there is no successor function. Thus the English system is mathe-
matically backward, compared with certain 'primitive' systems. This
is the reason why children ask 'What is the greatest number?' but it
does not prove that no-one can count; it proves only that aged,
civilised non-mathematicians do not count. This we knew already.

Objection 3. If anybody could count, then certainly mathemati-
cians would be able to count. But mathematicians pretend to count
by means of a system supposed to satisfy the so-called Peano axioms.
In fact, there cannot be any Peano axioms, since they were really in-
vented by Dedekind. Hence even mathematicians cannot count.
Furthermore, the piano has only 88 keys; hence, anyone counting
with these axioms is soon played out.

Reply to Objection 3. Dedekind's version [4] is indeed equivalent
to Peano's [5] although it was stated differently. Dedekind's paper is
still of great historical interest, and contains many suggestive ideas.
But Peano (G. for Giuseppe, not Grand) wus an influential innovator
in ideas, and in the mathematical notation of this century; he deserves
the renown for his beautifully stated axioms. He invented a system of
stenographic shorthand, but I am unaware of any musical works. In
his system (without shorthand) 89 is written

0+++++++++++++++++++++++++++++
+++++++++++++++++++++++++++++
+++++++++++++++++++++++++++++
++.

Objection 4. Furthermore, Peano's axioms imply a supposed well-
ordering, whereby there is a least number with a given property,
provided any number possesses the property. But the least number

14

that cannot be named in fewer than twenty-one English syllables has
just been named in fewer than twenty-one English syllables; hence
no number exists that cannot be named in fewer than twenty-one
English syllables. But there are at most 10,000 such syllables, hence
at most 21^10000 numbers. But then mathematicians cannot count to
21^10000+1.

Reply to Objection 4. This logical paradox is called the Richard
paradox. [6] It is a question of foundations of mathematics, rather
than mathematics itself; or, at least, I hope so. The reply is left to the
reader as an exercise. (This phrase always means that the writer can-
not do the problem himself.)

Objection 5. Only mathematicians really count; but they are dis-
embodied minds, hence a mathematician is not any body.

Reply to Objection 5. So how did I write this book?

Objection 6. Really to count is to count realiter; that is, by means
of things (rebus). But only children enjoy rebusses, and children do
not count in a serious argument. Moreover, counting by means of
things usually means counting fingers, which are not bodies but parts
of a body. Those who count on their fingers admit they cannot count
well.

Reply to Objection 6. On the contrary, counting on the fingers is
excellent exercise in finite cardinal arithmetic. This subject is treated
in a later section; at present we are concerned with counting per se.

Objection 7. Mathematicians assert that the set of numbers is a
non-empty infinite set. But that which is infinite has no parts, since a
part is divided from the whole by a boundary (fine). But only the
empty set has no proper subsets; i.e., no parts. Hence there are no
numbers.

Reply to Objection 7. A set is infinite in mathematics not if it is
infinite unconditionally, but (according to one definition) if the set
of cardinal numbers of finite subsets is unbounded as a set of natural
numbers. This is obvious for the set of natural numbers itself since
the cardinal number of

{0, 1, ..., n - 1}

is n. According to another definition, a set is infinite if the set of
natural numbers can be injected into it. The identity function is such
an injection for the set of natural numbers itself. Another possible

15

definition requires a proper injection from the set to itself: such is

n -> n+

for N[0].


PEANO'S AXIOMS

(0) N[0] is a set;
(1) 0 (epsilon) N[0];
(2) +: N[0] -> N[0] is a function;
(3) If 0 (epsilon) E (subset) N[0] and if (E)+ (subset) E then E = N[0].

(Note that + is written on the right of its argument; (3) is the in-
duction axiom.)

(4) + is injective (a+ = b+ -> a = b);
(5) 0 (is not (epsilon)) (N[0])+.

Axioms (0), (1), (2) just say that

{0} :-> N[0] ->[+] N[0]

is a counting system. The child's system
/---\
\|/ |
*0 -> 1 -> 2 --/

previouly mentioned is inductive but not injective; the Chinese
system of naming years, isomorphic to

*0 -> 1 -> 2 -> ... -> 58 -> 59 |-\
/|\ |
\--------------------------------/

is inductive and injective, but fails (5); non-inductive systems like

(1) *Away -> away -> with -> rum
/|\ |
| \|/
gum <- by

or like

16

(2) Never -> *say -> die! |-\
/|\ |
\------------/

I -> give |-\
/|\ |
\-- up <----/

provide numbers that are never used in counting.


I ANSWER THAT some people really count, namely mathematicians,
some savages, and some two-year-olds. Anyone counts who possesses
a counting system. Mathematicians have a universal counting system
N[0] called the natural numbers. But he who counts both universally
and naturally counts really. Also, some two-year-olds count as
follows:

one, two, three, three, three, ....

But this implies the existence of a counting system
/---\
\|/ |
*one -> two -> three --/,

since if s: X -> X is defined by

X = {one, two, three}
s: one -> two
two |-------\
\|/
three -> three

then s is a function and hence

{one} :-> X ->[s] X

is a counting system (this is discussed further below). Moreover,
many so-called 'primitive' languages have such a system.


QUESTION 2. Whether it is natural to use the natural numbers?

Objection 1. They would seem to be very unnatural, since they
involve sets and functions; these are abstractions and innovations,
probably irreligious, impractical, and disrespectful to the flag. +
--------------------------------------------------------------------------
+ In Britain, the queen.

17

Reply to Objection 1. Sets and functions are more natural than
arrangements of n like objects in a line without repetitions, in-
dependent variables, and such paraphernalia, [7] all of which belong
in the Natural History Museum because they used to be natural but
now they are history. These relics are equally abstractions. Mathe-
matics is not meant to be practical. Sets and functions are intensely
pious and loyal.

Objection 2. What is natural can be easily understood by children;
but only a genius could make sense of all this stuff.

Reply to Objection 2. Children aged twelve to eighteen can easily
learn the Schro:der-Bernstein theorem and the proof that finite fields
have prime-power order. [8]

Objection 3. The Peano axioms are already seventy years old; but
it is unnatural in science to hang on to what is passe' or old hat.

Reply to Objection 3. At present the Peano axioms are still found
to be useful.

Objection 4. The Peano axioms describe the 'natural numbers'
internally, rather than as a universal object in the category of
pointed functions. This would seem unnatural, since categories are
the mathematics of the future.

Reply to Objection 4. This is partly true; the future will tell.

Objection 5. The additive group of integers is basic, and includes
the natural numbers as well as the negative numbers. Why not start
with that group, and kill two birds with one stone?

Reply to Objection 5. You do it your way, I'll do it my way.

Objection 6. Everybody does arithmetic in decimal, or arabic,
numerals. This seems a perfectly good system, and does not require
all this abstraction. You are just making mathematics difficult!

Reply to Objection 6. First of all, there are no arabic numerals
since they were invented by the Hindus. [9] Secondly, they involve
the arbitrary choice of ten as a base. Mathematics requires consider-
ing numbers in infinite sets as well as one at a time, and for this pur-
pose numerals are insufficient. A numeral is a sequence of digits;
i.e., a function

N[0] -> {0, 1, 2, ..., 9}.

What does this mean if N[0] is not already available?

18

Objection 7. Before you study Numbers you should read Genesis,
Exodus, and Leviticus. [10]

Reply to Objection 7. I did: 'Look now toward heaven, and tell
the stars, if thou be able to number them: ...' [11] Since the number
of stars is not known and the naked eye not mentioned, it is natural
to establish N[0] before beginning the project.

Objection 8. Peano's axioms are clearly nonsense: Let f be a
bijection from the supposed N[0] to a set of filing cabinets. Lock
f(0) and put the key in f(1), lock f(1) and put the key in f(2), and
continue by induction. Now all the cabinets are locked, with their
keys locked inside their neighbours. That is obviously impossible.
Hence N[0] does not exist.

Reply to Objection 8. Filing cabinets are not mathematics.

Objection 9. Induction is nonsense, since it purports to prove that
all girls are blondes, as follows: It is sufficient to show that if one of
any n girls is a blonde, then they all are blondes. This is clear if
n = 1. With n >= 1, suppose the proposition is true for n. Then it is
also true for n + 1. Suppose one of n + 1 girls is a blonde. She is
either among the first n of the girls, or among the last n--say the
first n. These, by the hypothesis about n, are all blondes. Then at
least one of the last n is a blondes, and they, too, are all blondes.
Now all the n + 1 girls are blondes.

Reply to Objection 9. This is ineptly stated, since the same thing
can be proved more easily with hydrogen peroxide. Also, it does not
work if n + 1 = 2.

Objection 10. The natural numbers bore me to tears. But it would
seem unnatural to use anything uninteresting in mathematics, since
this subject is intensely interesting through and through.

Reply to Objection 10. All numbers are interesting, since the first
uninteresting number would be interesting. Or better; no number is
interesting but the system of all natural numbers is very interesting.


I ANSWER THAT it is very natural for mathematicians to use the natural
numbers, since without them they could not do mathematics. The
natural number system is a universal pointed function and so must be
natural, since universal objects are a natural idea.

If {x} :-> X ->[s] X (abbreviated X) and {y} :-> Y ->[t] Y are

19

pointed functions a morphism X ->[f] Y is a function such
that the squares of

{x} :-> X ->[s] X
| | [f]
\|/ \|/
{y} :-> Y ->[t] Y

commute (which means that x -> y and fs = tf). If X has the
property that from X to any system Y there is exactly one morphism,
then X is universally repelling and mathematicians universally find
X attractive.


AXIOM [12]. There exists a universally repelling pointed function

{0} :-> N[0] ->[+] N[0].


QUESTION 3. Whether 2 + 2 = 4?

Objection 1. When I say to a mathematician, 'You've come a long
way since you learned that 2 + 2 = 4, he always makes a wry face.
Maybe 2 + 2 is not 4?

Reply to Objection 1. The comments people make when they find
out you are a mathematician are always painful to hear. The average
man's notion of what a microbiologist or an anthropologist does is
relatively accurate compared with his absurd fancies about mathe-
matical work. Popular notions regarding the spirit of mathematical
inquiry are definitely stuck in the Babylonian era. This is especially
the case with respect to things like '2 + 2 = 4'; for this reason, the
latter is a principal mathematical shibboleth that others will do well
to avoid mentioning in our company.

Objection 2. '2 + 2 = 4' is an obvious fact of everyday life, but
everyday life is not mathematics. Hence in mathematics it would
seem that 2 + 2 is not 4.

Reply to Objection 2. If people choose to use addition in everyday
life, we do not mind. This makes mathematics neither more nor less
valid in its own sphere.

Objection 3. Solomon, a very wise king, says 'There be three things
which are too wonderful for me, yea, four which I know not: ....'
[13] Thus 2 + 2 would appear to be 3, and not four.

20

Reply to Objection 3. This author is not writing about mathe-
matics; moreover, he explicitly disclaims mathematical knowledge:
'four, which I know not'.

Objection 4. What is 2? Define your terms!

Reply to Objection 4. The number 2 will be defined by the above
equation 1 + 1 = 2, as soon as I get around to defining 1 and +.

Objection 5. This would seem to imply that two games of chess
equal a game of bridge. But chess is mathematical in nature, whereas
bridges are an engineering problem and merely an application of
mathematics.

Reply to Objection 5. We do not yet assert that two players plus
two other players make four players; this is adding things, which
goes beyond the scope of this section.

Objection 6. Furthermore, when people put two and two together,
admittedly they usually get four, but sometimes they get other
numbers. Hence we cannot be absolutely sure that 2 + 2 = 4. But
mathematics deals in absolute certainty, so '2 + 2 = 4' is not
mathematically valid.

Reply to Objection 6. Our assertion bears absolutely no relation to
the data of practical experience. Moreover, most people cannot
count, let alone add.

Objection 7. That 2 + 2 = 4 is a consequence of the so-called
associative law, obeyed by all monoids including that of the natural
numbers. But natural numbers have no organs of communication.
Hence they cannot form associations, groups, or societies. Hence it
would seem that 2 + 2 <> 4.

Reply to Objection 7. It is the mathematicians who group the
numbers, not the numbers themselves. But monoids have the
property that it makes no difference whether we group their ele-
ments one way or another, which property we call the associative
law.

Objection 8. It would be unnatural to add numbers, since we have
already seen that mathematicians make no claim to know what those
are in themselves. If you do not know what 2 is, bow can you know
how to add it to anything?

Reply to Objection 8. It is true that it is unnatural to add numbers.
The natural thing to do is to add, or compose, functions, and the
objector will be relieved to learn that this is what we shall do. Since

21

the functions are induced by the numbers in a natural way we may
speak, par abus de langage, of adding numbers.

Objection 9. 2 + 2 can be anything you like, since we may use +
for any law of composition. For instance, if you count

/---\
\|/ |
0 -> 1 -> 2 -> 3 --/

then probably you would say 2 + 2 = 3.

Reply to Objection 9. It is true that the symbol '+', or any symbol,
may be used for any law of composition; though '+' is usually
reserved for commutative ones. Those who count with 0, 1, 2, 3
would indeed be naturally led to the commutative monoid

3 | 3 3 3 3
2 | 2 3 3 3
0 | 0 1 2 3
---+------------
+ | 0 1 2 3

whereby 2 + 2 = 3; but this counting system, and this monoid, can
be obtained from N[0] by identifying all numbers, except 0, 1, and 2
with each other; hence 3 is just another name for 4 (as Solomon, of
course, realised).


I ANSWER THAT 2 + 2 = 4. Mathematicians have established the
associative law for the binary operation +, whereby

(a + b) + c = a + (b + c)

for any natural numbers a, b, and c. But by definition

2 = 1 + 1,
3 = 2 + 1,
and 4 = 3 + 1.
Hence 2 + 2 = 2 + (1 + 1) by definition
= (2 + 1) + 1 by the associative law
= 3 + 1 by definition
= 4 by definition.

The associative law will be made clear later.

22

2. Add astra per aspirin


This section is about addition. The fact that the reader has been told
this does not necessarily mean that he knows what the section is about,
at all. He still has to know what addition is, and that he may not yet
know. It is the author's fond hope that he may not even know it after
he has read the whole section. Though addition, in general, is a
special case of multiplication, here it is thought enough to consider
only the addition of numbers, a very special case.

In the dim ages of the half-forgotten past, in Babylon and ancient
Egypt, most people seem to have proceeded to learn to do what they
thought was adding after having mastered counting; and they then
'added' with the same numbers they 'counted' with. The childhood of
the man is said to mirror the childhood of the race, and among
children this traditional course is indeed still followed by many.
Should this be so, or ought we rather to add before we can count?
We no longer dwell in caves, and the fact that counting has already
been mentioned in this book before adding need hardly prejudice the
reader to think that it comes first.

'Generations have trod, have trod, have trod;' but pity the poor
ordinary sod, the present-day beneficiary of all their traditional
treading! You have only to pronounce in his hearing the word
'addition' to conjure up before his fevered imagination a frantic
nightmare. He sees the numerals in a black whose blackness has no
bottom to its depth:

$179.63
L~26'96
L~1066-1-5
1-0-1
205-12-71
----------
TOTAL: 13250-2-9 1/2

He knows for sure the sum is wrong. But how? Where? The digits
loom larger and nearer; he falls into the blackness; at last, mercifully,
he faints. Alas! Yet, in his benighted way, he too has entered

23

ploddingly on the great quest. He is trying, unconsciously perhaps
but struggling for all that, to make mathematics difficult. His best
effort is a nasty long sum; very well: let him whose mathematical
erudition is perfect cast the first aspersion.

With a little more mathematics, it is easy to produce more difficulty
with less effort; this principle of efficiency is basic, and is discussed
further in the Appendix. + There is no need to dirty our fingers with
sums--they are beneath our dignity. The horrid difficulties of infinite
series, of such things as

1 + 1/2 + 1/4 + ...

or

0'333...

--put these aside. There is a great lot of difficulty to be got out of
much, much simpler things--for the simpler the things a man gets
difficulty out of the better his mathmanship. Do you suppose there
is much difficulty in such a thing as

12
+25
---
= 37 ?

Indeed there is! To be honest, the previous discussion of 2 + 2 = 4
was mere fools' play. Let us be fair, and start with

1
+ 1
---
= 2

A bit too hasty, perhaps. The trouble starts with that '1'. Remember
that the experts are less than sanguine about giving any real, obvious
meaning to such a thing as 1. What is the real meaning of 1? Who can
say? It is impossible to agree on any meaning; and if the deepest
thinkers could agree on some idea of 1, that would be of no real help
at all. But let it pass. What can be meant by '+'?

Since it appears to be difficult beyond all difficulty for the greatest
philosophers to say what anything is 'in itself', mathematicians
customarily come to some sort of agreement about what they will
--------------------------------------------------------------------------
+ See p. 25, footnote.

24

get together and 'believe'+ about things. Since nobody has the
slightest idea--least of all the poor mathematicians themselves--
what the things are about which the mathematicians entertain their
so-called 'beliefs', who can blame them for their harmless fancy?
Like the world of a science-fiction story, a system of beliefs need not
be highly credible--it may be as wild as you like, so long as it is not
self-contradictory--and it should lead to some interesting difficulties,
some of which should, in the end, be resolved.

a + 0 = a = 0 + a. (M1)

This is a shorthand way of saying that if you add 0 to any number
whatever, you get the same number. Since nobody knows what a
number is, it might begin to appear as if this rule had a very limited
applicability. Supposing Mao Tse-Tung to be a number, for example,
one could write the sum

Mao Tse-Tung + 0 = Mao Tse-Tung.

On the other hand, if he is not a number, it does not say if you can, or
not. May we put the beloved chairman into our sums, or not? Is it a
friendly or an unfriendly act? Will he back us up if we do it? Will he
turn the cold shoulder and apologise to our head of state for our bad
behaviour? Is he really a number, or is it only propaganda? Naturally,
the reader shall not find out from me. Partly, what is involved here
is the 'belief':

Everything, even Chairman Mao, either is or is not a number. (S1)

There are a few other little pieces of etiquette, such as never writing
+ between any two things except two numbers, and that 0 is a
number. (The reader had better get used to the idea of not knowing
whether Mao is a number. But if he is not one, the reader must not
write him into any sums. =)

Part of what has just been said is contained in writing: *

+: N[0] X N[0] -> N[0];

This also tells you that adding two numbers produces a number; not,
--------------------------------------------------------------------------
+ The expression '"believe"', as distinct from 'believe', is used in this
book for something like 'pretend to pretend to believe'. See Hypocritic
Oath, Appendix. *

= The author is unaware is he has ever written Chairman Mao into a sum.
On the other hand, if he is not a number then I haven't.

* See p. 24.

25

for instance, a suit of cards--unless the suit of cards is a number. For
example, if it is the case (as is much rumoured of late) that Mao
Tse-Tung and Spiro Agnew are numbers, and if it is also the case
that

Mao Tse-Tung + Spiro Agnew = Hearts

(which Heaven forfend), then Hearts must be a number. Neverthe-
less, being himself no searcher of Hearts, the author does not wish
to insist on this conclusion; it is stated only tentatively and subject to
several conditions.


QUESTION 4. Whether 1 is a number?

Objection 1. Numbers are either politicians or suits of cards. But
1 is neither. Therefore 1 is not a number.

Reply to Objection 1. Maybe 1 is also a politician, a young lady, or
a suit of cards. I did not say it was not. The set of politicians, young
ladies, and suits of cards cannot be properly injected into itself; the
set of numbers can; therefore some numbers are none of these things.
Armies and political parties are given to claiming that numbers are
on their side; they cannot mean all numbers. The allegiance of the
divine and mystical 1 has often been claimed on both sides of a
dispute, but proofs are usually lacking.

Objection 2. Back when you were teaching us how to count, it was
made very clear that 1 was a number. There was no doubt about it
whatsoever. This would make it seem natural that now, when we are
doing addition, it should still be a number. So most likely you will
say that 1 is not a number, just to make us look silly.

Reply to Objection 2. This is an example of 'second-guessing'. It
looks as if I am about to do one thing, and therefore it is likely that
I may do the opposite. The difficulty is that I may go ahead and do
the first thing, thus second-guessing the second-guesser. So I have,
and this is one of the attractive things about mathematics: sometimes
the answers turn out to be just what you thought they would be.
Mathematics is a tricky subject.

Objection 3. If we say, 'A number of old men were affected with
apoplexy, and took a number of valerian drops,' we always mean

26

several old men and several valerian drops, and never one. Since a
number never means one, one is not a number.

Reply to Objection 3. The fact that we do not say 'a number' when
we mean '1' is a singularity of usage.


I ANSWER THAT 1 is indeed a number. There is no special reason why;
it is just a number because I say so.

Since 1 is a number, there is no harm in forming the sum '1 + 1';
and if we choose to call the resulting number 2, no-one can argue. Of
course, it is not fair to use the name '2' for the result of this sum if
that name already has another, possibly quite different meaning.
Before writing

1 + 1 = 2

it is necessary to divest oneself of all preconceptions about 1 and
about 2. In other words, it is fair to write it only in case it is meaning-
less. Goodbye to the idea that 2 is company! Adieu to death and
taxes, for they are no longer certain to be 2 things in this life that are
certain! We must be ignorant of such fancies. What is 2, but the other
end of an arrow?

(1, 1) ->

Let us be generous with ourselves. Unstinting, profligate, prodigal,
let us write

2 + 1 = 3
3 + 1 = 4
4 + 1 = 5
5 + 1 = 6

Unflinching in our audacity, with lunatic abandon let us dare to
inscribe:

6 + 1 = 7
7 + 1 = 8
8 + 1 = 9.

27

With the famous axiom of Associativity

a + (b + c ) = (a + b) + c (M2)

one may then prove

5 + 4 = 5 + (3 + 1) = (5 + 3) + 1 = (5 + (2 + 1)) + 1 =
= ((5 + 2) + 1) + 1 = ((5 + (1 + 1)) + 1) + 1 =
= (((5 + 1) + 1) + 1) + 1 = ((6 + 1) + 1) + 1 =
= (7 + 1) + 1 = 8 + 1 = 9.

What glorious certainty! What a happy ending! Does the reader
feel a bit relaxed, cosy, warm, satisfied? A common reaction. It may
even be worth the taxpayers' money, paying all those university
mathematicians; at least they are busy making sure of the facts of
arithmetic. Without them, who knows, perhaps 5 + 4 might slip and
become 6 instead of 9. It's good to know that things are all right,
after all. Or are they?

Far from it. Mathematical ideas, as they trickle down to the popu-
lar mind, become diluted. A sneaking suspicion is abroad that
mathematicians, some few evil ones at least, are casting doubt on
eternal verities, like 5 + 4 = 9 or 2 + 2 = 4. Fie! There is not a
grain of truth in it. The calumny is too low, too base. This vulgar
reaction is an excellent case in point for the novice mathsman to
sharpen his wits on. Popular imagination conjures up the myth of
the wicked professor who teaches that 2 + 2 is not 4 precisely be-
cause it cannot conceive the truth. The truth is much stranger, more
monstrous, more impressive. It is not scepticism about 5 + 4 = 9
that exists, but scepticism about 5, about 4, and about 9.

With a few brackets it is easy enough to see that 5 + 4 is 9. What
is not easy to see is that 5 + 4 is not 6. (Great delicacy and tact are
needed in presenting this idea in conversation, if the aim is, as it
should be, to bewilder and frighten the opponent. His level of
sophistication is very important. He may know all about it; then he
will utter some crushing reply, like 'So what else is new?' He may
even add, 'Just finding out about cyclic groups?'--or mention some
other concept you yourself have never heard of; if he does so, you
have lost the advantage and may not get out of it without a few

28

scratches on your own escutcheon. On the other hand he may be so
ignorant as to be impervious to doubt; you will be laughed at. The
idea is much more useful if the intention is merely to annoy; and is
often used for this purpose by children to teachers. One strong point
is the amount of time that can sometimes be wasted merely making
clear just what is meant; another is that it undermines authority,
since very few teachers can answer this question sensibly.)

When we said that

1 + 1 = 2

this meant only that 1 + 1 was to be called 2. It may perhaps be, for
all we know, that 1 + 1 is really 0, so that we have agreed to call 0
by the name '2'. In other words, it may be that 2 = 0. Who can tell?
It was agreed that all prejudices and preconceptions about 2 were to
be given up. So if we say 0 = 2 we are hardly saying that none are
company or that nothing is certain in life. The addition table below
satisfies the axioms so far; there are not many numbers, which
greatly simplifies all the effort expended in teaching arithmetic.

/--------------------------\
| 0 + 1 = 1 | 1 + 1 = 0 |
|------------+-------------|
| 0 + 0 = 0 | 1 + 0 = 1 |
\--------------------------/

It still remains to be settled whether

1 + 1 = 0.

All we know so far is that it cannot be settled at all if all that is
known about the system of numbers is that it is a monoid. Obviously,
the thing to do is to assume a universal property:

The+ monoid N[0] is universally repelling.

(It is self-evident that this has reference to the subcategory of the
category of all categories; but to obviate any doubt, the morphisms
are the functors.)

Note: The reader may well object to the use of monoids and not
groups. He should be reminded that this book is not a first text in
--------------------------------------------------------------------------
+ pointed, with 1.

29

algebra, and that the difficulties of negative numbers are sufficient
unto another section. It must be admitted that monoids are logically
prior.

If

S = {0, (alpha)} (0 <> (alpha))

is the monoid with table

/--------------------------------------------------------\
| 0 + (alpha) = (alpha) | (alpha) + (alpha) = (alpha) |
|------------------------+-------------------------------|
| 0 + 0 = 0 | (alpha) + 0 = (alpha) |
\--------------------------------------------------------/

then the unique monoid homomorphism

(psi): N[0] -> S

such that

1 -> (alpha)

satisfies

(psi)(1 + 1) = (psi)(1) + (psi)(1) = (alpha) + (alpha) = (alpha);

if now

1 + 1 = 0

then

0 = (psi)(0) = (psi)(1 + 1) = (alpha),

which is false. The contradiction establishes that

1 + 1 <> 0.

Back in section 1, when the subject of discussion was counting,
the phrase 'natural numbers' was mentioned several times. Now that
addition is all the rage, the very same phrase has cropped up.
Moreover, the set of numbers used has both times been written N[0].
But is this really legitimate? Is there really such a close relationship
between the counting system and the adding system? Granted that
many of us are wont to use the same words in ordinary discourse--
we say 'one, two, three, fiddle dee dee' as readily as 'two plus two is
four'--might not this be a convenient shorthand for a situation that

30

is perhaps, in reality much more complicated, and fraught with
difficulty?


QUESTION 5. Whether one should count with the same numbers he
adds with, up to isomorphism?

Objection 1. Of course you must use the same numbers in both
cases! Pshaw, when it comes down to practical applications, despite
all this falferal about morphisms and whatnot, people count things
to see how many they are. Then they put two sets of things alongside
of one another, and use addition to see how many they have got
altogether. Then they count again to see if the answer came out right.
Bring in two separate kinds of numbers, and all that becomes
impossibly difficult. Isomorphism has nothing whatever to do with
it. Ever heard of Steamomorphism? That's the place I go up to when
I want to do some arithmetic. Warm and cosy, not at all like up to
Isomorphism.

Reply to Objection 1. It is far beneath us to meddle with what any-
one finds it convenient to do when he is counting, or adding. All we
want is for everybody to be happy.

Objection 2. Mathematicians have the sort of orderly minds that
want the right tool for the job. It is certainly very unmathematical to
count with one and the same system that is used for adding. It is
like using the very same screwdriver both for putting in screws and
for pulling them out. And mark my words, serious trouble is going
to come when you run into transfinite numbers.

Reply to Objection 2. If anyone is superstitious, he may count with
finite ordinals and add with finite cardinals. It is possible to avoid
encountering those (shudder) transfinite numbers by resolutely not
counting, or adding, anything at all, but just counting and adding
pure and simple.

Objection 3. If you use a number for several different purposes,
who can blame it for becoming a little suspicious? It will begin to
resent such treatment. Especially in these days of 'bloodymindedness'
and continual unrest, we must guard against the development of
system-consciousness among numbers. There is the ever-present
danger that they will begin to ask themselves 'Was sind und was
sollen wir Zahlen?' Perhaps all numbers will desire to be perfect, or

31

at least prime, or even transfinite. Be kind, but firm, to your numbers;
write them on paper that is neither too fine nor too coarse. Do not
let them get together in large sets. Above all, give them a sense of
identification with one specific job.

Reply to Objection 3. Some people seem to like bringing politics
into mathematics. There is really nothing to fear from these radicals.
If all the partially ordered sets of the universe should one day become
discrete, thus discarding their chains, new structures could be
manufactured on demand.


I ANSWER THAT one may count with the very same numbers he adds
with.


PROPOSITION. If you can add, you can count.


Proof. In counting with the additive monoid N[0], we start at 0; after
saying n, we say n + 1; the few diehards still around who prefer to
begin counting at 1 instead of 0 must make the adjustment for once
and go along with the growing trend, because it will be much more
convenient in the present instance to start at 0. Thus we already have
a counting system.

n -> n + 1
{0} :-> N[0] -> N[0].

But having a counting system is not enough. What you must have in
order to assure success in all your counting endeavours is a real, true
initial counting system. So let

{x} :-> X ->[f] X

be any counting system. The set of all functions

X -> X

is easily verified to be a monoid X~. Hence there is (since adding is one
thing we feel absolutely confident about and can do blindfolded with
our hands tied behind our backs) a unique monoid homomorphism

N[0] -> X~

sending 1 -> f

32

which is written

n -> f.

Now it is possible to define a mapping

N[0] -> X

by writing

n -> f(x).

We could go on to show that n -> f(x) is the only morphism of
counting systems from N[0] to X. This would complete the proof that
you can count with the same numbers that you add with. In fact, the
very same set can serve for both counting and adding. The qualifying
phrase 'up to isomorphism' is not strictly necessary. But some of us
consider it a virtue to specify which among isomorphic systems is in
use, only when it is absolutely necessary; and to be as vague as
possible as often as possible. This is a feeling that merges into the
psychological. When two mathematicians discuss the natural num-
bers, do they both have in mind the same numbers, or are there two
distinct but isomorphic systems? It is not the sort of question to ask
very often, but perhaps one may be permitted to ask such a question
once in, say, a sidereal decade.

Addition, like mathematics, occurs on various cultural levels. A
perfect example is provided by a certain meal in a restaurant in
Athens. The diners at the table near the back are mathematicians.
According to the custom of the place, when they have finished the
waiter asks them what they have had, takes down the items on his
pad at dictation, affixing the prices, and adds up. At that point, for
some reason, one of the mathematicians remembers--'Oh, yes. And
besides all that, I also had a beer.' In such an eventuality, even in
Athens, a waiter will commonly add the price of one beer to the sum
already obtained and present the corrected bill. This waiter, instead,
tore up the incorrect bill and added up the whole meal again with the
extra beer included. When the diners explained what was the more
usual procedure in such cases, and suggested that it also produced the
correct sum, the man in question admitted that that might theor-
etically be as they said. But he still stuck fast to his own method.
'I have a restaurant to run; I am not a philosopher.'

33

On the lowest level of mathematical culture, adding up is a ritual,
part of the daily ritual of life; it affects the pattern of events somehow,
and any departure from the ritual pattern produces an aberration in
the course of events. Something might go wrong with the restaurant.
Most waiters live somewhere above the ritual or superstitious plane
of mathematical thought. They understand that the purpose of adding
the bill is a fair exchange of money for goods. Because they are
capable of considerations of this nature, tkey acquire a certain
arithmetical dexterity. Sometimes you have to watch them carefully.
This is the level of applied mathematics.

The mathematicians at the table are at the highest level, that of
pure mathematics. To them. the meal is a function

f: M -> N[0],

where M is the menu and f(m) is the number of items m consumed;
and the adding up is a homomorphism

(beta): N[0] -> N[0]

(the bill function) from the meal-monoid to the Greek currency
monoid (which is isomorphic to N[0]). The bill (beta)(f) for a meal f
is a natural number of drachmas. Since the extra beer is in itself a
small meal, and since

(incorrect hypothetical meal) + (extra beer) = (corrected meal)

in N[0], the additivity of (beta) implies that the usual procedure of
waiters is correct.

One very small difficulty is the commutative law. Many students
of ordinary arithmetic have noticed that

7 + 5 = 12;

and also that

5 + 7 = 12;

they may, indeed, have observed that the equation

a + b = b + a

holds true in a large number of cases in ordinary arithmetic. Now,
of course, ordinary arithmetic is a complicated subject--much more
complicated than the simplified version considered here. It involves

34

digital representations and quite a large number of ideas all mixed
up together. There are lots of numbers. How many people have
actually checked that

19667 + 543628 = 543628 + 19667?

And how many cases of the rule

a + b = b + a

must be verified in order to rule out even one single exception?
Obviously, no number of cases will suffice. The commutative law, as
any schoolchild knows today, does not hold for every conceivable
monoid. An example is a monoid in which 'addition'+ acts by
absorbing the right-hand (or dexter) element into the left-hand (or
sinister) one:

x + y = x

--an exception must be made if one of the elements is the neutral
element, since in any monoid this element must always be absorbed.
To be specific:

(gamma) | (gamma) (alpha) (beta) (gamma)
(beta) | (beta) (alpha) (beta) (gamma)
(alpha) | (alpha) (alpha) (beta) (gamma)
0 | 0 (alpha) (beta) (gamma)
---------+-----------------------------------
| 0 (alpha) (beta) (gamma)

(The table is read like a map, in analytic geometry fashion.)

To prove that the natural numbers commute, one need only
consider the centre; this is obviously a submonoid, and contains 1
if the centraliser of 1 is N[0]. Since the centraliser of 1 is a submonoid,
one now need only show that 1 is an element of this centraliser. This
follows from the fact that

1 + 1 = 1 + 1.

Now the proof is complete, depending only on the universal property
--------------------------------------------------------------------------
+ Admittedly, this behaviour is so queer that convention would prefer to
call it multiplication, not addition. This is because ordinary multiplication
of natural numbers is just as commutative as ordinary addition, but is more
complicated.

35

of N[0]; since the centre N contains 1, the left-hand (or sinister)
homomorphism in

1 -> 1
N[0] -> N -> N[0]

can be defined. The composition must be the identity, so

N = N[0]

and

a + b = b + a

for every pair of natural numbers a and b.

The addition of columns like

1
174
9
34
9
--
?

leads to the general associative laws, which are too deep to consider
as yet.


3. Subtraction


Marco (the waiter) Cameriere meets Pigritio (Piggy) Risorgiamento
in the automat. 'Say, Marco. You know that guy Watermaker?'

'Watermaker the differential geometer?'

'Yeah. Know something? He uses infinitesimals. Infinitesimals
aren't rigorous.'

'Aw, c'mon, Piggy. Everybody uses infinitesimals these days.
They can be made rigorous. They are rigorous.'

'They're not rigorous, Marco. Watermaker needs a lesson in
rigour.' Idly chopping his cigar in bits with poultry shears, Pigritio
emphasises the point. 'What he needs is rigour.'

The following day, a shocking announcement is made at the con-
ference on singularities at New Rochelle. 'Dr Watermaker is in

36

hospital, apparently with a shoulder wound caused by a sawn-off
shotgun or some similar instrument. Fortunately, we have been able
to fill in a replacement speaker. Professor P. Risorgiamento will talk
on "The treatment of obstructions".'

There are many practical applications of tbe idea of subtraction
in everyday life. The general idea may be used for shortening cigars,
for treating obstructions, for removing corns. At present, consider-
ation is restricted to subtraction of one natural number from an-
other. As this is accomplished in ordinary arithmetic, it is a com-
picated operation involving the borrowing of ones and some rather
fancy pencilwork, based on fossilised ritual technique.

A generally accepted idea about subtraction is that it works only
in one direction. For instance,

327
- 143
-----
???

has an answer, and

143
- 327
-----
???

has none. The only case when both 'a - b' and 'b - a' are con-
sidered to be subtractions that can be carried out with natural
numbers, and giving natural numbers as answers as well, is when a
and b are both exactly the same number:

a = b.

The standard line of argument is this: If you could take a from b and
still have some left over, that would mean that b was bigger than a.
Similarly, if b can be taken from a then a is bigger. But it is impossible
for each to be bigger than the other. So the subtraction cannot be
carried out in both directions. If that were all there was to it, the
question would be simple indeed; but, unfortunately, there is a flaw
in the reasoning. To say that one number is bigger than another is
no more or less than to say that the second can be subtracted from
the first, leaving something over. Hence, to say that each of two
numbers cannot be bigger than the other is to repeat the statement
that is to be proved. It is not correct in logic to prove something by

37

saying it over again; that only works in politics, and even there it is
usually considered desirable to repeat the proposition hundreds of
times before considering it as definitely established.

A slightly more mathematical idea is this: Suppose it were possible
to subtract a from b, and also possible to subtract b from a. To make
the idea realistic, write

a - b = p

and

b - a = q.

These equations give

a = b + p
b = a + q

and hence

a = b + p = (a + q) + p = a + (q + p)

If a is cancelled from the above equation, the result is

0 = q + p.

But then it is known that both q and p are 0, so a = b. That is just
what everyone has always thought--the subtraction works in
both directions only if the two numbers are the same number.

The last-mentioned argument appears so finical and precise that
it is in danger of being considered correct. It has one mistake, how-
ever.


QUESTION 6. Whether, in an equation setting the sum of two natural
numbers equal to another, if the same number appears on either side,
one may cancel it?

Objection 1. The word cancel comes from the Latin cancello, to
make like a lattice; to cut crosswise; to deface, rase, or cross out, to
cancel. This in turn is derived from cancelli, -orum, meaning lattices,
balisters or rails to compass in; windows, casements, or peeping-
holes; also little crevisses or crab-fishes. [14] 'And all that have not
fins and scales in the seas, and in the rivers, of all that move in the

38

waters, and of any living thing that is in the waters, they shall be an
abomination unto you...' [15] Since cancelli are an abomination,
one may not cancel.

Reply to Objection 1. The Latin word for 'scales' in the passage
quoted is squamas. Anything that has squamas is perfectly all right,
as the immediately preceding verse states explicitly. But e'crevisses or
crawfish are extremely squeamish, as any little girl will tell you. The
other meaning of cancelli, about ladders and banisters, directly
points to the idea of a ladder or scala, similarly constructed with
rails and used for scaling walls. Hence by another proof, cancelli
have scales.

Objection 2. If something is to be defaced or rased, it should not
have been written in the first place. But in mathematics, anything
may be written in a proof that is an axiom, a hypothesis, or that
follows from previous steps in the proof by modus ponens. If some-
thing should not have been written, it did not satisfy these conditions.
Hence any proof containing a cancellation also contains an error.

Reply to Objection 2. One does not actually need to cross out any-
thing in order to perform the abstract mathematical act of cancel-
lation. What would be done in a formal proof is to copy the equation
over, omitting the terms that are considered to be cancelled. If this is
too difficult, the actual cancellation may be done on a piece of
scratch paper and immediately burned. Then who will know?

Objection 3. Cancellation is not always correct. For example,
2 X 0 = 0 and 314209 X 0 = 0. Hence

2 X 0 = 314209 X 0.

If cancellation were always correct, then by cancelling 0 on both sides
we should obtain

2 = 314209.

But this is false. Hence it is incorrect to cancel in mathematics.

Reply to Objection 3. It is true that cancellation is not always
correct; but it is correct under the conditions stated in the question.
The example given is multiplication.


I ANSWER THAT one may cancel.

39

Suppose that x is a cancellable number; i.e., suppose that

x + p = x + q

never holds true unless

p = q

also holds true, no matter what the numbers p and q are. Suppose
that y is also cancellable. Since

(x + y) + p = x + (y + p)

from

(x + y) + p = (x + y) + q

one can infer that

y + p = y + q

and hence that

p = q.

It has just been proved that the sum of cancellable numbers is itself
cancellable. If

{x} :-> X ->[f] X

is any counting system whatsoever, it can be augmented by writing

X[0] = X (union) {x[0]}

with x[0] (is not (epsilon)) X, and by defining

f[0]: X[0] -> X[0]

so that f[0](y) = f(y) for y (epsilon) X and by f[0](x[0]) = x. Then the
morphism N[0] -> X[0] provides by restriction a morphism

(psi): N[0] + 1 -> X.

With X set equal to N[0], necessarily

(psi)(n + 1) = n;

which shows that

n -> n + 1

is injective and that 1 is a cancellable number. Now one may con-
clude that all numbers are cancellable.

40

Thus, the question whether two quite distinct and unequal num-
bers exist each of which can be subtracted from the other, either way
round you like, has definitely and positively been settled in the
negative. The answer is no. The next question would seem to be, are
there two numbers, neither of which can be subtracted from the
other, no matter how the problem is attempted? These would be two
numbers, neither of which is any bigger than the other, nor are they
the same number. The two numbers could not be compared in size;
they would be mutually incomparable.


QUESTION 7. Whether any two numbers are comparable?

Objection 1. The kinds of comparison--the absolute, compara-
tive, and superlative, are called degrees. Hence the question refers to
the numbers of degrees used in recording temperatures. But if the
temperature exceeds a certain level, mathematics is impossible, since
mathematicians require pencils and paper and these would ignite.
Hence only small numbers are comparable to other numbers.

Reply to Objection 1. A mathematician can deal with arbitrarily
large temperatures under the most comfortable working conditions,
simply by inventing new scales for the measurement of temperature.
This is done by means of a 'scaling factor' which converts one
degree Fahrenheit (or Celsius) into as many degrees as one likes of
the new scale.

Objection 2. Some numbers are so large that no-one can form an
idea of their magnitude. But mathematicians cannot deal with that of
which they can form no idea; hence to compare two such great
magnitudes is beyond them.

Reply to Objection 2. Mathematical statements about natural
numbers are really, or can be converted into, statements about the
system of natural numbers, obviating the necessity for actual con-
templation of huge magnitudes and other such arcane trash.

Objection 3. If any two numbers can be compared so that one is
greater, then there must be a greatest number n. But then n + 1 is
greater than n, contradicting the supremacy of n. Since this is non-
sense, there must be two mutually incomparable numbers.

41

Reply to Objection 3. It is inaccurate to infer that if any two things
of a certain kind are comparable then one of the things is greatest.
The argument of this objection shows correctly that no natural
number is greatest, despite which fact it is the case that any two of
them are comparable.


I ANSWER THAT any two numbers are comparable.

A number is subtractable, for the nonce, if it is comparable to any
other natural number whatsoever. If x is subtractable and if y is any
other number whatsoever, then either the subtraction 'x - y' can be
done, or else the subtraction 'y - x' can be done. (In 'x - y', the
number x is the minuend and y is the subtrahend. +) Suppose that each
of x and y is a subtractable number and that z is any other number
at all. If z can be subtracted from x then

x = z + p

whence

x + y = (z + p) + y = z + (p + y)

so that z can be subtracted from x + y. Otherwise, since x can be
subtracted from z, the equation

z = x + p

holds for suitable p. Since y is subtractable, either

y = p + q

or

p = y + r;

yielding either

x + y = z + q

or z = (x + y) + r.

Since all numbers of the form

1 + p,

together with 0, form a submonoid of N[0] containing 1, the number 1
is subtractable, and so are all natural numbers.
--------------------------------------------------------------------------
+ This terminology is of no use whatsoever.

42

4. The negative-positive diathesis


QUESTION 8. Whether negative numbers are fictions?

Objection 1. Negative numbers are fictions. The expression

'-3 + 4'

has a meaning, since it is just

'4 - 3'

written backwards. It is really possible to take 3 from 4. Since it is
not at all possible to take 4 from 3, the expression

'3 - 4'

is a mere fiction. In practice, it is sometimes convenient to use such
a fiction; the procedure for working out the sum is as follows: Since
no-one could ever take 4 from 3, change the numbers about and take
3 from 4. The answer is 1. But it would be wrong to say that 1 is the
answer to '3 - 4', since a fictional problem cannot have a true
numerical answer. Hence, the minus sign is affixed to the 1 and -1
is given as the answer. Since both sides of the equation

'3 - 4 = - 1'

are nonsense, the equation is true.

Reply to Objection 1. Good fiction is never nonsense. The mistake
here is in thinking positive numbers are more real or actual or con-
crete than negative numbers. Both are equally fictitious.

Objection 2. Negative numbers are not fictions. They represent
such things as debts (to take an example from the field of commerce).
Debts are only fictions if the borrower does not pay them; in that
case they are fictions because that is dishonest. But failure to pay
debts is punishable by law; hence it is wrong; and when something
is wrong, mathematicians need not pay attention to it.

Reply to Objection 2. Money is probably fictitious, too.


I ANSWER THAT negative numbers are fictions, but they are not
nonsense.

43

The system of all positive and negative numbers is the group of
integers. It is indeed possible to construct the group of integers more
or less along the lines of Objection 1. The danger that one encounters
is that of being led to take natural numbers as prior, or natural, and
negative numbers as secondary, or artificial. This is the diathesis that
must be avoided at all costs. There are already too many temptations
to fall into this grievous error. One of the worst is the usual method of
working out sums such as 3 - 4 + 2 or -17 + 4 - 3 + 21 - 13.
A better method, and one less prejudiced against the negative num-
bers, is suggested later, under division.

The widespread prejudice against the negative numbers is totally
unjustified by science. At some point, a mathematician had to point
to either 1 or -1 as being the distinguished element. He did so quite
arbitrarily, but ever since then the positive numbers have felt
superior--just because they have been placed on the right-hand (or
dexter) side. Their opposite numbers on the left-hand side of the
origin are pushed aside when selection is made for the important
jobs. They are considered to be abnormal, peculiar, inferior. How
unjust!

The integers form a group--as every undergraduate, even every
sharp third-former, knows, a monoid in which every element has an
opposite or inverse. Thus

-x + x = 0 = x + (-x).

In other words, a group is a category with one object in which every
morphism is an isomorphism. Just about the simplest category is O~,


/-\
| |i
O '


which has one object O and one morphism i: O -> O. Slightly more
fancy is I~:

44

/-\
| |iA
/-> A '---\
| |
g| |f
| |
| |
\-' B <---/
| |iB
\-/


with gf = iA, fg = iB. From O~ to I~ there exist exactly two functors
and no more, namely a: i -> iA and b: i -> iB. Then the equaliser
z: I~ -> Z of


a
O~ ----------> I~
| /|\
\--------------/
b


is such that


a
O~ ----------> I~
| |
| |
b| |z
| |
\|/ z \|/
I~ ----------> Z




45

commutes, and that if


a
O~ ----------> I~
| |
| |
b| |c
| |
\|/ \|/
I~ ----------> C~
c


commutes then there exists a unique functor d: Z -> C~ such that


/-- S~ --\
| |
z| |c
\|/ \|/
Z -----> C~
d


commutes.

Since z(A) = z(B), it is easily seen that Z is a monoid. The invertible
morphisms in Z form a submonoid and include z(f) and z(g); hence
Z is a group. This group is called the group of integers. The com-
position of the morphisms in Z is usually written additively: m + n
and not mn or m o n. The usual notation for z(iA) or z(iB)--they are
the same--is 0. It is necessary to choose, if the usual notation is
preferred or a distinguished element is desired, whether to use z(f)
or z(g) for 1. Whichever is 1, the other must of course be -1. There
are many ways of making this choice. One might, e.g., ask z(f) and

46

z(g) both to bring offerings. Then one could have respect unto one
of their offerings, and not have respect unto the other. [16] The
morphism unto whose offering you have not respect would then get a
mark set upon it, namely the minus sign. Or, one might flip a coin.

It is possible to get the natural numbers in a similar way. The
category I~ is replaced by K~,


/-\
| |iA
A '---\
|
|f
|
|
' B <---/
| |iB
\-/


There was a time when an axiom was a general ground or rule of
any art, and when the axioms of mathematics were supposed to be
incontrovertible propositions, so obvious as not to require any
proof. That day is past. [17] Nowadays one experiments with new
art forms, in mathematics as in other arts. A new art requires a new
basis, and one may say that at the present time an axiom is a propo-
sition so useful that it must be put beyond question. The truth of an
axiom, in any absolute sense, is not a matter of interest. Mathematics
has been called the art in which men do not know what they are
talking about and cannot tell if what they say is true. The ultimate
judge of the mechanical correctness of a mathematician's work is a
mathematical proof-checking machine; the ultimate judge of its
rightness and suitability is the opinion of the other mathematicians.
As an axiom on which to base the positive numbers and the integers,
which have in the past produced much harmless amusement and are
still widely accepted as useful by most mathematicians, some such
proposition as the following is sometimes considered as being
pleasant, elegant, or at least handy:


AXIOM: Equalisers exist in the category of categories.

47

It is worth emphasising that according to the approach outlined
above--or any other enlightened approach--there is nothing nega-
tive about the negative numbers. The notion is entirely false that
negative numbers are positive numbers turned the wrong way; or
rather, this notion, and the notion that positive numbers are negative
numbers turned the wrong way are equally slanderous. In the group
of integers there is absolutely nothing to tell which numbers are
positive and which are negative. Psychological tests under stringent
conditions of probabilistic and experimental rigour have shown that
subjects, being shown photographs of numbers, did so badly at
identifying the sign of the number that the correlation between the
answer given and the true sign of the number was nil. More sur-
prisingly, leftists and rightists showed no more aptitude at this exer-
cise than people of moderate directional tendencies. What we have
got is not that positive integers can be distinguished in some special,
magical way from other integers; we have got a pointed group

{1} :-> Z

in which the submonoid generated by 1 can of course easily be
distinguished.

On the other hand, it is usually taken for granted that the natural
numbers are part of the integers. The integer 1, the famous and dis-
tinguished element of the group of integers, is written exactly the
same as the distinguished natural number 1. Looking at the numbers
on the pages of this book, the man in the street will seldom ask, is it
48 the natural number, or 48 the integer that is being used to number
this page? Yet the mathematical system within which the page num-
bering is done is surely of the greatest importance. It would appear
to be wrong to do anything at all with numbers without paying close
attention to tbe question of what number system one is working in.
The numbers themselves, indeed, are of secondary importance. It is
the mathematical system that makes all the difference. Moreover,
natural numbers, as opposed to integers, have political overtones:
grocers, classicists, schoolteachers prefer natural numbers, while
technicians and pseudo-intellectuals like integers better. Failure to
make the distinction could lead, not only to confusion, but to a
delicate situation requiring the utmost tact, and a sprit of tolerant
give-and-take if the inevitable differences of opinion based on such

48

a wide mathematical discrepancy are to be avoided. After all, the
system of integers and even the natural number 0 are still regarded
in many quarters as contrivances of the devil.

Fortunately, the difficulty just mentioned is not entirely insur-
mountable. Popular culture, in this permissive age, has mellowed
toward integers; they are still regarded by many as effete intellectual
trash, but it is now felt more and more that effete intellectual trash
has its place. Being effete is increasingly regarded as just another way
of doing your thing. It is admitted by many that the world would be
a duller place without the prime number theorem.

When tempers run high on the question of page numbering in
integers as opposed to natural numbers, there is mathematical oil to
pour on the troubled waters. The fact is that there is a natural, down-
to-earth homomorphism from the naturals to the integers, and that
this homomorphism, sending as it does the natural number 1 to the
integer 1 and being so natural and injective as it certainly is, is just
perfect for being considered as a mere inclusion and nothing else.
In other words, one might as well pretend that some of the integers
actually are natural numbers, and that natural numbers are integers.
The integers one pretends this about are called positive (or non-
negative).




5. Multiplication


One of the heresiarchs of Uqbar declared mirrors and fatherhood to
be abominable, because they multiply the visible universe. [18]
Multiply it by what, one may wish to ask? Here we pass over such
questions belonging principally to heretical theology and the in-
vestigations of Tlo:nistas, just as we ignore the whole matter of the
multiplication table, since this belongs properly to division.

By 'multiplication', properly speaking, a mathematician may mean
practically anything. By 'addition' he may mean a great variety of
things, but not so great a variety as he will mean by 'multiplication'.
What, then, is the main difference between addition and multi-
plication? The most important difference is that addition is always

49

denoted by '+'. This is not quite true. Occasionally, for decorative
purposes primarily, one will see something like

/*\
.-----. * *
/ | \ /*-*-/*\-*-*\
|----+----| or \*-*-\*/-*-*/
\ | / * *
`-----' \*/


Multiplication may be denoted by X or by *, or by something more
monstrous, or by nothing at all. There are other, less important
differences. Usually, addition is commutative. Multiplication may
also be commutative. Addition is usually associative. So is multi-
plication (usually, but not always). The main thing to remember,
for the beginner, is that the things multiplied need not be numbers;
nor need multiplying things make them more multiple, or multiplex.
(It is advisable for the beginner not to try too hard to remember
isolated facts. On a first reading, or possibly for a few readings
thereafter, it may be desirable to forget temporarily everything that
the reader has just learnt.)

Bricks are made of clay which is brought to a plastic consistency
by the admixture of a suitable quantity of water, formed into shapes
established by custom and precedent, dried in the sun, and finally
baked. They differ from the loaves of Beaulieu Derrie`re in that their
dimensions never vary from day to day, and in that they are not in
general square or cubical. The brick, because of its shape and be-
cause of the fact that the ratio of its sides remains constant, may be
used for the computation of volumes. Bricks have, of course, other
uses. Who has not seen, outside the tent of a particularly cheap side-
show at a travelling carnival, an advertisement of the presence within
of that destructive creature, the red Irish batt? The uses of the brick
in cosmetics and in the art of preparing camels for the journey across
the desert are too universally known to merit discussion in a work of
this sort. Slightly more mathematical is the use of bricks in con-
nection with simplified versions of the three-body problem: by com-
puting parabolic trajectories in intersection, one may practise the
applied mathematical art of trashing. None of the foregoing uses, it
may fairly be objected, is a proper use of bricks; in all these instances
we see bricks used, yes--but they are not used qua bricks. When
bricks are used qua bricks, for constructional purposes, three pro-

50

perties at least which bricks enjoy are of great importance: bricks are
movable, rigid, and take up space. If it were not for these properties,
among others, bricks would be useless for building walls.

But are we safe in relying on these ideas? May we feel secure in our
houses, if we have allowed them to pass unquestioned? Is there
really such a thing as the volume of a brick, remaining unchanged
however the brick is displaced by rotation, translation, and the zig-
zag-and-swirl of Lawsonomic motion? [19] It hardly seems a blessed
hope. One way, of course, of dealing at least partially with this
frightening problem is the standard one, familiar to every school
algebraist. (Here we assume bricks whose sides are whole numbers.
Although this rules out the lovely bricks of nineteenth-century
Surrey malt factories, which had faces that were golden rectangles
[20], our result still holds good if applied to the more mundane bricks
of commerce.) The school algebraist will remind his hearer of the
commutativity and associativity of the ring Z. He will point out that
integers, when you are multiplying them, are functions, and that the
composition of functions is about the most associative thing we've
got in this sublunary vale. No doubt he will exhort you to remember
from your schooldays that the centre of a ring is a subring, and that
it contains 1. By a universal argument, he can then demonstrate
beyond chance of contradiction that the volume of a brick of integer
sides does not depend on how you turn the brick--even if you turn
it inside out. (On the question of turning cubes inside out see [21].)


(fg)h = f(gh)
/----------------------------------------------------\
| \|/
| ____ ____
| gh / \ f / \
| /--------------------> < > -------------> < >
| | \____/ \____/
| | Y Z
| | /|\ /|\
| | |g |fg
| | | |
____ ____ |
/ \ h / \ |
< > -----------------> < > ------------------/
\____/ \____/
W X


THE SIDES OF A BRICK ARE (ESSENTIALLY) HOMOMORPHISMS

51

That is, as I may have hinted, the old-fashioned, crotchety school-
boy approach to bricks. A more sensible, and not very different,
way of looking at the problem is this one: The volume of a brick of
dimensions (a, b, c) is the value at (a, b, c) of a trilinear symmetric
function

Z<3> -> Z

taking

(1, 1, 1) -> 1.

(The latter says that a brick whose dimensions are 1 inch in one
direction, again 1 inch in a perpendicular direction, and once more
1 inch in a third direction perpendicular to them both, is 1 cubic inch
--provided the measurements (1, 1, 1) are written down in the correct
order! Note that the non-existence of such bricks is entirely im-
material; indeed, the whole question only becomes intelligible if all
the bricks involved are themselves wholly immaterial.) Obviously,
exactly one trilinear function Z<3> -> Z sending (1, 1, 1) -> 1 exists;
just as obviously, the uniqueness makes it symmetric.


EXERCISE. Show that four-dimensional bricks behave in the same
way.


EXERCISE. A brick is thrown by an extraverted undergraduate
genius through the window of the mathematics common-room at the
University of Both Putfords, Devon. It strikes the head of the Senior
Lecturer in Analytic Manipulations and shatters, falling together on
the floor in such a way that the new volume of the brick exceeds that
of the mathematics building. [22] Indeed, it is now bigger than the
Putfords. Since the mathematics department is now a vacuous body
of men, what degree must be awarded posthumously to the student?
Submit detailed plans for archaeological excavations and recon-
struction of the villages and the new university, bearing in mind the
reduced density of the brick.


QUESTION 9. Whether 2 X 2 = 4?

Objection 1. Geometrically, multiplication is carried out by using

52

rectangles. For example, if a line segment 3 units in length is con-
structed perpendicular to another segment of length 2 units, and if
certain further lines parallel to these are constructed, one obtains a
rectangle divided into 6 unit squares. The fact that from a length of
2 units and a length of 3 units one obtains a rectangle of 6 units is the
reason for saying that 2 X 3 = 6. But from segments 2 units long
each, one obtains not a rectangle at all but a square. Thus one may
not say '2 X 2 = 4,' but only '2^2 = 4'; i.e. 'two, squared, is four'.

Objection 2. Moreover, 2 units multiplied by 2 units is not 4 units
but 4 squared units. But 4 squared is 16. Hence 2 X 2 = 16.

Reply to Objections 1 and 2. Squares are rectangles; in any case,
we are doing arithmetic, not geometry. Hence both arguments are
irrelevant. Besides, it is not at all obvious a priori that 16 is not
exactly 4.


I ANSWER THAT 2 X 2 = 4. The first step is to replace '2' by '1 + 1'.
Few mathematicians doubt that this can be done, but few except the
subtlest logicians can put up a good defence for it against determined
opposition, at least unless they are given adequate time to prepare a
defence. The principle involved is the replacement of one thing by
something equal to it. No doubt there may be contexts in which this
may safely be done; but how hasty can we be in assuming that we
have one here? Difficulties can certainly arise in the application of
the principle here employed. The standard example is [23]. Editions
of works by this author can easily be found in second-hand book-
stores that bear on the spine the words 'by the author of Waverley'.
It is not at all unlikely that someone, somewhere, somewhen has
looked at such a book and said, 'I wonder if the author of Waverley
is Scott?' Now, as a matter of fact, it is very definitely the case that
the author of Waverley is indeed Scott himself. That being the case,
by the application of the famous principle that equals may be sub-
stituted for equals, anyone whatever, be he ever so slow-witted, may
arrive at the conclusion that the speaker might equally well, and just
as truly, have said 'I wonder if Scott is Scott?'--or, 'I wonder if the
author of Waverley is the author of Waverley?' Some mockers have
the temerity to suggest that they, at least, do not intend to be caught
pondering such idiotic tautologies. So, if equals may not always be
substituted for equals, when may they be?'

53

The answer is, of course, that we shall have to invent a special
reason why one can substitute '1 + 1' for '2', at least just this once.
Or else, we must just do it and worry about it later, if at all.

Having got past this hurdle, we now have

2 X 2 = 2 X (1 + 1).

The brackets indicate that we add 1 and 1, and then take 2 times the
result. The next step is to say

2 X (1 + 1) = (2 X 1) + (2 X 1).

Here, the argument is that multiplication is a bilinear map,

Z X Z -> Z.

It is a triviality, arising from the definition of multiplication, that

2 X 1 = 2.

Since 2 + 2 = 4, we are done.


Note. One may also attack the problem via the natural isomorphism

n -> (1 -> n)

from tbe additive group Z to the additive group End(Z), whereby the
obvious ring structure of End(Z) may be transferred back to Z. It is
left to the reader to consider which of the two approaches may be the
more elegant.



6. Division


So far in this present work, it has not been possible to mention cer-
tain very interesting, natural numbers. At least, this is true on the
hypothesis that there are any interesting natural numbers. Students
of the question, whether there are any interesting natural numbers,
are divided on what answer to give. Some experts believe that all
natural numbers are interesting; others, perhaps more cautious of
hasty generalisation or less sanguine of temperament, hold the wiser
and more temperate opinion that no natural number at all is of any
interest. (This view is preferred by the author of the present work,
which should settle the matter.) Whether or not there are any in-
teresting natural numbers, then, it has not been possible to mention

54

any of them so far in this book; because so far in the book, no
means of mentioning natural numbers greater than 9 has been
established. (This statement is made rather informally, and is open
to all sorts of quibbling. Some people would say that it has not been
established what is meant by mentioning something. It would be hard
to show, perhaps, that no mention has been made herein of the ninth
reindeer in Santa Claus' team. A good lawyer would point to all the
occurrences of the letters 'red', in that sequence; the allusions to the
idea of nose; the fact that Carnap is a logician; that this book dwells
heavily on the subject of logic; that Carnap's first name is Rudolf.
Once it is established that Rudolph, the red-nosed reindeer, has been
mentioned, then it becomes clear that almost anything may have
been mentioned. In particular, anything nonexistent may have been
mentioned--Rudolph is known to be apocryphal, to be a later
accretion on the fixed body of mythos septentrionalis. [24] If you can
mention him, you can mention anything or anybody--even all those
nonexistent interesting natural numbers. And very likely they have
been mentioned. The good lawyer might get away with the argument
just outlined, just because nobody knows what it means to mention
a thing, anyway.)

Because of all this quibble, let us merely say that there has been up
to this point no way at all to mention numbers bigger than 9, except
by circumlocution; it was always possible to talk about the sum of
6 and 7, and even to show that this number was bigger than 9; what
was not possible was to write this number as 13. It is still not possible
to write it as 13; or rather, it is just as possible to write it as 13 as to
write it as 31 or 995. There is no system for writing these numbers.
So far as this book is concerned, we do not yet know how to write
them. Equally well, we do not know, mathematically and within the
system of this book, how to name them. We cannot yet count up to a
hundred, though we can count to 9.

The ability to count to a hundred is part of numeration, or the
study of number numbering. Traditionally, one learns the art of
numeration before the other arts of arithmetic. How this can be is of
course beyond the ability of the present author to account; but per-
haps one may suppose a latent ability in the human soul to count,
somewhat on the lines of the supposed ability latent in the human
soul to learn a human language. [25] After all, the old philosophers

55

[26] are said to have taught that man's soul is a number numbering
itself. In the study of mathematics, this pedagogical order must in
any case be reversed. In mathematical terms, if you are to learn to
give names to the numbers in something like the usual system, and
so to count aloud, you must first learn to divide; and that cannot be
done until you have already mastered addition, subtraction, and
multiplication. Fortunately, anyone who has carefully and atten-
tively read the preceding pages of this book has already mastered
the arts of adding, subtracting, and multiplying. (Human beings
appear to have had since the earliest times the ability to multiply;
indeed this faculty seems to extend even to the higher animals, though
with them it appears to be a seasonal activity. Division, strangely,
appears to be very much an acquired taste among the human race,
though oddly enough it is commonly observed under the microscope
as a natural activity of the lowest forms of life--bacteria, protozoa,
and such. The old philosophers do not appear to have considered the
protozoan soul, but if they did they might well think of it as a number
dividing itself.)

Before it will be possible to go any further, it is necessary to dis-
pose of a ridiculous quibble, which the author's mountainous wealth
of practical teaching experience teaches him is bound to arise at this
point. Little minds love to ask big questions, or what appear to them
as big questions; never stopping to reflect how trivial the answer
must be, if only the questioner would take the trouble to think it
through. Sometimes it is necessary for the writer of such a serious
work as the present one to call a halt in the consideration of matters
of real weight and interest and to remember how weak and frail are
the reasoning powers of his lowly readers. Someone, somewhere, has
asked, 'What about the numbers at the foot of the pages? You really
had no right to have them there, since in a work of this nature, they
have no meaning as yet.' The answer will be obvious, if the reader
will only think. It is true that we have already come to pages of the
book with numbers on them that are bigger than 9. In reading a
scientific treatise, however, it is well to remember that patience is a
virtue. One could have compressed all the material of this book, up
to and including the present section, into one page. Very often, in a
book, the first page of text has no number on it at all. One could
have done this simply by telling the printer to use very tiny type. It

56

would have been hard to read, of course. But this was not done. Why?
Because the publisher insisted. Publishers simply insist on numbering
their pages. The author protests, expostulates, threatens--all to no
avail. The author emphatically assures the troubled reader that
absolutely none of the mathematical material of this book depends
essentially on the numbering of the pages. If the pages were not
numbered everything herein contained would be as firm, as crystal
clear as it is now. And even if it were not so, the reader will soon learn
all about these numbers; they are to be thoroughly explained, and
any nagging doubts about them will be dispelled.

But above and beyond these rather trite and obvious remarks, it is
certain from principles of aesthetic economy that the numerals at
the foot of all the pages preceding this page cannot be empty of
meaning; since as we have just seen they do not mean numbers, we
may infer that they mean something else. Hence we have shown that
the numerals have meanings other than their numerical referents:
thus establishing the bases of the science of numerology. This is of
course a very interesting remark indeed, and having turned aside for
a moment to make it, we may now smugly return to the main subject
at hand.


QUESTION 10. Whether the ordinary method of numeration, used by
book publishers and other normal people for writing numbers,
makes sense?

Objection 1. The question is irrelevant to a mathematical discus-
sion, and does not belong in a book of the present sort. Publishers
and grocers, and other simple folk of that ilk, are free to follow the
rules of their respective callings. They may decorate their books and
vegetables with whatever arcane symbols they find suitable, whether
from reasons of custom, cunning, or artistic satisfaction. Such
decorations may quite legitimately include strings of denary digits,
like 666 (a number used to designate certain pages in long books, and
certain beasts). The approval of the mathematical societies is neither
demanded nor required.

Reply to Objection 1. In and of itself, of course, the use of digits
and the paraphernalia of numeration is not proscribed, and is open

57

to any class of people, no matter how base. Occasionally, however,
such uses have become connected by custom in the popular mind with
notions of a quasi-mathematical nature; such as, that a book con-
taining a page 666 must be a longish book, as books go; or that if a
revue advertises '28 beautiful girls 28' there is available a moderately
wide sample of feminine pulchritude. Now these uses of numbers are
so exceedingly various that only in mathematics can we hope to find
any underlying system explaining all at once the common idea present
in each of the ramifications of applied numeration--vegetable shop
numeration, strip-show numeration, lottery numeration, etc.

Objection 2. Let us suppose, then, that the question is admissible,
and that one may legitimately ask if there is sense in the ordinary
way of writing numbers by means of the digits 0, 1, 2, 3, 4, 5, 6, 7, 8,
9, all combined in artful arrays to form figures, such as 343, 78, or
5629. Then the overwhelming weight of the total experience of
civilisation, from the earliest days right up to the present, is that it
docs indeed make perfect sense. Otherwise, everything would have
broken down long ago.

Reply to Objection 2. This most absurd idea is unlikely to tempt
any but the rank mathematical outsider. The whole question clearly
turns on the age of civilisation. Now just here we must be ever so very
careful, as this is a mathematical discussion, and one or two readers
may not be expert palaeontologists. But the author has asked around
among his eminent palaeontological acquaintances, and almost all
of them agree that civilisation had a beginning; that there used to be
a time when there was no civilisation. There was even a time when
there was nobody around that counted. Now, in a discussion of this
kind, we cannot just come straight out with a non-mathematical
statement, like, 'People have only been able to count for a finite
length of time, and before that time there were no people able to
count (or no people at all, or no universe and no time for them to
count in, etc.)' because that is beyond our competence. We are not
supposed to know about all that. But the fact that lots of people do
say all that, and not just people, either, but people who are supposed
to know, does at least put the burden on the other fellow to show that
it is not so. So we may fairly assume that people have so far only had
a finite length of time to do their counting in. That will be one of the
cardinal points of our argument. Another cardinal point, which I

58

think you will find the most eminent psychologists, graphologists,
and linguists will grant, is that it takes a certain minimum length of
time to say, write, or think a number. Finally, physicists will admit
that it is quite customary to use the real numbers as a mathematical
model for the time axis. Since the reals obey the archimedean property,
one is led to the conclusion that only finitely many numbers exist that
have ever actually been considered in terms of a digital representation
and hence there is a biggest such number. Now of course the above
considerations come immediately to the mind of a mathematician,
and he does not need to be reminded of them. (It should be noted, of
course, that even if civilisation, including the use of numbers, has
always existed it does not necessarily follow that all the numbers
have been written and all the arithmetic problems done. Some of them
could still have been left out, by accident as it were.) Nor does he
deny that for practical purposes the experience of the ages provides
reasonable grounds for caballists and stratarithmeticians to continue
their work without the necessity of a theoretical study of mathematics.
Mathematically, however, this is no indication that the idea of digital
numeration makes sense. 'Probability and sensible proof, may well
serve in things naturall, and is commendable: In Mathematicall
reasonings, a probable Argument, is nothing regarded: nor yet the
testimony of sens, any whit credited: But onely a perfect demonstra-
tion, of truths certain, necessary, and invincible: universally and
necessarily concluded: is allowed as sufficient for an Argument
exactly and purely Mathematicall.' [27]

Objection 3. Granted that experience is not a sure and certain
guide in high theoretical matters. It may still be maintained on purely
mathematical grounds that the digital system of numeration is
mathematically correct, and far preferable to the effete intellectual
rubbish often set forth in pretentious articles and lectures, and in the
present preposterous book. If a numeral is given as a string of digits,
and a large number of people are each asked to add 1 to it so as to
get the next number, they will usually all get the same answer. This
shows that they have been taught a trick for adding 1 to any string
of digits; and if they have been taught a trick then such a trick must
exist. This means they have got a counting system.

Reply to Objection 3. What this argument shows is that the
ordinary system of arithmetic must involve some kind of counting

59

system. What it does not show is that it is a universal counting system.
It could be, for instance, that if you start counting at 0, and keep
going long enough, you will find yourself going round in circles,
coming back to the same old numbers that you have already counted
before. If you take a public opinion poll on this subject, only a small
percentage of the people asked will tell you that this is at all likely;
most will say they never heard of such a ridiculous idea, if you can
get them to understand the idea at all. A large number will suppose
that you are proselytising for a new Californian religion, or selling
encyclopedias. But public opinion is almost always wrong. Admit-
tedly, no-one has ever actually had the experience of counting until
he came back to the same numbers again, and recorded the fact. But
this might be because nobody ever went on long enough. To be
absolutely sure, one would have to do the counting on paper, which
is laborious. Then again, has anyone actually counted up to

19749382759345298724867432987875654578578945634543875328639
27497856?

If not, how do we know that in the system that is in popular use this
string of digits ever gets used? If there are any extra strings of digits
that do not in fact come into the counting when we start at 0 but are
left out no matter how long we go on trying to reach them, then the
system fails to be a universal counting system.

Moreover, even to say what a string of digits is one needs N[0]
since the strings of digits are the elements of

{0} (union)
U {0, 1, 2, 3, 4, 5, 6, 7, 8, 9} X {1, 2, 3, 4, 5, 6, 7, 8, 9}.
n (epsilon) N[0]

Objection 4. Perhaps it must be admitted that the ordinary num-
bers written with digits have their weak points from a theoretical
point of view. On the other hand, the natural numbers, being an
abstract system, have a disadvantage or two from a practical point
of view. As an example, take the numbering of pages in books.
Everybody knows how to number them in the good old everyday
system that all the publishers use. How does one number pages
mathematically?

60

Reply to Objection 4. It will be seen below that numbering things
is just as hard, or just as easy, whether you do it with the natural
numbers or with the numerals of commerce--except that if a book is
too long there may be technical difficulties in getting tbe numerals on
the pages towards the end of the book. There would not necessarily
be room to print the numerals on the pages. This difficulty does not
occur with natural numbers, which need not be printed on the pages
at all. For this reason the first printed books and many old manu-
scripts were numbered with natural numbers, causing certain ignor-
ant modern philologists to suppose that the pages were not numbered
at all; which is quite false and merely shows the decline in abstract
knowledge that has accompanied the spread of mere arithmetical
technique.

Nevertheless, the problem of numbering the pages in a book is a
very hard one indeed, and is treated in another place. The point here
is merely that the difficulty does not depend on using, or not using,
digits.


I ANSWER THAT the system of decimal, or denary, numeration+ is
quite all right. Mathematically it is sound, and the difficulty of
explaining how it works, which a priori would make it seem unlikely
that many people could be taught to use it, may safely be disregarded
since (rather amazingly) people do seem to be able to pick it up some-
how, and even to do arithmetic in it.

A string of digits has already been defined in reply to Objection
3. It is more convenient to write numbers backwards, and of course
we shall find it helpful to use the standard notation for an element of
the Cartesian product. Thus instead of 1975 we shall write (5, 7, 9, 1).
Note that empty strings like ( ) have been excluded. = It will now be
explained how to get s + 1, where s is a string of digits; in other
words the successor function will be defined. If s is a string, it is
useful to augment s by tacking on a 0 at the end (in ordinary terms,
the beginning): if s = (s[0], s[1], ..., s[n-1]) then s is replaced by
s' = (s[0], s[1], ..., s[n-1], 0) = (s[0]', s[1]', ..., s[n-1]', s[n]').

We then define s + 1 as follows: let k be the first natural number
such that 0 <= k <= n and such that s[k]' <> 9. If 0 <= j <= k replace
--------------------------------------------------------------------------
+ Also called decadic [28].
= See p. 158.

61

s[j]' by 0 = t[j]'; and replace s[k]' by s[k]' + 1 = t[k]'. If now t[n]' = 0,
remove t[n]', to get

s + 1 = t = (t[0], t[1], ..., t[n-1]) = (t[0]', t[1]', ..., t[n1]').

Otherwise take

s + 1 = t = t'.

It is clear that we have defined a function from the strings of
digits to themselves. In order to show that this system is mathematic-
ally sound, we must establish the universal property. To do so it is
necessary to map the denary numerals to the naturals homomorphic-
ally. The map is of course

n - 1
s -> (sum) s[j]X^j
j = 0

where X is 9 + 1 (the natural number). We shall be able to establish
that this is a homomorphism if we can show that the equation

n - 1
(sum) 9X^j + 1 = X^n
j = 0

holds for natural numbers.

But this is equivalent to the polynomial identity

n - 1
x^n - 1 = (x - 1) (sum) x^j
j = 0

evaluated at x = X. The remaining details of the proof are trivial.
Since the homomorphism just given is obviously surjective, it need
only be shown to be injective.

From the above discussion it is now clear that every number can
be written in denary notation; indeed, in only one way. Since

n - 1 n - 1
(sum) s[j]X^j = s[0] + X (sum) s[j]X^(j - 1)
j = 0 j = 1

n - 2
= s[0] + X (sum) s[j + 1]X^j,
j = 0

we see that it is possible to divide by X. If X is any natural number

62

bigger than 1, and if we replace {0, 1, 2, 3, 4, 5, 6, 7, 8, 9} by
{n (epsilon) N[0]: n < X} then all the above arguments apply, and it is
possible to divide by X. By this we mean, of course, division with quotient
and remainder.

It is not at all obvious why people are so fond of writing numbers
to base ten; in other words, why do they not use some other system
such as the binary or the duodecimal base? One explanation is that
in former times, when such customs became fixed, people were not
so terribly broad-minded and tolerant as they have lately become.
Those of us lucky enough to have twelve fingers, instead of being
praised for our cleverness and admired for our piano playing, were
shunned--despised--rejected. We became pariahs and outcasts. No
one would give us so much as the time of day. Now it was we, the
twelve-fingered supermen, who naturally counted by the duodecimal
system. When we wrote the numbers, they began

1, 2, 3, 4, 5, 6, 7, 8, 9, X, E, 10, 11, ...

(Note that the last few are commonly pronounced 'ten, eleven,
twelve, thirteen' by you starfish--as we call you among ourselves; and
you write them differently.) Only in the last few decades--or duo-
decades, it is roughly the same thing--have we been able to lift the veil
with any degree of safety. Now we are beginning to come out into
the open, and the more broad-minded of you starfish are beginning
to admit that we are infinitely superior to you, and deserve to take
over. By the end of the century it is my own private opinion, which I
seldom dare to reveal publicly, that those of you who do not knuckle
under will be ruthlessly exterminated. We have already disposed of
grosses and grosses of the most virulent decimalists.

If one does accept ten as the base for the expression of numbers in
digits, it is by no means immediate that one must use the numbers
0, 1, 2, 3, 4, 5, 6, 7, 8, 9 as the digits and not some other set of numbers.
In fact, the numbers

-4, -3, -2, -1, 0, 1, 2, 3, 4, 5

will also do as digits. In some ways, they do better. Let us agree to
write 1~, 2~, 3~, 4~ as being more compact than the usual -1, -2, -3,
-4. Then the number usually written 171 is 200 - 30 + 1, and hence
in the new system must not be written 171 at all; it must only be
written 23~1.

63

On the other hand, -171 is -200 + 30 - 1; that is, we must not
write -171, but only 2~31~.

The perceptive reader perceives that one may write both positive
and negative integers in this system without the encumbrance of
minus signs. The multiplication table is easier to learn too. One may
do division by the usual tricks, suitably modified, but only by the use
of an extra symbol, 5~.

Here is a worked example, showing how to divide the number
ordinarily written 4056 by 3:

145~2 quot.
-------
3)414~4~
3
-
11
12
--
14~
15~
---
14~
14~
---
0 rem.

It is now necessary to note that 145~2 = 1352. Note that in dividing
by 3, one must not accept 2 as a remainder; the only remainders
allowed are 1~, 0, and 1. The general rule for an odd divisor is that
the modulus of the remainder must be smaller than half the divisor.
Leaving for posterity this marvellous simplification and generalisa-
tion of the workings of ordinary arithmetic problems, + the author
passes on to the more mundane question below.


QUESTION 11. Does plain long division, as it was before all these
elegant variations were introduced, give the right answer?


I ANSWER THAT it does. The first step in ordinary long division, as
taught in school, says: 'To divide N by d, choose k so that

(10^(k-1))d <= N < (10^k)d.
--------------------------------------------------------------------------
+ Posterity will wish to know how to pronounce the new digits; the
suggestion that they be called ruof, eerht, owt, eno (and evif) seems to
me to have merit. It is due to L. Sabroc.

64

Taking N = n.10^(k-1) + s where 0 <= s < 10^(k-1), since

d <= n + s/10^(k-1) < 10d,

and since s/10^(k-1) is less than 1, the division of n by d gives a quotient
q and remainder r such that q is a digit and r < d. Take q as the first
digit of the answer, and continue the problem by dividing d into
r.10^(k-1) + s.' That is how one is told to do it; and by following
these instructions (which may sometimes be couched in slightly
different language) we have quite generally got good results. But is
it mathematically correct? Certainly--the mathematically inclined
reader has probably found the proof by now, but for the record one
remarks that r.10^(k-1) + s < d.10^(k-1), so that induction on k is
available.


QUESTION 12. Whether it is true that all numbers that will divide
exactly by 2 end in an even figure, 0, 2, 4, 6, 8?

There are numerous possible objections: a number has two ends,
the left-hand end and the right-hand end. In a Hebrew book the num-
bers are still written the right way around even though the writing in
the book is written backwards (and is in a foreign language). Not
only that, but the book itself begins at the end, not at the beginning.
Perhaps one should say that the number either begins or ends in an
even figure, depending on your point of view? Then there is the whole
question of numbers written to base eleven, and such. One mode of
proof is the following (the objections are left for the reader to dispose
of as he sees fit). Every even number 2a is a zero-divisor modulo ten.
By inspection, the only zero-divisors among residue classes modulo
ten are 0, 2, 4, 5, 6, 8. Thus we have shown that every even number
either ends in one of the digits we require it to end in, or else ends in
5. We must now show that no even number actually ends in 5. If 2a
ends in 5 then of course so does 10a, since 5 is idempotent modulo
10. But 10a ends in 0. Hence no even number ends in 5, and we are
done.





65

7. Casting out nines


Enneekbole, or the casting out of nines, is an ancient and interesting
method of checking or verifying the accuracy of an arithmetical
computation. Any addition, subtraction, or multiplication can be
checked by the method of casting out nines. That is, it is certainly
feasible to check the problem in this way. It was customary among
the ancient Romans to check up on the correctness of their decisions
in difficult matters by the method of inspecting the insides of birds,
to see what they had been eating. It would be feasible to try to find
out if sums had been worked correctly by opening up chickens.
Occasionally, in wild and woolly colonial places, if the general
storekeeper demanded payment in excess of what one suspected he
was reasonably entitled to, one might test his sincerity by purchasing
two hunting knives and offering him one of them. This might be
called checking an arithmetic problem by the method of opening up
general storekeepers. Although both of these methods have had their
day and for all we know may or may not become very fashionable
again, it would be expecting too much to demand a mathematical
proof of their effectiveness; too little is known, mathematically
speaking, about the inward parts of birds and merchants. Can the
same be said of the method of casting out of nines?

Just what is the casting out of nines? Let us consider the multiplica-
tion

14 X 11 = 154.

To cast the nines out of 14, we add all the digits of the number,
obtaining 1 + 4 = 5. Similarly for 11 we get 1 + 1 = 2. Casting
nines out of 154, we see that 5 + 4 = 9, so we cast out the 5 and the
4 (sending them, if you will, into a sort of limbus enneadum). This
leaves only the 1. Now instead of the problem we started with we
have got the simpler multiplication

5 X 2 = 1.

Since 5 X 2 is known to be 10, and since 1 + 0 is indeed 1, we see
that the original working was correct, and that indeed

14 X 11 = 154.

66

(It should be remembered that our present purpose is not to defend
the method of casting out nines, but merely to show how it is done.)
To make perfectly sure the method is understood, let us contemplate
another worked example:

3991
X 19
-----
75829

By casting out the two nines, the number 3991 becomes 3 + 1 = 4.
Similarly, 19 becomes 1. Casting out the nine, the answer 75829
becomes 7 + 5 + 8 + 2 = 22, and by casting nines out of this
it becomes 4. Since the problem has reduced to 4 X 1 = 4, it must
be that the answer was correct.

Here then is the general rule for checking any arithmetic problem
by the method of casting out nines: To cast nines out of a number,
be thorough. First cast out all the nines you can see--any nines that
occur as digits, and any sets of digits that add up to nine. Now add
up the digits of the number. But do not stop there, if you can help it.
If the new number so obtained has any nines in it, cast them out. If
there remains more than one digit, add up the digits. Continue in this
way, by repeated application of casting nines and summing digits,
until all that is left of the number is a single digit, which will not be
9 but may be 0. Now do this to every single number that occurs in
the problem. Finally, do the problem all over again, but using the
new numbers that you have obtained as replacements for the old
numbers by the method of casting out all the nines in the old numbers.
When you have got the answer to the new problem that you have done
with the new numbers, cast the nines out of it, too. Compare this
with the number that was obtained from the original answer by the
casting out of nines. They should be the same.

No-one has ever seriously asserted that by the casting out of nines
you will necessarily detect any error you might care to make in
arithmetic. For instance, 17 X 17 reduces to 8 X 8 or 64, and this
gives 6 + 4 which is 10, which gives 1. If someone, by some slip,
obtained the answer

17
X 17
----
100

67

then casting nines out of 100 gives 1 as well, and the method fails to
detect the error. It is not a perfect method for detecting any error, no
matter how gross. But there is the question of the accuracy of the
procedure when it seems to have detected an error.



QUESTION 13. Whether, when a problem has been checked by the
method of casting out nines, and the two answers disagree, the
problem is necessarily wrong?


I ANSWER THAT the method is completely accurate. The cyclic group
of order 9, which we write Z[9], can be taken as the set of natural
numbers {0, 1, 2, 3, 4, 5, 6, 7, 8} with addition modulo 9. If n is any
natural number, write S(n) for the digit sum of n, taken according to
the rules mentioned above--in other words, S(n) is n with the nines
cast out. One considers S as a map N[0] -> Z[9], and it is easy to see that

S(n + 1) = S(n) + 1

of course the + and the 1 on the right-hand side mean something
quite different from the + and the 1 on the left-hand (or sinister)
side. Hence by induction, the map S is a homomorphism. Since
S(1) = 1, it is the standard homomorphism from N[0] to Z[9], which
extends to the standard homomorphism from Z to Z[9]. Since the
latter is a ring homomorphism, not only do we have

S(m + n) = S(m) + S(n)

but also

S(mn) = S(m)S(n).

Note that the first equation above could have been proved directly,
if we had had at our disposal the methods of ordinary school addition,
with carrying and all the rest of it. This subject has been excluded
from the present work, as being somewhat too abstruse for treatment
in an elementary work of this kind.

How effective is the method of enneekbole? If an arbitrary wrong
answer is given to an arithmetic problem, the casting out of nines
will show up the error about 8/9ths of the time--in the sense that the

68

proportion p[N] of wrong answers such that 1 <= A <= N that are
shown up, taken among all wrong answers, obeys

(epsilon)~
A~ E~ A~ (N > M -> |p[N] - 8/9| < (epsilon)).
(epsilon) > 0 M (epsilon) N[0] N (epsilon) N[0]

From this it would appear to be obvious that to check an arithmetical
problem, one ought to look merely at the last digit, and see if that is
right. The reason for preferring this very simple method to casting
out nines is not only its greater simplicity, but also the fact that it
gives a greater chance of detecting an error: 9/10 as opposed to 8/9.
There may be special cases where the casting out of nines is to be
preferred, however. In the calculations just made, it was tacitly
assumed that the worker of the problem was just guessing. Although
this is not impossible, it is more likely that he has learned a few tricks
about working arithmetic problems. One of the most useful tricks
for faking up answers to arithmetic problems is to be sure and get
the first and last digits of the answer correct. This succeeds in giving
the appearance of a correct computation, without requiring much
effort; and anyone poking his nose over your shoulder will probably
not go further than checking the last digit of the answer. To counter
the fact that the problem-solver has probably thought such thoughts
as these, the person checking the problem will be on his guard, and
may find the method of casting out nines to be of use. Faking up a
wrong answer so that it passes the nines test is so laborious that it is
often easier simply to do the sum honestly.

For problems involving rather big numbers, tbe method of casting
out ninety-nines is useful; it is harder to do than the nines test, but it
exposes all but 1'010101...% of the wrong answers. To check

54679
X 2604
---------
142384116,

we take 5 + 46 + 79 = 130, then 1 + 30 = 31; and 26 + 04 = 30.
The product of 31 and 30 is 930, which gives 9 + 30 = 39. The
answer 142384116 gives 1 + 42 + 38 + 41 + 16 = 138, then 1 +
38 = 39. These agree, and so it is rather likely that the answer given
above is correct. It need hardly be noted that one may verify the

69

correctness of a calculation by casting out 999s, 9999s, etc., and
that these methods are successively more accurate. If the method of
casting out 9999999999s passes on the answer given to the problem
above, that it is absolutely certain that the answer is correct--except,
of course, that an error may have been made in doing the problem
by the method of casting out 9999999999s. This is quite possible,
since doing the problem by that method amounts, in the case under
consideration, to just doing the same problem all over again.































70

2

FACTORS AND FRACTIONS


1. Prime numbers


It is commonly stated that prime numbers have no factors except
themselves and 1; see, e.g., [29]. This notion is erroneous for two
reasons: one, that a prime number has four factors--if p is prime,
then p is a multiple of p, of -p, of 1, and of -1. That is, we can find
a number x such that

p = p.x,

or such that

p = (-p).x,
p = 1.x,

or even

p = (-1).x.

The same number cannot be used in each case, but there is always
such a number x that the equation is true, no matter which of the
four equations you are trying to solve. Note that this is true even if p
is not prime. Moreover, unless p is 0, the two numbers p and -p are
actually distinct: p <> -p. (If p = 0, the fact that p has four factors
is even more obvious simply by solving 0 = n.x, which can be done
for any value of n whatever--and in this case one can use the same
solution for every value of n.) There is only one exception to the rule
that every number has four factors at least: this exception is the

71

number 1--and of course its associate -1. This is because in order to
show that the four factors p, -p, 1, -1 are distinct we must have
six 'inequations':


p <> -p

< <> <
> <> >

1 <> -1


The cases p <> -p and 1 <> -1 have already been considered (it was
shown above that 1 <> 0), and the other four can easily be reduced to
two:

p <> 1

and

p <> -1.

Unfortunately, if these do not hold, in other words if p is 1 or -1,
then p has only two factors, since p, -p, 1, -1 are really only two
distinct numbers. The only number, then, that has no factors except
itself and 1 is -1. According to one standard work, which is very
popular, I regret to say, among schoolmasters, there is only one
single prime number. That is, of course, a very serious error. There
are lots of prime numbers. In fact the work cited mentions in the
very next sentence (I quote) 'For example, 7 is a prime number, ...'.
This is a direct contradiction of the definition, since the factors of
7 are 7, -7, 1, -1, all of which are distinct. (This means that none
of the four numbers is the same as any other.)

How is it possible for a writer of such distinction to make such an
elementary mistake as to define prime numbers incorrectly? Very
likely, what is involved here is not a serious misconception of what
prime numbers are--though that is the first thought that strikes one's

72

mind on reading this absurd definition--but the mere omission,
quite possibly by the typesetter and not the author, of some qualify-
ing phrase. Such a phrase would be: 'where associates are of course
to be identified'; or equally well, 'it goes without saying that if a = ub,
where u is a unit, then we must count a and b as the same factor'.
Since {1, -1} is the group of units in the ring of integers, we find
that the definition quoted above becomes more nearly satisfactory
once this qualifying phrase has been restored to it. The objection
remains that in the book to which we refer, no mention at all is made
of what ring we are dealing with. And of course the number 7 is
prime only if we consider it as an element of a suitable ring. If it is the
ring of integers that is involved--since the reader is forced to make a
guess, this is perhaps his shrewdest guess--then 7 is a prime. If one
is in the field of rationals, on the other hand, there are no primes.
But we are not out of the water yet. Let us look again at the defini-
tion of prime number in this text: 'Prime numbers have no factors
except themselves and 1.' This phrase occurs with the words 'Prime
numbers' in bold-face type in a list of similar phrases headed
'DEFINITIONS, OR WHAT THEY MEAN'. Hence we have no choice but to
take it as a definition in the fullest sense of the term; it must mean
not only that prime numbers have this property, but that numbers
having this property are prime numbers. Now as a matter of fact,
the number 1 has no other factors except itself and 1; which is only
another way of saying that the number 1 has no other factors except
the number 1. (It also has as a factor tbe number -1, but this is a
unit times 1, and so for this purpose is not to be distinguished from
the number 1; remember that this was the proviso that the typesetter
forgot to include.) Hence, according to the definition, the number 1 is
a prime. But that is not the case! The number 1 is not prime. There is
no way of escaping the responsibility for this error, and the reference
[29] must be condemned as a pretty awful book.

Let us turn now to the work [30] of another famous author, one
of almost equal distinction whose works also used to be popular in
the schools. They are now not so popular there as the previous
reference. Here, numbers are considered as being obtained by the
addition of units; it is not certain, by the way, that the word 'unit' as
used by this author has its ring-theoretic meaning. Some of the
terminology is rather archaic. But it is observed that some numbers

73

are obtained also by the addition of a quantity that is not a unit.
Thus, 10 is obtained by the addition of twos:

10 = 2 + 2 + 2 + 2 + 2,

as well as by the addition of fives:

10 = 5 + 5.

A prime number is then defined as a number that can only be obtained
by the addition of units. Note that it is not considered fair to say
that 7 can be obtained in different ways; one might wish to say that
we have both

7 = 1 + 1 + 1 + 1 + 1 + 1 + 1

and

7 = 7,

and we might say that the second equation shows 7 as obtained by the
addition of sevens, in fact by the adding up of a string of sevens so
short that it consists of only one seven. But Euclid (the author of the
work in question) does not go along with that. That is not allowed as
an addition. So since 7 can only be obtained by the addition of units,
it is a prime. Note that, for some strange reason, the unit is not
considered to be a number at all. Since 1 is not a number, the ques-
tion does not arise in this theory whether 1 is prime or not. (Even if it
did arise, it is not true that 1 can only be obtained by the addition of
units; it cannot be obtained by addition at all.) The other question,
that of identifying associates, is easily settled: these were always
identified by the ancient mathematicians, to the extent that they
never bothered with negative numbers at all.

A correct definition of prime numbers, then, and one that does
not require the absorption of ancient modes of thought, is this:
We require first of all to know what is a factor of a number. If two
numbers can be multiplied together to give a third number, each of
the two numbers is a factor of the third. A prime number, then, is a
number that has exactly two factors; for this purpose, we consider
two numbers to be the same if one of them is the product of the
other times a unit. (A unit is a number that has an inverse in the ring
under consideration; thus, -1 is a unit in the ring of integers, since
(-1).(-1) = 1, while 509 is not a unit, since 1/509 is not an integer.)
The two factors of a prime number are necessarily the number and 1.

74

The number 1 is not prime, since any factor of 1 is ipso facto a unit.
The number

4,294,967,297

is obviously not prime, since it is 641 X 6,670,417; on the other hand

170,141,183,460,469,231,731,687,303,715,884,105,727

is said to be prime. It is often stated that the numbers 2, 3, 5, 7, 11, 13
are prime. Let us consider the first case.



QUESTION 14. Whether 2 is a prime number?

Objection 1. It is clear that 2 cannot be a prime number. Anyone
looking at the list of primes above will see that 2 sticks out like a sore
thumb. All the other numbers in the list are odd; of all these numbers,
only 2 is even. Clearly, the number 2 does not belong in the list. As a
matter of fact, if you will take the trouble to find longer lists of prime
numbers going up far beyond 13, and even including all the prime
numbers that have ever been discovered, you will not find even one
more even number among them. This makes it all but certain that 2
is not prime.

Reply to Objection 1. This method of argument is common among
the increasing class of people who have been subjected to intelligence
tests. These tests often include questions in which the examinee is
asked to choose the object among a given class of objects that does
not belong, because it is different from all the others. For instance,
one may be shown an amoeba, a picture of John Bull, a hatter, a
diagram of a planet moving in an orbit according to the law of
epicycles, and a wine-bottle. The answer is clearly that the amoeba is
the odd man out, because all the others share the property of being
eccentric (or in the case of the wine-bottle, of causing eccentric
behaviour), whereas the amoeba is amorphous and so cannot have a
well-defined eccentricity. Or else, the amoeba is out because any of
the others is suitable for framing and hanging on the wall, whereas
this is not commonly done wth amoebae. Or because amoebae are
microscopic. At any rate, we are all agreed (if we are intelligent)

75

that the amoeba is the sore thumb. We mark box A and hastily
proceed to the next exercise in discernment.

Mathematical arguments, however, must be independent of dis-
cernment. The most insensitive person must be able to follow them.
The reason why 2 is the only even prime is that every other even
number is divisible by 2, as well as by itself and 1.

Objection 2. People have been contemplating the number 2 for a
very long time. Either you must accept the fact that nobody has ever
been able to break this number up into factors, and hence you must
admit that it is prime, or else you must say that the problem has
never resulted in a definite solution and that very likely no one will
ever know if 2 is prime.

Reply to Objection 2. Mathematically, it is indeed inconclusive
that nobody has ever succeeded in finding any numbers that divide
2 exactly, other than the numbers 1 and 2 which obviously do this;
this does not at all show that no such numbers exist. On the other
hand, it may be quite easy to show that there are no such numbers.
In mathematics, if one approach does not work, one must always try
another; according to the principle catus multifariam deglubitur.

I ANSWER THAT the number 2 is prime. It is easily seen that the only
numbers between 0 and 2, including 0 but excluding 2, are 0 and 1.
Thus the remainder left by any number on division by 2 is either 0 or
1. Hence the quotient ring

Z/2Z,

where 2Z is the ideal in Z generated by 2, has only the elements [0]
and [1], where these are the images of 0 and 1 under the canonical
quotient map. Since [1] must be the unit of this ring, every element of
this ring except [0] is a unit, and the ring is a field. As such, it has no
zero-divisors other than [0]. But looking back now at Z this shows
that if ab = 2, then one of a, b is an element of 2Z; i.e., is an even
number. In other words, we have either a = 2m or b = 2m; say the
former. By cancellation, mb = 1, so that both m and b are units.
(Or one may argue that since Z/2Z is a field, the ideal 2Z is maximal
in Z and hence prime, which implies that the generator is prime.)

The fact that the number 2 is prime is useful in many ways. It
prevents us, for instance, from indulging in the time-wasting habit,

76

to which the ignorant are so prone, of attempting to discover a proper
factor for the number 2. It shows that it would be very unwise to
program a computer to keep on trying numbers one after the other
until the computer should have found a divisor of 2 distinct from 1
and 2. It saves the trouble of looking for a rectangle with integral
sides and area 2.

Prime numbers have the most various applications to the world of
everyday life; I mention here but a few of them. One application
known to most of us is to the carving of roast beef, and is said to be
the discovery of Gnivri, the steward in the feasting hall of Stoattr
the Fat. Gnivri had been on raids in the Mediterranean in his younger
days, and had been found to be a handy fellow in the pillage of
libraries. Before decapitating the head librarian, he would usually
strike up a conversation on general topics--like most of the better
sort of Vikings, Gnivri was quite a pleasant man, really, and enjoyed
a good chat; besides, it helped to break the monotony of chop, chop,
chop. 'Vell,' Gnivri would say, 'it loaks as if the library is goaing to
need a nue head, hoa, hoa, hoa. Yoost tenk vot a lot of stoof you
moost haff lernt ven you voss reading oal dem boaks? Vy doan't you
teach it to may, hoa, hoa? Vot iss oal dett prame nomberce naun-
sanse?' Then he would listen for a while to a diverting lecture on
number theory, or whatever the head librarian's speciality might be,
occasionally breaking in with an appreciative word or to ask an
intelligent question, such as 'Iss dett soa?' It was from these little
diversions, these fleeting contacts with civilised learning, that Gnivri
picked up his bits and pieces of knowledge, including the first few
prime numbers. 'Tvoa, tray, femm, whew, ell vah!, ...' he would
mutter. Sometimes, as he was arranging the skulls of slain enemies
in a row to pick out the best ones for drinking cups, he would pick
out the prime ones first. Later, it was Gnivri, by now retired from
active service as a steward, who made the famous culinary discovery.
Today, if you go into any expensive restaurant worthy of the name,
you will see on the menu 'prime ribs of beef'. The restaurant that
advertises composite ribs of beef is unheard of; but it is well to
remember your number theory if you have the slightest reason to
suspect the honesty of the establishment or the reliability of the
chef's knowledge. 'I say, my good man,' you might wish to ask him
--but in French this is 'Je dis, mon bon homme'--'Just how many

77

prime ribs are there on a beef?' It is surprising how many chefs think
the number one rib is prime.



2. Finding prime factors of a number


If a factor of a number is prime, then it is a prime factor of the
number. For example, 28 = 7 X 4, and so both 7 and 4 are factors of
28, but 4 is not prime, so it is not a prime factor of 28. However, the
number 7 is prime, so it is a prime factor.


QUESTION 15. Whether every number other than +/-1 has a prime
factor?

This question is of great practical interest. If the answer is no, then
when the mathematics master in a school sets the children some
numbers to find the prime factors of as homework, they (or their
parents, whichever one does the homework) can bring up the easy
retort, 'Perhaps there are no prime factors. Some numbers haven't
got any, you know. Hadn't you better show us the existence of these
things you are asking us to find?' An organised use of this line may
be an unfair trick, but it is guaranteed to produce a diversion. The
maths master is not allowed to crack the whip these days; he is
expected to encourage an enquiring spirit. This makes it easy to
disrupt the maths class. On the other hand, the answer may be yes;
perhaps we are about to find that all numbers except 1 and -1 have
got prime factors. This too can be turned to good account. There are
two main approaches, and they should be used separately or succes-
sively, not both at once. The first approach is to take an eager-to-
learn attitude. 'Oh, Mr Samwise, I say! These prime numbers are
terribly interesting. Fantastic! Gosh, sir, is it really true that every
number has got prime factors?' The important thing, after this
rather soft, gooey beginning, is to hang on like the British bulldog.
Insist on a rigorous proof. If you have any suspicion that the master
is trying to sneak through a simplified explanation or an argument
that is merely plausible, the right thing to do would be to gently hint

78

that somewhat more is expected of him, perhaps in such words as
these: 'Oh, wonderful! I almost think I see. Could you be a bit more
explicit?' If this sort of tactic ever results in a real, rigorous proof, it
will unfortunately be necessary to show commensurate gratitude;
however, the time wasted and disruption caused are almost certain
to be worth it. The second approach, in case the answer is that
numbers have got prime factors, is to use this as leverage for not
finding them. What is the point of verifying case after case of a
general proposition that has already been established in full gener-
ality? If we know we can always do it, why bother?

Objection 1. It is very simple to see that a number bigger than 1
has got prime factors. If it is a prime number, then of course it is a
factor of itself since 1 times any number is the number again. If it is
not a prime number, then it breaks into two or more factors.

Reply to Objection 1. It is in the second case that this argument
does not pay sufficient attention to the context. A number might
have lots of factors, and none of these factors need necessarily be
prime. In such a case, the factors themselves would need to have lots
of factors, and none of these factors would be prime, either. It is
necessary to find some way of ascertaining that this is not what
happens.


I ANSWER THAT every number other than 1 and -1 has indeed got a
prime factor. This is obvious in the case of even numbers, because we
already know that 2 is a prime. But the proof will not make use of
this fact. (Try to find some other facts that the proof does not make
use of; after you read the proof, try again!) Since the number n
in question is not a unit, the set of its multiples

a~ = {xn: x (epsilon) z}

is not all of Z. Consider the class S~ of all proper ideals of Z containing
a~ as a subset; the set-inclusion relation (subset) makes S~ a partially
ordered set. Now consider any subclass of S~ with the property that
if b~, c~ (epsilon) C~ then either b~ (subset) c~ or c~ (subset) b~; the
union of C~ is trivially a
proper ideal of Z containing as a subset every ideal of C~ and also
containing a~ as a subset. By Zorn's Lemma, a proper ideal m~ of Z
exists that is maximal with respect to the property of being a proper
ideal of Z containing a~ as a subset. Hence m~ is maximal with respect

79

to the property of being a proper ideal; and hence is a prime ideal.

Now, in the ring Z every ideal has a generator. (This is explained
below under greatest common divisors.) The generator of a prime
ideal is prime; since n is in this ideal, we are done. It should be noted
that the existence of the maximal ideal mentioned above need not be
referred to Zorn's Lemma, for in the particular case of Z one may
rely on the fact that Z is Noetherian, and hence every class of ideals
has a maximal element. This again follows from the fact that every
ideal of integers has a generator. Zorn's Lemma, of course, is an
axiom of set theory, equivalent to the Axiom of Choice. The Axiom
of Choice is often explained in terms of drawers full of socks. Many
mathematicians do not like mentioning or even alluding to socks,
drawers, knickers, and such unmentionables in their books and
lectures, and this accounts for some of their shyness in using the
Axiom of Choice and related ideas. They try to get round it somehow.
In the same way, many ladies of delicate feelings refer to the part of
the leg of a roast chicken (or other bird that is carved at the table)
that is not the drumstick--the piece that connects the drumstick to
the chicken--as the second joint or the upper pulkeh. Thus they avoid
saying 'thigh'. In Arkansas, one does not say 'privy' or 'outhouse',
but 'State Capitol'. In this age of miniskirts, it is perhaps charming
that older customs of reticence have been preserved in specialised
areas--mathematics, carving, politics, to name a few; but the march
of progress moves inexorably forward, and these refinements can
only be regarded as frail relics of the past. The days when one could
get up from the dinner-table and leave the room at the mention of the
Axiom of Choice are gone with the wind, and we must accustom
ourselves to the new ways, the ways of tomorrow.



3. The greatest common divisor


The term 'greatest common divisor' is easily explained. This is a
noun phrase, and at the head we find the noun 'divisor'. This means
the same thing as 'factor'. Having said this, it is very important to
point out that these words have several meanings in popular usage,
and naturally as technical terms they must be treated with care. Not

80

just any of the popular meanings of the words will do. The word
'factor' in English often has the meaning 'Institor, negotiator,
procurator negotiorum; that is, an huckster, a foreman of a shop, one
that goes about with linen or woollen cloth'. If one spells the other
word 'deviser'--and this would itself be a great mistake--then one
might get the idea that the greatest common deviser, or the highest
common factor, as it is sometimes expressed, must be the most
important man in the House of Commons. This personage is often
chosen for his ability to devise, and he is usually accounted the great-
est or highest common person in the land. Indeed, divisions are not
at all uncommon in the House. It would be a serious mistake to
think that the greatest common divisor is the Prime Minister. The
greatest common divisor has nothing whatever to do with Parliament
or the Cabinet at all; it is erroneous even to think of the Composite
Ministers as common divisors. (This is the name sometimes given to
the ministers of Her Majesty's Cabinet who are not Prime.)

Furthermore, we must carefully guard against the idea that
common factors are in the habit of picking their noses in public, or
that they say 'Oy loike poineapple oice,' 'Haiw naiw, braiwn caiw,'
or indeed that they are guilty of any vulgarity whatsoever. Any
mental picture of a highest common factor drunkenly staggering up
the down escalator in a department store on Christmas Eve, or
ecstatically chanting 'Hare Krishna' in an awful accent while chewing
convolvulus seeds, would be misleading. It is not that kind of 'high'
that is meant.

By now it should be clear that the number of misconceptions about
greatest common divisors that are possible, while perhaps not
precisely infinite, is colossal.

This is why it may be desirable to consult a standard arithmetic
book if you want to know what the phrase means. There one is
likely to see 'The highest common factor or h.c.f. is the biggest of the
common factors of two or more numbers.' This approximates rather
closely to the right idea, but it has some flaws. After all, does it not
seem queer to take all the trouble of defining the highest common
factor of 'two or more numbers', and yet stop short by not even
discussing the highest common factor of a single number, taken all by
itself? It is even more ludicrous, though perhaps excusable, to leave
unstated what one means by the greatest common divisor of no

81

numbers at all. But these are trifling quibbles; the only change
necessary is to make it read, 'The h.c.f. (of a set of numbers) is the
biggest of the common factors of the numbers.' This allows the set of
numbers to be empty.

A more serious difficulty is with the word 'biggest'. There can be
no doubt that this means just what it says. We have no choice, in
the context of any known standard arithmetic text, but to interpret
one number to be bigger than a second if the second can be sub-
tracted from the first, leaving a positive answer. That is what bigger
always means. And a number is the biggest of all the common
divisors if it is bigger than every other common divisor. Given all this
knowledge, we may ask: what is the highest common factor of the
numbers 0 and 0? (There is no rule that says the two numbers may
not be the same. If someone insists that they may not be the same,
we can ask the same question in a different way: what is the highest
common factor of 0?) Now, you see, if you are to apply the rule
given in all the books, you have got to take all the divisors of 0; and
then look for the biggest. That seems easy. First of all, we shall
identify a number with any other number obtained from it by
multiplication by a unit. All the arithmetic books do this in this
context; it means ignoring negative numbers. We note that 0 is a
factor of 0, since 0 = 275 X 0. That is one factor of 0 out of the way.
Then also, the number 1 is a factor of 0, since 0 = 0 X 1. And since
0 = 0 X 2, the number 2 is a factor of 0. After a length of time
which varies from individual to individual, one may wake up to the
fact that every number is a factor of 0. Wonderful! We now can be
certain that we have found all the factors of 0. Now to find the
biggest. By this we obviously mean a number that is bigger than every
other number. Again, it is a variable length of time before people
wake up to the fact that there is no such number. If n is the highest
common factor of 0 and 0, then n + 1 is also a common factor, so n
is bigger than n + 1. This means that n + 1 can be subtracted from
n leaving a positive answer. This answer is -1. In fact, it can be
shown by trichotomy that -1 is not a positive number. Now the
trouble with all this is not, as it would appear, that 0 has not got a
highest common factor. The whole trouble is in the use of the word
'biggest'.

And of course it is a most puerile error to say that the highest

82

common factor is the biggest of the common factors. It is only the
youngest of children who make this silly mistake; they often confuse
height, size, age, and so on. If you ask an infant who is the tallest
child in the class at school, he is likely to say 'Robert, because he's
seven,' or something like that. The young child cannot distinguish
between height and age. Older folks like us know that they are
different concepts, but often go together. The correct definition is of
course this: the highest common factor is the highest of the common
factors of the numbers. Numbers, you see, are ranked in a hierarchy;
some of them are higher than others. Now the highest number in the
whole hierarchy is 0. One often wonders why a number with so
little apparent character was given this exalted position, but that is
not for us to say. It has always been that way, and it would appear
unlikely that any change will occur for some time yet to come. But
of course, the number 0 is not the biggest nor even the greatest of the
numbers; though he is high, he is little.

Now what, exactly, constitutes height among numbers? A number
is defined to be higher than any of the factors of the number. Now
that one has made this change in the definition of the h.c.f., it begins
to make sense.


QUESTION 16. Whether two numbers have got a highest common
factor?

Objection 1. Before someone came along and made the definition
explained above, one had no trouble finding highest common factors.
Take the numbers 4 and 8, for example. The only numbers that divide
both of them are 1, 2, and 4. Now if anybody gives you a set of
numbers, at least if it is a finite set, you can find the biggest of them.
In this case, 4 is bigger than 1 or 2. But if someone gives you a finite
set of numbers--take, for example, 2 and 3--can you find the highest
of them? In the case under consideration (2 and 3) neither of the
numbers divides evenly into the other without leaving a remainder.
So neither is higher than the other. Then which is the highest? It
seems that there is no highest number in this set.

Reply to Objection 1. It is true that a set of numbers need not
contain a highest. This sort of thing often occurs in hierarchies. Two
elements in a hierarchy (which is only a colourful name, here, for a

83

partially ordered set) need not be comparable; and among a set of
objects, there need not be a highest. There is no highest among 2 and
3, + but as a matter of fact there is no set of numbers that has 2 and 3
as its only common divisors, so it does not matter. All that is re-
quired is that the set of common divisors of a set of numbers should
have a highest element, and as we shall see this is the case.

Objection 2. The normal method of finding highest common
factors is to split both numbers into products of primes. Then it is
only necessary to multiply all the primes that go into both numbers.
For example, 30 = 2 X 3 X 5 and 42 = 2 X 3 X 7, so the h.c.f.
is 2 X 3, or 6. This seems so simple that one ought to be able to
devise a proof along these lines.

Reply to Objection 2. It is clear that 6 divides both 30 and 42, but
without experiment or further proof, it is not clear that 13 does not.
It is inexpedient to attempt to tackle the problem in this way.


I ANSWER THAT any set of numbers (two numbers, or more, or fewer)
has a highest common factor. Let x and y be the numbers. It is only
necessary to find a common divisor of x and y that can be written in
the form rx + sy, since any number a that divides exactly into both
x and y also divides exactly into rx + sy--we have

x = ka & y = ma -> rx + sy = (rk + sm)a.

A similar formulation of the problem works for any set of numbers:
instead of x and y, we can use x[i], where i runs through a suitable
index set. Thus we are looking for a generator for the ideal generated
by {x[i]} i (epsilon) I. If this ideal is {0}, the conclusion is obvious;
otherwise, the ideal has a least strictly positive element, and this clearly
generates.


PROBLEM. To find the h.c.f. of two numbers.

The normal method of doing this is mentioned above. It begins:
'Split each number up into its prime factors, by dividing in turn by
the prime numbers.' As we shall see later, there are infinitely many
--------------------------------------------------------------------------
+ Some purists would say, 'no higher among 2 and 3', or even perhaps
'between 2 and 3'. Since the number of elements in the set is irrelevant to
the argument, it is mathematically better to ignore the fact that there are
two; this leaves us with the superlative, which is the general case.

84

prime numbers, so that no table listing all of them actually exists.
Hence it is not a mere algorithm to divide a number in turn by the
prime numbers. Only a few of the prime numbers are known, so that
one would never know if one had left one out. The question whether
in fact numbers can be split up into prime factors, even theoretically,
has not yet been discussed here. Neither has the question whether,
supposing it is possible in an individual case to split both the numbers
into prime factors, the further instructions will really give the highest
common factor.

Fortunately, there is no need to despair. If (a, b) is the ideal
generated by a and b then (a, b) = (a - bq, b), and this gives us
Euclid's algorithm via well-ordering and ordinary division with
remainder. Thus to find the h.c.f. of 715 and 413, we write


715 - 413 = 302
413 - 302 = 111
302 - 111.2 = 80
111 - 80 = 31
80 - 31.2 = 18
31 - 18 = 13
18 - 13 = 5
13 - 5.2 = 3
5 - 3 = 2
3 - 2 = 1
2 - 1.2 = 0


The last number obtained on the right-hand side, not counting 0,
is the h.c.f. If both the numbers at the beginning were positive, this
method takes a finite number of steps; if the smaller of the numbers
has three digits you will be finished in less than twenty steps.

Now consider doing the same problem by the so-called normal
method of the school-books. First we must split 715 into primes (we
have as yet no assurance that this is possible, but it may be; and we
may not need any of the primes we have never heard of, such as 97
or 131,071). Since 715 ends in 5, it is divisible by 5; we find

715 = 5 X 143.

85

Now by casting out nines, 143 = 8 (mod 9) and hence is not a zero-
divisor (mod 9); therefore 3 does not divide 143. Since 143 does
not end in 0, 2, 4, 6, or 8, it is not divisible by 2; since it does not
end in 0 or 5 it is not divisible by 5. Since 143 leaves remainder
3 <> 0 on division by 7, it is not divisible by 7. Trying 11, we find
143 = 11.13. Hence

715 = 5 X 11 X 13.

Attacking 413 in a similar way, we find that 2, 3, and 5 do not divide
413--for the same reasons mentioned when 143 was under considera-
tion. Then we find that 7 divides 413, and in fact 413 = 7 X 59. Now
the list of prime numbers that the average user of a school arithmetic
book carries in his head is not long enough to take in 59; he will not
be sure that 59 is a prime, but he will also not be able to think of a
factor other than 59 or 1. If he knows a little elementary number
theory, he may know that if 59 is not prime; then since it is not a
perfect square (being congruent to 3 modulo 4, which perfect squares
are not) it has a prime factor less than its square root. Since 59 < 64,
and (root)64 = 8, he will reflect that such a prime will be at most 7.
There are four such primes: 2, 3, 5, and 7. By one trick or another,
or by actual division in digits, it can be verified that none of these
goes into 59 without leaving a remainder. Hence

413 = 7 X 59

and the second number has been factored into primes. None of these
primes (neither of them, if one has counted and found there are two)
is also one of those in the particular factorisation of 715 that happens
to have been found, though it would take some work to show that
some different way of writing 715 as the product of primes might not
contain 7 or 59. The directions in the school-book do not say explicitly
what to do in this case, but the teacher will just possibly know, and
the right thing to do is this: you must still multiply together all the
primes common to both factorisations. Now in this case there are
none, so we must multiply together no numbers at all. This is easy if
you know how to do it, and the answer is that the product of no
numbers at all, multiplied together, is 1. Hence the h.c.f. of 715 and
413 is 1.

86

QUESTION 17. Whether a number can be split into prime factors in
more than one way?

Objection 1. It would seem that any number that is not a power of a
prime, or even any number that can be split up into primes otherwise
than as a power of a prime, is writable in more than one way as a
product of primes. For example, 6 = 2 X 3 = 3 X 2. Here the
number 6 is written as the product of primes in two distinct ways.

Reply to Objection 1. This objection is not fair. The question did
not mean to consider 2 X 3 and 3 X 2 as different ways of writing 6;
but by the phrasing of the question, which is really far too vague,
we have let ourselves in for this difficulty. This question is very im-
portant, as all the children in civilised countries are taught (rightly
or wrongly) to find h.c.f.'s and lowest common denominators by
assuming that the answer is yes. It is sensible at least to know what
the question is that these little souls are assuming the answer to is yes.
The example in the objection shows that by a way of writing a number
as a product of primes, one does not mean a way of writing it as the
product of a sequence of primes. Two finite sequences, say (2, 3) and
(3, 2), can give the same number. Nor can the idea be that of the
product of a set of primes: the number 4 is the product of two 2's,
and the number 8 is the product of three 2's, but the set of primes
involved is in both cases the same set, namely {2}. The product of the
primes in this set is in both cases the same number, which is neither
4 nor 8 but 2. In fact the only numbers that can be written as the
product of a set of primes are those that are quadratfrei [31]. As a
matter of fact, the correct concept is that of a finite set with finite
multiplicities; what is asserted here is that if

L+(P,N[0])

is the set of functions f: P -> N[0] with the property that

E~ (epsilon)~ A~ p >= n[0] -> f(p) = 0
n[0] (epsilon) N[0] p (epsilon) P

then there exists a bijection N[1] -> L+(P, N[0]) such that if n -> f then

n = (product) p^f(p)
p (epsilon) P

Here P is the prime numbers and N[1] = N[0] (without) {0}. Why cannot this
be made plain in the schools?

87

I ANSWER THAT there is exactly one way to write a number as the
product of primes, once that concept is rightly understood. Consider-
ing the difficulty of formulating the question precisely, it is truly
amazing that this was so early discovered; even more amazing that
it is part of the arithmetic curriculum in the average school. If p is
prime and does not divide m, then it is clear from the definition of
prime numbers that one can find r, s such that 1 = rp + sm. For
example, the prime 11 does not go evenly into 100, and in fact

1 = (-99) X 11 + 100 X 100.

Once this is known, if p divides exactly into a product ab, then either
p divides into a or one can write

b = rpb + sab,

which implies p|b. By induction on (sum) f(p), we are done.
p (epsilon) P

The mathematical subject which we have been considering is all
about dividing one number by another number in such a way as to get
no remainder. It is easy to divide a number by another number, but
it is not so easy to do it in such a way as to get no remainder. This
task is by no means always easy; in fact, it is often impossible. This
fascinating subject, together with all its ramifications, is called
divisibility. People often devote their lives to a part of mathematics
that they have chosen for various characteristics. Of course, until
one has devoted his life to a part of mathematics, one cannot know
for certain what the characteristics of the particular field actually
are; and by the time one does know, it is often too late. One of the
major criteria a mathematician looks for in choosing a subject to
work in is uselessness. It is considered fun to do something really
useless. One American mathematician had devoted much of his
life's work to divisibility and related topics. In the autumn of a fruit-
ful sojourn on earth, he rested on his laurels, gratefully remembering
how useless his work had been. To fill the idle hours, he played with
the dial of a radio receiver of the kind that can get anything. Picking
up a ground-to-air communication, he heard this:

'Hey, Mac. How's divisibility up dere?'

'Pretty bad, Chet. How's divisibility on de ground?'

'Can't see a ting, Mac.'

88

'Neider can I. Guess I better come in.'

'Dat's right, Mac. Divisibility is one of de most important tings
when you're up in de air.'

On learning that his darling subject, on which he had spent the
best years of his life, had applications to aeronautics, the poor man
lost his mind, and became a folk-singer. His songs were based on the
Pythagorean idea that music is applied arithmetic, and had titles
such as 'A protest song based on the first five Mersenne primes, for
sackbut and thumb-piano', or 'A contrapuntal imprecation based on
65537, for ophicleide and calliope, in 17 parts for hoarse vocalists,'
and won world-wide acclaim.



4. Guess the next number


A great deal of what we learn at school is of little use in later life.
This is especially true of mathematics. Beyond the most basic
arithmetic, which does have a use in checking the bill in a restaurant,
there is very little that is ever used again except by specialists. A
knowledge of probability theory is handy for an undertaker, so that
he can work out when his customers are likely to need him; a little
topological group representation theory is not amiss if you happen to
end up a quantum mechanic, repairing other peoples' quanta when
they begin to wear out. But for most of us, most of our mathematics
moulders away slowly as the brain cells blink out, cell after cell, in
our heads. It never gets used.

Number guessing is an exception. Why is it not taught in the
schools? This is one branch of quasi-mathematical trickery that
everybody needs desperately, yet it is never found in the school texts.
By number guessing, I mean being able to answer those little riddles,
like the following.

Here is a little test. Do not be afraid of it; the questions are all of
the same type. Though they start out easy at first, pretty soon they
are going to get much, much harder. So keep calm. All you have to
do is guess the next number in the sequence. We give you, gratis,
some of the numbers, which shows how very kind we are, because
then we only ask for one number back from you. Sounds simple,

89

doesn't it? In case you still haven't caught on, the first example is
worked for you. Here goes:

(1) 8, 75, 3, 9, ____.

Now all you have to do is look at the numbers, and then in the
blank provided write in the number that seems to you logically ought
to go there. Now read the numbers again: eight, seventy-five, three,
nine, ... What was that you were about to say? Was it 17? Right!
The only number any sensible person would put there is 17. So we
write in the number 17 in the space provided, like this:

(1) 8, 75, 3, 9, 17.
--

That's all you have to do. Good luck!

Everyone has encountered one of these little tests, and we all know
how much depends on them--that place in a really good university,
that step up in the firm that has been hanging fire for years, that
membership of MENSA--the chance to look down on more and more
stupid people who cannot guess the next number.

Really, it is inexcusable that this art is not taught in every school.
The scientific fact is universally acknowledged that only intelligent
people can do these puzzles; moreover, nobody denies that there is a
crying need for intelligence in all areas of the national economy.
Hence, and one would think the inference would be obvious to any
person who can guess the next number even in the easy example we
saw just above, all that needs to be done in order to cure a vast
proportion of the world's ills is to teach everyone to guess the next
number. Because then, naturally, everyone would be intelligent. The
only case of which the writer knows in which so simple a remedy is
known to exist for so serious a social disease is that of mental illness.
It is universally accepted that by examining a man's handwriting,
trained graphologists can determine what is his mental character, and
in particular diagnose any tendencies to violence, mania, etc. Yet
nothing is done to teach sane handwriting! But, lest we deceive our-
selves, that might involve some difficulty. Good handwriting is not
acquired overnight. Fortunately, the ability to answer the sort of
puzzle that is the subject of this section can be acquired in anything
from a few seconds to an hour.

There is only one little snag. The people who set the intelligence
tests are a very special breed. After you learn how to answer the sort

90

of question we are considering here, you will be absolutely certain
that the answer you put down is correct; not so the fellow who made
up the test. He has his own way of looking at things. The fact is, by
his own intelligence test he may not be very intelligent. He may not
recognise a correct answer when he sees one. Perhaps that is why he
went into his particular line of work. Who ever thinks to question
the intelligence of a man whose very job is testing intelligence?
What occupation can you think of that makes you utterly safe from
the prying doubter who asks, 'Has he got a high I.Q.?' Quis exami-
nabit examinatores ipsos?



QUESTION 18. Whether there is a simple method, whereby one may
always give a logical answer to the sort of puzzle that says, 'Here are
some numbers; what number comes next?'; and whether it is easily,
painlessly, and rapidly learnt?

Objection 1. It would be very surprising if such a method existed.
The example mentioned earlier was 8, 75, 3, 9, ____. There does not
appear to be any relationship at all connecting those four numbers.
Now if one were a better mathematician, or had a better mind for
figures, granted one might be able to see how to connect them up;
and then one could see what number follows logically after the ones
already given. But surely this subject is simple only to the most
accomplished number-cruncher.

Reply to Objection 1. It is true that some knowledge of mathe-
matics is necessary before one can actually write down a 'relationship'
that will produce these numbers; but it is not all that stupendous a
task. Writing down such a relationship is a mere technical difficulty,
and if one is good at mathematics he can master the problem of
writing down such a relationship in a few minutes. Above all, though,
one must not forget that no intelligence test or puzzle of this sort
actually requires anyone to write down a mathematical formula.
It is not necessary even to understand what it is that connects the
numbers and makes them come out in that order. The trick which is
going to be explained gives you an easy, one might almost say magic
method for giving the right answer without understanding anything!

91

Objection 2. The idea that such questions can be answered easily is
absurd in the extreme. Great care is taken in the devising and arrang-
ing of the questions to test the examinee's abstract pattern-recogni-
tion level. While it is conceivable that an examinee will do slightly
better or slightly worse than he ought to do, it is a matter of experi-
ence that examinees tend to come out about the same if tested again.
Stupid people absolutely cannot do well on this test, except by a
once-in-a-million chance, like guessing all the right numbers on the
lottery.

Reply to Objection 2. It is true that very stupid people cannot do
this test; nor can very uneducated people. It is necessary to be able to
write a number, which illiterates and cretins cannot do. Moreover, it
is necessary to be able to understand that one ought to write a
number, and where. These are not very taxing demands; all the
abstract pattern-recognition that is needful is contained in them.
One of the abstract patterns that does, admittedly, have to be
recognised is the line on which one is meant to write the correct
answer.


I ANSWER THAT there is an easy, painless, and simple method for
always writing in a correct answer to one of these brain-twisters.
The following are some examples of the kind of thing one gets.

(1) 5, 4, 3, 2, ____.
(2) 2, 4, 6, 8, ____.
(3) 1, 3, 5, 7, ____.
(4) 1, 2, 4, 8, 16, ____.

And now here are some of the answers one is likely to get, with the
reasoning behind them.

Example 1. In order to see more clearly the relationship among
the numbers, let us draw a graph (Figure 1).

We see that the dots are on a straight line. It is logical to continue
the line, and if we do so, we see that the next dot should be at the
place marked with a circle. The technique is called extrapolation;
it is also used to determine the price of tobacco in the year A.D. 2000,

92


#------X------+------+------+------+-
# | | | | |
#------+------X------+------+------+-
# | | | | |
#------+------+------X------+------+-
# | | | | |
#------+------+------+------X------+-
# | | | | /+\
#------+------+------+------+-----+++
# | | | | \+/
===#====================================
#
Figure 1



# | | | | | /+\
#------+------+------+------+------+-----+++--
# | | | | | \+/
#------+------+------+------+------X------+---
# | | | | | |
#------+------+------+------X------+------+---
# | | | | | |
#------+------+------X------+------+------+---
# | | | | | |
#------+------X------+------+------+------+---
# | | | | | |
===#======X======================================
#
Figure 2


93

and to predict the future fortunes of transcendental meditation
among humanoids.) Thus the answer is

5, 4, 3, 2, 1.
-

Example 2. This time, if the numbers are plotted on a graph, we
do not get a straight line. Trying to think of something else to do, we
first take the logarithms to the base 2, and then plot the graph. These
are: log2 1 = 0, log2 2 = 1, log2 4 = 2, log2 8 = 3, log2 16 = 4.
Hence the graph will be as shown in Figure 2: a straight line; the
extrapolation gives 5, and taking 2^5 we get 32. Thus the answer is

1, 2, 4, 8, 16, 32.
--

Emphatically, this is not the method being presented here. Not
everyone can get the correct answer by this sort of argument. Before
explaining the sensible approach to the problem, it was thought
useful to remind the reader how he is expected to do the problem.
This approach is not systematic. Who could guess that one would
get a straight line by taking the logarithms of the numbers in the
fourth example? Why not take the exponential function of the
numbers, or the inverse tangent?

One easily proves by induction that if f is a polynomial function
of degree <= n, and if f(x[i]) = 0 for every integer i such that 0 <= i
<= n wherever (x[i]) [0 <= i <= n] is a finite sequence of real numbers in-
dexed by the set of integers i such that 0 <= i <= n with the property
that x[i] <> x[j], if i <> j, then f is identically zero. Moreover, the poly-
nomial function

(product) (x - x[i])
n i <> k
(sum) y[k] -----------------------
k = 0 (product) (x[k] - x[i])
i <> k

(under the same hypotheses about (x[i]) has degree <= n, and sends
x[i] -> y[i], for 0 <= i <= n; it must be the unique polynomial that does
so. This method provides a systematic method for solving our
problem, which gives the formula

1 + (7/12)x + (11/24)x^2 - (1/12)x^3 + (1/24)x^4

94

for Example (2). Thus if x = 0, then we get 1 + 0 + 0 + 0 + 0 = 1;
if x = 1 we get

1 + 7/12 + 11/24 - 1/12 + 1/24 = 2;

if x = 2 we get

1 + 14/12 + 44/24 - 8/12 + 16/24 = 4;

if x = 3 we get

1 + 21/12 + 99/24 - 27/12 + 81/24 = 8;

and so on.

If we apply this formula in the case x = 5, we get tbe answer

1 + 35/12 + 275/24 - 125/12 + 625/24 = 31.

It may be noted that by logarithms, as the example is often done, we
got not 31, but 32. Does this worry us? Not a bit! First of all, there
is not much difference between 31 and 32; the difference is only
32 - 31 = 1. Secondly, even if we are forced, in an unguarded
moment, to admit that 31 and 32 are not quite the same, there remains
the question: which answer is better anyhow,

1, 2, 4, 8, 16, 31, or
--
1, 2, 4, 8, 16, 32?
--

Which of these is really more logical, more true to a mathematical
way of thinking? Which really shows the greater pattern-recognition
facility? There can be no doubt that 31 is the better answer by a
long-to-middling chalk; and this for the simplest of reasons. It must
be admitted that if anyone should be so narrow-minded and curious
as to put down 32 as the number that follows 1, 2, 4, 8, 16, then he
had some reason for doing so. It is possible that he arrived at 32 in
any of several ways. He may have noticed immediately that the
numbers 1, 2, 4, 8, 16, if exchanged for the letters of the alphabet
that correspond to them, are just the initial letters of the words 'Alien
birds do have peculiar...'; and that these are the words of an old
Kentish proverb, the last word 'feathers' being omitted. Arranging
the alphabet in a circle with z next to a, and counting on past z to a,
which then becomes 27, we see that the letter f, with which the extra
word 'feathers' begins, gets the value 32, which is the right answer.
This is all well and good, and is a reasonable interpretation of the

95

problem showing remarkable skill at pattern-recognition. But one
can argue, of course, that the letter w, by its very name, is not really
a letter at all; it is merely a way of writing the letter u, or just perhaps
the letter v, when they are doubled. Hence this letter must be omitted
in the counting, and while 32 is a good approximation to the correct
answer, it is not as good as the true answer 31.

But it is also possible that the answer 32 was obtained in another
way. It could even have been obtained by the graphical method
explained earlier, in which by lucky chance someone thought of
taking the logarithm. This leads to the formula

a[n] = 2^n,

where of course the first number of the sequence is considered as the
zeroth or noughtth number of the sequence. On the other hand, our
own work, which was based on a system, gives

a[n] = 1 + (7/12)n + (11/24)n^2 + - (1/12)n^3 + (1/24)n^4.

Now both of these answers are backed up by perfectly good mathe-
matical formulae, and so both are perfectly logical. Of the two
answers, which is the better--both being correct? The answer is, of
course, that the second answer is to be preferred, because it is much
the simpler, and is easier to use, and is obtained by a more general
method.

On the other hand, some people may prefer the answer 32 and
the formula a[n] = 2^n, because here we have a homomorphism from
the additive monoid N[0] to the multiplicative monoid Z, or rather
End(Z). Since both answers are correct, it is a moot point whether
one of them can be preferred to the other. Perhaps it is best to leave
the question of preference to the metaphysicians. At least one can
say that the investigation of this example has led to one important
improvement in our point of view, if we had thought before that there
could be only one logical answer to the question. Following up this
suggestion of the muse, let us consider an arbitrary answer to the
problem

1, 2, 4, 8, 16, ____.

An arbitrary answer means any old answer at all. Think of a number.
It need not be a number between 1 and 100; it need not be less than
a million; it need not be short enough to write down, even if one is

96

allowed only one atom of any element at all for each letter. There is
no restriction on the number one is allowed to write down as the
answer; but to avoid unnecessary explanations we shall restrict it to
be an integer. Just write this number on the line provided. (This line
may be extended if it is not long enough; and if the number is too
long to actually write during the probable lifetime of the reader,
simply think of the number as written in.) Since I do not know what
number you have written in, let us call it a[5]. Since there is nothing
special about the number 5, let us assume that to an arbitrary finite
sequence

(a[0], a[1], ..., a[n-1])

the reader has added one more number, making it

(a[0], a[1], ..., a[n-1], a[n]).

Then the polynomial-generating formula that was mentioned earlier
will produce a formula that will not only interpret the given terms
a[0], a[1], ..., a[n-1], but will also include the new term added arbitrarily
by the reader in the interpretation. This gives us the clear and simple
rule that was sought, which would enable any fool to answer these
little puzzles with a minimum of difficulty. The rule is this: you
probably have a favourite number--most of us have one. If you do
not know what your favourite number is, try to find out or decide
somehow; perhaps an astrologer could help. Preferably, add 1 to this
number; but if you are unable to perform this calculation, it does
not matter greatly. Now take the number you have arrived at, and
whenever you are given one of these questions involving guessing
the next number in a sequence, use this number. (The addition of 1
to your favourite number is simply a device that makes it more
difficult to determine your character defects by analysing your
number. No technique by which a person's character may be found
out from his secret number is known to the author, but of course
someone may some day invent such a technique.) For example, here
is how you may answer the quiz we started with:

(1) 5, 4, 3, 2, 19.
--
(2) 2, 4, 6, 8, 19.
--
(3) 1, 3, 5, 7, 19.
--
(4) 1, 2, 4, 8, 16, 19.
--

97

There can be absolutely no question that the answer given here is
correct, and that it tends to show a very high level of understanding
of pattern-recognition. It must be admitted immediately that some
of the people who evaluate these tests are so poorly equipped,
patternwise, as to be unable to recognise the correctness of the
answer given above. Of course, if the object is to guess the number
chosen by the examiner as the one he intends to accept as correct,
then a study of parapsychology is more relevant than a mathematical
study of which answer is correct. Parapsychology is also called
E.S.P. or mind-reading, and is a much more useful art than is
mathematics. A good esper can pass an examination in any subject
at all, so long as the examiner can also pass his own exam. Still, even
E.S.P. is helpless if the examiner himself does not know the correct
answer.

The reader may be assured that in writing the following it is not
mere random whimsy that guides the author's unerring finger:

0, 0, 0, 0, 0, 0, 0, ....

There is something very definitely in the writer's mind when he sets
out that string of numbers. If this is regarded as to be completed by
the addition of one more number, then that number may as well be 0,
or 19. If the object, on the other hand, is to find out just what are the
numbers (all of them) in the infinite sequence that I happen to have
in mind, then even an accomplished mind-reader cannot tell you
very much about the numbers. He can tell you something: since
what I have in mind is that the term in the nth place is 0 if n = 0, 1,
or 3; and 0 if n >= 3 and if there exists no solution in integers to the
equation

a^n + b^n = c^n

such that abc <> 0; and 1 if n >= 3 and there does--since all that is
what I have in mind, he can tell you it. What he cannot tell you is
whether all the terms of the sequence are 0, or some of them are 1.
He cannot tell you that because the answer is unknown, not only to
him and to me but to everyone. He has no minds to read about the
answer in. That sort of question is a good one on any intelligence
test: namely, the sort of question that has never been answered, but
is known to have only one correct answer. At the opposite extreme
from this is the question that is known to have all possible answers

98

as correct answers; indeed it is strange that people not only get
great enjoyment out of answering such questions, but that only
intelligent people are able to answer them.




5. Fractions


Fractions are usually introduced to us in terms of pies. Suppose, for
instance, that seven greedy boys wish to eat three large apple pies--
it is only a supposition. One might try to solve the problem by the
methods of arithmetic in integers; this would not allow one to cut
the pies, a most inappropriate condition. It is very silly not to cut
pies; apple pies, especially, are eaten with a fork, and not with the
fingers, and are eaten more easily without being cut. If the problem
is to find out how to share the pies in case they may not be cut, it
ought really to be reformulated thus: Suppose that seven greedy
boys wish to eat three large oysters. The oyster, as everyone is well
aware, is never cut. (It is not true, by the way, that oysters may not
be chewed. They should be taken whole into the mouth, and one
must not bite off pieces; also, of course, one must never remove
partly chewed oysters from the mouth in order to inspect them.
Part of the fun of eating oysters is the tantalising thought of how
ghastly and sickening the insides must look, and once the insides
have been viewed all the spice goes out of the procedure for many
people. Then, too, if you look at the insides of your oyster it is just
possible that you will lose your nerve, and be unable to finish. That
would brand you as a coward and a chicken for life, and no woman
could love you after such craven behaviour.) In order to finish the
oysters without any of those inequities of distribution which the pure
mind of childhood finds so intolerable, the little lads realise the
necessity of a homomorphism Z -> Z sending 7 -> 3; putting, if you
will, the seven boys round the three oysters. Or else--since there are
seven of them, they may be excused for formulating the problem in
various forms at the start--one seeks an integer n such that under
the unique symmetrical bilinear map

Z X Z -> Z

99

that sends (1, 1) -> 1 one should have

(n, 7) -> 3

--or (7, n) -> 3, which is the same thing. These two suggestions
have been made already by the first two lads, when a third chimes in
to point out that they are of course equivalent. The equivalence is
soon proved by the fourth boy. The fifth then asks pensively, 'Then
in that case I wonder whether this proposition--for two equivalent
propositions may, I take it, sensibly be regarded as one only--
whether (I say) this proposition may be true; namely, that there
exists such a homomorphism (respectively, such an integer)?'

To the sixth little fellow (and we can hardly blame him) this
remark is hardly one that can be passed over unnoticed. 'You cannot
mean, Q.C.,' says he (the name of his playmate is Quintus Columus
MacIlhenny, his father having learned Latin while engaged in
agricultural pursuits at the State Prison Farm in Penalville, so out of
tender feelings his chums call him Q.C.), 'surely you are not thinking,
what I take to be an unwarranted generalisation of your conjecture;
namely, that whatever be the number of men, distinct, of course,
from 0, and whatever the supply of bivalves, they can eat them?
For in the case of two men required to consume a single oyster,
I perceive the impossibility of a solution--no doubt you have seen
already what I am thinking of.' This way of speaking--calling his
comrades men and not boys--was the custom in this group of
friends, which the speaker had agreed to follow afer quieting his
original misgivings with the thought that though they were not
actually, they were virtually grown men, just as in the view of the
ancient mathematicians the unit 1 was a number, not actually but by
virtue of being the generator of the natural numbers. Besides that,
he was aware that when they should have reached mature years, he
and his buddies would almost certainly make up for it by calling
themselves boys. In idle moments he would reflect that no doubt this
colloquial inversion had poetic truth as well, because of the line
'The child is father of the man', which one could not otherwise
justify.

Very much to their credit, the other little nippers owned up im-
mediately to their inability to follow him. They had not the slight-
est idea why two persons could not ingest a single mollusc without

100

the impropriety of dissection. Oh, of course they would not like to be
obliged to do it; they had some misgivings about the success of such
an experiment; if one were to place a bet, it would no doubt have to
go down on the side of failure. But that, they knew, proved nothing.

On the insistence of his pals, our sixth little chap unwillingly pro-
duced his proof. 'It is nothing, really; only don't you see that if a
homomorphism (which may as well be called n) sends 2 to 1, and if
2 is the homomorphism sending 1 to 2, then the composition, that is
2 followed by n, must be the identity? I know it may not be elegant,
but one sees that the restriction of n to the subgroup generated by
the integer 2 (which is after all nothing more than the image of the
other homomorphism) is already surjective onto Z. Clearly then n
itself cannot be injective; but we know that any non-zero homo-
morphism Z -> Z is injective (by consideration of the properties of
the natural numbers, or by some other means). I think that makes
it clear.'

Without any question, it was agreed that that did indeed make the
matter all too clear; the sunny faces of the little gang of companions
took on, as one, a hang-dog expression. Their brows furrowed.
Twenty-eight fingers were raised to scratch seven heads. One thought
was in seven minds--and how can that be, since nobody has ever
succeeded in slicing a thought so as to distribute it evenly among
seven? It was this: if it is not always possible to distribute a number
of oysters among a number of boys (for in their inmost thoughts the
boys recognised their juvenile status), then might it not also be
impossible in the very case they were now considering--with
gathering urgency as time drew on? One could not but admit that it
might be impossible; and then the question is, 'Ah! But we must
know for certain: is it?' For some time the company sat pensively,
seven hands under seven chins, staring at the oysters. The dust of the
country road settled in their hair, on their shoulders, on the open
shellfish before them. At length there was a small change in this
dismal scene; but it was not a happy one, nor one to be proud of.
Let us hope the passing travellers, if any came by at that moment,
were too preoccupied to notice. For if you had been present just
then, you would have seen a small tear that issued from the eye of
the seventh and last young hayseed. Only one--we must be swift to
point that out. It trickled down to the level of the lips, and had

101

almost mingled with the trail of saliva which there poured forth
when by an unconscious reaction the little pink tongue flicked out
and caught it gratefully.

The boy spoke with the calm of despair. 'Can't be done,' he said
manfully. 'You will all want to know the reason; I am sensible to
that, so here it is. I'll tell you what: let me put the argument to you
in the form of questions, and if you will, do you all answer me in
chorus.' On receiving their agreement to this mode of discourse,
number seven began his pitch. 'Tell me, is the number 3 prime, or
composite?'

The answer came in a sextet: 'Prime, though you ought really to
have asked, is it prime, composite, or a unit; those are the three
alternatives.'

'I stand corrected, fellows. You forgot to mention 0; that is a class
to itself. Still, the point is made. Very well then. In what way may
a prime number be the image under a homomorphism from the
integers to the integers of a number not a unit?'

'That can only happen in one way,' came the reply; there was no
polyphony, but simple unison, for these were simple country lads.
'The homomorphism must be an isomorphism.'

'You have struck the nail on the head with perfect orthogonality,
mates,' spoke the inventor of the demonstration. 'And which isomor-
phisms exist from the group of integers to itself?'

'There are two only, as generally recognised; that is to say,
negation, and the identity.'

'Right again!' cried the interlocutor, 'though we ought really to
have spoken of endomorphisms and automorphisms a moment ago;
nevertheless, the terms used will serve. Finally, you must tell me,
what are the images under both these isomorphisms, or automor-
phisms if you find the term preferable, of the number 7? Is either of
them 3?'

A gloomy tone of finality rang in the still, hot country air as the
answer came in chorus: 'Under the one, -7, and under the other, 7.
Neither of these numbers is 3.' Six more tears trickled down, and
were absorbed by six tongues. The seventh child made as if not to
notice.

The minds of the young are volatile, and not disposed to rest for
long on one thing. One of the boys remarked that they might have

102

noted, if their senses had not overpowered their other faculties (he
was referring to the sight of oysters, and to the Pavlovian reaction
of the saliva in flowing out of the corners of their mouths) that a very
similar occurrence was recorded in the Bible. The others recognised
immediately his reference to King David and the water which his men
procured for him at great danger through the enemy lines; because it
was generally thought that except for difficulties of division it might
have been shared out among the army, whereas in fact it had been
poured out on the ground. They said nothing of all this, but one
other of the lads remembered a bottle of root beer that had been
buried a year ago, and ought by now to be ready; and the clutch of
rattlesnake eggs that a third lad had been incubating in his pocket
was produced, and found to have begun to hatch through the good
offices of the sun, which had been beating down steadily on that area
of the boy's shirt during their deliberations. The oysters were added
to the writhing mass in the pocket, and the seven set off to procure a
spade, and soon mathematics was forgotten for a time.

Now if those boys had been of the bookish, studious kind, that
stays indoors all day in the summer to work out mathematics
problems, they might have applied the idea of division in the
Euclidean domain of integers, and come out with the fact that on
division by 7, the number 3 leaves quotient 0 and remainder 3; as it
was, being red-blooded fellows they solved this problem informally
and practically by eating no oysters each and feeding the remaining
three molluscs to the snakes. How many snakes there were does not
matter--if snakes eat oysters at that age, it is unlikely that they do so
with fastidious manners, insisting on equal apportionment and
swallowing whole. In any case the question did not occur to the
seven boys, and in their healthy outdoor way they could not care
less for such hair-splittings and refinements. They did remember,
however, to improve the snakes' appetite by the addition of a couple
of shakes of Tabasco Sauce, which Q.C. usually carried about with
him.

But apple pies--to return to the starting-point of the discussion--
are something quite different from oysters. Nobody loses his appetite
when he sees an apple pie cut--not if it is a good apple pie, not if the
physical properties of the crust and of the filling are what they should
be, not if the aroma emanating therefrom is the one we expect. No,

103

indeed. People do cut apple pies, and that is one of the reasons why
the integers alone have not been considered sufficient for all mathe-
matical purposes, even applied ones.



QUESTION 19. Whether there are fractions?

Objection 1. Clearly fractions involve the idea of breaking things
in pieces. Now it is clear that some things can be broken in pieces;
apple pies, or the hearts of unrequited lovers. But a mathematical
unit cannot be broken in pieces, since it is one and must remain so.
Hence fractions are mere physical things, without the ideal truth of
mathematics. Instead of fractions, the mathematician ought to be
talking about ratios. Ratios are relations between integers, and so
they have a mathematical existence.

Reply to Objection 1. This objection stems (aside from any
philosophical or metaphysical ideas that may lie behind it) from a
misconception of the role played in mathematics by the idea of 1. If
we think of 1 as something having an existence apart from the system
of integers itself, then we may assign to it absolutely certain qualities
that it has only by virtue of being one of the privileged class of
integers. Qua integer, the number 1 cannot be broken in pieces; qua
rational, it can. It is possible to think of rational numbers (or
fractions, as they are sometimes termed) as ratios, in a sense to be
made precise later. It is possible not to think of them thus.

Objection 2. A fraction is a number consisting of a top number
and a bottom number, which are called the numerator and denomi-
nator respectively. Now the number 0.111 is a fraction since it is big-
ger than 0 and less than 1. But it has no top and no bottom number,
and this is nonsense. The contradiction leads us to the conclusion
that fractions are absurd.

Reply to Objection 1. It is possible to devise an explanation that
will circumvent this apparent difficulty, and that is based on the
notion that a fraction has a top number and a bottom number, to use
the barbarous words of the objection. In fact, 0.111 is the same as
111/1000. The idea that fractions must be less than 1, or that numbers
less than 1 are fractions, is also mistaken. But the real error lies in
speaking of fractions as if they were things in themselves. We cannot

104

really speak of fractions at all; we can only speak of the field of
rational numbers. The field of rational numbers is sublime, like the
laeta arva, the delightsome, verdant fields of Elysium, the happy
hunting grounds, of which all men love to speak. The field of rationals
is a subject for poetry. Whoever attempts to speak of fractions is
bound to become entangled in barbarous expressions, to sink in a
miasmatic bog of barbarous inelegancies. Unclean! Unclean! Let
them cry as they walk through the streets, all those who mention the
unspeakable numerator and denominator.


I ANSWER THAT there are fractions, if you care to mention them;
and if you are sensible, you will not. What people usually are after
when they want fractions is something on the order of seven boys
eating three pies fairly. They want to have some kind of confidence
that it can be done, which has nothing at all to do with the stomachic
capacity of the lads; that we can take to be infinite. In general, one
would like to be able to divide. One would like to be able to say that
if n is an integer not 0, and if a is given, then a unique solution x
exists for the equation

nx = a.

To make everything as simple and neat as possible, one may also
wish to require that if x <> 0 and if nx = 0 then n = 0; and that
tells us that we are looking for an abelian group G such that

Z -> End(G)

(the natural map) shall be an injective monoid homomorphism from
the multiplicative monoid Z (without) {0}--a restriction is necessary to
eliminate 0, where before the restriction the natural map is of course
defined--to a submonoid of the multiplicative group

Aut(G)

of units in End(G).

Unfortunately, tbe problem as thus presented is too easy, since
there are non-isomorphic solutions. We are used to thinking of
fractions as well-determined entities, and this contains a small grain
of truth: we should hope to find that the group of fractions is well-
determined up to isomorphism. It is impossible to allow non-iso-
morphic objects to share the glory. There must therefore be found

105

some way of choosing among all the non-isomorphic entities one
that pleases us best.

It is already acknowledged that the sum of two integers is again an
integer, and hence is either a non-zero integer or 0. Since in any
group mn = nm -> m^(-1)n^(-1) = n^(-1)m^(-1), and since n(n^(-1)m) = m =
mnn^(-1) = n(mn^(-1)) likewise follows if m and n commute, giving
n^(-1)m = mn^(-1), it is easily seen that the subgroup of Aut(G) gener-
ated by the image of Z (without) {0} consists of all elements mn^(-1), with
m, n (epsilon) Im(Z (without) {0}), and is commutative. The equation

mn^(-1) + pq^(-1) = (mq + pn)n^(-1)q^(-1)

then shows that the sum of two automorphisms in this subgroup is
again such an automorphism, or is 0. Hence the subgroup is a field,
which we dignify with the name Q. It is easily verified that Q may
take the place of G, and that Q is a universally repelling object in the
category of fields of characteristic 0.

Hence the existence of fractions will be assured if only it can be
shown that a group satisfying the conditions imposed on G exists.
Such a group is the following: Let R be the ring

Z[X]
-----
(X^2)

of polynomials in a single indeterminate X over the ring of integers,
modulo the ideal generated by X^2. Let M be the monoid obtained
from the multiplicative monoid of this ring by excluding all elements
a + bx such that a = 0, where x = X + (X^2). Now if C is the
submonoid of 'constants', i.e. elements a + bx with b = 0, then
the relation

________
(a + bx)(c + dx) (epsilon) C

is an equivalence, where

______
c + dx = c - dx.

Let G be the quotient; i.e., the set of equivalence classes under this
equivalence. Because the equivalence is compatible with the monoid
structure, G is clearly a monoid with the structure inherited from M.

106

_
Since mm (epsilon) C for every element m of M, the monoid G is clearly a
group, and trivially abelian. Since

(a + bx)^n (defined as =) a + nbx

in M, the conditions which G is to satisfy are easily verified. Thus
fractions exist.

In the example given there were three pies and seven boys. Since
we have the isomorphism of Aut(Q) and Q together with the injec-
tion Z -> Q--which we shall always interpret as an inclusion--we
see that all we need to do is to take the field elements 7 and 3 and
form 3'7^(-1). Because fields are commutative this may be written as
3/7 without much danger, and this is what is often done. Thus we see
that some fractions have a numerator and a denominator.



QUESTION 20. Are whole numbers fractions?

I ANSWER THAT whole numbers are fractions, or rather that integers
are rational numbers. There is a curious distinction sometimes made
in arithmetic books between whole numbers and fractions, or between
proper and improper fractions, usually in connection with a notion
of something called 'mixed numbers'. Usually, these correspond to
various kinds of rational numbers written in various ways. Negative
rationals are frequently ignored or treated as curiosities. The whole
point about integers being rational numbers is that Z is a subring of
the field Q, in the sense that the natural injection is considered as an
inclusion.



QUESTION 21. Whether a fraction has a numerator and a deno-
minator?

I ANSWER THAT in a sense, a rational number has a numerator and a
denominator, and in a sense this is not so. There is no way of taking
a rational number and getting out of it a numerator and a denomina-
tor, unless one is willing to accept several numerators and several
denominators, or unless one is willing to study the subject known as

107

reduction to lowest terms. But on the other hand, if one is given the
numerator and denominator to begin with, and if the denominator
is not 0, then it is possible to get a rational number and only one out
of the numerator and the denominator. In civilised terminology,
one would distinguish between fractions on the one hand and
rational numbers on the other. A fraction is merely an element of
Z X (Z (without) {0}), whereas a rational number is an element of Q.

To find the rational number associated with a numerator N and a
denominator D, one simply maps (N, D) to Q, considered as End(Q),
by taking the product in the latter of the endomorphism associated
with the integer N and the inverse of the automorphism associated
with the integer D. By previous remarks and exercises, the resulting
map from fractions to rationals

Z (X Z (without) {0}) -> Q

is surjective. The fraction (N, D) is commonly written as N/D, and
this same symbol is often used for the rational number obtained
from the fraction. As fractions, 3/4 and 27/36 are distinct, or

3/4 <> 27/36;

whereas

3/4 = 27/36

if they are rational numbers. Fractions are of course rather ridiculous
and pointless objects, and both the idea and the word are best
forgotten. Rational numbers are lovely, civilised and useful things.
That may be why they are called rational numbers.



6. Calculations with fractions

(1) How to find a numerator and denominator for a rational number

We are aware that if the rational number is not 0, then it may be
considered in a natural way as an automorphism of the Z-module of
rational numbers. We also know that this group of automorphisms
is commutative, and is generated by the submonoid of (automor-
phisms associated in the natural way with elements of) Z (without) {0}.
Hence it can be written in the form ND^(-1), so that it is the rational

108

number associated with the fraction N/D by the civilising process
described above. If the rational number is 0, then it is also 0D^(-1) and
hence comes from numerator 0 and denominator D, where D can be
any integer except 0. Then N and D are a numerator for the rational
number.


(2) How to reduce a fraction to lowest terms

The fraction is barbarous, so first we must civilise it by the map

Z X (Z (without) {0}) -> Q.

Once a civilised rational number is available, it is possible to get to
work and do something. Let the rational number be q. As we know
the set D of integers d such that

dq (epsilon) Z

is not empty; trivial examination shows that D is an ideal. Since
D <> {0}, the ideal has a non-zero generator D. Taking

Dq = N,

we have found a particular numerator and denominator for the
rational number q. The fraction N/D is said to be reduced to lowest
terms. So is the fraction -N/-D, and these are the only fractions
reduced to lowest terms that give this rational number, as is obvious
from the uniqueness to within a unit of the single generator of an
ideal in a ring without non-trivial divisors of zero.

Consider the reduction to lowest terms of the fraction 6/8. If q is
the rational associated with this fraction, then it is easily seen that
dq (epsilon) Z <-> 3d (epsilon) 4Z. The proposition on the right-hand side
of this equivalence implies that 9d (epsilon) 4Z, and since in any case
8d = 4.2d (epsilon) 4Z we get by subtraction that a further consequence of
this is d = 9d - 8d (epsilon) 4Z. This last trivially implies that 3d is
again an element of 4Z, so that 3d (epsilon) 4Z <-> d (epsilon) 4Z. The ideal
D in question is therefore generated
by either 4 or -4, whichever one cares to take, and since (taking, say,
D = 4) we get 4q = 3, we come on the result that 6/8 reduced to
lowest terms is 3/4. The answer -3/-4 is just as good.


(3) How to deal with mixed numbers

A mixed number is a pair, consisting of an integer and a fraction
N/D in which one has 0 <= N <= D. The sum of the integer and the

109

rational associated with the fraction is another rational, called the
rational associated with the mixed number. The customary mode
of writing mixed numbers is most curious, and is also entertaining so
long as one is not called upon to use it. It is a sort of triangular
array, and involves three integers. There would be nothing odd about
this if these were always the first integer of the mixed number, and
the numerator and denominator of the fraction that forms the second
term in the ordered pair that is the mixed number. Those are the
three integers that one sensibly expects in writing the mixed number,
and they are sometimes the integers used. In case they are the integers
used, the mixed number (I, N/D) is simply written

N
I -.
D

But if I is 0, it is omitted, leaving N/D. If N is 0, both N and D are
omitted, leaving just I. These exceptions are rather simple. The
general exception in case I < 0 is that one must write

D - N
(I + 1)-----;
D

but to this exception there are numerous exceptions. If I + 1 is
0, it is replaced by a mere minus sign in this expression. If N = 0,
one ignores this exception and uses only the exception about what
to do if N is 0. The subject of mixed numbers is more nearly a branch
of generative grammar than of arithmetic or algebra.


(4) How to deal with improper fractions

A fraction N/D is proper if 0 <= N < D; otherwise it is improper.
The thing to remember about improper fractions is that no lady or
self-respecting working girl ever allowed an improper fraction to pass
her lips. Note that the fractional part of a mixed number is proper.
It is possible to express any rational as a mixed number, and the
availability of this alternative to the use of improper expressions of
rationals is a blessing to all right-thinking young ladies. It is quite
probable that the wish to avoid the improper has led to the invention
of mixed numbers; these are hard to account for otherwise. The
mixed number (I, N/D) represents the same rational as the fraction
(ID + N)/D, and by carefully observing the rules of the cryptogram,

110

one can write an improper fraction as a mixed number, or a mixed
number as a fraction.


EXERCISE. Write as fractions:

-7 17/52; -3/4;

write as mixed numbers:

54/-3; -9/-2.


(5) How to add fractions

The sum of two fractions is of course not well defined. Any fraction
or mixed number can be considered as the sum of two fractions or
mixed numbers if it represents the sum of the two rationals that are
represented by the two fractions or mixed numbers. Definite rules
for computing the sum of two rationals by representing them and
obtaining a fraction or mixed number representing the sum from the
two representations have, nevertheless, been devised. In ordinary
practice, despite their totally impractical mode of writing, mixed
numbers are more commonly used in such computations than frac-
tions. In order to explain how this is done, it is necessary to use a
special symbol (psi) to stand for any of I, N, or D that is not really
there; for example, we must take the mixed number 2 as 2 (psi)/(psi), and we
must take 3/4 as (psi) 3/4. If we wish to add

N N'
I- + I'--
D D'

there are several cases. If both I and I' are positive integers or (psi),
then the result is either

ND' + N'D
(I + I')---------
DD'

or else

ND' + N'D - DD'
(I + I' + 1)---------------
DD'

with certain remarks and provisos.

111

If in a mixed number both I and N are 0, then one of the exceptions
to the exceptions in the rule for writing mixed numbers which is
outlined above is that instead of writing nothing at all, which would
be (psi) (psi)/(psi) in our present temporary notation, one actually writes 0,
or in present notation 0 (psi)/(psi). This is the only case when the place of
I is filled by the integer 0 in writing mixed numbers. In computing I + I'
or any of the other sums and products, certain special rules must be
observed if any of I, I', N, N' is (psi).

(i) In addition, (psi) acts like 0; thus, (psi) + 3 = 3.

(ii) In multiplication, (psi)D = (psi) and N(psi) = N, which we may
express by saying that (psi) is absorbing on the left and neutral on the
right, like British toilet paper.

(iii) It is sometimes necessary to use D'D as denominator instead,
as in the case 0 + 3/4.

(iv) 0 + (psi) = (psi) + 0 = (psi).

In case one only of I, I' is a negative integer or is -, any of the
following forms may be the correct answer: (say I' < 0 or I' = -)

ND' - N'D
(I - I')---------
DD'

DD' - ND' + N'D
or (I - I' + 1)---------------
DD'

DD' + ND' - N'D
or (I - I' - 1)---------------
DD'

-ND' + N'D
or (I - I')----------
DD'

There are of course certain remarks and provisos; however, we only
touch on the subject here, and for a full treatment the reader is
referred to any school arithmetic text, where the case when both I
and I' are negative numbers or - may also be omitted, as it is here.

112

Fractions are also sometimes added as fractions, especially in the
case of proper ones. This is much easier to do. When fractions have
been added it is considered the done thing to put them in lowest
terms.

(6) How to multiply fractions

Since rational numbers are the elements of a ring of endomorphisms,
they may be multiplied by composition. It is this which shows us the
kind of application to everyday life the multiplication of fractions is
likely to have. The following is a very practical and concrete example;
like many problems in the elementary theory of fractions, it involves
the idea of sharing.

The headmaster of a provincial boys' school believes strongly in
the great value of practical experience in forming boys' minds. It is
this that has made up his mind to take the whole form to the zoologi-
cal gardens of a great metropolitan city. 'How else can they pick up
the rudiments of zoological gardening?' he asks himself--the ques-
tion is merely rhetorical and does not require an answer, but to
show his agreement with the wit and wisdom of the inquirer, the
headmaster chuckles softly to himself, and sighs. Alas! Little does he
know what lies in store. Because the school has been founded on
principles of freedom from fidgety nonsense, no books are kept and
it is unknown just what is the size of the form. When people ask how
many boys are in the various forms, Dr Chockle-Fervyn answers
mildly, 'Oh, various numbers--I think.' For the problem that presents
itself in a few moments, the fact that the number of boys is unknown
takes on great importance. The boys go to the zoo; they see animals;
they grow hungry. The tables in the cafeteria seat twelve. For each
tableful of boys, the Doctor brings 45 steak pies; not that he has been
so fussy as to ask for that number--he merely says to the attendant,
'A carton or so of hot steak pies, perhaps, please,' each time he sees
a table of unfed boys. (Afterwards, enquiries in the packing depart-
ment of Hengist & Horsa Olde Englishe Meat Products produced
the number 45. 'We always puts 45, your honour, sir. The sum of the
hexpounents on the proims bein' free, you see. Free dimensions, loik
spice. Glad to 'elp your worship.') When the boys have finished, it is
time to feed the lions. This is done, in the metropolitan zoo to which
the problem has reference, by opening the gates between the lion

113

house and the cafeteria; the fact being known in the metropolitan
area, only tourists are eaten on most days. Today, the only tourists
are the boys from Fervyn Towers of Learning. Old 'Chock-full-of-
vermin' is in the loo. The lions lick their whiskers and prowl smugly
back to the lion house. 'Can't say exactly,' was the headmaster's
reply to anxious parental enquiries after the hols commenced--and
those words, at least, were spoken true. 'Some of them may have got
lost at the "zoo", of course. An interesting question. Are you sure
you had a boy here?'

It must not be thought, however, that vagueness went with want of
curiosity or with pedagogical apathy in Dr C.-F.--far from it. He
went so far as to prevail on the head gardener of the zoological
garden to have the stomachs of 25 of the lions pumped, and the
contents sorted. 'Unheard of, sir,' was the initial reaction of this
official; but the good Doctor plied him with educational anecdotes,
and reminded him that the schools could not be expected to produce
the required quota of zoological gardeners unless the gardeners took
an interest in education. The curator (as he was also called) mollified.
'Better see to it myself, you know. Can't trust these young fellows
these days. Would the boys benefit by a personal report from myself?'
It was agreed that the results would be presented at assembly in the
Great Hall down at Pummidge in the near future. Briefly, it can be
stated that every 25 lions managed 18 boys among them.

'And now, lads,' said the Doctor after the vote of thanks, 'I have
a practical exercise for you in the multiplication of fractions. How
many steak pies were eaten by each lion? You know the facts, and you
should be able to visualise the problem. Think of pies inside boys,
and boys inside lions. For artistic expression, there is a prize for the
most vivid painting of the meal. Those doing military science will
show the optimum formations for the armies involved: lion, boy,
and pie. For social science....' Only the arithmetic problem need
detain us here. Let m be the rational number corresponding to the
meal in which the boys eat and the pies are eaten. Thinking of m as an
endomorphism of Q, we have m: 45 -> 12. If M is the endomorphism
whereby lions eat and boys are eaten, then M: 18 -> 25. Since these
are Z-module endomorphisms we may also write

18 X 45 ->[m] 18 X 12 ->[M] 25 X 12,

114

so that the composition Mm sends 810 -> 300. But the rational
represented by the fraction 10/27 sends 810 -> 30 X 27 to 300 =
30 X 10. If two endomorphisms of Q agree at a non-zero element of
Q, they agree on all of Q, since this amounts to stating that cancella-
tion holds in Q, and since Q is a field. Thus Mm sends 27 to 10, or
Mm = 10/27, or, as we easily see, 27/10 pies went into each lion.
In mixed numbers this is 2 7/10.



(7) How to divide fractions

Since every non-zero element of a field has an inverse, it is easy to
divide rational numbers. In order to divide one fraction by another,
turn the fractions into rational numbers and divide the rationals.
Then turn the answer back into any of the fractions that represent it.
Division of fractions is of use in those places where cannibalism
flourishes on a highly developed cultural level. This is because a good
side of porkh (pronounced porch) has got a history attached to it,
relating to the quantity of porkh and hamh (pronounced haunch)
eaten by the side of porkh in its earthly life. In other words, if you go
to a really good family butcher and if you are particular, you will
want to know how many men (for such purposes one does not call
them men but pork pighs, on the principle that man is long pig or
pigh, and his flesh is porkh or hamh)--how many men (to use that
expression) he himself has eaten. That is because the meat of cannibals
tastes better than the meat of heterophagi, or non-cannibals. A living
cannibal is never called a pork pigh in polite society. Here is a prob-
lem in division of fractions: if 7/5 pork pighs were eaten by 3/4 pork
pigh, how many pork pighs did the whole pork pigh eat? (Some
people, especially waiters in fancy restaurants, say 'hyperbolic ham'
instead of 'haunch'. Etymologically this is correct, since 'hyper-
bolic' means 'past all likelihood of truth'--it is a roundabout way of
hinting that it is not really just ordinary ham. The same is true of the
hyperbolic functions sinh and cosh, sometimes pronounced shin and
cosh; the pronunciation reveals the cannibalistic origins of these
functions; the cosh being a weapon used in preparing a market-ready
animal for slaughter, and the significance of the shin being clear
enough.)

115

In order to divide by 3/4, it is sufficient to invert and multiply. How
to invert 3/4? Since 3/4 is an automorphism sending 4 -> 3, its inverse
sends 3 -> 4; and hence is 4/3.


















116

3

ALGEBRA


1. The wonderful quadratic formula

The quadratic formula is an amazing discovery of the utmost utility
in solving the most multifarious practical problems, such as abound
in the best texts of school algebra. One example will suffice to illus-
trate its depth and range of application.

Muscular M. Boulangiaire, the baker of the picturesque village of
Beaulieu Derrie`re, makes three kinds of loaves in his shop (Figure
3). The first kind is a flat square of side x; the second is an ordinary
French loaf just x units long and one unit wide; and the third loaf--
it is really more a bun than a loaf--again a square one like the first,
but measuring one unit along each side. All loaves have exactly the
same height.

M. Boulangiaire knows just how much dough goes into the one-by-
one loaf, or bun. He also knows, of course, just how much dough he
has prepared on any particular day--in fact, this never varies; there
is always just enough dough for 113'006 buns. What does vary is the
number x, which tells how long the French loaves are and determines
the size of the x-by-x square. The baker's method of work is the
following: By enquiries among his customers, he determines what
numbers of the several loaves will be required for the morrow. Let
us suppose that today he has learned that tomorrow's demand will
be for seventeen variable-sided square loaves, for 5'310 ordinary

117



/---------------------\ /|\ /|\ /------\ /-------\ /|\
/| | | | | | /| | 1
| | /----\ | | | | | | | | |
| | | [] | | | | | | | \-------/ \|/
| | |---< | x x | | |/ /
| | | [] | | | | | | \--------
| | \----/ | | | | | <---1---->
| | | | | | |
| \---------------------/ \|/ \|/ |------|
|/ / | |
\---------------------- \------/
<----------x----------->


Figure 3. The three styles of loaf at Beaulieu Derrie`re.



loaves, and for half-a-dozen buns. Then he adjusts the variable x
accordingly, so as to make the dough come out just right, with none
left over. M. Boulangiaire learned mathematics at the E'cole
Paranormale, and is able to compute just how long to make his
loaves when he bakes tomorrow's bread for the village.

Can you help him? (Remember that, being French, he writes the
number 317,476'52 as 317.476,52.)

It is well known among French provincial bakers with a suitable
mathematics education that a quadratic equation has two solutions,
or roots; and that these solutions may be the same; they may one or
both be negative; and they may both be complex numbers. It is not
the custom to measure loaves of bread in complex numbers, nor even
in negative ones. This has on rare occasions caused difficulty at
M. Boulangiaire's shop. Indeed, one August when all the village was
away on vacation, Beaulieu suffered a mammoth influx of Crimean
Goths (Figure 4). These folk speak a dialect descended from that of
bishop Ulfilas, translator of the Bible into Gothic. Like Ulfilas they

118

/----------------
H ------> | /----------------
| | /----------------\
| | | |
F ------> | | | |
| | | |
| | | > < |
/-----/ |\
| /| - ====` '==== ||
| || - ---' `--- ||
| \| - | ||
\--/ - / \ |/
- /-\_/-\ -|
- - |
- ' ' - |
//-----=======--------\
| | /-\ |
| | /-|-|-\ |
|-| |-|


Figure 4. Crimean Goth eating bread. Note the straight, fair hair, H, and the
typically Arian tetragonocephalic forehead, F.


are Arian heretics and believe it a sin to eat any but square bread.
This is because Ulfilas' (lost) translation of Genesis has the error 'In
the shape (rather than 'sweat') of thy face shalt thou eat thy bread'.
The demand for the one-by-one buns alone was in fact more than the
dough on hand could supply, which made the baker a little pessi-
mistic, but he went ahead and worked out the quadratic equation.
The Goths moved on, leaving the good baker a sadder but wiser
man. The ensuing efforts of comparative philologists to trace the
further course of the linguistic group have so far proved fruitless.
Show that the quadratic equation had two purely imaginary roots
(at least one Goth ordered a variable-sided square loaf). Explain the
meaning of this answer in practical terms so as to be readily under-
stood by (i) a provincial baker with some school maths; (ii) a Crimean
Goth.

Our M. Boulangiaire is a rather indecisive man when he is deprived
of the moral support of algebra. Especially since the August catas-
trophe he has been bothered by the recurring thought that one day

119

in his daily algebra he might find two distinct, positive roots for the
bread equation. How would one choose in such a case? As far as he
can see, nothing in the mathematics would give one the slightest clue
about which of the two solutions to use. One could take either value
for x, and either way the dough would exactly suffice to fill all the
orders. What a terrible responsibility, to make the choice all unaided!
Can you put M. Boulangiaire's mind at ease?

The quadratic formula says that to find the roots of

ax^2 + bx + c = 0

you must take the square root of b^2 - 4ac; then the roots are

-b + (root)(b^2 - 4ac)
----------------------
2a

and

-b - (root)(b^2 - 4ac)
----------------------
2a

The formula involves no difficulty in itself so long as a <> 0. If
a = 0 then the formula is, of course, nonsense, since 0 is absorbing
in any ring and hence if 0^(-1) existed we should have 0 = 00^(-1) = 1.
By definition, (root)(b^2 - 4ac) is a number such that

{(root)(b^2 - 4ac)}^2 = b^2 - 4ac,

so that a trivial verification yields

-b + (root)(b^2 - 4ac) -b + (root)(b^2 - 4ac)
a(----------------------)^2 + b(----------------------) + c = 0.
2a 2a

-b - (root)(b^2 - 4ac)
The same thing works for ----------------------.
2a

Note that these formulas involve the use of the operation (root), known
as the square root operation.



QUESTION 22. Whether it is possible to take square roots of rational
numbers?

Objection 1. Patently, it is impossible to take square roots of
rational numbers, unless one is very lucky in the rational numbers he

120

uses. The square root of 4 certainly exists, and is 2, since 2 X 2 = 4.
The square root of 7 74/2601 is obviously 2 13/19. Let q be a rational number
such that q^2 = 2. Then writing q in lowest terms, we get the fraction
m/n, with n the generator of the ideal of integers consisting of those
integers x that make xq an integer. It is clear that m and n are relatively
prime, and that 2n^2 = q^2n^2 = m^2. The prime 2 must divide m^2;
hence 2 divides m and 2^2 divides m^2; hence 2 divides n. Since 2
divides m and n they are not relatively prime, and a contradiction
results.

Reply to Objection 1. The argument is correct, but all it shows is
that rational numbers may not have a rational square root. They may
still have some other kind of square root. In fact, all that is necessary
is to construct the ring of polynomials over Q in a single indeter-
minate, which we write Q[X], and to take the quotient ring over the
ideal generated by the polynomial X^2 - 2. If x is the coset of X, then
x^2 = 2 in this ring. Moreover no multiple of X^2 - 2 by a poly-
nomial is a rational number, so Q is injected into this quotient ring
in a natural way, and the injection may be thought of as an inclusion.
If we replace 2 by b^2 - 4ac, we get a ring in which the formula for
the roots of the quadratic equation makes sense.

Objection 2. The bakers may not think so. Going back to (root)2, and
this could easily be the answer given by the quadratic formula to a
problem involving a quadratic equation, can you justify this number?
If it is not rational, in what way is it a number at all? Can it be the
length of a loaf of bread? If it is an element of some algebraic
extension of the rational field, it would seem that it could not be a
length.

Reply to Objection 2. The geometric answer might be this: from
the vertex of an isosceles right triangle drop a perpendicular to the
hypotenuse as shown. The two smaller triangles are also isosceles,

/|\
/\|/\
/ | \
1 / | \
/ | \
/ | \
/ | \
/ | \
-----------------

121

and if the side of the larger triangle is 1 then it follows that the
hypotenuse is (root)2 by the fact that the perpendicular bisects the
hypotenuse and by similar triangles. But this argument is invalid,
since it is customary to establish geometry on the basis of an ordered
field and not vice versa. Let us try harder.

Clearly every element of the quotient ring Q[X]/(X^2 - 2) can be
written as a + bx, where x is the square root of 2, and where a and b
are rational. Suppose there exists a positive integer n and rational
numbers a[1], a[2], ..., a[n] and b[1], b[2], ..., b[n] such that

n
-1 = (sum) (a[i] + b[i]x)^2.
i = 1

Then since x is not rational,

n n
-1 = (sum) a[i]^2 + 2 (sum) b[i]^2.
i = 1 i = 1

Since squares are positive and since 2 = 1 + 1, this is impossible.

Moreover, Q[X]/(X^2 - 2) is a field since

(a + bx)^(-1) = (a - bx)(a^2 - 2b^2)^(-1)

if a and b are not both 0. It is well known that every real field has a
real closure; i.e., is a subfield of a real closed field algebraic over the
first field--this is an easy consequence of Zorn's Lemma. But a real
closed field has the property that a sum of squares has a square root
in the same field; and hence has a natural ordering. This is inherited
by Q[X]/(X^2 - 2). Since the two possible roots of 2 have opposite
sign, we may choose x to be the positive one. Then either x < 1 or
2 = x^2 >= x; hence, x < 3. It follows that the field in question is
archimedean. Any positive element of an archimedean ordered
field can obviously be considered as a length. Hence, (root)2 is a length.

Objection 3. No doubt, it is true that (root)2 is a length, and any
baker worth his salt would accept it as such after such a clear and
beautiful explanation. The argument involved one little trick, how-
ever, namely the fact that 2 = 1 + 1, and hence 2 is a sum of squares.
But -1 is not a sum of squares. Moreover, any field containing a
square root of -1 cannot possibly be ordered, since in an ordered
field squares are positive, whereas -1 is negative. Therefore, a
square root of -1 makes no sense.

122

Reply to Objection 3. It is quite true that no ordered field can
contain a square root of -1. Hence, since (root)(-1) cannot be a length
or the negative of a length, it is of no use to bakers. Mouldy wheat
is also of no use to bakers; nevertheless, mouldy wheat contains
ergot, which is of use to pharmacologists. [32] Moreover, mouldy
wheat exists. Therefore, it is possible that (root)(-1) may exist.



I ANSWER THAT square roots of rational numbers exist. If a rational
number q has no rational square root, then X^2 - q is irreducible
over Q; hence if b is a non-zero rational and if a is rational then
rational numbers c, d and a non-zero rational r exist such that

(X^2 - q) = (a + bX)(c + dX) + r,

which makes the quotient ring Q[X]/(X^2 - q) a field. If q is positive,
then this field can be ordered, since if q is positive then

mn
q = mn^(-1) = (mn)n^(-2) = (sum) (n^(-1))^2,
i = 1

which is a sum of squares. Otherwise, this field cannot be ordered. If
finding the roots of a quadratic equation involves taking a non-
ordered extension of the rationals, then the roots cannot be con-
sidered as lengths, nor can they be used to weigh bread-dough. The
reason is that a practical man is sufficiently imprecise to accept an
approximation of the desired measure between certain upper and
lower bounds. He knows that (root)2 lies between 1'4 and 1'5, or
between 1'4142 and 1'4143. He can only know that if the field is an
ordered field, in which 'between' makes sense. That is why bakers
never try to bake loaves of length (root)(-1), even approximately. The
question, 'Have I shaped this loaf a trifle too long or a smidgin too
short?' would be meaningless. In practice, bakers seldom feel
obliged to produce loaves of length (root)(-1), and this helps to keep them
in a happy, carefree frame of mind.

123

How to find the square root of a number

It must be assumed that the number is positive and rational. Other-
wise, it is hopeless to try to approximate its square root by rationals.


Method 1. Use the fact that

(root)q = exp((1/2)log q).

Writing q = 1 + p, we get

(1/2)log q = (1/2)(p - (1/2)p^2 + (1/3)p^3 - (1/4)p^4 + ...).

This is not valid if q >= 2. Take sufficiently many terms to make a
good approximation to (1/2)log q. Next, we have

(root)q = 1 + (1/2)log q + (1/8)(log q)^2 + (1/48)(log q)^3 + ....

Again, it is necessary to take several terms. As an example, this
method gives 1'098 as the square root of 1'21. This is in error by
0'002, since the square root is 1'1. Here is the work:

log 1'2l = 0'21 - (1/2)(0'21)^2 + (1/3)(0'21)^3 - (1/4)(0'21)^4 approximately,

and this is 0'21 - 0'02205 + 0'003087 - 0'0004862025, or

0'213087 - 0'0225362025 = 0'1905507975.

Half this number is 0'09527539875. It remains only to take the square
and cube of this last number, and to add

1 + 0'09527539875 + (1/2)(0'09527539875)^2 + (1/6)(0'09527539875)^3,

which will give the answer stated.


Method 2. Using a table of squares. To find the square root of q,
write

x^2 = q

and find the greatest integer n such that n^2 <= q. Put x = n + x', and
x- = 1/x'. Clearly, if x' = 0 then the square root is n and we are done.
If not then we have x- >= 1, and x- satisfies the new quadratic equation

(q - n^2)(x-)^2 - 2n(x-) = 1.

Now we may seek the greatest integer such that the left-hand side is
less than or equal to 1; call this integer n-, and write

x- = n- + 1/<2>x-.

124

Iterating, one gets a sequence of integers

(n, n-, <2>n-, ...).

From the sequence of fractions

(n, n + 1/n-, n + 1/(n- + 1/<2>n-), ...)

These fractions approximate the desired square root. As an example,
let us compute (root)3.

First write x^2 = 3. Since 1^2 = 1 < 3 but 2^2 = 4 > 3, we take n = 1.
Now (1 + 1/y)^2 = 3 gives 2y(y - 1) = 1.

Since 2.1(1 - 1) = 0 < 1 and 2.2(2 - 1) = 4 > 1, we put n- = 1
and write y = 1 + 1/z. This gives z(z - 2) = 2 and <2>n- = 2.
Putting z = 2 + 1/a, we get 2a(a - 1) = 1. It follows that y = a,
and

n- = <3>n-.

Thus the sequence of integers is

(1, 1, 2, 1, 2, 1, 2, ...).

This is periodic after a certain point, or in other words the 'hat'
operation '- obeys

'-<3> = '-.

As a matter of fact this will always happen, in the sense that there
exist a non-negative integer p and a positive q such that

'- = '-

.

A convenient way of writing the sequence of fractions that results is

1 + 1/(1 + 1/(2 + 1/(1 + 1/(2 + ...,

or

/-\ /-\
| | | |
| | | |
| | | |
| | | |
| | \-/
| | -------------------
| | /-\
| | | | 1
| | | | ---------
| | | | 1
| | | | 2 -----
\-/ \-/ 1 1/2...



125

These are called continued fractions. The most useful thing to do is
to start with the iteration of

y = 1 + 1/(2 + 1/y).

Taking y = 1 as a beginning, we get successively y = 1 1/3, y = 1 4/11,
y = 1 15/41, ..., and the corresponding values of x are x = 1 3/4, 1 11/15,
1 41/56, .... Note that

(1 41/56)^2 = 3 1/9409.

The sequence for (root)2 + 1 is (2, 2, 2, ...); for (pi) it is

(3, 7, 15, 1, 292, 1, 1, 1, 2, ...).

This explains why it is often stated that (pi) = 3 1/7.





2. The incommensurable of the incommensurable


Let us begin to consider an event which may or may not have hap-
pened to the reader in the course of his life to date, in a semblance
more or less as is to be described here; and one which may--or
possibly may not--befall him in the future. Most of us have a sort of
routine by which we trace out the day. A man will go to his daily
place of occupation, where the surroundings change little, or change
in an expected way; where he finds tbe people seem to understand
what he says with a fair degree of accuracy, and to show little sur-
prise at the man's own commonplace gestures and remarks. The sort
of thing that has just been described is not unusual. It may not hap-
pen all the time, but it happens at least a good part of the time; the
number of days that vary from the pattern are few, if any, and they
do not vary much from the pattern. The psychological effect of this
monotony is to produce a commonsense outlook, a reasonable,
matter-of-fact way of looking at the world.

Then WHAMMO! The illusion is suddenly shattered. You are creep-
ing, let us say, down the queue in the refectory at lunch-time, just as
you do every day. Nothing at all out of the ordinary has happened.

126

The level of hunger is normal for such a situation. There is a choice
of meat pie or spaghetti; you must speak.

'Meat pie, please.'

'Sorry, sir, no cheese today. Only meat pie or spaghetti.'

'I'll have meat pie, please.'

'I told you, sir. You'll have to take what we've got.'

'All right.' By now you are a wee bit puzzled, and mere acqui-
escence seems an easy way out of the difficulty.

'We only feed that to the rats, sir. If you'd like to come around the
back I'll see what I can do.' The assistant is plainly puzzled by your
obstinacy, but means to be cooperative. She leads the way behind the
counters, the ovens, into a corner, through a door, into a broom
closet. The door of the closet is shut on you. It is dark. A muffled
voice is heard outside.

'The spiders don't often come before teatime, sir. Shall I slide a
few lollies under the door? A nice lolly will help to keep off the peck-
ish feeling, you know.'

'No! Why am I in here? Let me out!'

'Oh, well. I'll have to ask the butcher. Are you sure you can't
wait? The knives are all dull, and there's still a lot of meat in the
freezer. Most of the customers prefer spiders if the knives are dull.'

'Help! Police!'

The door opens, and you are on a vast plain of tall grass. Nothing
is visible except a caravan of Tartars approaching in the distance. As
their voices grow nearer and nearer, the words of their chant become
distinctly audible:

'Meat pie! Meat pie! Meat pie! ...'

Something like this happens if you construct an isosceles right
triangle. There is a procession from rationality to irrationality. There
is an encounter, where by an encounter we mean something different
from a lunch-counter. There is a meeting-up with something that,
while perhaps not Wholly Other, is rather Other. There is the feeling
of the Absurd; there is the wish to cry, Help! Police! The encounter
may come when two circles meet. It may come in a variety of ways.
When it comes, we cross over from rationality to irrationality, and
enter a new domain of incommensurability.

The word 'surd' is popularly used to designate irrationality. This
usage has fallen out of favour with mathematicians, but is still found

127

among certain classes of pedagogues and is used in school text-
books. The basic meaning of the word is 'deaf', but it may mean
irrational or unresponsive. Some early etymologists derived the word
from 'sordid', probably incorrectly, because those whose ears are
dirty or sordid are unable to hear clearly. A surd is defined as

(1) any irrational solution of an equation

x^n = q,

where q is rational; or else

(2) any number obtained by a succession of rational operations
and extractions of roots, beginning with rational numbers; or

(3) any algebraic irrational.

Since surds were discussed before cases (2) and (3) were known to
be distinct, and before the question whether they might be distinct
had been much considered, it is not always clear which kind of surd
is meant. Here, as in many cases in mathematics, one has a usefully
expressive term which time and the development of the subject have
passed by. In the meaning (1), it sometimes says beautifully what one
would like to say about an irrational; namely, that nobody can give
a rational solution to the equation, so that in a sense the equation is
deaf and unhearing to the question we ask it, and also that we can
give approximate solutions, but that to do so is ugly and sordid. Just
as with prime numbers, one must not expect an ordinary dictionary
to tell the truth about surds. One dictionary [14] defines a surd num-
ber as one that cannot be squared. This would be ridiculous if it
meant what it seems to mean, namely that it is impossible to multiply
the number by itself, since every element of a ring may be multiplied
by every other. If the definition meant that, then the number 17
would be surd when we considered it as an element of the abelian
group Z, and rational as an element of the ring Z. That would be
absurd. In fact that is not what is meant, in this old-fashioned
terminology, when it is said that a surd is a number that cannot be
squared. What is meant is that a surd is a number q such that the
equation

x^2 = q

has no rational solution. This differs in only two respects from the
defintion (1): we must replace 2 by n, where n is an integer at least 2;

128

and we must call x, not q, the surd. This use of the word 'square' to
mean 'square root' derives from geometry.



3. The inconstructible of the inconstructible


Once it became thoroughly well known that surds were quite surd,
and that it was absurd to keep on asking them, 'Who are you?
Please answer in the form of the quotient of two integers (the second
being non-zero),' people cast about for other methods of obtaining
information. Certain surds, it was found, would respond if subjected
to a very tiresome, long-drawn-out process requiring lots of strong
light and the use of (1) a sharp pointed stick; (2) a pair of sharp
pointed sticks joined at the blunt ends and capable of being set at any
desired degree of opening at the angle thus formed; and (3) a stiff,
straight rod of suitable length. Numbers which became more com-
pliant under this mode of investigation were known as constructible
numbers. All this occurred in an age of the world when people were
far more cruel than they are now. In obtaining information of this
type, it was not thought to be particularly relentless or cruel to use
such tools, or to stretch the number out along a line segment and
so possibly to put it painfully out of shape.

It would perhaps not have been surprising, considering the lengths
to which people were prepared to go in representing surds geo-
metrically, or constructing them, as it was quaintly termed, if all the
surds had given way. But there were surds too obstinate to accept
such treatment. Not only would these surds not speak up and say
who they were, they would not even lie down meekly on a line seg-
ment with their feet (so to speak) up in the air. These brave numbers
won the grudging admiration of even the ruthless investigators them-
selves for their unflinchingly irrational behaviour under the most
brutal methods of the police state, which included bisection, in-
definite extension, rotation, translation, cutting by lines and circles--
but the reader must be spared the full force of this infamy. Suffice it
to say that some of those that withstood such treatment became
known as surds of the third degree. In the curious parlance of that
abominable era, they were inconstructible.

129

QUESTION 23. Whether a surd of the third degree is inconstructible?

Objection 1. We must never give up hope. Indeed, one hears daily
of someone who has found a new way of treating some surd--say
<3>(root)2--that seems, to the investigator at least, to promise success.
There is so much more time available now for scientific research than
ever before that it hardly seems plausible to suppose the problem will
not be broken in the near future.

Reply to Objection 1. It is not a problem any longer. It is quite
certain that surds of the third degree will never yield to the dreaded
ruler and compass. Time spent on them is time wasted utterly, and
beyond hope of return. No research of this kind can be called
scientific.

Objection 2. Clearly the methods used in the past for the treatment
of irrationality have been wrong; they are essentially those of the
snake-pit. As to the irrationals of the third degree, they have been
subjected to this sadistic cruelty for so long that they may quite
possibly be beyond help by now. But some of the higher algebraic
irrationals might be chosen for a more modern approach, such as
neo-Freudian analysis or prefrontal lobotomy.

Reply to Objection 2. Prefrontal lobotomy should be done only
under medical supervision, even as applied to irrationals. It, too,
involves the use of a pointed instrument. Under analysis, even
transcendental irrationals become relatively tractable, it is admitted.
It is quite true that irrationals of the third degree may become con-
structible if the methods of ruler-and-compass geometry are aban-
doned; if for instance, the geometer is allowed a device that draws
the curve

| X
| X
| X
| X
| XX
-----------X-X-+-X-X-----------
XX |
X |
X |
X | 3
X | y = x\/


130

tangent at its point of inflection to a given line at a given point.
It is difficult to understand why such a device is not provided in the
kit with the other supplies when young children first set out in
geometry.

Objection 3. It is easy to construct the cube root of a given number,
which is a surd of the third degree. Voila`:



C
/|\
/ |/\
/ | \
/ | \
/ | \
/ | \
/ | \
/\ | \
B --------+-------- D
\ |O
\ |
\ |
\ |
\ |
\ /|
\ |
A



Here AO = n (the given number), OD = 1, the triangles are similar,
and O is the centre of a circle with diameters AC and BD.

Reply to Objection 3. Unless n = 1, there is no such circle. If
n = 1, there is a simpler construction:



|---------|
O A



Here OA is the given length 1; it is also the cube root of that length,
since 1.1.1 = 1. But since 1 is rational, this is not a surd.

131

I ANSWER THAT it is impossible to construct any root of a cubic poly-
nomial

ax^3 + bx^2 + cx + d

unless one of the roots is rational. Thus no length that can be con-
structed by the methods of ruler-and-compass geometry can ever be
a solution of an equation

ax^3 + bx^2 + cx + d = 0,

where a <> 0 and where a, b, c, d are rational numbers; i.e., quotients
of integers.

In what ways is it possible to write the number 3 as the sum of a
finite series of positive integers? The following ways are certainly
possible ones:

3 = 3
3 = 2 + 1
3 = 1 + 1 + 1

If one requires further that the sequence be non-increasing, then it
can be shown that these possibilities are the only ones (this con-
dition implies, for example, the exclusion of '3 = 1 + 2', since 2 is
greater than 1 and that makes this an increasing series). Essentially
all that is required is the inequality

0 < 1 < 2 < 3,

together with the fact that no integer lies strictly between two succes-
sive members of this inequality.

From the fact just mentioned, it is possible to show that

ax^3 + bx^2 + cx + d

must satisfy one of the following conditions: either

(i) ax^3 + bx^2 + cx + d cannot be factored; i.e. cannot be
written as the product of polynomials of lower degree having rational
coefficients; or

(ii) this polynomial can be factored, and one of the factors is
linear; i.e.,

ax^3 + bx^2 + cx + d = a(x + t)(x^2 + ux + v)
or = a(x + r)(x + s)(x + t);

again, the coefficients are rational. Ths is clear because if the
factorisation is

132

k
(product) f[i](x)
i = 1

k
then (sum) deg(f[i]) = 3,
i = 1

and because by commutativity we can assume that the degrees do not
increase.

Thus, if the cubic equation has no rational root then the poly-
nomial is irreducible; henceforth, let

f(x) = ax^3 + bx^2 + cx + d

be irreducible. If p(x) is another polynomial, and if p(x) is not of the
form f(x)q(x) for some polynomial q(x)--all polynomials shall have
rational coefficients--then since the ring

Q[x]

of polynomials with rational coefficients in an indeterminate x is a
principal ideal domain, there exist polynomials a(x), b(x) such that

1 = a(x)f(x) + b(x)p(x);

in particular, this is true if p(x) is quadratic or linear. Obviously,
then, no root (alpha) of f(x) satisfies any quadratic equation with rational
coefficients. Moreover, it is clear that

Q[x]/(f(x))

is a field containing Q as a subfield. If Q[(alpha)) is the smallest subfield
of F that contains both Q and (alpha), where (alpha) is a root of f(x) in some
extension F of the rationals, then (f(x)) is in the kernel of

(psi): Q[x] -> Q[(alpha)]
x -> (alpha),

and the induced map

Q[x]
(phi): ------ -> Q[(alpha)]
(f(x))

is a surjective field homomorphism, and hence an isomorphism.
Hence (1, (alpha), (alpha)^2) is a basis of Q[(alpha)] as a vector space over
Q. But since

133

constructible numbers can be produced by a succession of rational
operations and square-root operations, the smallest subfield of F
containing such a number has even dimension over Q; we merely
use the fact that if each of the fields k~, l~, m~ extends the preceding and
is a finite-dimensional vector space over the preceding field, then the
dimension of m~ over k~ is the product of the dimension of m~ over l~ and
the dimension of l~ over k~. But 3 is not even.


How to solve a cubic equation

It is easy to solve cubic equations. Let the equation be

ax^ + bx^2 + cx + d = 0,

and suppose the roots are x[1], x[2], x[3], so that

ax^3 + bx^2 + cx + d = a(x - x[1])(x - x[2])(x - x[3]).

If (sigma)[1] = x[1] + x[2] + x[3],
(sigma)[2] = x[1]x[2] + x[2]x[3] + x[3]x[1],
(sigma)[3] = x[1]x[2]x[3]

are the elementary symmetric functions, then writing

f(z) = x[1] + x[2]z + x[3]z^2

we get

3x[1] = f(1) + f(w) + f(w^2),

where

f(w)^3 + f(w^2)^3 = 2(sigma)[1]^3 - 9((sigma)[1](sigma)[2]
- 3(sigma)[3]),
f(w)f(w^2) = (sigma)[1]^2 - 3(sigma)[2].

Hence a root is

1/(3a){-b + [-2b^3 + 9abc - 27(a^2)d + 3(root)(-3(a^2)(b^2)(c^2) +
+ 81(a^4)(d^2) - 54(a^3)bcd +
+ 12(a^2)(b^3)d + 12(a^3)(c^3))]^(1/3).2^(-1/3) + [-2b^3 + 9abc -
- 27(a^2)d -
- 3(root)(-3(a^2)(b^2)(c^2) + 81(a^4)(d^2) - 54(a^3)bcd +
+ 12(a^2)(b^3)d + 12(a^3)(c^3))]^(1/3).2^(-1/3)}.

This is known as the cubic formula; in taking the two cube roots that
are indicated in it, always remember that among the nine combin-
ations that result from choosing among three roots in each case, only
those three combinations are allowed for which the product of the
cube roots is 2^(1/3)a^(-2)(b^2 - 3ac).

134

EXAMPLE. A man wishes to build a belfry, which shall be square at
the base, and which shall be three dekacubits higher than it is wide.
The desired volume of the belfry is four cubic dekacubits. Please give
the required measurements.


Solution. Let x be the height of the belfry in dekacubits. Then the
width and thickness at the base are both (x - 3). The equation is
this

(x - 3)(x - 3)x = 4,

or

x^3 - 6x^2 + 9x - 4 = 0.

The formula gives

x = (1/3){6 + [432 - 486 + 108 + 3(root)(-8748 + 1296 - 11664 +
+ 10368 + 8748)]^(1/3).2^(-1/3) + [432 - 486 + 108 -
- 3(root)(-8748 + 1296 - 11664 + 10368 + 8748)]^(1/3).2^(-1/3)},

or 4. Thus, a belfry 4 dekacubits high, and square at the base with
breadth 1 dekacubit will have volume 4 cubic dekacubits. From this
it should be easy to compute the volume in decibels when the bells
are installed. The bells are to be installed, of course, on decibelisation
day, and special decibel coins will be issued by the vicar of the
parish to all those who attend the ringing-in. On the great day when
for the first time all measurements will be in cubic units, it was thought
especially fitting to offer a seemly entertainment centred on the
church and capable of competing, noisewise, with the most deafening
discotheque. Afterwards there will be a talk on cubic measure-
ments, the cubit equation and the cubit formula; it is hoped that
no-one present will be able to hear what it said. Here is a run-down
on cubit measurements; note that no conversion factor from
standard British measurements is necessary:

1000 cubic millicubits = 1 cubic centicubit
1000 cubic centicubits = 1 cubic decicubit
1000 cubic decicubits = 1 cubic cubit
1000 cubic cubits = 1 cubic dekacubit
1000 cubic dekacubits = 1 cubic hectacubit
1000 cubic hectacubits = 1 cubic kilocubit
1000 cubic kilocubits = 1 cubic myriacubit.

135

The cubic cubit is established as the volume of water displaced by
the forearm of the person doing the measuring. This may be estab-
lished by the method of the great Greek mathematician, philosopher,
and physicist Eureka, who when he had put his whole forearm into
the bathtub to test the temperature of the water shouted 'Archi-
medes!', or as we would say in English, 'Decibelisation!' Regarding
the above table, it is necessary to remark that the number 1000
appears only as a rough guide, which is valid in almost all cases,
and that the general rule in the case of people with missing fingers
or parts of fingers as well as people with more than ten fingers is to
use the cube of the actual number of fingers present at the time when
the measuring is done. The joints of the fingers are usually taken as
dividing the fingers into thirds, and if extra joints are present the
finger is counted as more than one whole finger. Thus if you have 1 1/3
fingers missing from working in a sawmill, and if you were born with
11 fingers, then you must use the conversion factor 903 8/27, so that in
your case that number of cubic kilocubits make a cubic myriacubit.

The linear cubit, or measurement of length after decibelisation day,
is defined as the length of any edge of a cube of water at a comfortable
temperature that has been displaced by your forearm. The cube of
water should of course be in Euclidean 3-space, and should not be
frozen to facilitate measuring the edge as that would change the
volume somewhat and hence also the length of the edge. The square
cubit is similarly the area of a face of such a cube of water.

Ordinarily, a linear cubit is called simply a lineit, and a square
cubit is a quadratit. A measurement of 'hypervolume' in four-
dimensional space is defined in a completely similar way, and is
termed a biquadratit, or more freely, a quartit. This is sometimes
shortened to quart; other abbreviations are frowned on, and it is
especially hoped that the unpleasant dratit will not replace the more
seemly and correct quadratit.

In getting your own cubit, use the arm that seems more natural.
It is quite unnecessary to sever the limb at the elbow; simply hold the
arm in the water so that the elbow line lies in the surface of the water.
Artificial limbs having well-defined elbows may be used exactly as if
they were real ones.

136

4. Pain de maison


Poor M. Boulangiaire, the French baker, remained long in a state of
hypostatic tension over his quadratic equations. The trouble was
about those imaginary roots. He knew they would not come up in his
own bakery, because of certain facts relating to the coefficients. No-
one in the village was more sensitive to the good man's anxieties than
his widowed sister Gasparde Mange. Indeed, she felt the weight of it
more than he, being one of those women who live on anxiety. It
could almost be said that Gasparde got fat on anxiety; at least, she
tended to eat when she worried, and she worried prodigiously. Her
brother, Cre'tain the baker, was the focus around which her worries
orbited. The year after the invasion of the Ostrogoths, Gasparde
determined that her body and mind could use a course of slimming,
and took off for London. Her previous holiday, a couple of lustra
ago, had also been in London; Gasparde's life centred on Beaulieu
Derrie`re, and she could not bear to leave the village more frequently
than that. Long ago, she had learned that the food in London was
inedible, and that she had nothing to worry about in the land where
everyone does his duty.

Gasparde's sojourns in the homeland of the Anglo-Saxons were a
spasm of relief for her from the crowded activity of the village. She
usually avoided anything to do with home--she did not look at the
bucolic Butter Marketing Board posters, the vegetable shops, the
police horses, the 'Pinta' advertisements. (What a strange word for
milk that was! But everything about the English was strange.) On all
the other journeys, her eyes had never tarried on a baker's window;
that of all things was to be forgotten. Yet somehow, this time, she
could not help it. How queer the bread looked! She knew, of course,
that it was not bread; English bakers had a secret art of making the
appearance of bread out of air and water. But, it went without saying,
they could not make it quite like real bread if they chose those in-
gredients. Whatever their process might be, it certainly did not in-
volve baking. Why, it would all evaporate in the heat--and you could
see for yourself that there was no crust.

It must have been just then, when she was ruminating in front of
the bakery window, that the idea came to Gasparde. When the

137

assistant came out to ask her to leave (her immense form had been
blocking most of the light trying to come into the shop through the
plate glass) she had a question. 'Excusez-moi. Zat olmo^st spherical
wan wiz minimal superficial airia, wat is he call?'

'Cottage loaf,' the young person had said. By means of her pocket
dictionary, Gasparde resolved the cryptic name: bread of house, it
meant; a curious name. Probably because the shape of the loaf,
which had a little knob on top, could not be imitated and was a
speciality of this bakery. But it was the shape, all huddled together
on the inside with no outside surface to speak of, that interested
Gasparde. There was something about that that poor Cre'tain back
home in the village might find very interesting. Feeling her appetite
rise again at the thought of the serious business of the bakery at
home, Mme Mange hurried to the window of a tea-shop for the
only infallible remedy: the sight of Englishmen eating mashed
potatoes, tinned spaghetti and baked beans on toast. Hunger
vanished immediately, and soon the widow's mind returned to her
favourite pursuit on holiday--discovering new twists in the un-
fathomable ways of the islanders.

That winter, she recalled her moment of enlightenment for Cre'tain.
'You know, it would solve all your problems; only, you must not
make it like two balls smashed together. It is more like a chignon.'

'-- I could never, never compute the volume of a chignon; even
the sphere would introduce a difficulty, because the number (pi) on
my slide-rule is out by several thousandths.' When they spoke to
each other, Cre'tain and Gasparde always began the first sentence
with a long dash, as everyone does who speaks good conversational
French. It is the first indispensable rule; whoever ignores it is exposed
as a conversational barbarian before he has uttered a word.

'-- No, no, Cre'tain. Listen to me, I am telling you. You will not
make them like that. Do you even make the ordinary loaves like the
other bakers? Yours are pure rectangular parallelepipeds. You will
make them cubes, of course.'

'-- What an idea! The only baker in France to make a cubical
loaf. You cannot imagine the joy of it! And the mathematical impli-
cations... Now I am glad that I could never find you a husband;
you can never know how you have eased my mind. Because of
course... '

138

'-- There is always at least one real root. It does not depend on
the supply of dough, on the orders... '

'-- Wait, Gasparde. What if no-one orders the variable cube? I
will be the laughing stock of the village--there would then be a
quadratic equation, not a cubic.'

'-- Do you think I have not thought of it? Do you still think, my
brother, that because I am a fat old woman I am stupid? Have I not
been thinking of it since the autumn? I promise always to eat the
large cube. I will place an order. I myself. Only because I am your
sister--who else would eat a huge lump of uncooked dough every
day? And you are wrong about being the only baker of cubes. That
you are already; only they are the small cubes with edge of length 1.
There are others besides you who bake by the quadratic equation.
But you will bake by the cubic equation; that will be your unique
distinction!'

Immediately, M. Boulangiaire reached for a pencil that adhered
to his left ear, and wrote on the surface of the table:

ax^3 + bx^2 + fx + c = p,

where the coefficients are a for the Anglo-Saxon loaves required,
b for the flat, square loaves preferred by the Goths (M. Boulangiaire
always called them Boches), f for the ordinary French loaves, c for
the little cubic loaves whose side was constant, and p for the total
amount of dough for which there was provision. Yes, it was going
to be a grand game. The villagers would not know what to make
of it. To questions about the new loaf he would only reply, 'It is
special--you would not like it.'



QUESTION 24. Whether the cubic equation will always provide a real
root, and in fact a positive one, so that the baker can begin to bake?

Objection 1. The coefficient a is the number of Anglo-Saxon
cubical loaves, preferably of large size so as to be soggy inside. This is
certain to be at least 1, since Gasparde has promised to consume one
loaf daily, and she of all women can always be relied on to eat what-
ever she says she will eat. No doubt she will want several. Moreover,

139

the supply p always provides for more of the little buns of unit side
than the requirement c, since the villagers are never that keen to have
them. Hence if f is the polynomial function whose value at x is the
left-hand side of the equation,

f(0) < p

and

lim f(x) = (infinity).
x -> (infinity)

Taking N very large and positive, and applying the Intermediate
Value Theorem to f on the interval from 0 to N, we see that

f(0) < p < f(N)

and hence there exists a real number (phi) such that 0 < (phi) < N and such
that f((phi)) = p. This is then a positive root.

Reply to Objection 1. This solution of the problem has already
occurred to M. Boulangiaire, but has been rejected by him for the
very good reason that it requires the Intermediate Value Theorem,
of which he was not given a rigorous proof at the E'cole Paranormale.
We, of course, know of a rigorous proof of the Intermediate Value
Theorem, but it is one that uses the idea of the existence of a univer-
sally attracting object in the category of archimedean fields, and the
fact that such a universally attracting object must necessarily contain
the least upper bound, or supremum, of a set of numbers that is not
empty and that possesses an upper bound. Not only would M.
Boulangiaire be unable at his present age to assimilate all the neces-
sary concepts to follow a proof, but he would be justifiably annoyed
at the introduction of analytical concepts in order to solve an
algebraic problem. 'You are employing a blast furnace in order to
toast marshmallows,' he would say in his shrewdly practical way.

_______
Objection 2. Replacing any root (alpha) by its complex conjugate (alpha),
we get

_______ _______ _______
a(alpha)^3 + b(alpha)^2 + f(alpha) + c
________________________________________ _
= (a(alpha)^3 + b(alpha)^2 + f(alpha) + c) = p,

_______
so that (alpha) is also a root. Hence the roots occur in conjugate pairs,
and since 3 is an odd number and there are three roots in all, one of
them must be equal to its complex conjugate and hence real.

Reply to Objection 2. In order to do the problem this way, it is
necessary to construct the complex number field, or some other field

140

containing all the roots of the equation, and possessing an alter-
_ _
nating automorphism k -> k; i.e., one such that k = k; the alter-
nation must have the property that the subfield it leaves invariant is a
real field. The study of automorphisms of extension fields and the
subfield that they leave invariant is Galois theory, and the study of
the complex numbers is analysis; hence, these methods are too high-
powered.


I ANSWER THAT there exists a real root of the cubic equation under
consideration, and that it is positive. This means that the baker will
always be capable of baking bread in such a way as to solve the
equation and just use up his dough--although he will have to
approximate and so may be a little bit out, the error and hence the
amount of dough left over or not available for finishing the last loaf
can be made as small as desired. Unfortunately, he will not in general
be able to construct x, the length of the French loaf and the edge of
the Gothic and Anglo-Saxon loaves, by ruler and compasses as he
used to do with the old method. When he had only a quadratic
equation to solve, it was possible to obtain the required dimension by
ruler-and-compass construction, and the baker greatly enjoyed the
process. This recreation will now have to be abandoned. A number
of good methods nevertheless exist by which the approximation can
be refined within acceptable bakery tolerances.

The point is to have a system of numbers in which it is possible to
do addition, multiplication, subtraction, and division, in which it is
possible to say whether a number is positive or negative in a way
consistent with the arithmetic of the system of numbers; the system
must be generous enough to include a positive number that corres-
ponds to the desired dimension x, where x is the solution of the
bread equation given by the orders of the various customers. That
will all be the case if there is a real extension field of the field of
rational numbers containing a root of the equation.

If the equation is reducible, this is obvious. Taking g(X) to be

aX^3 + bX^2 + fX + c - p,

and (alpha) to be a root of this polynomial, suppose -1 is a sum of squares

141

in Q((alpha)). We get the equation

n
-1 = (sum) p[i](X)^2 + k(X)g(X),
i = 1

where of course the p[i](X) have degree less than 3. Since k(X) there-
fore necessarily has degree 1, or is the zero polynomial, there exists
a rational number q such that k(q) = 0. Hence, -1 is the sum of
squares of a finite sequence of rational numbers. This is impossible,
and hence a real root exists.

The question remains whether any of the roots is positive. This
is clear if all roots lie in a real field, since their product is (p - c)/a,
which is a positive rational. Otherwise, the polynomial is reducible,
so that

aX^3 + bX^2 + fX + c - p = a(X - (alpha))(X^2 + BX + C).

Since B^2 >= 4C would imply that all roots of the equation lie in a real
field, it may be assumed that B^2 < 4C. Since c - p is negative, so
must be -(alpha)C, and it follows that (alpha) is positive.



How to find cube roots

In solving a cubic equation by the cubic formula, one is obliged to
take cube roots. This may be done by the method of approximation
by continued fractions to within any required degree of accuracy.
For example, to find <3>(root)2 we note that x^3 - 2 changes sign between
1 and 2; this leads to

x = 1 + 1/y,

and on substitution we find

y^3 - 3y^2 - 3y - 1 = 0.

The left-hand side changes sign between 3 and 4, which leads to

y = 3 + 1/z.

Similarly we get z = 1 + 1/w; continuing this we get

<3>(root)2 = 1 + 1/(3 + 1/(1 + 1/(5 + ...;

the computation taken thus far gives 29/23, the cube of which is

2 55/12167.

142

5. The Rule of Three


If a man and a half, all of whom speak Scots Gaelic as well as
English and have had 5 years of schooling, dig a hole and a half
while the temperature is 35 degrees Fahrenheit, smoking 22 cigars, eating
4 lb of preserved goose, reciting a Shakesperean sonnet, and taking
a day and a half to do the job, how long will it take one man, speak-
ing only Ancient Egyptian, Sumerian, and Chinook Trade Jargon
and having 3 years of schooling, to dig a single hole while the tem-
perature is 3 degrees Centigrade, smoking 11 cigars and eating 3 kilos of
moules marinie`re while saying off from memory all the poetic con-
tents of the Egyptian edition of The Lord of the Rings? [33]

The above is an example of a question in the Rule of Three. It is a
moderately complicated example of the genus, and the fact that it
can be solved at all is a tribute to the immense versatility of our con-
venient algebraic notation. A simpler problem is this: if John gets
10 hectathrills from accompanying to a ball a well-proportioned lady
of 25 years who is 5 feet tall, how many thrills does John get from
accompanying to the same ball a well-proportioned lady of 50 years
who is 10 feet tall? (The thrill is the basic measurement of joy; a
hectathrill is equal to 100 thrills.) Let us simplify still further, by sup-
posing that not only the gentleman, John, but also the age of his
young lady remains constant--whether she is 25 or 50 need not con-
cern us. The only thing that changes is her height. Then we note that
the problem is stated as a question. This is very unpleasant, because
it is not a proposition unless it is in the indicative mood. Hence we
make it a statement, using a letter (say O) to stand for the unknown
number of thrills. It now takes the form:

O thrills of John are to 10 feet of lady
as
100 thrills of John are to 5 feet of lady.

Here we have a problem in the standard rule of three, so called
because it has as its aim the solution of all problems in which one is
to find the fourth, unknown, term in a proposition in which three
terms are known. The more complicated examples, such as the one

143

previously mentioned involving a choice of cigars, languages,
delicatessen, etc., can only be treated once the main points of the
simple, classical case are fully understood. Before considering these,
it would be desirable to examine first why there is no such thing as a
Rule of Two, or a Rule of One. In a sense, of course, these things do
exist. A Rule of One might be thought of as a way of finding a single
unknown from a single given quantity, and a Rule of Two as a func-
tion of two variables. But these ideas do not fully mirror the structure
of the Rule of Three, which is not merely a matter of a function of
three variables. It is more nearly a matter of a class of functions
indexed by a parameter; two of the given numbers suffice to deter-
mine the parameter, and the third determines the particular value of
the function associated with the value thus determined for the para-
meter. Thus, no fewer than three given numbers can produce a
sufficiently rich structure. Rules associated with larger odd numbers
are then a matter of families of functions of several variables; i.e. the
Rule of 2n + 1 is simply produced by considering an index set A and
a function

(f[a]) a (epsilon) A: A -> Y,

where

n
X = X~ X[i]
i = 1

Hence, there can be no Rule of Two since 2 is even, and there can be
no Rule of One since n = 0 gives a Cartesian product over an empty
set of indices (from 1 to 0), so that X is a singleton {x}. Necessarily
the given term in the proposition is f[a](x), and this is just what we are
required to find--hardly an arduous task requiring a special rule for
its accomplishment.

In the case of John and the expandible-contractible ladies, it is
clearly presupposed that once we know how John reacts to a little
thing of 5 feet we can say just how a 10-foot version of the same
delicious morsel will hit him. Now various men react to different
women in various ways. For some, the thrill function reaches its
maximum value at 5 feet 2 inches. Clearly, what we must know from

144

the given information is John's thrill function. In other words,
the function

X -> Y

induced by A -> Y actually takes its values in the subset of Y

consisting of injections A -> Y. Thus, if two different men (John and
Xerxes, let us say) receive exactly the same thrillage from the com-
pany of a 5-foot woman, then they must also receive exactly the same
thrillage from a 10-foot woman. Otherwise, it is impossible to answer
the problem as set.

Let us suppose, in order to be able to do the problem, that a lady
h feet high produces in a man a number of thrills equal to

a sin(((pi)/10)h)

where a is the parameter. We easily determine in the present case that
for John the parameter a takes the value 100; hence it is easily com-
puted that he gets exactly no thrills from taking out a 10-foot lady-
friend.



6. Polynomials


In treating problems related to the Rule of Three, it is often assumed
for simplicity that the functions f[a] in the problem are all of the form

f[a](x[1], x[2], ..., x[n]) = a(x[1]^k[1])(x[2]^k[2])...(x[n]^k[n])

where the k[i] are integers, and where A, X[1], X[2], ..., X[n] are all the
same field. If we restrict this further to the case where the k[i] are
natural numbers, we reach a kind of function closely related to the
polynomial. Polynomials are sometimes divided into monomials (pro-
perly spoken, these should be mononomials), binomials, trinomials,
.... There seems to be no special name for polynomials that have no
terms at all, but otherwise these are just fancy ways of incorporating
into the name exactly the number of terms that exist in the poly-
nomial. Of these, only the first type, the monomial, deserves separate
study. Hence we begin the study of polynomials with that of mono-
nomials, or what might almost be called terms.

145

An example of a monomial is

(tau)(A^a)(B^b)(C^c)(D^d)(E^e)(F^f)...(Z^z)

where (tau) is likely to be an element of a ring, and where the exponents
a, b, c, ..., z are natural numbers. The letters A, B, C, ..., Z are
called letters or indeterminates. It can easily be shown that the
monomial just mentioned is a monomial in (at least) 26 letters; in
order to simplify slightly the questions which arise, we shall begin by
considering monomials in just one letter.

An example of a monomial in just one letter is aX. From this
monomial we can make a function, in a way which will be more fully
described below. If we assume that the thrill function with parameter
is obtained from this monomial, rather than from the function sin,
then we shall be able to compute that if John gets 100 thrills from
dancing with a lady 5 feet tall, then

100 = a.5,

so that a = 20. This shows that if John dances with a lady 10 feet
tall then he gets 200 thrills. (This approach is called linearising the
problem, and shows one of the many charming applications of this
special kind of problem.)

Another kind of monomial is X^a. Using this monomial we produce
a parametrically indexed family of functions, and the same problem
as before--if the reader will forgive the intrusion of one more worked
example of the Rule of Three--can again be worked on the hypoth-
esis that the thrill reaction of a man dancing with a lady depends on
the height of the lady according to the law

h -> h^a,

where h is her height and where a is a parameter which has to be
determined in the case of every individual man, since men do not all
react to ladies in the same way. In the present case, we get

100 = 5^a,

and since a is a natural number, this is impossible, since it leads to

5^a = 5^3 - 5^2,

which can be shown by computation to be false if a = 0, 1, 2, or 3,
and which is false for a >= 3 because

Z^(a-2) - Z + 1 = 0

146

has as its only rational roots a subset of {-1, 1}. Why can the prob-
lem not be done? There are two explanations: one, that the word
'lady' no longer has a definite meaning. In 1840, a self-respecting
servant girl could say, 'Why, if I were a lady, I should be delighted
to be the object of Captain Ainstruther's affectionate interest.' Then,
it was a matter of fact whether a woman was a lady or not. At
present, many women who are certainly not ladies are most insulted
if they are called women, and speakers of the language are divided
between those who call all women ladies, and those who call no
women ladies. Now in a mathematical exercise, it is of crucial import-
ance that all the words used have a definite, clearly defined meaning.
It is no good applying mathematics to a mixture of vague impres-
sions, compounded of the notions 'lady', 'gentleman', 'well-propor-
tioned', etc. All the concepts must rest on a firm logical basis, as
ball:

{x (epsilon) R<3>: |x| <= 1}

--one may gather that it is the 3-ball from the fact that the people
who are going to the ball are in a problem illustrating the Rule of
Three, but if it were some other odd number, one would simply take
the ball in the Euclidean space of the corresponding number of
dimensions. That is one reason why we cannot expect very startling
success in determining the thrillage. The second reason is that in this
case the image of 5 under X -> Y
is not a surjection A -> Y, and
that this is just as necessary as that the image be an injection.



QUESTION 25. Whether a polynomial is a kind of function, or not?

Objection 1. A polynomial is a kind of function. In fact, if you
start with a variable x, by which is meant a number that can be any-
thing it likes, and if there are some constants, by which are meant
some numbers that are stuck and cannot change at all, no matter how
hard they try, and if addition, subtraction, and multiplication are
allowed but not division, then one will get an expression of the form

a[n]x^n + a[n-1]x^(n-1) + ... + a[0].

Now if you let the variable x vary--which is to say, change all over

147

the place--then this quantity will also vary. As x varies independently
of any control, the quantity formed in the way described will vary in
a way that depends on the way x varies. Hence x is called the in-
dependent variable and y is the dependent variable. But where a
dependent variable depends on an independent variable, there one
has a function; and functions constructed according to this formula
are called polynomials.

Reply to Objection 1. It is bad form to mention variables; not only
that, but it is very imprecise to talk about numbers changing all over
the place, moving and heaving about like quicksand. It would be
incorrect to say that the thing presented here is a polynomial, and
just as imprecise to call it a function. It is by no means clear just
what it is.


I ANSWER THAT polynomials are not functions. The reason is that
polynomials can be used to obtain functions, but that two poly-
nomials may produce the same function without being the same
polynomial. Before explaining function, we must of course go back
to monomials, and indeed we must begin by considering monomials
without coefficients. There will be a set of indeterminates, or letters,
which are to be used; let this be S. Of course, if M is any monoid
then M|S| has the obvious monoid structure, and it is called the set of
monomials in the letters of S without coefficients, and with exponents
in M|S|. One ordinarily takes M to be the natural numbers N[0]. In terms
of notation, there is a trick to be learnt; to indicate the element of
M|S| that sends A[i] to n[i] for i and integer such that 1 <= i <= k, where
S = {A[1], A[2], ..., A[k]} and k is a positive integer, we write

(A[1]^n[1])(A[2]^n[2])...(A[k]^n[k])

Moreover, if A (epsilon) S and if A is sent to 1, it is permissible not to
write the 1; and if A is sent to 0 it is permissible to write neither the A nor
the 0; but this does not apply if S = {A}.

The obvious structure on the monoid N[0]|S| then makes it obvious
that (taking S = {X}) we have

(X^m)(X^n) = X^(m + n);

it is impossible to exaggerate the importance of this identity in com-
putations with polynomials.

148

Unfortunately, the above considerations of notation have the
effect of importing a certain difficulty into the study of the monomial.
Taking again the case S = {X} for simplicity, it is clear that the
expression X^n can mean two separate things: firstly, the function
S -> N[0] that sends X to n; secondly, the nth power of X = X^1,
which is the function S -> N[0] sending X to 1. The solution is not far
to seek--we merely map N[0] homomorphically to N[0]|S| by

n -> X^n,

which means two homomorphisms by the two interpretations of
'X^n', and observe that the two homomorphisms are not distinct,
making one homomorphism and hence one interpretation of 'X^n'.

By the same token, we see that there is an a priori danger of con-
fusion between AB on the one hand and AB on the other; or slightly
more generally between (X^m)(Y^n) and (X^m)(Y^n). It would of course be
somewhat discomfiting for the theory of polynomials if these were
really as distinct as one must suppose, without proof, they may be.
Therefore let us attempt to show, if possible, that they remain the
same, always restricting ourselves for simplicity to the case AB
versus AB.



QUESTION 26. Whether one may be certain that AB is the same as
AB?

Objection 1. This, surely, of all the ridiculous quibbles, of all the
effete nonsense, of all the crabbed contortions found in the present
compendium of cant, is the silliest and most abominably absurd! Is
a thing the same thing as itself, when it is mentioned under the same
name? Flogging is the answer to those who permit themselves to ask
such questions, if no more ignominious punishment is available.
And a remedial period of good, hard work could without detriment
follow, to be continued until a more socially oriented behaviour
pattern develops. Apiary work would do. The patient might be re-
quired to repeat, whenever stung, 'A bee is a bee is a bee.'

Reply to Objection 1. It may possibly seem, at first sight, that when
two things are mentioned under the same name they are the same

149

thing. If one is set the problem: Prove that Oberon is the King of the
Fairies, it is not surprising that there is some work to do--one may
have to find a fairy and say, 'Take me to your leader!' But on being
asked to prove that Oberon is Oberon, one reacts differently. Since
it is not clear how to proceed, even though the task hardly promises
to be one of extreme difficulty, one may feel inclined to bite one's
nails or to call for the police. It is not unnatural to react in this way.

The present case is in fact slightly deceptive. The conventions of
notation have produced two expressions that were intended to say
different things, and that were arrived at by different roads; and by
pure chance the two expressions have the same form. It is rather like
the two English sentences

'Fruit flies like a banana.'

and

'Fruit flies like a banana.'

The first of these sentences means that certain small flies, some of
which belong to the species Drosophila, and all of which feed on
fruit when they are in the larval stage, if they can get it, have among
the fruits for which they feel a preferential inclination a yellow,
elongated fruit of the genus and species Musa paradisiaca sapientum.
In other words, those fruit flies really go for a banana. The second of
the two sentences means something quite different; to wit, that
vegetable produce succeeding the flower passes through the air in a
manner resembling that in which a banana would pass through the
air. The first sentence could be tested as to veracity by presenting
some fruit flies with various kinds of food: a measuring tape, some
beefsteak tartare--the list can vary without restriction, except that
it must contain a banana. That way, you could tell if fruit flies like a
banana. To test the second proposition, one might use a wind tunnel,
or a large catapult, in order to study the aerobatic properties of
mangoes, lichees, pears, and (just possibly) aubergines and horse-
chestnuts, and to compare these with the aerobatic properties of a
banana. 'Fruit flies like a banana' is a proposition in entomological
gastronomy, whereas 'Fruit flies like a banana' is a proposition in
horticultural aerobatics. These two subjects, while they are important
intellectual disciplines demanding our utmost respect, are still in
their infancy, and unfortunately we cannot safely rely on them for
more than tentative, if hopeful, guesses as to whether fruit flies do

150

like a banana, or as to whether fruit does fly like a banana. The im-
portant point is that the two fields have as yet no common ground;
at present, no reputable interdisciplinary work has been done in
both at once; no joint degree has been taken in entomological
gastronomy as applied to horticultural aerobatics; no lecture en-
titled 'Fruit flight and the diet of Drosophila: bananism versus
pomegranitry' has been read. There is no connection between the
two sentences. If we are to prove that AB, understood in one sense,
is the same as AB, understood in a quite different sense, we shall not
do it just by remarking that they look the same.


I ANSWER THAT without doubt AB = AB. Consider the three
monomials involved in the equation: A, B, and AB. These are
monomials in a certain set of letters S, where S certainly contains A
and B--considered as letters, not as monomials--and where S may
possibly contain any further finite set of letters--say, all Russian
minuscules that normally precede the letter 'e' in a native Russian
common noun, all Hebrew letters that take daghesh forte but not
daghesh lene, and the Devanagari voiced aspirates together with
Latin majuscules not congruent to 3 modulo 4.

Now as a monomial,

A: A -> 1

and if (mu) (epsilon) S (without) {A} then

A: (mu) -> 0.

Displaying in a table this information for the monomial A with
similar information for the monomials B, AB,

| (mu) = A | (mu) = B | (mu) (epsilon) S (without) {A,B}
------------+------------+------------+----------------------------------
A: (mu) -> | 1 | 0 | 0
B: (mu) -> | 0 | 1 | 0
AB: (mu) -> | 1 | 1 | 0

we get that A~ A((mu)) + B((mu)) = AB((mu)).
(mu) (epsilon) S


Those who have mastered the art of forming monomials may, if
they are good, be admitted to the secrets of polynomials. These are

151

among the most important weapons in the arsenal of the mathe-
maticians, so that familiarity and skill in the handling of polynomial
manipulations is absolutely essential for passing examinations. The
most important skill of all, the sine qua non of the would-be alge-
braist, is the ability to recognise a polynomial on sight. This is not
as simple as it sounds. Recognising polynomiaIs on sight can be done
on various levels. At a rather lowish level, one may be shown this:

a[0] + a[1]X + a[2]X^2 + ... + a[n]X^n,

and one may be asked, 'Well, how about it? Is it a polynomial, or
isn't it?' The low-level answer is (depending on the dialect one
speaks),

'Yeah, sure I guess. Looks like one. Yeah, sure, that's a poly-
nomial, all right.'

Now, such an answer is neither too extremely bad nor on the other
hand is it as good as it might be. The worst answer (excluding some
very improbable sorts of answers, since there is not room here to
consider any but the most common possible responses) would be a
flat, definite No. After that, the next to the worst answer is a flat,
definite Yes. The present answer, a hesitant, fumbling Yes, has many
advantages. Hesitation is often a symptom of superior knowledge; in
fact, one may have rarely had the opportunity to see polynomials, so
that one is just barely able to tell when something looks like a poly-
nomial. In order to hide one's ignorance, one is inclined to be daring
and come out with a flat answer. But then one realises that experts
ofen hesitate. Experts have all sorts of reasons to think two ways on
a question, because they realise all the fearful complexities of a sub-
ject. By hesitating, one may get oneself mistaken for an expert. The
finishing touch on this approach is to use the all-important phrase,
'It all depends'. (For added effect, learn to say this in several foreign
languages and say it as if quoting from an eminent algebraist who
spoke the appropriate language, as 'Das kommt darauf an, as Gauss
might have said'. Gauss may be used for almost any subject.)

Rather better than this last answer is one that shows, if not any
great familiarity with polynomials, at least some idea about mathe-
matics; one may say, 'That all depends. Do you treat it as a poly-
nomial? Do you operate on it as you operate on a polynomial? If so,
it is a polynomial, as near as makes no difference. If not, then it is

152

certainly not a polynomial.' This could be extended further by, 'Is it
part of a system isomorphic to a system composed of polynomials?'
and words to that effect.

Of course if one can define a polynomial, it is a good thing to do so;
and in that way one can of course give the correct answer to the
question. Knowing what a polynomial is comes before all other
knowledge in the field of polynomials as an essential prerequisite.
Nothing could be simpler. First, polynomials have got monomial
expressions, and these to begin with are pure and sans coefficients.
Then, they have got to have a supply of coefficients from somewhere.
Since the coefficients are going to have to be added and multiplied
at some point, they are usually taken to come from a ring. Thus we
have the pure-monomial-expression monoid N[0]|S| and the ring (sigma).
The most obvious thing to do with these is to form the monoid
algebra of N[0]|S| over (sigma), which may be denoted (sigma)N[0]|S|, or for
simplicity (sigma)[S]. This consists of all functions in

(sigma)N[0]|S|

that take the value 0 everywhere except on a finite subset of N[0]|S|.
These functions are added in the obvious way, and are multipied by
convolution or Faltung. This means that if p and q are polynomials,
their product r has the property that

A~ r((mu)) = (sum) p(x)q((tau))
(mu) (epsilon) N[0]|S| x(tau) = (mu)

An obvious change of notation produces the more usual form; thus
when S = {X} we can write the polynomial in the form

a[0] + a[1]X + ... + a[n]X^n.

The reader will note that this change of notation bears an obvious
similarity to integral transforms. The verification that the sum of X
and Y, considered as polynomials in the letters X and Y, is exactly
the polynomial X + Y need not detain us; it is no harder than the
proof that the product of X and Y is XY, which has already been
explained.





153

Polynomial identities

One of the most basic identities for practical purposes in connection
with computations is

(X - Y)(X + Y) = X^2 - Y^2.

This is sometimes expressed in words: the sum times the difference
is the difference of the squares. Since we know how to multiply
polynomials (previous section), this should not offer any great
difficulty. Every term in the polynomial (X - Y) has degree 1; in
other words, the value of the function at an element of the monoid of
pure monomials is 0, unless the element of the monoid of pure
monomials is (X^h)(Y^i) with h + i = 1. Note that for simplicity it has
been taken for granted that all the polynomials are polynomials in
the letters X and Y, and that if this supposition is not made then the
last statement must be modified. Similarly, the polynomial (X + Y)
is a function taking the value 0 on the pure monomial (X^j)(Y^k) unless
j + k = 1. Note further that we have no idea what is meant by the
statement 'the function takes the value 0 at a monomial', since we
have no idea what this 0 is and are even unaware what ring 0 is the
zero element of; and note also that this makes no difference to the
argument. Putting

h + j = m
i + k = n,

we see that the product of the two polynomials has degree 2; that is,
their product is a function that must take the value 0 at (X^m)(Y^n)
except in case m + n = 2. This is because m + n = (h + j) +
(i + k) = (h + i) + (j + k) = 2 whenever even one term in the con-
volution sum is non-zero. Hence we know that the product takes the
form

aX^2 + bXY + cY^2,

and it remains only to compute the actual values of a, b, and c. In
order to compute a, it is necessary to observe only that if

/- -\ /- -\ /- -\
| 1 0 1 0 | | h | | 2 |
| 0 1 0 1 | | i | = | 0 |
| 1 1 0 0 | | j | | 1 |
| 0 0 1 1 | | k | | 1 |
\- -/ \- -/ \- -/

154

then (h, i, j, k) must differ from (1, 0, 1, 0) by an element of the
kernel of the homomorphism Z<4> -> Z<4> determined by the matrix;
hence is of the form

(1 + w, -w, 1 - w, +w),

where w is an integer. Since it is further required that h, i, j, k be
natural numbers, we must necessarily have

w = 0.

Hence the coefficient of X^2 is the product of the coefficient of X in
(X - Y) and the coefficient of X in (X + Y), or 1.1, giving

a = 1.

In a similar way, if (X^h)(Y^i) and (X^j)(Y^k) are to make a contribution to
the coefficient b of XY then we must have

(h, i, j, k) = (1 + w, -w, -w, 1 + w).

There are seen to be two solutions in naturals, namely (1, 0, 0, 1) and
(0, 1, 1, 0), and we get b = 0. Similarly c = -1 and we are done.

Note that the above product X^2 - Y^2 becomes X^2 + Y^2 over
rings of characteristic 2, where -1 = 1.

An exactly similar method may be used to show that many other
familiar polynomial identities hold. Alternatively, of course, one
may verify that the monoid algebra (sigma)[S] satisfies the ring axioms;
i.e., one may verify that the convolution product is associative:

(sum) ((sum) p(x)q((tau)))r((mu)) = (sum) p(x)
(delta) = i(mu) i = x(tau) xi = (delta)
((sum) q((tau))r((mu)))
(tau)(mu) = i

and that the 'constant polynomial' taking the value 1 on the neutral
monomial and taking the value 0 elsewhere is neutral for con-
volution. Distributivity over addition must also be verified.




7. What are brackets?


Nothing is so important in mathematics as knowing what brackets
are. In America and elsewhere abroad, brackets are parentheses; at
least the people in those places think that brackets are parentheses,

155

and call them parentheses. Doubtless this is but one instance of the
fact that, broadly speaking, the people abroad speak broadly.
Americans, especially speak so broadly they speak expansively; they
expand the word 'lift', a box for hauling people up and down in, into
'elevator', and they have applied the same principle to the word
'bracket'. This habit is known as American largesse.

The division of brackets is twofold. First, they are divided into
species, which are the round, the square, and the curly. Secondly, each
species is itself divided into two sexes, the left-hand (or sinister) and
the right-hand (or dexter). For example, ( is a left-hand bracket, and )
is a right-hand one.


A parenthesis on the sex life of brackets

Human beings live in a very different world from that inhabited by
brackets. Our world (the author is himself a human being) is a three-
dimensional one; that is, disregarding time and speaking locally, a
human being who refrains from extended space travel spends most
of his time in a homeomorph of R<3>. In a similar sense, we may call a
bracket a two-dimensionalite; he (or she) lives in R<2>. Just as we can
hardly imagine the life of a being in four-space, so can brackets only
dimly guess at us. A thoughtful bracket could certainly form the
abstract idea of the space we live in. An imaginative bracket could
perhaps people such a figmentary space with putative beings. A lucky
bracket might guess something about the division of our race into
sexes, and the shapes of our bodies. But surely no being so alien to
us could ever suppose our complex customs of marriage and court-
ship to be such as they are. Brackets, on those rare occasions when
they think of us at all, must picture us as some kind of huge, curly
bracket.

But are we in a better state with regard to our little fellow creatures?
Most of us see them often enough; how often do we try to think
what life is really like for them? Yet the rules of their game, the sys-
tem sanctioned by nature and society by which they regulate their
lives, is perhaps as complex for brackets as our own system is for us.

One misconception must be scotched at the start. Just because the
left- and the right-hand brackets look rather alike, people sometimes
think that no polarity exists between them. This is not true, as we shall
see. An especially dangerous idea is that a left bracket can, by leaping

156

upward on the page and turning a somersault, become a right
bracket--in short, that sex-change is possible for brackets. Again,
not true at all. The bracket remains on the line, and never tilts for-
ward or backward; so that he (or she) can never change sex. Motion
exists, but it is always a motion along the line of print.

Brackets marry. Unmarried brackets are extremely rare; indeed,
they are almost non-existent. (There is an exception in France.)
Brackets marry brackets of the opposite sex, but of the same species.
Again, one would like to say that this rule is universally true, but
there are (alas!) exceptions. These are as follows, and readers may
well wish to close their eyes: square and round brackets of opposite
sexes do marry. Unions of male or left-handed squares with female
rounds, and unions of female squares with male rounds are about
equally frequent; the couples are never very deliriously happy,
bcing separated by an interval; e.g., [a, b) on the one hand and (a, b]
on the other. No such mixed matings are ever indulged in by the
curlies. The name of marriage is too holy to dignify the unnatural
liaisons that occur chiefly in France, to the shame of that proud
nation: such monstrosities as [a, b[ and ]a, b] in which two brackets
of the same sex join horribly. Even harder for virtuous nations to
understand, perhaps, is the ridiculous sort of marriage that occurs
there between two brackets who, although they are of the same species
and of opposite sexes, have got their ro^les reversed: ]a, b[. Even in
France, it is only the square brackets that commit these abnormali-
ties; thus curly brackets are the purest of all in their marriage
behaviour.

Courtship as such is hardly known among brackets. Brackets are
mature almost immediately after 'birth', quite contrary to the rule
that long-lived creatures generally take a long time to grow up. They
marry just as quickly. There is almost never any doubt as to who is
married to whom. Thus in (a(bc)) the inner pair are married, and the
outer pair are married. The absolute fixity which reigns in the
relations of brackets might lead some human students to suppose
that love among the brackets is a very humdrum affair. As we all
know, courtship by the very uncertainty of its result adds a dimension
of suspense and anticipation to love. Many romantic novels play on
the theme of courtship; few carry on the story into married bliss. If
we examine our hearts, we see that we should not be satisfied if the

157

course of love were perfectly smooth. When we examine the life
story of a bracket, perhaps we ask, 'How can it interest them?' It
seems to us like utter monotony.

Nothing, in actual fact, could be further from the truth. The love
life of brackets is not static bliss, but dynamic anticipation. The male
and female brackets forever strive to come together, and forever
they are prevented. There is always something in the way. + Often one
sees couples for whom the gleam of hope must appear infinitesimally
tiny and distant, like a faint ray of light to a lost explorer of under-
ground caverns:

(x[1] + y[1] + z[1], x[2] + y[2] + z[2], ..., x[q] + y[q] + z[q]).

At other times, especially in these latter days, couples almost in
each other's arms:

f(.).

Indeed, a male bracket must feel much as did a knight in the days of
chivalry, who had pledged himself beyond recall to a lady far above
his station, to whom he could never hope to attain.

So all-pervasive is this feature of parenthetic love life that we
ought perhaps to speak of betrothal rather than marriage. Many
bracket couples not only cannot contrive to come together, but are
prevented even from seeing one another by older couples who stand
between the lovesick male and female bracket; while these older
couples may themselves be the objects of a similar chaperoning
operation performed by yet other couples:

((ab)((c(de))e))f.

Here the couple (de) separate another pair in (c(de)) and these
separate yet others, and so on. Only (ab) and (de) can see each other.

Looking on with pitying eye, if we are kind we may be seized with
a desire to interfere in the ro^le of Cupid. It would be possible for us,
from our god-like vantage, to unite the pairs, to bring to fulfilment
loves otherwise unrequited. We can twine two brackets in a true
lovers' knot, and from (ab) make

/----\
| ab |
\----/
--------------------------------------------------------------------------
+ See p. 61.

158

or from ((ab)((c(de))e))f make


/----------------\
| /----------\ |
/----\ | | /----\ | |
| ab | | | c | de | | e | f.
\----/ | | \----/ | |
| \----------/ |
\----------------/


Thus the two are made one, and the he-bracket may say to his she-
bracket, 'Borne of my borne, and fle`che of my fle`che!' It is possible
for us to do the thing; but is it wise? For what we have joined can no
longer be put asunder. What is written, is written, like the laws of
the Medes and the Persians. And it may begin to pall on the brackets.
Too much of a good thing may well be worse than none at all. Had
we not better leave well enough alone? For after all is said and done,
human knowledge of the love life of brackets is meagre.


The most important use of brackets is to show the order in which
operations must be performed. Thus, the above may indicate

F(F(F(a, b), F(F(c, F(d, e)), e)), f).

The function F is suppressed as being too obvious for mention, and
the simpler form involving only brackets results. There is a direct
approach which is used in Poland sometimes, whereby instead of
suppressing the functions one suppresses the brackets; for example,
one writes

FFFabFFcFdeef.

This way of writing things is said to be easy for machines to read.
One may also replace brackets by a numeral or mark representing the
depth of the deepest bracket at the point, as

ab///c/de//e////f.

However we choose to indicate the order of operations, we must
make it clear that one must perform either ab or de first, and that the
last step is the one that uses f.

It is clear that when brackets are used in this way, certain rules
must be obeyed if the result is to make sense. If the rules are not
followed a very unhappy situation develops in which brackets are

159

not married, or are not sure to whom they are married. A sensitive
reader soon notices when this has happened--we do not know quite
how; perhaps it is because of a kind of thought transfer from bracket
to man. For example, in writing

(ab)c)d)((ef)

the author has started some very sad brackets on the road of life.
In ((ef) we have, on the left, a sort of eternal best man who can never
be married and who is doomed to be part of an eternal triangle,
hanging about on the outskirts of the married couple (ef). It cannot
be good for any of them, although ) may be flattered by the extra
attention she gets. There are other heartrending situations present in
this example, and the author would never have created this woeful
little microcosm but for his desire to warn others not to make the
same mistake.

If we are to avoid such errors, we need a rule explaining when ex-
pressions obey the social rules so necessary for the smooth functioning
of domestic life among the brackets, and when the expressions
introduce friction into the well-lubricated workings of the bracket
household. We need to have a procedure for writing bracketed
expressions that will produce all the right expressions, and none of
the wrong ones. Since a bracketed expression can be obtained from
two smaller expressions by writing one down after the other, and then
surrounding the whole mess by a pair of brackets, we may write

S = A + (SS),

where A = a + b + c + ... + z, unless we wish to use further
letters; e.g., the 5000 most common characters of standard literary
Chinese. If we wish to use these they must be added into the sum for
A. Note also that the brackets are not brackets here, but generators
of a free semigroup on 28 generators (or 5028 generators if you
include the Chinese characters). It is 28 and not 26 because of the
two brackets. It is clear that all bracketed expressions are generated,
and only such expressions. Moreover, it is clear that no expression is
generated ambiguously (use induction on the number of letters in the
expression, not counting the brackets). It is extremely difficult to
expand S as a series in the semigroup on 28 generators with integer
coefficients, but as we have said the coefficients in such an expansion

160

are all 1. If, however, we delete ( and ) and replace A by a, we get

S = a + SS,

and if S = k[0] + k[1]a + k[2]aa + ... we see that this is equivalent to

(k[i]) * (k[i]) = ((delta)[i]<1>) + (k[i])
i (epsilon) N[0] i (epsilon) N[0] i (epsilon) N[0] i (epsilon) N[0]<9>

where * is the convolution product and where ((delta)[i]<1>) i (epsilon) N[0]
is (0, 1, 0, 0, 0, ...), the sequence that is 1 in the first place and 0 in
the zeroth and all other places. With q[i] = k[i] + (delta)[i]<0>, we get

(q[i]) * (q[i]) = 1/4 - a,

which gives us

2(2i - 3)!
k[i] = -------------
i~)2~(!~ - i!

by the binomial theorem. Thus there are 429 ways of inserting brackets
in an expression of length 8: for instance, one way of inserting
brackets in abcdefgh is

(a(b(c(d(e(f(gh))))))),

and there are 428 other ways.


















161

4

TOPOLOGY AND GEOMETRY


1. Into the interior


Topology begins at Ujiji, on the continent of Africa, where for the
first time the interior was penetrated by the explorer Stanley. If it
were not for Stanley, we should be as ignorant of the interior as we
were then, and topology would be dark to us--as dark as the thickest
jungle of the dark continent.

Though many people are acquainted with the outlines of the story,
not so many know the important details. We give here only as much
of it as is necessary for our purposes. [34] Stanley penetrated so deeply
and became so lost that an American reporter, Dr L. I. Presume,
was sent to find him. The American, as he approached Stanley,
extended his hand stiffly according to the custom of the time--people
in those days were still rather Victorian. There were no herald and
no butler to announce his arrival. After all, one was in the bush.
The inhabitants of Ujiji (called Bajiji) spoke Mujiji, not English.
Having no-one to announce his name, he announced it himself: 'Dr
Livingstone I. Presume!' This piece of American resourcefulness has
gone down in history, and has become a famous quotation.

But it is not this for which we mathematicians remember Dr
Presume. Indeed, we break off the history of exploration and dis-
covery at this point, not even pausing to recount the later death of
Stanley at Ujiji, and his mummification in the rays of the all-purging

163

sun; not dwelling on the forwarding of his remains to the isle of
Britain, where the skies are continually overcast with cloud, or on
their consequent putrefaction and interment. These things need not
detain us.

Dr Presume, being American, is usually called by his first name
Livingstone. It is not generally known, except to topologists, that
Livingstone approached Stanley along a net. A naked white man,
speaking only a variety of Swahili picked up from the apes with whom
this strange being consorted, had at first attempted to teach Living-
stone to approach Ujiji along a sequence, swinging from point to
point on a vine. This was the white apeman's own customary mode
of locomotion, acquired also from his simian companions. But
Livingstone found that his ten-gallon Stetson hat kept falling off in
the breeze. That was when the birthday boy (whose name was
Tarson, and who is thought to be related to the Tar-baby of American
Negro folklore) strung a net over the trees for Livingstone to ap-
proach Ujiji along. As he dangled from the net, swinging along
by his hands like a circus performer, Livingstone stretched, and was
taller when he arrived at Ujiji than when he had set out. There was
more Livingstone than before. For this reason convergence along a
net is sometimes called Moore-Smith convergence. +

The concept of the interior, and how a point of the interior may be
approached, is one of the great starting-points of topology. We hold
these truths to be self-evident:

(i) The interior of everything is everything;
(ii) The interior of the interior is the interior;
(iii) The interior of anything is a part of that thing;
(iv) The interior of the intersection is the intersection of the
interiors.

Perhaps these wlll be made more clear by a word of explanation.
Number (ii) says that if one considers, say, Africa, and then the
interior of Africa, and finally the interior of the interior of Africa,
the last two are the same place. Number (iii) says that the interior
of Africa is entirely African, so that the Spitsbergen archipelago is
not in the interior of Africa. To understand (iv), think of the United
--------------------------------------------------------------------------
+ Taking Smith as a sort of generalized name for Livingstone.

164

Kingdom and also of the island of Hibernia. These intersect in a
place called Northern Ireland. Its interior is the interior of Northern
Ireland. The interior of the United Kingdom and the interior of the
island of Hibernia then must also intersect in the interior of Northern
Ireland. The conditions may be expressed by writing

(i) X = X
(ii) A = A
(iii) A (subset) A
(iv) (A (intersection) B)
= A (intersection) B;

if we write AB for A (intersection) B and 1 for X we see that these conditions
make sense on any monoid, and that they say that A -> A is an
idempotent endomorphism of the monoid such that AA = A. In
order that (subset) should be a partial order, however, it is reasonable to
restrict attention to commutative monoids in which every element is
idempotent. For technical reasons, it is sometimes convenient to
consider the case in which the monoid M satisfies the further con-
dition that if (a[i]) i (epsilon) I is a family of elements then there exists
an element a (epsilon) M such that

A~ a[i]a = a[i]
i (epsilon) I

and such that ax = a whenever

A~ a[i]x = a[i].

We write a = V a[i], and note that if it exists, a is unique. In this
i (epsilon) I
case, it is obvious that the submonoid T of M consisting of elements
invariant under a -> a is itself closed under V, and that of course
the restriction of to this submonoid is the identity. One calls T
the topology of M. Very often, one takes an even more special case: the
set of all subsets of a set X is a commutative monoid in which every
element is idempotent (the operation being (intersection) and the neutral
element of course X), and so all the above remarks may be applied
to it. If we take A to be (psi) if a <> X, we get the indiscrete
topology ((psi), X). A space in the indiscrete topology is like an enormous
room. Everyone is lumped together, and nobody is housed off. In fact, the
indiscrete topology is the most un-housed-off topology there is. There
is no separation, no seclusion, and so if one is living in such a space

165

it is impossible to do anything discreetly. Lovers find indiscrete
spaces very tiresome indeed. But they are a paradise for exhibitionists.

Considering our own terrestrial orb for the moment as X, we
see that Ujiji is in the interior of Africa, but that Africa is not the
whole earth. Hence our world is not indiscrete. This will be no sur-
prise to lovers, who find this world a very pleasant place indeed,
especially in the spring. Nor (to leap ahead for a moment) is it
discrete--and none but hermits would have it so.



QUESTION 27. Whether it is possible to use nets in order to approach a
point in a topological space?

Objection 1. Nets are merely glorified sequences. But approaching
a point along a sequence is like marching on a pogo stick--you may
never get near your objective. Suppose topological space to consist
of the ordinal numbers up to, and including, the first uncountable
ordinal (omega). The open sets are the intervals [(alpha), (omega)] with
(alpha) < (omega); alternatively, the interior of A is [(alpha), (omega)] if
(alpha) is the first ordinal less than (omega) such that [(alpha), (omega)]
(subset) A, and is (psi) if no such (alpha) exists. If a sequence
((alpha)[n]) approaches (omega) it is clear that

(alpha)[n] = (omega)

from some point onwards. In other words, it is impossible to ap-
proach (omega) along a sequence except by being at (omega) almost from the
start. But that would be cheating. Hence, sequences are inadequate
for a proper stalking exercise in this kind of country, and no ex-
perienced ordinal-hunter would use them.

Reply to Objection 1. It is true that nets are glorified sequences,
and that sequences are woefully insufficient in many spaces. One
weeps to think how hard it would be to approach one's objective if
there were nothing better than sequences to use. It would be like
lion-hunting Masai style, with spear and leather shield. Fortunately,
weapons have been devised that make lion-hunting as easy as picking
up a sleeping pussy cat; they are, in essence, merely glorified spears.
The projectile is thrown more forcefully and accurately, and from
a greater distance. Similarly, nets can do what sequences cannot,
despite the fact that they possess the same basic principle of design.

166

Note carefully that the use of nets in approaching or converging
to a point in a topological space has nothing at all to do with the
use of nets in Roman circuses by fighters known as retiarii. These
did not hunt lions primarily, as is commonly thought, but were pitted
against an armoured swordsmen with a fish on his hat called a
murmillo. The retiarii carried tridents, which are of almost no use in
topology. The use of nets for catching armoured swordsmen is by
now almost a lost art, whereas thousands of undergraduates an-
nually engage in contests to determine which of them can best stalk
a point in a topological space using nets. The losers are fed to the
lions, in a quaint survival of the ancient life of the arena.

Objection 2. To approach a point along a net, the net must be
finer than the neighbourhood filter at the point. When Livingstone
set off for Ujiji on the net which his friend Tarson had provided for
him, he must have realised that even in the neighbourhood of that
savage place, the neighborhood filter could not be so coarse as his
net. Of course in the nineteenth century it was difficult to obtain
really fine filters in such heathen places (that was the main reason
why missionaries were sent out, bearing the message of modern
civilisation and comfort); but even so, it was inherently unlikely
that the wildest cannibal could drink his coffee without removing
the grounds with any device more effective than a fish-net.

Reply to Objection 2. It is an abuse of language to say that a net
is finer than a filter. Only another filter can be finer than a filter.
What is actually meant by the phrase is that a filter associated
with, or generated by, the net in question is finer than a certain filter.
It is also a mistake to talk about nets and filters as if they had holes
in them; or as if nets and filters were distinguished according to
whether the holes were larger or smaller than the holes in cheese-
cloth. But it is true that filters are a more refined concept than nets.

By the way, another possible conclusion may be avoided if it is
pointed out quite sharply that filters, in our sense, have nothing
whatever to do with philtres; thus, when Tarson is feeling very melan-
choly he may make use of philtres in wooing his girl friend Jane, but
he never lends philtres to American reporters who wish to approach
her. Philtres are not very fashionable in topology, unlike filters. The
two must never be confused. Putting a philtre through your girl friend
is very different from putting your girl friend through a filter. Like

167

any other substance, when she passes through a filter she comes out
in a refined state; whereas when a philtre passes through her the
probable effect on her behavior is not likely to be one that could
be described in terms of greater refinement.


I ANSWER THAT approaching points in a topological space along a
net is indeed perfectly possible.

Sometimes it is even possible to approach a point along a sequence.
We all know about the thirsty frog who wished to jump down the
well, but found that his strength was fast failing him. The well was
two feet away. His first jump took him one foot, but his second jump
carried him only six inches. Indeed, he found that after each jump,
he had only enough power left to jump half as far. Thus the total
distance he had jumped after n jumps was

2 - (1/2)^(n-1) feet.

By taking more jumps than a certain minimal number he could
ensure that he would come as close as desired to the well, and
that he would ever afterward remain that close to the well. (He
could never actually reach the well; but if we are to do experiments
on animals we must first harden our hearts to the promptings of pity.
It is of no interest to the investigating scientist if the frog feels
terribly dry, if his skin and throat become parched and his muscles
fatigued. No doubt frogs cannot really feel pain. In any case, if the
frog fell in the well he would surely drown.)

In a more general topological space than the real numbers, our
frog might find it necessary to proceed along a net indexed by a
more general ordered set than the set of natural numbers. So as to
avoid needless complexities, we may assume that the supremum of
any two elements in this ordered set always exists. We say that the
net (x[i]) i (epsilon) I converges to, or approaches the point x if for
every neighbourhood N of x one can always ensure that a term x[i] of the net is
in the neighbourhood N by requiring merely that i is greater than a
certain suitable element i[N] of the ordered set I, which must depend
on N only. If by a tail of the net we mean a net (x[i]) i (epsilon) I' obtained
by restricting the net to a subset I' of I of the form I' = {i (epsilon) I:
i >= j}, where of course j may be any element of I, then the net converges to
x if and only if any filter that contains the images of all the tails of

168

the net must necessarily contain the neighbourhood filter. If (x[N])
N (epsilon) N~ is a choice function for the neighbourhood filter N~ at x,
and if this is ordered in the natural way, then this net approaches x.



QUESTION 28. Whether Africa is connected to America?

Objection 1. It would appear obvious that America cannot be
connected to Africa. America is a continent; i.e., a vast extent of
land that one may run through without crossing the sea. [35] But, if
America and Africa were connected, then one could run about all
over America, cross over to Africa, and then run all over Africa. If
one could do that, then the union of both America and Africa,
together with the bit of land that one ran along, would amount to
a vast extent of land, and one would have run through it without
crossing the sea. That is, the union of these three land areas would
be a continent. But there are exactly seven continents, and no one of
them is a superset of two others. Hence America and Africa are
not connected.

Reply to Objection 1. According to the definition used, it is not
clear at all that America is a continent. The word 'vast' is defined
[14] as 'Wast, huge, hurly; wheady, wide, broad and large, misshapen,
illfavoured, ...'. It is perhaps doubtful if all these adjectives apply
to America. If they do apply, to what other parts of the earth do they
apply? Most countries are rather oddly shaped, and seem large if
one walks over them. But the condition about 'running through' the
vast extent of land is even harder to justify. The only sensible mean-
ing that can be given to this is that one is able to follow a path that
crosses every point of the land in question. It would not appear to be
sufficient that one could find a path that would pass within a short
distance of every point of America. The topological properties of
America are little known: is it, for instance, an open set? Is it
compact? Is it even a Borel set? And so it would appear difficult
to settle the existence of a continuous surjection [0, 1] -> America.

But let it be granted that America is a continent. Then Eurasia,
by the argument presented in the objection, is as much a continent as
any other. Hence a continent may indeed be a union of two others,

169

and the conventional total of seven continents is very wide of the
mark.

Objection 2. America is not even connected to itself, since 1914
when the Panama canal was completed, severing the North from the
South and putting asunder what was meant to stay together. How
then can America be connected to something else?

Reply to Objection 2. If X (union) Y and Y (union) Z are each connected, and
if Y (in this case, Africa) is not empty, then it is elementary that X (union)
Y (union) Z is connected, though X (union) Z may not be. But Africa has
at least one point; namely, Ujiji. Hence we cannot infer from the
argument in the objection that America and Africa are not connected.


I ANSWER THAT Africa and America are not connected. We may
safely take the surface of the earth to be a sphere S<2>. (This is known
to be only approximately true. We must first break down all natural
bridges, which of themselves make the earth a surface of high genus.
It is not known to the author how numerous these bridges are in
various parts of the earth, and they will present a difficulty if there
are any in certain parts of the planet, especially Greenland, Antarc-
tica, and the Eastern tip of Siberia, and certain small islands. The
restoration of the natural bridges, and the proof that even after
the restoration the Americas are still not connected to Africa, are
tasks left to the reader as exercises.) At least, we shall take the surface
of the planet to be a homeomorph of S<2>. Now it is well known that
neither the north pole v nor the south pole (sigma) lies in America or
Africa. Include the punctured sphere S<2> ~ {v, (sigma)} as the unit sphere in
R<3>; circumscribe about it the cylinder S<1> x (-1, 1) (subset) R<2> X R
= R<3>, which has the line through v = (0, 0, 1) and (sigma) = (0, 0, -1) as
its axis, and project the punctured sphere outward from the axis on to
the cylinder. The homeomorphism just described is, of course,
nothing but the usual way of defining the equiareal projection, a
kind of map. Mercator's projection would have done as well, but is
more difficult to describe. Now identify the punctured sphere with
the cylinder. Choose points (gamma) = Porto Praia in the islands of
Santiago, of the Capverdian chain, and (alpha) = Abkit or Anadyr (take
your choice). Then if (alpha)[1], (alpha)[2], (gamma)[1], (gamma)[2] are
points a little east and west of (alpha), (gamma), and if the projection on
S<1> sends these to a[1], a[2], c[1], c[2], a continuous function

170

S<1> -> [0, 1]

exists sending

a[1] -> 0
a[2] -> 1
c[1] -> 1
c[2] -> 0

and taking values that depend linearly on arc length over any interval
not containing any of these four points. (By an interval on S<1> we
mean a proper connected subset containing more than one point.)
This fact is not difficult to get, and can be obtained by noting that
S<1> is [0, 1] with 0 and 1 identified. It may be taken as a fact of
geography that by the composition of

S<2> ~ {v, (sigma)} -> S<1> X (-1, 1) -> S<1> -> [0, 1]

we have sent all of Africa to 0 and all of America to 1. Restricting
to Africa (union) America and corestricting to {0, 1}, we have a surjection.


2. The insides


People often talk as if they had insides. When a man is feeling queasy,
uneasy, under the weather, when he suffers from neuritis or neuralgia,
he may say mournfully, 'Oh! There is something the matter with my
inwards; I am not quite right inside; I am the victim of intestine
strife.' What rubbish! How does he know that he has any insides?
Just because human beings ordinarily possess livers, guts, gall-
bladders, and other such slimy organs, does not go any distance
toward proving that these things are inside the human beings. In
many fairy stories, people's insides are represented as being outside
the people. A little boy may be entrusted with the care and keeping
of his mother's heart, and may carry it about with him. How do we
know that this is merely an extraordinary fancy? Why do we so
easily suppose that it cannot happen to us? Perhaps everyone's
heart is outside him. Perhaps my heart is outside me, and my fingers

171

and eyes, the paper I am writing on, the reader and the book he is
reading are all inside me. Or perhaps the book and I are inside the
reader, and his stomach is outside him. Who can tell?

But even worse, why do we suppose that there is a distinction of
the world into two parts, one outside us and one inside? Ah, the
reader may say, now there you go too far. I am willing to entertain
the ridiculous notion that my insides are really outside me, and my
outsides are inside me. But now you ask me to think that there is no
difference between the inside and the outside; that my stomach and
the book I am reading are both on the same side, the only side;
that there are not two sides but one side. It is too much. I cannot
imagine it.

Yet, that is the situation we are faced wth. It is not asserted that a
human being has not got an inside and an outside; all that is asserted
is that there is a problem. And it is a topological problem. Now a
human being is topologically a difficult object to study. He is always
moving about; that is, he is constantly changing shape. A man with
his fingers in his ears has two more handles than a man standing
with arms and legs outstretched. If the alimentary passage is not
blocked, and if we make certain other simplifying assumptions, we
may consider the man to be composed of the skin and the lining of
the oesophagus, stomach, and intestines. It must be emphasised that
in this view, the bones, muscles, brains, etc. are not part of the man
thus considered. He is a surface; and it would help our deliberations
if he would be so kind as to keep open his mouth and his anus. The
breezes must pass freely through the tunnel. Then a man is a torus;
to be exact, he is a two-torus T<2> = S<1> X S<1>. He is the product of two
circles. More than that, he is a torus in three-space R<3>. Has he got an
inside and an outside? To put it another way, if we consider all the
rest of the universe besides the man, which consists of the guts, heart,
liver, brain, blood, etc. of the man together with the Eiffel Tower, the
Big Dipper or Ursa Major, Grant's tomb, and whatever else exists
besides the man himself--if we consider all that, is there a sensible
way of dividing it into two persons that shall be separated from one
another by the man? Is there a division of the waters under the firma-
ment from the waters above the firmament? Is the liver inside the
man and the Eiffel Tower outside--or vice versa? Later we shall
return to this question. Before we go on to simplify the problem

172

sufficiently to make it tractable, let us pause to consider a related
question, namely that of a man who has closed his mouth and his anus.
This type of man is a sphere, as we shall see later. What about him?
Has he got an inside and an outside?

Here is an easier problem:



QUESTION 29. Whether a circle has an inside and an outside?

Objection 1. Consider a hoop or loop floating in space after being
jettisoned by a lunar module. The hoop is a circle, but it is not a
subset of any particular surface. The remainder of the universe, after
the hoop has been deleted, is connected. Hence the circle has no
inside and outside.

Reply to Objection 1. It is necessary to stipulate that the circle lies
in some surface.

Objection 2. Consider a circle drawn on an inner tube, of the sort
used to inflate the tyres of automobiles. If the tube is cut along the
circle, it may or may not fall in two pieces. For a very small circle,
the tube will fall in two pieces (Figure 5a), but for a circle drawn
right round the tube, when the tube is cut it remains in one piece
(Figure 5b). The second kind of circle has no inside.


_________________ _________________
/ \ / / | \
/ _______ \ / \__|___ \
/ \/ \/ \ / \/ \/ \
\ \_______/ / \ \_______/ /
\ | A | B / \ /
\____|___|________/ \_________________/

(a) (b)



Figure 5



173

Reply to Objection 2. How true! It is clear that

(S<1> X S<1>) ~ (S<1> X {a}) = S<1> X (S<1> ~ {a}),

which is homeomorphic to S<1> X (0, 1), and the natural equivalence
of {0, 1} with ({0, 1}) proves (via ({0, 1}) = {0, 1} =
{0, 1}) that the product of connected spaces is connected. Hence if
a circle is to have an inside and an outside, it must not be drawn on
an inner tube.

Objection 3. The Equator is a circle running round the middle of
the earth. The Arctic Circle also runs round the earth, but not round
its middle. It is usually held that Santa Claus lives inside the Arctic
Circle, and that the Patagonians live outside the Arctic Circle. But
on which side of the equator does Santa Claus live? On the inside,
with the Patagonians outside, or is it they who live inside the
Equator?

Reply to Objection 1. There are many ways to answer the question.
Some people would say that if you walk round the Arctic Circle
going eastward then Santa Claus lives on the inside, but if you walk
round it going westward then he lives on the outside and the Pata-
gonians live on the inside. This is doubtless true, since anyone who
begins to walk round the Arctic Circle will certainly drown, freeze,
or be eaten by polar bears long before he finishes the journey. Other
experts hold that Santa Claus does not exist, but does live inside the
Arctic Circle. Still others would say that there are two sides of the
Arctic Circle, but that it is impossible to tell which of them is in and
which is out. This is the best answer, together with the stipulation
that the earth is a sphere [9] and that it is required to draw the
circle on a plane, and only on a plane. Note that the sphere, being a
closed, bounded subset of R<3>, cannot be a plane R<2>, since the latter
is not compact.



I ANSWER THAT a circle drawn in a plane has an inside and an outside
(Figure 6). This is sometimes hard for people to understand, for
various reasons. Some people think that if the circle is to have an
inside and and outside, there must be a door. Houses have insides, out-
sides, and doors; if the circle has an inside and an outside, where is
the door? Before we proceed to show that a circle has an inside and

174




______
/ \
/ \
< inside > outside
\ /
\______/

Figure 6




/
/----/ ---\ /-----------\
| | | |
| | | |
| | | |
| | | |
| | | |
\-----------/ \-----------/
(a) (b)

Figure 7



and outside, we had better dispense with this difficulty; otherwise, the
reader's mind may be insufficiently prepared for the truth, and unable
to assimilate it fully. In Figure 7 we have two houses. House (a) has
the door open; house (b) has the door shut. We are looking at a
floor plan. The reader should see immediately that house (a) has no
inside; it is porous, permeable, open; anybody can walk through the

175

door without committing a felony. There is no separation. The
walls and door of the house do not divide the world into an inside
and an outside. Now consider house (b). The door is shut, and that
is why it does not show in the figure. We shall return to this point
presently.

Now for simplicity in proving the circle has an inside, let it have
radius 1. Let P, P' be any two points of the plane, and let r, r' be the
distances |OP|, |OP'|. Since one side of a triangle is shorter than
the sum of two others, |r - r'| <= |PP'|, which shows that the
mapping

R<2> -> [0, (infinity))

defined by

P -> |OP|

is continuous. Since

| r r' |
| ----- - ------ | <= |r - r'|,
| 1 + r 1 + r' |

the mapping

f: [0, (infinity)) -> [0, 1]

defined by

r
r -> -----
1 + r

is continuous; notice that f is also strictly increasing, and that
f(1) = 1/2. If S<1> is our circle, we get a map

R<2> ~ S<1> -> [0, (infinity)) -> [0, 1].

By corestriction, we may get one

R<2> ~ S<1> -> [0, 1/2) (union) (1/2, 1]

and composing this with the obvious map

[0, 1/2) (union) (1/2, 1] -> {0, 1}

we are done with the major part of the problem; namely, to show
that the circle divides the plane into two pieces. Note that the same
argument works on the sphere, so that it is true that the Arctic
Circle (or the Equator) divides the surface of the globe into two parts.
This is because the sphere can be regarded as the one-point compacti-
fication of the plane, and all maps extend suitably. But beware! As

176

we saw, it is not enough in the case of the sphere to show that there
are two pieces, or components, because even so one is unable to
distinguish which of the pieces is the inside and which is the outside.
Can we do so in the present case?

Consider the two pieces. The piece containing O (the centre of the
circle) will be shown to be the inside, and the other piece will be
shown to be the outside. Thus, among other things we shall learn
that the centre of the circle is inside the circle. Write C[O] for the com-
ponent of the plane (with the circle taken out) that contains O, and
____
C[O] for the union of this set with the circle itself. Then it is to be
____
proved that C[O] consists of all the points either inside the circle or on
the circle: in any case, it is known that

____
C[O] = {P: |OP| <= 1}.

The other component forms in a similar way

_____________
C[(infinity)] = {P: |OP| >= 1}.

and it is to be shown that this is the set of all points either outside the
circle or on the circle.

If P is a point of the circle and if 0 <= r <= 1, there is exactly one
point P' on the line segment OP at a distance r from O; this provides
a mapping

____
S<1> X [0, 1] -> C[O]

which is easily verified to be a continuous surjection. By Tychonoff,
____ _______
it follows that C[O] is compact. For n (epsilon) N[0], let B[n](O) be the
closed ball of radius n at O. Then

____________ _______
C[(infinity)] (subset) R<2> = U~ {B[n](O): n (epsilon) N[0]},

_____________
but C[(infinity)] is not a subset of a union of any finite subcollection of
this collection. Hence there is an essential difference between the com-
ponent containing the centre of the circle and the other component.
It is therefore reasonable to call one of these components the inside
of the circle and the other the outside. For some reason, convention
has settled on the term 'inside' for the component of the plane, when
the circle has been cut out, that has the centre of the circle in it. Like
all conventions, this one is not easy to explain rationally. Why should
not the centre be said to be outside the circle, and why should not we

177

designate all the points of the plane very far from the centre as points
inside? In other words, why is the conventional terminology not
the one shown in Figure 8? There appears to be no satisfactory


______
/ \
/ \
< outside? > inside?
\ /
\______/


Figure 8


answer to this question. There is no real reason why one should
prefer to say that the centre of the circle is inside, and not outside.
However, if most people prefer to say that the centre is inside the
circle, it is good manners to acquiesce; nothing is ruder than to find
occasion for dispute in every inconsequential alternative. Really it
would be best if we could all agree to speak of the compactside and
the countercompactside of the circle. Thus, a man with indigestion
could say, 'I've got to see a doctor. There's something seriously
wrong with me compactside.' An economical housewife buying a
chicken might say to the butcher, 'And please could I have the com-
pactsides for the cat?' A prisoner in a maximum security institution
might mutter to his compactmates on the night before he expected
to get 'sprung': 'Youse won't see me tomorrow. Tomorrow I'm
gonna be countercompactside.'

But not so swiftly! So far it is only the circle which one knows to
have a compactside and a countercompactside (or an inside and
an outside, to use the terminology of stodgy convention). Prisons,
chickens, and men are much more complicated things, by their very
nature. It is now time to consider whether men have insides. Chickens
can be taken to be pretty much the same for our purposes as men;
certainly, in a mathematical discussion, a chicken is often as much

178

use as a man and makes comments of about equal intelligence. As
we know, there are really two very distinct types of human being:
mouth and anus open, and mouth and anus closed (Figure 9). (The
acute reader will note that it is possible, in general, to consider the
further cases that occur when one of the apertures of the alimentary


_______
/ o o \
/ @ * @ \
<={ --- }=>
\ |=o=| /
\__:|___/
(a)

_________________
/ o o \
/ @ ___*___ @ \
/ \/ \/ \
\={ \_______/ }=/
\ |=o=| /
\_______:|________/

(b)



Figure 9. Two essential types of men.


passage is open and the other closed. It is usually difficult to dis-
tinguish one of these cases from the other, especially in political
oratory; the resulting surface is in both of these cases topologically
the same. Technically, it is an identification space of a torus. (Only
one of the possible ways of including such a topological space in R<3>
is capable of being human; the other is the skin of a ring sausage
(Figure 10). Neither the ring sausage nor the orator is a manifold,
and the difficulties of this case force us to ignore it.)

179



,--.-.----.
/ :__)_ \
|___/ \ \
( \ \ _\
\_.'\ ____)/ )
\___\/ `. /
`------' ______________
(a) /------. _____\
/ \/_ \
\........'`.|..../
\_______`._|___/

(b)





Figure 10. (a) Ring sausage; (b) political orator.


The first essential type of man is the Quiet Man. His motto:
Silence is Golden. He is not given to flatulence or loquaciousness.
His sign is the Sun. The second essential type of man is the Loud
Man. His motto: Speech is Silver. He is flatulent, loquacious, and
given to overeating. He thinks and acts little, but speaks much. His
sign is Saturn, the planet with rings. Most men oscillate between
these two types.

Loud Men, abstractly, are capable of much more interesting com-
binations than Quiet Men. They may be linked together in chains. A
simple configuration of this type is shown in Figure 11.

Because of the theoretical possibility of this configuration, biolo-
gists have searched assiduously for an example. Their quest has been
fruitless, so far as authenticated records show. Loud Men, as such,
exist in abundance; linked Loud Men appear to be very rare. Why

180


###
/__/
&& < <
| \_____________> \___
|__|____ | |@|
| | |o| ###
### _|_|_________|_|_ |_|
\__\___/__@__o__*__o__@__\__| |
\______ | _____|
| :| |
/_________:|________\
___ / __________________ \
&__/\/_/ | | | | \ _\
__\/ / /|_|_________|_| | |__
&____|\_/ / < < |____&
\___/ \ \
\ \
(a) ###



(b) ________________
___/ ________ \
/ ______/ __#_ \ \
| / ___/ o/ \ \ \ |
/ / / | | \ \ | &| |
| \# \__| _|_/ #/ | | |
\ \___o___o___/ | &| /
\___ \ o\____/ / /
\__ \____#___/ /
\__________/



Figure 11. (a) Linked Loud
Men; (b) the same, simplified.

181

is this? Topological biology seems to have provided at least a partial
answer. We know from modern embryology that all men begin as
Quiet Men; that is, a man is at first a tiny sphere in the womb of his
mother. The potentiality of being Loud is developed somehow in
utero. If it can be shown that two fully developed Loud Men, not
already linked, can only become linked by a major surgical operation,
then the question, Why do we see no linked Loud Men? will have
been reduced to the two further questions: Why are twin Loud Men
never born linked; and, Why do no Loud Men choose to undergo a
surgical operation so as to become a pair of linked Loud Men?



QUESTION 30. Whether two men who customarily maintain the alimen-
tary tract in an open and ventilated state, by never shutting the mouth
or anus (hereafter called Loud Men), can become linked like the
links of a chain--so that each man passes down the oesophagus of the



____________
/ ______&_ \
/ / _______\ \_____
/ / / ______\ \_ & >
\ \/ / \ / /
\ / / \/ /
// / / /\
/# / \ / / _\
/ /\ \______/ / \ #\
\ \ \__________/ / /
\ \ / /
\ \___________/ /
\_@__o__*__o__@_/


Figure 12. Knotted Loud Men.



182

other, out through the vent, and back again--without a major
surgical operation?
(Note: It is helpful to refer to Figure 11.)


I ANSWER THAT two Loud Men, not yet linked, can only become so
through surgery.

The two Loud Men, not yet linked, will for simplicity be taken to
be geometrical tori. (This is a big assumption; in particular it assumes
that no Loud Man is tied in knots, as in Figure 12.) Similarly for the
linked Loud Men. The beginning and end of the hypothetical linking
thus appear as in Figure 13.

Having made this assumption, it is possible to define a belt of a
Loud Man. The man is generated as a surface of revolution by
revolving a circle, not intersecting the axis of revolution; the circle


___________
/ _______ \ ________
/ / \ \ / ____ \______
\ \_______/ / / / / \ o\ __ \
\___o___o___/ | | / / | | \ \
M[1] | | \ \_|__|__/ /
___________ | | \___o___o___/ M[2]
/ _______ \ \ \____/ o/
/ / \ \ M[1] \________/
\ \_______/ /
\___o___o___/ After
M[2]
Before



Figure 13. Hypothetical linking of Loud Man.




183

traced by any point of this circle as it is revolved about the axis is a
belt of the man. It will be assumed that, after linking, the belts are
in their right places again; i.e., that they are still belts. It can be shown
that this assumption is fair. It can also be shown that it can be assumed
without loss of generality that the whole procedure whereby the two
men are linked can be carried out without moving or changing man
M[2] at all. He can remain fixed while M[1] is twisted and tortured.

Suppose it were possible to link the two men without cutting,
tearing, or any other surgery. Take C[1] and C[2] to be belts which the
two men are wearing. Then the belts have also been linked without
surgery, and indeed without unbuckling them. The question has been
reduced to this: Can a belt be linked to a stationary belt without
breaking or unbuckling? By hypothesis, we have a homotopy

C[1] X [0, 1] -> R<3> ~ C[2]

in which the image of C[1] X {0} is a geometric circle not linked with
C[2], and the image of C[1] X {1} is a geometric circle linked with C[2].
If Q is the centre of C[1] X {0} in R<3> ~ C[2], then taking Q as origin of
coordinates and 1, say, as the radius of C[1], the map

(x, y, t) -> (tx, ty)

makes C[1] X {0} homotopic to {Q} in R<3> ~ C[2]. Hence there exists a
homotopy shrinking C[1] X {1} to a point. Taking cylindrical co-
ordinates in R<3> ~ C[2], with the axis of rotation about which C[2] may
be generated by rotating a point taken as the z-axis, we note that

(r, (theta), z) -> (r, z)

is a continuous function from R ~ C[2] to [0, (infinity)) X R ~ {(1, 0)}. Com-
posing this map with the homotopy that shrinks C[1] X {1} to a point
in R<3> -> C[2], we infer that a circle can be shrunk to a point in a plane
punctured at the centre of the circle; which is as much as to say that
the identity map

S<1> -> S<1>

is homotopic to a trivial map. But it is well known that the relevant
group is Z, and is generated by the homotopy class of the identity
map.

This conclusion, that Loud Men cannot be linked without surgery,
is one of theoretical interest, but it has for most of us little practical
importance. Two linked Loud Men would no doubt be as loud as

184

two unlinked ones, and no louder. If there is any practical inference to
be drawn from the theoretical fact that Loud Men cannot be linked,
it is one which brings a comforting thought to the mind, such as it is,
of every member of the Loud Fraternity. Since he cannot be linked,
the danger is averted that (relapsing momentarily into quiet and
closing his mouth) his link-mate might bite him through from mouth
to anus. Such a calamity is then impossible. If it could happen, it
would clearly change the toroidal Loud Man into a very quiet,
spherical Quiet Man (Figure 14).

What looks like a halo is the toroidal soul ascending whence it came.
His grave must have a marker: let us put, 'May he ever rest in peace.'


__________
/__________\
((__________))
\__________/

_________
/ -- -- \
/#-___*___-#\
/ / |:::| \ \
\ \ !|:::|! / /
\ \!|:::|!/ /
\&!_____!&/
! !


/---\
|:::| Endoderm
\---/

/---\
| ! | Mesoderm
\---/


Figure 14. Transformation of toroidal Loud Man.



185


3. Geometry


Civilised man lives close by the brink of the water. To the great
civilisations that developed along the banks of the great rivers--the
Indus, the Euphrates, the Cam--can be traced much of that artistic
and scientific heritage which adorns modern life. The earliest
recorded discovery of a useful art occurred in Eden, near the head
of the Euphrates. This was the discovery of both nakedness and its
complementary opposite, clothes; both of which arts remain im-
portant industries in the great cities of today. The art of falling was
discovered by a river-valley man, Isaac Newton, although the actual
discovery took place at Grantham, Lincolnshire, and not at Cam-
bridge (the centre of the great Cam civilisation). This event, known
as the Fall of the Apple, occurred much later than the other, the
Fall of Man. Before either of them occurred the Fall of Lucifer,
which was not an art, though it became the subject-matter of several
great works of art. From the Fall of the Apple was developed the
great art of Potential Theory.

In the valley of the Indus a great civilisation developed, to which
we owe Sanskrit and the Indo-European language. The most impor-
tant word in this language is lox, or, if you prefer, Lakshmi. She is
the goddess of smoked salmon and bagels. A bagel is a torus, and
has been encountered already in the chapter on topology. It is eaten
with lox, and is a topological group R<2>/Z<2>. Along with lox, bagels,
and language came the great Aryan race; to them we certainly owe
the discovery of the square. It has long been unknown whether the
Aryans developed the torus, or bagel, from the square, or vice versa.
Which came first? Square enthusiasts point to the shape of the
Aryan head as an indication that they must have been forced at an
early date to contemplate this figure. At first they would have found
it natural to bake their bread in this shape; later, because the torus
is obtained by an identification of opposite edges of a square, it
naturally occurred to their rather simple, straightforward minds to
make bagels, or toroid bread as a kind of elegant variation. That is
how one segment of opinion reconstructs the story, which must
nevertheless in its more intimate details remain dark for us. It is

186

noteworthy that one branch of the Aryan race, the famous Aryan
heretics, persevered in making and eating exclusively square bread,
and eschewed bagels. (Bagels are hard to chew, but it is foolish to
eschew them). The Aryan race were also interested in the wheel,
which they called a wheel-wheel, whirr-whirr, a GLGL, a quirquel, or
a circle. It is possible that the wheel was developed from the bagel;
that when the early Aryans began to think about the bagel, they
realised that it was a compact topological group and began searching
for other compact topological groups, and that this search brought
them on the trail of the wheel, or circle. It cannot be overlooked that
others derive the bagel not from the square but from the circle.
When the Aryans heard of Tychonoff's theorem (the product of
compact spaces is compact) they applied it to the wheel, and got a
new compact group:

BAGEL = WHEEL X WHEEL.

So much for the Indus Valley and the immense heritage we possess
as its successors.

The Yellow River civilisation gave us the I Ching, or index of
permutations, from which developed the symmetric group on n letters
and the non-conservation of parity, and was responsible for the
Chinese Remainder Theorem, space travel, dragons, Pekingese dogs,
and fortune cookies.

The Nile Valley has a civilisation so ancient that the name given
to the process of suspended animation discovered on the banks of
this river is mummification, and the subjects of it, who were provided
with picture-books of instructions and light entertainment against
their future reanimation, were designated as mummies. These names
make it absolutely certain that the Nile civilisation dates from before
the Fall of Man, when the distinction between mummies and daddies
was discovered. Pyramids, another Egyptian invention, were prin-
cipally a tourist attraction, like the Eiffel Tower. They were copied
in America, where they were used for heart transplant operations,
flower shows, and football games. Pyramids were also influential in
the history of Greek geometry. Among all the ideas of the Nilesiders
--mummies, men with bird's faces, cats with women's faces, the
feeding of non-mummies to crocodiles--only the pyramid is of
mathematical significance, and this only through the Greeks.

187

The Greeks were a Hellenic people; hence, although they were
good at mathematics, poetry, politics, architecture, navigation, and
many useful arts, they were overly fond of beautiful women. Fond of
measurement, they measured feminine pulchritude in Hellens. Rather
oddly for a civilised people, they based their unit of measurement on
the curious idea that the more beautiful was a woman, the larger
was the number of ships her face could launch, and that the largest
number of ships that could be so launched was 1000, or 10^3. This
quantity of beauty was called one Hellen. The practical result of
this idea was that the Greek women spent much time on the seaside
toughening their facial features by pushing boats into the water,
and in the end the Hellenes lost interest in women and invented
Platonic love. At the same time, the huge fleet of ships thus launched
made the Hellenes a nautical nation, so that they did much sailing.
The sailors would often bring home little trinkets, among which were
small replicas of the great pyramids, something like little Eiffel
Towers cast in lead that visitors bring home from Paris as souvenirs.
These knicknacks had far-reaching effects on later Greek thought, as
we shall see. Most important of all, doubtless, was the simple habit
of spending time at the beach. At first one wandered down to the
beach for the idle amusement of watching the ladies at their competi-
tive launching. But the men, to pass the hours, would chat lying on
the sand, and scratch crude drawings in its smooth surface. Later,
when Platonism had replaced the mythical Hellen at the centre of the
Greek imagination, and feminine beauty no longer played the leading
ro^le, the beach remained a gathering point. A preoccupation with
form, with the shape of things, remained; but now it was abstract
forms that were contemplated, and geometric shapes were drawn in
the sand. Geometry dawned. Her day is not yet spent.

One of the most difficult of the geometric concepts invented by the
geometric Greeks was the geometric line. A line was said to have no
thickness or breadth, and to be generated by a moving point. A
point was said to have no length, breadth, or thickness, and to be
indivisible. Thus, we are forced to interpret a point as a singleton
{x}; i.e., a universally attracting object in the category of sets. That
is simple enough. But how are we to interpret the line?



188

QUESTION 31. Whether, if three distinct points lie on a line, one of them
must lie between the other two?

Objection 1. A point has no extent; i.e., no length, breadth, or
thickness. That which has no length obviously cannot lie, sit, or
stand. Hence, a point cannot lie on a line.

Reply to Objection 1. It is quite clear that what is meant by the
statement, 'a point lies on a line', is not that the point reclines there,
but that there exists a monomorphism from the singleton to the line,
and that this monomorphism is among those considered to be inclu-
sions. Alternatively, there exists a map f from the line to itself such
that for any map g from the line to itself

fg = f.

It is merely another way of saying that the relation of incidence
holds between the point and the line.

Objection 2. This would seem obvious. By the Krein-Milman
theorem, the closed convex hull of {P, Q, R} being clearly compact
and convex, it is the closed convex hull of its extreme points. But
since only two of P, Q, R can be extreme, the other is in the closure
of their convex hull--which is as much as to say, in their convex
hull itself.

Reply to Objection 2. It is true that by the Krein-Milman theorem,
each of P, Q, R lies in the closed convex hull of the extreme points
of the closed convex hull of {P, Q, R}; but the objection assumes
without proof that this set can have only two extreme points. If it
has three, the argument fails.

Objection 3. You could try moving the points around. Keep two of
the points fixed firmly in their right places, but let the third point
move freely on the line. If the third point is between the other two,
it cannot go on for ever, either in one direction or the other, without
bumping into one of the other points; if it is not between the other
two, then the moving point can go on for ever in at least one direc-
tion, and never encounter either of the other points. Let one of the
points, which we may call P, be supposed (as a hypothesis) not to
lie between the other two. Move P toward Q with a large velocity,
so that when P strikes Q, the particle Q will continue in the same
direction with the same velocity. If Q never strikes the third point R,

189

then the whole line has been swept out without encountering R,
and hence R does not exist. It follows that Q is between P and R.

Reply to Objection 3. The objection is stated in terms with a
pseudo-particle-dynamical flavour, but we need not insist on look-
ing at the objection in this light; if we did, it would of course have
to be rejected as non-mathematical. The idea of a moving point is a
highly sophisticated one by the time it is given a mathematical
expression. The billiard-ball behaviour of Q when struck by P in-
volves differential equations, and the points are even treated as
impermeable to one another. Worst of all, the objection assumes
without proof that a line has only two directions. Just what is the
justification for this bold assumption?

I ANSWER THAT it is the case that if three points on a line are distinct,
then one lies between the two others. The general idea of the proof
is very simple. Since P, Q, R (the three points) lie on a line, we may
draw arrows from P to Q, from Q to R, and from R to P. It is
necessary to show that either two of these arrows go to the right,
or two of them go to the left. Since the line is an affine space, there
exists a way of associating to each pair of points a vector or arrow:
to (A, B) we associate the vector AB. Since the line is by Euclid's
definition a one-dimensional affine space, the vectors in question
lie in a one-dimensional space--i.e., module--over a field of charac-
teristic 0--i.e., characteristic (infinity)--which possesses an archimedean
order. (It was the great Eureka who discovered that lines have got
to have an archimedean order. This was on one of the many occa-
sions when Eureka was singing in the bath. Since he had very bad
pitch combined with good volume, Eureka's singing was always in-
distinguishable from his shouting.) We cannot be sure, and need not
worry about, just which archimedean ordered field F is being used,
but since it contains (root)2, it cannot be the universally repelling archi-
medean ordered field Q, though it certainly contains Q. Since
PQ + QR + RP = PP = 0 by the axioms of affine spaces, and since
the vector space, being a one-dimensional free module over F, is
isomorphic as an abelian group to the abelian group structure of F,
the fact that none of PQ, QR, RP is 0 enables us to conclude that by
the use of a suitable permutation (in the symmetric group on
{P, Q, R}) it can be arranged that two of PQ, QR, RP are positive

190

without loss of generality. These can be PQ, QR, so that Q is
between P and R.


(Tri)gonometry

One of the most important elementary subjects under the general
heading of geometry is gonometry. The origins of trigonometry,
which is the most elementary form of gonometry and is followed by
tetragonometry, pentagonometry, hexagonometry, etc., are to be
found in the great Ur-city, Ur. The Ur-civilisation was a great river-
civilisation, and invented trigonometry, which is therefore a useful
art. Its citizens, as we know from plastic art, wore beards in the
shape of rectangular parallelepipeds, and since Ur was situated on
the Euphrates, we know that they wore clothes and divided their
children into girls and boys. The boys had beards, and the girls had
none. The great interest in triangles that was such a great part of the
cultural life of Ur is evidenced by their writing, in which all letters
were formed of triangles. This is called cuneiform. Even little bearded
boys of the lowest social orders ran about the streets of the great
Ur-city studying triangles. Because of their characteristic beards,
these children were called Ur-chins. Nowadays, the name remains
despite the fact that few Urchins any longer possess the beard to
which it refers.

Clearly, if one is to do gonometry, which is the measurement of
angles, one requires some means of measuring angles. Hence the
question whether the Ur-menschen could have constructed a pro-
tractor was long a thorn in the side of the historian (see Figure 15).



QUESTION 32. Whether the Ur-menschen (as the inhabitants of Ur
were called) could have constructed a protractor?

Objection 1. It is well known that the Ur-menschen counted in
sixties. Since it is easy to construct an angle of 60 degrees, they
would naturally have done so, and this angle would then have been
divided in sixtieths, or angles of one degree. But an angle of one
degree is the basis of the protractor. Hence the Ur-menschen con-
structed a protractor.

191

Reply to Objection 1. This merely shows that it would come
naturally to the Ur-mensch to construct a protractor. But an Ur-
mensch is far from being a Natur-mensch. The latter was not invented
until all the Ur-menschen had disappeared. Hence the intersection
of Ur-menschen and Natur-menschen is empty, and no Ur-mensch
always acted naturally. Hence it is possible that no Ur-mensch
constructed a protractor.



\|______
| _/.|
| |@ > _ _
|_/ (__ \\//
\ __| |##/
_|_|____/ /
| || __/
|/_/|||
/ /_|||
\ \/,-|
|\##:_|
| |\ \
/ / \ \_/\
|___> |___/



Figure 15. Ur-chin, studying angle.



I ANSWER THAT no man has at any time constructed a protractor.
The cosine of 40 degrees obviously satisfies

8 cos^3(40 degrees) - 6 cos(40 degrees) + 1 = 0,

which clearly makes the construction of a protractor an impossibility.

192

How then did the men of Ur measure their angles? Possibly they
knew that one can integrate

x dt
(integral) ---------------
1 (root)(1 - t^2)

for 1 >= x >= -1, and that when this (obviously continuous mono-
tonic) function is inverted, it is capable of providing material for a
homomorphism from the topological group of real numbers to the
(multiplicative) topological group of non-zero complex numbers
whose image is the circle. If so, the measurement of angles would
become a triviality for them. But there is no record of a theory of
divergent Riemann integrals among the Urmen, and perhaps they
did it via Taylor's series.

The line, the circle, the angle--to these basic ideas of geometry
must be added a fourth.


Area of a rectangle

Almost everything is made in the shape of a rectangle. Two of the
most obvious examples are a book and a farmer's field. Why this
should be is difficult to say, but it is worthy of attention that the
farmer ploughs his field in furrows, and that the writer writes and the
reader reads the book in lines, which are very like furrows. The
author has never understood why books are not written even more
plough-fashion than they are. Why not write like this?

Oh, what are the zeros of zeta of s?
!ssefnoc ylurt I, uoy llet nac ydoboN

It would be so much easier to read--the eye would not travel back
to the beginning of the line emptyhanded, as it were, but would get
some reading done on the way back. If farmers ploughed their fields
as inefficiently as we read, we should all likely starve. However that
may be, the consideration of furrows only partly helps to explain
why everything is a rectangle. All it explains is why everything is a
product of something times a unit interval. Perhaps, in some dark
half-conscious way, the farmer is trying to prove that the fence on
one side of his field is homotopic to the fence on the other. Who
knows? In any case, we have not explained why his field is not an
infinite cylinder, or why this book is not a punctured disc.

193

Why is it that when a man has a rectangle, the first thing he does
with it is to calculate its area? Is it because he knows how to get
the right answer? That hardly seems likely.



QUESTION 33. Whether it be true that the area of a rectangle whose
sides have lengths B and H respectively, is the product of those
numbers, BH?

This is one question of which every intelligent person ought to be
able to guess the right answer, at least after he has read the correct
definition of 'rectangle'. A rectangle is a geometric figure composed
of two pairs of parallel line segments, perpendicular to each other,
such that each line segment meets each of the perpendicular line
segments in its endpoints. Basically then, a rectangle is made up of
four lines. And yet people say that the area of a rectangle is the
product of base times height!

I ANSWER THAT the area of a rectangle is never the product of base
times height. This is because the base of a rectangle is the length of
a line segment, and so cannot be 0; neither can the height be 0. But
the area of a rectangle is always 0, since it is obvious that a line seg-
ment has area 0, and hence so has the union of four line segments.
If we are to speak of areas, we ought to speak of the area of the
convex hull of a rectangle, or the inside of a rectangle. The fence
round the farmer's field is a rectangle, perhaps; the field is its convex
hull.



QUESTION 34. Whether the area of the convex hull of a rectangle,
whose sides have lengths B and H, is BH?

Objection 1. The correct answer is not Yes; neither is it No. The
correct answer is Maybe. For example, the rectangle that has base
1 and height 1 might be said to have area 1; but it would make just
as much sense to say that it had area 197 X 54. This would, of
course, necessitate giving a rectangle of base 2 and height 2 exactly
the area 790 X 16.

194

Reply to Objection 1. The objection is perfectly correct, except
that it has already been mentioned (tacitly) in the question, which
asked what was 'the area of the convex hull of a rectangle, normalised
in the obvious way, needless to say'. Since it was needless to say that
the area was normalised in the obvious way, the words 'normalised
in the obvious way, needless to say' were omitted from the question.
It is very important that the question be phrased as it is here, because
if normalisation is not mentioned then one may indeed give the
unit square the area 197 X 54.

Objection 2. The answer Yes would seem to be obvious, since the
Jacobean determinant of the linear transformation sending ortho-
gonal basis elements e[1], e[2] to Be[1], He[2], respectively is merely the
determinant of the linear transformation itself, namely BH. It is
well known that under a transformation obeying suitable conditions
areas are multiplied by the Jacobean determinant, so that the rect-
angle of sides B and H which is the image of the unit square under
this transformation has area BH.

Reply to Objection 1. This fact is only well known after the area
of a rectangle is well known. Hence this objection is illogical.


I ANSWER THAT the area of a rectangle is base times height, at least
if the rectangle has sides parallel to the coordinate axes. (It is
difficult to see what are the base and height of a rectangle, if the
rcctangle is at an angle to the coordinate axes, anyway.)

The area of a one-by-one rectangle is taken as 1. This is purely for
simplicity's sake; if anyone disagrees, it will be easy to alter the areas
of all rectangles so that the one-by-one rectangle has area equal to
whatever suits one's fancy. It is not assumed that the area of any
one-by-one rectangle whatever is 1; that would be rather bold.
However, it will be necessary to assume the existence of at least one
such rectangle with that area.

If a rectangle is moved about, its area remains the same. It is
desirable to move the rectangles gently, as they are not rigid figures.
Preferably, they are to be moved by an affine transformation of the
form

c + I,

where of course I is the identity endomorphism of the vector space

195

R<3>. It is of course far from obvious that the area of a rectangle does
remain the same when it is moved, since not every measure on a
locally compact topological group is Haar measure. Clearly, one of
the tacit words in the question was Haar: the question asked about
the Haar area of a rectangle with sides B and H. (The words 'convex
hull', the reader will notice, have become tacit by now.)

If a rectangle is (the boundary of the union of the convex hulls of)
two other rectangles joined together, and if the two other rectangles
do not overlap (so that their convex hulls have disjoint interiors)
then its area is the sum of theirs. This is because the words 'finitely
additive' appear tacitly in the question.

It is obvious that if axes have been chosen through two adjacent
sides of the basic unit rectangle, the relation given holds so long as
the vertices of the rectangles have rational coordinates. The rather
mild assumption that countable additivity or continuity from above
at (psi) holds for rectangles suffices to finish the job. It has not been
shown that rectangles really have areas; what has been shown is that
if they have areas which are as described in the question (mostly by
tacit expressions hidden therein) then the areas are as stated. The
question still remains: are areas of rectangles all nonsense?

















196

BIBLIOGRAPHY

1. The Shorter Oxford English Dictionary on Historical Principles
prepared by William Little, H. W. Fowler, and J. Coulson,
revised and edited by C. T. Onions, vols. 1 and 2, Oxford, 1933.

2. Reich, Peter A., The finiteness of natural language, Language,
vol. 45 (1969), pp. 831-843.

3. Melanesian Pidgin Phrase-Book and Vocabalary with Gram-
matical Introduction, Baltimore (Linguistic Society of America),
1943.

4. Dedekind, Richard, Was sind und was sollen die Zahlen?,
Braunschweig, 1888.

5. Peano, Giuseppe, Formulaire de Mathe'matique, 3rd edn., Paris,
Gauthier-Villars, 1901.

6. Richard, Jules, Les principes des mathe'matiques et le proble`me
des ensembles, Acta Mathematica, vol. 30 (1906).

7. Durell, Clement Vavasor, Advanced Algebra, vol. 1, London
(G. Bell & Son), 1932.

8. Papy, Mathe'matique Moderne, vol. 3, Brussels (Didier), 1965.

9. Ball, Walther William Rouse, A Short Account of the History of
Mathematics, 2nd edn., London (MacMillan), 1893.

10. ben Amman, Moses, Torah, Eng. tr. in The Holy Bible, Author-
ised King James Version, London (Collins).

205

11. ben Amman, Moses, Beresith, Eng. tr. the first book of Moses,
called Genesis, in the Holy Bible, Authorised King James
Version, London (Collins).

12. Lawvere, F. W., The category of categories as a foundation for
mathematics, in Proceedings of the Conference on Categorical
Algebra, La Jolla, 1965, Berlin (Springer), 1966.

13. ben David, Solomon, Misle, Eng. tr. The proverbs, in The Holy
Bible, Authorised King James Version, London (Collins).

14. Littleton, Adam, Linguae Latinae Liber Dictionarius Quadri-
partitus. A Latine Dictionary, In Four Parts, London, 1684.

15. ben Amman, Moses, Vaicra, Latin tr. by St Jerome Liber
leviticus in Biblia Sacra Vulgatae Editionis, Turin (Pomba),
1840.

16. Dilling, Elizabeth, (Mrs Albert W. Dilling), The Roosevelt Red
Record and its Background, Chicago (Dilling), 1936.

17. Heath, Sir Thomas L., The Thirteen Books of Euclid's Elements,
Translated from the Text of Heiberg, with Introduction and
Commentary, 2nd edn. revised with additions, Cambridge Uni-
versity Press, 1926.

18. Borges, Jorge Luis, Ficciones, vol. 5 of Obras Completas de
Jorge Luis Borges, Buenos Aires (Emece'), 1967.

19. Lawson, Alfred, Lawsonomy, vol. 1, Detroit (Humanity Bene-
factor Foundation), 1935.

20. Sunshine, I., editor, Handbook of Analytical Toxicology, Cleve-
land (The Chemical Rubber Company), 1969.

21. Klarner, David A., A combinatorial formula involving the Fred-
holm integral equation, Journal of Combinatorial Theory, vol. 5
(1968), pp. 59-74.

22. Banach, Stefan, and Tarski, Alfred, Sur la de'composition des
ensembles de points en parties respectivement congruentes,
Fundamenta Mathamticae, tom 6 (1924) pp. 244-277.

23. Scott, Sir Walter, Waverley; or, 'Tis Sixty Years Hence, volumes
1 and 2 in The Waverly Novels, 48 volumes, London (Archibald
Constable), 1895-96.

24. Moore, Clement Clarke, A Visit from St. Nicholas.

25. Chomsky, Noam, Language and Mind, New York (Harcourt
Brace & World), 1968.

206

26. Appleton, Reginald Bainbridge, The Elements of Greek Philo-
sophy from Thales to Aristotle, London (Methuen), 1922.

27. Dee, John, introduction to Rudd, Captain Thomas, Euclides
Elements of Geometry the first VI books: in a compendious
form contracted and demonstrated, London, 1657.

28. Gauss, Carl Friedrich, Disquisitiones Arithmeticae, in vol. 1 or
Werke, herausgegeben von der ko:niglichen Gesellschaft def
Wissenschaft zu Go:ttingen, 1870.

29. Hix, Muriel, The Awful Mathematician's Book, London (Wolfe),
1965.

30. Euclid, Elementa Liber VI, in Euclidis Opera Omnia, ed. by
I. L. Heiberg and H. Menge, Leipzig (Teubner), 1883.

31. Hardy, Godfrey Harold and Wright, Edward Maitland, An
Introduction to the Theory of Numbers, 3rd edn., Oxford Univer-
sity Press, 1954.

32. Watts, Alan Wilson, An Outline of Zen Buddhism, London
(Golden Vista Press), 1932.

33. Tolkien, John Ronald Reuel, The Lord of the Rings, three
volumes, 2nd edn., London (George Allen and Unwin), 1966.

34. Chambers's Encyclopaedia, vol. 13, Spain--Tsushima, New
Edition, London (George Newnes), 1959.

35. Nouveau Petit Larousse Illustre', Dictionnaire Encyclope'dique,
e'dition spe'ciale re'alise'e pour les cinquante ans de l'ouvrage de
Claude et Paul Auge', Paris (Larousse), 1952.















207