Anda di halaman 1dari 156

DOVER BOOKS ON

ADVANCED MATHEMATICS
ToPOLOGY, John G. Hotking and Gail S. Young. (65676-4) S7.9S
Itml.OOUCT1ON TO UNEAJI. ALOI!llRA, Marvin Marcus and HcnryJc Mine.
(6569S-O) $6.95
NIJMIID. TIlEORY AND ITS HISTORY, Oystcin Ore. (6S62()'9) 58.95
MATItEMATlCAL MrntODS OF OPERATIONS iU:sEARCH, Thomas L Santy.
(6570).5) $12.95
A SURVEY OF NUMERJCAL MATlIEMATlC'S, David M. Young and Robert Todd
Gregory. (65691-8, 65692-6) Two·volume sel $27.90
DlITRlBtmONTllEOII.Y AND TRANSfORM ANALYStS, A.H. Zemanian. (65419.
6) $8.95
Gaol£lw.lZED INTEGRAL TRANSfORMATlOss, A.H. Zemanian. (65375-7)
S7.9S
THE ORIGlh'S OF THE INRNITfSIMAL c..u,alws, Margaret E. Ilaron.
(65371-4) $7.95
INTltODlK'TlON TO NUMERICALAAALys\s. F. B. Hildebrand. (65363.) S I ].95
MAntEMATlCAL EcoNOt.tIC'S, Kelvin ulOcasler. (65391·9) 59.95
FIFTY OlALLEN(j[NO PROBLfMS IN PROBABIlITY Wtnl SoLUT10NS, Frederick
Mosteller. (65)55·2) $3.95
HANDBOOK. OF MAntEMATlCAL FUNCTIONS, Milton Abrnrnowilz and Irene
A. Slegun. (61272-4) $21.95
TENSOR ANALYStS OS MANIFOlDS, Richard L Bishop and Samuel I.
Goldberg. (64039·6) 55.95
VECTOR AND TENSOR ANALYSIS ....ml
. AI'l'UCATIONS, A.1. Boriscnko and I.E.
Tarapov. (63833-2) 56.00
TilE HISTORY or TIlE CALCULUS AND ITS CONCEPTUAL DEvELOI'MENT. Carl B.
Boyer. (605094) $6.95
PRINClI'W OF STATISTICS, M.G. Bulmer. (63760-3) 55.95
TilE TIlEORY OF SMNORS, Bie Cartan. (64070-1) 55.00
ADVANCED NU�IDER TIIEORY, Harvey Cohn. (64023.X) 56.00
STAnmcs MANUAL, Edwin L Crow, Francis Davis, nnd Margaret
Maxfield. (60599-X) 56.00
CoMPUTABlUIT AND UNSOLVABIUIT, Martin Davis. (61471.9) 56.95
Asn.U'"TOTIC METitoDS IN ANALYSIS, N. O. de I1ruijn.
PJ.OBL.f;MS [N GROUP TIlEORY, John D. Dixon. (6IS74-X) 54.95
TilE MATIIEM Ancs or G.ut£S Of STRATEGY. Melvin
55.00
AsVMFTOnC ExfANSIOSS, A. Erd�lyi. (60318-0) 53.50
CoMPI..£)( VARIABILS: tiARMON1C AND ANALYTIC FUNCTIOSS, Francis J.
Aanigan. (61388-7) 57.95
A HISTORY OF GREEK MATIlI!MATICS, Sir Thomas Healh.
(24073·8, 24074-6) Two- volu me sct 521.90
DWl'£R[NllAL EQuATIONS: A MOD�1lN AI'rRoAol, Harry Hochsladt.
(61941·9) 57.50

(ronlinUf'fi on back flap)


THEORY OF SETS
By E. Kamke
Thi� i5 the c1C3r<_'S1 ,mil simplest introduction yet wriucn to the
[[,,:ory of St:ts. Making LISt: "r the diilCO\'cr;c5 of GnlOr, Russell,
Wcicrstr�SJ. Zcrmclo, Ilcrn5lcin. Dcdcldm!. amI other malhemal;·
ciatu. it :In:tlyzl.'S (;OnCCpls .md principks and offers innumerable
cxarnplcs. ils emphasis is all fUII!!Juu,nt31s :Ind the pn;sclllalion ;5
casil)' wmptdlcllsih1c to n::ulcrs with sollle college algclll":t. IIUI
special subdivisions. such a5 the Ihl"Ory of SCtS of points, :,re
considered.

The COl1tci1!S include Tlulimcnl5 (first cbssific;u;01l5. suhs<:u, SIIIIlS.


imcl'Scclion of 5<:15. noncnulllcrahlc SCIS. elc.)= arbitrary scu and
their cardinal numhers (extensions of numher (oncepl. cqni\'aicu{c
of scu. SUI1lS :md products of tWO a",1 many cardinal numocn, et(.):
ordered SCtS and thcir order I)'pes; and well,ordered set� and their
ordinal !lumbers (:tdtlitio" and mnltiplication of onli,,;,1 nlUt,hen.
transfinite induction. produr,t� ;''''\ 1><;I"'el1l of ordinal IIlImhe!!.
well·ordering IlteUl'etn. wcll·ordcring of cardinal and fJl'tiinal num·
bers, etc.).

"Exceptionall), well writtell," SriIOO/ Sritma: ",,,/ MIllilcIIIIl/ir.S,


",\ "er), riue bool," MIl/h�"w/ics Trllclu:r. "Of real scr\'ice to logi.
cians 3tLtI philosophers who h:I\'c hitherto had no acccS!t 10 a concise
and accnrate introduction to lite general tlt�'()ry of leIS," Philo·
sophical Utrlliew.
Erst ETtgJish Ir.Itlslatiou, translated for Do\'(:r (from the second
German ctlilion) hy Fretlcricl lIagemihl. "ii + ]4'lpl" 5Ya x 8.
]'aperhollnd,

ISBN 0-486-60141-2
90000

$4.95 INUSA
(conlinuedfrom front flap)

THE THEORY OF MATRICES IN NUMElUCAL ANALYSlS, Alston S. Householder.


(61781-5) $6.95
MA11fEMATICAL FOUNDATIONS OF INFORMATION THEORY, A.I. Khinchin.
(60434-9) $3.50
INTROOUCJORY REAL ANALYSIS, A.N. Kolmogorov and S.V. Fomin.
(61226-0) $7.95
SPfClALFUNcnONS ANDTHEI1l APPUCATIONS, N.N. Lebedev. (60624-4) $6.95
WORKED PROBLEMS IN APPUED MATHEMATIcs, N.N. Lebedev, I.P. Skalskaya
and V.S. UOyand. (63730-1) $8.50
FUNDAMENTAL CoNC1!P1'S OF ALoEBRA, Bruce E. Meserve. (61470-0) $6.50
FUNDAMENTALCoNC1!P1'SOFGEOMETRY, Bruce E. Meserve. (63415-9) $7.95
TABLES OF INDEFINITE IrmoRALS, G. Petit Bois. (60225-7) $6.95
MATRICES AND 1iv.NSFORMATION5, Anthony J. Pettofrezzo. (63634-8) $3.95
PROBABIU1Y THEORY: A CoNCISE CoURSE, Y.A. Rozanov. (63544-9) 54.50
ORDINARY DIFFEREI'mAL EQUATIONS AND STABIUTY THEORY: AN
INTRODUcnON, SCd A. SAnchez. (63828-6) $4.95
LINEAR ALoEBRA, Georgi E. Shilov. (63518-X) $7.95
PROB1J!MS IN PROBABIU1Y THEORY, MATHEMATICAL STATImCS AND THEORY
OF RANDoM FUNcnoNS, A.A. Sveshnikov. (63717-4) $8.95
TENSOR CALcuws, J.L Synge and A. Schild. (63612-7) 57.00
CALcvLus OF VARlATIONS WITH APPuCATIONS TO PHYsiCS AND ENGINEERING,
Robert Weinstock. (63069-2) 57.50
1NTR0DUcnON TO VECTOR AND TENSOR ANALYSIS, Robert C. Wrede.
(61 879-X) $8.95

Paperbound unless otherwise indicated. Prices subject to change without


notice. Available at your book dealer or write for free catalogues to Dept.
23, Dover Publications, Inc., 31 East 2nd Street. Mineola, N.Y. 11501.
Please indicate field of interest. Each year Dover publishes over 200 books
on fine art. music, crafts and needlework, antiques, languages, literature,
children's books, chess, cookery, nature, anthropology, science,
mathematics, and other areas.
Manufactured in the U.S.A.
Theory of Sets

BY DR. E. KAMKE

PROFESSOR OF I(
\ ATHElIATICS,

TRANSLATED BY FREDERICK BAGEMIHL

ASSISTANT PROFESSOR OF MATHElI(ATICS, UNIVERSITY OF ROCHESTER

NEW YORK

Dover Publicatioru, Inc.


Copyright © 1950 by Dover Publications. Inc.
All rights reserved under Pan American and
International Copyright Conventions.

Published in Canada by General Publishing Com·


pany, Ltd., lIO Lesmill Road, Don Mills, Toronto,
Ontario.
Published in the United Kingdom by Constable
and Company, Ltd.

This Dover edition, first published in 1950, is an


English translation of the second German edition of
Mengenlehere by E. Kamke.

Standard Book Number: 486·60H/-2

Library 0/ CongreJS Catalog Card Number: 50-7144

Manufactured in the United States of America


Dover Publications, Inc.
180 Variek Street
New York, N.Y. 10014
Contents

INTRODUC1'ION . • . . . • • • • . vii

CHAPTER I. TilE RUDIMENTS OF SET TIfEORY

§ 1 . A First Classification of Sets . . . . . . . . . 1


§ 2. Three Remarkable Examples of Enumerahle Sets 2
§ 3. Suhset, Sum, and Intersection of Sets; in Particular,
of Enumerable Sets . . . . . . . ,I)
§ 4. An Example of a Nonenumerahle Set . . . . . 9

CHAPTJ.;R II. AIlIII'rRAlty SETS ... ND TIIEIIl CARDISAL NUl'tIDEItS

§ 1. Extensions of the Number Concept 12


§ 2. Equivalence of Sets . . . . . . . . . . . . . ' 13
§ 3. Cardinal Numbers. . . . . . . . . . . . . . 17
§ 4 . Introductory Remarks Concerning the Scale of
Cardinal Numbers . . . . . . . . 20
§ 5. F. Bernstein's Equivalence-Theorem . . 22
§ 6. The Sum of Two Cardinal Numbers . . 25
§ 7 . The Product of Two Cardinal Numbers 28
§ 8 . The Sum of Arbitrarily Many Cardinal Numbers 32
§ 9 . The Product of Arbitrarily Many Cardinal Numbers 37
§ 1O . The Power . . . . . . . . . . . . . . . 41
§ 11 . Some Examples of the Evaluation of Powers 47

CHAPTER III. ORDERED SETS ... ND TIIEIIl ORDER TYPES

§ 1. Definition of Ordered Set . 52


§ 2. Similarity and Order Type . . . 55
§ 3. The Sum of Order Types. . . . 58
§ 4 . The Product of Two Order Types 6 1
§ 5. Power of Type Classes 66
§ 6. Dense Sets . . . 69
§ 7. Continuous Sets . . . 73
1/
vi
CHAPTER IV. WELL-ORDERED SETS AND THEm ORDINAL
NUMBERS

§ 1. Definition of Well-ordering and of Ordinal Number 79


§ 2. Addition of Arbitrarily Many, and Multiplication
of Two, Ordinal Numbers ... . .. .. . 81
§ 3.Subsets and Similarity Mappings of Well-ordered
Sets ...' . .. . . . . . . .. 82
§ 4.The Comparison of Ordinal Numbers 85
§ 5. Sequences of Ordinal Numbers . . . 89
§ 6.Operating with Ordinal Numbers .. 93
§ 7.The Sequence of Ordinal Numbers, and Trans-
finite Induction . .. . . .. .. .. .. . 98
§ 8.The Product of Arbitrarily Many Ordinal Numbers 100
§ 9. Powers of Ordinal Numbers 104
§10.Polynomials in Ordinal Numbers .. .. . 107
ill.The Well-ordering Theorem . . . . . .110•

§12. An Application of the Well-ordering Theorem 115


§13.The Well-ordering of Cardinal Numbers . . 119
§14.Further Rules of Operation for Cardinal Numbers.
Order Type of Number Cl asses .. 121
§15.Ordinal Numbers and Sets of Points . 126

CONCLUDING REMARKS 135

BIBLIOGRAPHY • • • 139

KEY TO SYMBOLS 141

INDEX • • . • 143
Introduction

THE THEORY OF SETS, which was founded by G. Cantor (1845-


1918) and already developed by him into an admirable system,
is one of the greatest creations of the human mind. In no other
science is such bold fonnation of concepts found, and only the
theory of numbers, perhaps, contains methods of proof of
comparable beauty. It is no wonder, then, that everyone who
studies the theory of sets is indescribably fascinated by it.
Over and above that, however, this theory has become of the
very greatest importance for the whole of mathematics. It has
enriched nearly every part of mathematics, and lent it a new
appearance. It has given rise to new branches of mathematics,
or at least first rendered possible their further development,
such as the theory of sets of points, the theory of real functions,
and topology. Finally, the theory of sets has had particular
influence on the investigation of the foundations of mathe­
matics, acting in this respect, as well as through the generality
of its concepts, as a connecting link between mathematics and
philosophy.
Here we shall present the basic features of the general theory
of sets. The theory of point sets will be merely touched upon.
CHAPTf:n I

The Rudimmlts 0/ Set Theory


1. A First Classification of Sets
What is a set? By a set 911 we arc to understand, :l('(�ording
to G. Cantor, "a collection into a whole, of definite, well­
distinguished objects (called the 'elements' of 9)l) of our per­
ception or of our thought."
For example, the prime numbers between 1 and 100 consti­
tute a set of 25 elements; the totality of even numbers, a set
of infinitely many clements; the vertices of a square, a set of
4 elements; the points of a circle, a set of infinitely many
elements.
For a set, the order of succession of its elements shall not
matter, provided that nothing is said to the contrary. Thus,
e.g., the setl 11, 2, 31, consisting of the clements 1, 2, 3, is
the same set as 13, 1, 21 or 12, 3, 1). Furthermore, the same
element shall not be allowed to appear more than once. The
number complex 1 , 2, 1 , 2, 3, conscquently, becomes a set only
after deleting the repeated elements.
Two sets 911 and 91 arc said to be ('qual, in symbols: 9Jl = ��,
if they contain the same elements; i. c., if c\'ery clement of iDl
is also an clement of 91, and, conversely, every clement of 91
also belongs to IDt For example, p, 2, 31 = {3, 1,21. IDl � 9�
shall mean that IDl is not equal to 91. If m is an clement of IDl,
we write 7n E IDl (read: m is an clement of IDl), whereas In EE 9Jl
shall denote that 7n is not an element of IDl.
A first, coarse classification of sets distinguishes them into
finite and infinite sets, according us the sets do, or do 1I0t,
contain a finite number of clements. From the infinite sets, the
set of natural numbers, whose elements we may think of as
being given in their natural order of succession 11, 2, 3, . . I, .

·Scts are frequently designated by enclosing their clements in hraces.


2
is especially singled out, and is called an enumerable set. More
generally, an infinite set gn is said to be enumerable if, and only
if, it can be be written as a sequence {ml , m2 , ma , Jj i. e.,
• • •

if, and only if, to every element m of the set, a natural number
can be made to correspond in such a manner, that to every
clement of the set corresponds precisely one natural number,
and to every natural number corresponds precisely one element
of the set.
The letters in all the printing presses on the earth form a
linite set, although the number of elements is "very large."
The same is true of the number of volumes in the "universal
library" of K Lasswitz's,2 which is so large, that the librarian,
were he to dash along the row of books even with the velocity
of light, would arrive at the last volume only after 1O , .••u • •
82
years.
If a set is finite or enumerable, we shall say that it is at most
enumerable. If it is neither finite nor enumerable, it will be
called nonenumerable.

2. Three Remarkable Examples of Enumerable Sets


G. Cantor proved, already in one of his first papers on bet
theory, the enumerability of two sets which hardly seem at
first glance to P088CSS this property.

THEOREM 1. The set of all rational numbers is enumerable.


Proof: Let us first deal only with the positive rational num­
bers. We can imagine to be written down in order of magnitude,
first, all whole numbers, i. e., all numbers with denominator Ii
then, all fractions with denominator 2j then, all fractions with
denominator 3; etc. There arise in this manner the rows of
numbers

'See his book, "Traumkristalle".


If we write down the numbers in the order of succession
indicated by the line drawn in (leaving out numbers which
have already appeared), then every positive rational number
certainly appears, and also only once. The totality of these
rational numbers is thus written as a sequence
1, 2, l, i, 3, 4, i, I, i, ... .
If we denote this sequence by (r, , r2 , rl , •}, then obviously
• •

(0, -r, , r, , -T2 , T2 , • } is the set of all rational numbers,


• •

and the enumerability of this set is herewith established.


For Cantor's second theorem, which asserts the enumerability
of an even "more extensive" class of numbers, let us recall the
definition of algebraic number. A number of this sort is under­
stood to be one which is a root of a polynomial
f(x) = a .x" + a,,_,x"-' + . . . + a,x + ao ,
where a" � 0 and all a.'s are integral rational numbers. Alge­
braic numbers include, among others, all rational numbers and
all roots of such.
THEOREM 2. The set of aU algebraic numbers is enumerable.
Proof:' Let f(x) be a polynomial of the kind just describcd�

aWe confine oUl1I8lvea here to real algebraic numbers. The theorem and
proof are also valid, however, for complex algebraic numbers.
4
We may suppose, moreover, without loss of generality, that
an > O. Define the "height" of the polynomial as the positive
number
II = n + a� + I a._I I + ... + I al I + I ao I·
The height is obviously an integer � 1. The same height is
possessed by only a finite number of polynomials, because
n � h and every 1 at 1 � h. Therefore to every height corre­
spond only a finite number of algebraic numbers. This makes
it possible to write the set of all algebraic numbers as a sequence.
First we write down all the algebraic numbers yielded by the
height 2. Since the only polynomials of height 2 are x and 2,
we obtain the sole number o. The polynomials of height 3 are
x2, 2x, x + 1, x - 1,3. These give the new roots -1 and +1.
The new roots arising from the polynomials of height 4 are,
in order of magnitude: -2, -}, +1, +2. Height 5 yields
-3, -i - t vis -O, -1 0, t - } 0, I, i, -! + -

{vis, t vl2, 0, ! + t vis, 3; etc. Thus, by allowing the


height to run through the series of natural numbers, and writing
down the newly arising, finitely many algebraic numbers corre­
sponding to each value of the height, we obtain a sequence of
distinct algebraic numbers. Since every polynomial has a
height, all nlgf'hruic numbers appear in the sequence. This
completes the proof of the theorem.
That the concept of enumerable set also leads to valuable
results concerning functions is shown by
THEOREM 3. Every fundion f{x) which is monotonic in an
interval a ::; x � b is discontinuous at an at most enumerable
number 01 point8 01 thi8 interval.
Proof: It suffices to carry out the proof for monotonically
increasing functions. Let f(x) be such a function. It is dis­
eontinuous at a point t if, and only if,·
uW = f(.t + 0) - f(t - 0) > 0,

4A.'1 UHUILI. M/(� - 0) dCllote tho left-hand limit and/(E + 0) tho right­
hand limit of I(z) at Lim point E.
5
where we agree that f(a - 0) = f(a),f(b + 0) = feb). If
a < �I < �2 < ... < �p < b,
and XI , • • • , Xp_1 are numbllrs in the intervals

�. < x. < �'+I ,


and if, finally, Xo = a, Xp = b, then
f (x.) - !(X._I) � f(�. + 0) - f(�. - 0) = O'(�.),
so that
p p

feb) - f(a) = L <f(x.)


•-1
- f(x.-I» � L
.-1
O'(�.) .

If, now, the �. are numbers with O'(�. ) > lin, it follows that
p < nU(b) - !(a»;
i. e., the' number of points of discontinuity � with O'(E) > lin
has a fixed upper bound. Hence, at most a finite number of
points of discontinuity � with O'(E) > lin belong to the interval
a 5 X 5 b, even when the end points have been taken into
account.The totality of points of discontinuity can therefore
be written as a sequence, by first writing down the finite
number of points of discontinuity � with O'(�) > I , next the
newly arising ones with a(�) > i, then the newly acquired
ones with a(E) > i, etc.Since to each point of discontinuity
� there belongs a definite positive number aW, every point of
discontinuity appears in this sequence, and indeed precisely.
once. Thus our assertion is proved.

3. Subset, Sum, and Intersection of Sets, in Particular, of


Enumerable Sets
The considerations which led in §2 to the enumerability of
the set of all rational numbers, furnish, over and above this,
a general theorem which is often useful for establishing the
enumerability of a set. Before we formulate this theorem, we
shall introduce several ne,,- concepts which will be encountered
continually in what follows.
6
Let a set on be given. The set \n is called a subset or part
of Ril, in symbols: 91 c on, if every element of m is at the same
time an element of on; i. e., if n E 91 always implies n E ID1.
E. g., according to this, every set is a subset of itself, and, in
fact, it is called an improper subset. On the other hand, tn is
called a proper subset of IDl, in symbols: 91 C IDl, if 91 is a subset
of on and m � on. The set of those elements of on which do
not belong at the same time to the subset tn, is termed the
residual set 9l belonging to tn, or the complement of m with
respect to IDl; in symbols: 9l = IDl - tn. So that this definition
may be valid also for the case in which 91 is an improper subset
of IDl, we introduce an ideal set, the so-called null set or empty
set, which we denote by 0.1 This is analogous to the intro­
duction, in geometry, of infinitely distant points as ideal points,
and the introduction, at one time, of zero into the number
system, as an ideal number.II The null set shall be classed with
the finite sets. It shall be a subset of every set, and, in particular,
of itself.
2 a
Example8. The boundary points x + y = 1 form a proper
' 2
subset of the points of the circle x + y � 1; its complement
2
consists of the interior points x + 11 < 1 of the circle. Tbe
set of rational numbers is a subset of the set of all real numbersj
the complementary set consists of the irrational numbers.
Let there be given an enumerable set on = I m1 , ma , ma , . . • I ,
and suppose that 0 C tn � on. Then, in the sequence on, there
is a first element m•• which belongs to tnj let n1 = m•• . This
is followed again by a first element, m•• , in on, which belongs
to 91; let 7lJ = m•• ; etc. The process does or does not terminate,
according as 91 is finite or not. Since on contains all the elements
of 91 (and perhaps even more), the possibly terminating sequence
In1 , 7lJ , na , ···1 comprises precisely the elements of 91.
Therefore m is finite or enumerable. Hence, we have

lOne must distinguish the empty set 0 from the sot. {OJ containing tho
number 0 88 an element.
IHere we IU'e thinking of the historical development of the number con­
cept, not, say, of the role whioh the number 0 plaYIl in an axiomatic con­
struction of the number lIystem.
7
THEOREM 1. Every subset of an enumerable set is at most
enumerable.
E. g., according to this theorem, the set of all algebraic
numbers which satisfy irreducible equations of prime degree is
enumerable as a part of the set of aU algebraic numbers. The
set of positive fractions which can be represented as the sum
of three biquadrates of rational numbers is enumerable as a
subset of the set of all rational numbers.
The union or (logical) sum e of finitely or infinitely many
sets is understood to be the set of those clements which belong
to at least one of the sets. The union of an at most enumerahle
number of sets ml , IDl2 , IDl3 , • • . is written in the form

or

The intersection, logical product, or product of the first kind,


of arbitrarily many sets, is understood to be the set of those
elements which belong to each of the aforesaid sets. For the
intersection, Xl, of two sets IDlI and IDl2 , we write
� = �(IDl I , rol2) or � = 9)11·9)12 ;
and for the intersection of at most enumerably many sets,
...
or � = IDl1·9)1, • . • = n IDlt .
l-l

The second notation, which will be employed here by prefer­


ence, offers the advantage that rules of operation for the slim
and product of sets then hold, which arc similar to those for
the sum and product of numbers. For example,
IDl(� + 0) = ID1.� + m·,Q.
In forming sums, null sets are left out, and the sum of null
sets shall be the null set. If, in a product, one of the sets is
empty, we put � = OJ likewise if the sets have no element in
common. If the intersection of every pair of sets under con­
sideration is 0, the sets are called mutually exclusive or dis­
junct.
8
Examples. If IDl = I I, 2, 3, ···1 and m = {5, 7, 91, then
ID2 + m = ID'l and ID1·m = m. If ID1 is the set of even integers
and m is the set of odd integers, then ID1 + m is the set of all
integers and ID1·m = O. The two circles in Fig. 1 have as sum
the entire hatched region, and as product, the crosshatched
region. The enumerllbly many sets I I, 2, 3, 4, • . . J, 1 2, 3, 4, . I,
. .

13, 4, ... I, '" have the first set as their sum, and their inter­
section is O.
The theorem alluded to at the beginning of this f3 now reads
a s follows:
THEOREM 2. Thc sum oj at 'most enumerably many scts, each oj
which is at most enumerable, is likewise at most enumerable.
ProoJ: If we are dealing with two sets, ID1 = {ml , m2 , • . I, ,

9l = {nl , n2 , '" I, which are finite or enumerable, we can


form the sequence ml , nl , m, , n2 , • • • so far as the given sets
furnish elements as contributions. If we strike out in this
sequence every element that coincides with a preceding one,
we have obviously written the set ID1 + m as a sequence and
proved the enumerability of the sum. In this way the proof
for finitely many sets can always bP. carried out by "sliding
together" the given sets.
If there are, however, enumerllbly many sets, ID11 , ID12 , • . • ,
let ID'l" have the elements mjll , m..2 , • • • • We first write down,
one under another, the sets given as sequences:

and then write down the elements anew in the order of succes­
Ilion indicated by the arrows:
9

leaving out elements that have already appeared. We thus get


a sequence containing all the clements that occur in any of
our ID1,.'s, thereby proving the assertion.
The principle of this proof, which is obviously that already
employed on p. 3 and is also involved in the proof on p. 4 is
called the first (or Cauchy) diagonal-method.
Example8. In the plane, the set of all points whose coordinates
are both rational numbers is enumerable. For if the rational
abscissa x, say, is hcld fixed, the points with this fixed abscissa
and arbitrary rational ordinate y form, according to §2, Theorem

Fig. 1.

1, an enumerable set. If we now permit x also to run through


all the rational numbers, we have enumerably many enumerable
sets, which thus all together yield once more an enumerable
set. It follows, likewise, in space, that the set of all points
whose coordinates are all rational numbers is enumerable. In
particular, it is a consequence of this, together with Theorem 1,
that the set of all cubes with sides of length 1 and vertices with
integral coordinates is enumerable. Further, in the plane, the
set of all circles with rational centers and radii is enumerable.
4. An Example of a Nonenumerable Set
According to what has preceded, there is still the possibility
of all infinite sets being enumerable. The distinction of sets
into enumerable and nonenumerable ones becomes significant
only after the existence of nonenumerable sets has been demon­
strated. We therefore prove the following
THEOREM. The .'tet of all rcal1lltmllers in tile interval 0 � x � 1
is 1l0nClIll7neralJle.
Proof: The proof is also carried out by means of II. so-called
diagonal method; let us call it the second (or Cantor) diagonal-
10
method. On account of 13, Theorem I, it is sufficient to demon­
strate the nonenumerability of the set 0 < :I: � 1. Every
number 0 < :I: � 1 can be written as a nontenninating decimal
fraction O'ala,a, • • • (e. g., I = 0.499 ... , 1 = 0.999 • . . ),
and, in fact, in a unique manner. If, now, the numbers in the
interval 0 < :I: � 1 were enumerable, they could be written
as a sequence of nonterminating decimal fractions


(1)

From the principal diagonal indicated by the line, we now form


the nontenninating decimal fraction O'allauaaa • • • , and from
this one we construct a new decimal fraction by replacing each
digit a.. by a different one, b. , where the latter is taken different
from O. We have then for the resulting decimal fraction, d =
0·b1btb• • . • , also 0 < d � 1. The decimal fraction for d does
not tenninate because of the absence of zeros. It should there­
fore have to coincide in all digits with one of the decimal
fractions in (1). That, however, according to our construction
of d, is certainly not the case, because d surely differs from
each nth decimal fraction in (1) in the nth digit. This completes
the proof of the theorem.
From this theorem we obtain the following very remarkable
Corollary. The Bet 01 aU transcendental number8' is non­
enumerable.
For if the set were enumerable, the addition of the set of
algebraic numbers, which, according to §2, Theorem 2, is
enumerable, would again result in only an enumerable set. This
fA number is caned transcendental, if it. is not an algebraic number.
Hermite, in 1873, proved that. tho number e is transccndental, and Linde­
mann, in 1892, established the transccndentality of the Dumber 11'.
11
means that the set of all real numbers, and hence, a fortiori,
the subset of real numbers lying in the interval 0 < z :5 1,
would be enumerable. This would, however, contradict the
theorem just proved.
CBAPrER II

Arbitrary Sets and Their Cardinal Numbers

1. Extensions of the Number Concept


Up to now we have classified sets into finite, enumerable,
and nonenumerable sets. To the last class belong simply all
those sets that remain after the first two classes have been
split off. One can raise the question, whether it is not possible
to further subdivide the class of nonenumerable sets. G. Cantor
gave this question, in truly ingenious fashion, the following
tum: Can the concept of natural number be generalized in !mch
a manner, that every set is assigned one of these generalized
"numbers" for the "number of its elements", as it were?
Should this be possible, there would result immediately a classi­
fication of infinite sets, too, according to the "number of their
elements." Clearly these new "numbers" would have to be
something quite novel. It is therefore useful to recall how it is
otherwise customary in mathematics to introduce new numbers.
The number concept has, of course, been extended several
times in the evolution of mathematics. The first extension
consisted in the introduction of fractions. In a rigorous intro­
duction of the rational numbers, one must, by all means,
abandon the primitive method of "dividing a whole into 0.
certain number of parts", and instead of this, proceed as
follows: One considers, as a new kind of "number", & pair of
natural numbers, a, b, which, with this interpretation, shall be
written in the form alb, in conformity with the usual manner
of writing fractions. This, however, would not yet accomplish
altogether what one would like to have. For, the number pairs
alb, 2a/2b, 3a/3b are all to be regarded merely as different
representations of the same rational number. The correct in­
troduction of the rational number can therefore occur as
follows: In the first place, one makes the stipulation that every
pair of number pairs po/pb and qa/qb, composed of natural
12
13
numbers, shall be regarded as equivalent. In the second place,
one agrees that a rational number shall be u nderstood to be
an arbitrary representative chosen from a class consisting of
equivalent number pairs. In deriving the rules of operation for
rational numbers, one must always make sure that these rules
of operation are independent of which representative of the
rational number is chosen in any given instance.
Nowadays the irrational number is usually introduced by
means of nests of intervals. A nest of intervals is a sequence
of closed intervals

(rl ,81), (r2' 82), (r3' 83) , ... ,


where the r�'s and 8�'S arc rational numbers such that always

and 8� - r" -+ 0 as n -+ co. Two nests of intervals, consisting


of the interval sequences (r.. , 8..) and fT.. , &,,), are regarded as
"equivalent," if r" � 8" and r.. � i" for all n. Analogously to
what was stated ahove, an irrational number or, in general, a
real number is an arbitrary representative taken from a class
composed of equivalent nests of intervals. Here, too, in de­
riving the rules of operation for real numbers, one must make
certain that thcy arc independent of which reprcsentative of
the real number is choscn in any particular instance.
Let us bear these things in mind for the introduction of
cardinal numbers.

2. Equivalence of Sets
A second point of departure for Cantor's extension of the
number concept is, as opposed to the first, of a very simple
nature. A child who is unable to count can nevertheless de­
termine whether there are, for example, just as many chairR
as persons present in a room. He need only have each person
take a scat on a chair. By this act, pairs are formed, each pair
consisting of one pm"MOIl Ilnd one chair. One can also say that
t.he chairs and persons ure mudc to correspond to cach other
in a one-t.o-one manner; i. c., in such a way that precisely ono
14
chair corresponds to each person, and exactly one person corre­
sponds to each chair. This primit.ive procedure, however, can
be carried over to arbitrary sets too, and it then leads to a
concept which corresponds to that of the "same number" of
elements in the case of finite sets. We accordingly set down the
following
Definition. A set IDl is said to be equivalent to a set m, in 81j1T1bols:
IDl '" 91, iJ it is possible to make the elements oj 91 correspond to
the elements oj IDl in a one-to-one manner; i. e., iJ it is possible
to make correspond to every element m of IDl an element n of 91
in such a manner that, on the basis of this correspondence, to
every element of IDl there corresponds one, and only one, element
of 91, and, conversely, to every element of 91, one, and only one,
element of [It Imtead oj saying one-to-one correspondence, we
also speak, for brevity, of a mapping. The empty set shall be
equivalent only to itself.
As with all definitions of this kind (equality, similarity, etc.),
it is to be required of the concept of equivalence, that it be
reflexive, symmetric, and transitive; i. e., that it possess the
following properties:
a) IDl '" IDl; i. e., every set is equivalent to itself;
{3) IDl '" 91 implies 91 '" IDl;
'Y) if IDl '" m and m '" �, then IDl '" �.
These three laws indeed fonow immediately from the defini­
tion.
Since the concept of equivalence is of really fundamental
importance, we shall illustrate it by a series of examples to
which we shall also occasionally refer later on.
a) The sets of points of the intervals1 (O, 1), (O, 1), (0, 1),
(0, 1) are equivalent to each other.
First let us prove that (0, I) '" (0, I). We denot.e the points
of the first of the two intervals by x, and those of the second,
by y, and set up the following correspondence:
I{a, b) denotes a closed interval; (a, b), an open interval; (a, b) and (a, b),
hal/-open intervals. These are intervals in which both end points belong,
no end point belopgs, the left-hand end point belongs, the right-hand end
point belongs, to the respective interval.
15

y = � - x for! < x � 1; then we have ! � y < 1;


y = I - x for t < x � !; then we have 1- � II < !;
y = i -x for 1 < x � 1; then we have 1 � y < 1;
etc.

It is evident that to every x of the first interval there is


hereby made to correspond one, and only one, point y of the
second interval, and, conversely, to every y, one, and only
one, x. This proves our assertion.
One shows in an analogous fashion, that (0, 1) ,....., (0, 1).
From this it follows finally, however, that also (0, 1) ,....., (0, I).
For we can associate the point 0 of the first interval with the
point ° of the second, and set up a one-to-one correspondence
between the remaining points of the two intervals n the
manner prescribed in the preceding paragraph.
b) For the points of any two finite intervals 31 and 32 we
have invariably 31 ,....., 32.
It suffices to sho",.· that one of these two intervals, say 31 , is
equivalent to one of the intervals mentioned in a) . Choose this
interval so, that if it be applied to a line parallel to 31 , it be
circumstanced exactly as 31 with regard to the appurtenance
of its limit points to the interval. Then it is easy to bring about
a one-to-one correspondence between the points of the two
intervals by means of a central projection, as indicated in
Figure 2.

Fig. 2.

c) A half-line and an entire line are equivalent to an interval.


For, a half-line can be obtained by central projection, from
16
an interval erected at right angles to it and open above (Fig. 3);
and a full line, in a similar manner, from a bent, open interval
(Fig. 4).
d) The interval (0, I) is often designated as the continuum.
The results established above then state t.hat all intervals, half­
lines, and complete straight lines are equivalent to one another
and, in particular, to the continuum.
e) Two finite sets are obviously equivalent if, and only if,
they contain equally many elements.
f) All enumerable sets are equivalent to each other, but not
to any finite set.

Fig. 3.

Fig. 4.

The first follows from the fact that every enumerable set is,
according to its definition (p. 2) , equivalent to the set
{I, 2, 3, . . . }. The second is obvious.
g) An infinite set can be equivalent to one of its proper
subsets. This is shown by the result d) or the equivalence of
the sets {I, 2, 3, .. . I and 12, 4, 6, ... I. Infinite sets thus
exhibit in this respect an entirely different behavior from that
of finite sets (cf. e». This property can therefore also be used
to fix the distinction between finite and infinite sets, inde­
pendently of any enumeration.
h) There exist infinite sets which are not equivalent to each
other; namely, as follows from ch. I, §4, the continuum on
the one hand and the enumerable sets on the other. It is con­
ceivable, however, that all nonenumerable sets might be
17
equivalent to one another. In that case, the introduction of the
concept of equivalence would bring about no new classification
of sets. The following result is therefore of fundamental im­
portance:
i) There exist infinite sets which are neither enumerable nor
equivalent to the continuum. An example of such It set is the
set of all real functions defined in the interval (0, 1).
The proof of this is obtained by a procedure which is again
essentially the second diagonal method (cf. p. 9). The con­
cept of function which is taken here as basis is Dirichlet's
general concept of a function. According to this, a function
y = J(x) is defined in an interval, if, with every x of the interval,
there is associated a well-determined number y. Further, two
functions are to be designated as distinct, even if they possess
different functional values at merely a single point of the in­
terval. Now it is clear not only that there are infinitely many
distinct functions in the interval (0, I), but also that there
exist nonenumerably many. For, at the point x = ° alone,
the functions can already assume the nonenumerably many
values between a and 1. Thus, there remains to be proved only
that the set of functions is not equivalent to the continuum.
Suppose, now, that the set of functions were equivalent to the
continuum. Then it would be possible to make the functions
J(x) and the points z of the interval (0, 1) correspond in a
one-ta-one manner. Denote by J,(x) the function thus assigned
to the point z. Now construct a function g(z) in the interval
o 5 z 5 I, with the property that, at every point z, g(z) �
J,(z). (This can be accomplished in various ways.) Since g(z)
is also a function defined in the interval (0, I), it must coincide
with anJ (x); hence, in particular, g(u) = J.(u). This, however,
..

is excluded by the definition of g(z). The assumption that the


set of our functions is equivalent to the continuum has thus
led to a contradiction.

3. Cardinal Numbers
The introduction of the heralded "numbers" offers no diffi­
culty now after the preliminaries encountered in §l and §2.
18
Corresponding exactly to the definition of the rational number
given on p. 13, we set down the following
Definition 1. By a cardinal number of a power m we mean an
arbitrary representative IDl of a class of mutually equivalent sets.
The cardinal number or power of a set IDl will alBo be denoted by
I IDl I· The notatian I IDl I accordingly means Bimply that the set IDl
may be replaced by any set equivalent to it.
Abstract as this definition may sound, it does nevertheless
accord precisely with the grammatical sense of the words
"cardinal number" in the case of finire sets. For, the number
of elements-and the cardinal numbers of the language serve
to designate this number-is detennined, for finite sets, by
counting; e. g., the number 4, by counting 1, 2, 3, 4. That is,
the number 4 is represented by the set (I, 2, 3, 4). In virtue
of §2, e, this coincides exactly with the definition of cardinal
number given above. Consequently, for a finite set, the cardinal
number introduced above is precisely the number, n, of its
elements, and shall also be denoted simply by n.
Of infinite sets we know as yet but a few classe s of equivalent
sets, namely: the enumerable sets, which may be represented
by the set {I, 2, 3, • . . 1, say; further, those sets which are
equivalent to the continuum; and those that are equivalent to
the set of functions in §2, i. Let us introduce the abbreviations
4, c, f for the cardinal numbers of these sets. If, finally, we
assign the cardinal number 0 to the empty set, we are familiar
thus far with the cardinal numbers

0, 1, 2, 3, " ' ; tt, c, f.

It is necessary now to decide when one of two cardinal num­


bers shall be regarded as the smaller of the two. This is ac­
complished by
Definition 2. A set IDl i8 said to have a smaller cardinal number,
or a lesser power, than a set 9l, in symbo18: I IDlI < I 9l1, if, and
only il, IDl is equivalellt to a subset of 91, but 91 is equivalent to
no subset of IDl.
This definition makes an agreement concerning the cardinal
19
numbers of two sets, on the basis of certain relations between
the sets themselves. Therein lies the assumption that these
relations are preserved if the sets in them arc replaced hy
equivalent sets. Now, for the definition to have a meaning,
this must first be demonstrated. '1'0 this end, let £JR ,...., £JR and
9l ,...., m. Further, suppose that £JR '""" 911 for a subset 911 of m,
but that 9l is equivalent to no subset of £JR. Since 9l '" 9l,
there is a one-to-one correspondence between the elements of
9l and those of m. A subset ml of m is thereby made to corre­
spond to the elements of 911 . Then ml '" 911 '" £JR '""" £JR, and
hence 91. '" rot Now we have to show merely that m is equiv­
alent to no subset of rot Were this not the case, the procedure
just employed would prove 9l to be equivalent to a subset of
rol, contradicting the hypothesis. The justification of the defini­
tion is thus established.
It is c1e� that the above definition, for finite cardinal num­
bers, yields the custoJaarJ "less than't concept.
Beyond this, hOWEl'fer, it is desirable that the familiar laws
governing inequalitie� involving finite numbers hold also for
cardinal numbers. 'VII refer to the following three rules:
a} If m < 0 and 0 < p, then m < p.
(3) If m, n are any two cardinal numbers, then at most one
of the three relations m < 0, m = n, 0 < m holds; and hence,
in particular, if m :::; n and n :::; m, then invariably m = n.
'Y} If m, n are any two cardinal numbers, then at least one
of the three relations In < n, lU = n, n < lUsubsists.
If both (3) and 'Y} are true, it follows that precisely one of
these three relations bolds.
To a): Let IDl, m, � be representatives of the cardinal num­
bers m, n, p. Then, by hypothesis, there exist sets 911 and �.
such that

Because of the last equ ivalence, there is a mapping of the set


91 on �• . By means of this mapping, m. in particular is mapped
on a subset �2 of � • . Th.'n 91. '""" �2 , and hence also IDl '" �2 •
20
Now we have merely to show that � is equivalent to no subset
of IDt If we had say � '" IDll C IDl, however, then, because of
IDl '" inl , there would exist a set in, C inl such that IDl, '" 91, ,
and hence � ,..., 912 , which would contradict n < p.
To (J) : A(�c()rding t.o the delinit.ion of /lless tha.n", the middle
one oC tho rCliati()llS fJ) cannot hoM l:IimultRncously with any
one of the other two. If, however, one of the other two relations
subsists, say nt < n, and these two cardinal numbers are
represented by the sets rol, 91, then IDl is equivalent to a subset
of in. But then we cannot have n < m.
To 'Y): This proposition is indeed true also, but can be
proved only much later (p. 1 19) and with the help of the well­
ordering theorem. It will therefore be necessary in the mean­
time, when operating with cardinal numbers, to take care that
'Y) is not used.
StipUlate, finally, that m > n shall mean the same as n < m.
This is sometimes convenient for calculation.

4. Introductory Remarks Concerning the Scale of Cardinal


Numbers
If we apply Definition 2 of §3 to finite cardinal numbers, we
obtain, as already remarked on p. 19, nothing new. If we
proceed to the cardinal number of infinite sets, or, as we say
briefly, to transfinite cardinal numbers, we can establish the
following for those cardinal numbers \\ith which we are familiar :
a) If m is any transfinite cardinal number, and n is finite,
then m > n.
For if IDl and 91 are representatives of m and n, rol contains
a subset equivalent to in, namely, every subset of n elements.
On the other hand, IDl, being an infinite set, is equivalent to
no subset of the finite set 91.
b) For every transfinite cardinal number m, m � Q. Thus
the number Q is the smallest transfinite cardinal n umber.
This follows from the fact that every infinite set contains· an
enumerable subset.
c) Since a S c and a ;,& c, we have Q < c. Whether there are
any cardinal numbers between a and c is not known to this
21
day, despite the most strenuous efforts to settle this question,
which is the substance of the so--called continuum problem.2
d) Q < C < f.
Because of c) we have merely to prove that c < f. If (£ de-­
notes the continuum, and ij the set of functions in the interval
(0, 1), then G: is equivalent to a subset of ij. For, the functions
which are constant, and, in fact, equal to a value c in the
interval (0, 1), constitute a subset of tj, and this subset is
equivalent to (£. Hence, we need only show that tj is equivalent
to no subset of (£. This follows once again from the method of
proof given in §2, i, but can also be inferred, at the conclusion
of §5, from Bernstein's equivalence theorem.
Up to now we know only three transfinite cardinal numbers.
It is therefore natural to inquire whether there are any others.
The answer is: There are infinitely many transfinite cardinal
numbers. This follows immediately from the fact that, given
any transfinite cardinal number, there exists a larger one; and
this fact can be proved, moreover, in the following sharper
form:
THEOREM. For every set IDl, the set U(IDl) oj all its subsets has
a greater cardinal number than IDl.
For finite sets, this statement is trivial. For oxo.mple, the
set of all subsets of the empty set is the set 10 } , i. e., a set
having one elemeRt, so that it has the cardinal number 1,
whereas the empty set has the cardinal number O. For the set
{ a } , U( ( a } ) = { O, { a } J , so that the aRsHrtion here is that
1 < 2. It is easy to see, in general, that for a set consisting of
a finite number, say n, of elements, the theorem states that
n < 2".
The general proof will include the case of finite sets. Let
ID'l(m) denote a set with the elements m, nnd likewise let
ID'l( {m}) be the set which results from IDl(m.) by replacing every
element m hy the set 1 m } . Since invariahly f 'IM I C IDl, we have

tThc detailed execution of an idea, due to D. Hilhcrt (Math. Anll. 95


(1926), 161 ft.), Cor a proof, still presents great difficulties.
22
IDl«m l ) � U(IDl} and at the same time IDl( {mJ ) ,.... IDl(m}.
According to this, IDl is equivalent to a subset of U(IDl), and
all that remains to be proved is that U(IDl} is equivalent to no
subset of IDl. To this end, we shall show, in a manner after the
second diagonal method, that, if Uo I"J IDlo for any two sets
Uo C U(IDl} and IDlo C IDl, then no C U(IDl) . From this it then
follows that U(IDl) is equivalent to no subset of IDl.
Suppose, then, that
IDlo C IDl, Uo C U(IDl), IDlo """ Uo •

Then there exists a mapping of IDlo on no . Keep this mapping


fixed, and let u = .,,(m} denote that element u of no which is
made to correspond, under this mapping, to the element m of
IDlo Since every element u of no is a subset of IDl, it is meaningful

to ask whether a given element m of IDlo is also an element of


a given u. Now let u be the set consisting of those elements
m of IDlo , each of which is not contained as an element in the
corresponding u = .,, (m}. This set U, which is a (possibly empty)
subset of IDl, does not appear as an element i n no .
In fact, if it were an clement of Uo , an element m of IDlo
would correspond to u under the mapping fixed at the be­
ginning. Then precisely one of the two cases m E u or m EE u
would have to hold. The first case cannot occur, according to
the definition of U. In the second case, m would not be contained
in the u corresponding to it; but then, by the definition of u,
m would have to belong to u all the same. The assumption
u E no thus invariably leads to a contradiction. Hence,
Uo C U(9,n) , and this completes the proof of the theorem.
5. F. Bernstein's EquilJalence-Theorem
If two sets, IDl and 91, are given. precisely one of the following
two CIlSCS can occur:
u) IDl is equivalent to a subset. 91, , of 91;
II} IDl is equivalent to no subset of 91.
Likewise, exactly one of the following two possibilities can
take place:
a} 91 is equivalent to a subset, IDl, , of IDl;
23
fJ) 91 is equivalent to no subset of rot
There are four conceivable combinations of these two pairs
of cases, viz., aa, a{3, ba, b{3. In the two middle cases we have,
according to p. 18, Def. 2, I rol l < I 91 1 , I rol l > I �n I , re­
spectively. The occurrence of the last case, b{3, would mean
that rol and 91 are not comparahle. As we have already re­
marked on p. 20, this case cannot take place, but the proof
of this fact must be deferred until later. The case aa remains;
and for it we have the following theorem, which was conjectured
already by G. Cantor, and proved by F. Bernstein :
EQUIVALENCE THEOREM. If each. of two 8ets, rol and 91, is
equivalent to a subset of the other, then rol '" 91.

This theorem can be reduced to the following proposition:


(P) If rol is equivalent to a subset rol2 , then rol is also equiva­
lent to every set roll "between" rol and rol2 , i . e., to every set
IDlI with the property that rol, C roll C IDl.
In fact, suppose that roll , 911 are subsets of rol, 91 such that
IDl '" 911 , 91 '" IDlI Then, by means of a mapping resulting

from the last equivalence, the set 91 is mapped on roll , and


hence, in particular, the subset 911 i s mapped on a subset,
roll ' of roll . Thus, rol3 C IDlI C IDl and IDl '" 911 '" IDl2 . Conse­
quently, according to (P), roll '" IDl; and since roll '" 91, we
have also IDl '" m.
For the proof of (P) we may obviously assume that IDl2 C
IDll C IDl. IDl is an infinite set, because IDl2 '" IDl. For con­
venience, set
IDlI = W, IDlI - rol2 = 58, rol - IDlI = �.
Then proposition (P) reads as follows:
(P*) If �, 58, (£ are disjunct sets, then
W + 58 + � /"OJ W implies � + 58 + � ,...., � + 58.
Now, according to the hypothesis that W + 58 + (£ '" W,
there exists a mapping I(J of the set W + 58 + � on W. Let
� 1 , 581 I <il be the subsets of W which are the images of W, 58,
24
(£ (cf. the hy all means very crude visualization through Fig.
5). Then
(In) �(. + �. + (£. = �l,
( I I» �( ,...., �(. J � ,...., �. , (£ ""'"' (£. J

2(1 , �I J (£1 arc disjunct.

A l B I q I
A, IB, I q , I
A: 18.11l,1
A, lsJ4J
I

Fig. 5.

Since � is mapped by II' on WI J it follows from (180) that the


subsets �I , 5.81 , �I of � are mapped by tp on subsets W2 J .sat ,
(£2 of �rJ . Then we have
(2a) 212 + .sa2 + �2 = � I ,
(2b) 211 I"'oJ �2 ' .sal I"'oJ .sa2 J (£1 ""'"' cr2 J

212 J .sa2 J �2 arc disjunct.

The next step leads to three sets 2la , .sa3 , cr3 with
(3n) 2(3 + .sa3 + 6:3 = 21a J

(3b) 2(2 ,..., 2(3 J .saa I"'oJ .sa3 J (£2 '" cr3 ,

2(3 J 583 J <S3 arc disjunctj


et cetera. Due to the fact that 2( '" 211 '" 2(2 ,..., • . • J the process
does not terminate. Note especially the equivalence
(I)
arising from (lb), (2b) , . . . .
If wc now sct SD = �1 · �2 · 213 . . . (1l may be empty), th en
�l + 58 + <S = SD + 58 + � + 581 + (£1 + 582 + U2 + . . . ,

Yl + 58 = ID + 58 + crl + 581 + (£2 + 582 + (Ia +


25

Here, on the right-hand side of the first equation, all the terms
are mutually exclusive, and the same is true of the second
equat.ion. Hem�e, � + � is mapped on �{ + � + IS, nnd the reby
(1'*) is pI'oYed, if we succeed in cstu.hli8hing a mapping bet.\\'t:'('U
e\'cry I,erm 011 \.IlC right-hand �ide of !.h e Iin5t equation, and
the term directly below it in the second equation, Tho existence
of this mapping, however, is ensured by (I) .
An immediate consequence of the equivalence theorem is t.he
following
Corollary. If [n is equivalent to a subset of m, then I 9.)l I :::;; I £Il l .

6. The Sum of Two Cardinal Numbers


We must now develop further the operations on cardinal
numbers. We shall first define addition.
A child, if asked to determine how many balls are three balls
and two balls, will combine these two sets into a single one,
and then ascertain the number of elements of this union of
the two sets. With abstract finite sets, this method must be
applied with caution. For, each of the sets I I , 2, 3 } , { 1, 2, 4}
has three elements. Their union, however, i s { I, 2, 3, 4} . This
set has only four elements, and so it cannot be used for de­
termining the sum of the numbers 3 and 3. On the other hand,
the method is always applicable to finite sets, if the sets repre­
senting the numbers are disjunct.
A suitable generalization of this method gives rise to the
following general definition of the sum of two cardinal numbers.
Let the cardinal numbers m and n be represented by the dis­
junct sets3 [n and 9t Form their union, e 9,R + £It, and then
=

stipulate that m + n = I e I.
For this definition to have a meaning, it is necessary to show
that one arrives at the same cardinal number I e I by starting

3Two cardinal numbers m and n can always be represented by disj unct


sets. For, if 9Jl(m) and men) are sets having cardinal numbers m and n, it is
sufficient to replace every element m of the first set by ml or (m, 1 ), and
every clement n of the second set by n2 or (n, 2). The sets obtained in this
manner arc certainly disjunct, and they too represent the cardinal numbers
m and n.
26
from two arbitrary sets rol, i which are equivalent to rol, 91,
and which, of course, are likewise mutually exclus!ye. In fact,
the mappings resulting from Wl '" IDl and 91 � 91 t.!!en also
furnish a mapping of the set m + 91 on the set IDl + m.
Since the relations

IDl + 91 = m + rol,

(IDl + 91) + � = IDl + (91 + �)


are valid for arbitrary sets, we have
a) m + n = n + m (commitatuve law) j

(j) (m + n) + p = m + (n + p) (associative law).'


'Y) Further, if

then invariably

For if the four cardinal numbers are represented by the four


disjunct sets roll , IDl2 I 91. , 912 , then by hypothesis there exist
subsets it S;;; 911 and i2 S;;; 912 such that IDl. '" ml and IDl. '"
ia . Since the sets are disjunct, the mappings resulting from
these equivalences then also map the set IDlI + IDl. on ml + i2 .
Consequently,

IDlI + IDl2 '" i. + 912


and 911 + 9la S;;; 911 + 912 , and therefore, by the corollary of
the equivalence theo�m (p. 25), ml + ma :::; nl + n2 •

Remark 1. We cannot, however, always infer from ml < n ,


and ma :5 n2 , that ml + rna < nl + n2 , because, e.g., for
every finite cardinal number n we have n < 0 and a :5 0, but
(cf. several lines below) a + n = a + a.

4()n the basis of the associative law, a finite sum of more than two car­
dinal numbers can also be introduced. We shall postpone this generaliza­
tion, as well as a much stronger one, until 18.
27
Remark 2. Whether the relations ml < nl , m2 < n2 invariably
imply
ml + ma < UI + U2
cannot be decided until later (p. 123) .
The following rules are valid for the addition of the simplest
cardinal numbers:
a) a +n= a, a + a = a.
For, the cardinal numbers n and a can be represented by
( 1, 2, . . . , n } , { n + I, n + 2, . . . ) , respectively, and the
sum of these sets is the set of natural numbers, which has
power a.-To derive the second rule, represent the first term
by the set of odd integers, and the second by the set of even
integers. The sum of these sets is the enumerable set of all
integers.
b) c + n = c, c + a = c, c + c = c.
To prove the last equation, represent the first term by the
points of the interval (0, 1), the second by (1, 2). Since (0, 1 ) +
(I, 2) = (0, 2), which again has power C, the third equation is
proved. From this one, the other two follow with the help of ,.).
For we have

c = c+0 � {� t :} � c + c = c.
Remark. Whether m + m = m is true for every transfinite
cardinal number m cannot be settled until later (p. 122) .
c) For every transfinite cardinal number m, m + a = m .
'fo demonstrate this, let m and a be represented by the dis·
junct sets IDl and �. Since IDl is an infinite set, it contains an
enumerable subset sa. � + sa is also enumerable, and can
therefore be mapped on 58; and, trivially, IDl - 58 can be
mapped on itself. ( � + sa) + om - sa) is thereby mapped
on sa + om - sa), i. e., IDl + �( is mapped on IDl, which proves
the assertion.
Remark. It follows from c) that the set of transcendental
28
numbers has power c. For if IDl is the set of transcendental
numbers, and � is the set of algebraic numbers, then IDl + �(
is the set of all real numbers, and therefore has power c. Conse­
quently, I IDl I = / IDl / + / � / = I IDl + � / = c.
7. The Product of Two Cardinal Numbers
For a child, mUltiplication is a repeated additionj e. g.,
3 · 4 = 4 + 4 + 4 . In order to represent each term, one has to
write down four clements, say a, b, c, d, do this once more,
then once again, and finally determine the total number of
elements written down. Here the first a must be distinguished
from the second and third. This distinction can be made by
attaching an index, or by forming the sets {I , al , {2 , a i ,
13 , a } . We shall proceed with arbitrary sets i n an analogous
fashion, and we first make the following
Definition 1. Let IDl and 91 be two disjunct scts with elements
m, n, ,·espectively. By the combinatoriaZ product, or product of the
second kind, of the two sets IDl and 91, in symbols: ID1 X In, we
shall mean the 8et of all sets 1m, n} that can be formed from the
elements m of ID1 and n of 91. If at least one oj the sets IDl, 91 is
empty, we 8tipulate that IDl X In = O.
If, e. g., IDl = I I, 2 , 3 } and 91 = 14 , 5 1 . then
9Jl X In = { I I, 4 1 , { I , 5 1 , {2 , 4 } , 12 , 5 } , 13 , 4 1 , 1 3 , 5 I l ·
For finite sets IDl and In, the cardinal number of IDl X m is
obviously equal to the product of the cardinal numbers of ID1
and 91. Hence we set down next
Definition 2. Let thc cardinal numbers m and n be repre8ented
by the di8junct sets ID1 and 91. Then the product m · n shall equal
I IDl X ln l· _ _ _ _

For disjunct sets IDl, In and IDl, In, if IDl '" IDl and 91 ...... In,
then IDl X 91 '" IDl X Ulj i. e., the product m · n is independent
of the particular representatives of the cardinal numbers m
and n. The proof of this is left to the reader.
In practice, the following form of the definition is often more
convenitmt:
Definition 211.. Let tite cardinal number8 m > 0 and n > 0 be
29
represented by the (not necessarily disjunct) sets IDl(m) and �(n).
Form the set � oJ all element pairs (m, n), where the first place
always contains an element oJ IDl, and the second, an element oJ
�. Then m · n shall equal 1 � I . If at least one oJ the cardinal
numbers lU, n is zero, we 8et m · n = O.
That this definition gives the same result us thc pl'cvious
one follows from the fact that disjunct sets are obtained from
IDl(m) and m en) by attaching the index 1 to every m and the
index 2 to every n, and that the set of all element pairs (m, n)
is equivalent to the set of all sets I ml , n2 1 .
It is obvious that IDl X � = 91 X ID1, and hence

a) m·n = n·m (commutative law).


We have, further, (ffil X 9l) X � ,....., IDl X (�l X �). To See
tp,is, for every triple of clements m, n, p of these sets simply
let the element { 1m, n } , p } of the first combinatorial product
correspond to the element {m, In, p ! } of the second. Conse­
quently,
(j) (m · n) · 1l = m · (n ' ll) (associative law) .
For disjunct sets ID1, �, �, i t is clear that om + �) X � =
ID1 X � + 91 X �. Hence,
'Y) (m + n) 1l = mil + nil (distributive law) .
8) From
it follows that
m, · m2 � n, · n2 '
For if the four cardinal numbers are represented by the four
mutually exclusive sets ID11 , ID12 , �, , �2 , there exist subsets
i, c 911 and i2 f;;; �2 Buch that rol l "" §'i1 and IDl2 ,....., i2 .
But then obviously
£In, X ID1, ,....., 911 X ml
and 911 X itll C 91, X 912 I so that
1 £In, X £In, 1 � 1 911 X 912 I ·
30
Remark. Here again we cannot infer from m, < n. , ml 5 nl ,
that m. m2 < lt1n2 . Whether this inequality is implied by
m . < " . , lth < lt2 , cannot be decided until p. 123.
We shall now become acquainted with the beginning of the
multiplication table for cardinal numbers.
a) n · n = n for every finite number n ¢ 0; n · n = n.
The second rule is derived as follows: Let a be represented
by the set { I , 2, 3, • • . } . Then, according to Definition 2&, the
product a · a is represented by the set of all number pairs (a, b),
i. e., by the set

(2, 4)

(3, 3) (3, 4)

(4, 3) (4 , 4)

which can again be enumerated in the indicated manner by


the first diagonal method. From this it follows, further, using
a), that a 5 n · a 5 a · a = a; i. e., the first rule is also valid.
b) n · e= c for every finite number n ¢ 0; a · e = c.
To prove the second rulc, represent a by the set { I , 2, 3, . . . I ,
and c by the point set (0, 1). Then the combinatorial pra<lolct
is the set of all sets I n, x } , where n is an arbitrary naturru
number, and x is an arbitrary point of the interval (0, 1) . If
each of the sets I n, x} is made to correspond to the point
n + x, we obtain a mapping of the combinatorial product on
the half-line which begins at 1, and this half-line is a point set
having power c. This proves the second rule. The first follows
from the fact that c 5 n · e 5 a · e = c.
c) c · e = c.

For the proof, let e be I'Qpresented by the numbers of the


interval (0, 1). Then e · c is represented by the set of number
31
pairs (:1:, y) with 0 < :I: =:;; 1, 0 < y =:;; 1. The assertion is proved
as soon as we show that the set of these number pairs can be
mapped on the interval 0 < z :5 1. For this demonstration,
represent each of the three numbers x, y, z by a nonterminating
decimal fraction. Since the representation of our numbers by
means of such decimal fractions is a unique one, it is necessary
to show merely that the set of decimal-fraction pairs :1:, y can
be mapped on the set of decimal fractions z. Let the pair of
decimal fractions :1:, y be given; e. g.,

x = 0 .3 06 007 8 9

y = 0.00 1 2 3 004 6
In these we mark off numeral complexes, by going always up
to the next nonzero numeral, inclusive. Now we write down
the first :t-Complex, after that the first y-complex, then the
second x-complex, foUowed by the second v-complex, etc.
Since neither :I: nor y exhibits only zeros from a certain point
on, the process can be continued without ending, and gives rise
to a nonterminating decimal-fraction; in o11r example, to

z = 0.3 001 06 2 007 3 8 004 9 6 ... .


The decimal fraction arising in this manner shall be made to
correspond to the number pair x, y. Every number pair X, y
thus determines precisely one z, and we see at the SaIne time,
that the complexes of X, V also yield just the complexes of z.
From this it follows that, conversely, every z determines exactly
one number pair X, V, and, in fact, precisely that pair which
gave rise to z; we have only to mark off the complexes in z,
and arrange them alternately so as to form two decimal frac­
tions. A one-to-one correspondence between the number pairs
x, V and the numbers z is thereby established, and this com­
pletes the proof of the rule.
Rule c) contains a fact which at first appears downright
paradoxical. If we interpret the number pairs X, y as rectangular
coordinates of points of the plane, then it follows from c) that
the points of a square can be mapped on the segment (0, 1).
32
What is more, we can represent the factors c by the set of all
real numbers. Then the number pairs x, y yield, in fact, the
points of the whole plane, and we can now infer from rule c)
that the entire plane can be mapped on the segment (0, 1).
This means that there exist functions f(z) and g(z) such that
the number pairs x = f(z), y = g(z) yield the coordinates of
all the points of the plane as z ranges over the interval (0, I),
and that every point of the plane is thereby obtained only
once. Thus, the dimension number of a geometrical configura­
tion need not be preserved under such mappings. The dimen­
sion number is, however, invariant under one-to-one mappings,
if these are, in addition, continuous. For a general proof of this
theorem, cf. L. E. J. Brouwer, Math. Ann. 70 (191 1), 161-1 65.

8. The Sum of Arbitrarily Man.y Cardinal Numbers

Let there be given a nonempty set .re(k), and to every element


k of this set let there correspond a set IDl• • We shall put this
more briefly by saying: Let there be given a set complex M'(IDl.)
arising from .re(k) .6 A cardinal-number complex �(m.) is de­
fined analogously. The union of all the sets IDl. of the set com­
plex �omt) shall be denoted by EIEll rol• . This is a generaliza­
tion of the notation introduced on p . 7 only for enumerably
many sets.
The sum of arbitrarily many cardinal numbers can now be
defined as follows:
Let M'(m,) be a cardinal-number complex arising from the
set �(k). Let every cardinal number m. be represented by a
set IDl. in such a manner, that all the IDl.'s are mutually ex­
clusive.o Then the sum of the cardinal-number complex shall
be defined as

E m. = I E IDl. I·
.ea ,ea

�Since the same set may correspond to distinct elements k, a set com pla
need not be a Bet or sets.
'It is sufficient, in order to obtain such a representation, to furnish the
elemell� m of the set !IJ', with the index k.
33
It again follows immediately from the representation of the
cardinal numbers by disjunct sets IDl. , that the sum of the
cardinal numbers is independent of their particular representa­
tives. Further, the above definition ohviously cont.ains as a
special case the one in §6.
The three rules of operation which bold fOl' Mums of two sets
can also be generalized. With regard to the commutative law,
it is to be noted that the fact that the sum is independent or
the order of succession of the tenns is already included in the
definition, since no order of succession of the elements of the
set $f(k) is specified. The commutative law therefore assumes
a fonn which differs somewhat from the usual one.
THEOREM 1 (commutative law). Two cardinal-number com­
plexe8 $f(rn..) and �(nl) have the same 8um, if every cardinal
number lJ appearing in the one complex also appears, and with the
same multiplicity, in the other compkx. Preci8ely formulated, thi8
mean8 the following: Let .ret(k) denote the 8ub8et of ekments k
of .fi'(k) to which the same cardinal number lJ corresponds. Then
the different .re,(k) 's are mutually exclusive. Define the sets �t(l)
in like manner. Then
� SMk) = .fi'(k) , � �,(l) = �(l) .
t ,
It is assumed, now, that, for every p,

(1) .fi', (k) '" �,(l),


and it i8 as8erted that
� m.. = � nl '
..ea Ie e

Proof: Represent all the cardinal numbers m.. , corresponding


to the clements k of .fi'(k), by disjunct sets rol.. j and all the
cardinal numbers nl , corresponding to the elements l of �(l),
by disjunct sets 911 Because of (1), there is a. one-to-one corre­
.

spondence between the elements k and l of each pair of sets


.fi',(k) and �,(l). I(eep this mapping fixed. Suppose an element
I corresponds to the element k. Then to these elements corre­
spond sets rol. and 911 having the same power p. These sets
34
therefore can also be mapped on one another. This, however,
determines a one-to-one mapping of the set Eie. ID'l, on the
set E, e� 91, . For if m is an arbitrary clement of the first of
these two sets, it belongs to a well-determincd ID'l_ , and conse­
quently a well-determined n of the second sum corresponds to
it in a fully unique manner; and conversely.
In connection with the general formulation of the associative
law, the following is to be noted. In its simplest form, this law
asserts that (m + n) + Il = m + (n + Il), i. e., that, for the
addition of three cardinal numbers, it is immaterial how this
set of three elements is decomposed into subsets of at most
two elements, in order to calculate the sums successively. Since
we now have a general definition of sum for arbitrarily many
terms, the associative law can immediately be given the form
(m + n) + Il = m + n + Il, n + ( m + p) = m + n + lJ,
etc. Thus we arrive at
THEOREM 2 (associative law). To every element l oj a set �(l),
let there correspond a set .re,(k). These seta .re,(k) are a8sumed to
be mutuaUy exclmive. Let .re(k) = E ,et te,(k), and to even} k
oj .re(k) let there correspond a cardinal number m• . Then a cardinal­
number complex .reem.) and a set oj cardinal-number compleU8
.re,(m.) are defined. The a8sertion, now, is that
L L m. =
iL
m • .
,EI! .ell ea
ProoJ: Once again we represent the cardinal numbers m. by
the disjunct sets ID'l. . Then L.ea , ID'li represents the cardinal
number LiEa. m. , and the assertion now follows immediately
from the obviously correct equation
L L ID'l L
, e e .ea. .
= ID'l• •
lea
THEOREM 3. Let .reem.) and .re(n.) be two cardinal-number
compleu8 arUring Jrom the set .re(k), and let m, S n. Jor every k.
Then
35
Proof: Represent the cardinal numbers m. by disjunct sets
roll , and the cardinal numbers n. by disjunct sets 91• . We
have to show that L: rol. is equivalent to a subset of L: 91• •
This follows immediately, however, from the fact that, by
hypothesis, every rol. is equivalent to a subset of 91• . Keep
fixed a mapping of the set rol. on a part of 91. , resulting
from this equivalence. Then, because of the mutual exclusive­
ness of the sets roll , and that of the sets 91. , this gives at once
a mapping of the set L: rol. on a subset of 1: m• .
Let us prove another inequality concerning cardinal numbers.

THEOREM 4. Let Jr(ml) be a cardinal-number complex which


contains no greatest cardinal number. Then, Jor every cardinal
number m of the complex,

1: m" > m .
lEIl

Proof: If we replace by 0 all the terms in the sum of the com­


plex except one, m, , then Theorem 3 states that

L: m. 2:: m, .
•E I

If, for some l, the equality sign held here, we should have, for
this lo , that every m, :5 m , Thus there would be a greatest
• .

among the cardinal numbers of the complex, namely, m ,. ,


contrary to hypothesis.
Remark. Note the following: It is not assumed in Theorem 4 ,
that to every cardinal number of the complex there is a. still
greater cardina.l number within the complex. The hypothesis
of the theorem would still be fulfilled if, e. g., the complex
contained a cardinal number which was not comparable with
any other number of the complex.

THEOREM 5. For every set 9R of cardinal numbers there exists a


cardinal number which is greater than every cardinal number in 9Jl.
Proof: If there is a greatest cardinal number among the
clements of IDl, then, by the theorem on p. 21, there exists a
36
cardinal number greater than this one. If there is no greatest
cardinaJ number among the elements of rol, the assertion follows
from Theorem 4.
From Theorem 5 follows a fact which is of very great im­
portance for the theory of sets. According to this theorem,
namely, there is no set which contains all cardinal numbers;
the "set of all cardinal numbers" is therefore a meaningless
concept. This comes as a great surprise after the procedure we
have been following thus far, because up to now it has been
regarded as self-evident that the formation of sets is subject to
no restrictions. The result just obtained shows that such an un­
restrained formation of sets may lead to contradictions. A dis­
cussion of this first antinomy, as well as others, will be post­
poned until we have a more extensive fund of the results nnd
arguments of the theory of sets. Cf. the concluding remarks on
p. 135.
The product of two cardinal numbers f· m can now be re­
garded also as a "sum of f cardinal numbers m". For if the set
,fi'(k) has power ll, and if to every element of .fi'(k) we make
correspond the same cardinal number m, the sum of this com­
plex is precisely a sum of r equal terms m. According to the
definition of sum, the m's corresponding to the different ele­
ments k must be represented by mutually exclmlive sets. This
can he effected hy taking an arbitrary set !lJl(m) with power
m and with no element in common with .fi', and letting corre­
spond to every element k the set !lJl( I k, m } ) . Then the set
Liea rol(/k, m } ) represents, on the one hand, the sum of r
terms m, and, on the other hand, is equal to .fi' X rol, so that
it also represents the cardinal number r · m.
The proof of the equation a · c = C on p. 30 really rests on
this connection between addition and multiplication. Further,
it now follows immediately from the results of §7, that a sum
of enumerably many ones gives the cardinal number a, and a
sum of c ones yields the cardinal number c. Finally, a sum of
enumerably many eardinaJ numbers a is also equal to n ' a = a.
From these results we obtain, e. g.: 1 + 2 + 3 + . ,
. = aj
for we have
37

a = 1 + 1 + 1 + ··· � 1 + 2 + 3 + ···

� a + a + a + . . , = a · a = a.
Moreover,

= (t + 2 + a + . . . ) + (n + n + n + . . .)

+ (c + c + c + . . . )

= a + a · a + a · c = a + a + c = c.
9. The Product of Arbitrarily Many Cardinal Numbers
Since the definition of product will again be bused on the
combinatorial product, we first generalize the latter notion.
Let a set !r(k) be given, and to every one of its elements k
let there correspond a set IDl• . These sets IDl. arc assumed to
be disjunct. From the set of sets .ft'(IDl.), form a new set by
replacing every set 9.n. by an arbitrary one of its elements m• •
The sets .ft'(m.) arising in this manner shall be taken as elements
of a new set, called the combinatorial product, or product of
the second kind, of the sets IDl. ; in symbols: x fI.ea 9.n• . If
at least one of the sets IDl. is empty, the product shall be O.
If, e. g., !r is enumerable, and 9.n1 = la, b l , IDl2 = I e, d l ,
9.n3 = 1 3 \ , M. = 14 \ , ' " , then the combinatorial product is
{ la, c, 3, 4, . . \ ,
. la, d, 3, 4, . . . \ ,

1 b, c, 3, 4, . . . \ , I b , d, 3, 4, ...\1.
Note that the elements of the union of given sets are invariably
elements of these sets, whereas the elements of the combinatorial
product are always sets of elements of the given sets.
Now let $t(rn.) be a given cardinal-number complex arising
from !r(k) ; i. e., to every clement k of �(k) let there correspond
a cardinal number rnA . Represent each of the cardinal numbers
111, hy u. set IDl. , and in such a manner, that the 9.n.'s arc
38
mutually exclusive. Then the product of the cardinal-number
complex shall be defined as

IT m. = , xn rol. , .
l E Il IE Il

It is clear that this definition includes as a special case the


one given in §7 for two factors. We must now show that the
product of the cardinal numbers is independent of their partic­
ular representatives. To this end, let the IDl.'s be sets which
are equivalent to the rol.'s, and keep fixed a one-to-one corre­
spondence between their clements m and m which arises from
this equivalence. In each Sf(m.), which is an clement of the
combinatorial product of the rol.'s, replace every element m
by the element 'iii corresponding to it. The result is an element
Sfcm.) of the combinatorial product of the rol.'s, and if we let
it correspond to .!r(m.), we have the desired equivalence be­
tween the two combinatorial products.

ThEOREM 1 (commutative law). Two cardinal-number com­


plexes Sf(m) and � (n) have the same product, if every cardinal
number appearing in the one complex also appears, and with the
same multiplicity, in the other complex. Preci8ely formulated,
this means tile foUowing: Let .fiI,{k) denote ti,e 8ubset of elements
k of � (k) to which the same cardinal number IJ corresponds. Then
the different Sf.(k)'s are mutuaUy exclusive. Define the sets �,(l)
in like manner. Then
L .re.(k) = �(k), L {!.(l) = �(l) .
, •

It i8 assumed, now, that, lor every IJ,

(1) .re. (k) ,..., �,(l),


and it is asserted that
n m. = n n, .
lEa IE t

Proof: Represent all the cardinal numbers m. , corresponding


to the elements k of .re(k) , by disjunct sets rol. ; and all the
cardinal numbers 0, , corresponding to the clements l of {!(l),
39
by disjunct sets \)1, . Because of (I), there is 11 ooo·to-onc
correspondence between the clements k nnd l of ench pnir of
sets .Il',(k) and �. (l). Keep this mnpping fixed. To the clement
k, let there correspond nn clement l = �(k). To these clements
then correspond sets ml� and 91 , of the same power. 'l'hcoo sets
cnn therefore be mnpped on one nnother. Under this mapping,
let the clement n l = 1J-(mt) correspond to the clement. flit .

l\eep this mapping fixed too. Now, in cvery clement 5I'(m,) of


the combinatorial product. of the ID1.'s, replace ench ml by
111 = y,(ml). Since every k belongs to exactly one .!r.(k), and
every I belongs to exactly onc !!,(l). we obtain in this manner
precisely one clement. i'!(n,) of the combinntorinl product of
the \)1/5. These elements .Il'(m.) and 12(11,) !lfC now made to
correspond to ea.ch other. This determines a mapping betwccn
the eombinnlorilll products of the IDll's nnd the 91,'s, and hence
the I\SSCrtion is provcd.

'l'tIEOHE�t 2 (nssodlLtive Inw) . ']'0 cvcry ClClllCllt l of a set �(l) ,


let therc correapond a ad .I?,(k). Thesc sets Sl', (k) arc CUlSUmed to
be mutually c:rciusivc. [.Jet .I?(k) :L'EI ,!l',(k), alUl to even} k
",.

of .Q'(k) let there corrc8pomi a cardinal numbcr IIIl . Then a


cardinal�numbcr complcz R(ntl) GIld a act of cardi,lal-number
complexes .fl', ( l11l) arc (lcfincd. 'J'hc asscriioll, 1I0W, £8 that

(2) n n IIIl = n
l EI
1111 •

H l ' IEt,
Proof: Represent the cardinal numbers Tn, by disjulld sets
ID'I• . Then the product niH, III, is represented by the set
with thc clements .fl'.(1II,). For every l E 13 we get a combi­
natorial product with these clements, and the sets which we
obtain for the different l's nre mutually exclusive. Hence, the
left-hnnd side of (2) is represcn�d by II set. whose clements nrc
13(.fl',(m,»), i. c., by a set whose clements result. from 13(l) by
replacing cvery l by n. set .I?,(1II.). Each 13(.I?,(ml» contains
cxaeUy one clement of every im, and therefore de�rmines
I

precisely one clement .fi'(ml) of the combinatorial product of


all the ID1,'s. Let this clement Sl'(m,) correspond, now, to
�(.I?,(m,» . The set which represcnts tho loft-IHmd side of (2)
4()
is thereby mapped on the combinatorial product of all the
IDl.'s, and the validity of (2) is thus established.
The general form of the distributive law has hardly any
application, and will therefore not be formulated here.

THEOREM 3. lA:t �(m.) and .re(n.) be two cardinal-number com­


plezes arising from the Bet �(k), and let m.. S n. for every k. Then

IT

ea
m. S IT n.. .
..ea

Proof; Represent the caTdinal numbers mt by disjunct sets


IDl. , and the n.'s by disjunct sets m• . We are to show that the
combinatorial product of the IDl.'s is equivalent to a subset of
the combinatorial product of the iTh's. By hypothesis, every
IDl. is equivalent to a subset �.. of 91.. . This gives rise to a
mapping of IDl.. on m• . Keep this mapping fixed. To each ele­
ment �(m..) of the combinatorial product of all the IDl..'s we
now let correspond that se(1i.) which results from se(m..) by
replacing every m. by the ii.. corresponding to it. The combi­
natorial product of the IDl.'s is thereby mapped on the combi­
natorial product of the �.'s, i. e., on a subset of the combina­
torial product of the m.'s, which proves the assertion.
The principal application of the general definition of product
is to introduce the power. We conclude this paragraph by
listing merely a few simple consequences of, and remarks con­
cerning, what has preceded; their justification may be left to
the reader.
a) A sum I)f cardinal numbers is equal to zero, if, and only
if, each of the terms is zero.
b) For every cardinal number m we have m + 0 = m, and
o is the only number which possesses this property for every
number m.
c) A product of cardinal numbers is equal to zero, if, and
only if, at least one of the factors is zero.
d) A product of cardinal numbers is equal to 1, if, and only
if, each of the factors is equal to 1.
e) For every cardinal number m we have m · l = m, and I
is the only number which possesses this properLy for every
number III.
f) One might. try to define subtraction and division of cardinnl
numbers, in order to proceed from these perhaps even to IUl

introduction of negative !lnd fmctional cardinal numbers. For


naturnl numbers, Lho difference n - 111, or the quotient 71/111,
is defmed lUI the solution, if it exists, of the equation m +
x = 11, or m·x = 11, rcspec�ively, (the latter only for 1» ;o'! 0),
This dcfinitioli is possible herc becausc cach of these equations
hus at most one solution in the domain of natural numbers.
If one attempt.s to imitate this in the domain of cllrdinal num­
bel"8, it fails, bccnusc these equations can have actually in­
finitely mnny solutions in the domain of cnrclinal numbers, ns
the following examples show:

(l + 1 = a+ 2 "" (l +3 = =
f1 ;

= a·2 .,. f1 · 3 = fl .

10. The Power


The introduction of the power, which leads to some quite
interesting result-s, will be carried out first according to t.he
pattern of ordinary nrithmelic. Let f ,.e 0 and III be t.wo arbi­
trary cardinal numbers. Then m t is defined IlS a product. of
f factors UI. When formulated ill greater detail, this means the
following: Lct. there be givell n sct 5l'(k) having cnrtiinlll numher
f. With every clement k we IIssociztte the Slime cardinal number
m. Then the power 1lI r shalJ be the product of the cardin!!l­
number complex .11' (111); i. e., m t = nuc� Itt. For lit ,.e 0, we
put mO = 1 .
I n particular, according to this definitioll, 1 t
'
I nnd 0 = 0 =

for r ¢ o.
'
It W8.S showil on p. 38 thnt Ill docs not depend on the
pnrtieular representative of the number III. That the power is
!Ilso ilHiepcndent of the particulnr representative of the cardinal
38.
number r follows directly from the commutative law on p.
G. Cantur introduced the power in a somewhat different.
manner, which is independent of the gcneral producL. This
42
definition is perhaps more transparent, and is particularly
suited for many applications. We shall arrive at it here by
specialization from our general definition.
According to our definition we may proceed as follows: Let
IDl(m) be a set having cardinal number m. We form new sets
IDl«k, m» from it by replacing, for an arbitrary, fixed element
k E -fi', every clement m of IDl by the pair of elements (k, m).
The sets IDl. = IDl(k, m) thus obtained are mutually exclusive
for the different k's, and their combinatorial product repre­
'
sents the power m . The combinatorial product is formed,
however, in such a manner, that in .re(k) every element k is
replaced by an arbitrary clement of IDlIl , i. e., by an arbitrary
pair of clements (k, m). Here the m's need by no means he
different for different k's, because for different k's the dis­
similarity of the pairs of elements is already taken care of
by the prefixed k. The sets .!1«k, m» arising in this way con­
stitute the clements of the combinatorial product of the IDlIl's.
Every .R'«k, m» can, however, also be described thus: With
every clement k of Jt we associate an arbitrary clement m of
IDl. We have here a process which is analogous to the con­
struction of a function y = lex); for, the function is also defined
by a correspondence of certain values y to the values x. This
association of elements m with the elements of ore Cantor calls
a covering of the set .re with the set IDl. The set of all such
coverings of the set Jt with the set IDl is called the covering
set .re I IDl. The power m ' can accordingly be defined also as
follows: Cover a set .re having cardinal number f, with a set
IDl having cardinal number m. The covering set ore I IDl then
represents the cardinal number m ' .
For example, represent the cardinal numbers r = 2 and
m = 3 by the sets .fi' = { I, 2 1 and IDl = { I, 2, 3 1 . Then the
covering set is

{ {(I, I ), (2, 1» ) , { (I, 1), (2, 2» ) , ({ I , 1), (2, 3» ) ,


1 ( 1 , 2), (2, I) } , {(I, 2), (2, 2) 1 , {(I, 2), (2, 3» ) ,
{(1, 3), (2, I) } , {(I, 3), (2, 2» ) , ( 1 , 3), (2, 3) I L
43
and i t has cardinal number 9 = 32• The point here is to cover
the first place with one of the numbers 1 , 2, 3, and, inde­
pendently of this, to make an analogous covering of the second
place. This gives 32 cases, as is easily seen even without actually
executing the covering. In general, for finite cardinal numbers,
m I yields the ordinary concept of power; of this one can con­
vince oneself in the above manner, or it can be inferred from
the fact that in this case the product of the cardinal numbers
coincides exactly with the elementary product.
a) With the aid of the power concept, the cardinal number oj
the set U(9l) oj all subsets of a given set �n can now be given
immediately; it is I U(9l) I = 2", where n = 1 91 1.
For let 91. b e a given subset of 91, and cover 91 with the set
(O, 1 ) in such a manner, that "ith every element of 91 is
associated the number 0 or 1 according as this element docs or
does not appear in the subset 911 • The covering set 91 I \0, 1 )
thus obtained is obviously equivalent to U,91) , and has cardinal
number 2".
b) From n) and the theorem on p. 2 1 it follows at once that
2" > n,

and this also holds, trivially, for 11 = o.


We shall now prove the customary rules of operation for
powers.
c)
Actually we shall need this rule in the following more general
form:
Cl) If, with every element 1 of a set �(l) , there is associated
a cardinal number fl ¢ 0, then for every cardinal number m
we have
Z I,

(1) n m l' =
m lE a •
lE t

This follows directly from Theorem 2 on p. 39, if, for the


sets �, (k) there, we choose sets having cardinal number fl'
and put every m. = m . For then, according to the definition
44
of power, the left-hand side of equation (2) on p. 39 is jJ,lst
the left-hand side of (1), and the right-hand side of (2) is the
right-hand side of (1).
d) for f ¢ o.
This rule too follows from Theorem 2 on p. 39 . For, first,
we take � = ( 1, 2 ) . and for .!Mk) and ,ff, (k) we take two dis­
junct sets, each of cardinal number " and associate with each
k of .ftl the cal'dinal number ml , and with each k of .fi'2 the
cal'dinnl number mt . Then equation (2) on p. 30 states thaL
(2)

where Jf = Jfl + �2 Next, on the basis of the relation .lt l .......,


�2 , we choose a mapping of the sets ,ff. and .ft2 on each other.


From each pair of corresponding elements of the two set� we
form a set (k, k* J , where k E �. and k* E .re2 Then Ltea.

( k, k* ) = .re. The set ,ff. now plays the role of � in Theorcm 2,


and every (k, k* ) plays the role of a sri . To every k E Jfl
corresponds the cardinal number ml , and to every k* E .fi2
corresponds the cardinal number mt , so that equation (2) on
p. 39 now states that
(3)

The assertion d) is a consequence of (2) and (3).


e) (ml) ' = ml•1 for r, I ¢ O.
This follows immediately from c.), if we take � to be a set
of cardinal number I and put every fl = f.
f) If r ¢ 0, then m. :::; m2 implies m� :::; m! .
This is a direct consequence of Theorem 3 on p. 40, since
we have a product of f factors on each side.
g) Ir Q < fl � f2 ' then mi. :::; ml'.
For, let �. and �2 be representatives of the cardinal numbers
fl and f2 , and assume that m ¢ O. From the hypothesis, it
follows that �I '" �2 for a subset �2 C �2 'I and f2 can therefore

be represented by two sets .reI and .re2 , where .fi. C oft, . If


45
.R'3 = jl', -
.ff'l has cardinal number '1 , then tl + tJ t, . =

On account of f), I " :5: m l ,. Hence, multiplication by m "


" "
gives m :::;; m . 1 , m" , =

The result contained in b) is capable of n far-reaching


generalization:

J. KONIG'S TIU:OIH':�1. To every clement k of a aivel! set .It(k),


m. , III 8uch that in­
let there correspond ilLlo cardinal 'lUmbers
variably m. < It, . Then

L: 111. < fT. n • .


lEt .Et

Pruuf: Let the cardinal J1llmbcr� II. he represcnted by tile


disjullct set.s ill• . Since 111. < 111 I
the lilt'S can be represented
by subsets ID1. e m• . Finally, put 91. = 91. - �n. J
nnd denote
the clements of �m. , 1Jl. , �n, by m. , 11. , T• • Then

n n. � 1 "n
lEI
91. I ·
lEt

First. we prove that L IDl. is equivalent. to a subset. of


)(0 = xn 9l. J frum which it t.hen follows that.

(1) L: III. :::; IT II, .

Let any clement In E L 911. be givcn, nnd suppose t.hat it


belongs tu �m, , Slly. With t.his clement. we arc going to associate
1111 clement. of x n . Every clement of x n is 1\ set. 5l'(n.). We
now choose a .R'.. = R(n.) in sueh a manner, that the element
Itl E I)�, appearing in .R' is equal to III, and t.hat n. for k j<!. l is
equal to all r• . We let. t.his $1'.. correspond to Ill . Then, to
'
different nt S belong different. ft ..'s. ConsequenUy, L: IDlI is
mupped on a subset of xn, and t.his proves (4).
Now we have to show that. the equality sign in (4) docs not
hold. This is accomplished by showing that every set 'll c xn
which is equivalent to L 1m. is II ]Jroper subset of xn , and
we prove this by means of a genemlization of the second
diugullul-method (d. p. 9). Let IU ......, :E IDl. lLiHI Il.\ c xn.
Theil c\'ery WI.. is mapped 011 a part, tI. , of IU, llild L .. c.t

�l.. = 'P. Denote by ft.(n.) the clements belonging to tI. (,I!'..


46
has nothing to do ,,;th the Jr defined previously). IDl" '" �. I
..

and consequently to every m" E ID1. there corresponds exactly


one .n'" E \}3" in a one-to-one manner, and hence also a "diagonal
clement" n" of the set �,,(n.) in a unique (but perhaps no
longer one-to-onc) manner. Thus the set of distinct diagonal
clements n" which, for fixed h, can appear in the Jr,,(nJ:)'s, has
no greater cardinal number than I IDl. 1 = mIl < u" • Denote
by 91: the set of these different diagonal elements nit • Then
91: c m. , and hence m: = 91" - 91: is not empty. We now
choose an arbitrary rt E mt, and thereby obtain such an
element rt for every h E �. With these clements we form the
set .sl(rt) . This set is an element of xTI, but not of �. For,
every clement of \}3 belongs to exactly one \}3" ; and .R (rt)
differs from every �,,(nJ:) E �A , because the elements with
the index h E � in �A(n.) and .It(r:) are certainly distinct.
Hence wo have indeed \}3 C xTI, which completes the proof
of the theorem.
If we choose, in particular, In = 1 and n = 2, then, for t =
I sr I, the theorem yields
f = L 1 < TI 2 = 2 ',
.el lEI

which is inequality b) once more.


H we ehoose for � an enumerable set, and put
� = ml + m2 + m, + . . . J
p :::o m. · m2 · m3 • . .

for
o < nt. < 1n2 < rna < . . . ,

then the theorem gives

rnl + rn2 + rna + . . . < rn2 · rna ' nt, . . .

and consequently

� < p = rnl • m2 • rna • • • ::; 8".


47
By exponentiation it follows from this that

and hence
s" == Il",

II, Some Examples of the Evaluation of Powers

We shall first evaluate a few powers, and then interpret the


results in various ways.
a) m" < a, a
"
== Q, c
" = c for every finite cardinal number
n F- O. The first relation is trivialj the other two follow from
the equations Q . Q == a and c · c = c obtained on pp. 30 and 3 1 .
P) 2" = 10" = n" = Q" = c" = c (n ;;:: 2).
The fact that 10" = c follows from the definition of c and
from b) on p. 27. For, 10" is represented by the set of all
coverings of the enumerable set { I, 2, 3, . . . } with the numerals
0, 1, 2, , . . , 9. Each individual covering gives rise to an infinite
decimal fraction 0 ' ala2/la , . , , and the set of all these infinite
decimal fractions is obviously equivalent to the covering set.
These infinite decimal fractions fall into two classes. The first
class consists of those in which numerals different from 0
appear again and again; and the second class, of the rest. The
first class is a set which is equivalent to the points of the in­
terval (0, 1); and the second is enumerable, because it repre­
sents only rational numbers. Hence, 10" = c + a = c.
One can prove in a similar manner, that 2" = c and n" = c,
by representing these powers by covering sets, and the indi­
vidual coverings by dyadic or n-adic fractions. Suppose 2" = c,
say, to 'be proved in that way. Then, as we shall see imme­
diately, n" = C can be proved in yet another manner.
'
The fact that c = c now results from 2" = c according to
the rules of operation as follows: We have
'
c = (2')' = 2' " == 2' == c.
From this we get, for n ;;:: 2,
48

c = 2' � {::} � . c· = c.
Thit; proves all the relations in /3) .
From this it follows, e. g., that if in the sequence of cardinal
number:; 01 , 02 , 03 , • • • we have invuriably 2 � Ok � c, then
0 . ' 02 ' 03 • • • = c.
For,
c = 2' � a. · n2 · a� . . . � c· = c.
Ben('(', in particular,
a! = 1 · 2 · 3 . . . = c.
Now we give, to begin with, two interpretations of the niles
in a).
a) Let IDl be a finite or an enumerable set. From the elements
m of ID1, form, for a fixed number n, all n-termed complexes
(ml , m 2 , • • • , m.) . Here the m.. 's need by no means be distinct
clements of IDl. We thus obtain the coverings of the set
( I , 2, . . . , n ) with the set IDl. This covering set has cardinal
number I IDl I" � a. The set of all n-termed complexes for
n = 1 , 2, 3, . . . then has cardinal number :E I IDl I " � a · a = a.
"
If, in particular, we take IDl to be an "alphabet", then we
can say (Hausdorff [ I ], p. 61) : "From an 'alphabet', i . e.,
a finite set of 'letters', we can form an enumerable set of finite
complexes of letters, i. e., 'words', some of which, of course,
like abracadabra, are meaningless. If further elements, such as
punctuation marks, printing spaces, numerals, notes, etc . , are
added to these letters, it is clear that the set of all books,
catalogues, symphonies, operas, is also enumerable, and would
remain enumerable if we were to employ even an enumerable
number of symbols (but for each complex only a finite number) .
On the other hand, if we have a finite number of symbols,
and we limit the complexes to a certain maximum number of
clements by declaring words of more than one hundred letters
and books of more than one million words inadmissible, then
these sets become finite. And if we assume, with Giordano
Bruno, the existence of an infinite number of heavenly bodies
49
with speaking, writing, and music-making inhabitants, then it
follows with mathematical certainty, that on infinitely many
of these heavenly bodies the same opera must be performed,
with the same libretto, the same names of the composer, the
librettist, the members of the orchestra, and the singers."
b) Space of n dimensions contains "just as many" points as
the segment (0, I). For, every point in n-dimensional space is
determined by a complex of n numbers Xl , X, , • • • , X" Each •

of these number complexes is a covering of the numbers 1 ,


2 , . . . , n with real numbers. The set of all these coverings has
cardinal number c· = c.
From the results i n fJ) we can infer the following:
c) Between 0 and 1 there are "just as many" dyadic frac­
tions as there are decimal fractions or n-adic fractions, and the
set of all dyadic fractions between 0 and 1 comprises "just as
many" elements as the set of all dyadic fractions. This follows
from the fact that the set of n-adic fractions is always equivalent
to the set of all coverings of an enumerable set with the num­
bers 1, 2, . . . , n, whether we consider all of these fractions or
only those lying between 0 and 1; and that this covering set
has cardinal number n° = c, so that its cardinal number is
independent of n .
d) A point in space of enumerably many dimensions is under­
stood to be a system of enumerably many numbers Xl , X2 ,
Xs , • • •• The totality of all points in space of enumerably
many dimensions is "just as large" a set as the set of all points
of the interval (0, 1). For, the first set is the set of all coverings
of an enumerable set with the set of all real numbers, and i t
therefore has power CO = c.
e) If we understand a laUice point in space of enumerably
many dimensions to be a system of whole numbers XI , X2 ,
Xa , • • • , then we can say that the "number" of all lattice

points in space of enumerably many dimensions "coincides"


with the "number" of all points in space of enumembly many
dimensions. For, the first set is the set of all coverings of an
enumerable set with an enumerable set, and it therefore has
cardinal number n° = c.
50
f) The set of all real, continuous functions has the same
power as the set of aU real, analytic functions, be these func­
tions defined only in an interval or for all values of the argu­
ment; each set has power c (Borel).
For, each of the sets contains, in particular, the constant
functions. Their power, however, is c, since they can assume
an arbitrary real value. Therefore the power of each of the
two sets of functions is at least c. To obtain an upper estimate,
we need consider only the continuous functions, because the
class of analytic functions is contained in the class of con­
tinuous functions. Now a continuous function is already known
if it is known at merely an everywhere dense set of points'
(where, however, it cannot, of course, be defined completely
arbitrarily), say at the rational points. But the set of all func­
tions defined at the rational points can be regarded as the set
of aU coverings of the rational numbers with the real numbers,
and this covering set has cardinal number c" = c. The cardinal
number of each of our sets is thus �c and � c, and is therefore
equal to c.
Remark. The same result is obtained if the functions are
allowed to assume only the values of a certain interval.
Concerning powers with exponent c, we have
2· = n· = It' = c· = f (n � 2) .
For, f was defined as the cardinal number of the set of all real
functions in the interval (0, 1). This set, however, is the set of
all coverings of the continuum with the set of all real numbers,
so that its cardinal number can also be represented by c' . This
proves the last of the above relations. From this follows
2· = 2" " = (2")' = c' = f,
and hence

f = 2· � {::} � c· = f,
which proves all the equations in 'Y).
'A set of points is said to be everywhere dem6 on a straight linc, if every
subinterval contains at least. one point or the set.
51
Now we infer, e. g. :
g) The set of aU real functions which assu me only the values
o and 1, has the same cardinal number as the set of all real
functions of one variable. For, the first set of functions has
cardinal number 2 t = f.
h) The set of all real functions of enumerahly many variables
has the same cardinal number as the set of real functions of a
single variable. For, the points of space of enumerably many
dimensions constitute a set of cardinal number c. Therefore the
set of all real functions of cnumerably many variables has
cardinal number c' = f.
As a simple application of the rules of operation, we note, in
addition, that
r = fO = ft = f (n ;?! 1 ).
CHAPTER III

Orderetl Sets and Their Order Type.Ii

1. Definition of Ordered Set


In the preceding chapter, the concept of "number" was
extended, and in such a manner, that transfinite numbers were
added to the finite ones. These numbers, finite and transfinite,
were called cardinal numbers. Now already in connection with
the ordinary process of counting finite sets, every number plays
a double role. Grammar, too, distinguishes numbers into
cardinal numbers (fundamental numbers) and ordinal numbers
(order numbers) . For with finite sets, numbers can also be
used to determine a certain order of succession of the elements.
For instance, one can stipulate that in the set l a, b, e) , the
first element shall be e, the second, a, and the third, b. It is
then convenient to write the set in the form Ie, a, b} . Which
of two distinct elements precedes the other is hereby deter­
mined. Such determinations will now be carried over to arbi­
trary sets.
Sets for whose elements a certain order of succession has been
prescribed play an important role in applications. If, e. g., a set
of real numbers is given, then of every pair of distinct numbers
of this set, one is the smaller of the two. Any set of real numbers
can thus be ordered in an entirely natural fashion according
to the magnitude of its elements. An order for the points on a
segment suggests itself in just as natural a manner.
Since the clements of sets need by no means be numbers, or
points of a segment, and; furthermore, even if the elements
are numbers, it is not at all necessary for them to be ordered
according to magnitude (e. g., the order of succession l a, 1, 2 }
can be prescribed), the concept of ordered set must b e formu­
lated quite abstractly, We thus arrive at
Definition 1 . A set ID1 is called an ordered set, provided that a
relation, denoted, say, bY the symbol -<, subsists between every
52
53
pair of di8tinct elements a and b of the 8et, and only between dis­
stinct elements, and 8atisfies the following two conditions:
1. If a ¢ b, then either a -< b or b -< a.
2. If a -< b and b -< c, then invariably a -< Cj i. e., the relation
i8 transitive.
We shall regard the empty set, as well as a set consisting of
a single element, as orderedj likewise, a set consisting of two
elements, provided that an ordering relation is given for its
two elements a, b, according to which precisely one of the two
relations a -< b, b -< a holds.
The relation a >- b shall mean the same as b -< a. Finally,
the symbol a -< b is read "a precedes b"j and the symbol a >- b,
"a succeeds b".
This last terminology is by no means intended to give the
symbol an intuitive meaning; it serves merely to facilitate
comprehension. Moreover, the meaning of the symbol -< may
change from set to set. For instance, in the case of sets of
numbers, it may coincide at times with the symbol < , and at
other times, with the symbol > . A few examples will serve
to clarify the concept of ordered set.
a) The set of natural numbers 1 1 , 2, 3, . . . } , ordered ac­
cording to increasing magnitude. Here -< coincides with < .
b) The set of natural numbers, ordered according to de­
creasing magnitude. We shall indicate thi� order by writing
I · . . , 3, 2, I } . Here -< coincides with > .
c) The set of all i ntegers, ordered according to increasing
magnitude. This order is symbolized by
I · · · , - 3, - 2, - 1, 0, 1 , 2, 3, . . . }.

Here -< coincides with < .


d) The set of integers, but this time in the order indicated
by 10, I , - 1 , 2, - 2, · · · } i i. e., the nonnegative integers are
ordered according to increasing magnitude, and each negative
integer comes immediately after the positive integer having
the same absolute value. The symhol -< in this cn.se coincides
in part with < , in part with > .
e) t o, 2, 4 , . , 1 , 3, 5, . . . } i i . c., first come the even, nOIl-
. .
54
negative integers ordered according to increasing magnitude,
and then the odd integers.
f) fO, 2, 4, , 5, 3, 1 1 ; i. e., first come the even, nonnegative
• . .

integers ordered according to increasing magnitude, then the


odd integers ordered according to decreasing magnitude.
g) { I, 2, 3, . . • , j-, f, j, .
. . , i, I, -t,
.
. , " , 1 ; i. e., the
.

positive rational numbers ordered according to increasing de­


nominator, and, for the same denominator, according to in­
creasing numerator.
h) The set of all real numbers in the interval (0, 1) ordered
according to decreasing magnitude.
The question whether every given set can be ordered will
not be decided until later (p. 1 10). It is clear that every finite
set can be ordered.
We have obviously
THEOREM 1 . Let IDl be an ordered set. Then, by means of the
ordering principle given for IDl, every subset of g:n i8 alBo ordered.
Hence, the 8Ub8ets of ordered sets can always be regarded as ordered.
Definition 2. Two ordered 8ets g:n and 9l are 8aid to be equal, in
81J71Ibols: g:n = 91, if they contain the same elements, and if the
ordering relation a -< b valid for g:n invariably implies the relation
a *-< b valid for 9l. Conversely, then, it follows obviously Uwt
a .*-< b alway8 implies a -< b.
Thus, the sets a) and b) are to be considered as distinct in
the sense of ordered sets. Without regard to an order, however,
they afe equal.
Finally, we introduce the following terms: If, for three
elements a, b, c of a set, we have a -< b -< c, then we shall say
that b lies between a and c. If an ordered set ID'l contains an
element a such that for every b E g:n with b ¢ a the relation
a -< b holds, then a will be called a first element of the set IDl.
If g:n contains an element c such that for every b E IDl with
b ¢ c the relation c >- b holds, then c will be called a last
element of IDl.
THEOREM 2. An ordered set contains at most one fir8t elc/nent
and at most One last element.
55

Proof: If the set contained, e. g., two first elements a and a,


we should have a -< a as well as a -< a, which is impossible.
An ordered set, however, need have neither a first nor a last
clement. Examples of this are the sets c} and h}.

2. Similarity and Order Type


For ordered sets, those mappings oC one set on another will,
naturally, be oC importance, whieh, to put it briefly, preserve
the order of succession of the elements. In this way we arrive
at the following sharpening of the equivalence concept:
Definition 1. A n ordered set IDl with the ordering relation -< is
said to be similar to an ordered set 91 with the ordering relation *-<,
in symbols: IDl � 91, if the set IDl can be mapped on the set 91 in
such a manner, that if ml , � are any two elements of IDl, and
nl , n2 are the corresponding elements of 91, then m, -< m2 always
implies n, *-< n2 It follows then obviously that n, *-< n2 always

implies that, for the corresponding elements, tnl -< 7n2 • Such a
mapping i8 called a similarity mapping.
The four fundamental properties result immediately from
the definition:
ex) IDl � IDl, i. e., every ordered set is similar to itself.

fJ) IDl � 91 implies 91 � gn.


'Y) If IDl � 91 and 91 � �, then IDl � IP.
8) IDl � 91 implies IDl ,...., 91.
Let us first clarify the concept of similarity by a few examples.
a) If an ordered set is mapped on itself, it is by no means
necessary that every element correspond invariably to itself.
If, e. g., IDl is the set of numbers in the interval (0, 1) ordered
2
according to increasing magnitude, then the function y = x
maps this interval on itself. No point corresponds to itself,
however, under this similarity mapping.
b) An ordered set can be similar to a proper subset. For
example, { I, 2, 3, . . . } � { 2, 3, 4, . . . } , and the set ( I, 2, 3, . . . I
is similar to the set of prime numbers if these are ordered ac­
cording to increasing magnitude, because there are infinitely
many prime numbers.
c) The set of rational numbers can be ordered in such a
56
manner, that it is similar to the ordered set { I , 2, 3, . . . } .
This is accomplished by the order of tho rat-ional numhers
defined on p. 3. Morc generally, we have

THJ.;oRf�l\I 1 . If a 8ft 9? i.� fqll;ml1cnt to all o rdered set rol, thcn


9l can. be urdacd in lweI! a 'W(/!J Ihat 9l � 9Jl.
Proof: If ml , m2 are two clements of 9,)'l, and if tnl -< m, ,

then, for the corresponding elements nl , 112 of 91, we stipulate


that nl -< 112 This defines an ordering relation for the elements

of 9l; and the relation is transitive within the set 91, because,
by hypothesis, it possesses this property in the set IDt According
to Definition 1 on p. 52, the set 91 is ordered by this relation.
The following theorem is sometimes useful for investigating
the similarity of given sets.

THEonJolM 2. If two sets are bimilar, then eithel' both possess


first (last) eleme1lts or neither of the two sets has such an element.

Proof: If one of the sets has a first element mo , then for


every other element m of this set we have mo -< m. Let no
denote that element which corresponds to mo under the simi­
larity mapping of the one set on the other, and likewise let n
be the element corresponding to m. Then, according to Defini­
tion J , no -< n for every n � no ; i. e., the set 91 also has a first
clement.
d) The sets a) and h) on p. 53 are not similar, becau:;e a)
has a first element and b) does not. Further, with the aid of
Theorem 2 follows, e. g., the dissimilarity of the sets a) and
c) , b) and c), a) and f), b) and f), h) and f), h) and a) .
e) If we consider the numbers of the two intervals (0, 1)
and \0, I), ordered in each case according to increasing magni­
tude, then the two sets are dissimilar. Consequently, under a
mapping of the two equivalent sets on one another, the order
of succession of the elements in one of the two sets must of
necessity be disturbed, as, for example, on p. 1 5.
f) If IDl is the set of real numbers in the interval (0, 1 ) , and
in is the set of all real numbers, each set ordered according to
increasing magnitude of its elements, then IDl � 91. This follows
57
from the mapping given in c) on p. 1 5, which is evidently a
similarity mapping.
g) Let IDl Le the set of all points P (x, y) i n the plane with
rectangular coordinates .r., y. Order this set 11 0 that

P (x , , V.) -< P(X2 , V2) Cot' x, < X2 as well as COl' x, = X2 , V, < Y2 •

This relation is obviously transitive, so that it actually,defines


an order for the set gn. Further, let 91 be the set of points
P (x, V) of the square 0 < x < 1 , 0 < y < 1, ordered according
to the same rules. Then gn � 91. This is seen immediately by
mapping the sides of the square similarly on the complete co­
ordinate axes, in virtue of f) . It will turn out later (p. i5),
by the way, that the set gn is not similar to the interval (0, 1 ) .
With the a i d of the concept of similarity, w e can n o w once
more introduce a new kind of number according to the pattern
of ch. II, §3:
Definition 2. By an order type J.& we mean an arbitrary repre·
sentative gn 01 a class 01 mutually similar ordered sets. The order
type 01 an ordered set gn will sometimes be denoted by I gn I . The
notation I gn l accordingly means simply that the ordered set gn
may be replaced by any ordered set similar to it.
Since all sets that are similar to each other are also equivalent
to each other, all sets possessing the same order type have the
same cardinal number tooj i. e., Ignl = 1 91 1 always implies
I gn I = 1 91 1 · Consequently, to every order type J.& belongs a
cardinal number, which we shall denote by I J.& I . Thus, J.& = II
implies I J.& I = I II I·
If two finite sets are equivalent, then it i s obvious that, no
matter how their elements are ordered, they are always similar
too. Hence, to every finite cardinal number there corresponds
precisely one order type, which, for brevity, will be denoted by
the appropriate cardinal number. In the case of finite sets, then,
cardinal number and order type coincide.
The order type of the set of natural numbers, ordered ac­
cording to increasing magnitude, is denoted by w:

w = . ( 1, 2, 3, . . . ) 1.
58
If, on the other hand, these numbers are ordered according to
decreasing magnitude, then the order type of the set is denoted
by *w:
*w = . 1 · · · , 3, 2, I } · .

In general, *p. denotes that order type which results from p.


when the order of succession of the elements is reversed, i. e. ,
when a new ordering relation is derived from the relation -<
defined for the original set, by setting *-< equal to >-.
We speak here not of order numbers, but of order types, and
for the following reason. If we wish to compare two distinct
order types, we must say, corresponding to the comparisoQ. of
cardinal numbers introduced on p. 18, that the order type J.&
is smaller than the order type " if a representative ID1 of Po is
similar to a subset of a representative 9l of p. Let us now con­
sider the order types w and *w. Every subset of the set
I . . , 3, 2, I } which represents the order type *w, has a last
.

element, and therefore cannot be similar to the set { I, 2, 3, . . . ,


which represents the order type w. Likewise we see that the
first set cannot be similar to any subset of the second either.
These order types are therefore not comparable. Order types
thus lack an important property possessed by numbers. For
this reason the term order number is not applied at this point.

3. The Sum of Order Types


For the purpose of defining addition of order types, we shall
first introduce an ordered addition of ordered sets.
Definition 1. Let IDl and 9l be two disjunct, ordered sets. We
define an ordering relation for the elements s of their union � as
foUows: Let 81 and s� be two elements of the union. If they both
belong to IDl, we let 81 -< 8, or 82 -< 81 in the union, accordiny as
lIte fir8t or second of the8e relations hold8 for these element8 in IDl.
If both elements belong to 91, then the relation which is valid for
them in 9l shall again hold also in the union. But if one of the
elements belong8 to IDl, and the other belong8 to 9l, say 81 E IDl,
s, E 9l, then we let 81 -< s� in the union. The ordering relation
thus defined for the union i8 obviously transitive, and it therefore
59
orcIcrs e . The ordered sum ID1 + 91 of the ordtred �cts ID1 and 91 is
now understood to be the union @5, ordered in tllC 111 mller indicated.
IVe speak hr.re of an ordr.rcd atlditi()n of the 8('tS ))l amI 91. It 1'8
obt·iolls thai, wlu'n d('alillg IVitil ordered addilion, f,c slim � + 91
is to b,. dislinguished from I},(' sum 9� + �l)�.
For example, if the sums are ord£'r(,ci Imm�,
{ I , 2, 3 1 + ( -1, 5, G, 7 \ = { I , 2, 3, 4, 5, ti, 7 ! ,
hut
{ 4 , 5, 6, 7 } + { I , 2, 3 1 = 1 4, 5, 6, 7. 1 , 2, 3 1 .
Definition 2. To obtain the sum 01 tlte two order types jJ and ",
represent them by two disjunct sets IDl amI 91 a,ul form tlte o/'dered
sum e = ID1 + 91. Then we stipulate that jJ + " = 1 6 1 •
It is again easy to see that the sum jJ + " is independent of
the particular representatives of the two ol'(ler types. FurLher,
from the definition of sum, it follows that
a) I p. + I = 1 1.1 1
II + I " I·
It i s clear that the commutative law holds for the ucMition
of finite order-types. For transfinite order-typcs, however, ruldi­
tion need not be commutative. For if
ID1 = 1 1, 2, 3, . . . } and 91 = 10 J ,
then
91 + ID1 = 10, 1 , 2, 3, . . . J , ID1 + 91 = 1 1 , 2, 3, . . . , OJ ,
so that 1 + w = w, but w + 1 � w, because the order type
w + 1 has a last element, whereas w does not.
We take a few more examples to illustrate addition of order
types:
a) n + w = w.

For we have
n = I { I, 2, . . . , n } I, W = I In + 1, 11 + 2, n + 3, . . . I I,
and hence
n + w = I 1 1, 2, .
. . I n, n + 1, . . . 1 1 = W.
60
b)
For,
*CII + 11. = , f ' " , 11. + 2, 11. + I J ' + , f 11., ••• , 2, 1 } '
= II··· , 11. + 1, 11., • • • , I J '.
c) CII, CII + 1, CII + 2, . • . are distinct order types. For if we
had CII + m = CII "+ 11., with, say, 0 � m < 11., then (d + m would
have to have an element that is followed by precisely 11. - 1
elements, which is not the case.
d) One shows in a similar manner that *CII, 1 + *CII, 2 + *CII,
• . . are distinct order types.
e) The order type *CII + CII belongs, e. g., to the ordered set
f. . , - 3, - 2, - 1, 0, I , 2, 3,
• •
Jj the order type (d + *c.J, to
. • •

the set { I, 2, 3 , , -3, -2, - I J .


. • .

Although the commutative law ordinarily does not hold for


order types, at least the associative law is valid:

(J) (p + II) + r = P + (II + r).


This follows directly from the definition of addition. For if
the order types are represented by the mutually exclusive sets
ID'l, m, �, then for the ordered addition ot these sets we have
obviously OIn + m) + � = ID1 + (m + �).
From this we get, using the results obtained previously:

f) 11. + CII + (d = (11. + (d) + CII = CII + c.J ;

c.J + 11. + CII = c.J + (11. + c.J) = c.J + CII ;

CII _ . CII + 11. = c.J + '" + n.

g) *", + 11. + (d = (*Co) + 11.) + (d = *Co) + Co);

and
are distinct order types.
Addition can also be extended to arbitrarily many order
types. We begin with
61
Dejinition 3. Let an ordered set �(k) be given, and to every
element k let there C01Te8pOnd an ordered set m. , where the m.'s
are mutually exclusive. Then we order the union E5 = L:.e. IDl.
in the following manner: If two elements 81 and s, of E5 belong to
the same mJo , then we let SI -< 8, or s, -< 81 in E5, according as
the fir8t or 8econd relation is valid for these elements in m. . If,
however, 81 and 8, belong to different m.'8, say 81 E m• • , 8, E m•• ,
then we let SI -< 8, or 8, -< 81 in E5, according as kl -< k, or
k, -< kl . The ordered sum L:.e. rot. , now, shall equal the union
E5, ordered as indicated.
Definition 4. Let an ordered set �(k) be given, and to every
element k of this set let there C01Tespond an order type P. i in other
words, let there be given an ordered complex of order types, arising
from the ordered set .re(k). Represent the order types p., by mutually
exclusive ordered sets m. . Then the sum of this complex of order
types shall be.

where the addition of the sets on the right-hand Bide is to be carried


out as an ordered addition.
It is easy to see once more that this sum is independent of
the particular representatives of the order types. Further, we
have obviously:

'Y) I L:
e p.. I = L: I p.. I ·
t l t E ll

Here, too, of course, the universal validity of the commutative


law is out of the question. It is true that there is an associative
law here which goes beyond fj), but we shall not set it up,
because it is not needed in what follows.

4. The Product of Two Order Types


If, in particular, the same order type IJ is llSSociatcd with
every ,. in §3, Definition 4, we arrive at
Dejiuition 1. Let Utere be given two order types " ¢ 0 and p..
Represent the order type " by the Bet .re(k), and with every k associate
62
the order type IJ. Then the product IJ · " shall equal the BUm oJ the
complex oj order types thus determined; i. e.,

For " = 0 we put Jl."K = o.


That the product is independent of the particular repre­
sentative of IJ was already stated on p. 6 1 . It is also easy to
sec that it is independent of the particular representative of IC
too; this will, moreover, follow once more, presently. The
product can also be defined as follows:
Definition 2. Let there be given two order types IC � 0 and
IJ � o . Represent " by .fi' (k) and IJ by IDl(m). Form the set (If aU
pairs of elements (k, m), and drfine a1l order Jor these pairs by
stipulating that
(kl , ml ) -< (k2
, m2) for kl -< k2 as well as Jor kl = k2 ,
ml -< m2 Let e denote tlie set oj pairs oj elements (k, m) ordered

in this manner. e will also be called the ordered product of the


second kind, IDl X .If (note tlie order in which the factors are written
down). Then the product Jl." IC sliall equal l e i . If IC = 0 or IJ = 0,
we put Jl." IC = 0. The names lexicographical order and order
according to first differences readily suggest themselves for the order,
just described, oj tlie pairs oj elements (k, m).
This second definition follows immediately from the first.
The set .If is the same in both . In the first definition we associate
with every k E .If a set IDl. with order type IJ, and all the 9Jl.'s
must be mutually exclusive. This can be done by taking a set
IDl with order type IJ, and replacing the elements m of IDl, for
every k, by (k, m). If we now associate with every k the set
IDl«k, m», the requirement of Definition 1 is fulfilled. The
union of the IDl « k, m»'s is precisely the set denoted by e5 in
Definition 2, and the order prescribed for its elements in Defini­
t ion 2 is also that prescribed in Definition l .
I t is immediately evident from Definition 2, that the product
of t he OI·del" types is indepemlent of their particular repre­
sentatives. Further, it follows directly from each of the two
definitions, that
63

)
ex

For multiplication there is no commutative law either. For if


we first choose I( = W, " = 2, then 2 · w is represented, e. g., by
the ordered set {ai , b, , a2 , b2 , aa , b3 , • • • 1 . and is therefore
equal to w. If we then choose I( = 2, " = w, then Cd ' 2 is rep­
resented, e . g., by { I , 3, 5, . . . ; 2, 4, 6, . . . I . This set is
certainly not similar to the first one, because the second con­
tains elements (e. g., the element 2) which arc precedcd by
infinitely many elements, whereas no such elements appear in
the first set. Hence, w · 2 � 2 · w. With products too, then, one
must pay attention to the order of succession of the factors.
Under these circumstances it would undoubtedly be more
agreeable if the product defined above were denoted by 1( ' "
instead of " ' 1(. Nevertheless, the notation introduced above
has become accepted, and a change of notation at this time
would lead to great confusion in the literature. Consequently,
one must always bear in mind that when considering the
product W I(, the number pairs (k, m), and not (m, k), are to
be ordered according to first differences.
We do have, however, for multiplication, as for addition, the
associative law, which is so important for calculation:
(3)
For let the order types be represented by sets with the cle­
ments m, n, p. Then, according to Definition 2, th('< order type
(W I') "IT is represented by the set of clements (p, (n, m», where
(PI , (nl , m l » -< (P2 , (nz , m2»
if PI -< p � , or PI = = nl , m l -< m2 .
P2 , n l -< n2 , or PI = P2 , nl
Likewise, " . (I' ' 11') is represented by the set of elements
«p, n), m), where
«P I , n l) , ml) -< « P2 , n2), m2)
once more under the conditions just citcd. If we now associate,
for the same elements 1n, n, p, the clement (p, (n, m» with
the clement «p, n), m), this defines a similarity correspondenco
64
between the sets representing (p , p)1r and "(P ' 1r), which proves
the assertion.
The distributive law does not always hold. It does not always
hold, namely, when the first factor is a sum. For example,
(Col + 1) · 2 = Col + 1 + Col + 1 = Col + Col + 1 = ", · 2 + I ,
and this differs from
", · 2 + 1 · 2 = Col · 2 + 2,
because the last order type has a next to the last element,
whereas the first order type does not.
The distributive law is valid, however, if the second fa<�tor
is a sum; i. e. ,
'Y) ,,(II + 11') = "II + ,,'If'.

For, let the order types be represented by the sets 9R(m),


men), �(p), where m · � = O. Then ,,(II + 11') is represented by
the set of all elements (q, m), where q is an arbitrary one of
the clements n, p, and the relation

(ql , ml) -« (q2 , m,)


holds if

or

or

or
ql = q2 , ml -« m, •

N ow it is evident that the same ordered set is obtained as


representative of "II + ,,11'.
It is customary to denote by � the order type of the set of
all real numbers ordered according to increasing magnitude,
65
and by fI, the set of all ra.tional numbers ordered in ana.logous
fashion. Further, the set of real numbers of an open interval,
and the set of rational numbers of an open interval, ordered
according to magnitude, have the respective order types � and
fI too. This is shown by the mapping y = %/(1 + I % I) of the
real numbers % on the interval I y I < 1. The real numbers of
the intervals
(O, 1) , (0, 1 ), (O, 1 )
accordingly have the order types
1 + �, � + I, 1 + � + 1,
and the rational numbers lying in these intervals have the order
types
1 + fI, fI + I, 1 + fI + 1.
If we represent � + 1 by the interval (0, I), and a second
term � by the interval (I, 2), we obtain
� + I + X = X.
By a similar juxtaposition of the intervals which represent the
terms, it follows that
1 + � = (1 + X)n = (1 + �)""

x + 1 = (� + l)n = (X + 1) · *""

X = (X + 1)", = ( 1 + X) · *"' ;
and, analogously,

1 + fI = (1 + fI)n = ( 1 + fI)""

'1 + 1 = ('1 + l)n = ('1 + 1) · *""

fI = (fl + 1)", = ( 1 + fI) ' *""


For the order type 'I, we have, moreover:
66

This is seen immediately by putting, e. g., intervals with irra­


tional end points in a row.
It follows from the validity of the associative law, tha.t we
can speak of a product of a finite number of order types 1-1, 'Il,
• • • 1-'.. without inserting parentheses, because, according to the
associative law, it is immaterial where they stand. The proof
of the associative law shows at the same time, that Definition
2 of the product can be extended so as to provide a. general
dcfinition for a product of 11 order types. If the same order
type is chosen for all the factors of the product, we get the
power p." of an order type with a finite exponent. From a) then
follows:
&) For finite n, , p.'" = , p. ' ''.
The power e,l can be represented, e. g . , by the ordered set
( 1 , 2, 3, . , . ; i, i, i, . . . ; 1, J, -i, . . . j · · · 1

of positive rational numbers.


}t'or the extension of the product to an infinite set of order
types, consult Hausdorff [1], p. 147 or (2), p. 73. This extension
requires the theory of well-ordered !rots.

5. Power 0/ Type Classes


I t docs not yet follow from the results derived thus far, that
to every cardinal number m there corresponds at least one
order type I-' with ' p. ' = m, or, in other words, that every set
can be ordered. Consequently, if we denote by �m the set of
distinct order types which belong to a given cardinal number 111,
and call this set the type class belonging to m, we must at
first reckon with the possibility that �.. = O. From above,
however, we can already obtain an estimate for the power of
the type class. For there is the following
THEOREM 1 . For the type class X.. belonging to the cardinal
m we have· ' �.. I � 2 .. · .. . '
?lllmber
'Actually I to! I = 2" ; see Hausdorff ( I I, p. 455.
67
Proof: Let IDl he an arbitrary, but then fixed, set wit.h cardinal
number nt. Let II he any order type of XCI Then it is always .

possible to order IDl so that IDl acquires the order type II


(Theorem 1 on p. 56) . The set of all the diffC'rent order types
of the class �m is therefore a subset. of the sct. of all possible
orderings of our set IDl. If, now, we have any onlel"ing of £In,
we can form the set of all pairs of elements (m, m) of IDl for
which the relation m -< m holds under the ordering which is
heing considered. The set of these pairs of elements is a subset
of the set of all pairs of elements (m, m) which can be formed
from the elements of the set IDt The set of all possible orderings
of the set IDl, and a fortiori the type class Xm , therefore has
at most the power which belongs to the set of all subsets which
can he formed from the set of all pairs of dement s of rot Now
the set of all pairs of elements of £Ill has (sec the definition on
2
p. 42) cardinal number m , and eonsequently the set of aU
subsets of this set has cardinal number 2m•m, which proves the
assertion.
It follows, in particular, from the theorem, for m = a, that

I l:' I 5 2· ·' = 2· = c.
For this case, however, it is easy to show that the equality
sign must actually hold :

TUEOREM 2. For the class X. of order types of enumerable sets


I X. I = c.

Proof: There remains to be shown merely that I l:. I � c.


For this purpose we make use of the order type " = *w + w,
which has neither a first nor a Inst clement. With nny sequence
of natural numbers

(1)

w e form, b y addition, the order type

(2) or = al + " + a2 + " + aa + II +


We shall show thnt distinct sequences ( 1 ) always give rise to
distinct order types (2). Let
68
(3)
also Le Il. scquence of na.tural numbers, and with it form the
order type
fJ = vJ + " + ba + " + b3 + " + . . . .

Suppose that a = fJ. Then we are to show that the sequences


(1) and (3) coincide. This will follow by induction from the
following auxiliary consideration:
Let � J , {Il and 91 1 , 911 be ordered sets. Then the relations

1
�I � �" 911 � 91,
follow from
{IJ + 921 � tt, + 91, ,

(a) 911 and 92, have no first element,

@I and tt2 are finite sets,

{
n.q well ns from

92 1 + <II � 91, + <I, ,


(b)
I mil = I 92. 1 = II .

For consider case (a). Under a similarity mapping of �J + �nJ


on <i, + 91. , no element of 91, can correspond to an element
of ttl , because every element of (!I is preceded by only a finite
number of elements i n �I + 911 , whereas every element of
91, is preceded by infinitely many elements in i, + 91, . It
follows, likewise, that no element of 911 can correspond to an
element of <!, Consequently �I is mapped on @, and 91J is

mapped on 91, . In case (b) too, under a similarity mapping of


911 + ttJ on 91, + <!2 ' an element el of @I cannot correspond
to an element 11, of 912 ; otherwise the subset 911 �eceding
eJ in 911 + {II would have to be mapped on a part 91. of the
elements of 912 , which is impossible, because 911 has no last
element, whereas m, does. In like manner we see that no
element of NI can correlpond to an element of E, . Henco, also
69
in this case, E, is mapped on E2 and N, is mapped on Nt ,
Q. E.D.
If we now apply (a) to the equal sums a and p, we get
a , = b,
and
I' + � + II + a, + . . · = 1I + b2 + II + b3 + · · · ,
so that, by (b),
02 + II + aa + . . . = b2 + II + b, +
and hence, again by (a),
O2 = b2 and I' + a3 + II + ' " = II + b3 + I' + . . . ,
etc. Thus the two sequences (1) and (3) do in fact coincide.
Since the totality of distinct sequences (1) gives rise exclu­
sively, as we have seen, to distinct order types (2), and since
(2) implies that
I a I � Q . Q = Q,
st. has at least the power of the set of all sequences (1) that
can be formed from the natural numbers. The power of these
sequences of numbers is Q" = c, and this proves the theorem.

6. Dense Sets
In this paragraph and the next, we shall take up several
classifications of order types according to their structure. It is
an interesting fact that many concepts formed at first only for
point sets can be carried over in general to ordered sets. This
will give us means for proving in many cases the dissimilarity
of ordered sets. It will also be possible to describe the structure
of the set of all rational numbers on a line, and, above all, that
of the continuum, solely with the aid of the concept of set.
By the border elements of an ordered set we shall meun its­
first and last dements, if it has such clemelltl!.
An ordered set is called unbordered, if i� is not empty t\ncl
has no border clements.
70
An unbordered set is always infinite, because in every finite
set there is a first as well as a last element.
THEOREM 1. An unbordered set can be similar only to an un­
bordered 8et. Formulated more briefly (c/. Definition 2 on p. 57) :
if lIn i8 unbordered, then 80 is ,lIn'.
This is merely another formulation of Theorem 2 on p. 56.
Examples of unbordered sets are: The interval (0, 1); the
entire number axis; the rational numbers in the interval (0. 1);
the type ."" + "".
If two elements ma and m, of an ordered set IDl have the
property that no element of lIn lies between them, then t hey
are called neighboring (consecutive) elements. If these elements
satisfy the relation ma -< m� , then ma is called the immediate
predecessor of m2 , and m2 is called the immediate successor
of ml . An ordered set is said to be dense, if it contains at least
two clements and no neighboring elements.
A dense set is always infinite, because every finite set (:on­
taining at least two elements has also neighporing elementti.
THEOREM 2. A den8e 8et can be similar only to a dense 8et; i. e.,
if IDl i8 dcn8e, then so i8 IIDlI.
Proof: Let IDl � 91, and suppose that 91 contains two neigh­
boring elements na , n2 . Then there can be no element of lIn,
either, between the elements ma and m, of IDl which correspond
to n a and n2 under a similarity mapping of 91 on lIn, since other­
wise there would also have to be an element of m between na
and n2 •

Examples. Dense sets are: The numbers of the interval (0, 1),
ordered according to magnitude; the rational numbers in the
interval (0, I), in their natural order; the set (0, 1) + 2. Sets
which are not dense arc: { I, 2, 3, . . . } ; the rational numbers
when arranged in a sequence, sincc to every element there now
corresponds an immediate successor; the set (0, 1) + 2, be­
cause the elements 1 and 2 are neighboring.
If we confine ourselves to enumerable dense sets, there is the
following important
71
THEOREM 3. AU unbordered, dense, enumerable seta are simill1r
to one another. (G. Cantor)

Proof: Let IDl(m) and m(n) be two sets of the kind mentioned
in the theorem. Since they are enumerable, they can be written
as sequences, whereby their given order may, of course, be
disturbed. When written as sequences, let them be

IDl = I m ' , m il, m il' , . . . } and m = In ', n il, n ' '', . . } .


.

We have to show that the sets IDl and m with their original
orders can be mapped on each other. To this end, let m1 = m'
correspond to any element nl of m (sa� nl = n') .
We then look for the element in m - I nd with smallest
upper index, and call this element n, . In m, now, either nl � n,
or nl >- n, . According as the first or second is the case, we
choose in IDl an element m, with ml -< m, or ml >- m, i this
is possible because IDl is unbordered. We associate this m,
with n, .
Now we go over once more to the set IDl, and in IDl I ml , m, 1
-

we choose the element with smallest upper index, calling this


element m3 It may appear before. after, or between the ele­

ments ml and m, under the given order of IDl. In each case,


since m is unbordered and dense, there is an element na which
has the same position relative to nl and n, as ma has relative
to the elements ml and m, .
N ow we consider m again, choose in m I nl , n, , n3 l the
-

element with smallest upper index, call it n. , and select an


element m. of ID1 which has the same position relative to the
three elements ml , m, , ma as n. has relative to the three
elements nl , n, , n3 .
In this way, every element of IDl, i. e., of IDl, and every
element of m. i. e., of m, is chosen. Thus. the set IDl is mapped
on the set m, without disturbing the order of succession of the
elements. For let ml -< m, be two arbitrary elements of IDl;
and let nl and n, be the corresponding elements of m. Then
the correspondence betw�en one of the pairs of elements ml ,
nl and m, , n , is set up, according to our procedure. later than
72
that between the other pair; let us assume that it takes place
later for the second pair than for the first. Now in making this
correspondence between m, and n2 we make sure that the
clements ml , m2 have the same relative position in IDl as the
elements nl , n2 have in 9l. Hence we have actually IDl � 9l.
By means of Theorem 3, the order type 'I of the set of all
rational numbers in their natural order is described, and this
without reference to these numbers themselves. For, according
to Theorem 3, the order type 'I belongs to every ordered set
IDl possessing the following properties:
a) IDl is enumerable;
(3) IDl contains neither a first nor a last element;
'Y) IDl is dense, i. e., between every pair of its elements there
is at least one further element of IDl.
Theorem 3 can be generalized to
THEOREM 4. The four types
'I, 1 + "1, 'I + I, 1 + 'I + 1

are the only distinct order types of dense, enumerable sets. That
is, every dense, enumerable set is similar to the naturally ordered
set of rational numbers in one of the intervals (0, 1), (0, I), (0, I),
(0, 1).
Proof: By deleting its border clements, every enumerable,
dense set can be made into one of the sets treated in Theorem
3, i . e., into a set of type 'I. If we now replace the border de­
ments, the assertion follows.
The method of proof of Theorem 3 also yields
'lHEOREM 5. Given any dense set and any ordered enumerable
set, U,ere is always a subset of the former which is similar to the
latter.
Proof: Let the given dense set be made into an unbordered
dense set IDl by leaving out any border clements, and let it
be the given ordered, enumerable set. We have to show that
IDl contains a subset which, under the order prescribed for it
by IDI, is similar to �(. Let us write the set �I agnin as Il. sequence

it = { aI , at , aa , . I ,
. .
73
disturbing, possibly, the given order. With 41 we associate an
arbitrary element ml of ID'l. With 4, we associate an element
m, of ID'l which has the same position relative to ml as 41 has
to 41 under the order given by ! for these elements. Then we
associate with 4a an element ma which has the same position
relative to ml , m, as 43 has to 41 , 4:1 ; etc. It is seen immediately
from the proof of Theorem 3, that all this is possible and leads
to our objective, since the present circumstances are actually
simpler than before.
The preceding theorems imply the following:
a) The set of all terminating decimal fractions > 0 (or
terminating dyadic or n-adic fractions) has order type fl. This
follows from Theorem 3, because this set is enumerable, un­
bordered, and dense.
b) For every finite or enumerable order type " � 0, we
have fI" = fl.
For, represent " by .re(k) and fI by 9l(r). Then fI" is repre­
sented by the lexicographically ordered set of pairs of elements
(k, r). Let (k, To) be one of these pairs of clements. Since m is
unbordered, there is an TI -< To nnd an T2 >- To Then also

(k, TI) -< (k, To) -< (k, T2) ' Suppose, further, that (k, , TI) -<
(kz , T2)' Then there are two possibilities. If k, -< k2 , choose
an Ta -< T2 and form (k2 , Ta) . If, however, k, = k2 and TI -< T2 ,
then, since m is dense, there is an TI -< rJ -< r2 , and we again
form (k2 , Ta) . In each case the newly formed pair of clements
lies between the given ones. The set of pairs of clements (k, m)
is thus un bordered and dense and, finally, also enumerable,
because I fI" I = I fI / I " I 5 Q2 = Q. This set therefore has
order type '1.
This generalizes a part of the results on p. 66; in particular,
3
it follows that 'I ' '1 = 'I, and hence also '1 = '1; etc.
c) Likewise: For every finite or enumerable order type " � 0,
we have (1 + '1)" = 1 + 'l or '1 according as " has a first element
or not. Thus, in p arti cular, (1 + fI)7] = 'I, (1 + fI) 2 = 1 + fl.
7, Continuous Sets
Let there ue given an ordered set ID'l. A suuset � of IDl is
called an initial part of ID'l, if 4 E � implies that every element
74
preceding a in the set IDl also belongs to W. The initial part �
shall always be ordered in the order prescribed by IDl. A subset
.8 of IDl is called a remainder of IDl, if z E .B implies that every
element succeeding z in the set IDl also belongs to .8. The re­
mainder too shall always be ordered in the order prescribed
by IDl.
If, e. g., m(r) is the set of all rational numbers in the natural
order, then the set of rational numbers r for which r < 2,
constitutes an initial part, and so does the set of rational num­
bers for which r � 2. Likewise, a remainder of m is given by
the set of all r > 2 as well as by the set of r � 2. The empty
set is both an initial part and a remainder of every ordered set.
A decomposition of an ordered set IDl is a representation of
i t in the form IDl = W + .8, where � is an initial part and .B
is a remainder of IDl, W ·.8 = 0, and the sum is, of course, an
ordered sum.
A decomposition IDl = � + .8 of an ordered set IDl into two
nonempty sets W and .8 is called
a jump, if W has a last and .8 has a first element;
a cut, if W has a last and .8 has no first element, or
if W has no last and .8 has a first element;
a gap, if W has no last and .8 has no first element.
For example, if the real numbers in the following intervals
are in natural oraer,
(0, 1) + (2, 3) is a jump,
(0, 1) + (I , 2) is a cut,
(0, 1) + ( I , 2) is a gap.

In the domain of the naturaUy ordered rational numbers r,


with notation that is self-explanatory,
I r :S 0 I + I r > 0 I is 8. cut,
whereas
IT < y2} + IT > y2} is a gap.
75
The existence of jumps in an ordered set js synonymous with
the existence of neighboring clements. A dense set, conse­
quently, is a set without jumps; it has only cuts or gaps.

Tm:OUE!.1 I . If olle of two mutually simnllr sets has jUllll)S, or


cuts, or gaps, thCIl the same holds also for the other ut. lIenCfJ,
(fiVfm ally one of these three properties, all .�ets harrillg Ihe same
order tYI)e eilher do or do not posse8s this pr01)crly.
Proof: This follows from the f,let that, under a similarity
mapping, since it preserves the order of succession of the
clements, every decomposition goes over into a decomposition,
and a first or 11151. clement of 1\ remainder or an initial parL,
rcspcct.i\'cly, is mapped agltin into !;uch 1m clement.
From this it now follows, e. g., that the set of all points
P(z, y) in the plane, ordered in g) on p. 57, is not similar to
the naturally ordered set (0, 1). For, the first set has gaps, one
of which is obtained, e. g., by taking �( to be the set of points
P(x, y) for which x < 0, find denoting by .8 the remainder,
determined by �(, of the set. Since the set (0, 1) has no gaps,
this set is not similar to the first.
An ordered set, each of whose deeomposit.ions is 1\ cut, is
said to be continuous. In ot.her words, a set is continuous, if
it contains at least \'\\'0 clements, and has neit.her jumps nor
gaps.
Accordingly, 11 dense set is contiuuous, if it has no gnps. The
set of rational numbers is not continuous, no lllatter how i t is
ordered. For, every continuous set is dcnsc, and, by p. 72,
there arc only four types, 1/, I + 1), 1) + I , I + 1) + I , of dense
enumerable sets, each of which hns gaps, as WllS shown in an
example above. On the other hu.nd, the set of real numbers in
their nu.tural order is continuous.
The conccpts which wc hu.vc introduced here go hack es­
scntially to Dedekind. With their aid, the irmtioll:l.1 numbers
can, as \\"e shall 11011' sketch briefly, be int roduced as follows�
Suppose that up to now only the rational numbers have becn
knowil. With every mtional lHullhcr we ClIlI llssociate a dcfinite
point of a given line by means of one of" the well-known cle-
76
mentary geometrical constructions. But it is not possible,
conversely, to associate a rational number with every point of
this line. For if we construct (Fig. 6) an isosceles right triangle

U I Z

Fig. 6.

with legs of unit length, and mark off the length of the hy­
potenuse on the given line, beginning at the origin, we obtain
a point whose distance z from the origin must satisfy the
2
equation Z = 2. Such a number z, however, does not exist in
the domain of rational numbers. But if one wishes to be able
to pursue the study of analytic geometry, it is necessary not
only that a point correspond to every number, but also that
a number correspond to every point on the line. It is therefore
necessary to fill in the "gaps" in the domain of rational numbers
(jumps, of course, do not appear in the set of rational numbers).
And this takes place according to Dedekind, by introducing
the "gaps" as a new kind of number, for which the rules of
operation are then defined in a suitable manner. When this has
been done, the set of naturally ordered numbers is a continuous
set just as the set of points on a straight line. A mapping oC
the set of numbers on the set of points, preserving the order of
their respective elements, as is required in analytic geometry,
is now possible.
If Dedekind's reflections are carried over to general ordered
sets, we arrive at the following two theorems:
ThEOREM 2. Every continuous set IDl contains a 8ubset oJ type
A, i. e., a subset which is Bimilar to the set oJ real number8 in their
natural order.
ProoJ: According to p. i2, Theorem 5, IDl contains a subi>Ct
m oC type '1. The set �l hus gaps. Let

(1) m = � + .B
77
be a decomposition of m which determines a gap in m. Then
IDl contains an element m which, in ID1, succeeds all elements
of �l and, at the same time, precedes all elements of .8. For
let i be the totality of elements of IDl which belong to � or
precede an element of �, and let B be the totality of elements
of IDl which belong to .8 or succeed an element of .8. If i + B
were a decomposition of IDl, IDl would have a �p, c�ntrary to
the assumption that IDl is continuous. Hence, !l + .8 lacks at
least one element m of IDl, and, according to the construction
of i + B, i and precedes B, so that it
this element succeeds
also succeeds � and precedes .8.
Every gap (1) thus determines at least one element m of IDl.
For every gap (1), we choose one of these elements m and keep
it fixed. Denote by 91 the set m after it has been supplemented by
the addition of these gap elements m. This set, as a subset of
the ordered set IDl, is also ordered, and is similar to the set of
all rational numbers together with their gap elements in the
domain of real numbers. In other words, according to Dede­
kind's introduction of the real numbers, 91 is similar precisely
to the set of all real numbers, which proves the theorem.
Just as it was possible to describe, on p. 72, the order type
of the set of rational numbers without referring to the rational
numbers themselves, the analogue is now possible also for the
set of real numbers, as the following theorem shows:

THEOREM 3. Let IDl be an ordered set with the Jollowing prop-


ertie8:
a) IDl u unbordered;
(3) IDl u continuous;
'Y) IDl contaim an enumerable set � such that between every pair
oJ elements oJ IDl there is at least one element oj U; i. e., there is an
enumerable set !l which is dcme in IDt
Then IDl Itas order type ).; i. e., IDl is similar to tlte naturally
ordered set oJ all real numbers.
Proof: The set � is unbordered, because before and after
every element of � there are still clements of IDl, and hence,
due to "Y), also clements of �. Likewise, it follows from (3) and
78
'Y) that W is dense too. Consequently, W has order type 'I, so
that it is similar to the set of all rational numbers. Every gap
in the set �( detcnnines, according to the proof of Theorem 2,
at least one element m of IDl, and, because of 'Y) , not more than
one element of IDl. This, however, implies the assertion. For,
the set of rational numbers can be mapped on �1, whereby
with every gap in the domain of rational numbers there is
associated precisely one gap of the set �, and, consequently,
also precisely one element of IDl. We have thus obtainro a
similarity mapping of the set of all real numbers on the set ID1.
P. Franklin (Transactions of the American Mathematical
Society, vol. 27 (1925), pp. 91-100) has made some interesting
investigations in connection with our theorems of the last two
paragraphs. Suppose that in each of the intervals 0 < x < 1
and 0 < y < 1 we have an enumerahle set which is denre in
that interval. This means that ill every subinterval, however
small, of the interval (0, I), there is at least one point of the
enumerable set. It follows from Theorem 3 on p. 7 1 , that
both sets can be mapped similarly on each other. It is then
further evident from the considerations made in connection with
t.he last two theorems, that there exists a continuous monotonic
function y = J(x) which maps the interval 0 < x < 1 (' .In­
tinuously on the interval 0 < y < I , and, at the same time,
maps one of the given sets precisely on the other. Franklin,
now, has shown that this function J(x) can be chosen to be
actually an analytic function, and has obtained, besides other
interesting results, the following. In the interval (0, 1), let
there be given a strictly monotonic and continuous function
which maps the interval (0, 1) on the interval (0, 1) again.
Then this function can be approximated arbitrarily closely by
a function which is analytic in the interval and which assumes
rational values precisely for rational x's.
CHAPTER IV

Well-ordered Sets mul Their Ordinal Numbers

1. Definition of Well-ordering and of Ordinal Number


Ordered sets generally lack a property which is possessed by
every set of nu.tul'al numbers. In every set of natural numberR,
namely, there is u. smu.llcstj i. c., in every subset of the set of
nu.turu.l numbers, ordered according to increa.sing magnitude of
its elements, there is a first element. This property of the natural
numbers is used in many proofs (cf. especially §7 in this con­
nection), without, perhaps, one's being always clcarly con­
scious of it. It is therefore to be expected that the corresponding
property is also of importance for arbitrary ordered scts. Ac­
cordingly, we set down
Definition 1. An ordered set IDl is said to be well-ordered, ir�
itself, as well as every nonempty subset 01 IDl, has a first element
under the order prescribed lor its elements by �. TIle empty sct
is also regarded as well-ordered.
Examples. Every finite set is well-ordered, and so is the set
1 1, 2, 3, . . . ) . The set { . . . , 3, 2, 1 1 is not well-ordered, be­
cause it has no first element. The set of real numbers in the
interval (0, 1 ), in their natural order, is not well-ordered. The
set itself, to be sure, has a first element, but the subset (0, 1)
does not. It follows, likewise, that no set of order type 'I ,
I + 'II, 'II + 1 , or I + 'II + 1 , is well-ordered. On the other hu.nd,
e. g., the set oC positive rational numbers in the order

(1) ( 1 , 2" 3 • . • ,• 1 .3. �,


� , 'fJ JI. · · · ,· 1. , 3,
"I :t tJ . . . •, • • • )

is well-ordered. For let any subset oC this set be given. This


subset contains fractions with smallest denominator, and among
these there is one fraction with smallest numerator. This
fraction, then, is the first element of the subset.
79
80
THEOREM 1. In every well-ordered 8et, every element which i8
1/ot a last element has an immediate 8ucce880r.
Proof: Let m, be an element of the well-ordered set IDl. The
set of clements coming after m, in IDl is a subset of IDl, and
therefore, according to the definition of well-ordered set, if
it is not empty it has a first clement � . Then m, -< � , and
there is no element of IDl between these two elements, according
to the definition of � .
Thus, the well-ordered sets certainly do not belong to the
dense or the continuous sets considered in the two preceding
paragraphs.
Further, the following two facts follow immediately from the
definition:
THEOREM 2. Every subset 91 of a weU-ordered 8et IDl i8 also
well--ordered (under the order prescribed Jor it by IDl).
THEOREM 3. Every ordered 8et which i8 similar to a well-ordered
set i8 itself well-ordered.
According to Theorem 3, of the sets having any given order
type, either aU or none arc well-ordered. This fact renders
possible the following
Definition 2. By an ordinal number we mean an order type
which i8 represented by well-ordered 8eU. Small Greek letters wiU
be used to denote ordinal number8 as weU as order types.
In the case of finite sets, cardinal number, order type, and
ordinal number coincide. With infinite sets, to a cardinal num­
ber there may correspond many order types. Whether there
are always ordinal numbers among these, i. e ., whether every
set can be well-ordered, will not be decided until later (ill).
I t follows already from previous considerations of ours, that
there are ordinal numbers corresponding to enumerable sets.
For according to the examples on the preceding page, the set
f 1, 2, 3, . . . } and the set (1) are well-ordered. The first set
has order type CIJ, �d the second, by p. 66, has order type CIJ�.
Hence, these two order types are also ordinal numbers. We
shall see in the next paragraph how infinitely many ordinal
numbers can already be formed from CIJ alone.
81
2. Addition of Arbitrarily Many, and Multiplication of Two,
Ordinal Numbers
On p. 61 we dealt with ordered addition of ordered sets. If
the ordered sets appearing there are replaced by well-ordered
sets, we arrive at

THEOREM 1. Let .R'(IDlt) be a well-ordered set of mutually ex­


clusive well-ordered sets; i. e., let .re(k) be a well-ordered set, and
with every element k of this set, associate a well-ordered set IDlt ,
where all these IDl.'s arc taken to be mutuaUy exclusive. Then we
assert that the ordered sum ® = Lie. IDlI; of these sets, formed
according to Definition 3 on p. 61, is itself a well-ordered set.
Proof: Let � be an arbitrary nonempty subset of ®. We have
to show that � has a first element. To this end we consider the
elements k E � such that elements of roll; actually appear in
�. These k's constitute a subset of the well-ordered set �,
and they therefore have a first element ko Further, the ele­

ments of ID'lt. which appear in � form a subset of the well­


ordered set roll. , and they therefore also have a first element
mo •This element, however, according to the order of ® de­
termined in Definition 3 on p. 61, has the property that it
precedes all other elements in �.
No new definition is necessary for the addition of ordinal
numbers, since ordinal numbers are merely special order types,
so that the definition of sum given in ch. III, §3 holds for them
too. By specializing that definition somewhat, we obtain

THEORE!'t1 2. Let sr(Il.) be a complex of ordinal numbers arising


from a well-ordered set �(k)-briefty, a well-ord6red complex of
ordinal numbers. Then its ordered sum (f = Llee lit is also an
ordinal number.
Proof: According to Definition 4 on p. 61, (f = ,Llee rolt',
where the ID'lt'S are mutually exclusive sets which represent
the Pt's. Since the Pt'S are now ordinal numbers, the roll'S arc
well-ordered. ore is also well-ordered, so that, according to
Theorem I, the set sum is well-ordered, and hence it does indeed
represent an ordinal number.
82
What is important here is that well-ordered addition of
ordinal numbers invariably leads again to ordinal numbers. In
particUlar, the sum of two ordinal numbers is always an ordinal
number.
If the terms of the well-ordered complex of ordinal numbers
arc all the same, Theorem 2 gives rise to

THEOREM 3. The product of two, and hence also of finitely


many, ordinal numbers is again an ordinal number. In particular,
every power oj an ordinal number is an ordinal number, provided
that the exponent is a finite number.

From this it now follows that (II ' n and ",n arc ordinal num­
bers too. Further, Theorem 2 shows that every "polynomial"

whose coefficients a.. are finite numbers, is also an ordinal


number.

3. Subsets and Similarity Mappings of Well-ordered Sets


Corresponding to the theorem that every transfinite set con­
tains an enumerable subset, we have the following theorem for
well-ordered sets:

THEOREM 1. Every transfinite well-ordered set contains a subset


whose ordinal number is (II.

Proof: IDl contains a first element mo ; further, according to


p. 80, an immediate successor ml j then an m2 which imme­
diately succeeds ml j etc. The process does not terminate, be­
cause IDl is transfinite, and therefore leads to an enumerable
sequence of elements

( 1) � = { mo , m l , m2 , · · · 1
having ordinal number (II, and the elements in this sequence
appear in the same order of succession as they do in .AI.
1.'01' every transfinite well-ordered set IDl, there is u similadty
mapping of 911 on one of its subsets 9l, such that, if m and n
arc corresponding elements of 91l and 9l, and we consider t heir
83
position in IDl, m -< n infinitely often. For let us form the set
(1), as in the preceding proof. This set is an initial part of IDl.
Let the corresponding (possibly empty) remainder be 8, so
that IDl = � + .8. Now put

m = i + ,8 .

The sets �{ and � are mapped similarly on each other by asso­


ciating each m" of the first set with the m,, + 1 of the second set.
Further, we let every element of .B correspond to itself. This
gives a similarity mapping of IDl on 91 under which every
element m" of IDl corresponds to that clement of 91 which just
succeeds m" in IDt
The reverse, however, - and this is a very important result
for the theory of well-ordered sets-cannot occur. For we have

THJo.:oRJo.:M 2. Let there exist a similarity mapping between the


well-ordered set IDl and its (proper or 1,lnproper) subset 91, de­
noting corresponding elements of IDl and 91 by m and n. If we
consider the position of these elements iu the set IDl, then we have
invariably m = n or m -< n, but /lever n -< m (Zermelo) .

Proof: Assume the theorem to be false. Then the set of pairs


of elements m, n for which n -< 111 , is not empty. The n's with
n -< m constitute a subset of the well-ordered set 91, and they
therefore have a first element rio . For the element mo corre­
sponding to no we have no -< 1110 The element no , however, is

an element of IDl too, and will in this role be denoted also by


1111• To this element 111 1 there corresponds an element n l of
91. Thus, with the elements mo , 111 1 of IDl there are associated
the elements no , nl of 91. Since m l = 110 -< mo , and the simi­
larity mapping does not disturb the order of succeSSion, we
have also 111 -< lit! , i. e., 111 -< ml Hence, no is not the first

Cleml'flt of 9l to precede the corresponding elcm(mt of IDl: the


elem('nt 1/ . -< 1/" already has this property. Our assumption
that the Ilsscrtion is false has thuM led to the desired cbntra­
dictiun.
Fl'om thiM we now obtain
84
Tm:OREM 3. ElJery welt-ordered set IDl can be mapped similarly
on itself only by the identity mapping. IJ IDl is well-ordered, and
IDl � 91, then there 18 only one similarity mapping oj IDl on 91.

ProoJ: IDl is a subset of itself. Let m and m be two corre­


sponding elements under a mapping of IDl on itself. Then,
according to Theorem 2, we never have m -( m or m -( m,
so that invariably m = m. This now gives the second assertion.
For if 91 could be mapped on 1m in two ways, it would be
possible, by interposing 91, to map IDl on itself in two ways,
which is impossible.
We have already considered a particular kind of subset of
an ordered set, namely, the initial part. We shall now alter the
terminology a little. Let IDl be a well-ordered set, and let m
be one of its elements. Then the subset of elements of IDl which
precede the element m will be called the segment IDl.. deter­
mined by m. The segment belonging to the first element of
IDl shall be the empty set. The complement of IDl.. in IDl is
called the remainder determined by m, and will be denoted
by 1m - IDl.. . The segment determined by m does not contain
the element m itselfj this element belongs, rather, to the re­
mainder determined by m. From this we immediately get

THEOREM 4. IJ m i8 a 8egment oj the well-ordered 8et � , and


then IDl i8 also a segment oj �.
IDl is a segment oj the 8et 91,
THEOREM 5. Let 1m.. and 1m" be two distinct segments oj the
same wel£..ordered set 1m. Then one oj the two segments is a 8eg­
ment oj the other.

ProoJ: Let m � n, so that, say, m � n. Then, for every


p E ID1 for which p -( m, we have a Jortiori p -( n, which yields
the assertion.

THEOREM 6. Under a similarity mapping oj a wel£..ordered set


IDl on a well-ordered 8et 91, elJeTY segment oj IDl is transJormed
into a 8egment 01 m.

Prool: Let 1m .. be a segment of IDl, and let the element n of


85
m correspond to the element m of IDl. Then the clements of
IDl which precede the element m go over precisely into the
clements of 91 which precede the clement n.
From Theorem 2 we obtain further:

THEOREM 7. No well-ordered set IDl is similar to allY oj its


8egments, or to a segment of a 8ubset. Two distinct 8egments oj the
8ame well-ordered 8et are never similar to each other.
PrE!!!: If, for some segment IDl.. of the subset IDl C IDl, we
had IDl.. � IDl, the element m of the set IDl would correspond
to an element of IDl.. , i. e., to an element preceding it in IDl,
which, by Theorem 2, is impossible. If, now, IDl.. and ID1.. are
two distinct segments of the same set, then, according to
Theorem 5, one of them must be a segment of the other, and
consequently the second assertion follows from the first.

4. The Comparison of Ordinal Numbers


For order types, and consequently for ordinal numbers, it is
clear when they are equal, but the relations "less than" and
"greater than" have not yet been defined. This will now take
place for ordinal numbers (but not even now for order types).
Let p., II be the ordinal numbers of the well-ordered sets
IDl, 91. Then we say that p. < JI and, at tho same time, also
II > p., if IDl is similar to a segment of m.

On account of §3, Theorem 6, this definition is independent


of the particular representatives of the ordinal numbers I" and II.
As for rules governing inequalities, we first obtain almost
immediately the following two:
a) If I" < II and II < 11", then p. < '" (transitive law).
For let the three ordinal numbers be represented by the sets
IDl, 91, �. Then the assumptions state that IDl is similar to a
segment m. of 91, and 91 is similar to a segment �p of �. Ac­
cording to §3, Theorem 6, the similarity mapping of 91 on �.
also maps 91.. on a segment of �, 80 that IDl too is similar to a
segment of �.
(3) For every pair of ordinal numbers 1", II, at most one of the
three relations
86
}I. < 1', }I. = 1', }I. > I'
is valid.
For if w£> hud simultaneously IJ < I' and Il = 1', we should
also have Il < Il, i. e., a well-ordered set would be similar to
on(' of it.s segments, which iR impossible a.ccording to §3,
Thcomm 7. Likewise it follows that Il = I' Ilnd }I. > I' cannot
hold simultaneously either. Fino.lly, if we had Il < I' and at
the same time }I. > 1', it would follow from a) that p. < p.
again, so that this case too cannot occur.
We have now gotten just as fo.r with ordinal numbers as
with co.rdinal numbers. For both kinds of numbers, namely,
the question is still open, whether it is also true that invariably
at least one of the relations {3) holds, i. e., whether every two
numbers of the same variety arc always comparable with one
another. For ordinal numbers, this question can now be de­
cided. From this will then follow the decision for cardinal
numbers too, but, to be sure, only on the basis of the very
dcep well-ordering theorem.
Two ordinal numbers which arc represented by segments of
one and the same set are certainly comparable. For, according
to §3, Theorem 5, either both segments are the same, or one
of them is a segment of the other. The general proof is executed
with the aid of a certain normal representation of a given
ordinal number by means of ordinal numbers themselves.
Let p. > 0. Then the set .!fiji , which ,..ill also be used fre­
quently in the future, is understood to be the set of all ordinal
numbers which are less than }I.. If }I. = 0, we put .!fio = 0. I
If P. > 0, .!fiji is not empty, because then the ordinal number
° always belongs to this set. We have, for example,

SlB" = t o, 1, 2, " ' , n - l ) i .!fi .. = t o, 1, 2, . . . ) .


THEOREM 1 . The elements oj the set .!fiji are comparable with

IThe reader should observe that. neither in this definition nor in the
proofs of tbe following t.wo theorems do we make use of a comparability
theorem. It is just. this theorem, rather, which is to be proved by means of
the ensuing considerations.
87
one another. If, as is accordingly possible, we order the ordinal
numbers appearing in m.. according to increasing magnitude, the
set m3.. has ordinal number p. .
Proof: Let IDl be a well-ordered set with ordinal number p..
Let " be any element of m" , i . e., let " < p. . Then, by the
definition of Ille88 than", " is the ordinal number of a segment
of IDt All ordinal numbers appearing in m .. are thus ordinal
numbers of segments of the same set, and hence, by what was
said on the preceding page, these ordinal numbers are all com­
pamble with one Mother. It is clear that the ordinal number
of an arbitrary segment of IDl is also contained in the set m,. .
Now let " be an clement of m,. , and let IDl.. be the segment
of IDl which belongs to this ordinal number. We associate the
ordinal number " with the clement m. By what has preceded ,
this defines a mapping of the set m.. on the set rot But this
mapping is also a similarity mapping. For if ", < "2 , then
IDl ... is a segment of IDl ... , so that m, -< m2 • Consequently
m.. � IDl, and this completes the proof of the theorem.
With this theorem we have obtained a certain normal repre­
sentation of ordinal numbers. In all operations, now, where a
well-ordered set may be replaced by MY set similar to it, we
can always replace the given set by the set m,. , where p. is
the ordinal number of the given set. This is often done in
proofs. We first prove in this manner the comparability theorem
for ordinal numbers:

THEOREM 2. A ny two ordinal numbers p., v satisfy at least one,


and consequently, according to P) , invariably precisely one, of tile
three relations
p. < v, p. = v, p. > v;
i. e., two ordinal numbers are alway8 comparable with one another.
Proof: We make use of the sets m.. and m • . Let :1) be their
intersection, i. c., the set of ordinal numbers which are both
<p. and < v. :1), as a subset of the well-ordered set m. , is well­
ordered, and therefore has an ordinal number 6. We shall now
show that 6 is comparable with p., and that 6 S p.. In demon-
88
strating this, we may n.ssumc that 8 � "c. Then � is a proper
subset of SID" , and
m.. = � + OID" - I) .
Here I) it! Ull initial part of SID" . For let a E ID and {j E
(5lB1l ID). Then a and {j, as clements of ml.. , are comparable.
-

They are distinct, however, so that either a < P or {j < a.


If we had {j < a, then, since a < "c, 1', we should also have
P < "c, 1', so that p too would belong to �, which is not con­
sistent with the definition of {j. Consequently we have always
a < p, i. e., � is an initial part of 5&.. . The subset of those

elements of 5lB.. which do not belong to � has a first element


�, and � is the segment of SID.. determined b! �. Sin�e this
segment, by Theorem 1 , has ordinal number {j, 8 = {j, and
hence 8 < "c. This is valid under the assumption that 8 � "c.
Hence, invariably 8 5 "c, and analogously we obtain 8 5 1'.
Now we cannot have 8 < "c, 8 < " simultaneously, because this
would imply that 8 E ID, so that } would have to belong to
the segment of SID,. determined by {j = 8, which is impossible.
Therefore only the following combinations are possible:
8 = "c, 8 = 1', so that Il = II;

8 = "c, 8 < 1', so that Il < 1';


8 < Il, 8 = 1', so that Il > 1';

i . e., the case that Il and " are not comparable does not occur.
From this follows immediately the comparability of cardinal
numbers of well-ordered sets:
THEOREM 3. Let m and n be two cardinal numbers which can
be represented by weU-ordered sets. Then precisely one of the three
relations
m < n, m = n, m > n

holds.
Proof: On account of (j) on p. 19, we have only to show that
at least one of these relations holds. Let m, n be represented by
89
the well-ordered sets IDl, 91 having ordinal numbers � , v. If
� = v, then a fortiori m = n. If, however, � < v, say, then IDl

is similar to a segment of 91, i. e., IDl is equivalent to a subset


of m. Hence, by the Corollary on p. 25, m � n.
According to this, the comparability of two arbitrary cardinal
numbers is ensured, if we can show that every set can be well­
ordered. The general comparability will, as a matter of fact,
be proved in this way, and this is also the only method up to
now by which comparability in general has been demonstrated.
For the time being, however, we shall deal further with ordinal
numbers, and we first derive two simple facts concerning in­
equalities.
THEOREM 4. Let tile set IDl be a subset of the well-ordered set
m. Then their ordinal numbers satisfy the relation � � v.

Proof: If the assertion were false, the comparability theorem


would yield � > V, so that 91 would be similar to a subset,
which is impossible by §3, Theorem 7.
THEOREM 5. Two ordinal numbers satisfy the relation � < v if,
and only if, there i8 an ordinal number r > 0 such that � + r = II.

Proof: Let � and II be represented by the sets IDl and m. If


� < II, then IDl is similar to a segment � of 91. For the remainder
.8 of 91 we have � + .8 = 91. Hence, if .8 has ordinal number
r, � + r = II. Here r > 0 because .8 is not empty. Conversely,
suppose that � + r = II and that r is represented by the set
.8 which has no elements in common with IDl. Then IDl + .8
has ordinal number v = � + r, and � < v because IDl is a
segment of IDl + .8.

5. Sequences of Ordinal Numbers


Let there be given a well-ordered complex of ordinal num­
bers; i. e., to every element k of a well-ordered set �(k), let
there correspond an ordinal number �. Such a complex of
.

ordinal numbers will also be called a sequence of ordinal num­


bers. In particular, a sequence of ordinal numbers will be
called an increasing sequence, if kl -< k2 always implies
90
Pi, < p" j a decreasing sequence, if .k. � � invariably implies
Pi, > Pi, • Concerning the fonner we have

THEOREM 1 . Every set 01 ordinal numbers, il ordered according


10 illCl'casil/g magnitude, is well-ordered, al/d can therefore be
wrillctt (IS a" illcrca8il/g sequcl/cc.

Proo/: We have merely to show that every nonempty subset


m of the given set of ordinal numbers has a smallest element.
I A!t P be no arbitrary ordinal number in m. If p. is not already
the smallest ordinal number in 5ID, form the well-ordered set
ml• . The intersection m · m,. can then be well-ordered, and is
not empty, so that it therefore contains a smallest ordinal
number, which is obviously also the smallest ordinal number
in 5ID.
Concerning decreasing sequences of ordinal numbers, we have

THEOREM 2. Every decreasing sequence 01 ordinal numbers con­


tains only a finite number 01 ordinal numbers.

Proo/: Suppose that the sequenee were not finite. Then, by


Theorem 1 on p. 82, the set �(k) defining the sequence would
contain a subset ( k. , k2 , ks , · · · 1 of ordinal number til. The
ordinal numbers p. . , P2 , Pa , • • • associated with these elements
would have no smallest, because they d ecrease, contradicting
Theorem 1 .

THEOREM 3 . To every set 5ID oj ordinal numbers p., there is an


ordinal number which is greater than every ordinal numbu p. in
mj and, in Jact, there is a definitc next larger ordinal number P.

Proal: We may assume that mJe) is an increasing sequence.


From 5ID(p.) we form a new set 5ID of ordinal numbers by re­
placing every element P of 5ID by P + 1 . Let (I be the sum of
the elements of the well-ordered set m = m(p + 1). Then

every Po + 1 � (I,

as is seen from §4, Theorem 4 by passing from the ordinal


numbers to their representatives. Hence, by §4, Theorem 5,
91
every IJ < tT,

which provcs the first part of the assertion.


If, now, tT is not yet the next larger ordinal number relative
to the set 5ID, i. e., if there are still ordinal numbers which arc
larger than every IJ and smaller than tT, then all such ordinal
numbers are contained in the set aB.. . Since this set is well­
ordered, the subset of m.. consisting of those numbers which
are greater than every IJ, has a smallest element ", Q.E.D.
This theorem now yields a result which is similar to the one
established on p. 36 for cardinal numbers. To every set of
ordinal numbers, namely, there is, according to our theorem, a
still greater ordinal number. Accordingly, the "set of all ordinal
numbers" is a meaningless conccpt. We shall return to this so­
called Burali-Forti paradox too in our concluding remarks. In
the meantime, we must note that it is not permissible to
operate with the set of all ordinal numbers. And the proofs
given thus far have been carried out with this in mind. Indeed,
it is for this reason that we were always careful to operate
with the set $,. , i. e., with the set of all ordinal numbers less
than a given ordinal number, instead of with the set of all
ordinal numbers, which might at first, perhaps, scem more
natural.
We shall also derive the following theorem at thiH timo,
although it will not be used until later:
THEOREM 4 . Let m be an enumerable set of ordinal numbers 0/
enumerable sets. Then the ordinal number which immediately
succeed8 m is also an ordinal number of an enumerable sct.
Proof: Let the elements of m again be IJ. Then

I tT I = E I IJ I � Q ' Q = Q .
OlE a

Consequently, for the ordinal numbC!f which immediately suc­


ceeds � we also. huve I " I � a.
THEOREM 5. Let m bc a set of ordinal numbers IJ, and let !ill
be a set 0/ ordinal numbers � 8UCI� that to every IJ there is a � � IJ.
92
Let 1', ; be 1M ordinal numbers immediately succeied ng /J, ji, re­
spectively. Thm p � 1'.
Prool: v is greater than every ii, and is therefore also greater
than every p.. 1', however, is the smallest number which is
greater than every p.. Hence, ii � 1'.

If I' i s an arbitrary ordinal number, the following two cases


are conceivable:
a) To I' there corresponds a next smaller ordinal number, i. e.,
a /J < I' such that no ordinal number lies between /J and 1'.

By §4, Theorem 5, there exists a number r such that /J + r = 1',

and r then must necessarily be the smallest positive ordinal


number, which is 1; i. e., /J + 1 = 1'. In this case we write also
/J = I' - I, and I' is called an isolated number or an ordinal
number of the first kind.
b) To I' there corresponds no next smaller ordinal number.
In this case we write also I' = lim,.<. /J, and I' is called a limit
number or an ordinal number of the second kind. The limit
numbers represent precisely those well-ordered f1ets without a
last element, as consideration of the set sm, shows.
By a fundamental sequence we mean an increasing sequence
of ordinal numbers containing no greatest. Theorem 3 can be
applied to every fundamental sequence. This means that to
the numbers of a fundamental sequence there corresponds a
next larger ordinal number, and this number, according to the
definition of fundamental sequence, is certainly not equal to
any of the numbers of the sequence increased by 1 . But it is
still conceivable at first, that the concept "next larger number
after the numbers of a fundamental sequence" does not com­
pletely coincide with the concept of limit number. That they
actually do coincide, however, is shown by

THEOREM 6. An increasing 8equence 01 ordinal numbers de­


termines a limit number as the next larger number in the sen8e
01 Theorem 3, iI, and only if, it is a fundamental 8equence. This
nezt larger number 01 the lundamental sequence sm i8 then called
the limit of the fundamental sequence, and is denoted by " =
lim"ellll /J.
93
Proof: If the number II which, by virtue of Theorem 3, im­
mediately succeeds the increasing sequence 5ID{I') , is a limit
number, then 5ID is a fundamental sequence. For if 5ID contained
a greatest number ii, there would be additional numbers be­
tween the limi t number II and ii, contrary to the definition of
II. On the other hand, let II be determined by a fundamental

sequence 5ID (I') . We are to show that, for every number p < '"
there are additional ordinal numbers between p and II. Now
there is actually always a I' in 5ID such that p. > p; for otherwise
we should have invariably I' :::;; p, and hence, since there is
no greatest number among the p's, invariably I' < p, and
consequently " :::;; p.

6. Operating With Ordinal Numbers


Since we have a comparability theorem for ordinal numbers,
we can set up much more far-reaching rules of operation for
them than for order types. We shall now derive these rules.
a) If I' < II, and p is an arbitrary ordinal number, then
a) p + I' < p + II;
fJ) I' + p :::;; II + p;
'Y) P ' I' < P ' II for p > 0;
8) p op :::;; " ' p.
Proof: From the assumption follows the existence of a r such
that I' + r = v. Consequently,
a) P + v = P + (I' + r) = (p + 1') + r > p + 1' ;
'Y) P ' v = p(1' + r> = PI' + pr > PI'·
To prove the two remaining rules, represent the ordinal num­
bers 1', v, P by the sets IDl, 9l, 9t Since I' < v, we can assume
that IDl C 9l. Further, let m · m = O. Then, for the ordered
sums and products,
IDl + m C m + m, IDl X m C m X m.
These relations, however, together with §4, Theorem 4, yield
assertions fJ) and a).
Rules (3) and a) cannot be sharpened, becauSb, e.g.,
1 + Col = 2+ Col and 1 · Col = 2 · Col.
94
b) If p < JI, p < u, then p + p < JI + ui and if 0 < p � JI,
o < p < (I, then pp < JlU.
Proof; We have
JI + U � p + U > P + P
and
JIt1 � P(l > pp.

The converses of rules a) obviously read as follows:


c)
a)
{orif pp + P << JIP + JI}, then P < Jli
+ P + P
(3) P + p = p + JI implies p = Vi

'Y)
{if pp << PJI}, then P < Jl i
or PP JlP
6) if PP = PJl, P > 0, then P = JI.
For ordinal numbers it is also possible to define a subtraction
and division, where, to be sure, essentially one-sided subtrac­
tion and division come into question, due to the absence of the
commutative law for addition and multiplication.
Let P < JI. Then, according to §4, Theol'em 5, there is an
ordinal number r such that p + r = JI, and, by c) /3), there is
only one such number r. This number can be designated as
the difference r = -p + JI of the two numbers JI and p, where
then, e. g., - 1 + JI is to be distinguished from JI 1. -

O n the other hand, the equation E + p = JI , with p < JI, is


not always solvable. For example, the equation E + 1 = CAl
cannot be solved, because the left-hand side represents an
order type with a last element, whereas the right-hand side
represents an order type without a last element.
The following result leads to division:
d) For any two ordinal numbers a > 0 and p, there is pre­
cisely one pair of ordinal numbers E and " such that
p = a'1 + E, E < a.
Proof: Put (3 = p + 1 . Then
a{3 = a(p + 1 ) � p + 1 > p.
95
Represent the ordinal numbers a, {3 by the sets �, .sa with the
clements a, b. Then a{3 is the order type of the ordered product
of the second kind � X 5.8, i. e., of the set of pairs of clements
(b, a) ordered according to first differences. This is a well­
ordered set, because it represents an ordinal number. Since
p. < a{3, � X 58 contains a segment of order type p.. Let this
segment be determined by the element (bo , ao) . Under the
order determined for � X sa, this segment consists of precisely
those elements (y, x) for which y -< bo in sa or y = bo , x -< all
in �. The element bo determines a segment in sa, and hence, an
ordinal number '1i and ao determines a segment in �, and conse­
quently, an ordinal number �. The set � X 58 can now be
represented as an ordered sum of two terms, as follows: First
come all elements (y, x) with y -< bo and arbitrary x, then all
elements of the form (bo , x) wit-h x -< ao The first part is a

set having ordinal number a'1, the second has ordinal number
�. Thus p. = a'1 + �, with � < a.
N ow the numbers �, '1 are also uniquely determined. For let
a'1l + �1 = a 'l2 + �2 • .

If = '12 , then �I = �2 follows from c) (3). But if '1 1 � '12 ,


'11
say '11 < '12 , then '11 + 1 =:; '12 , and the assumed equation
therefore yields 01'11 + �I � a'1l + a + �2 Hence, by c), •

�I � a + �2 , which contradicts �I < a. This second case,


consequently, does not occur, and this establishes the unique­
ness.
In the result just obtained, '1 plays the role of the quotient,
and �, that of the remainder, in division. We can, of course,
repeat the process, gaining thereby a Euclidean algorithm in
the domain of ordinal numbers:
96
The process terminates after a finite number of steps. This is
a consequence of Theorem 2 on p. 90, since the a,,'s form a
decreasing sequence. Hence, there is a natural number n such
that a" = O.
If, in d), we choose, in particular, " = w, then for every
ordinal number we get a representation
p = W'1 + �, � < w,

i. e., one with a finite number as remainder. If p is a limit


number, then p = W'1.
For " = 2 it follows that every ordinal number is either of
the form 2'1 or 2'1 + 1, i. e., either "even" or "odd". For in­
stance, w is an even number, because (cf. p. 63) w = 2w.
Every limit number v is divisible by each of the numbers I,
2, , W; i. e., for every 0 < " � w, v can be represented in
. . •

the form v = "'1. For according to d), there is a representation


JI = "'1 + �, � < ".

If E here were > 0, v would be no limit number.


We take this opportunity to point out that the "set of all
even ordinal numbers" is also a meaningless concept. For if
this set had a meaning, there would exist an ordinal number v
greater than every even ordinal number. But then v or JI + 1
would be even and at the same time greater than every even
ordinal number.
Concerning limit numbers in particular, there is the following
rule:
e) Let v be a limit number, and let " be any kind of ordinal
number. Then " + v is also a limit number, and, if " ¢ 0,
"V is a limit number too. In fact,

(1) " + lim p = lim (" + p), " . lim p = lim ("p) .
,,<. .<. ,.<. JI<.

Proof: If the ordinal numbers " + p and "P are ordered ac­
cording to increasing p's, we obtain increasing sequences with­
out a last element, and these sequences therefore define limit
numbers.
97
Now let
13 = lim (a + II) '
According t o the definition of v, certainly a + v > a + every
II. Hence, also a + v > every (a + II), so that
(2) a + v � p.
Since, on the other hand, 13 > a, there is a 'Y > 0 such that
a + 'Y = p. Then, by the definition of 13,
a + 'Y = 13 > every (a + II),
so that
a+ 'Y > a + every II,
and hence
'Y > every II, and consequently 'Y � Vj therefore a + v � a +
'Y = p. This, in connection with (2), yields the first of the
relations (1) .
To prove the second of these relations, put
It = lim (all).

By the definition of v, certainly av > a ' (every II) . Hence, also


av > every (all), so that av � p.
On the other hand, 13 can be represented in the fonn
13 = a-y + a, a < a.
Since It > every all, it follows that
every all < a-y + a < a('Y + I},
and therefore
every II < 'Y + 1 .
Consequently, II S 'Y + I, and hence, since II i s a limit number,
II < 'Y + I, so that II S 'Yj
the�-efore all S a-y � p. This, in connection with all � 13, proves
the rest of the assertion.
98

7. The Sequence of Ordinal Numbers, and Transfinite


Induction
We return once more to the considerations on p. 80; i n
part.icular, t o Theorem 1 und the set mll' consisting of the
ordinal numbers which are less than 1'. According to Theorem
] , every well-ordered set 9)l is similar to a set mll' . This means
that thE! clements m of ihe set 9)2 cun be associated with tl\l�
ord inul !lumbers P < I' in slIch u. one-to-one m lUlIW I', t.hat, if
1111 , till me any two plement.s, 111 1 -< 1n2 i f, and only if, [UI' t.he
cOI"l'esponding ordinal numbers PI , P2 , We have also PI < P2 '
In other words-this will make the fact seem more familiar­
we can count off the elements of every well-ordered set of
ordinal number IJ with the aid of the ordinal numbers < /J, a
fact which of course is trivial for sets with a finite number of
clements. For example, the set l a, b, c, d} having ordinal
number 4 can he counted with the ordinal numhers 0, 1, 2, 3.
If we keep this counting process for finite sets in mind, it seems
hopeless at fi rst to obtain an analogous counting process for
infinite sets. All the more highly must we value the accom­
plishment of Cantor's, who, by the i ntroduction of \\"ell­
ordered sets and ordinal numbers, overcame these difficulties.
It will be well to note the first numbers of a set ml. for
sufficiently l arge IJ.
First come the finite numbers, ordered according to in­
creasing magnitude, After the set of all finite ordinal numbers
we find the first transfinite ordinal number. It is, according to
Theorem 1 on p . 86, precisely the order type of the set of
all ordinal numbers preceding it; in this ease, therefore, the
order type of the set 1 0, 1 , 2, . . . / ; i. e., w. This ordinal number
is essentially different from those preceding it. For, w has no
immediate predecessor; it is a limit number.
To w, as to every ordinal Ilumber, there corresponds an im­
mf>dil1te successor w + 1 ; then comes w + 2; etc. We arc thus
led to the sequence of ordinal numbers

0, 1 , 2, . . . , w. w + 1 , w + 2,
99
Since this sequence has order type ", + '" = ", · 2, the number
coming next after this sequence is ", · 2. This ordinal number is
followed by ", · 2 + 1 , "" 2 + 2, . . . , and after all these comes
", · 3j etc. Thus the beginning of the sequence of ordinal num­
bers is
0, 1 , 2, " ' , "" '" + 1 , '" + 2, . . . , ", · 2, ", · 2 + 1, . . . ,

"" n, "" n + 1 , . . .
This set has order type ",2 . Therefore the next ordinal number
is ",2 . This is succeeded by ",2 + 1 , ",2 + 2, . . . , ",2 + "', . . . ,
more generally, all ordinal numbers of the form ",2 + ", . n, + no ,
where no and n, are finite ordinal numbers. Now the set of all
the ordinal numbers considered thus far consists of two ge.
quences of order type ",2 placed one after the other, and there­
fore as a whole it represents the ordinal numher ",2 . 2, so that
t.his is tho next ordinal number. If we cont.inue in this manner,
we obtain all ordinal numbers which ean he writ.t.en in the
form of Ilpolynomials"
(1) ",. · n. + ",. -' ·n._, + . . . + "" n, + no
with finite numbers k and n, . The ordinal number which ap­
pears after all these numbers (1) can no longer be expre�d in
terms of "', at least not until products with infinitely many
factors or arbitrary powers of ordinal numbers have been
introduced, which will take place in the next paragraph. This
will require transfinite induction, which we shall discuss now.
Suppose that we wish to prove let U� saY' the binomial
formula
"

(a + b)" = :E (:) a"-·b·


.-0

for natural numbers n, or the general validity of an assertion


A (n) set up for every natural number n, with the aid of ucom-
plete (or mathematical) induction" or Hinference from n to
n + 1" . Then, u.s if! well-known, we proceed as follows:
Pirst, we prove the validity of the proposition A (u) for the
100
smallest number for which the truth of A (n) is affirmed, say
for n = O.
Second, we prove that if the proposition A (n) holds for all
numbers n < no , it is also true for no .
Third, from this now follows the universal validity of the
proposition A (n). For if A (n) were false for some n, there
would be a smallest such number, call it no . On account of
the first step, no > O. A (n), then, is true for the n which pre­
cedes no , and consequently, by the second step, also for no
itself. The assumption "A (n) is false" has thus proved to be
untenable.
Corresponding to this method of proof we can now also pro­
ceed with a&�rtions A (I') which refer to an arbitrary ordinal
number 1'.
First, we prove the validity of A (I') for the smallest ordinal
number I' that comes into question, say I' = O.
Second, we prove that if the proposition is true for aU ordinal
numbers I' < 1'0 , then it is also true for 1'0 •

Third, from this follows the universal validity of the propo­


sition A (1') . For suppose that A (I') were false for some p.
Consider the well-ordered set flB;;+ , , and in this take the subset
of numbers I' for which A (p) is false. This set, as a subset of
a well-ordered set, has a first element 1'0 which, on account of
the first step, is > 0. The set of numbers I' < 1'0 is therefore
not empty. According to the definition of 1'0 , A (I') is true for
all numbers I' < 1'0 , and hence, by the second step, also for
1'0 itself. The assumption "A (1'0) is false" has thus proved to
be untenable. Consequently, by the logical lsw of the excluded
middle, the assertion A (p) is universally valid.
This procedure is frequently called transfinite induction, be­
cause it is now no longer restricted to finite sets. It can be
used for definitions as well M for proofs. Examples of this are
contained in the next paragraph.

8. The Product of Arbitrarily Many Ordinal Numbers


Up to now we havc defined multiplication of only a finite
number of ordinal numbers, in particular, of two ordinal
101
numbers. The definition of a product of arbitrarily many
factors will take place in the same way that We proceeded from
a product of two factors to one of finitely many factors, viz.,
by the successive performance of multiplications of two factors,
where now, of course, there will be an additional limit process.
First, however, the analogue will take place for the sum, not­
withstanding the fact that the sum of arbitrarily many terms
has already been defined. On the basis of the earlier definition
of sum, we shall derive two properties of sums; and then we
shall show that, conversely, the sum can already be defined by
means of these two properties.
Let a well-ordered complex �(p) of ordinal numbers be
given, where � has ordinal number Ie. By Theorem 1 on p. 86,
the set � is similar to the set SID. , if the elements p < Ie of the
latter set are ordered according to increasing magnitude. The
given complex of ordinal numbers can therefore also be de­
fined so that to every ordinal number p < Ie there corresponds
an ordinal number p, •

We now keep fixed the ordinal number Ie as well as the


p, > 0 associated with each p < Ie. We then consider all sums
L,<� p, , where the terms are ordered according to increasing
magnitude of the p's, and we prove:
(I) If 0 < X =:;; Ie, and if X is not a limit number, then
L p, + PH = L p, •

,<A-l p<A
(II) If 0 < X =:;; Ie, and if X is a limit number, then Lp< ). p, is
a limit number too, and
lim L p, = L p, •
or<1- ,<or ,<A
Relation (I) follows immediately (rom Definition 4 on p. 61,
i( the I','S appearing in (I) are represented by mutually exclu­
sive sets.
To prove (II), represent the I','S likewise by well-ordered
disjunct sets IDl, . It is clear that the sums Lp < or 1', , ordered
according to increasing (T (or all (T < X, form an increasing
102
sequence, and, indeed, one without a last element, because �
is a limit number. Hence, in any case,
lim L 1', S L 1',. •
,<1 ,<, ,<1
It now remains to be sho� that there is no ordinal number
between the left- and right-hand members of this inequality;
in other words, that, for every a < Lp< l 1',. , also a < lim.d
Lp<. 1', Let, then, a < L,<l I'p Then a is the ordinal number
• •

of a segment of L,.< l IDlp This segment is determined by nn


clement which belongs to the set IDl, , say. Since X is a limit


number, we have also T + 1 < X and a is the ordinal number
of a segment of E,< ,+I finp Consequently,

a < L
p <r+1
1', < lim L 1', ,
.<1 ,<.
which proves the assertion.
With the help of properties (I) and (II), we can now also
define the sum of arbitrarily many ordinal numbers, by means
of transfinite induction, as follows, if the sum of two ordinal
numbers has been defined. We put

(a) 1(0) == OJ

further, if 1(P) has already been defined for all p < X:


(P) I(X) == I(X - 1) + 1'1-1
if X is not a limit number, and
I(X) = lim I(p)
,<1
if >. is a limit number.
Herewith 1(>') is defined for every ordinal number >. S Ie,

and, by what has preceded, we have precisely

I(X) =
p <l
L 1',. •

This definition is obviously nothing but, c. g., the HUm


definition
103
a + b +c+d = \ ((a + b) + c) + d l
carried over to arbitrary well-ordered complexes of terms.
According to the pattern just carried out, a product of an
arbitrary well-ordered complex of ordinal numbers is now de­
fined as follows: Let the complex be defined again so that to
every A < IC there corresponds an ordinal number II). and the
Il).'s are ordered according to increasing magnitude of the A's.
The definition of a product of two factors is assumed to be
known. Keeping fixed the numhers introduced so far, the
product TI p <). lip is defined for all A ::;; /C as follows: Put
(a) 1(0) = 1 ;
further, if I(p) is known already for all p < A,
(b) I(A) = I(A - 1 ) ' 11).-1 if " is not a limit number, and
(c) I(A) = lim p <). I(p) if A is a limit number.
Then, for all " ::; /C, /(A) is defined inductively, and shall be
the product
TI lip = I(A) .
p <).
It is obvious that

1( 1 ) = Ito , 1(2) = /J9 ' 11 1 , . . . , /(n) = /J9 ' Il , ' " 11,,- 1 ,

so that this definition coincides with the one given earlier for
a finite product.
We shall not derive rules of operation for the general product
here, because we shall not need them. We must, however,
point out an important difference between the two kinds of
product. If we form, e. g. , the infinite product 2 · 2 · 2 . . . with
enumerably many factors, this product, by our definition, is
equal to
l im 2" = the next number after 1 2°, 2', 22, . . . I ,
.. < ..

i. c., it equals w. In this connection, the 2's here arc to he in­


t.erpreted as ordinal numbers. But if we interpret the product
as a product of cardinal numbel'8, we obtain 2· = c . Thus,
whereas for finite products of ordinal numbers we had invariably
\ "' . p I = I fA 1 ' \ II \, this is by no means the ca.<;c here, since
1 04
I ", 1 = Q < c. The same is true for the product 2 · 3 · 4 • • • • As
a product of ordinal numbers, it is
lim nl = ",;
-<-

as a product of cardinal numbers,


2 · 3 ·4 . . · � 2" = c.

We must therefore note that for infinite products of ordinal


numbers it is possible to have
1 IT 1', 1 � IT 1 p, I ·
,<. p< .
This is due to the fact that products of infinitely many ordinal
numbers are, to be sure, again ordinal numbers, on the basis
of the definition of such products, but they are no longer
defined as order types of products of sets.
9. Powers of Ordinal Numbers
The power I'a is now once more defined as a product of equal
factors. By specializing the general definition of the product,
we obtain the following definition of the power: Put
(a) 1(0) = 1;
further, if I(p) is already defined for all p < A,
(b) I(A) = I(A - 1) · 1' if A is not a limit number, and
(c) IfA) = lim,< l /(p) if A is a limit number.
Then we stipulate that l = I(A).
For example, according to this we have
2- = 2 · 2·2 . . . = "' ; "," = "' .'" . . . = lim ",- .
Moreover,
1 + '" = "',

'
1 + '" + ",
'
= '" + ",' = ",( 1 + "') = ", ,

. . . . . , . . . , .
105
Consequently, by p. 101 (II), also

1 + (&) + (&)' + (&)s + . . . = (&)'" •

We can now continue the investigation on p. 99. There we


left open the question of how to denote the ordinal number
which comes after all polynomials

(1 ) r. = (&)·n. + (&).- In'_l + . . . + (&)nl + no •

For every r. , certainly (&)' .:S; r. < (&).+1 . Hence, by p. 91,


Theorem 5, the fundamental sequence of the r.'s has the same
limit as the fundamental sequence 1, (&), (&)2, (&)S, , viz., the
• • •

limit (&)" . Thus, after the polynomial (1) come the ordinal
numbers (&)", (&)" + 1, ., (&)" . (&)",
. • • • • •

Since the commutative law for the product does not hold,
we lose one of the usual laws of operation for powers. For it
can happen that (p"Y F pP,,', as is shown, e. g., by II = P = 2,
I' = (&). We do have, however,
a) p ap� = p "' +iJ,
b) (pa ) � = 1'''''.
Prool: For fixed I' and a, the rules are proved by induction
for every fJ. For fJ = 0 they are trivially correct. Let us assume,
then, that they have already been verified for all 'Y < fJ. If
fJ is not a limit number, then, by the definition of the power,
and the rule which we have assumed is true for fJ 1, -

p a .p� = p0(p'-I . p) = (p"'p'-l) .p

If fJ is a limit number, fJ = lim 'Y , then, bearing in mind p. 96,


e),

ISincc a) is used to prove b), we must first carry out the proof of a) and
then that of b). In the text, however, both proors are given together in
order to save space.
106
� .. ' �/J = � .. , I lID ' (�..�'!') = I 1m
' �'!' = I lID ' � ,,+ '!'

_ " ' l Im'!'


- �
Thus, e.g"
I " " .. .. '
\w ) = W ', (w" ') " = III •

In the fundamental sequence


" Qt ' w-
W , w , w ,

the exponents form a fundamental sequence having the limit


w", Therefore the above fundamental sequence has the limit
Cl if ( ", til )
W = w ,

We can continue in this manner, obtaining finally the funda­


mental sequence
.. ..
1 , w , w " , w .. .. , w " ,

whose limit will be denoted by Eo ' According to the definition


of the power,
w• • = l'1m I w, III .. , w.. .. " • . } = Eo •

Cantor called every solution of the equation w· = E an E­


number.
As to inequalities for powers, we mention the following :
c) If � :s; " and a � 0, then �" :s; ,,".
Proof: The asse rtion is true for a = O. Suppose that it is
true already for all "I < a . Then, if a is not a limit number, i t
follows from p . 105, a) that

and if a is a limit number, a = lim "I, it follows from p 91 ,


Theorem 5 that
1 07
d) If a < fj and 110 > 1, then 110" < 11.'.
Proof: The assumption implies the existence of a number
'Y > 0 such that a + 'Y = fj. But then

e) For 110 > 1 and a � 1, 110" � a.


Proof: The assertion is true for a = 1. Let it be true already
for all 1 :S fj < a. Then it is also true for a. For flUpp0!o!e that.
a iR not It limit numhOl" Then, on aceollnt of a � 2,

= (a - 1) + (a - 1) � (a - 1) + 1 = a.

If, however, a is a limit number, a = lim fj, then

110 a = lim II.' � lim fj = a.

10. Polynomials in Ordinal Numbers


Ordinal numbers exhibit a behavior similar to that of the
natural numbers in several other respects. Just as every natural
number can be represented as a decimal, i. e., in tenDS of
powers of 10, more generally, in tenDS of powers of an arbitrary
natural number b > 1, for ordinal numbers we have

THEOREM 1 . Let fj > 1 be an arbitrary ordinal number. Then


every ordinal number f > 0 can be represented in one, and only
one, way in the form
( 1) f = fj "'Y + fj " ''Y1 + . . . + fj " .'Y. ,

with a > al > . . . > a. � 0 and 0 < 'Y, 'YI , • • • I 'Y.. < fj j i . e .,
a8 a finite BUm of power8 with coejJicienU which are smaller than
the base.
Proof: The existence of the representation is established as
follows: According to §9, e), fjr+ l > f. Hence, there is a smallest
number 0 such that fj' > f. Here 0 is not a limit number, be­
cause otherwise, for every " < 0, we should have successively
108
11 + 1 < 8, {J,+1 S f, {J' < f, and finally {J' = lim {J' S t,
which contradicts {J' > f.
Bince 8 is not a limit number, it has an immediate predecessor
a = 8 - 1 , and by the definition of 8,

(2) {J" S r < (J.".

According to p. 94, d), there exist numbers "I, fl such that


f = {Ja'Y + f. , f. < {J a .

Because of (2), it follows that 0 < {Ja'Y < {J ,,+l = {Ja{J, and
hence 0 < "I < {J. The process can be repeated with tl , pro­
vided that f. > O. We thus obtain a chain of equations

I
t = {Ja'Y + tl ,

(3) �. � �:''Y' + t2 �
. .

with the inequalities


t � {Ja > tl � {Ja. > t2 � • • • ,

from which follows a > al > ai > . . . as well as 0 < "I,


"I. , < {J. Bince a decreasing sequence contains only a
• . •

finite number of terms, the process tenninates after a finite


number of steps. This means that there exists a f.+l = O.
The representation then follows from equations (3).
The uniqueness of the representation is proved as follows:
For every number f of the form (1), it is easy to see that
f < fJ"". If we had
;
(4) t = {J"'Y + {J"''Y. + . . . = {J � + {J"'�l + . . .

and a < cr, then a + 1 � cr, and (4) would imply

{J.+ ' > r � {J- � {J-",

which is impossible. Therefore a = a. If we now had :Y > 1,


109
say 'Y = 'Y + �, then a leading term would drop out in (4),
giving
{3 a ''Y I + ' . ' = {3 a=
'Y + , . , ,
which is impossible again, because 0:1 < 0:.
The theorem says, e. g., for {3 = 2, that every ordinal number
r can be represented in the form
a
r = 2 + 2 a , + . . . + 2 a -,
Moreover (p. 1 04), w = 2". In such a representation, therefore,
the highest exponent that occurs has by no means to be smaller
than the number represented.
We obtain an especially important example for {3 = w . It
follows then that every ordinal number e can be represented
in the form
a a
e = w c + W 'CI + . , ' + W 'hc" ,
where the c..'s are less than w, so that they are finite numbers.
If we add terms with zero coefficients, we obtain for every
ordinal number e a representation
e = . . . + w"x .. + . . . + w'x, + WXI + X/I
� W aX ,
= * £,.., a
a

where the star before the E indicates that this sum is ordered
according to decreasing powers of the base w, and where the
xeo's are integers � O. This representation only apparently
contains infinitely many terms. Actually only a finite number
of nonzero terms occur. If we agree that the representation
shall begin with a nonzero term, then the representation is
uniquely determined by e. Now let e, '1 be two arbitrary ordinal
numbers, having the representations
e = * E w ax" , "
'1 = *E W Ya .
a a
Then we can form their so-called "natural sum" (G . Hessen­
berg)
O' (�, '1) = E w " (x" + Ya)
a
110
which also contains only a finite number of nonzero coefficients.
Since the natural sum is a sum of ordinal numbers, it cer­
tainly represents an ordinal number; and obviously o'(�, ,,) =
0'(", �) because the commutative law of addition holds for
natural numbers. But this natural sum has nothing to do with
the sum � + " of the two ordinal numbers �, ". For example,
for � = "', '1 = ",2 + I, we have
�+" = '" + ",2 + 1 = ",(1 + "') + 1 = ",2 + I ,

O'(�, ,,) = ",2 + '" + I,


and these three results are different from one another.
In analogous fashion, by formal multiplication, one can also
form natural products.
I�ater on we shall need the following property of tho natural
sum:
THEOREM 2. For a given ordinal number (, the equation
(5) o'(�, ,,) = r
has only a finite number 01 solution8 �, .".
PJ'ool: The numLer r can be represented in the Corm
( = *L ","Z.. .
..

If equation (5) is to hold, we must have


z.. = x.. + y,.
for every a. This equation, for every a, hIlS exactly 1 + z ..
solutions. Since, however, there are but finitely many z.. � 0,
there exist only finitely many solution systems x.. and y.. ,
and hence, only a finite numher of solutions E, ".
11. The WeU-ordering Theorem
Earlier already we raised the question whet.her every set
can he well-ordered. We have already seen that an affirmative
111

answer to this question is also of the greatest importance for


the theory of cardinal numbers. It would show, at the same
time, that every set can be represented as an ordered set. This
question too was left open.
Cantor regarded the possibility of well-ordering every set, as
a logical necessity. The following argument can be given for
this opinion: If a set IDl is given, pick out an element mo and
put it in first position. Then choose an arbitrary element ml
in 9)1 - Imo l . and put it in second position . Then select in
any manner an clement m2 in IDl - I mo , mi l , and assign it
the next place. Continue in this fashion. If the process termi­
nated in any manner before the set IDl was entirely exhausted,
we would still have a nonempty subset of IDl left, and could
again pick out an element of this subset and place it after all
the elements already chosen. The process would thus continue
nevertheless, contrary to our assumption; it can therefore come
to an end only when the set IDl has been used up entirely.
This argument, of course, merely makes plausible the possi­
bility of well-ordering every set. Think only of the complicated
structure which a well-ordered set can have, in particular, of
the many successive limit numbers, of which there need by no
means be only an enumerable number.
An actual proof of the well-ordering theorem was first given
by E. Zermelo (Mathematische Annalen, vol. 59(1904), pp. 514-
516; a second proof in Mathematische Annalen, vol. 65(1908),
pp. 107-128). In the first of the two proofs, the significance of
the individual steps can be followed better than in the second
proof, and we shall therefore reproduce the first of the two
proofs here. In the proof, we do not first well-order the whole
set, but rather proceed from well-ordered subsets of the set IDl
which is to be well-ordered. Such subsets can be constructed
according to the simple procedure just described, and we cer­
tainly arrive at finite and enumerable well-ordered subsets of
IDl. To be able to get beyond these well-ordered subsets, they
are, to be sure, subjected to an additional condition. Starting
from these special well-ordered subsets, it is then possible to
well-order the entire set. The selection of elements from IDl
112
plays a role in the proof. Now, however, the selections are not
made successively, as in the argument presented beforei on the
contrary, at the start an element is chosen from every non­
empty subset of IDt How this is to take place remains unde­
cided. Having made these preparatory remarks, we now prove
the
WELL-ORDERING THEOREM. Every set 9» can be well-ordered.
Proof: That every finite set can be well-ordered has already
heen mentioned. We therefore assume that 9» is an infinite
set, and we break up the proof into a series of steps.
I. Let the elements of 9» be denoted by m, and let U(ID1)
be the set of all subsets of IDl. In every nonempty subset 91 of
IDl we choose in any manner an element n and associate it
with the set Ul. We call this element the "distinguished" ele­
ment of 9l and denote it also by
n = .p(9l).
Here it is by no means neceBBary that distinct subsets 9l
have distinct distinguished elements. This correspondence be­
tween elements 110 and the subsets 91 is also designated as a
covering of the set U(ID1), and remains fixed throughout the
entire proof.
II. For the set IDl and the established covering of U(IDl) with
distinguished elements, a nonempty well-ordered subset r of
IDl shall be called a r-sequence, if for every element c of r
and for the segment r. of r determined by c (r. may be 0,
hut invariably ID1 - r. � 0) the relation
'P(9)> - r.) = c

holds.
By means of this definition, the auxiliary condition for well­
ordered sets announced at the beginning is introduced. If we
have a r-sequence, we now no longer need to choose any
element from ID1 -r. ; such an element is already given with
r. , viz., c.
r-sequences always exist. For if, e. g., ?no = .p(9)>), then
113
obviously ( mo l i s a r-sequence. But w e can say even more:
Every r-sequence has mo as first element. This is true because
for the first element c of a r-sequence we have r. = 0, and
hence c =- .p(1m) = mo .
Further, if ml is the distinguished element of 1m - { mo } ,
then obviously (mo , m d i s a r-sequence. Moreover, every
r-sequence with at least two elements has ml as second ele.
ment. For if c is the second element, then, since the first element
is mo ,

r. = ( mo l , and hence c = .p(1m - (mo l > = ml •

For the two smallest r-sequences we have thus established


the fact that they are segments of every r-sequence which
contains at least the same number of elements. We have, in­
deed, quite generally, the following result:
III. Of two distinct r-sequences it is invariably the case that
one is a segment of the other.
For let r, r·· be two distinct r-sequences. Then, in any
case, since well-ordered sets are always comparable, one of
them, say r, is similar to a segment r· of the other. We have
now to show that r � r· here implies actually r = r·. In
other words, if the element of r· associated with an element
c E r is denoted by c·, we have to prove that invariably
c = c·. That this is true for the first element of r has already
been shown in II. It follows in general by means of transfinite
induction. Suppose that the assertion is correct for all c -< Cl •

Then the segments r .. and r�. coincide. Then, since

cf = .p(1m - r�.),

also Cl = cr, which proves the assertion.


From this immediately follows:
IV. If two r-sequenees have an element c in common, then
their segments determined by c coincide too. And from this
we get the following: If two r-sequences have the two clements
a and b in common, then either a -< b in both sequences or a >- b
in both sequences.
V. The union 2; of all r-sequences can be ordered.
114
For i f a , b are two distinct elements of }; , they belong to two
r-sequcnces, say r and r*. If r ¢ r*, then, according to III,
one of them, say r, is a segment of the other. Both elements
a, b then belong to r*. By virtue of the order of the elements

in r*, one of the order relations a -< b or a >- b is determined


for a, b. By the second part of IV, this determination is inde­
pendent of the particular set r* to which the elements belong.
This order relation shall now hold for a, b in }; too. This de­
termines an order for };, provided that we can show that the
order relation is transitive. Let

a -< b, b -< c

for three elements of };. Let r determine the order of a, b, and


r· the order of b, c. Then, by IV, r * also contains the segment
rb , and hence, in particular, the element a. Since the order
of two elements is determined by every r-sequence containing
them, that of a, c, in particular, is also determined by r*, i. e.,
a -< c, as was asserted.
VI. The set }; is actually well-ordered by the order just de­
termined.
We have to show that every nonempty subset };* of � has
a first element. Let c* be an arbitrary element of };*, where
for the proof we may assume that c* is not already the first
element of I*. The element c* belongs to a set r*. Let a* be
an arbitrary element which precedes c· in I*. Then, according
to the determination of the order in I, a* precedes c* in a
certain r-sequence, and hence, by IV, also in r* . The elements
which precede c* in I* thus form a subset of the well-ordered
set r* , and there is therefore a first element among them, Q.E.D.
VII. The set I is actually a r-sequence.
For let };. be the segment determined by an element c of
I. The element c belongs to a set r, and determines in this
set a segment r • .According to the argument presented in
VI,· every element of };. belongs to r. , and the converse follows
from the ordering of �. Hence, �. = r. , and consequently

tp(IDl - I.) = tp(IDl - r.) = c, Q.E.P.


115
VIII . IDl = 2:, and hence IDl is well-ordered by the given
procedure.
:For if we had 2: C IDl, IDl - 2: would be a llonempt.y subset
of rol, and would possess a distinguished clement z . Then
� + {zJ would also be a well-ordered set; hore the sum is
meant to be an ordered sum, so that z comes after all the
elements of 2:. For every segment r. of 2: + (z l , determined
by nn element c of this set,

fP(rol - r.) = C;

because for c ¢ z this is true by VII, and for c = z, by the


definition of z. Thus � + { z } would also be a r-sequence,
contradicting the fact that �, by definition, contains every
r-sequence.
From the well-ordering theorem it follows, e. g., that the
continuum ean be well-ordered, although it has been impossible
thus far to produce a specific well-ordering of the continuum.
The well-ordering theorem is just a pure existence-theorem.
With regard to subsequent critical remarks, note that,
starting from thc given set rol, only the following sets were
newly constructed: subsets of IDl, the power set U(rol), and
subsets of U(rol) . The formation of supersets of rol was thus
restricted to the formation of the power set.

12. An Application of the Well-ordering Theorem


Before we discuss the effects of the well-ordering theorem on
the theory of sets, let us treat an interesting application of
this theorem, which goes back to G. Hamel (Mathematische
Annalen, vol . 60 (1905), pp. 459-462).
Cauchy considered the problem of determining a function
f(x) such that
(l) f(x + y) = f(x) + f(y) ·
The function ex obviously has this property, for every constant
c;and if only continuous functions are sought, there are no
other solutions of (1) . For, put f( l ) = c. Then from (1) we get
116
1(2) = 1(1 + 1) = 1(1) + 1(1) = c + c = 2c,

f(3) = f(2 + 1) = f(2) + 1(1) = 2c + c = 3c,


amI in general, for every natural number n,
(2) fen) = nco

Further, I(n + 0) = I(n) + 1(0), and hence 1(0) = o. Conse­


(l\1(\ntly
I( - n) + I(n) = 1(0) = 0,
so that (2) holds for every integer n. Moreover, analogous con­
siderations for every integer m
> 0 show that

fen) = �m . �) m.��) ,=

and hence

so that
f(T) = TC
for every rational T. For a continuous function, however, this
implies that I(x) = ex.
The question remained open, whether the functional equa­
tion (1) had other solutions besides those just found, if dis­
continuous functions were allowed. Hamel, now, has shown
with the help of the wel1-ordering theorem, that there are then
indeed still further solutions. The proof runs as follows:
In the domain of real numbers, a set S8(b) is called a. basis
if it has the following two properties:
a) For any finite number of ba.sis elements b, , b. , • • • , b.. ,
and arbitrary rational numbers TI , T, , • • • , T. not all zero,
we have never
117
b) For every number z ¢ 0 there are among the basis ele­
ments a finite number of basis numbers b. , , b ... , and,
• . •

corresponding to these, rational numbers T. , • • • , T.. all


different from zero, such that

First we have to show that such a basis exists. This is ac­


complished by building up the basis as a well-ordered set by
means of transfinite induction. The starting-point is the well­
ordering
m = l A , B, C, . . . \

of the set of all real numbers. The number A shall belong to


the basis if, and only if, A ¢ O. Suppose that it has already
been decided lor all the numbers of the segment m.l' , which
numbers shall belong to the basis and which shall not. Then
X shall belong to the basis if, and only if, no equation of the
kind described under a) subsists between X and any finite
number of basis numbers already determined. In this way
a set sa is defined by transfinite induction. sa, as a subset of
the well-ordered set m, is well-ordered by m.
This set 58 is a basis. For suppose that an equation of the
sort described under a) held for some finite number of basis
numbers. 'I'hen one of these finitely many basis numbers
would be the last to have a nonzero coefficient, and could then
not be admitted into the basis, contrary to our assumption.
Condition a) is thus fulfilled. That b) too is fulfilled is seen as
follows: A given number z ¢ 0 is either a basis number b or
not. In the first case, b) is fulfilled trivially by the equation
z = b. In the second case, an equation of the kind described
in a) subsists between z and the basis numbers preceding it
in 91. On account of property a) already established for the
basis numbers, z has a nonzero coefficient in this equation. The
equation can therefore be solved fOl' z, and if we leave out the
terms with zero coefficients, assertion b) is obtained.
The representation which exists, according to b), for every
118
number z � 0, i s uniquely determined b y z. For i f there were
two distinct representations, subtraction of the two equations
would result in a contradiction of a) .
Now we define the function I(x) first only for the basis num­
bers, and this is done in an arbitrary manner. Further, we let
1(0) = 0. If x � 0 is an arbitrary real number, it can be repre­
sented in terms of a finite number of basis numbers with
rational coefficients, in the form
x = L r.b • .
..
Here we shall also admit terms whose coefficients are zero; the
uniqueness of the representation is then obviously st.ill valid.
Now we put
I(x) = :E rt/(b.) .
Then, if

is a representation of 'U in terms of basis elements,


x + Y = L (r. + 8.)b.
is such a representation of x + V, and, ag a matter of Cact, due
to the uniqueness, precisely the rcprel:!entation. Consequently,
I(x + Y) = L (r. + 8.)/(b.) = :E r../(b.) + L 8./(b.)

= I{x) + I{y) .
The function I is thus a solution of tho functional equation. It
is a discontinuous solution if we put, say, I(b.) = 0, I(ba) = 1
for two basis elements b. , b. ; for if I were a continuous solu­
tion, the argument presented at the beginning shows that we
should have to have I(b.): I(b,) = b. : b2 , which is not the CBBe.
Further relevant papers are cited, e. g., by Kamke, Jahres­
bericht der Deutschen Mathematiker-Vereinigung, vol. 36
(1927), pp. 145-156.
119
13. The Well-ordering of Cardinal Numbers
With the well-ordering theorem we immediately obtain from
Theorem 3 on p . 88:
THEOREM 1. If nt, n are two arbitrary cardinal numbers, pre­
cisely one of the three f'clations
m < 11, m = 11, m > n
holds; i. e., any two carditlal numbers are comparal)lc with each
other.
Since the transitive law holds for the relations of magnitude
of cardinal numbers, every set of cardinal numbers can be
ordered according to increllSing magnitude of its elements.
This ordered set, however, will prove to be actually well­
ordered . For this and similar considerations concerning cardinal
numbers, i t is practical to represent the cardinal numbers by
means of ordinal numbers. According to the well-ordering
theorem, every cardinal number can be represented by a well­
ordered set; to put it briefly: every cardinal number m can be
represented by an ordinal number 11-. The totality of ordinal
numbers II- which can represent a cardinal number m is desig­
nated as the number class3 .8(m). This number class .8(m) con­
tains, as does every class of ordinal numbers, a smallest ordinal
number, which is called the initial number of .8 (m), or the
initial number belonging to m.
THEOREM 2. Every transfinite initial number is a limit num ber .

Proof: If some transfinite initial number Were not a limit


number, it would be immediately preceded uy an ordinal num­
ber ", and the initial number would he " + 1 . But then v and
the initial number v + 1 would have the same cardinal num­
ber, i. e . , " + 1 would not be the smallest ordinal number in
its number class.
It is now possible to carry over §5, with the exception of

'Here, then, the m in parentheses by no means represents the clements


of .8.
120
Theorem 4, to cardinal numbers. In particular, the analogue
of Theorems 1 and 3 reads as follows:

THEOREM 3. EVel1/ set � of cardinal numbers m, ordered ac­


cording til increasing magnitude, i3 well-ordered. There exists a
cltfdinal number which u greater than every cardinal number m
in �, and, in fact, there u a definite next larger cardinal number n.

Proof: For the first part of the assertion we have to show


that every nonempty subset �, of ore contains a smallest cardinal
number. This is immediately clear, however, if every cardinal
number is represented by the initial number belonging to it.
Among these initial numbers then there is, as in every non­
empty set of ordinal numbers, a smallest one, and the corre­
sponding cardinal number is the smallest cardinal number in
�, .
That there exists a cardinal number which is greater than
every cardinal number in �, was proved already on p. 35.
Here we have merely to show, therefore, that there is a smallest
cardinal number of this kind. Let iii be a cardinal number
which is greater than every m in .re. If iii is not already itself
the smallest cardinal number of this sort, form the set of
cardinal numbers which are smaller than m and at the same
time greater than every m in .re. This set contains, as does
every nonempty set of cardinal numbers, a smallest element,
and this proves the second part of the assertion.
If the number n has no immediate predecessor, n is again
called a limit number and we write also

n = l im m .
.. <-

If the set of all transfinite cardinal numbers which are less


than a given cardinal number n, is ordered according to in·
creasing magnitude, then we denote the smallest (transfinite)
cardinal number by Ko " the next, by N, , etc. In general, every
cardinal number receives as index the ordinal number of the

'K co aleph, first letter of the Hebrew alphabet.


121
set of cardinal numbers which precede the cardinal number in
question. In particular, therefore, Wt' have invariably
Nil < N. for Ii < 11.
The initial number which belongs to Nil is denoted by w" •

The smallest transfinite cardinal number was denoted before


by a, and hence No = a. The power of the continuum is also
denoted by N without an index. Certainly, then, N ;::: N • . Tlw
question whether the equality sign holds here i8 prech;ely the
continuum problem. Further, Wo is the smallest transfinite limit
number, and is therefore the same as the w introduced earlier.

14. Further Rules 01 Operation lor Cardinal Numbers. Order


Type 01 Number Classes
Already i n ch. II we proved the formulas a · a = a and a · c = c.
These are special cases of the following general rule:
(1)
Proof: The initial number which belongs to N,. i s a limit
number, and therefore, by p. 96, is divisible by Wj i. e., it can
be represented in the form w" = W · II. If we set 1 11 I = n, then
accordingly
N,. = No n,
and hence indeed
NoN,. = NoNon = Non = N" .
Likewise we already proved in ch. II that a · a = a and
c · c = c. These are special cases of the following general rule:

(2) N! = Nil (G. Hessenberg).


Proof: The set of ordinal numbers � < WII has order type w" ,
according to p . 86, Theorem 1 , and is therefore a set of power
N,. . Consequently, the cardinal number N! can be repre­
sented by the set of pairs (�, ,,) of all ordinal numbers � < WI' ,
" < w.. • We shall now show that, on the other hand, this set
122
has at most the cardinal number N• . For this purpose we fonn
the nat.ural polynomial (p. 109)
t = O'(�, 'I) = *L w·(x. + Y.)
.
lor � < w" , 'I < WI' • l�or such tL pair of numben;, let
O'(�, 'I) = w" (x + y) + . . . ,
starting with the highest term. Since only a finite number of
nOllzcro l,tlrms appear, it follows that

(a) O'(�, 'I) < w" (x + y + 1).


Since w" hus a coefficient ¢ 0 in at least one of the representa­
tions
� = *L w·x. ,
*E w·y. , 'I =
and since � < WI' ,
'I < we have also w" < w Accordingly,
w,, ' • .

since w. is an initial number, I w" I < I w" l, and hence, since


x + y + 1 is a finite number,
I w"(x + y + 1) I < I Wil l·
Consequently, from (a) follows also
r = O'(�, 'I) < w" •

We therefore certainly obtain all number pairs �, 'I by solving


the equation O'(�, 'I) = t for �, 'I, for every r < w� But for •

every t this equation has only a finite number of solutions


(p. 1 10, Theorem 2) . Therefore the cardinal number of the set
of all solutions for all t < WI' is at most No · N.. = N,. . Hence,
N! � N. , from which the assertion follows.
From these two fundamental rules we now obtain:
(3) For Nil :5 N. we have NI' + N, = NII · N. = N. ,

and hence, in particular, N. + N. = N • .

For, N. � N.. + N. :5 2N. � N. ,


and N. � N.. N. � N! · = � • .
123

(4) For N, < N. and N. < N. we have


N, + N,. < N. + N. and N, · N. < N,, · N • .

For let N, :::; N,. , say . 'fhen, by (3 ) ,


N, + N,. = N, · N,. = N" < N • .

(5) L:,s,. N, = Nil , and, if p is a limit number, also


E N, = N" .

For invariably N,. :::; E,s. N, :::; N. · N. = N,. . If J.' is n


limit number, then, on the one hand,

E N, :::; N" ,
, <"
and, on the other hand, for every " < p, also " + 1 < p, so
that E,<p N, � N'+ I > N. ; i. c., thc sum is gr('M.cr t,han every
cardinul number which is less than N" Th('sc two I'csul!'s then
.

yield the assertion in this case too.

(6) N: = N. for finite 11 > o.

This follows immediately from (2) .


In this formula, it is not pussilJln, i n gcnnral, t.o repl:we the
exponent n by a transfinite cardinul numlJer, such as No , c . g.
It is tl1Je that N-· = N, and, for every N. N:', =

But N:· > No , and generally, for an enumeralJle fundamental


sequence of cardinal numbers with the limit N. ,

N:· > N. ,
as follows from p . 46 with the aid of (5) .
(7)
For, Il.<:cording to p . 43, b) and the hypothesis,
2-' � N'H � N" ,
124
and hence

(8)

Because 2·' > No , i. e., 2·' � Nl , implies that


N = 2·' 5 N�' 5 (2·') · ' = 2·' = N.
For the proofs of two additional aleph relations, we need the
following
Lemma. Let �(pl:) be a welkJrdered set of ordinal numbers P. ,
I .fi' I 5 N"-1 , and every I p. 1 5 Nil-I . Then there exists an ordinal
number (f < CIJ,. such that every P. < (f.
Proof: Put (f = l:.ea (P. + 1). Then (f � every P. + 1,
and hence (f > every P. , and I (f I 5 l:.ea I p" + 1 I 5
N,-1 N"-1 = N"-1 . Consequently, since CIJ" is the initial number
of .B(N,,), also (f < CIJ" •

(9) N:' = Nil N:':1 , if p is not a limit number (Hausdorff).


Proof: For p 5 II it follows from (7) that

Multiplication by Nji yields


u., u v., u v.,
n$l = "....'11 == ""�$I-l .

li'or p 5 II, this already proves the assertion. Let us assume,


then, that p > II, and first prove that
(IJ) N: ' 5 l: I (f I ·'
�< ...

if p is not a limit number. Let IDl be the set of ordinal numbers


P < CIJ,. , and let m be the set of ordinal numbers < CIJ. ; the
elements of m will also be denoted by 1:. For these sets, since
p is not a limit number, we have

('Y) I 911 \ = I CIJ.. I = N,. , I m I = I CIJ. I = N , 5 N,.-1 ,


125

and

( �) IpI < I w" I, so that I p i ::; N,,_, for p E IDl.

The left-hand side of ({J) ca.n be represented by the covering


set (cf. p. 42) 91 I IDt Every element of 91 1 9R is n set 91«n, p».
For a fixed element of this sort, it follows from the lemma,
because of (or) and ( �), that all p's of this element belong to a
set ml, with tT < w" Therefore the element 91«n, p» also

appears in the covering set 91 1 ml, for a certain tT < w" Tho •

covering £.ct 91 I ID2 is thus contained in the totality of all


covering sets 91 1 ml. arising from all tT < w" This fact implies •

({J).
From ({J) now follows, since, in the sum, every tT < w.. and
hence I tT I ::; N.. _ , ,
N: ' ::; I w.. I · N:.:, = N"N:':, ::; N"N: ' = N: ' ,
which proves (9) .
(10) For p ::; II + n, where n denotes an integer � O, we have
(F. Bernstein) .

Proof: For p ::; 1', the assertion follows from (7) after multi­
plication by N.. . The other cases follow by mathematical
(finite) induction. For if the assertion is alwady true for n,
then, for p = II + n + 1 in ({J),
I tT 1-' = 2- ' I tT I,
and hence it follows from ((J) that
N: ' ::; 2- 'N! = 2- 'N" ::; N: 'N" = N: ' .
The number class iHN.) was defined on p. 1 1 9 as the set of
those ordinal numbers which have power N" . The number
class is thus a set of ordinal numbers, and we can therefore
inquire as to its order type and its power. The answer is fur­
nished by the following

THEOREM. The number clas8 .8(N..) , ordered according to in-


126
cretuing magnitude of its elements, has order type (oJ,.. 1 , and
therefore has cardinal number NII•1 •
Proof: The clement.s of .B(K,,) arc the ordinal numbers JI with
(oJ,. :::;; JI < (oJ,. H '

Hence, in the sense of addition of w U-ordered sets, L


sm.. ,. + .B(N,.) = sm ..,.. . .

If s is the ordinal number of .B(N..), this equation implies


(e) "'"+ r = "',..1 and KII + I r I = N,.+! .
From this follows I r I :::;; Nil. 1 But, on account of (4), we

cannot have I r I < Nil" . Therefore I r I = Nil. 1 , so that


s � "',, + 1 , and hence, because of (e), we have indeed r = (oJ,,+ I •
Cantor called the set of all finite ordinal numbers the first
number-class, and the set of ordinal numbers with cardinal
number No , the second number-class. We just showed that
this number class has power NI . The continuum problem can
accordingly be formulated also as follows: Does the second
number-class have power N?

15. Ordinal Numbers and Sets of Points


Cantor made use of ordinal numbers in order to define, for
sets of points, derived sets of arbitrarily high order. We shall
restrict ourselves here to point sets which lie in the plane, be­
cause the matters can be visualized best in the plane. The
following considerations, however, can be carried over im­
mediately to space.
By an e-neighborhood or briefly a neighborhood of a point
Po we mean the set of all points P whose distance PPo from
Po is less than ej in other words, the interior of the circle with
center Po and radius E. If a point set IDl is given, a point P,
irrespective of whether it belongs to IDl or not, is called a limit
point of IDl, if every neighborhood of P contains infinitely
mnny points of rol, or, what amounts to the same thing, if
every neighborhood of P contains at least one point which
127
belongs to IDl and differs from P. For example, if IDl is the set
of all points, both of whose coordinates are rational numbers,
then every point of the plane is a limit point of IDl. On the
other hand, the set of points, both of whose coordinates arc
integers, has no limit points. Likewise, no finite set has a limit
point.
By the first derived set IDl' of a set IDl we mean the set of
limit points of IDl, provided that such points exist. If IDl has
no limit point, we put IDr = O. The second derived set IDl "
of IDl shall be the derived set of IDl', so that IDl" = (IDl') ' . By
continuing this definition, we can thus ascend to derived sets
IDllnl of arbitrary finite order. By transfinite induction, how­
ever, we can ascend also to derived sets gnl.) whose order "
is an arbitrary ordinal number. If the derived sets are already
defined for all I/o < ", we put

{ (IDlI'_ll)', if " is not a limit number;


gnl,) -
- the intersection of all IDlI.."s for all I/o < /I, if " is a
limit number.

Accordingly, we have, e. g.,


IDl I .. ) =
IDl' . IDl" . IDl'" " ' , IDll w+11 = (IDl I .. »,.

Actually, however, there is no point in ascending to derived


sets of arbitrarily high order. For, as Cantor proved, the process
automatically terminates with an ordinal number of the second
number-class .B(No) in the sense that all subsequent derived
sets are identical with one another. The proof of this theorem
is the main goal of the following discussion. For the proof
we have to introduce a few more concepts from the theory of
point sets.
First let us sharpen the concept of limit point: I f a point. set
IDl is given, a point P, irrespective of whether it belongs to IDl
or not, is called a condensation point of rol, if every neighbor­
hood of P contains a nonenumerable number of points of IDt
If, e. g., IDl is the set of points on the number axis which
128
correspond to the irrational numbers, then every point of the
number axis is a condensation point of IDl.
Regarding the existence of limit and condensation points,
there are the following two theorems:

THEOREM 1 (Bolzano-Weierstrass) . Every bounded5 infinite


point set IDl has at ieast one limit point.
Proof: Since IDl is bounded, IDl lies completely within some
closed square £1. Subdivide the latter into four congruent
squares, and consider each of these as closed. Since IDl i� in­
finite, at least one of these four subsquares also contains in­
finitely many points of IDl; let £11 denote such a square. �ow
we again quarter £) 1 , and denote by £12 one of the quarters
containing infinitely many points of IDl; etc. We thus obtain
a sequence of nested squares whose sides tend to the limit O.
This nest of squares therefore determines a point P, and this
is a limit point of 9)1, because every neighborhood of P cer­
tainly contains one of the squares O n completely, and the latter
contains infinitely many points of IDl.
Note that the proof asserts nothing as to whether P belongs
to 9)1 or not.

'l'Ht;OREM 2. Every 1/ollCll umerablc point set ID1 has at least one
condensation point.
Proof: First, let IDl be bounded. Then the assertion follows
by replacing the words "limit point" and "infinitel, many" in
the preceding proof, by "condensation point" and "non­
enumerably many". If IDl is not bounded, we divide the plane
up into the squares

a ::; x < a + 1, b ::; y < b + I,

where a and b run through all the integers. At least one of

thrsc squares eontains nonenumerably many points of IDl, be­


('ause otherwise 9R, as the union of the subsets of IDl lying in
t he enumerably many squares, would be at most enumerable,

'A point set is said w be bounded, if it lies entirely withiD Borne square.
1 29

contrary to hypothesis. But then the part of the theorem already


proved again yiclds at least one eondcnsation point.
Thc concepts " limit point" and "condensation point" are
closely rclated to two other coneepts. Between a poi n t. ['et 9J?
and its first derived set ID1', th e following relations, among
others, are concei" ablc:
u) [ll' C IDl or b) un C IDl'.
Both ca�':'l aetually occur too. On the numbm' axis, the Het
to, 1, }, 1. · · · 1 furnishes an example of case 11.) , and th e 8l�t
of rational numbers is an example of the second casc. Wc arc
therefore entitled to make the following definition:
A point set IDl is said to be closed, if un' C un. It is said to
be dense in itself, if un c un' and un ¢ O. It is said to be per­
fect, if IDl' = un and IDl ¢ O.
Every closed interval is a closed, and actually a perfect, set.
Every open interval is dense in itself, but not closed, and
therefore also not perfect. Every finite set is closed, because
its derived set is the empty set, and the latter is a subset of
every set. Every finite set, howc\'cr, is not dense in itself, and
therefore also not perfect.
THEOREM 3. The intersection :l) of arbitrarily many closed sets
IDl is closed.

Proof: If �' = 0, the assertion is trivially correct. Suppose,


then, that D is a Hmit point of 1). Then every neighborhood of
D contains infinitely many points of ID, and hence also in­
finitely many points of every un. This means that D is a limit
point of every un, and hence, since the un's are closed, a point
of every un. D therefore is also u point of ID; i. e., ID is closed.
THEOREM 4. Every derived set [ll e.) of a point set un is closed.
Proof: We first prove the asscrtion for the first derived set,
where we may assume that the next derived set un" ¢ O.
Now let P" be any point of un". We have to show that P"
belongs to un', i. e., that P" is a limit point of un. In fact, given
E > 0, there exists at least one point P' of rol' in the �E-neigh-
1 30
borhood of pll, because pll is a limit point of IDl'. Since P' is
a limit point of IDl, there arc infinitely many points of IDl in
the �E-neighborhood of P', and hence a fortiori, in the E-neigh­
horhoml of PII. From the fact that tho next derived set of every
set h; closed, however, the gem:ral validity of the theorem follow8
hy transfinite induction. For we just proved the theorem for
the first derived set. Suppose that it has already been proved
for all IDl 'l'l JS with J.& < II. If II is not a limit number, then ml,),
a..'i the derived set of IDl" - I l , is closed, according to the part
just proved. If, however, II is a limit number, then IDe') is the
intersection of all the preceding derived sets, i . e., the inter­
section of closed sets, and the assertion then foUows from
Theorem 3.
The derived set is the set of limit points. For condensation
points, the analogue of Theorem 4 reads as follows:
THEOREM 5. For every nonenumerable set IDl, the set of its
condensation points is perfect.
Proof: Let 58 be the set of condensation points of IDl. We
shall first prove that 58 is closed, and in doing so we may as­
sume that 58' ¢ O. Every neighborhood of a point V' of $'
contains at least one point of 58. This neighborhood, therefore,
by the same argument as in the preceding proof, contains also
nonenumerably many points of IDl. Consequently, 1" belongs
to 58, i. e., 58 is closed. On the other hand, 58 is also dense in
itself, i. e., every neighborhood of an arbitrary point V of 58
contains at least one other point of 58. For if this were not the
case, there would exist an E-neighborhood of V containing no
other point of 58. By successively halving E, we construct a.
sequence of concentric circles about V, whose radii tend to the
limit O. These circles determine a sequence of circular rings
II , l2 , • • • According to our assumption, no circular ring

contains a point of 58, i. e., every circular ring contains at most


enumerably many points of IDl. Since the sum of the circular
rings, however, gives the E-neighborhood of V except for the
point V itself, the E-neighborhood of V would also contain only
enumerably many points of IDl, i. e., V would not be a con­
densation point of IDl, 'contrary to our assumption.
131
The converses of the two thcorems just proved are also true;
i. e., it can be shown that every closed set of points can be
regarded as the derived set of some set, and that every perfect
set coincides with the set of its condensation points. We shall
not, however, make any use of these facts. On the other hand,
we do need
THEOREM 6. Every perfect set of points ha3 power N.
Proof; The set of all the points of the plane has power K,
and the power of the given perfect sot IDl is therefore at most
N. We have merely to prove, then, that I lIn I � N. Since IDl
is a perfect set, it certainly contains infinitely many points.
Let Po , PI be two points of IDl. Describe mutually external
circles !to , .It I about these two points. Since IDl is dense in
itself, there are two points POG , POI in io which belong to IDl.
About each oC these two points we again describe mutually
external circles !too , ROI , which lie within .fi'o and have radii
t.lmt arc at most equal to half the radius of �t'o . We do the
same in il , and the circles obtained there are denoted by
i lo , JIll • We can continue in this fashion. In every circle,
about two points of IDl we always describe circles which lie in
the preceding circle, are mutually external, and have radii tha.t
are at most equal to half the radius of the preceding circle.
The new circles receive as indices the index of the preceding
circle with an added 0 or 1 . We thus oLtain, for every termi­
nating dyadic fraction
d.. = O ' alaa • • • all , a circle !t.......... .

If
d", = a ' al ' " a..a,,+ l ' " a.. ,
then .fi'. . . . .... is contained in .fi'. . . . . .. , and the diameters of the
circles Sf. . .. ... tend to the limit a as n -HI') .
With every infinite dyadic fraction (which may, from a
certain point on, l�ontain nothing but zeros) O ' alaaa3 • • • ,
there is thus asso ciated a sequence of nestell circles whkh,
because of the limit relation for their radii, determine precisely
one point P belonging to all the circles of this sequence. Since
every circle contains points of IDl, P is a limit point of IDli and
132
due to the fact that IDl is closed, P is actually a point of IDl .
Further, from the way in which the circles were constructed,
it follows that two distinct dyadic fractions determine two
distinct points of IDl. The set IDl therefore contains a subset
which is equivalent to the set of all dyadic fractions o · a.¥3
. . . . Since the set of these dyadic fractions has power 2 - · = K,
/ IDl / � K, and the theorem is proved.
After these preparations, we now tum to the proposition
formulated above on p. 127, and we prove a general theorem
from which the aforementioned proposition easily follows.
THEOREM 7 (R. Baire). For every ordinal nllmber Q of the first
or second number CUus, let a closed set IDl a be defined, and let

IDl� C IDla for Q < {3.


Then there exists an ordinal number 'Y oj the first or second number
class, such that
6
IDl .. = IDly Jor every Q � 'Y.

Proof: Let Q be a number for which a 8 > Q exists such that


IDl , C IDl a , and in fact let 8 be the smallest number of this
kind. Among the defined IDl's which satisfy this relation for
this 8 and which are equal to IDl.. , there is one whose index is
smallest, and we shall assume that this IDl is already IDl.. to
begin with. Then IDla contains a point P which docs not belong
to IDl. , and hence, since the sets are closed, 1m.. contains a
neighborhood of P which docs not belong to IDl, . There is
therefore a circle fa which contains P but no point of IDl, , and
which has a rational radius as well as a center whose coordinates
are both rational numbers.
Let Q < {3 be two ordinal numbers of the first or second
number class, for which fa and f� exist. Then fa � f6 . For
there is then a smallest number 8 > Q for which rol, C IDla ,
and 8 =:; {3 . Since fa contains no point of rol, , and since IDl., C
IDl, , fa contains no point of IDl/f . But f/f contains at lenst one
point of roliJ , and therefore f. ¢ f6 .
aFor further theorems of a similar nature, sec nausdorff (2), p. 170.
1 33
According to the last example on p. 9, there are only an
enumerable number of circles with rational radii and centers
whose coordinates are rational numbers. Consequently, there
exist also merely an enumerable number of the above fa's; i. e.,
the indices of the fa's form an enumerable set of ordinal num­
bers of the first or second number class. By p. 9 1, Theorem 4
there is then an ordinal number 'Y of the first or second number
class, such that a < 'Y for each of these a's. Thus, for every
a � 1', no fa exists; i. e., for these a's, IDl a = IDly .
From Baire's theorem now follows
THEOREM 8 (Cantor-Bendixson). Let IDl be a closed set. Then
there exists a number 'Y oJ Ute first or second number cla8s, such
that IDl'Y ) is either empty or perfect. Further, IDl ' a ) = IDl' Y) Jor
a � 1', and IDl' Y ) = 0 or :> 0 according as IDl is cnllmerab� or
not. In every case, IDl can be represented in the Jorm IDl = IDl ' Y ) +
m, where m is at most enumerab� and m · IDl' Y ) = O.
ProoJ: Since, according to Theorem 4, every derived set is
closed, it follows from the definition of derived set, that the
derived sets IDl ' '' ) satisfy the assumptions made concerning the
IDla's in the preceding theorem. lIence, there exists a number
'Y of the first or second number class, such that IDl' a ) = IDlly)
for a � 'Y . IDl' Y), as a derived set, is closed; but it is also either
dense in itself or empty, because otherwise IDl ' Y ) would con­
tain a point which was not a limit point of IDl ' Y ' , and this
point would therefore disappear if we went over to the next
derived set, so that we should have IDl ' Y + !) ¢ IDl'Y). Conse­
quently, IDl ' Y ) is either empty or perfect. If IDl ' Y ) :::> 0, and
hence is perfect, then, since IDl'Y) C IDl, IDl contains a perfect
subset, and therefore, by Theorem 6, IDl is nonenumerable.
Conversely, if IDl is nonenumerable, then, according to Theorems
2 and 5, IDl contains a perfect subset which then belongs also
to all the derived sets, so that IDl ' Y ' :::> O.
Further, m is at most enumerable. For let m.. be the set of
those points of m which have a uist9.nc� � 1 /r. from every
point of IDl ' Y ) . Then m" is a closed set, and is also enumerable.
For if 9l" were not enumerable, application of that part of our
134
theorem which has already been proved would show that 91.
contained a perfect set, which IDlIT) would then have to con­
tain in addition. Since every single point of the set m has a
positive distance from IDl'T) (this follows from the fact that
IDlIT) is closed), m is the sum of all the m,,'s for n = 1, 2, 3,
. • .

91 is thus the sum of cnumerably many sets, each of which is


at most enumerable, so that 1)1 itself is at most enumerable.
Finally, m · IDl' T ) = 0, and hence, since m C IDl and therefore
m i T) C IDl I T ) , a /oniori also m · m'T) = O.
For Cantor, the motive for the investigations described here
was the continuum problem, which, in the more general form
of whether there always exists another cardinal number be­
tween N. and 2 1 " is the central problem of the theory of sets.
The cardinal number N can be represented by the points of
the interval (0, 1). If NI < N, the interval (0, 1) must contain
a point set of cardinal number NI The Cantor-Bendixson
.

theorem shows now that such a set, if it exists at all, i. e., if


NI < N, at all events cannot be a closed set. For from
IDl = IDlIT) + m

it follows, since IDl' T ) is either empty or perfect, and hence, by


Theorem 6, has either cardinal number 0 or N, that every
closed set has a cardinal number which is either �No or = N.
Concluding Re11UJTks

We have already occasionally pointed out several contradic­


tions which arise in pursuing the concepts employed in the
theory of sets. These contradictions were connected with the
following concepts :
a) The set of all cardinal numbers (p. 36).
b) The set of all ordinal numbers (Burali-FortPs Paradox,
p. 9 1 ) .
c ) The set o f all even ordinal numbers (p. 96).
To these the following paradoxes can be added :
d) The set of all sets which do not contain themselves as
elements (Zermelo-Russcll).
For this set rol, as for every set, only the following two cases
are conceivable : either rol contains itself as an element, Or it
does not. The first case cannot occur, bccause otherwise !In
would have an element, namely, rol, which contained itself as
an element. The second assumption, however, also leads to a
contradiction. For in this case, rol would be a set which did not
contain itself as an element, and for this vcry reason, IDl would
be contained in the totality of aU such sets, i. e., in IDl.
e) The set of all sets.
For if IDl is this set, it would have to contain the set of all its
subsets. The set of all subsets of IDl, however, has a greater
power than IDl.
Here we have aided ourselves by erecting danger signs before
the paradoxical notions. All proofs have been carried out so as
to avoid those ideas which have hitherto been recognized Soc;
self-contradictory.
Such a procedure is justifill.ble if the purpose is to acquaint
someone for the first time with the results and methods of proof
peculiar to the theory of sets. But a rigorous and final construc­
tion of set theory cannot be attained in this way. For we should
have to fear the possible existence of more paradoxes (in fact, a
135
136
recipe is given below for the construction of additional para­
doxes).
In the paradoxes listed above, it strikes one that the word
"all" appears in each of them. AB a result, the set e), if it is
meaningful, is a set which contains itself as an element. The
same holds for the set d). For if IDZ is any set whose elements are
exclwively sets which do not contain themselves as elements,
then IDZ cannot be an element of itself. This argument has noth­
ing to do yet with the word "all". Not until the second part of
the argument given in connection with d) is a conclusion drawn
from the occurrence of the word "all " , this inference being that
the set d) would have to contain itself as an element, which then
leads to a contradiction.
These two samples lead one to conjecture that all sets which
contain themselves as elements are, as a matter of course, se1£­
contradictory concepts, and are therefore inadmissible. In fact,
up to now no sets are known which contain themselves as ele­
ments and which everyone would regard without hes'tat�,)D as
meaningful sets. Also, Cantor's definition of a set as "a collection
into a whole of definite, well-distinguished objects" must be
interpreted to mean that something new is created by this act
of collecting, so that a set can never equal one of its elements.
If we do not admit sets, then, which contain themselves as
elements, paradoxes d) and e) fall to the ground, since the sets
are inadmissible according to this criterion. The elimination of
the remaining paradoxes by means of this criterion, however, is
not possible in a satisfactory manner. One can say that cardinal
numbers and ordinal numbers are nothing but sets, and that in
order to arrive, e. g., at the set .re of all cardinal numbers, one
would first have to form the set of all sets, which has already
been recognized as meaningless , and then pick out a representa­
tive from every class of mutually equivalent set.�, in order to
obtain the elements of .re. But one is by no means forced to take
this path : In order to arrive at cardinal and ordinal numbers, we
do not have to consider sets of an arbitrary sort, bllt, on the
contrary, may restrict ourselves to sets with well-defined mathe­
matical objects as elements. For t.he first few ordinal numbers ,
1 37
such representations by means of the set 1 0, I , 2, . . . } arc
given on p. 86.
If, in cases a) to c), we order the set according to increasing
magnitude of its elements, then the contradiction becomes im­
mediately apparent if we ascribe an ordinal number to this
fundamental sequence. There is, however, an analogue of this
for ordinary sequences of real numbers in analysis. A monotonic­
ally increasing sequence of real numbers converges only if it is
bounded, i. e., if its elements are less than a fixed real number.
If one were to assu me that every monotonically increasing
sequence of real numbers converged to a real number, one would
soon arrive at contradictions. It appears to bc similar in the
theory of sets. For not only do the antinomies a) to e) disappear
when we admit as elements of sets only such seL", ordinal num­
bers, and cardinal numbers as arc bounded above by a fixed
cardinal number, but we see also that paradoxes always arise
if we collect into a set any sets, cardinal numbers, or ordinal
numbers which are not bounded above by a fixed cardinal num­
ber. Thus we have here a recipe for constructing additional
paradoxes. We can form, e. g., the set of all ordinal numbers
which are divisible by some fixed ordinal number a, or the set
of all initial numbers, or the set of all sets which have a la'>t
elcment, etc. All these sets prove to be self-contradictory.
Whereas the given principle removes all hitherto known
paradoxes of set theory,' this solution nevertheless cannot be
regarded as satisfactory, because a mere argument by analogy,
such as led us here to our principle, posseses
s no force of con­
viction, but can only give us hinL" as to the direction in which,
perhaps, the solution of the problem is to be sought. One is
reminded of the contradictions which the history of mathe­
matics shows attended unscrupulous operation with infinite
series and sequences, and which were subsequently removed by

'In the development of the theory of sets, some other paradoxes, to be


sure, besides those mentioned above play a role. But these do not belong
basically to set theory, but rather to general logie, and their resolution is
therefore also to be expe cted only from this direc�ion. For this reason they
may be disregarded here.
1 38
clarifying the concept of number and the circumstances under
which a sequence or seriE'.5 can represent a number. What is of
foremost importance, then, is the production of the scales of
cardinal and ordinal numbers in a rigorolls fashion. If this is
attempted by means of a constructive principle, it turns out.
(cf. Enzyklopiidie, II , Heft 2, article 5, no. 27) that certain
principles of construction lead to a very comprehensive domain
mJ1 of ordinal numbers (and hence also of cardinal numbers) ,
but that, in accordance with our resolution above of the para­
doxe�, no ordinal number of the domain mJ l is the ordinal num­
her of the set of aU ordinal numbers of this domain. If we extend
the construction principles in a suitable manner, we obtain an
even more comprehensive domain mJ l I of ordinal numbers, but
once again none of the ordinnl numbers defined thus far corre­
sponds to the set of all ordinal numbers of this domain. This
observation repeats itself on continuing such a constructive
development. It corresponds exactly to the aforementioned fact
that we encounter paradoxes if we consider boundless sets of
cardinal or ordinal numbers.
For further critical remarks concerning the foundations of the
theory of sets, the reader is referred to Enzyklopiidie, II , Heft 2,
article 5, nos. 1 1-16.
Bibliography

For more extensive studies of the whole field of the theory of sets, the
reader is referred to:

Hausdorff, F., [II Grundzilge der Afengenlehre, 1st edition (reprint), New
Yorle, 1949.
--- , 121 lIfengenlehre, 3d edition (reprint), New York, 1944.
Sicrpiilllki, W., /.efons Bur lea nonWres tran.�.finiB, Paris, W28.
--- , HypothUe du continu, WIU'Sz&w&-Lw6w, 1934.

The reader who is more interested in the development of the principles


should consult:
Fr&Clikd, A., [1]
Einleilung in die Mengenlehre, 3d edition (reprint), New
York, 1946.
---" 121 Zehn Vorlesungen uber die Grurnllegllng der !Ifengenlehre,
Leipzig and Berlin, 1027.

Wu mention, in addition, the following reference works:


Enzyklopddie der MathemaliBchen iVissenschaften, Band I" lIeft 2, 2d
edition, Leipzig and Berlin, 1939.
Pascal, R, Repertorium der hiiheren Matfiemalik, vol. 1, 2d edition, parts
1 and 3, Leipzig and Berlin, 1910, 1928.
Sehoenftics, A., Die Entwickelung der Lehre lIOn den Punklmannigfaltigkeiten,
part 1 [Jahresbericht. der Deutschen Mathematiker-Vereinigung, vol. 8
(1900) ], part 2 [Jahresbcricht der Deutschen Mat.hematiker-Vereinigunl!,
lId supplementary volume (1908)], Leipzig.
Sehoenfiics, A., and Hahn, H., Entwickelung der Alengenlehre und ihrer
Anu-endungen, first half: AUgemeine Theorie der unendlidlen Mengen und
Theone der Punklmengen, by A. Sehoenftics, Leipzig and Berlin, 1913.

139
Key to Symbols

e, 1 ""', 14 I: ID'l. , 7, 61
t, 1 �, 55 .re I ID'l, 42
C, G l ID'l l, 18 Q, c, f, 18
C, G l ID'll, 57 ml" , 86
-<, 52 IDHn, 7 '1, 65
>-, 53 ID1 X 91, 28, 62 >., 64
< , 18, 85 IT IDl. , 7 w, 57
> , 20, 85 x I1 ID'l.. , 37 ·w, 58

141
lllllex

.·\ddition, ordered, 59, 01 Dilll:ollni process, lil'llt , !l


,\IKdJr.lic number, 3 - -, !'i"'Clln,l, !!
:\ulIlylic function, 50 Di�jllllcl, 7

IIl1llia of lIio real numbc�. I Hi EI" IIWIII, fi,-,;t, .''>1


IIcLWL'Cll, 5 1 -, llIlIl , 5 t
Horder clement, O() E[ell1enI8, eOIl3l)CIII["I.', 70
-, ndJ;hboriu!;, 70
CRrdillnl n\Ullbct, 18 Empl.y !!Ct. G
ClLrtlinnl-numbcr complex, 32 ElIIlIllers\.JIc, 2
ClLnlinal numbcr of point lieU, -19 &Iuiv!\[encc theorem, 23
Cllrdinnl lIumbcl'l:l, rulcs of OllCrll.- &lul\" I[cnt., '4
lion for, 1 2 1 EudidcRII hl"orithm, 95
Close,1 in\er"t1I, 1-1
- !!Ct., I2'J hUII\lIl1lenllll llCquence, !J2
Comhinntorinl product, 28, 37
Comllfl.rll.bility thoorem for cnrtlinill GIIII, 7 1
muubcnl, 88, I I!)
- - - ordinsl lIumbcl'lJ, 87 "Iell[ IICt, G
Complementsry IICt, G Inili!li part, 73
Complex, :12 - lIum\x'r, I I!)
-, ordel\!d, of ordcr Iypc�, til inlcN(!Cliol1, 7
-, \\'cllo()rlicr<!d, of ordin(li lIum- isolillcd number, !J2
iJCI'lJ, 81
CondCII!SfItion point, 127 Jump, i I
Coutinuous functions, SO
- set, 75
ContinuuIII, 1(;
- problem, 21, 12(; Lexicographici'll order, {i2
Counting of sets, 98 Limit number, !)2, Oil, 120
Covelin]:, ,12 - point, 120
- set, 42
Cut, i-I �1!lPIJinJ;, ...
-, Himilarity, 55
Decomposi\ion, '
j .! �lulu!ll1y exclusive, 7
Dellsc, 50, ,0, i8
- in it;SClf, 1211 Nillural pmtlud, 110
Derl\'ed set, 12; - Hum, 1(f.I
144
Neighborhood, 126 Remainder, 74, 84
Null set, 6 Residual set, 6
Number class, 1 19, 126
Segment, 84
Open interval, 1 4 Sequence of ordinal numbers, 89, 98
Order according to first differences, - - - -, decreasing, 90
62 - - - -, increasing, 89
- type, 57 Similar, 55
Ordered set, 52 Subset, 6
Ordinal number, 80 -8, set of, 21, 43

- -s, evell, odd, 96 SucLoeeds , 53


Sum of cardinal numoors, 25, 32
Perfect, 129 - - order types, 61
Points of discontinuity, 4 - - ordinal numbers, 81, 10 1
Powers of cardinal numbers, 4 1 , 47 - - sets, 7
- - order types, 66 -s of powers of ordinal numbers,

- - ordinal numbers, 104 107


- - sets, 18
Precedes, 53 Transcendental numbers, 10, 27
Product of cardinal numbers, 28, 37 Transfinite cardinal number, 20
- - order types, 62 - induction, 99
- - ordinal numbers, 100 Type class, 66
- - the first kind, 7
- - - second kind, 28, 37 Union, 7
-, ordered, 62
Well-ordered set, 79
Real functions, 51 Well-ordering theorem, 1 1 2
A CATALOG O F S E L EC T E D

DOVER BOOKS
I N S C I E N C E A N D M A T H E M ATICS

OJ
A CATALOG OF S E LECTED

D OVER B O O KS
IN SCI ENCE A N D MATHEM ATICS

QUALITATIVE THEORY OF DIFFERENTIAL EQUATIONS, V.V. Nemytskii


and V.V. Stepanov. Classic graduate-level text by two prominent Soviet mathe­
maticians covers classical differential equations as well as topological dynamics
and erqodic theory. Bibliographies. 52!1pp. 5� x BIt 65954-2 Pa. $ 1 0.95

MATRICES AND LINEAR ALGEBRA, Hans Schneider and George Phillip


Barker. Basic textbook covers theory of matrices and its applications to sy�tems of
linear equations and related topics such as determinants, eigenvalues and differen­
tial equations. Sumerous exercises. 4!12pp. 5� x Bli. 66014-1 Pa. $B.95

QUANTUM THEORY, David Bohm. This advanced undergraduate-level tt'xt


presents the quantum theory in terms of qualitative and imaginative concepts,
followed by specific applications worked out in mathematical detail. Preface.
Index. 655pp. 5" x Bli. 65969-0 Pa. $ 1 0.95

ATOMIC PHYSICS (Sth edition), Max Born. Nobel laureate's lucid treatment of
kinetic theory of gases, elementary particles, nuclear atom, wave-corpuscles, atomic
structure and spectral lines, much more. Over 40 appendices, bibliography. 495pp.
5* x Bli. 659B4-4 Pa. S 1 1 .9:'

ELECTRONIC STRUCTURE AND THE PROPERTIES OF SOLIDS: TIle


Physics of the Chemical Bond, Walter A. Harrison. Innovative text offers basic
understanding of the electronic structure of covalent and ionic solids, simple
metals, transition metals and their compounds. Problems. 1980 edition. :'82pp.
6� x 9'4. 66021-4 Pa. S 1 4.95

BOUNDARY VALUE PROBLEMS OF HEAT CONDUCTION, M. Necali


(hisik. Systematic, comprehensive treatment of modern mathematical methods of
solving problems in heat conduction and diffusion. Numerous example� and
problems. Selected references. AppendiL-es. 505pp. 5* x Sli. 65990-9 Pa. S 1 1 .95

A SHORT H ISTORY OF CHEMISTRY (!lrd edition), J.R. Partington. Classic


exposition explores origins of chemistry, alchemy, early medical chemistry, nature
of atmosphere, theory of valency, laws and structure of atomic theory, much more.
428pp. 5* x Bli. (Available in U.S. only) 65977- 1 Pa. $ 1 0.95

A HISTORY OF ASTRONOMY, A. Pannekoek. Well-balanced, carefully rea·


soned study covers such topics as Ptolemaic theory, work of Copernicus, Kepler,
Newton, Eddington's work on stars, much more. Illustrated. References. 52 1 pp.
5* x 81i. 65994-1 Pa. SI 1.9)

PRINCIPLES OF M ETEOROLOGICAL ANALYSIS, Walter J. Saucier. Highly


respected, abundantly illustrated classic reviews atmospheric variables, hydro­
statics, static stability, various analyses (scalar, cross-section, isobaric, isentropic,
more). For intermediate meteorology students. 451pp. 6� x 9\4. 65979-8 Pa. $ 1 !!.95
CIIT/II_OC; Of' DO VEIl 1l00KS

RELATIVITY. T I I E R .\IOIWNtU,llC'i ANn COSM O LOG Y , Rir.!wd C 1'01·


m:l1l. L:mdrn:'rk sludy eXlends IhennocirnamiD to 'pui'l!. gener.,l ldati, il)'; abo
applications of rdath'hlic mrd,anics. Ihermod)namiD to cosmologir.11 models.
501PIl. , '
; ' " 8i 6!J:S83·11 P". 5 1 1 .95

AI'PUt:1l AN,\ LYSIS. Qunclht) L..1ncIOs. C1as�ic work on an'lly�is :111<1 (k-sign of
finit.: luoo::e$$C� 101 apprm.:irnaling solution of analytical 1)lOh1cm�. Aigehraic
.
l�l ualions. rn:Hrir es. h;mnonie anal)·sis. quadr.uult· mClhods. much more. ;iS9pp.
5''' 8�. 6S(i56·X P:•. 5 1 1 .9:'

SI'ECIAL REL.,\TIVITY FOR P I I l'SICISTS. G. Strphenson and CW. "ilmislet.


Conci� r1cgaru aCCOUIlI for non�puia1iSls, LOICI1lI IlOlnsiorma.ion. npliGl1 "nd
d)'l,:unical applir.uion>. mort'. Bibliography. 108pp. 5 , ,, 81i. (i55 19·9 I'a. U.9:'

INTRODUCTION TO ANAl.YSIS. Muwcli ItoM·nlidH. UuuSIIall)·dear. :teees·


sible coverage of �I Ihrory. 11':11 numher system. mClric spaces. rolllinuou�
fllnClions. Riemann imegr. uion. rnultiple intl'gl-ah. more. Wide r.Ulgeof plOblems.
Ulldergr:ldu;ue leu:l. Bibliogr.lphy. 2';·IP II. :.,,, 8". 6S038·' l'a. 57.00

INTRODUCnON TO QUANTUM :o.IECIIAN1CS Wuh AppliC:ltioIl5IO Chem·


hlry. Linus Paulin!; II: E. IIright Wilson. Jr. Classic lIIuicrgmdu:He text hy Nobel
Prile ",jnllel" :Ol 'plics quanl1HU mcd,ani<:s to chemiGll :llld ph)·,jc:Ll pmhlfnlS.
Numemus tahles ami figures enh,mce the tCXl, Chapler hibliographid.. ,\pren·
dieel. Index. -I68I)P. 5"" Sn. &1871·0 l';t. 59.95

ASY�[ PTOTIC EXI'ANSIONS O.' INTt:GRALS. N or",al) Uki�l('in S: R ich:lId A.


I bndelsman. neSt imroduclion 10 irnpcJI Ian I field with :.ppliC' uiuns
.. in a ,,:11icty of
�ientific disciplines. Nc,,' pld"ee. Problems. Di:'grams. T:obles. lIihlio!;laphy.
Index. -!-ISpp. 5)\" 81t 65082·0 l'a. S 10.95

"'ATIiDIATIC.'i ,\PI'U ED TO CONTINUUM MECHANICS. Ltt A. &gd.


An'IIY/.n model� of f[" id flow andsolid ddOlln:ltiOl" FOI upper·leo'el l"atll. scienc('
;Iud engineering Situltim. 608pp. ';\" 8". 65�(i9·2 l'a. 512.9';

f.Lt:�IENTS OF REAL ANALYSIS. [hvid 1\. Splechn. Classit text COVCl"S


fumbmcnl,11 concepls. re,'] number system. point sel�. r UnCtions of:t Ical ,·ariabic..
I'ourin !-Cries. lIlurll more. 0"1'1 :'00 t'xelclSl'S. 3:'2pp. 5!!1" 8\,. 6r,3SS·-I I'a. 58.%

1'1,1 YSICAL PRINCIPLES OF"J"tl E QUANTU �I 1'1-1EOR Y. WelllCi Heisenberg.


Nobel L.1111ca1C discusses qu;ulIum Iheory. IInce,,,,inIY. ,,'al'C mn;h;miQ. "'OIk of
Dirac. Schroedin.l:cl. CampIOn. Wilson. Eilmei". etr. 184pp. ';\" IlIi.
6011�·7 1':0. 5,1.9:'

INTRODUcrORY RE/\L ANALYSIS. ,\.N. "0Irn0.l:0I01·. S.\'. Fomin. Tr.U1s,


I,no:d by Rich:ud A. Silverman. Sdf,colHained. e...:nl), [);Icetl irmod"elinn 10 re:.1
,Ind lunelional :Ulalysi). Somc�.;o ploblcms. -I03P1). ';\" 8�. 61226·0 1';1. 57.9:'

I'IWIILDIS ,\NIl SOLUTIONS IN QUANTUM CU EM ISTR Y ANI) 1'1-1 \'SICS.


Charles S. Johmon. JI. ;Ind Leo: G. 1"t1:1eI!ic:n. UnuSII:tlly "ark-d prohlems. delailed
wlUlions in CO\'cT:lge 01 'tllalHllm mechallics. wavc mcchanic,. angnl ..r mOllleli'
turn . molecular SI>t<:tros.;op)'. $Cutering lI,cory. more. 280 prohlem. plus 139
supplcrneruary exerciso. ·1l0PJl. 6li " !)IA. 65Z3G·X 1':1. 510.95
CATALOG OF DO VER B OOKS

THE FOUR-COLOR PROBLEM: Assaults and Conquest. Thomas L. Saaty and


Paul G. Kainen. Engrossing. comprehensive account of the century-old combina­
torial topological problem. its history and solution. Bibliographies. Index. 1 1 0
figures. 22Spp. 5" x 81S. 65092-8 Pa. $6.00

CATALYSIS IN CHEMISTRY AND ENZYMOLOGY. William P. Jencks.


Exceptionally clear coverage of mechanisms for catalysis. forces in aqueous
solution. carbonyl- and acyl-group reactions. practical kinetics. more. 864pp.
Mi x Sli. 65460·5 Pa. $18.95

PROBABILITY: An Introduction. Samuel Goldberg. Excellent basic text covers set


theory. probability theory for finite sample spaces. binomial theorem. much more.
560 problems. Bibliographies. 522pp. 5" x Sli. 65252- 1 p.... $7.95

LIGHTNING. Martin A. Uman. Revised. updated edition of classic work on the


physics of Iighming. Phenomena. terminology. measurement. photography.
spectroscopy. thunder. more. Reviews recent research. Bibliography. Indices.
520pp. 5" x SIt 64575-4 Pa. $7.95

PROBABILITY THEORY: A Concise Course. Y.A. Rozanov. Highly readable.


self-contained introduction covers combination of events. dependent events.
Bernoulli trials. etc. Translation by Richard Silverman. 1 48pp. 5" x 81t
65544-9 Pa. $4.50

THE CEASELESS WIND: An Introduction to the Theory of Atmospheric Motion.


John A. Dullon. Acclaimed text integrates disciplines of mathematics and physics
for full understanding of dynamics of atmospheric motion. Over 400 problems.
Index. 97 illustrations. 640pp. 6 x 9. 65096-0 Pa. S 16.95

STATISTICS MANUAL. Edwin L. Crow. et al. Comprehensive. practical


collection of classical and modem methods prepared by U.S. Naval Ordnance Test
Station. Stress on use. Basics of statistics assumed. 2SSpp. 5" x 8li.
60599·X Pa. $6.00

WIND WAVES: Their Generation and Propagation on the Ocean Surface. Blair
Kinsman. Classic of oceanography offers detailed discussion of stochastic processes
and power spectral analysis that revolutionized ocean wave theory. Rigorous. I ucid.
676pp. 5" x 81S. 64652- 1 Pa. $ 1 4.95

STATISTICAL METHOD FROM THE VIEWPOINT OF QUALITY CON­


TROL. Walter A. Shewhart. Important text explains regulation of variables. uses
of statistical control to achieve quality control in industry. agriculture. other areas.
192pp. � x 8li. 65252-7 Pa. $6.00

THE INTERPRETATION OF GEOLOGICAL PHASE DIAGRAMS. Ernest G.


Ehlers. Clear. concise text emphasizes diagrams of systems under nuid or
containing pressure; also coverage of complex binary systems. hydrothermal
melting. more. 288pp. 6li x 91t 65589-7 Pa. $8.95

STATISTICAL ADJUSTMENT OF DATA. W. Edwards Deming. Introduction to


basic concepts of statistics. curve filling. least squares solution. conditions without
parameter. conditions containing parameters. 26 exercises worked out. 271pp.
5" x SIt 646S5·S Pa. $7.95
CATALOG OF DOVER nOOKS

CHALLE:-<GING MATI-IEIIIATICAL PROBLEIIIS WITI-I ELBI I'.. N TARY


SOLUTIONS, A.III. y:,glom allll l,,\I. Yaglom, O"er 170 ch:,llenging pwblcmsorl
pJOb�hililY lheory. comhin�lOri�1 �nal)'$is, poinlS and lines, lopology, COlWtK
polH�nm, m:lny Olher IOpin. Solulions. TOlal of H5pp_ 5S. x 81i. Two-'·ol. .l-el.
Vol. I 65536·9 1':1. $:
, .95
Vol. J[ 65537·7 I':t. $5.9:'

FWI'Y CHALLENGING PROBLEMS IN PROBABILITY WITH SOLU·


TIONS, Frederick Mosteller. Remarkable puulcls, gr.lded in difficulty, ilIuStr.lle
eknu:nWry and adl'anced aspects of probahililY. J)el:li1ed SOlulions. 8Spp. 5S K 8!-i.
fi5:i!i5·2 Pa. $�.95

EXPERIMENTS IN TOPOLOGY, Slephen Barr. Classic, lil'ely expl�nalion of


olle of Ihe b)'ways of malhema,ics. Klein bOllles, Mocbim sllillS, projecli"" planes,
10:']) colorillg, IlrobJem of lh" Koenigsberg IJlidgcs. much mu,c. ,lcscriht-d Wilh
daril)' and wit. ·13 figures. 2 l Opp. !is x 8ij. 25933·1 I'�, .1-1,95

RELATIVITY IN ILLUSTRATIONS. Jacob T. Schwam. Clear nOIHt'Chnical


lrealmelH makes It!:llil'ilY mort ac«'S�iblc ,han cI'cr beforc. O"cr ij() drawings
iJluSIrOltc concepls murcciearly than tC);1 alone. Only high school gcometry net-ded.
Bibliography. 12Spp, 6\; x 9\!. 25965·X I'a. $5.95

AN INTRODUCnO:-< TO ORDINARY DIFFERENTIAL EQUATIONS, £:111


A. Coddint:lon. A lilUrough and !>),stelllalic lirs! cours<; in elemCllIary diflclenti:ll
eiJuatiom for umlcrgradualcs in ,"a,hCll1;llics and science. wi! h mall)' eKctdses alld
ptoblcms (wilh answers). IndcK. 301pp, 5;" x 8"i. .�J.l2·9 1'3, $7.9.'i
6;

FOURIER SERIES AND ORTHOGONAL FUNCTIONS, I-iarry F. J)�l'is. An


incisil'e lexl combining 'heory aud practical cKamplc 10 inuodnCt' Fourier s<;ries,
or1hogonal lunct ions and applications of ,he Fourier melhod to boundarY'I'll"c
problems. 570 cKClciscs. ,\nswCIs and 1101"5. ·116J>Jl. 5S x Sli. 6:'973·9 J'a. $8.9.'i

THE THOERY OF IIRANCHING PROCESSES, Theodore t:. lIarris. Fi'SI


sysl�m;llic, cOlllprel,cnsil'e IleatlllCnI of branching (i.e. nluhipliGllil'e) processcs
ami their :'ppli(:aliom. Gahou·W:uson model. /II;ukol' hr:lllching proccs!>t."5,
clCCllOII·pholOll cucade, man)' ocher IOpics. RigOious proofs. lIibliogr:lphy.
2·IOPI'· [,lO x 8\; 65952·6 1':1.. $6.9:'

AN INTRODUCnON TO ALGEBRAIC STRUcrURES, Joseph Lmdin.


Sup",b .l-elf·r.omaiut'(l teXl covers "abslr:lcl ;tlgehr.j' · : SCIS :nul numbcrs, Iht'OI)' of
gTOUpS, theory of rings, much more. Numcrous well·chosen cK:unples, exercises.
2·17pp. !ili x 8">, 659·10·2 1'3. $6.9�

GAMES AND J)ECISIONS: InlroduClion aud Critical SUrVC)', R. Duncan Luce


aud I-Ioward Raiff:l. Superb nOIHechniGlI introduction 10 game 'henr)", primarily
applied 10 social !dences. Utility Ihenry, lno·sum gamcs, n·pt:rson gamcs,
dl'Cision'making, much 1I100e. lIibliograph)'. 50<Jpp. !is x Sli, 6.'i9'13·7 I'a. $10.95

I'''r�• ...b,UI IO rI,,"'g� w,th""1 '101l(�.


Avai1ahl�:11 )'0\1' hook d<:.lel 01 ..-rit< for II"" Malhemalic> and Scient:( CIf;,Jog I<> I)"pl. Gl.
I)(,.-,:r Pl1bliauiom, Inc., 31 East 2ndSt., M'nffl13. N. �'. 11501. Do,·�, p"bli.hn mOle Ihall 17�
books �,,<h l.-:tl Oil scien�r, deme""'I)' �"d �d,'anced malhern:"'a.. biology, """ie. :trl, I;'e"lI)"
I" "o'l',
SO(";:,J Kiw«"S and olher areU.

Anda mungkin juga menyukai