418

The Classical Fields

Embed Size (px)

Citation preview

Page 1: The Classical Fields
Page 2: The Classical Fields

ENCYCLOPEDIA OF MATHEMATICS AND ITS APPLICATIONS

FOUNDED BY G.-C. ROTA

Editorial BoardP. Flajolet, M. Ismail, E. Lutwak

Volume 112

The Classical Fields:Structural Features of the

Real and Rational Numbers

Page 3: The Classical Fields

ENCYCLOPEDIA OF MATHEMATICS AND ITS APPLICATIONS

FOUNDING EDITOR G.-C. ROTAEditorial BoardP. Flajolet, M. Ismail, E. Lutwak

40 N. White (ed.) Matroid Applications41 S. Sakai Operator Algebras in Dynamical Systems42 W. Hodges Basic Model Theory43 H. Stahl and V. Totik General Orthogonal Polynomials45 G. Da Prato and J. Zabczyk Stochastic Equations in Infinite Dimensions46 A. Bjorner et al. Oriented Matroids47 G. Edgar and L. Sucheston Stopping Times and Directed Processes48 C. Sims Computation with Finitely Presented Groups49 T. Palmer Banach Algebras and the General Theory of *-Algebras I50 F. Borceux Handbook of Categorical Algebra I51 F. Borceux Handbook of Categorical Algebra II52 F. Borceux Handbook of Categorical Algebra III53 V. F. Kolchin Random Graphs54 A. Katok and B. Hasselblatt Introduction to the Modern Theory of Dynamical Systems55 V. N. Sachkov Combinatorial Methods in Discrete Mathematics56 V. N. Sachkov Probabilistic Methods in Discrete Mathematics57 P. M. Cohn Skew Fields58 R. Gardner Geometric Tomography59 G. A. Baker, Jr., and P. Graves-Morris Pade Approximants, 2nd edn60 J. Krajicek Bounded Arithmetic, Propositional Logic, and Complexity Theory61 H. Groemer Geometric Applications of Fourier Series and Spherical Harmonics62 H. O. Fattorini Infinite Dimensional Optimization and Control Theory63 A. C. Thompson Minkowski Geometry64 R. B. Bapat and T. E. S. Raghavan Nonnegative Matrices with Applications65 K. Engel Sperner Theory66 D. Cvetkovic, P. Rowlinson and S. Simic Eigenspaces of Graphs67 F. Bergeron, G. Labelle and P. Leroux Combinatorial Species and Tree-Like Structures68 R. Goodman and N. Wallach Representations and Invariants of the Classical Groups69 T. Beth, D. Jungnickel, and H. Lenz Design Theory I, 2nd edn70 A. Pietsch and J. Wenzel Orthonormal Systems for Banach Space Geometry71 G. E. Andrews, R. Askey and R. Roy Special Functions72 R. Ticciati Quantum Field Theory for Mathematicians73 M. Stern Semimodular Lattices74 I. Lasiecka and R. Triggiani Control Theory for Partial Differential Equations I75 I. Lasiecka and R. Triggiani Control Theory for Partial Differential Equations II76 A. A. Ivanov Geometry of Sporadic Groups I77 A. Schinzel Polymomials with Special Regard to Reducibility78 H. Lenz, T. Beth, and D. Jungnickel Design Theory II, 2nd edn79 T. Palmer Banach Algebras and the General Theory of *-Algebras II80 O. Stormark Lie’s Structural Approach to PDE Systems81 C. F. Dunkl and Y. Xu Orthogonal Polynomials of Several Variables82 J. P. Mayberry The Foundations of Mathematics in the Theory of Sets83 C. Foias et al. Navier–Stokes Equations and Turbulence84 B. Polster and G. Steinke Geometries on Surfaces85 R. B. Paris and D. Kaminski Asymptotics and Mellin–Barnes Integrals86 R. McEliece The Theory of Information and Coding, 2nd edn87 B. Magurn Algebraic Introduction to K-Theory88 T. Mora Solving Polynomial Equation Systems I89 K. Bichteler Stochastic Integration with Jumps90 M. Lothaire Algebraic Combinatorics on Words91 A. A. Ivanov and S. V. Shpectorov Geometry of Sporadic Groups II92 P. McMullen and E. Schulte Abstract Regular Polytopes93 G. Gierz et al. Continuous Lattices and Domains94 S. Finch Mathematical Constants95 Y. Jabri The Mountain Pass Theorem96 G. Gasper and M. Rahman Basic Hypergeometric Series, 2nd edn97 M. C. Pedicchio and W. Tholen (eds.) Categorical Foundations98 M. Ismail Classical and Quantum Orthogonal Polynomials in One Variable99 T. Mora Solving Polynomial Equation Systems II

100 E. Olivieri and M. Eulalia Vares Large Deviations and Metastability102 L. W. Beineke et al. (eds.) Topics in Algebraic Graph Theory103 O. Staffans Well-Posed Linear Systems105 M. Lothaire Applied Combinatorics on Words106 A. Markoe Analytic Tomography107 P. A. Martin Multiple Scattering108 R. A. Brualdi Combinatorial Matrix Classes110 M.-J. Lai, L. L.Schumaker Spline Functions on Triangulations111 R. T. Curtis Symmetric Generation of Groups

Page 4: The Classical Fields

ENCYCLOPEDIA OF MATHEMATICS AND ITS APPLICATIONS

The Classical FieldsStructural Features of the Real and Rational Numbers

H. SALZMANNUniversity of Tubingen

T. GRUNDHOFERUniversity of Wurzburg

H. HAHLUniversity of Stuttgart

R. LOWENTechnical University of Braunschweig

Page 5: The Classical Fields

cambridge university pressCambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo

Cambridge University PressThe Edinburgh Building, Cambridge CB2 8RU, UK

Published in the United States of America by Cambridge University Press,New York

www.cambridge.orgInformation on this title: www.cambridge.org/9780521865166

C© H. Salzmann, T. Grundhofer, H. Hahl and R. Lowen 2007

This publication is in copyright. Subject to statutory exceptionand to the provisions of relevant collective licensing agreements,

no reproduction of any part may take place withoutthe written permission of Cambridge University Press.

First published 2007

Printed in the United Kingdom at the University Press, Cambridge

A catalog record for this publication is available from the British Library

ISBN 978-0-521-86516-6 hardback

Cambridge University Press has no responsibility for the persistence or accuracy of URLs for externalor third-party internet websites referred to in this publication, and does not guarantee that any content

on such websites is, or will remain, accurate or appropriate.

Page 6: The Classical Fields

Contents

Preface page ixNotation xiii

1 Real numbers 11 The additive group of real numbers 22 The multiplication of real numbers, with a digression on fields 143 The real numbers as an ordered set 224 Continued fractions 285 The real numbers as a topological space 32

Characterizing the real line, the arc, and the circle 34Independence of characteristic properties 44Subspaces and continuous images of the real line 51Homeomorphisms of the real line 57Weird topologies on the real line 64

6 The real numbers as a field 707 The real numbers as an ordered group 758 The real numbers as a topological group 81

Subgroups and quotients 83Characterizations 86A counter-example 93Automorphisms and endomorphisms 95Groups having an endomorphism field 96

9 Multiplication and topology of the real numbers 10010 The real numbers as a measure space 10411 The real numbers as an ordered field 11212 Formally real and real closed fields 12213 The real numbers as a topological field 13514 The complex numbers 140

v

Page 7: The Classical Fields

vi Contents

2 Non-standard numbers 15421 Ultraproducts 15422 Non-standard rationals 15823 A construction of the real numbers 15924 Non-standard reals 162

Ordering and topology 164η1-fields 166

25 Continuity and convergence 17026 Topology of the real numbers in non-standard terms 17327 Differentiation 17528 Planes and fields 177

3 Rational numbers 17931 The additive group of the rational numbers 17932 The multiplication of the rational numbers 18533 Ordering and topology of the rational numbers 19334 The rational numbers as a field 20735 Ordered groups of rational numbers 21636 Addition and topologies of the rational numbers 22137 Multiplication and topologies of the rational numbers 228

4 Completion 23541 Completion of chains 23642 Completion of ordered groups and fields 23943 Completion of topological abelian groups 24844 Completion of topological rings and fields 264

5 The p-adic numbers 27851 The field of p-adic numbers 27952 The additive group of p-adic numbers 28553 The multiplicative group of p-adic numbers 29254 Squares of p-adic numbers and quadratic forms 29555 Absolute values 30056 Valuations 30657 Topologies of valuation type 31658 Local fields and locally compact fields 322

Page 8: The Classical Fields

Contents vii

6 Appendix 33561 Ordinals and cardinals 33562 Topological groups 34063 Locally compact abelian groups and Pontryagin duality 34464 Fields 350

Hints and solutions 360References 383Index 399

Page 9: The Classical Fields
Page 10: The Classical Fields

Preface

The rational numbers, the real numbers, the complex numbers and thep-adic numbers are classical fields. These number systems are the topicof this book.

The real numbers, which are basic and indispensable for most partsof mathematics, comprise several rich and intimately interwoven struc-tures, namely the algebraic structure as a field, the topological structureand the ordering structure. Each of these structures, as well as theirparticular blend, is beautifully adapted to the intended use of numbers(for counting, computing, taking measurements, comparing sizes andmodelling physical space and time). It is the purpose of this book toconsider these structures separately, and to analyse the interaction andthe interdependencies between these structures. The real numbers arecharacterized in various categories by simple abstract properties. Eachof these characterization results is a possible answer to the question:why exactly are the real numbers so fundamentally important?

The ordering and the topology of the real numbers are rooted deeply inour geometric intuition about points on a line. The algebraic operationsof addition and multiplication describe the isometries and the similari-ties of the one-dimensional geometry of a line. (In fact, one-dimensionalgeometry becomes interesting only by imposing some additional struc-ture on the set of points of the only line.)

Apart from the real numbers, we also treat the rational numbers (inChapter 3) and the p-adic numbers (in Chapter 5). The complex num-bers are considered in Section 14 (to some extent also in Section 13),and Chapter 2 deals with non-standard numbers. We study the struc-tural components of each of these fields and their interactions; we alsodescribe the pertaining automorphism groups and typical substructuresand quotients.

ix

Page 11: The Classical Fields

x Preface

Of course, also finite fields are classical number systems. However,finite fields are purely algebraic objects, they have no interesting orderingor topology, and their algebraic features are closely related to numbertheory. Therefore finite fields appear only incidentally in this book.

∗ ∗ ∗The first and longest chapter considers the field R of real numbers.

We study the additive and the multiplicative group of R, and then R asan ordered set, as a topological space, as a measure space, as an abstractfield and as a topological field. The additive group (R, +) is consideredas an ordered group and as a topological group. Algebraic properties ofthe field R lead to the Artin–Schreier theory of formally real fields.

The last section of Chapter 1 treats the complex numbers C; manystructural features of C can be inferred from the description of C =R(√−1) as a quadratic field extension of R. However, the existence

of discontinuous field automorphisms and of unexpected subfields is apeculiar property of C.

According to Pontryagin, R and C are the only topological fields thatare locally compact and connected; see Theorem 13.8.

It is not our main task to construct the real numbers R, we rather takethem for granted (constructions can be found in the books mentionedat the end of this preface). Still, we describe constructions of R inSection 23 (by means of an ultrapower of the field Q of rational numbers)and in 42.11 and 44.11.

Non-standard numbers are the theme of Chapter 2. These numbersystems can be constructed easily via ultrapowers. Contrasting R withits non-standard counterpart ∗R sheds additional light on the particularrole of R. The additive and the multiplicative groups of R and ∗R areisomorphic, and ∗R(

√−1) ∼= C (see 24.2, 24.4 and 24.6), but R and∗R are not isomorphic as fields, and R and ∗R are quite different astopological spaces. The natural embedding of R into ∗R leads to somebasic notions of non-standard analysis.

In Chapter 3 we treat the system Q of rational numbers in a simi-lar way as the real numbers. The different structural components of Qare less tightly related; in particular, the additive group and the mul-tiplicative group of Q are quite different. As Q is the field of fractionsof the ring Z of integers, number theory plays an important role in thischapter. In many respects R is simpler than Q; for example, R has onlytwo square classes and an easy theory of quadratic forms, whereas Q hasinfinitely many square classes and a rich theory of quadratic forms (seeCassels 1978). Moreover, the natural topology of R is locally compact

Page 12: The Classical Fields

Preface xi

and connected, in contrast to the topology of Q. This explains why inthis book we treat first R and then Q, despite the fact that R can beobtained by completing Q.

A field is said to be complete, if its additive group is complete withrespect to a given ordering or topology. In Chapter 4 we discuss rele-vant completion procedures. First we complete chains, and then orderedgroups and fields. Next we construct the (essentially unique) comple-tion of a topological abelian group. A complete ordered group is alsocomplete in the topology determined by the ordering (see 43.10). Theresults are finally applied to topological rings and fields.

In Chapter 5 we deal with the p-adic numbers Qp as relatives of thereal numbers. Indeed, Qp can be obtained by completing Q with re-spect to the p-adic metric; this metric reflects the divisibility by powersof the prime p. As a consequence, the p-adic topology has a rather al-gebraic flavour; note that the non-zero ideals of the ring Zp ⊂ Qp ofp-adic integers form a neighbourhood basis at 0 for the topology of Qp.This topology is locally compact and totally disconnected. We con-sider the additive and the multiplicative group of Qp, and we study thesquares of Qp. The field Qp cannot be made into an ordered field (com-pare 54.2). Like R, the field Qp admits no automorphism except theidentity (see 53.5).

The properties of Qp are placed in a more general context by con-sidering absolute values, valuations and topologies of valuation type inSections 55, 56, 57. The last section of Chapter 5 deals with field ex-tensions of Qp and with the classification of all locally compact (skew)fields. We prove that R and Qp are the only non-discrete locally compactfields that contain Q as a dense subfield (58.7).

Note that R and Qp are encoded in the additive group (Q, +), andhence in the semigroup (N, +) of positive integers: the field Q is theendomorphism ring of (Q, +) (see 8.28), and R and the fields Qp arethe completions of Q with respect to the absolute values of Q (com-pare 44.11).

In an Appendix we collect some facts on ordinal and cardinal numbersand on topological groups, we summarize the duality theory of locallycompact abelian groups, and we present basic facts and constructions offield theory.

Most sections end with a few exercises, of different character anddegree of difficulty. The chapter ‘Hints and solutions’ provides a solutionor at least a clue for each exercise.

∗ ∗ ∗

Page 13: The Classical Fields

xii Preface

There is a vast literature on number theory, and there exist manybooks which deal with the real numbers and the rational numbers. Sev-eral of these texts explain the successive construction of the number sys-tems N, Z, Q, R and C; typical examples are Dedekind 1872, Landau

1930, Cohen–Ehrlich 1963, Feferman 1964. The historical develop-ment is presented in Flegg 1983, Ehrlich 1994 and Lopez Pellicer

1994. The three volumes by Felscher 1978/79 on ‘Naive Mengen undabstrakte Zahlen’ emphasize logical and set-theoretic aspects. The clas-sical division algebras H (Hamilton’s quaternions) and O (octonions)are treated in Salzmann et al. 1995 Chapter 1, Ebbinghaus et al.1991 Part B; see also Conway–Smith 2003, Baez 2002, Ward 1997.None of these books has much overlap with the present text.

Our book is based on lectures given by H. Salzmann in Tubingen in1971/72 and on a two-volume set of lecture notes (prepared by R. Lowenand H. Hahl) with the title ‘Zahlbereiche’. These lecture notes had beenavailable for a short while in mimeographed form: Salzmann 1971 andSalzmann 1973.

We would like to thank Nils Rosehr for technical support, and JoachimGrater for helpful discussions. We are grateful to the friendly staff ofCambridge University Press for their professional help and advice inpublishing this book.

The authors

Page 14: The Classical Fields

Notation

As usual, N, Z, Q, R, C denote the natural, integer, rational, real andcomplex numbers, respectively. By convention, 0 /∈ N and N0 := N∪{0}.

We use := for equality by definition and � for equivalence by defini-tion. The symbols ∧ and ∨ are the logical connectives ‘and’ and ‘or’.

We write A ∼= B if two structures A and B are isomorphic, and A < B

if A is a proper substructure of B.The notation X ≈ Y means that the topological spaces X and Y are

homeomorphic (compare 5.51).

Fq finite field of order q

P prime numbersRpos positive real numbers (as a multiplicative group)Ralg real algebraic numbersSn sphere of dimension n

T := R/Z torus (1.20)L, L+ long line, long ray (5.25)C Cantor set (5.35, as a topological space)H quaternions (Section 13, Exercise 6, and 34.17)Qp p-adic numbers (44.11 and Chapter 5)Zp p-adic integers (51.6)

a | b a divides b

gcd(a, b) greatest common divisor of a, b

[ c0; c1, c2, . . . ] continued fraction (4.1)

2S power set of S⋃S union {x | ∃S∈S : x ∈ S } of a system S of sets⋂S intersection {x | ∀S∈S : x ∈ S } of a system S of sets

xiii

Page 15: The Classical Fields

xiv Notation

id identity mappingim imageY X set {f | f : X → Y } of mappingsFix set of all fixed elementssupp support (5.61, 64.22)×i Si Cartesian product of sets (with structure) Si

O ordinal numbers (Section 61)cardS cardinality of a set S (61.8)ℵ0 cardinality of Nℵ = 2ℵ0 cardinality of R (1.10)

Cn cyclic group of order n

Cp∞ Prufer group (1.26)Sym S symmetric (permutation) group of the set S

GLnF general linear group of Fn

PGL2F projective quotient of GL2F (11.16, 64.19)H = H(R) homeomorphism group of R (5.51)H(Q) homeomorphism group of Q (33.12)AutX automorphism group of X

Autc X group of continuous automorphisms of X

EndX endomorphisms of X

Endc X continuous endomorphisms of X

Hom(X, Y ) homomorphisms X → Y

Cs centralizerA∗ character group of A (63.1)⊕

i Gi direct sum of groups Gi (1.16)

E|F field extension F ⊆ E (64.1)[E : F ] degree of E|F (64.1)F+ additive group of F

F× multiplicative group (of units) of F

F � set {x2 | 0 �= x ∈ F } of squares of F×

F � algebraic closure of F (64.13)char(F ) characteristic of F (64.4)trdeg transcendency degree (64.20)tr traceGalF E, AutF E Galois group, relative automorphism group (64.17)F [t] polynomial ringF (t) field of fractions of F [t]

Page 16: The Classical Fields

Notation xv

F [[t]] ring of power series (64.22)F ((t)) field of Laurent series (64.23)F ((t1/∞)) field of Puiseux series (64.24)F ((Γ)), F ((Γ))1 fields of Hahn power series (64.25)

lim B limit of a filterbase B (43.2)C minimal concentrated filter (43.15)

Non-standard objects (Chapter 2):

SΨ ultrapower (21.4)∗Q non-standard rationals (Section 22)∗R non-standard reals (Section 24)T ∗R as a topological space (Section 24)hΨ, ∗h extension of a map h (21.10, 25.0)◦a standard part of a (23.9)x ≈ y y − x is infinitely small (23.9, Sections 25–27)

Page 17: The Classical Fields
Page 18: The Classical Fields

1

Real numbers

This chapter is devoted to various aspects of the structure of R, the fieldof real numbers. Since we do not intend to give a detailed account ofa construction of the real numbers from the very beginning, we need toclarify the basis of our subsequent arguments. What we shall assumeabout the real numbers is that they form an ordered field whose order-ing is complete, in the sense that every non-empty bounded set of realnumbers has a least upper bound. All these notions will be explainedin due course, but presumably they are familiar to most readers. It iswell known and will be proved in Section 11 that the properties justmentioned characterize the field of real numbers.

Historically, a satisfactory theory of the real numbers was obtainedonly at the end of the nineteenth century by work of Weierstraß, Can-tor and Dedekind (see Flegg 1983, Ehrlich 1994 and Lopez Pel-

licer 1994). Starting from the rational numbers, they used differentapproaches, namely, Cauchy sequences on the one hand and Dedekindcuts on the other.

In Sections 42 and 44, we shall actually show how to obtain the realnumbers from the rational numbers via completion. Another construc-tion in the context of non-standard real numbers will be given in Sec-tion 23. We mention also the approach of Conway 1976, whose ‘surrealnumbers’ go beyond non-standard numbers. These ideas were carriedfurther by Gonshor 1986, Alling 1987; see also Ehrlich 1994, 2001and Dales–Woodin 1996.

Several methods have been proposed for constructing R directly fromthe ring Z of integers, without using the rational numbers as an inter-mediate step; compare Section 6, Exercise 2 (which is related to Faltin

et al. 1975; see also Section 51, Exercise 3, for the p-adic analogue). de

Bruijn 1976 defines the ordered additive group of real numbers via cer-

1

Page 19: The Classical Fields

2 Real numbers

tain mappings f : Z→ {0, 1, . . . , b−1}, with the idea that a non-negativereal number is represented by

∑n∈Z f(n)b−n. A’Campo 2003 considers

all maps f : Z → Z which are ‘slopes’ (or ‘near-endomorphisms’) inthe sense that {f(x + y) − f(x) − f(y) | x, y ∈ Z} is a finite set, andhe constructs R by identifying two slopes f, g if f − g has finite image;see also Arthan 2004, Grundhofer 2005 for details pertaining to thisconstruction.

1 The additive group of real numbers

The first feature of the field R to be examined is its additive group(R, +). The following is the essential fact about this group (which actu-ally characterizes it): (R, +) is a vector space over the rational numbers,and the dimension of this vector space is given by the cardinality of thereal numbers.

We shall derive various consequences from this rational vector spacestructure; in particular, we consider subgroups and quotient groups andcharacterize them as groups where possible. The axiom of choice will beused in many places in this section because we rely on the existence ofbases for infinite dimensional vector spaces.

We do not at this point go into the obvious question as to what canbe said about the additive structure of the rational numbers themselves.This will be deferred to Section 31.

1.1 The additive group of real numbers From the construction ofthe real numbers, we take the following facts for granted.(a) The real numbers under addition form an abelian group denoted

(R, +), or briefly R+, with neutral element 0.(b) By repeated addition of the multiplicative unit 1 we exhaust the set

N = {1, 2, 3 . . . } of natural numbers; together with their additiveinverses and 0 they form the infinite cyclic group of integers, whichis a subgroup Z+ ≤ R+.

Here, the word ‘cyclic’ means ‘generated by a single element’, namely,by 1 (or by −1). In other words, Z+ is the smallest subgroup of R+

containing 1. That this subgroup is infinite is a consequence of theordering: from 0 < 1 one obtains, by induction, that 0 < n < n + 1 forall natural numbers n.

We start our investigation of the group R+ by examining its sub-groups, starting with the smallest possible ones. First, we turn to ageneral consideration.

Page 20: The Classical Fields

1 The additive group of real numbers 3

1.2 Cyclic subgroups of arbitrary groups In an additively writtengroup (G, +), we consider the multiples ng := g + · · ·+ g of an elementg ∈ G, where n ∈ N is the number of summands. In addition, we set0g := 0 and zg := (−z)(−g) for negative integers z. The order #g of g

is, by definition, the smallest natural number k such that kg = 0; we set#g =∞ if no such number k exists.

Fix g ∈ G. The mapping defined by ϕ(z) := zg is a homomorphismϕ : (Z, +) → (G, +) which maps Z onto the smallest subgroup 〈g〉 con-taining g. The kernel kerϕ equals kZ if k = #g is finite, and kerϕ = {0}otherwise. It follows that 〈g〉 is isomorphic to the factor group Z/ ker ϕ,and thus to the cyclic group Ck if k ∈ N and to Z+ otherwise. Clearly,all homomorphisms Z→ G arise in this manner (just set g := ϕ(1)).

The group G is said to be torsion free if all its elements except 0have infinite order; equivalently, if all its non-trivial cyclic subgroupsare infinite or if all non-trivial homomorphisms Z→ G are injective.

1.3 Theorem: Cyclic subgroups of R+ The group homomorphisms

Z+ → R+ are precisely the maps ϕs : z �→ zs for arbitrary s ∈ R.

Every non-zero element r ∈ R generates an infinite cyclic subgroup

〈r〉 = Zr ∼= Z, and the group R+ is torsion free.

Proof It suffices to remark that, by virtue of the distributive law, themultiple zs can also be obtained as z · s, using the multiplication of realnumbers. Now the distributive law shows that ϕs is a homomorphism;if r �= 0, then ϕr is injective by the absence of zero divisors. Insteadof these arguments, one could use the same reasoning (based on theordering) as in 1.1. �

1.4 Theorem Every non-trivial subgroup H ≤ R+ is either cyclic or

dense in R.

Proof Let r ∈ R be the infimum (the greatest lower bound) of the set{h ∈ H | h > 0}. If r �= 0 and r ∈ H, then the cyclic group Zr coincideswith H. Indeed, every g ∈ H belongs to the interval [zr, (z + 1)r[ forsome z ∈ Z, which implies that g− zr ∈ [0, r[∩H and g = zr. (We haveused, somewhat informally, the fact that R+ is an Archimedean orderedgroup; compare 7.4 and 7.5.)

On the other hand, if r = 0 or r /∈ H, then r is a cluster point ofH, and H contains pairs with arbitrarily small differences. Since H isa group, it contains those differences as elements. Now if g ∈ ]0, ε[ ∩H,then every closed interval of length ε contains some integer multiple zg,

Page 21: The Classical Fields

4 Real numbers

z ∈ Z, and H is dense. (This argument used the Archimedean propertyagain.) �

Having treated completely the subgroups generated by a single ele-ment, we turn now to subgroups generated by pairs of elements. As apreparation, we need the following simple result of number theory.

1.5 Bezout’s Theorem The greatest common divisor gcd(m,n) of

two integers is an integral linear combination of m and n, i.e., there are

integers x, y such that gcd(m,n) = xm + yn. In particular, if m,n are

relatively prime, then 1 is a linear combination: 1 = xm + yn.

Proof The subgroup I = {xm + yn | x, y ∈ Z} ≤ Z+ is generated byany element d ∈ I � {0} of smallest absolute value. This is shown byapplying the Euclidean algorithm: an element z ∈ I can be written asz = bd + r with b ∈ Z and |r| < |d|; it follows from the equation thatr ∈ I, and we must have r = 0 by the definition of d.

Now if I = 〈d〉, then d divides m and n and is a linear combination ofthe two; hence every common divisor of m,n divides d. — We remarkthat d, x and y can be computed explicitly by making full use of theEuclidean algorithm. This is needed, for example, in order to computemultiplicative inverses in the field Z/pZ, where p is a prime. �

Now we apply this fact to subgroups of R.

1.6 Theorem (a) A subgroup 〈a, b〉 ≤ R+ generated by two non-zero

elements a, b is cyclic if, and only if, the quotient a/b is a rational

number. More precisely, if a/b = m/n with m,n relatively prime,

then 〈a, b〉 = 〈b/n〉.(b) If a/b is irrational, then 〈a, b〉 is dense in R.

(c) The additive group Q+ of rational numbers is locally cyclic, i.e., any

finite subset generates a cyclic subgroup.

Proof (a) If a/b = m/n, then xa+yb = b(xm+yn)/n for x, y ∈ Z. FromBezout’s Theorem 1.5 it follows that b/n ∈ 〈a, b〉. On the other hand,both b and a = ba/b = bm/n are integer multiples of b/n. Conversely,if a and b belong to a cyclic group 〈c〉, then a = xc and b = yc for someintegers x, y, and a/b = x/y.

(b) follows from (a) together with 1.4, and (c) is obtained by repeatedapplication of (a). �

There is a more general (and less easy) result behind the densityassertion 1.6b. See 5.69 for a statement and proof of this theorem dueto Kronecker.

Page 22: The Classical Fields

1 The additive group of real numbers 5

Our next aim is the characterization of R+ given in 1.14. It uses thenotion of divisible group, which we introduce first.

1.7 Definition: Divisible groups A group (G, +) is said to be divis-ible if for every g ∈ G and every n ∈ N there is h ∈ G such that nh = g.Here, the integer multiple nh is taken in the sense of 1.2. An abeliangroup that is both divisible and torsion free is uniquely divisible, i.e.,the element h is uniquely determined by g and n. Indeed, g = nh = nh′

implies n(h − h′) = 0, and then h − h′ = 0 because G is torsion free.The unique h satisfying nh = g will then be denoted g/n.

In particular, we have:

1.8 Theorem The group R+ is uniquely divisible. �

1.9 Theorem A uniquely divisible abelian group G carries a unique

structure as a rational vector space.

Proof It follows from the vector space axioms that multiplication by ascalar z ∈ Z is the one defined in 1.2, and then multiplication by thescalar 1/n is the operation introduced in 1.7. This proves uniqueness.On the other hand, we can always introduce on G the structure of avector space over Q in this way. �

The structure of uniquely divisible non-abelian groups is more com-plicated: Guba 1986 shows that there exists a group of this kind thatis generated by two elements.

Any bijection between given bases of two vector spaces over the samefield extends to an isomorphism between the spaces, thus a vector spaceover a given field is determined up to isomorphism by the cardinality ofa basis, i.e., by its (possibly transfinite) dimension; compare Exercise 1.Hence we can characterize the uniquely divisible group R+ if we deter-mine its dimension as a vector space over the rationals. Before we cando this, we need to determine the cardinality of R itself.

1.10 Theorem The set R has cardinality ℵ := card R = 2ℵ0 > ℵ0.

Proof We use the fact that every real number between 0 and 1 has aunique binary expansion

∑n∈N cn2−n with cn ∈ {0, 1} and cn = 0 for

infinitely many n. By 61.14, the last condition excludes only countablymany coefficient sequences (cn)n∈N, and there remain 2ℵ0 admissible se-quences. Thus, card [0, 1[ = 2ℵ0 and, since R decomposes into countablymany intervals [m,m + 1[, it follows that card R = ℵ0 · 2ℵ0 = 2ℵ0 ; see61.12. According to 61.11, we have 2ℵ0 > ℵ0.

Page 23: The Classical Fields

6 Real numbers

There is a slightly faster way to establish the equation card[0, 1] = 2ℵ0

using the Cantor set C = {∑∞ν=1 cν3−ν | cν ∈ {0, 2}}; compare 5.35ff.

Indeed, C is a subset of the unit interval, and on the other hand, C mapsonto [0, 1] via

∑∞ν=1 cν3−ν �→∑∞

ν=1 cν2−ν−1. �

One naturally wonders what is the precise relationship between thetwo cardinalities card Q = ℵ0 and card R = ℵ; are there any othercardinalities in between or not? The question thus raised is known asthe continuum problem; see 61.17 for a brief introduction.

1.11 The real numbers as a rational vector space The vectorspace structure on R+ that we determined in 1.9 can be described moreeasily. In fact, the product qr of a scalar q ∈ Q and a vector r ∈ R isjust their product as real numbers. This follows from uniqueness of thevector space structure by observing that multiplication in R does definesuch a structure.

What is the dimension of this vector space? The cardinality of abasis B ⊆ R cannot exceed that of the space R itself. To obtain a lowerestimate for card B, we count the finite rational linear combinations of B.Every linear combination is determined by a finite subset of Q×B. By61.14, the set of all these subsets has the same cardinality as Q×B itself.Thus we have card B ≤ card R = ℵ ≤ card(Q × B) = max{ℵ0, cardB};for the last equality use 61.12. This proves the following.

1.12 Theorem Any basis of the vector space R over Q has the same

cardinality as R itself, that is, dimQ R = ℵ = card R. �

Incidentally, we have proved that the concept of transfinite dimensionis meaningful in the given situation, independently of Exercise 1. Ex-amining the proof, we see that countability of Q is essential. In fact wehave shown that a basis of any infinite-dimensional rational space V hasthe same cardinality as V itself. This is not true for real vector spaces.

1.13 Hamel bases A basis of the rational space R is usually referredto as a Hamel basis. No one has ever written down such a basis, norprobably ever will. Yet such bases have several applications. Theypermit, for example, the solution of the functional equation f(x + y) =f(x) + f(y); this is precisely what Hamel invented them for (Hamel

1905, compare also Aczel 1966). Moreover, Hamel bases allow one toconstruct subsets of R that behave strangely with respect to Lebesguemeasure; see 10.8ff. One should therefore keep in mind that the existenceof Hamel bases depends on the Axiom of Choice (AC); compare theintroduction to Section 61. A large proportion of the subsequent results

Page 24: The Classical Fields

1 The additive group of real numbers 7

in this section therefore need AC, namely 1.14, 1.15, 1.17, 1.19, 1.24. Inaddition, 1.23 uses AC (directly).

In spite of the elusiveness of Hamel bases, there are many ways ofconstructing large Q-linearly independent sets of real numbers. A simpleexample is the set of all logarithms of prime numbers (Exercise 6). Morerefined techniques yield sets that are uncountable (Brenner 1992). Thisis surpassed by von Neumann 1928 and Kneser 1960, who show thatcertain sets of cardinality ℵ are even algebraically independent over Q(as defined in 64.20). Kneser’s set consists of the numbers∑

n≥1 2−�nn+s�, 0 ≤ s < 1,

where �t� denotes the largest integer not exceeding t ∈ R. Other exam-ples of this kind are given by Durand 1975 and Elsner 2000.

There is a nice survey on this topic by Waldschmidt 1992, and thereare several other contributions by the same author. Finally, we mentionLaczkovich 1998, who shows that there are Q-linearly independentsubsets of R that are Gδ-sets (that is, intersections of countably manyopen sets).

Observe that the only information about R that we needed in orderto compute the dimension was the cardinality of R itself. Thus we haveproved the following.

1.14 Characterization Theorem An abelian group is isomorphic to

the additive group R+ of real numbers if, and only if, it is torsion free

and divisible and has the same cardinality as R. �

1.15 Consequences A few surprising (at first sight) consequencesof the characterization are worth pointing out. The additive group ofany finite dimensional real vector space, e.g., of Rn or of C, satisfiesthe conditions that characterize R+. Thus, all these vector groups areisomorphic. By contrast, two real vector spaces of different finite dimen-sions are not isomorphic, and their additive groups are not isomorphicas topological groups; indeed, every additive map between such vectorspaces is Q-linear (by 1.9), hence R-linear if it is continuous; alterna-tively, we could use Theorem 8.6, which implies that Rn contains closeddiscrete subgroups isomorphic to Zn, but no closed discrete subgroupisomorphic to Zn+1.

A Hamel basis has many subsets of the same cardinality, giving rise tomany vector subspaces that are isomorphic to R both as rational vector

Page 25: The Classical Fields

8 Real numbers

spaces and as additive groups. These subgroups may also be obtainedas factor groups; just factor out a complementary vector subspace.

Now we determine the cardinality of the set G of all subgroups of R+:the set 2B of all subsets of a Hamel basis B injects into G, because asubset C ⊆ B generates a vector subspace VC , and C = VC ∩B. Hencethe cardinality 2ℵ of 2B is a lower bound for cardG. But 2ℵ is also thecardinality of the set of all subsets of R, which is an upper bound forcardG. Thus we see that there are 2ℵ = 22ℵ0 distinct subgroups of R+.(This does not say anything about the number of isomorphism types ofsubgroups, which depends on how many cardinalities there are betweenℵ0 and ℵ; compare 61.17.)

We proceed to examine decompositions of R+ as a direct sum. Firstwe present the basic notions of direct sum and direct product.

1.16 Definition: Direct sums and products Given a family ofadditively written groups Gi (abelian or otherwise), indexed by a set I,we form a group

×i∈I Gi ,

called the direct product of the Gi, as follows: the elements of this groupare the indexed families (gi)i∈I such that gi ∈ Gi, and the group opera-tion is defined componentwise, i.e., (gi)i∈I + (hi)i∈I = (gi + hi)i∈I .

At the moment, we are more interested in the subgroup formed bythose families (gi)i∈I that satisfy gi = 0 with only finitely many excep-tions. This group is called the direct sum of the Gi and denoted⊕

i∈I Gi .

Of course, a difference between sums and products exists only if I isinfinite; compare, for example, 1.30. Note that every summand (orfactor) Gi is isomorphic to a subgroup of the direct sum or product,respectively. Likewise, the sum or product of the same Gi taken overany subset of the index set is contained in the total sum or product,respectively.

In the special case where Gi = G for all i, we refer to the directproduct as a power of G; its elements can be thought of as functionsI → G. We use the simplified notation

×i∈I G = GI and⊕

i∈I G = G(I) .

If V is any vector space over a field F and B ⊆ V is a basis, then everyelement of V has a unique representation as a finite linear combination

Page 26: The Classical Fields

1 The additive group of real numbers 9

v =∑

b∈B fbb, thus fb = 0 with finitely many exceptions. We can viewv as an element (fb)b∈B of a direct sum of copies of the additive groupF+, indexed by the set B. Therefore, we have an isomorphism

V + ∼=⊕b∈B F+b ,

where F+b = F+ for all b. In particular,

1.17 Theorem The group R+ is a direct sum of ℵ copies of Q+. �

The above decomposition cannot be refined any further, as the follow-ing theorem shows.

1.18 Theorem The group Q+ cannot be decomposed as a direct pro-

duct of two non-trivial subgroups G, H.

Proof Any two non-zero elements a/b ∈ G, c/d ∈ H have a non-zerocommon multiple cb(a/b) = ca = ad(c/d) ∈ G ∩ H, which is a contra-diction to G ∩H = {0}. �

1.19 Theorem The group R+ admits a decomposition as a direct

sum of indecomposable subgroups. This decomposition is unique up to

isomorphism.

Proof Existence is obtained from 1.17 and 1.18; it remains to proveuniqueness. Suppose we have two decompositions of the specified kind.Unique divisibility of R+ implies that every summand is uniquely di-visible; remember that addition is done componentwise. Hence, everysummand is a vector space over Q. Indecomposability implies that thesummands are in fact one-dimensional vector spaces. Choosing a non-zero element from each summand of one decomposition, we construct aHamel basis of R+. The two bases so obtained can be mapped onto eachother by a vector space isomorphism. This shows that the two directsums are isomorphic by an isomorphism that maps the summands of oneonto those of the other. �

The remainder of this section is devoted to the study of a prominentfactor group of R+, the torus group T = R+/Z. We shall return to this in5.16 and in Section 8. First we show why this group plays an importantrole both in analysis and in geometry.

1.20 Theorem The following three groups are isomorphic:

(a) The factor group T = R+/Z(b) The multiplicative group S1 of complex numbers of absolute value 1(c) The group SO2R of rotations of the plane R2.

Page 27: The Classical Fields

10 Real numbers

Proof The exponential law ez+w = ezew for the complex exponentialfunction implies that the map ϕ : t �→ e2πit is a group homomorphism ofR+ into the multiplicative group of complex numbers. From |ϕ(t)|2 =ϕ(t)ϕ(t) = ϕ(t)ϕ(−t) = 1 we see that ϕ(t) ∈ S1. The Euler relationeix = cos x + i sinx implies that ϕ maps R onto S1 and that kerϕ = Z.It follows that R+/Z ∼= S1.

By definition, SO2R consists of the real orthogonal 2 × 2 matrices ofdeterminant 1 (or of the linear maps of R2 defined by these matrices).For z ∈ S1, let γ(z) be the R-linear map C → C defined by w �→ zw.This map preserves the norm |w| and, hence, the scalar product. Thematrix of γ(eix) with respect to the basis 1, i of C is(

cos x − sinx

sinx cos x

);

its determinant is 1. From the fact that the first column vectors exhaustS1 it follows that we have constructed a surjective map γ : S1 → SO2R.It is immediate from the definition that γ is an injective group homo-morphism. This completes the proof. �

The reader might wonder why this group is called a torus. The nameoriginally refers to the direct product S1 × S1, which is a topologicaltorus (doughnut). More generally, the product of n ≥ 1 copies of S1 iscalled an n-torus, so S1 itself is the 1-torus.

We shall now examine the structure of T in a similar way as we didfor R+. As a tool, we need the concept of an injective abelian group.

1.21 Definition An abelian group S is said to be injective if everyhomomorphism of abelian groups G → S extends to any abelian groupH containing G. What we need here is the following consequence of thedefinition.

1.22 Theorem Let S be an injective subgroup of an abelian group G.

Then S is a direct summand of G, i.e., there is a subgroup T ≤ G such

that the direct sum S⊕T is isomorphic to G via the map (s, t) �→ s + t.

Proof By injectivity, the identity map of S extends to a homomorphismρ : G→ S (a retraction). Define T := ker ρ and observe that ρ ◦ ρ = ρ.Therefore, τ(g) := g − ρ(g) ∈ T for every g ∈ G, and the identityg = ρ(g)+ τ(g) shows that g �→ (ρ(g), τ(g)) is an isomorphism of G ontoS ⊕ T with inverse (s, t) �→ s + t. �

1.23 Theorem An abelian group is injective if, and only if, it is divis-

ible.

Page 28: The Classical Fields

1 The additive group of real numbers 11

Proof (1) Suppose that S is injective and let s ∈ S, n ∈ N. Define ahomomorphism nZ → S by n �→ s. By injectivity, this extends to ahomomorphism α : Z → S, and we have nα(1) = α(n) = s, hence S isdivisible.

(2) Conversely, let S be divisible, and let ϕ : G → S be a homomor-phism. We have to extend ϕ to a given abelian group H containing G.We consider the set of all pairs (N, ρ) consisting of a group N betweenG and H together with an extension ρ : N → S of ϕ. We say that (N, ρ)precedes (N ′, ρ′) in this set if N ⊆ N ′ and ρ′ extends ρ. It is easily seenthat the ordering defined in this way satisfies the hypothesis of Zorn’sLemma (compare Section 61), so there is an extension ψ : M → S of ϕ

that cannot be extended any further. We shall show that in case M �= H,further extension is possible; the conclusion then is that M = H.

Consider any element h ∈ H � M . If 〈h〉 intersects M trivially, thenwe can extend ψ to the (direct) sum M + 〈h〉 by setting ψ(zh) = 0 forall integers z. If the intersection is non-trivial, then M ∩ 〈h〉 = 〈kh〉 forsome k ∈ N. The image ψ(kh) = s is already defined, and we chooseψ(h) ∈ S such that kψ(h) = s; this is possible by divisibility. It is easyto check that this yields a well-defined homomorphism M + 〈h〉 → S

extending ψ. �

We are now ready to examine the structure of the torus group.

1.24 Theorem There is an isomorphism T = R+/Z ∼= R+ ⊕ (Q+/Z).

Proof There exists a Hamel basis B such that 1 ∈ B. The remainderB′ := B � {1} has the same cardinality as B, hence the rational vectorspace R generated by B′ is isomorphic to R. We have R+ ∼= R+ ⊕Q+,and factorization modulo Z ≤ Q yields the result. �

1.25 Consequences By 1.15, we may substitute (Rn+1)+ ∼= R+ forthe left summand of T, and we obtain after regrouping that

T ∼= (Rn)+ ⊕ T .

We continue by examining Q/Z; compare also Section 31. First weremark that Q/Z is a torsion group, i.e., every element has finite order,because every rational number has some multiple belonging to Z.

Given any abelian group G and a prime p, the p-primary componentGp ≤ G is defined as the set of all elements of G whose order is a powerof p. This is in fact a subgroup; indeed, if #g = pk and #h = pl, thenpmax(k,l)(g + h) = 0, hence #(g + h) is a power of p. Let us examine theprimary components of G = Q/Z.

Page 29: The Classical Fields

12 Real numbers

1.26 Prufer groups The primary components of the group G = Q/Zare called Prufer groups. We denote them by:

Cp∞ := (Q/Z)p .

Consider the coset ab−1 + Z of a rational number ab−1, where a, b arerelatively prime. The coset belongs to Cp∞ if, and only if, there is apower pk such that pkab−1 is an integer, that is, if, and only if, thedenominator b is a power of p. It follows that Cp∞ is the union (or thedirect limit) of the groups Cpk = (p−kZ)/Z ∼= Z/pkZ, the cyclic groupsof order pk:

Cp∞ =⋃

k∈N Cpk .

We note the following consequence of this representation. If a subgroupG ≤ Cp∞ contains elements of arbitrarily large order, then G is theentire group. If the orders of the elements of G are bounded, then G isone of the subgroups Cpk . In particular, every proper subgroup of Cp∞

is cyclic and is uniquely determined by its order.

1.27 Theorem Cp∞ is a divisible group. Moreover, multiplication by

a natural number q not divisible by p defines an automorphism of Cp∞ .

Proof In order to divide an element of Cpk by p, we have to use Cpk+1 ;we have ap−k = p(ap−k−1) for a ∈ Z. Division by q takes place withinthe group Cpk ; indeed, multiplication by q is an injective endomorphismof that finite group, hence an automorphism. Passing to the union overall k, we obtain an automorphism of Cp∞ . �

1.28 Theorem There is a decomposition Q+/Z ∼=⊕p Cp∞ , where the

sum is taken over all primes.

Proof We have to show that every element x = ab−1 + Z has a uniquerepresentation x = gp1 + · · · + gpk

, where gpi∈ Cpi

∞ . To prove exis-tence, it suffices to write ab−1 as a sum of fractions with prime powerdenominators. This can be done by induction. If b = plq, where p isa prime not dividing q, write up−l + vq−1 = (uq + vpl)(plq)−1 and useBezout’s Theorem 1.5 to find u, v such that uq + vpl equals a.

The proof of uniqueness reduces quickly to proving that x = 0 hasonly the trivial representation. Writing −gp1 = gp2 + · · · + gpk

, we seethat the order of the left-hand side is a power of p1 while the order ofthe right-hand side is a product of powers of the remaining primes. Thisshows that both sides are zero if the given primes are all distinct, anduniqueness follows by induction. �

Page 30: The Classical Fields

1 The additive group of real numbers 13

Virtually the same proof shows that every abelian torsion group splitsas the direct sum of its primary components (Exercise 2). The followingresult can be obtained from 1.27 together with 1.28, or directly from thefact that Q+/Z is an epimorphic image of the divisible group Q+.

1.29 Corollary The group Q+/Z is divisible. �

This group will be studied more closely in Section 31.

1.30 Theorem There is an isomorphism×p Cp∞ ∼= R+ ⊕⊕p Cp∞ ,

where p ranges over all primes.

Proof By definition, the direct sum of all Prufer groups is contained intheir direct product. Both groups are divisible, hence injective (1.23),and 1.22 yields a decomposition×p Cp∞ = R ⊕⊕p Cp∞ . We shall usethe characterization 1.14 of R+ in order to show that R ∼= R+. Beinga summand of a divisible group, R is divisible. Moreover, R is torsionfree. Indeed, consider any element x = (gp)p ∈ R�{0}, where gp ∈ Cp∞ .Then x does not belong to the direct sum of the Cp∞ , hence gpi

is non-zero for an infinite sequence of primes pi. If the order of x were a finitenumber n, then all pi would divide n, which is impossible.

It remains to check that R has the right cardinality. The product of allPrufer groups is a Cartesian product of countably many countable sets,hence its cardinality is ℵℵ0

0 = 2ℵ0 ; see 61.15. On the other hand, theright summand is the countable group Q/Z; see 1.28. From×p Cp∞ =R⊕Q/Z we get 2ℵ0 = ℵ0 · cardR = cardR by 61.12. �

1.31 Corollary The torus group R+/Z is isomorphic to the direct

product×p Cp∞ of all Prufer groups.

Proof Combine the results 1.24, 1.28, and 1.30. �

We conclude this section by looking at automorphisms of R+.

1.32 Theorem The automorphism group Λ = Aut R+ consists of the

Q-linear bijections of R, and cardΛ = 2ℵ > ℵ.Proof If λ is an automorphism of R+ and if r ∈ Q, then λ(r ·x) = r ·λ(x)because R+ is uniquely divisible; compare 1.9. Therefore, λ is Q-linear.

A Hamel basis B of R has the same cardinality ℵ as R itself; see 1.12.Therefore, B and R have the same transfinite number card SymB = 2ℵ

of permutations; compare 61.16. Every permutation of B gives rise toan automorphism of R+, which is, of course, a permutation of R. Thus,we have 2ℵ = card SymB ≤ cardΛ ≤ card Sym R = 2ℵ. The last part ofthe assertion follows from 61.11. �

Page 31: The Classical Fields

14 Real numbers

1.33 Corollary The automorphism group Aut T of the torus group

has the same cardinality 2ℵ as Aut R+.

Proof By 1.25, we have T ∼= R+ ⊕ T. Thus, every automorphism of Rextends to an automorphism of T. We infer that 2ℵ ≤ card Aut(R+) ≤card Aut T ≤ card Sym T = 2ℵ. �

This corollary should be contrasted with the fact that the torus grouphas only two continuous automorphisms; see Theorem 8.27. We inferthat the torus group has uncountably many discontinuous group auto-morphisms.

Exercises(1) Suppose that V is a vector space having two infinite bases B and B′. Showthat card B = card B′.

(2) Prove that every abelian torsion group splits as the direct sum of all itsprimary components.

(3) Show that R+ contains subgroups isomorphic to R+ × Zn for arbitraryn ∈ N. Show moreover that these groups are not isomorphic to R+.

(4) The set H of hyperplanes in the rational vector space R has cardinalitycardH = 2ℵ: there are more hyperplanes than one-dimensional subspaces.

(5) The Cantor set {P∞ν=1 cν3−ν | cν ∈ {0, 2}} (compare 5.35) is not con-

tained in any proper subgroup of R.

(6) The numbers log p, p a prime in N, are linearly independent over Q.

(7) Let F be a field of characteristic 0. Show that the additive group F+ hasno maximal subgroup.

(8) Let r ∈ R � Q. Does the subgroup Z + rZ of R+ admit an automorphismof order 5 ?

(9) Determine the isomorphism type of the factor group R/Q of additivegroups.

2 The multiplication of real numbers, with a digression onfields

The multiplicative group of real numbers has ‘almost’ the same struc-ture as the additive group. This fact will be established quickly (2.2),and after that we conduct a systematic search for similar phenomenain other fields. This will culminate in the construction of a field whosemultiplicative group is actually isomorphic to the additive group of realnumbers (see 2.11). The construction makes use of formal power se-ries, which will be treated more systematically later in this book. Somereaders may therefore prefer to skip this topic on first reading.

Page 32: The Classical Fields

2 The multiplication of real numbers, with a digression on fields 15

By R× we denote the multiplicative group (R�{0}, ·) of real numbers.It contains two notable subgroups, the cyclic group {1,−1} ∼= C2 oforder 2 and the group R×

pos of positive real numbers. A real numberr �= 0 is uniquely expressed as a product of sign r ∈ {1,−1} and itsabsolute value |r| ∈ R×

pos. This proves the following.

2.1 Theorem There is a direct sum decomposition R× = R×pos⊕C2. �

The following theorem is the structural interpretation of the functionalequation exp(x + y) = exp(x) exp(y).

2.2 Theorem The exponential function is an isomorphism R+ ∼= R×pos.

Therefore, R× ∼= R+ ⊕ C2. �

The close relationship between the additive and the multiplicativegroup of real numbers exhibited by 2.2 is so important that it seemsworthwhile to look systematically for other fields sharing this property.(As always in this book, fields are commutative by definition.)

Let us consider a few examples, starting with the complex numbers.There, the exponential function is not injective. In fact, the multiplica-tive group C× contains a group isomophic to R+/Z (see 1.20), which hasa large torsion subgroup Q+/Z (see 1.24ff). Hence, C× is not isomorphicto the torsion free group C+ or to C+⊕C2, nor to any subgroup of thesegroups.

The rational numbers form another negative example. Indeed, theadditive group Q+ is locally cyclic (1.6c), and Q× is not. (In fact, Q×

is a direct sum of C2 and a countably infinite number of infinite cyclicgroups; compare 32.1.)

Both examples seem to indicate that the isomorphism 2.2 is a rarephenomenon for fields in general. However, the exact answer dependson how we make our question precise. The following result answers arather coarse form of the question.

2.3 Proposition There is no (skew) field F such that the additive

group F+ is isomorphic to the entire multiplicative group F×.

Proof We prove this by counting involutions. An involution in a groupis an element g of order 2, i.e., g �= 1 = g2 in multiplicative notation org �= 0 = 2g written additively.

The non-zero elements of the additive group F+ of a field F all havethe same order, depending on the characteristic charF ; compare 64.4.This common order is equal to char F if char F is a prime, and infinite ifcharF = 0. Thus, F+ contains involutions only if charF = 2, and thenall elements except 0 are involutions.

Page 33: The Classical Fields

16 Real numbers

On the other hand, if s ∈ F× is an involution, then (s − 1)(s + 1) =s2 − 1 = 0. This equation has either no solution s �= 1 (if charF = 2)or one such solution (if charF �= 2). In both cases, the numbers ofinvolutions do not match. �

This was disappointing, so we modify our question and ask for pairs offields F , G such that F+ is isomorphic to G×. To formulate the answerin the finite case, we need a number theoretic notion.

2.4 Definition: Mersenne primes A prime number of the formp = 2k − 1, k ∈ N, is called a Mersenne prime. Of course, 2k − 1 isnot always prime. A necessary (but not sufficient) condition is that theexponent k be prime. Indeed, for k = mn, we have

2k − 1 = (2k−n + 2k−2n + · · ·+ 2k−mn)(2n − 1) .

More information on Mersenne primes, including references, will be givenin 32.10.

2.5 Example Let q be a prime power. We denote by Fq the uniquefinite field of order q (for a proof of uniqueness see Cohn 2003a 7.8.2,Jacobson 1985 p. 287 or Lang 1993 V 5.1). We have

F+p∼= F×

p+1

if p = 2 or if p is a prime such that p+1 is a power of 2; in the latter case,p is a Mersenne prime. Indeed, the conditions ensure that fields of thegiven orders exist; moreover, the two groups have the same order, andthey are both cyclic (the multiplicative group of a finite field is alwayscyclic; see Exercise 1 of Section 64).

2.6 Theorem The pairs Fp, Fp+1 where p = 2 or p is a Mersenne prime

are the only pairs of fields F , G such that char F �= 0 and F+ ∼= G×.

Proof Let p = charF , and note that F+ contains at least p−1 non-zeroelements, all of order p, while G× cannot have more than p−1 elementsof order p (the solutions x �= 1 of xp − 1 = 0). Now F+ ∼= G× impliesthat F = Fp and G = Fp+1. �

In order to formulate a first result for the case char F = 0, we needthe following.

2.7 Definition We say that a multiplicative group (A, ·) has uniqueroots if the mapping a �→ an is a bijection of A for all n ∈ N. Thisis just the equivalent in multiplicative language of the notion of uniquedivisibility (1.7). We say that a field has unique roots if its multiplicative

Page 34: The Classical Fields

2 The multiplication of real numbers, with a digression on fields 17

group has this property. Note that only fields of characteristic two canhave unique roots, because (−1)2 = 12 in any field.

2.8 Proposition Given a field G, there exists a field F of characteristic

zero such that F+ ∼= G× if, and only if, (char G = 2 and) G has unique

roots.

Proof If charF = 0, then F is a vector space over its prime field Q,hence the additive group F+ is uniquely divisible. Thus, a necessarycondition for the existence of F is that G has unique roots. Conversely,assume that this is the case. Then G× is isomorphic to the additivegroup of a rational vector space V ; see 1.9. Both this vector space andits additive group are determined, up to isomorphism, by the cardinalitycardB of a basis.

If this cardinality is finite, say cardB = n, then we take F to be analgebraic extension of degree n over Q (e.g., F = Q(n

√2)) to ensure that

F+ ∼= G×. If cardB is infinite, then we use a purely transcendentalextension F = Q(T ) (compare 64.19), where card T = cardB. In orderto prove that F+ ∼= V + ∼= G×, we have to show that dimQ Q(T ) =cardT if T is infinite. This is proved in 64.20. �

Proposition 2.8 raises the question as to which uniquely divisiblegroups occur as the multiplicative groups of fields. This question isanswered completely by Contessa et al. 1999 5.3 and 5.5; they showthat a uniquely divisible abelian group A is the multiplicative group ofsome field if, and only if, the dimension of A as a vector space over Q isinfinite. These groups also occur as the additive groups of fields, as wehave shown in 2.8.

Here we shall be content to give examples of fields G having uniqueroots. The following lemma allows us to obtain roots in the power seriesring F [[t]], which is defined in 64.22.

2.9 Lemma Let F be a field. If m ∈ N is not a multiple of the

characteristic of F , then every element a ∈ 1 + tF [[t]] admits an mth

root c ∈ 1 + tF [[t]], that is, cm = a.

Proof Put c0 = 1. Then cm0 = 1 ≡ a mod t; in general, a congruence

x ≡ y mod tn means that tn divides x− y in the ring F [[t]].Assume that we have found elements c0, c1, . . . , ck−1 ∈ F such that

pmk−1 ≡ a mod tk, where pk−1 :=

∑k−1i=0 cit

i. Then

pmk−1 ≡ a + btk mod tk+1

for some b ∈ F . By our assumption on m, we can define ck ∈ F by

Page 35: The Classical Fields

18 Real numbers

mck = −b. We show that the polynomial pk := pk−1 + cktk satisfiespm

k ≡ a mod tk+1.We have pk−1 ≡ c0 ≡ 1 mod t, which implies pm−1

k−1 ≡ 1 mod t andpm−1

k−1 tk ≡ tk mod tk+1. Therefore the following congruences mod tk+1

hold:

pmk = (pk−1 + cktk)m ≡ pm

k−1 + mckpm−1k−1 tk ≡ a + btk − bpm−1

k−1 tk ≡ a .

Now the formal power series c :=∑

i≥0 citi satisfies the congruences

cm − a ≡ pmk−1 − a ≡ 0 mod tk for every k ∈ N, hence cm − a = 0. �

We shall now apply this lemma to fields of Puiseux series F ((t1/∞))as introduced in 64.24, in order to produce examples of fields havingunique roots. A Puiseux series is a formal sum

a =∑

i≥k aiti/n ,

where n ∈ N, k ∈ Z, and the coefficients ai belong to a given field F .

2.10 Theorem If F is a field (necessarily of characteristic 2) which has

unique roots, e.g., F = F2, then the field F ((t1/∞)) of Puiseux series

has unique roots, as well.

In particular, the multiplicative group of F ((t1/∞)) is isomorphic to

the additive group of some other field, by 2.8.

Proof (1) Each non-zero Puiseux series can be written uniquely as aproduct atr(1+ b), where a ∈ F×, r ∈ Q and b is a Puiseux series whichinvolves only powers of t with positive exponents, i.e., b ∈ t1/nF [[t1/n]]for some n ∈ N.

The Puiseux series of the form 1 + b = 1 +∑

i≥1 biti/n with bi ∈ F

form a subgroup of the multiplicative group F ((t1/∞))× (note that thegeometric series

∑i≥0 bi = (1+b)−1 makes sense and shows that (1+b)−1

is again of that form; compare also 64.24), and F ((t1/∞))× is the directproduct of F×, tQ and this subgroup.

Therefore it suffices to examine these three factors separately. Theelement a ∈ F× has unique roots by our hypothesis, and tr/m is theunique mth root of tr. It remains to show that 1 + b = 1 +

∑i≥1 bit

i/n

has a unique mth root for every m ∈ N, and we may assume that m isa prime number.

(2) If m �= 2 = char(F ), then the existence of such a root is a con-sequence of Lemma 2.9, since the subring F [[t1/n]] of F ((t1/∞)) is iso-morphic to the power series ring F [[t]]. For m = 2 we compute directlythat (1 +

∑i≥1

√bit

i/(2n))2 = 1 + b.

Page 36: The Classical Fields

2 The multiplication of real numbers, with a digression on fields 19

(3) For the proof of uniqueness, it suffices to show that 1 = (1 + b)m

implies b = 0. Assume that b �= 0 and let k := min{ i ∈ N | bi �= 0} ≥ 1.By binomial expansion we have

1 = (1 + b)m = 1 +(m1

)b + · · ·+ bm = 1 + mbktk/n +

∑i>k cit

i/n

with ci ∈ F . We obtain mbk = 0 �= bk, hence m = 2 and 1 = (1 + b)2 =1 + b2, a contradiction to b �= 0. �

2.11 Corollary The multiplicative group of the field F = F2((t1/∞))is isomorphic to the additive group R+.

Proof By 2.10, the group F× is divisible and torsion free. MoreoverFN

2 ⊆ F ⊆ FQ2 , hence card F = 2ℵ0 = card R. Now 1.14 gives the

assertion. �

Using different methods (compare 2.12), this was shown by Contessa

et al. 1999 5.6 (Moreover they prove that, in contrast, the additive groupof rational numbers is not isomorphic to the multiplicative group of anyfield, loc. cit., 5.3).

Furthermore the fields F = F2((Qn)) and F = F2((R)) of Hahn powerseries (see 64.25) have the property that F× ∼= R+.

The field F2((t1/∞)) of Puiseux series is not an algebraic extension ofF2, as t is transcendental. In fact, a proper algebraic extension of F2

contains finite subfields distinct from F2, hence it contains non-trivialroots of unity. However, we show next that it is possible to obtain fieldshaving unique roots by algebraic extension from fields containing noroots of unity. The result is taken from Contessa et al. 1999 4.3.

2.12 Theorem Let L be a field of characteristic 2 that does not contain

any non-trivial roots of unity. If L is not the prime field F2, then there

is an algebraic extension field M of L that has unique roots.

For example, the field L may be any purely transcendental extensionof F2.

Proof The problem is to adjoin roots of all degrees for every a ∈ L�{0, 1}without adjoining any root of unity. Consider an algebraic closure L�

and the collection M of all fields F with L ≤ F ≤ L� such that F doesnot contain any non-trivial root of unity. The union of every chain inMbelongs to M. By Zorn’s Lemma, there is a maximal element M ∈ M,and we proceed to show that M contains a pth root of a for every primep and every a ∈ M . We may assume that a /∈ {0, 1}. Our claim is aconsequence of maximality together with the following lemma. �

Page 37: The Classical Fields

20 Real numbers

2.13 Lemma Let M be a field of characteristic 2 that does not contain

any non-trivial roots of unity. Suppose that the polynomial q(x) = xp−a

has no root in M for some prime p and some a ∈M � {1}, and consider

the algebraic extension M(ϑ) by a root ϑ of q(x). Then M(ϑ) does not

contain any non-trivial roots of unity.

Proof (1) Suppose that p = 2 and that for some b, c ∈ M , the elementη = b+ cϑ ∈M(ϑ) satisfies ηk = 1, where k > 1. Then η2 = b2 + c2ϑ2 =b2 + c2a ∈ M as we are in characteristic 2. Since η2k = 12 and sinceM does not contain any roots of unity other than 1, we conclude thatη2 = 1 and, hence, that η = 1. For the remainder of the proof, we mayassume that p is odd.

(2) We claim that the polynomial q(x) is irreducible over M (compareCohn 2003a 7.10.8 or Lang 1993 VI §9). The roots of q(x) in M �

are of the form ζνϑ, where ζ0, ζ1, . . . , ζp−1 are the pth roots of unity inM �. If q(x) splits over M as q(x) = r(x)s(x), then r(x) is a productof some linear factors x + ζνϑ, hence the constant term b ∈ M of r(x)has the form δϑμ, where δp = 1 and 0 < μ < p. We have bp = aμ,and Bezout’s Theorem 1.5 provides m,n ∈ Z such that mμ + np = 1;this yields a = amμ+np = bmpanp, hence bman ∈ M is a root of q(x), acontradiction.

We have shown that the degree [M(ϑ) : M ] equals the prime p, hencethe degree formula 64.2 implies that there are no fields properly betweenM and M(ϑ).

(3) Suppose that M(ϑ) contains some root of unity η �= 1. Let k > 1be the multiplicative order of η. Then xk − 1 =

∏k−1i=0 (x − ηi). Thus,

M(η) ⊆ M(ϑ) is the splitting field of the polynomial xk − 1 and isa normal extension field of M ; see 64.10. Now step (2) implies thatM(η) = M(ϑ); therefore, q(x) splits into linear factors in M(ϑ), and thelinear factors are as shown in step (2). It follows that M(ϑ) containsall pth roots of unity ζ0, ζ1, . . . , ζp−1. We have ζi �= 1 for some index i,otherwise xp − 1 = (x − 1)p in M [x], which leads to the excluded casep = 2. As before, we conclude that M(ζi) = M(ϑ), but the minimalpolynomial of ζi over M divides xp−1 + xp−2 + ... + 1 and hence hasdegree less than p, a contradiction. �

The reader may feel that we still have not treated the ‘right’ question.We are looking for fields whose behaviour is similar to that of the realnumbers – but R+ is not isomorphic to R×. To come closer to the realcase, we should consider ordered fields F (necessarily of characteristiczero) whose additive group is isomorphic to the multiplicative group

Page 38: The Classical Fields

2 The multiplication of real numbers, with a digression on fields 21

of positive elements. For the notion of an ordered field, compare 11.1.We shall not determine all fields having the above property, but thegroups appearing as F+ and F×

pos for the same ordered field F will bedetermined up to isomorphism. The following result was presented byG. Kaerlein at a conference at Bad Windsheim, Germany, in 1980.

2.14 Theorem For an abelian group (H, +), the following conditions

are equivalent.

(a) H is uniquely divisible and the dimension of H as a rational vector

space (compare 1.9) is infinite.

(b) H is a direct sum of infinitely many copies of Q+.

(c) There is an ordered field F such that H ∼= F+ ∼= F×pos.

Proof (1) We know from Section 1 that conditions (a) and (b) are equiva-lent, so we have to show that they are necessary and sufficient for (c).

(2) Necessity. An ordered field F has characteristic 0, hence F is avector space over its prime field Q, and F+ is uniquely divisible. We haveto show that dimQ F is infinite. Now F×

pos∼= F+ has unique roots. This

implies that F contains elements that are algebraic over Q of arbitrarilylarge degrees, and dimQ F is an upper bound for these degrees (see 64.5),hence dimQ F is infinite.

(3) Sufficiency. As in the proof of 2.8, we find a purely transcendentalextension Q(T ) whose additive group is isomorphic to H, and we proceedto turn Q(T ) into an ordered field. We use a total ordering of thetranscendency basis T and define a lexicographic ordering on monomials:For t1 > · · · > tk and integers ni ≥ 0, mi ≥ 0, we set tn1

1 . . . tnk

k >

tm11 . . . tmk

k if, and only if, there is an index i0 such that ni = mi for i < i0and ni0 > mi0 . A polynomial in the indeterminates t ∈ T is positive bydefinition if the coefficient of the largest monomial is positive. A quotientof two polynomials is said to be positive if either both polynomials arepositive or both are negative. The proof that this makes the field offractions Q(T ) an ordered field is left to the reader.

From the fact that the field Q(T ) is ordered, we deduce that it isformally real; see 12.1. By 12.16, it has an algebraic extension F that isreal closed. Now 12.10(ii) together with 7.3 yields that F×

pos has uniqueroots. The cardinality of F is the same as that of Q(T ); see 64.5. Againthe arguments of the proof of 2.8 show that F+ ∼= H ∼= F×

pos. �

Remembering that the real exponential function is monotone, we maysharpen the previous question once again. We can ask for ordered fieldssuch that F+ and F×

pos are isomorphic as ordered groups. This leads

Page 39: The Classical Fields

22 Real numbers

to the notion of exponential fields. The question, which belongs to therealm of ordered fields, will be discussed in Section 11; see 11.10.

Exercises(1) Supply the details for the proof of 2.14 by proving that the definition giventhere turns a purely transcendental extension Q(T ) into an ordered field.

(2) The group R+ has infinitely many extensions by the cyclic group C2 oforder 2. Only one of them is commutative.

3 The real numbers as an ordered set

There is an ordering relation (or order) < on the set of real numbers,which makes R a chain (or totally ordered set). This means that forr, s ∈ R, precisely one of the relations r < s, s < r, s = r holds. Inaddition, we have the usual ‘transitivity’ property of an ordering: r < s

and s < t together imply r < t. The following examples explain ournotation for intervals in a chain C:

[a, b[ = { c ∈ C | a ≤ c < b}]a, [ = { c ∈ C | a < c} .

We note that the ordering induced on an interval [n, n + 1[ betweenconsecutive integers coincides with the lexicographic ordering obtainedby comparing binary expansions.

The following notions will be used to characterize the chain R of realnumbers and its subchains Z (the integers) and Q (rational numbers).In a later chapter, this will be applied in order to characterize the topo-logical space of real numbers; see 5.10.

3.1 Completeness, density, separability A chain C (or its ordering)is said to be complete if every non-empty subset B ⊆ C which is boundedabove has a least upper bound (or supremum) supB. The chain of realnumbers is complete. In fact, the usual way to obtain the real numbersis by completion of the rational numbers. Compare 42.11 and 44.11.The idea of completion will be studied thoroughly in Chapter 4.

A subset A of a chain C is said to be coterminal if A contains lowerand upper bounds for every c ∈ C. A coterminal subset A ⊆ C is saidto be weakly dense in C if every c ∈ C satisfies c = inf{a ∈ A | c ≤ a}and c = sup{a ∈ A | a ≤ c}. Thus A is weakly dense in the chain C if,and only if, for all c1 < c2 in C there exist elements a1, a2, a3, a4 ∈ A

such that a1 ≤ c1 ≤ a2 < a3 ≤ c2 ≤ a4. This implies that every closed

Page 40: The Classical Fields

3 The real numbers as an ordered set 23

interval [c, d] of C with c < d contains an element of A. For the purposesof the present section, the latter condition would suffice, but in Sections41 and 42 we shall need the notion of weak density as defined here.

A subset A ⊆ C is called strongly dense if for all pairs c < d inC, the set A contains elements x, y, z such that x < c < y < d < z.For example, the rational numbers form a strongly dense subset of R.Observe that a strongly dense subset of C is infinite unless C has onlyone element. Note also that strong density implies weak density.

If A is strongly dense in C, then C has no extremal (smallest orlargest) element and does not contain any gaps, i.e., there is no emptyopen interval ]c, d[. If A is only weakly dense, then extremal elementsmay exist, but they have to belong to A, and gaps ]c, d[ may exist, butthen c and d must belong to A. On the other hand, if ]c, d[ is non-empty,say x ∈ ]c, d[, then [x, d[ must contain an element of A.

A chain C is called weakly or strongly separable if it has a weakly orstrongly dense subset A, respectively, which is at most countable. Thedensity of Q implies that R is strongly separable.

3.2 Topology induced by an ordering Every chain C is a topo-logical space in a natural way. The topology induced by the ordering ororder topology of C is the smallest topology such that all open intervals(including ]a, [ and ] , a[) are open sets. In this topology, a set is openif, and only if, it is the union of (perhaps infinitely many) open intervals.

Let us compare the notions of density for chains to their topologicalanalogues. In a topological space X, density means that every non-empty open set of X contains an element of A. In a chain, an openinterval can be empty. Therefore, a subset can be dense in the topologyinduced by the ordering without being strongly dense in the chain. Forexample, the pair of chains Z ⊆ Z is topologically dense and weaklybut not strongly dense, and the pair ]0, 1] ⊆ [0, 1] is topologically densebut not even weakly dense. Conversely, weak density implies topologicaldensity, because only non-empty open intervals are involved. Thus wehave the implications

strongly dense ⇒ weakly dense ⇒ topologically dense,

none of which is reversible.A topological space X is said to be separable if it has a countable dense

subset A, and similar remarks as above hold for this notion. Thus, Z isa separable topological space, but not a strongly separable chain; it is,however, weakly separable.

Page 41: The Classical Fields

24 Real numbers

The following is a standard fact.

3.3 Proposition For a chain C, the following two conditions are equiva-

lent.

(1) The topology induced by the ordering of C is connected.

(2) The chain C is complete, and each pair c < d in C defines a non-

empty interval ]c, d[.

The condition about intervals is satisfied if C is strongly dense initself, but it does not exclude the existence of a largest (or a smallest)element.

Proof Suppose that condition (2) is violated; we show that C is notconnected. If the interval ]c, d[ is empty, then C is the disjoint unionof the two open sets ] , d[ and ]c, [, and is not connected. On the otherhand, if C is not complete, let A ⊆ C be a bounded subset without asupremum. Then the set B of all upper bounds of A is open, as well asthe union U of all intervals ] , a[, a ∈ A. The sets U and B are disjointand cover C.

Next suppose that (2) is satisfied; we show that C is connected. Thuswe assume that C is the disjoint union of two non-empty open setsU , V , and we derive a contradiction. There are elements a ∈ U andb ∈ V , and we may assume that a < b. We consider the bounded setB obtained as the union of all intervals ] , u[, where u ∈ U and u < b.We show that every open interval I containing s = sup B meets bothU and V , contradicting the assumption that one of these disjoint opensets contains s.

We prove that I ∩ V is non-empty. We may assume that s < b, andthen the non-empty set ]s, b[ does not contain any u ∈ U ; or else, ]s, u[would be contained in B, contrary to the definition of s.

Finally, I contains some interval ]c, s], which in turn contains someelement x ∈ B. By definition of B, there is an element u ∈ U such thatx < u < b; in fact, u ≤ s because we have seen that ]s, b[ ⊆ V . Thus,u ∈ I ∩ U . �

We say that map ϕ : C → C ′ between two chains preserves the order-ing, or is order-preserving , if for all x, y ∈ C such that x ≤ y one hasϕ(x) ≤ ϕ(y). Note that such a map need not be injective. An order-preserving map ϕ is injective precisely if x < y implies ϕ(x) < ϕ(y).

An order-preserving bijection ϕ : C → C ′ is called an order isomor-phism; if such a bijection exists, then the chains C and C′ are calledorder isomorphic.

Page 42: The Classical Fields

3 The real numbers as an ordered set 25

The following characterization of the chain Q is due to Cantor 1895§9; compare Hausdorff 1914 p. 99, Birkhoff 1948 p. 31. It is acrucial step in the proof of the subsequent characterization of R.

3.4 Theorem A chain A is order isomorphic to the chain Q of rational

numbers if, and only if, A is countable and strongly dense in itself.

Proof Clearly, Q has the properties mentioned. Conversely, supposethat A has the same properties. We use enumerations a1, a2, . . . andq1, q2, . . . of A and Q, respectively, in order to construct an order iso-morphism f : A→ Q via induction.

We set out by defining f1(a1) = q1 and A1 = {a1}, Q1 = {q1}. Wedefine two sequences of subsets Am ⊆ A and Qm ⊆ Q, both of cardinalitym, and order isomorphisms fm : Am → Qm. Alternatingly, we performtwo different inductive steps, each time adding single elements aμ andqν to the sets Am and Qm, respectively, and at the same time extendingfm to an isomorphism Am+1 → Qm+1. If m is odd, we choose ν to bethe smallest number such that qν /∈ Qm. By strong density of A, there isan element aμ ∈ A�Am such that the desired extension is possible, andwe insist that μ be chosen minimal in order to make a definite choice. Ifm is even, we interchange the roles of A and Q; in other words, we beginby choosing aν /∈ Am with ν minimal, and then select qμ such that theextension of fm is possible and μ is minimal. This alternating strategytogether with the minimality condition for ν ensures that every elementof A and of Q is used at some point, hence by taking unions we obtaina bijection A→ Q which preserves the ordering. �

The basic idea of this proof can be used to show that the chain Qembeds into every chain A that is dense in itself. The direction of themap f is of course reversed, and it is not necessary to alternate the rolesof A and Q. Instead of an enumeration of A, one uses a well-ordering.

3.5 Theorem The following conditions are mutually equivalent for a

chain R having more than one element.

(a) R is order isomorphic to the chain R of real numbers.

(b) R is strongly separable and complete.

(c) R is strongly separable and the topology induced by the ordering of

R is connected.

Proof We know that (b) and (c) are properties of the real numbers, andwe proved in 3.3 that (b) and (c) are equivalent. We shall finish theproof by showing that a complete strongly separable chain R can bemapped order isomorphically onto the chain R. By strong separability,

Page 43: The Classical Fields

26 Real numbers

R contains a countably infinite chain A that is strongly dense in R and,hence, in itself. By 3.4, there is an order isomorphism f : A → Q. Weshow how to extend f to an order isomorphism R→ R.

Given r ∈ R, let Ar := {a ∈ A | a ≤ r } and choose b ∈ A suchthat b > r. Then f(Ar) < f(b) is a bounded subset of the completechain R, and we may define f(r) := sup f(Ar). This map extends thegiven one, and it has an inverse, defined in the same way; note thatr = sup Ar. Moreover, f preserves the ordering. Indeed, if r < s, thenf(Ar) ⊆ f(As), hence f(r) = sup f(Ar) ≤ sup f(As) = f(s). �

Note that the same arguments prove that a chain is isomorphic to asubchain of R if, and only if, it is weakly separable; compare Birkhoff

1948 p. 32.

3.6 Definition A chain is said to be homogeneous, if its group ofautomorphisms is transitive, that is, if every element can be mappedonto every other one by an automorphism.

Examples of homogeneous chains are Z (the integers), Q (rationalnumbers), and R (real numbers). For two of these examples, we havethe following characterization, taken from Birkhoff 1948 III.8.

3.7 Theorem Let C be a chain that is complete, homogeneous, and

weakly separable as defined in 3.1. Then C is order isomorphic to the

chain Z of integers or to the chain R of real numbers, or C has only one

element.

Proof (1) If C has a smallest or largest element, then C is a singletonby homogeneity. Henceforth, we assume that this is not the case.

(2) Assume that every interval ]a, b[ with a < b is non-empty. Thiscase leads to the real numbers. Indeed, ]a, b[ contains a closed interval,and weak separability of C implies strong separability, whence we mayapply Theorem 3.5.

(3) If a < b and ]a, b[ = ∅, then every element of C has an upperneighbour and a lower neighbour, by homogeneity. Inductively, we mayconstruct an isomorphism of Z onto a subchain Z ⊆ C by insisting thatthe ‘upper neighbour’ relation be preserved. This subchain cannot bebounded, because its supremum or infimum could not have any loweror upper neighbours, respectively. Therefore, every c ∈ C must coincidewith its least upper bound in Z. It follows that Z = C, and the proofis complete. �

We remark that R contains other subchains with rather strange prop-erties: van Mill 1992 constructs an example of a homogeneous subchain

Page 44: The Classical Fields

3 The real numbers as an ordered set 27

X ⊆ R which is dense in itself and rigid in the sense that the identity isthe only automorphism fixing a point; the automorphism group of X isthen sharply transitive on X.

A striking characterization of the real line is obtained in Gurevich–

Holland 1981; compare also Glass 1981. They prove the existence ofa formula in the elementary language of groups such that R is the onlyhomogeneous chain whose automorphism group satisfies this formula.

There is a similar result for the chain of rational numbers, but witha difference. The chain of real numbers is obtained from the chain ofrational numbers by completion (see 42.11), hence the automorphismgroup of the chain Q can be identified with the group of all automor-phisms of R that leave Q invariant. The same argument holds for thechain R�Q of irrational numbers, hence the chains Q and R�Q cannotbe distinguished by properties of their automorphism groups.

3.8 Souslin’s problem A weakly separable chain C satisfies Souslin’scondition (also called the countable chain condition, Souslin 1920):there is no uncountable set of pairwise disjoint, non-empty open intervalsin C. Indeed, each of those intervals would contain an element of a fixedcountable dense subset of C.

It has been conjectured that this condition is equivalent to topologicalseparability (Souslin’s hypothesis; see Birkhoff 1948 III.8, Alvarez

1999). Then we would obtain a variation of the preceding character-ization of R (Theorem 3.7) with the Souslin condition in place of weakseparability. It turned out soon that Souslin’s hypothesis cannot bedecided on the basis of a ‘weak’ set theory. Later, it was even shownthat there are models of set theory satisfying the axiom of choice and thegeneralized continuum hypothesis (see 61.16), where Souslin’s hypothesisis false (Tennenbaum 1968 and Rudin 1969).

The Souslin hypothesis is equivalent to the non-existence of Souslintrees, which are defined as follows. A Souslin tree is a partially or-dered set (T,≤) of cardinality ℵ1 such that (1) for each x ∈ T , the set{y ∈ T | y < x} is well-ordered with respect to ≤, and (2) if X ⊆ T

is a chain (totally ordered by ≤) or an antichain (containing no pair of≤-comparable elements), then Y is at most countable. For more infor-mation on the Souslin problem, see Devlin–Johnsbraten 1974. Inhis article, Felgner 1976 treats the Souslin hypothesis in the contextof ordered algebraic structures. Among other things, he proves thatan abelian ordered group that is dense in itself and satisfies Souslin’scondition is separable. He also obtains some negative results.

Page 45: The Classical Fields

28 Real numbers

Exercises(1) Leaving the given ordering unchanged, split each real number x into a pair{x, x} with x < x. Compare the properties of the chain R of the split realswith those of R.

(2) Consider the set R2 with lexicographic ordering, that is, (x, y) < (u, v)means that x < u or (x = u and y < v). Is there an injection R2 → R whichpreserves the ordering?

4 Continued fractions

Continued fractions are a useful tool for a better understanding of theembedding of Q into R, the approximation of irrational numbers byrational ones, the topology of R � Q, and related phenomena. Thebooklet by Khinchin 1964 gives an easy introduction to the subject;see also Chapter 5 of Niven 1956, Chapter X of Hardy–Wright 1971,or the more recent text by Rockett–Szusz 1992.

4.1 Definition For each real number ζ, the process (�) defined by theconditions

ζ = ζ0 = c0 + 1/ζ1, c0 ∈ Z, 1 < ζν = cν + 1/ζν+1, cν ∈ N for ν > 0

uniquely determines the (finite or infinite) sequences of integers cν andreal numbers ζν . If ζ /∈ Q, then none of the ζν is rational, and theprocess (�) does not terminate. For ζ ∈ Q � Z, however, ζ1 = a1/a2

is a quotient of two coprime natural numbers and the cν are given bythe Euclidean algorithm aν = cνaν+1 + aν+2 with aν > aν+1 ∈ N.Hence (�) stops at some finite index n with ζn = cn, and one obtains arepresentation of ζ as a finite continued fraction

[ c0; c1, c2, . . . , cn ] := c0 +1

c1 +1

c2 +1

· · ·+ 1cn

= pn/qn ,

where pn ∈ Z and qn ∈ N are integers without common factor. Forζ /∈ Q it will be seen below that the rational numbers pn/qn converge toζ in a particularly nice manner.

The definition of [ c0; c1, c2, . . . , cn ] makes sense also if cn ≥ 1 is anarbitrary real number.

4.2 Observe that [ c0; c1, . . . , cn+1 ] is obtained from [ c0; c1, . . . , cn ] bywriting cn + 1/cn+1 in place of cn. Let pro forma p−1 = 1, q−1 = 0

Page 46: The Classical Fields

4 Continued fractions 29

and put p0 = c0, q0 = 1. Then the following recursion formulae can beproved by induction:

pν+1 = cν+1pν + pν−1, qν+1 = cν+1qν + qν−1

and pν+1qν − pνqν+1 = (−1)ν .(†)

In fact, if (†) is true for smaller indices, then

pν+1

qν+1=

(cν + c−1ν+1)pν−1 + pν−2

(cν + c−1ν+1)qν−1 + qν−2

=pν + pν−1/cν+1

qν + qν−1/cν+1=

cν+1pν + pν−1

cν+1qν + qν−1.

The last part of (†) follows inductively from the first. It shows that pν

and qν are indeed coprime.Equivalently,(

pν pν−1

qν qν−1

)=(

c0 11

)(c1 11

)· · ·(

cν 11

). (‡)

The determinant of (‡) yields the last part of (†) once more. For lateruse, we rewrite this as

pν/qν − pν−1/qν−1 = (−1)ν−1/qνqν−1 .

Note that for ν ≥ 1 the qν form a strictly increasing sequence. Moreover,again by induction, qν/qν−1 = [ cν ; cν−1, . . . , c1 ].

The recursion formulae can also be expressed as follows: if

| c1, c2, . . . , cν | =

∣∣∣∣∣∣∣∣∣∣∣∣∣

c1 1−1 c2 1

−1 c3 1. . . . . . . . . . . . . . . . . . . . . . . .

−1 cν−1 1−1 cν

∣∣∣∣∣∣∣∣∣∣∣∣∣,

then pν = | c0, c1, . . . , cν | and qν = | c1, c2, . . . , cν |.4.3 Convergence If ζ ∈ R � Q, if the cν are defined by (�), andif [ c0; c1, . . . , cn ] = sn, then s2ν < ζ < s2ν+1 : in fact, we haveζ = [ c0; c1, . . . , cn−1, ζn ], and continued fractions are monotone in eachvariable, increasing for even indices, decreasing for odd indices. In 4.2it has been shown that s2ν+1 − s2ν = 1/q2ν+1q2ν < q−2

2ν ≤ (2ν)−2 con-verges to zero. Hence the sν converge alternatingly to ζ, and one maywrite ζ as an infinite continued fraction ζ = [ c0; c1, c2, . . . ].

For an arbitrary sequence (cν)ν ∈ Z × NN, let sν = [ c0; c1, . . . , cν ] =pν/qν as before. Then again |sν+1 − sν | < q−2

ν , and the approximating

Page 47: The Classical Fields

30 Real numbers

fractions sν converge to some real number γ. If p/q is any approximatingfraction, then 0 < |qγ−p| < q−1, as we have seen. Thus 1 and γ generatea dense subgroup of (R, +) and γ is irrational; see 1.6a. Moreover, theprocess (�) applied to γ produces the sequence (cν)ν . This follows by in-duction, using the obvious fact that [ c0; c1, c2, . . . ] = c0 +[ c1; c2, . . . ]−1.

4.4 Example If ζ =√

3, then ζ1 = (√

3 − 1)−1 = (√

3 + 1)/2, ζ2 =2(√

3−1)−1 =√

3+1, ζ3 = ζ1, ζ4 = ζ2, and hence ζ = [ 1; 1, 2, 1, 2, 1, . . . ],s7 = 97/56, s7 −

√3 < 10−4. Conversely, let z = [ 0; 1, 2, 1, 2, . . . ]. Then

z = 1/(1 + 1/(2 + z)) = (z + 2)/(z + 3), hence (z + 1)2 = 3.

4.5 Remark The example is a typical case of the following theorem:A continued fraction c = [ c0; c1, c2, . . . ] is finally periodic if, and only if,

Q(c) is quadratic over Q; see Khinchin 1964 §10 or Rockett–Szusz

1992 Chapter III for proofs.

4.6 Theorem (Hurwitz 1891) One of any three consecutive approx-

imating fractions satisfies even |ζ − pν/qν | <(√

5 q 2ν

)−1.

For a proof see Khinchin 1964 §7 Theorem 20 or Rockett–Szusz

1992 p. 80; compare also Benito–Escribano 2002 and Exercise 3.The approximation of real numbers by rational numbers has been

studied by many mathematicians, including Thue, Siegel, Dyson, Gel-fond and T. Schneider. The following famous result says that only trans-cendental numbers admit many rational approximations that are betterthan 4.6. This result is due to Roth 1955 (see also Roth 1960 and, fora detailed proof, Cassels 1957 Chapter VI).

4.7 Theorem (Roth 1955) For any real algebraic number α and any

real number μ > 2, there are only finitely many rational numbers p/q

such that |α− p/q| < q−μ.

In contrast, for μ ≤ 2 and every irrational number α there are infinitelymany rational numbers p/q such that |α − p/q| < q−μ. This can beobtained for instance by expanding α into a continued fraction; see 4.3.

4.8 Ordering For infinite continued fractions a and b, the alternatinglexicographic ordering

a < b �{

aμ < bμ ∧ μ ∈ 2Z

aμ > bμ ∧ μ /∈ 2Z, μ = μ(a, b) = min{ν ≥ 0 | aν �= bν },

gives the usual ordering of the irrational numbers; it can be extendedto the rationals by writing [ c0; c1, . . . , cn ] = [ c0; c1, . . . , cn,∞ ], where

Page 48: The Classical Fields

4 Continued fractions 31

R <∞. Thus, mapping [ c0; c1, c2, . . . ] to its limit is a homeomorphismof the set of all infinite continued fractions onto the set I of the irrationalreals with respect to the order topologies.

4.9 Proposition A lower bound for the distance of an irrational

number ζ and an approximating fraction is given by |ζ − pν/qν | >

q−1ν (qν + qν+1)−1.

Proof Let ζ = [ c0; c1, c2, . . . ]. If ν is an even integer, then pν/qν =[ c0; c1, . . . , cν ] < ζ = [ c0; c1, . . . , cν , ζν+1 ]. The interval [ pν/qν , ζ ] con-tains the number [ c0; c1, ..., cν , cν+1 + 1 ], which by 4.2(†) is equal to(pν+1 + pν)/(qν+1 + qν), and the assertion follows. The odd case istreated analogously. �

Now we explain how continued fractions can be used to describe thetopology of I = R�Q.

4.10 Topology Consider the set of all infinite continued fractions asa subspace of the (Tychonoff) power ZN, where Z carries the discretetopology. This means that a typical neighbourhood of [ c0; c1, c2, . . . ]consists of all x = [x0; x1, x2, . . . ] such that xκ = cκ for κ ≤ m andsome natural number m. If m is even, this is equivalent to

[ c0; c1, . . . , cm ] < x < [ c0; c1, . . . , cm−1, cm + 1 ] ;

analogous relations hold for odd m. Therefore, ZN induces the sametopology as the ordering, and I is homeomorphic to Z × NN. As bothZ and N are countable, discrete spaces and hence Z ≈ N, the space I isalso homeomorphic to NN.

4.11 Metric The discrete topology on the νth factor of ZN is given bythe metric dν that has only the values 0 and 1. The space ZN is theneasily seen to be complete with respect to the metric d =

∑ν 2−νdν ;

note that d induces the product topology on ZN. Together with 4.10,this yields a complete metric for the subspace I of R.

Remark A more general sort of continued fractions (with numeratorsbν instead of 1) is discussed in Beardon 2001.

Exercises(1) Let [ 0; c1, . . . , cν ] = pν/qν with integers pν , qν ∈ N and ν ≥ 4. Thenqν ≥ 5 · (3/2)ν−4 and qν−1qν ≥ 6 · (5/2)ν−3.

(2) Show that pνqν−2 − pν−2qν = (−1)νcν .

(3) Prove that one of any two consecutive approximating fractions for anirrational number ζ satisfies |ζ − p/q| < (2q2)−1.

Page 49: The Classical Fields

32 Real numbers

5 The real numbers as a topological space

The chain R can be endowed with the topology induced by the ordering(see 3.1), and thus becomes a topological space. The open intervals forma basis of this topology; in other words, every open set is a union of openintervals. The same topology can be generated by the metric d(x, y) =|x − y|. This means that the open balls Uε(y) = {x ∈ R | d(x, y) < ε}for all y ∈ R and all ε > 0 form a basis, as well.

Like every metric space, R is a Hausdorff space, i.e., every two pointspossess disjoint neighbourhoods. In particular, each point r ∈ R formsa closed set, and R is a T1-space. In passing, we mention some strongerconsequences of being metric, namely, normality and paracompactness;see, for example, Dugundji 1966 Section IX.5.

Density of the rational numbers in the chain R (see 3.1) implies thateach point in Uε(y) belongs to some ball Ur(q) contained in Uε(y) andhaving rational centre q and rational radius r. In other words, we havefound another basis for the topology of R consisting of countably manyopen sets. This property is stronger than separability (existence of acountable dense subset in the topological sense, see 3.2; the rationalnumbers form such a set). Whenever we speak of the real numbers asa topological space without further specification, we mean this ‘natural’topology. We record the properties just recognized.

5.1 Proposition The topological space R of real numbers has a count-

able basis; in particular, R is separable. �

We call a topological space compact , if every covering of the space byopen sets contains a finite subcovering; the Hausdorff separation prop-erty is not required.

In order to determine the compact subsets of R, we characterize thechains that are compact with respect to their order topologies. We calla chain bounded if it has a smallest and a greatest element. A chain thatis both bounded and complete is unconditionally complete, i.e., everynon-empty subset has a least upper bound and a greatest lower bound.

5.2 Theorem A chain C is compact with respect to the topology

induced by the ordering if, and only if, it is unconditionally complete.

Proof Suppose that C is compact. In search of a least upper bound of anon-empty set A ⊆ C, we consider a set of open subsets of C, consistingof the intervals ] , a[ for all a ∈ A together with the intervals ]b, [ for allb ≥ A. Note that an element of C not contained in any of these intervalsis a least upper bound of A. Therefore, we may assume that we have

Page 50: The Classical Fields

5 The real numbers as a topological space 33

constructed an open cover of C. There is a finite subcover, and then infact C = ] , a0[ ∪ ]b0, [. However, neither of the two intervals containsa0, which is a contradiction. Similarly, A has a greatest lower bound.

Now suppose that C is unconditionally complete and let U be anopen cover of C. We have to show the existence of a finite subcover.The smallest element s ∈ C belongs to some U ∈ U , and U contains aninterval [s, x[. A fortiori, this interval can be covered by finitely manysets from U , and the non-empty set X of all x′ sharing this property hasa least upper bound y. This element, too, belongs to an open interval I

contained in some set V ∈ U . That interval I is contained in X, becausewe can add V to some finite cover of [s, t[ for t ∈ I ∩ X. It followsthat y ∈ X and that y must be the greatest element of C, or else wewould have a contradiction to the definition of y. We have proved thatC = [s, y] is covered by finitely many sets from U . �

A topological space is said to be locally compact if every point hasarbitrarily small compact neighbourhoods.

5.3 Theorem A set A ⊆ R of real numbers is compact if, and only if,

it is closed and bounded. In particular, R is locally compact, but not

compact.

Proof For an unbounded set A, the sets A ∩ ]−n, n[, n ∈ N, form anopen cover without a finite subcover. If A is not closed and x is aboundary point not belonging to A, then the open cover formed bythe sets A � [x− 1/n, x + 1/n], n ∈ N, lacks a finite subcover. Thiscompletes the first half of the proof.

Conversely, let A be bounded and closed. Then A is a closed subsetof some interval I = [x, y], which is compact according to 5.2. Let U bean open cover of A. Each U ∈ U is the intersection of some open setU ′ ⊆ I with A, and the sets U ′, U ∈ U , together with I � A, form anopen cover U ′ of I. A finite subcover of U ′ yields a finite subcover of A

by taking intersections with A. �

Next we turn to the connectedness properties of R. A topologicalspace is said to be locally connected if every point has arbitrarily smallconnected neighbourhoods. In the following theorem, the term intervalrefers to all sorts of intervals, closed, open, and half open, bounded orunbounded. In particular, R itself is considered as an interval.

5.4 Theorem The connected subsets of R are precisely the intervals.

In particular, R is connected and locally connected.

Page 51: The Classical Fields

34 Real numbers

Proof Completeness of the chain R, together with strong density, impliesthat every interval is connected; see 3.3. Conversely, a set A that isnot an interval contains elements a, b separated by some x /∈ A, i.e.,a < x < b. Then A = (A ∩ ] , x[) ∪ (A ∩ ]x, [) is not connected. �

5.5 Connected components By definition, a connected componentof a topological space X is a connected subset that is maximal, i.e., notcontained in a bigger one. Since the closure of every connected set isconnected, all connected components are closed. Distinct components donot intersect, or else their union would be connected. Hence every spaceis the disjoint union of all its connected components. If X happens tobe locally connected, then the connected component containing a pointx contains every connected neighbourhood of x, hence the componentsof such a space are open sets. As a consequence, we obtain a descriptionof the open subsets of R.

5.6 Theorem Every open subset of R is a disjoint union of at most

countably many pairwise disjoint open intervals.

Proof An open subset U ⊆ R is locally connected, just like R itself.Therefore, all connected components of U are open sets. By 5.4, thesesets are (open) intervals, and they are pairwise disjoint (being connectedcomponents). Each of these intervals contains a rational number. Inview of their disjointness, this implies that there can be only countablymany of them. �

Characterizing the real line, the arc, and the circle

We remark first that the real line R is homeomorphic to any open inter-val. For instance, a homeomorphism

R ≈ ]−π, π[

is given by the function arctan. Apart from the real line, we shall beconcerned with some other spaces. The arc is the closed unit intervalI = [0, 1] ⊆ R or any space homeomorphic to it; also the half openinterval [0, 1[ will be of interest to us. The circle is the subset of R2

given by

S1 = { (x, y) ∈ R2 | x2 + y2 = 1}.Our aim is to characterize these spaces among all topological spaces,using the properties we have detected so far. Besides their apparent

Page 52: The Classical Fields

5 The real numbers as a topological space 35

similarity, we have more reasons to treat the spaces together. The in-tervals are subspaces of the real line, and, as we shall show later (5.16),the circle can be considered as a quotient space of R.

Some of the results we prove are ‘best possible’ in the sense that everyrelaxation of conditions leads to counter-examples; we shall demonstratethis in the next subsection. Other results are not best possible. Theneed to optimize the setup as a whole made it impossible to optimizeevery single result, so that in some cases there will be a superfluoushypothesis. There, we shall give references to the literature to make upfor the defects. Before we can prove theorems, we need some auxiliarynotions and results. For better understanding, the reader is advised tovisualize each of these in the cases X = R and X = [0, 1].

5.7 Definition A subset A in a connected topological space X is saidto be separating , if X �A is disconnected. In particular, we speak abouta separating point x when A = {x} is separating. Other points are saidto be non-separating, or we call them end points. We say that a point x

separates two other points if these lie in different connected componentsof X � {x}.

Every disconnected space admits a separation, that is, the space canbe represented as a disjoint union of two non-empty open subsets U, V .Somewhat loosely, we speak of ‘a separation X = U ∪ V ’. We note thatin a locally connected space, every connected component is open. Theunion of all other components is open, too, hence every component U ofa locally connected space X gives rise to a separation X = U ∪ V .

5.8 Lemma Suppose that X is a connected space, A ⊆ X is any subset,

and X � A = U ∪ V is a separation.

(a) If A is connected, then U ∪A is connected.

(b) If A is closed, then A contains the boundary of U relative to X.

(c) If A = {a} is a singleton and X is a T1-space, then U = U ∪ {a},and this set is connected.

Proof Suppose that A is connected and U ∪A = S ∪ T is a separation.The connected set A must be contained in S or in T , let us say, in S;then T ⊆ U . Therefore, T is closed and open in each of the followingspaces: in U ∪A, in U (which is closed and open in X �A) and hence inX � A, and finally in the union of these spaces, X; compare Exercise 1.This contradiction proves (a).

If A is closed, then U and V are open in X, and assertion (b) follows.Finally, (a) and (b) together imply (c). �

Page 53: The Classical Fields

36 Real numbers

The following shows a first glimpse of how the ordering of the real linereflects in its topology.

5.9 Lemma Consider two points x, y in a connected T1-space X, and

let X � {x} = Ux ∪ Vx and X � {y} = Uy ∪ Vy be separations such that

y ∈ Ux and x ∈ Uy. Then Vx ⊆ Uy and Vy ⊆ Ux, and no other inclusions

occur among these sets.

Proof By Lemma 5.8, the closure V x = Vx ∪ {x} is a connected set notcontaining y, hence it is contained in Uy or in Vy. Since x /∈ Vy, onlythe first possibility can occur. Likewise, Ux contains Vy. The inclusionis proper since y ∈ Ux � Vy, hence Ux is not contained in either of thetwo sets Uy, Vy. �

We are ready for the first characterization result. It is easy to dis-tinguish between the spaces ]0, 1[ ≈ R, [0, 1], and ]0, 1] (by the numberof non-separating points). Therefore, it is quite enough to characterizethe three spaces together, as is done in the following theorem due toKowalsky 1958. Other, similar results will follow.

5.10 Theorem A topological space X is homeomorphic to one of the

spaces ]0, 1[ ≈ R, [0, 1], or ]0, 1] if, and only if, it has at least two points

and has the following four properties.

(i) X is separable.

(ii) X is a T1-space, that is, all points of X are closed.

(iii) X is connected and locally connected.

(iv) Among any three connected proper subsets of X, there are two

which fail to cover X.

Proof (1) The necessity of the first three conditions has already beenshown (5.1, 5.4). A connected proper subset of any of the three typesof intervals is a subinterval which has a non-trivial bound on at leastone side. Among any three such subintervals, some two are boundedon the same side and, hence, fail to cover the entire interval. In theremainder of this proof, we shall be concerned with the sufficiency ofthe four conditions.(2) For each x ∈ X, the closure of any connected component C of X�{x}equals C ∪ {x} and is connected.

By (ii), the complement X � {x} is open, and we may assume thatC �= X � {x}. By local connectedness, there is a separation X � {x} =C ∪D. The assertion now follows from 5.8c.(3) The complement of a point in X has at most two connected compo-nents.

Page 54: The Classical Fields

5 The real numbers as a topological space 37

Here we use condition (iv). If X�{x} has more than two components,we choose C1, C2 among them and denote the union of all the othersby Y . Then we obtain three proper subsets, any two of which cover X,by taking X � C1, X � C2, and X � Y . By 5.8c, each of these sets isconnected, contrary to (iv).

Applying (iv) to three connected point complements, we obtain thenext observation.(4) At most two points have connected complements.

In particular, there are separating points, because the connected T1-space X cannot have only two points.(5) The connected components of X � {x} will be called x-parts forbrevity. For a separating point x, the two x-parts form a separationof X � {x}. If x is an end point, then also {x} will be considered asan x-part, so that each point x defines precisely two x-parts. If x, y ∈X are two distinct points, then we denote the x-part containing y byT (x, y), and the other x-part by A(x, y). The letters T and A here standfor ‘pointing towards y’ and ‘pointing away from y’, respectively. Theinclusion relations among these sets will eventually lead to an orderingrelation on X. The following holds trivially for end points and followsfrom 5.9 in the case of separating points.(6) If x, y ∈ X are distinct points, then A(x, y) ⊆ T (y, x) and A(y, x) ⊆T (x, y), and no other inclusion relations hold among the x-parts and they-parts.(7) Two parts defined by distinct points will be called compatible if oneof them contains the other. (6) may be rephrased by saying that, givenany x-part P and a point y �= x, there is always precisely one y-partcompatible with P . We choose an arbitrary point e ∈ X and an e-part Pe. For every other point x, we let Px denote the unique x-partcompatible with Pe. We claim:(8) The parts Px are all mutually compatible.

For x, y, e distinct from each other, we have to show that Px and Py

are compatible. We know that both of them are compatible with Pe. IfPx ⊆ Pe ⊆ Py, then no argument is needed. The difficult cases are thatPx and Py are both contained in, or both contain, Pe.

We assume that none of Px, Py contains the other, and we look atthe two sets M1 = Px ∪ P �

y and M2 = Py ∪ P �x , where Y � denotes the

complement of Y ⊆ X. Both sets are non-empty and connected (forinstance, Px must meet P �

y ). Moreover, M1 is a proper subset of X, orelse we have Py ⊆ Px. Likewise, we also obtain that M2 is a propersubset.

Page 55: The Classical Fields

38 Real numbers

If Px, Py ⊆ Pe, then the three sets M1, M2 and Pe contradict condi-tion (iv). If Pe ⊆ Px, Py, then the same is true for M1, M2 and P �

e , andthis proves (8).(9) We have shown that the set P of all parts Px, x ∈ X, is totallyordered by inclusion. We have a bijection x �→ Px from P onto X

(injectivity follows from (2)), and hence we can carry the total orderingto X:

x ≤ y � Px ⊆ Py .

From (6) it follows that x < y ⇔ y /∈ Px ∪ {x} ⇔ x ∈ Py � {y}.(10) The topology of X coincides with the topology obtained from theordering ≤.

The order topology is generated by the intervals of type ]x, [ and ] , x[,where x ∈ X is a separating point. By (9), these intervals coincide withthe connected components of the complement X � {x}, and these areopen sets of X because X is a locally connected T1 space. This provesthat the identity is a continuous map from X with the given topologyto X with the order topology.

Conversely, we show that every open neighbourhood U of x ∈ X

with respect to the given topology contains a neighbourhood of x withrespect to the order topology. We may assume that U is connected. Bycontinuity of the identity map, U is also connected with respect to theorder topology. In every order topology, a connected set must be aninterval. Suppose that the interval U contains one of its end points, y.If y �= x, then we may replace U by U � {y}. If U = ]a, x] and x is notan end point of X, then X = ] , x] ∪ ]x, [ is disconnected. This showsthat U is open in the order topology and proves step (10).(11) It remains to determine the isomorphism type of X as a chain.Here we use the separability of X for the first time. Any end points ofX will be deleted at first. After the remainder of X has been identifiedas ]0, 1[, the end points may be restored as smallest or largest elementsof the chain.

We know that the order topology is connected, and we shall show thatthe chain X is strongly separable. Then condition (c) of the character-ization theorem 3.5 is satisfied and our proof is complete. Separabilityof the order topology implies that every non-empty open interval con-tains an element of some countable set A ⊆ X. By connectedness, andbecause we deleted end points, the open intervals ] , x[, ]x, [ and ]x, y[for x < y are non-empty and contain elements of A. This means strongseparability of the chain X. �

Page 56: The Classical Fields

5 The real numbers as a topological space 39

The conditions characterizing R according to the last theorem requireinformation about every connected subset of X, and this may be askinga bit much. We therefore give now another characterization that usesinformation on the connected components of point complements only.This result was proved directly in the lecture notes Salzmann 1971,without using Kowalsky’s theorem 5.10.

5.11 Theorem A topological space X is homeomorphic to one of the

spaces ]0, 1[ ≈ R, [0, 1], or ]0, 1] if, and only if, it has at least four points

and has the following four properties.

(i) X is separable.

(ii) X is a T1-space, that is, all points of X are closed.

(iii) X is connected and locally connected.

(iv′) Among any three points of X, there is one which separates the other

two (see 5.7).

Proof We show that a space X with these four properties satisfies theconditions of 5.10.

(1) The closure of any connected component C of X � {x} is C =C∪{x}, as in step (2) of the last proof. From (iv′) it follows immediatelythat there are at most two non-separating points. Next, we prove thatX � {x} has at most two components. Indeed, if C1, C2, C3 are distinctcomponents, then no point x1 ∈ C1 can separate x2 ∈ C2 from x3 ∈ C3,because these points belong to the connected set C2 ∪ C3, which doesnot contain x1.

(2) The last fact allows us to introduce the same notation A(x, y)and T (x, y) as in the last proof (step (5)). We also have the propertyA(x, y) ⊆ T (y, x) obtained from 5.9. We conclude that the only possi-bility to cover X by an x-part together with a y-part is to use T (x, y)together with T (y, x).

(3) We show that X satisfies condition (iv) of 5.10. A proper con-nected subset A of X is contained in an open x-part, where x /∈ A.Hence we may assume that the three sets appearing in condition (iv)are three open parts with respect to points x, y, z. Two open parts withrespect to the same point do not cover X, hence we may assume thatwe have three distinct points. If condition (iv) is violated, then step (2)shows that each of the three given parts contains both of the pointsdefining the other parts. But by (iv′), one of the points, say y, separatesthe other two, i.e., x and z lie in different parts with respect to y. Thiscontradiction completes the proof. �

Page 57: The Classical Fields

40 Real numbers

If compactness is assumed, then the separation conditions (iv) or (iv′)can be weakened considerably. This is based on the following auxiliaryresult, taken from Christenson–Voxman 1977.

5.12 Lemma Suppose that p is a separating point of a compact, con-

nected T1-space X and let X �{p} = U ∪V be a separation. Then both

U and V contain at least one non-separating point of X.

Proof We assume that all points of U are separating and aim for acontradiction. For each x ∈ U , choose a separation (Ux, Vx) such thatp ∈ Vx. By 5.9 we have Ux ⊆ U . The set U = {Ux | x ∈ U } ispartially ordered by inclusion. By Hausdorff’s maximal chain principle(or equivalently, by Zorn’s lemma, see the beginning of Section 61),there exists a maximal subchain W ⊆ U . We shall show (by an indirectargument) that

⋂W = ∅ and then derive a final contradiction from thatresult.

We set W = {x ∈ U | Ux ∈ W } ⊆ U . For z ∈ ⋂W and w ∈ W ,we have z ∈ Uw ⊆ U . Hence, we may form Uz, and we shall show thatUz ⊆ Uw � {z}. This tells us that W ∪ {Uz} is a chain larger than W,contrary to maximality. We know that z ∈ Uw, whence the connectedset Vw∪{w} is contained in X �{z} = Uz∪Vz. The former set meets Vz

(in p), hence it is contained in Vz and, in particular, w ∈ Vz. ThereforeUz ∪ {z} is contained in X � {w} and meets Uw (in z), so that finallyUz ⊆ Uw � {z}.

We have just seen that⋂W = ∅, and we deduce next that the chain of

closed sets {Uw ∪ {w} | Uw ∈ W } has empty intersection, too — whichcontradicts compactness. An element belonging to that intersection canonly be a point w ∈W . The set Uw is not a smallest element of W, andthere is a proper subset Uv of Uw. Then w /∈ Uv ∪ {v}, and our chain ofclosed sets has empty intersection. �

5.13 Theorem A topological space X of cardinality ≥ 2 is homeo-

morphic to the arc [0, 1] if, and only if, X is compact, T1, separable,

connected and locally connected, and X contains at most two non-

separating points.

Proof From Lemma 5.12, we deduce that (i) there are exactly two non-separating points a, b ∈ X, and that (ii) every other point x ∈ X �{a, b}separates a from b. It follows that no proper connected subset cancontain both a and b, hence the space has property (iv) and the assertionfollows from 5.10. �

Page 58: The Classical Fields

5 The real numbers as a topological space 41

5.14 Theorem A topological space X of cardinality ≥ 2 is homeo-

morphic to the circle S1 if, and only if, X is compact, T1, separable,

connected and locally connected, and every pair of distinct points sepa-

rates X.

Proof (1) According to Lemma 5.12, there exist two non-separatingpoints p, q ∈ X. Let A be a connected component of X � {p, q}. SinceX is locally connected, there is a separation (A,B) of X �{p, q} = (X �

{p}) � {q}. Using Lemma 5.8a, we conclude that A ∪ {q} is connected.Likewise, A ∪ {p} and, hence, A = A ∪ {p, q} are connected.

(2) If A is separated by each point a ∈ A, then there is a homeomor-phism [0, 1]→ A sending {0, 1} to {p, q}. We cannot use 5.13 directly inorder to show this, as A might not be locally connected at p or at q. ButA itself is connected and locally connected, and each a ∈ A separates p

from q; this follows from 5.12, applied to the space A, which is compactby 5.8. Therefore, no proper connected subset of A accumulates at bothp and q, hence A has property (iv), and 5.10 shows that A ≈ ]0, 1[.

We identify A with ]0, 1[, and we show next that p and q are attachedto A in the right way. We have seen that for each a ∈ ]0, 1[, there is an(open) separation A � {a} = U ∪ V such that p ∈ U and q ∈ V . Eachconnected component of ]0, 1[�{a} is contained in U or in V . Hence, wemay adjust the notation such that U = {p} ∪ ]0, a[ and V = ]a, 1[∪ {q}.Therefore, the standard neighbourhoods of 0 and 1 correspond to opensets of A. Conversely, we note that K = A�U = [a, 1[∪{q} is compact.If W is an open neighbourhood of q, then K � W is a compact subsetof [a, 1[, and hence W contains a standard neighbourhood.

(3) If B is a connected component of X � {p, q} that is not separatedby some b ∈ B, then all other connected components C satisfy theassumptions of step (2), or else there is a point c ∈ C such that C � {c}is connected; denoting by E the union of all sets D obtained from theremaining components, we infer that X�{b, c} = (B�{b})∪(C�{c})∪E

is connected, contrary to our hypothesis. It follows that C ≈ [0, 1] in thiscase. However, this implies that the connected components of C � {c}all intersect B � {b} (in p or in q), and that X � {b, c} is connected.This is impossible.

(4) We have shown in (3) that B ≈ [0, 1] for each connected compo-nent of X � {p, q}, where the points 0, 1 correspond to p, q. Clearly,the hypothesis implies that there are precisely two such components,and mapping each of the two intervals onto one half of the circle in acompatible way we obtain a homeomorphism X → S1. �

Page 59: The Classical Fields

42 Real numbers

5.15 Remarks and variations The last two results hold for metricspaces without the assumption of local connectedness. Note that a com-pact space is metrizable if, and only if, it is a Hausdorff space with acountable basis for its topology. In this form, 5.13 was given by Moore

1920; see Christenson–Voxman 1977 9.A.8 for a proof in modern lan-guage. The methods are very similar to the ones used here in the proofof 5.10, but it seems to be impossible to obtain all these results simulta-neously. The variation of 5.14 without local connectedness is an exercisein Christenson–Voxman 1977. Our proof is more complicated sincewe need local connectedness when we apply 5.13.

Kowalsky 1958 proves a weak form of 5.13. A characterization of thereal line in the spirit of 5.13 is given by Ward 1936. He requires thatthe complement of any point has exactly two connected components,and that the whole space is metric, separable, and locally connected.This result is certainly not true without the condition of metrizability;compare Example 5.23. The real line has also been characterized interms of the cardinalities of the boundaries of connected open subsets;see Jones 1939. His result implies Kowalsky’s theorem 5.10 in the caseof Hausdorff spaces.

In the context of manifolds, we shall now present another characteri-zation of R and of the circle.

5.16 The circle as a quotient of R There is a surjective map p :R → S1 defined by p(t) = (cos 2πt, sin 2πt). If we identify R2 with C,the set of complex numbers, via (x, y) �→ x + iy, then we may writep(t) = e2πit. The map p relates the topologies of R and S1 in a veryefficient way. We say that p is an identification map or that S1 carriesthe quotient topology with respect to p, and this means that a subset ofS1 is open if, and only if, its inverse image under p is open in R. The‘only if’ part of this statement expresses the well known continuity of p.The ‘if’ part follows from the fact that the restriction of p to any openinterval of length 1 is a homeomorphism onto its (open) image, as canbe seen from the existence of complex logarithms.

This last fact also shows that S1 is a one-dimensional manifold (or1-manifold for short). A Hausdorff space is said to be an n-manifoldif each point has an open neighbourhood homeomorphic to ]0, 1[n (or,equivalently, to Rn). We shall prove the following.

5.17 Theorem Every separable, connected one-dimensional manifold

is homeomorphic to R or to S1.

Page 60: The Classical Fields

5 The real numbers as a topological space 43

We remark that separability is essential. In 5.25 below, we shall in-troduce the ‘long line’, a space which is a non-separable 1-manifold (see5.26). The proof of 5.17 requires some auxiliary notions and results; itwill be given after 5.20. First recall that an arc in a topological spaceX is a subset A ⊆ X homeomorphic to the closed interval [0, 1]. Thenon-separating points of A (corresponding to 0 and 1) are called the endpoints of the arc.

5.18 Lemma Suppose that a topological space X contains an arc A

with end point a and an arc B with end point b �= a. If the intersection

of the two arcs is non-empty, then their union contains an arc with end

points a, b.

Proof Let f : [0, 1] → A be a homeomorphism which sends 0 to a. Bycompactness, there is a smallest t ∈ [0, 1] such that f(t) ∈ B. Thenthe subarc f([0, t]) together with the unique subarc C of B containingf(t) and b form the desired arc from a to b. To show that we really getan arc, we replace f on the interval [t, 1] with a parametrization of C

running from f(t) to b. — This argument will be used tacitly in thefuture. �

5.19 Definition A topological space X is said to be arcwise connectedif every two distinct points x, y ∈ X are ‘joined’ by an arc A ⊆ X

with end points x, y. In general, the maximal arcwise connected subsetsof X are called the arc components of X. It follows from 5.18 thattwo arc components are either disjoint or equal. If a space is locallyarcwise connected (every point has arbitrarily small arcwise connectedneighbourhoods), then every arc component of this space is open. Theunion of the other arc components is open, as well, hence arc componentsare open and closed in this case. In particular, a connected, locallyarcwise connected space is arcwise connected.

5.20 Lemma: Domain invariance property of 1-manifoldsA 1-manifold Y contained in another 1-manifold X is always open.

Proof Every point y ∈ Y has a neighbourhood U ≈ R in X, and U

contains a neighbourhood V ≈ R with respect to Y . The connectedsubset V ⊆ U is an interval by 5.4; more precisely, an open interval as itis homeomorphic to R. This shows that y is an interior point of Y . �

We are now ready to prove 5.17. We shall invoke the characterizationtheorem 5.11 in this proof. This will save us some of the technicallabours involved in a direct proof such as Christenson–Voxman 1977

Page 61: The Classical Fields

44 Real numbers

Chapter 5A. In particular, separability is exploited in a very simple way.We believe that our proof nonetheless gives appropriate insight.

Proof of 5.17. Every manifold is locally arcwise connected, hence thegiven space X is arcwise connected by 5.19. We distinguish between twocases.

Case I: There are points u, v ∈ X that can be joined by two distinctarcs A,B. Choose a point lying on only one of the arcs, say x ∈ A � B.Then A contains exactly two minimal arcs joining x to points b, c ∈ B.These two arcs have only the point x in common. Together with thesubarc of B joining b to c, they form a subset Y ⊆ X homeomorphicto S1. The compact 1-manifold Y is closed in X and also open by 5.20,hence X = Y ≈ S1.

Case II: Every pair of distinct points x, y ∈ X is joined by a uniquearc having end points x, y. We know that X has all the characteristicproperties of R required in 5.11, with the possible exception of the sep-aration property (iv′). We end our proof by showing that X does havethe separation property. Observe that the other spaces characterized bythese conditions (closed and half open intervals) are not manifolds.

Suppose that an arc A in X joins the points a, b, and that x is a thirdpoint on the arc. Then we claim that x separates a from b. If not,then a, b belong to the same connected component U ⊆ X � {x}. Weknow that U is connected, open in X, and locally arcwise connected.According to 5.19, there is an arc in U joining a to b, contrary to theassumption of Case II.

Now let three points x, y, z ∈ X be given. If one of them lies on thearc joining the other two, then we can finish our proof by applying theprevious paragraph. Assume therefore that this does not happen andconsider the two arcs A and B joining x to y and to z, respectively.According to 5.18, the union A∪B contains an arc C joining y to z, andC � {y, z} is open in X by 5.20. But, following the arc A from x, theremust be a first point where we hit C, and this point is distinct from y

and z. This is a contradiction, and the proof is complete. �

Independence of characteristic properties

Our next aim is to prove independence of the conditions characterizing Raccording to 5.11. To do this, we shall exhibit examples of spaces thathave three of the properties but not the fourth. Moreover, in all ourexamples the complement of any point will have exactly two connected

Page 62: The Classical Fields

5 The real numbers as a topological space 45

components. For instance, we are not satisfied with X = R2 as anexample of a space violating only the separation property (iv′).

It is relatively easy to give an example Xiii that is (connected but) notlocally connected, and has all the other properties. One takes the set ofreal numbers and defines the space Xiii by refining the topology of R asfollows. A set A is open by definition if, and only if, each point a ∈ A hasa neighbourhood U in the usual topology of R such that U∩Q ⊆ A. Theverifications are left to the reader. Compare also counter-example 68 inSteen–Seebach 1978.

A space Xii without the T1 property but satisfying the other condi-tions is obtained from the subspace ] ,−1] ∪ {0} ∪ [1, [ ⊆ R by a slightmodification of the topology. The change consists in the omission ofcertain open sets, namely, those that contain one of the points 1 or −1without containing 0. In particular, {0} is not a closed subset. Theother verifications are again left to the reader.

The remaining examples are more interesting and will be consideredin detail. In order to construct a space that violates only the separa-tion property (iv′), we begin by constructing an auxiliary space thatis countable but resembles the circle in many respects. This space istaken from the beautiful book of topological counter-examples Steen–

Seebach 1978 (counter-example 61). Originally, it was described byGolomb 1959 and Kirch 1969.

5.21 Example We construct a topological space with point set N, thenatural numbers (0 excluded). In order to do so, we specify a basisB for the topology, i.e., the open sets will be the unions of arbitrarilymany sets from B. There is a condition that we have to observe if this issupposed to work. The intersection of two basis sets has to be a unionof other basis sets. Our basis consists of all sets of the form N∩ (a+bZ),where a, b ∈ N are relatively prime, that is, gcd(a, b) = 1, and b is squarefree, i.e., no square > 1 divides b. It is not necessary to make use of thepossibility b = 1. Then we can always assume that a < b, and write thecorresponding basis set as a + bN0, where N0 = N ∪ {0}. Let us notethat every element of this set is relatively prime to b.

We verify the condition for a basis in the case of two basis sets a+bN0

and c+dN0. We consider a point x in the intersection, x = a+bu = c+dv

and look for a basis set containing x and lying within the intersection.Such a set is given by N∩(x+eZ), where e is the least common multiple ofb and d, which is square free. The choice of x implies that x is relativelyprime to both b and d and, hence, to e.

Page 63: The Classical Fields

46 Real numbers

5.22 Proposition The space N with the topology constructed above

is separable (being countable), connected and locally connected, and is

a Hausdorff space, but not metrizable. Moreover, every point in this

space has a connected complement.

Proof We show first that the basis sets A = a + bN0 ∈ B are connected.Thus we assume that A = S ∪ T where S, T are open and non-empty,and we show that S ∩ T �= ∅. There are basis sets s + bcN0 ⊆ S andt + bdN0 ⊆ T ; the special form is enforced since these sets are containedin A. We set u = cd; then b and u are relatively prime, because both bc

and bd are square free. By 1.5, there are integers f, z such that

a + bz = uf .

This yields a + b(z + nu) = u(f + nb) for every n ∈ N, and we see thatwe may choose f and z to be positive. The displayed equation showsthat uf belongs to A and, hence, to S or to T . The roles played byS and T in the construction of this element are interchangeable, henceit does not matter which of the two sets it belongs to – say uf ∈ T .There exists a neighbourhood uf + bgN0 ⊆ T of this element, where1 = gcd(uf, bg) = gcd(cdf, bg) = gcd(c, g). Since s and uf belong to A,we have

uf − s = be ∈ bZ .

We use 1.5 again to write cx = e+gy, where x, y ∈ Z; as before, we maytake both these integers to be positive. We finally obtain the elements + bcx = s + be + bgy = uf + bgy ∈ S ∩ T .

We may conclude that N is locally connected, and we show next thatevery point z ∈ N has a connected complement. As a by-product, thiswill imply that N itself is connected and, hence, not metrizable (notethat the distance function on a connected metric space has to take acontinuum of values). Given x, y ∈ N � {z}, we propose to constructtwo basis sets A = N ∩ (x + pZ) and B = N ∩ (y + qZ) not containing z

such that A ∩ B �= ∅. Then A ∪ B is connected and contains x and y,proving that N � {z} is connected. At the same time, we will obtain theHausdorff property (just knowing that x ∈ A ⊆ N � {z}).

We choose p and q to be distinct primes such that p > |x − z| andq > |y − z|. This is possible because there are infinitely many primes(see 32.7), and it ensures that z /∈ A ∪ B. Once more by 1.5, there arepositive integers m,n such that x − y = qm − pn, and then x + pn =y + qm ∈ A ∩B. �

Page 64: The Classical Fields

5 The real numbers as a topological space 47

5.23 Example We construct a new space Xiv (looking like a brush)from the ‘countable circle’ (N with the topology defined in 5.21) byattaching a bristle to each of its points. Formally, Xiv = N× [0, 1[ withthe topology defined by the following basis. For n ∈ N and an open setU ⊆ ]0, 1[ (not containing 0 !), the set {n}×U is a basis set, and for anopen set A ⊆ N and ε > 0, the set A × [0, ε[ is another one. The proofthat this is the basis of a topology having the properties stated in thenext proposition is left to the reader.

5.24 Proposition The space Xiv constructed above is a connected

and locally connected, separable Hausdorff space. The complement of

any point has exactly two connected components. However, a point in

N × {0} can never separate two other points belonging to that subset.

Our next aim is to construct a space Xi having all characteristic prop-erties of R except separability. In the proof of Theorem 5.10, only thelast step made use of separability. Therefore, the space we are lookingfor must be a non-separable chain endowed with the order topology. Thestarting point of the construction is the observation that Z× [0, 1[ withthe lexicographic ordering is a chain isomorphic to R, and one tries toreplace Z with an uncountable chain that is not too big.

For this purpose, we use the first uncountable ordinal ω1. For informa-tion on ordinals, see the Appendix 61 and 61.7 in particular. We needthe properties that ω1 is an uncountable well-ordered chain and thata subset of this chain is countable if, and only if, it is bounded. Theverification of the latter property is the only occasion where we need toknow that every ordinal can be identified, somewhat paradoxically, withthe set of all smaller ordinals (61.6), and that every set M of ordinalshas a least upper bound in the ordered class of all ordinals, obtainedsimply as the union

⋃M (61.5). The elements of ω1 are precisely allcountable ordinals (61.7), and it follows that every bounded set in ω1 iscountable. Conversely, if we have a countable set of countable ordinals,then their union is again a countable set (61.13) and an ordinal, henceit is an element of ω1. Henceforth, we shall ignore the original meaningof the ordering relation for ordinals and write < for this relation.

5.25 Example: Alexandroff’s long line The long line L will becomposed of the long ray L+ and its mirror image −L+, the detailsbeing as follows. The long ray is defined as the chain

L+ = ω1 × [0, 1[ ,

Page 65: The Classical Fields

48 Real numbers

with the lexicographic ordering relation

(μ, s) < (ν, t) � (μ < ν) ∨((μ = ν) ∧ (s < t)

).

For brevity, we shall write 0 instead of (0, 0) ∈ L+. For each x ∈ L+,we introduce a mirror image −x, such that x �→ −x becomes a bijectivecorrespondence L+ → −L+ = {−x | x ∈ L+ }. The sets L+ and −L+

are supposed to be disjoint except for 0 = −0. For x < y in L+ we set−y < −x < y and thus turn L = L+ ∪ −L+ into a chain. This chain isthe long line. Both L and L+ are endowed with the topology induced bytheir ordering, and both L and L+ � {0} can serve as our example Xi ofa space having all the characteristic properties of R except separability,as it will turn out later.

The involutory map ι : L → L that interchanges x and −x is ananti-automorphism of the chain L and hence a homeomorphism of thelong line. However, appearances can be deceptive, and if ν + 1 denotesthe successor of an ordinal, then the map (ν, s) �→ (ν + 1, s) from L+

into itself is not continuous. Indeed, if ν runs over all natural numbers,then both (ν, 0) and (ν + 1, 0) converge to (ω0, 0), but (ω0, 0) is notfixed by the map. For similar reasons, we refrained from describing allof L by a single lexicographic product; this would not allow for a simpledescription of the map ι.

The long line is a more prominent example than those discussed so far.It was first presented in Alexandroff 1924. The existence of severalnon-isomorphic analytic structures both on L and on L+ was shownin Kneser–Kneser 1960. More recently, Nyikos 1992 constructed2ℵ1 smooth structures on each of them, whose tangent bundles are nothomeomorphic.

5.26 Proposition Every interval [a, b] in the long line L is order iso-

morphic and, hence, homeomorphic to the unit interval [0, 1] ⊆ R.

Proof We treat the case a = 0 in detail. After that one sees, using theorder isomorphism ι introduced in 5.25, that [−b, b] = [−b, 0] ∪ [0, b] ∼=[−1, 0] ∪ [0, 1] ⊆ R, and the proposition follows. First we obtain an iso-morphism of the open intervals, ]0, b[ ∼= ]0, 1[, using the characterizationof the chain R given in 3.5b. Such an isomorphism uniquely extends tothe closed intervals.

If b = (ν, t), then the interval [0, ν] of ordinals (that is, the ordinalν + 1) is countable, and ([0, ν] × Q) ∩ ]0, b[ is a countable, stronglydense subset of ]0, b[. It remains to be shown that the given interval iscomplete.

Page 66: The Classical Fields

5 The real numbers as a topological space 49

So let a bounded subset A ⊆ ]0, b[ be given, A ≤ (η, s) < b, andconsider the projection p1 : A → ω1 : (ν, t) �→ ν. The well-ordering ofω1 implies that p1(A) has a least upper bound ξ ≤ η in ω1. If ξ /∈ p1(A),then the least upper bound of A is supA = (ξ, 0) ≤ (η, s). If ξ ∈ p1(A),then we look for supA in the complete interval [(ξ, 0), (ξ + 1, 0)] ∼= [0, 1]and find supA ≤ (η, s). Greatest lower bounds are treated in the samemanner. �

5.27 Definition A topological space X is said to be n-homogeneous,n ∈ N, if any sequence x1, . . . , xn of n distinct points of X can bemapped onto any other such sequence, y1, . . . , yn, by a homeomorphismof f : X → X. To be precise, it is required that f(xi) = yi for each i.

The real line R is 2-homogeneous (but not 3-homogeneous): usingaffine maps x �→ ax + b, every pair of real numbers can be mapped ontoevery other pair (but a homeomorphism cannot move the middle pointof a triple to a position where it does not separate the images of theother two). We show that the long line shares this property of R.

5.28 Corollary The long line L is 2-homogeneous.

Proof Let two pairs of points x1, x2 and y1, y2 be given as in 5.27. Byapplying the map ι from 5.25, if necessary, to one or to both pairs, wemay arrange that x1 < x2 and y1 < y2. We find an open interval I ⊆ Lcontaining the four points. By 5.26, this interval I is homeomorphic tothe real interval ]0, 1[ ≈ R. The remarks preceding the corollary showthat I admits a homeomorphism f sending xi to yi, and f is monotonedue to our special arrangement of the two pairs of points. It follows thatf extends to a homeomorphism of L that induces the identity on thecomplement of I. To solve the original task, this map f may have to bemultiplied on one side or on both by the map ι. �

5.29 Corollary Both the long line and the long ray satisfy all condi-

tions of 5.11 except separability.

Proof The two cases are very similar, and we concentrate on the longline itself. Every chain is a Hausdorff space in the induced topology.The long line is the union of all its intervals [−a, a], which are connectedby 5.26. This shows connectedness of L, and local connectedness is evenmore obvious. The separation condition (iv′) again holds in every chain.We note that, as with our other counter-examples, the complement ofany point has exactly two connected components. Finally, L is notseparable; this follows from the following stronger assertion. �

Page 67: The Classical Fields

50 Real numbers

5.30 Proposition Every countable subset of the long line is bounded.

Proof This is an immediate consequence of the fact that ω1 has the sameproperty; see the remarks preceding 5.25. �

The following is perhaps the most striking feature of the long line.

5.31 Proposition If f : L → R is a continuous map, then there are

a, b ∈ L such that f is constant on ] , a] and on [b, [ .

Proof We concentrate on the restriction of f to L+. For a set X of realnumbers, we define the diameter diam X = sup{ |x − y| | x, y ∈ X } ∈R∪{∞}. Our first claim is that for each n ∈ N we can find cn ∈ L suchthat diam f([cn, [) ≤ 1/n. Suppose that this is impossible. Then wecan find a positive real δ and an increasing sequence {ak | k ∈ N} in L+

such that |f(ak)− f(ak+1)| ≥ δ for each k. The sequence is bounded by5.30, hence it is contained in a subchain isomorphic to [0, 1] by 5.26. Inthis subchain, every increasing sequence converges, but f(ak) does notconverge — a contradiction to continuity.

The sequence { cn | n ∈ N} has an upper bound c ∈ L, too, andwe conclude that diam f([c, [) ≤ diam f([cn, [) ≤ 1/n for each n. Thismeans that f is constant on [c, [. �

Properties 5.30 and 5.31 of the long line are actually compactnessproperties. We make that explicit after the following.

5.32 Definition A topological space X is said to be pseudo-compactif every continuous map f : X → R produces a compact image f(X).The space X is said to be sequentially compact if every sequence in X

has a convergent subsequence.For further reading on notions of compactness, we recommend the

book Christenson–Voxman 1977. We mention that a compact spaceis both sequentially compact and pseudo-compact, and that a sequen-tially compact space is countably compact , that is, every countable opencover {Ui | i ∈ N} has a finite subcover. Indeed, if the first n sets Ui

fail to cover, we select a point xn that is not covered. If this happens forevery n, there results a sequence, which has a convergent subsequence— but the limit cannot belong to any Ui, a contradiction.

5.33 Proposition The long line L is not compact, but it is locally com-

pact, sequentially compact, countably compact, and pseudo-compact.

Proof The long line is covered by all its open intervals ]a, b[, and thiscover has no finite subcover. Thus, L is not compact. On the other hand,

Page 68: The Classical Fields

5 The real numbers as a topological space 51

every interval [a, b] is compact according to 5.26. This implies localcompactness of L. By 5.30, every sequence is contained in a compactinterval and, hence, has an accumulation point. Countable compactnessfollows from sequential compactness (5.32). For a continuous real-valuedmap f , we learn from 5.31 that some compact interval has the sameimage as all of L, which implies peudocompactness. �

5.34 Corollary There is no metric generating the topology of the long

line.

Proof In metric spaces, sequential compactness is the same as compact-ness; see, for example, Christenson–Voxman 1977 3.A.11 p. 83. Moreeasily we can show that a sequentially compact metric space is separa-ble and then apply 5.29. So suppose that X is a sequentially compactspace. If, for every n ∈ N, there is a finite subset Fn ⊆ X such thatevery point of X lies at distance < 1/n from some member of Fn, thenthe union of all Fn is a countable (61.13) dense set. If no finite set hasthe properties required for some particular Fn, then by induction we finda sequence {xk | k ∈ N} such that each xk is at distance ≥ 1/n from allits predecessors. Such a sequence has no convergent subsequence. �

As a last remark on the long line, we note that it does not satisfy theSouslin condition (compare 3.8): there are uncountably many disjointopen intervals {ν} × ]0, 1[ in L.

Subspaces and continuous images of the real line

We recall that one important continuous image of the real line is thecircle; see 5.16. Several characterizations of the circle have been given(5.14, 5.17). Among subsets of the real line, a class that has receivedspecial attention is formed by the totally imperfect subsets, i.e., setsthat do not contain any uncountable closed set. We shall not treat thistopic but refer to the literature (Miller 1984, 1993).

We continue by discussing the Cantor set (which is closed and un-countable). At first sight, this looks like a rather far-fetched examplehaving weird properties. However, it turns out that the Cantor set is ofcrucial importance in our context. We show that it can be used in theconstruction of space-filling curves, and this is closely related to the roleit plays in the proof of the Hahn–Mazurkiewicz theorem characterizingthe continuous images of the arc. Furthermore, the continuous imagesof the Cantor set itself form an important class of topological spaces.

Page 69: The Classical Fields

52 Real numbers

In order to define the Cantor set, we use the triadic expansion of realnumbers x ∈ [0, 1] :

x =∑∞

n=1 an3−n ,

where an ∈ {0, 1, 2}. This expansion is not unique; typically, 0.10 =0.02, where we write, as in the usual decimal notation, x = 0.a1a2a3 . . .

for the above expansion, and periods are indicated by overlining.

5.35 Definition The Cantor set C ⊆ R is defined as the set of allx ∈ [0, 1] admitting a (unique) triadic expansion using only 0 and 2 asdigits. The following alternative description may be more familiar; weleave it to the reader to prove the equivalence with our first description.

One starts with the unit interval I = I0 and forms I1 = I �]13, 2

3

[=

[0, 13 ]∪ [23 , 1]. In other words, I1 is obtained by deleting the open middle

third from the unit interval. We continue in the same way, formingIn+1 by deleting the open middle thirds from each of the 2n intervalsconstituting In. This gives a descending chain of compact subsets of I,and the (compact, non-empty) intersection of this chain is the Cantorset. Since the intervals forming In are of length 3−n, we see immediatelythat the total length ( 2

3 )n of In tends to zero, and we have obtained twoimportant features of C (compare 10.6):

5.36 Proposition The Cantor set C ⊆ [0, 1] is a compact set of

Lebesgue measure zero. �

Note, however, that the measure is not an invariant property of thetopological space C; compare 10.6(c). Compactness follows once morefrom our next result, where the space {0, 2} is taken with the discretetopology.

5.37 Proposition The Cantor set is homeomorphic to the compact

product space {0, 2}N.

Proof We show that a homeomorphism is obtained by reading the co-efficient sequence of the triadic expansion with digits an ∈ {0, 2} as anelement of the product space. The product topology is defined using abasis, where a typical basis element consists of all sequences (ak)k witha fixed initial part a1, . . . , an. The corresponding points of I form theopen and closed set C ∩ (

∑nk=1 ak3−k + [0, 3−n]). This proves that we

have a homeomorphism. Since {0, 2} is a compact space, it follows thatthe product space is compact, as well. �

5.38 Corollary The Cantor set C is homeomorphic to its own square

C × C.

Page 70: The Classical Fields

5 The real numbers as a topological space 53

Proof There is a homeomorphism {0, 2}N → {0, 2}N×{0, 2}N defined by‘opening the zipper’, that is, the sequence a1, b1, a2, b2, . . . is mapped tothe pair of sequences (a1, a2, . . . ; b1, b2, . . . ). �

5.39 Proposition There is a continuous surjection C → I = [0, 1].

Proof Such a map may be defined by reading the coefficient sequence ofa triadic expansion as the coefficient sequence of a dyadic one, replacingthe digit an by an/2. �

5.40 Proposition There is a continuous surjection I → I2 of the arc

onto a square.

Proof We use the homeomorphism ϕ : C → C2 of 5.38 and the surjectionρ : C → I of 5.39. We form the surjection ρ× ρ : C2 → I2 sending (x, y)to (ρ(x), ρ(y)). The composition (ρ× ρ) ◦ ϕ : C → I2 is continuous andsurjective. Its domain is a closed subset of I, and it can be extendedto all of I by Tietze’s extension theorem; see Christenson–Voxman

1977 4.B.8. In our concrete situation, it is not necessary to invoke thattheorem, since it suffices to define the map on each of the middle thirdintervals deleted from I in the course of the construction of C. This canbe done directly using the convexity of the unit square I2; one simplymaps each of these intervals to the segment joining the images of its endpoints, using a suitable affine map x �→ ax + b.

A different construction of a surface-filling curve is given in Exer-cise 5. �

5.41 Remarks Clearly, the last result can be generalized; every powerIn, n ∈ N, is a continuous image of the arc. But this is only a shadowof what can be proved: there is a comprehensive theorem, the Hahn–Mazurkiewicz theorem, which characterizes exactly the spaces that arecontinuous images of the arc; see 5.42 below. The basic idea is the sameas for the proof of 5.40; in order to obtain a continuous surjection of I

onto a space X, one first constructs a surjection C → X and then extendsit over I. Therefore, the characterization of the continuous images of Cis a key ingredient; see 5.43 below. The extension is not as easy as inthe last proof. In order to define it on one missing interval, one uses thetopological characterization of the arc 5.13, but in the form that doesnot involve local connectedness; see 5.15. We refer the interested readerto Christenson–Voxman 1977 for details and proofs, but we state theprecise form of the results.

Page 71: The Classical Fields

54 Real numbers

5.42 Theorem A Hausdorff space is a continuous image of the arc

if, and only if, it is a compact, connected and locally connected metric

space. �

Such spaces are called Peano continua in honour of G. Peano andhis space-filling curves. In the course of the proof, it is shown thatevery compact, connected and locally connected metric space is arcwiseconnected. We might mention here that, of course, the circle is a specialcase; compare 5.16. We return to our systematic study of the Cantorset by stating the result mentioned in 5.41.

5.43 Theorem Every compact metric space is a continuous image of

the Cantor set. �

The Cantor set plays a prominent role also in lattice theory. Indeed,the collection of all subsets of C that are both open and closed formsa lattice, which is a countable, atomless Boolean algebra (atomless, be-cause C has no isolated points). This Boolean algebra is determinedup to isomorphism by the properties mentioned; usually, this fact isreduced, via the Stone representation theorem, to the topological char-acterization of C given in 5.48; see Sikorski 1964 §9C p. 28; comparealso §14B p. 44. An elementary, direct proof is suggested in Exercise 9.

The Cantor set is markedly more homogeneous than the real line; com-pare the remark following 5.27. In Droste–Gobel 2002, it is shownthat the homeomorphism group (compare 5.51 below) of C is quite largein the sense that it cannot be obtained as the union of a countableascending chain of proper subgroups. In fact, the size of the homeo-morphism group is such that it contains an isomorphic copy of everycountably infinite group; see Hjorth–Molberg 2006. Moreover, thehomeomorphism group of the Cantor set is a simple group. Togetherwith similar results about the space of rational numbers and the spaceof irrational numbers, this is proved by Anderson 1958. Here we arecontent to prove the following.

5.44 Proposition The Cantor set is n-homogeneous for each n ∈ N.

Proof (1) First we prove that {0, 2}N ≈ C is 1-homogeneous. Let (an)n

and (bn)n be elements of {0, 2}N. For each n ∈ N, there is a bijectionfn of {0, 2} such that fn(an) = bn. Applying fn to the nth term of eachsequence in {0, 2}N, we obtain a bijection f which maps the sequence(an)n to (bn)n. In order to verify continuity of f we have to check,according to the universal property of products, that the nth member

Page 72: The Classical Fields

5 The real numbers as a topological space 55

dn of the sequence f(x) depends continuously on the sequence x. Thisis true, since dn = fn(cn), where cn is the nth member of x.

(2) From the second description of the Cantor set (as the intersec-tion of a descending chain of compact sets In) it follows that everypermutation of the intervals forming some particular In gives rise to ahomeomorphism of the Cantor set. One has to observe that for eachof these intervals J , the intersection J ∩ C is open in C and is a scaledversion of C itself, and that any two of these sets are homeomorphic bya translation.

Now let two finite sequences x1, . . . , xk and y1, . . . , yk of distinct pointsin C be given. Choose n so large that no two of the xi or of the yi occupythe same interval in In. Permute the intervals so that the ith intervalcontains the image of xi for i = 1, . . . , n and use a homeomorphism ofJ ∩ C as in step (1) for each of these intervals, so that xi is sent to thesmallest element in J ∩ C. Together, this yields a homeomorphism g ofC onto itself. There is a similar homeomorphism h for the yi, and thequotient h−1 ◦ g sends xi to yi for each i. �

5.45 Definition A topological space X is said to be totally disconnectedif every connected component of X is reduced to a point. Equivalently,X contains no connected subsets except points and the empty set.

A space X is said to be perfect if it has no isolated points, that is, nosingleton {x} ⊆ X is an open set.

5.46 Proposition A 2-homogeneous topological space is either con-

nected or totally disconnected.

Proof If A ⊆ X is connected and contains two distinct points a, b, thenany two points x, y ∈ X belong to a connected set f(A) ⊆ X, where f

is a homeomorphism of X sending {a, b} to {x, y}. It follows that X isconnected. �

5.47 Proposition The Cantor set is totally disconnected and perfect.

Proof The Cantor set is 2-homogeneous and not connected, hence it istotally disconnected. The definition of the product topology on {0, 2}Nmakes it clear that no open set can consist of a single point. �

There is a nice characterization of the Cantor set on the basis of theproperties that we have proved. Again we record this result withoutproof; see Christenson–Voxman 1977 6.C.11.

Page 73: The Classical Fields

56 Real numbers

5.48 Theorem Every compact, totally disconnected, perfect metric

space is homeomorphic to the Cantor set. �

We turn now to the subspace I = R � Q ⊆ R, the space of irrationalnumbers. One feels that there is an immense gap separating the spacesR and I. Clearly, I contains no non-trivial interval of the real line, henceit is totally disconnected by 5.4. On the other hand, it was shown in4.11 that I admits a complete metric generating the topology. The proofrelies on the result 4.10, which we repeat here, and which will allow usto prove an unexpected, close relationship between I and R (5.50).

5.49 Theorem The space I = R � Q is homeomorphic to the product

space NN. �

5.50 Theorem There is a continuous bijection I = R � Q→ R.

Proof (1) We shall construct (in step (2) below) a continuous bijectionα : NN → [0, 1[. Together with the homeomorphism β : I → NN of 5.49this yields a continuous bijection γ = β ◦α : I→ [0, 1[. This is not quitewhat we want; we have to get rid of the element 0 in the image of thismap. This can be done by showing that I � {γ−1(0)} is homeomorphicto I. Indeed, Q∪{γ−1(0)} is chain isomorphic to Q by 3.4, and the proofof 3.5 shows that any isomorphism of these chains extends to a chainautomorphism ϕ of R, and ϕ induces the homeomorphism we need.

A more direct construction is this: The space I is the disjoint unionof the open subsets I ∩ [n, n + 1], n ∈ Z. Under the homeomorphismI→ Z× NN constructed in 4.10 using continued fractions, this intervalcorresponds to {n} × NN, hence it is homeomorphic to I itself. Thus,each of these intervals can be mapped continuously and bijectively onto[n, n + 1[, and together these maps yield the desired continuous bijectionI→ R.

(2) In order to construct α, we read an element (an)n ∈ NN as aninstruction for the definition of a sequence of digits bn ∈ {0, 2} : twoconsecutive digits equal to 0 in this sequence shall be separated by an

digits equal to 2. Thus, the zero digits bn = 0 are those whose index hasthe form n = a1+a2+· · ·+ak−1, for some k ∈ N. This defines a map ψ ofNN into the Cantor set C ≈ {0, 2}N (see 5.37), in fact, a homeomorphismof NN onto ψ(NN) because both spaces carry the product topology. Theimage ψ(NN) consists precisely of the sequences having infinitely manydigits equal to 0, and on this set the surjection C → [0, 1] constructed in5.39 induces a continuous bijection onto [0, 1[. �

Page 74: The Classical Fields

5 The real numbers as a topological space 57

We remark that this theorem together with 5.42 allows us to map theirrationals continuously onto any Peano continuum.

The space I is clearly not homeomorphic to its complement Q (thetwo spaces even have different cardinalities; more about the topology ofQ will be said in Section 33). It is possible, however, to exhibit subsets ofR that are homeomorphic to their complements. One can even constructa situation where both complementary sets are homogeneous, and thiscan be done in essentially different ways; see van Mill 1982a, 1982b.

Homeomorphisms of the real line

5.51 The homeomorphism group We have frequently used thenotion of homeomorphism in this section. In the particular case of ahomeomorphism of the topological space R onto itself, we are dealingwith a bijective map f : R → R such that both f and the inverse mapf−1 are continuous.

Various characterizations of homeomorphisms are known as standardfacts of real analysis (but note that they do not hold for general topologi-cal spaces). For instance, it follows from the intermediate value theoremthat an injective continuous map of R into itself preserves or reversesthe ordering (i.e., it is strictly monotone or antitone). Conversely, it isimmediate that every order-preserving or order-reversing bijection mapseach open interval onto an open interval, hence its inverse map is contin-uous. Therefore, the homeomorphisms of R are precisely the bijectionswhich preserve or reverse the ordering.

The same arguments, put together differently, show that a continuousbijection of R preserves or reverses the ordering and therefore has acontinuous inverse. In other words, the continuity of f−1 need not berequired in the definition of homeomorphisms of the real line. This istrue for maps between any two connected manifolds of the same dimen-sion. The proof of this general result is much harder and uses Brouwer’stheorem on the invariance of domain; see, for example, Dugundji 1966XVII.3.1.

Clearly, the homeomorphisms of R form a group under composition,which will be denoted H(R) or, briefly, H. It has a normal subgroupof index two, formed by the monotone (or order-preserving) bijectionsand denoted H+(R) or H+. In other words, the subgroup H+ ≤ H hasprecisely two cosets, consisting of the monotone and antitone bijections,respectively.

Page 75: The Classical Fields

58 Real numbers

Our aims are to show that both the fixed point free elements andthe involutions form a conjugacy class, and that every element of His a product of at most four involutions. Moreover, we determine thecommutator subgroups and we describe all normal subgroups of H andof H+. There are very few of them, thus the two groups are very closeto being simple. The material is taken from Fine–Schweigert 1955.The more difficult proofs will be only sketched, the reader being re-ferred to the original article for full detail. A more special result aboutsubgroups consisting of fixed point free homeomorphisms will also beincluded (5.57).

We remark that we treat H as an abstract group, without a topology,although there is a topology making H a topological transformationgroup of R, namely, the compact-open topology; see Dugundji 1966Chapter XII. We also point out that we often omit the symbol for com-position in H and write fg instead of f ◦ g.

5.52 Definition: Special homeomorphisms Any element i ∈ H oforder two will be called an involution. We shall show that each involutionis conjugate to the standard involution i0 defined by i0(x) = −x. (Recallthat two elements g, h of a group G are said to be conjugate if there isan f ∈ G such that h = f−1gf . This is an equivalence relation.) By atranslation of R we mean an element t ∈ H without fixed points, thatis, t(x) �= x for all x. The name will be justified when we show thateach translation is conjugate to the standard translation t1 defined byt1(x) = x+1. Applying the intermediate value theorem to t− id, we seethat a translation satisfies either t(x) < x for all x or x < t(x) for all x.Accordingly, we call t a left translation or a right translation.

The long line L does not admit any translations; compare Exercise 8.

5.53 Theorem Every involution i ∈ H is conjugate to the standard

involution i0, where i0(x) = −x. In particular, i reverses the ordering.

Proof Choose a ∈ R arbitrarily. The involution i interchanges a and i(a)and therefore reverses the ordering. Moreover, i maps the interval withend points a, i(a) onto itself. Applying the intermediate value theoremto the map x �→ i(x) − x on this interval, we see that there is a fixedpoint x0 of i. Replacing i by its conjugate i′, where i′(x) = i(x+x0)−x0,we reduce the problem to the case x0 = 0.

Define a homeomorphism f ∈ H+ by f(x) = x for x ≤ 0 and f(x) =−i′(x) for x ≥ 0. Then f−1i0f = i′ may be verified by showing that−f(x) = f(i′(x)) both for x ≤ 0 and for x ≥ 0. �

Page 76: The Classical Fields

5 The real numbers as a topological space 59

5.54 Theorem Every translation (fixed point free homeomorphism)t ∈ H is conjugate to the standard translation t1, where t1(x) = x + 1.

In particular, t preserves the ordering.

Proof The idea is similar to that of the last proof. If t is a left transla-tion, we pass to t′(x) = −t(−x); in other words, we conjugate t by i0.Thus, we may assume that t is a right translation. Then the two-endedsequence { tn(x) | n ∈ Z} is strictly monotone. A least upper boundof this sequence would be a fixed point of t, hence the sequence is un-bounded above and, likewise, unbounded below. Thus, R can be writtenas the union of the intervals In = [tn(0), tn+1(0)] for all n ∈ Z. Clearly,t induces a homeomorphism of In onto In+1. We set f(x) = tn(x − n)for n ≤ x ≤ n + 1, n ∈ Z. This defines a homeomorphism f ∈ H, andthe conjugation identity f(t1(x)) = t(f(x)) is easily verified. �

We remark that the conjugating element in the last theorem may beantitone. If we consider conjugacy in the group H+, then the first step(conjugation with i0) in the proof of 5.54 is not needed. Hence, thesame proof shows that there are two classes of translations, the righttranslations and the left translations.

5.55 Definition A one-parameter group of homeomorphisms of R ora flow on R is a group homomorphism r �→ gr from the additive groupR into H. Thus, we have the identity gr+s = grgs = gsgr. Note thatgr = gr/2gr/2 is always monotone. Conversely, we have the following.

5.56 Theorem Every monotone homeomorphism g ∈ H+ can be em-

bedded as g1 in a flow gr, r ∈ R.

Proof The proof is easy in the case of a translation g = t. Write g =f−1t1f as in 5.54 and set gr = f−1trf , where tr(x) = x + r is thestandard flow. In general, a monotone g ∈ H has a fixed point set F ,and the complement of F splits up into at most countably many disjointopen intervals (the connected components of R � F ). On each of theseintervals, g induces a translation, to which the previous argument may beapplied. The resulting flows combine into one global flow which inducesthe identity on F . �

The following result is taken from Salzmann 1958. Its proof intro-duces an ordering on a certain group of homeomorphisms. The entiregroup H cannot be made into a chain in a reasonable way, but it doeshave partial orders. The system of all partial orders has some strikingproperties; compare Holland 1992.

Page 77: The Classical Fields

60 Real numbers

5.57 Theorem Let G ≤ H be a group of homeomorphisms acting

freely on R, that is, every element g ∈ G�{id} is fixed point free. Then

G is isomorphic, as an abstract group, to a subgroup of the additive

group R. In particular, G is commutative.

Proof Every element of G�{id} is either a left or right translation. Fordistinct elements g, h ∈ G, we set g < h if h−1g is a left translation.This means that the evaluation map g �→ g(x0) of G onto an arbitraryorbit G(x0) = {g(x0) | g ∈ G} is an order isomorphism. This definitionmakes G an ordered group, that is, g < h implies fg < fh and gf < hf

for all f, g, h ∈ G.Moreover, this ordered group is Archimedean, which means that for

any two elements g, h ∈ G such that id < g < h there exists a naturalnumber n such that h < gn. To verify this, just observe that the sequencegn(x0) tends to infinity for every x0 (use 5.54 or look into the proof ofthat result).

Now every Archimedean ordered group is isomorphic to a subgroup ofthe ordered group of real numbers with addition. We shall give a proofof this embedding theorem in 7.8. �

5.58 Warning The last result should not be misunderstood. It doesnot say that a freely acting group of homeomorphisms of R acts like asubgroup of R considered as the group of standard translations x �→ x+r.If G does not happen to be cyclic, then it is true that G is a densesubgroup of R by 1.4, and every orbit G(x) is order isomorphic to thisdense subgroup of R. This does not allow us to conclude that G(x) isdense in R. (If this is the case, then the action of G on R is indeed thestandard action.) For counter-examples and for a complete classificationof the groups G as transformation groups, see Lowen 1985.

5.59 Theorem In the groupH of homeomorphisms of R, every element

is a product of no more than four involutions. The number four is

best possible. Every monotone homeomorphism is a product of two

translations (in the sense of 5.52).

Proof (1) The standard translation x �→ x + 1 is the product of the twoinvolutions x �→ −x and x �→ 1−x. Since every translation t is conjugateto this one (5.54), it follows that t is a product of two involutions as well.

(2) We show that every monotone homeomorphism g is a product oftwo translations (and, hence, of four involutions). On each connectedcomponent of the complement of its fixed point set, g induces a transla-tion. Since all translations are conjugate, we see that g is conjugate to

Page 78: The Classical Fields

5 The real numbers as a topological space 61

a homeomorphism h such that |h(x)−x| is bounded by 1. Then t = ht2is fixed point free, where t2(x) = x + 2. Thus h = tt−1

2 is a product oftwo translations, and this carries over to the conjugate g of h.

(3) If h ∈ H is antitone, consider the antitone homeomorphism g =t−1h and note that g(x) < h(x) for all x. The antitone continuous mapdefined by i(x) = min{g(x), g−1(x)} is an involution i ∈ H. Indeed, ifg(x) ≤ g−1(x), for instance, then gg(x) ≥ x, hence ii(x) = i(g(x)) = x.For all x we have i(x) ≤ g(x) < h(x); therefore, t = ih is fixed pointfree, and h = it is a product of three involutions.

(4) Let h(x) = x for x ≤ 0 and h(x) = 2x for x ≥ 0. If h is a productof fewer than four involutions, then their number must be two, because h

is monotone. Suppose that h = ij is a product of two involutions. Thenthe conjugate ihi = ji = h−1 has fixed point set Fihi = i(Fh) = i(] , 0]).This set is an interval of the form [a, [. But clearly h and h−1 shouldhave the same fixed point set, a contradiction. �

The elements of H that can be written as a product of two involutionshave been characterized by Young 1994. Moreover, he shows that thereare three involutions that generate a dense subgroup of H.

5.60 Theorem The centre of H is trivial, and the commutator sub-

group of both H and H+ is H′ = H+.

Proof (1) The map ha defined by ha(x) = 2x−a has a ∈ R as its uniquefixed point. Any element g ∈ H commuting with ha must fix a.

(2) All commutators hgh−1g−1 are monotone. Conversely, we showthat every (right) translation is a commutator of two monotone homeo-morphisms. Since the translations generate the group H+ by 5.59, allassertions about commutator groups follow. So let t be a right transla-tion. By 5.54, the two right translations t and t2 are conjugate; in fact,the remark following 5.54 shows that t2 = f−1tf for some monotonehomeomorphism f . The desired relation t = t−1f−1tf follows. �

Our next aim is to show that the only normal subgroups of H+ arethe trivial ones {id} and H+ itself, and the groups H−∞, H∞ and Hc

which we define next.

5.61 Definition We continue to denote the fixed point set of h ∈ Hby Fh. The closure R � {Fh} of the set of non-fixed points is called thesupport of h.

The groups H−∞, H∞ and Hc consist of those homeomorphisms of Rwhose support is bounded below or bounded above or compact, respec-tively. Thus, h belongs to H−∞ or to H∞ if, and only if, Fh contains

Page 79: The Classical Fields

62 Real numbers

some interval ] , a] or [a, [, respectively, and Hc = H−∞ ∩H∞. Each ofthese groups is a normal subgroup of H+, and Hc is also normal in Hitself.

5.62 Lemma Every proper normal subgroup N ≤ H+ is contained in

H−∞ or in H∞.

Proof (1) We assume that N contains elements whose support is un-bounded below and others whose support is unbounded above. Ouraim is to show that then N contains a right translation. This impliesthat N contains all right translations (5.54) and their inverses, the lefttranslations. Finally, 5.59 shows that that N = H+.

(2) We examine the tools available for our construction. Every mono-tone homeomorphism h ∈ H+ has a fixed point set Fh, whose comple-ment is a disjoint union of open intervals. On each interval I of this kind,h induces a (left or right) translation. By passing to a conjugate fhf−1,f ∈ H+, we may manipulate the fixed point set, which becomes f(Fh).Instead of doing this, we can conjugate the translation induced on aninterval I and replace it with any other translation having the samedirection (left or right), see Theorem 5.54. This can be done simulta-neously for all intervals I; the conjugating homeomorphisms togetherdefine a monotone homeomorphism of R that induces the identity onthe set Fh.

Our strategy in applying these tools is to construct, in this order,elements of N whose fixed point set is bounded above, elements withcompact fixed point set, elements with at most one fixed point, andfinally translations.

(3) Assume that the support of h ∈ N is unbounded above. We shallconstruct an element g ∈ N whose fixed point set is bounded above.If h itself does not have this property, then there is an unbounded,increasing sequence of open intervals In, n ∈ N, on which h inducestranslations in the same direction (say, right translations). The gapbetween In and In+1 is a closed interval Jn, which contains fixed pointsand perhaps some other points that may be moved to the left or to theright. Note that Jn is mapped onto itself. Using the tools exhibitedin step 2, we may construct a conjugate f of h having ‘gap intervals’Jn = [2n − 1/5, 2n + 1/5] and satisfying ϕf (x) = |f(x) − x| > 2/5 forall x ∈ [2n + 1 − 1/5, 2n + 1 + 1/5], all n ∈ N. Note that ϕf (x) < 2/5for x ∈ Jn. Now we form the conjugate f ′ = t−1ft1, for which the rolesof the odd and even numbers are interchanged. It is easily verified thatthe product g = f ′f moves every x ≥ 2 to the right.

Page 80: The Classical Fields

5 The real numbers as a topological space 63

(4) If the support of the element g constructed in step 3 is unboundedbelow, then we can repeat the procedure of step 3 in a symmetricalfashion and obtain an element with compact fixed point set. The samecan be achieved if the support of g is bounded below: by assumption,there is an element whose support is unbounded below, and we can doeverything over again to obtain an element g′ whose fixed point set isbounded below and whose support is bounded above. Then the fixedpoint set of gg′ is compact.

(5) We have obtained an element k ∈ N that induces translationson two unbounded intervals ] , a[ and ]b, [. If these translations havethe same direction, then the ideas of step 3 can be applied to obtaina translation belonging to N . If the directions are not the same, thenit is possible to construct an element with precisely one fixed point bycombining the ideas of step 3 with those of the next step 6. The detailsof this argument are left to the reader (and can be found in Fine–

Schweigert 1955).(6) Suppose that l ∈ N fixes precisely one point, say 0, and moves all

x < 0 to the right and all x > 0 to the left. After taking the inverse, ifnecessary, this is what we get from step 5. Using the tools of step 2, weobtain a conjugate l′ of l that fixes the point 1 and satisfies l(x) < l′(x)for all x. Then l−1l′ is a right translation, and this ends the proof. �

5.63 Theorem (a) The group Hc of all homeomorphisms of R with

compact supports is simple, that is, it has only the trivial normal

subgroups: {id} and Hc itself.

(b) The only non-trivial normal subgroup of H−∞ and of H∞ is Hc.

(c) The non-trivial normal subgroups of H+ are precisely H−∞, H∞,

and Hc.

(d) The non-trivial normal subgroups of H are precisely Hc and H+.

Proof (1) For u < v in R ∪ {−∞,∞}, we introduce the group Hu,v

consisting of all h ∈ H that induce the identity on ] , u] ∪ [v, [. Thesegroups are all isomorphic to H+. An isomorphism ψ : Hu,v → H+

is obtained by ‘conjugation’ with a homeomorphism f : ]u, v[ → R.Clearly, the groups Hu,v for u, v ∈ R are all conjugate.

(2) If N < Hc is a normal subgroup and h ∈ N is not the identity,then there is a minimal closed interval [u, v] ⊆ R such that h ∈ Hu,v.The appropriate isomorphism ψ takes N ∩ Hu,v to a normal subgroupM of H+. The support of ψ(h) is unbounded above and below, and weconclude from 5.62 that M = H+ and that N contains Hu,v. Hence, it

Page 81: The Classical Fields

64 Real numbers

contains Hc, which is the union of all conjugates of Hu,v. This provesassertion (a); the proof of (b) is entirely similar and is omitted.

(3) A proper normal subgroup of H+ is a normal subgroup of H−∞ orof H∞ by 5.62. Hence, assertion (c) follows from (b). Finally, considera normal subgroup N < H which is not contained in H+. Then there isan antitone homeomorphism h ∈ N . Given any right translation t, thecommutator h−1(t−1ht) belongs to N . Now h−1(t−1ht) = (h−1t−1h)tis a product of two right translations, and such a product is again aright translation. Hence, N contains all translations and their products,which exhaust all of H+; see 5.54 and 5.59. �

A characterization of the group H has been mentioned after 3.7. Com-pare also Section 6 of Fine–Schweigert 1955. For more general resultsabout simplicity of homeomorphism groups, we refer to Epstein 1970and references given there.

In relation to the structure of the homeomorphism group H, there aresome more results in the literature that we would like to mention.

The first such result, due to Abel 1980, is about the subgroup P ≤ Hconsisting of all monotone, piecewise linear bijections and contrasts thesimplicity results aboutH. In fact, Abel shows that the group P containsan uncountable collection of normal subgroups.

The following example given by Cohen–Glass 1997 is of interest: if p

is an odd prime, then the elements f, g ∈ H defined by f(x) = x+1 andg(x) = xp generate a free subgroup of H. Bennett 1997 constructs afree subgroup of H with n free generators, for every n ∈ N. Actually, Hcontains a free subgroup with a free generator set having the cardinalityof the continuum. For a proof, see Blass–Kister 1986.

Banach proved the following strange property of the group of ordinarytranslations of the real line. There is a decomposition of R into disjointsets M1,M2, each having the same cardinality as R, such that for eachx ∈ R and each i ∈ {1, 2} the cardinality of the symmetric difference(Mi ∪ (Mi +x))� (Mi ∩ (Mi +x)) is strictly smaller than that of R. Fora recent proof, see Abel–Misfeld 1990, where it is also shown that thecorresponding statement for Q is false.

Weird topologies on the real line

Despite their strange properties, the topologies described in this subsec-tion occur quite naturally. The first of them represents a pathology thatcauses unavoidable trouble in the theory of Lie groups (see 5.66, 5.68).

Page 82: The Classical Fields

5 The real numbers as a topological space 65

5.64 Definition: The torus Recall the surjective local homeomor-phism p : R → S1 introduced in 5.16. (A local homeomorphism is amap of topological spaces f : X → Y such that each x ∈ X has anopen neighbourhood that is mapped homeomorphically onto some openneighbourhood of f(x).) We consider the product space T2 = S1 × S1,which can be embedded in R3 as the surface of a doughnut, and thesurjective map

q = p× p : R2 → T2 : (x, y) �→ (e2πix, e2πiy) .

If I, J ⊆ R are any two open intervals of length 1, then the restriction ofq to the open square S = I×J is a homeomorphism onto its image q(S),and q(S) is open in T2. This follows from the corresponding propertiesof p; compare 5.16. We conclude that T2 is a 2-manifold and that thesurjection q is continuous and open. In particular, T2 carries the quotienttopology with respect to q. From the periodicity of p it follows that theinverse images q−1(u) of the points u ∈ T2 are precisely the cosets of thesubgroup Z2 ≤ R2. Since q is a group homomorphism, it follows thatT2 may be considered as the factor group:

T2 = S1 × S1 = R2/Z2 .

5.65 Example: The torus topology on R The restriction of themap q (see 5.64) to the subgroup R(1,

√2) is injective because

√2 is

irrational. Therefore, we can identify R as a set with its image underthe map t �→ q(t, t

√2). The torus topology is defined as the subspace

topology induced on this copy of R by the torus T2. We shall write T

for the set of real numbers with this topology.We note that T ⊆ T2 is a dense subset; this follows from a well known

theorem of Kronecker, which asserts that even the cyclic subsemigroupof T2 generated by q(

√3,√

6) is dense. We shall prove Kronecker’stheorem in 5.69. Each coset of T + v arises as the image of T underthe homeomorphism u �→ u + v of T2, hence T + v is dense, as well. Inparticular, the complement of T is dense.

5.66 Remark The torus is a topological group (addition and subtrac-tion are continuous, compare Section 8) and, in fact, a Lie group (thegroup operations are even differentiable). The continuity of group op-erations carries over from T2 to the subgroup T , i.e., T is a topologicalsubgroup. The study of this subgroup is forced on us by Lie theory.Indeed, the map q can be viewed as the exponential map from the Liealgebra R2 onto T2, and T is the exponential image of the Lie subalgebra

Page 83: The Classical Fields

66 Real numbers

R(1,√

2) ≤ R2. It follows from the fundamental theorems of Lie theorythat T is a Lie group itself, but unfortunately, since T is not closed inT2, the topology making T into a Lie group is not the topology inheritedfrom T2. This is a widespread phenomenon among Lie groups; comparealso 5.68 below.

5.67 Proposition The torus topology on R is homogeneous and con-

nected, but not locally connected and not locally compact. All points

of T are non-separating, and T is locally homeomorphic to R×Q.

Proof Since T is a topological group (5.66), the maps u → u + v arehomeomorphisms of T for each v ∈ T , and homogeneity follows. Thecontinuity of the surjection q : R→ T shows that T is connected.

The next two in our list of properties follow from the local homeo-morphism to R × Q, which is obtained as follows. The inverse imageA = q−1(T ) equals V + Z2, where V = R(1,

√2). The map q is a

local homeomorphism from this set onto T because q : R2 → T2 is alocal homeomorphism (5.64). Therefore, it suffices to consider A. Us-ing a real vector space decomposition R2 = V × W , we obtain thatA = V × (W ∩A), and W ∩A ∼= A/V is a non-cyclic subgroup of W ∼= Rgenerated by two elements (namely, the elements corresponding to twogenerators of Z2, compare 1.6). By 1.4 or 1.6b, this subgroup is dense,and by 3.5 it is chain isomorphic and, hence, homeomorphic, to the setof rationals, Q.

It remains to be shown that no point separates T . By homogeneity,it suffices to do this for the neutral element 0 ∈ T . The proof of thedensity of T ⊆ T2 shows in fact that q(P (1,

√2)), where P = ]0, [, is

dense in T2. In particular, this connected set is dense in T � {0}, whichshows the connectedness of the latter set. �

5.68 Remark: Retrieving the original topology of R The factthat T is locally homeomorphic to R × Q allows us to reconstruct thestandard topology of R. In a neighbourhood homeomorphic to R × Q(every open set in R×Q contains such a neighbourhood), the sets of theform R × {r}, r ∈ Q, are precisely the arcwise connected components.Under the map q, these sets correspond to open intervals in the realline. Therefore, a method to reconstruct the original topology is to de-fine the arcwise connected components of all open sets of T to be open.This method works for non-closed Lie subgroups of Lie groups in gen-eral; see Gleason–Palais 1957, Hofmann–Morris 1998 Appendix 4,Stroppel 2006 §39.

Page 84: The Classical Fields

5 The real numbers as a topological space 67

This ends our treatment of the torus topology, but before we startanother topic we want to supply a statement and proof of Kronecker’sTheorem, which we relied on in 5.65. In the course of the proof it willbecome clear in which sense the theorem generalizes the result 1.6b onsubgroups of R+ having two generators. The proof given here is takenfrom Bourbaki 1966 VII §1.3. A very direct proof is given by Adams

1969 4.3. We formulate Kronecker’s theorem (in its first version) for then-dimensional torus Tn = Rn/Zn. Let q : Rn → Tn denote the quotientmap.

5.69 Kronecker’s Theorem Let w = (w1, . . . , wn) ∈ Rn be given and

suppose that the numbers 1, w1, . . . , wn are linearly independent over Q.

Then the semigroup generated by q(w) is dense in Tn.

Proof (1) First we look at the cyclic group generated by q(w) ratherthan the semigroup; in other words, we consider also multiples zq(w)with negative z ∈ Z. By continuity of q, it suffices to show that thegroup H ≤ Rn generated by w together with ker q = Zn is dense. (Inthe case n = 1, this is the result 1.6b.)

If H is not dense, then we apply the classification of all closed sub-groups of Rn; see 8.6. It implies that there is a basis v1, . . . , vn of Rn

such that the closure H consists of all linear combinations v =∑n

i=1 rivi

satisfying ri ∈ Z for i > i1 and ri = 0 for i > i2, where i1 ≤ i2 ≤ n.We have i1 < n since H is a proper subgroup of Rn, and i2 = n be-cause Zn ≤ H generates the real vector space Rn. In particular, thereis a non-trivial linear form f(v) = rn on Rn which takes only inte-ger values on H. The standard basis vectors e1, . . . , en belong to Zn,hence the values f(ei) are integers, not all equal to zero. We find thatf(w) =

∑ni=1 wif(ei) ∈ Z, and this contradicts our assumption of linear

independence.(2) We have shown in particular that the integral multiples of q(w)

accumulate at 0 ∈ Tn. This implies (by continuity of subtraction in Tn)that there is a strictly increasing sequence of natural numbers nk suchthat nkq(w)→ 0 in Tn. Now we use a metric d on Tn. Suppose we haved(zq(w), t) < ε for some negative integer z and t ∈ Tn. By continuityof addition, (z + nk)q(w) → zq(w). Thus, we can find nk > −z suchthat d((z + nk)q(w), zq(w)) < ε, and then d((z + nk)q(w), t) < 2ε asdesired. �

We use this opportunity to rephrase Kronecker’s Theorem in terms ofDiophantine approximations.

Page 85: The Classical Fields

68 Real numbers

5.70 Kronecker’s Theorem, second version Let w1, . . . , wn be real

numbers such that 1, w1, . . . , wn are linearly independent over Q. For

any numbers v1, . . . , vn ∈ R and ε > 0 there exist integers k > 0 and

z1, . . . , zn such that for each i ∈ {1, . . . , n} we have

|vi − kwi − zi| < ε .

Proof Let w = (w1, . . . , wn) and v = (v1, . . . , vn). Let U ⊆ Rn be theopen set U = {x − z | x ∈ Rn, z ∈ Zn, ‖v − x‖ < ε}. Here we use thenorm ‖y‖ = max{|y1|, . . . , |yn|} on Rn. With respect to the quotientmap q : Rn → Tn, the set U is saturated, that is, U = q−1q(U), andtherefore q(U) is open in Tn. According to 5.69, the set q(U) containssome multiple kq(w), k ∈ N. This means exactly what is stated in theassertion. We remark that these arguments can be read backwards toshow that the present version of the theorem implies the earlier, grouptheoretic one. �

5.71 Uniform distribution The following sharper result is due toBohl, Sierpinski and Weyl: under the assumptions of 5.70, the sequencekw ∈ Rn, k ∈ N, is uniformly distributed modulo 1. This means thatfor every box B(ε) = {y ∈ Rn | ‖y − v‖ < ε}, the proportion ofelements kw contained in B(ε) + Zn among the first N members of thesequence converges to the box volume (2ε)n as N → ∞. For a proofand more information, see Hlawka et al. 1991 Proposition 4 on p. 26or Kuipers–Niederreiter 1974.

5.72 Example: The topologies of semicontinuity Let X be atopological space. A lower semicontinuous map f : X → R is a map suchthat for each b ∈ R the inverse image of ]b, [ is an open set. This propertycan also be described as continuity with respect to the topology on Rgenerated by a basis consisting of all the intervals ]b, [ with b ≥ −∞.In fact, the basis coincides with the topology in this case, and the opensets form a chain with respect to inclusion.

From this it follows that the topology of lower semicontinuity is con-nected (even every subset is connected), and the subspaces [x, [ arecompact. Points are not closed sets, hence there is no metric generatingthis topology. There is a countable basis: just restrict b to be a rationalnumber. The ordinary topology of R is the coarsest one that containsthis topology together with its mirror image, the topology of upper semi-continuity. This expresses the familiar fact that a map is continuous if,and only if, it is both upper and lower semicontinuous.

Page 86: The Classical Fields

5 The real numbers as a topological space 69

5.73 Example: The Sorgenfrey topology The Sorgenfrey line S

is the set R endowed with the topology generated by all intervals of theform ]a, b]. For a metric space X, continuity of a map f : S → X meansthat the left-sided limit f(s−) exists for each s ∈ S and coincides withf(s). The Sorgenfrey line obviously is separable, but the topology doesnot have a countable basis. Indeed, a basis must contain at least oneinterval ]a, b] for each b ∈ R. A separable metric space has a countablebasis, consisting of all balls having rational radius, and centre belongingto a given countable dense set; compare 61.13. Consequently, S is notmetrizable.

The sets ] , x] and ]x, [ are open in S, and S is not connected. How-ever, S can be embedded as a dense subspace of some connected andlocally pathwise connected space; see Fedeli–Le Donne 2001. The se-quence (1/n)n has no convergent subsequence. Every open set containsa translate of this sequence (up to finitely many elements), hence noneighbourhood of any point is sequentially compact and S is not locallycompact.

In Sorgenfrey 1947, this topology on R was introduced as an exam-ple of a paracompact (and hence normal) space whose Cartesian squareS × S is not normal; compare also van Douwen–Pfeffer 1979.

5.74 Note We mention some other results pertaining to our presenttopic. In Baker et al. 1978, it is shown that there are many topologieson the real line that produce the same set of convergent sequences asthe standard topology. The infimum of these topologies is the cofinitetopology (precisely the complements of all finite sets are open); thistopology itself does not have the property mentioned. In Guthrie et al.1978 a maximal connected topology on R is constructed that is finer thanthe standard topology. van Douwen–Wicke 1977 give an example of atopology on R having an especially strange cocktail of properties. Finallywe mention the article Ch’uan–Liu 1981 on topologies making R+ intoa topological group; we shall return to this in Section 8.

5.75 Note: Ordering and topology on R Following the generalpattern of this book, the reader might expect to see a section on therelationships of the ordering and the topology on the real line, suchas the one on ordering and group structure which follows (Section 7).However, there is no such section, and this is due to the results proved inthe present section, which say that not only does the ordering determinethe topology (3.1), but also the topology determines the ordering upto reversal (that is, up to interchanging < with >); this was the key

Page 87: The Classical Fields

70 Real numbers

step in the proof of the characterization theorem 5.10. One cannotexpect something new to arise from combining two structures that areessentially the same. Compare, however, 5.51.

Exercises(1) Consider a topological space Z = X ∪ Y . A subset A ⊆ X ∩ Y which isopen (closed) in both X and Y is open (closed) in Z.

(2) Consider the split reals (Section 3, Exercise 1) with the topology inducedby the ordering and compare the topological properties of this space to thoseof R.

(3) Prove the properties of the counter-examples described before 5.21 and in5.23.

(4) Show that the space I = R � Q of irrational numbers is n-homogeneousfor each n ∈ N.

(5) Divide the triangle T with sides of length 3, 4 and 5 by an altitude into twosimilar triangles T0 and T1, where T0 is the smaller one. Proceed by induction:divide each Tk by an altitude into similar triangles Tk0, Tk1, the smaller onebeing Tk0. Write each real number in the interval [0, 1] as a binary expansionc = 0.c1c2 . . . , put c|ν = c1c2 . . . cν , and map c onto Tc =

T

ν Tc|ν . Show thatthis map is well-defined and is a continuous surjection ϕ : [0, 1] → T .

(6) Give an example of a subchain S of R such that the order topology of Sis different from the subspace topology.

(7) Give an example of a subspace of R which is complete with respect to acompatible metric, but not complete with respect to the ordering.

(8) Show that every continuous, increasing self-map of the long line L has afixed point.

(9) Use back and forth induction as in 3.4 to show that any two countable,atomless Boolean algebras are isomorphic. (A Boolean algebra is the samething as a complemented distributive lattice.)

(10) Let I = R � Q. Show that the subset S := Q2 ∪ I2 of R2 is arcwiseconnected (Ren 1992).

6 The real numbers as a field

The real numbers equipped with their addition and multiplication forma field. The field structure has already appeared explicitly in this book(Section 2), and was used implicitly on several occasions, for example,in Section 3. The present section will focus on properties concerning thefield structure alone. There will be other sections on R as an ordered field(Section 11) or as a topological field (Section 13). In order to emphasizethe roles played by ordering and topology, we treat these aspects inseparate sections although the ordering and, hence, the order topology

Page 88: The Classical Fields

6 The real numbers as a field 71

of the real number field are completely determined by the algebraicstructure (the positive real numbers are precisely the non-zero squares).

Following the general pattern of this book, we begin by examiningthe precise relationship between addition and multiplication in a field;in other words, we scrutinize the field axioms; compare also Section 64.

6.1 The field axioms A skew field consists of two group structures(F,+), with neutral element 0, and F× = (F � {0}, ·), with neutralelement 1, having almost the same underlying set. Multiplication has tobe defined for all pairs in F × F , and the two structures are connectedby the two distributive laws

a(x + y) = ax + ay

(x + y)a = xa + ya .

Usually, commutativity of addition x + y = y + x is stipulated as anadditional axiom, but this can be proved by computing (1+x)(1+ y) intwo ways, using the left and right distributive laws one after the other,in different orders.

Commutativity of multiplication xy = yx is another matter and holdsonly for a subclass of all skew fields, the fields, which of course includethe real numbers. The standard example of a proper (non-commutative)skew field is provided by the quaternions H; see Section 13, Exercise 6,or 34.17.

It is often intriguing to beginners that the neutral element 0 has tobe excluded from the multiplicative group, which means essentially thatdivision by zero is banned. This is unavoidable because we do not wantto accept fields consisting of 0 alone: the distributive law enforces that0x = 0 for all x (indeed, 0x = (0 + 0)x = 0x + 0x); if we allow divisionby 0, we end up with 1 = 00−1 = 0 and x = 1x = 0x = 0 for all x. (Asa ring, {0} is usually accepted, but not as a field.)

To highlight the importance of the distributive laws, we interpret themas follows. We consider the left multiplication maps λa : x �→ ax, a ∈ F .The first distributive law asserts that λa is an endomorphism of theadditive group (F,+), and the second one says that the map sendinga to λa is a homomorphism from (F,+) into the semigroup End F ofall endomorphisms of (F,+). A similar statement, with the roles of thedistributive laws interchanged, holds for the right multiplication mapsρa : x �→ xa.

While students are sometimes tempted to divide by zero, ignoringthe fundamental lack of symmetry between 0 and 1, the much stronger

Page 89: The Classical Fields

72 Real numbers

asymmetry of the distributive laws usually goes unnoticed, and no onemisses the ‘dual distributive law’ (ab) + c = (a + c)(b + c) among thefield axioms. This law does hold in Boolean algebras (where one usuallywrites ∨ and ∧ for the operations of addition and multiplication), butthere we do not have additive or multiplicative inverses. In the presenceof the other field axioms, the dual distributive law becomes contradic-tory: evaluating (1+1)(0+1) in two ways one would obtain that 1 = 0.

Interesting generalizations of the skew field axioms are obtained ifone dismisses the associative laws of addition and multiplication andthe distributive laws wholly or in part. From the group axioms, oneretains that the equations ax = b, xa = b for a �= 0 and a + x = b,x + a = b for arbitrary a have unique solutions. This leads to notionssuch as nearfield, quasifield, semifield. These structures are importantin geometry because the weak axioms, together with a condition on theunique solvability of certain types of ‘linear’ equations, still ensure thatF×F with lines defined by the equations of type y = ax+b or x = const.forms an affine plane, obeying Euclid’s parallel postulate. A detailed dis-cussion and numerous examples can be found in Salzmann et al. 1995.Nearfields (lacking only one distributive law) are also important becauseof their relationship with sharply 2-transitive permutation groups; seeDixon–Mortimer 1996 7.6.

6.2 Independence of the field operations (a) The additive groupof real numbers does not determine the multiplication of the field R.Indeed, the field of complex numbers has an additive group isomorphicto that of the real numbers (see 1.15), and C× �∼= R×, because C× has alarge torsion subgroup Q+/Z (compare 1.20 and 1.24).(b) Conversely, suppose that F is a field such that F× ∼= R× ∼= C2×R+;compare 2.2. If char F = 0, then F+ ∼= R+ by 1.14. (We shall return tothis case in part (d) below.) If charF = p is a prime, then p �= 2 andF× ∼= R× contains the multiplicative group F×

p of the prime field, hencep = 3.(c) Now we can produce an example of a field F such that F× ∼= R×

and F+ �∼= R+. Using model theory, an example with similar propertiesis constructed by Contessa et al. 1999 6.2. As in 2.10, we consider thefield of Puiseux series

F = F3((t1/∞)) ;

compare 64.24. We have char F = 3, hence all non-zero elements ofF+ have order 3, and F+ �∼= R+. Every series a ∈ F× can be writtenuniquely as a = a0t

r(1 + b), where r = v(a) ∈ Q and a0 ∈ F×3 . Here, v

Page 90: The Classical Fields

6 The real numbers as a field 73

is the natural valuation of F , as in 64.24. It follows that 1 + b belongsto the subgroup F1 = {1 + c ∈ F | v(c) > 0} ≤ F×, and we obtain adirect product decomposition of the multiplicative group

F× ∼= F×3 × tQ × F1.

As in the proof of 2.10, it can be shown using Lemma 2.9 that F1 hasunique roots; in steps (2) and (3) of the proof, the prime 2 has to bereplaced with 3. This is possible since F has characteristic 3, whence(∑

i≥k citi/n)3 =

∑i≥k c3

i t3i/n. It follows that F× ∼= C2×H, where the

group H has unique roots.We want to apply 1.14 in order to show that H ∼= R+, hence we

need to determine card H = card F . The inclusions FN3 ⊆ F ⊆ FQ

3

show that cardF = 2ℵ0 = card R (compare 1.10), and 2.2 implies thatF× ∼= C2 × R+ ∼= R×.

See 14.7 and 34.2 for analogous questions about the fields C and Q.(d) Finally, we give an example of a field F such that F+ ∼= R+ andF× ∼= R× are separately isomorphic to their counterparts in R butF �∼= R as a field. See 24.4f for other examples with these properties,and 14.7 for an analogous example concerning the field C.

The field R may be described as an algebraic extension of a purelytranscendental extension Q(T ); in other words, T is a transcendencybasis of R over Q; compare 64.20. Now let S ⊆ T be a proper subsethaving the same cardinality as T . Then Q(S) ∼= Q(T ), and this fieldhas the same cardinality as its algebraic closure, C; see 64.16. From theinclusion S ⊆ T we obtain a proper inclusion of the algebraic closuresQ(S)� ⊆ Q(T )� = C (proper, because elements of T � S do not belongto Q(S)�; compare 64.20), and we define F to be the field

F = Q(S)� ∩ R

of all real numbers that are algebraic over Q(S). The additive group F+

is uniquely divisible (since charF = 0) and has the same cardinality asR, hence F+ ∼= R+ by 1.14. The multiplicative group is a direct productF× = 〈−1〉 ×F×

pos, where F×pos consists of the positive elements. By our

construction, the latter group has unique roots, and 1.14 shows that itis isomorphic to R+. By 2.2, this implies that F× ∼= R×. As a field,however, F is not isomorphic to R. Indeed, the positive elements of F

are precisely the non-zero squares, hence a field isomorphism F → Rwould be an order isomorphism (compare 11.7), but F is a dense propersubchain of R, hence F is not complete.

Page 91: The Classical Fields

74 Real numbers

6.3 Subfields of R Every subfield F ≤ R contains the element 1 and,hence, the subfield generated by 1, the prime field Q of R. Thus F

may be described as an extension field of Q; compare Section 64. Easyexamples are the fields Q(

√2) = {q + r

√2 | q, r ∈ Q}, and the simple

transcendental extensions Q(t) (e.g., t = π), consisting of all rationalexpressions p1(t)/p2(t) ∈ R, where p1, p2 are polynomials with rationalcoefficients and p2 �= 0. Other nice subfields include the field R ∩Q� ofall real algebraic numbers, the Euclidean closure and the Pythagoreanclosure of Q; see 12.8. A remarkable fact is that the vector space di-mension of R over any proper subfield is infinite, as a consequence ofTheorem 12.15.

Following our general scheme, we look for homomorphic images of thefield R, but we do not find much. This is because, for every homomor-phism ϕ : F → H between fields, the kernel ϕ−1(0) is an ideal of F ,and since F is a field, its only ideals are {0} and F ; compare 64.3. Thismeans that ϕ is a monomorphism (so the image field is nothing new)or ϕ = 0. Of course, in the latter case ϕ(1) �= 1, so ϕ is usually notconsidered to be admissible as a homomorphism between fields.

For endomorphisms ϕ : R→ R, we can say even more, but we need theinextricable tie between the field structure and the ordering of R, whichwe shall study systematically later on. At first, it seems conceivable thatR is isomorphic to some proper subfield of itself or that R has non-trivialautomorphisms, but none of this is true; the following result is due toDarboux 1880.

6.4 Rigidity Theorem (Darboux) The only non-zero endomorphism

of the field R is the identity. In particular, the automorphism group of

R is trivial: Aut R = {id}.Proof A non-zero endomorphism ϕ is injective by the preceding remarks.From ϕ(1) �= 0 and ϕ(1)2 = ϕ(12) we conclude that ϕ(1) = 1. It followsthat ϕ induces the identity on Q.

Now we use the fact that the positive real numbers are precisely thenon-zero squares, and its consequence, that every endomorphism sendspositive numbers to positive numbers and, hence, preserves the ordering;compare 11.7. If there is a number r such that ϕ(r) �= r, then there isa rational number q between r and ϕ(r), and ϕ cannot preserve theordering of the pair q, r. �

A skew field without any non-trivial automorphisms is said to be rigid ,thus we have proved that R is rigid. A well known non-rigid field is the

Page 92: The Classical Fields

7 The real numbers as an ordered group 75

field of complex numbers with its conjugation automorphism. There aresubfields of R admitting the same type of automorphism, for example,the field Q(

√2) has the automorphism q + r

√2 �→ q − r

√2. The simple

transcendental extension Q(t) has automorphisms sending the element t

to (at+b)/(ct+d), where ad−bc �= 0; see 64.19 and references given there.Apart from non-trivial automorphisms, the field of complex numbers hasproper (i.e., non-surjective) endomorphisms; see 14.9. We remark herethat C has rigid extension fields. In fact, Dugas–Gobel 1987 showthat every field can be embedded in an extension field with prescribedautomorphism group.

Exercises(1) The field R has an uncountable, strictly increasing family of subfields.

(2) We denote by Z[[x]]� := {P

n≥0 anxn | an ∈ Z ∧ ∃C,k∈N∀n∈N |an| ≤ Cnk }the set of all integral power series with polynomially bounded coefficients.Show that Z[[x]]� is a subring of the integral power series ring Z[[x]] and that

the quotient ring Z[[x]]�/(1 − 2x) is isomorphic to the field R.(See also Section 51, Exercise 3.)

7 The real numbers as an ordered group

We take it as a known fact that the real numbers with their usual addi-tion and their usual ordering form an ordered group in the sense of thefollowing definition.

7.1 Definition An ordered group is a set G equipped with two struc-tures, a group structure (G, +) and a chain structure (G, <), related bythe law of monotonicity

x < y =⇒ x + c < y + c and c + x < c + y ,

which holds for all x, y, c ∈ G. Even though we write + for the operationin G (because we think of the example R+), we do not assume that G

is commutative.The law of monotonicity admits an interpretation reminiscent of the

one we gave for the distributive laws in a skew field, see 6.1. We considerthe left and right addition maps λc, ρc : G→ G, defined by λc(x) = c+x

and ρc(x) = x + c. The law of monotonicity expresses that these twomaps are endomorphisms (order-preserving maps) of the chain (G, <).Observing that, for instance, λcλ−c = λ−cλc = idG, we see that these

Page 93: The Classical Fields

76 Real numbers

maps are in fact automorphisms (order-preserving bijections). Observ-ing moreover that λy−x(x) = y, we obtain the following proposition. Inpassing, we also note that the inversion map g �→ −g is antitone in everyordered group.

7.2 Proposition The chain underlying an ordered group is homoge-

neous, i.e., it has a transitive automorphism group. �

Thus the group influences the chain underlying an ordered group.Conversely, if 0 < x ∈ G, then x < x + x by monotonicity, and aneasy induction shows that the multiples nx with n ∈ N form a strictlyincreasing sequence. In particular, they are all distinct, and the sameholds for x < 0 by a symmetric argument; these remarks prove thefollowing.

7.3 Proposition Every ordered group is torsion free. �

Conversely, a torsion free group which has a (transfinite) descendingcentral series with intersection {0} can be ordered; see Neumann 1949b,Prieß-Crampe 1983 I §4 Satz 14. In particular, this holds for torsionfree abelian groups and for non-abelian free groups.

We mention that every divisible ordered abelian group is isomorphicto an initial subgroup of the additive group of Conway’s surreal numbers;see Ehrlich 2001.

The next definition introduces a particularly important feature of theordered group of real numbers.

7.4 Definition An ordered group G is said to be Archimedean if forany two elements a, b ∈ G such that 0 < a, there is a natural numbern ∈ N such that b ≤ na.

Examples of Archimedean ordered groups are common enough; com-pare 7.5 below. For examples of non-Archimedean ordered groups, see7.14 and the exercises.

For brevity, we call an ordered group with a complete underlying chaina completely ordered group.

7.5 Theorem Every completely ordered group is Archimedean. In

particular, the ordered group of real numbers is Archimedean.

Proof If G is not Archimedean, then there is a positive a ∈ G suchthat the increasing sequence na, n ∈ N, is bounded. There exists c =sup Na = sup(Na + a), but monotonicity implies that sup(Na + a) =(sup Na) + a, which is a contradiction. �

Page 94: The Classical Fields

7 The real numbers as an ordered group 77

7.6 Proposition Let G be an ordered group.

(a) If G has no smallest positive element, then for every positive g ∈ G

and every n ∈ N, there exists a positive h ∈ G such that nh < g.

(b) If G is Archimedean and has a smallest positive element e, then G is

the infinite cyclic group with generator e, endowed with the natural

ordering.

Proof (a) If no smallest positive element exists and 0 < g ∈ G is given,we find an element x with 0 < x < g. We write g = x + y and find apositive element z < min(x, y). Then 2z < g, and assertion (a) followsby induction.

(b) Suppose that e is a smallest positive element and that G is Archi-medean. The cyclic group Ze generated by e exhausts G because foreach g ∈ G there is an n ∈ Z such that ne ≤ g < (n + 1)e, and theng = ne or else 0 < g − ne < e. Clearly, we have ne < me if, and only if,n < m. �

The following Proposition is the first step in proving Holder’s embed-ding theorem 7.8.

7.7 Proposition Every Archimedean ordered group is commutative.

Proof If G is Archimedean and not commutative, then there are positiveelements a, b ∈ G such that a+b < b+a. We may write b+a = a+b+c,and in view of 7.6 we may assume that there is a positive d ∈ G with2d < c. We find non-negative integers m,n satisfying md ≤ a < (m+1)dand nd ≤ b < (n+1)d, and we set k = m+n+2. Then kd < a+ b+ c =b + a < kd, a contradiction. �

The following important embedding theorem is essentially due toHolder 1901 §12.

7.8 Theorem (Holder) Every Archimedean ordered group G is order-

isomorphic to some subgroup of the ordered group R.

Proof We follow Prieß-Crampe 1983 I §3 Satz 4; compare also Blyth

2005 10.16. We may assume that G is not trivial. Fix an arbitrarilychosen positive element e ∈ G. For each a ∈ G, we define two subsetsof Q, the ‘lower set’ L(a) and the ‘upper set’ U(a) such that these twosets form a Dedekind cut of Q. This means that Q = L(a) ∪ U(a) andthat q ∈ L(a), r ∈ U(a) implies q < r. The definition is

L(a) = { mn| m ∈ Z, n ∈ N,me ≤ na}

U(a) = { uv | u ∈ Z, v ∈ N, ue > va} .

Page 95: The Classical Fields

78 Real numbers

If me ≤ na and ue > va, then mve ≤ nva < nue and mv < nu. Thisproves that L(a) < U(a). That Q is the union of the two sets is a directconsequence of their definition. Moreover, L(a) and U(a) are not empty,since G is Archimedean. Thus we have obtained a Dedekind cut, andwe define f(a) = supL(a) = inf U(a), the real number determined bythis cut.

We show next that L(a) + L(b) ⊆ L(a + b), whence f(a) + f(b) ≤f(a + b). Together with a similar result for upper sets, this will implythat f is a group homomorphism. So let me ≤ na and ue ≤ vb; then(mv + nu)e ≤ nva + nvb = nv(a + b) as desired; note that we usedcommutativity (7.7) for the last equation.

It remains to be shown that f is injective. So let 0 < a in G. Bythe Archimedean property, we may choose n ∈ N such that e < na, andthen 1/n ∈ L(a) shows that f(a) > 0. �

7.9 Corollary Every non-trivial completely ordered group is isomor-

phic to one of the ordered groups Z or R.

Proof A completely ordered group G is Archimedean (7.5) and isomor-phic to a completely ordered and, hence, closed subgroup of R by 7.8.On the other hand, every subgroup of R is either cyclic or dense (1.4),and the assertion follows. �

7.10 Corollary If G is an ordered group which is isomorphic to R as

a chain, then G is isomorphic to R as an ordered group. �

In general, the connection between an ordered group and its underly-ing chain is much looser. For example, Alling–Kuhlmann 1994 showthat for any α > 0, there exist 2ℵα non-isomorphic ordered divisibleabelian groups of cardinality ℵα that are all isomorphic as chains.

The opposite of 7.10 is also far from true; namely, there are manynon-isomorphic ordered groups that are isomorphic to R+ as groups.For example, every hyperplane of the rational vector space R (compare1.11ff) is an ordered group (with the ordering induced from R) and isnot completely ordered, but Archimedean. A non-Archimedean orderingon the group R+ can be obtained using the isomorphism R+ ∼= R+⊕R+

of 1.15 and the lexicographic ordering on the latter group, defined by

(x, y) ≤ (x′, y′) � x < x′ or x = x′ ∧ y ≤ y′ .

Another example is given by the non-standard rationals; see 22.4 and22.8. See 7.14 and Alling–Kuhlmann 1994 for refinements of theseconstructions.

Page 96: The Classical Fields

7 The real numbers as an ordered group 79

Subgroups of the ordered group R abound; compare 1.15. Quotientgroups, on the other hand, are scarce. Indeed, the kernel of an epimor-phism ϕ : R → G of ordered groups is an interval since ϕ is monotone,hence ϕ = 0 or ϕ is an isomorphism.

In contrast to the huge automorphism group of the additive group R+

(see 1.32), the automorphism group of the ordered group R is rathertame:

7.11 Theorem The endomorphisms of the ordered group R are pre-

cisely the maps ϕa : x �→ ax, where 0 ≤ a ∈ R. Thus, the semi-

group End(R, +, <) is isomorphic to the multiplicative semigroup of

non-negative real numbers.

Proof Let ϕ be an order-preserving endomorphism. If ϕ(r) = 0 forsome r �= 0, then ϕ[0, r] = {0} by monotonicity, and then ϕ = 0 = ϕ0.Therefore, we may assume that ϕ is injective. We set a = ϕ(1) > 0 anddefine an order-preserving endomorphism ψ = ϕa−1ϕ; then ψ(1) = 1,and we have to show that ψ = id. Unique divisibility of R+ (see 1.8)implies that ψ(q) = q for all q ∈ Q. If ψ(r) �= r for some r ∈ R, thenthere is a rational number between ψ(r) and r (see 3.1), and we obtaina contradiction to monotonicity of ψ. �

The preceding result is a special case of the next one. Nevertheless,we include two independent proofs, as both of them are instructive.

7.12 Theorem For any subgroup G ≤ R+, the order-preserving group

homomorphisms ϕ : G→ R+ are precisely the maps ϕa : g �→ ag, where

0 ≤ a ∈ R.

Proof As in the previous proof, we only have to consider injective ho-momorphisms ϕ. For any two positive elements g, h ∈ G and m,n ∈ Nwe have the equivalence

m

n≤ h

g⇐⇒ mg ≤ nh ⇐⇒ mϕ(g) ≤ nϕ(h) ⇐⇒ m

n≤ ϕ(h)

ϕ(g).

In other words, the real numbers h/g and ϕ(h)/ϕ(g) are both equalto the supremum of the same set of rational numbers. Hence, the twonumbers are equal, and we infer that ϕ(g)/g = ϕ(h)/h is a positiveconstant a ∈ R. This shows that ϕ(g) = ϕa(g) for 0 < g, and it followsthat this equality holds for all g ∈ G. �

The exponential map is a group isomorphism R+ ∼= R×pos (see 2.2),

and is monotone. Hence R+ and R×pos are isomorphic ordered groups,

and we rephrase the previous result in terms of multiplication as follows.

Page 97: The Classical Fields

80 Real numbers

7.13 Corollary For any subgroup H ≤ R×pos, the order-preserving

group homomorphisms H → R×pos are precisely the maps h �→ ha, where

0 ≤ a ∈ R. �

The following result on the number of orderings of (R, +) will be usedlater for constructions of real closed fields and of involutions in Aut C;see 12.14 and 14.16.

7.14 Theorem Let G be the group R+, let ℵ = 2ℵ0 = cardG, and

let c ∈ {ℵ0,ℵ, 2ℵ}. Then there exist precisely 2ℵ isomorphism types of

ordered groups (G, <) such that card Aut(G, <) = c.

Proof There exist only 2ℵ (ordering) relations on the set R, hence itsuffices to find at least 2ℵ isomorphism types of ordered groups (G, <)as required. We construct rational vector spaces G of cardinality ℵ withsuitable orderings, rather than taking G = R+ directly; compare 1.17;the constructions use the axiom of choice.

First we treat the case c = ℵ0. The rational vector space R contains2ℵ hyperplanes (see Section 1, Exercise 4). Take for G any of thesehyperplanes, and endow it with the ordering induced from the naturalordering of R. By 7.12, each hyperplane is order-isomorphic to at most ℵother hyperplanes. Therefore we obtain 2ℵ isomorphism types of orderedgroups (G, <); compare 61.13a. Again by 7.12, the automorphism groupAut(G, <) consists of all maps x �→ rx with 0 < r ∈ R and rG = G.Clearly F := {r ∈ R | rG = G} ∪ {0} is a subfield of R, and G is avector space over F , hence also the quotient group R/G ∼= Q is a vectorspace over F . Thus cardF = ℵ0 = card Aut(G, <).

Next we consider the case c = ℵ. Choose any subfield F of R suchthat cardF = ℵ = [R : F ]; such subfields can be obtained by splittinga transcendency basis of R over Q into two parts of cardinality ℵ (see64.20). There exist 2ℵ subspaces G of R considered as a vector spaceover F . As in the previous paragraph, we use 7.12 to conclude that weobtain 2ℵ isomorphism types of non-trivial ordered groups (G, <). Herewe have card Aut(G, <) = ℵ; see 7.12 and observe that each map x �→ rx

with 0 < r ∈ F is an automorphism of (G, <).Now we deal with the case c = 2ℵ. Let (G, <) be one of the ordered

groups constructed so far, with card Aut(G, <) ∈ {ℵ0,ℵ}. We denote byG(R) the direct sum

⊕r∈R G endowed with the lexicographic ordering.

Then cardG(R) = ℵ by 61.14 and 61.13. The automorphism group ofthe ordered group G(R) contains the Cartesian power×r∈R Aut(G, <),which has cardinality 2ℵ, hence card AutG(R) = 2ℵ.

Page 98: The Classical Fields

8 The real numbers as a topological group 81

We claim that distinct isomorphism types of ordered groups (G, <)give distinct isomorphism types of ordered groups G(R); this impliesthat there are enough isomorphism types as required. For each positiveelement g ∈ G(R), the two sets Ag :=

⋃n∈N {x ∈ G(R) | −ng ≤ x ≤ ng }

and Bg :=⋂

n∈N {x ∈ G(R) | −g < nx < g } are convex subgroups.The ordered quotient group Ag/Bg is called a jump of G(R). Now g

has finite support, and we denote by r be the smallest real numberwith gr > 0. We obtain Ag = {x ∈ G(R) | xs = 0 for s < r } andBg = {x ∈ G(R) | xs = 0 for s ≤ r }, since (G, <) is an Archimedeanordered group. Therefore each jump determined by a positive elementg of G(R) is isomorphic to (G, <). �

Exercises(1) The vector group R2 becomes a non-Archimedean ordered group whenendowed with the lexicographic ordering defined by (a, b) < (a′, b′) � a < a′,or a = a′ and b < b′.(2) A non-commutative, non-Archimedean ordered group is obtained by takingon R2 the lexicographic ordering and the group operation (a, b) + (c, d) =(a+ c, b+ ead) of the semidirect product R � R; compare 9.4f. Generalize thisto arbitrary semidirect products of ordered groups.

(3) The set {2m3n | m, n ∈ Z} is dense in the set of positive real numbers.

(4) Let c ∈ R � Q and A = Z + Zc. Describe the structure of the group ofall those automorphisms of the group A which preserve the ordering inheritedfrom R.

8 The real numbers as a topological group

A topological group should be a group (G, +) carrying a topology com-patible with the group structure. We could model the compatibilityconditions after the conditions for ordered groups as expressed in theremark following 7.1. In other words, we would require that left or rightmultiplication with any fixed element defines a homeomorphism of G.This leads to the notion of a semi-topological group, which is too weakfor most purposes; compare 62.1 and the subsequent remarks. (Butat least, this shows that the topological space underlying a topologicalgroup is homogeneous.) Instead, we define topological groups as follows.

8.1 Definition Let (G, +) be a group and let τ be a topology on G.Then (G, +, τ) is called a topological group if addition (g, h) �→ g+h is acontinuous map G×G→ G and inversion g �→ −g is a continuous mapG→ G. Usually, it will be assumed that τ is a Hausdorff topology.

Page 99: The Classical Fields

82 Real numbers

If we equip R+ with the cofinite topology (a proper subset is closedif, and only if, it is finite), then we obtain a semi-topological group withcontinuous inversion, but addition R× R→ R is discontinuous. On theother hand, R+ with the Sorgenfrey topology (5.73) is a para-topologicalgroup, which means that addition is continuous in two variables, butinversion need not be (and in our example, is not) continuous.

8.2 Proposition R+ is a topological group with respect to the usual

topology on R.

Proof If x, x′, y, y′ ∈ R and |x − x′| < ε, |y − y′| < ε, then the triangleinequality gives |x + y − (x′ + y′)| < 2ε and | − x− (−x′)| < ε.

The same proof works for the additive group of any (skew) field withabsolute value (as defined in 55.1). �

We remark that local compactness of the group R+ is the most im-portant of its topological properties in the context of topological groups.Other relevant properties are connectedness, separability, and the factthat the topology is induced by a metric.

We shall see shortly that every ordered group is a topological group(8.4), and this will prove 8.2 once more in view of Section 7. While thisproof is less trivial than the one just given, its first step, which followsnext, will facilitate the construction of topological groups in general.The objective is to clarify what is needed for a semi-topological groupto be a topological group.

8.3 Proposition Let (G, +) be a group with a topology such that

(i) the left and right addition maps λc : x �→ c + x and ρc : x �→ x + c

of G are homeomorphisms,

(ii) addition (x, y) �→ x + y is continuous at (0, 0), and

(iii) inversion x �→ −x is continuous at 0.

Then G is a topological group.

Proof Writing the inversion map i as i = λ−ciρ−c, we see that it iscontinuous at c. Similarly, the decomposition (x, y) �→ (−c+x, y−d) �→−c + x + y − d �→ x + y of the addition shows that it is continuous at(c, d). �

8.4 Proposition Every ordered group (G, +, <) is a topological group

with respect to the topology induced by the ordering.

Proof We verify the conditions of 8.3. According to the remark follow-ing 7.1, the maps λc and ρc are order-preserving bijections and, hence,

Page 100: The Classical Fields

8 The real numbers as a topological group 83

homeomorphisms with respect to the topology induced by the ordering.Likewise, inversion is an order-reversing bijection and, again, a homeo-morphism.

Given a neighbourhood ]a, b[ of 0, we define c = min{−a, b}; thenI := ]−c, c[ ⊆ ]a, b[. We are looking for an interval J = ]−d, d[ suchthat J + J ⊆ I, thus proving continuity of addition at (0, 0). If c is asmallest positive element, then J := I will do. If not, choose d such that0 < 2d < c, applying 7.6. Again, J + J ⊆ I holds and G is a topologicalgroup by 8.3. �

Subgroups and quotients

As pointed out in Section 62, the adequate notion of substructure fortopological groups is that of a closed subgroup. This includes all opensubgroups (62.7), and the assumption of closedness implies that cosetspaces inherit a natural Hausdorff topology from the group (62.8).

The topological group of real numbers has very few closed subgroupsin spite of the abundance of subgroups in general (1.15). We recordthis fact in the following proposition, which is merely a restatement ofTheorem 1.4.

8.5 Proposition The closed subgroups of the topological group R+

are precisely {0}, the cyclic groups rZ ∼= Z, 0 �= r ∈ R, and R itself. �

In particular, the topological group R does not have any non-trivialcompact subgroups, nor any ‘small’ subgroups (that are contained ina preassigned neighbourhood of the neutral element). These propertiesplay an important role in the theory of locally compact groups. We shallshow in 8.19 that the property 8.5 characterizes R among all locallycompact, connected groups.

We complement the last result by its generalization to the vectorgroups Rn, which are topological groups as well (same proof as 8.2).

8.6 Theorem Let H be a closed subgroup of a vector group Rn.

Then there is a basis v1, . . . , vn of Rn such that H consists of all linear

combinations∑n

i=1 rivi satisfying ri ∈ Z for i > i1 and ri = 0 for i > i2,

where i1 ≤ i2 ≤ n. In particular, H is isomorphic to the direct sum

Ri1 ⊕ Zi2−i1 .

Proof By 8.5, the intersection of H with a one-dimensional subspaceRv of Rn is either all of Rv or infinite cyclic or trivial. Let U ≤ Rn

be the largest vector space contained in H and choose a vector space V

Page 101: The Classical Fields

84 Real numbers

complementary to U . Now H is the direct sum of U and H ∩ V , hencethe problem is reduced to the case that H contains no vector space �= 0.In this case, H is discrete, or else we find non-zero elements hk ∈ H

tending to 0, and we could assume that the vectors wk := hk‖hk‖−1

converge to some vector w with ‖w‖ = 1. As in 8.5, an argument similarto the proof of 1.4 would then show that the set {z‖hk‖ | k ∈ N, z ∈ Z}is dense in R, and this implies that Rw ⊆ H, a contradiction.

Take an element u ∈ H � {0} such that ‖u‖ is minimal, and considerthe image G ∼= H/〈u〉 of H in Rn/Ru. This is again a discrete group.Otherwise, there are elements gk ∈ H � Ru and real numbers rk suchthat gk−rku→ 0. Adding suitable integers to rk, we could arrange thatthe set {rk | k ∈ N} is bounded, contradicting the discreteness of H.Using induction over n, we may assume that G consists of all integralcombinations of some independent vectors v2, . . . , vs ∈ Rn/Ru. If wechoose an inverse image vi ∈ H for each vi, we get a set v1 = u, v2, . . . , vs

having the properties asserted in the theorem. �

Next we turn to quotients of the topological group R+. Only quotientsmodulo closed subgroups inherit a reasonable structure, and 8.5 saysthat there is only one possibility up to isomorphism.

8.7 The circle group as a topological quotient group of R+ Thecircle S1 has been introduced in Section 5 as the set of all vectors oflength 1 in R2. Identifying R2 with the field C of complex numbers, wesee that S1 is a closed subgroup of the multiplicative group C×. Theformulae (x + iy)(u + iv) = xu − yv + i(xv + yu) and (x + iy)−1 =(x2 + y2)−1(x − iy) show that multiplication and inversion in C× arecontinuous, and C× is a topological group containing S1 as a closedsubgroup.

Now remember the surjective map

p : R→ S1 : t→ e2πit

introduced in 5.16. It is a group homomorphism by the exponentiallaw ez+w = ezew, and the topology of S1 is the quotient topology withrespect to this map; see 5.16. It follows that S1 is isomorphic to thetopological quotient group of R modulo ker p = Z, as introduced in62.11.

As an abstract group, the quotient R/Z appeared first in 1.20 and wasdenoted T. The symbol abbreviates the name torus group, which is pri-marily used for the topological group S1 and for the higher dimensionaltori S1×· · ·×S1 = Tn (where n is the number of factors); compare 5.64.

Page 102: The Classical Fields

8 The real numbers as a topological group 85

Originally, the name torus refers to the topological space S1 × S1, thedoughnut surface. Because of the isomorphism

S1∼= T = R+/Z,

we shall not be too strict in distinguishing between the group S1 ≤ C×

and the factor group T = R/Z, but at least for the next proposition andits proof, we shall be careful about this point.

8.8 Proposition (a) The closed subgroups of the torus group S1∼= T

are precisely the groups S1 and the finite cyclic groups generated by

the roots of unity e2πi/n, where n ∈ N.

(b) The only connected subgroups of S1 are S1 itself and the trivial

group.

(c) The open set {z ∈ S1 | Re z > 0} does not contain any non-trivial

subgroup of S1.

Proof (a) Let G ≤ S1 be a closed subgroup. Then p−1(G) is a closed sub-group of R; compare 8.7. Moreover, this subgroup contains the number1 ∈ ker p. Now the assertion follows from 8.5.

(b) If G ≤ S1 is a non-trivial connected subgroup, then p−1(G) con-tains a non-trivial connected subset and, hence, contains some neigh-bourhood of 0. It follows that p−1(G) = R.

(c) If G ≤ S1 is a non-trivial subgroup contained in the right half plane,then the closure G is one of the groups listed in (a) and is contained inthe closed right half plane. Such a group does not exist. �

8.9 Theorem Let G be a non-trivial (Hausdorff ) topological group.

If there is a continuous epimorphism R → G, then G is isomorphic to

the torus T as a topological group or isomorphic to R as a group. In the

second case, the topology of G may be coarser than the topology of R;

compare 5.66.

Proof The kernel of the given epimorphism ϕ is a closed subgroup ofR because G is a Hausdorff space. By 8.5, it follows that either ϕ isinjective, which implies that G ∼= R as a group, or after rescaling (usingϕ(rt) instead of ϕ(t)) we may assume that ker ϕ = Z. Then ϕ inducesa group isomorphism ψ : R/Z → G such that ϕ = ψp, where p isthe canonical epimorphism R → R/Z. Moreover, it follows from thedefinition of the quotient topology that ψ is continuous. Indeed, thepreimage ψ−1(U) of an open set U ⊆ G has open preimage under p

because ϕ = ψp is continuous, and this property characterizes ψ−1(U)as an open set.

Page 103: The Classical Fields

86 Real numbers

Now ψ is a continuous bijection of the compact space R/Z ≈ S1 ontothe Hausdorff space G, hence ψ is a homeomorphism, and an isomor-phism of topological groups. �

Characterizations

The following result shows that the topology of R determines the groupstructure. We shall return to this point later (8.16).

8.10 Theorem Let G be a topological group whose underlying topo-

logical space is homeomorphic to R. Then G is isomorphic to R as a

topological group. In particular, G is abelian.

Proof We show that there is an ordering on G inducing the given topo-logy and making G an ordered group. Since G is connected by assump-tion, the ordering is complete (3.3), and the assertion is a consequenceof 7.9.

We define the ordering on G such that the given homeomorphismG → R becomes an isomorphism of chains. We have to show that theleft and right addition maps λc and ρc of G are monotone; compare 7.1and the subsequent remarks. Since G is a topological group, we knowthat these maps are homeomorphisms. Hence, they are either monotoneor antitone; compare 5.51. Any antitone homeomorphism f of R has afixed point (apply the intermediate value theorem to f − id). For c �= 0it follows that λc and ρc are indeed monotone. �

We shall push the last result as far as possible in 8.21 and 8.22. Herewe want to prove a similar result for the torus group. This will be aneasy application of some standard results about covering maps, whichwe present first. A good reference is Greenberg 1967 §§5 and 6.

8.11 Definition (a) A surjective continuous map p : E → B of topo-logical spaces is called a covering map if every point b ∈ B has an openneighbourhood U such that p−1(U) is a disjoint union of open subsetsUi, i ∈ I, each of which is mapped homeomorphically onto U by p.Neighbourhoods like U will be called special neighbourhoods, and the Ui

are called the sheets over U .(b) A topological space X is said to be simply connected if X is path-

wise connected and if every loop can be contracted in X. This meansthat, given a continuous map f : [0, 1]→ X satisfying f(0) = f(1) = x0,there is a continuous map F : [0, 1]× [0, 1]→ X such that F (s, 0) = f(s)for all s and F (s, t) = x0 whenever s ∈ {0, 1} or t = 1.

Page 104: The Classical Fields

8 The real numbers as a topological group 87

The standard example of a simply connected space is Rn, where themap F is most easily defined by F (s, t) = tx0 + (1 − t)f(s). The mostprominent example of a covering map is the quotient map p : R→ S1 of8.7. Every proper open subset U ⊆ S1 is a special neighbourhood; thisfollows from the existence of the complex logarithm function.

Instead of going into the details of the theory of covering spaces, weshall just quote the one result that we need. For a proof, see Greenberg

1967 loc. cit.

8.12 Theorem Let p : E → B be a covering map, and h : X → B

a continuous map with X simply connected. If x ∈ X and e ∈ E are

points such that h(x) = p(e), then there is a unique continuous map

h : X → E sending x to e and satisfying h = ph.

E

X B�p

��

��eh

�h

The map h is called a lift of h over p. We apply this result to topo-logical groups:

8.13 Theorem (a) Let (G, +) be a topological group with neutral ele-

ment 0 and p : E → G a covering map with E simply connected.

Given e ∈ E such that p(e) = 0, there is a unique group structure

on E with neutral element e such that E is a topological group and

p is a group homomorphism.

(b) If H is a simply connected topological group, then every continuous

group homomorphism ψ : H → G lifts to a unique continuous group

homomorphism ψ : H → E, that is, ψ = pψ.

(c) Every continuous group endomorphism ϕ : G→ G lifts to a unique

continuous group endomorphism ϕ : E → E, that is, pϕ = ϕp.

E E

G G

�eϕ

�p

�p

�ϕ

Proof (a) Let h : E × E → E be given by h(a, b) = p(a) + p(b). It iseasy to verify that E × E is simply connected, hence there is a uniquelift h : E × E → E sending (e, e) to e. This map is going to be the

Page 105: The Classical Fields

88 Real numbers

addition in E, that is, a + b = h(a, b). Similarly, inversion in E isdefined as a lift of inversion in G. The group axioms are all verified bythe same method, exploiting the uniqueness part of 8.12. For example,consider the associative law. It says that the two maps E ×E ×E → E

defined by f1(a, b, c) = (a + b) + c and by f2(a, b, c) = a + (b + c)agree. Now the corresponding maps in G do agree, and f1 is a lift of(a, b, c) �→ (p(a) + p(b)) + p(c), similarly for f2. Thus we have two liftsof the same map, which agree at the point (e, e, e), so they are equal.

(b) The existence of ψ is immediate from 8.12 (for x and e, take theneutral elements). The homomorphism property is again translated intoequality of two lifts, namely of (a, b) �→ ψ(a)+ ψ(b), which lifts the map(a, b) �→ ψ(a)+ψ(b), and (a, b) �→ ψ(a+ b), which lifts (a, b) �→ ψ(a+ b).Again, the two lifts are equal because they agree at (e, e).

(c) This follows if we apply (b) to ψ = ϕp. �

Now we return to the torus group.

8.14 Theorem Every topological group homeomorphic to S1 is iso-

morphic as a topological group to T.

Proof Using the covering map p : R → S1 and 8.13, we may define atopological group structure on R such that p becomes a continuous groupepimorphism. The result follows from 8.9; note that p is not injective,so we are dealing with the first case of 8.9. �

8.15 Theorem The only topological groups whose underlying spaces

are connected 1-manifolds are R and T.

Proof The essential step is to show that the space of such a group G isseparable; then we can apply the classification of connected, separable1-manifolds (5.17) and our results 8.10 and 8.14.

Let U be an open neighbourhood of 0 homeomorphic to R. ThenV := U ∩ (−U) = −V is a separable open set. The union W of the setsnV = V + · · · + V = {v1 + · · · + vn | vi ∈ V } with n ∈ N is an opensubgroup, hence also closed (62.7, all cosets are open), and W = G byconnectedness. Now W is separable: first, V contains a countable denseset A, giving rise to a countable dense set An = A×· · ·×A ⊆ V n; underaddition, An is mapped to nA, hence nA is dense in nV , and the unionof all sets nA, n ∈ N, is still countable by 61.13. �

The last result can be obtained as a corollary of deep results in thetheory of locally compact groups. It has been proved by Montgomery,Zippin, and Gleason (see Montgomery–Zippin 1955 4.10 p. 184) that

Page 106: The Classical Fields

8 The real numbers as a topological group 89

every locally Euclidean topological group is a Lie group, that is, it ad-mits an analytic structure such that the group operations (addition andinversion) are analytic maps. Lie groups can be classified using theirLie algebras, which are linearizations of the group structures. Then 8.15reduces to the rather easy classification of 1-dimensional Lie groups.

8.16 Independence of group structure and topology We have justseen that the topology of R and of T determines the group structure upto isomorphism. The converse is far from true, and we give a numberof examples of non-homeomorphic topological groups whose underlyinggroups are isomorphic to T or to R.

(1) The group R+ is isomorphic to any vector group Rn (see 1.15), andRn is a topological group. Unlike R, the topological space Rn, n > 1,remains connected after deletion of a point, hence the two spaces arenot homeomorphic.

(2) R with the torus topology 5.65 is a (non-closed) subgroup of thetopological group T2 (compare 5.66), and hence is again a topologicalgroup. Proposition 5.67 amply demonstrates that this topology is nothomeomorphic to the usual one.

(3) Let A be a subspace of the rational vector space R such that A hasthe same (transfinite) dimension as R itself. Then the additive group A+

is isomorphic to R+; compare 1.15. For example, A could be a hyper-plane. In any case, A is a (non-closed) subgroup of the topological groupR, hence A is a topological group. The topology of A is totally discon-nected and not locally compact, because both A and its complement aredense in R.

(4) By 1.17, the group R+ is a direct sum of ℵ copies of Q+. Thedirect sum is a subgroup of the direct product, which is a topologicalgroup with respect to the product topology, and the direct sum is atopological group with respect to the induced topology.

(5) The solenoid Q∗d introduced in 8.29 is compact, connected, and

isomorphic to R as a group; see 8.32 and 8.33.(6) It was shown in 1.25 that T ∼= Rn ⊕ T as groups. Of course, both

are topological groups, but one of the spaces (T) is compact, and theother is not.

(7) A systematic investigation of group topologies on R is presentedin Ch’uan–Liu 1981, leading to the result that there are exactly 22ℵ

non-isomorphic topological groups (ℵ = card R) that are algebraicallyisomorphic to R. Among these, countably many are compact, and sim-ilar results are given for topologies with other special properties.

Page 107: The Classical Fields

90 Real numbers

By contrast, the topology on the additive group of integers is uniquelydetermined under mild additional conditions, as we show now.

8.17 Theorem The discrete topology is the only locally compact Haus-

dorff topology making the additive group Z of integers (or, in fact, any

countable group) into a topological group.

Proof Our group is a union of countably many singletons {x}, which areclosed by assumption. At least one of the singletons must contain a non-empty open subset. This follows from the Baire category theorem, whichholds in locally compact spaces, see Dugundji 1966 XI.10.3 p. 250. Weconclude that some (and, hence, every) singleton is open. �

Theorem 5.15 characterizes the topological group R, and we shall con-tinue by proving a number of other characterization theorems. The firstone (8.18) is due, in the form presented here, to Morris 1986. The proofuses the theorem of Mal’cev–Iwasawa, and also the theorem of Peter–Weyl, which says that each compact group has enough (63.11) contin-uous unitary representations. A complex matrix A is called unitary, ifAAT = 1; thus a one-dimensional continuous unitary representation isthe same thing as a character.

8.18 Theorem Let G be a non-discrete, locally compact Hausdorff

topological group, and suppose that all proper closed subgroups of G

are discrete. Then the topological group G is isomorphic to the additive

group R of real numbers or to the torus group T.

Proof We denote by G1 the connected component of G, that is, thelargest connected subgroup; note that G1 is closed, because the closureof a connected set is connected (compare 62.13). We infer that eitherG1 = G, or G1 is discrete and therefore trivial. In the second case,G has a neighbourhood basis at 1 consisting of compact open propersubgroups (see 62.13 and Exercise 3 of Section 63); these subgroupsare discrete, so G is discrete, which is a contradiction. Hence G isconnected. The theorem of Mal’cev–Iwasawa (62.14) then shows in viewof our hypothesis that G is either compact or isomorphic to R. It remainsto consider the case where G is connected and compact.

We claim that G contains an element g of infinite order. The Peter–Weyl theorem yields a non-trivial continuous unitary representation ϕ :G → GLnC; see Hofmann–Morris 1998 2.27, 2.28 p. 44, Stroppel

2006 14.33 (compare also Hewitt–Ross 1963 22.12, 22.13, Pontrya-

gin 1986 §32, §33, Adams 1969 3.39). The corresponding trace mapG → C : g �→ trϕ(g) is continuous and not constant, since tr ϕ(g) = n

Page 108: The Classical Fields

8 The real numbers as a topological group 91

implies ϕ(g) = id (note that ϕ(g) can be diagonalized and has eigen-values of absolute value 1). If all elements g ∈ G have finite order,then the eigenvalues of ϕ(g) are roots of unity. Hence the connected set{ trϕ(g) | g ∈ G} �= {n} consists of finite sums of roots of unity and istherefore countable, which is a contradiction.

An element g ∈ G of infinite order generates a dense subgroup of G;otherwise the closure of 〈g〉 would be a proper compact subgroup, hencediscrete by assumption, and thus finite, which is absurd. We concludethat G is abelian.

Thus by Theorem 63.13, there exists a non-trivial character χ : G→ T(in fact, the representation ϕ can be chosen to be irreducible; then ϕ

is one-dimensional, hence χ = ϕ is such a character). The connectedimage χ(G) is all of T, and the kernel of χ is a proper closed subgroup,hence discrete and finite. It is straightforward to show that G is locallyhomeomorphic to T, hence a compact 1-manifold. Now 5.17 and 8.14imply that G ∼= T as topological groups. �

As a consequence, we show that the property 8.5 (all non-trivial properclosed subgroups are infinite cyclic) is in fact a characterizing propertyof R.

8.19 Corollary Let G be a non-discrete, locally compact Hausdorff

topological group, and suppose that all proper closed subgroups of G are

cyclic. Then the topological group G is isomorphic to the additive group

R of real numbers or to the torus group T. If all non-trivial proper closed

subgroups of G are infinite cyclic, then G ∼= R is the only possibility.

Proof The hypothesis of 8.19 implies that of 8.18, because a closed cyclicsubgroup of G discrete by 8.17. The second statement is easily deducedsince the torus contains non-trivial finite subgroups (in fact, we obtaina proof for the second statement in 8.19 simply by omitting the secondparagraph of the proof for 8.18). �

The next result is due to Montgomery 1948. We need the notion oftopological dimension. By definition, a topological space X has (cover-ing) dimension dimX ≤ n if every finite open cover of X has a refine-ment containing at most n+1 sets with non-empty intersection. A goodreference on dimension is Nagami 1970; compare also Salzmann et al.1995 Section 92 and Theorems 93.5 to 93.7.

8.20 Theorem Let G be a locally compact, connected topological

group. If G is one-dimensional and not compact, then G is isomorphic

to the topological group R.

Page 109: The Classical Fields

92 Real numbers

Proof The theorem of Mal’cev–Iwasawa (62.14) shows that G is homeo-morphic to C × Rk, where C ≤ G is a (maximal) compact subgroup.This implies that k = dim Rk ≤ dim G, hence k = 1 since G is not com-pact. It follows that dimG = dimC +1; this is not as trivial as it seems,see Nagami 1970 42-3 for a proof; compare also Salzmann et al. 199592.11. Now dimC = 0 implies that C is totally disconnected. On theother hand, connectedness of G ≈ C × R implies that C is connected.Finally, we conclude that C = {0}, and then 8.10 or the theorem ofMal’cev–Iwasawa shows that G ∼= R as a topological group. �

The next result should be compared with 8.10. Instead of G ≈ R wemake only weak assumptions concerning connectivity. Yet we obtain arather strong assertion.

8.21 Theorem Let G be a Hausdorff topological group. If G is con-

nected but G�{0} is not, then there is a continuous group isomorphism

ϕ : G → R. Thus, G can be obtained from the topological group R by

refining the topology.

Proof (1) We claim that H = G � {0} has precisely two connectedcomponents (later we shall use this fact to define an ordering on G). LetH = U∪V be a separation as defined in 5.7, that is, the sets U and V arenon-empty, open and disjoint. If both U and V are connected, then thereis nothing to prove. If not, then one of the two sets admits a separation,and H = U∪V ∪W with non-empty, disjoint open sets U, V,W . Applying5.8c to U and V ∪W , we obtain that U = U ∪ {0} is connected. Selectconnected components CU , CV , CW of U, V,W , respectively. Then theproduct CV × U is connected, and continuity of addition and inversionimplies that CV +(−U) is connected, too. This set is contained in H andcontains the connected component CV , hence CV + (−CU ) ⊆ CV . Bysymmetry, we also have CU +(−CV ) ⊆ CU and, after taking inverses onboth sides, CV +(−CU ) ⊆ −CU . These results show that the intersectionCV ∩ (−CU ) is not empty. Both these sets are connected components ofH, hence they are equal. The same holds with W in place of V , whichis a contradiction.

(2) Repeating the argument (1) with two sets instead of three, we seethat the connected components of H are interchanged by inversion, sowe denote them by C and −C. We define x ≤ y � y−x ∈ C = C ∪{0},and we claim that this turns G into an ordered group. First we have toshow that the relation ≤ transitive, i.e., that x ≤ y and y ≤ z togetherimply x ≤ z. The connected set C+C ⊆ H contains C, hence C = C+C

Page 110: The Classical Fields

8 The real numbers as a topological group 93

and C = C + C, and transitivity follows. From C ∩ −C = {0} we inferthat x ≤ y and y ≤ x together imply x = y, and ≤ is an orderingrelation.

Suppose that a ≤ b and x �= 0. Then b + x − (a + x) = b − a showsthat a + x ≤ b + x. In order to prove that x + a ≤ x + b, i.e., thatx + (b − a) − x ∈ C, we need to know that x + C − x ⊆ C. Nowx + C − x is a connected subset of H, and contains an element of C

(namely, x or −x, whichever belongs to C). Therefore, x + C − x ⊆ C

and x + C − x ⊆ C. Thus, we have defined an ordered group.(3) We compare topologies: the sets C, C + a and −C + b are open in

the given topology of G, hence the intersection ]a, b[ = (C+a)∩(−C+b)is open, too. Therefore, the identity mapping from G with the giventopology to G with the order topology is continuous, and the ordertopology is connected. By 3.3, it follows that the ordering is complete,and 7.9 shows that the ordered group G is isomorphic to Z or to R;however, the first possibility is excluded by connectedness. �

Groups G as described in 8.21 do exist; we give an example in 8.23.One additional hypothesis, however, suffices to exclude these possibili-ties, as we prove next.

8.22 Theorem If G satisfies the hypothesis of 8.21 and is locally

compact, then G is isomorphic to the topological group R.

Proof Any compact subgroup C ≤ G is mapped by ϕ to a compactsubgroup of R. Using injectivity of ϕ together with 8.5, we infer thatC = {0}. Now the theorem of Mal’cev–Iwasawa (62.14) implies thatG is homeomorphic to Rk for some k. For k > 1, Euclidean space Rk

is not separated by any point, hence k = 1 by our hypothesis. In thiscase, the theorem of Mal’cev–Iwasawa (or 8.10) asserts that G ∼= R as atopological group. �

A counter-example

We show that Theorem 8.21 cannot be improved in the sense that G

is isomorphic to R as a topological group. The example given in theproof of 8.23 below is due to Livenson 1937. See also 14.9 for an evenstronger result.

8.23 Theorem There is a non-closed subgroup G ⊆ R2 such that G,

but not G � {0}, is connected in the topology induced by R2, and such

that G as a topological group is not isomorphic to R.

Page 111: The Classical Fields

94 Real numbers

Proof We obtain G as the graph of a certain discontinuous group ho-momorphism ψ : R → R, i.e., G = { (x, ψ(x)) | x ∈ R}. The homo-morphism is chosen in such a way that G meets the boundary of almostevery open set of R2.

There is a countable basis for the topology of R2, hence there exist(at most) ℵ = 2ℵ0 open subsets of R2. A Hamel basis B of R as a vectorspace over Q also has cardinality ℵ (see Theorem 1.12), hence there isa surjective map f : B → U onto the following collection U of open setsin R2.

For an open set U ⊆ R2, define U = p(U) ∩ p(R2 � U), where p :R2 → R is the projection onto the first factor, p(x, y) = x. The set U

belongs to U if U �= ∅. In other words, the condition for U is that someline x = const (a parallel to the y-axis) meets both U and R2 � U and,hence, meets the boundary ∂U . In fact, this happens for all x in theopen set U .

Now assume that b ∈ B and U = f(b) ∈ U . Then bQ contains a non-zero element of U , and we may choose ψ(b) ∈ R such that Q(b, ψ(b))intersects ∂U . In this way, we obtain a homomorphism ψ : R → Rwhose graph G meets the boundary of every U ∈ U . An immediateconsequence is that G is dense in R2.

The y-axis (x = 0) separates R2 and meets G only in the point 0 =(0, 0), hence G � {0} is disconnected.

In order to show that G is connected, we consider the intersectionC = G ∩ H with the half plane H = { (x, y) | x > 0}. We shall showthat C is connected; then also C ∪ {0} ⊆ C and G = C ∪ {0} ∪ −C areconnected.

Suppose that C is disconnected. Then there exist open subsets U, V

of H such that C ⊆ U ∪ V and U ∩ V ∩ C = ∅. The latter conditionimplies that U ∩ V = ∅, because C is dense and U ∩ V is open. Thenmoreover U ∩ V = ∅ because U is contained in the closed set H � V .Now consider the case U ∈ U . Then ∂U contains an element of C; thiselement is not contained in U , hence it belongs to V ∩ U , which is acontradiction. On the other hand, if U /∈ U , then p(U)∩p(V ) ⊆ U = ∅.Moreover, p(U) and p(V ) are open sets, but their union p(U) ∪ p(V ) =p(C) = { (x, 0) | x > 0} is connected, which is a contradiction.

It remains to be shown that G is not isomorphic to R as a topologicalgroup, and this is very easy. In the rational vector space R, every one-dimensional subspace is dense with respect to the usual topology. SinceG is dense in R2, this is far from true in G. �

Page 112: The Classical Fields

8 The real numbers as a topological group 95

Automorphisms and endomorphisms

To formulate the appropriate notion of automorphisms of a topologicalgroup G, we require that automorphisms are at the same time homeo-morphisms of the underlying topological space and group homomor-phisms. Automorphisms in this sense form a group denoted Autc G.

8.24 Proposition The automorphisms of the topological group R are

precisely the multiplication maps ϕa : x �→ ax, where 0 �= a ∈ R. In

particular, the automorphism group Autc R is isomorphic to R×.

Proof Everything is similar to 7.11, only a = ϕ(1) may be any non-zeronumber, and ϕ may be antitone; compare 5.51. �

8.25 Definition The endomorphisms of an abelian group G form aring EndG called the endomorphism ring of G. Multiplication of thisring is the composition of maps and addition is defined by (ϕ + ψ)(x) =ϕ(x) + ψ(x).

Note that commutativity of G is indispensable; the homomorphismproperty of id + id requires that x + x + y + y = x + y + x + y holds asan identity in G.

If G is a topological group, then the continuous endomorphisms forma subring, denoted Endc G, of EndG.

8.26 Theorem The continuous endomorphisms of the topological

group R are precisely the maps ϕa for arbitrary a ∈ R. In particu-

lar, the ring Endc R of continuous endomorphisms is isomorphic to the

field R.

Proof Each endomorphism ϕ of R+ is Q-linear, by the unique divisibilityof R+ (1.8, 1.9). Let a := ϕ(1). Then ϕ(x) = ϕ(1 · x) = a · x for x ∈ Q.As Q is dense in R, the continuity of ϕ implies that this equation holdsfor all x ∈ R. Thus ϕ = ϕa. �

For a stronger statement, see Exercise 4.

8.27 Theorem The ring Endc T of continuous endomorphisms of T is

isomorphic to the ring Z of integers, and Autc T ∼= C2 is cyclic of order 2.

Proof By 8.13c, every continuous endomorphism ϕ of T lifts to a con-tinuous endomorphism ϕ of R, that is, ϕp = pϕ. From this equation itfollows that ϕ maps ker p = Z into itself. From 8.26, we know that ϕ

has the form ϕa for some a ∈ R, and a = ϕa(1) must be an integer.Conversely, if a ∈ Z then ϕa can be ‘pushed down’ to an endo-

morphism of T, sending x ∈ T to xa. This gives the isomorphism

Page 113: The Classical Fields

96 Real numbers

Endc T ∼= Z, and the automorphisms are the invertible elements in thisring, that is, the maps ϕa for a ∈ {1,−1}. �

In the remainder of this section, we consider topological groups whoseendomorphism rings are skew fields.

Groups having an endomorphism field

Let G be a locally compact, connected abelian topological group allof whose non-trivial continuous endomorphisms are invertible, so thatEndc G is a skew field. We shall show that there are exactly two suchgroups G. To warm up, we treat the analogous question for groupswithout topology (or with discrete topology); the result 8.28 is due toSzele 1949; see also Fuchs 1973 111.1. In fact, this is more thanan exercise: the connected case will be reduced to the discrete caseusing character theory. Thus the reader should absorb the definition ofcharacter groups from Section 63, and then refer to that section as needarises.

8.28 Theorem Let G be an (abstract) abelian group. The ring EndG

is a skew field if, and only if, G ∼= Cp is cyclic of prime order or G ∼= Q+ is

the additive group of rationals. We have field isomorphisms EndCp∼= Fp

if p is a prime, and End Q+ ∼= Q.

Proof (1) Suppose that EndG is a skew field, and let F be its (commu-tative) prime field, i.e., the smallest subfield of EndG. Then F ∼= Fp isof prime order, or F ∼= Q. The group G is a vector space over EndG

and, hence, over F ; the product of a scalar ϕ ∈ EndG and a vectorg ∈ G is defined by applying ϕ to the group element: ϕg = ϕ(g). Everyvector space of dimension more than 1 has plenty of non-invertible en-domorphisms, which are also endomorphisms of the underlying abeliangroup. (Here we used the existence of bases and, hence, the Axiom ofChoice.) Thus G is isomorphic to the additive group of F .

(2) Conversely, let ϕ be an endomorphism of Cp. The order of asubgroup divides the order of the group, hence the only subgroups ofCp are {0} and Cp itself. Either kerϕ = {0} and ϕ is bijective, orker ϕ = Cp and ϕ = 0.

Now let ϕ be an endomorphism of Q+. If n ∈ N and ϕ(nq) = 0, thennϕ(q) = 0 and ϕ(q) = 0. Thus, kerϕ is divisible. The only divisiblesubgroups of Q+ are {0} and Q+ itself, hence ϕ = 0 or ϕ is injective. Inthe latter case, ϕ(Q) is a non-trivial divisible subgroup of Q+, and ϕ isbijective. �

Page 114: The Classical Fields

8 The real numbers as a topological group 97

8.29 Theorem There are only two connected, locally compact, Haus-

dorff abelian groups G such that the ring Endc G of continuous endomor-

phisms is a skew field, namely the topological group R and the character

group Q∗d, where Qd denotes the additive group of rationals with discrete

topology. We have field isomorphisms Endc R ∼= R and Endc(Q∗d) ∼= Q.

The group Q∗d belongs to the class of so-called solenoids. It will be

featured in 8.32 below.

Proof Suppose that G is a group satisfying the assumptions of the theo-rem and that Endc G is a skew field. In our abelian case, the theorem ofMal’cev–Iwasawa asserts that G ∼= C ⊕ Rn, where C is a compact con-nected (abelian) group; see 63.14. There are only two cases where theexistence of non-invertible endomorphisms ϕ �= 0 is not obvious, namely,C = {0} and n = 1, i.e., G ∼= R, and n = 0, i.e., G is compact. In thefirst case, we know that Endc G ∼= R is a field (8.26).

In the remaining case, G compact, we shall apply character theory.It suffices to determine the character group G∗, because the Pontryaginduality theorem asserts that G ∼= G∗∗; see 63.27. By 8.30 below, Endc G

is anti-isomorphic to Endc(G∗) as a ring, and G∗ is a discrete groupaccording to 63.5. Thus, all endomorphisms of G∗ are continuous andwe may apply 8.28; we find that G∗ ∼= Cp (a cyclic group of primeorder) or G∗ ∼= Qd. The first possibility entails that G = C∗

p is a finiteHausdorff space, hence disconnected.

The second possibility leads to G = Q∗d, which is a connected group

because Qd has no compact (i.e., finite) subgroups other than {0}; com-pare 63.30. Moreover, Q∗

d is compact according to 63.5, and Endc(Q∗d) ∼=

Endc(Qd) = End Q+ ∼= Q by 8.28; note that an anti-isomorphism is anisomorphism in this case. �

8.30 Theorem Let G be a locally compact abelian group and G∗ its

character group. Then the ring Endc G is anti-isomorphic to the ring

Endc(G∗).In fact, let γ∗ denote the adjoint of γ ∈ Endc G as defined in 63.3.

Then γ �→ γ∗ is bijective and satisfies (γδ)∗ = δ∗γ∗ and (γ+δ)∗ = γ∗+δ∗.

Proof The identities follow directly from the definition of the adjoint.Our notation differs from 63.3, where γ acts from the right and γ∗ actsfrom the left. Here, all maps act from the left, which enforces the reversalof factors in the first identity.

We have to show that γ �→ γ∗ is bijective. We use the duality theorem63.27 and identify A∗∗ with A via the duality isomorphism. This means

Page 115: The Classical Fields

98 Real numbers

that we define γ∗ for γ ∈ Endc G by 〈γ(g), χ〉 = 〈g, γ∗(χ)〉 and σ∗ forσ ∈ Endc(G∗) by 〈g, σ(χ)〉 = 〈σ∗(g), χ〉; here, g ∈ G and χ ∈ G∗. Nowwe have 〈γ(a), χ〉 = 〈a, γ∗(χ)〉 = 〈γ∗∗(a), χ〉 for every χ ∈ G∗ (and forfixed a). Since G has enough characters (compare 63.12), it follows thatγ = γ∗∗, which proves bijectivity. �

To end this section, we compute the character groups of R and of somerelated groups (see also Exercise 5 of Section 52), and finally we give adescription of Q∗

d.

8.31 Proposition (a) The topological group R is self-dual, R∗ ∼= R.

(b) The character group T∗ is isomorphic to Z, the discrete group of

integers. An isomorphism Z → T∗ is given by n �→ χn, where

χn : T→ T sends z ∈ T to its power zn.

(c) The character group Z∗ is isomorphic to the torus group T. An

isomorphism T → Z∗ is given by z �→ χz, where χz : Z → T sends

the integer n to zn.

Proof (1) We begin with the simpler parts. As a group, T∗ is simplythe additive group of the ring Endc T, which is isomorphic to the ring ofintegers (8.27). The topology of T∗ is discrete by 63.5. Of course, thiscan be seen directly: the endomorphisms of T are of the form z �→ zn,n ∈ Z, and if a sequence of such maps converges (pointwise convergencesuffices), then it is finally constant. This proves assertion (b), and (c)follows by the duality theorem 63.27. Again, this can be verified directly,a character χ of Z is determined by the image χ(1) ∈ T, and this definesthe isomorphism Z∗ → T.

(2) Theorem 18.3b shows that for every character χ : R → T thereis a unique continuous endomorphism χ : R → R such that χ = pχ,where p : R → T is the quotient map, p(t) = e2πit; note that we aretreating T as a multiplicative group once more. According to 8.26, wehave χ(t) = rt for some real number r, and χ(t) = e2πirt =: χr(t).Clearly, χr+s = χrχs, and only the topology of R∗ ∼= R+ remains to bedetermined.

It suffices to check the continuity of the group homomorphism r �→ χr

and of its inverse at the neutral elements 0 ∈ R and 1 ∈ T, respectively.So let C ⊆ R be compact and let U ⊆ T be a neighbourhood of 1. Wewant to determine ε > 0 such that |r| < ε implies that e2πirc ∈ U for allc ∈ C. This is possible because (r, t) �→ e2πirt is continuous. Conversely,let ε > 0 be given. We have to specify a compact set C ⊆ R and aneighbourhood U of 1 in T such that χr(C) ⊆ U implies |r| < ε. We

Page 116: The Classical Fields

8 The real numbers as a topological group 99

let C = [0, 1] ⊆ R and choose a sufficiently small U and a local inverseLog : U → R of p such that Log(1) = 0 (a modification of the complexlogarithm). If χr(C) ⊆ U , then χr = Log χr holds on C. Continuityof Log implies that rC = χr(C) ⊆ ]−ε, ε[ if U is small enough. Thisimplies that |r| < ε. �

8.32 The solenoid Q∗d We shall describe the character group of the

rationals in various respects. For proofs and for further information,we refer to Hewitt–Ross 1963 10.12-10.13 and 25.4-25.5. Being thecharacter group of a discrete group, the solenoid Q∗

d is compact (63.5).Since Q does not have any non-trivial compact subgroups, the solenoidis connected (63.30).

An explicit description can be derived from the following description ofQ: the set of rationals can be obtained as the union of the sets (n!)−1Z,n ∈ N. More abstractly, this can be expressed by saying that Q is theinductive limit of the sequence

Z2−→Z

3−→Z4−→Z

5−→Z −→ . . . ,

where the homomorphism marked n is given by z �→ nz. Via Pontryaginduality, the inductive limit is transformed into the projective limit of thedual sequence; see Hofmann–Morris 1998 7.11(iv). The dual sequenceis given by

T2←−T

3←−T4←−T

5←−T←− . . . ,

where the arrow marked n maps t ∈ T to tn if we think of T as amultiplicative group. The projective limit can be described explicitly:in the product group TN, the projective limit is the subgroup

Q∗d = { (tn)n∈N | tn ∈ T, tnn = tn−1 } ⊆ TN .

There are other solenoids whose construction differs from the one aboveby the sequence of numbers assigned to the arrows. Some of the othersolenoids have elements of finite order (see Exercise 6 of Section 52), butQ∗

d does not, because Q is divisible; compare 63.31.All solenoids have in common that they are important examples of

compact non-Lie groups. They contain compact, totally disconnectedsubgroups with quotient groups isomorphic to T. Locally, a solenoidis homeomorphic to the product of an arc with the Cantor set; seeHewitt–Ross 1963 10.15 or Hofmann–Morris 1998 8.23 (equivalenceof (3) and (8)).

The following fact was observed by Halmos 1944.

Page 117: The Classical Fields

100 Real numbers

8.33 Proposition As a group, the solenoid Q∗d is isomorphic to R+.

Proof Since Q is uniquely divisible (in other words, divisible and torsionfree), the character group Q∗

d has the same properties; see 63.31 and63.32. Thus, Q∗

d is a rational vector space, see 1.9, and it remains todetermine the cardinality of a basis or of Q∗

d itself (1.14).There are at most ℵ characters χ : Qd → T, as χ is determined by the

sequence of values χ(1/n) ∈ T, n ∈ N. Since card T = card R = 2ℵ0 (see1.10), the set of these sequences has cardinality (2ℵ0)ℵ0 = 2ℵ0 = card R.

Conversely, the standard character Qd → Q/Z ≤ R/Z = T may befollowed by any automorphism of Q/Z, and it is a consequence of 1.28together with 1.27 that there are ℵ such automorphisms. �

Exercises(1) It is known that for any two sets A, B ⊆ R of positive Lebesgue measure,the sum A+B contains a non-trivial interval; compare 10.7. Prove that thereexists even a meagre set S of Lebesgue measure zero such that S + S = R.

(2) With respect to Sorgenfrey’s topology on R (5.73), addition is continuous,but not the map x �→ −x.

(3) Fill in the details for examples 8.16(3) and (4). In particular, show that Ais not locally compact and that the direct product of (any number of) copiesof Q+ is a topological group.

(4) Endow the ring of continuous endomorphisms Endc R with the compact-open topology (see 63.2). Show that Endc R ∼= R as a topological field.

9 Multiplication and topology of the real numbers

First we verify that the multiplicative group R× = R�{0} is a topologi-cal group. In fact, we show slightly more, so that the result will combinewith 8.2 to prove that R is a topological field.

9.1 Theorem Multiplication (x, y) �→ xy is continuous on R×R, and

inversion R× → R× is continuous. In particular, R× is a topological

group.

Proof Consider (x + g)(y + h) = xy + gy + hx + gh. Given ε ∈ ]0, 1], setm = max{|x|, |y|, 1} and δ = ε/(3m). If |g|, |h| < δ, then |gy+hx+gh| <mδ+mδ+δ2 < ε. This yields continuity of multiplication. For x �= 0 andε > 0, choose δ = min{ε|x|2, |x|}/2. If |x− y| < δ, then |y| ≥ |x|/2 and|y − x| < ε|xy|. Therefore, |1/x− 1/y| < ε, and inversion is continuous.

The same proof works for the additive group of any (skew) field withabsolute value (as defined in 55.1). �

Page 118: The Classical Fields

9 Multiplication and topology of the real numbers 101

9.2 Proposition The multiplicative group Rpos of positive real num-

bers is isomorphic to R+ as a topological group, and R× is isomorphic

to the direct product R+ × C2 as a topological group.

Proof In 2.2 we described a group isomorphism ϕ : R× → R+ ⊕ C2 =R+×C2 using the exponential function. Explicitly, ϕ is given by ϕ(r) =(ln |r|, r|r|−1). This map is a homeomorphism. �

Our goal in this section is to demonstrate that the topology of R×

does not determine the multiplication; in fact, there is precisely one (non-commutative) group structure other than the usual one which makes R×

a topological group. Instead of constructing this group in an abstractway, we prefer to show that it arises naturally as the isometry group ofthe metric space of real numbers.

9.3 Definition: The affine group of the real line We consider themappings R→ R defined by

τa,b : x �→ a + bx, a, b ∈ R, b �= 0 .

They are bijective and form a group Aff R, the affine group of the realline. Multiplication in this group is composition of maps, and is givenby

τa,bτc,d = τa+bc, bd .

The map τ0,1 is the identity, and inversion is given by τ−1a,b = τ−a/b,1/b.

Using the parameters a, b, we may consider Aff R as a subset of R2, andthe topology induced by R2 makes Aff R a topological group. We mightadd that the map Aff R × R → R given by (τ, r) �→ τ(r) is continuous,so Aff R is a topological transformation group acting continuously on R.

Sending the map τa,b to the matrix(

1 0a b

), we may represent Aff R

as a group of matrices, and the action on R corresponds to(1 0a b

)(1x

)=(

1a + bx

).

From a purely group theoretic point of view, Aff R is best describedas a semidirect product.

9.4 Semidirect products Suppose that the group (G, ·) contains anormal subgroup A and a (not necessarily normal) subgroup B such thatA∩B = {1} and A ·B = G. For b ∈ B, let ψb be the automorphism of A

induced by the inner automorphism ϑb : g �→ bgb−1 of G. Then b �→ ψb

Page 119: The Classical Fields

102 Real numbers

is a homomorphism B → AutA, and the map ab �→ (a, b) ∈ A×B is anisomorphism of G onto the semidirect product A �ψ B, defined by

A �ψ B := A×B as a set, and (a, b) · (c, d) := (a · ψb(c), b · d) .

To verify the isomorphism, it suffices to note that (ab)(cd) = a(bcb−1)bd.Conversely, given two groups A, B and a homomorphism ψ : B →

AutA, the above formula defines a semidirect product group. Thesemidirect product reduces to a direct product if, and only if, ψb = id forall b. Another equivalent condition is that also B is a normal subgroupof G.

Comparison with 9.3 shows that Aff R is a semidirect product R � S

of the normal subgroup R ∼= R+ of translations, given by b = 1, withthe subgroup S ∼= R× defined by a = 0,

Aff R ∼= R � S ∼= R+ � R× .

Note that the group operation in R appears as addition in 9.3, and thatψb(c) = bc. Hence, the multiplication in the semidirect product hasto be written as (a, b)(c, d) = (a + bc, bd). The action of S on R byconjugation (which determines the structure of the semidirect product)is precisely the action of Aut R on R; compare 8.24.

9.5 Definition: The motion group M of the real line We singleout a subgroup of Aff R by restricting b in 9.3 to {1,−1}. We call thisgroup M the motion group of R, that is, M = {x �→ a ± x | a ∈ R}.This is a semidirect product,

M ∼= R+ �ψ C2,

where C2 is identified with the subgroup {1,−1} ≤ S and −1 acts onR ∼= R+ by inversion, that is, ψ−1(r) = −r. Other examples of abeliangroups extended by inversion are the finite dihedral groups Dn

∼= Cn�C2

(the symmetry groups of the regular n-gons, n ≥ 3) and the orthogonalgroup of the plane, O2R ∼= (SO2R)�C2. The proof is left as an exercise.

We note that every element of M is either a translation τa,1 or areflection τa,−1. More precisely, the involution τa,−1 reflects the realline R about the point a/2. We leave it as an exercise to prove that Mconsists precisely of the isometries of the standard metric on R. This iswhat the name motion group expresses.

Obviously, M is homeomorphic to R×, and M is not abelian, so itsgroup structure differs from that of R×. It is easier to compare the twogroups if we write R× ∼= Rpos × C2 and M ∼= Rpos � C2. Then the

Page 120: The Classical Fields

9 Multiplication and topology of the real numbers 103

multiplication in the semidirect product becomes

(r, e)(s, f) = (rse, ef) .

Our next result is that these two topological groups are the only onesthat can live on the topological space R � {0}.9.6 Theorem Every topological group (G, ·) homeomorphic to R× is

isomorphic as a topological group to R× ∼= R+ × C2 or to the motion

group M ∼= R+ � C2.

Proof (1) Let R ≈ R be the connected component of G containing theneutral element 1. Then R is a normal subgroup of G (see 62.13), andR ∼= R+ ∼= Rpos as a topological group by 8.10 and 9.2. We may assumethat in fact (R, ·) = (Rpos, ·). The only other connected component ofG must be a coset Rs.

(2) The factor group G/R has order 2, hence s2 ∈ R. Consider theinner automorphism ϑs : g �→ sgs−1 induced by s. The square ϑ2

s isthe inner automorphism induced by s2 ∈ R; it induces the identity onthe abelian group R. Using 8.24, we infer that there are precisely twopossibilities for the automorphism of R induced by ϑs: it is either theidentity or inversion.

(3) If ϑs induces inversion, then the equation s2 = ϑs(s2) = s−2

shows that s2 = 1 and 〈s〉 ∼= C2. Now 9.4 shows that we obtain anisomorphism G→ Rpos � C2 of topological groups by sending r ∈ R to(r, 1) and rs ∈ Rs to (r,−1).

(4) Suppose that ϑs induces the identity on R. Choose an elementr ∈ R such that r2 = s2. We set s′ = sr−1; then (s′)2 = sr−1sr−1 =ϑs(r−1)s2r−1 = 1, and ϑs′ = ϑsϑr−1 induces the identity on R. Thenthe same map as in step 3 is an isomorphism with respect to the directproduct multiplication on Rpos × C2. �

Exercises(1) Show that all automorphisms of the topological group R× have the formx �→ xa, where a ∈ R×, and deduce that Autc R× ∼= R×. (How should xa bedefined for x < 0 ?)

(2) Prove that the motion group M introduced in 9.5 consists precisely of theisometries, i.e., of the maps preserving the standard metric d(x, y) = |x − y|on R.

(3) Verify the semidirect product decompositions of dihedral groups and ofthe orthogonal group stated in 9.5.

Page 121: The Classical Fields

104 Real numbers

10 The real numbers as a measure space

A measure μ on a (metric) space M associates with some subsets of M

non-negative real numbers, measuring their ‘sizes’, in such a way thatμ is additive on finite or countable families of pairwise disjoint sets.If M carries a group structure, μ is usually required to be translationinvariant. Generally, it is not possible to extend μ consistently to theclass of all subsets of M ; see 10.9–11. The members of the largestdomain M ⊆ 2M to which μ can be extended are called measurable.In the real case, the pair (M, μ) is unique up to a scalar factor. Thevery complicated nature of M is illustrated by results 10.15ff. For anintroduction to measure theory see Halmos 1950 or Cohn 1993.

10.1 Definition A σ-field (or σ-algebra) on a space M is a subsetS ⊆ 2M such that:(a) ∅ ∈ S and (S ∈ S⇒M � S ∈ S)(b) If Sν ∈ S for ν ∈ N, then

⋃ν∈N Sν ∈ S (and

⋂ν∈N Sν ∈ S by (a)).

The σ-field 〈D〉σ generated by a set D ⊆ 2M is the intersection of allσ-fields in 2M containing D. (Note that 2M itself is a σ-field.)

In a metric space M , the open sets and the closed sets generate thesame σ-field B; it is called the Borel field of M , and each element of B

is called a Borel set .In this section, M will usually be the space R with its ordinary metric.

Since each open set in R is a union of countably many intervals, the Borelfield of R is generated by the open or the closed intervals or the compactsets as well.

10.2 Definition A measure space (M,S, μ) consists of a metric spaceM , a σ-field S containing the Borel field of M , and a σ-additive functionμ : S→ [ 0,∞ ] satisfying μ(∅) = 0 and μ(S) <∞ for each bounded setS ∈ S. Precisely, μ is called σ-additive if

μ(⋃

ν∈N Sν

)=∑

ν∈N μ(Sν) (∗)for any union of pairwise disjoint sets Sν ∈ S, ν ∈ N.

The measure μ is said to be regular if

μ(S) = inf{μ(O) | S ⊆ O ∧ O is open}for each S ∈ S, and μ is complete if T ⊂ S ∈ S and μ(S) = 0 implyT ∈ S.

If (R,S, μ) is a measure space, then μ is called (translation) invariant,if μ(S + t) = μ(S) for all S ∈ S and t ∈ R.

Page 122: The Classical Fields

10 The real numbers as a measure space 105

10.3 Lebesgue measure There is a unique regular, complete, andtranslation invariant measure λ on R such that λ([0, 1]) = 1. Thismeasure is known as Lebesgue measure, and the sets in the correspondingσ-field M are the measurable sets.

We sketch a proof for this assertion and characterize the sets in M.Since λ is assumed to be additive, regular, and invariant, λ([ a, b ]) =b − a for each interval of rational length and then for all intervals (byregularity). Each open subset O of R is the disjoint union of its connectedcomponents, and these are (at most countably many) open intervals;compare 5.6. Thus, λ(O) is necessarily the sum of the lengths of theconnected components of O. Since each compact set C is a difference ofa closed interval and an open set, also λ(C) is well-defined. By

λ(X) = inf{

λ(O)∣∣ X ⊆ O ∧ O is open in R

}we can define an outer measure λ : 2R → [ 0,∞ ], and λ is obviouslytranslation invariant. Moreover, λ is easily seen to be σ-subadditive:λ(⋃

ν∈N Xν

) ≤ ∑ν∈N λ(Xν). The inner measure λ is defined dually,approximating X from within by closed or compact sets. It satisfiesλ(⋃

ν Xν

) ≥ ∑ν λ(Xν) for any countable family of pairwise disjointsets Xν . By the very definition, λ(X) ≤ λ(X) for all X ⊆M .

The only way to enforce regularity consists in choosing the right σ-algebra. We put

X ∈M � 0 = inf{

λ(O � X)∣∣ X ⊆ O ∧ O is open in R

}and λ(X) = λ(X) for X ∈M. By construction, λ is translation invari-ant.

10.4 Theorem (R,M, λ) is a regular complete measure space.

A proof consists of several steps, most of which are easy consequencesof the definitions:(a) If Xν ∈ M for ν ∈ N, then X =

⋃ν Xν ∈ M. In fact, for each

ε > 0 there are open sets Oν and Uν such that Xν ⊆ Oν ⊆ Xν ∪ Uν

and λ(Uν) < ε2−ν . Put U =⋃

ν Uν . Then X ⊆ ⋃Oν ⊆ X ∪ U , andλ(U) < 2ε is arbitrarily small.(b) If X ∈M and O is open in R, then X ∩O ∈M.(c) If X ∈M and λ(X) < ∞, then infC λ(X � C) = 0, where C variesover all compact subsets of X. In particular, X ∈M implies R�X ∈M

for each bounded set X. Together with (a) and (b) this shows that M

is a σ-field.

Page 123: The Classical Fields

106 Real numbers

(c′) Step (c) can be rephrased as follows: if λ(X) <∞, then X ∈M if,and only if, λ(X) = λ(X).(d) If X =

⋃ν Xν is a countable union of pairwise disjoint measurable

sets Xν , then λ(X) < ∞ if, and only if,∑

ν λ(Xν) < ∞. In this caseλ(X) ≤ ∑ν λ(Xν) ≤ λ(X), and (c′) shows that X ∈ M and λ(X) =∑

ν λ(Xν). Thus, λ is σ-additive on M.(e) By definition, λ is regular. If λ(X) = 0 and Y ⊆ X, then Y ∈ M.Therefore, λ is also complete. �

10.5 Uniqueness If (R,S, μ) is a measure space, and if μ satisfies theconditions imposed upon λ in 10.3, then S = M and λ = μ.

Indeed, λ and μ agree on intervals and hence on the Borel field B.Since μ is regular, μ(S) = λ(S) for each S ∈ S. Moreover, S is containedin a Borel set B ∈ B (in fact, an intersection of countably many opensets) such that μ(S) = μ(B) and λ(B � S) = 0. This implies S ∈ M.We will show that S = M. Any set N with λ(N) = 0 is contained in aBorel set of measure 0. Because μ is complete, it follows that N ∈ S.An arbitrary set X ∈M can be written as X = B ∪N with B ∈ B andλ(N) = 0. Hence X ∈ S. �

10.6 Examples (a) The countable field Ralg of real algebraic numbershas measure λ(Ralg) = 0. Argument (a) of the proof of 10.4 shows thatRalg has open neighbourhoods O of arbitrarily small measure λ(O), andthe complement R � O consists entirely of transcendental numbers.

(b) The Cantor set C = {∑∞ν=1 cν3−ν | cν ∈ {0, 2}} has the measure

λ(C) = 0. In fact, C is obtained from the interval [0, 1] by repeatedlyremoving ‘middle thirds’; see 5.35. Thus, λ(C) ≤ (2/3)n for any n ∈ N.Since card C = ℵ = 2ℵ0 , it follows from completeness that card M = 2ℵ.Note that C + C = [0, 2] has positive measure (Nymann 1993).

(c) A modified construction leads to a set which is homeomorphicto C but has positive measure: one starts again with [0, 1], instead ofmiddle thirds, however, in the ν-th step one removes an open intervalof length ν−2λ(J) from the middle of each remaining closed interval J .This yields a compact set C2 of measure

∏∞ν=2(1−ν−2) = 1/2, the value

of the infinite product being obtained as follows:∏nν=2 (ν2 − 1)/ν2 = (n− 1)! (n + 1)! /2(n!)2 = (n + 1)/2n .

10.7 Differences If λ(S) > 0, then S−S = {s− t | s, t ∈ S } contains

some open neighbourhood of 0.

Page 124: The Classical Fields

10 The real numbers as a measure space 107

Proof Since S contains a compact subset of positive measure, we mayassume that S itself is compact. There is an open neighbourhood O ofS such that λ(S) ≥ 5

6 · λ(O). The set O is a union of disjoint intervals.Therefore, O contains an interval J with λ(S ∩ J) ≥ 4

5 · λ(J). PutT = S∩J and λ(J) = 5ε, and assume that |x| < ε and x /∈ T −T . Thenx+T and T are both contained in an ε-neighbourhood of J , hence in aninterval K of length 7ε, and (x+T )∩T = ∅, but λ(x+T ) = λ(T ) ≥ 4ε.This contradicts additivity and shows that ]−ε, ε[ ⊆ S − S. �

The fact that R is a vector space over Q yields many examples of non-measurable sets. For any Hamel basis B (see 1.13), the lattice L = LB

consists of all numbers∑

b∈B nb · b with nb ∈ Z and nb = 0 for all butfinitely many b ∈ B. We will also consider an arbitrary hyperplane H

of R over Q.

10.8 Inner measure Each lattice L and each hyperplane H has inner

measure 0. More generally, λ(G) = 0 for any proper additive subgroup

G < R.

Proof Since G is a group, G−G = G. If λ(G) > 0, then G−G containsan open neighbourhood of 0 by 10.7, but such a neighbourhood generatesthe full group R. �

10.9 Proposition A hyperplane H intersects an arbitrary open interval

J in a set of positive outer measure. In particular, H ∩ J /∈M.

Proof Assume that λ(H ∩ J) = 0. Then λ(H ∩ J) = 0 by 10.8. Thereare countably many elements hν ∈ H such that

⋃ν(J + hν) = R. Note

that (H∩J)+hν = H∩(J +hν). Thus, λ(H∩J) = 0 implies λ(H) = 0.Because R/H ∼= Q, the set R is a union of countably many cosets of H,and it would follow that λ(R) = 0. �

10.10 Proposition Any lattice L = LB in the rational vector space Rhas outer measure λ(L) =∞. Therefore, L /∈M.

Proof The elements 12b with b ∈ B represent uncountably many distinct

cosets of L in R. Hence one cannot argue as in the previous proof.The definition of λ shows that λ(r · S) = r · λ(S) for any positivenumber r. Obviously, R =

⋃n∈N (n!)−1L, and hence ∞ ≤ λ(R) ≤∑

n (n!)−1·λ(L) = e · λ(L). �

The following analysis shows that a lattice L is non-measurable in avery strong sense.

Page 125: The Classical Fields

108 Real numbers

10.11 Theorem If L is a lattice in R and U is open, then λ(L∩U) =λ(U), while λ(L ∩ U) = 0.

Proof By 10.8, it suffices to prove the first claim. Two distinct elementsof a basis of R over Q generate a dense subgroup of R; see 1.6b. Inparticular, L = R.(a) We show that for each ϑ ∈ ]0, 1[ and for each ε > 0, there is someopen interval J of length λ(J) < ε such that λ(L ∩ J) > ϑ · λ(J).

In fact, λ(L) =∞ and subadditivity of λ imply that there is a compactinterval C with λ(L ∩ C) := m > 0. By definition of λ, we can findan open neighbourhood O of L ∩ C satisfying λ(O) < m + r with anarbitrarily small number r > 0. Let 2r < (1 − ϑ) ·m. The open set O

can be covered by intervals Jν of length less that ε such that λ(O) ≤∑ν λ(Jν) < m + 2r.Suppose that λ(L∩Jν) ≤ ϑ ·λ(Jν) for all ν; then m ≤∑ν λ(L∩Jν) ≤

ϑ · (m + 2r) < m, a contradiction.(b) We may assume that U is an open interval. Let J be chosen accordingto step (a). Because L is dense in R, there are finitely many elementst1, . . . , tk ∈ L such that the intervals J + tκ ⊆ U are pairwise disjointand that λ(U)− k · λ(J) < ε. It follows that

λ(L ∩ U) ≥ λ(⋃k

κ=1 L ∩ (J + tκ))

=∑k

κ=1 λ((L ∩ J) + tκ

)= k · λ(L ∩ J) > k · ϑ · λ(J) ,

hence λ(L ∩ U) > ϑ · (λ(U)− ε) is arbitrarily close to λ(U). �

Remark With the same arguments, an analogous result can be provedfor hyperplanes.

Intuitively, both meagre sets (i.e., countable unions of nowhere densesets) and sets of Lebesgue measure zero (null sets) are small. The twonotions are quite independent of each other. This is illustrated by thefollowing astonishing result:

10.12 Proposition R is a union of a meagre set M and a null set N .

Proof Let Q = {rν | ν ∈ N} and put Jκν = {x ∈ R | |x− rν | < 2−κ−ν }for κ ≥ 1. Then Oκ =

⋃∞ν=1 Jκν is an open neighbourhood of the dense

set Q, and R � Oκ is nowhere dense. The set N =⋂

κ Oκ is a null set(since λ(Oκ) ≤ 21−κ) and R � N is meagre. �

We mention a kind of duality between meagre sets and null sets (fora proof see Oxtoby 1971 §19):

Page 126: The Classical Fields

10 The real numbers as a measure space 109

10.13 Theorem (Erdos) Under the continuum hypothesis 2ℵ0 = ℵ1

there exists an involutory mapping ϕ : R → R such that λ(X) = 0 if,

and only if, ϕ(X) is meagre.

Traditionally, any union of countably many closed sets is called anFσ-set.

10.14 Measurable sets A set S ⊆ R is measurable if, and only if,

there is an Fσ-set F and a null set N with S = F ∪N .

Proof If S is measurable, so is X = R � S. By definition of M, the setX is contained in an intersection D of countably many open sets suchthat λ(D � X) = 0. Hence F = R � D is an Fσ-set and S � F = D � X

has measure 0. �

In order to illustrate the high degree of complexity of the σ-field M

of the measurable sets, we introduce the notions of Borel classes and ofSouslin sets. Starting with the compact or the open sets, all Borel setscan be obtained by alternatingly forming countable unions and countableintersections. For X ⊆ 2R, let Xσ be the set of all unions and Xδ the setof all intersections of countable subfamilies of X.

We shall now use ordinal numbers; refer to Section 61 for notions andnotation.

10.15 Borel hierarchy Denote by O = O0 the set of all open subsetsof R, and put Oν =

⋃μ∈ν(Oμ)δσ.

By definition, Oν ⊆ B for each ordinal ν. Because any countablesubset of the least uncountable ordinal number Ω has an upper boundin Ω (see 61.5), we have (OΩ)δσ = OΩ. Since closed sets are contained inOδ, it follows inductively that the complement of any set in Oμ belongsto Oμ+1. Hence B = OΩ.

10.16 Borel sets We have cardB = ℵ = 2ℵ0 < cardM.

Proof By 61.10, we have cardXσ, cardXδ ≤ (cardX)ℵ0 . Since R hasa countable basis, there are only ℵ0

ℵ0 = ℵ open sets. By induction itfollows that cardOν = ℵ for ν ∈ Ω. Now 2ℵα > ℵα by 61.11, hencecard Ω = ℵ1 ≤ ℵ, and with 61.13(a) we obtain cardOΩ = ℵ.

Because each subset of the Cantor set C (see 10.6(b)) is a null set,cardM = 2ℵ > ℵ. �

10.17 Theorem For ν ≤ Ω, the Borel classes Oν are strictly increasing.

A proof can be found in Hausdorff 1935 §33 I; compare also Miller

1979.

Page 127: The Classical Fields

110 Real numbers

10.18 The Souslin operation Let F be the set of all finite sequencesof elements of N. For any sequence ν ∈ NN of natural numbers, putν|κ = (νι)ι≤κ ∈ F. Consider an arbitrary family T ⊆ 2R. Any map-ping F → T, i.e., any choice of sets Tν|κ ∈ T determines a Souslin set⋃

ν∈NN

⋂κ∈N Tν|κ over T; note that the union is to be taken over un-

countably many sequences ν.The family of all Souslin sets over T is denoted by TS. The choice

Tν|κ = Tκ ∈ T shows that Tδ ⊆ TS, similarly Tν|κ = Tν1 ∈ T givesTσ ⊆ TS . An argument related to general distributivity shows that(TS)S = TS ; see Hausdorff 1935 §19 I, Jacobs 1978 Chapter XIIIProposition 1.3, or Rogers–Jayne 1980 Theorem 2.3.1.

Let S = OS denote the collection of the Souslin sets over the opensubsets of R, and write C for the class of all compact sets.

10.19 Proposition The family S contains all Borel subsets of R.

Moreover, CS = S.

Proof By 10.15, the Borel field B is the smallest family of subsets of Rsuch that O ⊆ B and Bδ = Bσ = B, and 10.18 shows that Sδ,Sσ ⊆SS = S. Hence B ⊆ S. Moreover, C ⊆ Oδ and O ⊆ Cσ imply CS ⊆ OS

and S ⊆ (Cσ)S ⊆ (CS)S = CS. �

10.20 Analytic sets By definition, an analytic set is the image of

some continuous map ϕ : NN → R. Each analytic set is contained in S.

Proof It is tacitly understood that NN carries the product topology. Atypical neighbourhood Uκ(ν) of ν ∈ NN consists of all sequences μ ∈NN such that μ|κ = ν|κ. For a fixed ν, continuity of ϕ implies thatϕUκ(ν) is contained in some ε-neighbourhood Vε(ϕ(ν)), where the ε = εκ

converge to 0 as κ increases. Choose Oν|κ as an εκ-neighbourhood ofϕUκ(ν). Then Oν|κ ⊆ V2ε(ϕ(ν)) and

⋂κ∈N Oν|κ = {ϕ(ν)}. The proof is

completed by taking the union over all ν ∈ NN. �

10.21 Theorem Each Borel set and even each Souslin set is analytic.

Proof By 4.10, the space I of irrational real numbers may be identifiedwith NN.(a) Iν = I∩[ν, ν+1] is homeomorphic to I. If ϕν : Iν → R is continuous,then there is a continuous map ϕ : I→ R with ϕ|[ν,ν+1] = ϕν . Obviouslyϕ(I) =

⋃ν ϕν(Iν). Hence a union of countably many analytic sets is

analytic.(b) The set I = NN contains a copy of the Cantor set C = {0, 1}N and

Page 128: The Classical Fields

10 The real numbers as a measure space 111((cν)ν �→

∑ν cν2−ν

): C → [0, 1] is a continuous surjection. It extends

to a continuous map I → [0, 1] by normality of I. Hence each closedinterval is analytic, and (a) implies that open sets are analytic.(c) Let ϕν : I→ R be continuous, and assume that

⋂ν∈N ϕν(I) = E �= ∅.

The set

D ={

x = (xν)ν ∈ IN∣∣ ∀ν ϕν(xν) = ϕ1(x1)

}is closed in IN, and (x �→ ϕ1(x1)) : D → E is a continuous surjection.Now D is a retract of IN, because IN = NN×N ≈ NN = I and each closedsubset of NN is a retract (see Exercise 7 or Engelking 1969). Hence,(x �→ ϕ1(x1)) extends to IN, and the intersection E of countably manyanalytic sets is analytic. 10.15 proves the first claim.(d) By the last part of 10.19, Souslin sets can be obtained from compactsets instead of open sets, hence each Souslin set is of the form A =⋃

ν∈NN

⋂κ∈N Cν|κ with compact sets Cν|κ. We may assume, moreover,

that κ < λ implies Cν|κ ⊇ Cν|λ. Write Cν|κ =⋃

i∈N C iν|κ, where each

C iν|κ is contained in some interval of length κ−1. Infinite distributivity

gives⋂

κ∈N Cν|κ =⋃

ι∈NN

⋂κ∈N C ικ

ν|κ. Therefore the Cν|κ may be chosenof diameter at most κ−1. We define a set D by ν ∈ D �

⋂κ∈N Cν|κ =

{xν} �= ∅. The map ϕ : D → A : ν �→ xν is a continuous surjection.Moreover, D is closed in NN, since ν /∈ D implies Cν|κ = ∅ for someκ ∈ N, and then μ /∈ D for all μ with μ|κ = ν|κ. Again, D is a retractof NN and ϕ extends to a continuous map NN → A. �

10.22 Proposition The family of all Souslin sets in R has cardinality

cardS = ℵ.Proof Since R has a countable basis for the topology, card O = ℵ = 2ℵ0 .The set F of all finite sequences ν|κ is countable by 61.14. Each Souslinset is determined by a choice of open sets Oν|κ, that is, by a map F→ O.There are ℵℵ0 = ℵ such maps (61.13b). �

10.23 Theorem There is a Souslin set X in R such that R � X /∈ S.

In particular, X is not a Borel set.

Proof By 61.14, the set F of all finite sequences ν|κ is countable, hencethere is a bijection F→ N denoted by (ν|κ �→ ν(κ)). The map (ν �→ ν)is injective and defines a bijection of NN onto its image N ⊆ NN. EachSouslin set is now of the form

⋃n∈N⋂

κ∈N On(κ). Choose a countablebasis B = (Bμ)μ∈N for the open sets and note that Bσ = O. For ξ ∈ NN

put S(ξ) =⋃

n∈N⋂

κ∈N Bξn(κ) . Then {S(ξ) | ξ ∈ NN } = BS = S.

Page 129: The Classical Fields

112 Real numbers

We identify NN with the space I of all irrational numbers in R asin 4.10. Let X ⊆ R be defined by ξ ∈ X � ξ ∈ S(ξ). Assume thatR � X = S(ξ) for some ξ ∈ NN. Then ξ ∈ R � X ⇔ ξ ∈ S(ξ)⇔ ξ ∈ X,a contradiction. Therefore, R � X is not a Souslin set.

In order to show that X ∈ S, let Tκ = { ξ ∈ I | ξ ∈ Bξκ }, and notethat ξ ∈ Bξκ

if, and only if, there is some μ ∈ N such that ξ ∈ Bμ andμ = ξκ. Now { ξ | ξκ = μ} is open in NN, hence Tκ is open in I, andtherefore Tκ ∈ Oδ. Thus X =

⋃n∈N⋂

κ∈N Tn(κ) ∈ OS. �

10.24 Souslin sets are measurable The set M of all Lebesgue mea-

surable sets satisfies MS = M. Hence S ⊆ M, even S ⊂ M, since

cardS = ℵ < cardM.

Proofs of this standard result can be found in Saks 1937 Chapter IITheorem 5.5, or Rogers–Jayne 1980 Corollary 2.9.3. They are some-what technical and will be omitted here.

Finally we mention the following result by Laczkovich 1998:

10.25 Every analytic proper subgroup of the reals can be covered by

an Fσ null set.

Exercises(1) In the notation of 10.15, show that O1 = O.

(2) Represent Q as a Souslin set.

(3) Any sublattice S in the intersection of finitely many hyperplanes in therational vector space R has infinite outer measure.

(4) There are many σ-fields between S and M.

(5) The modified Cantor set C2 (defined in 10.6c) has non-measurable subsets.

(6) A map ϕ : R → R is said to be measurable if, and only if, each open set Ohas an inverse image ϕ−1(O) ∈ M. If ϕ is measurable, then the inverse imageof each Borel set is measurable, but not necessarily the inverse image of eachset X ∈ M.

(7) If A is closed in NN, then there is a continuous map ρ : NN → A withρ|A = idA. (Such a map is called a retraction.)

11 The real numbers as an ordered field

In this section we show that the field R can be made into an ordered fieldin a unique way. Conversely, ordering and addition of the real numbersdetermine the multiplication up to isomorphism. After a brief surveyon exponential functions in ordered fields we characterize those orderedfields which are isomorphic to subfields of R.

Page 130: The Classical Fields

11 The real numbers as an ordered field 113

11.1 Definition An ordered skew field is a skew field F equippedwith the additional structure of a chain (F,<) such that the followingmonotonicity laws hold. For all a, b, c ∈ F such that a < b, we stipulatethat

a + c < b + c

ac < bc if 0 < c .

From these axioms it follows that bc < ac if c < 0; indeed, from c < 0 weobtain 0 < −c by the first monotonicity law, and the second one yieldsa(−c) < b(−c). This implies 0 < ac− bc, and the claim follows.

The first law says that the additive group of F becomes an orderedgroup in the sense of 7.1. Note that (F,+) is commutative by 6.1,hence we do not need to stipulate the monotonicity of left addition.There is a similar lack of symmetry in the second law although at thismoment the multiplication in F is not assumed to be commutative. Weshall show in 11.5 that the missing law comes for free. If F happensto be commutative, then the second law alone implies that the positiveelements a > 0 form an ordered group under multiplication.

In any ordered skew field F , we define the absolute value of a ∈ F by

|a| = max{a,−a} .

This absolute value satisfies the triangle inequality

|a + b| ≤ |a|+ |b| ,which is proved by adding the inequalities εa ≤ |a| and εb ≤ |b|, whereε ∈ {1,−1}, to obtain ε(a + b) ≤ |a|+ |b|.11.2 Remark The monotonicity law for multiplication implies thatevery square a2, 0 �= a ∈ F , is positive: a2 > 0. In particular, it is notpossible to make the complex numbers into an ordered field, becauseboth −1 = i2 and 1 = 12 would be positive, resulting in the contradiction0 = −1 + 1 > 0. Moreover, in every ordered skew field, we have 1 =12 > 0 and 1 + · · ·+ 1 > 0, hence the characteristic of an ordered skewfield is always zero.

An ordering on a field F is introduced most conveniently by specifyingthe set of positive elements. The properties that this set has to satisfyare collected in the following definition.

11.3 Definition A domain of positivity in a skew field F is a subgroupP ≤ F× of index 2 in the multiplicative group which is closed under

Page 131: The Classical Fields

114 Real numbers

addition. In other words, P does not contain zero and forms a groupunder multiplication, and we have P + P ⊆ P and F× = P ∪ aP foreach a /∈ P ∪ {0}.11.4 Proposition (a) Every ordered skew field (F,<) defines a domain

of positivity by

P = {x ∈ F | 0 < x}.(b) Conversely, every domain of positivity P in a skew field F defines

an ordering which makes F an ordered skew field by the rule

a < b⇔ b− a ∈ P .

(c) The operations in (a) and (b) are inverses of each other.

Proof (a) The monotonicity laws imply that P ≤ F× is a subgroup whichis closed under addition. Moreover, the product of any two negativeelements is positive, hence the index of this subgroup is at most 2.Finally, −1 is negative, so the index is exactly 2.

(b) Being a group under multiplication, P contains 1 but not 0, henceit does not contain −1 since P is closed under addition. It follows thatF× = P∪−P (disjoint union), and this implies that for distinct elementsa, b ∈ F , precisely one of the relations a < b, b < a holds. Moreover, therelation < is transitive: a < b < c implies that c−a = (c− b)+(b−a) ∈P + P . Thus, we have an ordering relation.

Monotonicity of addition holds because (b + c) − (a + c) = b − a. Ifa < b and 0 < c, then bc− ac = (b− a)c ∈ PP = P . If a < b and c < 0,then bc− ac ∈ P · (−P ) = −(PP ) = −P .

(c) This assertion is easily verified. �

11.5 Remark We can now explain the lack of symmetry in the mono-tonicity laws for multiplication: the domain of positivity is a subgroup ofF× of index two, hence a normal subgroup (the unique non-trivial rightcoset must coincide with the unique non-trivial left coset). This impliesthat cb− ca = c(b− a) ∈ cP = Pc if a < b, hence ca < cb⇔ ac < bc; inother words, monotonicity of left multiplication holds automatically.

We take it for granted that the field of real numbers is an orderedfield with respect to the usual ordering (see 42.11 or Section 23), and weproceed to show that this is the only way of making R into an orderedfield. We need a fact known from elementary analysis.

11.6 Lemma Every positive real number has a square root in R.

Page 132: The Classical Fields

11 The real numbers as an ordered field 115

Proof It suffices to find a root of a > 1 since c2 = a implies (c−1)2 = a−1.Then the non-empty set B = { b ∈ R | 1 ≤ b ∧ b2 < a} is bounded aboveby a, hence it has a supremum c. For every n ∈ N, there exists b ∈ B

such that c − n−1 < b, hence (c − n−1)2 < a ≤ (c + n−1)2. Passing tothe limit as n→∞, we obtain that c2 = a. �

11.7 Corollary There is only one way of ordering the field of real

numbers. The same is true of the rational number field.

Proof Every domain of positivity Q ⊆ R contains the set S of non-zerosquares, and the usual domain of positivity P is contained in S, as wehave just seen (11.6). Now Q and P have the same finite index in R×,hence they are equal.

The reasoning is different for the rational numbers. There, one seesthat 1 = 12 > 0, hence all n ∈ N are positive by monotonicity, and then1/n and, finally, m/n for m ∈ N are positive. �

Not only is the ordering of R uniquely determined by the field struc-ture, as we have just seen, but also the multiplication is determined (upto isomorphism) by the ordered additive group. This follows from thecharacterization of the ordered field R given in 42.9 and 42.11 (R is theonly completely ordered field). Here we give a direct proof.

11.8 Proposition If R is an ordered skew field with the usual addition

and ordering and some multiplication ∗, then this ordered skew field is

isomorphic to R with the usual structure.

Proof For a > 0, the distributive law together with the monotonicitylaw tells us that left multiplication x �→ a ∗ x is an automorphism ofthe ordered additive group. According to 7.11 or 7.12, this implies thatthere exists some (necessarily positive) b ∈ R such that a ∗ x = bx forevery x. Taking x = 1, we see that b = a ∗ 1, and we have

a ∗ x = (a ∗ 1)x

for all x. We proved this only for a > 0, but the distributive law gives(−a) ∗ x = −(a ∗ x), hence the above equation holds for all a. Takingx = e, the neutral element of the multiplication ∗, we obtain that a =a ∗ e = (a ∗ 1)e and a ∗ 1 = ae−1, the inverse being taken with respectto the usual multiplication. Finally, we get

a ∗ x = ae−1x .

It follows that e = (1 ∗ 1)−1 is positive. We have (ae) ∗ (xe) = (ax)e,

Page 133: The Classical Fields

116 Real numbers

and the map a �→ ae is an isomorphism of R with the usual structureonto the given ordered skew field. �

Property 11.7 of the real numbers is special: in general, two orderedfields can be non-isomorphic as ordered fields although they are isomor-phic as fields; they can even be isomorphic as chains at the same time.In other words, a given field structure and a given chain structure cansometimes be put together in essentially different ways. This is shownby the following examples.

11.9 Examples We construct three ordered fields such that no two ofthem are isomorphic as ordered fields, but their underlying fields andchains are separately all isomorphic. Two of the ordered fields will beArchimedean (this refers to the ordered additive group; compare 7.4),but not the third one.

The fields will all be isomorphic to the simple transcendental extensionQ(t) of the field of rational numbers; compare 64.19. Taking t to be atranscendental real number, we obtain a subfield Q(t) ⊆ R, and thissubfield inherits an Archimedean ordering from R. The field Q(t2) ⊆ Ris again a simple transcendental extension and inherits an Archimedeanordering, but a hypothetical order isomorphism ϕ : Q(t2)→ Q(t) has tofix 1 and, hence, all rational numbers (remember that the additive groupof R is uniquely divisible, 1.8). Since ϕ preserves the ordering and Q ⊆ Ris dense, ϕ would be the identity by 11.13 below, and Q(t) = Q(t2). Butt does not belong to Q(t2), because the substitution t �→ −t induces theidentity on Q(t2).

The third ordered field is obtained by taking the usual ordering onthe subfield Q ⊆ Q(t) and putting Q < t. The extension to an orderingon Q(t) is enforced: t < t2 < t3 < . . . implies that a polynomial p(t) ispositive if, and only if, the coefficient of the highest power of t is positive,and a quotient of two positive or two negative polynomials is positive,etc. We do get an ordered field, which is not Archimedean since N < t.See also Exercises 1 and 3 of Section 12.

It remains to explain why the three ordered fields have isomorphicunderlying chains. In fact, the chains are all countable (by 61.14) andstrongly dense in themselves, as shown by an easy verification. Theisomorphism follows from these properties by Theorem 3.4.

11.10 Exponential fields This is a variation of the theme consideredin Section 2, adapted to ordered fields. We have seen in 2.14 that thereare many ordered fields whose additive group is isomorphic to the mul-tiplicative group of positive elements. The real numbers have a much

Page 134: The Classical Fields

11 The real numbers as an ordered field 117

stronger property, however. Namely, the group R+ is isomorphic asan ordered group to the multiplicative group R×

pos of positive elementsthrough the exponential function. An ordered field F is called an expo-nential field if there is an isomorphism of ordered groups e : F+ → F×

pos,and any such isomorphism is called an exponential function. Sometimesadditional ‘growth conditions’ are imposed in order to make the functione more similar to the real exponential function; compare 56.16.

The Archimedean exponential fields (i.e., those whose additive or-dered group is Archimedean) are comparatively simple. First of all,every Archimedean ordered field is isomorphic to an ordered subfield ofthe real numbers; see 11.14 below. Then the homomorphism propertytogether with unique divisibility shows that e(x) = e(1)x for every ra-tional number x. Since Q ⊆ R is dense and e preserves the ordering,this equation holds for every x. It follows that the rational numbers,for example, are not an exponential field; compare Section 1, Exercise 6.Next, it is not hard to see that every subfield of R is contained in aunique minimal exponential subfield of R; compare Exercise 1.

The study of non-Archimedean exponential fields, initiated by Alling

1962, is much harder. The monograph Kuhlmann 2000 gives an impres-sive view of its present state, focusing on extensions of the real numberfield which admit an extension of the ordinary exponential. A naturalcandidate for such an exponential field appears to be a Hahn power se-ries field R((Γ)) with exponents taken from a divisible ordered abeliangroup Γ, as described in 64.25. Such fields can be used, like the Puiseuxfields in 2.10, to obtain examples with a multiplicative group isomorphicto the additive group of some other field. However, it was shown inKuhlmann et al. 1997 that such a field is never exponential unless Γ istrivial. (These fields do, however, admit a non-surjective logarithm; seeKuhlmann 2000.) On the other hand, Alling 1962 and Kuhlmann

2000 give many examples of non-Archimedean exponential fields; suchexamples can be obtained as countable unions of Laurent series fields.Other examples of non-Archimedean exponential fields include the non-standard real numbers (see 24.3) and a large class of examples given byAlling 1962, which we describe next.

11.11 Non-Archimedean exponential fields constructed fromrings of continuous functions Let X be a completely regular topo-logical space (for example, X = N or X = R), and consider the ringC(X) of continuous real-valued functions on X with pointwise additionand multiplication of functions. If we set f < g whenever f(x) < g(x)

Page 135: The Classical Fields

118 Real numbers

for all x ∈ X, then C(X) becomes a partially ordered ring. We want toform the quotient of C(X) modulo one of its maximal ideals M (com-pare 64.3). It can be shown that M is convex , that is, 0 ≤ f ≤ u ∈ M

implies f ∈ M , whence the ordering on C(X) carries over to the fieldF = C(X)/M and turns it into an ordered field; see Gillman–Jerison

1960 Theorem 5.5. The field F contains the real numbers as a subfield(namely, the residue classes of the constant functions). If F is Archime-dean, then it follows that it is isomorphic to R; compare our discussionof Archimedean exponential fields above. The question is how we canobtain non-Archimedean examples.

We should avoid using the ideals {f | f(p) = 0} where p ∈ X, becausethey are the kernels of the evaluation maps f �→ f(p) and their quotientfields are isomorphic to R. If X is not compact, then any maximalideal containing the ideal of all functions with compact support is not ofthis type. Now we assume moreover that C(X) contains an unboundedfunction u; this excludes pseudo-compact spaces like the long line Lintroduced in 5.25. It is plausible that then for some ideal M the residueclass u+M is an upper bound for all classes of constant functions, henceC(X)/M is non-Archimedean. A proof is given in Gillman–Jerison

1960 5.7.For the construction of an exponential we also need the fact that the

zero set f−1(0) of a function determines the maximal ideals M contain-ing f . More precisely, f belongs to M if, and only if, its zero set meetsthe zero set of every member of M ; see Gillman–Jerison 1960 2.6.

Using the ordinary exponential function exp, we obtain a mapping e :f �→ exp ◦f of C(X) into itself, which is a homomorphism (C(X), +)→(C(X), ·). For u ∈ M , also e(u) − 1 belongs to M because it has thesame zero set. It follows that e(f + u) − e(f) = e(f)(e(u) − 1) ∈ M ,and e induces a map on the quotient field F = C(X)/M , again denotedby e. On constant functions, e agrees with the ordinary exponential.The map e preserves the ordering on C(X); in particular, the functionsexp ◦f are all positive, hence we have constructed an order-preservinghomomorphism (F,+) → (Fpos, ·) extending the real exponential. Thisis in fact an isomorphism, because for f > 0 we have exp ◦ log ◦f = f .

We remark that according to Gillman–Jerison 1960 13.4, the ex-ponential fields constructed here are all real closed; compare Section 12.

Our next goal is a characterization of the ordered subfields of R; com-pare 11.14. We need the following notions in order to express the char-acteristic properties.

Page 136: The Classical Fields

11 The real numbers as an ordered field 119

11.12 Definition: Properties of pairs of ordered fields Let F ⊆ G

be a pair of ordered fields, that is, G is an ordered field and F is a subfield,endowed with the induced ordering.

We say that the pair is Archimedean or cofinal if every g ∈ G isdominated by some f ∈ F , that is, g ≤ f . (In particular, G is anArchimedean ordered field if the pair Q ⊆ G is Archimedean, where Qis the prime field of G; compare 11.7.)

We say that the pair is dense if F is dense in the chain G. Notethat an ordered field is strongly dense in itself in the sense of 3.1, henceit does not matter how we define density of chains, and it suffices torequire that for every pair a < b in G there is an element f ∈ F suchthat a ≤ f ≤ b.

The pair F ⊆ G is called order-rigid if every order automorphism ofG that fixes every f ∈ F is the identity.

Finally, we say that the pair is algebraic if G is an algebraic extensionof the field F ; compare 64.5.

The following result about the relationships among these propertiesis taken from Baer 1970a. We remind the reader that our fields arecommutative.

11.13 Proposition Let F ⊆ G be a pair of ordered fields. Then the

following implications hold among the properties defined in 11.12.

algebraic

Archimedean order-rigid

dense

���������

����

�����

There is no implication between ‘dense’ and ‘algebraic’ in either direc-

tion. Consequently, none of the four implications shown is reversible.

About the mutual relationship of the properties ‘Archimedean’ and‘order-rigid’ we know nothing.

Proof (1) We show that an algebraic pair is Archimedean. Let g ∈ G

be given; if g ≤ 1, we have nothing to show, hence we assume thatg > 1. The pair is algebraic, hence g satisfies an equation a0 + a1g +· · · + an−1g

n−1 + gn = 0, where ak ∈ F . After dividing by gn−1, weobtain g = −(an−1 + · · ·+a1g

2−n +a0g1−n), and the triangle inequality

yields |g| ≤ |an−1|+ · · ·+ |a0| ∈ F .

Page 137: The Classical Fields

120 Real numbers

(2) Suppose that the pair is algebraic; we show that it is order-rigid.Let α be an automorphism of G fixing each element of F . As before,every g ∈ G satisfies a polynomial equation p(g) = 0 with coefficients inF , and α has to permute the roots of p in G. There are at most n = deg p

roots, and α has to respect their ordering. This is only possible if α fixesall roots.

(3) If the pair is dense, then it is order-rigid. Indeed, if an automor-phism α of the ordered field G fixes all elements of F and moves g ∈ G,then there is some f ∈ F lying between g and α(g), and α fixes f ; thisis impossible.

(4) Again if the pair is dense, let g ∈ G be given; there is f ∈ F suchthat g ≤ f ≤ g + 1, and this shows that the pair is Archimedean.

(5) The pair Q ⊆ R is dense, hence also Archimedean and order-rigid,but not algebraic.

(6) We describe a pair which is algebraic but not dense, namely, thepair R((t2)) ⊆ R((t)) of Laurent series fields; the elements of R((t)) areLaurent series

∑k≥n rktk, where n ∈ Z and rk ∈ R; see 64.23. The

subfield R((t2)) consists of those series where rk = 0 whenever k is odd.This pair of fields is algebraic; indeed, the variable t (viewed as thepower series with r1 = 1 and rk = 0 in all other cases) satisfies thepolynomial equation x2 − t2 = 0 over R((t2)). Every element of R((t))can be split into odd and even powers, and thus can be written in theform

∑k≥n rkt2k + t ·∑k≥n skt2k. This shows that R((t)) is a quadratic

extension field of R((t2)).It remains to define an ordering on the field R((t)) in such a way that

it becomes an ordered field and that the subfield R((t2)) is not dense.This is done by specifying P = {∑k≥n rktk | n ∈ Z, rn > 0} as thedomain of positivity. For every positive r ∈ R, the power series t andr − t are positive, hence 0 < t < r. We have t < 2t, but no elementof R((t2)) lies between t and 2t (hence the pair of fields is not dense).Indeed, from t ≤ g =

∑k≥n rkt2k and rn �= 0 it follows that n ≤ 0, while

g ≤ 2t implies that n ≥ 1. �

11.14 Theorem Let F be an ordered field. Then the following asser-

tions are mutually equivalent.

(a) F is a subfield of R endowed with the induced ordering.

(b) The prime field Q ⊆ F is dense.

(c) For every subfield S of F , the pair Q ⊆ S is order-rigid.

(d) F is Archimedean, i.e., the pair Q ⊆ F is cofinal.

Page 138: The Classical Fields

11 The real numbers as an ordered field 121

Proof (1) We prove the implications (a)⇒ (b)⇒ (c)⇒ (d)⇒ (a). Thefirst implication is obvious, and the second one has been proved in 11.13.

(2) Assume that F contains an element t > Q. We claim that the pairof subfields Q ⊆ Q(t) of F is not order-rigid. We know from 11.13 thatQ(t) is a transcendental extension of Q. The map sending p(t)/q(t) top(t+1)/q(t+1) is an automorphism of Q(t) which fixes every element ofQ; compare 64.19. Moreover, the map preserves the ordering. Indeed,we have 1 < t < t2 < t3 < . . . , hence a polynomial in t is positive if,and only if, the coefficient of the highest power of t is positive. Thiscoefficient remains unchanged if we substitute t + 1 for t.

(3) If F is Archimedean as a field, then its additive group is an Archi-medean ordered group, hence there is an injective homomorphism ofordered groups α : F+ → R+. Composing α with an automorphismof R+ we may arrange that α(1) = 1, and then α induces the identityon the prime field Q ⊆ F . We claim that α is a homomorphism withrespect to multiplication (and hence an isomorphism of ordered fieldsF ∼= α(F ) ⊆ R). Let λf : F → F be left multiplication by f ∈ F×, thatis, λf (g) = fg. Similarly, we have λα(f) : α(F ) → R. The maps λα(f)

and α ◦ λf ◦ α−1 coincide on the dense subset Q ⊆ α(F ) because thegroup R+ is uniquely divisible. Both maps are order-preserving if f > 0and order-reversing if f < 0. This implies that the maps coincide, andthe proof is complete.

The implication (d) ⇒ (a) is proved again in 42.12, using the conceptof completion. �

11.15 Corollary Every completely ordered field is isomorphic to R.

Proof By 7.5, such a field F is Archimedean, and then F is isomorphicto a subfield of R by 11.14. This subfield is closed by completeness.Compare also 42.9. �

11.16 Transcendental extensions Part (2) of the proof of 11.14suggests a general question. If F is any field and F (t) is a simple trans-cendental extension, then the group of all automorphisms of F (t) fixingF elementwise is AutF F (t) = PGL2F ; see 64.19. This group consistsof the maps defined by

t �→ at + b

ct + d,

where a, b, c, d ∈ F and ad − bc �= 0. If F is an ordered field, we askwhich of these maps preserve the ordering on F (t) defined by F < t.

Page 139: The Classical Fields

122 Real numbers

The answer is that precisely the automorphisms defined by t �→ at + b

with 0 < a are order-preserving. We leave the proof as Exercise 2.

Exercises(1) Construct a countable subfield F of R such that exp induces an isomor-phism of (F, +) onto the multiplicative group of positive elements of F .

(2) Prove the claim in 11.16 on order-preserving automorphisms of simpletranscendental extensions.

(3) Use Theorem 11.14 in order to give a quick proof of Proposition 11.8.

(4) Show that (the additive group of) an ordered field F is Archimedean if,and only if, the multiplicative group of positive elements is Archimedean.

12 Formally real and real closed fields

In this section we study abstract fields which can be made into orderedfields (as defined in 11.1). Real closed fields are maximal (in a suitablesense; see 12.6) with respect to this property; they are characterized bycertain algebraic properties, which are familiar from the field R of realnumbers. We prove some fundamental results due to Artin and Schreieron real closed fields.

More information on formally real and real closed fields can be foundin Prestel 1984 (including model-theoretic aspects), Scharlau 1985Chapter 3, Jacobson 1989 Chapter 11, Lang 1993 Chapter XI, Pfis-

ter 1995, Bochnak et al. 1998, Prestel–Delzell 2001 Chapter 1and Lam 2005 Chapter VIII.

12.1 Definition A field F is said to be formally real , if −1 is not asum of squares in F .

Each ordered field F is formally real, since a2 ≥ 0 > −1 for everya ∈ F . We prepare the proof of the converse assertion (12.3).

12.2 Proposition Let F be a field, let a ∈ F× and let S ⊆ F be a

subset with the following properties:

S + S ⊆ S, S · S ⊆ S, x2 ∈ S for every x ∈ F, and S ∩ {−1, a} = ∅.Then there exists an ordering < on F such that a < 0 ≤ S.

Proof Let M = {T | S ⊆ T ⊆ F, T + T ⊆ T, T · T ⊆ T, −1 �∈ T }; theelements ofM are so-called preorderings of F . If T ∈M and b ∈ F �T ,then T − bT ∈ M; indeed, T − bT contains S and is closed underaddition and multiplication (since b2 ∈ S ⊆ T ), and the assumption

Page 140: The Classical Fields

12 Formally real and real closed fields 123

−1 ∈ T − bT implies −1 = t′ − bt with t′, t ∈ T , hence t �= 0 andb = t−2t(1 + t′) ∈ ST (1 + T ) ⊆ T , a contradiction. In particular,S − aS ∈M.

The union of any chain inM belongs toM, hence the same is true forthe non-empty subset Ma := {T ∈ M | S − aS ⊆ T }. Zorn’s Lemmaimplies the existence of a maximal element P ∈ Ma. By construction,P contains S and the element 0− a · 1 = −a. In view of 11.4b it sufficesto show that P � {0} is a domain of positivity in F . This task reducesto proving that P ∪ −P = F .

Let b ∈ F � P . Then P ⊆ P − bP ∈ Ma, hence P = P − bP by themaximality of P . In particular −b ∈ P , and therefore b ∈ −P . �

12.3 Corollary (Artin–Schreier) Each formally real field F can be

made into an ordered field (F,<).

Proof Let S be the set of all sums of squares of F . Then S · S ⊆ S,since (

∑i a2

i )(∑

j b2j ) =

∑i,j(aibj)2, and −1 �∈ S. Thus 12.2, applied

with a = −1, yields the assertion. �

12.4 Corollary Let F be a field with characteristic distinct from 2.

An element a ∈ F× is a sum of squares in F if, and only if, 0 < a for

every ordering < of F .

In other words: the intersection of all domains of positivity of F

consists precisely of all non-zero sums of squares of F .

Proof Each non-zero sum of squares is positive with respect to anyordering of F . For the converse, we consider two cases.

If F is formally real, and if a ∈ F does not belong to the set S of allsums of squares in F , then 12.2 implies that a < 0 for some ordering <

of F .If F is not formally real, then −1 is a sum of squares in F , and the

equation 4a = (a + 1)2 + (−1)(a− 1)2 shows that every a ∈ F is a sumof squares; furthermore F has no ordering in this case. �

12.5 Corollary A field F has a unique ordering if, and only if, the

non-zero sums of squares form a domain of positivity of F .

Proof If these sums of squares form a domain of positivity P , then P

is the only domain of positivity of F . Conversely, if F has a uniqueordering <, then 12.4 implies that {a ∈ F | 0 < a} consists of allnon-zero sums of squares. �

Page 141: The Classical Fields

124 Real numbers

Now we consider formally real fields which are maximal in the follow-ing sense.

12.6 Definition A field F is said to be real closed , if F is formallyreal, but no proper algebraic extension of F is formally real.

The field R and the field of all real algebraic numbers are examplesof real closed fields (by 12.13 and 12.11). By 24.6, the field ∗R of non-standard real numbers is a real closed field with non-Archimedean or-dering (12.7); compare also Exercise 4.

Let Γ be a divisible ordered abelian group. Then the field C((Γ)) ofHahn power series is algebraically closed (see 64.25), hence R((Γ)) is realclosed by 12.10 below. Different isomorphism types of ordered abeliangroups Γ give different isomorphism types of fields R((Γ)), because theordered group Γ is the value group of the natural ordering valuation(56.16, 64.25) determined by the unique ordering (12.7) of R((Γ)).

See 12.14, 12.18 and Ribenboim 1992, 1993 for more examples. Wemention that every real-closed field is isomorphic to an initial subfieldof Conway’s surreal numbers; see Ehrlich 2001.

12.7 Lemma Each real closed field F has a unique ordering <, and

this ordering is given by 0 < a⇔ a = b2 for some b ∈ F×.

Proof By 12.3 the field F has some ordering <. It suffices to show thateach element a ∈ F with 0 < a is a square in F . Otherwise we couldconsider the quadratic field extension F (

√a), which is not formally real

(as F is real closed). Then there exist elements xj , yj ∈ F such that

−1 =∑

j(xj + yj√

a)2 =∑

j(x2j + ay2

j ) + (∑

j 2xjyj)√

a .

As√

a �∈ F we infer that −1 =∑

j(x2j + ay2

j ) ≥ 0, which is absurd. �

12.8 Euclidean and Pythagorean fields A field F is said to beEuclidean if the non-zero squares of F form a domain of positivity of F

(equivalently, F is formally real and for each a ∈ F×, either a or −a isa square); then F has a unique ordering, which is defined as in 12.7.

The field R is Euclidean; see Lemma 11.6. The Euclidean closure E

of Q in R consists of all real algebraic numbers that can be reachedby iterated quadratic extensions F (

√c) with 0 < c ∈ F , starting from

F = Q. A point (x, y) of the real plane R2 can be constructed withruler and compass, starting from the the unit square, if, and only if,(x, y) ∈ E2; see Martin 1998 Chapter 2.

We mention a related concept: a field F is called Pythagorean, ifeach sum of squares of F is a square in F . The Pythagorean closure

Page 142: The Classical Fields

12 Formally real and real closed fields 125

P of Q in R consists of all real algebraic numbers that can be reachedby iterated quadratic extensions F (

√1 + c2) with c ∈ F , starting from

F = Q (compare Scharlau 1985 p. 52 or Lam 2005 VIII.4). A point(x, y) ∈ R2 can be constructed with ruler and dividers, starting from theunit square, if, and only if, (x, y) ∈ P 2; see Martin 1998 Chapter 5 orHilbert 1903 §§36f (constructions by ‘Lineal und Eichmaß’).

We remark that the field P is the splitting field in R of the set of all ra-tional polynomials which are solvable by real radicals; see Pambuccian

1990.A formally real field F is Euclidean if, and only if, F has no formally

real quadratic extension; see Becker 1974 Satz 1 or Lam 2005 VIII.1.7.By a result of Whaples, which was rediscovered by Diller and Dress(see Lam 2005 VIII.5.5 p. 269), a field F with characteristic not 2 isPythagorean precisely if F has no cyclic Galois extension of degree 4.See also Exercise 9.

The field P is a proper subfield of E; see Exercise 5. Each real closedfield is Euclidean, by 12.7, but the fields E and P are not real closed(since they are proper subfields of the ordered field of all real algebraicnumbers). See Ribenboim 1992 Section 6 or Lam 2005 VIII for moreexamples of Euclidean or Pythagorean fields.

A real closed field F is isomorphic to a subfield of R if and only if theunique ordering of F is Archimedean; see 11.14. The following resultgeneralizes the rigidity theorem 6.4 for R.

12.9 Corollary (i) Let E be an algebraic extension field of a field F . If

E is real closed, then AutF E = {id}, i.e., each field automorphism

of E which fixes each element of F is trivial.

(ii) Each real closed subfield F of R is rigid, i.e., EndF = {id}.Proof (i) By 12.7 the field E has a unique ordering, hence every elementϕ ∈ AutF E preserves this ordering. Furthermore ϕ leaves invariant thefinite set f−1(0) of roots of each polynomial f ∈ F [x], whence ϕ = id;compare step (2) of the proof of 11.13.

(ii) By 12.7 the field F has a unique ordering; this ordering coincideswith the ordering induced by R. Each endomorphism ϕ of F preservesthis unique ordering. Furthermore ϕ fixes each element of the primefield Q. Since Q is dense in R and in F , we infer that ϕ = id. �

Concerning 12.9(ii), we remark that Shelah 1983 constructs rigid realclosed fields which are not subfields of R (using suitable set-theoreticassumptions).

Page 143: The Classical Fields

126 Real numbers

Now we show that real closed fields are characterized by certain alge-braic properties of R; see 12.10, 12.12 and 12.15.

12.10 Theorem (Artin–Schreier) The following conditions for a

field F are equivalent.

(i) F is real closed.

(ii) F is Euclidean, and each polynomial in F [x] of odd degree has a

root in F .

(iii) The element −1 is not a square in F , and the field F (√−1) is

algebraically closed.

Proof (i) implies (ii): By 12.7, F is Euclidean. We show that eachpolynomial f ∈ F [x] of odd degree n has a root in F . It suffices toconsider irreducible polynomials f . We proceed indirectly and choose acounter-example f of minimal degree n = deg f . Then n > 1 and f hasno root in F . Let c be a root of f in an algebraic closure of F . ThenF (c) is a proper algebraic extension of the real closed field F , henceF (c) = {g(c) | g ∈ F [x], deg g < n} is not formally real. Therefore

−1 =∑

j gj(c)2

for suitable polynomials gj ∈ F [x] with deg gj < n, and we infer that c isa root of the polynomial h := 1+

∑j g2

j . Hence the minimal polynomialf of c divides h. The degree of h is even (consider leading coefficientsand use the ordering) and strictly smaller than 2n, hence the quotienth/f has odd degree deg h − n < n. Since n was chosen to be minimal,h has a root b in F , hence −1 =

∑j gj(b)2. This means that F is not

formally real, a contradiction.(ii) implies (iii): By (ii), the element −1 is not a square in F , hence

F (i) := F (√−1) is a quadratic field extension of F ; it remains to show

that F (i) is algebraically closed. We consider an irreducible polynomialf ∈ F (i)[x]. Let E be the field obtained from F (i) by adjoining all rootsof f (in an algebraic closure of F (i)). Then E is a Galois extensionof F (note that F has characteristic zero, hence E|F is separable by64.11), with a finite Galois group G (see 64.8, 64.10). The field of allfixed elements of a Sylow 2-subgroup H of G is an extension of F ofodd degree k = [G : H]; see 64.18. By (ii) and by the existence of aprimitive element (64.12) we have k = 1, hence G is a 2-group. TheGalois group Gi of the Galois extension E|F (i) is a subgroup of G,hence also a 2-group. In the following paragraph we show that Gi istrivial; then E = F (i), thus F (i) has no proper algebraic extension andis therefore algebraically closed.

Page 144: The Classical Fields

12 Formally real and real closed fields 127

If Gi is not trivial, then Gi has a subgroup H of index 2 (because thecomposition factors of any finite 2-group are cyclic of order 2; see alsoSylow’s theorem in the form of Jacobson 1985 1.13 p. 80). The fixedelements of H form a quadratic extension of F (i). However, by (ii) eachelement of F = {a ∈ F | a > 0}∪ {0}∪ {a ∈ F | −a > 0} is a square inF (i), and we show now that each element of F (i)�F is a square in F (i);indeed, if a, b ∈ F with b �= 0, then a2 + b2 > 0 and hence a2 + b2 = c2

with 0 < c ∈ F×, which gives c > a; thus 2(c − a) = d2 with d ∈ F×,and then a + ib = (b/d + id/2)2. Hence F (i) has no quadratic extension(note that −1 �= 1 in F ), which is a contradiction.

(iii) implies (i): Since F (i) = F (√−1) is algebraically closed, each

element a + ib with a, b ∈ F is of the form (c + id)2 with c, d ∈ F .Thus a = c2 − d2 and b = 2cd, which implies a2 + b2 = (c2 + d2)2. Weconclude that each sum of squares in F is in fact a square in F . Now −1is not a square, hence not a sum of squares, whence F is formally real.Furthermore the only proper algebraic extension F (i) is not formallyreal, thus F is real closed. �

12.11 Corollary Let F be a subfield of a real closed field E. Then the

field F �E of all elements of E which are algebraic over F is a real closed

subfield of E.

Proof Let a ∈ F be positive with respect to the unique ordering (12.7)of E. Then the polynomial x2−a has a root in E, and this root belongsto F �E . Hence the non-zero squares of F �E form a domain of positivityof F �E , whence F �E is Euclidean. Furthermore, if f ∈ F �E [x] is apolynomial of odd degree, then f has a root in E by 12.10; again thisroot belongs to F �E. Thus F �E is real closed by 12.10. �

12.12 Corollary A field F is real closed if, and only if, F admits an

ordering < such that the intermediate value theorem holds for polyno-

mials f ∈ F [x], i.e., a < b and f(a)f(b) < 0 imply that f(c) = 0 for

some c ∈ F with a < c < b.

Proof Let F be real closed. By 12.7, the field F has a unique orderingand each positive element is a square. We may assume that the leadingcoefficient of the given polynomial f is 1. By 12.10(iii), f decomposesinto factors x − c of degree 1 and quadratic factors x2 − 2dx + e =(x − d)2 + e − d2 with e − d2 > 0. Each irreducible quadratic factorx2 − 2dx + e assumes only positive values on F . Therefore the change

Page 145: The Classical Fields

128 Real numbers

of the sign of f between a and b occurs also at some factor x − c of f ,whence c is a root of f with a < c < b.

Conversely, assume that the intermediate value theorem holds forpolynomials. Let a ∈ F with a > 0. The intermediate value theo-rem for the polynomial x2−a implies that a is a square of some elementof F (between 0 and max{a, 1}). Hence F is Euclidean. Furthermoreeach polynomial f =

∑nj=0 ajx

j ∈ F [x] of odd degree n assumes posi-tive and negative values, as we show now. We may assume that an = 1.Choose b ∈ F larger than 1 +

∑j |aj |, where |a| := max{a,−a}. Then∑n−1

j=0 ±ajbj ≤ ∑n−1

j=0 |aj |bn−1 < bn for any combination of signs ±,hence f(b) > 0 > f(−b). Thus f has a root in F by the intermediatevalue theorem, and 12.10 implies that F is real closed. �

12.13 Corollary The field R of real numbers is real closed and the

field C = R(√−1) of complex numbers is algebraically closed.

Proof The intermediate value theorem holds for real polynomials, by 5.4and the continuity of polynomials. Hence R is real closed by 12.12 (orby 12.10). Now 12.10 shows that C is algebraically closed. �

The fact that C is algebraically closed is traditionally called the Fun-damental Theorem of Algebra. The proof given above is due to Artin;see Fine–Rosenberger 1997 or Ebbinghaus et al. 1991 Chapter 4,for other proofs.

Now we construct plenty of real closed fields (to be used in Theorem14.16).

12.14 Theorem Let ℵ = 2ℵ0 = card R.

(i) There exist precisely 2ℵ isomorphism types of real closed fields (of

cardinality ℵ) with Archimedean ordering (12.7). All these fields

are rigid.

(ii) There exist precisely 2ℵ isomorphism types of real closed fields F

with cardF = ℵ and card AutF = 2ℵ.

Proof (i) Real closed subfields of R are rigid by 12.9. According to 11.14,each Archimedean ordered field is isomorphic to a subfield of R, hencethere exist at most 2ℵ isomorphism types as in (i). Now we construct2ℵ isomorphism types.

Using Zorn’s Lemma, choose a transcendency basis T of R over Q.Then R is algebraic over the purely transcendental field Q(T ), hencecard R = card Q(T ) by 64.5. The set T is infinite, as R is uncountable.Hence card T = card Q(T ) = card R = ℵ; see 64.20. For each subset

Page 146: The Classical Fields

12 Formally real and real closed fields 129

S ⊆ T we write Q(S)� for the algebraic closure in C of Q(S), and wedenote by FS := R∩Q(S)� the real part of Q(S)�, as in 6.2d. Then Q(S)�

has degree 2 over FS , since i ∈ Q� ⊆ Q(S)� and FS is the field of fixedelements of the complex conjugation map γ, restricted to Q(S)�. Notethat FS is a real closed field by 12.11, hence it has a unique ordering(defined by the squares of FS); see 12.7.

We claim that the fields FS with S ⊆ T are mutually non-isomorphic.To prove this, we observe that the unique ordering of FS is the restrictionof the ordering of R. Each field isomorphism ϕ : FS → FS′ , whereS, S′ ⊆ T , preserves the unique ordering and fixes each element of Q.The strong density (3.1) of Q in R implies that ϕ is the identity, whenceFS = FS′ and S = T ∩ FS = T ∩ FS′ = S′.

This shows that there are 2ℵ isomorphism types of real closed fieldswith Archimedean ordering. If we admit only subsets S with cardS =cardT = ℵ, then cardFS = ℵ, and we still obtain 2ℵ isomorphism types,since card{S | S ⊆ T, cardS = cardT } = 2ℵ by 61.16.

(ii) The additive group of any field F as in (ii) is isomorphic to (R, +);see 1.14. Hence there exist at most 2ℵ isomorphism types of such fields(note that card(RR×R) = ℵℵ2

= ℵℵ = 2ℵ by 61.12 and 61.13b). Now weconstruct that many isomorphism types.

For every divisible ordered abelian group Γ, the real Hahn power serieswith countable support form a real closed field R((Γ))1; see 64.25. Ifcard Γ ≤ ℵ, then card R((Γ))1 = ℵ, since card R((Γ))1 ≤ card(R× Γ)N =ℵℵ0 = ℵ; see 61.13b. Different isomorphism types of ordered abeliangroups Γ give different isomorphism types of fields R((Γ))1, because theordered group Γ is the value group of the natural ordering valuation(56.16) of R((Γ))1. (See Alling–Kuhlmann 1994 Corollary 4.1 for arelated result.)

Now let Γ be the ordered group (R, +, <) where < is an orderingof (R, +) such that card Aut Γ = 2ℵ. By 7.14 there are 2ℵ isomorphismtypes for Γ, hence we have that many isomorphism types of fields R((Γ))1with cardinality ℵ. Moreover, each (non-trivial) automorphism of theordered group Γ induces a (non-trivial) field automorphism of R((Γ))1,and this implies that R((Γ))1 has precisely 2ℵ field automorphisms (notmore, since ℵℵ = 2ℵ by 61.13b). �

There exist many isomorphism types of skew fields which have di-mension 4 over a subfield isomorphic R; this amazing remark is due toGerstenhaber–Yang 1960 p. 462; see also Deschamps 2001. In-deed, for any field F as in 12.14, the quaternion skew field HF =

Page 147: The Classical Fields

130 Real numbers

F + Fi + Fj + Fij with centre F can be defined in the usual fash-ion; compare 34.17. Then F (i) = F + Fi ∼= C = R + iR by 12.10 and64.21, and HF = F (i)+F (i)j. Hence HF has dimension 4 over a subfieldisomorphic to R.

These copies of R are not contained in the centre F of HF , hence thereis no contradiction to the classical result of Frobenius, which implies thatHamilton’s quaternion skew field H is the only skew field of dimension4 over its centre R (see 58.11 or the remarks after 13.8).

The following characterization result is a remarkable achievement ofArtin and Schreier; the proof given below uses simplifications due toLeicht 1966 and Waterhouse 1985 (see also Lam 2005 VIII.2.21 forthe special case of fields F of characteristic 0).

12.15 Theorem (Artin–Schreier) Let F be a field. Then the alge-

braic closure F � of F has finite dimension n over F if, and only if, F is

real closed (then n = 2) or algebraically closed (then n = 1).

Proof If F is real closed, then F � has degree 2 over F by 12.10. Nowwe assume that F � �= F has finite degree over F . We have to show thatF is real closed; this is achieved by verifying condition 12.10(iii), seesteps (3) and (4). Let p denote the characteristic of F .

(1) We claim that F �|F is a Galois extension.Clearly F �|F is normal (64.9), so by 64.17 we have to show that F �|F

is separable. This holds for p = 0 (see 64.11), so let p > 0. Then x �→ xp

is a field endomorphism, hence F p = {ap | a ∈ F } is a subfield of F ,and (F �)p = F � since F � is algebraically closed. Furthermore the degree[F � : F p] = [(F �)p : F p] ≤ [F � : F ] is finite, hence F = F p.

The equation F = F p means, by definition, that F is perfect. Thisimplies that every algebraic extension of F is separable; indeed, if an ir-reducible polynomial f ∈ F [x] has multiple roots, then f and its deriva-tive f ′ have a common divisor. Thus f ′ = 0, hence f =

∑i aix

pi withai ∈ F . Since F is perfect, we have ai = bp

i with suitable bi ∈ F , hencef =∑

i bpi x

pi = (∑

i bixi)p, a contradiction to the irreducibility of f .(2) We claim that the characteristic p does not divide the order of the

Galois group G of F �|F .Otherwise G contains an element g of order p > 0 (by Sylow’s theo-

rem, compare Jacobson 1985 1.13 p. 80), and F � is a (cyclic Galois)extension of degree p of the field E = Fix g of fixed points of g. Thetrace map tr = 1 + g + · · · + gp−1 is an E-linear endomorphism of F �

into E. By Dedekind’s independence lemma (see Jacobson 1985 4.14

Page 148: The Classical Fields

12 Formally real and real closed fields 131

p. 291 or Cohn 2003a 7.5.1 or Lang 1993 VI.4), the powers gj with0 ≤ j < p are linearly independent. In particular, tr is not the trivialendomorphism. Thus tr(F �) = E, and we can choose a ∈ F � such thattr(a) = −1. The element

b :=∑p−1

j=1 jgj(a)

satisfies g(b) = b − tr(a) = b + 1, hence gj(b) = b + j for each j ∈ N.Thus b, b + 1, . . . , b + p − 1 is the orbit of b under 〈g〉. The equation∏

0≤j<p(x − j) = xp − x holds in E[x], since both sides have the sameroots (as a consequence of Fermat’s little theorem). Substituting x − b

for x we infer that

f(x) :=∏

0≤j<p

(x− b− j) = (x− b)p − (x− b) = xp − x + (−b)p + b

is the minimal polynomial of b over E (compare Section 64, Exercise 2).Its constant term (−b)p + b ∈ E can be written as (−b)p + b = tr(c) withc ∈ F �, and c = d − dp for some d ∈ F �, as F � is algebraically closed.We infer that (−b)p + b = tr(d− dp) = tr(d)− tr(d)p, hence tr(d) ∈ E isone of the roots b + j of f(x). This is a contradiction, as b �∈ E.

(3) We claim that G is a 2-group (hence p �= 2 by step 2) and thati =√−1 �∈ F .

Let q be any prime divisor of |G|. Then G contains an element g oforder q. Let E be the field of fixed elements of H := 〈g〉 in F �. ThenF �|E is a Galois extension of degree q with Galois group H (see 64.18).The irreducible polynomials in E[x] have degree 1 or q (use the degreeformula 64.2 and the fact that F � is algebraically closed). Hence xq−1 =(x− 1)(xq−1 + · · ·+ x + 1) splits over E into factors of degree 1. Sinceq �= p by step (2), we infer that xq − 1 is separable. Thus E contains aroot of unity ζ with ζq = 1 �= ζ.

As a consequence, the extension F �|E can be written in the formF � = E(r) where r is a root of some irreducible binomial f(x) = xq−c =∏

h∈H(x − h(r)) with c ∈ E. This well known fact (see Cohn 2003a7.10.7 or Lang 1993 VI.6.2) may be proved as follows. Every element

r :=∑q−1

j=0 ζjgj(a)

with a ∈ F � satisfies ζg(r) = r. By Dedekind’s independence lemma(see step (2) above) we can choose a ∈ F � such that r �= 0. Theng(r) = ζ−1r �= r, hence r /∈ E and therefore F � = E(r). Moreoverg(rq) = g(r)q = (ζ−1r)q = rq, hence c := rq ∈ E, and f(x) = xq − c isthe (irreducible) minimal polynomial of r over E.

Page 149: The Classical Fields

132 Real numbers

The algebraically closed field F � contains an element s with sq = r.The norm N(s) :=

∏h∈H h(s) belongs to E (as N(s) is fixed by H) and

satisfies the relation

N(s)q = N(sq) = N(r) = (−1)qf(0) = (−1)q+1c .

If q is odd, then N(s) ∈ E is a root of f , which is a contradiction tothe irreducibility of f over E. Hence q = 2 and N(s)2 = −c. Assumingi ∈ F we obtain that iN(s) ∈ E, and iN(s) is a root of f(x) = x2 − c.This contradiction shows that i �∈ F .

(4) It remains to show that F (i) = F �. Otherwise we could repeat theproof given above with F (i) instead of F , which in step (3) would leadto the contradiction i �∈ F (i). �

The following two results prepare the definition of real closures (12.18).

12.16 Proposition Each formally real field F has an algebraic exten-

sion which is real closed.

Each ordered field (F,<) has a real closed algebraic extension E such

that the unique ordering of E described in 12.7 induces on F the given

ordering <.

Proof Let F be formally real and let F � be an algebraic closure of F .The set {E | F ⊆ E ⊆ F � ∧ E is formally real} contains maximalelements by Zorn’s Lemma; each of these maximal elements is a realclosed field. This proves the first assertion.

Now let (F,<) be an ordered field. We claim that the subfield F ′ :=F (√

a | 0 < a ∈ F ) of F � is formally real. Otherwise we considerrelations of the form

−1 =∑

i cib2i ,

where 0 < ci ∈ F and each bi belongs to a subfield F (√

a1, . . . ,√

an)with 0 < aj ∈ F ; relations of this form exist, since we may take ci = 1.We choose such a relation with minimal n, and we write bi = xi +yi

√an

with xi, yi ∈ F (√

a1, . . . ,√

an−1); then

−1 =∑

i ci(x2i + any2

i ) + (∑

i 2cixiyi)√

an .

Since√

an �∈ F (√

a1, . . . ,√

an−1) by the minimality of n, we infer that−1 =

∑i cix

2i +∑

i ciany2i , a contradiction to the minimality of n.

By the first assertion, F ′ has an algebraic extension E which is realclosed. Each positive element a ∈ F is a square in F ′, hence the uniqueordering of E induces on F the given ordering <. �

Page 150: The Classical Fields

12 Formally real and real closed fields 133

12.17 Theorem (Artin–Schreier) Let F1, F2 be ordered fields, and

let Ei be a real closed algebraic extension of Fi such that the unique

ordering of Ei induces on Fi the given ordering, for i = 1, 2. Then each

order-preserving isomorphism ϕ : F1 → F2 has a unique extension to an

isomorphism ϕ : E1 → E2, and ϕ is order-preserving.

The uniqueness statement follows from 12.9(i) (consider the quotientof two isomorphisms ϕ). For the more difficult existence of an extensionϕ, see Prestel 1984 3.10, Scharlau 1985 Chapter 3 §2, Lang 1993XI.2.9, Jacobson 1989 Theorem 11.4 p. 656, Bochnak et al. 1998 1.3,Cohn 2003a 8.8.13, Prestel–Delzell 2001 1.3, or Lam 2005 VIII.2.We shall refer to 12.17 only in the following remarks and in 24.25.

12.18 Remarks on real closures A real closure of an ordered field(F,<) is defined to be a real closed algebraic extension field E such thatthe unique ordering of E induces on F the given ordering <. By 12.16,each ordered field (F,<) has a real closure E, and 12.17 says that E isuniquely determined up to isomorphisms of (F,<).

For example, the field Ralg of all real algebraic numbers is the realclosure of Q. The real closure of the Laurent series field Q((t)), orderedsuch that 0 < t < q for each positive q ∈ Q, is the field Ralg((t1/∞))of all Puiseux series over Ralg; this is a consequence of 12.10, sinceadjoining i =

√−1 to Ralg((t1/∞)) gives the field Calg((t1/∞)), which isalgebraically closed (see 64.24). The field R(t), taken with its orderingP0,+ as in Exercise 3, has as real closure the field R((t1/∞))alg of all realPuiseux series which are algebraic over R(t) (by Exercise 4 and 12.11).More examples can be found in Ribenboim 1992, 1993.

Despite the uniqueness statement above, an algebraic closure F � ofF may contain several copies of E. We describe an example of thisphenomenon (see also 14.15 for a plethora of embeddings of R into C).

Let F1 = Q( 3√

2) ⊆ R. Then F1∼= F2 := Q(ζ 3

√2), where ζ �= 1 = ζ3

is a root of unity of order 3. The field isomorphism α : F1 → F2 withα( 3√

2) = ζ 3√

2 extends to an automorphism α of the algebraic closureQ� of Q (see 64.15). By 12.11, the field E = Q�∩R = Ralg is real closed,hence a real closure of F = Q, and the same holds for the isomorphicfield α(E). Since ζ is not real, ζ 3

√2 ∈ α(E) � E, hence E and α(E) are

distinct copies of E in Q�.The existence and the uniqueness of the real closure of an ordered field

(i.e. a field with a given ordering) can be proved without using Zorn’sLemma (which appears in the proof of 12.16) or the axiom of choice;see Sander 1991, Lombardi–Roy 1991 and Zassenhaus 1970. In

Page 151: The Classical Fields

134 Real numbers

contrast, the existence of an ordering on every formally real field (12.3,12.2) cannot be proved without Zorn’s Lemma or one of its equivalents;see Lauchli 1962.

12.19 Hilbert’s 17th problem At the International Congress ofMathematicians in 1900 at Paris, Hilbert proposed a famous list of 23open problems. In his 17th problem, he asked whether each polynomialf ∈ R[x1, . . . , xn] with f(x) ≥ 0 for all x ∈ Rn is a sum of squares ofrational functions. This was proved by Artin in 1927, using the theoryof ordered fields and real closures, and Pfister proved in 1967 that f isactually a sum of 2n squares of rational functions. For details see Ja-

cobson 1989 11.4, Scharlau 1985 Chapter 3 §3, Prestel 1984 Theo-rem 5.7, Pfister 1995 Chapter 6, Bochnak et al. 1998 6.1.1, 6.4.18or Prestel–Delzell 2001. These results triggered the development ofreal algebra and real algebraic geometry; compare also Powers 1996.

We remark that rational functions are unavoidable in this context. Infact, the polynomial x4y2+x2y4−3x2y2+1 is non-negative on R2, but nota sum of squares of polynomials; compare Choi–Lam 1977, Bochnak

et al. 1998 6.3.6, 6.6.1 or Lam 2005 p. 519 for more examples.

Exercises(1) Let (F, <) be an ordered field, and let t be transcendental over F . Showthat < has precisely one extension to an ordering of F (t) such that a < t foreach a ∈ F .

(2) The field Q(√

2) has precisely two orderings.

(3) Show that the field R(t) of rational functions over R has domains of posi-tivity Pr,+ and Pr,− for r ∈ R ∪ {∞} with the following properties: for eachpositive real ε, we have r < t < r + ε with respect to Pr,+, and r − ε < t < rwith respect to Pr,−; moreover R < t with respect to P∞,+, and t < R withrespect to P∞,−. Show that R(t) has no other domains of positivity.

(4) The field R((t1/∞)) of all real Puiseux series is a real closed field withnon-Archimedean ordering.

(5) Show thatp

1 +√

2 ∈ E � P and 4√

2 ∈ E � P in the notation of 12.8.More generally, show that a ∈ E belongs to P if, and only if, a is totally real(that is, all roots of the minimal polynomial of a over Q are real).

(6) The field R has no maximal subfield.

(7) The field R contains an uncountable, strictly increasing family of realclosed subfields.

(8) Let F be a field such that the degrees of the irreducible polynomials inF [x] are bounded. Show that F is real closed or algebraically closed (hencethese degrees are bounded by 2).

(9) A field is Euclidean if, and only if, it is Pythagorean with a unique ordering.

Page 152: The Classical Fields

13 The real numbers as a topological field 135

13 The real numbers as a topological field

In this section we introduce the concept of topological fields, whichmeans that we add a compatible topology to the field structure. Weanalyse the interplay between topology and field structure, and in 13.8we obtain a characterization of R and C due to Pontryagin.

13.1 Definition A topological ring is a ring R with a topology suchthat(i) the additive group R+ is a topological group, and(ii) the multiplication R2 → R : (x, y) �→ xy is continuous.

Then the topology is called a ring topology of R. A topological (skew)field is a (skew) field R with a ring topology such that(iii) the inversion R× → R× : x �→ x−1 is continuous.Then F× is a topological group, and the topology is called a (skew) fieldtopology of F (see 13.5 for a slightly different definition).

13.2 Examples (a) The real field R, endowed with its usual topology,is a topological field. This fundamental fact is proved in 8.2 and 9.1.

More generally, every ordered field (F,<) as defined in 11.1 is a topo-logical field with respect to its order topology. Indeed, the additivegroup is easily seen to be a topological group (compare 8.2), and fromthe identity xy − ab = (x− a)y + a(y − b) we infer that the multiplica-tion is continuous (as in 9.1); now in view of x−1 = a−1(xa−1)−1 (seealso 8.3) it suffices to prove the continuity of inversion at 1, and for anygiven r ∈ F with 0 < r ≤ 1 the inequalities −r/2 < 1 − x < r/2 imply0 < x−1 < 2, hence −r < x−1 − 1 < r.

(b) Let F be a (skew) field with an absolute value | | : F → R, asdefined in 55.1. Then the definition d(x, y) = |x−y| yields a metric d onF which describes a (skew) field topology of F . This can be proved bysimilar considerations as in (a) and in 8.2, 9.1 (only the properties of anabsolute value are needed). In particular, the field C of complex num-bers is a topological field with its natural topology, since that topologyis derived from an absolute value (see Section 14). The same remarkapplies to the skew field H of Hamiltons’s quaternions; see Exercise 6.

Let F be a (skew) field with a valuation v : F → Γ∪{∞}, as defined in56.1; here Γ is an ordered abelian group. Then we obtain a (skew) fieldtopology of F by taking the sets {x ∈ F | v(x−a) > γ } with γ ∈ Γ as aneighbourhood base at a ∈ F ; the required arguments are again similarto the arguments in (a), see Warner 1989 Theorem 20.16 or Shell 19902.1 p. 15. We show only that inversion is continuous at 1: if 0 ≤ γ ∈ Γ,

Page 153: The Classical Fields

136 Real numbers

then v(x − 1) > γ entails v(x) = v(x − 1 + 1) = min{v(x − 1), 0} = 0and v(1− x−1) = v((x− 1)x−1) = v(x− 1)− v(x) > γ.

This valuation topology is always totally disconnected, because thesets {x ∈ F | v(x) > γ } are subgroups of F+ (since v is a valuation)which are open (by definition) and closed (by 62.7).

(c) Let F be a topological field, and let E be a (skew) field extensionof F of finite (left or right) degree n. We endow E with the producttopology with respect to any basis of E, considered as a (left or right)vector space over F . Then E is also a topological (skew) field, as weshow now.

The extension E is isomorphic (via right or left regular representationx �→ xa or x �→ ax) to a subfield of the ring F n×n of all n× n matricesover F . This matrix ring Fn×n, endowed with the product topologyof F n2

, is a topological ring with continuous inversion; indeed, matrixmultiplication is bilinear in each matrix entry and therefore continuous;the continuity of matrix inversion may be inferred from Cramer’s rule,which gives the matrix entries of A−1 as rational functions of the entriesof A; compare Shell 1990 7.2.

In particular, this shows again that C and H are topological (skew)fields with their natural topologies.

Several elementary properties of topological fields depend on the fol-lowing easy observation. (Results 13.3–13.7 hold also for skew fields.)

13.3 Homogeneity Lemma If F is a field with a ring topology, then

the affine maps x �→ ax + b with a �= 0 form a doubly transitive group

of homeomorphisms of F .

13.4 Corollary Let F be a field with a ring topology which is not in-

discrete. Then F is a completely regular space, in particular a Hausdorff

space, and F is either connected or totally disconnected.

Proof Since the topology is not the indiscrete one, we find an open setO and points a /∈ O, b ∈ O. By 13.3 we can cover F � {a} by images ofO under homeomorphisms fixing a, hence F � {a} is open in F . ThusF is a T1-space, and 62.4 says that F is completely regular. The secondassertion follows from 13.3 and 5.46. �

According to Shakhmatov 1983, 1987, there exist field topologieswhich are not normal.

13.5 Lemma Let F be a field with a T1-topology which renders F+

and F× topological groups. Then F is a topological field.

Page 154: The Classical Fields

13 The real numbers as a topological field 137

Proof By assumption, F× is open in F . Therefore the identity xy =(x+1)y− y implies that the multiplication is continuous on F ×F×, bysimilar identities also on F× × F and at (0, 0), hence everywhere. �

13.6 Lemma Let F be field with a compact ring topology which is not

indiscrete. Then F is finite and discrete.

Proof Using 13.3 we find a proper open subset W of F with 0 ∈W . Nowx · 0 = 0 for every x ∈ F , hence there exist open neighbourhoods Ux, Vx

of x and 0 respectively such that Ux · Vx ⊆ W . By compactness, F is aunion of finitely many sets Ux. The intersection V of the correspondingsets Vx is a neighbourhood of 0 with F · V ⊆ W . As F is a field andW �= F , we infer that V = {0}. Hence the topology is discrete, and F

is finite by compactness. �

The usual topology of R does determine the field structure, as thefollowing result shows.

13.7 Proposition Let F be a field with a ring topology such that F is

homeomorphic to R. Then F and R are isomorphic as topological fields.

Proof By 8.10 there exists an isomorphism ϕ : R+ → F+ of topologicalgroups, and we may assume that ϕ(1) = 1; compare 8.24. Then theunique divisibility of R+ implies that ϕ(Q) is the prime field of F , andthat ϕ|Q is multiplicative. Since Q is dense in R, we infer that ϕ ismultiplicative, hence an isomorphism of (topological) fields. �

The result of Kiltinen quoted in 13.10 below shows that the usualfield topology of R is far from being uniquely determined by the abstractfield R. However, additional assumptions on the topology do give stronguniqueness results:

13.8 Theorem (Pontryagin) Let F be a field with a Hausdorff ring

topology which is locally compact and connected. Then F is isomorphic

as a topological field to R or to C, endowed with their usual topologies.

Proof According to the structure theorem 63.14 for locally compactabelian groups, F+ is isomorphic as a topological group to a directsum Rn ⊕ C, where C is a compact (connected) group. In fact, C isthe largest compact subgroup of F+, because Rn has only one (trivial)compact subgroup; see 8.6. The multiplications x �→ ax with a ∈ F×

are automorphisms of F+, and these automorphisms act transitively onF � {0}, leaving C invariant. In view of 13.6 this implies that C = {0},hence F+ is isomorphic as a topological group to the vector group Rn.

Page 155: The Classical Fields

138 Real numbers

In particular, F+ is torsion free, whence F contains a copy Q of thefield of rational numbers. The topological closure Q of Q is the one-dimensional subspace spanned by 1, hence Q is isomorphic to R, evenas a topological field. Clearly F has finite dimension n over Q, hence F

is algebraic over Q ∼= R. Now C is algebraically closed (see 12.13) and2-dimensional over R, and this implies that F ∼= R or F ∼= C. �

Actually Pontryagin 1932 allowed also skew fields, and he obtainedthe skew field H of Hamilton’s quaternions (see Exercise 6 for a def-inition) as the only further possibility. In fact, the arguments abovestill apply, since Q ∼= R is contained in the centre, and then H is theonly further possibility by a well known algebraic result of Frobenius;see 58.11 or Jacobson 1985 7.7, Ebbinghaus et al. 1991 Chapter 8,Palais 1968 or Petro 1987.

13.9 Corollary (i) Let τ be a Hausdorff ring topology of the field R.

If τ is locally compact and connected, then τ coincides with the usual

topology of R.

(ii) Let τ be a Hausdorff ring topology of the field C. If τ is locally

compact and connected, then τ is the image of the usual topology of Cunder some field automorphism of C.

Proof Part (ii) holds by 13.8, and (i) follows from 13.8 and 6.4. �

Connectedness in 13.9 may be replaced by non-discreteness in bothcases, as we shall see in 58.8. The assumption of local compactnesscannot be omitted, since there exist strange (even locally) connectedfield topologies on R; see 14.10. Compare also Exercise 11.

13.10 Existence results Each infinite field has a field topology whichis neither discrete nor indiscrete; in fact, Kiltinen 1973 proved thateach infinite field F of cardinality c has 22c

distinct field topologies.Waterman–Bergman 1966 and Mutylin 1966 construct for any

field F an arcwise connected (metrizable) field topology on the fieldF1 = F (xt | t ∈ R) of rational functions in the indeterminates xt suchthat F is a discrete subfield of the connected field F1 (see also Wi

‘es�law

1988 Chapter 10); in fact, the set of indeterminates xt is homeomorphicto R. This shows that there exist connected fields of prime characteristic.

This was improved on by Mutylin 1968, who showed that any fieldF is a discrete subfield of a complete, metrizable, connected field E; inparticular, there exist complete, metrizable, connected fields of primecharacteristic. For the proof, one defines a (rather complicated) metricon F1 which gives a ring topology on F1 such that the completion F1 is

Page 156: The Classical Fields

13 The real numbers as a topological field 139

a ring with an open group of units, and then E is obtained as a quotientof F1 modulo a maximal ideal.

Mutylin 1968 proves also that the field C of complex numbers is aproper subfield of a complete, metrizable field E such that E induceson C the usual topology (see Section 14). He defines a suitable (compli-cated) ring topology on F1 = C(x), and then E is again obtained as aquotient of the completion F1.

Finally we mention a result of Gelbaum, Kalisch and Olmsted whichsays that every Hausdorff ring topology contains a (weaker) Hausdorfffield topology; see Warner 1989 Theorem 14.4, Wi

‘es�law 1988 I §3

Theorem 3 or Shell 1990 3.2.

Topological fields will appear again in later chapters, especially inSection 58, which deals with the classification of all locally compact(skew) fields.

A first introduction to topological fields can be found in Pontryagin

1986 Chapter 4, a fuller account is given by Warner 1989; see alsoWi

‘es�law 1988, Shell 1990. Some arguments remain valid without the

associative and distributive laws; see Grundhofer–Salzmann 1990.

Exercises(1) Show that the non-zero ideals of Z form a neighbourhood base at 0 for aring topology on Q which is not a field topology. Is there a ring topology ofR which is not a field topology?

(2) Every topological field whose underlying topological space is a 1-manifoldis isomorphic to R.

(3) For transcendental real numbers s and t, the subfields Q(s) and Q(t) ofR are algebraically isomorphic. Find necessary and sufficient conditions on sand t for the two fields to be isomorphic as topological fields. (It is an openproblem if this is the case for s = e, t = π.)

(4) Show that every non-empty open subset of R contains a transcendencybasis of R over Q.

(5) Let F be a field with characteristic distinct from 2. Then every grouptopology of F+ which renders the inversion of F× continuous is a field topologyof F .

(6) Show that the set H of all matrices

a −bb a

«

with a, b ∈ C, endowed with

the usual matrix operations, is a skew field; this is called the skew field ofHamilton’s quaternions. Show that R is the centre of H and that dimR H = 4.Show also that ϕ(A) = | det A| defines an absolute value ϕ : H → R (see 55.1)which induces on H the natural topology (as a real subspace of C2×2).

(7) Define real numbers by c1 = c =√

2 and cn+1 = ccn for n ∈ N. Show thatthe sequence (cn)n converges and find the limit.

Page 157: The Classical Fields

140 Real numbers

(8) For any sequence a = (an)n ∈ RN, we define its Cesaro transform Ca =(cn)n as the sequence of the arithmetic means cn := n−1 Pn

j=1 aj . Show thatCa converges to s if a converges to s. Find a sequence a such that Ca doesnot converge, but CCa does.

(9) Let an ∈ R with 0 < an < 1 for n ∈ N. ThenP∞

n=1 an < ∞ if, and onlyif,

Q∞n=1(1 − an) > 0.

(10) Show that every arcwise connected topological field is locally connected.

(11) Construct an arcwise connected Hausdorff ring topology on R which isdistinct from the usual topology. Is there also a field topology of R with theseproperties?

14 The complex numbers

Only one section of this book deals with the field C of complex numbers.This is due to the fact that many properties of C = R(i) can be inferredfrom those of R, as we show below.

First we consider the (topological) additive and multiplicative groupof C, and then we focus on the abstract field C. We also exhibit someunusual features of C, such as the existence of many discontinuous fieldautomorphisms and of strange subfields.

The set C = {a + bi | a, b ∈ R} with the usual addition and multi-plication of complex numbers (in particular, i2 = −1) is a field. Indeed,since the polynomial p = x2 + 1 is irreducible in the real polynomialring R[x], the quotient ring C := R[x]/pR[x] is a field, and we arrive atthe above description of C if we take i as an abbreviation for the cosetx + pR[x] (compare 64.8).

The mapping γ : C→ C : z = a + bi �→ γ(z) := z := a− bi, called thecomplex conjugation, is a field automorphism of C of order 2.

The definitions |a + bi| = √a2 + b2 for a, b ∈ R, d(x, y) = |x − y| forx, y ∈ C give a metric d on C that describes the usual topology of C.In fact, 1, i is a basis of C, considered as a vector space over R, and d

describes the product topology, hence C is homeomorphic to the productspace R× R.

With this topology, C is a topological field; this can be verified directly,but it is also a special case of 13.2(c), and of 13.2(b) as well, since|z| = √zz is an absolute value of C (as defined in 55.1).

Another method to introduce C is to define it as the set of all 2 × 2

matrices(

a −b

b a

)with a, b ∈ R; compare Exercise 6 of Section 13. It is

easy to verify that this ring C is a topological field, see 13.2(c); in thisdescription, the square | |2 of the absolute value is just the determinant.

Page 158: The Classical Fields

14 The complex numbers 141

The topological additive group C+ is isomorphic to R×R. The topo-logical multiplicative group C× is isomorphic to R× T, via the isomor-phism z �→ (|z|, z/|z|), which introduces polar coordinates in the plane;compare 1.20.

For the abstract groups the following holds.

14.1 Proposition As an abstract group, C+ is isomorphic to R+, and

C× is isomorphic to the torus group T.

Proof The group C+ ∼= R+⊕R+ is divisible and torsion free, and C hasthe same cardinality as R by 61.12, hence C+ ∼= R+ by 1.14. Further-more C× ∼= R× T ∼= T by 1.25. �

The topological groups C+ and R+ are not isomorphic, because Z+Zi

is a closed subgroup of C+ that is not cyclic; see 1.4. Therefore theusual topology of C cannot be induced by an ordering of the group C+:otherwise 3.3 and 7.9 would imply that C+ and R+ are isomorphic asordered groups and then also as topological groups, which is not true.

The relation i2 = −1 shows that there is no ordering that would becompatible with the multiplicative group C×; see 11.2. In particular, Cis not formally real (12.1). Therefore the concept of ordering will notplay any role in this section.

Now we consider the influence of the topology on the additive and onthe multiplicative group of C. The following results 14.2 and 14.4 aredue to von Kerekjarto 1931; we remark that the group L2 in 14.2 isa subgroup of index 2 in the affine group Aff R appearing in 9.3.

14.2 Proposition Let G be a topological group such that the under-

lying topological space is homeomorphic to R2. Then G is isomorphic

as a topological group to the abelian group R⊕R or to the non-abelian

group L2 := { ( 1 0a b ) | a, b ∈ R, b > 0}.

Proof This is an easy consequence of the theory of Lie groups and Liealgebras: such a group G is a Lie group by the solution of Hilbert’s 5thproblem; see Montgomery–Zippin 1955 4.10 p. 184 (for groups definedon surfaces, Hilbert’s 5th problem is somewhat easier; see Salzmann

et al. 1995 96.24). Furthermore G is simply connected, and the simplyconnected Lie groups (of dimension n) are classified by the real Liealgebras (of the same dimension n).

There exist only two Lie algebras of dimension 2, hence there are onlytwo possibilities for G; since R ⊕ R and L2 are clearly distinct, thiscompletes the proof. �

Page 159: The Classical Fields

142 Real numbers

14.3 Corollary Let A be a topological abelian group whose underlying

topological space is homeomorphic to C. Then A is isomorphic to C+

as a topological group.

Proof This is an immediate consequence of 14.2. We indicate a moredirect argument, which relies on the Splitting Theorem for locally com-pact abelian groups (63.14). This theorem implies that A is isomorphicas a topological group to Rn ⊕ C with a compact group C. It sufficesto show that C is trivial; then the assertion follows, because in Rn thecomplement of a point is connected but not simply connected only ifn = 2.

By a topological result of Kramer 2000, R2 is homeomorphic to aproduct Rn ×C only if C is homeomorphic to some space Rk, and thenk = 0 by the compactness of C.

Alternatively, we infer that C is simply connected (as A is simplyconnected), and we show now that every compact, simply connectedabelian group C is trivial. By 8.13b every character χ : C → T lifts toa continuous group homomorphism χ : C → R, that is, χ = p ◦ χ wherep : R→ T = R/Z is the canonical covering map as in 8.7. The compactsubgroup χ(C) of R is trivial (see 8.5), hence χ and χ = p◦ χ are trivial.Thus C admits only the trivial character; by 63.13, C is trivial. �

Each endomorphism of the group C+ is Q-linear, since C+ ∼= R2 isuniquely divisible (1.8, 1.9). Thus each continuous endomorphism ofthe topological group C+ is R-linear, as Q is dense in R. The ring of allcontinuous endomorphisms of C+ is therefore the endomorphism ring ofthe real vector space C = R2, and hence is isomorphic to the ring ofall real 2 × 2 matrices, and the group GL2R of all real invertible 2 × 2matrices is the automorphism group of the topological group C+.

The influence of the topology on the multiplicative group is evenstronger than in the real case (9.6); it is not necessary to assume com-mutativity, as we show now.

14.4 Corollary Let G be a topological group such that the underlying

topological space is homeomorphic to C � {0}. Then G is isomorphic to

C× as a topological group.

Proof The mapping p : R2 → G = C � {0} defined by ϕ(r, s) = er+is isa covering map (8.11), and C is simply connected. Hence we obtain aunique structure of a topological group on R2 with neutral element (0, 0)such that p becomes a group homomorphism; see 8.13a. The kernelp−1(1) = {0} × 2πZ is a discrete normal subgroup of that connected

Page 160: The Classical Fields

14 The complex numbers 143

group R2, hence contained in the centre (Exercise 4). By 14.2 there areonly two possible group structures on R2, and the centre of the groupL2 is trivial. This shows that we can identify our group operation onR2 with the usual addition of vectors. Doing so, we can maintain thatp−1(1) = {0} × 2πZ, because the automorphism group GL2R of R2

acts transitively on the non-zero vectors and on the non-trivial cyclicsubgroups of R2. Therefore G = p(R2) is isomorphic to R2/p−1(1) ∼=R⊕ (R/2πZ) ∼= R× T ∼= C× as a topological group. �

14.5 Proposition The continuous endomorphisms of the topological

group C× are precisely the maps

C× → C× : exp(r + is) �→ exp(ra + i(rb + ns))

where a, b ∈ R and n ∈ Z. The ring of all these endomorphisms is

isomorphic to the ring of all matrices

(a 0b n

)with a, b ∈ R, n ∈ Z.

Proof Let p : R2 → C�{0} be the covering map used in the proof of 14.4,and let α be a continuous endomorphism of C×. By 8.13 there existsa lift α : R2 → R2, that is, a continuous endomorphism α of R2 withα ◦ p = p ◦ α. This equation implies that the kernel p−1(1) = {0}× 2πZis invariant under α. As we have mentioned above, α is in fact R-linear.With respect to the basis 1, i of C = R2 the linear map α is describedby a matrix as in 14.5. Conversely, each map as in 14.5 is a continuousendomorphism of C×.

As an alternative, one could use the isomorphism R+ ⊕ T → C× :(r, s+2πZ) �→ exp(r + is) of topological groups, where T = R/2πZ. Forevery continuous endomorphism α of R⊕T, the image of the restrictionα|{0}⊕T is contained in T by compactness (and 8.5), hence by 8.27 thisrestriction is a map of the form (0, s + 2πZ) �→ (0, ns + 2πZ) for somen ∈ Z. The restriction α|R⊕{0} consists of an endomorphism of R and of acharacter of R, hence by 8.26 and 8.31a it is a map (r, 0) �→ (ra, rb+2πZ)for some a, b ∈ R. Assembling the pieces we obtain α(r, s + 2πZ) =(ra, rb + ns + 2πZ), hence the assertion. �

The real exponential function relates the additive and multiplicativegroups of R very closely; see 2.2. The complex exponential function(which was used in the proofs of 14.4 and 14.5) is a continuous grouphomomorphism exp : C+ → C×, which is surjective but not injective:by the periodicity of exp, the kernel is the discrete subgroup 2πiZ.

Page 161: The Classical Fields

144 Real numbers

Results 14.1 and 1.24 show that there exist group monomorphismsC+ → C×, but we prove now that none of these is continuous.

14.6 Proposition The continuous group homomorphisms of C+ into

C× are precisely the maps

C+ → C× : r + is �→ exp(ar + bs + i(cr + ds))

where a, b, c, d ∈ R. As a consequence, there exist no continuous group

monomorphisms of C+ into C×.

Proof The complex exponential function exp : C+ → C× is a coveringmap as defined in 8.11. By 8.13 each continuous group homomorphismϕ : C+ → C× lifts to a continuous group endomorphism ϕ of C+, thatis, ϕ = exp ◦ϕ. The maps ϕ arising here are precisely the R-linearendomorphisms of C+ ∼= R2, and these endomorphisms are described bythe real 2× 2 matrices.

Assume that ϕ is injective. Then the linear map ϕ is injective, hencebijective. Thus (ϕ)−1(2πi) is a non-trivial element of the kernel of ϕ =exp ◦ϕ, which is absurd. �

In the rest of this section, we study the abstract field C. It is a basicfact that C is algebraically closed; see 12.13, Fine–Rosenberger 1997or Ebbinghaus et al. 1991 Chapter 4 for various proofs of this result,the so-called the Fundamental Theorem of Algebra. Each algebraicallyclosed field of characteristic 0 with the same cardinality as C is isomor-phic to C; this is a special case of a result of Steinitz; see 64.21.

The simple transcendental extension C(t) and its algebraic closureC(t)� have the same cardinality as C (see 64.16 and 64.19), hence Cis isomorphic to C(t)�; see also 64.24 and 64.25. This implies that Cadmits uncountably many field automorphisms; compare 64.19. But wecan do much better; see 14.11 below (or Section 64, Exercise 4).

14.7 Addition and multiplication do not determine the field ofcomplex numbers The structure of the field C is not determined bythe isomorphism types of its multiplicative group and its additive group(see 6.2c, d and 34.2 for analogous questions on R and Q). Indeed, weshow that there exists a field F with F× ∼= C× and F+ ∼= C+ as groups,but F �∼= C as fields.

We call a Galois extension solvable, if its Galois group is solvable. LetF = C(t)solv be the union of all finite solvable Galois extensions of C(t)in some fixed algebraic closure of C(t); this union is actually a field; seeExercise 8. The field F is called the maximal prosolvable extension of

Page 162: The Classical Fields

14 The complex numbers 145

C(t), because GalC(t)F is a projective limit of finite solvable groups, butwe do not need this fact. We have card F = card C (see 64.16), henceF+ ∼= C+ by 1.14.

We claim that the multiplicative group F× is divisible. Let a ∈ F×,n ∈ N, and let p(x) =

∏kj=1(x − aj) ∈ C(t)[x] be the minimal poly-

nomial of a over C(t). Then each aj belongs to F . The splitting fieldEa := C(t)(a1, . . . , ak) ⊆ F of p(x) over C(t) is a solvable Galois exten-sion of C(t), by Exercise 8. Let E be the splitting field of p(xn) overC(t). Then E|C(t) is a Galois extension (by 64.10), and E is also thesplitting field of p(xn) =

∏i(x

n − ai) over Ea. Thus E|Ea is a so-calledKummer extension, and it is easy to see that the Galois group of E|Ea

is abelian: each element of this Galois group permutes the n-th roots ofaj by multiplication with an n-th root of unity ζj ∈ C ⊆ Ea (compareJacobson 1985 4.7 Lemma 2 p. 253, Morandi 1996 11.4 or Cohn

2003a p. 442). Hence the Galois extension E|C(t) is solvable (by themain theorem of Galois theory; see 64.18(ii)). We conclude that E ⊆ F ,hence F contains an n-th root of a.

Now we show that F× ∼= C×. The torsion subgroups of F× and C×

coincide. Being divisible, this torsion subgroup is a direct factor of bothF× and C×; see 1.22, 1.23. The complementary subgroups are divisibleand torsion free of the same cardinality, hence isomorphic by 1.14.

It remains to show that the field F is not isomorphic to C. Thesplitting field E of the polynomial x5 − 5tx + 4t ∈ C(t)[x] is a Galoisextension of C(t) with the non-solvable Galois group S5; see Malle–

Matzat 1999 I.9.4 (in fact, every finite group is a Galois group overC(t); see Malle–Matzat 1999 I.1.5 or Volklein 1996 Section 2.2.2).Thus E is not contained in F , hence F is not algebraically closed.

For several constructions below in this section, the following lemma iscrucial.

14.8 Basic Lemma Every transcendency basis T of C over Q has

cardinality ℵ = 2ℵ0 = card C.

Proof C is algebraic over the purely transcendental field Q(T ), hencecard C = card Q(T ) by 64.5. The set T is infinite, as C is uncountable.Hence T has the same cardinality as Q(T ) and as C; see 64.20. �

14.9 Corollary The field C admits precisely 2ℵ proper field endomor-

phisms with mutually distinct images; thus C has precisely 2ℵ proper

subfields which are isomorphic to C.

Page 163: The Classical Fields

146 Real numbers

Proof Using Zorn’s Lemma we choose a transcendency basis T of Cover Q; see 64.20. For each S ⊂ T with cardS = cardT , we have a fieldisomorphism Q(T ) → Q(S) which maps T bijectively onto S. Since Cis an algebraic closure of Q(T ), this field isomorphism extends by 64.15to a field endomorphism ϕS of C whose image ϕS(C) is algebraic overQ(S). We have T ∩ ϕS(C) = S, because T is algebraically independent.Hence the images of endomorphisms ϕS with distinct sets S as above aredistinct. Furthermore card{S | S ⊂ T, cardS = cardT } = 2ℵ by 14.8and 61.16. For an upper bound, we remark that 2ℵ is the cardinality ofthe set CC; see 61.10 and 61.15. �

Of course, none of the subfields in 14.9 is topologically closed in C.Indeed, R is the only proper subfield of C which is closed with respectto the usual topology (since the prime field Q is dense in R).

14.10 Dense connected subfields Dieudonne 1945 constructed aset S ⊆ C which is algebraically independent over Q such that the fieldQ(S) is locally connected (hence S is uncountable); the constructionis related to the proof of Theorem 8.23; see Shell 1990 Chapter 9,Wi

‘es�law 1988 Chapter 9, Warner 1989 Exercise 27.7, Bourbaki 1972

VI §9 Exercise 2 p. 470 for details.Then Q(S) is connected (compare 13.4) and as an abstract field not

isomorphic to, hence distinct from, R and C (see Exercise 5). In fact,Q(S) is not contained in R, as R has no proper connected subfield; see5.4. Hence Q(S) is a proper, dense, connected and locally connectedsubfield of C.

The algebraic closure Q(S)� of Q(S) in C is isomorphic to C by 64.21,hence Q(S)� has a subfield R which is isomorphic to R and containsQ(S). This field R is connected, locally connected (as Q(S) is locallyconnected and dense in R) and dense in C, but not complete (as R isnot closed in C).

This shows that the field R, hence also C, has a strange connected andlocally connected field topology; compare also Exercise 11 of Section 13.(We remark that Kapuano 1946 constructed an algebraically closedsubfield F of C with F ∩ R = Q� such that F has small inductivedimension 1.)

In contrast, Baer–Hasse 1932 Satz 6 (see also Warner 1989 Exer-cise 27.8) showed that R and C are the only arcwise connected subfieldsof C. This implies that C contains no arc consisting of elements whichare algebraically independent over Q (adjoin such an arc to Q and useExercise 5).

Page 164: The Classical Fields

14 The complex numbers 147

14.11 Theorem The group Aut C of all field automorphisms of C has

cardinality 2ℵ = 22ℵ0. The group Sym C of all permutations of C is

involved in Aut C, i.e., Sym C is isomorphic to a quotient of a subgroup

of Aut C. Hence every group of cardinality at most ℵ is involved in

Aut C. Furthermore Aut C has a subgroup which is free with a free

generating set of cardinality ℵ.Proof Using Zorn’s Lemma we choose a transcendency basis T of Cover Q; see 64.20. Each permutation of T extends (uniquely) to a fieldautomorphism of the purely transcendental field Q(T ). Since C is analgebraic closure of Q(T ), every automorphism of Q(T ) extends to anautomorphism of C; see 64.15. Thus the symmetric group Sym T isinvolved in Aut C, and SymT is isomorphic to Sym C by 14.8. Thisimplies that card Aut C = card Sym C = 2ℵ; see 61.16. (We observe thatSym C is not isomorphic to a subgroup of Aut C, as Aut C contains noelement of order 3; see 14.13(i)).

Every group G is isomorphic to the subgroup {x �→ xg | g ∈ G} of thesymmetric group Sym G. If card G ≤ ℵ, then by 61.8 we find an injectivemap G → C, hence Sym G is isomorphic to a subgroup of Sym C. Thisshows that G is involved in Aut C.

Now we take G to be a free group with a free generating set of car-dinality ℵ. Then G has cardinality ℵ; see 61.14. As G is involved inAut C, we find a subset S of Aut C with cardS = ℵ such that S is afree generating set modulo some subgroup of Aut C. Hence S is a freegenerating set for the subgroup 〈S〉 of Aut C. �

The following proposition says that continuity is a rare phenomenonin Aut C.

14.12 Proposition The identity and the complex conjugation z �→ z

are the only field endomorphisms of C which are continuous with respect

to the usual topology of C. Furthermore these are also the only field

endomorphisms of C which leave R invariant.

Proof Each field endomorphism α fixes 1 (by definition), hence α fixeseach element of Q. Continuity implies that α fixes each real number, asQ is dense in R; the same conclusion holds if α(R) ⊆ R; see 6.4. Fromthe equation α(i)2 = α(i2) = −1 we infer that α(i) ∈ {i,−i}, hence α isthe identity or complex conjugation. �

Every discontinuous field automorphism α of C is extremely discon-tinuous, in the following sense: for any arc J ⊆ C, the image α(J) isdense in C; for proofs see Salzmann 1969, Kallman–Simmons 1985

Page 165: The Classical Fields

148 Real numbers

and Kestelman 1951. The weaker assertion that α(R) is dense in Cis easy to prove: otherwise the topological closure of α(R) is a properclosed subfield of C, hence equal to R; then 6.4 implies that α is R-linear,and hence continuous.

Now we study maximal subfields of the field C and involutions in thegroup Aut C.

14.13 Theorem (i) The non-trivial elements in Aut C are either invo-

lutions or elements of infinite order.

(ii) If α ∈ Aut C is an involution and F = Fix α := {z ∈ C | α(z) = z },then F is a real closed field, C = F (i) and α(i) = −i.

(iii) The maximal subfields of C are precisely the subfields F such that

[C : F ] = 2. Assigning to an involution α ∈ Aut C the corresponding

field Fixα of fixed elements gives a bijection between the set of all

involutions in Aut C and the set of all maximal subfields of C.

(iv) Two involutions in Aut C are conjugate if, and only if, the corre-

sponding fields of fixed elements are isomorphic (as fields).(v) Distinct involutions in Aut C do not commute, their product has

infinite order.

Proof (i) If α ∈ Aut C has finite order n, then n = [C : Fixα] by 64.18;hence (i) follows from 12.15 (we require this result only for the field C;this special case of 12.15 has a shorter proof, as C has characteristic 0).

(ii) By 64.18 we have [C : F ] = 2, hence F is real closed by 12.15, and12.10 yields the assertion.

(iii) If F is a maximal subfield of C, then C = F (a) for every a ∈ C�F .If a were transcendental over F , then F < F (a2) < C, a contradiction.Thus the degree [C : F ] is finite, in fact equal to 2 by 12.15.

This shows that the maximal subfields of C are precisely the subfieldsF with [C : F ] = 2. By Galois theory (64.18), each of these subfieldsis the field of fixed elements of a unique involution in Aut C, and eachinvolution in Aut C gives such a subfield.

(iv) If two such involutions α, β are conjugate under ϕ ∈ Aut C, thenϕ induces an isomorphism between Fixα and Fixβ. Conversely, a givenisomorphism ϕ : Fix α→ Fixβ extends to an automorphism ϕ of C (by64.15 or more directly by (ii)), and ϕ ◦ α ◦ ϕ−1 is then an involutionfixing the elements of ϕ(Fix α) = Fixβ, and hence equal to β by (iii).

(v) If α, β ∈ Aut C are involutions, then α(β(i)) = α(−i) = i by (ii).Thus αβ is not an involution, again by (ii), hence it is trivial or of infiniteorder by (i). �

Page 166: The Classical Fields

14 The complex numbers 149

14.14 Proposition Let α ∈ Aut C be an involution and let F = Fixα

be the field of fixed elements of α.

(i) The centralizer Cs α := {β ∈ Aut C | αβ = βα} of α is isomorphic

to the direct product 〈α〉 ×AutF .

(ii) The group AutF is torsion free, and it is trivial if F is isomorphic

to a subfield of R.

Proof By 14.13(ii) we have C = F (i) and α(i) = −i. Each ϕ ∈ AutF

extends uniquely to an automorphism ϕ ∈ Aut C that fixes i; indeed,ϕ is given by ϕ(a + bi) = ϕ(a) + ϕ(b)i for a, b ∈ F . Clearly the groupK := {ϕ | ϕ ∈ AutF } is a subgroup of Csα.

Each β ∈ Cs α acts on F and on {i,−i}, and β(i) = i implies thatβ = β|F . Hence K is precisely the kernel of the action of Csα on the set{i,−i} = {i, α(i)}, whence Cs α = 〈α〉 ×K ∼= 〈α〉 ×AutF . This provespart (i).

Concerning (ii), we observe that the group Aut F is torsion free by14.13(i, v). Furthermore F is real closed by 14.13(ii). Hence by 12.9(ii)AutF is trivial if F is isomorphic to a subfield of R. �

14.15 Corollary The complex conjugation γ defined by γ(z) = z has

2ℵ conjugates in Aut C, and C contains 2ℵ maximal subfields that are

isomorphic to R. The centre of the group Aut C is trivial.

Proof Since Aut R is trivial (by 6.4), we infer from 14.14 that γ generatesits own centralizer. Hence γ has 2ℵ conjugates in Aut C (see 14.11),and the centre of Aut C is trivial. By 14.13(iii) the field C contains 2ℵ

maximal subfields that are isomorphic to R. �

Actually, we can do much better (or worse), as we show now.

14.16 Theorem (i) The group Aut C contains 2ℵ = 22ℵ0conjugacy

classes of involutions α with Cs α = 〈α〉; each of these conjugacy classes

has cardinality 2ℵ. The field C contains 2ℵ mutually non-isomorphic

maximal subfields that are isomorphic to a subfield of R; in fact 2ℵ

copies of each of these maximal subfields.

(ii) The group Aut C contains 2ℵ conjugacy classes of involutions α

with card Cs α = 2ℵ. The corresponding fields of fixed elements are

maximal subfields of C that are not isomorphic to a subfield of R; in

fact, such a field is real closed with a non-Archimedean ordering.

Proof Recall from 14.11 that card Aut C = 2ℵ.By invoking Theorem 12.14, we obtain many isomorphism types of

real closed fields F of cardinality ℵ. The algebraic closure F � of such a

Page 167: The Classical Fields

150 Real numbers

field F is isomorphic to C; see 64.21. Hence C contains a maximal sub-field isomorphic to F , and our results 14.13, 14.14 apply. In particular,distinct isomorphism types of fields F lead to distinct conjugacy classesof involutions in Aut C, and the cardinality of such a conjugacy class isalso the number of copies of maximal subfields of C that are isomorphicto F .

(i) By 12.14(i) there exist 2ℵ isomorphism types for F such that theunique ordering (12.7) of F is Archimedean. This leads to 2ℵ conjugacyclasses of involutions α ∈ Aut C with Cs α = 〈α〉, and each of theseconjugacy classes has cardinality |Aut C : 〈α〉| = card Aut C = 2ℵ.

(ii) By 12.14(ii) there exist 2ℵ isomorphism types of real closed fieldsF of cardinality ℵ such that card AutF = 2ℵ. The maximal subfieldsobtained here are not isomorphic to a subfield of R, because their uniqueorderings are not Archimedean; see 11.14. �

The field ∗R of non-standard real numbers is isomorphic to a maximalsubfield of C, since ∗R(i) = ∗C is algebraically closed by 21.7, and∗C ∼= C by 64.21 and 24.1. For the cardinality of Aut(∗R) see 24.27.

More information on (involutions in) automorphism groups of alge-braically closed fields can be found in Baer 1970b, Soundarara-

jan 1991, Schnor 1992. Assuming the continuum hypothesis (61.17),Evans–Lascar 1997 prove that every automorphism of Aut C is aninner automorphism.

The automorphisms of C which fix i form a subgroup of index 2 inAut C. More generally, for each Galois extension E of Q there is a fi-nite quotient group of Aut C (and of Aut Q�) which is isomorphic to theGalois group of E|Q; see 64.18. It is an old conjecture that every finitegroup arises in this fashion. Inverse Galois theory is working towardsa proof of this conjecture, which has been verified in many special in-stances, for example, for all solvable groups and for many simple groups;see Malle–Matzat 1999, Volklein 1996.

The group Aut C acts on the field Q� of all algebraic numbers, in-ducing (by 64.15) on Q� the group Aut Q�, which is a (profinite) groupof cardinality ℵ. The kernel of this action is the normal subgroup ofall automorphisms of C which fix each algebraic number. According toLascar 1992, 1997 this kernel is a simple group (of cardinality 2ℵ, see14.11). This unusual simple group is torsion free by 14.13(i, ii). Fromthis point of view, Aut C consists of a large simple subgroup, with thecomparatively small group Aut Q� (the absolute Galois group of Q) ontop.

Page 168: The Classical Fields

14 The complex numbers 151

At this inconspicuous place of the book, we dare to indulge briefly ingeometric applications.

14.17 The Euclidean plane If we identify the plane R2 with C asindicated at the beginning of this section, then the absolute value |z| of apoint z ∈ C is the Euclidean distance of z from 0. The multiplicativity ofthe absolute value |z| = |z| implies immediately that all maps z �→ az+b

and z �→ az + b with a, b ∈ C and |a| = 1 are Euclidean isometries; forexample, the map z �→ z is the reflection at the axis R, and z �→ az is arotation with centre 0; compare the proof of 1.20.

In fact, we have described all Euclidean isometries, and the complexaffine group

Aff C = {z �→ az + b | a ∈ C×, b ∈ C} ∼= C+ � C×

(see 9.3, 9.4) is a subgroup of index 2 in the group Aff C � 〈z �→ z〉 ofall similarity transformations of the Euclidean affine plane R2.

This close connection between the field C and the Euclidean plane R2

allows to deal with plane Euclidean geometry in an algebraic fashion,using complex numbers. Many examples for this approach can be foundin Hahn 1994, Ebbinghaus et al. 1991 Chapter 3 §4, and Schwerdt-

feger 1979; see also Connes 1998, Geiges 2001, Schonhardt 1963,Hofmann 1958.

14.18 Complex affine spaces Let n ∈ N. The complex affine spaceof dimension n is the lattice of all affine subspaces of the complex vectorspace Cn. If A ∈ GLnC and α ∈ Aut C, then the semi-linear bijection(A,α) : Cn → Cn defined by

(A,α)

⎛⎜⎝z1

...zn

⎞⎟⎠ = A

⎛⎜⎝α(z1)...

α(zn)

⎞⎟⎠for z1, . . . , zn ∈ C is clearly an automorphism of this lattice, or in otherterminology: a collineation of the affine space Cn. In fact, for n ≥ 2the semi-linear group ΓLnC := { (A,α) | A ∈ GLnC, α ∈ Aut C} =GLnC � Aut C is the group of all those collineations of this affine spacewhich fix the vector 0, by the fundamental theorem of affine geometry,but we do not need this fact.

Our aim here is to prove the following amazing result.

14.19 Theorem For each n ∈ N the semi-linear group ΓLnC contains

a subgroup G with the following properties.

Page 169: The Classical Fields

152 Real numbers

(i) G is transitive on the set of all ordered bases of Cn.

(ii) The identity is the only element of G that is continuous.

(iii) G is a free group with a free generating set of cardinality ℵ = 2ℵ0 .

Proof According to 14.11, Aut C contains a free subgroup H = 〈S〉,where S is a free generating set of cardinality ℵ. Choose any surjectionS → GLnC : s �→ As, and define f(s) := (As, s) ∈ ΓLnC. The mappingf : S → ΓLnC extends to a group homomorphism f : H → ΓLnC, sinceS is a free generating set of H. We show that the image G := f(H) hasall the required properties.

Applying f(s) = (As, s) to a basis of the rational vector space Qn hasthe same effect as applying only As. Since we have chosen a surjections �→ As, we obtain all bases of Cn by varying s ∈ S, hence (i) holds.(By the subsequent paragraph, the set f(S) is a free generating set ofG, so G is never sharply transitive on the bases).

Let 1 �= h ∈ H. Because the free group H contains no involution(see Exercise 9), h and (id, h) are discontinuous by 14.12. Furthermoref(h) = (A, h) for some A ∈ GLnC, and the equation f(h) = (A, h) =(A, 1)(id, h) shows that f(h) is discontinuous, hence (ii) holds. Thediscontinuous map f(h) cannot be the identity. Hence the group epi-morphism f : H → G is injective, and G is isomorphic to the free groupH, whence (iii). �

In the proof above, the structure of the group GLnC did not matterat all; we have used only the fact that GLnC is a group of cardinality atmost ℵ consisting of homeomorphisms. This means that we obtain simi-lar examples by substituting for GLnC any group of homeomorphismsof Cn with a reasonable transitivity property (note that the homeomor-phism group of Cn has cardinality ℵ, since Cn has the countable densesubset Q(i)n).

In particular, we can consider subgroups of the full affine group ofall collineations x �→ γ(x) + a with γ ∈ ΓLnC, a ∈ Cn; by choosingthe subgroup {x �→ x + a | a ∈ Zn } ∼= Zn, we obtain a free group ofdiscontinuous collineations which induces on Qn precisely the translationgroup Zn.

Glatthaar 1971 considered the complex projective plane and thegroup GL3C, which is transitive on ordered quadrangles. We remarkthat Tits 1974 11.14 uses a similar construction of free groups in thecontext of algebraic groups over fields that are purely transcendentalover a subfield.

Page 170: The Classical Fields

14 The complex numbers 153

Exercises(1) Show that each non-cyclic discrete subgroup of C+ is of the form Za + Zb,where a, b ∈ C are R-linearly independent.

(2) The circle S1 is the largest compact subgroup of C×.

(3) Show that each continuous group homomorphism ϕ : C× → C+ is of theform ϕ(z) = a log |z| with a fixed complex number a ∈ C.

(4) Each discrete normal subgroup N of a connected topological group G iscontained in the centre of G.

(5) Neither R nor C is purely transcendental over any proper subfield.

(6) Show that a ∈ C is contained in some maximal subfield of C if, and onlyif, the field Q(a) if formally real.

(7) Show that { (a+ b, a+ ζb, a+ ζ2b) | a ∈ C, b ∈ C× } is the set of all equilat-eral triangles in the Euclidean plane R2 = C, where ζ ∈ C is a root of unity oforder 3. Use the existence of discontinuous field automorphisms of C to provethat the following geometric notions cannot be expressed just by equilateraltriangles: collinearity of points, parallelity, orthogonality and congruence ofsegments determined by two points (see also Beth–Tarski 1956).

(8) Let F ⊆ Ei ⊆ F � be fields such that Ei|F is a finite solvable Galoisextension (as defined in 14.7), where i = 1, 2. Show that the compositeE1E2 := F (E1 ∪ E2) is a finite solvable Galois extension of F .

(9) Show that a free group contains no element of order 2.

Page 171: The Classical Fields

2

Non-standard numbers

This chapter has a threefold purpose: to introduce ultraproducts andconstruct the real numbers via an ultrapower of Q, to study the structureof an ultrapower ∗R of the reals in a similar way as this has been donefor R in Chapter I, and to illustrate how the non-standard extension ∗Rmay be used to prove results on R in an easy way.

We will exclude, however, the many deeper results on non-standardnumbers which require a discussion of a hierarchy of formal languagesnecessary for a development beyond the simplest notions of non-standardanalysis; see, e.g., Hurd–Loeb 1985, Lindstrøm 1988, Keisler 1994,or Goldblatt 1998. We need the axiom of choice.

21 Ultraproducts

Ultraproducts can be defined for any infinite index set. Later on, onlynatural numbers will be used as indices. The notion of an ultrafilter isessential for the construction.

21.1 Filters For any set N write 2N = {S | S ⊆ N }. A subset Φ ⊆ 2N

is called a filter on (or over) N if the following three conditions hold:

∅ /∈ Φ, A, B ∈ Φ⇒ A ∩B ∈ Φ, (A ∈ Φ ∧ A ⊆ B ⊆ N)⇒ B ∈ Φ .

The union of any chain of filters is again a filter. Therefore, by Zorn’sLemma, each filter on N is contained in a maximal one.

21.2 Ultrafilters A filter Ψ over N is called an ultrafilter , if Ψ hasone of the following equivalent properties:(a) Ψ is a maximal filter on N ,(b) for each A ⊆ N either A or its complement A� belongs to Ψ,(c) whenever A ∪B ∈ Ψ, then A ∈ Ψ or B ∈ Ψ.

154

Page 172: The Classical Fields

21 Ultraproducts 155

Proof Obviously, (b) ⇒ (a), and (b) is a special case of (c) becauseN ∈ Ψ. If A ∪B ∈ Ψ, then (A ∪B)� = A� ∩B� /∈ Ψ, and (b) ⇒ (c).

Assume that (a) holds; we prove (b). If A ∩X = ∅ for some X ∈ Ψ,then X ⊆ A� ∈ Ψ; if A ∩X �= ∅ for each X ∈ Ψ, then

〈Ψ, A〉 ={

Y ⊆ N∣∣ ∃X∈Ψ A ∩X ⊆ Y

}is the filter generated by Ψ and A, and A ∈ Ψ by maximality. �

Equivalently, an ultrafilter Ψ over N may be considered as a finitelyadditive probability measure μ : 2N → {0, 1}, the null sets being exactlythe sets A /∈ Ψ. An extensive treatment of ultrafilters can be found inComfort–Negrepontis 1974.

21.3 Free ultrafilters Obviously, each one-element set {a} generatesan ultrafilter of the form 〈a〉 = {X ⊆ N | a ∈ X }. These so-calledprincipal ultrafilters are uninteresting.

The non-principal or free ultrafilters Ψ can be characterized by eachof the following equivalent properties:(a)⋂

Ψ = ∅ (if⋂

Ψ = A �= ∅, then Ψ =⋂

a∈A〈a〉 is either principal ornot maximal),

(b) Ψ does not contain any finite set (if there is a finite set in Ψ, then⋂Ψ �= ∅),

(c) Ψ contains the cofinite filter on N (consisting of all sets with finitecomplement).

21.4 Ultraproducts Let P =×ν∈N Sν ={

(xν)ν

∣∣ ∀ν∈N xν ∈ Sν

}be the product of a family of non-empty sets Sν . If Ψ is any ultrafilteron N , two sequences are called equivalent modulo Ψ if they agree onsome set of the filter, formally

(xν)νΨ(yν)ν � {ν ∈ N | xν = yν } ∈ Ψ .

The equivalence class of x = (xν)ν will be denoted by x/Ψ, sometimessimply by x. The quotient P/Ψ is called an ultraproduct of the Sν ;generally, its structure depends on the choice of Ψ.

If Sν = S for ν ∈ N , we denote the ultraproduct by SΨ and speak ofan ultrapower . There is a canonical injection s �→ (s)ν/Ψ of S into SΨ.

Remarks (1) It is often convenient to express a statement of the form{ν ∈ N | A(ν)} ∈ Ψ by a phrase like ‘condition A(ν) holds for almostall ν ∈ N ’. Since Ψ contains all cofinite sets, this use of the expression‘for almost all’ extends the traditional one. Property 21.2(b) shows thatx �= y if, and only if, xν �= yν for almost all ν.

Page 173: The Classical Fields

156 Non-standard numbers

(2) If Ψ = 〈n〉 is a principal ultrafilter, then the projection πn : P → Sn

induces an isomorphism P/Ψ ∼= Sn. For this reason, it will be tacitlyassumed in all constructions that Ψ is a free ultrafilter.(3) If S is a finite set, then condition 21.2(c), inductively applied toa union of cardS sets, implies that the canonical injection S → SΨ issurjective.

21.5 Chains An ultraproduct of chains (i.e. linearly ordered sets) isagain a chain. In fact, if x �= y, then either xν < yν or xν > yν foralmost all ν. (Use 21.2.)

21.6 Fields Assume that in the definition 21.4 each Sν is a field. Bythe properties of an (ultra-)filter, the operations x + y = x + y andx · y = xy are well-defined. Obviously, (P/Ψ, +, ·) is a ring. If x �= 0,we may assume that xν �= 0 for each ν. Then (x−1

ν )ν/Ψ is a multiplica-tive inverse of x, and P/Ψ is even a field. Analogous assertions holdfor other algebraic structures, because all first-order statements (i.e.,statements which do not involve quantifiers for set variables, compareChang–Keisler 1990 1.3) carry over to ultraproducts.

21.7 Theorem: Algebraically closed fields An ultraproduct of

algebraically closed fields is algebraically closed.

Proof Let P =×ν∈N Sν be a product of algebraically closed fields. Fork > 0 and κ = 0, 1, . . . , k consider coefficients cκ = (cκν)ν/Ψ ∈ P/Ψ. Wemay assume that ckν �= 0 for each ν ∈ N . Since Sν is algebraically closed,there is some xν ∈ Sν such that

∑kκ=0 cκνxκ

ν = 0. Then x = (xν)ν/Ψ

satisfies∑k

κ=0 cκxκ = 0. �

An analogous result holds for real closed fields, since these fields canbe characterized by a first-order property; see 12.10.

21.8 Example An ultraproduct F =×p∈P Fp/Ψ of all Galois fields ofprime order has the following properties:(a) charF = 0,(b) each element in F is a sum of two squares,(c) the element −1 is a square in F if, and only if, the set{p ∈ P | p ≡ 1 mod 4} belongs to Ψ,

(d) F has cardinality ℵ = 2ℵ0 ,(e) the algebraic closure F � of F is isomorphic to C,(f) F � is a proper subfield of×p∈P F �

p/Ψ.

Proof (a) One has n · 1 �= 0 in Fp for almost all p ∈ P, hence n · 1 �= 0for all n ∈ N.

Page 174: The Classical Fields

21 Ultraproducts 157

(b) Put Q = {x2 | x ∈ Fp }. If p is odd, then Q has (p+1)/2 elements,and so has c−Q for any c ∈ Fp. Therefore, Q and c−Q intersect, andhence there are a, b ∈ Fp with a2 + b2 = c (see also 34.14). By theconstruction of F , each c ∈ F is a sum of two squares.

(c) Recall that F×p is a cyclic group of order p− 1 (Section 64, Exer-

cise 1). Hence −1 is a square in Fp if, and only if, p−1 ∈ 4N. The claimis now an immediate consequence of the definition of an ultraproduct.

(d) is a special case of the following Theorem 21.9.(e) is true for any field satisfying (a) and (d); see 64.21 and 64.20.(f) Choose cp ∈ F �

p such that cp has degree sp over Fp and the sp areunbounded. Then c = (cp)p/Ψ is not algebraic over F : if

∑mμ=0 aμcμ = 0

for aμ ∈ F , then∑m

μ=0 aμpcμp = 0 and sp ≤ m for almost all p ∈ P. �

21.9 Theorem: Cardinality If Ψ is a free ultrafilter over a countable

set, say over N, and if the cardinalities of the sets Sν satisfy the condi-

tions ν < cardSν ≤ ℵ, where ℵ := 2ℵ0 , then card (×ν∈N Sν/Ψ) = ℵ.Proof For each of the ℵ functions f = (fν)ν ∈ {0, 1}N define a mapf� : N → N by f�(n) =

∑ν<n fν2ν < 2n. Choose s(μ, ν) ∈ Sν for

μ = 0, 1, . . . , ν such that s(μ, ν) �= s(μ′, ν) whenever μ �= μ′. Put fν =s(f�(hν), ν), where hν is the largest integer such that 2hν −1 ≤ ν. (Thiscondition implies that f�(hν) ≤ ν, thus it guarantees that fν is defined.)

If f �= g, then fκ �= gκ for some κ and f�(n) �= g�(n) for all n > κ.Now ν ≥ 2κ+1 implies hν > κ, f�(hν) �= g�(hν) and fν �= gν . Hencef and g represent different elements of the ultraproduct P/Ψ, and ℵ ≤card(P/Ψ) ≤ cardP ≤∏ν cardSν ≤ ℵℵ0 = 2ℵ0 by 61.13b. �

21.10 Extension of functions Let an ultrafilter Ψ over N be given.Any map h : S → T has a canonical extension hΨ : SΨ → TΨ definedby hΨ((sν)ν/Ψ) = (h(sν)ν)/Ψ or, in short, hΨ(s) = h(s). The extensionwill often also be denoted by h without danger of confusion.

Exercises

(1) Discuss an ultraproduct E =×ν∈N Fpν /Ψ in a similar way as the field Fin example 21.8.

(2) Put h(x) = x3. Show that hΨ is an automorphism of the chain RΨ.

(3) Determine the image of sinΨ : RΨ → RΨ .

Page 175: The Classical Fields

158 Non-standard numbers

22 Non-standard rationals

In this section, the structure of the ultrapower QΨ with respect to a givenfree ultrafilter Ψ over N will be studied in a similar (but less detailed)manner as the structure of R has been discussed in Chapter 1. Thespecial choice of Ψ will not play any role (compare 24.28), and QΨ willalso be denoted by the customary name ∗Q. The properties of ∗Q willbe compared with those of Q or R.

Relevant results on Q can be found in Chapter 3, most of them arequite familiar. By P we denote the set of all primes in N, as usual.The inclusions P ↪→ N ↪→ Z ↪→ Q give rise to natural embeddings∗P ↪→ ∗N ↪→ ∗Z ↪→ ∗Q. Statements for which no proof is given followdirectly from the definition 21.4.

22.1 Fact ∗Q is an ordered field of cardinality ℵ, and ∗Q is the field

of fractions of its subring ∗Z. Fractions can be added and multiplied in

the same way as in Q.

The assertion that card ∗Q = ℵ is a special case of Theorem 21.9.From Lagrange’s theorem (34.18) we infer the following result.

22.2 Each positive element in ∗Z and hence also in ∗Q is a sum of at

most four squares.

22.3 Corollary ∗Q has only one ordering which satisfies the mono-

tonicity laws.

Notation For positive numbers, we write a" b � ∀n∈N n · a < b. Theabsolute value |c| of c is defined by the condition 0 ≤ |c| ∈ {c,−c}, asusual, and c is said to be infinitely small or infinitesimal if |c| " 1.

22.4 Additive group ∗Q+ is a vector space of dimension ℵ over the

subfield Q, hence ∗Q+ ∼= R+, and ∗Q+ is not locally cyclic.

Proof The dimension is the cardinality of a basis b of ∗Q over Q. Since∗Q is the union of all subspaces spanned by finitely many elements inthe basis, ℵ = card ∗Q = card b. In detail, any finite-dimensional vectorspace over Q is countable. Hence b is infinite. By 61.14, the basis b hasexactly card b finite subsets. Thus, ∗Q is covered by card b countablesubspaces, and card ∗Q = card b. �

22.5 Primes Let 1 �= p = (pν)ν/Ψ. Then p is a prime element in ∗Nif, and only if, pν is a prime in N for almost all ν. Hence ∗P is the set of

all prime elements in ∗N. �

Page 176: The Classical Fields

23 A construction of the real numbers 159

22.6 Multiplication ∗P generates a proper subgroup of the multi-

plicative group ∗Q×pos of the positive elements in ∗Q.

Proof Consider P = {pκ}κ in its natural ordering. Put qν =∏

κ≤ν pκ

and q = (qν)ν/Ψ. If s = (sν)ν/Ψ is a product of m elements in ∗P, then,for almost all ν, the component sν has only m prime factors, and s �= q.

22.7 Proposition ∗Q×pos is a non-Archimedean ordered group.

Proof If 0 < c" 1, then (1 + c)n ≤ (1 + 1n )n <

∑ν 1/ν! < 3. �

22.8 Ordering The group ∗Z+ is a non-Archimedean ordered group

of cardinality ℵ. The shift z �→ z + 1 is an automorphism of the chain∗Z, it maps each element onto its immediate successor. There is an

embedding ϕ of the first uncountable ordinal Ω = ω1 into ∗N, but ∗N is

not well-ordered.

Proof For an inductive definition of ϕ, let ϕ(0) = 0 and define ϕ(ν + 1)as the successor of ϕ(ν); if σ ∈ Ω is not of the form ν + 1, the countableset {ϕ(ν) | ν < σ } has an upper bound s ∈ ∗N (see 24.11), and ϕ(σ)may be chosen as s + 1. The shift z �→ z + 1 maps N into itself andinduces an automorphism of the chain M = ∗N � N. Hence M has nosmallest element. �

Exercises(1) Let Q� and QΨ be embedded into (Q�)Ψ in a natural way. Show thatQ� ∩ QΨ = Q.

(2) The continued fraction for√

2 gives a sequence of rational numbers sν

converging to√

2 such that s2ν <√

2 < s2ν+1. Let s = (sν)ν/Ψ ∈ QΨ. Showthat s2 < 2 or s2 > 2 depending on the choice of Ψ.

(3) The number of primes p ≤ n is denoted by π(n). The prime numbertheorem says that n−1π(n) log n converges to 1 for n → ∞; see, e.g., Hardy–Wright 1971 Theorem 6 p. 9 or Zagier 1997. Write pν for the νth primein N and use the prime number theorem in order to show that (ν)ν/Ψ �(pν)ν/Ψ � (ν2)ν/Ψ.

23 A construction of the real numbers

Axiomatically, the real numbers are characterized as a completely or-dered field (as defined in 42.7). There are several ways in which R canbe obtained by completion of Q. Most familiar are the following two.

Page 177: The Classical Fields

160 Non-standard numbers

(I) First the ordered set Q is completed, say by Dedekind cuts, and thenthe algebraic operations are extended. The crucial step is to show thatthe resulting ring has no zero divisors.(II) All sequences converging to 0 form a maximal ideal M in the ring C

of all rational Cauchy sequences. The ordering of Q is extended to C/M ,and one has to show that this quotient is order complete (see 23.7).Details and related methods can be found in Chapter 4.

The procedure of Meisters–Monk 1973 to be described in this sec-tion is particularly transparent. Compared to (I) and (II), it has theadvantage that the algebraic operations and the ordering relation areextended simultaneously. If R is the ring of all finite elements (see 23.1)in ∗Q and N is the (maximal) ideal of all infinitesimals (Section 22),then R/N turns out to be a completely ordered field, hence R ∼= R/N .The proof is independent of Section 22, it uses only 21.4 and 21.5.

A proper subring R of a field F is said to be a valuation ring , ifa ∈ F � R implies a−1 ∈ R; compare 56.4.

23.1 The set R = {a ∈ ∗Q | ∃n∈N |a| < n}, consisting of all ‘finite’

elements, is a valuation ring.

Proof Sum and product of finite elements are obviously finite again, sothat R is a ring. If a /∈ R, then 1 " |a| and a−1 is infinitesimal, hencea−1 ∈ R. �

23.2 The infinitesimals form a maximal ideal N of R, and R�N = R×

is the group of units in R.

Proof By definition, N is an ideal in R. If a−1 /∈ R, then |a|−1 # 1,and a ∈ N . Therefore, each element in R � N has an inverse in R, andany proper ideal in R is contained in N . �

23.3 With the ordering induced by ∗Q, the ring R is an ordered ring,

and the ideal N is convex (that is, if a, b ∈ N and a < c < b, then

c ∈ N). �

23.4 Corollary The residue field F := R/N is an ordered field.

Proof The canonical projection R → R/N preserves the ordering: infact, 0 < c /∈ N implies N < c + N by the convexity of N . �

23.5 Lemma For a given k ∈ N, any element b ∈ R has a representation

b = (bν)ν/Ψ such that |bν − bμ| < k−1 for all μ, ν ∈ N.

Page 178: The Classical Fields

23 A construction of the real numbers 161

Proof Since b is finite, there is some m ∈ N such that |b| < m and hence|bν | < m for almost all ν. Divide the interval [−m,m] into finitely manysubintervals of length less than k−1. By property 21.2(c) of ultrafilters,there is one among these subintervals, say J , such that K = {ν ∈ N |bν ∈ J } ∈ Ψ. According to the definition of the equivalence relation Ψ,the bμ with μ /∈ K may be chosen arbitrarily in the interval J withoutchanging the element b. �

23.6 Proposition The canonical inclusion κ : Q ↪→ ∗Q (by constant

sequences) induces an embedding Q ↪→ F with strongly dense image.

Proof Since κ : Q ↪→ R and Qκ ∩ N = {0}, the map κ induces an em-bedding Q ↪→ F . Consider elements x, y ∈ R with x < y and y−x /∈ N .Then there exists k ∈ N such that y − x > 3k−1. By 23.5, we mayassume that x = (xν)ν/Ψ with |xν −xμ| < k−1 for all μ, ν ∈ N and thaty satisfies an analogous condition. Choose μ such that yμ − xμ > 3k−1

and put c = (xμ + yμ)/2 ∈ Q. Then x + N < c < y + N . �

23.7 Theorem As an ordered field, F is complete.

Proof Let ∅ �= M ⊂ F , and assume that M has an upper bound in F .Then T = { t ∈ Q | M < t} and S = {s ∈ Q | s < T } are not empty,and (S, T ) is a Dedekind cut: in fact, if z ∈ Q and z /∈ S, then t ≤ z forsome t ∈ T , hence M < z and z ∈ T .

There is a least integer t0 ∈ T , and s0 = t0−1 ∈ S. Define inductivelysν ∈ S and tν ∈ T as follows: put mν = (sν + tν)/2 and let

sν+1 =

{mν if mν ∈ S

sν if mν ∈ T, tν+1 =

{tν if mν ∈ S

mν if mν ∈ T.

Then tν − sν = 2−ν . Consider in R the elements s = (sν)ν/Ψ andt = (tν)ν/Ψ, and note that by construction s < T and S < t. Fromtν − sν = 2−ν it follows that t − s ∈ N . Thus, s and t represent thesame element r ∈ F , and S ≤ r ≤ T . The fact that Q is strongly densein F implies that r is a least upper bound of M in F : if r < x for somex ∈ M , there exists t ∈ Q with r < t < x, but then t ∈ T and x < t.This contradiction shows M ≤ r. If M ≤ u < r, then M < t < r forsome t ∈ Q. This implies t ∈ T and r ≤ t, again a contradiction. �

Remark Suppose that R is given. By mapping each rational sequencec = (cν)ν which converges to r ∈ R to the element c + N , the identitymapping of Q extends to an isomorphism R ∼= F .

Page 179: The Classical Fields

162 Non-standard numbers

There is an analogous construction starting with the order completereal field R instead of Q. The finite elements in the ultrapower ∗R = RΨ

also form a valuation ring

R ={

x ∈ ∗R∣∣ ∃n∈N |x| < n

},

and its maximal ideal n = R � R× consists of the infinitesimals: n ={x ∈ R | |x| " 1}.23.8 Proposition The canonical injection κ : R ↪→ ∗R (via constant

sequences) induces an isomorphism R ∼= R/n, and R = R + n.

Proof By definition, R = Rκ is cofinal in R and R∩n = {0}. Therefore,κ induces an inclusion of R into the ordered field R/n. Assume thatthere is an element a ∈ R such that a + n /∈ R. Since R is complete andcofinal in R, the set { t ∈ R | t < a} has a least upper bound s ∈ R.This implies s− ε < a < s + ε for all positive ε in R, hence a− s ∈ n incontradiction to the choice of a. �

23.9 Terminology By 23.8, each finite element b belongs to a uniqueresidue class r + n or monad with r = ◦b ∈ R; the element ◦b is knownas the standard part of b. The equivalence relation x ≈ y � ◦x = ◦y ⇔y − x ∈ n will play an important role.

Exercises(1) Show directly that R/N is an Archimedean ordered field.

(2) A Euclidean field is an ordered field in which each positive element is asquare, and the Euclidean closure E of Q is the smallest Euclidean subfield ofR; see 12.8. Compare ∗E and the Euclidean closure K of ∗Q in ∗R.

(3) A Pythagorean field is a field in which the sum of two squares is againa square. By Exercise 5 of Section 12, the Pythagorean closure P of Q isproperly contained in E. What can be said about ∗P ?

24 Non-standard reals

In several respects, the field R is simpler than Q. A similar observa-tion can be made with regard to ultrapowers of R and Q. Usually, anultrapower of R is denoted by ∗R, the specific ultrafilter used in theconstruction is not explicitly referred to. This is partially justified byTheorem 24.26 below. The properties of the algebraical and topologicalstructures underlying ∗R will be discussed according to the scheme of

Page 180: The Classical Fields

24 Non-standard reals 163

Chapter 1. Algebraically, R and ∗R share many features, the topologyof ∗R, however, is not so nice.

24.1 Fact ∗R is an ordered field of cardinality card ∗R = card R = ℵ.This follows immediately from 21.5 and R ↪→ ∗R = RN/Ψ without

using Theorem 21.9. �

24.2 Addition The additive group of ∗R is a rational vector space of

dimension ℵ, hence it is isomorphic to R+.

Remark By 22.4 we have ∗Q+ ∼= ∗R+, but ∗Q is a proper subfield of∗R because R < ∗R and 2 is not a square in ∗Q.

24.3 Exponentiation By 21.10, there is a function exp : ∗R→ ∗R, andexp is an isomorphism of ∗R+ onto ∗R×

pos (which preserves the ordering).Its inverse is the extended logarithm. The assertions are immediateconsequences of the definitions. Together with 24.2 we get

24.4 Corollary The multiplicative group ∗R× is isomorphic to R×.

In spite of 24.2 and 24.4, however, the fields ∗R and R are not isomor-phic (since ∗R has a non-Archimedean ordering).

24.5 Ordering In the natural (non-Archimedean) ordering of ∗R (theone inherited from R), the positive elements are exactly the squares, i.e.,∗R is a Euclidean field. Hence this is the only ordering relation on ∗Rwhich satisfies the monotonicity laws.

Proof According to 21.10, the absolute value on R extends to ∗R asfollows: if x = (xν)ν/Ψ, then |x| = (|xν |)ν/Ψ. Obviously, the latterterm is a square. The uniqueness of the ordering is also a consequenceof the next theorem. �

Remember from 12.15 that a field F of characteristic 0 is real closedif a quadratic extension of F is algebraically closed.

24.6 Theorem ∗R is a real closed field, and ∗C is isomorphic to C.

Proof The field ∗C is algebraically closed by 21.7. Its elements can bewritten in the form (aν + ibν)ν/Ψ = a + ib. Hence ∗C = ∗R + ∗Ri. Thefirst result follows also from the fact that being a real closed field is afirst-order property; compare Section 21. By 64.20 and 64.21, any alge-braically closed field of characteristic 0 and cardinality ℵ is isomorphicto C. �

Page 181: The Classical Fields

164 Non-standard numbers

Ordering and topology

Remember notation and results discussed at the end of Section 23. Inparticular, R = R + n is a valuation ring in ∗R. Associated with R isthe valuation v : ∗R× → Γ+ ∼= ∗R×/R×. The relation v(x) ≤ v(y) �yR ⊆ xR turns the value group Γ into an ordered group; compare 56.5.For convenience, put v(0) =∞ > Γ.

The sets Vγ = {x ∈ ∗R | v(x) > γ } with γ ∈ Γ can be taken asneighbourhoods of 0, as in 13.2. On the other hand, the open intervalsform a basis of a natural topology τ on ∗R. It is easy to show thatthe topology defined by the Vγ coincides with the order topology τ (seeExercise 7 below).

In the sequel, some properties of the topological space T := (∗R, τ) willbe described. Like any ordered space, T is normal; compare Birkhoff

1948 Chapter III Theorem 6, Bourbaki 1966 Chapter IX §4 Exercise 5or Lutzer 1980. Whereas each homeomorphism of R preserves or re-verses the ordering, the topology of T does in no way determine theordering of ∗R; see 24.14. The ordering of ∗R is discussed in Di Nasso–

Forti 2002.

24.7 Proposition: Homogeneity The space T is doubly homoge-

neous, i.e., the homeomorphism group is doubly transitive on T; see also

33.13.

Proof Since ∗R is a topological field by 13.2(a), each non-constant mapx �→ ax + b is a homeomorphism. �

24.8 The space T is totally disconnected: the only non-empty connected

sets are the points.

Proof If some connected set would contain more than one point, thenany two points would lie in a connected set by 24.7, and T would beconnected (Dugundji 1966 Theorem V.1.5), but R and each of its cosetsis open, hence also closed. �

24.9 The (canonically embedded) subspace R is discrete, and R is closed

in T.

Proof The infinitesimal ideal n is open and (r + n) ∩ R = {r} for eachr ∈ R. Each monad without its standard point is open and so is thecomplement of R. The union of these sets is the complement of R. �

24.10 T is a union of uncountably many pairwise disjoint open sets.

Hence T is not separable and not a Lindelof space.

Page 182: The Classical Fields

24 Non-standard reals 165

Proof In fact, T is the union of all cosets of the open set n. �

24.11 Proposition Any countable subset of ∗R has an upper bound

in (∗N, <).

Proof Consider a sequence xκ = (xκν)ν/Ψ ∈ ∗R, choose zν ∈ N suchthat xκν ≤ zν for κ ≤ ν, and put z = (zν)ν/Ψ. Then xκ ≤ z for each κ:in fact, zν ≥ xκν for all ν ≥ κ. �

24.12 Corollary The space T is not first countable, i.e., there are no

countable neighbourhood bases. In particular, T is not metrizable. Each

convergent sequence in T is finally constant.

Proof Suppose that the open intervals ]c− aκ, c + aκ[ with κ ∈ N forma neighbourhood base at c ∈ T. If b = d−1 > a−1

κ for all κ ∈ N,then the interval ]a − d, a + d[ is properly contained in all the givenneighbourhoods.

If a sequence of elements cn ∈ T � {0} converges to 0, then the |cn|−1

are bounded by some s ∈ T and |cn| ≥ s−1 > 0, a contradiction. �

24.13 Proposition T is not locally compact, not even locally pseudo-

compact.

Proof A space X is called pseudo-compact, if each continuous functionf : X → R has a bounded and then even a compact image. If X islocally pseudo-compact, then each point has a pseudo-compact closedneighbourhood (since f(A) ⊆ f(A) by continuity). Because each con-tinuous function f on a closed interval in T has a continuous extensionf : T → R with the same range as f , local pseudo-compactness of T

would imply the existence of a pseudo-compact interval [a, b] and theneach closed interval would be pseudo-compact by 24.7. Consider nowan infinite element c such that R ⊂ [−c, c]. Since T is normal, eachcontinuous function on the closed subspace R can be extended to [−c, c],but continuous functions on R need not be bounded. �

24.14 Homeomorphisms Each permutation of the subset R of T is

induced by a homeomorphism of T.

Proof The open and closed monads can be permuted arbitrarily byhomeomorphisms of R, and (r + n) ∩ R = {r}. �

24.15 Proposition The value group Γ is a non-Archimedean group.

Page 183: The Classical Fields

166 Non-standard numbers

The proof follows easily from 24.11: let R < x−1 and choose z suchthat x−n < z−1 for all n ∈ N. Then 0 < x−nz < 1, z ∈ xnR, andn · v(x) = v(xn) ≤ v(z). �

Remark In contrast with 24.5, the topology of an infinite field is neveruniquely determined: a field of cardinality c admits 22c

different fieldtopologies (Kiltinen 1973). Only additional assumptions restrict thenumber of topologies as well as the structure of the field. The mostprominent among such assumptions is local compactness. In fact, up tofinite extensions, all locally compact topological fields can be describedquite explicitly; see 58.7 or Weil 1967 Chapter I. The following resultshows that the field ∗R has a different structure than any locally compactfield.

24.16 There is no locally compact field topology on ∗R.

Proof By 58.8, every formally real, locally compact field is isomorphicto R. �

The Sorgenfrey topology on R has all half-open intervals ]a, b] as abasis; it is a source of interesting examples; see 5.73. The Sorgenfreytopology σ on ∗R is defined analogously. It is strictly finer (stronger)than τ , since ]a, b[=

⋃x∈]a,b[ ]a, x] ∈ σ but ]a, b] /∈ τ . The following

statements, for which no proofs are given, can easily be verified.

24.17 Addition in (∗R, +, σ) is continuous, but not the map x �→ −x.

24.18 Homogeneity The homeomorphism group of (∗R, σ) is transi-

tive on ordered pairs.

24.19 The monads are open and closed in (∗R, σ). Hence assertions

24.8–12 are true for (∗R, σ) as well as for T.

η1-fields

According to Hausdorff 1914 p. 180/1, a chain K< is called an η1-set,if for any two subsets A,B of K such that A < B and A ∪B is at mostcountable, there is an element c ∈ K with A < c < B. (Here, A or B maybe empty.) Chains with this property, and η1-fields in particular, havebeen studied a good deal; compare Prieß-Crampe 1983 Chapter IVand Dales–Woodin 1996. Noteworthy are the uniqueness results onη1-structures of cardinality ℵ1. The following result provides examples.

Page 184: The Classical Fields

24 Non-standard reals 167

24.20 Theorem If Ψ is an ultrafilter over N and if F is any subfield

of R, then (F Ψ, <) is an η1-field.

This can be proved by a slight modification of the argument in 24.11:write A = {aκ}κ with aκ = (aκν)ν/Ψ and analogously B = {bλ}λ. Theassumption A < B implies that ∀κ,λ≤μ aκν < bλν for almost all ν. Aneasy induction over μ shows that we may assume aκν < bλν for all κ, λ

and ν. Let maxκ≤ν aκν < cν < minλ≤ν bλν . Then c = (cν)ν/Ψ satisfiesthe assertion. �

24.21 Cardinality Any η1-chain K< has cardinality at least 2ℵ0 .

Proof Since K is dense in itself, Q can be embedded as a subchaininto K; see the remark following 3.4. Each of the ℵ irrational numberss ∈ R � Q determines a Dedekind cut (As, Bs) in Q. By assumption,there is an element cs ∈ K with As < cs < Bs. If s < t, then At ∩ Bs

contains some element d, and cs < d < ct. Hence (s �→ cs) : R � Q→ K

is an injective map. �

24.22 A strongly dense subchain D of an η1-chain K is itself an η1-

chain.

Proof If A and B are countable subsets of D with A < B, then repeatedapplication of the definition shows that there is more than one elementand hence an interval J in K between A and B. Since D is dense, J

contains a point of D. �

The following uniqueness theorem has already been proved by Haus-

dorff 1914 p. 181; see also Prieß-Crampe 1983 Chapter IV §2.

24.23 Any two η1-chains K and L of cardinality ℵ1 are isomorphic as

chains.

This result will not be used in the sequel. It can be proved by transfi-nite induction, mapping alternately an element of K to L and vice versa.Details are left to the reader. 24.21 shows that η1-chains of cardinalityℵ1 exist only if the continuum hypothesis 2ℵ0 = ℵ1 is assumed, i.e., if2ℵ0 is the smallest uncountable cardinal; compare 61.17.

Combining the idea of the proof with Artin–Schreier Theory (see Sec-tion 12 or Jacobson 1989 Chapter 11), one obtains (Erdos et al. 1955):

24.24 Uniqueness Theorem Any two real closed η1-fields F and F ′

of cardinality ℵ1 are isomorphic as fields and hence as ordered fields.

In the proof of the theorem the following is needed.

Page 185: The Classical Fields

168 Non-standard numbers

24.25 Lemma Any η1-field F contains a strongly dense set T of alge-

braically independent elements over Q.

Proof by transfinite induction. Because of 24.21, the field F has cardi-nality ℵα ≥ 2ℵ0 . The set of all open intervals in F can be written as{Jκ | κ < ωα }, where ωα is the smallest ordinal such that cardωα = ℵα.Consider an ordinal number σ < ωα, and assume that algebraically inde-pendent elements tρ ∈ Jρ have been found for ρ < σ. Let Eσ be the fieldof all elements in F which are algebraic over Q(tρ)ρ<σ, and note thatcardEσ = ℵ0 cardσ < ℵα. By homogeneity, card Jσ = card[−1, 1] = ℵα.Hence there is an element tσ ∈ Jσ � Eσ, and {tρ}ρ≤σ is algebraicallyindependent by construction. �

Proof of Theorem 24.24. By 24.25, there exist strongly dense maximalalgebraically independent sets (transcendency bases) T ⊂ F and T ′ ⊂ F ′

of cardinality ℵ1. These bases shall be considered as well-ordered setsof order type ω1 (with respect to an ordering different from the oneinherited from F or F ′ respectively). Denote the real closure of A inF by [A]F , and similarly for F ′. Let σ < ω1 and assume that elementstρ ∈ T have been selected for ρ < σ. Put Eσ = [Q(tρ)ρ<σ]F and notethat Eσ is countable because σ has only countably many predecessors.Suppose that we have an isomorphism ϕσ : Eσ

∼= E′σ which maps tρ to

t′ρ for ρ < σ. Since Eσ is real closed, ϕσ preserves the ordering, and sinceE0 is the real closure of Q, the isomorphism ϕ0 exists and is uniquelydetermined.

We will distinguish between even and odd ordinal numbers as follows:define λ to be even if λ is a limit ordinal, i.e., if λ �= ξ + 1 for eachordinal ξ, and let ξ + 1 be odd (even) if ξ is even (odd). It follows from61.5 that each ordinal is of the form λ + μ with λ a limit ordinal andμ ∈ ω. Hence each ordinal is either even or odd. The map ξ �→ ξ + 1is an order-preserving bijection of the even ordinals onto the odd ones,and { ξ ∈ ω1 | ξ is even} is order-isomorphic to ω1.

If σ is even, let tσ be the first element in T � {tρ}ρ<σ. Because T isalgebraically independent, tσ /∈ Eσ, and Eσ = A ∪ B with A < tσ < B.Applying the map ϕσ (which preserves the ordering) to A and B, weobtain sets A′ and B′ such that A′ < B′ and A′∪B′ = E′

σ is a countablesubset of E′

σ ∪ T ′. This set is dense in the η1-field F ′. Hence E′σ ∪ T ′ is

an η1-chain by 24.22, and there exist elements t′ ∈ T ′ with A′ < t′ < B′.The first of these elements (in the well-ordering of T ′) is taken as t′σ.By construction, a < s := tσ ⇔ aϕσ < s′ := t′σ for all a ∈ Eσ. Thiscondition is necessary and sufficient for ϕσ to have an order-preserving

Page 186: The Classical Fields

24 Non-standard reals 169

extension ϕσ : Eσ(tσ) ∼= E′σ(t′σ) mapping tσ onto t′σ. In fact, since Eσ

is real closed, each monic polynomial f(s) ∈ Eσ[s] splits uniquely into aproduct of linear factors s−cν and some quadratic factors (s−pμ)2 +qμ

with qμ > 0. In the ordering induced by F on Eσ(s) the element f(s) ispositive if, and only if, the number of linear factors with s < cν is even,and this condition is preserved by ϕσ. The isomorphism Eσ[s] ∼= E′

σ[s′]of chains then extends in a natural way to the fields of fractions. Finally,ϕσ can be extended to an isomorphism ϕσ+1 of the real closures; see12.17. This concludes the inductive step in the even case.

If σ is odd, proceed analogously with the roles of F and F ′ inter-changed and all maps reversed. Because the first remaining elementsof T or T ′ are selected alternately, both bases will be exhausted by theinduction. Thus,

⋃σ∈ω1

Eσ = [Q(T )]F = F , and the common extensionof the ϕσ yields an isomorphism ϕ : F ∼= F ′. �

Let Ralg = Q� ∩ R denote the real closed field of all real algebraicnumbers.

24.26 Corollary If the continuum hypothesis 2ℵ0 = ℵ1 is assumed,

and if Φ and Ψ are arbitrary free ultrafilters over N, then RΦalg∼= RΨ as

ordered fields. In particular, any two ultrapowers of R over a countable

index set are isomorphic.

Proof By 24.20 and the argument of 24.6, both RΦalg and RΨ are real

closed η1-fields of cardinality 2ℵ0 ; compare also Hatcher–Laflamme

1983. �

In fact, the continuum hypothesis implies that every ultrapower of Ras in 24.26 is isomorphic to the field R((Γ))1 of Hahn power series withcountable support (as defined in 64.25), where Γ = R((S))1 is the Hahngroup over Sierpinski’s ordered set S; see Dales–Woodin 1996 4.30p. 88.

A slight modification at the beginning of the proof of 24.24 showsthat the field ∗R has at least ℵ1 automorphisms, in contrast with thefact that R is rigid.

24.27 Corollary If 2ℵ0 = ℵ1, then Aut ∗R is uncountable.

The proof will be formulated for the field ∗Ralg. If 1 " t ∈ ∗Ralg,then t is transcendental over Ralg because tn > f(t) for any polynomialf over Ralg of degree < n. Fix an element t0 # 1. For each of the ℵ1

different elements t′ # 1 there is an isomorphism τ : Ralg(t0) ∼= Ralg(t′)

Page 187: The Classical Fields

170 Non-standard numbers

which preserves the ordering and maps t0 to t′. Exactly as before, τ canbe extended to an automorphism ϕ of ∗Ralg. �

Regarding other algebraic structures, we mention the following by-product of results in model theory.

24.28 Uniqueness If the continuum hypothesis 2ℵ0 = ℵ1 is assumed,

and if S is an algebraic structure of cardinality at most ℵ = ℵ1, then

any two ultrapowers of S over a countable index set are isomorphic.

A proof can be found in Potthoff 1981 Corollary 23.6 or Chang–

Keisler 1990 6.1.9.

Exercises(1) The ultrapower RΨ can also be written in the form RN/M where M is amaximal ideal in the ring RN. Express M by Ψ and vice versa.

(2) Show that exp : ∗C+ → ∗C× is surjective, and determine the kernel.

(3) Define ba in ∗R for an arbitrary positive element b.

(4) The field C has an involutory automorphism fixing the elements of a sub-field F ∼= ∗R.

(5) As a chain, ∗R is isomorphic to each of its open intervals, but not to thechain R.

(6) As a real vector space, ∗R does not have a countable basis.

(7) Show that the valuation topology of ∗R has a basis consisting of the openintervals.

25 Continuity and convergence

By embedding the real numbers R into R = R + n instead of identifyingR with R/n, arguments involving infinitely small entities, which wereused before the invention of ‘ε and δ’, could be made precise. Of course,R is only a ring; if the numbers are required to form a field, one hasto admit infinitely large elements also and work in the field of fractions∗R = R(R � {0})−1.

Several notions of elementary analysis can be dealt with very easilywithin this framework; see for example 25.1 and 25.7 for the relation ofsimple and uniform continuity. Our discussion will be restricted to suchparts of non-standard analysis which do not presuppose some routine informal logic and model theory.

Extensions of functions as defined in 21.10 will be used continually,often in the following form.

Page 188: The Classical Fields

25 Continuity and convergence 171

25.0 Definition Let Ψ be a given ultrafilter over N. Consider any maph : D → R with D ⊆ R. Write ∗D = DΨ = {c = c/Ψ | c = (cν)ν ∈ DN }and put ∗h(c) = (h(cν))ν/Ψ = h(c). Then ∗h is a well-defined map suchthat the diagram

D R

∗D ∗R

�h

� ��

∗h

is commutative. If there is no danger of confusion, ∗h will also be denotedby h. Recall that we write x ≈ y if x− y is infinitely small.

25.1 Continuity A function f : R→ R is continuous at a ∈ R if, and

only if,

∀x∈∗R

(x ≈ a⇒ ∗f(x) ≈ ∗f(a)

).

Consequently, the composition of continuous functions is again contin-

uous.

Proof (a) Put x = (xν)ν/Ψ. Then x ≈ a if, and only if, for each m ∈ Nthe condition |xν − a| < m−1 holds for almost all ν. If f is continuousat a in the usual sense, then for each n ∈ N there is some m such that|xν − a| < m−1 implies |f(xν)− f(a)| < n−1. Therefore, ∗f(x) ≈ ∗f(a).

(b) Assume that f is not continuous at a. Then there is some n ∈ Nsuch that for each ν ∈ N there exists some xν with |xν − a| < ν−1 and|f(xν)− f(a)| ≥ n−1. In particular, x ≈ a, but |∗f(x)− ∗f(a)| /∈ n. �

25.2 Convergent sequences A sequence s : N → R converges to

t ∈ R if, and only if, ∗s(m) ≈ t for each infinite m ∈ ∗N.

Proof Suppose that for each n ∈ N there is some μ ∈ N such that|s(ν) − t| < n−1 for all ν > μ. Consider m = (mν)ν/Ψ ∈ ∗N � N.Then {ν ∈ N | mν ≤ μ} /∈ Ψ and mν > μ for almost all ν, hence|∗s(m)− t| < n−1.

If the sν do not converge to t, then there is an element n ∈ N and foreach ν ∈ N some mν > ν such that |s(mν) − t| ≥ n−1. It follows thatm = (mν)ν/Ψ ∈ ∗N � N and that ∗s(m) �≈ t. �

25.3 Example We show that n√

n converges to 1: put s(n) = n√

n− 1.Then (s(n) + 1)n = n and, by the binomial expansion,

(n2

)s(n)2 < n− 1

or |s(n)| <√

2n−1. Hence ∗s(m) ≈ 0 for each m ∈ ∗N � N.

Page 189: The Classical Fields

172 Non-standard numbers

25.4 Corollary A real function f is continuous at a ∈ R if, and only

if, f(a + sν) converges to f(a) whenever the sν converge to 0.

Proof We may assume that a = 0 = f(a), and we write sm = (smν)ν/Ψ

for m = (mν)ν/Ψ. If f is continuous at 0, then sm ≈ 0 implies f(sm) ≈ 0for each infinite m ∈ ∗N. On the other hand, any element in n can bewritten in the form sm with m = (ν)ν/Ψ. �

25.5 Monotone convergence If s : N→ R is an increasing sequence

and if ∗s(m) ≈ t for some m ∈ ∗N � N and t ∈ R, then s converges to t.

Proof For each n ∈ N, the assumption on m = (mν)ν/Ψ implies mν ≥ n

for almost all ν, hence s(n) ≤ ∗s(m) and s(n) ≤ t. If r < t and r ∈ R,then r < s(mν) for almost all ν. Since s is increasing, r < s(n) ≤ t forall but finitely many n ∈ N. �

25.6 Cauchy criterion A sequence s : N → R converges if, and only

if,

m,n ∈ ∗N � N =⇒ ∗s(m) ≈ ∗s(n) .

Proof If s converges to t ∈ R in the ordinary sense, then ∗s(m) ≈ ∗s(n)by 25.2. Suppose now that s is not a Cauchy sequence in the ordinarysense. Then for some k ∈ N, there are arbitrarily large mν and nν ∈ Nwith s(nν)− s(mν) > k−1, and ∗s(m)− ∗s(n) /∈ n. �

25.7 Uniform continuity A real function f is uniformly continuous

on R if, and only if,

∀x,y∈∗R

(x ≈ y ⇒ ∗f(x) ≈ ∗f(y)

).

Proof The arguments are quite similar to those in 25.1. Assume that f

is uniformly continuous. Then for each n ∈ N there is some m such that|yν − xν | < m−1 implies |f(yν) − f(xν)| < n−1, and the conclusion of25.7 is true.

If f is not uniformly continuous, then there is some n ∈ N such thatfor each ν ∈ N there are elements xν and yν with |yν − xν | < ν−1 and|f(yν)−f(xν)| ≥ n−1. Now x = (xν)ν/Ψ and y = (yν)ν/Ψ satisfy x ≈ y

but not ∗f(x) ≈ ∗f(y). �

Note If one of the variables in 25.7 is restricted to R, then, accordingto 25.1, the condition characterizes ordinary continuity of f .

Page 190: The Classical Fields

26 Topology of the real numbers in non-standard terms 173

25.8 Corollary Any continuous function f : [a, b] → R is even uni-

formly continuous.

Proof For real t ≤ a or t ≥ b let f be constant. Consider x, y ∈ ∗Rwith x ≈ y. If x, y ∈ [a, b], then x and y are finite, and x ≈ y implies◦x = ◦y = c ∈ R, hence ∗f(x) ≈ f(c) ≈ ∗f(y) by 25.1. If x < a ≤ y

and x ≈ y, then y ≈ a, as n is convex, and ∗f(x) = f(a) ≈ ∗f(y) bycontinuity. �

25.9 Ring of continuous functions Sum and product of continuous

functions are continuous. If f is continuous at a and f(a) �= 0, then

g = 1/f is also continuous at a.

Proof The first part is an immediate consequence of 25.1 and the factthat n is an ideal in R.

For the second part, note that ∗f(x) ≈ f(a) implies ∗f(x) /∈ n and∗g(x) ∈ R. Hence ∗g(x)− g(a) =

(f(a)− ∗f(x)

)∗g(x)g(a) ∈ n. �

ExercisesThese exercises should be dealt with in the setting of this section.

(1) Which sequences s : N → {0, 1} are convergent?

(2) Show that (x �→ x2) : R → R is not uniformly continuous.

(3) Let f : R → R be any function. If ∗f is constant on each monad, then f isconstant.

26 Topology of the real numbers in non-standard terms

A few basic properties of the topological space R (with the usual topo-logy) will be dealt with via the embedding R ↪→ ∗R. Remember thatthe topology τ defined in Section 24 induces on R the discrete topology(24.9), so that R is not to be considered as subspace of ∗R.

If A ⊂ R, write ∗A = AN/Ψ and note that ∗(R � A) = ∗R � ∗A (by21.2). For other notation see the last part of Section 23.

26.1 Open sets A set A ⊂ R is open in R if, and only if, A + n ⊆ ∗A.

Proof If a ∈ A and A is open in R, then (a− r, a + r)∩R ⊆ A for somepositive real number r, and a + n ⊆ (a− r, a + r) ⊆ ∗A.

If A is not open, then there is an element a ∈ A which does not belongto the interior of A, and some sequence of numbers cν ∈ R�A convergesto a. It follows that c = (cν)ν/Ψ /∈ ∗A and c− a ∈ n. �

Page 191: The Classical Fields

174 Non-standard numbers

26.2 Closure The closure of a set A in R is given by A = (∗A+n)∩R =◦(∗A ∩R).

Proof If t ∈ R and t ∈ A, then there is a sequence of numbers aν ∈ A

which converges to t, and a = (aν)ν/Ψ ∈ ∗A ∩ (t + n). Conversely,t ∈ R � A = O implies t + n ⊆ ∗O ⊆ ∗(R � A) by 26.1. �

26.3 Continuity If f : R → R is continuous, then each open set

U ⊆ R has an open pre-image O = f←(U), moreover, f(A) ⊆ f(A) for

each A ⊆ R. (In fact, each of these equivalent conditions characterizescontinuity of f .)

Proof By continuity, x ≈ a ∈ O implies ∗f(x) ≈ f(a) ∈ U . Since U isopen, 26.1 gives ∗f(x) ∈ ∗U , and this equivalent with x ∈ ∗O. Moreover,if a ≈ x ∈ ∗A, then f(a) ≈ ∗f(x) ∈ ∗(f(A)) and hence f(a) ∈ f(A). �

26.4 Boundedness A set A ⊆ R is bounded if, and only if, ∗A ⊆ R.

Proof By definition, ∗A contains infinite elements if, and only if, A isunbounded. �

26.5 Compactness A is compact if, and only if, for each x ∈ ∗A there

is an element a ∈ A with x ≈ a, in short, if ◦(∗A) = A.

This is proved by combining 26.2 and 26.4. �

26.6 Corollary Compactness is preserved by continuous maps.

For a proof, note that f(∗A) = ∗(f(A)) and apply 25.1. �

26.7 Intersections If Aκ+1 ⊂ Aκ for κ ∈ N, and if each Aκ is compact

and non-empty, then D =⋂

κ∈N Aκ �= ∅.Proof Choose aν ∈ Aν and put a = (aν)ν/Ψ. Then a ∈ ∗Aκ for each κ

(since aν ∈ Aκ for ν ≥ κ), and the standard part d of a is in Aκ by 26.5.Hence d ∈ D. �

26.8 Accumulation points A real number t is an accumulation point

of A ⊆ R if, and only if, there is an element a ∈ ∗A such that t ≈ a and

t �= a.

Proof According to 26.2, the condition is equivalent with t ∈ A � {t},and this is the standard definition of an accumulation point. �

Page 192: The Classical Fields

27 Differentiation 175

26.9 Corollary (Bolzano–Weierstraß) Each bounded infinite subset

A of R has an accumulation point.

Proof The assumptions imply A ⊂ ∗A ⊆ R, and then the standard partof any element a ∈ ∗A � A is an accumulation point of A. �

Exercises(1) Use 26.1 in order to show that any union of open sets is open and that the

intersection of two open sets is open. Show also that A ∪ B = A ∪ B.

(2) Deduce from results in this chapter that Q is dense in R.

(3) Show that the intersection of countably many dense open subsets of R isdense in R (but there are open neighbourhoods of Q in R of arbitrarily smallLebesgue measure).

(4) A continuous function on a closed interval assumes its maximum.

(5) If C = [a, b] ⊂ R and if f : C → R is continuous, then ∗f assumes itsmaximum at a standard point of ∗C.

27 Differentiation

Hardly any topic is better suited for non-standard methods than differen-tiation. Reasoning in the extension ∗R of R, several heuristic argumentsdealing with infinitely small numbers can be put on a sound foundation.

For the sake of simplicity, only real-valued functions defined on anopen and connected domain D ⊆ R will be considered.

27.1 Characterization A function f : D → R is differentiable at

a ∈ D in the usual sense with derivative f ′(a) ∈ R if, and only if,(∗f(a + h)− f(a))h−1 ≈ f ′(a) for each h ≈ 0, h �= 0 . (∗)

Proof Assume that f is differentiable at a. This means that for eachn ∈ N there is an m ∈ N such that 0 < |s| < m−1 implies

|(f(a + s)− f(a)) s−1 − f ′(a)| < n−1 .

If 0 �= h = (sν)ν/Ψ ≈ 0, then 0 < |sν | < m−1 for almost all ν ∈ N, andthe assertion follows.

Suppose conversely that f is not differentiable at a or that f ′(a) �=t ∈ R. Then there exists a sequence (sν) which converges to 0 suchthat |(f(a + sν) − f(a))s−1

ν − t| ≥ ε for some positive ε ∈ R. Lettingh = (sν)ν/Ψ, we have h ≈ 0, but

(∗f(a + h)− f(a))h−1 �≈ t. �

Page 193: The Classical Fields

176 Non-standard numbers

27.2 Continuity If f is differentiable at a, then f is also continuous

at a.

Proof From equation (∗) we obtain ∗f(a + h)− f(a) ≈ f ′(a) · h ≈ 0 forh ≈ 0. �

27.3 Corollary: Chain Rule If f and g are differentiable and if the

composition g ◦ f is defined in a neighbourhood of a, then (g ◦ f)′(a) =g′(f(a)) · f ′(a).

Proof Let a �= x ≈ a. Then(∗f(x)− f(a)

)(x− a)−1 ≈ f ′(a) and hence

∗f(x)− f(a) ≈ 0. Therefore(∗(g ◦ f)(x)− (g ◦ f)(a)

)(∗f(x)− f(a))−1 ≈

g′(f(a)) whenever ∗f(x) �= f(a). �

27.4 Products If f and g are differentiable at a, then

(f · g)′(a) = f ′(a) · g(a) + f(a) · g′(a) .

Proof Let x ≈ a and x �= a. Then g(x) ≈ g(a) by 27.2, and the identity(f · g)(x) − (f · g)(a) = (f(x) − f(a))g(x) + f(a)(g(x) − g(a)) togetherwith 27.1 gives the assertion. �

The following kind of Cauchy criterion does not require knowledge ofthe derivative in advance.

27.5 Differentiability A function f : D → R is differentiable at a ∈ D

if, and only if, x, y ≈ a �= x, y implies

(∗f(x)− f(a))(x− a)−1 ≈ (∗f(y)− f(a))(y − a)−1 . (∗∗)

Proof For x ∈ R put q(x) = (f(x) − f(a))(x− a)−1. Assume first that∗q(z) is infinitely large for some z = (zκ)κ/Ψ ≈ a. Then for each n ∈ Nthere are infinitely many κ ∈ N with 0 < |zκ − a| < n−1 and q(zκ) ≥ n.Among the zκ define inductively a sequence of elements xν convergingto a such that q(xν+1) ≥ q(xν) + 1, and let yν = xν+1. Then x, y ≈ a

and ∗q(y) ≥ ∗q(x) + 1 �≈ ∗q(x).Hence (∗∗) implies that ∗q(z) ∈ R. Denoting the standard part of

∗q(z) by f ′(a), differentiability follows from 27.1. �

27.6 Continuous derivative A function f : D → R is continuously

differentiable if, and only if, the following condition holds:

∀a∈D

∃d∈R

∀x,y∈∗R

[x ≈ a ≈ y �= x⇒ (∗f(y)− ∗f(x)

)(y − x)−1 ≈ d

](†)

Page 194: The Classical Fields

28 Planes and fields 177

Proof For simplicity, write q(x, y) = (f(y) − f(x))(y − x)−1, and let x

be the smaller one of the two distinct elements x and y. Represent x inthe usual way as x = (xν)ν/Ψ etc.

Let f be differentiable and x, y ≈ a. By the mean value theorem,we find real numbers zν between xν and yν such that q(xν , yν) = f ′(zν)and hence ∗q(x, y) = ∗f ′(z). The convexity of monads gives z ≈ a, andcontinuity of f ′ shows that ∗f ′(z) ≈ f ′(a); see 25.1.

By 27.1 or (∗∗), condition (†) implies that f is differentiable at eachpoint a ∈ D. Assume that for every ε > 0 there is some ν ∈ N suchthat |q(x, y)−f ′(a)| < ε for any two distinct real numbers x, y in a ν−1-neighbourhood of a; then f ′ is continuous at a. Suppose now that thisis not true. Then for some ε > 0 and for each ν ∈ N there are elementsxν and yν with a− ν−1 < xν < yν < a+ ν−1 and |q(xν , yν)− f ′(a)| ≥ ε.Consider x = (xν)ν/Ψ and y = (yν)ν/Ψ. By construction, x < y andx ≈ y ≈ a, but |q(x, y)− f ′(a)| ≥ ε. This contradicts (†). �

27.7 Examples Let p(x) = xn. If x ≈ a, then (xn − an)(x − a)−1 =∑nν=1 xn−νaν−1 ≈ n · an−1 = p′(a).Let s(x) = sinx. The power series expansion shows immediately

that limx→0 x−1s(x) = 1. Hence h ≈ 0 implies s(h)h−1 ≈ 1 = s′(0).Moreover, cos h ≈ 1 and (1 − cos h)h−1 = s(h)2h−1(1 + cos h)−1 ≈ 0.Therefore, (s(x + h) − s(x))h−1 = (cos x sin h − sinx(1 − cos h))h−1 ≈cos x = s′(x).

Exercises(1) Show that f ′(a) = 0 if f is differentiable and has a local minimum at a.

(2) If f is differentiable, f(a) = 0, and g = 1/f , then g′(a) = −f ′(a)/f(a)2.

28 Planes and fields

Any (commutative) field F coordinatizes a Pappian affine plane withpoint space F 2 such that the lines are described by linear equations.This correspondence between fields and planes is compatible with theultraproduct construction. Thus, all results in this chapter have a geo-metric interpretation.

28.1 A projective plane P consists of a point set P and a set L of linesL ⊂ P such that any two distinct points belong to a unique line, and,dually, any two lines intersect in a point. By deleting from P a line andall of its points, one obtains an affine plane.

Page 195: The Classical Fields

178 Non-standard numbers

If the well known Pappos theorem holds, then we speak of a Pap-pian plane. Precisely the Pappian affine planes can be represented by a(commutative) field in the way indicated above.

The lines of an affine or projective plane D = (P,L) are determinedby the collinearity relation λ on P 3 which holds exactly if three pointsbelong to the same line. Using the notion of collinearity, ultraproductsof planes can conveniently be defined as follows.

28.2 Ultraproducts of planes Given affine or projective planes Dν =(Pν , λν) and a free ultrafilter Ψ over N, the relation λ holds for threepoints pκ = (pκν)ν/Ψ in the ultraproduct of the Pν if λν(p1ν , p2ν , p3ν)holds for almost all ν ∈ N. Ultraproducts of affine or projective planesare again affine or projective planes, and the same is true for Pappianplanes, because all first-order properties carry over to ultraproducts.

28.3 If the field F coordinatizes a (Pappian) plane D, then any ultra-

power DΨ is coordinatized by FΨ.

28.4 If Ψ is a free ultrafilter over N, then the ultrapower QΨ of the

projective plane Q over the rational field admits a homomorphism (i.e.,a collinearity preserving mapping) onto the real projective plane.

Proof As in Section 23, let R denote the valuation ring of all finiteelements in QΨ. Each point in QΨ has homogeneous coordinates zκ ∈ R

such that at least one of the zκ is a unit in R×. The canonical projectionR→ R/N ∼= R induces a homomorphism onto the real projective plane(because no coordinate triple is mapped to 0). �

A Euclidean plane is an affine plane over a Euclidean field (see 12.8)with an orthogonality relation ⊥ for lines, orthogonality having the fa-miliar properties. (Remember that lines given by y = ax and y = bx areorthogonal if, and only if, ab = −1.)

28.5 Euclidean planes An ultrapower ∗E = EΨ of the ordinary Eu-

clidean plane E over the real numbers is a Euclidean plane.

28.6 Example Let C be the parabola in E with the equation y = x2,let p = (a, a2) ∈ C, and put ∗C = CN/Ψ. Then ∗C = { (x, x2) | x ∈ ∗R}.We show that the tangent at p to C consists of the standard parts ofthe points on a line through p and some other point q ∈ ∗C infinitelyclose to p. Put q = (u, u2) where u ≈ a. The line connecting p and q

has the equation y = (a + u)(x − a) + a2. If s ≈ x, t ≈ y and s, t ∈ R,then t ≈ 2a(s− a) + a2, hence t = 2a(s− a) + a2.

Page 196: The Classical Fields

3

Rational numbers

This chapter treats the different kinds of structures on the field Q of ra-tional numbers (algebra, order and topology) and various combinationsof them in the same way as it was done in Chapter 1 for the field R ofreal numbers.

We sometimes profit from the fact that Q is embedded in R so thatwe can use results from Chapter 1. In doing so, we take the field R forgranted. Constructions of R from Q are presented in Chapter 2 (via anultrapower of Q) and in Chapter 4 (by completion of Q).

31 The additive group of the rational numbers

Under the usual addition, the rational numbers form a group (Q, +), orbriefly Q+. Being a subgroup of R+, this group has already been studiedto some extent in Section 1, together with the factor group Q+/Z. Wecontinue the investigation of these groups, we characterize them in theclass of all groups, and we study their endomorphism rings.

31.1 Definition A group G is called locally cyclic, if the subgroup〈a1, a2, . . . , an〉 generated by finitely many elements a1, a2, . . . , an of G

is always cyclic, that is, this subgroup may be generated in fact bya single element of G. By induction, this is equivalent to the propertythat the subgroup generated by any two elements is cyclic. In particular,every locally cyclic group is abelian.

31.2 For 0 �= b ∈ Q the subgroup 〈b〉 of Q+ generated by b is isomorphic

to 〈1〉 = Z+, as there is an automorphism of Q+ mapping b to 1.

Indeed, such an automorphism is given by r �→ r/b; it induces anisomorphism of 〈b〉 onto Z. �

179

Page 197: The Classical Fields

180 Rational numbers

31.3 Characterization theorem The group Q+ is torsion free, divis-

ible and locally cyclic, and is, in fact, the only non-trivial group having

these properties (up to isomorphism).

Proof It is clear that Q+ is torsion free and divisible (for the definitionsof these notions, the reader is referred to 1.2 and 1.7). In 1.6(c), it isshown that Q+ is locally cyclic.

Conversely, let A be any torsion free, divisible and locally cyclic groupwhich is not reduced to the neutral element. Then A is abelian, and thefirst two properties imply that A can be made into a rational vectorspace; see 1.9. As A is locally cyclic, any two elements are containedin a 1-dimensional Q-linear subspace. Thus the dimension of A over Qis 1, and the group A is isomorphic to Q+. �

According to 31.2 and 31.3, the factor group of Q+ by any non-trivialfinitely generated subgroup is isomorphic to the factor group Q+/Z.

31.4 The group Q+/Z is a torsion group, which means that every el-ement has finite order. In fact, it is the torsion subgroup of the torusgroup R+/Z, i.e., the subgroup consisting of all elements of finite or-der. Indeed, the real numbers x having an integer multiple nx ∈ Z areprecisely the rational numbers.

31.5 The group Q+/Z is a divisible, locally cyclic torsion group.

Proof The group Q+/Z is the homomorphic image of Q+ under thecanonical projection Q+ → Q+/Z and therefore inherits the propertiesof being divisible and locally cyclic from Q+; see 31.3. �

In the sequel, we shall find out to what extent the properties of 31.5suffice to characterize Q+/Z. For this, the following description will behelpful.

31.6 Let p be a prime. The p-primary component of an abelian groupis the subgroup consisting of all elements whose order is a power of p.The p-primary component of Q+/Z is the so-called Prufer group (see1.26)

Cp∞ = { mpn + Z | m ∈ Z, n ∈ N} ⊆ Q+/Z .

In 1.28 it has been shown that

Q+/Z =⊕

p∈P Cp∞ ,

where the sum is taken over the set P of all primes, just as every abeliantorsion group is the direct sum of all its primary components; compare

Page 198: The Classical Fields

31 The additive group of the rational numbers 181

Exercise 2 of Section 1. As a consequence, we may infer that each Prufergroup Cp∞ is divisible since Q+/Z is; see also 1.27 for a direct proof ofthis fact.

31.7 Lemma Let A be a divisible group. If A contains a non-trivial

element of finite order and if the prime p divides this order, then A

contains a subgroup isomorphic to Cp∞ .

Proof We use additive notation, although A is not required to be abelian.By assumption, there is an element 0 �= x1 ∈ A such that px1 = 0. SinceA is divisible, we may choose inductively elements xn for 2 ≤ n ∈ N suchthat px2 = x1, px3 = x2, . . . , pxn+1 = xn. Then pn−1xn = x1 �= 0 andpnxn = 0, so that the order of xn is pn. The map

Cp∞ → A : mpn + Z �→ mxn

(for m ∈ Z, n ∈ N) is a well-defined injective homomorphism; indeed,mxn = 0 if, and only if, the order pn of xn divides m, in other words, if,and only if, m/pn ∈ Z. The image of this homomorphism is a subgroupof A isomorphic to Cp∞ . �

31.8 Characterization Theorem Up to isomorphism, the divisible,

locally cyclic torsion groups are precisely the direct sums⊕p∈P′ Cp∞ for arbitrary subsets P′ ⊆ P .

Proof (1) We show that such a direct sum has the asserted properties.The Prufer groups are divisible torsion groups (see 31.6 and 1.27), hencethe same holds for any direct sum of Prufer groups. Moreover, eachdirect sum as above is a subgroup of Q+/Z =

⊕p∈P Cp∞ and therefore

inherits the property of being locally cyclic from Q+/Z.(2) Now let A be an arbitrary divisible, locally cyclic torsion group.

Let P′ be the set of all primes dividing the order of some group element.We claim that the group B =

⊕p∈P′ Cp∞ is isomorphic to A. In order

to prove this, we construct a homomorphism ϕ : B → A and show thatit is bijective.

For each prime p ∈ P′, there exists an injective homomorphism ϕp :Cp∞ → A by 31.7, and we define ϕ by

ϕ((cp)p∈P′) =∑

p∈P′ ϕp(cp) .

Here, we use the notation for elements of direct sums introduced in 1.16,and the fact that A is abelian (31.1). If c = (cp)p∈P′ is distinct from 0,then some component cp is non-zero. Multiplying the image with the

Page 199: The Classical Fields

182 Rational numbers

product r of the orders of all other components cq for q ∈ P′ � {p}, weobtain rϕ(c) = rϕp(cp), and this is distinct from 0 since the order ofϕp(cp) is prime to r.

We have shown that the subgroup ϕ(B) of A is isomorphic to B.Assume that an element a ∈ A � ϕ(B) exists. Then all prime factorsof the order of a belong to P′, hence ϕ(B) contains an element b havingthe same order as a. The subgroup of A generated by a, b is cyclicand finite by assumption, but a finite cyclic group contains at most onesubgroup of any given order, hence 〈a〉 = 〈b〉. This contradiction showsthat ϕ(B) = A.

We remark that in step (2) above, one could also use Exercise 2 ofSection 1. �

31.9 Corollary A group is isomorphic to Q+/Z if, and only if, it is

a divisible, locally cyclic torsion group and if every prime p divides the

order of some group element.

Proof Since Q+/Z =⊕

p∈P Cp∞ , this follows immediately from 31.8;indeed, from the direct sum in 31.8 the set P′ may be retrieved as beingthe set of primes dividing the orders of elements of the group. �

We now turn to endomorphisms of Q+ and Q+/Z. Recall that the setEndA of endomorphisms of an abelian group A is a ring in a naturalway; its addition is obtained from the operation of the group A, andmultiplication is the composition of maps.

31.10 The endomorphism ring of Q+ is End Q+ = {x �→ rx | r ∈ Q};as a ring, it is isomorphic to the field Q.

Remark In 8.28, we have seen that the only other abelian groups whoseendomorphism rings are fields are the cyclic groups of prime order.

Proof For an endomorphism ϕ of Q+ and for m ∈ Z, n ∈ N we havenϕ(m/n) = ϕ(n ·m/n) = ϕ(m) = mϕ(1), hence ϕ(m/n) = mϕ(1)/n,so that ϕ(x) = rx where r = ϕ(1). Conversely, it is immediate that forarbitrary r ∈ Q this is an endomorphism of Q+. It is also clear thatϕ �→ ϕ(1) is a ring isomorphism of End Q+ onto Q. �

31.11 Theorem The endomorphism ring of Q+/Z is isomorphic to the

following subring of the direct product R :=×k∈N Z/k!Z of the factor

rings Z/k!Z, with componentwise addition and multiplication:

End(Q+/Z) ∼= { (mk + k!Z)k ∈ R∣∣ ∀k∈N mk+1 ≡ mk mod k!

}

Page 200: The Classical Fields

31 The additive group of the rational numbers 183

Proof (i) A rational number m/n for m ∈ Z, n ∈ N can be written asm(n− 1)!/n!, hence Q+/Z is the union of the subgroups

C1/k! := (1/k!)Z/Z

for k ∈ N, which form an ascending chain of subgroups.Every endomorphism ϕ of Q+/Z induces an endomorphism ϕk of the

subgroup C1/k! since this subgroup consists precisely of those elementsof Q+/Z whose k!-fold multiple is the neutral element, and this propertyis preserved by endomorphisms.

The endomorphisms ϕk of C1/k! induced by ϕ satisfy

ϕk+1|C1/k! = ϕk . (1)

Conversely, if a sequence of endomorphisms ϕk ∈ EndC1/k! for k ∈ Nsatisfies this compatibility condition (1), then they can be merged intoan endomorphism ϕ of Q+/Z which induces them all.

(ii) Since C1/k! is a cyclic group, there exists, for given m ∈ Z, aunique endomorphism μk,m of C1/k! mapping the generator (k!)−1 + Zto m(k!)−1 + Z; this endomorphism maps every element to its m-foldmultiple. The surjective map

Z→ EndC1/k! : m �→ μk,m (2)

is clearly a ring homomorphism with kernel {m | m(k!)−1 ∈ Z} = k!Z,so that by factorization this ring homomorphism yields a ring isomor-phism

Z/k!Z→ EndC1/k! : m + k!Z �→ μk,m . (3)

(iii) For a sequence (mk + k!Z)k ∈ R, the corresponding sequenceof endomorphisms ϕk := μk,mk

∈ EndC1/k! satisfies the compatibilitycondition (1) if, and only if, mk+1 − mk belongs to the kernel k!Z ofthe homomorphism (2), i.e., if mk+1 ≡ mk mod k!. As explained instep (i), this is equivalent to the existence of a unique endomorphism ofQ+/Z inducing all the endomorphisms μk,mk

. Mapping such sequencesto the resulting endomorphisms of Q+/Z we obtain a homomorphism ofthe ring of all such sequences onto End(Q+/Z) which is induced by theisomorphisms (3) and therefore is an isomorphism itself. �

31.12 Remarks (1) The additive group of End(Q+/Z) can also bedescribed as the group Hom(Q+/Z, T) of homomorphisms to the torusgroup T = R+/Z, since Q+/Z is the torsion subgroup of T (see 31.4) andthus is mapped into itself by every homomorphism to T.

Page 201: The Classical Fields

184 Rational numbers

(2) The situation of 31.11 can be handled very efficiently by usingthe notion of direct and inverse limits of directed systems of homo-morphisms. For an introduction to these, the reader may consult, forexample, Fuchs 1970 Sections 11, 12 p. 53ff, Engler–Prestel 20055.1 or, in a more general context, Jacobson 1989 2.5 p. 70ff.

The ring used in 31.11 to describe End(Q+/Z) is the inverse limitlim←− k

Z/k!Z of the following system of homomorphisms:

Z/k!Z→ Z/j!Z : m + k!Z �→ m + j!Z for j ≤ k ∈ N . (1)

We shall indicate a duality between this and the fact that the groupQ+/Z =

⋃k∈N

1k!Z/Z =

⋃k∈N C1/k! can be considered as the direct limit

lim−→ kC1/k! of the system of inclusion homomorphisms

C1/j! → C1/k! for j ≤ k ∈ N ; (2)

see Fuchs 1970 Section 11 Example 1 p. 56. The system (1) is dualto the system (2) in the sense that it is obtained from (2) by applyingthe functor Hom( , T). Indeed, similarly as in Remark (1) and us-ing arguments from the proof of 31.11, one sees that Hom(C1/k! , T) =EndC1/k!

∼= Z/k!Z. According to a general fact (see Fuchs 1970 Theo-rem 44.2 p. 185), one may conclude that End(Q+/Z) = Hom(Q+/Z, T)coincides with

Hom(lim−→ kC1/k! , T) = lim←− k

Hom(C1/k! , T) ∼= lim←− kZ/k!Z ,

thus proving 31.11 in a more categorical setting.(3) Another description of End(Q+/Z) = Hom(Q+/Z, T) using the

p-adic numbers will be given in 52.10.

31.13 Corollary The ring End(Q+/Z) has the cardinality 2ℵ0 of the

continuum.

Proof Rather than deriving this from the description of End(Q+/Z) in31.11, we follow another line. We describe Q+/Z as the direct sum⊕

p∈P Cp∞ of all Prufer groups; see 1.28. For every subset S ⊆ P thereis an endomorphism of the direct sum mapping the element (ap)p to(a′

p)p where a′p = ap if p ∈ S and a′p = 0 otherwise. Since there are

infinitely many primes (see 32.7ff) we have thus constructed a subset ofEnd(Q+/Z) having the cardinality 2ℵ0 of the power set of P. On theother hand, End(Q+/Z) is contained in the set of all maps of Q+/Z intoitself, and since Q is countable, this set has cardinality ℵℵ0

0 = 2ℵ0 aswell; see 61.15. Thus, our assertion is proved. �

Page 202: The Classical Fields

32 The multiplication of the rational numbers 185

Exercises(1) No proper subgroup of Q+ is isomorphic to Q+.

(2) Show that Q+ has no maximal subgroup.

(3) Are the subgroups of Q+ generated by 2Z and by 3Z isomorphic?

(4) Show that Q+ has 2ℵ0 mutually non-isomorphic subgroups.

32 The multiplication of the rational numbers

The multiplicative group of non-zero rational numbers will be denotedby Q×. The positive rational numbers form a subgroup Q×

pos of Q×.We show that Q×

pos is a free abelian group with the prime numbers asa free set of generators. The rank of this group, that is, the cardinalityof a free set of generators, is ℵ0, since there are infinitely many primes;various proofs of this truly classical result will be presented at the endof this section (see 32.7ff.). Before this, the homomorphisms betweenQ× or Q×

pos and Q+ will be determined. We shall see in particular thatthe additive and the multiplicative group of Q are radically different,contrary to the situation in the field of real numbers (compare Section 2).

32.1 Proposition The multiplicative group Q×pos of positive rational

numbers is the direct sum of the subgroups 〈p〉 = {pm | m ∈ Z} ∼= Zgenerated by the prime numbers p:

Q×pos =

⊕p∈P〈p〉 ∼= Z(P) , and Q× = {1,−1} ⊕Q×

pos∼= C2 ⊕ Z(P) .

Proof Recall from 1.16 that the elements of the direct sum Z(P) of copiesof Z indexed by the set P of primes are the maps α : P → Z with theproperty that the set of primes p such that α(p) �= 0 is finite. For sucha map, the product

∏p∈P pα(p) has only finitely many factors different

from 1 and therefore is a well-defined element of Q×pos. It is easy to verify

that the map

Z(P) → Q×pos : α �→∏p∈P pα(p)

is a homomorphism. It is bijective, since every natural number canbe written as a product of prime powers in a unique way. Under thisisomorphism, the summands of the direct sum Z(P) correspond to thesubgroups 〈p〉 generated by the primes p ∈ P. Thus, Q×

pos is the directsum of these subgroups.

The map {1,−1}⊕Q×pos → Q× : (ε, r) �→ εr clearly is an isomorphism.

Page 203: The Classical Fields

186 Rational numbers

A free abelian group is a direct sum Z(I) of copies of Z, and the rankof Z(I) is the cardinality of I; compare Fuchs 1970 Section 14 p. 72ff.Granting for the moment that there are infinitely many primes (whichwill be proved in 32.7ff.), in other words, that the cardinality of P is ℵ0,the result of 32.1 can thus be expressed as follows.

32.2 The group Q×pos is a free abelian group of rank ℵ0. �

In order to study homomorphisms defined on the group Q×, whichaccording to 32.1 may be represented as a direct sum, we will use thefollowing general facts.

32.3 Homomorphisms of direct sums We consider a family ofabelian groups Ai indexed by the elements of a set I, and a furtherabelian group B; here, these groups shall be written additively. Givena homomorphism ϕi : Ai → B for every index i ∈ I, we may define amapping ∑

i∈I ϕi :⊕

i∈I Ai → B : (ai)i∈I �→∑

i∈I ϕi(ai) ;

indeed, in the sum on the right only finitely many terms are differentfrom 0, since the same holds for the ai, by definition of a direct sum. Itis straightforward to see that this map is a homomorphism. Conversely,every homomorphism ϕ :

⊕i∈I Ai → B is obtained in this manner if

one considers the summand Ai as a subgroup of the direct sum in theobvious way and takes ϕi to be the restriction of ϕ to Ai. Thus, weobtain a bijection

×i∈I Hom(Ai, B)→ Hom(⊕

i∈I Ai, B) : (ϕi)i∈I �→∑

i∈I ϕi ,

and a straightforward verification shows that this bijection is a homo-morphism and, hence, an isomorphism.

32.4 Theorem The additive group of the endomorphism ring End Q×pos

is isomorphic to the direct product (Q×pos)

P of countably many copies of

Q×pos indexed by P:

(End Q×pos, +) ∼= (Q×

pos)P , and (End Q×, +) ∼= C2 × (Q×)P .

In particular, both groups have the cardinality of the continuum.

Remarks Again, we take it for granted that there are infinitely manyprimes.

The given isomorphisms are not suited for describing the multiplica-tion of these endomorphism rings; see Exercise 4.

Page 204: The Classical Fields

32 The multiplication of the rational numbers 187

Proof By 32.1, the group Q× is the direct sum {1,−1}⊕Z(P) of {1,−1}and of copies of Z. Thus by 32.3, End Q× is the direct product

Hom({1,−1}, Q×)× (Hom(Z, Q×))P (1)

of Hom({1,−1}, Q×) and of copies of Hom(Z, Q×).Since the only element of Q× of order 2 is −1, the only homomor-

phisms of {1,−1} to Q× are the constant homomorphism mapping both1 and −1 to 1, and the inclusion homomorphism; thus

Hom({1,−1}, Q×) ∼= C2 .

The additive group Z is generated by 1, therefore every homomorphismϕ : Z→ Q× satisfies ϕ(m) = ϕ(m · 1) = ϕ(1)m, and

Hom(Z, Q×) = {m �→ rm | r ∈ Q � {0}} ∼= Q× ;

recall that addition of homomorphisms is defined by multiplication ofthe values in Q×. Analogously Hom(Z, Q×

pos) ∼= Q×pos. The assertion

on End Q× now follows from (1), and for End Q×pos it is obtained in the

same way by disregarding the summand {1,−1}.The cardinality of both direct products is ℵℵ0

0 = 2ℵ0 , the cardinalityof the continuum (compare 61.15). �

32.5 Theorem Hom(Q×, Q+) ∼= Hom(Q×pos, Q

+) ∼= (Q+)P, in particu-

lar, these groups have the cardinality of the continuum.

Proof The proof is analogous to the preceding arguments: from 32.1 and32.4 we infer that Hom(Q×, Q+) ∼= Hom({1,−1}, Q+)×Hom(Q×

pos, Q+)

and Hom(Q×pos, Q

+) ∼= (Hom(Z, Q+))P. Because the group Q+ is torsionfree, the set Hom({1,−1}, Q+) contains only the trivial homomorphism,and Hom(Z, Q+) = {m �→ mr | r ∈ Q} ∼= Q+. �

In 2.2 we noted that the additive group of the real numbers is iso-morphic to the multiplicative group of positive real numbers (via theexponential function). In this respect, the rational numbers are quite dif-ferent, as can already be seen from the different cardinalities of End Q+

and End Q×pos (31.10 and 32.4). The difference is even more drastic.

32.6 Proposition The only homomorphism of Q+ into Q× is the

constant map x �→ 1.

Proof The idea is to use the divisibility of Q+ on the one hand and toshow that, on the other hand, the only element of Q× having an n-throot for every n ∈ N is 1.

Page 205: The Classical Fields

188 Rational numbers

For a homomorphism ϕ : Q+ → Q×, for 0 �= x ∈ Q and n ∈ N onehas ϕ(x) = ϕ(n · x/n) = ϕ(x/n)n, so that ϕ(x) has an n-th root in Q×

for every n ∈ N.We now show that 1 is the only element of Q× having this property.

Under the isomorphism Q× ∼= C2 × Z(P) of 32.1, an element with thisproperty is mapped to an element of the direct sum which is divisible byevery n ∈ N (the group operation of the direct sum being addition). Nowin the summands C2 and Z, the only element which has this divisibilityproperty clearly is 0, which proves our assertions. �

Existence of infinitely many primes

As announced, we now give several proofs of the following classical result:

32.7 Theorem There are infinitely many primes.

32.8 The proof from Euclid’s Elements This is the oldest proofof which we know. One shows that, given any finite list p1, p2, . . . , pn

of primes, there is a further prime not contained in the list. Let m =∏nk=1 pk. The number m + 1 has a prime factor p, and p cannot be one

of the primes pk (k = 1, . . . , n) of the list, else p would be a divisor of m

and hence also of 1, which is impossible. �

32.9 Proof by the number of non-quadratic factors For a naturalnumber n, we consider its unique decomposition into prime powers, andfor every prime p appearing with an odd power we split off a factor p.In this way, we may represent n uniquely in the form n = fn · s2

n, wherefn, sn ∈ N and fn is not divisible by any square, in other words, everyprime is contained at most once as a factor in fn. The numbers s2

n andfn will be called the quadratic and the non-quadratic factor of n.

For a natural number m ≤ n, we obviously have sm ≤√

m ≤ √n, sothat there are at most

√n possible different quadratic factors of numbers

≤ n. Hence, for these numbers there must be at least√

n different non-quadratic factors fm. Let π(n) be the number of primes ≤ n. We obtainthat 2π(n) ≥ √n, since fm with m ≤ n contains each prime ≤ n at mostonce. Therefore 4π(n) ≥ n and π(n) ≥ lnn/ ln 4, in particular π(4k) ≥ k.Hence, there exist infinitely many primes. �

The following three proofs use sequences of special numbers.

32.10 Proof by the Mersenne numbers These are the numbers ofthe form 2p − 1, where p is a prime number; see 2.4. Let q be a prime

Page 206: The Classical Fields

32 The multiplication of the rational numbers 189

factor of 2p − 1; we show that p divides q− 1. This then proves that forevery prime number p there is a greater prime number q, so that theremust be infinitely many of them.

We consider the multiplicative group F×q = Fq � {0} of the prime

field Fq with q elements (in other words, of the factor ring Z/qZ). Themultiplicative group has q − 1 elements. As q divides 2p − 1, so that2p ∈ 1 + qZ, we infer that p is the order of the element 2 = 1 + 1 ∈ F×

q .Now the order of an element of a group divides the order of the group;thus p divides q − 1. �

Remark Many Mersenne numbers are prime numbers themselves; it isas yet unknown, however, if there are infinitely many prime Mersennenumbers.

For a prime p < 5 000 the corresponding Mersenne number is prime if,and only if, p = 2, 3, 5, 7, 13, 17, 19, 31, 61, 89, 107, 127, 521, 607, 1279,2203, 2281, 3217, 4253, 4423. Much larger prime Mersenne numbershave been found, such as 225 964 951− 1 (Nowak 2005), 230 402 457− 1 and232 582 657 − 1 (Cooper and Boone 2005, 2006). A current record is kepton the internet site www.mersenne.org.

32.11 Proof by the Fermat numbers These are the numbers

Fn = 2(2n) + 1

for n ∈ N ∪ {0}. We shall show that any two of these numbers arerelatively prime. Considering their prime factors, one may concludethat there must be infinitely many prime numbers.

More precisely, we shall prove the following recursion formula:∏n−1k=0 Fk = Fn − 2 .

From this it follows immediately that two Fermat numbers Fk and Fn

for k < n are relatively prime: If d is a common divisor, then by therecursion formula d also divides 2, and since the Fermat numbers areodd, we conclude that d = 1.

The recursion formula is proved by induction on n. We have F1− 2 =3 = F0. Assuming the formula to be valid, one obtains

∏nk=0 Fk =

(Fn − 2) · Fn = (2(2n) − 1)(2(2n) + 1) = 2(2n+1) − 1 = Fn+1 − 2. �

Remarks Fermat had conjectured all Fermat numbers to be prime.Now the numbers F0 = 3, F1 = 5, F2 = 17, F3 = 257, F4 = 65 537indeed are prime; but until today no further Fermat number has beenfound that is prime.

Page 207: The Classical Fields

190 Rational numbers

Let us prove for example that F5 is not prime: 641 = 5·27+1 = 54+24,so that mod 641 one obtains the congruences 54 ≡ −24 and 5 · 27 ≡ −1.Raised to the 4-th power, the latter yields 54 ·228 ≡ 1. Using the formercongruence, one obtains−24·228 ≡ 1 or, in other words, F5 = 232+1 ≡ 0,and 641 divides F5.

One of the instances in which the Fermat numbers are significant isthe following. A regular n-gon can be constructed by ruler and compassif, and only if, in the decomposition of n into primes only 2 and primeFermat numbers occur, and the latter occur only with exponent 1. See,for example, Jacobson 1985 Theorem 4.18 p. 274.

32.12 Proof by the Fibonacci numbers These are the numbers fk

(for k ∈ N ∪ {0}) defined recursively by

f0 = 0; f1 = 1; fk+1 = fk + fk−1

for k ∈ N. By induction on k ∈ N we show that

fn = fk · fn−k+1 + fk−1 · fn−k for k < n ∈ N . (1)

This is clear for k = 1. Assuming (1) to be true for a certain k and alln > k, we obtain for k + 1 instead of k and n > k + 1 the fact thatfk+1 · fn−(k+1)+1 + f(k+1)−1 · fn−(k+1) = fk+1 · fn−k + fk · fn−k−1 =fk · fn−k + fk−1 · fn−k + fk · fn−k−1 = fk · fn−k+1 + fk−1 · fn−k = fn.Also by induction on k, it is easy to see that the greatest common divisorof two consecutive Fibonacci numbers is 1:

gcd(fk, fk+1) = 1 for k ∈ N . (2)

Next we generalize this to obtain

gcd(fk, fn) = fd where d = gcd(k, n) for k, n ∈ N . (3)

For the proof, we may assume k ≤ n and proceed by induction on n.For n = 1 and k = 1, the assertion is clear. Now let 2 ≤ N ∈ N, andassume (3) to be true for all k ≤ n ∈ N such that n < N and hence for allk, n ∈ N smaller than N . We prove (3) for N instead of n and for k ≤ N .If k = 1 or k = N , the assertion is trivial. For 1 < k < N , we obtain from(1) and (2) that gcd(fk, fN ) = gcd(fk, fk−1 · fN−k) = gcd(fk, fN−k);by the induction hypothesis, it follows that gcd(fk, fN ) = fD whereD = gcd(k, N − k) = gcd(k, N).

We now use (3) to show that for every finite list p1 = 2, p2 = 3,p3 = 5, p4 = 7, . . . , pn of n pairwise different prime numbers arranged

Page 208: The Classical Fields

32 The multiplication of the rational numbers 191

by their size, there is a prime number not contained in the list. We doso by considering the Fibonacci numbers

f3 = 2 = p1, f4 = 3 = p2, fp3 = 5 = p3, fp4 , . . . , fpn.

By assertion (3), these are relatively prime. If all the prime factorsof these n numbers were from the above list containing just n primenumbers, then fpk

would have to be prime for 3 ≤ k ≤ n, and fpk= pk

by reasons of order. But this is obviously false (p4 = 7 �= 13 = f7). �

Remark The Fibonacci numbers are related to the approximation ofζ = (

√5 + 1)/2 by continued fractions; see Section 4. The numbers cν

for this approximation according to 4.1 are easily found to be cν = 1 forall ν ∈ N ∪ {0}, and the numbers pν and qν , according to the recursionformulae (†) in 4.2, are just the Fibonacci numbers pν = fν+2, qν = fν+1.Hence, the quotients pν/qν = fν+2/fν+1 of two consecutive Fibonaccinumbers converge to (

√5 + 1)/2.

32.13 Proof by the logarithm function For 1 ≤ x ∈ R, let π(x)be the number of primes ≤ x. Let us number the primes in ascendingorder: P = {p1, p2, p3, . . . }. We now estimate the following integral fromabove using a step function:

lnx =∫ x

1

1t

dt ≤∑

n∈N, n≤x

1n

.

The sum on the right-hand side may be estimated from above by∑

n−1

where summation is extended over all n ∈ N having only prime divisors≤ x. This sum in turn is the Cauchy product of geometric series

∏p∈P, p≤x

∞∑k=0

1pk

=∏

p∈P, p≤x

p

p− 1=

π(x)∏n=1

pn

pn − 1.

By the rough estimate n + 1 ≤ pn it follows that

pn

pn − 1= 1 +

1pn − 1

≤ 1 +1n

=n + 1

n

and hence

lnx ≤π(x)∏n=1

n + 1n

= π(x) + 1 .

In particular, limx→∞ π(x) = ∞, since the same is known for the loga-rithm, so that there must be infinitely many primes.

Page 209: The Classical Fields

192 Rational numbers

In fact, we have obtained an estimate for π(x) which is better than theestimate in 32.9. The growth of π(x) has been studied intensively; thevery deep prime number theorem says that limx→∞ π(x) ln x/x = 1; see,for example, Hardy–Wright 1971 Theorem 6 p. 9 or Zagier 1997. �

32.14 A topological proof We define a topology on N by declaringa subset open if it is the union of subsets of the form (a + bZ) ∩ N fora, b ∈ N. In order to see that this is a topology indeed, we have to showthat the intersection of two open sets is open; for this, it suffices to verifythat for a, b, a′, b′ ∈ N the intersection (a + bZ) ∩ (a′ + b′Z) ∩N is open.For an element c of this intersection, one easily verifies that

c ∈ (c + bb′Z) ∩ N ⊆ (a + bZ) ∩ (a′ + b′Z) ∩ N .

Thus, this intersection contains an open neighbourhood of every elementand hence is open.

For b ∈ N � {1}, the set N � bN =⋃b−1

k=1(k + bZ) ∩ N is open in thistopology, so that bN is closed.

Now, if the set P of prime numbers were finite, then N�{1} =⋃

p∈P pNwould be closed and the singleton {1} would be open. But in this to-pology, all non-empty open sets are infinite. Thus, we have obtained acontradiction.

We remark that in the same way we could have used the more so-phisticated topology on N described in 5.21. For the present purpose,however, the topology used here does the trick just as well. �

Exercises(1) The multiplicative group Q× has uncountably many subgroups which areisomorphic to Q×.

(2) Which subgroups of Q× can be embedded into Q+ ?

(3) The group Q× has uncountably many factor groups isomorphic to Q×.

(4) Consider the infinite matrices (zpq)p,q∈P over Z indexed by the set P ofprimes such that for each q ∈ P the corresponding column (zpq)p has only

finitely many non-zero entries. These matrices form a ring Z(P)×P with theusual addition and multiplication of matrices. Show that Z(P)×P is isomorphicto the endomorphism ring End Q×

pos. Compare this with 32.4.

(5) Show that there are infinitely many primes of the form 3n−1 and infinitelymany of the form 4n − 1.

(6) Show that there are infinitely many primes of the form 3n + 1.

(7) Show that there are infinitely many prime elements in the ring J = Z+ iZ.

Page 210: The Classical Fields

33 Ordering and topology of the rational numbers 193

33 Ordering and topology of the rational numbers

As an ordered set, Q has already been studied in Section 3. Accord-ing to Cantor’s characterization 3.4, every totally ordered countable setwhich is strongly dense in itself is order isomorphic to Q. This resultwill be used for the study of the topological space Q. Characteriza-tions of Q as a topological space will be proved. A few results are givenabout the fact that algebraic irrational numbers are relatively hard toapproximate by rational numbers, starting with Liouville’s classical re-sult on this subject. Transitivity properties of the group Aut(Q, <) oforder-preserving bijections and of the homeomorphism group H(Q) ofQ are presented. The section closes with references concerning normalsubgroups of Aut(Q) and the simplicity of H(Q).

33.1 The topology of Q (a) The usual topology on Q is the ordertopology obtained from the structure of Q as an ordered set. Basic opensets in this topology are the open intervals ]a, b[ ⊆ Q for a < b ∈ Q. Anarbitrary subset is open if, and only if, it is a union of open intervals.

(b) This topology can also be described as the topology induced onQ by the order topology on R. In order to see this, one has to verifythat for real numbers r < s ∈ R the set ]r, s[R ∩ Q is open in the ordertopology of Q. (The notation ]r, s[R is used to indicate that we meanan interval in R.) Of course, this is an interesting question only if r ors are not rational. In general, for every x ∈ ]r, s[R ∩ Q we find rationalnumbers ax, bx ∈ Q such that r < ax < x < bx < s, since Q is densein R, and then x ∈ ]ax, bx[ ⊆ ]r, s[R ∩ Q, which shows that the latterset is open in the order topology of Q. The fact that Q is dense in R isessential for this argument, as the example in (d) below shows.

(c) The topology of Q is also induced by the metric d(a, b) = |b− a|.Indeed, the interval ]a, b[ is just the ball of radius (b − a)/2 centred at(a + b)/2.

(d) We demonstrate by an example that there are subsets of Q whichare homeomorphic to Q in their order topology, but not in the topologyinduced from R, so that these two topologies do not coincide. Consider

M = ]−∞,−1[ ∪ {0} ∪ ]1,∞[ ⊆ Q .

There is an order-preserving bijection

M → Q : x �→

⎧⎪⎪⎨⎪⎪⎩x + 1 if x < −1

0 if x = 0

x− 1 if x > 1,

Page 211: The Classical Fields

194 Rational numbers

so M endowed with the order topology is homeomorphic to Q via thismap. But in the topology induced from R, the point 0 is isolated in M ,whereas Q does not have isolated points.

33.2 We consider a few topological properties of Q illustrating theporosity of this space.

(a) In Q every point has arbitrarily small neighbourhoods with empty

boundaries.

Indeed, such neighbourhoods may be chosen of the form U := ]a, b[ :=]a, b[R∩Q where a < b are irrational real numbers. This set is open (see33.1(b)), but also closed since the complement in Q is ]−∞, a[ ∪ ]b,∞[and thus is open, as well. In other words, U equals its closure U and itsinterior intU . This is equivalent to saying that the boundary of U , i.e.the set U � intU is empty.

Using the notion of (small) inductive dimension ind, Property (a) maybe expressed as ind Q = 0. Furthermore, it implies:

(b) The space Q is totally disconnected, i.e. the only connected subsets

are the singletons.

To see this, let A be a subset containing two different points a, b. LetU be a neighbourhood of a in Q having empty boundary such that b /∈ U .Then U is open and closed at the same time, as we have seen in (a),and A = (A∩U)∪ (A∩ (Q � U)) is the union of two non-empty disjointsubsets of A which are open in A, so that A is not connected.

33.3 Theorem For every n ∈ N, the Cartesian product space Qn is

homeomorphic to Q.

This will be obtained by showing that Q is homeomorphic to a certainspace of sequences; a homeomorphism onto Cartesian products will thenbe established by the technique of mixing sequences.

Specifically, we shall consider the space NN of all sequences of naturalnumbers. This is a Cartesian product of countably many copies of N.The factors N of this Cartesian product will be endowed with the discretetopology (in which all subsets are open), and we consider the producttopology on NN.

Now let Per NN be the subset consisting of the finally periodic se-quences, that is, of the sequences (an) with the property that for certainn0, p ∈ N

an+p = an for all n > n0 . (1)

33.4 Theorem The space Per NN is homeomorphic to Q.

Page 212: The Classical Fields

33 Ordering and topology of the rational numbers 195

Proof of Theorem 33.3 by Theorem 33.4. Granting 33.4 for the moment,we can readily obtain 33.3. It suffices to establish that the space Per NN

is homeomorphic to its Cartesian square (Per NN)2; then Theorem 33.3follows by induction. Now the ‘mixing map’

((a1, a2, a3, . . . ), (b1, b2, b3, . . . )) �→ (a1, b1, a2, b2, a3, b3, . . . )

is a homeomorphism of Per NN × Per NN onto Per NN, as can be verifiedeasily; for a period of the mixed sequence one may take twice the productof periods of the original sequences. �

Proof of Theorem 33.4. (1) The set Per NN is countable. Indeed, theset of sequences satisfying equation (1) of 33.3 for fixed n0 and p iscountable, so that Per NN is a union of countably many countable sets;now use 61.13.

(2) We shall also consider the Cartesian product Z × NN with theproduct topology (starting from the discrete topology on Z and N) anduse the homeomorphism Z× NN → R � Q established in 4.10:

ϕ : Z× NN → R � Q : (a0, a1, a2, . . . ) �→ [a0; a1, a2, . . . ] (1)

mapping a sequence to the corresponding continued fraction. We recallthe definition of a continued fraction from 4.1 and 4.2. For a sequence(a0; a1, a2, . . . ) ∈ Z×NN, that is a0 ∈ Z and an ∈ N for n ∈ N, one mayform the fractions

[a0; a1, . . . , ak] := a0 +1

a1 +1

a2 +1

· · ·+ 1ak ·

Then, for k →∞, the sequence ([a0; a1, . . . , ak])k∈N converges to a limitζ ∈ R � Q, and the limit is approached in an alternating way from theleft- and the right-hand side:

[a0; a1, . . . , a2k] < [a0; a1, . . . , a2k+2] < ζ <

< [a0; a1, . . . , a2k+3] < [a0; a1, . . . , a2k+1] . (2)

The limit ζ is denoted by

[an]n∈{0}∪N = [a0; a1, a2, . . . ] := limk→∞

[a0; a1, . . . ak] .

The sequence (an)n∈{0}∪N is called the expansion of ζ into a continuedfraction (it is unique).

Page 213: The Classical Fields

196 Rational numbers

(3) The topological space Per NN is obviously homeomorphic to thesubspace {0} ×Per NN ⊆ Z×NN and therefore, by the homeomorphismϕ in equation (1), homeomorphic to

P := ϕ({0} × Per NN) ⊆ ]0, 1[R .

We show that P is dense in the real interval ]0, 1[R. Let 0 < x < y < 1.We choose an irrational number ζ ∈ R � Q such that x < ζ < y.If [0, a1, a2, . . . ] is the expansion of ζ into a continued fraction, thenaccording to equation (2) there is k ∈ N such that

x < [0; a1, . . . a2k] < ζ < [0; a1, . . . , a2k+1] < y .

The number η := [0; a1, . . . , a2k, a2k+1, a2k+1, . . . ] is the image of a pe-riodic sequence of period 1 under ϕ, so that η ∈ P . Again by equation(2), we see that

x < [0; a1, . . . a2k] < η < [0; a1, . . . , a2k+1] < y .

Thus, for any two different real numbers x, y ∈ ]0, 1[R, there is indeedan element η ∈ P lying strictly between them.

(4) In particular, the topology of P (induced from R) coincides withthe order topology of P as an ordered set; this can be obtained bythe same argument as for Q in 33.1(b), the relevant point being thedensity property proved in step (3). Furthermore, because of this densityproperty and since P is countable by step (1), we know from Theorem 3.4that P is isomorphic as an ordered set to Q, so that with the ordertopology P is homeomorphic to Q. �

33.5 Remarks (1) In the proof above, the set P := ϕ({0} × Per NN)of numbers between 0 and 1 whose expansion into a continued fractionis periodic served as a link between the topological space Per NN andthe ordered set Q. The set P consists in fact of all irrational solutionsbetween 0 and 1 of quadratic equations with rational coefficients; seethe references in 4.5.

(2) The same method as in the preceding proofs shows of course thatR�Q is homeomorphic to its own Cartesian square. This is even simplersince instead of Per NN one just uses the whole space Z × NN, which ishomeomorphic to R � Q.

33.6 Example One might think that a phenomenon as in 33.3 shouldonly occur for spaces of (small inductive) dimension 0, since one ex-pects a decent dimension function to be additive for Cartesian products.

Page 214: The Classical Fields

33 Ordering and topology of the rational numbers 197

However, this is not true; decent dimension functions are well-behavedonly with decent spaces. The following example is discussed in Erdos

1940, see also Engelking 1978 Example 1.5.17 p. 47 or Pears 1975Chapter 4 Example 1.8 p. 153.

Consider the Hilbert space �2R of square summable real sequences,and let �2Q be the subset consisting of the sequences having rationalentries only. Then ind �2Q = 1. Again with the technique of mixingsequences one sees that �2Q is homeomorphic to its Cartesian square(�2Q)2 and, hence, to (�2Q)n for all n ∈ N.

We now prove a characterization of the topological space Q amongthe subspaces of R.

33.7 Theorem A subset of R is homeomorphic to Q in the induced

topology if, and only if, it is countably infinite and has no isolated points.

This implies that there are many subsets of R that are homeomorphicto Q, but not isomorphic to Q as an ordered set. As a concrete example,we mention the set {∑n cn3−n | cn ∈ {0, 2} ∧ ∃k∀n≥k cn = ck } ⊆ C ofend points of the intervals used in 5.35 to define the Cantor set C.Proof (compare Sierpinski 1920): (1) It is clear that every topologicalspace homeomorphic to Q has the asserted properties.

(2) Now let Q be a countable subset of R without isolated points.We may assume that Q is contained in the interval ]0, 1[, since thisinterval is homeomorphic to R (for instance, via the homeomorphismx �→ (2x− 1)/(1− |2x− 1|)).

Since an interval is uncountable, it must meet the complement of Q.Now every open subset of ]0, 1[ contains an interval whose ends arerational, and by choosing a point not in Q from each of these intervals,one obtains a countable dense subset D ⊆ ]0, 1[ such that D ∩ Q = ∅.We enumerate the elements of the countable subsets Q and D in anarbitrary way:

Q = {q1, q2, . . . }, D = {d1, d2, . . . } . (1)

(3) To every finite binary sequence i = (i1, i2, . . . im) ∈ {0, 1}m, wherem ∈ N, we now associate an interval

Ii = [ai, bi] ⊆ [0, 1]

and elements pi ∈ Q, di ∈ D in such a way that the following conditionsare satisfied:

Page 215: The Classical Fields

198 Rational numbers

(3a) We have Qi := Q ∩ Ii �= ∅, but ai, bi /∈ Q.

(3b) If c is the first element of D encountered in the enumeration (1)such that

[0, c] ∩Q �= ∅ �= [c, 1] ∩Q ,

then

{I(0), I(1)} = {[0, c], [c, 1]} ;

more precisely, I(0) is to be the one of the two intervals on the

right-hand side containing q1, and the other one is I(1).

(3c) The element pi is the first element of Q encountered in the enumer-

ation (1) which is contained in Qi; in particular, by (3b) we have

p(0) = q1.

(3d) The element di is the first element of D encountered in the enumer-

ation (1) which is contained in the interior of Ii and such that

[ai, di] ∩Q �= ∅ �= [di, bi] ∩Q .

(3e) For i = (i1, i2, . . . im) we define i′ = (i1, i2, . . . im, 0) and i′′ =(i1, i2, . . . im, 1). Then

{Ii′ , Ii′′} = {[ai, di], [di, bi]} ;

more precisely, Ii′ is the one of the two intervals on the right-hand

side containing pi,

pi ∈ Ii′ ,

and the other one is Ii′′ .

These intervals and elements are constructed by recursion on m, therecursion step being already specified by the stated properties. For therecursive construction of the elements di in (3d), we note that the openinterval Ii �{ai, bi} = ]ai, bi[ intersecting Q non-trivially in fact containsinfinitely many elements of Q, as Q has no isolated points.

We now establish certain properties of these elements and sets. Thefollowing is clear from the construction.

(4) For finite binary sequences i = (i1, i2, . . . im) ∈ {0, 1}m and j =(j1, j2, . . . jn) ∈ {0, 1}n such that m ≤ n one has

Qi ∩Qj �= ∅ ⇐⇒ ik = jk for k = 1, . . . m ⇐⇒ Qi ⊇ Qj ,

pi = pj ⇐⇒ ik = jk for k = 1, . . . ,m, and jk = 0 for k > n .

Page 216: The Classical Fields

33 Ordering and topology of the rational numbers 199

Now we prove:(5) For every infinite binary sequence (iν)ν∈N ∈ {0, 1}N the diameter

of Q(i1,...im) tends to 0 for m→∞.

Assume that this is not the case. Then there is δ > 0 and an infinitesubset M ⊆ N such that for m ∈M the set Q(i1,...im) contains elementsum, vm satisfying |um−vm| > δ. We also may assume that the sequences(um)m∈M and (vm)m∈M converge to elements u, v ∈ [0, 1]. Then we have|u − v| ≥ δ. Let d ∈ D such that u < d < v. Now, by construction,the elements d(i1,...im) ∈ D are pairwise different for m ∈ N, hence, ifm is sufficiently large, d(i1,...im) appears later than d in the enumerationof D. By the properties of d(i1,...im), this implies that Q(i1,...im) liesentirely on one side of d, but for sufficiently large m ∈M , the elementsum, vm ∈ Q(i1,...im) converging to u and v must be on different sides ofd because of u < d < v. This contradiction proves our claim (5).

(6) For every q ∈ Q there is a finite binary sequence i ∈ {0, 1}m such

that q = pi, and the length m of i may be chosen arbitrarily large.

Indeed, by a recursive construction using (3e), one may construct aninfinite binary sequence (iν)ν∈N ∈ {0, 1}N such that every initial part(i1, . . . , im), m ∈ N, of this sequence satisfies

q ∈ I(i1,...,im) . (2)

According to (3c), the elements p(i1,...,im) appear not later than q in theenumeration (1) of the elements of Q. Hence, they describe only finitelymany elements of Q, and there is p ∈ Q such that p = p(i1,...,im) ∈Q(i1,...,im) for infinitely many m. In view of equation (2), it follows fromassertion (5) that p and q are arbitrarily close to each other, i.e., q = p

as claimed in (6).(7) We now consider another subset Q′ of R which is countable and

has no isolated points, like Q. As with Q, we may assume up to homeo-morphism that Q′ ⊆ ]0, 1[. We apply the same constructions as aboveto Q′ instead of Q and we obtain elements p′i and subsets Q′

i having thesame properties with respect to Q′ as the pi and Qi have for Q. Weshow that we obtain a well-defined continuous map

ϕ : Q→ Q′ by setting ϕ(pi) = p′i for i ∈ {0, 1}m, m ∈ N .

This will suffice to finish our proof, for by exchanging the roles of Q andQ′ one then obtains a continuous inverse of ϕ, so that ϕ is a homeomor-phism of Q onto Q′.

Now, ϕ is well-defined by (4), and by (6) it is defined for all elementsof Q. In order to prove continuity, let m ∈ N, i = (i1, . . . im) ∈ {0, 1}m

Page 217: The Classical Fields

200 Rational numbers

and ε > 0. By (5), we find n ∈ N with n ≥ m such that for the finitesequence in = (i1, . . . , im, 0, 0, . . . , 0) ∈ {0, 1}n of length n starting withi and continuing with zeros, the diameter of Q′

in is less than ε. Nowpi = pin by 4), and Qin is a neighbourhood of pin , hence continuity ofϕ in this point will be clear if we establish that ϕ maps Qin into Q′

in .By (6), an element p ∈ Qin may be written as p = pj for a finite binarysequence j of length at least n. By (4) the first n entries of in and of jcoincide, so that ϕ(p) = p′j ∈ Q′

j ⊆ Q′in .

(8) In particular, applying this procedure to Q instead of Q′, oneobtains a homeomorphism of every such subset Q onto Q. �

Via standard topological tools, 33.7 yields the following purely topo-logical characterization of Q.

33.8 Corollary Every regular countable topological space satisfying

the first axiom of countability and having no isolated points is homeo-

morphic to Q.

Proof Let Q be such a space. Since it is countable and satisfies the firstaxiom of countability, is also satisfies the second axiom of countability.Being regular, it is therefore metrizable, according to Urysohn’s theo-rem; see Dugundji 1966 IX 9.2 p. 195. Since Q is countable and sincea singleton is 0-dimensional, Q has dimension 0, and hence is homeo-morphic to a subset of R; see Hurewicz–Wallman 1948 Theorem II 2p. 18 and Theorem V 3 p. 60. Now one may apply 33.7. �

Remark Theorem 33.4, which we proved independently, and Theorem33.3, which we obtained via 33.4, could instead be derived as easy corol-laries of Theorem 33.8. Indeed, the proofs of Theorems 33.7 and 33.8 donot rely on these former results. In the case of 33.3, this is the approachof Neumann 1985, where variants of 33.7 and 33.8 may also be found.The proof of 33.4 given here, based on Cantor’s characterization 3.4 ofQ as an ordered set, is constructive and more concrete.

33.9 Approximation of irrational numbers by rationals Everyirrational number may be approximated by rational numbers, but thereare differences in how easily this can be done. These differences playan important role in number theory. In certain instances, they may beused to tell algebraic numbers from transcendental numbers.

We recall that a real (or complex) number is said to be algebraic ofdegree d ∈ N if it satisfies a polynomial equation of degree d with integercoefficients. A number which is not algebraic of any degree is said to betranscendental.

Page 218: The Classical Fields

33 Ordering and topology of the rational numbers 201

One of the starting points of the subject of approximation by rationalnumbers was the following.

Theorem (Liouville 1844) For every irrational algebraic real number

a of degree d there exists an ε > 0 such that for every rational numbermn , m ∈ Z, n ∈ N ∣∣∣a− m

n

∣∣∣ > ε

nd.

This means vaguely that irrational algebraic real numbers cannot beapproximated all too closely by fractions whose denominators are rela-tively small.

Proof Let f be a polynomial function of degree d with integer coefficientssuch that f(a) = 0. If a = a1, a2, . . . , ad are the complex roots of f andc ∈ Z is the highest coefficient of f , then f(x) = c ·∏d

ν=1(x − aν). Letm/n be a rational number such that |a−m/n| ≤ 1. Then∣∣f(m

n )∣∣ = |c| ∣∣a− m

n

∣∣∏dν=2

∣∣aν − mn

∣∣ ≤≤ |c| ∣∣a− m

n

∣∣∏dν=2(|aν |+ 1 + |a|) = C · ∣∣a− m

n

∣∣ ,where C > 0 is a constant. Now, since the coefficients of f are integers,ndf(m/n) is an integer, and hence nd |f(m/n)| ≥ 1 except if m/n isone of the roots aν . With these possible exceptions we get from theseestimates that |a−m/n| ≥ C−1n−d. The irrational number a is notamong these exceptions. Hence, if ε is chosen small enough to excludethese finitely many exceptions and such that ε < C−1, then the assertionis true. �

The following strong version of Liouville’s theorem in which the de-pendence on the degree of the algebraic number has disappeared can beobtained immediately from Roth’s theorem 4.7.

Theorem Let k > 2 be a real number and a an irrational algebraic

number. Then there is ε > 0 such that for every rational number m/n,m ∈ Z, n ∈ N ∣∣∣a− m

n

∣∣∣ > ε

nk.

For more information, we refer the reader to Hardy–Wright 1971Chapter XI, Baker 1975 Chapter 7 and Baker 1984 Chapter 6 Sec-tions 3ff. A more elementary introduction may be found in Niven 1961Chapters Six and Seven. Results of this type also have applications toDiophantine equations; see, for example, Lang 1971 Section 11 p. 671.

Page 219: The Classical Fields

202 Rational numbers

The preceding discussion motivates the study of the following set ofreal numbers (with k = 3).

33.10 Example For 0 < ε < 1, let

Aε =⋃

m,n∈Nm<n

]mn− ε

n3 , mn

+ εn3

[and A =

⋂0<ε<1

Aε =⋂k∈N

A1/k .

The elements of A are numbers that can be approximated rather wellby fractions with relatively small denominators, the exponent 3 beinga measure for the degree of approximability. More precisely, A consistsof those real numbers a between 0 and 1 for which the statements ofLiouville’s theorem with d = 3 and of the second theorem in 33.9 withk = 3 do not hold. In particular, Liouville’s theorem (which we haveproved) says that any irrational elements of A cannot be algebraic ofdegree ≤ 3; they even must be transcendental by the second theorem in33.9 (which we did not prove). But of course

Q ∩ ]0, 1[ ⊆ A ⊆ ]0, 1[ .

We shall establish the following properties of A.

(i) A is uncountable and hence contains irrational numbers.

(ii) A has Lebesgue measure 0.

Proof (i) We use a topological tool, the so-called Baire category argu-ment; see Dugundji 1966 XI.10.1. We assume that A is countable toobtain a contradiction, and enumerate A = {an | n ∈ N} in an arbitraryway. For n ∈ N, the set A1/n is open and dense in [0, 1], since it containsQ ∩ ]0, 1[, and so is then A1/n � {an}. By Baire’s theorem, applied inthe compact space [0, 1], the intersection of these sets is dense, as well,but⋂

n∈N(A1/n � {an}) =⋂

n∈N A1/n � {an | n ∈ N} = A � A = ∅.This contradiction proves (i).

(ii) Basic information about the Lebesgue measure μ may be foundin Section 10 or in Bauer 1968 §6 pp. 31ff, Halmos 1950 §15 pp. 62ff,Gordon 1994 Chapter 1. Since the measure of an interval is its lengthand since μ is σ-subadditive, the measure of Aε can be estimated asfollows:

μ(Aε) ≤∑

m<n∈N

μ]

mn− ε

n3 , mn

+ εn3

[=∑n∈N

(n− 1) · 2ε

n3≤ 2ε ·

∑n∈N

1n2

.

As the series on the right is convergent, this can be made arbitrarilysmall by choosing ε small enough. From μ(A) ≤ μ(Aε) we infer thatμ(A) = 0. �

Page 220: The Classical Fields

33 Ordering and topology of the rational numbers 203

In particular, A provides another example of a set of real numberswhich has Lebesgue measure 0 without being countable. We knew from5.36 that the Cantor set C has these properties.

As in the proof of (i), a Baire category argument will be used to provethe following result.

33.11 Proposition The set Q cannot be obtained as the intersection

of countably many open subsets of R.

Proof Assume that Q is the intersection of a countable family of opensubsets On, n ∈ N. Since Q is dense in R, so are the On. The setsR�{q} for q ∈ Q are a countable family of open and dense subsets of R,as well. By Baire’s theorem, the intersection

⋂n∈N On ∩

⋂q∈Q(R � {q})

would have to be dense in R, as well; but this intersection is empty.Hence, our assumption is false. �

Order-preserving automorphisms and homeomorphisms

33.12 Let Aut(Q, <) denote the set of all order-preserving bijections ofQ onto itself. Under composition, Aut(Q, <) is a group, and the sameis true for the set H(Q) of homeomorphisms of the topological space Qonto itself. Aut(Q, <) is a subgroup of H(Q), since the topology of Q isthe order topology.

33.13 Lemma For each n ∈ N, the group Aut(Q, <) acts transitively

on the set of subsets of Q having n elements. (In fact, this is true for

any ordered skew field instead of Q.)

Proof It suffices to show that for a finite strictly increasing sequencea1 < a2 < · · · < an, there is an order-preserving bijection α mapping k

to ak for k ∈ {1, 2, . . . , n}. In fact, α can be chosen as a piecewise affinemap; compare the proof of 33.14 below. �

There are analogous results for certain infinite subsets. In order toavoid technicalities in the case of non-Archimedean ordered fields, weformulate such a result just for Q.

33.14 Lemma For any strictly increasing unbounded infinite sequence

a1 < a2 < a3 . . . of rational numbers there is an order-preserving bijec-

tion of Q which maps k to ak for all k ∈ N.

Page 221: The Classical Fields

204 Rational numbers

Proof The definition

α(r) =

{a1 + r − 1 if r ≤ 1

ak + (r − k)(ak+1 − ak) if r ∈ [k, k + 1] with k ∈ N

yields a bijection α with the required properties. �

33.15 Corollary The groups Aut(Q, <) and H(Q) have the cardinality

2ℵ0 of the continuum.

Proof These groups consist of mappings of Q into itself and hence havecardinality at most ℵℵ0

0 = 2ℵ0 ; see 61.15. By constructing an injectionNN → Aut(Q, <), we shall show that the cardinality of Aut(Q, <) is atleast ℵℵ0

0 = 2ℵ0 , and the same applies to H(Q) ⊇ Aut(Q, <).An injection NN → Aut(Q, <) is obtained by mapping a sequence

(mν)ν∈N of natural numbers to an order-preserving bijection of Q whichsends k ∈ N to

∑kν=1 mν according to 33.14. �

There are other aspects under which Aut(Q, <) appears to be quitelarge; for instance, this group cannot be obtained as the union of anycountable chain of proper subgroups; see Gourion 1992.

33.16 Corollary For each integer n ∈ N, the topological space Q is

n-homogeneous, which means that the homeomorphism group H(Q) is

transitive on the set of ordered n-tuples.

Remark Note the difference between this result and 33.13. In thisrespect, Q contrasts with R, which is not n-homogeneous for n ≥ 3; see5.27. There are far more homeomorphisms of Q than order-preservingor order-reversing bijections, as 33.16 shows.

Proof We have to show that for any n ∈ N and for any two n-tuples(a1, a2, . . . , an), (b1, b2, . . . , bn) of distinct elements of Q there is a homeo-morphism mapping aν to bν for ν = 1, . . . , n. Choose a rational numberε > 0 so small that the intervals of length ε

√2 centred at the aν and the

bν are mutually disjoint. These intervals are open and closed in Q, sincetheir end points are irrational. Hence any map which permutes theseintervals, induces homeomorphisms between them and fixes all otherelements of Q is a homeomorphism of Q onto itself. �

We have just used rational intervals with irrational end points and thefact that they are both open and closed in Q. For such subsets, of whichthese intervals are only the tamest examples, the following transitivityproperty of H(Q) can be proved quite generally.

Page 222: The Classical Fields

33 Ordering and topology of the rational numbers 205

33.17 Theorem Let C and D be non-empty proper subsets of Q which

are both open and closed in Q. Then there is a homeomorphism of Qmapping C onto D.

Proof Since C �= ∅ is open, it is (countably) infinite and has no isolatedpoints. The same is true for the complement Q � C, which is openand closed as well, and for D and Q � D. By 33.7, all these sets arehomeomorphic to Q, so that there are homeomorphisms γ : C → D andγ′ : Q � C → Q � D. As their domains of definition are both open,they may be put together to yield a homeomorphism ϕ : Q→ Q whoserestrictions to C and to Q � C are γ and γ′, respectively. �

33.18 Notes More on H(Q) as a permutation group may be found inNeumann 1985. Among other things he shows that there are homeo-morphisms of Q which permute Q in one single infinite cycle (this hadalready been proved by Abel 1982), and even that the conjugacy classesof such homeomorphisms form a set which has the cardinality of thecontinuum. Furthermore, the cycle types of homeomorphisms of Q areinvestigated. Truss 1997 determines the conjugacy classes in H(Q) ofelements of certain cycle types. Mekler 1986 characterizes the count-able subgroups of H(Q) as permutation groups.

Normal subgroups of the automorphism groups

The property proved in 33.13 that Aut(Q, <) is transitive on subsets hav-ing two elements is quite important structurally. In Chapter 2 of Glass

1981 automorphism groups of totally ordered sets with this property arestudied systematically. Under further assumptions, such automorphismgroups have only few normal subgroups, which may be determined ex-plicitly (loc. cit. Theorem 2.3.2 p. 65ff). We formulate the result in thespecial case of Q. Let

Λ(Q) and P(Q)

be the subgroups of Aut(Q, <) consisting of all order-preserving bijec-tions α of Q onto itself which fix every element of some interval ]−∞, r](for α ∈ Λ(Q)) or [r,∞[ (for α ∈ P(Q)), respectively. It is easy to seethat these are normal subgroups of Aut(Q, <).

33.19 Theorem (Glass 1981) The only non-trivial proper normal

subgroups of Aut(Q, <) are Λ(Q) and P(Q) and their intersection, and

the latter is the only non-trivial proper normal subgroup of both Λ(Q)and P(Q). �

Page 223: The Classical Fields

206 Rational numbers

33.20 Simplicity of H(Q) and other groups of homeomorphismsAnderson 1958 proved that the following groups of homeomorphismsare simple, which means that they have no non-trivial proper normalsubgroups:

• the group of all homeomorphisms of Q, of R � Q, and of the Cantorset C (5.35)

• the group of orientation preserving homeomorphisms of the 2-sphereand the 3-sphere

and some more. In the case of spheres, the restriction to the group oforientation-preserving homeomorphisms has to be made since this is anormal subgroup of the group of all homeomorphisms.

Although the spaces considered are rather different, Anderson suc-ceeds in developing a unified method working for them all. For each ofthem, he considers a specific family of subsets which are chosen in sucha way that these families share certain properties with regard to homeo-morphisms, and then he works with these properties, only. In the caseof Q, this family consists of all subsets that are both open and closed.As we have noted in the proof of 33.17, all these subsets are homeomor-phic to each other. This, together with the ensuing transitivity propertystated in 33.17, are the basic facts about these subsets that are neededto establish the properties used by Anderson.

Anderson then shows that for each of these groups and any two non-trivial elements γ, δ there are six elements γ1, . . . , γ6 such that γ =∏6

ν=1 γ−1ν δγν . Then, if N is a normal subgroup and δ ∈ N � {id}, it

follows that γ ∈ N. Since γ was arbitrary, N is the entire group.

Exercises

(1) The group (Q, +) has a proper subgroup S such that the chain (S, <) isisomorphic to the chain (Q, <).

(2) The chain (Q, <) has 2ℵ0 automorphisms.

(3) Are the sets Q2 and Q3, taken with their respective lexicographic orderings,isomorphic as chains?

(4) Let Γ = Aut(Q, <) and H = H(Q) as in 33.12. Is Γ a normal subgroup ofH ? Determine the index of Γ in H.

(5) A countable metric space is not complete.

(6) Compare the Sorgenfrey topology σ on Q (see 5.73) with the ordinarytopology. Show that σ has a countable basis and that (Q, σ) is metrizable (incontrast to the Sorgenfrey topology on R).

Page 224: The Classical Fields

34 The rational numbers as a field 207

34 The rational numbers as a field

The field Q is the unique infinite prime field. We discuss some of itsdistinctive features, in particular some results on sums of squares.

Let F be any field, and consider the ring homomorphism ϕ : Z → F

with nϕ = n · 1. Either Zϕ is finite, in which case Zϕ ∼= Z/pZ for somenatural number p, and p is a prime (because F has no zero-divisors and(hk) · 1 = (h · 1)(k · 1) by distributivity), or ϕ is injective and we mayidentify n · 1 with the integer n. In this case, the smallest subfield of F

(the prime field) consists of all fractions m/n with m,n ∈ Z and n �= 0,and we have

34.1 If the smallest subfield F0 of an arbitrary field is infinite, then F0

is isomorphic to Q.

Without referring to a given field, Q can be described as the field offractions of Z as follows: define an equivalence relation ∼ on Z × N by(a, b) ∼ (c, d) � ad = bc, write a/b for the equivalence class of (a, b), anddefine addition and multiplication in the familiar way. Then a/b = ab−1

and Q = {ab−1 | a ∈ Z ∧ b ∈ N}.34.2 Remarks on addition and multiplication The field Q is es-sentially determined by its additive group: if F is a field with F+ ∼= Q+,then by 31.10 there is an isomorphism σ : Q+ → F+ which maps 1 tothe unit element e of F = (F,+, ∗); now the distributive law impliesthat σ(nq) = nσ(q) = (ne)∗σ(q) = σ(n)∗σ(q) for n ∈ N, q ∈ Q, whenceσ is an isomorphism of fields.

In contrast, Q is not determined by its multiplicative group: we denoteby F = F3(x) the field of fractions of the polynomial ring F3[x]. ThenQ× ∼= F× by 32.1, because F× is the direct product of {±1} with thedirect sum of the cyclic groups generated by the countably many monicirreducible polynomials in F3[x]. Another example is the field F = Q(x);replacing x by a finite or countable set of indeterminates gives moreexamples. If a is any real algebraic number, then Q(a)× ∼= Q×; this is aconsequence of a result of Skolem; see Fuchs 1973 Theorem 127.2.

Similar remarks apply to the ring Z: the additive group does essen-tially determine the multiplication; indeed, upon choosing one of the twogenerators of Z+ as the unit element, the multiplication is determinedvia the distributive law. On the other hand, there are many rings R

such that the multiplicative semigroup (R, ·) is isomorphic to (Z, ·), likethe polynomial rings F3[x] or Z[x].

See 6.2c, d and 14.7 for analogous remarks on the fields R and C.

Page 225: The Classical Fields

208 Rational numbers

We write Q� = {x2 | x ∈ Q× } for the multiplicative group of non-zerosquares of Q. Then 32.2 implies immediately

34.3 Written additively, the factor group Q×/Q� is a vector space over

F2 of dimension ℵ0.

This contrasts sharply with the real case, where R×/R� has order 2.Next, we study the set

S = Q� + Q�

consisting of all sums of two non-zero squares.

34.4 Lemma The following identity of Diophant holds in every com-

mutative ring:

(a2 + b2)(c2 + d2) = (ac± bd)2 + (ad∓ bc)2 . �

34.5 Proposition The set S is a subgroup of Q×, and Q� < S < Q×.

Proof (a) As 1 = (35)2 +( 4

5)2, we have 1 ∈ S and hence Q� = 1 ·Q� ⊆ S.

The well known fact that 2 /∈ Q� gives Q� �= S.(b) Lemma 34.4 implies that S ·S ⊆ S∪Q� = S. Since s−1 = s(s−1)2

for s ∈ S, the set S is indeed a subgroup of Q×.(c) Assume that 3 ∈ S. Then there are numbers a, b, c ∈ N with

a2 + b2 = 3c2 and a, b �≡ 0 mod 3. This would imply 1 + 1 ≡ 0 mod 3,which is a contradiction. Consequently, S �= Q×. �

A rational number q belongs to S if, and only if, it can be writtenin the form q = (a2 + b2)c−2 with a, b, c ∈ N. In order to describe S itsuffices therefore to determine the set {a2 + b2 | a, b ∈ N}. This will bedone in the next steps; see also Exercise 5.

34.6 Lemma If a2 + b2 = m with a, b ∈ N, and if p is an odd prime

dividing m, then p ≡ 1 mod 4 or a ≡ b ≡ 0 mod p.

Proof The numbers a and b may be considered as elements of the fieldFp = Z/pZ. If b ∈ F×

p and c = ab−1, then c2 = −1 ∈ F×p . It follows that

the order p− 1 of F×p is a multiple of 4. �

34.7 Theorem For an odd prime number p the following conditions

are equivalent:

(a) p ≡ 1 mod 4(b) p = x2 + y2 for some rational numbers x and y

(c) p = a2 + b2 for a unique (unordered) pair of natural numbers a, b.

Page 226: The Classical Fields

34 The rational numbers as a field 209

Proof (b) is weaker than (c), and (a) follows from (b) by Lemma 34.6.We show that (a) implies (c).

If (a) holds, then F×p is a cyclic group (see Section 64, Exercise 1) of

order 4k. Hence there is an integer c with c2 ≡ −1 mod p and c canbe chosen between 0 and p/2. Consider the continued fraction p/c =[ q0; q1, . . . , q� ] as described in Section 4. Starting with r−1 = p andr0 = c, the qν are given by the Euclidean algorithm rν−1 = qνrν + rν+1,ending with r� = (p, c) = 1. If s−1 = 1 and if sν denotes the numeratorof [ q0; q1, . . . , qν ], then s0 = q0 and sν+1 = qν+1sν + sν−1; see 4.1. Byinduction we show that

p = rνsν + rν+1sν−1 and r2ν + s2

ν−1 ≡ 0 mod p ;

indeed, p = r0s0 + r1s−1, and the assertion for ν implies that p =(qν+1rν+1 + rν+2)sν + rν+1sν−1 = rν+1(qν+1sν + sν−1) + rν+2sν =rν+1sν+1 + rν+2sν . Thus the first claim is proved; in particular wehave r2

νs2ν ≡ r2

ν+1s2ν−1 mod p. Now r2

ν + s2ν−1 ≡ 0 mod p implies that

r2νs2

ν ≡ −r2νr2

ν+1 mod p and then r2ν+1 + s2

ν ≡ 0 mod p.There is an index κ such that rκ <

√p < rκ−1, and sκ−1 <

√p in view

of rκ−1sκ−1 < p. Since r2κ + s2

κ−1 ≡ 0 mod p and 0 < r2κ + s2

κ−1 < 2p,existence of a and b follows, and only uniqueness remains to be shown.Assume that p = u2 + v2 for another pair of numbers u, v ∈ N. Then(au + bv)2 + (av − bu)2 = p2 by Diophant’s identity 34.4. Since c−1 ≡−c mod p and p = a2 + b2, notation can be chosen in such a way thata/b ≡ c ≡ u/v mod p. This implies av − bu ≡ 0 ≡ au + bv mod p.Because au + bv > 0, it follows that au + bv = p and bu − av = 0.Eliminating v gives (a2 + b2)u = ap and therefore u = a and v = b. �

34.8 Remarks Note that sν is the determinant |q0; q1, . . . , qν | as in 4.2.Utilizing the formal properties of these determinants, a related prooffor the existence of a and b has been given by Smith 1855; comparealso Wagon 1990. A completely different, nice and short proof is dueto Zagier 1990; see Elsholtz 2003 for a detailed discussion. Theequivalence of (a) and (c) has already been proved by Fermat.

34.9 Corollary A natural number n =∏

p pvp is a sum of two squares

of integers if, and only if, vp is even whenever p ≡ 3 mod 4.

Proof The condition on the vp is sufficient. This follows from 2 = 12+12

and Theorem 34.7, together with Diophant’s identity 34.4.The condition is also necessary. If a prime p ≡ 3 mod 4 divides n =

a2 + b2, then p divides a and b by 34.6, and p2 can be cancelled. �

Page 227: The Classical Fields

210 Rational numbers

34.10 Remarks In general, Diophant’s identity 34.5 yields differentrepresentations, for example, 13 · 29 = 112 + 162 = 192 + 42.

As will be shown in 34.22, the number of distinct representations ofany given n ∈ N can be determined. It is thus possible to count thelattice points in the interior of a circle. If the radius of the circle issufficiently large, the number of lattice points approximates the area.This will lead to the famous series π/4 = 1 − 1/3 + 1/5 − 1/7 + − . . . ;see 34.23.

For a characterization of the sums of two positive squares of integerssee Exercise 5.

We put P− = {p ∈ P | p ≡ −1 mod 4} and P+ = P � P−.

34.11 Proposition Both sets P+ and P− are infinite.

Proof A slight variation of Euclid’s proof 32.8 applies to P−: if m is anyproduct of finitely many primes in P−, then 2m+1 ≡ −1 mod 4. Hence2m + 1 has at least one prime divisor q ≡ −1 mod 4, and q is differentfrom each factor of m.

Assume that P :=∏

p∈P+p is finite. By 34.6, any odd prime factor p

of P 2 +1 satisfies p ≡ 1 mod 4 and hence divides P , a contradiction. �

34.12 Written additively, both groups Q×/S and S/Q� are countably

infinite vector spaces over F2.

Proof Remember from 34.5 that Q� < S. Therefore each coset in one ofthe groups has an integer representative of the form z = ±∏ν p vν

ν withpν ∈ P and vν ∈ {0, 1}, vν = 0 for almost all ν. Note that 2 ∈ S. In thefirst case, it follows from 34.7 that z can be chosen as ±∏p∈Q p for somefinite set Q ⊂ P−; in the second case, 34.9 implies that z =

∏p∈R p with

R ⊂ P+. Moreover, these representatives are unique (use Lemma 34.6in the first case). By 34.11 both groups are infinite. �

34.13 Pythagorean triples Let a, b, c ∈ N such that gcd(a, b) = 1and a is odd. Then a2 + b2 = c2 if, and only if, a = u2 − v2, b = 2uv,

and c = u2 + v2 for some u, v ∈ N.

Proof Assume that a2 + b2 = c2. Then b is even and c is odd (or elsea2 ≡ b2 ≡ 1 mod 4 and c2 ≡ 2 mod 4 which is impossible). Hence thenumbers r = b/2, s = (c + a)/2, t = (c − a)/2 belong to N. Moreover,gcd(s, t) = gcd(c, a) = gcd(a, b) = 1 and st = r2. Consequently, thereare u, v ∈ N with s = u2 and t = v2, and u and v have the requiredproperties. The converse is obvious. �

Page 228: The Classical Fields

34 The rational numbers as a field 211

For comparison and later use we show the following.

34.14 Theorem In any finite field Fq each element is a sum of two

squares (including 0).

Proof (a) If q is even, then x �→ x2 is a homomorphism of the additivegroup F+

q and hence a field automorphism.(b) In the odd case, put Q = {x2 | x ∈ Fq }, S = Q + Q, and

S× = S � {0}. Then Q× ≤ S× ≤ F×q since S× is contained in the finite

group F×q and S× is a semigroup by Diophant’s identity 34.4. Note

that |F×q : Q×| = 2 because the map x �→ x2 has a kernel of order 2. If

S× = Q×, then S is a subgroup of order (q+1)/2 in F+q , a contradiction.

Therefore S× = F×q .

Alternative proof: For each c ∈ Fq the set c−Q has (q+1)/2 elements.Consequently, Q ∩ (c − Q) �= ∅, and there are elements x, y ∈ Fq suchthat x2 = c− y2. �

34.15 Remark If a, b ∈ F×q , then {ax2 + by2 | x, y ∈ Fq } = Fq by an

obvious modification of the last proof.

Fermat claimed and Lagrange proved that each natural number is asum of at most four integer squares. First we note

34.16 A natural number of the form 4en with n ≡ 7 mod 8 cannot be

written as a sum of at most three squares.

Proof If x is odd, then x = 4k ± 1 and x2 ≡ 1 mod 8. Hence n cannotbe obtained as a sum of three odd squares or of one odd and a few evensquares. If 4en = a2 + b2 + c2, we may assume that a and b are odd andc is even, but then a2 + b2 + c2 ≡ 2 mod 4. �

Quaternions are a useful tool for dealing with sums of four squares;compare Ebbinghaus et al. 1991 Chapter 7, Salzmann et al. 1995 §11,Conway–Smith 2003, and Section 13, Exercise 6.

34.17 Review: Quaternions The (real) quaternions form a skewfield H with centre R and basis 1, i, j, k over R such that i2 = j2 = −1and k = ij = −ji. There is an anti-automorphism of H mapping x =x0+x1i+x2j+x3k onto the conjugate x = 2x0−x, and Nx = xx = xx =∑

ν x2ν ∈ R. Obviously, N(xy) = (Nx)(Ny). This identity expresses a

product of sums of four squares again as a sum of four squares.Note that the quaternions with integer coefficients form a subring G

of H. If x, y ∈ G and if the coordinates satisfy xν − yν ∈ mZ, we write

Page 229: The Classical Fields

212 Rational numbers

x ≡ y mod m for convenience. This implies xz ≡ yz mod m for anyz ∈ G, also x ≡ y mod m and Nx ≡ Ny mod m.

34.18 Theorem (Lagrange 1770) Each natural number n is a sum

of four integer squares (including 0).

Proof By 34.7 and the identity in 34.17, it suffices to prove the assertionfor prime numbers p ≡ 3 mod 4. Theorem 34.14 implies that a2+b2+1 ≡0 mod p for some integers a, b with 0 ≤ a, b < p/2. Hence there is asmallest number m ∈ N such that mp is a sum of four squares, i.e. thatmp = Nx for some x ∈ G. From |a|, |b| < p/2 it follows that m < p/2.We have to show that m = 1.

Assume that m ≥ 2. There is a quaternion c ∈ G with c ≡ x mod m

and |cν | ≤ m/2 for all coordinates. 34.17 gives

cx ≡ Nc ≡ Nx ≡ 0 mod m .

Consequently, Nc = hm for some h ∈ N and h ≤ m (as |cν | ≤ m/2).Moreover, since cx ≡ 0 mod m, there exists a quaternion z ∈ G suchthat cx = m · z. Taking norms and dividing by m2, we find Nz = hp,and minimality of m implies h = m. This is only possible if |cν | =m/2 = k ∈ N for all ν, and then x ≡ 0 mod k. On the other hand, byminimality of m and because 0 < m < p/2, the xν are relatively prime,and m = 2. Hence Nx = 2p ≡ 2 mod 4 and exactly two of the xν areodd, say x1 and x3. We put

y0 = 12 (x2 + x0), y1 = 1

2 (x3 + x1), y2 = 12 (x2 − x0), y3 = 1

2 (x3 − x1)

and we form a quaternion y ∈ G with coordinates yν . Then we haveNy = 1

2Nx = p. �

34.19 Remarks (a) If we start with a2 +1 ≡ 0 mod p and restrict x, c

and z to numbers in G∩C, then the assumption m > 1 yields h ≤ m/2.This contradicts minimality of m and gives a new proof of 34.7.

(b) Note that 41 is not a sum of four positive squares, and that 103is not a sum of four distinct squares.

(c) Result 34.18 is an immediate consequence of the following moredifficult result, which is proved in Grosswald 1985 Chapter 4 andSmall 1986; compare also Halter-Koch 1982 §1, Greenfield 1983and Conway 1997 p. 137/9.

Theorem (Legendre 1798) Any natural number which is not of the

form 4en with n ≡ 7 mod 8 can be represented as a sum of at most three

squares.

Page 230: The Classical Fields

34 The rational numbers as a field 213

Most other proofs of 34.18 are similar to the one given above; see, forexample, Grosswald 1985 Chapter 3 §3, Small 1982, or Rousseau

1987.Conway and Schneeberger have found the following remarkable gen-

eralization of 34.18; see Conway 2000 and Bhargava 2000:

Theorem If a positive definite quadratic form given by a symmetric

integral matrix (in any number of variables) represents each of the nine

numbers 1, 2, 3, 5, 6, 7, 10, 14, 15, then it is universal, i.e., it represents

every natural number.

Write r4(n) = card{x ∈ G | Nx = n} for the number of distinctordered quadruples of integers such that the sum of their squares isequal to n. We mention a formula for r4(n).

Theorem (Jacobi) r4(n) = 8∑{d ∈ N | d �≡ 0 mod 4 ∧ d | n} .

For a proof see Hardy–Wright 1971 Theorem 386, Andrews et al.1993 or Hirschhorn 1987.

34.20 Corollary Each positive rational number is a sum of four non-

zero squares in Q. Some numbers in Qpos cannot be represented as a

sum of at most three rational squares.

Proof Write m/n = (mn)n−2 and apply Lagrange’s theorem to mn.Using any Pythagorean triple, a sum of fewer than four squares can alsobe written as a sum of exactly four positive squares. The second claimfollows from 34.16; for example, 7 is not a sum of three squares. �

The the subring Z(i) of C will help to make the proof of Theorem34.22 more transparent.

34.21 Gaussian integers The ring J := Z(i) = Z + Zi is a unique

factorization domain. If c ∈ J and cc = p is a prime in N, then c and c

are prime in J.

These well known facts can be proved by means of the Euclideanalgorithm with respect to the ordinary norm. The first part is a standardresult and can be found in most algebra texts (e.g. Herstein 1975 3.8).Assume now that c = uv with u, v ∈ J and uu > 1. Then uu divides theprime p, hence vv = 1 and v is a unit. �

The number r2(n) = card{ (x, y) ∈ Z × Z | x2 + y2 = n} of latticepoints on a circle of radius

√n is expressed in the next result in terms

of δ±(n) = card{ t ∈ N | t ≡ ±1 mod 4 ∧ t | n}.

Page 231: The Classical Fields

214 Rational numbers

34.22 Theorem Put δ(n) = δ+(n)− δ−(n). Then r2(n) = 4δ(n).

Proof Consider ρ(n) = card{ (a, b) ∈ N×(N0) | a2+b2 = n}. Obviously,r2(n) = 4ρ(n), and we have to show that ρ = δ. This will be done inseveral small steps.

(a) Note that n = a2 + b2 ⇔ 2n = (a + b)2 + (a− b)2. Hence ρ(2n) =ρ(n) = ρ(2en). Because δ depends only on the odd divisors of n, itsuffices to study odd numbers n.

(b) The assertion is true for primes: in fact, if p ≡ 1 mod 4, thenp = a2+b2 for exactly two ordered pairs of natural numbers by Theorem34.7, and δ(p) = δ+(p) = 2. If p ≡ −1 mod 4, then p is not a sum of twosquares (34.6), and δ+(p) = δ−(p) = 1.

(c) If a2 + b2 = n = mq for a prime q ≡ −1 mod 4 and some m ∈ N,then q | a, b by 34.6. Consequently, ρ(mq2) = ρ(m). On the other hand,δ(m) = −δ(mq) = δ(mq2). In the case q2 � m, it follows that ρ(n) = 0,and one has δ+(n) = δ−(n).

We may assume, therefore, that n is a product of primes pν ≡ 1 mod 4.(d) In this case, δ can easily be calculated: if p ≡ 1 mod 4, then

δ(pκ) = κ + 1 is just the number of all divisors of pκ. Moreover,gcd(m,n) = 1 implies δ(mn) = δ(m)δ(n). It remains to show that ρ

has analogous properties.(e) Let ρ(n) = card{ (a, b) ∈ N2 | a2 + b2 = n ∧ (a, b) = 1} be the

number of primitive solutions, and put ρ(1) = 1. By the very definition,ρ(n) =

∑d2|n ρ(d−2n).

(f) For a prime p ≡ 1 mod 4, the equation a2 + b2 = p has onlyprimitive solutions, and ρ(p) = ρ(p) = 2 by theorem 34.7. In step (g)we will show that ρ(pκ) = 2 for each κ ∈ N. With (e) this implies foreven as well as for odd κ that ρ(pκ) = κ + 1.

(g) Suppose that pκ = u2 + v2 with gcd(u, v) = 1, and let p = cc,where c ∈ J. By 34.21, the elements c, c are primes in J and J is aunique factorization domain. Hence w = u + vi is of the form cκ−λcλ

(up to a unit iμ). If λ �= 0, κ, then cc = p divides w. This means thatw = pz for some z ∈ J, and p | u, v, which is a contradiction. Therefore,up to a unit, w = cκ or w = cκ and the pair {u, v} is essentially unique.Consequently, ρ(pκ) = 2.

(h) Let n =∏r

ν=1 pκνν = u2 +v2 with pν ≡ 1 mod 4 and gcd(u, v) = 1.

We have to show that ρ(n) = 2r. Consider a prime divisor cν of pν

in J and put w = u + vi. By the argument of step (g) it follows thatw = iμ

∏rν=1 cκν

ν , where cν ∈ {cν , cν}. As only positive solutions arecounted, this yields exactly 2r pairs (u, v).

Page 232: The Classical Fields

34 The rational numbers as a field 215

(i) Assume, finally, that gcd(m,n) = 1 and that m · n has only primefactors p ≡ 1 mod 4. Then ρ(m · n) = ρ(m) · ρ(n) by (h). The formulaρ(m) =

∑d2|m ρ(d−2m) from (e) implies immediately that ρ(m · n) =

ρ(m) · ρ(n). �

34.23 Corollary (Leibniz 1673) π4 = 1− 1

3 + 15 − 1

7 + 19 −+ . . . .

Proof The number of lattice points in a circle of radius r is

γ(r) = card{ (x, y) ∈ Z2 | x2 + y2 ≤ r2 } = 1 + 4∑

n≤r2 ρ(n).

The unit squares centred at these lattice points approximate the areaof the circle. More precisely, (r − 1)2π ≤ γ(r) ≤ (r + 1)2π, andlimr→∞ r−2(γ(r) − 1) = π. By the last theorem, this is equivalent toπ/4 = limr→∞ r−2

∑n≤r2 δ(n).

This sum can be determined by counting how often each odd numberd divides some natural number n ≤ r2. We obtain∑

n≤r2 δ(n) =[r2]− [3−1r2

]+[5−1r2

]−+ . . .

= r2 − 3−1r2 +− · · · − (4k − 1)−1r2 + ψ(r) .

(As customary, [x] denotes the greatest integer z ≤ x.) It is easy to finda bound for the error |ψ(r)|: if only the first 2k terms of the upper sumare considered, the error is at most

[(4k + 1)−1r2

]because the terms of

the sum decrease and have alternating signs. For the same reason, thetwo sums differ by less than k. Choose k such that r ≤ 2k < r+2. Then|ψ(r)| < (2r)−1r2 + r/2+1 = r +1. Hence r−2|ψ(r)| converges to 0 andπ/4 =

∑∞ν=0(−1)ν(2ν + 1)−1.

See also Hilbert–Cohn-Vossen 1932 §6.1. For the problem of find-ing the precise order of magnitude of γ(r)− r2π, see Grosswald 1985Chapter 2 §7. �

34.24 Multiplicative quadratic forms Diophant’s identity 34.4 orthe equation N(xy) = (Nx)(Ny) for quaternions are special cases ofmore general results mainly due to Pfister.

Let V ∼= Fn be a vector space over a field F of characteristic �= 2.Then a quadratic form ϕ on V given by ϕ(x) = xAxt (with a matrixA ∈ Fn×n) is said to be multiplicative if there exists a vector z =(z1, . . . , zn) with rational functions zν ∈ F (x, y) such that the equationϕ(x) · ϕ(y) = ϕ(z) holds in F (x, y); if z depends linearly on y for eachvector x, then ϕ is called strictly multiplicative. It is a rare phenomenonthat z is even a bilinear function of (x, y); in that case we say that ϕ isa composition form.

Page 233: The Classical Fields

216 Rational numbers

Let 〈a1, . . . , an〉 denote the form defined by the diagonal matrix A =diag(a1, . . . , an). By definition, the tensor product of two forms is givenby the tensor product of the corresponding matrices.

Proofs of the following theorem can be found in Pfister 1995; seealso Scharlau 1985 Chapter 2 §10.

Theorem (1) For each choice of a1, . . . , ak ∈ F×, the 2k-dimensional

form 〈1, a1〉 ⊗ · · · ⊗ 〈1, ak〉 is strictly multiplicative.

(2) Every anisotropic multiplicative form is equivalent to a form

〈1, a1〉 ⊗ · · · ⊗ 〈1, ak〉.(3) Any isotropic form is multiplicative.

(4) Non-degenerate composition forms exist only in the dimensions

n ∈ {1, 2, 4, 8}.The last statement (4) is essentially due to Hurwitz 1898; see Lam

2005 V Theorem 5.10 for a proof. For early developments compareEbbinghaus et al. 1991 Chapter 10 §1.

Exercises(1) Show that Q has non-isomorphic quadratic field extensions with isomorphicadditive and isomorphic multiplicative groups.

(2) In the rational plane Q2, the circle with the equation x2 +y2 = 3 is empty.

(3) Let C = { (x, y) ∈ R2 | x2 + y2 = 1} denote the unit circle in R2. Showthat C ∩ Q2 is dense in C.

(4) Determine the rotation group Δ of the rational unit circle C ∩ Q2.

(5) Let n =Q

p pvp ∈ N be a sum of two squares of integers. Show that n is asum of two positive squares of integers if, and only if, v2 is odd or vp > 0 forsome prime p ≡ 1 mod 4.

(6) Verify Jacobi’s theorem 34.19 in the case n = 30.

(7) If a, b ∈ J := Z+Zi and b = 0, there exists z ∈ J such that N(a−bz) < Nb(Euclidean algorithm in J).

35 Ordered groups of rational numbers

The additive group Q+ and the multiplicative group Q×pos of positive

rational numbers are ordered groups in the sense of 7.1 with the usualordering; these ordered groups will sometimes be denoted by Q+

< and(Q×

pos)< for brevity. It is easy to see that they are both Archimedean(compare 7.4 for this notion).

Note that in the multiplicative case it makes no sense to include thenegative rationals since an ordered group is always torsion free (7.3).

Page 234: The Classical Fields

35 Ordered groups of rational numbers 217

In the ordered group Q+< the group structure determines the ordering

(see 35.1). For Q×pos this is not true (see 35.2).

In the case of real numbers, one knows for the isomorphic orderedgroups R+

< and (R×pos)< that, conversely, the ordering determines the

group operation: it follows from completeness that an ordered groupwhose underlying ordered set is R (with the usual ordering) is isomorphicto the ordered group R with the usual addition and the usual ordering;see 7.10. For rational numbers, an analogous result cannot be expected;we shall present several counter-examples of increasingly bad behaviour.

Concerning automorphisms of these structures, we shall prove thatthe ordered group (Q×

pos)< is rigid, that is, it has no automorphismexcept the identity. The proof uses quite deep results from the theoryof transcendental numbers.

35.1 Proposition The only ordering relations on Q making Q+ an

ordered group are the usual ordering and its converse.

Proof Let ≺ be such an ordering relation. We may assume that 0 ≺ 1; ifnot, we replace ≺ by the converse ordering. By monotonicity of additionand transitivity of an ordering relation, it follows that 0 ≺ n for alln ∈ N. For 0 < x ∈ Q, there is m ∈ N such that mx ∈ N, so that0 ≺ mx and hence 0 ≺ x (since x ≺ 0 would imply mx ≺ 0). Bymonotonicity, it follows that −x ≺ 0. Thus, the positive elements withrespect to ≺ are precisely the positive rational numbers in the usualsense, but the ordering of an ordered group is determined by the set ofpositive elements. �

35.2 Example There are ordering relations making Q×pos a non-Archi-

medean ordered group.

Since the usual ordering on Qpos is Archimedean, this shows in par-ticular that there are essentially different ways of turning Q×

pos into anordered group, in contrast with 35.1.

Instead of constructing non-Archimedean ordering relations on Q×pos,

we may do so on the isomorphic group Z(N), the direct sum of infinitelymany copies of the group (Z, +) indexed by N; see 32.1. The elementsof Z(N) are the maps α : N→ Z such that there exist only finitely manyn ∈ N with α(n) �= 0. We define the lexicographic ordering ≺ on Z(N):For different elements α, β : N→ Z of Z(N) and

m := min{n ∈ N | α(n) �= β(n)} , let α ≺ β ⇐⇒ α(m) < β(m) ,

and β ≺ α otherwise. It is straightforward to verify that this is an

Page 235: The Classical Fields

218 Rational numbers

ordering relation and that Z(N) with this ordering relation is an orderedgroup. However, it is not Archimedean; indeed, if εk : N → Z is thecharacteristic map of {k} mapping k to 1 and the other integers to 0,then 0 ≺ ε2 and n · ε2 ≺ ε1 for all n ∈ N.

We now present several counter-examples showing that in the orderedgroups Q+

< and (Q×pos)<, the ordering does not determine the group

structure. We do so by constructing ordered groups such that the un-derlying ordered sets are order isomorphic to Q and hence to Qpos (withthe usual ordering), but they are not isomorphic to Q+

< or to (Q×pos)<.

Note that Q and Qpos are order isomorphic according to 3.4 since Qpos

is countable and strongly dense in itself.

35.3 Example The set

Q(√

2) ={

a + b√

2∣∣ a, b ∈ Q

}is a subgroup of the additive group of real numbers, hence it is an orderedgroup with the addition and the usual ordering of real numbers. It iscountable and, as an ordered set, it is strongly dense in itself, hence itis order isomorphic to Q by 3.4.

The group (Q(√

2), +) is, however, not isomorphic to Q+, since thelatter group is locally cyclic (1.6c, 31.3), but the former is not: 1 and√

2 generate a subgroup of (Q(√

2), +) isomorphic to the group Z × Z,which is not cyclic.

The group (Q(√

2), +) is not isomorphic to Q×pos either, since it is

clearly divisible (in the sense of 1.7), but Q×pos is not, being isomorphic

to a direct sum of copies of Z by 32.1.

35.4 Example The set Q × Q with componentwise addition and thelexicographic ordering ≺ defined by

(r1, r2) ≺ (s1, s2) ⇐⇒ r1 < s1 or (r1 = s1 and r2 < s2)

is an ordered group. As an ordered set, it is order isomorphic to Q by3.4 since it is countable and strongly dense in itself.

As a group, Q×Q is isomorphic to the group Q(√

2) considered in theprevious example via the isomorphism (a, b) �→ a + b

√2. As explained

there, these groups are not isomorphic either to Q+ or to Q×pos. So, Q×Q

is another example of an ordered group not isomorphic to Q+< nor to

(Q×pos)< whose underlying ordered set is isomorphic to Q. This example

is essentially different from 35.3, although the underlying groups areisomorphic. Indeed, it is straightforward that Q(

√2) is an Archimedean

Page 236: The Classical Fields

35 Ordered groups of rational numbers 219

ordered group, whereas Q × Q is not Archimedean: 0 = (0, 0) ≺ (0, 1),but n · (0, 1) = (0, n) ≺ (1, 0) for all n ∈ N.

35.5 Example We finally present an example of a non-commutativeordered group such that the underlying ordered set is order isomorphicto Q. Being non-commutative, this group cannot be isomorphic to Q+

nor to Q×pos.

The underlying group will be the subgroup A+1 (Q) of GL2R consisting

of all matrices(

1 0a α

)with a ∈ Q and α ∈ Qpos. It is isomorphic to the

group of order-preserving affine transformations Q→ Q : x �→ a + αx ofQ; compare 9.3 and 9.4, where such groups are considered for the fieldR instead of Q (and without restrictions concerning the ordering).

We endow the group A+1 (Q) with an ordering relation by carrying the

lexicographic ordering on Qpos ×Q to A+1 (Q) via the bijection

Qpos ×Q→ A+1 (Q) : (α, a) �→

(1 0a α

).

One may easily verify that A+1 (Q) with this ordering relation is an or-

dered group. The underlying ordered set is order isomorphic to Qpos×Qand hence to Q by 3.4, as in the previous examples. Clearly, A+

1 (Q) isnot commutative; e.g., for 0 �= a ∈ Q and 1 �= α ∈ Qpos, the matrices(

1 00 α

)and(

1 0a α

)do not commute.

Next, we consider the automorphisms of the ordered groups Q+< and

(Q×pos)<, that is, group automorphisms which preserve the ordering. For

Q+< it is easy to single them out from the endomorphisms of Q+, which

have been determined in 31.10.

35.6 Aut Q+< = {x �→ rx | r ∈ Qpos } ∼= Q×

pos.

For (Q×pos)<, the analogous question is much more difficult to solve.

The result is the following.

35.7 Theorem The only order-preserving monomorphisms of Q×pos

into itself are the mappings x �→ xn for n ∈ N.

Corollary The ordered group (Q×pos)< has no automorphism except

the identity.

By 7.13, an order-preserving monomorphism of Q×pos into itself is of

the form x �→ xt for some real number t > 0. Therefore, Theorem 35.7is equivalent to the following.

Page 237: The Classical Fields

220 Rational numbers

35.8 Theorem Let t be a positive real number such that rt ∈ Qpos for

all r ∈ Qpos. Then t is a natural number.

Assume that, for some natural number n, the map x �→ xn is even anautomorphism of Q×

pos. Then the inverse mapping x �→ x1/n is also anautomorphism. By 35.8, we conclude that 1/n is a natural number, aswell, hence n = 1, which establishes the corollary of Theorem 35.7.

Theorem 35.8 shall be derived from the following powerful result intranscendental number theory, which we use without proof; see 35.11 forreferences.

35.9 Theorem Let c1, c2, d1, d2, d3 be complex numbers such that c1, c2

are linearly independent over Q as well as d1, d2, d3. Then at least one of

the six numbers ecμdν for μ ∈ {1, 2}, ν ∈ {1, 2, 3} is transcendental, that

is, not a solution of a polynomial equation with rational coefficients. �

In our application of this theorem, linear independence will be grantedby the following fact:

35.10 Lemma The set { ln p | p ∈ P} of logarithms of prime numbers

is linearly independent over Q.

Proof Consider n distinct primes pν , 1 ≤ ν ≤ n, and rational numbersaν with

∑nν=1 aν · ln pν = 0. Write aν = mν/d where mν ∈ Z and

d ∈ N (a common denominator of the aν). Then 0 =∑n

ν=1 mν · ln pν =ln(∏n

ν=1 pmνν ) and

∏nν=1 pmν

ν = 1. By uniqueness of prime factor decom-position, it follows that mν = 0 and hence aν = 0 for all ν. �

Proof of Theorem 35.8. Let t > 0 be a real number such that rt ∈ Qpos

for all r ∈ Qpos. It suffices to show that t must be rational, for then, wemay argue as follows: Write t = m/n with natural numbers m,n, then(2t)n = 2m. This shows that in the decomposition of the rational number2t as a product of positive and negative powers of finitely many primes,the exponent k of the prime 2 satisfies kn = m, so that t = m/n = k isa natural number, as asserted.

In order to prove that t is rational, we distinguish two cases.Assume first that t and t2 are linearly independent over Q. This is

where transcendental number theory comes in. Take any three differentprimes p1, p2, p3. By 35.10, the numbers ln p1, ln p2, ln p3 are linearlyindependent over Q. By Theorem 35.9, therefore, at least one of thenumbers e(ln pν)·tμ

= ptμ

ν for μ ∈ {1, 2}, ν ∈ {1, 2, 3} is transcendentaland a fortiori not rational. But this is a contradiction to the hypothesison t.

Page 238: The Classical Fields

36 Addition and topologies of the rational numbers 221

In the remaining case, t and t2 are linearly dependent over Q, so thatat + bt2 = 0 for some rational coefficients a, b not both equal to 0. Ifb = 0, then a �= 0 and t = 0; if b �= 0, then t = 0 or t = −a/b. Thus t isrational, as asserted. �

35.11 Remarks (1) Theorem 35.9 was discovered independently bySiegel, Lang and Ramachandra. Proofs may be found in Lang 1966Chapter II §1 Theorem 1 p. 8 and Baker 1975 Theorem 12.3 p. 119.A survey on the subject with a sketch of proof is Lang 1971, see inparticular 1.6 p. 638 and pp. 640ff. Theorem 35.8 is a special case of acorollary derived from Theorem 35.9 in Lang 1966 Chapter II.

(2) The arguments presented here leading to Theorem 35.7 are takenfrom the lecture notes Salzmann 1973. Practically the same approachcan be found in Glass–Ribenboim 1994, together with various gen-eralizations. They point out that instead of 35.9 the following easierstatement suffices for the proof of 35.8 and, hence, of 35.7:

If 0 < t ∈ R � Q, then at least one of the three numbers 2t, 3t, 5t isnot rational (in fact, transcendental).As to a proof of the latter statement, Glass–Ribenboim 1994 cite Hal-

berstam 1974, who shows that 2t, 3t, 5t ∈ N implies t ∈ N and mentionsthat his proof together with some results on algebraic numbers yieldsthe transcendency statement above.

36 Addition and topologies of the rational numbers

The additive group Q+ endowed with the usual topology is a topologicalgroup (see Definition 8.1), since it is a subgroup of the topological groupR+; see 8.2. However, there are quite different group topologies on Q+,that is, topologies making Q+ into a topological group. An importantclass of such topologies is obtained from the p-adic metrics on Q whichwe shall present. Furthermore, using continuous characters of Q+, weshall construct a plethora of group topologies on Q+ that cannot beobtained from a metric. Before going into these general constructionsof group topologies, we shall ascertain that there are non-discrete grouptopologies on Q+ that are strictly coarser than the usual topology.

36.1 Example Consider the factor group R/Z of the additive group Rof real numbers; it is a topological group when endowed with the quotienttopology, which makes the canonical projection homomorphism

q : R→ R/Z : x �→ x + Z

Page 239: The Classical Fields

222 Rational numbers

continuous and open (see 62.6 and 62.11). The continuous homomor-phism

w : Q+ → R/Z : r �→ q(r√

2)

is injective since for 0 �= r ∈ Q the irrational number√

2r is not anelement of the kernel Z of q. The image group Q = w(Q) ≤ R/Z is atopological group with the subspace topology inherited from R/Z. Thetopology τ on Q induced by w, whose open sets are the preimages ofopen sets of R/Z, makes w a homeomorphism of Q onto Q and makesQ+ a topological group.

The topology τ is coarser than the usual topology, since w is con-tinuous. In fact, it is strictly coarser; more precisely, we shall see that(Q+, τ) is not isomorphic as a topological group to Q+ with the usualtopology. For this, we consider cyclic subgroups of Q+. In the usualtopology the cyclic subgroups are not dense, whereas we shall find acyclic subgroup which is dense in the topology τ . The subgroup 〈1,

√2〉

of R generated by 1 and√

2 is dense by 1.6(b). The image subgroupq(〈1,

√2〉) therefore is dense in R/Z; since 1 belongs to the kernel of q,

this subgroup is just the cyclic group generated by q(√

2) = w(1), inother words, the image of the subgroup Z of Q+ under the isomorphism(Q+, τ)→ Q of topological groups induced by w. Via this isomorphism,it follows that Z is dense in the topology τ .

The usual topology on Q is induced by the metric d(a, b) = |b − a|(33.1(c)). With respect to this metric, Q+ is even a metric group in thefollowing sense:

36.2 Definition A metric group is a group G together with a metric d

on G which is invariant under right and left translations in the group G,that is (in additive notation) for all a, x, y ∈ G we have d(a+x, a+y) =d(x, y) = d(x + a, y + a).

It is immediate that every metric group is a topological group withrespect to the topology induced by the metric.

Next, we shall construct other invariant metrics on Q+.

36.3 The p-adic metric Fix a prime number p. Every non-zerorational number x can be written in the form

x = pe(x) a

b

with integers e(x) ∈ Z and a, b ∈ Z � {0} such that a and b are notdivisible by p. Here e(x) is uniquely determined by x, as a consequence of

Page 240: The Classical Fields

36 Addition and topologies of the rational numbers 223

uniqueness of prime factorization. Note that e(−x) = e(x) and e(x) ≥ 0if x ∈ Z � {0}. We assert for x, y ∈ Q � {0} that

e(xy) = e(x) + e(y) (1)

e(x + y) ≥ min{e(x), e(y)} if x + y �= 0 (2)

e(x + y) = min{e(x), e(y)} if e(x) �= e(y) . (3)

Assertion (1) is immediate. For (2) and (3), in case x + y �= 0, writey = pe(y)c/d with c, d ∈ Z � pZ, and assume e(x) ≤ e(y). Then

x + y = pe(x)

(a

b+

pe(y)−e(x)c

d

)= pe(x) g

bd

where 0 �= g = ad+pe(y)−e(x)bc ∈ Z, so that e(x+y) = e(x)+e(g) ≥ e(x),which proves (2). If e(x) < e(y), then g is not divisible by p, so thate(g) = 0, and we have obtained (3).

We now define the p-adic absolute value | |p on Q by |0|p = 0 and

|x|p = p−e(x)

for 0 �= x ∈ Q. According to (1), (2) and (3), this absolute value satisfies

|xy|p = |x|p|y|p (1′)

|x + y|p ≤ max{|x|p, |y|p} (2′)

|x + y|p = max{|x|p, |y|p} if |x|p �= |y|p (3′)

for x, y ∈ Q. We note that (3′) could also be deduced from (2′) by ageneral argument; see 55.3.

The inequality (2′) is called the ultrametric property. It implies theweaker triangle inequality |x + y|p ≤ |x|p + |y|p. One verifies easily thattherefore the map dp : Q×Q→ [0,∞[ defined by

dp(x, y) = |x− y|pis a metric on Q, the p-adic metric, which is translation invariant. Itinduces a topology on Q, the p-adic topology . With this topology, Q isa topological field; this can be shown by exactly the same arguments asfor R with the standard absolute value; see 8.2 and 9.1.

In the p-adic topology, a sequence (xn)n∈N converges to 0 if, andonly if, |xn|p tends to 0 in the usual topology, in other words, if the p-exponent e(xn) of xn tends to ∞. A typical example of such a sequenceis (pn)n∈N (which in the usual topology tends to∞ itself). In particular,the subgroup Z of Q+ is not discrete in the p-adic topology, in contrastwith the usual topology.

Page 241: The Classical Fields

224 Rational numbers

36.4 The endomorphisms of Q+ are the maps x �→ rx for r ∈ Q;see 31.10. Except for the zero endomorphism, they are automorphisms.They are continuous not only in the usual topology, but also in the p-adictopology, since by the multiplicative property (1′) of the p-adic absolutevalue one has dp(rx, ry) = |r|pdp(x, y).

But the automorphisms are not continuous when considered as mapsof Q with the p-adic topology into Q with the usual topology or theq-adic topology for a different prime q �= p. Indeed, pn tends to 0 in thep-adic topology, but for r �= 0 the sequence (rpn) does not converge to 0in the ordinary topology or in the q-adic topology, since |rpn|q = |r|q �= 0is constant.

36.5 Corollary For different primes p and q, the p-adic topology, the q-

adic topology and the usual topology on Q+ give rise to non-isomorphic

topological groups. The same is true for the subgroup Z+. �

Thus, we have found countably many essentially different ways ofmaking Q+ into a metric group. But we may go much further.

36.6 Construction From the p-adic metrics, we shall construct con-tinuously many translation invariant metrics on Q+ such that the topo-logical groups obtained from the corresponding topologies on Q+ arepairwise non-isomorphic.

Let X be an infinite set of primes. The p-adic distances dp(x, y) oftwo fixed rational numbers x, y assume only finitely many values, since|x − y| has only finitely many prime factors, and it is easily seen thatthe convergent sum

dX(x, y) =∑

p∈X 2−p · dp(x, y)

defines a metric dX on Q. This metric is obviously translation invariant,since the metrics dp are, so that the corresponding topology makes Q+

into a topological group. Let X = {pn | n ∈ N}; then the sequence(p1p2 . . . pn)n converges to 0 with respect to the metric dX .

Now we consider another infinite set Y of primes, and we assume thatq ∈ Y � X. The image r(p1p2 . . . pn)n of the sequence (p1p2 . . . pn)n

under an automorphism x �→ rx, r �= 0, of Q+ does not converge to 0with respect to dY , since

dY (r(p1p2 . . . pn)n, 0) ≥ 2−q · |r(p1p2 . . . pn)n|q = 2−q · |r|q > 0

is bounded below by a non-zero constant. Thus, the group Q+ with thetopology obtained from the metric dX is not isomorphic as a topologicalgroup to Q+ with the topology obtained from dY .

Page 242: The Classical Fields

36 Addition and topologies of the rational numbers 225

Calling two invariant metrics on Q+ essentially different if the result-ing topological groups are non-isomorphic, we thus have obtained thatthere are at least as many essentially different invariant metrics on Q+

as there are infinite subsets of the set of primes, that is, continuouslymany, since the set of primes is countably infinite (32.7).

In the sequel we shall define further group topologies on Q+ using thefollowing general procedure.

36.7 Construction Let G be a group (written additively), T a topo-

logical group (written multiplicatively), and X a non-empty set of ho-

momorphisms of G to T . Let τX be the topology on G induced by X,

that is, the topology having the set of preimages of open sets of T under

elements of X as subbasis. Then the following statements hold:

(a) (Universal property) Let Y be any topological space. Then a map

f : Y → G is continuous with respect to the topology τX if, and

only if, the composition χ ◦ f : Y → T is continuous for all χ ∈ X.

(b) With the topology τX, the group G is a topological group.

(c) A neighbourhood base of 0 in G in the topology τX is given by

the intersections⋂

χ∈Φ χ−1(V ) where Φ ⊆ X varies over the finite

subsets of X and V over a neighbourhood base of 1 in T .

Proof (a) Clearly the maps χ ∈ X are continuous with respect to τX.Thus, continuity of f implies continuity of χ ◦ f for all χ ∈ X.

Assume, conversely, the continuity of the maps χ◦f with χ ∈ X. Thenfor an open subset U of T the preimage f−1(χ−1(U)) = (χ ◦ f)−1(U) isopen in Y . Thus, the elements of the subbasis of the topology τX haveopen preimages under f , so that f is continuous.

(b) One has to show that the maps

α : G×G→ G : (x, y) �→ x + y and ι : G→ G : x→ −x

are continuous with respect to the topology τX and the product topologyon G×G. For the topological group T , the corresponding maps

α′ : T × T → T : (w, z) �→ wz and ι′ : T → T : z → z−1

are known to be continuous. For χ ∈ X, the product map χ×χ : G×G→T × T : (x, y) �→ (χ(x), χ(y)) is continuous. The fact that χ : G → T

is a homomorphism means that χ ◦ α = α′ ◦ (χ × χ) and χ ◦ ι = ι′ ◦ χ.Hence, all these maps are continuous. By the universal property (a), itfollows that α and ι are continuous.

(c) is clear. �

Page 243: The Classical Fields

226 Rational numbers

We now specialize the preceding construction to G = Q+.

36.8 Construction Let T = S1 ≤ C× be the 1-dimensional torusgroup. For s ∈ R, the map

χs : Q→ T : x �→ e2πisx

is a homomorphism of Q+ to T (a character of Q+). For a non-emptysubset H ⊆ R, let

XH = {χs | s ∈ H } .

Consider the topology τH = τXHon Q induced by XH as described

in 36.7. With this topology, Q+ is a topological group.For a finite subset F ⊆ H and ε > 0, let

Uε,F (0) ={

x ∈ Q∣∣ ∀s∈F |e2πisx − 1| < ε

}.

According to 36.7(c), these subsets form a neighbourhood base of 0 inthe topology τH if F varies over the finite subsets of H and ε assumesarbitrarily small positive values.

If H is countable, then we have a countable neighbourhood base withε = 1/n for n ∈ N, and the topology τH might be (and in fact is) inducedby a metric. If H is uncountable, then τH is a non-metric group topologyon Q+ by 36.11(b) below.

36.9 If H �⊆ Q, then τH is a Hausdorff topology on Q.

Proof Since Q+ is a topological group in this topology, it suffices toshow that the intersection of all neighbourhoods of 0 is {0} (see 62.4).In the notation of 36.8, this intersection is⋂

ε>0 Uε,H(0) ={

x ∈ Q∣∣ ∀s∈H e2πisx = 1

}.

If s ∈ H � Q, then sx /∈ Q for 0 �= x ∈ Q, so that e2πisx �= 1, and theintersection above consists of 0 only. �

36.10 Lemma Let B be a basis of R as a vector space over Q (a Hamel

basis, see 1.13) such that 1 ∈ B, and F a finite subset of B. Then for

u ∈ B � (F ∪ {1}) and δ, ε ∈ ]0, 1[, the basic 0-neighbourhoods specified

in 36.8 satisfy

Uε,F (0) �⊆ Uδ,{u}(0) and Uε,F (0) ∩ Z �= {0} .

Proof By the second version 5.70 of Kronecker’s Theorem, one mayfind integers zs for s ∈ F ∪ {u} and an integer k such that ks − zs is

Page 244: The Classical Fields

36 Addition and topologies of the rational numbers 227

arbitrarily close to 0 for s ∈ F and ku − zu is arbitrarily close to 1/2.Now e2πi(ks−zs) = e2πiks, e0 = 1 and e2πi·1/2 = eπi = −1. Thus, bycontinuity of the exponential function, the integers zs and k may bechosen such that |e2πiks− 1| < ε for s ∈ F and |e2πiku− (−1)| < ε. Thefirst inequality means that k ∈ Uε,F (0), whereas the second inequalitytogether with δ, ε < 1 implies that 0 �= k /∈ Uδ,{u}(0). �

36.11 Theorem Let B be a basis of R as a vector space over Q such

that 1 ∈ B, and let H ⊆ B � {1}. Then the following assertions about

the topology τH on Q introduced in 36.8 hold.

(a) The topology τH is not discrete.

(b) If H is uncountable, then 0 does not have a countable neighbourhood

base in the topology τH . In particular, this topology is not induced

by any metric.

(c) For distinct subsets H and H ′ of B � {1}, the topological groups

(Q+, τH) and (Q+, τH′) are not isomorphic.

Remark The basis B has the cardinality of the continuum; see 1.12.

Proof (a) Every neighbourhood of 0 contains a basic neighbourhoodUε,F (0) for a finite subset F ⊆ B and 0 < ε < 1 and hence an elementdifferent from 0 by Lemma 36.10.

(b) Assume that 0 has a countable neighbourhood base. Then thereis a countable neighbourhood base consisting of basic neighbourhoodsof the form Uεn,Fn

(0) with finite subsets Fn ⊆ H and εn ∈ ]0, 1[, n ∈ N.As H is uncountable, there exists u ∈ H not contained in any of the Fn.By 36.10 the neighbourhood Uδ,{u}(0) for 0 < δ < 1 does not containany of the neighbourhoods Uεn,Fn

(0), contradicting our assumption thatthese form a neighbourhood base.

(c) A group automorphism of Q+ is of the form x �→ rx for 0 �= r ∈ Q;see 31.10. Assume that there is u ∈ H ′ � H. Then we show that theautomorphism x �→ rx is not continuous as a map (Q, τH) → (Q, τH′).This then proves that the topological groups (Q+, τH) and (Q+, τH′) arenot isomorphic.

More specifically, we show for 0 < δ < 1 that Uδ,{u}, which is aneighbourhood of 0 in the topology τH′ , does not contain the image ofany basic neighbourhood Uε,F (0) in the topology τH , for ε > 0 and afinite subset F ⊆ H. The image of the latter neighbourhood is easilyseen to be r · Uε,F (0) = Uε,r−1F (0). That this set is not contained inUδ,{u} is obtained by applying 36.10 to the basis {1, u}∪r−1 ·(B�{1, u})instead of B. �

Page 245: The Classical Fields

228 Rational numbers

36.12 Corollary The cardinality of the set of isomorphism classes of

Hausdorff topological groups with Q+ as underlying group is 22ℵ0.

Proof Let c be the cardinality in question. Any topology on Q is givenby specifying the open sets, that is, by a subset of the power set 2Q.Hence, c ≤ 22ℵ0 , the cardinality of the power set of 2Q.

Let B be a basis of R as a vector space over Q such that 1 ∈ B. Then(B � {1}) ∩ Q = ∅, so that the topology τH constructed from a subsetH ⊆ B �{1} according to 36.8 makes Q+ a Hausdorff topological groupby 36.9 and 36.7. Now B has the cardinality 2ℵ0 of the continuum (see1.12 and 1.10), and the same holds for B � {1}. Therefore assertion (c)of Theorem 36.11 shows that c ≥ 22ℵ0 . The converse estimate has beenestablished initially. �

Exercises(1) Let S = {0} be a subspace of the rational vector space R, and considerS ≤ R+ as a topological group with the topology inherited from R. ThenS ∼= Q if, and only if, each automorphism of S is continuous.

(2) Determine the open and the closed subgroups of Q+ in the usual topology.

37 Multiplication and topologies of the rational numbers

The multiplicative group Q× of non-zero rational numbers endowed withthe usual topology is a topological group, a subgroup of the topologicalgroup R× studied in Section 9.

The positive rational numbers form an open subgroup Q×pos of Q×. It

will be shown (37.3) that in fact this is the only proper open subgroup.We then establish that the only automorphisms of the topological groupQ× are the identity and inversion (37.4), by reducing the question tothe analogous question about the ordered group (Q×

pos)<, which hasbeen discussed in Section 35. The proof uses approximation of irrationalnumbers by continued fractions.

Q× also becomes a topological group when endowed with the p-adictopology introduced in 36.3. The topology induced by the p-adic to-pology on certain cyclic subgroups of Q× will be studied using someelementary number theory. The result will be employed to show that fordifferent primes p and q, the topological groups obtained by the p-adicand the q-adic topology on Q× are not isomorphic.

Page 246: The Classical Fields

37 Multiplication and topologies of the rational numbers 229

37.1 Proposition A subgroup of Q×pos is either cyclic or dense (with

respect to the ordinary topology).

Proof Let A ≤ Q×pos be a subgroup which is not cyclic. Then A is dense

in R×pos (and then a fortiori in Q×

pos) by 1.4, since R×pos is isomorphic to

R+ as a topological group (see 2.1). �

37.2 Example For two different primes p and q, the subgroup pZqZ of

Q×pos generated by p and q is not cyclic and hence dense.

Proof The subgroup in question is the image group of the homomor-phism Z2 → Q× : (m,n) �→ pmqn. This homomorphism is injective dueto uniqueness of prime factorization. Thus, pZqZ is isomorphic to thegroup Z2, which is not cyclic. �

37.3 Theorem The only proper open subgroup of Q× is Q×pos.

Proof An open subgroup A is also closed; see 62.7. The open subgroupA ∩ Q×

pos cannot be cyclic, so that A ∩ Q×pos is dense in Q×

pos by 37.1.Hence A ∩ Q×

pos = Q×pos since it is closed, in other words Q×

pos ⊆ A. Ifmoreover A contains any negative number, then A = Q×. �

37.4 Theorem The only automorphisms of the topological groups Q×

and Q×pos are the identity and the inversion map x �→ x−1.

Proof An automorphism α of the topological group Q× maps the onlyproper open subgroup Qpos onto itself. Moreover, α is determined by itsrestriction on Qpos since Q× = Qpos · {1,−1}. Thus it suffices to provethe theorem for an automorphism τ of the topological group Q×

pos.We shall reduce the problem to the analogous problem for the ordered

group Q×pos, which has been studied in Section 35. To this end, we shall

show that τ preserves or reverses the ordering. In the first case τ is theidentity by Theorem 35.7. In the second case we compose τ with theinversion map to obtain an automorphism which preserves the orderingand hence, again by 35.7, is the identity, so that τ is the inversion map.(We point out that the proof of 35.7 depends on the difficult theorem35.9 from transcendental number theory.)

In order to prove that τ preserves or reverses the ordering, we comparethe images of two arbitrary distinct prime numbers p and q. For this, weuse the techniques of continued fractions introduced in Section 4, in par-ticular the following facts: every irrational number ζ may be expandedinto an infinite continued fraction [c0; c1, c2, . . . ] with c0 ∈ Z and cν ∈ N

Page 247: The Classical Fields

230 Rational numbers

for ν ≥ 1; the approximating finite continued fractions

[c0; c1, c2, . . . , cn] = pn/qn

(where pn, qn ∈ N are relatively prime) converge to ζ, and the numberspn, qn have the following properties:(i) the sequence (qn)n∈N is strictly increasing(ii) pnqn+1 − pn+1qn ∈ {1,−1}(iii) |ζ − pn/qn| < 1/(qnqn+1) < 1/q2

n

(for (i) and (ii), see 4.2(†), and 4.3 for (iii)).Now let p and q be different primes. The ratio (ln p)/(ln q) is irra-

tional, for if it could be expressed as a fraction m/n of natural num-bers m,n, then ln pn = n ln p = m ln q = ln qm, so that pn = qm,contradicting unique prime factorization. Let pn/qn be the fractionsapproximating ζ = (ln p)/(ln q) as described above. From (iii) we in-fer that |(ln p)/(ln q) − pn/qn| < 1/q2

n, and we obtain |ln (pqn/qpn)| =|ln pqn − ln qpn | < (ln q)/qn. By (i), 1/qn tends to 0 for n→∞, so thatlimn→∞ pqn/qpn = 1. This property is preserved under a continuousmultiplicative homomorphism, hence limn→∞ τ(p)qn/τ(q)pn = 1 . Bytaking logarithms we infer that

limn→∞ (qn ln τ(p)− pn ln τ(q)) = 0 .

Since the sequence (qn) is strictly increasing by property (i), one haslimn→∞ 1/(qn ln τ(q)) = 0 as well. By multiplying, we obtain

limn→∞

(ln τ(p)ln τ(q)

− pn

qn

)= 0 , hence

ln τ(p)ln τ(q)

= limn→∞

pn

qn=

ln p

ln q.

Thus t := (ln τ(p))/(ln p) = (ln τ(q))/(ln q) is a real constant indepen-dent of the primes p, q. Exponentiation gives τ(p) = pt for all primes p.Since every positive rational number is a finite product of (positive andnegative) powers of primes and since τ is multiplicative, it follows thatτ(x) = xt for all x ∈ Qpos. Now it is clear that τ preserves or reversesthe ordering. �

Remark The fact that every automorphism of the topological groupQ×

pos preserves or reverses the ordering, which in the proof above hasbeen obtained by using continued fractions, can be established in a moresystematic way by invoking the concept of completion of topologicalgroups. Here, we anticipate results from Section 43. The completion ofthe topological group Q×

pos is R×pos (see 43.11), and every automorphism

Page 248: The Classical Fields

37 Multiplication and topologies of the rational numbers 231

can be uniquely extended to the completion (43.24). Now R×pos is iso-

morphic to the additive group R+, both as a topological group and asan ordered group; indeed, the exponential function is an isomorphismR+ → R×

pos and a homeomorphism and it preserves the ordering. Thusit suffices to see that every automorphism of the topological group R+

preserves or reverses the ordering; but it is well known that this is truemore generally for every continuous bijection (essentially due to the in-termediate value theorem).

Multiplication of rational numbers and p-adic topologies

We consider, for a prime p, the p-adic topology on Q which was intro-duced in 36.3 via the p-adic absolute value | |p. As we remarked there,Q is a topological field with this topology. In particular, the multiplica-tive group Q× with the topology induced by the p-adic topology is atopological group.

The following number-theoretic lemma, which we shall use as a techni-cal tool in the sequel, improves upon Theorems II and III in Birkhoff–

Vandiver 1904.

37.5 Lemma Let p be a prime and n, u, v ∈ Z such that u �= v and

u ≡ v �≡ 0 mod p.

(i) If p �= 2 or if u ≡ v mod p2 or if n is odd, then

|un − vn|p = |u− v|p · |n|p .

(ii) In the remaining case, with p = 2, u �≡ v mod 4 and n even, one has

|un − vn|2 = |u + v|2 · |n|2 .

Proof We may assume that n ∈ N, as |−n|p = |n|p and |u−n − v−n|p =|un − vn|p, in view of |u|p = 1 = |v|p.

(i) For 2 ≤ m ∈ N, we obtain by binomial expansion

um − vm

u− v=

(v + u− v)m − vm

u− v

= mvm−1 +(m2

)vm−2(u− v) +

∑mj=3

(mj

)vm−j(u− v)j−1

≡ mvm−1 +(m2

)vm−2(u− v) mod p2

≡ mvm−1 mod p .

Assume that m �≡ 0 mod p. Then |(um − vm)/(u− v)|p = 1 = |m|p, and

Page 249: The Classical Fields

232 Rational numbers

this is trivially true for m = 1, as well. For n = mpk one then has∣∣∣∣un − vn

u− v

∣∣∣∣p

=∣∣∣∣ un − vn

um − vm

∣∣∣∣p

=k−1∏i=0

∣∣∣∣∣umpi+1 − vmpi+1

umpi − vmpi

∣∣∣∣∣p

and |n|p = |p|kp. Thus it suffices to prove (i) in the special case n = p.In this case, p �= 2 or u ≡ v mod p2, hence

(p2

)vp−2(u − v) ≡ 0 mod p2.

The above considerations (for m = p) show that (up − vp)/(u − v) ≡pvp−1 mod p2, so that |(up − vp)/(u− v)|p = p−1 = |p|p, which proves(i) for n = p.

Assertion (ii) follows from (i): the assumptions of (ii) imply thatu2 − v2 = (u + v)(u− v) ≡ 0 mod 4 and |u2 − v2|2 = |2(u + v)|2. Henceby (i) we have |un − vn|2 = |(u2)n/2 − (v2)n/2|2 = |u2 − v2|2 · |n/2|2 =|u + v|2 · |n|2 . �

Let 1 �= a ∈ Qpos and endow the subgroup 〈a〉 = aZ ∼= Z+ of Q× withthe p-adic topology. One may ask when aZ (with the p-adic topology) isisomorphic as a topological group to the group Z+, also endowed withthe p-adic topology, or when aZ is discrete in the p-adic topology. Notethat the p-adic topology on Z+ is not discrete; see the end of 36.3.

37.6 Theorem Let p be a prime, and a = u/v where u �= v are

relatively prime positive integers. Then the following assertions hold.

(i) The group aZ is discrete in the p-adic topology if, and only if, the

prime p divides u or v.

(ii) The groups aZ and Z+, both endowed with the p-adic topology, are

isomorphic as topological groups if, and only if, u ≡ v �≡ 0 mod p.

(ii′) In particular, if u �≡ 0 �≡ v mod p, then the subgroup (ap−1)Z gen-

erated by ap−1 and the group Z+, both endowed with the p-adic

topology, are isomorphic topological groups.

(iii) Assume that u �≡ 0 �≡ v mod p, and consider another prime q and

another positive rational number b. If the group aZ with the p-adic

topology and the group bZ with the q-adic topology are isomorphic

as topological groups, then necessarily q = p.

(iv) If p �= q, then the group aZ with the p-adic topology is not isomor-

phic to the group Z+ with the q-adic topology.

Proof (i) First we make the following Observation: if p divides u, thenit does not divide v, so that an = un/vn tends to 0 for n → ∞ in thep-adic topology.

Assume now that aZ is not discrete in the p-adic topology. Then thereis an unbounded sequence (mν)ν∈N of integers such that amν converges

Page 250: The Classical Fields

37 Multiplication and topologies of the rational numbers 233

p-adically to am for some m ∈ Z, and amν−m = amν a−m converges to 1.By a mere change of notation we may assume that amν converges to 1.Moreover we may assume that mν > 0 for all ν (by passing to inverses, ifnecessary) and that mν tends to ∞. Then our initial observation showsthat p cannot divide u. Since a and a−1 generate the same subgroup ofQ×, we also obtain that p cannot divide v if aZ is not discrete. Thus,we have proved the ‘if’ part of (i) by contraposition.

The ‘only if’ part will be postponed and proved after the proof of (ii).(ii) If the groups Z+ and aZ are isomorphic as topological groups with

the p-adic topology, then in the first place aZ is not discrete, so that bythe part of (i) that is already proved we know that u �≡ 0 �≡ v mod p.

Specifically, an isomorphism maps the generator 1 of Z to one of thegenerators a or a−1 of aZ, and since the inversion map is an automor-phism of the topological group aZ with the p-adic topology, there is anisomorphism Z+ → aZ of topological groups mapping 1 to a and hencem ∈ Z to am. In other words, the group isomorphism

Z+ → aZ : m �→ am (1)

is a homeomorphism with respect to the p-adic topology. Now the se-quence (pμ)μ∈N converges to 0 in the p-adic topology; see the end of 36.3.Hence, apμ

converges to 1. In other words, apμ − 1 = (upμ − vpμ

)/vpμ

converges p-adically to 0. Hence, for μ sufficiently large, p dividesupμ − vpμ

. Since by Fermat’s little theorem up ≡ u mod p, it followsthat 0 ≡ upμ − vpμ ≡ u− v mod p, that is, u ≡ v mod p. Thus the ‘onlyif’ part of (ii) is proved.

Assume now conversely that u ≡ v �≡ 0 mod p. We show that thegroup isomorphism (1) is a homeomorphism with respect to the p-adictopology on both groups. Since |a|p = 1 = |v|p we obtain from 37.5 form,n ∈ Z and k := |m− n| that

|am−an|p = |ak−1|p = |(uk−vk)v−k|p = |uk−vk|p = |m−n|p · |u±v|p .

This shows that the bijection (1) is a homeomorphism, as |u± v|p �= 0.Thus, the ‘if’ part of (ii) is also established.

By Fermat’s little theorem again, up−1 ≡ 1 mod p if u �≡ 0. Hence,(ii′) immediately follows from (ii).

Thus if u �≡ 0 �≡ v mod p, then a(p−1)Z is not discrete in the p-adictopology, hence the subgroup aZ containing this non-discrete subgroupis not discrete. This is the contraposition of the ‘only if’ part of (i).

(iii) As in the proof of (ii), one sees that if aZ with the p-adic topologyand bZ with the q-adic topology are isomorphic topological groups, then

Page 251: The Classical Fields

234 Rational numbers

the map am �→ bm for m ∈ Z is an isomorphism. This map then inducesan isomorphism of the subgroup (a(p−1)(q−1))Z with the p-adic topologyonto the subgroup (b(p−1)(q−1))Z with the q-adic topology. By (ii′), thefirst group is isomorphic to Z+ with the p-adic topology. In particular,both subgroups are not discrete. By (i) and (ii′) again, the secondsubgroup is isomorphic to Z+ with the q-adic topology. Thus, the p-adic and the q-adic topology on Z give rise to isomorphic topologicalgroups. By 36.5, it follows that p = q.

(iv) If p divides u or v, then aZ is discrete in the p-adic topology by (i);in particular, it cannot be isomorphic to Z+ with the q-adic topology.

If u �≡ 0 �≡ v mod p, and if aZ with the p-adic topology is isomorphicto Z+ with the q-adic topology, then by (ii′) these groups contain cyclicsubgroups isomorphic to Z+ with the p-adic topology. Thus, there existsm ∈ Z � {0} such that the sequence (pμm)μ∈N converges to 0 in the q-adic topology, which means that, for μ sufficiently large, pμm is divisibleby arbitrarily large powers of q. This is possible only if q = p. �

37.7 Corollary For distinct primes p and q, the p-adic topology, the q-

adic topology and the usual topology on Q× give rise to non-isomorphic

topological groups. The same is true for Q×pos.

Proof First, Q×pos contains cyclic subgroups which are not discrete in the

p-adic topology according to 37.6. In contrast, every cyclic subgroup ofQ× is discrete in the ordinary topology, so that Q×

pos and Q× with theordinary topology cannot be isomorphic as topological groups to theircounterparts with the p-adic topology.

Now assume that for primes p and q there is an isomorphism ϕ of Q×pos

(or Q×) with the p-adic topology onto Q×pos (or Q×, respectively) with

the q-adic topology. We show that then p = q, so that our corollary isproved. Let m ∈ N � pZ. Then ϕ(m2) = ϕ(m)2 ∈ Q×

pos. Via ϕ, thesubgroup (m2)Z of Q×

pos with the p-adic topology is isomorphic to thesubgroup ϕ(m2)Z of Q×

pos with the q-adic topology. Since m2 ∈ N � pZ,as well, 37.6(iii) says that indeed q = p. �

Exercises(1) Multiplication defines a map μ :

L

p∈PpZ → Q×

pos from a subspace of a

product of discrete groups into R. Is μ or μ−1 continuous?

(2) The topological groups Q+ and Q×pos have isomorphic character groups.

Page 252: The Classical Fields

4

Completion

In Chapter 1, the basic properties of the real numbers were taken forgranted. In Chapter 3, devoted to the rational numbers, we sometimesused the fact that they are embedded in the real numbers. In the presentchapter, we shall discuss standard procedures which allow us to constructthe domain of real numbers from the domain of rational numbers by so-called completion. (A non-standard procedure was already presented inSection 23.)

The rational numbers are not complete, either with respect to theirordering (there are non-empty bounded sets which have no supremumwithin the rational numbers), or as a topological group (there are Cauchysequences of rational numbers which do not converge to a rational num-ber). Completion processes remedy these defects by a cautious enlarge-ment which supplies the missing suprema or limits, without introducingnew incompleteness problems. Corresponding to the two facets of incom-pleteness of the rational numbers, there are two types of completion, onefor ordered structures and, more specifically, for ordered groups, and an-other one for certain topological structures, in particular for topologicalgroups. These completion principles will be presented here. (We donot, however, discuss completion of metric spaces or, more generally, ofuniform spaces without algebraic structure.)

For the rational numbers, both kinds of completion, the completionof Q as an ordered group and its completion as a topological group, willlead to the same mathematical object (up to isomorphism), the additivegroup R of real numbers, with its ordering and its topology. In the caseof an ordered field and of a topological field, one may also extend themultiplication to the completion of the additive group. In particular,this allows us to endow R with the structure of an ordered field and ofa topological field in a canonical way.

235

Page 253: The Classical Fields

236 Completion

41 Completion of chains

In this section we shall discuss the completion of chains, as a basis forthe next section on the completion of ordered groups.

41.1 Basic notions We recall from 3.1 that a chain is called completeif every non-empty subset A that is bounded above has a least upperbound, the supremum supA. In a complete chain C, every non-emptysubset B that is bounded below has a greatest lower bound, the infimuminf B. Indeed, the set A := {x ∈ C | x ≤ B } of all lower bounds ofB is non-empty and bounded above (by any element of B), and it isan immediate consequence of the definitions that supA is the greatestlower bound of B.

A subset D of a chain C is called coterminal if for every c ∈ C thereare x, y ∈ D such that x ≤ c ≤ y.

A subset D of C is called weakly dense if it is coterminal and for everyc ∈ C we have sup{d ∈ D | d ≤ c} = c = inf{d ∈ D | c ≤ d}. Clearly,this is equivalent to saying that every element of C is the supremum ofsome non-empty subset of D and the infimum of some non-empty subsetof D. Yet another way of putting this is that for all c1 < c2 in C thereare d1, d2, d3, d4 ∈ D such that d1 ≤ c1 ≤ d2 < d3 ≤ c2 ≤ d4. In 3.1 and3.2, this was compared with other density notions. In particular, weakdensity is stronger than topological density of D in C, which means thatevery non-empty open interval of C contains an element of D.

An ideal of a chain C is a non-empty subset I with the property thatfor every element a ∈ I, every element smaller than a is also containedin I, that is, I contains the interval ] , a] = {x ∈ C | x ≤ a}.

As in Section 3, a map ϕ : C → C ′ between two chains is said topreserve the ordering (or to be order-preserving) if for all x, y ∈ C suchthat x ≤ y one has ϕ(x) ≤ ϕ(y).

41.2 Definition A completion of a chain C is a complete chain C

together with an order-preserving injective map ι : C → C whose imageι(C) is weakly dense in C.

41.3 Construction Every chain C has a completion.

Proof (1) We call an ideal of C admissible, if it is bounded above andcontains its supremum if this supremum exists, and we define C to bethe set of all admissible ideals of C.

The ordering relation on C will be inclusion of sets. The simplestadmissible ideals are the intervals ] , c] for c ∈ C; we may therefore

Page 254: The Classical Fields

41 Completion of chains 237

embed C into C by the map

ι : C → C : c �→ ] , c] ,

which clearly is injective and preserves the ordering.The admissible ideals can be obtained in the following way. For an

arbitrary ideal I of C, we define

I =

{I ∪ {sup I} if I has a supremum

I otherwise.

One verifies that this is an ideal again. Obviously, if I has a supremum,then

sup I = sup I , so that ¯I = I . (1)

Moreover, one verifies for ideals I, J that

I ⊆ J =⇒ I ⊆ J . (2)

If the ideal I is bounded above, then I is admissible in view of equa-tion (1), and an admissible ideal A satisfies A = A. Thus, the admissibleideals are precisely the ideals I obtained from an ideal I which is boundedabove:

C = { I | I ideal of C and bounded above} .

(2) We remark that for an admissible ideal A the pair (A,C � A)constitutes what is called a Dedekind cut of the chain C. We shall notmake further use of this notion.

(3) It is easy to see that C is a chain. Indeed, if I �= J are any idealsof C and if J � I �= ∅, say, we obtain for a ∈ J � I that I ≤ a since I isan ideal and then I ⊆ J since J is an ideal.

(4) Now it will be shown that the chain C is complete. Let A ⊆ C

be bounded above, that is, A consists of admissible ideals all containedin some admissible ideal W . It is immediate from the definition thatthe union U :=

⋃A =⋃{A | A ∈ A} is an ideal again; it is bounded

above since it is contained in W which is bounded above. Clearly, U

is an element of C which is an upper bound of A. We show that it isthe supremum of A, which means that it is contained in every admis-sible ideal W containing all the elements of A. Clearly U ⊆ W ; usingequations (2) and (1), we conclude that U ⊆W = W .

(5) Finally we verify that ι(C) is weakly dense in C. Let A,B ∈ C

such that A ⊆ B, A �= B. An element b ∈ B � A is an upper boundof A, but there is still another bound u ∈ C of A such that u < b

Page 255: The Classical Fields

238 Completion

(else b = sup A ∈ A). Finally, let v ∈ C be an upper bound of B

and a ∈ A. Then ] , a] ⊆ A ⊆ ] , u] ⊂ ] , b] ⊆ B ⊆ ] , v], in other wordsι(a) ⊆ A ⊆ ι(u) ⊂ ι(b) ⊆ B ⊆ ι(v). This shows that ι(C) is weakly densein C, according to one of the alternative descriptions of this notion in41.1. �

In order to show that essentially this is the only way to construct acompletion of a chain, we study some properties of completions withrespect to extensions of order-preserving maps. It is instructive to dis-tinguish the roles played by the two constituents of the notion of acompletion, completeness and density.

41.4 Lemma Let ι : C → L be an order-preserving injective map

between chains such that ι(C) is coterminal in L, and let R be a complete

chain. Then every order-preserving map ϕ : C → R admits an extension

over ι, in other words there exists an order-preserving map ϕ : L → R

such that ϕ ◦ ι = ϕ.

Proof For x ∈ L, the preimage ι−1(] , x]) of the interval ] , x] ⊆ L is non-empty and bounded above in C since ι(C) is coterminal in L. Hence,ϕ(ι−1(] , x])) is bounded above in R, so that the map

ϕ : L→ R : x �→ supϕ(ι−1(] , x]))

is well-defined. It is immediate that ϕ has the stated properties. �

41.5 Lemma Let ι : C → L be an order-preserving injective map

between chains such that ι(C) is weakly dense in L. Then the identity

map ψ = id is the only map ψ : L→ L that preserves the ordering and

satisfies ψ ◦ ι = ι.

Proof Let ψ be such a map. The condition ψ ◦ ι = ι says that ψ inducesthe identity map on ι(C). Now suppose that x �= ψ(x) for some x ∈ L,say x < ψ(x) (the other case is treated analogously). Then x /∈ ι(C).By weak density, there is c ∈ C such that x < ι(c) < ψ(x), and thenψ reverses the order of the elements ι(c) = ψ(ι(c)) and x, which is acontradiction. �

41.6 Theorem: Universal property of completions Let Ci for

i = 1, 2 be a chain, and ιi : Ci → Ci a completion of Ci.

Then for every order-preserving bijection ϕ : C1 → C2 there is a

unique extension ϕ : C1 → C2 of ϕ to the completions, that is an order-

preserving map such that ϕ ◦ ι1 = ι2 ◦ ϕ. Moreover, the extension ϕ is

bijective.

Page 256: The Classical Fields

42 Completion of ordered groups and fields 239

Proof Such a map ϕ is obtained by applying 41.4 to ι2 ◦ ϕ : C1 → C2.We replace ϕ by ψ := ϕ−1 and permute the roles of ι1 and ι2; thisgives an order-preserving map in the opposite direction ψ : C2 → C1

such that ψ ◦ ι2 = ι1 ◦ ψ. The composition ψ ◦ ϕ : C1 → C1 satisfiesψ ◦ ϕ ◦ ι1 = ψ ◦ ι2 ◦ϕ = ι1 ◦ψ ◦ϕ = ι1, so that ψ ◦ ϕ = id by 41.5. Againby permuting roles one obtains analogously that ϕ ◦ ψ = id. Thus, ϕ isbijective with inverse map ψ. This holds for every pair of extensions ϕ

and ψ; hence these extensions are unique. �

41.7 Corollary: Uniqueness of completion Let ι : C → C and

ι : C → C be completions of a chain C. Then there is a unique order-

preserving bijection ϕ : C → C such that ϕ ◦ ι = ι.

Proof This is the special case of 41.6 with C1 = C = C2 and ϕ = id. �

The following will provide a useful tool for handling elements of thecompletion without knowing the completion explicitly.

41.8 Lemma Let ι : C → C be a completion of a chain C, and

X, Y non-empty subsets of C that are bounded above. Then sup ι(X) ≤sup ι(Y ) if and only if every upper bound u ∈ C of Y is an upper bound

of X as well.

Remark Note that the upper bounds u are restricted to C. Thesuprema of ι(X) and ι(Y ) exist in the completion.

Proof We remark that for u ∈ C the condition sup ι(X) ≤ ι(u) is equiva-lent to X ≤ u. The ‘only if’ part is then clear. For the ‘if’ part, weconsider the set U(X) = {u ∈ C | X ≤ u} of all upper bounds of X

in C. Since ι(C) is weakly dense in C, the supremum of ι(X) is the in-fimum of the set { ι(u) | u ∈ C, sup ι(X) ≤ ι(u)}, which is just ι(U(X))by the initial remark. Hence, if by assumption U(Y ) ⊆ U(X), thensup ι(X) = inf ι(U(X)) ≤ inf ι(U(Y )) = sup ι(Y ). �

42 Completion of ordered groups and fields

We now turn to the completion problem for ordered groups. (The defini-tion and basic properties of an ordered group may be found in Section 7.)When constructing the completion of the underlying chain of an orderedgroup as in 41.3, one may try to extend the group operation to the com-pletion and hope to obtain an ordered group again. Now according to7.5, a completely ordered group is Archimedean, so this programme can

Page 257: The Classical Fields

240 Completion

only be successful for Archimedean groups. Furthermore, recall thatby 7.7 an Archimedean ordered group is commutative. This is why allgroups considered here will be written additively.

42.1 Definitions An ordered group is called completely ordered if theunderlying chain is complete. A completion of an ordered group G isa completely ordered group G together with an order-preserving groupmonomorphism ι : G → G whose image ι(G) is weakly dense in G (sothat ι : G→ G is a completion of the chain underlying G).

The following lemma gives a hint as to how the group operation of anordered group should be extended to the chain completion.

42.2 Lemma Let X, Y be non-empty subsets of an ordered group G.

(a) If Y has a supremum in G, and if z ∈ G satisfies X + Y ≤ z, then

X + sup Y ≤ z.

(b) If the suprema of X and Y exist in G, then the same holds for

X + Y , and sup(X + Y ) = supX + sup Y .

Proof (a) The assumption X + Y ≤ z implies that Y ≤ −x + z for allx ∈ X so that sup Y ≤ −x + z and hence X + sup Y ≤ z.

(b) It follows that X ≤ z − supY and hence, if sup X exists, thensupX ≤ z− supY , sup X + sup Y ≤ z for all upper bounds z of X + Y .On the other hand, it is clear that sup X+sup Y is itself an upper boundof X + Y , hence it is the least upper bound. �

42.3 Construction Every Archimedean ordered group G has a com-

pletion.

Proof (1) Let ι : G → G be any completion of G as a chain. Using theaddition of G, we define an addition on G as follows.

For a ∈ G, let A(a) := {x ∈ G | ι(x) ≤ a}. Since ι(G) is weakly densein G, we have

a = sup ι(A(a)) .

Since ι(G) is coterminal in G, the subset A(a) is bounded above in G.Now consider a second element b ∈ G. Since G is an ordered group,A(a) + A(b) is bounded above. With view to 42.2, we define

a + b := sup ι(A(a) + A(b)) .

It is clear that this addition satisfies the monotonicity law required foran ordered group since for c ∈ G the relation a ≤ c translates intoA(a) ⊆ A(c).

Page 258: The Classical Fields

42 Completion of ordered groups and fields 241

(2) We now prove more generally that this addition satisfies the fol-lowing equation for arbitrary non-empty subsets X, Y ⊆ G that arebounded above:

sup ι(X) + sup ι(Y ) = sup ι(X + Y ) (1)

Let a = sup ι(X), b = sup ι(Y ). Then X ⊆ A(a), Y ⊆ A(b), so thatsup ι(X + Y ) ≤ sup ι(A(a) + A(b)) = a + b.

As to the converse inequality sup ι(X + Y ) ≥ sup ι(A(a) + A(b)),according to 41.8, it suffices to show for z ∈ G that

X + Y ≤ z =⇒ A(a) + A(b) ≤ z . (2)

Let x′ ∈ A(a). If ι(x′) < a = sup ι(X), there is x ∈ X with x′ < x.If ι(x′) = a = sup ι(X), then x′ = sup X. Hence there is x ∈ G suchthat x′ ≤ x and x ∈ X or x = sup X. Likewise, for y′ ∈ A(b), there isy ∈ G such that y′ ≤ y and y ∈ Y or y = sup Y . By 42.2, it follows thatx′ + y′ ≤ x + y ≤ z, which proves the implication (2) and equation (1).

(3) If in equation (1) we take X and Y to be singletons, we see at oncethat ι is a homomorphism with respect to the addition defined on G. Ifone of X, Y is the singleton {0}, it follows that ι(0) is a neutral elementfor the addition in G. Moreover, this addition is associative, because thegroup G is associative: For a, b, c ∈ G, repeated use of equation (1) givesa + (b + c) = sup ι(A(a)) + sup ι(A(b) + A(c)) = sup ι(A(a) + (A(b) +A(c))) = sup ι((A(a)+A(b))+A(c)) = sup ι(A(a)+A(b))+sup ι(A(c)) =(a + b) + c.

(4) Up to now, we have not used the hypothesis that G is Archimedean.This will be crucial for the existence of inverses in G. For a ∈ G, considerthe set A(a) defined above and

A′(a) = { t ∈ G | t ≤ −A(a)} .

We show that a′ := sup A′(a) satisfies a+a′ = ι(0), which by equation (1)is equivalent to sup ι(A(a) + A′(a)) = ι(0).

It is clear from the definitions that A(a) + A′(a) ≤ 0, which impliesthat sup ι(A(a)+A′(a)) ≤ ι(0). In order to prove the converse inequality,we need to show, according to 41.8, that 0 ≤ e for every upper bounde ∈ G of A(a) + A′(a). Now for x ∈ A(a), t ∈ A′(a) we have x + t ≤ e

by the choice of e. Thus, t − e ≤ −x, and since this holds for allx ∈ A(a), t ∈ A′(a), we have shown that A′(a) − e ⊆ A′(a). Inductionyields A′(a)−ne ⊆ A′(a) for all n ∈ N, so that for x ∈ A(a), t ∈ A′(a) weobtain n(−e) ≤ −t−x. Since G is Archimedean, it follows that −e ≤ 0,so that e ≥ 0, which was to be shown. �

Page 259: The Classical Fields

242 Completion

42.4 Theorem: Universal property of group completions Let

ι : G → G and ι′ : H → H be completions of ordered groups G, H.

Then for every order-preserving group isomorphism ϕ : G → H the

unique extension ϕ : G→ H of ϕ to the completions according to 41.6,

that is the order-preserving map satisfying ϕ ◦ ι = ι′ ◦ ϕ, is a group

isomorphism.

Proof For elements a, b ∈ G there are non-empty subsets X, Y ⊆ G

bounded above in G such that a = sup ι(X) and b = sup ι(Y ). Theorder-preserving extension ϕ is a bijection by 41.6; therefore ϕ respectssuprema, e.g., ϕ(a) = sup ϕ(ι(X)). Now we can use 42.2 to obtain thatϕ(a+b) = ϕ(sup(ι(X)+ι(Y ))) = sup ϕ(ι(X +Y )) = sup ι′(ϕ(X +Y )) =sup(ι′(ϕ(X))+ ι′(ϕ(Y ))) = sup ι′(ϕ(X))+sup ι′(ϕ(Y )) = sup ϕ(ι(X))+sup ϕ(ι(Y )) = ϕ(sup ι(X)) + ϕ(sup ι(Y )) = ϕ(a) + ϕ(b). �

42.5 Corollary: Uniqueness of group completions Let ι : G→ G

and ι : G → G be completions of an ordered group G. Then there

is a unique order-preserving group isomorphism ϕ : G → G such that

ϕ ◦ ι = ι.

Proof This is the special case of 42.4 with H = G and ϕ = id. �

Finally, we study the completion of Archimedean ordered fields. Forthe definition of an ordered skew field, see 11.1; the additive group ofan ordered skew field F is an ordered group, and the set P of positiveelements is a subgroup of the multiplicative group and an ordered group,as well (11.5). The characteristic of F is zero (see 11.2), thus the primefield of F is isomorphic to Q and consists of the elements m · (n · 1)−1

for m,n ∈ Z, n �= 0, which we write as m/n for short.

42.6 Lemma If the additive group of an ordered skew field F is Archi-

medean, then the same holds for the group P of positive elements under

multiplication, and F is commutative.

In this case, we call F an Archimedean ordered field .

Proof For 1 < a ∈ F , we have −1 + a > 0. Since the additive group isArchimedean, there is n ∈ N such that n(−1 + a) > 1, that is −1 + a >

1/n, a > 1 + 1/n. It follows for m ∈ N that am > (1 + 1/n)m =1+m/n+ · · · > m/n. Now for 0 < b ∈ K, by the Archimedean propertyagain, we may choose m in such a way that m/n = m · 1/n > b, so thatfinally am > b. This proves that P is Archimedean (and is part of asolution of Exercise 4 in Section 11).

Page 260: The Classical Fields

42 Completion of ordered groups and fields 243

By 7.7, it follows that P is commutative. Hence F× = P × {±1} iscommutative, as well. �

42.7 Definitions An ordered field is called completely ordered if theunderlying chain is complete. A completion of an ordered field F isa completely ordered field F together with an order-preserving fieldmonomorphism ι : F → F whose image ι(F ) is weakly dense in F

(so that ι : F → F is a chain completion of F ).

42.8 Construction Every Archimedean ordered field F has a comple-

tion.

Proof (1) Let ι : F → F be a group completion of the additive groupof F , which is an ordered group; such a group completion exists by42.3. By 7.5 and 7.7, the ordered group F is Archimedean and thereforecommutative; its group operation will be denoted and referred to asaddition.

By identifying the elements of F with their images under ι, we mayassume that F ⊆ F and that ι is the set theoretic inclusion map. Thiswill simplify notation.

We shall extend the multiplication of F to F in order to obtain amultiplication of F which makes F an ordered field. One way wouldbe to write down the multiplication explicitly for suprema of subsetsof ι(F ), similarly as for the addition of an Archimedean ordered groupin 42.3. Here, we choose an alternative way. We use the completionof the ordered group P of positive elements of F , which is a subgroupof the multiplicative group; the completion of P as a chain is just theset of positive elements of F . After extending the multiplication to thewhole of F we use the universal property 42.2 of group completions as aconceptual method for the necessary verifications in order to show thatthe completion is again an ordered field. Now for the details.

(2) The set P = {a ∈ F | a > 0} is a complete chain, since F

is, and P ⊆ P is weakly dense in P . Hence, P is the completion ofthe chain P . By 42.6, the group P is Archimedean, so that is has acompletion as an ordered group according to 42.3. In other words, thereis a multiplication on P extending the multiplication of P . (Of course,this multiplication can be established explicitly as in step (2) of theproof of 42.3 by declaring, for non-empty subsets X and Y of P whichare bounded above, sup X · supY to be sup(X · Y ). However, one canjust as well rely on the existence and uniqueness of the chain completionand the group completion; see 41.7 and 42.5.)

Page 261: The Classical Fields

244 Completion

We now extend the multiplication of P to the whole of F in severalsteps, respecting the multiplication on F . For 0 < a ∈ F , that is, a ∈ P ,we consider the map

λa : F → F : b �→

⎧⎪⎪⎨⎪⎪⎩ab if b > 0

0 if b = 0

−a(−b) if b < 0 .

The restriction of λa to P is an order-preserving bijection of P into itself,since P is an ordered group. For the same reason, and since the bijection

μ : F → F : b �→ −b

reverses the ordering (being the inversion map of the additive group of F ,which is an ordered group), λa induces an order-preserving bijection of−P onto itself. Thus we obtain that λa is an order-preserving bijectionof the whole of F onto itself.

For a < 0, we define

λa := μ ◦ λ−a ,

obviously an order-reversing bijection of F onto itself. Using the mapsλa, we construct the multiplication of F as follows:

ab :=

{λa(b) if a �= 0

0 if a = 0 .

It is clear from the definition of λa for a > 0 that this multiplicationextends the multiplication of P so that no confusion will arise from usingthe same notation for both multiplications.

Next, it will be shown that

(−a)b = −ab = a(−b) (1)

for all a, b ∈ F . For a = 0, this is trivial. For a �= 0, equation (1) can betranslated into

μ ◦ λa = λ−a = λa ◦ μ (2)

and will be proved in this form. For a > 0, the first equality is just thedefinition of λ−a; for a < 0, by definition, λa = μ ◦ λ−a, from which thefirst equality follows by composition with the involutory map μ. Thesecond equality for a > 0 is obtained directly from the definitions of λa

and of λ−a; for a < 0, we may then conclude that λa ◦μ = μ ◦λ−a ◦μ =μ ◦ λa = λ−a.

Page 262: The Classical Fields

42 Completion of ordered groups and fields 245

(3) We now show that F � {0} with the multiplication defined aboveis a commutative group. Recall first that P is a group under multiplica-tion and is commutative, being a completely ordered group and henceArchimedean; see 7.5 and 7.7. By equation (1), the group propertiesand commutativity carry over directly from P to F � {0}; indeed (1)says that F � {0} is the direct product of P and of the group {1,−1}.

In order to establish that F is a field it remains to prove distributivity.Then F is an ordered field as is immediate from the stated monotonicityproperties of the maps λa.

(4) For u ∈ F , the restriction of λu to F is the map F �→ F : x �→ ux

obtained from the original multiplication in F . This is so because themultiplication of P extends that of P ⊆ F and since equation (1) holdsfor the multiplication of F . Thus, the multiplication of F extends thatof F , and the inclusion map F → F is a homomorphism with respect toboth multiplication and addition.

(5) For two positive elements u, v ∈ F both the maps λu+v and λu+λv

are order-preserving extensions of the order-preserving bijection F → F :x �→ (u+ v)x = ux+ vx. By the universal property 41.6 of completions,these two extensions coincide, which means that

(u + v)a = ua + va (3)

for a ∈ F . For u > v > 0 it follows that (u − v)a + va = ua, so that(u− v)a = ua− va. From this and equation (3) and using equation (1)one obtains that equation (3) holds for arbitrary u, v ∈ F , which meansthat for fixed a ∈ F �{0}, the set Fa is a subgroup of the additive groupof F and that the bijection

λa : F → F : b �→ ab = ba

induces an isomorphism of the additive group (F,+) onto (Fa,+).For a > 0, the map λa is order-preserving. Since P is weakly dense

in P , we infer that Fa is weakly dense in F , so that F is a completionof both the ordered groups F and Fa. By the universal property 42.4of group completions, λa is a group automorphism of (F ,+), being anextension of an order-preserving group isomorphism F → Fa.

For a < 0, we thus know that λ−a is an automorphism of the additivegroup. The same is true for the inversion map μ, since addition iscommutative; hence λa = μ ◦ λ−a is an automorphism.

Thus the distributive law a(b + c) = ab + ac holds for all a, b, c ∈ F .(Again, the case a = 0 needs a separate, but trivial verification.) All theproperties of a field completion are now established. �

Page 263: The Classical Fields

246 Completion

42.9 Uniqueness theorem Up to isomorphism, there is exactly one

completely ordered field. In other words, for any two completely ordered

fields F1 and F2, there is an order-preserving field isomorphism of F1

onto F2.

Remark An ordered field F is a topological field with the topologyinduced by the ordering. If F is completely ordered, F is also completeas a topological field (43.10). The converse does not hold in general, butis true for Archimedean ordered fields; see 43.29. Thus, Theorem 42.9may be rephrased as follows. Up to isomorphism, there is exactly oneArchimedean ordered field which is complete as a topological field. AnArchimedean ordered field is complete as a topological field if and onlyif every Cauchy sequence converges. Recall that order completenessimplies the Archimedean property. For a comprehensive discussion ofthese matters, see Prieß-Crampe 1983 Chapter III Section 1, Blyth

2005 Section 10.2 or Dales–Woodin 1996 Section 3.

The proof of Theorem 42.9 will make use of the following lemma.

42.10 Lemma In an Archimedean ordered field, the prime field is

weakly dense.

Remark The assertion of Lemma 42.10 is contained in Theorem 11.14.But there, the field of real numbers is involved; also, the assertion of42.10 is obtained rather indirectly there, for the sake of further relatedresults. Here, we give a direct proof which avoids the use of real numbers,as one of the aims of the present section is the construction of the realnumbers; see 42.11.

Proof Let F be an Archimedean ordered field and a ∈ F . Since F isArchimedean, every element lies between two elements of the prime field.Hence it suffices to show that for all a, b ∈ F such that a < b there isan element q of the prime field such that a < q < b. We may assumethat b is positive; if not, we apply the order-reversing bijection x �→ −x.Furthermore, we may assume that a ≥ 0; else we replace a by 0.

Then, since F is Archimedean, there is n ∈ N such that n·1 > (b−a)−1,so that 1/n < b−a, and likewise there is m ∈ N such that (m+1)·1/n ≥ b.If m is the smallest natural number with this property, then m · 1/n < b

and m · 1/n ≥ b − 1/n > a. Thus, m · 1/n is an element of the primefield lying between a and b. �

Proof of Theorem 42.9. There is a completely ordered field, for instancethe field completion according to 42.8 of the ordered field Q.

Page 264: The Classical Fields

42 Completion of ordered groups and fields 247

Now let F1 and F2 be completely ordered fields; we show that they areisomorphic. Let Q1, Q2 be the prime fields of F1 and F2. Since a com-pletely ordered field is Archimedean, these prime fields are weakly denseby 42.10, so that F1 and F2 are field completions of Q1 and Q2. Theprime fields are isomorphic to Q, and the natural bijection ϕ : Q1 → Q2

mapping the element m/n = m · (n · 1)−1 of Q1 to the correspond-ing element of Q2 is order-preserving and a field isomorphism. By theuniversal properties 41.6 and 42.4 of completions, ϕ extends to an order-preserving bijection ϕ : F1 → F2 which is a group isomorphism of theadditive groups of F1 and F2.

The set P1 of positive elements of F1 is a group under multiplication.It is also completely ordered, and the set Q1 ∩P1 of positive elements ofQ1 is weakly dense in P1. The same holds for the group P2 of positiveelements of F2, so that P1 and P2 are group completions of Q1 ∩ P1

and Q2 ∩ P2, respectively. The natural map ϕ : Q1 → Q2 induces anorder-preserving group isomorphism between the groups Q1 ∩ P1 andQ2 ∩ P2. The extension P1 → P2 of this map obtained by restrictingϕ therefore is an isomorphism (with respect to multiplication) by theuniversal property 42.4 of group completions. Since every product ofnon-zero elements of F1 can be reduced to a product of positive elementsup to sign, it follows that ϕ respects multiplication not only on P1,but throughout F1. We conclude that ϕ is an order-preserving fieldautomorphism. �

42.11 Construction of the real numbers On the basis of the unique-ness theorem 42.9, we may now define the field R of real numbers to beany completely ordered field. Since this determines R only up to iso-morphism, one may wish to specify a concrete construction of such acompletely ordered field. For instance, one may define the underlyingchain to be the completion of Q obtained by the explicit constructiondescribed in 41.3. This chain is the underlying chain for a group comple-tion of (Q,+) obtained by 42.3, which will be the additive group of R.On this completely ordered group, a multiplication is established accord-ing to the construction in 42.8, which finally produces the completelyordered field R.

In Chapter 1 of this book, the known properties of the real numberswere used without further justification, but these properties just expressthe fact that R is a completely ordered field, or can be derived easilyfrom this. Sometimes it is even helpful to recall that this comprisesall we know about R. For instance, 11.8 can be seen as an immediate

Page 265: The Classical Fields

248 Completion

corollary of the uniqueness theorem 42.9 for completely ordered fields;in Section 11, it was proved differently. Another example is Corollary42.12 below.

In the present section, we have sometimes used arguments from Chap-ter 1, in particular the simple facts about ordered groups and orderedfields from Sections 7 and 11. We have been careful, however, to useonly such arguments which in Chapter 1 are derived directly from thedefinitions of these structures, without making use of the real numbers.So, our conclusive construction of the real numbers based on these ar-guments does not suffer from a vicious circle.

For a different approach to the completion of ordered fields and theconstruction of the ordered field R avoiding many of the verifications in42.8 see Banaschewski 1998.

In the next two sections, topological completion methods are discussedwhich offer an alternative construction of R from Q. Another possibilityfor the construction of R using non-standard methods has already beendescribed in Section 23; see also the introduction to Chapter 1.

42.12 Corollary Every Archimedean ordered field admits an order-

preserving isomorphism onto a subfield of R endowed with the induced

ordering.

Proof The completion of an Archimedean ordered field is (isomorphicto) the field of real numbers by 42.9. �

Remark In Theorem 11.14, this is the implication (d) ⇒ (a). Theproof there uses the same ideas as here in an ad hoc manner.

43 Completion of topological abelian groups

A metric space is said to be complete if every Cauchy sequence in thisspace has a limit point. Recall that a Cauchy sequence is a sequence(xν)ν with the property that the diameters of the sequence ‘tails’ {xν |ν ≥ n} become arbitrarily small for large n ∈ N. Such a notion can onlybe expressed if one has the possibility of comparing the size of subsets atdifferent places of the space. A metric offers such a possibility, whereasthe notion of a topological space does not.

In a topological group, however, the concept of a Cauchy sequencecan be formulated adequately, since the neighbourhoods of an arbitraryelement are obtained by translation from the neighbourhoods of theneutral element. For topological groups which do not have countable

Page 266: The Classical Fields

43 Completion of topological abelian groups 249

neighbourhood bases, however, Cauchy sequences are not sufficient toexpress the property of completeness; instead of sequences, one has touse nets or, as we do here, filterbases.

In this section, we show that every Hausdorff abelian topological groupG can be embedded as a dense subgroup into a Hausdorff abelian topo-logical group G which is complete in the sense that every Cauchy filter-base has a limit point; G will be called the completion of G.

Our interest is not so much the completion of topological groups butrather the completion of topological fields, the completion of their ad-ditive groups being the first step; see Section 44. For this reason, wedeal with abelian topological groups only. It is not difficult, however, totreat completion of topological groups in general on the same lines; seeWarner 1989, Stroppel 2006 and the hints in 43.27.

43.1 Filterbases and filters A filterbase on a set X is a non-emptyset B of non-empty subsets of X such that the intersection of any twomembers of B contains a member of B. As we know from 21.1, a filteris a filterbase F such that every subset of X containing a member of F

belongs to F, as well. In particular, the intersection of any two membersof F then belongs to F. For a filterbase B, the smallest filter containingB as a subset is called the filter generated by B; it consists of the subsetsof X containing a member of B.

For filterbases A, B on X, we say that A is finer than B and writeA < B if every member of B contains a member of A. Equivalently, thismeans that the filter generated by B is a subset of the filter generatedby A. We say that A and B are equivalent if A < B and B < A; this istrue if, and only if, the filters generated by A and B coincide.

A sequence (xν)ν in X determines a filterbase consisting of all ‘tails’{xν | ν ≥ n} for n ∈ N. In this way, filterbases generalize the notion ofa sequence.

43.2 Convergence of filterbases (a) If X is a topological space, wesay that a filterbase B on X converges to x ∈ X, and write B → x,if every neighbourhood of x contains a member of B. With the termsintroduced above, this may be expressed by saying that B is finer thanthe filter Vx of neighbourhoods of x. The same then holds for the filtergenerated by B. The point x is also called a limit point of B. A filterbaseA which is finer than B then converges to x as well.

(b) A sequence (xν)ν in X converges to x in the usual sense if, andonly if, the filterbase consisting of all ‘tails’ {xν | ν ≥ n} for n ∈ Nconverges to x.

Page 267: The Classical Fields

250 Completion

(c) If X is a Hausdorff space, then a filterbase B on X convergingto x ∈ X has no other limit point except x. We may therefore writex = lim B to characterize this situation.

Indeed, assume that there were a second limit point y ∈ X. Then,since B would be finer than Vx and finer than Vy at the same time,every neighbourhood of x would intersect every neighbourhood of y,which contradicts the Hausdorff separation property.

(d) If ϕ : X → Y is a continuous map between two topological spaces,and if B is a filterbase on X which converges to x ∈ X, then it isstraightforward that ϕ(B) = {ϕ(B) | B ∈ B} is a filterbase on Y whichconverges to ϕ(x).

In what follows, G will be an abelian topological group, written addi-tively, with neutral element 0.

43.3 Concentrated filterbases (a) A filterbase C on G is said to beconcentrated if for every neighbourhood U of 0 there is C ∈ C such thatC − C ⊆ U . (Commonly, such filterbases are also called Cauchy filter-bases.) It is clear that the filter generated by a concentrated filterbase C

is concentrated, as well, and so is, more generally, any filterbase whichis finer than C.

(b) A sequence (xν)ν in G is called a Cauchy sequence if for everyneighbourhood U of 0 there is n ∈ N such that for all μ, ν ∈ N satisfyingμ ≥ n and ν ≥ n one has xμ − xν ∈ U . This just means that thefilterbase consisting of all ‘tails’ {xν | ν ≥ n} for n ∈ N is concentrated.

43.4 Neighbourhood filters Let a ∈ G. Continuity of the differencemap G×G→ G : (x, y) �→ x− y at (a, a) says that for every neighbour-hood U of 0 there is a neighbourhood N of a such that N −N ⊆ U , inother words, the neighbourhood filter Va is concentrated.

It follows that a filterbase which converges to some element is con-centrated, since it is finer than the neighbourhood filter of this element.The converse is not true, in general. This motivates the following notion.

43.5 Definition: Completeness The topological group G is calledcomplete if each concentrated filterbase converges. Then, in particular,every Cauchy sequence converges; see 43.3(b) and 43.2(b).

It is clear from the definition that a closed subgroup of a completegroup G is complete, as well.

Trivially, a group G with the discrete topology is a complete topo-logical group. Indeed, a filterbase C on G is concentrated if and only ifsome element of C is a singleton {x}, and then C converges to x.

Page 268: The Classical Fields

43 Completion of topological abelian groups 251

43.6 Lemma A filterbase C on G converges to c ∈ G if, and only if, it

is concentrated and c ∈ ⋂{C | C ∈ C} (the intersection of the closures

of the members of C).In particular, a Cauchy sequence in G converges to c if, and only if, c

is an accumulation point of the sequence.

Proof Assume that C converges to c. We have already noted that aconvergent filterbase is concentrated. Every neighbourhood of c containsa member of C, which in turn has non-empty intersection with everymember C ∈ C, so that c ∈ C.

Conversely, assume that C is concentrated and let c be an element ofthe intersection above. Every neighbourhood N of c contains a neigh-bourhood of the form U + c where U is a neighbourhood of 0. Let V

be a neighbourhood of 0 such that V + V ⊆ U . Since C is concen-trated, there is C ∈ C such that C − C ⊆ V , and (V + c) ∩ C �= ∅since c ∈ C. For x ∈ (V + c) ∩ C, we infer that C − x ⊆ V andC ⊆ V + x ⊆ V + V + c ⊆ U + c ⊆ N . Thus, C→ c. �

43.7 Corollary Let C, D be two filterbases on G such that C < D. If

C converges to c ∈ G and if D is concentrated, then D converges to c,

as well.

Proof By 43.6, we have c ∈ ⋂{C | C ∈ C}. Since C < D, clearly⋂{C | C ∈ C} ⊆ ⋂{D | D ∈ D}. Hence c belongs to the latterintersection, as well, and the assertion follows from 43.6. �

43.8 Lemma Let C be a concentrated filterbase on G. If there is a

member C0 ∈ C with compact closure C0, then C converges.

Proof By the properties of a filterbase, C has the finite intersectionproperty in the sense that any collection of finitely many members of C

has non-empty intersection. The same then is true for the set {C ∩C0 |C ∈ C} of closed subsets of C0. By compactness, therefore, the wholeset has non-empty intersection, which clearly equals

⋂{C | C ∈ C}. Ifthe filterbase C is concentrated, then it converges to every point of thisintersection by 43.6. �

43.9 Corollary A locally compact topological group is complete.

Proof Let U be a compact neighbourhood of 0, and V a neighbourhoodof 0 such that −V + V ⊆ U . A concentrated filterbase C contains amember C0 such that C0 − C0 ⊆ V and hence C0 ⊆ V + c for c ∈ C0.

Page 269: The Classical Fields

252 Completion

For x ∈ C0, the intersection (V +x)∩C0 is not empty, and for an elementd ∈ (V +x)∩C0 we find that x ∈ −V +d ⊆ −V +V + c ⊆ U + c. HenceC0 is contained in the compact subset U + c and therefore is compactitself. By 43.8, we conclude that C converges. �

43.10 Completeness of ordered groups Let G be an ordered group(see 7.1) that is complete as such. It is a topological group with thetopology induced by the ordering; see 8.4. Note that G is commutativeby 7.7.

Proposition If an ordered group G is complete as an ordered group,

it is also complete as a topological group.

Remarks (1) The converse need not be true; see 43.29.(2) The result applies in particular to the additive group of R, if we

think of R as the completion of the ordered field Q; see 42.11.

Proof The assertion is an immediate consequence of 43.9 since G islocally compact according to 5.2.

Because of the importance of our result for the structure of R, we giveanother, more straightforward proof. Let C be a concentrated filterbaseon G. For 0 < a ∈ G, there is C ∈ C such that C − C ⊆ ]−a, a[.For c ∈ C, then, C ⊆ ]−a + c, a + c[. Thus, C has members which arebounded (in the chain G) and hence have a least upper bound and agreatest lower bound, since G is order complete. For bounded membersC,C′ ∈ C, we have inf C ′ ≤ supC, for else C ∩C ′ = ∅, contradicting theproperties of a filterbase. Thus the set { supC | C ∈ C, C bounded} isbounded below and has a greatest lower bound

a := inf{ supC | C ∈ C, C bounded} .

We show that C converges to a. Let C ∈ C; by 43.6 if suffices to showthat a ∈ C. For 0 < ε ∈ G there is a bounded member B ∈ C suchthat a ≤ supB < a + ε. We may assume that B ⊆ C by making B

smaller if necessary. There is b ∈ B such that supB − ε < b ≤ supB,and then a − ε < b < a + ε. Since b ∈ C, this shows that C intersectsevery neighbourhood of a, so that a ∈ C. �

As we have remarked at the beginning of the preceding proof, thecomplete ordered group R is locally compact; see also Theorem 5.3.Since the multiplicative group R�{0} and the subgroup Rpos of positivereal numbers are open in R, they are locally compact as well. Thus,Corollary 43.9 implies the following.

Page 270: The Classical Fields

43 Completion of topological abelian groups 253

43.11 Examples The two groups R � {0} and Rpos are complete

topological groups. For R � {0}, this might seem strange on first sightsince there is a ‘hole’. Completeness of R � {0} as a topological groupdoes not mean completeness as a metric space with the usual metricinherited from R.

Concentrated filterbases are preserved under continuous homomor-phisms and, more generally, under maps of the following type.

43.12 Uniform continuity A map ϕ : G → H between topologicalgroups (not necessarily a homomorphism) is said to be uniformly con-tinuous if for every neighbourhood V of 0 in H there is a neighbourhoodU of 0 in G such that for all x ∈ G one has ϕ(U + x) ⊆ V + ϕ(x).

This is obviously satisfied if ϕ is a continuous homomorphism: chooseU in such a way that ϕ(U) ⊆ V .

If ϕ is uniformly continuous and C a concentrated filterbase on G,then the filterbase ϕ(C) on H is concentrated, as well. Indeed, for aneighbourhood V of 0 in H choose a neighbourhood U of 0 in G asabove and a member C ∈ C such that C − C ⊆ U . Then for all c ∈ C

one has C ⊆ U + c and consequently ϕ(C) ⊆ ϕ(U + c) ⊆ V + ϕ(c), sothat ϕ(C)− ϕ(C) ⊆ V .

In particular, the image sequence of a Cauchy sequence in G underthe uniformly continuous map ϕ is a Cauchy sequence again.

43.13 Definition An embedding of a topological group G into a topo-logical group H is a monomorphism ϕ : G→ H which induces a homeo-morphism of G onto the image group ϕ(G) endowed with the topologyinduced from H.

43.14 Definition A completion of a topological group G is a completetopological group G together with an embedding ι : G → G such thatι(G) is dense in G. A Hausdorff completion is a completion that is aHausdorff group. Of course, a Hausdorff completion will only exist forHausdorff topological groups.

For every abelian topological group G, a complete abelian topologicalgroup will be constructed below, which will give a completion if G is aHausdorff group. The elements of the complete group constructed fromG will be certain concentrated filters on G; after all, if such a filter doesnot converge in G, why not add it to the elements of G and let it convergeto itself? We do not use all concentrated filters, however, as differentfilters may converge to the same element. We focus on concentratedfilters of a special type which we discuss now.

Page 271: The Classical Fields

254 Completion

43.15 Minimal concentrated filters Let G be a topological group,

V0 the filter of neighbourhoods of 0, and C a concentrated filterbase

on G. Then V0 + C := {U + C | U ∈ V0, C ∈ C} is a concentrated

filterbase, as well.

Let C be the filter generated by V0 + C (which is concentrated, too).

Then C < C; moreover, if D is any concentrated filterbase such that

C < D, then D < C.

Remarks The last statement, when expressed for a concentrated filterF instead of the filterbase D, says that C < F implies C ⊆ F. In otherwords, C is the smallest filter among the concentrated filters F such thatC < F holds.

Note that the existence of filters having this minimality property isguaranteed by the explicit construction of C, not just by a transfiniteprinciple like Zorn’s lemma.

When working with sequences or nets instead of filterbases, an analo-gous construction is not possible; it would amount to producing a small-est subnet from a Cauchy net.

Proof For C1, C2 ∈ C there is C ∈ C such that C ⊆ C1 ∩ C2. ForU1, U2 ∈ V0 it is then clear that (U1 ∩U2) + C ⊆ (U1 + C1)∩ (U2 + C2).This shows that V0 + C is a filterbase.

In order to verify that V0+C is concentrated, let U be a neighbourhoodof 0. There is a neighbourhood V of 0 and a member C ∈ C such thatV +V −V ⊆ U and C−C ⊆ V . Then (V +C)−(V +C) = V +C−C−V ⊆V + V − V ⊆ U . Thus the filterbase V0 + C is concentrated, and so isthe filter C generated by it.

For U ∈ V0 and C ∈ C clearly C ⊆ U + C; hence C < V0 + C < C.For a concentrated filterbase D such that C < D we have to show that

every element U +C ∈ V0+C ⊆ C contains a member of D. Find D ∈ D

such that D−D ⊆ U . Since D contains a member of C, it intersects C;let d ∈ C ∩D. Then D ⊆ U + d ⊆ U + C. �

43.16 Lemma Let C and D be concentrated filterbases on a topological

group. If C < D, then C = D. In particular,C = C.

Proof If C < D, the minimality principle of 43.15 says that D < C,which implies that C < D. Since C < C, it follows that C < D. Thesame minimality principle yields D < C; thus the two filters D and C

are equivalent as filterbases and hence coincide, which proves the firststatement. The second statement follows since C < C. �

Page 272: The Classical Fields

43 Completion of topological abelian groups 255

43.17 Neighbourhood filters again For an element x of a topo-

logical group, the set consisting of just the singleton {x} obviously is

a concentrated filterbase, and {{x}} = Vx, the neighbourhood filter of

the element x. Moreover, a concentrated filterbase C converges to x if,

and only if, C = Vx.

Proof By definition, {{x}} is the filter which is generated by the filter-base V0 + {x}, which is just the neighbourhood filter Vx. A concen-trated filterbase C converges to x precisely if it is finer than Vx; thesecond statement now follows from 43.16. �

More generally, the filters constructed in 43.15 can be perceived asdistinguished representatives of equivalence classes of filterbases for acertain equivalence relation which we now elucidate.

43.18 Lemma For two concentrated filterbases C1,C2 on a topological

group the following statements are equivalent.

(i) C1 = C2

(ii) C ′1 ∩ C′

2 �= ∅ for all C′1 ∈ C1, C

′2 ∈ C2

(iii) The set C := {C1 ∪ C2 | C1 ∈ C1, C2 ∈ C2 } (which clearly is a

filterbase again) is concentrated.

Proof It is clear that C is a filterbase since C1 and C2 are; and obviously,(i) implies (ii).

(ii) implies (iii): For a neighbourhood U of 0 let V be a neighbourhoodof 0 such that V − V ⊆ U . Since C1 and C2 are concentrated, thereare C ′

1 ∈ C1, C′2 ∈ C2 such that C ′

1 − C ′1 ⊆ V and C′

2 − C′2 ⊆ V .

By (ii), there is an element c ∈ C ′1 ∩ C′

2. Then C ′1 ∪ C′

2 ⊆ V + c and(C ′

1∪C ′2)−(C ′

1∪C′2) ⊆ V +c−(V +c) = V −V ⊆ U . There are C1 ∈ C1

and C2 ∈ C2 such that C1 ⊆ C ′1, C2 ⊆ C′

2, and (C1∪C2)−(C1∪C2) ⊆ U ,as well. This proves (iii).

(iii) implies (i): Clearly C1 < C, and C < C by 43.15. By 43.16, hence,C1 = C. In the same way one obtains C2 = C, and (i) is proved. �

Now we provide two technical tools.

43.19 Lemma Let ι : G→ G be an embedding of a topological group

G into a topological group G such that ι(G) is dense in G. Then G is

complete if only for every concentrated filterbase C on G, the filterbase

ι(C) on G converges.

Remark This criterion makes it easier to prove completeness: one doesnot have to consider all concentrated filterbases on G, but only thosewhich come from filterbases on G via ι.

Page 273: The Classical Fields

256 Completion

Proof By V0 we denote the filter of neighbourhoods of 0 in G. Let D

be a concentrated filterbase on G; we have to show that it converges.It is finer than the filterbase V0 +D formed according to 43.15, which

is concentrated, as well. A member V + D of this filterbase for V ∈ V0,D ∈ D is a neighbourhood of every element of D. Since ι(G) is dense inG, the intersection (V +D)∩ ι(G) is non-empty, and (V0 +D)∩ ι(G) :={ (V + D) ∩ ι(G) | V ∈ V0, D ∈ D} is a filterbase which is finer thanV0 + D. Since ι−1 : ι(G) → G is a continuous homomorphism byassumption, the filterbase

ι−1((V0 + D) ∩ ι(G)) = { ι−1((V + D) ∩ ι(G)) | V ∈ V0, D ∈ D}on G is concentrated. By assumption, its image filterbase, which is just(V0 +D)∩ ι(G), converges in G. Since it is finer than V0 +D, the latterconverges as well by 43.7.

Thus the filterbase D, which is finer than V0 + D, converges too. �

The next result will help to extend algebraic properties to a comple-tion. The formulation covers a sufficiently general situation for later usein the completion of topological rings.

43.20 Lemma Let (G, ∗) and (G′, ∗′) be topological spaces with con-

tinuous binary operations, and ι : G→ G′ a continuous homomorphism

such that ι(G) is dense in G′. Moreover, G′ is assumed to be a Hausdorff

space.

(i) If the operation ∗ on G is associative or commutative, then so is the

operation ∗′ on G′.(ii) If e is a neutral element of (G, ∗), then ι(e) is a neutral element of

(G′, ∗′).(ii′) Assume in situation (ii) that there is a continuous map ν : G → G

such that for x ∈ G the image ν(x) is an inverse of x in the sense

that x ∗ ν(x) = e. Assume furthermore that there is a continuous

map ν′ : G′ → G′ such that ν′ ◦ ι = ι ◦ ν.

Then for every x′ ∈ G′ the image ν′(x′) is an inverse of x′ as well,

that is x′ ∗′ ν(x′) = ι(e).

Proof (i) Let ∗ be associative. We want to prove that ∗′ is associativeas well, which means that the two continuous maps

G′ ×G′ ×G′ → G′ : (x′, y′, z′) �→ (x′ ∗′ y′) ∗′ z′G′ ×G′ ×G′ → G′ : (x′, y′, z′) �→ x′ ∗′ (y′ ∗′ z′)

coincide. Since G′ is Hausdorff, the set of coincidence is closed. There-fore it suffices to show that the two maps coincide on the dense subset

Page 274: The Classical Fields

43 Completion of topological abelian groups 257

ι(G)× ι(G)× ι(G). But this is associativity on ι(G), which holds sinceassociativity of G is transported by the homomorphism ι. In exactly thesame way one proves that ∗′ is commutative if ∗ is.

(ii) Assume that e is a neutral element of (G, ∗). To prove that ι(e)is a neutral element of (G′, ∗′) means to show that the continuous mapG′ → G′ : x′ �→ x′ ∗′ ι(e) coincides with the identity map. As in (i), itsuffices to have coincidence on the dense subset ι(G); but this is clearsince ι is a homomorphism.

(ii′) Here, again, we deduce the coincidence of the two continuousmaps G′ → G′ : x′ �→ x′ ∗′ ν(x′) and G′ → G′ : x′ �→ ι(e) from the factthat they coincide on the dense subset ι(G), as can be easily seen: indeed,for x ∈ G one has ι(x)∗′ν′(ι(x)) = ι(x)∗′ ι(ν(x)) = ι(x∗ν(x)) = ι(e). �

43.21 Construction Every abelian topological Hausdorff group has

an abelian Hausdorff completion.

Proof (1) Let G be an abelian topological group (it need not be a Haus-dorff group for the moment). We shall construct a completion G expli-citly. The underlying set will consist of the minimal concentrated filterson G constructed in 43.15:

G = { C | C concentrated filterbase on G} .

By 43.17 there is a natural map

ι : G→ G : x �→ {{x}} = Vx .

(2) A topology on G will be defined using the following sets as a basis.For A ⊆ G, let

A = { C ∈ G | A ∈ C} .

We remark that by construction every member of a minimal concen-trated filter C contains an open subset. Hence A is non-empty if, andonly if, A contains an open subset; for instance, x is contained in theinterior of A if, and only if, Vx ∈ A.

One verifies directly for a subset B ⊆ G that A ∩B = A ∩ B sincethe elements of G are filters and not just filterbases. Thus, there is atopology on G having { A | A ⊆ G} as a basis. A neighbourhood baseof C ∈ G is given by { B | B ∈ C}.

The space G is a Hausdorff space whether G is a Hausdorff space ornot. Indeed, for C, D ∈ G, if C �= D then by 43.18(ii) there are C ∈ C

and D ∈ D such that C ∩D = ∅, and C, D are disjoint neighbourhoodsof C and D, respectively.

Page 275: The Classical Fields

258 Completion

(3) Properties of ι : G → G. The map ι is continuous. For this,it suffices to show that the preimage of every basic open set A is open.Now we have seen above that ι−1(A) = {x ∈ G | Vx ∈ A} is the interiorof A, which is open.

In order to see that ι(G) is dense in G, we verify that every non-emptybasic open set A intersects ι(G). As remarked above, the interior of A isnot empty, and for an element x of the interior one has ι(x) = Vx ∈ A.

The map ι is injective if G is a Hausdorff group. We show that then thebijection G→ ι(G) induced by ι is a homeomorphism. As ι is continuous,it remains to show that this map is open. For an open subset U ⊆ G

the image set ι(U) = {Vx | x ∈ U } = {Vx | U ∈ Vx } = U ∩ ι(G) is anopen subset of ι(G) in the topology induced from G.

(4) Group structure on G. We now define an addition which makes G

an abelian topological group. For C, D ∈ G, it is clear that

C⊕ D := {C + D | C ∈ C, D ∈ D}is a filterbase. We verify that this filterbase is concentrated since C, D

are. Indeed, for a neighbourhood U of 0 in G let V be a neighbourhoodof 0 such that V +V ⊆ U . There are C ∈ C, D ∈ D such that C−C ⊆ V

and D−D ⊆ V . Then (C+D)−(C+D) = C−C+D−D ⊆ V +V ⊆ U ;here we have used that G is abelian.

Now we may define addition on G by

C + D := C⊕ D ,

and we show that the map ι : G→ G is a homomorphism. For x, y ∈ G

we have to verify that Vx ⊕Vy = Vx+y. By 43.16, it suffices to showthat Vx ⊕Vy < Vx+y. By continuity of addition in G, for W ∈ Vx+y

there are neighbourhoods U ∈ Vx, V ∈ Vy such that U + V ⊆ W ; thisis our claim.

Next we prove that addition in G is continuous. Let A be a basicneighbourhood of C + D, which means that A ∈ C + D. By definition ofthe addition of G, there are C ∈ C, D ∈ D and a neighbourhood U of 0 inG such that U +C+D ⊆ A. The sets C, D are neighbourhoods of C andD, respectively. Continuity of addition is proved if for all X ∈ C, Y ∈ D

we show that X + Y belongs to the given neighbourhood A of C + D.Now C ∈ X, D ∈ Y, so that U +C +D ∈ X+ Y. Since U +C +D ⊆ A,it follows that A ∈ X + Y, in other words, that X + Y ∈ A, indeed.

By the general lemma 43.20(i, ii), addition on G inherits associativityand commutativity from the addition on G, and ι(0) is a neutral elementof G.

Page 276: The Classical Fields

43 Completion of topological abelian groups 259

Thus, in order to prove that G, like G, is a topological abelian group,we finally have to show that every element C ∈ G (for a concentratedfilter C on G) has an inverse which depends continuously on C. Since G

is abelian, the map

ν : G→ G : x→ −x

is a homomorphism, and therefore −C = {−C | C ∈ C} is concentrated(43.12). Since ν permutes the set of neighbourhoods of 0 in G, it is clearfrom the construction in 43.15 that −C = −C; in particular, this is anelement of G again. We verify that the map

ν : G→ G : C �→ −C

is continuous. It suffices to show that the preimage ν−1(A) of a basicopen set A for A ⊆ G is open. Now

C ∈ ν−1(A) ⇐⇒ −C ∈ A ⇐⇒ A ∈ −C ⇐⇒ −A ∈ C ⇐⇒ C ∈ −A ,

so that ν−1(A) = −A is a basic open set, again.The two maps ν and ν are linked via ι in the sense that ι ◦ ν = ν ◦ ι.

For x ∈ G, indeed, ι(ν(x)) = V−x consists of the neighbourhoods of −x,and the involutory homeomorphism ν exchanges the neighbourhoods ofx and of −x, so that V−x = −Vx = ν(ι(x)).

From these properties of ν, we infer by the general lemma 43.20(ii′)that ν(C) = −C is an inverse of C in G; and we have shown that itdepends continuously on C. Thus, G is an abelian topological group.

(5) Completeness. In order to prove that the topological group G iscomplete, we use the criterion 43.19. Let C be a concentrated filterbaseon G; we show that ι(C) converges in G to C, or equivalently, that forevery A ∈ C the basic neighbourhood A of C contains a member of ι(C).According to the construction of C, there are C ∈ C and a neighbourhoodU of 0 in G such that U + C ⊆ A. In particular, A is a neighbourhoodof every element c ∈ C, that is A ∈ Vc. Hence, ι(c) = Vc ∈ A for allc ∈ C, so that ι(C) ⊆ A.

Thus G has every property of a completion of G. �

The following is a technical preparation for our further study of com-pletions.

43.22 Special filterbases for dense embeddings Let ι : G→ G bean embedding of a topological group G into a topological group G suchthat ι(G) is dense in G. For x ∈ G, we consider the neighbourhood filterVx of x. For N ∈ Vx, the intersection N ∩ ι(G) is non-empty since ι(G)

Page 277: The Classical Fields

260 Completion

is dense, and the same is true for the preimage ι−1(N). Hence we havea filterbase

Vx ∩ ι(G) := { N ∩ ι(G) | N ∈ Vx }on G, which converges to x, since it is finer than Vx; in particular, itis concentrated. Hence, when Vx ∩ ι(G) is considered as a filterbase onι(G), it is concentrated, as well. Furthermore, we have a filterbase

Wx := ι−1(Vx) = { ι−1(N) | N ∈ Vx }on G. Now ι−1(N) = ι−1(N ∩ ι(G)) for N ⊆ G, so that

Wx = ι−1(Vx) = ι−1(Vx ∩ ι(G)) .

This filterbase is also concentrated, since by assumption ι−1 : ι(G)→ G

is continuous and hence uniformly continuous; see 43.12.Conversely, ι(ι−1(N)) = N ∩ ι(G)) for N ⊆ G; hence

ι(Wx) = ι(ι−1(Vx)) = Vx ∩ ι(G)→ x .

43.23 Theorem: Extensions of uniformly continuous maps Let

G and G be topological groups together with an embedding ι : G → G

such that ι(G) is dense in G, let H be a complete Hausdorff topological

group, and let ϕ : G→ H be a uniformly continuous map.

Then ϕ has a unique continuous extension ϕ : G→ H over ι, that is,

a continuous map such that ϕ ◦ ι = ϕ, and ϕ is uniformly continuous,

as well. This holds in particular if ϕ is a continuous homomorphism; in

this case, the extension ϕ is also a homomorphism.

Proof (1) We first prove uniqueness of the extension, using the filterbasesobtained in 43.22 from the neighbourhood filters Vx of elements x ∈ G.Since Vx∩ι(G)→ x, continuity implies that ϕ(Vx∩ι(G))→ ϕ(x). Nowϕ(Vx ∩ ι(G)) = ϕ(ι(ι−1(Vx))) = ϕ(ι−1(Vx)), so that

ϕ(ι−1(Vx))→ ϕ(x) . (1)

Thus, since in Hausdorff spaces filterbases have at most one limit point,ϕ is uniquely determined.

(2) We prove that a map ϕ : G → H having property (1) for everyx ∈ G is uniformly continuous, without using any other property of ϕ.

Let V1 be a neighbourhood of 0 in H. We claim that there is aneighbourhood U of 0 in G such that ϕ(U + x) ⊆ V1 + ϕ(x) for all x ∈ G.Choose a neighbourhood V2 of 0 in H such that −V2+V2+V2 ⊆ V1. Sinceϕ is uniformly continuous, there is a neighbourhood V of 0 in G such

Page 278: The Classical Fields

43 Completion of topological abelian groups 261

that ϕ(V +x) ⊆ V2 +ϕ(x) for all x ∈ G. As ι induces a homeomorphismG→ ι(G), there is a neighbourhood U1 of 0 in G whose preimage underι satisfies ι−1(U1) ⊆ V . Let U2 be a neighbourhood of 0 in G such thatU2 − U2 ⊆ U1; in addition, we may assume U2 to be open in G.

Now let x ∈ G and y ∈ U2 + x. By (1), there is a neighbourhoodN of y in G such that ϕ(ι−1(N)) ⊆ V2 + ϕ(y) and N ⊆ U2 + x. Fory ∈ ι−1(N) ⊆ ι−1(U2 + x) then ϕ(y) ∈ V2 + ϕ(y). In the same way,for x instead of y, we find x ∈ ι−1(U2 + x) such that ϕ(x) ∈ V2 + ϕ(x).Then ι(y) − ι(x) ∈ U2 − U2 ⊆ U1, so that y ∈ ι−1(U1) + x ⊆ V + x

and ϕ(y) ⊆ V2 + ϕ(x). Hence ϕ(y) ∈ −V2 + ϕ(y) ∈ −V2 + V2 + ϕ(x) ⊆−V2 + V2 + V2 + ϕ(x) ⊆ V1 + ϕ(x). Since y ∈ U2 + x was arbitrary,we have obtained that ϕ(U2 + x) ⊆ V1 + ϕ(x), which is our claim withU = U2.

(3) In order to prove that an extension ϕ with the required propertiesexists, we use (1) to define a map ϕ. Indeed the filterbase in (1) isconcentrated, since ι−1(Vx) is concentrated (see 43.22) and since ϕ isuniformly continuous. In the complete Hausdorff topological group H,this filterbase converges to a unique element, which is defined to be theimage point ϕ(x) of x, for every element x ∈ G. According to step (2),the map ϕ is uniformly continuous.

We now verify that ϕ ◦ ι = ϕ. For x ∈ G, by continuity of ι, theneighbourhood filter Vx of x is finer than ι−1(Vι(x)), so that ϕ(Vx)is finer than ϕ(ι−1(Vι(x))). By (1), consequently, ϕ(Vx) → ϕ(ι(x)).On the other hand, clearly Vx → x and hence ϕ(Vx) → ϕ(x). Thusϕ(ι(x)) = ϕ(x), again since H is a Hausdorff group.

(4) Now let ϕ be a continuous homomorphism. Then ϕ is uniformlycontinuous, with a continuous extension ϕ as above. We show that inthis case ϕ is also a homomorphism, by the density argument employedin 43.20. The continuous maps G × G → G : (x, y) �→ ϕ(x + y) andG× G→ G : (x, y) �→ ϕ(x) + ϕ(y) coincide on the dense subgroup ι(G),since ι and ϕ are homomorphisms; hence they coincide everywhere, andϕ is a homomorphism. �

43.24 Corollary: Universal property of completions Let Gi for

i = 1, 2 be a Hausdorff topological group, and ιi : Gi → Gi a Hausdorff

completion of Gi.

Then every continuous homomorphism ϕ : G1 → G2 extends to a

unique continuous map ϕ : G1 → G2 such that ϕ ◦ ι1 = ι2 ◦ ϕ, and ϕ is

a homomorphism.

If ϕ is an isomorphism of topological groups, then so is ϕ.

Page 279: The Classical Fields

262 Completion

Proof Existence and uniqueness of ϕ is an immediate consequence of43.23 with G = G1, G = G1, H = G2, and the map ι2 ◦ ϕ instead of ϕ.

Now assume that ϕ is an isomorphism, and let ψ : G2 → G1 be theunique extension of the continuous homomorphism ψ := ϕ−1 : G2 → G1

satisfying ψ ◦ ι2 = ι1 ◦ ψ. The homomorphism ϑ := ψ ◦ ϕ : G1 → G1

is the extension of the identity map idG1 : G1 → G1 in the sense thatϑ ◦ ι1 = ι1 ◦ idG1 , but the same holds also for the identity map of G1

instead of ϑ. Hence, again by uniqueness of extensions, we obtain thatψ◦ϕ = id

bG1. Likewise ϕ◦ψ : G2 → G2 is the extension of idG2 , and hence

equals the identity map of G2. Thus the continuous homomorphism ψ isinverse to the homomorphism ϕ, which therefore is an isomorphism. �

The universal property above, applied in the special case G1 = G2 andϕ = id, immediately shows that Hausdorff completions are essentiallyunique.

43.25 Corollary: Uniqueness of completion Let ι : G → G and

ι : G→ G be Hausdorff completions of a Hausdorff topological group G.

Then there is a unique isomorphism ϕ : G → G of topological groups

such that ϕ ◦ ι = ι. �

The following is a slight refinement.

43.26 Corollary Let ι : G → G be a Hausdorff completion of a

Hausdorff topological group G. Then the topology of G is minimal

among the topologies which make G a Hausdorff topological group and

ι an embedding.

Proof Let G be obtained from G by endowing it with such a topologythat is coarser. We have to show that this topology is in fact the originaltopology of G. Since the topology of G is coarser, ι(G) is still densein G. We show that G is complete. It suffices to restrict attention to aconcentrated filterbase C on G and to show that ι(C) converges in G; see43.19. Now ι(C) converges in G, hence it converges in G since the latterhas a coarser topology. Thus, ι : G → G is a completion of G, as well.We apply 43.24 in the special case G1 = G2 = G and ϕ = id : G → G

obtaining that there is a unique continuous homomorphism ϕ : G → G

such that ϕ ◦ ι = ι. But of course the identity map id of G is such ahomomorphism, so that ϕ = id. The last assertion of 43.24 finally saysthat this is an isomorphism of topological groups, so that the topologiesof G and G coincide. �

Page 280: The Classical Fields

43 Completion of topological abelian groups 263

The following will be used in Section 44 in order to identify comple-tions.

43.27 Lemma Let (Gi)i∈I be a family of complete topological groups.

Then the direct product group ×i∈I Gi with the product topology is

also a complete topological group.

Proof For k ∈ I, let πk :×i∈I Gi → Gk : (xi)i∈I �→ xk be the canonicalprojection. For x, y ∈ ×i∈I Gi, the projection πk(x + y) = πk(x) +πk(y) depends continuously on x and y, and πk(−x) = −πk(x) dependscontinuously on x, so that the maps (x, y) �→ x + y and x �→ −x arecontinuous. Thus×i∈I Gi is a topological group.

Now let C be a concentrated filterbase on×i∈I Gi. Then for i ∈ I, thefilterbase πi(C) is concentrated as well by 43.12 since πi is a continuoushomomorphism. By the completeness assumption, πi(C) converges to anelement ci ∈ Gi. We show that C converges to c = (ci)i∈I ; then we haveproved that×i∈I Gi is complete. A neighbourhood U of c contains aneighbourhood of the form×i∈I Ui where Ui is a neighbourhood of ci

and I ′ = { i ∈ I | Ui �= Gi } is finite. For i ∈ I ′, there is Ci ∈ C such thatπi(Ci) ⊆ Ui. Let C be a member of C such that C ⊆ ⋂i∈I′ Ci. Then forall i ∈ I we have πi(C) ⊆ Ui, and C ⊆×i∈I Ui ⊆ U . �

43.28 Completion of non-commutative topological groups Ifthe topological group G is not commutative, then one has to distinguishbetween left concentrated filterbases and right concentrated filterbases.

A filterbase C on G is said to be left concentrated (or right concen-trated) if for every neighbourhood U of 0 there exists C ∈ C such that−C + C ⊆ U (or C − C ⊆ U), respectively. (In Definition 43.3, theright concentrated filters were simply called concentrated.) A filterbaseis called bilaterally concentrated if it is both left and right concentrated.The topological group is called left, right or bilaterally complete if everyleft, right or bilaterally concentrated filterbase, respectively, converges.Note that under inversion left concentrated filterbases are mapped toright concentrated filterbases and vice versa, so that a topological groupis left complete if, and only if, it is right complete.

Dieudonne 1944 gives an example of a Hausdorff topological groupadmitting a right concentrated filterbase C whose image under inversionis not right concentrated, so that C itself is not left concentrated.

Every Hausdorff topological group has a bilateral completion which isessentially unique; see Warner 1989 Theorems 5.9 p. 35 and 5.2 p. 32.With arguments as in the remark on p. 33 following the latter theorem

Page 281: The Classical Fields

264 Completion

and using loc. cit. Theorem 4.4 p. 26 one can conclude that a Hausdorfftopological group with a right concentrated filterbase which is not leftconcentrated (as in the example above) has no right completion.

43.29 Completion of ordered groups as topological groups LetG be an ordered group. Considering it as a topological group (see 8.4),we may ask about its completion.

If the ordered group G is Archimedean, then it has a completionι : G → Gord as an ordered group; see 42.3. The complete orderedgroup Gord with the topology induced by the ordering is also completeas a topological group (43.10). As ι(G) is weakly dense in Gord, it istopologically dense. Thus, in the Archimedean case, ι : G → Gord is acompletion of G also as a topological group.

Now what if G is not Archimedean? In Banaschewski 1957 Section 4pp. 56ff it is proved that G has a completion G→ Gtop as a topologicalgroup, which can be retrieved within the completion ι : G → Gord

of G as an ordered set. Note that Gord is not a group if G is notArchimedean, since complete ordered groups are Archimedean; see 7.5.But Gord is a semigroup, as can be seen from the proof of 42.3, and Gtop

can be obtained as the largest subgroup of this semigroup containingι(G). Thus, Gtop carries an ordering, and it turns out that its topologyis just the topology induced by the ordering and that Gtop is an orderedgroup. If G is not Archimedean, Gtop cannot be complete as an orderedgroup. In particular, this shows that there are ordered groups that arecomplete as topological groups but not as ordered groups.

44 Completion of topological rings and fields

Now, the completion of topological abelian groups will be used for theadditive group of a topological ring. Its multiplication can be extendedto the completion in such a way that a topological ring results.

When this is applied to a topological (skew) field F , the resultingcomplete topological ring F will not necessarily be a (skew) field again.However, we shall see that F is a (skew) field if for instance the topologyof F comes from an absolute value, like the p-adic absolute value on Q;see 36.3, and 55.1 for the general notion.

The completion of the field Q with the usual topology (coming fromthe usual absolute value) is the topological field R. The completion ofthe field Q with the p-adic topology is the field Qp of p-adic numbers,which will be discussed in detail in Sections 51–54.

Page 282: The Classical Fields

44 Completion of topological rings and fields 265

44.1 Definition A topological ring R is said to be complete if itsadditive group (R, +) is a complete topological group.

44.2 Definition Let R be a topological ring. A ring completion (orcompletion for short) of R is a complete topological ring R together withan embedding ι : R→ R of topological rings (that is, a monomorphismof rings which induces a homeomorphism R → ι(R)) such that ι(R) isdense in R.

We shall presently see that for a Hausdorff topological ring such a ringcompletion always exists. First, we need some some technical prepara-tions.

44.3 Lemma Let C be a concentrated filterbase on a topological ring.

Then for every neighbourhood U of 0 there is C ∈ C and another neigh-

bourhood V of 0 such that V C ⊆ U and CV ⊆ U .

Proof By continuity of the operations, there are neighbourhoods U1, U2

of 0 such that U1 + U1 ⊆ U and U2U2 ⊆ U1. Since C is concentrated,there is C ∈ C such that C−C ⊆ U2. Choose c ∈ C; then C ⊆ U2+c. Bycontinuity again, there is a neighbourhood V of 0 such that V c ⊆ U1 andcV ⊆ U1, and we may assume that V ⊆ U2. Then V C ⊆ V (U2 + c) ⊆U2U2 + U1 ⊆ U1 + U1 ⊆ U , and similarly CV ⊆ U . �

44.4 Lemma Let C and D be concentrated filterbases on a topological

ring. Then CD := {CD | C ∈ C, D ∈ D} is a concentrated filterbase, as

well (where of course CD = { cd | c ∈ C, d ∈ D }).Proof It is clear that CD is a filterbase. We have to show that for aneighbourhood U of 0 there are C ∈ C, D ∈ D such that CD−CD ⊆ U .Choose a neighbourhood V of 0 such that V + V ⊆ U . By 44.3, thereare C ∈ C and D ∈ D and a neighbourhood W of 0 such that CW ⊆ V

and WD ⊆ V . By making C and D still smaller, if necessary, we mayobtain that also C − C ⊆ W and D − D ⊆ W . Then CD − CD ⊆(C − C)D + C(D −D) ⊆WD + CW ⊆ V + V ⊆ U. �

44.5 Construction Every Hausdorff topological ring R has a Hausdorff

ring completion ι : R→ R. If R is commutative, then so is R.

Proof (1) The additive group of R is a Hausdorff abelian topologicalgroup and therefore has a group completion ι : R→ R such that R is anabelian Hausdorff group; see 43.21. We shall define a multiplication on R

which makes R a topological ring and such that ι is a ring monomorphism(and hence an embedding of topological rings).

Page 283: The Classical Fields

266 Completion

The image ι(R) is dense in R. As in 43.22, we may therefore representthe elements of R by special filterbases in R. For x′ ∈ R, let Vx′ be theneighbourhood filter of x′ in R. Then

Vx′ ∩ ι(R) = {N ′ ∩ ι(R) | N ′ ∈ Vx′ }is a filterbase on ι(R) ⊆ R which converges to x′, and

Wx′ = ι−1(Vx′) = ι−1(Vx′ ∩ ι(R))

is a concentrated filterbase on R.(2) For a further element y′ ∈ R, we consider the product filterbase

Wx′Wy′ on R, which according to Lemma 44.4 is concentrated. Since ι

is uniformly continuous, the image filterbase ι(Wx′Wy′) is concentratedas well (43.12) and hence converges in the complete group R. We definea multiplication on R by

x′y′ := lim ι(Wx′Wy′) .

(Recall that in the Hausdorff space R the limit is unique.)We first prove that ι : R → R is a homomorphism with respect to

multiplication. Since ι induces a homeomorphism R→ ι(R), the neigh-bourhoods of an element x ∈ R are the inverse images of neighbourhoodsof ι(x) in R. In our notation this says that for x, y ∈ R the filterbasesWι(x) and Wι(y) are just the neighbourhood filters of x and y in R. Thecontinuity of multiplication in R implies that the filterbase Wι(x)Wι(y)

converges to xy. Hence, by continuity of ι, it follows that

ι(xy) = ι(lim(Wι(x)Wι(y))) = lim ι(Wι(x)Wι(y)) = ι(x) · ι(y) . (1)

(3) We now show that the multiplication defined in (2) is continuous.For fixed x′

0, y′0 ∈ R, let N ′ be a neighbourhood of x′

0y′0. Since R is

regular (62.4), there is a neighbourhood N ′1 of x′

0y′0 such that N ′

1 ⊆ N ′.By definition of x′

0y′0 there are X0 ∈ Wx′

0and Y0 ∈ Wy′

0such that

ι(X0Y0) ⊆ N ′1. Let X ′

0 ∈ Vx′0

and Y ′0 ∈ Vy′

0be neighbourhoods such

that X0 = ι−1(X ′0), Y0 = ι−1(Y ′

0), and choose smaller neighbourhoodsX ′ ∈ Vx′

0and Y ′ ∈ Vy′

0such that X ′

0 is a neighbourhood of all pointsof X ′ and that Y ′

0 is a neighbourhood of all points of Y ′. Then forx′ ∈ X ′ and y′ ∈ Y ′ we have X0 ∈ Wx′ and Y0 ∈ Wy′ . Consequently,the product x′y′ = lim ι(Wx′Wy′) is contained in the closure ι(X0Y0) by43.6 and hence in N ′

1 ⊆ N ′. Thus, the neighbourhoods X ′ of x′0 and Y ′

of y′0 satisfy X ′Y ′ ⊆ N ′, and continuity of the multiplication is proved.

(4) It remains to show that the multiplication of R has the propertiesof a ring multiplication, associativity and distributivity over the addition

Page 284: The Classical Fields

44 Completion of topological rings and fields 267

of R. This is obtained from the corresponding properties of the multi-plication of the ring R by continuity via the usual density argument,which we have already used in 43.20. Associativity is proved there, andalso that the multiplication of R is commutative if the multiplication ofR is, and that the image ι(1) of the unit element 1 of R is a unit elementof the multiplication of R. Distributivity is proved in the same way.

Thus, R is a ring, in fact topological ring, since the additive group isa topological group, and multiplication is continuous by step (3). More-over, ι is a ring homomorphism: it is an embedding of topological groupsof the additive group of R into that of R, and it respects multiplication;see equation (1). The other properties of a ring completion are clear. �

44.6 Theorem: Universal property of ring completions For

i = 1, 2, let Ri be a Hausdorff topological ring and ιi : Ri → Ri a

Hausdorff ring completion.

Then every continuous ring homomorphism ϕ : R1 → R2 extends to

a unique continuous map ϕ : R1 → R2 such that ϕ◦ ι1 = ι2 ◦ϕ, and ϕ is

a ring homomorphism. If ϕ is an isomorphism of topological rings, that

is, a ring homomorphism and a homeomorphism, then so is ϕ.

Proof A ring completion is in particular a group completion of the ad-ditive group. Thus, existence of the continuous extension ϕ follows fromthe analogous universal property 43.24 of group completions, and we alsoobtain that ϕ is a homomorphism of the additive groups, and a homeo-morphism if ϕ is. That ϕ respects multiplication, as well, is proved by adensity argument again, analogously to step (4) of the proof of 43.23. �

The universal property above, applied in the special case R1 = R2 andϕ = id, immediately gives that Hausdorff ring completions are essentiallyunique:

44.7 Corollary: Uniqueness of ring completion Let ι : R → R

and ι : R→ R be Hausdorff ring completions of a Hausdorff topological

ring R. Then there is an isomorphism ϕ : R → R of topological rings

such that ϕ ◦ ι = ι. �

The process of ring completion may of course be applied to a topo-logical skew field F , and one might hope that the completion F , whichis a topological ring, is even a topological skew field, but this is not truein general. The problem is not the continuity of the multiplicative in-version map, but more fundamentally the fact that the completion maycontain elements which are not invertible. In 44.12 we shall see that the

Page 285: The Classical Fields

268 Completion

completion may even contain zero divisors. For the more subtle case of atopological field whose ring completion is an integral domain yet still nota field, see Heckmanns 1991 and Warner 1989 Exercises 13.2–13.14pp. 101ff.

Here is a criterion for the completion of a topological skew field to bea skew field again. Recall from 13.4 that a topological skew field is aHausdorff space if it is not indiscrete.

44.8 Theorem Let F be a topological skew field whose topology is not

the indiscrete topology, and ι : F → F its Hausdorff ring completion.

Then F is a skew field if, and only if, the following condition holds in F :

For every concentrated filterbase C on F which does not converge to 0,

the filterbase C−1 := { (C � {0})−1 | C ∈ C} is concentrated as well.

If this holds, then the inversion map F � {0} → F � {0} : x′ �→ x′−1

is continuous, so that F is a topological skew field.

Remarks (1) If the filterbase C does not converge to 0, then by 43.6there is C ∈ C such that 0 /∈ C. In particular we have for every C ∈ C

that C � {0} �= ∅, so that in the statement of the theorem C−1 can beformed without problem and is a filterbase.

(2) The condition of Theorem 44.8 is satisfied for topological skewfields of type V , as defined in 57.5 (see Exercise 6 of Section 57). Theseinclude the skew fields whose topology is described by an absolute value(see Exercise 1 of Section 57); this case will be treated explicitly in 44.9.

Proof (1) For simplicity, we denote the unit element ι(1) of F by 1. Asbefore, we shall use filterbases within F to represent elements of F inthe following way. For x′ ∈ F , let Vx′ be the neighbourhood filter of x′

in F . As explained in 43.22,

Wx′ = ι−1(Vx′) = { ι−1(N ′) | N ′ ∈ Vx′ }is a concentrated filterbase on F whose image filterbase ι(Wx′) convergesto x′.

(2) We first prove that the stated condition is sufficient for F to bea skew field. We have to show that every element x′ ∈ F � {0} has amultiplicative inverse. Let C be any concentrated filterbase on F suchthat the image filterbase ι(C) converges to x′ (one could use the filterbaseWx′ above, but it does not matter which). C does not converge to 0,or else ι(C) would have to converge to ι(0) = 0. By assumption, thefilterbase C−1 on F is concentrated again, and so is the image filterbaseι(C−1), since ι is uniformly continuous. Hence ι(C−1) converges to an

Page 286: The Classical Fields

44 Completion of topological rings and fields 269

element y′ of the completion F . We shall show that x′y′ = 1. By 44.4,the product filterbase CC−1 is concentrated, as well. We verify thateach of its members contains 1. Indeed, for C1, C2 ∈ C the intersectionC1 ∩ C2 contains a further member C ∈ C, and 1 ∈ C · (C � {0})−1 ⊆C1 · (C2 �{0})−1. According to 43.6, therefore, CC−1 converges to 1. Bycontinuity of ι and of multiplication it follows that 1 = lim ι(CC−1) =lim ι(C)·ι(C−1) = lim ι(C)·lim ι(C−1) = x′y′, so that y′ is a multiplicativeinverse of x′.

(3) For the second part of the proof, let F be a skew field. We wantto show that the condition stated in the theorem holds. Let C be aconcentrated filterbase on F which does not converge to 0. The imagefilterbase ι(C) is concentrated again and thus converges to an element x′

of the completion F . Since the embedding ι induces a homeomorphismof F onto ι(F ), we can be sure that x′ �= 0 = ι(0), or else C = ι−1(ι(C))would have to converge to ι−1(ι(0)) = 0, contrary to our assumption. Asabove, there is a concentrated filterbase D on F whose image filterbaseι(D) converges to the inverse x′−1. We consider the product filterbaseCD; for continuity reasons, its image ι(CD) = ι(C) · ι(D) converges tox′x′−1 = 1. Again it follows that CD converges to 1 in F , since ι is anembedding. In order to prove that C−1 is concentrated, let U be a neigh-bourhood of 0 in F . By continuity of addition, there is a neighbourhoodV of 0 such that V + V ⊆ U . By 44.3, there is a neighbourhood N

of 0 and a member D ∈ D such that DN ⊆ V . By continuity of theoperations in F , there are neighbourhoods W0 of 0 and W1 of 1 suchthat 0 /∈ W1, W0W

−11 ⊆ V and W−1

1 −W−11 ⊆ N . Furthermore, we

may choose D so small that D − D ⊆ W0. Since CD converges to 1,there are C ∈ C and D′ ∈ D such that CD′ ⊆ W1. Replacing D andD′ by a member of D contained in their intersection, we may assumethat D′ = D. Then (C � {0})−1 − (C � {0})−1 ⊆ DW−1

1 − DW−11 ⊆

(D −D)W−11 + D

(W−1

1 −W−11

) ⊆W0W−11 + DN ⊆ V + V ⊆ U .

(4) It remains to prove continuity of the inversion map (still under theassumption that F is a skew field). By 8.3 it suffices to show continuityat 1. Let N ′

1 be a neighbourhood of 1 in F×. By continuity of multi-plication, there exists a neighbourhood N ′

2 of 1 such that N ′2N

′2 ⊆ N ′

1.Since inversion is continuous on F×, there exists a neighbourhood U

of 1 in F× such that U−1 ⊆ ι−1(N ′2). As ι is an embedding, we find

a neighbourhood Q′1 of 1 in F× such that Q′

1 ∩ ι(F ) ⊆ ι(U). Then(Q′

1 ∩ ι(F ))−1 ⊆ N ′2. Again, there is a neighbourhood Q′

2 of 1 such thatQ′

2Q′2 ⊆ Q′

1 and Q′2 ⊆ N ′

2. We now show that Q′2−1 ⊆ N ′

1, which finishesthe proof.

Page 287: The Classical Fields

270 Completion

Let x′ ∈ Q′2. Since ι(F ) is dense, the neighbourhood Q′

2x′ of x′

meets ι(F ), hence there exists y′ ∈ Q′2 such that y′x′ ∈ ι(F ). Then we

have y′x′ ∈ Q′1 ∩ ι(F ) and y′ ∈ N ′

2, which gives (x′)−1 = (y′x′)−1y′ ∈(Q′

1 ∩ ι(F ))−1y′ ⊆ N ′

2N′2 ⊆ N ′

1. �

44.9 Completion of fields with absolute value Let F be a fieldwith an absolute value, that is, a map ϕ : F → [0,∞[ ⊆ R whichsatisfies the following properties for x, y ∈ F : ϕ(x) = 0 if, and only if,x = 0 (definiteness), ϕ(x + y) ≤ ϕ(x) + ϕ(y) (triangle inequality), andϕ(xy) = ϕ(x)ϕ(y) (multiplicativity); see 55.1. Here, R is meant to bethe completion of Q as an ordered field; see 42.11.

A metric d on F is defined by d(x, y) = ϕ(x − y) for x, y ∈ F . Withthe topology induced by this metric F is a topological field; this can beshown exactly as in the special case of the standard absolute value onR; see 8.2 and 9.1.

Theorem Let ι : F → F be the Hausdorff ring completion of a field

F with an absolute value ϕ. Then F is a topological field. In fact, its

topology comes from an absolute value on F which is an extension of ϕ.

Proof (1) That F is a topological field can be verified via the criterionin Theorem 44.8. Let C be a concentrated filterbase in F which does notconverge to 0; one has to show that C−1 is concentrated, as well. Let U

be a neighbourhood of 0 in F and ε > 0 such that Bε(0) = {x ∈ F |ϕ(x) < ε} ⊆ U . According to 43.6, there is C ∈ C such that 0 /∈ C, inother words there is δ > 0 such that for all c ∈ C one has ϕ(c) ≥ δ. SinceC is concentrated, we may assume that C − C ⊆ Bδ2ε(0). For c, d ∈ C

then ϕ(c−1 − d−1) = ϕ(d−1(d − c)c−1) = ϕ(d−1) · ϕ(d − c) · ϕ(c−1) =ϕ(d)−1 · ϕ(d− c) · ϕ(c)−1 < δ−1δ2εδ−1 = ε, so that C−1 − C−1 ⊆ U .

(2) We now extend the absolute value of F to F . For simplicity, weshall write ι(0) = 0 and ι(1) = 1. We observe that the absolute value ϕ isuniformly continuous, as a direct consequence of the triangle inequality.By 43.10, the additive group R is complete. According to 43.23, we mayextend ϕ to a uniformly continuous map ϕ : F → R such that

ϕ(ι(x)) = ϕ(x) (1)

for all x ∈ F . Since ι(F ) is dense in F , the image ϕ(F ) is containedin the closure of ϕ(ι(F )), and since [0,∞[ is closed in R, it follows thatϕ(x′) ≥ 0 for all x′ ∈ F , as well.

(3) We prove that ϕ satisfies the triangle inequality. This amounts toverifying that the set T = { (x′, y′) ∈ F × F | ϕ(x′ +y′) ≤ ϕ(x′)+ ϕ(y′)}

Page 288: The Classical Fields

44 Completion of topological rings and fields 271

equals F × F . Since ι is a ring homomorphism and since ϕ satisfiesthe triangle inequality, equation (1) shows that T contains ι(F )× ι(F ),which is a dense subset of F × F . By continuity of addition and of ϕ

it is clear that T is closed in F × F , hence T = F × F , indeed. In thesame way, one obtains that ϕ is multiplicative since ϕ is: one verifiesthat { (x′, y′) ∈ F × F | ϕ(x′y′) = ϕ(x′)ϕ(y′)} = F × F .

(4) Next we show that ϕ also is definite. It is clear that ϕ(0) = 0.Assume that ϕ(x′) = 0. Since ι(F ) is dense in F , there is a concentratedfilterbase Wx′ on F such that ι(Wx′) converges to x′; see 43.22. Weshow that Wx′ is finer than the neighbourhood filter V0 of 0 in F . Thenι(Wx′) is finer than the filterbase ι(V0), which converges to ι(0) = 0, sothat ι(Wx′) converges to both x′ and 0, and x′ = 0 as F is a Hausdorffspace. In order to show that Wx′ is finer than V0, let U ∈ V0. Thereis ε > 0 such that Bε(0) ⊆ U . Since ι(Wx′) converges to x′ and sinceϕ(x′) = 0, there is W ∈ Wx′ such that all w ∈ W satisfy ϕ(ι(w)) < ε,by continuity of ϕ. Because of equation (1) this means W ⊆ Bε(0) ⊆ U .

(5) Due to steps (3) and (4), ϕ is an absolute value on F . The topologydefined by the corresponding metric d metric makes F a topological field,as well. We compare this metric topology with the topology of F ascompletion of F , the ‘completion’ topology. Since ϕ is continuous in thecompletion topology, the metric topology is coarser. Equation (1) saysthat ι is an isometric embedding of (F, d) into the metric space (F , d).It now follows from 43.26 that the metric topology coincides with thecompletion topology. �

44.10 Completion of fields with absolute values in terms ofCauchy sequences In the situation discussed above in 44.9, the com-pletion F may be described using Cauchy sequences in F . Since thetopologies of the topological fields F and F are defined by the transla-tion invariant metrics d and d, the notion of a Cauchy sequence can beexpressed in the usual way by these metrics. It is an easy special case of44.4 that the set C of all Cauchy sequences of F is closed under termwiseaddition and multiplication and hence is a ring. The isometry ι mapsa Cauchy sequence of F to a Cauchy sequence of F . By completeness,the image sequence has a (unique) limit in F . Thus we have a map

C → F : (xν)ν �→ lim ι(xν) . (1)

Since ι is a continuous ring homomorphism and since the ring operationsin F and F are continuous, the map (1) is a ring homomorphism, as well.It is surjective; indeed, since ι(F ) is dense in F , every element of F is

Page 289: The Classical Fields

272 Completion

the limit of a sequence in ι(F ), which is a Cauchy sequence with respectto the metric d, and hence is the image sequence of a Cauchy sequenceof F , as ι is an isometry. For the same reason, a Cauchy sequence of F

converges to 0 if, and only if, its image sequence converges to 0. Hence,the kernel of the map (1) is the ideal Z consisting of the sequences of F

converging to 0. The map (1) factors through the factor ring C/Z andyields a ring isomorphism

C/Z → F : (xν)ν + Z �→ lim ι(xν) . (2)

In order to have a self-contained description of the completion of F interms of Cauchy sequences, we prove once more that C/Z is a field.

The sequence all of whose terms are 1 is the unit element of C. ACauchy sequence (xν)ν which does not converge to 0 cannot even accu-mulate at 0 (see 43.6). Hence there are only finitely many ν such thatxν = 0, and the sequence (yν)ν defined by yν = xν if xν �= 0 and yν = 1 ifxν = 0 is still a Cauchy sequence which does not accumulate at 0. It hasno zero terms any more, and clearly (xν)ν +Z = (yν)ν +Z. If we ascer-tain that the inverse elements y−1

ν form a Cauchy sequence again, then itis clear that (y−1

ν )ν is an inverse of (yν)ν in C, and (xν)ν +Z = (yν)ν +Z

has the inverse (y−1ν )ν + Z.

The following argument is a Cauchy sequence version of step (1) in theproof of Theorem 44.9. The multiplicativity of absolute values impliesthat ϕ(y−1

μ −y−1ν ) = ϕ(y−1

ν (yν−yμ)y−1μ ) = ϕ(yν)−1 ·ϕ(yν−yμ)·ϕ(yμ)−1.

The real numbers ϕ(yν) are bounded away from 0, since (yν)ν does notaccumulate at 0, hence their inverses ϕ(yν)−1 are bounded. Thus theabove equation shows that ϕ(y−1

μ − y−1ν ) is arbitrarily small if μ, ν are

sufficiently large, since (yν)ν is a Cauchy sequence. Thus, indeed, (y−1ν )ν

is a Cauchy sequence, as well.Furthermore, we pull back the absolute value ϕ from F to C/Z using

the ring isomorphism (2). For a Cauchy sequence (xν)ν continuity of ϕ

implies that ϕ(lim ι(xν)) = lim ϕ(ι(xν)) = limϕ(xν). Thus, the abso-lute value Φ on C/Z corresponding to the absolute value ϕ on F underthe ring isomorphism (2) is given by

Φ((xν)ν + Z) = limϕ(xν) .

If one prefers, one may verify directly that this defines an absolute valueon C/Z.

With the topology defined by this absolute value, C/Z is a topologi-cal field which is isomorphic to F by the map (2). In particular, C/Z is

Page 290: The Classical Fields

44 Completion of topological rings and fields 273

complete. Under the isomorphism (2), the embedding ι : F → F corre-sponds to the map F → C/Z : x→ (x)ν +Z, where (x)ν is the constantsequence all of whose terms are equal to x. With this embedding, C/Z

can be viewed as the completion of the topological field F .We remark that above (a countable form of) the axiom of choice has

been used without mention to show that the map (2) is surjective. Thiscould be avoided by using concentrated filters instead of Cauchy se-quences in an analogous way.

If one employs C/Z to construct the completion of F , as is often done,then the axiom of choice is needed for proving that C/Z is complete. Incontrast, our proof of 44.9 does not use the axiom of choice.

44.11 Examples (1) The real numbers The ordered field R is atopological field with the ordering topology (see 8.2 and 9.1), and it is acomplete topological field (by 43.10). In 42.11, we have obtained R as thecompletion of the Archimedean ordered field Q of rational numbers. Inparticular, Q may be considered as a subfield of R which is weakly dense(see 41.1 about this density notion for chains). This implies that Q istopologically dense in R. Thus, the topological field R is the completionof Q as a topological field with the usual topology.

(2) The p-adic numbers Let p be a prime number. The completionof the topological field Q endowed with the p-adic topology (compare36.3) is the field Qp of p-adic numbers. This field will be discussed inSections 51–54.

(3) Laurent series fields (1) Let F be a field. The elements ofthe Laurent series field F ((t)) are those formal series

∑ν∈Z ξνtν with

ξν ∈ F whose support {ν ∈ Z | ξν �= 0} is bounded below; see 64.23.If n is a lower bound of the support, we also write the above Laurentseries as

∑ν≥n ξνtν . Addition of Laurent series is defined termwise, and

multiplication is the formal Cauchy product.The so-called power series are the elements of F ((t)) of the form∑ν≥0 ξνtν (the Laurent series whose support is contained in {0} ∪ N).

These power series form a subring F [[t]] of F ((t)).Let A := {∑n

ν=1 ξνt−ν | n ∈ N, ξν ∈ F } be the F -linear span of{ t−n | n ∈ N} in F ((t)). The map

A× F [[t]]→ F ((t)) : (a, b) �→ a + b (1)

is a group isomorphism with respect to addition.We endow F ((t)) with a topology. F [[t]] can be considered as the

Cartesian product F {0}∪N since the power series∑

ν≥0 ξνtν is just a

Page 291: The Classical Fields

274 Completion

suggestive way of writing the sequence (ξν)ν∈{0}∪N. We consider thediscrete topology on F , the product topology on F [[t]], the discretetopology on A and finally the product topology on A × F [[t]]. Thetopology on F ((t)) will be the topology which makes the map (1) ahomeomorphism.

Since discrete groups are complete (43.5), the additive group of F ((t))is a complete topological group by 43.27.

If F is finite, then the Cartesian product F [[t]] = F {0}∪N is compactby Tychonoff’s theorem; hence F ((t)) is locally compact.

(2) Next, we give another description of the topology of F ((t)) usingthe valuation

v : F ((t)) � {0} → Z : v(∑

ν∈Z ξνtν)

= min{ν ∈ Z | ξν �= 0} .

The term valuation means that v(x + y) ≥ min{v(x), v(y)} for x, y ∈F ((t))× such that y �= −x and v(xy) = v(x) + v(y) for x, y ∈ F ((t))×;compare 56.1. These properties imply that the map ϕ : F ((t))→ R withϕ(x) = 2−v(x) for x �= 0 and ϕ(0) = 0 is an absolute value on F ((t))satisfying the ultrametric property ϕ(x + y) ≤ max{ϕ(x), ϕ(y)}, whichis stronger than the ordinary triangle inequality.

Let d be the metric on F ((t)) defined by d(x, y) = ϕ(x − y). Withthe topology induced by this metric (the so-called valuation topology,compare Section 56), the field F ((t)) is a topological field as in 44.9. Form ∈ N, the Laurent series whose values under v are at least m constitutea neighbourhood

{∑ν≥m ξνtν | ξν ∈ F } = tmF [[t]]

of 0 which is an ideal of F [[t]]. Moreover, the ideals tmF [[t]] with m ∈ Nform a neighbourhood base of 0 in F [[t]] and in F ((t)) for the valuationtopology; see also 13.2(b). This neighbourhood base is also a neigh-bourhood base of 0 for the topology defined initially in step (1) usingproduct topologies. Since the additive group F ((t)) is a topologicalgroup for both topologies, it follows that the two topologies coincide.

(3) The polynomial ring F [t] is the subring of F [[t]] consisting ofthe power series with finite support. From the definition of the pro-duct topology, it is clear that F [t] is dense in F [[t]]. By continuity ofmultiplication in the topological field F ((t)), it follows that the ringF [t, t−1] =

⋃n≥0 t−nF [t] is dense in

⋃n≥0 t−nF [[t]] = F ((t)). The field

F ((t)) contains (an isomorphic copy of) the field of fractions F (t) of F [t]as a subfield which in turn contains F [t, t−1] and hence is dense, as well.Thus, F ((t)) is the field completion of F (t).

Page 292: The Classical Fields

44 Completion of topological rings and fields 275

Here, we must of course consider F (t) as a topological ring with thetopology induced by the topology of F ((t)). This topology is defined bythe restriction to F (t) of the absolute value ϕ, which comes from thevaluation v. For a polynomial f ∈ F [t], the value v(f) ∈ N ∪ {0} is thelowest exponent of t appearing in f . For a non-zero polynomial g, weinfer by the properties of a valuation that v(f/g) = v(f)− v(g). This isthe valuation vt of F (t) constructed as in 56.3(c) using the irreduciblepolynomial p = t ∈ F [t]. Thus, F ((t)) is the completion of F (t) for thetopology defined by this valuation.

(4) Another prominent valuation on F (t) is the degree valuation v∞given by v∞

(f/g)

= −deg f +deg g for polynomials f, g ∈ F [t]; compare56.3(c) again. We shall see that the unique field automorphism ϑ :F (t) → F (t) mapping t to 1/t and fixing every element of F satisfiesvt ◦ϑ = v∞. Thus, ϑ is an isomorphism of topological fields of F (t) withthe topology defined by vt onto F (t) with the topology defined by v∞.

Indeed, for f ∈ F [t], we have ϑ(f) = f(1/t). The polynomial tdf(1/t)with d = deg f has non-zero absolute term, hence vt(tdf(1/t)) = 0 andconsequently vt(ϑ(f)) = vt(f(1/t)) = −vt(td) = −d = v∞(f). Thisshows that the valuations vt ◦ ϑ and v∞ coincide on F [t], hence theycoincide on the field of fractions F (t).

The isomorphism ϑ of topological fields which we have just establishedextends to an isomorphism of the completions of these topological fieldsaccording to 44.7. Thus, F ((t)) can be considered as completion of F (t)not only for the topology defined by vt, but also for the topology definedby v∞. Note that these topologies do not coincide: the sequence (tν)ν∈N

converges to 0 in the topology defined by vt, but not in the topologydefined by v∞.

44.12 Example: A topological field whose completion has zerodivisors We consider the 2-adic and the 3-adic topology on Q definedby the 2-adic and the 3-adic absolute value | |2 and | |3; see 36.3. Witheach of these topologies, Q is a topological field. We endow the Cartesianproduct Q×Q with the product topology of the 2-adic topology on thefirst factor and the 3-adic topology on the second factor. This topologyon Q × Q will be called the hexadic topology. With componentwiseaddition and multiplication Q×Q is a topological ring.

The diagonal D = { (x, x) | x ∈ Q} ⊆ Q × Q with the topologyinduced by the hexadic topology is a topological field (as a field, it isisomorphic to Q). We shall see that the ring completion D of D has zerodivisors.

Page 293: The Classical Fields

276 Completion

For n ∈ N, n→∞, the rational numbers

xn =(2/3)n

1 + (2/3)n=

11 + (3/2)n

converge to 0 in the 2-adic topology and to 1 in the 3-adic topology.Indeed, the definition in 36.3 yields |(2/3)n|2 = 2−n → 0, so that |xn|2converges to 0 in the 2-adic topology. In the same way, |(3/2)n|3 =3−n → 0, hence xn converges to 1 in the 3-adic topology. The sequence(xn, xn)n∈N in D converges in Q × Q to (0, 1) /∈ D. Exchanging theprimes 2 and 3, we obtain a sequence (yn)n∈N in Q converging to 1 inthe 2-adic topology and to 0 in the 3-adic topology, and the sequence(yn, yn)n∈N in D converges in Q×Q to (1, 0) /∈ D. Thus, the closure ofD in the hexadic topology of Q×Q contains (1, 0) and (0, 1).

In order to determine the completion D, we embed Q × Q with thehexadic topology into the Cartesian product Q2×Q3 of the completionsof Q for the 2-adic and the 3-adic topology; this Cartesian product is atopological ring with componentwise addition and multiplication, and itis complete by 43.27, since the factors are complete. The closure D of D

in Q2×Q3 is a subgroup of the additive group of Q2×Q3; see 62.5. Thelatter is a topological vector space over Q with componentwise scalarmultiplication, and D is a linear subspace. For r ∈ Q, we obtain thatr · D ⊆ rD ⊆ D, so that D is a Q-linear subspace, as well, but as wehave seen above, D contains (0, 1) and (1, 0) and hence Q × Q. SinceQ×Q is dense in Q2×Q3, we infer that the diagonal D of Q×Q embedsas a dense subset into Q2 ×Q3 (this fact is generalized in Exercise 2 ofSection 55). Thus, the ring completion of the field D is D = Q2 × Q3.Since Q2 and Q3 are fields (see 44.11 Example (2) and 44.9), the zerodivisors of this ring are the elements of (Q2 × {0}) ∪ ({0} ×Q3).

44.13 Additive versus multiplicative completeness of topologi-cal skew fields One may ask if the multiplicative group F× of a com-plete topological skew field F is right complete as a topological group.(See 43.28 for the notions of right, left and bilateral completeness of non-commutative topological groups.) Completeness of F is expressed usingthe additive group, whereas the question about completeness of F× isconcerned with right concentrated filterbases on the group F×, withmultiplication as group operation. We do not know if such filterbasesare necessarily concentrated in the additive group.

A positive answer can be given if one asks for bilateral completenessof F× instead of right completeness; see Warner 1989 Theorem 14.11p. 111. He proves more generally that if in a complete topological ring R

Page 294: The Classical Fields

44 Completion of topological rings and fields 277

the group R× of invertible elements is open and if inversion is continuouson it, then R× is bilaterally complete. In the special case of topologicalfields (with commutative multiplication), there is no distinction betweenleft, right and bilateral completeness. The argument of Warner 1989loc. cit. can be adapted to show that the multiplicative group of a lo-cally bounded complete topological skew field (see 57.1) is even rightcomplete.

Conversely, it is rather trivial that a filterbase on the multiplicativegroup of a topological field which is concentrated in the additive groupneed not be concentrated in the multiplicative group. To see this, wetake a topological field F whose topology is neither discrete nor theindiscrete topology and whose multiplicative group is complete, and weconsider the neighbourhood filter V0 of 0 in F . It converges to 0. Sincethe topology of F is not discrete, {0} is not a neighbourhood of 0, sothat every neighbourhood V of 0 intersects F×. Hence V0 ∩ F× ={V ∩ F× | V ∈ V0 } is a filterbase which still converges to 0 and so isconcentrated in the additive group. However, as a filterbase of F×, itcannot be concentrated; else, by completeness, it would have to convergeto an element of F×, but instead it converges to 0.

Page 295: The Classical Fields

5

The p-adic numbers

The idea of p-adic numbers is due to Hensel, who was inspired by localpower series expansions of meromorphic functions (see Warner 1989p. 469f, Ebbinghaus et al. 1991 Chapter 6 and Ullrich 1998). Wetreat the p-adic numbers as relatives of the real numbers. In fact, com-pletion of the rational field Q with respect to an absolute value leadseither to the reals R or to a field Qp of p-adic numbers, where p is aprime number (see 44.9, 44.10, 51.4, 55.4), and these fields are locallycompact. We consider the additive and the multiplicative group of Qp

in Sections 52 and 53, and we study squares and quadratic forms overQp in Section 54. It turns out (see 53.2) that the additive and the mul-tiplicative group of Qp are locally isomorphic, in the sense that someopen (and compact) subgroup of Q×

p is isomorphic to an open subgroupof Q+

p ; this is similar to the situation for R.Comparing R and Qp, it appears that the structure of Qp is dominated

much more by algebraic and number theoretic features. A major topo-logical difference between the locally compact fields R and Qp is the factthat R is connected and Qp is totally disconnected (51.10). Moreover,Qp cannot be made into an ordered field (54.2).

In Sections 55–58 we put the fields Qp in the context of general topo-logical fields: we study absolute values, valuations and the correspondingtopologies. Section 58 deals with the classification of all locally compactfields and skew fields. If Q is a dense subfield of a non-discrete locallycompact field F , then F ∼= R or F ∼= Qp for some prime p; see 58.7.

A notable omission in this chapter is Hensel’s Lemma on fields F

which are complete with respect to a valuation v; this lemma allows tolift roots of suitable polynomials in the residue field Fv to roots in F , by amethod related to Newton approximation. We could have used Hensel’sLemma on several occasions (compare Robert 2000 1.6, Gouvea 1997

278

Page 296: The Classical Fields

51 The field of p-adic numbers 279

3.4, Lang 1970 II §2, Greenberg 1969 Chapter 5), but we chose todeviate from well-trodden paths and to find out how far we can getwithout it. See also Ribenboim 1985, Prieß-Crampe–Ribenboim

2000 and Engler–Prestel 2005 p. 20f.

51 The field of p-adic numbers

Here we construct the field Qp of p-adic numbers (51.4) and the ring Zp

of p-adic integers (51.6), and we derive some basic properties of Qp andof Zp. Everything hinges on the p-adic absolute value | |p : Q→ Q andits extension | |p : Qp → Q.

51.1 Definition Fix a prime number p. We recall some facts from36.3. Every non-zero rational number x can be written in the form

x = pn · ab

with integers n, a, b ∈ Z such that a and b are not divisible by p. Thep-adic absolute value | |p on Q is defined by

|x|p = p−n

for 0 �= x ∈ Q, and |0|p = 0. The mapping | |p is multiplicative andultrametric, i.e. |xy|p = |x|p|y|p and |x + y|p ≤ max{|x|p, |y|p} for allx, y ∈ Q; see 36.3. (Hence | |p is a non-Archimedean absolute value onQ as defined in 55.1.) The multiplicative group Q× is the direct productof {±1} and of the free abelian group generated by all prime numbers(see 32.1), and the p-adic absolute value is just the projection Q× → pZ

onto the factor pZ of Q×, followed by inversion pn �→ p−n.The ordinary absolute value |x| = max{x,−x} on Q is often associated

with the ‘prime p =∞’ (but Conway 1997 argues that it is associatedrather to the ‘prime p = −1’). These definitions give the product formula

|x| ·∏p∈P |x|p = 1

for all x ∈ Q×. One could also define |x|p to be 2−n instead of p−n, butthen the product formula would become slightly more complicated.

51.2 Metric and topology As in 36.3, we define on Q the p-adicmetric dp by dp(x, y) = |x − y|p for x, y ∈ Q. This is in fact an ultra-metric, i.e., we have dp(x, z) ≤ max{dp(x, y), dp(y, z)} for x, y, z ∈ Q, asa consequence of the ultrametric inequality for | |p. Ultrametric spaceshave some unexpected geometric properties: all triangles are isosceles,and spheres have many centres. The metric topology defined by dp is

Page 297: The Classical Fields

280 The p-adic numbers

the p-adic topology of Q. One defines p-adic convergence and p-adicCauchy sequences in the usual manner. For example, limn→∞ pn = 0 inthe p-adic topology.

The metric space (Q, dp) is not complete; indeed, the integers an =∑nk=0 pk2

or an =∑n

k=0 pk! form p-adic Cauchy sequences (since inboth cases |an − am|p ≤ p−min{m,n}) which have no limit in Q; seeProposition 51.11.

The following observation is useful when forming the completion of(Q, dp).

51.3 Lemma Let (an)n ∈ QN be a p-adic Cauchy sequence.

(i) Then the sequence (|an|p)n is a Cauchy sequence in R.

(ii) If the sequence (an)n does not converge to 0 in the p-adic topology,

then |an|p is finally constant (and non-zero).

Proof (i) follows from the inequality∣∣|x|p − |y|p∣∣ ≤ |x− y|p, which is a

consequence of the triangle inequality |x|p = |x−y+y|p ≤ |x−y|p + |y|p.(ii) The values of | |p form the set pZ ∪ {0}, and this closed subset of

R has zero as its only accumulation point. �

51.4 Definition We recall the construction of the p-adic completion ofQ from 44.9, 44.10. Let C(p) be the set of all p-adic Cauchy sequencesin Q, and let Z(p) be the set of all sequences in Q with p-adic limit 0.Then C(p) is a ring (with component-wise addition and multiplication),and Z(p) is an ideal in C(p) (since Cauchy sequences are bounded; see51.3). As shown in 44.10, the quotient ring

Qp := C(p)/Z(p)

is a field, the field Qp of p-adic numbers; the field property depends onthe observation that a sequence in C(p) � Z(p) cannot accumulate at 0;see 43.6 or 51.3(ii).

Mapping q ∈ Q to the constant sequence (q, q, . . . ) gives a naturalembedding Q → Qp, as in 44.10. We identify Q with its image in Qp,i.e. with the prime field of Qp.

By 44.10 or 51.3(i), the ring C(p) admits the map ϕ : C(p) → R :ϕ(a) = limn |an|p, which is multiplicative and ultrametric. Furthermore,ϕ is constant on each additive coset of Z(p), since a ∈ C(p), b ∈ Z(p)imply ϕ(b) = 0 and ϕ(a) = ϕ(a + b − b) ≤ ϕ(a + b) ≤ ϕ(a). Thus weobtain a well-defined extension of | |p to Qp = C(p)/Z(p) by defining

|a + Z(p)|p = limn→∞ |an|p for a ∈ C(p) .

Page 298: The Classical Fields

51 The field of p-adic numbers 281

This extension is again multiplicative and ultrametric; it is the p-adicabsolute value of Qp; compare 44.9, 44.10. The corresponding metricdp

(a + Z(p), b + Z(p)

)= |a − b|p = limn→∞ |an − bn|p on Qp extends

the p-adic metric on Q and defines a metric topology on Qp, the p-adictopology of Qp.

From 44.9, 44.10 we know that Q is dense in Qp, and that Qp iscomplete. So far, we have proved the following result.

51.5 Theorem For each prime p there exists a field extension Qp of Qand a map | |p : Qp → pZ ∪ {0} with the following properties:

(i) The map | |p is multiplicative and ultrametric and extends the p-

adic absolute value of Q.

(ii) The field Q is dense in Qp and Qp is complete (with respect to the

p-adic metric dp(x, y) = |x− y|p on Qp). �

We point out that Qp is determined uniquely by the properties statedin 51.5, in a strong sense: every field with the same properties is isomor-phic to Qp by a unique isomorphism, and this isomorphism preservesthe absolute values. Indeed, the inclusion of Q in any field F with theseproperties preserves the p-adic absolute value and has therefore a uniqueextension to an isomorphism Qp → F ; see 44.6. This means that onecan derive all results about Qp from 51.5, without recourse to the actualconstruction of Qp. The rest of this section relies only on 51.5.

We remark that Qp endowed with the p-adic topology is a topologicalfield (as defined in 13.1); this is a special case of 13.2(b) and of 44.9. Weuse this remark often without mentioning.

51.6 Definition Multiplicativity together with the ultrametric inequal-ity implies that the set

Zp :={

x ∈ Qp

∣∣ |x|p ≤ 1}

is a subring of Qp, the ring of p-adic integers. (It is the valuation ringof the valuation belonging to | |p; see Section 56.) Since |p|p = p−1, wehave

pnZp ={

x ∈ Qp

∣∣ |x|p ≤ p−n}

={

x ∈ Qp

∣∣ |x|p < p−n+1}

for all n ∈ Z. This shows that each set pnZp is simultaneously open andclosed in Qp.

51.7 Lemma The ring Zp has a unique maximal ideal, namely pZp ={x ∈ Zp | |x|p < 1}. Furthermore Zp = Z + pnZp for each n ∈ N.

The integers 0, 1, . . . , pn−1 are representatives for the additive cosets of

Page 299: The Classical Fields

282 The p-adic numbers

pnZp in Zp, and Zp/pnZp∼= Z/pnZ as rings. In particular, the quotient

ring

Zp/pZp∼= Z/pZ = Fp

is isomorphic to the finite field Fp with p elements.

Proof The set Zp � pZp = {x ∈ Zp | |x|p = 1} is the group of all unitsof Zp, hence the ideal pZp is the largest proper ideal of Zp.

Since Q is dense in Qp, the intersection Q ∩ Zp is dense in the openset Zp. Thus, for every x ∈ Zp, we find a, b ∈ Z with b not divisible by p

such that |x− ab−1|p < 1. Because Fp = Z/pZ is a field, we find b′ ∈ Zwith 1− bb′ ∈ pZ. Hence |ab−1 − ab′|p = |ab−1(1− bb′)|p < 1 and

|x− ab′|p ≤ max{|x− ab−1|p, |ab−1 − ab′|p} < 1 ,

which shows that x−ab′ ∈ pZp, hence x ∈ Z+pZp. Thus Zp = Z+pZp,and an easy induction gives Zp = Z + pnZp for n ∈ N. The descriptionof pnZp in 51.6 shows that Z ∩ pnZp = pnZ. We infer that the quotientZp/pnZp = (Z + pnZp)/pnZp

∼= Z/(Z ∩ pnZp) = Z/pnZ is representedby the integers 0, 1, . . . , pn − 1. �

51.8 Proposition We have Zp ={∑∞

n=0 cnpn∣∣ cn ∈ {0, 1, . . . , p−1}}

and

Qp ={∑∞

n=k cnpn∣∣ k ∈ Z, cn ∈ {0, 1, . . . , p− 1}} .

Each of these series converges, and the coefficients cn are determined

uniquely up to omitting some leading zeros.

Proof We have Qp =⋃

n≥0 p−nZp, hence it suffices to prove the state-ment about Zp. For each x ∈ Zp there exists by 51.7 a sequenceof integers xk ∈ {0, 1, . . . , pk − 1} such that x ∈ xk + pkZp. Sincexk ≡ xk+1 mod pk for all k ∈ N, we can write xk =

∑k−1n=0 cnpn with

integers cn ∈ {0, 1, . . . , p − 1}, where n ∈ N0. By construction we have|x − xk|p ≤ p−k for all k ∈ N, hence x = limk xk =

∑∞n=0 cnpn. The

coefficients cn are uniquely determined by x ∈ Zp, since c0 is the repre-sentative of x modulo pZp, c1p represents x − c0 modulo p2Zp, etc; ingeneral, ckpk represents x−∑k−1

n=0 cnpn modulo pk+1Zp.Finally, each of these series converges in Qp, because the partial

sums∑k

n=0 cnpn form a p-adic Cauchy sequence in Z for arbitrarycn ∈ {0, 1, . . . , p− 1}, as a consequence of the estimate

∣∣∑bn=a cnpn

∣∣p≤

maxa≤n≤b |cnpn|p ≤ p−a. �

Page 300: The Classical Fields

51 The field of p-adic numbers 283

51.9 Corollary The ring Zp is the closure of N and of Z in the p-adic

topology of Qp.

Proof Clearly N ⊆ Zp, and N is dense in the closed set Zp by 51.8. �

51.10 Proposition The sets pnZp with n ∈ N form a neighbourhood

base at 0 for the p-adic topology of Qp. Each of these sets, in particular

Zp, is compact and totally disconnected. Qp is locally compact, totally

disconnected, and not discrete.

Proof The first statement holds since pnZp = {x ∈ Qp | |x|p ≤ p−n }.The relation limn→∞ pn = 0 shows that Qp and Zp are not discrete.Thus it suffices to prove that the ring Zp is compact and totally discon-nected.

The series representation in 51.8 gives a natural bijection of Zp ontothe Cartesian power {0, 1, . . . , p − 1}N0 . Transferring the p-adic metricdp via this bijection to {0, 1, . . . , p− 1}N0 , we obtain the metric d with

d(x, y) = p−min{n|xn �=yn}

for distinct sequences x, y ∈ {0, 1, . . . , p − 1}N0 . We endow the finiteset {0, 1, . . . , p − 1} with the discrete topology; then d describes theproduct topology of {0, 1, . . . , p−1}N0 , which is compact (by Tychonoff’stheorem) and totally disconnected (since any projection of a non-emptyconnected set is a singleton). Hence also Zp is compact and totallydisconnected.

Alternatively, one can derive the compactness of Zp from that factthat it is complete and totally bounded; see Gouvea 1997 3.3.8. �

We mention that each set pnZp is in fact homeomorphic to the productspace {0, 1}N, hence homeomorphic to Cantor’s triadic set; see 5.48 (forp = 2 we have just shown this; compare also Exercise 8). FurthermoreQp is homeomorphic to Cantor’s triadic set minus a point, for everyprime p; see Christenson–Voxman 1977 6.C.11, Hewitt–Ross 19639.15 or Witt 1975.

Finally, we characterize the p-adic series which represent rational num-bers; the result is similar to the corresponding result for decimal expan-sions.

51.11 Proposition A p-adic number x =∑∞

n=k cnpn, with coefficients

cn ∈ {0, 1, . . . , p− 1}, is a rational number if, and only if, the sequence

(cn)n≥k is finally periodic:

x ∈ Q ⇐⇒ ∃s,t∈N ∀m≥t cm = cm+s .

Page 301: The Classical Fields

284 The p-adic numbers

Proof We may assume that k = 0. If the sequence (cn)n≥0 is finallyperiodic, then using s and t as above we can write x = a +

∑∞n=0 bpsn

with a ∈ Z and b = ctpt + ct+1p

t+1 + · · · + ct+s−1pt+s−1 ∈ Z. Since∑m−1

n=0 psn = (1 − psm)/(1 − ps) converges p-adically to 1/(1 − ps) form→∞, we infer that x = a + b/(1− ps) ∈ Q.

Conversely, let x =∑

n cnpn with cn ∈ {0, 1, . . . , p − 1} be rational,say x = a/b with a, b ∈ Z and b �= 0. For m ≥ 0 we define

zm := b∑

n≥m cnpn = a− b∑m−1

n=0 cnpn ∈ Z .

Since |zm|p ≤ p−m, the integer zm is divisible by pm, and for m ≥ 0 wehave the following estimate of ordinary absolute values

|p−mzm| = |p−ma− b

m−1∑n=0

cnpn−m| ≤ |a|+ |b|∑n≥1

(p− 1)p−n = |a|+ |b| ,

which does not depend on m. Thus the set {p−mzm | m ≥ 0} ofintegers is finite, hence p−tzt = p−t−szt+s for suitable numbers s, t ∈ N.We conclude that

b∑

n≥t cnpn = zt = p−szt+s = b∑

n≥t+s cnpn−s = b∑

n≥t cn+spn,

and the uniqueness of the coefficients (51.8) implies that cn = cn+s forall n ≥ t. �

51.12 Other constructions of Qp The ring Zp may be constructedalso as the inverse limit of the finite rings Z/pnZ with n ∈ N; see Serre

1973, Borevich–Shafarevich 1966, Neukirch 1992 II.2.5, Hewitt–

Ross 1963 §10 or Robert 2000 1.4.7. This construction is a modernversion of Hensel’s approach.

By 52.9 the endomorphism ring of the Prufer group Cp∞ is isomorphicto Zp, which gives another possibility to define Zp (and then Qp as well,as the field of fractions); compare Luneburg 1973 p. 113.

Exercise 3 below gives direct algebraic definitions of Zp and Qp.

Each field Qp is isomorphic to a subfield of C, since the algebraicclosure Q �

p of Qp is isomorphic to C by 64.21. One can show that Q �p

has dimension ℵ0 over Qp; see 58.2.Analysis in the locally compact field Qp has many aspects that are

quite different from the usual analysis in R. For details we refer thereader to the books by Robert 2000, Koblitz 1977, Gouvea 1997 andto Burger–Struppeck 1996, and we mention only the following resultof Mahler. A map f : N0 → Zp is continuous with respect to the p-adic

Page 302: The Classical Fields

52 The additive group of p-adic numbers 285

topology on N0 if, and only if, the values an :=∑n

k=0(−1)k(nk

)f(n− k)

tend to 0 as n→∞. Moreover, if this holds, then f has a unique contin-uous extension g : Zp → Zp, given by the series g(x) =

∑n≥0 an

(xn

)in

the binomial polynomials(

xn

):= x(x−1) · · · (x−n+1)/n! and

(x0

):= 1.

For a proof see Cohn 2003a 9.3.8, Robert 2000 4.2 or Cassels 198612.7.

Exercises(1) Show that the series

P∞n=0 pn converges in Qp to (1 − p)−1.

(2) Express as a p-adic series: −1 ∈ Qp, 1/10 ∈ Q3.

(3) Show that the ring Zp is isomorphic to the quotient Z[[x]]/(x − p) of theformal power series ring Z[[x]] modulo the ideal generated by x−p. Show alsothat the field Qp is isomorphic to Z((x))/(x − p), where Z((x)) denotes thering of integral Laurent series (see 64.23).(For the real analogue, see Exercise 2 of Section 6.)

(4) Let a ∈ Z � pZ. Show that the sequence of powers an contains a subse-quence which converges in Qp to 1.

(5) A seriesP

n an converges in Qp if, and only if, limn→∞ an = 0. Deducethat a series in Qp may be rearranged arbitrarily, without changing the con-vergence or the value of the series.

(6) A sequence (an)n in Qp converges if, and only if, limn |an+1 − an|p = 0.

(7) The topological closure in Qp of the set P of primes is (Zp � pZp) ∪ {p}(use Dirichlet’s theorem on primes, which says that an arithmetic progressioncontains infinitely many primes if it contains two coprime numbers).

(8) Let k ∈ N. Define a map f : {0, 1, . . . , k}N → {0, 1}N by replacing eachentry i < k in a sequence from {0, 1, . . . , k}N by i ones followed by a singlezero, and by replacing each occurrence of k by k ones. Show that f is ahomeomorphism of these two product spaces.

(9) Find all elements a ∈ Q×p such that the function f : N → Qp given by

f(n) = an is continuous with respect to the p-adic topology on N.

(10) Let f ∈ Q[x] be a polynomial with rational coefficients which assumes onN only integer values, i.e. f(N) ⊆ Z. Show that f(Z) ⊆ Z.

(11) For every k ∈ N, the p-adic numberP

n≥0 nkpn belongs to Q.

52 The additive group of p-adic numbers

In this section, we consider Q+p and Z+

p as abstract groups and as topo-logical groups. As usual, p denotes a prime number. The Prufer groupsCp∞ introduced in 1.26 will play an important role.

52.1 Proposition For each prime p the additive group Q+p is isomor-

phic to R+ (but compare 52.4 below).

Page 303: The Classical Fields

286 The p-adic numbers

Proof Qp has characteristic zero, hence Q+p is divisible and torsion free.

Furthermore card Qp = card R by 51.8, hence Q+p∼= R+ by 1.14. �

The topological groups R+ and Q+p are both locally compact, but Q+

p

is totally disconnected, since the subgroups pnZp with n ∈ N form aneighbourhood base at 0; see 51.10.

52.2 Lemma For each n ∈ Z, the quotient group Q+p/pnZp is iso-

morphic to the Prufer group Cp∞ . Moreover, one has the following

isomorphisms of abstract groups:

Q+p/Z ∼= Cp∞ ⊕ (Zp/Z) ∼= R/Z and Z+

p/Z ∼=×q∈P�{p} Cq∞ .

Proof For the first assertion it suffices to consider the case n = 0, sincex �→ pnx is an automorphism of Q+

p . Denote by R := Z[p−1] = {ap−m |a ∈ Z,m ∈ N} the subring of Q generated by p−1. Then R ∩ Zp = Z,and the series representations in 51.8 show that Qp = R + Zp. HenceQp/Zp = (R + Zp)/Zp

∼= R/(R ∩ Zp) = R/Z, and R/Z is the p-primarytorsion subgroup Cp∞ of Q/Z as defined in 1.26.

For the second assertion we derive from Qp = R + Zp and R∩Zp = Zthe equation

Qp/Z = (R/Z)⊕ (Zp/Z) = Cp∞ ⊕ (Zp/Z) ,

hence Zp/Z ∼= (Qp/Z)/Cp∞ . From 52.1 and 1.31 we infer that Qp/Z ∼=R/Z ∼=×q Cq∞ . This product of all Prufer groups Cq∞ has only onesubgroup isomorphic to Cp∞ , namely its p-primary torsion subgroup,hence Zp/Z ∼=×q∈P�{p} Cq∞ . �

52.3 Proposition (i) The proper open subgroups of Q+p are precisely

the compact groups pnZp with n ∈ Z.

(ii) The closed subgroups of Q+p are precisely the Zp-submodules of

Q+p , i.e. the subgroups {0}, Qp and pnZp with n ∈ Z.

(iii) Each non-zero ideal of the ring Zp is of the form pnZp with n ∈ N0.

Proof (i) By 51.10 each proper open subgroup U of Q+p contains pmZp

for some m ∈ N; and U/pmZp is a proper subgroup of Qp/pmZp, whichis isomorphic to Cp∞ by 52.2. Each proper subgroup of Cp∞ is finiteand uniquely determined by its order pk; see 1.26. Hence U = pm−kZp.

(ii) Each closed subgroup U is a Zp-submodule by the density (51.9)of Z in Zp. If U �= {0}, then the Zp-submodule U is open, and theassertion follows from (i).

(iii) is a consequence of (ii). �

Page 304: The Classical Fields

52 The additive group of p-adic numbers 287

We remark that the groups Z+p and Q+

p are the only non-discretelocally compact abelian Hausdorff groups with the property that everynon-trivial closed subgroup is open; see Robertson–Schreiber 1968.Moreover, the groups Z+

p are the only infinite compact Hausdorff groupssuch that all non-trivial closed subgroups are topologically isomorphic(see Morris–Oates-Williams 1987), and these groups are also theonly non-discrete locally compact Hausdorff groups such that all non-trivial closed subgroups have finite index (see Dikranjan 1979, Morris

et al. 1990).

52.4 Corollary Let p and q be distinct primes. Then the topological

groups Q+p and Q+

q are not isomorphic.

Proof This is a consequence of 52.3(i): the quotient pnZp/pn+1Zp∼=

Zp/pZp of two neighbours in the chain of all proper open subgroups ofQ+

p is cyclic of order p (see 51.7). One could also use 52.3(i) and 52.2,or 52.3(i) and 52.6 below. �

Incidentally, 52.4 provides another proof for the fact (36.5) that thegroup Q+ with the p-adic topology is not isomorphic to Q+ with theq-adic topology (using the existence and uniqueness of completions; see43.21 and 43.25).

Now we focus on properties of the group Z+p .

52.5 Definition Modifying the concept of divisibility (1.7), we saythat an abelian group A+ is p-divisible, if pA = A; this means that eachelement of A is of the form pa for some a ∈ A.

52.6 Lemma The group Z+p has no p-divisible subgroup except {0},

and Z+p is q-divisible for every prime q �= p.

Proof If U ≤ Z+p is p-divisible, then U = pnU ⊆ pnZp for each n ∈ N,

hence U ⊆ ⋂n pnZp = {0}; see 51.6. Furthermore, |q|p = 1, so q is aunit of the ring Zp, and qZp = Zp. �

In particular, the group Z+p is reduced, i.e., {0} is the only divisible

subgroup of Z+p . Lemma 52.6 shows that the abstract groups Z+

p withp ∈ P are mutually not isomorphic; in fact, more is true.

52.7 Corollary Let p and q be different primes. Then every group

homomorphism ϕ : Zp → Zq is trivial.

Proof The image ϕ(Zp) = ϕ(qZp) = qϕ(Zp) is a q-divisible subgroup ofZ+

q , hence trivial by 52.6. �

Page 305: The Classical Fields

288 The p-adic numbers

52.8 Proposition Each endomorphism of the group Z+p is continuous

and of the form z �→ az for some a ∈ Zp. The endomorphism ring

End Z+p is isomorphic to the ring Zp via the isomorphism

End Z+p → Zp : α �→ α(1) .

Proof By 51.10 the sets pnZp with n ∈ N form a neighbourhood baseat 0. If α ∈ End Z+

p , then α(x+ pnZp) = α(x)+ pnα(Zp) ⊆ α(x)+ pnZp

for each x ∈ Zp, hence α is continuous (in fact, non-expanding withrespect to the p-adic absolute value).

One has α(z) = α(z ·1) = zα(1) for each integer z ∈ Z. This equationholds also for all z ∈ Zp, because Z is dense in Zp by 51.9 and becauseα and the multiplication of Zp are continuous. Hence the evaluationmap α �→ α(1) is injective. It is also surjective, since z �→ za is anendomorphism of Z+

p for every a ∈ Zp. �

Now we determine the endomorphisms, the automorphisms and thecharacters of the Prufer group Cp∞ ; see Section 63 for general infor-mation on characters. Recall that the endomorphism ring of the finitecyclic group Cpn is isomorphic to the ring Z/pnZ (see Lang 1993 IITheorem 2.3).

52.9 Theorem The endomorphism ring of the group Cp∞ is isomorphic

to the ring Zp, even as topological rings, if we endow EndCp∞ with the

compact-open topology derived from the discrete topology on Cp∞ .

The automorphism group of Cp∞ is isomorphic to the group Z×p ={

x ∈ Qp

∣∣ |x|p = 1}

of units of Zp (see 53.3).The character group C ∗

p∞ = Hom(Cp∞ , T), as usual endowed with the

compact-open topology, is isomorphic to the topological group Z+p .

Proof Each a ∈ Zp yields an endomorphism αa of Qp/Zp = Cp∞ simplyby multiplication: αa(x + Zp) := ax + Zp. The ring homomorphismα : Zp → EndCp∞ : a �→ αa is injective, since αa = 0 implies aQp ⊆ Zp,hence a = 0.

Now we show that α is also surjective. Let ϕ be an endomorphism ofCp∞ = Qp/Zp. This group can be written as the union of the groupsZpp

−n/Zp with n ≥ 0. The unique (cyclic) subgroup Zpp−n/Zp of order

pn in Qp/Zp is generated by p−n + Zp and invariant under ϕ, hence wefind integers zn ∈ {0, 1, . . . , pn − 1} with ϕ(p−n + Zp) = znp−n + Zp foreach n ≥ 0. Applying ϕ to the relation p(p−n−1 + Zp) = p−n + Zp inQp/Zp we obtain zn+1 ≡ zn mod pn for n ≥ 0, hence zn+1 = zn + cnpn

with suitable coefficients cn ∈ {0, 1, . . . , p − 1}. Now the p-adic integer

Page 306: The Classical Fields

52 The additive group of p-adic numbers 289

a :=∑∞

n=0 cnpn satisfies αa(p−m + Zp) = (∑m−1

j=0 cjpj)p−m + Zp =

zmp−m + Zp = ϕ(p−m + Zp) for each m ≥ 0, hence αa = ϕ.For a discrete topological space X, the compact-open topology of

XX coincides with the ‘point-open’ topology, which is just the pro-duct topology. Since we endow Cp∞ with the discrete topology, thecompact-open topology of EndCp∞ has a subbasis consisting of the sets{ϕ | ϕ(x + Zp) = y + Zp } with x, y ∈ Qp. Under α−1 such a set cor-responds to {a ∈ Zp | ax ∈ y + Zp }, which is open in Zp. Hence thebijection α : Zp → EndCp∞ is continuous, in fact a homeomorphism, asZp is compact.

The image of any group homomorphism Cp∞ → T = R/Z consistsof elements whose order is a power of p, hence that image is containedin the p-primary torsion subgroup Cp∞ of T. Thus Hom(Cp∞ , T) =Hom(Cp∞ , Cp∞) is just the additive group of EndCp∞ . �

52.10 Some infinite Galois groups For a prime p and n ∈ N wedenote by Wp,n = { exp(2πia/pn) | a ∈ {0, 1, . . . , pn − 1}} the groupof all complex roots of unity of order dividing pn. The automorphismgroup Aut Q(Wp,n) of the cyclotomic field Q(Wp,n) is isomorphic to thegroup of units (Z/pnZ)×, via restriction to the cyclic group Wp,n oforder pn; compare Cohn 2003a 7.7.5 or Lang 1993 VI Theorem 3.1.The multiplicative group Wp :=

⋃n Wp,n is isomorphic to Cp∞ , and the

field Q(Wp) =⋃

n Q(Wp,n) is an infinite Galois extension of Q. Eachfield automorphism of Q(Wp) restricts to a group automorphism of Wp.The restriction Aut Q(Wp) → Aut(Wp) is injective, and also surjectiveby the above remark on Aut Q(Wp,n). Thus we infer from 52.9 that

Aut Q(Wp) ∼= AutCp∞ ∼= Z×p .

For the group-theoretic structure of Z×p see 53.3.

Denote by A := Q(exp(2πiQ)) = Q(⋃

p Wp) the field generated byall complex roots of unity. The Kronecker–Weber Theorem (compareCassels 1986 10.12, Lang 1970 X §3 or Neukirch 1992 V.1.10) im-plies that A|Q is the largest Galois extension of Q with abelian Galoisgroup. The Galois group Aut A is isomorphic to the Cartesian product×p Aut Q(Wp); see Ribenboim 1999 11.2. From the previous paragraphwe obtain that

Aut A ∼=×p∈P Z×p = Z×

is isomorphic to the group of units of the product ring Z :=×p Zp (seealso Weil 1967 XIII §4 Corollary 2).

Page 307: The Classical Fields

290 The p-adic numbers

This product ring Z is isomorphic to the ring End(Q+/Z) consid-ered in 31.11 and 31.12: by 1.28, we have the direct decompositionQ/Z ∼= ⊕p Cp∞ , and each primary component Cp∞ is invariant under

every endomorphism, hence End(Q+/Z) ∼=×p∈P EndCp∞ ∼=×p∈P Zp

by 52.9. The ring Z may also be described as the completion (44.5) ofZ with respect to the topology generated by all non-zero ideals.

In fact, infinite Galois groups carry a natural compact, totally dis-connected (or profinite) group topology, and the isomorphisms obtainedhere can be read as isomorphisms of topological groups; see Ribenboim

1999 11.1, 11.2.For prime numbers p, q let Fp,q :=

⋃n∈N Fpqn (in a fixed algebraic clo-

sure F �p of Fp). The automorphism group of the field Fp,q is isomorphic

to Z+q (Exercise 7). The algebraic closure F �

p is generated by its subfieldsFp,q with q ∈ P, and one can show that

Aut F �p∼=×q∈P Z+

q ;

see Ribenboim 1999 11.2 pp. 316–320. The product group appearinghere is the additive group of the ring Z as above, and the Frobeniusautomorphism x �→ xp generates a dense cyclic subgroup of Aut F �

p ;compare Bourbaki 1990 V.12.3, Neukirch 1992 IV.2.

Finally we show that the compact group topology of Zp is uniquelydetermined by the abstract group Z+

p ; the following result is due toSoundararajan 1969; see also Corwin 1976, Kallman 1976.

52.11 Theorem The only locally compact Hausdorff group topologies

of Z+p are the natural p-adic topology, which is compact, and the discrete

topology.

Proof For every n ∈ N the map x �→ pnx : Zp → Zp is continuous withrespect to any group topology τ on Zp. If τ is a compact Hausdorfftopology, then the image pnZp is compact and hence closed in Zp withrespect to τ . By 51.7 the index |Zp/pnZp| = |Z/pnZ| = pn is finite,and pnZp is also open with respect to τ . The p-adic topology τp ofZp has {pnZp | n ∈ N} as a neighbourhood base at 0. Therefore theidentity id : (Zp, τ) → (Zp, τp) is continuous at 0, hence everywhere.This bijective map between two compact Hausdorff spaces is also closed,hence a homeomorphism, which means that τ = τp.

Now let τ be a non-discrete locally compact Hausdorff group topologyof Z+

p . We shall show that τ is compact, which completes the proof.Lemma 52.6 implies that the divisible groups Rn with n > 0 do not

Page 308: The Classical Fields

52 The additive group of p-adic numbers 291

occur as subgroups of Z+p , hence the Splitting Theorem 63.14 entails

that (Zp, τ) has a compact open subgroup C �= {0}. We are going touse the character group C∗ as defined in Section 63 and the Pontryaginduality C∗∗ ∼= C; see 63.20.

Since Zp and C are torsion free, the character group C∗ is divisible by63.32, and C∗ is discrete by 63.5. If C∗ were torsion free, then C∗∗ ∼= C

would be divisible, again by 63.32; but this would contradict 52.6. ThusC∗ contains a torsion element of prime order q, hence by divisibilityalso a Prufer subgroup Π ∼= Cq∞ (compare 31.7). By 1.23 and 1.22this group Π is a direct summand of C∗, hence Π∗ is a subgroup (evena direct summand) of C∗∗ ∼= C (see 63.8). From 52.9 we know thatΠ∗ ∼= Z+

q , and 52.7 implies that q = p.We conclude that C contains a subgroup P ∼= Π∗ ∼= Z+

p . Each (ab-stract) isomorphism Zp → P is a non-zero endomorphism of Zp, hencethe image P has finite index in Zp by 52.8. Thus C ⊇ P has finite indexin Zp, whence τ is compact. �

Exercises

(1) Show that Q+p/Q ∼= R+ as abstract groups.

(2) Each subgroup of finite index in Z+p is of the form pnZp with n ∈ N0, and

Z+p/pnZp is cyclic of order pn.

(3) The elements of finite order in Zp/Z form a subgroup isomorphic to thedirect sum of all Prufer groups Cq∞ with q ∈ P � {p}.(4) The composition of the natural maps Qp → Qp/Zp

∼= Cp∞ → C× isthe function χ : Qp → C× with χ(

P

k ckpk) = exp(2πiP

k<0 ckpk) for ck ∈{0, 1, . . . , p − 1}. Show that χ is a continuous homomorphism of Q+

p into

the multiplicative group C×, and determine the kernel and the image of χ.Conclude that χ is a non-trivial character of Q+

p (the Tate character).

(5) Let F be a locally compact field which is neither discrete nor indiscrete,let χ1 : F+ → R/Z be a non-trivial character, and define χa by χa(x) = χ(ax)for a, x ∈ F . Show that the map a �→ χa is an isomorphism of the topologicalgroup F+ onto the character group of F+. In particular, the character groupof Q+

p is isomorphic to Q+p .

(6) The p-adic solenoid Sp can be defined as the topological quotient groupSp := (R × Zp)/{ (z, z) | z ∈ Z}. Show that Sp is a compact connectedHausdorff group, and determine the torsion subgroup of Sp.

(7) Let p, q be prime numbers. Show that the automorphism group of the fieldS

n Fpqn is isomorphic to Z+q .

Page 309: The Classical Fields

292 The p-adic numbers

53 The multiplicative group of p-adic numbers

Here we consider the structure of the multiplicative group Q×p of non-zero

p-adic numbers. It turns out (in Theorem 53.2) that Q+p and Q×

p haveopen subgroups which are isomorphic (as topological groups), similarlyas for R; compare 9.2. Furthermore we show in 53.5 that Qp has no fieldendomorphism apart from the identity (like R; see 6.4).

Let p always denote a prime number.

53.1 Proposition (i) For each n ∈ N, the set Un := 1 + pnZp is a

multiplicative subgroup of Q×p , and Un is open and closed in Qp.

Furthermore Un/Un+1∼= F+

p∼= Cp.

(ii) The field Qp contains a root of unity ζ of order p − 1, and the set

{0, 1, ζ, ζ2, . . . , ζp−2} is a system of representatives for Zp/pZp.

(iii) The group Q×p is the direct product Q×

p = 〈p〉 × 〈ζ〉 × U1 of the

three subgroups 〈p〉, 〈ζ〉 and U1, and Z×p = {x ∈ Qp | |x|p = 1} =

〈ζ〉 × U1.

Proof (i) Clearly the natural ring epimorphism Zp → Zp/pnZp �= {0}restricted to Z×

p = Zp � pZp is multiplicative, hence the kernel Un =1 + pnZp of that restriction is a subgroup of Q×

p . By 51.6, the subgroupUn is open and closed in Qp.

The mapping f : Un → Z+p/pZp defined by f(1 + pnx) = x + pZp for

x ∈ Zp is an epimorphism of groups, since f maps (1 + pnx)(1 + pny) =1+ pn(x+ y + pnxy) onto x+ y + pnxy + pZp = x+ y + pZp. The kernelof f is 1 + pnpZp = Un+1, hence Un/Un+1

∼= Zp/pZp∼= F+

p by 51.7.(ii) Since F×

p is cyclic (Section 64, Exercise 1), we find an integeri ∈ {1, 2, . . . , p − 1} which represents a generator of F×

p = (Z/pZ)×.For each n ∈ N the group of units of the finite ring Z/pnZ has orderpn − pn−1 (see Lang 1993 II.2), and i represents such a unit, henceip

n−pn−1 ≡ 1 mod pn and

ipn ≡ ip

n−1mod pn .

This means that the two p-adic series representing ipn

and ipn−1

, respec-tively, have the same initial coefficients c0, c1, ..., cn−1. Thus there existsa sequence of integer coefficients ck ∈ {0, 1, . . . , p− 1} with

ipn ≡∑n−1

k=0 ckpk mod pn

for each n ≥ 0.We define ζ ∈ Zp by ζ =

∑k≥0 ckpk (thus ζ = limn→∞ ip

n

); notethat ζ �= 0, as c0 = i. Modulo pn+1Zp we have ζ ≡ ip

n

and thereforeζp−ζ ≡ ip

n+1− ipn ≡ 0 for each n ∈ N, whence ζp−ζ = 0 and ζp−1 = 1.

Page 310: The Classical Fields

53 The multiplicative group of p-adic numbers 293

As ζ + pZp = i + pZp has multiplicative order p− 1 in (Zp/pZp)× ∼= F×p

(see 51.7), we infer that ζ is a root of unity of order exactly p− 1.Other methods to find ζ can be based on Hensel’s Lemma or Newton’s

rule; see Cohn 2003a p. 326, compare also Serre 1973 p. 16 or Weil

1967 I.4 Theorem 7 p. 16.(iii) The group epimorphism | |p : Q×

p → pZ has the kernel {x ∈ Qp ||x|p = 1} = Z×

p , hence Q×p = 〈p〉 × Z×

p . We infer from (ii) that Z×p is

the disjoint union of the cosets ζj + pZp = ζj(1 + pZp) = ζjU1 with0 ≤ j ≤ p− 2 (note that ζZp = Zp). Hence Z×

p = 〈ζ〉 × U1. �

The following result is crucial for relating the multiplicative group Q×p

to the additive group of p-adic numbers. Often this is achieved usingthe series for the exponential function exp (see Exercise 2). The proofbelow employs a variant of the usual exponential function: we use thebasis 1+p instead of e (see Warner 1989 22.2f for exponential functionswith arbitrary basis).

53.2 Theorem If p �= 2, then U1 = 1 + pZp ≤ Q×p is isomorphic to Z+

p

as a topological group. For p = 2, we have U1 = 1 + 2Z2 = 〈−1〉 × U2,

and U2 = 1 + 4Z2 is isomorphic to Z+2 as a topological group.

Proof The mapping β0 : Z→ U1 : z �→ (1+p)z is a group monomorphismof Z+ into the multiplicative group U1. Moreover β0 is continuous whenZ is provided with the p-adic topology, since |(1 + p)z − 1|p ≤ p−1|z|pby 37.5. As U1 = 1 + pZp is compact, β0 has a unique extension toa continuous group homomorphism β : Zp → U1; see 43.9, 43.23. (ByExercise 3, this extension is given by β(x) = (1 + p)x :=

∑n≥0

(xn

)pn,

but we do not need this series representation.)The kernel of β is a closed subgroup of Z+

p . Furthermore β is injectiveon the dense subset N, hence the kernel cannot be open in Zp. Thus by52.3(i, ii) the kernel is trivial and β is injective on Zp.

The image β(Zp) is the topological closure of the cyclic group β(Z) =(1 + p)Z. We have |U1/Un+1| = pn by 53.1(i). If p �= 2 and n ≥ 1, then37.5 shows that

|(1 + p)pn−1 − 1|p = |p|p|pn−1|p = p−n ,

hence (1 + p)pn−1 �∈ 1 + pn+1Zp = Un+1 (see also Exercise 4). Thus(1 + p)Un+1 has order pn in U1/Un+1 and is therefore a generator ofU1/Un+1. Since the subgroups Un with n ≥ 1 form a neighbourhoodbase at 1 in U1, this implies that 1 + p generates a dense subgroup ofU1, whence β(Zp) = U1. By compactness β is a homeomorphism.

Page 311: The Classical Fields

294 The p-adic numbers

It remains to deal with the case p = 2. Here (1 + p)Z = 3Z is notdense in U1 (in fact, 3U3 is an involution in the elementary abeliangroup U1/U3

∼= (Z/8Z)× of order 4). We have U1 = 〈−1〉 ×U2, becauseZ2 = 2Z2∪(1+2Z2) by 51.7, hence U1 = 1+2Z2 = (1+4Z2)∪(3+4Z2) =U2∪−U2. We replace β0 by the mapping γ0 : Z→ U2 : z �→ (1+4)z andobtain by extension as above a continuous monomorphism γ : Z2 → U2

of groups. The image γ(Z2) is the topological closure of 5Z in U2. From37.5 we infer that

|(1 + 4)2n−1 − 1|2 = 2−(n+1)

for n ≥ 1, hence (1 + 4)2n−1 �∈ 1 + 2n+2Z2 = Un+2 (see also Exercise 4).

Thus 5 represents an element of order 2n in the quotient group U2/Un+2,which has order 2n by 53.1(i). Therefore 5 generates a dense subgroupof U2, and γ : Z2 → U2 is surjective. �

53.3 Corollary Let ζ be as in 53.1, and write Cn for a cyclic group of

order n. If p �= 2, then

Q×p = 〈p〉 × 〈ζ〉 × (1 + pZp) ∼= Z× Cp−1 × Z+

p and Z×p∼= Cp−1 × Z+

p .

For p = 2, we have

Q×2 = 〈2〉 × 〈−1〉 × (1 + 4Z2) ∼= Z× C2 × Z+

2 and Z×2∼= C2 × Z+

2 .

The isomorphisms can be read as isomorphisms of topological groups,

where the cyclic factors are discrete.

Proof This is a direct consequence of 53.1 and 53.2. �

53.4 Corollary Let p and q be distinct primes. Then the abstract

groups Q×p and Q×

q are not isomorphic. In particular, the fields Qp and

Qq are not isomorphic.

Proof By 53.3 and 52.6, Q×p contains an infinite (in fact, uncountable)

r-divisible subgroup only for the prime r = p.Alternatively, 53.3 shows that the torsion subgroup of Q×

p (i.e. thegroup of all roots of unity of Qp) has order p−1 or 2; then it remains toconsider the case p = 2, q = 3. By 54.1, the group Q×

3 consists of foursquare classes, but Q×

2 has eight square classes. �

53.5 Theorem Let α : Qp → Qq be a ring homomorphism, with

α(1) = 1. Then p = q and α is the identity. In particular, the identity

is the only non-zero field endomorphism of Qp.

Page 312: The Classical Fields

54 Squares of p-adic numbers and quadratic forms 295

Proof The kernel of α is a proper ideal, hence α is injective. The imageα(Q×

p ) ∼= Q×p contains a subgroup isomorphic to Z+

p by 53.2. But Q×q

has such a subgroup only if q = p; see 53.3 and 52.6.Results 53.2, 53.3 and 52.6 imply that U1 = 1 + pZp is the set of

all non-zero p-adic numbers that admit r-th roots for each r ∈ N notdivisible by p:

1 + pZp ={

x ∈ Q×p

∣∣ ∀r∈N�pN∃y x = yr

}=⋂

r∈N�pN(Q×

p )r .

As α is multiplicative, we infer that α(1+pZp) ⊆ 1+pZp. Since α is alsoadditive, we obtain α(pZp) ⊆ pZp and then α(pnZp) ⊆ pnZp for n ≥ 1.(See Exercise 5 for an alternative proof of the inclusion α(Zp) ⊆ Zp.)This shows that α is continuous at 0, hence everywhere (by additivity).

Each rational number is fixed by α, and Q is dense in Qp by 51.5,whence α is the identity. �

Exercises(1) Show that Z×

p/U1∼= F×

p and Un∼= Z+

p for n ≥ 2.

(2) Determine the domain of p-adic convergence of the exponential seriesexp x =

P

n≥0 xn/n!.

(3) Consider the binomial coefficient`

xn

´

= x(x − 1) · · · (x − n + 1)/n! with

n ∈ N0 as a polynomial in x, where`

x0

´

:= 1. Show that the binomial series

β(x) = (1 + p)x :=P

n≥0

`

xn

´

pn converges for x ∈ Zp and defines a continuoushomomorphism β : Zp → U1 of groups.

(4) Let n ∈ N. Show that (1 + p)pn−1 ≡ 1 + pn mod pn+1 for p = 2, and

(1 + 4)2n−1 ≡ 1 + 2n+1 mod 2n+2.

(5) Show that Zp = {x ∈ Qp | 1 + px2 is a square in Qp } for p = 2, and thatZ2 = {x ∈ Q2 | 1 + 2x3 is a cube in Q2 }. Use these facts to give anotherproof of 53.5.

(6) Let p = 2 and n ∈ N � pN. Show that a non-zero p-adic numberpk P

m≥0 cmpm with c0 = 0 is an nth power in Q×p if, and only if, n divides k

and c0 is an nth power modulo p.

54 Squares of p-adic numbers and quadratic forms

For any field F we denote by F � := {x2 | x ∈ F× } the group of allnon-zero squares of F , and we call each coset aF � ∈ F×/F � a squareclass of F . The field R of real numbers has only two square classes, thepositive and the negative real numbers. For p-adic numbers the situationis slightly more complicated. As usual, p is a prime number.

Page 313: The Classical Fields

296 The p-adic numbers

54.1 Theorem For p �= 2 we have Q�p = 〈p2〉 × 〈ζ2〉 × (1 + pZp), and

Q×p/Q�

p consists of four square classes, with representatives 1, ζ, p, ζp,

where ζ is a root of unity of order p − 1 as in 53.1. Another system

of representatives is given by 1, r, p, rp, where r ∈ Z is not a square

modulo p.

For p = 2 we have Q�2 = 〈4〉× (1 + 8Z2), and Q×

2/Q�2 consists of eight

square classes, with representatives ±1, ±2, ±3, ±6; also ±1, ±2, ±5,

±10 are representatives.

Proof For p �= 2, this product decomposition of Q�p follows directly from

the decomposition of Q×p in 53.3; note that 1 + pZp

∼= Z+p is 2-divisible

by 52.6. For p = 2, result 53.3 implies that Q�2 = 〈4〉 × U2

2 , whereUn = 1 + 2nZ2, and 53.1(i) yields |U2/U3| = 2. Hence U2

2 ⊆ U3, and wehave in fact equality, as U2

∼= Z+2 by 53.2 and |Z2/2Z2| = 2 by 51.7.

The systems of representatives are obtained by combining coset rep-resentatives of factors of Q�

2 in the corresponding factors of Q×p . �

54.2 Corollary The field Qp is not formally real, hence it cannot be

made into an ordered field.

Proof An ordered field has only ±1 as roots of unity. By 53.1(ii), itremains to consider the two cases p = 2, 3. Using 54.1 (or 53.2 and 52.6)we see that −2 = 1 + 3(−1) is a square in Q3 and that −7 = 1 + 8(−1)is a square in Q2 (more generally, 1− p3 ∈ Q�

p for any prime p). Hencenone of these fields is formally real; compare 12.3.

We give a second, topological proof. Assume that P is a domain ofpositivity (11.3) of Qp. Then both P and −P are unions of square classesof Qp, hence open in Qp by 54.1. Thus P ∪{0} is topologically closed inQp and therefore contains the infinite sum

∑n≥0(p− 1)pn = −1, which

is a contradiction (alternatively, P ∪ {0} contains N and its topologicalclosure Zp ⊇ Z; see 51.9). �

54.3 Corollary If p �= 2, then Qp has precisely three quadratic field

extensions, namely Qp(√

a) with a ∈ {p, r, pr}, where r ∈ Z is not a

square modulo p. The field Q2 has precisely seven quadratic extensions,

namely Q2(√

a) with a ∈ {−1,±2,±3,±6}.Proof Every quadratic field extension of Qp is of the form Qp(

√a) with

a ∈ Qp � Q�p . Hence 54.1 shows that we have listed all quadratic exten-

sions of Qp. Furthermore, an element b ∈ Q×p is a square in Qp(

√a) if,

and only if, b belongs to the square class of 1 or of a in Qp. This showsthat the listed extensions are distinct. �

Page 314: The Classical Fields

54 Squares of p-adic numbers and quadratic forms 297

Corollary 54.3 is a special case of the following general fact: the fieldQp has only finitely many field extensions of given degree n, for eachn ∈ N; see 58.2.

Now we consider sums of squares in Qp and, more generally, quadraticforms over Qp.

54.4 Lemma Let a ∈ Q×p and Sa := {x2 − ay2 | x, y ∈ Qp }. If a is a

square in Qp, then Sa = Qp. If a is not a square in Qp, then Sa � {0} is

a subgroup of index 1 or 2 in Q×p .

Proof If a is a square, then Sa = {x2 − y2 | x, y ∈ Qp } contains allelements of the form (y + 1)2− y2 = 2y + 1 with y ∈ Qp, hence Sa = Qp

(this argument is a special case of 54.8 below).The equation (x2 − ay2)(z2 − aw2) = (xz + ayw)2 − a(xw + yz)2,

which is related to Diophant’s identity (34.4), shows that Sa is closedunder multiplication. Sa contains all squares of Q×

p , hence the formulax−1 = x(x−1)2 implies that Sa � {0} is a subgroup of Q×

p .Clearly −a ∈ Sa. If −a is not a square, then Sa contains at least two

square classes, hence by 54.1 the index of Sa � {0} in Q×p is at most 2,

provided that p �= 2. If −a is a square, then Sa = {x2 + y2 | x, y ∈ Qp }does contain a non-square (and therefore at least two square classes), aswe show by an indirect argument: otherwise Sa = {x2 | x ∈ Qp } wouldbe closed under addition, which is a contradiction, as p is not a squareby 54.1.

It remains to deal with the case p = 2. By 54.1 it suffices to considerthe seven cases a = −1,±2,±3,±6. Now 1,−a, 1 − a, 4 − a ∈ Sa, andone easily verifies in each of these seven cases that 1,−a, 1 − a, 4 − a

represent at least three distinct square classes (Scharlau 1985 p. 188and Conway 1997 p. 120 enumerate the square classes in Sa ⊆ Q2 foreach value a). Since |Q×

2/Q�2 | = 8, the subgroup Sa � {0} has index at

most 2 in Q×2 . �

Let a ∈ Qp be a non-square. Then Sa is the set of all norms ofthe quadratic Galois extension Qp(

√a)|Qp. Improving on 54.4, one can

show that Sa � {0} is always a subgroup of index 2 in Q×p , and one

can describe this subgroup in terms of the Hilbert symbol (which is acertain bilinear form on the vector space Q×

p/Q�p over F2); see Serre

1973 p. 19/20, Borevich–Shafarevich 1966 Chapter 1 §6, Cassels

1978 p. 56 or Lam 2005 VI.2.

54.5 Corollary If p ≡ 1 mod 4, then every p-adic number is a sum of

two (non-zero) squares of Qp.

Page 315: The Classical Fields

298 The p-adic numbers

Proof With ζ as in 54.1, we have −1 = ζ(p−1)/2. Hence −1 is a squarein Qp, and S−1 = Qp by 54.4. The formulae a2 = (3a/5)2 +(4a/5)2 and0 = 1 + (−1) show that every square in Qp is a sum of two non-zerosquares. �

In the following we consider quadratic forms f(x) =∑n

i=1 aix2i in

n variables over a field F with coefficients ai ∈ F . We say that f isisotropic over F , if f(x) = 0 has a solution x ∈ F n with x �= 0.

54.6 Lemma Let p �= 2. Then each quadratic form ax2 + by2 + cz2

with a, b, c ∈ Qp and |a|p = |b|p = |c|p is isotropic over Qp.

Proof Apart from the trivial case a = b = c = 0, we may assume that|a|p = |b|p = |c|p = 1, that is, a, b, c ∈ Zp � pZp. Then a + pZp andb + pZp are non-zero elements of Zp/pZp

∼= Fp; see 51.7. By 34.15 wefind integers x, y ∈ Z such that

ax2 + by2 ∈ −c + pZp = −c(1 + pZp) .

As p �= 2, the set 1 + pZp consists of squares of Q×p by 54.1, hence

ax2 + by2 = −cz2 for some z ∈ Q×p . �

The following result shows a remarkable contrast between Qp and R.To mention just one of the consequences of 54.7, there is no octonion(Cayley–Dickson) division algebra over Qp (see Lam 2005 X.2 p. 327).

54.7 Theorem Each quadratic form∑n

i=1 aix2i with ai ∈ Qp and n ≥ 5

is isotropic over Qp.

Proof We may assume that n = 5 and ai ∈ Q×p . By a substitution

xi �→ pn(i)xi we can achieve that |ai|p ∈ {1, p} for 1 ≤ i ≤ 5. Thenat least three of the absolute values |ai|p coincide, and 54.6 gives theassertion for p �= 2.

For the rest of the proof let p = 2. We show by ad hoc arguments thatf(x) =

∑5i=1 aix

2i is isotropic over Q2. If the five coefficients ai belong

to five different square classes of Q×2 , then by 54.1 the quotient of two

of them, say ai and aj , belongs to the square class of −1, i.e., −aia−1j

is a square in Q2, and then the forms aix2i + ajx

2j and f are isotropic.

Hence we may assume that a1 = a2 = 1 after changing f by a factor.The quadratic form x2

1 +x22 takes the values 1 (hence also the squares

−7 and −15; see 54.1), 1+1 = 2, 1+4 = 5, 1+9 = 10. If a3x23+a4x

24 takes

one of the values −1,−2,−5,−10, then the forms x21 + x2

2 + a3x23 + a4x

24

and f are isotropic. Hence we may assume that a3x23 + a4x

24 avoids the

square classes of −1,−2,−5,−10. Lemma 54.4 implies that a3x23 +a4x

24

Page 316: The Classical Fields

54 Squares of p-adic numbers and quadratic forms 299

takes as values at least four square classes of Q2, hence it takes thevalues 1, 2, 5, 10; compare 54.1.

Thus the sum x21+x2

2+a3x23+a4x

24 takes the values 2+5 = 7,−7+2 =

−5,−7+5 = −2 and −15+5 = −10, in addition to the values 1, 2, 5, 10taken by x2

1 +x22. As −7 is a square, 7 belongs to the square class of −1.

In view of 54.1 this shows that x21 +x2

2 +a3x23 +a4x

24 takes values in any

square class of Q×2 , in particular it takes the value −a5. This implies

that f is isotropic. �

Quadratic forms∑n

i=1 aix2i over Qp are well understood; the non-

degenerate ones (all ai �= 0) are classified by the number n of variables,the square class of

∏i ai and the Hasse (–Minkowski) invariant; see

Cassels 1978, Scharlau 1985 and Lam 2005 for more information.The following general result says that non-degenerate isotropic qua-

dratic forms are surjective:

54.8 Lemma Let F be a field with characteristic distinct from 2, and

let f(x) =∑n

i=1 aix2i be a quadratic form with ai ∈ F×. If f is isotropic

over F , then f : F n → F is surjective.

Proof By assumption, f(x) = 0 for some non-zero vector x ∈ Fn, sayx1 �= 0. We have f(tx1 + 1, tx2, . . . , txn) = t2f(x) + a1(2tx1 + 1) =a1(2tx1 + 1) for every t ∈ F , hence f is surjective. �

54.9 Corollary Let a1, a2, a3, a4 ∈ Q×p . Then the quadratic form

f : Q4p → Qp with f(x) =

∑4i=1 aix

2i is surjective.

Proof Let c ∈ Qp. By 54.7 the quadratic form f(x) − cx25 is isotropic,

hence f(x) − cx25 = 0 for some non-zero vector (x, x5) ∈ Q5

p. If x5 �= 0,then f(x−1

5 x) = c. If x5 = 0, then x �= 0 and f is isotropic, hencesurjective by 54.8. �

54.10 Corollary For p �= 2, every element of Qp is a sum of three

squares of Qp. Every element of Q2 is a sum of four squares of Q2.

Proof By 54.6 the form x2 + y2 + z2 is isotropic over Qp for p �= 2,hence surjective by 54.8. The assertion about Q2 is a consequence of54.9 (with ai = 1). �

The last corollary shows again that the fields Qp are not formally real(54.2). The fact that every p-adic number is a sum of four squares ofp-adic numbers may also be inferred from Lagrange’s theorem (34.18) bya topological argument: the set S := {a2 + b2 + c2 + d2 | a, b, c, d ∈ Qp }

Page 317: The Classical Fields

300 The p-adic numbers

is the union of {0} and of finitely many square classes of Qp, hence S istopologically closed in Qp by 54.1; furthermore N ⊆ S by 34.18, henceZp ⊆ S by 51.9, and Qp =

⋃n≥0 p−2nZp ⊆ S.

Exercises(1) Find all primes p such that −1 is a square in Qp.

(2) For p = 2, a p-adic number pr(c0 + c1p + c2p2 + · · · ) with r ∈ Z, ci ∈

{0, 1, . . . , p − 1} is a square in Qp if and only if r is even and c0 is a squaremodulo p. For p = 2 the squares are characterized by the condition that r iseven and c0 + 2c1 + 4c2 ≡ 1 mod 8.

(3) Characterize the square-free integers d ∈ Z such that Q(√

d) is isomorphicto a subfield of Qp.

(4) Show that −1 is not a sum of three squares in Q2; in fact, x2 + y2 + z2

represents all square classes of Q2 except the square class of −1.

(5) Let f(x) =Pn

i=1 aix2i be a quadratic form with integer coefficients ai ∈ Z

and n ≥ 5. Then f is isotropic over the finite ring Z/mZ for every integerm ≥ 2.

55 Absolute values

In this section, we define the general concept of absolute values, and weprove two results of Ostrowski which determine all absolute values of thefield Q of rational numbers, and all fields with an Archimedean absolutevalue.

55.1 Definition An absolute value ϕ of a field (or skew field) F isa non-trivial homomorphism ϕ : F× → Rpos of multiplicative groupswhich is subadditive, which means that

ϕ(x + y) ≤ ϕ(x) + ϕ(y)

for all x, y ∈ F× with x + y ∈ F×. It is convenient to define ϕ(0) = 0;then ϕ(xy) = ϕ(x)ϕ(y) and ϕ(x + y) ≤ ϕ(x) + ϕ(y) for all x, y ∈ F .

An absolute value ϕ is said to be Archimedean, if ϕ(n ·1) > 1 for somen ∈ N. Otherwise, we have ϕ(n · 1) ≤ 1 for all n ∈ N, and then we saythat ϕ is non-Archimedean.

The multiplicative group Rpos is torsion free, hence every absolutevalue ϕ satisfies ϕ(r) = 1 whenever r ∈ F is a root of unity. In particularϕ(−1) = 1, and this gives ϕ(−x) = ϕ(x) for every x ∈ F . If thecharacteristic of F is not zero, then the non-zero elements of the primefield of F are roots of unity, hence every absolute value of F is non-Archimedean.

Page 318: The Classical Fields

55 Absolute values 301

55.2 Examples The ordinary absolute values on Q, R or C are definedby |x| = max{x,−x} for x ∈ R and by |z| = (zz)1/2 = (a2 + b2)1/2 forz = a + ib ∈ C, a, b ∈ R. These are examples of Archimedean absolutevalues. The p-adic absolute value | |p on Q and its extension to Qp arenon-Archimedean absolute values; compare 36.3, 44.9 and 51.4. Theseare the main examples in the context of this book.

Further examples of non-Archimedean absolute values can be obtainedfrom valuations; see 56.2 and 56.3. See also 58.5, Section 23 and 44.11.

Non-Archimedean absolute values satisfy a stronger version of subad-ditivity:

55.3 Lemma Let F be a skew field. A non-trivial homomorphism

ϕ : F× → Rpos is a non-Archimedean absolute value if, and only if, ϕ is

ultrametric, that is,

ϕ(x + y) ≤ max{ϕ(x), ϕ(y)} for all x, y ∈ F .

In this case, ϕ(x) �= ϕ(y) implies that ϕ(x± y) = max{ϕ(x), ϕ(y)}.Proof The ultrametric inequality implies by induction that ϕ(n · 1) =ϕ(1 + 1 + · · · + 1) ≤ ϕ(1) = 1 for every n ∈ N, whence ϕ is non-Archimedean.

Conversely, suppose that ϕ is a non-Archimedean absolute value. Letϕ(x) ≤ 1. Then ϕ(x + 1) ≤ 2, hence m := sup{ϕ(x + 1) | ϕ(x) ≤ 1}is finite, and ϕ(x + y) ≤ m · max{ϕ(x), ϕ(y)} for all x, y ∈ F by themultiplicativity of ϕ. Furthermore ϕ(x) ≤ 1 implies

ϕ(x+1)3 = ϕ(x3 +1+3x(x+1)

) ≤ m ·max{ϕ(x3 +1), ϕ(x+1)} ≤ m2 ,

as ϕ(3) ≤ 1 and ϕ(x3) ≤ 1. Thus m3 = sup{ϕ(x+1)3 | ϕ(x) ≤ 1} ≤ m2,hence m ≤ 1, and ϕ is ultrametric.

As an alternative, we infer from ϕ(x) ≤ 1 that

ϕ(x + 1)n = ϕ(∑n

k=0

(nk

)xk) ≤∑n

k=0 ϕ(x)k ≤ n + 1

for each n ∈ N. Taking n-th roots and using limnn√

n + 1 = 1 we seethat m = sup{ϕ(x + 1) | ϕ(x) ≤ 1} ≤ 1, hence ϕ is ultrametric.

Finally, let ϕ be ultrametric and ϕ(x) > ϕ(y). In view of ϕ(y) =ϕ(−y) it suffices to compute ϕ(x + y). We have ϕ(x) = ϕ(x + y − y) ≤max{ϕ(x + y), ϕ(y)}, so this maximum is ϕ(x + y), and we infer thatϕ(x + y) = ϕ(x). �

By 55.3 the ultrametric character of an absolute value is revealedalready in the prime field.

Page 319: The Classical Fields

302 The p-adic numbers

If ϕ is an absolute value, then the function ϕs defined by ϕs(x) =ϕ(x)s is also an absolute value for 0 < s ≤ 1 (see Exercise 4), and if ϕ

is non-Archimedean, then one can even admit all s > 0.The following theorem determines all absolute values of the field Q of

rational numbers.

55.4 Theorem (Ostrowski) Let ϕ be an absolute value of Q. If ϕ

is Archimedean, then ϕ = | |s for some real number s with 0 < s ≤ 1.

If ϕ is non-Archimedean, then ϕ = | | sp for some prime number p and

some s > 0.

Proof First, let ϕ be non-Archimedean. Then S := {n ∈ Z | ϕ(n) < 1}is an additive subgroup of Z (see 55.3), and S is not trivial, becauseϕ(Q×) �= {1} by definition. Thus, S = pZ for some p ∈ N, and weclaim that p is a prime number. Otherwise p = ab with natural numbersa, b /∈ pZ = S. But then ϕ(a) = 1 = ϕ(b), a contradiction to ϕ(a)ϕ(b) =ϕ(p) < 1.

We can now write ϕ(p) = p−s with a real number s > 0, and we showthat ϕ = | | sp . Every prime number q �= p satisfies q /∈ pZ = S, henceϕ(q) = 1 = |q| sp . This shows that ϕ and | | sp coincide on all primes. Theprimes together with −1 generate the multiplicative group Q×, as wehave seen in 32.1. Therefore the two multiplicative functions ϕ and | | spare equal.

Now let ϕ be Archimedean. We consider integers a, b ∈ N with b ≥ 2,and we define C := max{ϕ(c) | c = 0, 1, . . . , b − 1}. For every naturalnumber n there is a unique m ∈ N0 with bm ≤ an < bm+1, hencem ≤ n · log a/ log b. From the representation

an =∑m

k=0 ckbk

of an with basis b, with digits ck ∈ {0, 1, . . . , b− 1}, we infer that

ϕ(a)n ≤m∑

k=0

ϕ(ck)ϕ(b)k ≤ (m + 1)C max{1, ϕ(b)}m

≤ (n · log a/ log b + 1)C max{1, ϕ(b)}n log a/ log b .

Taking the n-th root and observing that limnn√

nα + β = 1 for α ≥ 0,we obtain

ϕ(a) ≤ max{1, ϕ(b)}log a/ log b

for all a, b ∈ N with b ≥ 2. Since ϕ is Archimedean, we have ϕ(a) > 1for some a, hence ϕ(b) > 1 for each b ≥ 2, and our inequality gives

Page 320: The Classical Fields

55 Absolute values 303

ϕ(a)1/ log a ≤ ϕ(b)1/ log b for all a, b ∈ N with a, b ≥ 2. By symmetry, weconclude that ϕ(a)1/ log a is a constant es > 1, for some s > 0. Thus,

ϕ(a) = es log a = as = |a|s

for every a ∈ N, which implies that ϕ = | |s. �

Theorem 55.6 below describes all Archimedean absolute values of R.The non-Archimedean absolute values of R defy classification; indeed,for every transcendency basis T of R over Q, there are many possibilitiesto extend a p-adic absolute value of Q to Q(T ) (compare Exercise 5 ofSection 56), and then further extensions to R are possible (see Jacob-

son 1989 9.9, Lang 1993 XII §3 or Warner 1989 Theorem 26.6). Wemay also infer from 56.15 that the p-adic absolute value of Qp extendsto the algebraic closure Q �

p ; now Q �p∼= C by 64.21, and each of the

many embeddings of R into C (see 14.15, 14.9) gives a non-Archimedeanabsolute value on R.

55.5 Remark: Metric and topology Every absolute value ϕ of a fieldF yields a metric d which is defined by d(x, y) = ϕ(x− y) for x, y ∈ F .The corresponding topology is a field topology of F ; see 13.2(b).

The following theorem describes all fields with an Archimedean abso-lute value: they are just the subfields of the complex field C.

55.6 Theorem (Ostrowski) Let F be a field with an Archimedean

absolute value ϕ. Then F is isomorphic to a dense subfield of R or C,

and there exists a monomorphism ι of F into C and a real number s

with 0 < s ≤ 1 such that

ϕ(x) = |ι(x)|s for every x ∈ F ,

where | | is the ordinary absolute value of C.

Proof Since ϕ is Archimedean, F has Q as its prime field, and therestriction ϕ|Q× is not identically 1, hence an Archimedean absolutevalue of Q. Theorem 55.4 implies that ϕ|Q = | |s with 0 < s ≤ 1.Therefore the completion of Q with respect to ϕ|Q is isomorphic to thefield of real numbers; see 44.11 (1). Thus the topological closure of Q inthe completion F of F with respect to ϕ is a copy R of the real numbers.By 44.9, the topology of F is described by an absolute value ϕ : F → Rthat is a continuous extension of ϕ, and ϕ coincides on R with | |s, asQ is dense in R.

We show below that every element a ∈ F satisfies a quadratic equationover R. Since C is the only proper algebraic extension of R, this implies

Page 321: The Classical Fields

304 The p-adic numbers

F = R or F = C, and then the embedding of F into its completion F

gives the assertion of 55.6 (note that ϕ(ζ) = 1 for every root of unity ζ,hence ϕ is uniquely determined by ϕ|Q).

Let a ∈ F . We essentially follow Neukirch 1992 II 4.2 and definea continuous map f : C → R by f(z) = ϕ(a2 − (z + z)a + zz) forz ∈ C; note that z + z, zz ∈ R ⊆ F . The triangle inequality yieldsf(z) ≥ ϕ(zz)− ϕ((z + z)a)− ϕ(a2) = |zz|s− |z + z|sϕ(a)− ϕ(a2), hencelimz→∞ f(z) = ∞. Therefore f attains its lower bound m, and thepreimage f−1(m) ⊆ C is non-empty and compact. Choose z0 ∈ f−1(m)with |z0| = max |f−1(m)|. The real polynomial

q(x) := x2 − (z0 + z0) x + z0z0

satisfies q(x) ≥ 0 for every x ∈ R, since (z0 − z0)2 ≤ 0. If m = 0, thenq(a) = 0.

Now we assume that m > 0 and aim for a contradiction. Let μ =(m/2)1/s. The real polynomial q(x) + μ is strictly positive on R, henceit has roots w,w ∈ C. These roots satisfy |w|2 = ww = z0z0 +μ > |z0|2,hence f(w) > m by our choice of z0.

For any odd number n ∈ N, the real polynomial g(x) = q(x)n + μn ofdegree 2n is strictly positive on R, and g(w) = 0. Thus we have the realfactorization g(x) =

∏nj=1(x

2 − (wj + wj)x + wjwj), where w1 = w, w1,w2, w2, . . . , wn, wn are the complex roots of g(x). Substituting a for x,we obtain g(a) =

∏nj=1(a

2 − (wj + wj)a + wjwj) ∈ F and therefore

ϕ(g(a)) =∏n

j=1 f(wj) ≥ f(w) ·mn−1 .

On the other hand,

ϕ(g(a)) ≤ ϕ(q(a))n + ϕ(μ)n = f(z0)n + μsn = mn + (m/2)n .

We infer that f(w) ≤ m(1+2−n). This holds for all odd numbers n ∈ N,hence f(w) ≤ m, a contradiction to f(w) > m. �

For other proofs of 55.6 see Jacobson 1989 9.5, Ebbinghaus et al.1991 Chapter 8 §4, Bourbaki 1972 VI.6.6 p. 410 and Cassels 1986Chapter 3. The related Theorem of Gelfand and Mazur says that everyextension field of R which is complete with respect to a norm is isomor-phic to R or C (for the definition of a norm ϕ on a field, replace themultiplicativity in 55.1 by submultiplicativity: ϕ(xy) ≤ ϕ(x)ϕ(y)). Forproofs see Ebbinghaus et al. 1991 Chapter 8 §4, Ribenboim 1999 Theo-rem 3 p. 37, Lang 1993 XII.2, Engler–Prestel 2005 1.2.4, Bourbaki

1972 VI.6.4 p. 407 or Warner 1989 26.10 p. 262 and the remarks on

Page 322: The Classical Fields

55 Absolute values 305

pp. 499–501. The last two references allow also skew fields, and thenHamilton’s quaternions H appear as the only further possibility.

The following characterization of absolute values will be used in 57.4and 58.5.

55.7 Lemma (Artin’s trick) Let F be a skew field and ϕ : F → Rbe a multiplicative map such that ϕ(0) = 0, ϕ(1) = 1 and ϕ(x + y) ≤2 max{ϕ(x), ϕ(y)} for all x, y ∈ F . Then ϕ is an absolute value of F .

Proof We have ϕ(x1+· · ·+x2k) ≤ 2k max{ϕ(x1), . . . , ϕ(x2k)} for xi ∈ F ,as an easy induction on k shows; in particular ϕ(2k · 1F ) ≤ 2k. Thisimplies that ϕ(m · 1F ) ≤ 2m for all m ∈ N (one needs to add at most m

zeros in order to write m · 1F as a sum of 2k terms xi ∈ {0, 1F }). Forx ∈ F and n = 2k − 1 we compute by binomial expansion that

ϕ(1 + x)n ≤ 2k max0≤i≤n ϕ((ni

))ϕ(xi)

≤ 2k+1∑n

i=0

(ni

)ϕ(x)i = 2(n + 1)(1 + ϕ(x))n .

Taking the root of order n = 2k−1 we obtain for k →∞ that ϕ(1+x) ≤1 + ϕ(x) for all x ∈ F . If x, y ∈ F with 0 �= ϕ(x) ≥ ϕ(y), then

ϕ(x + y) ≤ (1 + ϕ(yx−1))ϕ(x) = ϕ(x) + ϕ(y) .

As ϕ is not constant, multiplicativity implies that ϕ(x) = 0 ⇔ x = 0.Hence ϕ is an absolute value of F . �

Exercises

(1) Two absolute values ϕ1 and ϕ2 of a field F define the same topology, if,and only if, ϕ1(a) < 1 ⇔ ϕ2(a) < 1 for a ∈ F , and this occurs precisely ifϕ1 = ϕs

2 for some positive real number s.

(2) Let ϕ1 and ϕ2 be absolute values of a field F that induce different topolo-gies on F . Show that the diagonal { (a, a) | a ∈ F } is dense in the productspace F1 × F2, where Fi denotes the set F endowed with the topology in-duced by ϕi. (This is a special case of the Approximation Theorem of Artin–Whaples.)

(3) Assume that ϕ is a non-Archimedean absolute value on a field F , and let

bϕ be the continuous extension of ϕ to the completion bF (compare 44.9). Show

that bϕ( bF ) = ϕ(F ).

(4) If ϕ is an absolute value on a field F and 0 < s ≤ 1, then also the map ϕs

defined by ϕs(a) = ϕ(a)s is an absolute value on F .

Page 323: The Classical Fields

306 The p-adic numbers

56 Valuations

We define general valuations (which are sometimes called Krull valu-ations), and we obtain some basic results of valuation theory. For morecomprehensive accounts of valuation theory see Cohn 2003a Chapter 9,Jacobson 1989 Chapter 9, Ribenboim 1999, Warner 1989 Chapter V,Bourbaki 1972 VI and Engler–Prestel 2005.

56.1 Definition Let Γ be a non-trivial ordered abelian group. A valu-ation v of a field (or skew field) F with value group Γ is an epimorphismv : F× → Γ of groups which satisfies

v(x + y) ≥ min{v(x), v(y)}for all x, y ∈ F× with x + y ∈ F×. If Γ ∼= Z, one speaks of a principalvaluation (or of a discrete rank 1 valuation). In that case, one identifiesΓ with the ordered group (Z, <), and one calls each a ∈ F with v(a) = 1a prime element (or a uniformizer) of v.

By convention, Γ is written additively. Often one defines v(0) = ∞,where∞ is a new element with γ <∞ and∞+γ = γ+∞ =∞+∞ =∞for each γ ∈ Γ. Then the map v : F → Γ ∪ {∞} satisfies v(xy) =v(x) + v(y) and v(x + y) ≥ min{v(x), v(y)} for all x, y ∈ F .

Since Γ is torsion free (7.3), we have v(x) = 0 whenever x ∈ F isa root of unity. In particular, v(−1) = 0, hence v(−x) = v(x) forall x ∈ F . Furthermore it is easy to show that v(x) �= v(y) impliesv(x + y) = min{v(x), v(y)}; compare the proof of 55.3.

56.2 Absolute values and valuations Let ϕ be a non-Archimedeanabsolute value of a field F , as defined in 55.1. Then v(x) = − loga ϕ(x)gives a valuation v of F , for every basis a > 1, and the value groupv(F×) ⊆ R is an Archimedean ordered group.

Conversely, let v be a valuation of a field F such that the value group Γ

is Archimedean. By 7.8 we may consider Γ as a subgroup of the orderedgroup R+, and then ϕ(x) = a−v(x) defines a non-Archimedean absolutevalue ϕ of F , for every real number a > 1.

Thus a valuation with Archimedean value group is essentially the samething as a non-Archimedean absolute value (an unfortunate clash of ter-minology, which is sometimes avoided by saying ‘real valuation’ insteadof ‘valuation with Archimedean value group’). In this sense, valuationsgeneralize non-Archimedean absolute values.

The negative sign (and the corresponding switch from max to min;compare 55.3 and 56.1) is motivated by the idea that v(x) should de-scribe a degree of divisibility of x, as in the following examples.

Page 324: The Classical Fields

56 Valuations 307

56.3 Examples (a) We fix a prime number p. The p-adic valuation vp

of Q is defined by

vp

(pn · a

b

)= n

for n ∈ Z and integers a, b ∈ Z that are not divisible by p (see 36.3). Anextension to Qp may be defined by the same formula, for a, b ∈ Zp �pZp.These principal valuations correspond to the p-adic absolute values of Qand Qp, via vp = − logp | |p; compare also 51.1, 51.4, 56.2 and 56.7.

(b) Let R be a unique factorization domain, F its field of fractions,and fix an irreducible element p ∈ R. Every element x ∈ F× has theform

x = pn · ab

with n ∈ Z and a, b ∈ R � pR. The integer n is uniquely determinedby x, and it is easy to verify that v : F× → Z : x �→ n is a principalvaluation of F with prime element p. Example (a) is the special casewhere R = Z.

(c) We describe some valuations which are relevant in algebraic geo-metry. Let F = k(x) be a simple transcendental extension of k (compare64.19). Then F is the field of fractions of the polynomial ring k[x], whichis a unique factorization domain. Hence for each irreducible polynomialp ∈ k[x], example (b) yields a principal valuation vp of F defined by

vp

(pn · a

b

)= n

for n ∈ Z and polynomials a, b ∈ k[x] which are not divisible by p.The degree of polynomials leads to another valuation v∞ of F , the

degree valuation, which is is given by

v∞(a

b

)= −deg a + deg b

for non-zero polynomials a, b ∈ k[x]. The field F = k(x) is also the fieldof fractions of its subring k[x−1] ∼= k[x], and the irreducible elementx−1 ∈ k[x−1] gives a valuation vx−1 on F with prime element x−1. Theequation

a = a′ · (x−1)− deg a

with a′ = ax− deg a ∈ k[x−1] shows that v∞ coincides with vx−1 .In fact, we have just described all valuations of k(x) which are trivial

on k (see Exercise 2). If k is algebraically closed, then all the irreduciblepolynomials p ∈ k[x] have degree 1, hence in this case the valuations of

Page 325: The Classical Fields

308 The p-adic numbers

k(x) which are trivial on k correspond to the elements of the ‘projectiveline’ k ∪ {∞} over k.

In the special case p = x − c ∈ k[x], the integer vp(f) describes thebehaviour near c of the rational function f ∈ F = k(x) in the followingsense: f vanishes at c of order n precisely if vp(f) = n > 0, and f has apole at c of order −n if, and only if, vp(f) = n < 0.

(d) The completion of k(x) with respect to the topology defined bythe valuation vx as in (c) is the field k((x)) of Laurent series over k,and the extended valuation maps

∑n anxn to min{n ∈ Z | an �= 0};

moreover, the valuation v∞ leads to the same completion. See 44.11 forthese facts.

(e) The non-standard rationals ∗Q and reals ∗R admit valuations thatare obtained essentially by factoring out infinitesimals; see Sections 23,24 (and use 56.5). Compare also 56.16.

Now we describe the valuations on a field F in terms of certain sub-rings of F .

56.4 Definitions A proper subring R of a field F is called a valuationring of F , if F = R ∪ {r−1 | 0 �= r ∈ R}.

Let v be a valuation on a field F with value group Γ. The ring R :={x ∈ F | v(x) ≥ 0} meets the requirements of 56.4, because v : F× → Γ

is not trivial and v(x) < 0 implies v(x−1) = −v(x) > 0. We call R thevaluation ring of v.

Then U := {x ∈ F | v(x) = 0} is the group of units of this ring R,and the set M := R � U = {x ∈ F | v(x) > 0} of non-units is an ideal,the unique maximal ideal of R. Thus the quotient Fv := R/M is a field,the residue field of the valuation v. Furthermore v : F× → Γ induces agroup isomorphism F×/U → Γ.

The p-adic valuation vp of the field Q has the valuation ring Rp :={ab−1 | a ∈ Z ∧ b ∈ Z � pZ} with the unique maximal ideal pRp andthe residue field Fp; see 51.7.

56.5 Proposition Let R be a valuation ring of a field F . Then there

exists a valuation v of F such that R = {x ∈ F | v(x) ≥ 0}.Proof The set U = {r ∈ R | r �= 0 ∧ r−1 ∈ R} is the group of unitsof R, and the factor group Γ := F×/U is not trivial, as R �= F . Thedefinition

xU ≤ yU ⇐⇒ x−1y ∈ R

for x, y ∈ F× renders Γ an ordered group; indeed, the relation ≤ is a

Page 326: The Classical Fields

56 Valuations 309

total ordering, because x−1y �∈ R implies y−1x ∈ R. A coset xU ∈ Γ is(strictly) positive precisely if x ∈ R � U .

We write Γ additively, and we denote its neutral element U by 0Γ, toavoid confusion with 0 ∈ F . The canonical map v : F× → Γ = F×/U isa group epimorphism, and R = {0}∪{x ∈ F× | v(x) ≥ 0Γ }. In order toshow that v is a valuation, let x, y ∈ F× with v(x) ≤ v(y) and x+y �= 0.Then x−1y ∈ R, hence 1 + x−1y ∈ R and v(1 + x−1y) ≥ 0Γ. We inferthat

v(x + y) = v(x(1 + x−1y)) = v(x) + v(1 + x−1y) ≥ v(x) .

This shows that v(x + y) ≥ min{v(x), v(y)} for all x, y ∈ F×. �

56.6 Proposition Two valuations v : F× → Γ and w : F× → Δ of

a field F have the same valuation ring if, and only if, there exists an

isomorphism α : Γ→ Δ of ordered groups such that w = α ◦ v.

In this situation v and w are called equivalent (see also 56.11).

Proof If w = α ◦ v, then v and w have the same valuation ring. Con-versely, if v and w have the same valuation ring R, then the group U ofunits of R is the common kernel of the epimorphisms v : F× → Γ andw : F× → Δ. Hence the quotient α of the induced group isomorphismsΓ← F×/U → Δ satisfies w = α ◦ v. Since v(R � U) and w(R � U) arethe sets of positive elements of Γ and Δ, respectively, we infer that α isan isomorphism of ordered groups. �

Extending Theorem 55.4 we now determine all valuations of the fieldQ of rational numbers.

56.7 Theorem Every valuation v of Q is equivalent to a p-adic valu-

ation vp = − logp | |p for a unique prime number p.

Proof By 56.1 we have v(Z) ≥ 0, and the restriction of v to Z � {0} isnot trivial, as v is not trivial on Q×. Hence P := {z ∈ Z | v(z) > 0}is a non-zero proper ideal of Z, in fact a prime ideal, because x, y ∈ Z,v(xy) > 0 implies v(x) > 0 or v(y) > 0. Thus P = pZ for a unique primenumber p. In particular, v(z) = 0 for all z ∈ Z�pZ. By multiplicativity,v and vp have the same valuation ring Rp = {ab−1 | a ∈ Z ∧ b ∈ Z�pZ},hence they are equivalent by 56.6.

These valuation rings Rp are mutually distinct, because the quotientof Rp modulo its unique maximal ideal pRp has cardinality p. �

Page 327: The Classical Fields

310 The p-adic numbers

56.8 Theorem Let F be a field and v : F× → Γ a valuation with

valuation ring R. Then the following conditions are equivalent:

(i) R is a maximal subring of F .

(ii) R contains only one proper non-zero prime ideal.

(iii) The value group Γ is Archimedean.

Proof (i) implies (ii): Let I be a proper non-zero prime ideal of R. Weshow that I coincides with the (unique) maximal ideal M of R; clearlyI ⊆M . Let 0 �= a ∈M . Then a−1 �∈ R, hence the ring R[a−1] generatedby R and a−1 is F , by (i). If 0 �= b ∈ I, then b−1 can be written as afinite sum b−1 =

∑0≤k≤n rka−k with elements rk ∈ R. We infer that

an = bb−1an = b∑

k rkan−k ∈ I, hence a ∈ I, as I is a prime ideal. Thisshows that M ⊆ I.

(ii) implies (iii): Let 0 < γ ∈ Γ and I := {a ∈ F | ∀n∈N v(a) > nγ }.The defining properties of v imply that I is an ideal of R. In fact, I is aprime ideal of R, because r1, r2 ∈ R � I entails v(ri) ≤ niγ for i = 1, 2with suitable integers ni ∈ N, hence v(r1r2) ≤ (n1 + n2)γ and r1r2 /∈ I.The preimages of γ under v belong to the maximal ideal of R, but notto I. Hence (ii) implies that I = {0}, which means that each a ∈ F×

satisfies v(a) ≤ nγ for some n ∈ N. Thus Γ is Archimedean.(iii) implies (i): Each c ∈ F � R satisfies v(c) < 0. For each a ∈ F

there exists by (iii) an integer n ∈ N such that nv(c) ≤ v(a). Hencev(ac−n) ≥ 0, so ac−n ∈ R. Therefore a = ac−ncn belongs to the subringR[c] generated by R and c. Thus R[c] = F , and (i) holds. �

56.9 Extension Theorem Let v be a valuation of a field F and let E

be an extension field of F . Then v has an extension to a valuation of E

(with a possibly larger value group).

Proof Let Rv = {x ∈ F | v(x) ≥ 0} be the valuation ring of v andMv = {x ∈ F | v(x) > 0} its maximal ideal. First we construct avaluation ring of E.

We consider all pairs (R,M) consisting of a subring R of E withR ⊇ Rv and a proper ideal M of R with M ⊇ Mv, and we order thesepairs by double inclusion: (R,M) ≤ (R′,M ′) ⇔ R ⊆ R′ ∧ M ⊆ M ′.Given a chain of such pairs, by taking unions we obtain again a pair ofthis type (note that always 1 /∈ M). Hence, by Zorn’s Lemma, thereexists a maximal pair (R,M).

We claim that R is a valuation ring of E. We have R �= E, sinceM ⊇ Mv �= {0} is a proper ideal of R. Furthermore, we assume thate /∈ R and e−1 /∈ R for some e ∈ E and aim for a contradiction. Then

Page 328: The Classical Fields

56 Valuations 311

the ring R[e] generated by R and e is strictly larger than R. If theideal M ′ := {∑i aie

i | ai ∈ M } generated by M in R[e] were proper,then (R[e],M ′) > (R,M), which contradicts the maximality. HenceM ′ = R[e], which implies that we have an equation

1 = a0 + a1e + · · ·+ amem (∗)with ai ∈ M . Similarly, since R[e−1] is strictly larger than R, we haveanother equation

1 = b0 + b1e−1 + · · ·+ bne−n

with bi ∈M . We may assume that m ≥ n by symmetry, and that m+n

is chosen to be as small as possible. If we now multiply equation (∗) by1− b0 and substitute for (1− b0)em from the second equation, we obtainan equation for e of the form (∗), but with m replaced by m − 1. Thisis a contradiction, which proves that R is a valuation ring of E.

By construction we have Rv ⊆ R ∩ F and Mv ⊆ M . FurthermoreF � Rv = (Mv � {0})−1 ⊆ (M � {0})−1 = E � R is disjoint from R,hence Rv = R ∩ F . By Proposition 56.5 there exists a valuation w ofE with valuation ring R. We have just shown that the valuation ringR ∩ F of the restriction w|F coincides with the valuation ring Rv of v.Using 56.6 we obtain an isomorphism α : w(F×) → v(F×) with v =α ◦w|F . By set-theoretic considerations, there exists an ordered abeliangroup B containing v(F×) as an ordered subgroup and an isomorphismβ : w(E×) → B extending α. Then β ◦ w is the desired valuation of E

extending v. �

More information on extensions of valuations can be found in thereferences mentioned at the beginning of this section; see also 56.15 andExercise 5.

The topology τv induced on a field F by a valuation v : F× → Γ

is defined by taking the sets {x ∈ F | v(x − a) > γ } with γ ∈ Γ asa neighbourhood base at a ∈ F . By 13.2(b), τv is always a totallydisconnected field topology of F .

We say that F is complete with respect to a valuation v, if the topo-logical field (F, τv) is complete (in the sense of 44.1).

By 56.9 each p-adic valuation of Q extends to R; this gives (infinitelymany) totally disconnected field topologies on R. However, none of thethese valuation topologies on R is locally compact or complete (other-wise R would contain a copy of the completion Qp of Q, which is acontradiction to 54.2).

Page 329: The Classical Fields

312 The p-adic numbers

56.10 Theorem Let v and v′ be valuations of a field F , with respective

valuation rings R and R′. Then the following conditions are equivalent.

(i) The valuations v and v′ induce the same topology τv = τv′ on F .

(ii) The subring generated by R ∪R′ is a proper subring of F .

Valuations v, v′ which satisfy one (hence both) of the properties in56.10 are often called dependent.

Proof (i) implies (ii): If τv = τv′ , then R′a ⊆ R for some a ∈ F×,hence RR′ ⊆ RRa−1 ⊆ Ra−1. The subring S generated by R∪R′ is theadditive subgroup generated by RR′, and Ra−1 is an additive group. Weconclude that S ⊆ Ra−1. Moreover, Ra−1 = {x ∈ F | v(x) ≥ −v(a)} isa proper subset of F , hence S is a proper subring.

(ii) implies (i): The subring S generated by R∪R′ is a valuation ringof F ; denote by w the valuation defined by S; see 56.5. It suffices toshow that τv = τw, because symmetry implies then that also τv′ = τw.

Since S �= F , we can pick a ∈ F�S. Then a �∈ R, hence a−1 ∈ R. Eachnon-zero s ∈ S satisfies s−1a �∈ R (otherwise a = s(s−1a) ∈ SR = S,which is a contradiction), hence a−1s ∈ R and s = a(a−1s) ∈ aR. ThusR ⊆ S ⊆ aR and bR ⊆ bS ⊆ baR for each b ∈ F×. This says that τv

and τw have the same neighbourhoods of 0, whence τv = τw. �

56.11 Corollary Let v and v′ be valuations of a field F , with respective

valuation rings R and R′. Assume that v is a principal valuation, or

more generally that the value group of v is Archimedean. Then v and

v′ induce the same topology τv = τv′ on F if, and only if, v and v′ are

equivalent (that is, R = R′; see 56.6).

Proof By 56.8 the valuation ring R is a maximal subring of F , hencethe assertion follows from 56.10. �

The following result characterizes the valuation topologies which arelocally compact.

56.12 Theorem Let v : F× → Γ be a valuation of a field F . Then

the valuation topology τv is locally compact if, and only if, Γ ∼= Z, the

residue field Fv is finite, and F is complete with respect to v.

Proof R = {a ∈ F | v(a) ≥ 0} is the valuation ring of v, and the setM = {a ∈ F | v(a) > 0} is the maximal ideal of R.

Assume that τv is locally compact. Then (F+, τv) is complete by 43.9.The closed sets aR with a ∈ F× form a neighbourhood base at 0. Bylocal compactness, one of these sets aR is compact, hence R = a−1aR

Page 330: The Classical Fields

56 Valuations 313

is compact. Since M is an open additive subgroup of R, the quotientFv = R/M is finite. Each set Iα := {a ∈ F | v(a) > α} with α > 0 is anopen additive subgroup of R, hence R/Iα is finite. Since Iβ/Iα ⊆ R/Iα

for 0 ≤ β ≤ α and since v : F× → Γ is surjective, we deduce thateach interval {β ∈ Γ | 0 ≤ β ≤ α} with α > 0 is finite. ThereforeΓ is Archimedean (as Γ is torsion free) and contains a smallest positiveelement, which is a generator of Γ (compare the proof of 1.4). This showsthat Γ is cyclic.

For the converse implication, we assume that F is complete with re-spect to a principal valuation v : F× → Z with finite residue field Fv.It suffices to show that the valuation ring R is compact, because thesets aR with a ∈ F× form a neighbourhood base at 0. The set R isclosed in F and therefore complete. Let π be a prime element of v,i.e., v(F×) = 〈v(π)〉. Then M = πR, and multiplication by π inducesisomorphisms of πn−1R/πnR onto πnR/πn+1R for all n ∈ N. Henceeach quotient ring R/πnR with n ∈ N is finite. Since

⋂n πnR = {0} we

obtain an injective homomorphism

ι : R→∏n∈N R/πnR : r �→ (rn)n , rn := r + πnR ,

of additive groups (even of rings). Endow each finite group R/πnR withthe discrete topology. Then the product group

∏n R/πnR is compact

by Tychonoff’s theorem. Moreover, ι is a homeomorphism of R onto itsimage ι(R), because the sets πkR form a neighbourhood base of 0 inR and their images ι(πkR) = ι(R) ∩ { (xn)n | xn = 0 for n ≥ k } forma neighbourhood base of 0 in ι(R). The image ι(R) is complete, henceclosed in the compact product space. Thus ι(R) is compact, and R aswell. �

The next result says that each field appearing in 56.12 admits onlyone principal valuation (up to equivalence); more general results can befound in Ribenboim 1999 3.W p. 103 and Warner 1989 32.24.

56.13 Theorem Let F be a field which is complete with respect to a

principal valuation v : F× → Z with finite residue field. Then v is, up

to equivalence, the only principal valuation of F .

Proof Let R be the valuation ring of v, let M be the maximal ideal ofR, let Fv = R/M denote the finite residue field of v, and let p be thecharacteristic of Fv. Then F× ∼= Z×R× ∼= Z×F×

v × (1 + M); comparethe proof of 53.1 and see also Exercise 3. Moreover, M is open andclosed in F , hence complete.

Page 331: The Classical Fields

314 The p-adic numbers

Let m ∈M . The group homomorphism z �→ (1 + m)z of Z+ into themultiplicative group 1 + M is continuous when Z is provided with thep-adic topology, because (1 + m)pn ∈ 1 + Mn+1 for each n ∈ N (thisis easily verified by induction on n, using the fact that p · 1F ∈ M ; seealso 37.5). As 1 + M is complete, the mapping z �→ (1 + m)z extendsby 43.23 to a continuous group homomorphism of Z+

p into 1 + M . Thisimplies that the multiplicative group 1 + M admits roots of order r foreach prime r �= p; see 52.6. (In fact, 1+M becomes a Zp-module in thisfashion; compare Weil 1967 p. 32.)

Fix any prime number r > |Fv|. Then R× = F×v × (1 + M) is the

largest r-divisible subgroup of F× ∼= Z × R×. (Compare the proof of53.5 for a related argument.)

If R′ is the valuation ring of another principal valuation v′ of F , thenF× ∼= Z × R′× (by Exercise 3). Thus the largest r-divisible subgroupR× of F× is contained in R′× ⊆ R′. Now R is additively generated byR× = R � M (as M is a proper additive subgroup of R), hence R ⊆ R′.Since R is a maximal subring of F by 56.8, this implies that R = R′,whence v and v′ are equivalent (see 56.6). �

56.14 Corollary Up to equivalence, the fields Qp and Fq((x)) admit

only one principal valuation.

Proof This is a consequence of Theorem 56.13; note that Fq((x)) iscomplete by 44.11. �

Let p and q be distinct primes. By 56.9 the q-adic valuation of Qextends to a valuation w of Qp. By 56.14, such an extension w cannotbe a principal valuation.

56.15 Theorem Let F be a field which is complete with respect to a

principal valuation v : F× → Z with finite residue field, and let E be

a field extension of F of finite degree n. Then v has an extension to

a valuation w of E, and the extension w is unique up to equivalence.

Moreover, w is again a principal valuation, its residue field is finite, and

E is complete with respect to w.

If E|F is a Galois extension of finite degree n with Galois group G,

then w(a) = v(N(a))/n for a ∈ E, where N : E → F : a �→ ∏γ∈G γ(a)is the norm map of E|F .

Proof Extensions of v to E exist by 56.9. For every extension w of v,the valuation topology τw renders E a topological field, hence also atopological vector space over the topological field F (with the valuation

Page 332: The Classical Fields

56 Valuations 315

topology τv). Since F is locally compact by 56.12, we infer from 58.6(i)that (E, τw) is locally compact. Now 56.12 implies that w is a principalvaluation with finite residue field and that E is complete with respectto w. Result 56.13 shows that w is unique up to equivalence.

Thus if E|F is a Galois extension and γ ∈ G, then w and w ◦ γ

are equivalent, hence w = w ◦ γ. Therefore nw(a) =∑

γ∈G w(γ(a)) =w(∏

γ∈G γ(a)) = w(N(a)).We mention that more direct proofs of more general versions of 56.15

can be found in many books; see Cohn 2003a 9.5.3 and 9.2.7, Jacobson

1989 Section 9.8 Corollary p. 583, Warner 1989 26.7 p. 260, Ebbing-

haus et al. 1991 Chapter 6 Theorem 11, Cassels 1986 Chapter 7 Theo-rem 1.1, Serre 1979 Chapter II §2, Ribenboim 1999 5.A p. 127 or thepaper Lenstra–Stevenhagen 1989. �

56.16 Natural ordering valuations Let (F,<) be an ordered fieldwhich is not Archimedean; then the prime field Q is not cofinal in F ;compare 11.12. Hence R = {a ∈ F | −q ≤ a ≤ q for some q ∈ Q}is a proper subring of F . In fact, R is a valuation ring of F with themaximal ideal M = {a ∈ F | −q ≤ a ≤ q for each positive q ∈ Q}. Thecorresponding valuation v of F is called the natural valuation of (F,<).The value group Γ = v(F×) = F×/R× is the group of Archimedeanclasses of the additive ordered group F+, and the residue field Fv = R/M

is an Archimedean ordered field. See Section 23 for examples.An ordered field (F,<) is called exponentially closed, if there exists

an isomorphism e : F+ → {a ∈ F | a > 0} of ordered groups (as in11.10) with the additional property that 1 + 1/n < e(1) < n for somen ∈ N. Alling 1962 Theorem 3.1 gives a characterization of all non-Archimedean ordered fields which are exponentially closed, using thenatural valuation and Hahn power series (as defined in 64.25).

Exercises(1) A field F has a valuation (or an absolute value) if, and only if, F is notalgebraic over any finite field.

(2) Let k be any field, and let v be a valuation of the simple transcendentalextension field k(x). If v is trivial on k, then v is equivalent to one of thevaluations vp or v∞ defined in 56.3.

(3) Let F be a field with a valuation v and valuation ring R. Show that vis a principal valuation if, and only if, R is a principal ideal domain; in thiscase F× = πZ × R× for each prime element π of v. Assume further that Fis complete with respect to v and that the residue field Fv = R/M is finite,where M = πR is the maximal ideal of R; then R× ∼= F×

v × (1 + M).

Page 333: The Classical Fields

316 The p-adic numbers

(4) Let E|F be an algebraic field extension and v a valuation of E. Show thatthe restriction v|F is a valuation (i.e. not trivial).

(5) Let F be a field with a valuation v : F× → Γ, let δ be an arbitrary elementof (an ordered group extending) Γ and define w on the polynomial ring F [x]by w(

P

j ajxj) = minj(v(aj) + jδ), where aj ∈ F . Show that w extends to a

valuation of the field of fractions F (x) of F [x].

(6) A maximal subring R of a field F is a valuation ring of F if, and only if,F is the field of fractions of R, i.e. F = {rs−1 | r, s ∈ R, s = 0}.(7) Let k be a field and Γ a non-trivial ordered abelian group. Construct afield F with a valuation v : F× → Γ such that Fv

∼= k.

(8) The field Qp is neither a purely transcendental extension nor a finite alge-braic extension of any proper subfield.

(9) The skew field H of Hamilton’s quaternions admits no valuation. Thep-adic valuation vp of Q extends to the rational quaternions only for p = 2.

57 Topologies of valuation type

In this section we characterize those ring topologies on a field F whichare induced by an absolute value or by a valuation of F ; Theorem 57.7says that these are precisely the topologies of type V (defined in 57.5).

57.1 Definition Let R be a topological ring. A subset B ⊆ R is calledbounded , if for each neighbourhood U of 0 there exists a neighbourhoodV of 0 such that V B ⊆ U and BV ⊆ U . This means that B can bemade small by multiplication with sufficiently small elements (see also57.2(iii)). Clearly every subset of a bounded set is bounded.

The topological ring R and also the topology of R is called locallybounded , if R contains a non-empty open bounded set; by 57.2(ii) thisis equivalent to assuming that some neighbourhood of 0 is bounded.

Each valuation and each absolute value of a field defines a locallybounded field topology; in fact, a set B ⊆ F is bounded with respect toan absolute value | | of F precisely if |B| is a bounded subset of R. SeeExercise 4 for examples of field topologies which are not locally bounded.

57.2 Lemma Let R be a topological ring.

(i) Each finite and each compact subset of R is bounded.

(ii) If B,B′ ⊆ R are bounded, then also B ∪ B′, B + B′ and BB′ are

bounded.

(iii) Let R be a field with a non-discrete ring topology. Then B ⊆ R is

bounded if, and only if, for each neighbourhood U of 0 there exists

an element a ∈ R � {0} with aB ⊆ U .

(iv) A field R with a ring topology such that R is bounded is either

indiscrete or discrete (compare 13.6).

Page 334: The Classical Fields

57 Topologies of valuation type 317

Proof (i) Let U be a neighbourhood of 0 in R. Since 0 · x = x · 0 = 0,there exist neighbourhoods Vx and Wx of 0 with Vx(x + Wx) ⊆ U and(x + Wx)Vx ⊆ U . Assume that C ⊆ R is compact or finite. ThenC ⊆ ⋃{x + Wx | x ∈ C0 } for some finite set C0 ⊆ C. Hence V :=⋂{Vx | x ∈ C0 } is a neighbourhood of 0 with V C ⊆ U and CV ⊆ U .

(ii) is a consequence of the continuity of addition and multiplication.(iii) If B is bounded, then the condition in (iii) is satisfied, since each

neighbourhood of 0 contains a non-zero element a. For the converse,let U be a neighbourhood of 0. By continuity of the multiplication wefind a neighbourhood W of 0 with WW ⊆ U , and by assumption wefind a �= 0 with aB ⊆ W . The neighbourhood V = Wa of 0 satisfiesV B = (V a−1)(aB) ⊆WW ⊆ U .

(iv) If such a field R is not indiscrete, then R contains an open set U

with 0 ∈ U �( 1; see 13.4. Since R = aR �⊆ U for each a ∈ R× we inferfrom (iii) that R is discrete. �

57.3 Definition An element a of a topological ring is called topologi-cally nilpotent , if the sequence of powers an converges to 0.

Non-zero topologically nilpotent elements exist in each field F witha topology defined by an absolute value | | : F → R or by a valuationv : F× → Γ with values in an Archimedean ordered group Γ.

Topologically nilpotent elements play a role in the characterizationtheorems 57.4 and 57.7.

57.4 Theorem (Shafarevich, Kaplansky) Let F be a field with a

Hausdorff ring topology. Then the following are equivalent.

(i) The topology of F is induced by an absolute value | | : F → R.

(ii) The set N := {a ∈ F | limn→∞ an = 0} of all topologically nilpo-

tent elements of F is a neighbourhood of 0 with N �= {0}, and

(F � N)−1 is bounded.

Proof It is easy to see that (i) implies (ii): N = {a ∈ F | |a| < 1} and(F � N)−1 = {a ∈ F | |a| ≤ 1} in the situation of (i), and both sets arebounded neighbourhoods of 0.

Now assume (ii). Then the subset

E := {a ∈ F× | a /∈ N ∧ a−1 /∈ N } = (F � N) ∩ (F � N)−1

of (F � N)−1 is bounded. We have N ⊆ {0}∪ (F � N)−1, as 0 �= a ∈ N

and a−1 ∈ N would lead to the contradiction 1 = ana−n → 0 · 0 = 0(here we need the Hausdorff property). Hence N is bounded. Moreover,

F = N ∪ E ∪ (N � {0})−1 .

Page 335: The Classical Fields

318 The p-adic numbers

The inclusion NN ⊆ N is a consequence of the commutativity ofthe multiplication. If am ∈ N for some m ∈ N, then a ∈ N , becauselimn→∞ amn = 0 entails limn→∞ amn+j = 0 for 0 ≤ j < m, hencelimn→∞ an = 0. This shows that yN ⊆ E for each y ∈ E; otherwisewe find n ∈ N with y±n ∈ N , hence y±1 ∈ N , and we have reached acontradiction.

We claim that NE ⊆ N . Since N and E are bounded, NE is boundedby 57.2(ii), hence aNE ⊆ N for some a ∈ F× by 57.2(iii). If x ∈ N

and y ∈ E, then xn ∈ aN for sufficiently large exponents n, hence(xy)n = xnyn ∈ aNyN ⊆ aNE ⊆ N and xy ∈ N .

Now we can show that E is a subgroup of F×: if x, y ∈ E and x−1y /∈E, then (x−1y)ε ∈ N for some ε ∈ {±1}, hence y = x(x−1y) ∈ xN ⊆ N

or x = y(x−1y)−1 ∈ yN ⊆ N , which is absurd.The set R := E ∪N is a bounded neighbourhood of 0. From 57.2 we

infer that R + R is bounded, and that (R + R)b ⊆ R for some b ∈ F×.The factor group Γ := F×/E is not trivial, because N �= {0}. The

definition

xE < yE � x−1y ∈ N

for x, y ∈ F× renders Γ an ordered group; the relation < is well-definedsince NE = N , and < is a total ordering, because x−1y �∈ N impliesy−1x ∈ E ∪ N . In fact, (Γ, <) is an Archimedean ordered group: ifx, y ∈ F× and yE is strictly positive, then y ∈ N , hence the elementsyn and x−1yn converge to 0, which entails x−1yn ∈ N and xE < ynE

for all sufficiently large integers n.According to Theorem 7.8 we can embed the ordered group (Γ, <)

into the ordered group (R+, <). Using the automorphisms of (R+, <)(compare 7.11), we find a monomorphism α : Γ → R of ordered groupswith α(bE) ≤ log 2. Now we define the mapping ϕ : F → R by

ϕ(x) = exp(−α(xE))

for x ∈ F× and ϕ(0) = 0. Clearly ϕ is multiplicative, and ϕ(1) = 1. Weclaim that

ϕ(x + y) ≤ 2 max{ϕ(x), ϕ(y)}for x, y ∈ F . It suffices to consider the case where ϕ(y) ≤ ϕ(x). Thenα(yE) ≥ α(xE) and x−1y ∈ N∪E = R, hence (x+y)b = x(1+x−1y)b ∈x(R + R)b ⊆ xR. This implies α((x + y)E) + log 2 ≥ α((x + y)bE) ≥α(xE) and ϕ(x + y) ≤ 2ϕ(x). By Artin’s trick 55.7, ϕ is an absolutevalue of F .

Page 336: The Classical Fields

57 Topologies of valuation type 319

By the assumptions on N , the sets aN = {x ∈ F | ϕ(x) < ϕ(a)} with0 �= a ∈ F form a neighbourhood base at 0 for the given topology of F .Therefore ϕ induces this topology. �

57.5 Definition Let F be a field with a ring topology. One says thatF and also its topology are of type V (or locally retrobounded), if theset (F � U)−1 is bounded for each neighbourhood U of 0.

It suffices to check this condition for the elements U of a neighbour-hood base at 0 (since subsets of bounded sets are bounded). See alsoExercise 5.

Here the symbol V stands for valuation; indeed, the topologies inducedby valuations or absolute values are of type V (Exercise 1). Theorem 57.7essentially says that also the converse statement is true.

57.6 Lemma Let F be a field. Then each ring topology of type V on

F is a locally bounded field topology.

Proof We may assume that F is a Hausdorff space; otherwise the topo-logy of F is indiscrete by 13.4, hence a locally bounded field topology.

First we show that inversion is continuous. Let U be a neighbourhoodof 0. By continuity of addition, there exists a neighbourhood W of0 with W + W ⊆ F � {−1}. As (F � U)−1 is bounded, there is aneighbourhood V ⊆ W with V (F � U)−1 ⊆ W and (F � U)−1V ⊆ W .We infer that −1 �∈ V + V (F � U)−1, hence 1 + v �∈ −v(F � U)−1 and(1+ v)−1 = 1+(1+ v)−1(−v) �∈ 1+(F �U) for each v ∈ V . This showsthat (1 + V )−1 ⊆ 1 + U . We conclude that inversion is continuous at 1,hence everywhere (see 8.3 or 13.2(a)).

Now we show that F is locally bounded. We can choose a neighbour-hood U of 0 with UU ⊆ F � {1} by the continuity of multiplicationat 0. If 0 �= x ∈ U , then x−1 �∈ U , hence x ∈ (F � U)−1. The set(F � U)−1 is bounded, since we have a topology of type V . Therefore,U ⊆ {0} ∪ (F � U)−1 is bounded; see 57.2(i, ii). �

There exist field topologies which are locally bounded, but not oftype V ; see Exercise 4.

57.7 Theorem (Kowalsky–Durbaum, Fleischer) Let F be a field

with a Hausdorff ring topology of type V .

(i) If 0 is the only topologically nilpotent element of F and if F is not

discrete, then the topology of F is induced by a valuation of F with

a value group which is not Archimedean.

(ii) If F contains a non-zero topologically nilpotent element, then the

topology of F is induced by an absolute value of F .

Page 337: The Classical Fields

320 The p-adic numbers

This result says that the non-discrete Hausdorff ring topologies oftype V on a field F are precisely the topologies induced by the ab-solute values of F , together with the valuation topologies of F (valu-ations with Archimedean value groups define the same topologies asnon-Archimedean absolute values; see 56.2).

Proof (i) We shall construct a valuation ring R of F such that R is abounded neighbourhood of 0. This proves (i), because then the sets aR

with a ∈ F× form a neighbourhood base at 0 (see 57.2(iii)), hence thegiven topology of F coincides with the topology induced by the valu-ation v associated with R; the value group v(F×) cannot be Archime-dean (otherwise each element with positive value would be topologicallynilpotent).

By 57.6 there exists a bounded neighbourhood V of 0, hence also aneighbourhood W of 0 with WV ⊆ V . The set D := {a ∈ F | aV ⊆ V }contains W and hence is a neighbourhood of 0. Moreover, D is bounded,since D ⊆ V v−1 whenever 0 �= v ∈ V , and V v−1 is bounded by 57.2(ii).As F has type V , also (F �D)−1 is bounded, hence U(F �D)−1 ⊆ D forsome neighbourhood U of 0. We pick any non-zero element c ∈ U ∩D,we put C := {a ∈ F | aD ⊆ c−N0D } and we define R to be the additivesubgroup of F generated by C. Note that c−1 ∈ C.

Then R is a subring of F , in view of CC ⊆ C. Moreover DD ⊆ D,hence D ⊆ C ⊆ R, whence R is a neighbourhood of 0. By the choice ofc we have (F � R)−1 ⊆ (F � D)−1 ⊆ c−1D ⊆ c−1C ⊆ CC ⊆ C ⊆ R.Thus R is a valuation ring of F , provided that R �= F . We complete theproof by showing that R is bounded, which implies R �= F according to57.2(iv).

We claim that the set c−N0 is bounded. Since V is bounded, the setsV a with a ∈ F× form a neighbourhood base at 0. If cn ∈ V a for somen ∈ N, then cn+1 ∈ c V a ⊆ V a (as c ∈ D), hence cm ∈ V a for all m ≥ n.By assumption, c is not topologically nilpotent, hence cN ∩ V a = ∅ forsome a ∈ F×. Thus c−N is contained in the set (F � V a)−1, which isbounded as F has type V ; hence c−N0 is bounded.

By 57.2(ii) also c−N0D is bounded, hence C is bounded as well in viewof C ⊆ c−N0Dd−1 for each non-zero d ∈ D. We infer again from 57.2(ii)that (C ∪ −C) + (C ∪ −C) is bounded, hence

(C ∪ −C) + (C ∪ −C) ⊆ b−1C

for some non-zero element b ∈ C by 57.2(iii). An easy induction showsthat the sum of 2n sets C ∪−C is contained in b−nC, hence R ⊆ b−N0C.

Page 338: The Classical Fields

57 Topologies of valuation type 321

Now C = C+{0} ⊆ b−1C, hence bC ⊆ C, and we infer as in the previousparagraph that b−N0 is bounded (using the bounded set C instead of V ).Thus R ⊆ b−N0C is bounded by 57.2(ii).

(ii) Let t ∈ F× be topologically nilpotent. By 57.6 we find a boundedneighbourhood V of 0. As in the proof of (i) above, we infer that D :={a ∈ F | aV ⊆ V } is a bounded neighbourhood of 0, hence the setstnD with n ∈ N form a neighbourhood base at 0. Each element of tD istopologically nilpotent, as DD ⊆ D. Thus the set N of all topologicallynilpotent elements of F is a neighbourhood of 0. Since F has type V ,the set (F � N)−1 is bounded, and assertion (ii) follows from 57.4. �

57.8 Corollary A non-discrete Hausdorff ring topology τ of a field

F is defined by a valuation of F if, and only if, τ has type V and the

additive group F+ contains a bounded open subgroup.

Proof Each valuation topology of F has type V (by Exercise 1), and{a ∈ F | v(a) > 0} is a bounded open subgroup of F+. The converseimplication is a consequence of Theorem 57.7, because Archimedeanabsolute values are excluded if there exists a bounded open subgroup U

of F+: there is an element u �= 0 in U , hence Z · 1F ⊆ Uu−1 is boundedby 57.2(ii), but |Z · 1F | is unbounded for every Archimedean absolutevalue | | of F . The topology of a non-Archimedean absolute value isalso induced by a valuation; see 56.2. �

The field Q of rational numbers has many field topologies (see Exer-cise 4 or 13.10), but only few of them are of type V .

57.9 Corollary The non-discrete Hausdorff ring topologies of type V

on Q are the usual topology and the p-adic topologies, where p is a prime

number.

Proof This is a consequence of 57.7, 56.7 and 55.4. �

In fact, all locally bounded field topologies of Q can be described:each topology of this type is the supremum of finitely many topologiesinduced by absolute values of Q (compare Exercise 4). See Warner

1989 and Wi‘es�law 1988 for proofs and for more information on locally

bounded fields and fields of type V ; compare also Shell 1990 Chapter 4.According to Mutylin 1968 Theorem 5 (see also Wi

‘es�law 1988 9.4),

R and C are the only locally bounded extension fields of R that induceon R the usual topology.

Page 339: The Classical Fields

322 The p-adic numbers

Exercises(1) Each valuation and each absolute value of a field F induces on F a fieldtopology of type V . The order topology of each ordered field is of type V .

(2) Let F be a field with a non-discrete ring topology, and let U be a boundedneighbourhood of 0. Show that the sets aU with a ∈ F× form a neighbourhoodbase at 0.

(3) Let τ be a non-discrete Hausdorff ring topology on a field F , let ∞ ∈ Fand let τ∞ be the topology on F∞ := F ∪ {∞} obtained from τ by adding allsets F∞�B where B is closed in F and bounded. Show that τ∞ is a Hausdorfftopology if, and only if, τ is locally bounded, and that τ is of type V if, andonly if, the inversion x �→ x−1, 0 �→ ∞, ∞ �→ 0 is continuous at ∞.

(4) For any non-empty set X of prime numbers, we define a metric dX on Qby dX(a, b) =

P

p∈X 2−p · |a − b|p (compare 36.6). Show that the topologyτX defined by dX is always a field topology of Q, that τX is locally boundedprecisely if X is finite, and that τX has type V precisely if card X = 1.

(5) A topological field F has type V if, and only if, for every neighbourhoodW of 0 there exists a neighbourhood U of 0 such that x, y ∈ F and xy ∈ Uimplies that x ∈ W or y ∈ W .

(6) The completion of a topological field of type V is again a topological field.

(7) Let v be a valuation of a field F , let a = 0 be a topologically nilpotentelement of F , and Ra := {x ∈ F | ∀n∈N nv(x) + v(a) > 0}. Show that Ra isa maximal subring of F , and deduce that the valuation topology of v is alsodefined by a valuation of F with an Archimedean value group.

58 Local fields and locally compact fields

Here we briefly discuss the concept of local fields, and we explain howone can classify all locally compact skew fields. In Theorem 58.7 weshall see that the local fields together with R and C are precisely thelocally compact fields which are neither discrete nor indiscrete.

In this section, a locally compact (skew) field is assumed to be endowedwith a non-discrete Hausdorff topology.

58.1 Local fields Usually a local field is defined as a field F which iscomplete with respect to a valuation v : F× → Z with a finite residuefield (see Section 56). These fields can be described more explicitly, asfollows: either F is isomorphic to the field Fq((t)) of Laurent series overa finite field Fq (see 64.23 and 44.11), or F is an extension of finite degreeof a p-adic field Qp (see 56.3(a), 56.15 and 58.2). Indeed, by 56.12 thelocal fields are precisely the locally compact fields with valuation, and58.7 gives the assertion (compare also Jacobson 1989 Theorem 9.16p. 577 and Neukirch 1992 II.5.2).

A local field F has only one principal valuation v : F× → Z; see 56.13.We call this valuation the natural valuation of F . As in 56.1, the prime

Page 340: The Classical Fields

58 Local fields and locally compact fields 323

elements π of F are defined by the condition v(π) = 1. Each primeelement of F generates the unique prime (in fact, maximal) ideal in thenatural valuation ring of F .

Many properties of Qp can be generalized to arbitrary local fields,as the books by Cassels 1986 and Serre 1979 show. However, themultiplicative group Fq((t))× is isomorphic to Z×Cq−1× (Z+

p )N, whereq is a power of the prime p (see Weil 1967 II.3 Proposition 10 p. 34),in contrast to the analogous result 53.3 on Q×

p .

58.2 Field extensions and automorphisms Let F be a local fieldwith its natural valuation v, and let E be a field extension of finite degreen of F . Then E is also a local field and v has a unique extension to E

which is the natural valuation w of E; see 56.15.The index e = |w(E×) : v(F×)| of the two value groups is called the

ramification index of E|F ; this terminology reflects the fact that eachprime element of F is a product of e prime elements of E. Furthermorethe residue field Fv is canonically embedded into Ew, and the degreef = [Ew : Fv] is called the residue degree of E|F . These numbers arerelated via the equation n = ef ; for proofs see, for example, Cohn

2003a 9.5.1, Jacobson 1989 p. 591, Serre 1979 p. 29 or Borevich–

Shafarevich 1966 Chapter 4 §1 Theorem 5 p. 262.A field extension E|F as considered above is called unramified , if

e = 1; this means that the value groups coincide and that the residuefields have the same degree n as E|F . If f = 1, then E|F is said to betotally ramified ; this means that the residue fields coincide and that thevalue group v(F×) has index n in w(E×).

Let F be a local field with residue field Fq (so q is a prime power), andlet n ∈ N. Then F has a unique unramified extension Fn of degree n,namely the splitting field of xqn − x over F , and Fn|F is a cyclic Galoisextension; see Reiner 1975 5.10 and 5.11, Serre 1979 III §5 Theorem 2p. 54 or Weil 1967 I §4 Corollaries 2 and 3 p. 18/19. If F = Fq((t)),then Fn = Fqn((t)).

Let F be a local field, and let π be a prime element of F . Adjoin-ing to F a root of the polynomial xn − π gives a totally ramified fieldextension of degree n over F , for any n ∈ N. In fact, the totally rami-fied extensions of F are precisely the extensions generated by roots ofEisenstein polynomials over F , i.e., polynomials xn +

∑n−1i=0 aix

i withv(a0) = 1 and v(ai) ≥ 1 for all i; see Serre 1979 I §6(ii) p. 19, Cas-

sels 1986 Theorem 7.1 p. 133, Ribenboim 1999 4.H p. 116, Lang 1970Chapter II Proposition 11 or Robert 2000 2.4.2.

Page 341: The Classical Fields

324 The p-adic numbers

Each local field F of characteristic p �= 0 is isomorphic to Fq((t)) forsome power q of p; indeed, each prime element π of F is transcendentalover the prime field Fp (otherwise π would be a root of unity), henceFp(π) ∼= Fp(x) with the valuation vx as in 56.3(c). Therefore F containsthe completion Fp((π)) ∼= Fp((x)) (compare 43.9 and 44.11(3)), and theextension F |Fp((π)) is finite-dimensional (of degree at most [Fv : Fp])and unramified, hence F ∼= Fp((x))n = Fpn((t)); see also Ribenboim

1999 Chapter 7 Theorem 1 p. 196.One can show that Qp has only finitely many extensions of given de-

gree (see 54.3 for the case of quadratic extensions); indeed, the Eisensteinpolynomials of fixed degree form a compact topological space X ⊂ Qp[x],and by Krasner’s Lemma the isomorphism type of the field Qp[x]/(f) isa locally constant function of f ∈ X; see Robert 2000 3.1.6, Weil 1967p. 207f, Reiner 1975 33.8 or Lang 1970 Chapter II Proposition 14. Forlocal fields of positive characteristic, the analogous statement is not true;see Exercise 3. Enumerations of extensions of Qp of low degree can befound in Pauli–Roblot 2001 and Klaas et al. 1997 IX; see also theonline database under http://math.asu.edu/∼jj/localfields/.

Theorem 56.13 implies that each field automorphism of a local fieldis an isometry with respect to the natural valuation, hence continuous.The automorphism group G of a finite field extension of Qp fixes eachelement of Q and of Qp = Q, hence G is finite. The Galois groups ofGalois extensions of local fields are always soluble; see Serre 1979 IV §2Corollary 5 p. 68 or Ribenboim 1999 9.Q p. 254. Thus each polynomialequation with coefficients in Qp is solvable by radicals over Qp.

Let q be a power of the prime p. The automorphism group G of Fq((t)),which coincides with the automorphism group of the power series ringFq[[t]] by 56.14, is notoriously complicated. The kernel N of the actionof G on Fq[[t]]/(t2) is called the Nottingham group. The quotient G/N isfinite, and N may be described as the set of all power series t+

∑i>1 ait

i

where ai ∈ Fq, with substitution as the group operation. This group N isa finitely generated pro-p-group which contains every finite p-group andevery finitely generated pro-p-group as a closed subgroup; see Leedham-

Green–McKay 2002 12.4.11, 12.4.16.Note that Fq((t)) admits many proper field endomorphisms; for ex-

ample, one can map t to tk for each k ≥ 2.

58.3 Local–global principles The fields R and Qp, p a prime, are thecompletions of Q with respect to the absolute values of Q; see 55.4, 44.11,51.4. Vaguely speaking, a local–global principle relates the behaviour of

Page 342: The Classical Fields

58 Local fields and locally compact fields 325

mathematical objects over Q to their behaviour over R and over thelocal fields Qp for all primes p. (The finite extensions of Q and of Fp(x)are often called global fields.) For example, a rational number a is asquare in Q if, and only if, a is a square in R (i.e. a ≥ 0) and in Qp

for each prime p (just observe that the p-adic value vp(a), as defined in56.3(a), is even if a is a square in Qp).

By considering the quadratic form ax2−y2, we see that this example isa very special case of the following famous theorem of Hasse–Minkowski:a quadratic form f =

∑i aix

2i with rational coefficients ai is isotropic

(as defined in Section 54) over Q if, and only if, f is isotropic over R andover Qp for each prime p (in fact, it suffices to consider only finitely manyprimes p). See Serre 1973, Borevich–Shafarevich 1966, Scharlau

1985, Cassels 1978 or Lam 2005 for proofs and extensions. Further-more, Hensel’s Lemma may be regarded as a local–global principle (seethe introduction to Chapter 5).

For Diophantine equations of higher degree (and for systems of morethan two quadratic forms), such a general local–global principle doesnot hold. For example, the cubic form 3x3 + 4y3 + 5z3 is isotropicover R and over Qp for each prime p, but not over Q (Selmer 1951; seeCassels 1986 Chapter 10 Lemma 9.1 or Borevich–Shafarevich 1966Chapter 1 7.6). For more examples see Exercise 4 and Cassels 1986Chapter 4 3bis. Still, the failure of a local–global principle can sometimesbe measured (via a Galois cohomological reformulation), thus salvagingsome connection between local and global behaviour; see the survey byMazur 1993.

58.4 Locally compact skew fields In the rest of this section we con-sider the classification of all locally compact skew fields. The connectedskew fields of this type are R, C and H; see 13.8 or 58.11. The fieldsFq((x)) and Qp, endowed with their natural valuation topologies, areexamples of locally compact disconnected fields (by 44.11 and 51.10);more generally, each local field is also an example (by 13.2(c) or 56.12).In fact, 13.2(c) shows that each skew field with finite (right or left) di-mension over a local field F is a locally compact skew field with respectto the product topology obtained from F .

The classification results 58.7 and 58.9 say that these remarks describeall locally compact skew fields.

There are two approaches to achieve this classification. The first ap-proach applies to locally compact skew fields F which are totally dis-connected: the study of the compact (open) subrings of such a skew

Page 343: The Classical Fields

326 The p-adic numbers

field F shows that the (unique) maximal compact subring R of F is avaluation ring (in a non-commutative sense); see Jacobson 1989 Sec-tion 9.13 for a good exposition. In fact, the maximal ideal of R is theset M = {x ∈ F | xn → 0} of all topologically nilpotent elements of F ,and R = F � (M � {0})−1.

Another approach, which can be found in Weil 1967 Chapter I (andin Bourbaki 1972 VI.9, Robert 2000 Appendix to Chapter 2), treatsthe connected and the totally disconnected case simultaneously; here oneuses a Haar measure of the additive group F+ for a direct constructionof an absolute value on F , as follows.

58.5 Lemma Let F be a locally compact skew field. Then there exists

a unique function | | : F → R such that

μ(aX) = |a|μ(X)

for each a ∈ F , each compact subset X ⊆ F and each Haar measure μ

of F+.

(i) This function | | : F → R is multiplicative and continuous, and

|1| = 1.

(ii) The balls Br := {x ∈ F | |x| ≤ r } with 0 < r ∈ R are compact and

form a neighbourhood base at 0 for F .

(iii) Some power | |s of | | is an absolute value of F inducing the given

locally compact topology of F .

(iv) Each discrete sub-skew-field F0 of F is finite.

Proof Recall that F carries a non-discrete Hausdorff topology. By defi-nition, a Haar measure of F+ is a translation invariant measure μ definedon all Borel subsets of F such that μ(X) > 0 if X �= ∅ is open in F .Each Haar measure is regular in the sense of 10.2 on (the σ-algebra gen-erated by the) compact sets; see Halmos 1950 §64 Theorem I p. 288 andp. 230. Like each locally compact group, F+ admits a Haar measure μ,and μ is determined uniquely up to a positive factor; see Halmos 1950§58 Theorem B p. 254 and §60 Theorem C p. 263. (For constructions ofthe Haar measure without using the axiom of choice see Hewitt–Ross

1963 §15, Loomis 1945 and Bredon 1963.) Moreover, μ(X) < ∞ foreach compact subset X of F ; see Halmos 1950 p. 255 or Hewitt–Ross

1963 15.8 and 11.24.Since x �→ ax is an automorphism of the topological group F+ for

0 �= a ∈ F , the assignment X �→ μ(aX) is also a Haar measure ofF+, hence μ(aX) = rμ(X) with a positive real factor r = |a|, which isindependent of X and μ.

Page 344: The Classical Fields

58 Local fields and locally compact fields 327

(i) The function | | is multiplicative, as an immediate consequence ofits definition, and |1| = 1. Now we show that | | is upper semicontinuous:let C be a compact neighbourhood of 0 in F , let a ∈ F and ε > 0. Sinceμ is regular (10.2), there exists an open set U in F with aC ⊆ U andμ(U) ≤ μ(aC) + ε. Furthermore we find a neighbourhood V of a withV C ⊆ U . Each x ∈ V satisfies xC ⊆ U , hence

|x| = μ(xC)/μ(C) ≤ μ(U)/μ(C) ≤ |a|+ ε/μ(C) .

Therefore | | is upper semicontinuous.In particular, | | is continuous at 0. Furthermore |x| = |x−1|−1 for

x �= 0, hence | | is also lower semicontinuous everywhere on F×, andtherefore continuous on F×.

(ii) Let V be a compact neighbourhood of 0 in F . There exists acompact neighbourhood W of 0 such that V W ⊆ V , as V is bounded by57.2(i). Since 0 is not isolated in V ∩W and | | is continuous, we canchoose a ∈ V ∩W such that 0 < |a| < 1. Then an ∈ V for all n ≥ 1, byinduction on n. From (i) we infer that 0 is the only accumulation pointof the sequence of powers an. This sequence is contained in the compactset V , hence it has the limit 0.

Now we show that each ball Br is compact. Let x ∈ Br � V . As anx

converges to 0, there exists a smallest integer n ∈ N with anx ∈ V . Thusanx ∈ V � aV , which implies |a|nr ≥ |anx| ≥ inf |V � aV | > 0. Thisgives an upper bound N for n which is independent of x. We concludethat Br is contained in the compact set V ∪⋃n≤N a−nV . FurthermoreBr is closed by (i), hence compact.

In order to show that these balls form a neighbourhood base at 0,we proceed indirectly and consider a neighbourhood U of 0 such thatB1/n �⊆ U for each n ∈ N. We obtain a sequence bn ∈ B1/n � U , whichhas an accumulation point b in the compact set B1. Each ball B1/n

contains bn, bn+1, bn+2, . . . (since these balls form a chain), hence b ∈B1/n. Thus b ∈ ⋂n∈N B1/n = {0} and b = 0, which is a contradiction,since bn �∈ U for all n ∈ N.

(iii) Let M := max |1 + B1|. The multiplicativity of | | implies that|x + y| ≤ M max{|x|, |y|} for all x, y ∈ F , since |x + y| = |1 + yx−1||x|and yx−1 ∈ B1 for 0 �= |x| ≥ |y|. (If M = 1, then | | is an ultrametricabsolute value; compare 55.3.)

Now choose s ∈ R such that Ms ≤ 2. Then the multiplicative functionϕ = | |s satisfies ϕ(x + y) ≤ 2 max{ϕ(x), ϕ(y)}. By Artin’s trick 55.7,this function ϕ is an absolute value of F , and by (ii) it induces the giventopology on F .

Page 345: The Classical Fields

328 The p-adic numbers

(iv) Since F0 is discrete, we infer |F0| = {0, 1} from (iii). Moreover,the additive group F0 is closed in F : if a sequence of elements an ∈ F0

converges to a ∈ F , then an+1 − an ∈ F0 converges to 0 and is finallyconstant, hence a ∈ F0. Thus F0 is a closed, discrete subset of thecompact ball B1, hence finite. �

The following proposition is a basic principle of functional analysis.

58.6 Proposition Let V be a Hausdorff topological vector space over

a locally compact skew field F . Then the following hold.

(i) Each finite-dimensional subspace S of V is closed in V and car-

ries the product topology obtained from any isomorphism with F d,

where d = dimS.

(ii) If V is locally compact, then V has finite dimension over F .

Proof To fix the notation, we consider V as a left vector space over F

(with scalars on the left). For vector spaces V with countable neigh-bourhood bases, the nets appearing in the proof for (i) can be replacedby sequences.

(i) We use the following property of F , which is a consequence of58.5(i,ii): if (an)n is a net in F such that the net (|an|)n converges to∞in the one-point compactification of R, then the net

(a−1

n

)n

convergesto 0 ∈ F ; this means that the inversion a �→ a−1 on F× extends to acontinuous involution on the one-point compactification F ∪{∞}. (Thus58.6 holds for complete fields F of type V , by Exercise 3 of Section 57.)

First we show by induction on d that each d-dimensional subspaceof V is closed in V ; this is obvious if d = 0. Let d > 0, let H bea subspace of dimension d − 1 and v ∈ V � H. By induction, H isclosed in V . In order to show that 〈H, v〉 = H ⊕ Fv is closed in V ,we consider a convergent net hn + anv → w ∈ V with hn ∈ H andan ∈ F ; we have to show that w ∈ 〈H, v〉. If the real numbers |an|are unbounded, then the scalars a−1

n accumulate at 0, hence the vectors−a−1

n (hn + anv) + v = −a−1n hn ∈ H accumulate at 0 ·w + v = v �∈ H, a

contradiction. This contradiction shows that the real numbers |an| arebounded, hence the scalars an accumulate at some a ∈ F (by 58.5(ii)),and the vectors (hn + anv)− anv = hn ∈ H accumulate at w − av. Weobtain w − av ∈ H and w ∈ 〈H, v〉.

Let H be a closed subspace of V and v ∈ V � H. We claim that thelinear bijection

λ : H × F → 〈H, v〉 : (h, a) �→ h + av

is a homeomorphism (then an easy induction completes the proof of (i)).

Page 346: The Classical Fields

58 Local fields and locally compact fields 329

Clearly λ is continuous; it suffices to verify the continuity of λ−1 at 0. Forthis we consider a net hn+anv converging to 0 with hn ∈ H and an ∈ F .We have to show that an → 0 (which entails hn → 0). Proceedingindirectly, we assume that the scalars an do not converge to 0. Then thereal numbers |a−1

n | do not converge to ∞, hence they have a boundedsubnet. Thus the scalars a−1

n have an accumulation point b in F (by58.5(ii)). We infer that the vectors −a−1

n (hn + anv) + v = −a−1n hn ∈ H

accumulate at b · 0 + v = v �∈ H, a contradiction, as H is closed in theHausdorff space V .

(ii) Let C be a compact neighbourhood of 0 in V , and pick a ∈ F with0 < |a| < 1. Then the powers an converge to 0, and the sets anC forma countable neighbourhood base of 0 in V (by 58.5(i, ii)). The compactset C is covered by finitely many sets vi + aC, 1 ≤ i ≤ d, where vi ∈ V .Let S = 〈v1, . . . , vd〉 be the (finite-dimensional) subspace generated byv1, . . . , vd. Then C ⊆ S + akC for each k ∈ N, as an easy inductionon k shows. Let v ∈ V . The vectors anv converge to 0, hence for allsufficiently large n we have anv ∈ C ⊆ S +a2nC and v ∈ S +anC. Thisshows that S is dense in V , and (i) implies S = V . �

If F is commutative, then the crucial fact 58.5(iii) that the topologyof F is induced by an absolute value may also be inferred from Theorem57.4; see Wi

‘es�law 1988 6.2 Lemma 3 p. 153. This suffices for the

following two results.

58.7 Theorem Each locally compact field F is a field extension of

finite degree of R, Qp or Fp((x)), where p is a prime. Thus the locally

compact fields are precisely the fields R, C and the local fields.

Proof By 58.5(iii), the topology of F is described by an absolute valueϕ of F .

First we consider the case where F has characteristic zero. ThenQ ⊆ F , and the restriction ϕ|Q is not trivial by 58.5(iv), hence anabsolute value of Q. Using Ostrowski’s enumeration (55.4) of all absolutevalues of Q, we conclude that ϕ|Q is a power of the usual absolute valueor of a p-adic absolute value. The topological closure Q of Q in F iscomplete, hence isomorphic to R or Qp as a topological field; see 44.11.Lemma 58.6 shows that F has finite dimension over Q.

Now we consider the case where F has characteristic p > 0. ThenFp ⊆ F . We can choose x ∈ F with 0 < ϕ(x) < 1. Then x is nota root of unity, hence x is transcendental over Fp, and the restrictionof ϕ to Fp(x) is obtained from the degree valuation of the field Fp(x)

Page 347: The Classical Fields

330 The p-adic numbers

of rational functions; see 56.2 and 56.3(c). The topological closure ofFp(x) is complete (43.9), hence isomorphic to the Laurent series fieldFp((x)); see 44.11. Lemma 58.6 shows that F has finite dimension overthe closure of Fp(x). �

58.8 Corollary The real field R is the only locally compact field which

is formally real. The complex field C is the only locally compact field

which is algebraically closed.

Proof Both statements follow from 58.7, because Qp is not formally realby 54.2, and a local field is never algebraically closed; see 58.2. �

A locally compact ring topology on a (skew) field is in fact a (skew)field topology; see 62.4 (or 13.4) and 62.2.

Thus Corollary 58.8 together with 6.4 shows that the abstract fieldR has only one locally compact ring topology (apart from the discreteand the indiscrete topology), namely the usual one, and that the (non-trivial) locally compact ring topologies of C are the images of the usualtopology under field automorphisms. This means that the connectednessassumptions in 13.9 can be replaced by non-discreteness. Similarly, eachlocal field has only one non-trivial locally compact ring topology (by58.5, 56.12 and 56.13).

58.9 Theorem Let F be a locally compact skew field. Then the centre

Z of F is not discrete, hence Z is isomorphic to R, C or to a local field,

and F has finite dimension over Z.

Proof Let P be the prime field of F . There exists an element a ∈ F

with |a| �= 0, 1; see 58.5. The field P (a) is not discrete, hence F hasfinite (right or left) dimension n over the topological closure P (a) by58.6. Clearly P (a) is a field (not necessarily contained in Z). Now ageneral algebraic result on skew fields implies that F has dimension atmost n2 over its centre Z; see Cohn 1995 3.1.4 p. 95. As F is infinite,we infer that Z is infinite, hence not discrete by 58.5(iv). This showsthat Z is one of the fields appearing in 58.7.

As an alternative, one can prove directly that the centre Z containsan element a with |a| �= 0, 1; see Weil 1967 I §4 Proposition 5 p. 20,Bourbaki 1972 VI.9.3 or Warner 1989 27.3 and 27.5. �

There remains the algebraic problem to classify all skew fields F offinite dimension over a centre Z as in 58.9. If Z ∈ {R, C}, then F = H isthe only proper skew field arising, by a classical result of Frobenius (see,for example, Jacobson 1985 7.7, Ebbinghaus et al. 1991 Chapter 8

Page 348: The Classical Fields

58 Local fields and locally compact fields 331

§2, Palais 1968 or 58.11 below). For local fields Z, this problem wassolved by Hasse in 1931; for each local field Z, there are infinitely manypossibilities for F , to be described now.

58.10 Cyclic algebras Let Z be any field, and let E|Z be a cyclicGalois extension of degree n ∈ N. Choose a generator γ of the Galoisgroup GalZE and c ∈ Z×. The cyclic algebra [E|Z; γ, c] determined bythese data is the associative Z-algebra defined by the following proper-ties: it is an n-dimensional (left) vector space

⊕n−1i=0 Ebi over E with

a basis 1, b, . . . , bn−1 consisting of powers of some element b, and themultiplication satisfies the rules bn = c and bx = xγb for all x ∈ E.

Denote by NE|Z ≤ Z× the group of all non-zero norms of E|Z. Onecan show that [E|Z; γ, c] is always a simple algebra of dimension n2 overits centre Z, that [E|Z; γ, c] is Z-isomorphic to [E|Z; γ, c′] precisely ifcNE|Z = c′NE|Z , and that [E|Z; γ, c] is a skew field if n is the order ofcNE|Z in the norm factor group Z×/NE|Z ; see Reiner 1975 30.4 and30.7, Jacobson 1989 8.5 or Scharlau 1985 Chapter 8 §12.

For example, the cyclic algebra [C|R; γ, c] with c ∈ R× is isomorphicto the skew field H of Hamilton’s quaternions if c < 0, and isomorphicto the ring End R2 = R2×2 of all real 2× 2 matrices if c > 0.

For local fields Z there are many possibilities. The unique unramifiedextension Zn of degree n over Z is a cyclic Galois extension of Z (see58.2), the norm factor group Z×/NZn|Z is cyclic of order n, and eachprime element π of Z represents a generator of Z×/NZn|Z (see Jacobson

1989 p. 610 or Reiner 1975 14.1). Let q be the order of the finite residuefield of Z, and let γ ∈ GalZZn be the distinguished generator whichinduces the Frobenius automorphism x �→ xq on the finite residue fieldof Zn (equivalently, ζγ = ζq for some root of unity ζ ∈ Zn of orderqn − 1). Then the cyclic algebra

Zn,r := [Zn|Z; γ, πr]

is a skew field (of dimension n2 over its centre Z) if the integer r is primeto n. The ϕ(n) integers r ∈ {1, . . . , n} prime to n give ϕ(n) skew fieldsZn,r which are mutually not isomorphic as Z-algebras. We remark thatthe skew field Zn,r contains each field extension of Z of degree n as amaximal subfield; see Serre 1979 XIII.3 Corollary 3 p. 194 or Reiner

1975 31.11.Choosing for γ another generator leads to the same skew fields Zn,r,

since [E|Z; γr, cr] is Z-isomorphic to [E|Z; γ, c] for each integer r primeto n; see Reiner 1975 30.4(i) or Scharlau 1985 Chapter 8 12.4.

Page 349: The Classical Fields

332 The p-adic numbers

58.11 Theorem (Pontryagin, Jacobson, Hasse) Let F be a locally

compact skew field. Then F is isomorphic to a cyclic algebra over a

locally compact field. More precisely, F is isomorphic to R, C, H or to

a cyclic algebra Zn,r, where Z is a local field, n ∈ N and r ∈ {1, . . . , n}is prime to n.

Proof By Theorem 58.9, such a skew field F has finite dimension overits centre Z. General results on skew fields say that this dimension is asquare n2, and that the maximal subfields of F are precisely the fields E

with Z ≤ E ≤ F and [E : Z] = n; see, for example, Cohn 2003b 5.1.12or Reiner 1975 Theorem 7.15.

If Z = C, then F = Z, as C is algebraically closed. If Z = R �= F ,then E ∼= C and n = 2. The complex conjugation γ ∈ AutE is inducedby an inner automorphism of F , by the Skolem–Noether Theorem (seeJacobson 1989 Theorem 4.9 p. 222). Thus we find b ∈ F× with bx =xγb for each x ∈ E. We have F = E ⊕ Eb, as b /∈ E. Furthermoreb2 commutes with E and b, so b2 = c ∈ Z× = R×. If c > 0, thenc = r2 with r ∈ R, hence (b + r)(b − r) = b2 − c = 0 and b = ±r ∈ Z,a contradiction. This shows that c < 0, thus the cyclic algebra F =[E|Z; γ, c] is isomorphic to H.

Now let Z be a local field. In this case, the crucial step is to provethe existence of a maximal subfield E of F such that the extension E|Zis unramified; for this we refer to the two proofs given in Serre 1979XII §1 and §2. Now E|Z is a cyclic Galois extension (compare 58.2),and the Skolem–Noether Theorem yields an element b ∈ F× such thatconjugation by b induces on E a generator γ of the Galois group GalZE.This implies that the elements 1, b, . . . , bn−1 are linearly independentover E (compare Jacobson 1989 Theorem 8.8 p. 478, Scharlau 1985Chapter 8 Theorem 12.2 p. 317, or Cohn 1995 Theorem 3.5.5 p. 124).For dimension reasons we obtain F =

⊕n−1i=0 Ebi. Furthermore, the

element bn commutes with E and b, hence bn ∈ Z. This shows that F isa cyclic algebra [E|Z; γ, bn] which is isomorphic to one of the skew fieldsZn,r described in 58.10.

Other proofs of Theorem 58.11 can be found in Weil 1967 I §4 Propo-sition 5 p. 20 and Jacobson 1989 Theorem 9.21 p. 607. �

We remark that the locally compact (skew) field topology of a skewfield F as in Theorem 58.11 is uniquely determined (as a consequenceof the results 58.9, 13.9 and 58.6) except if F = C (compare 14.11 and14.12).

Page 350: The Classical Fields

58 Local fields and locally compact fields 333

58.12 Brauer groups For an arbitrary field Z, we denote by B(Z)the set of all Z-isomorphism types of skew fields D of finite dimensionover their centre Z. One can make B(Z) into an abelian group, usingthe tensor product:

D1 + D2 = D3 ⇐⇒ D1 ⊗Z D2∼= Dk×k

3 for some k ∈ N .

Then Z ∈ B(Z) is the neutral element, and the additive inverse ofD is −D = Dop, the skew field with the opposite multiplication, asD ⊗Z Dop ∼= Zk×k with k = dimZ D. The Brauer group B(Z) is al-ways a torsion group. See Cohn 2003a 5.4, Cohn 2003b 5.2 and 5.5.4,Jacobson 1989 4.7 and Theorem 8.12, or Reiner 1975 Section 28 andTheorem 29.22 for details and for more elegant descriptions.

If Z is algebraically closed, then B(Z) = {0}, and the same holdsfor each finite field Z by Wedderburn’s theorem. The Brauer groupB(R) = {R, H} of the real numbers is a cyclic group of order 2, by aresult of Frobenius (compare 58.11), and the same is true for every realclosed field (see 12.10).

Now let Z be a local field. Then

B(Z) = {Zn,r | n ∈ N, r ∈ {1, . . . , n} is prime to n}

by 58.11. One can show that Zn,r +Zn,s = Zn,r+s in B(Z); see Reiner

1975 Theorem 30.4 or Scharlau 1985 Chapter 8 Theorem 12.7. Thisimplies that the so-called Hasse invariant, i.e., the mapping

B(Z)→ Q/Z : Zn,r �→ rn

+ Z ∈ Q/Z ,

is an isomorphism B(Z) ∼= Q/Z, for each local field Z; see Reiner 1975Theorem 30.4, Jacobson 1989 Theorem 9.22 or Serre 1979 XIII §3.

For global fields like Q, the situation is slightly more complicated,there is an exact sequence

0→ B(Q)→ B(R)⊕⊕p B(Qp)→ Q/Z→ 0 ,

which originates from a local–global principle for the splitting of skewfields and from a non-trivial relation between the Hasse invariants of thecompletions D⊗Qp of a skew field D ∈ B(Q). One can show that eachskew field of finite dimension over Q is a cyclic algebra (over a finiteextension Z of Q; compare 58.10); see Reiner 1975 Section 32 or Weil

1967 XIII §3 and §6.

Page 351: The Classical Fields

334 The p-adic numbers

Exercises(1) Show that the polynomial f = (xp − 1)/(x− 1) is irreducible over Qp anddefines a totally ramified extension of Qp of degree p − 1.

(2) Let c ∈ Qp be a non-square and F = Qp(√

c ). Show that F |Qp is unram-

ified precisely if vp(a2 − c) is even for each a ∈ Qp. Deduce that Q2(√

3 )|Q2

is ramified, and that Q2(√

5 )|Q2 is unramified.

(3) Show that F2((t)) has infinitely many separable quadratic extensions (con-sider the splitting fields of polynomials x2 + tnx + t with n ∈ N).

(4) Show that the polynomial (x2 − 2)(x2 − 17)(x2 − 34) has a root in Qp andin Fp for each prime p (but not in Q).

(5) Let E|Z be a Galois extension of degree 2, let γ be the generator of GalZEand c ∈ Z×. Show that the cyclic algebra [E|Z; γ, c] is isomorphic to the Z-algebra which consists of all matrices

x cyγ

y xγ

«

with x, y ∈ E. Show directly that these matrices form a skew field precisely ifc = xxγ for all x ∈ E.

(6) Let Z = Fq((x)) for some prime power q, and let r ∈ {1, . . . , n} be primeto n. Show that the cyclic algebra [Zn|Z; γr, x] as in 58.10 is isomorphic tothe skew Laurent series field Fqn((t)) with the multiplication rules tn = x and

ta = aqr

t for a ∈ Fqn .

Page 352: The Classical Fields

6

Appendix

The first section of the appendix collects a few facts on ordinal andcardinal numbers. Then we deal with topological groups, and we sum-marize the duality theory of locally compact abelian groups. Finally wepresent basic facts and constructions of field theory.

61 Ordinals and cardinals

This section is based on the system NBG (von Neumann–Bernays–Godel) of set theory, having as its primitive notions classes as ob-jects, a predicate set for certain classes, and the element relation ∈ ;see Dugundji 1966 Chapter I §8, Rubin 1967, or Smullyan–Fitting

1996. Each element of a class is a set, and so is each subclass of a set.The class of all sets is partially ordered by the relation ∈. The axiom offoundation stipulates that each ∈-descending sequence of sets is finite.

In this system, the class O of all ordinals or ordinal numbers consistsof all sets which are linearly ordered by the relation ∈, formally

ν ∈ O � ∀σ∈ν

∀ξ

[(ξ ∈ σ ⇒ ξ ∈ ν) ∧ (ξ ∈ ν ⇒ ξ ∈ σ ∨ ξ = σ ∨ σ ∈ ξ)] .

By the axiom of foundation, each ordinal number is even well-ordered(which means that every non-empty subset has a smallest element). Anywell-ordered set is order isomorphic to a unique ordinal. The well-ordering principle says that every set can be well-ordered, it impliesthat there are ordinal numbers of arbitrary cardinality.

Note that the well-ordering principle is equivalent to the axiom ofchoice (any Cartesian product of non-empty sets is non-empty) and toZorn’s lemma (if each chain in a partially ordered set S has an upperbound in S, then there exists at least one maximal element in S). This is

335

Page 353: The Classical Fields

336 Appendix

an immediate consequence of Hausdorff’s maximal chain principle (eachchain in S is contained in a maximal chain). Conversely, the maximalchain principle can be obtained by applying Zorn’s lemma to the set ofall subchains of S, ordered by inclusion. The well-ordering principle alsofollows easily: Consider the set S of all injective maps ϕ : ξ →M whereξ is an ordinal number. Write ϕ ≤ ψ if ψ is an extension of ϕ. Any chainin S< has a common extension, i.e., an upper bound. By Zorn’s lemmathere is a maximal element σ : μ → M in S. The map σ is necessarilysurjective and σ carries the well-ordering from μ to M . A proof of thewell-ordering principle from the axiom of choice is given in Dugundji

1966 Chapter II §2. For a full account of the axiom of choice and relatedtopics see Rubin–Rubin 1985.

61.1 If σ ∈ ν ∈ O, then σ ∈ O.

Proof If ξ ∈ η ∈ σ, then successively η ∈ ν and ξ ∈ ν. The possibilitiesξ ∈ η ∈ σ ∈ ξ or ξ ∈ η ∈ σ = ξ contradict the axiom of foundation.Hence ξ ∈ σ.

If ξ, η ∈ σ, then ξ, η ∈ ν and therefore ξ ∈ η or ξ = η or η ∈ ξ. �

61.2 If μ, ν ∈ O and μ ⊂ ν (proper inclusion), then μ ∈ ν.

Proof Let α be the ∈-minimal element of the non-empty set ν � μ. Weshow that α = μ: in fact, ξ ∈ α⇒ ξ ∈ μ by minimality of α. Conversely,ξ ∈ μ⇒ ξ ∈ α, because α /∈ μ and hence ξ �= α and α /∈ ξ. �

61.3 Any non-empty classM⊆ O has a minimum⋂M∈ O.

Proof⋂M = δ is a set. If ξ ∈ σ ∈ δ, then σ ∈ μ and hence ξ ∈ μ

for each μ ∈ M. This means that ξ ∈ δ. Similarly, it follows that δ islinearly ordered by ∈. Thus, δ ∈ O. If δ ⊂ μ for each μ ∈M, then δ ∈ δ

by 61.2, but this contradicts the axiom of foundation. �

61.4 Corollary The class O itself is well-ordered by ∈.

Proof If μ, ν ∈ O, then μ∩ν = min {μ, ν}, thus O is linearly ordered. �

61.5 If M is a subset of O, then⋃M = supM∈ O.

Proof (a) Because M is a set, so is σ =⋃M (by an axiom of NBG).

If ξ ∈ η ∈ σ, then η ∈ μ for some μ ∈ M. Consequently, ξ ∈ μ andthen ξ ∈ σ. If ξ, η ∈ σ, then there are μ, ν ∈ M such that ξ ∈ μ andη ∈ ν. Since M is linearly ordered, we may assume that μ ⊆ ν. By thedefinition of O it follows that ξ and η are comparable, and σ ∈ O.

Page 354: The Classical Fields

61 Ordinals and cardinals 337

(b) If μ ∈ M, then μ ⊆ σ and hence μ = σ or μ ∈ σ. Therefore, σ isan upper bound, even a least upper bound, since ξ ∈ σ implies ξ ∈ μ forsome μ ∈M. �

61.6 Explicit description Each ordinal number is the set of all its

predecessors. The immediate successor of ν is ν + 1 := ν ∪ {ν}. �

61.7 Small ordinals The first ordinal numbers are 0 = ∅, 1 = {0},2 = {0, 1}, and the subsequent finite numbers. The smallest infiniteordinal number ω = ω0 is the set of all finite ordinals, it is followed byω + 1, ω + 2, . . . , ω + ω = ω · 2, . . . , ω · 3, . . . , ω · ω = ω2, . . . , ω3, . . . ,ωω = 2ω, . . . , ωωω

= 3ω, . . . , ωω, . . . . All these ordinal numbers arecountable, because a (countable) sequence of countable ordinals has acountable upper bound. The first uncountable ordinal number Ω = ω1

is the set of all countable ordinals. �

61.8 Cardinal numbers A class of bijectively equivalent sets maybe considered as a cardinal number. It is more convenient, however, toselect a canonical element of the class to denote the cardinal number: ineach class of bijectively equivalent ordinals there is a smallest one, the so-called initial ordinal, or cardinal, of this class. The infinite cardinals forma well-ordered subclass C ofO, and there is a (unique) order isomorphism(α �→ ωα) : O → C. In contexts where the relevant property is size ratherthan ordering, the ordinal ωα is usually designated by ℵα.

The cardinality (or cardinal number) cardM of a set M is defined tobe the smallest element α ∈ O such that there exists an injective mapμ : M → α. Thus ℵ0 = card N. Moreover M ⊆ N implies card M ≤cardN .

For more details see Dugundji 1966 Chapter II §7, Rubin 1967,Bachmann 1967, Levy 1979, or Ciesielski 1997. Ordinals and cardi-nals in the system ZF of Zermelo and Fraenkel are presented in Chapter 1of Holz et al. 1999. We mention the Theorem of Cantor–Bernstein (orSchroder–Bernstein): If there are injections ϕ : X → Y and ψ : Y → X,then there exists also a bijection σ : X → Y (for proofs see Rauten-

berg 1987).

61.9 Addition and multiplication We define cardM + cardN =cardM ∪N if M ∩N = ∅, and cardM · cardN = cardM ×N .

61.10 Powers We denote by MX the set of all maps ϕ : X → M , asusual. Exponentiation of cardinal numbers is defined by cardM card X =cardMX . Note that card 2X = 2card X .

Page 355: The Classical Fields

338 Appendix

Attention: the least upper bound ωω of all ordinal numbers ωn withfinite n is to be distinguished from the set ℵℵ0

0 ; the first one is countable,the other is not.

61.11 Theorem If cardM > 1 and X �= ∅, then cardMX > cardX.

Proof Obviously, cardX ≤ cardMX . Assume that there is a bijectivemap (x �→ ϕx) : X →MX . For each x ∈ X choose an element ψ(x) ∈M

with ψ(x) �= ϕx(x); this requires the axiom of choice. Then ψ �= ϕx forany x, a contradiction. �

The following theorem plays an important role in determining thecardinality of several sets constructed in this book.

61.12 Squares For α ∈ O we have ℵ 2α = ℵα.

Proof (a) One has ℵ 20 = (cardω)2 = card(ω · ω) = ℵ0 by 61.7.

(b) It can easily be verified (compare Smullyan–Fitting 1996 Chap-ter 9 §8) that the definition (α, β) ≺ (γ, δ) �

(α∪β < γ∪δ) ∨ (α∪β = γ∪δ ∧ α < γ) ∨ (α∪β = γ∪δ ∧ α = γ ∧ β < δ)

yields a well-ordering ≺ of O×O. Since each well-ordered set is orderisomorphic to a unique ordinal, the relation ≺ determines an injectivemap κ : O×O → O.(c) Assume that there is a least α such that ℵ 2

α > ℵα. Then α > 0by step (a). We have ωα×ωα =

⋃μ∈ωα

μ×μ. From the minimality ofα we infer that cardμ×μ = cardμ and hence κ(μ×μ) ∈ ωα. There-fore, κ(ωα×ωα) =

⋃μ∈ωα

κ(μ×μ) ⊆ ωα and ℵ 2α = ℵα contrary to the

assumption.For a different proof see Dugundji 1966 Chapter II Theorem 8.5;

compare also Bachmann 1967 §28 and Deiser 2005. �

61.13 Corollary (a) If S =⋃

ι∈K Mι, where cardK ≤ ℵα and

cardMι ≤ ℵβ for each ι ∈ K, then cardS ≤ ℵα · ℵβ = ℵmax{α,β}.(b) For each finite n > 0 we have ℵn

α = ℵα, and ℵℵα = (2ℵ0)ℵα =2ℵ0ℵα = 2ℵα . �

61.14 Corollary If M is an infinite set, and if F consists of finite

sequences in M (or of finite subsets of M), then cardF ≤ cardM .

Proof If card M = ℵα, then there are ℵnα = ℵα sequences of length n in

M , and at most ℵα subsets of size n. Hence card F ≤ ℵ0 · ℵα = ℵα. �

Page 356: The Classical Fields

61 Ordinals and cardinals 339

61.15 Theorem If β ≤ α + 1 then ℵℵα

β = 2ℵα .

Proof We have ℵβ ≤ ℵα+1 ≤ 2ℵα by 61.11. Hence 2ℵα ≤ ℵℵα

β ≤(2ℵα)ℵα = 2ℵα . �

61.16 Corollary If Sym M denotes the group of all permutations of

M (i.e. bijections M → M), then card SymM = 2card M . Similarly,

card{S | S ⊆M ∧ cardS = cardM } = 2card M .

Proof Let c := cardM and partition M into c pairs. There are 2c

permutations of order 2 mapping each pair to itself. Therefore, 2c ≤card SymM ≤ cc = 2c.

Partition M into two subsets A and M � A of cardinality c. Then2c = card{S | A ⊆ S ⊆M } ≤ card{S | S ⊆M ∧ cardS = c} ≤ 2c. �

61.17 The continuum problem By the well-ordering principle, wehave card R = 2ℵ0 = ℵα for some ordinal number α. Cantor’s theorem61.11 shows that α > 0, but otherwise does not give any clue how thevalue of α might be determined. See Baumgartner–Prikry 1977 fora survey of the history of this so-called continuum problem. Cantor andHilbert conjectured that 2ℵ0 = ℵ1 (continuum hypothesis, CH). Thisis the first case of the generalized continuum hypothesis (GCH) whichasserts that 2ℵα = ℵα+1 for each α ∈ O. See also Stillwell 2002.

Without using ordinals, GCH may be expressed as follows: if m, n

denote classes of infinite bijectively equivalent sets, then m ≤ n < 2m

entails that m = n. This form of GCH implies the axiom of choice;compare Specker 1954, Gillman 2002 or Cohen 1966 Chapter 4 §12p. 148ff.

By the work of Godel 1940 and Cohen 1966 it has become clearthat the continuum hypothesis can neither be proved nor disproved onthe basis of the usual axioms of set theory. More precisely, if the systemNBG including the axiom of choice is consistent, then it is not possi-ble to derive a contradiction by assuming in addition the continuumhypothesis or its negation. Specifically, each assertion 2ℵ0 = ℵn with0 < n ∈ ω is consistent with the usual set theory, but not 2ℵ0 = ℵω.Analogously, the continuum hypothesis is independent of the axiom sys-tem ZF of Zermelo and Fraenkel; compare Woodin 2001. Moreover,GCH is consistent with ZF as well as with NBG; see the survey byMartin 1976 or the more recent treatment of the continuum problemby Smullyan–Fitting 1996, in particular p. 185 and Chapter 19 §6.

Page 357: The Classical Fields

340 Appendix

Exercises(1) Show that Ω2 is well-ordered by the lexicographic ordering.

(2) Show that R has only ℵ = card R countable subsets.

(3) Prove the following theorem of Konig: if card Aι < card Bι for ι ∈ K, thenP

ι∈K card Aι <Q

ι∈K card Bι.

(4) Use Konig’s theorem to prove that 2ℵ0 = ℵω.

62 Topological groups

The additive and multiplicative groups of the fields considered in thisbook are topological groups of a very special nature. For better un-derstanding and in order to avoid repetition of arguments, these groupsought to be looked at in the context of topological groups in general. Weintroduce a few basic notions of the theory and collect some simple facts.For more details see Pontryagin 1986, Bourbaki 1966 Chapter III,Hewitt–Ross 1963 Chapter II or Stroppel 2006.

62.1 Definitions Assume that (G, ·) is a group and that (G, τ) is atopological space. Then G = (G, ·, τ) is called a semi-topological groupif multiplication is continuous in each variable separately (i.e., if left andright multiplication with a fixed element are homeomorphisms).

G is said to be a para-topological group if multiplication is contin-uous in both variables simultaneously. If, moreover, inversion is alsocontinuous, then G is a topological group.

Remarks (R, +) with Sorgenfrey’s topology (5.73 or 24.17) is a para-topological, but not a topological group. Any infinite group with thecofinite topology (each neighbourhood has a finite complement) is semi-topological, but not para-topological.

62.2 Proposition A regular, locally compact semi-topological group

is a topological group.

This is Theorem 2 in Ellis 1957. The conclusion is true, in fact,under rather weak hypotheses; see Wu 1962 and Solecki–Srivastava

1997. �

The set Vε = { (x, y) | x, y ∈ R ∧ |y − x| < ε} is a surroundingof the diagonal in R×R. The filter (see 21.1) generated by all thesesurroundings (with ε > 0) is a uniformity U on R. A function f : R→ Ris uniformly continuous if f−1(U) ⊆ U. These notions can be generalizedto arbitrary topological groups as follows.

Page 358: The Classical Fields

62 Topological groups 341

62.3 If U denotes a neighbourhood of 1 in the topological group G, thenthe surroundings VU = { (x, y) ∈ G×G | yx−1 ∈ U } of the diagonal inG×G generate a uniformity on G, the right uniformity ; it makes rightand left multiplications of G uniformly continuous (see Bourbaki 1966Chapter III §3 no.1, or Hewitt–Ross 1963 Chapter II §4). Inversioninterchanges the right and the left uniformity. For a fuller treatment ofuniform structures on groups see Roelcke–Dierolf 1981.

62.4 Regularity If G is a topological group and if V denotes the filter

of all neighbourhoods of 1, then⋂

V = {1} implies that G is a T1-space

(i.e., points are closed in G), and then G is even completely regular (but

not necessarily normal).

Proof By homogeneity, each point has a neighbourhood that does notcontain the element 1. Hence G � {1} is open. Because the groupoperations are continuous, there are arbitrarily small neighbourhoods ofthe form UU−1 in V, and U ⊆ UU−1 shows that G is regular; see alsoBourbaki 1966 Chapter II §1 Proposition 3; according to Bourbaki

1966 Chapter IX §1 no.5, the space G is even completely regular. Theexistence of non-normal regular topological groups has been shown byMarkov 1945; compare Hewitt–Ross 1963 8.10–12. �

62.5 Subgroups Let G be a topological group. If H is a subgroup of

the group G, and if H is the topological closure of H in G, then H is

also a subgroup of G.

Proof The map γ = ((x, y) �→ xy−1) is continuous on G×G. SinceH×H = H×H and H is a group, γ(H×H) ⊆ γ(H×H) ⊆ H. �

62.6 Quotients Let H be a subgroup of a topological group G. Thequotient topology on the coset space G/H = {xH | x ∈ G} is charac-terized by the property that the canonical projection η : G → G/H :x �→ xH is continuous and open. In fact, if U is open in G, then sois the preimage UH =

⋃h∈H Uh of η(U), and this means that η(U) is

open in G/H; see also Hewitt–Ross 1963 Chapter II, 5.15–17.

The adequate notion of a substructure of a topological group is thatof a closed subgroup rather than just a subgroup.

62.7 Any open subgroup H of a topological group is also closed.

Proof The complement of H is a union of cosets of H, hence open. �

Page 359: The Classical Fields

342 Appendix

From now on, we assume that all topological groups are T1-spaces.

62.8 Proposition If H is a closed subgroup of a topological group G,

then G/H is a Hausdorff space. In fact, G/H is even regular.

Proof If x is in the open complement of H, then, by continuity of thegroup operations, there are neighbourhoods U of 1 and V of x such thatU−1V ∩ H = ∅ = V H ∩ UH, and 62.6 shows that η(V ) and η(U) aredisjoint neighbourhoods of x and 1, respectively. For the last assertion,see Hewitt–Ross 1963 Chapter II 5.21. �

62.9 Metric If G is a topological group with a countable neighbour-

hood base at 1, then G is metrizable, and there exists a metric which is

invariant under right multiplications.

For proofs see Bourbaki 1966 Chapter IX §3 or Hewitt–Ross 1963Chapter II §8.

62.10 Normality A locally compact group is paracompact and hence

is a normal space.

Proof A topological space is called paracompact if each open cover hasa locally finite refinement. If V is a compact symmetric (V = V −1)neighbourhood of 1 in G, then H =

⋃n∈N V n is an open subgroup of G.

Being a countable union of the compact sets V n, the group H is a Lin-delof space and hence is paracompact; see Dugundji 1966 Chapter VIIITheorem 6.5. Obviously, G, as a union of disjoint cosets of H, is alsoparacompact. By a standard result, paracompact spaces are normal;compare Dugundji 1966 Chapter VIII Theorem 2.2. �

62.11 Factor groups If N is a closed normal subgroup of the topo-

logical group G, then G/N is a topological group.

Proof Recall from 62.6 that the canonical epimorphism η : G→ G/N iscontinuous and open. If UV −1 ⊆W for open subsets U, V,W ⊆ G, thenη(U)η(V )−1 ⊆ η(W ). This shows the continuity of the group operationsin G/N ; see also Husain 1966 §24. �

62.12 Proposition If K is a connected closed subgroup of the topo-

logical group G, then G is connected if, and only if, G/K is connected.

Proof If G is connected, so is its continuous image G/K. If G is notconnected, then there exists a continuous map ϕ of G onto the discretespace {0, 1}, and ϕ is constant on each (connected) coset xK. Hence

Page 360: The Classical Fields

62 Topological groups 343

ϕ induces a continuous surjection ψ : G/K → {0, 1}, and G/K is notconnected. �

62.13 Connected component If G is a topological group and K

is the largest connected subset of G containing the element 1, then K

is a closed normal subgroup of G, the connected component , and G/K

is totally disconnected. If G is locally compact, then dimG/K = 0(compare Exercise 3).

In fact, K×K is connected, hence its continuous image KK−1 is con-tained in K, and K is invariant under inner automorphisms. Moreover,K is closed in G; see Dugundji 1966 Chapter V Theorem 1.6. By62.12, the connected component of G/K consists of one element, andthis implies that G/K is totally disconnected. For the last claim seeHofmann–Morris 1998 E8.6. �

Finally, we mention a fundamental structure theorem.

62.14 Theorem (Mal’cev–Iwasawa) Let G be a locally compact,

connected topological group.

(1) Then there exists a maximal compact subgroup C, and each maximal

compact subgroup of G is connected and conjugate to C. Moreover, each

compact subgroup of G is contained in a maximal one.

(2) There are closed subgroups Sμ∼= R such that the multiplication map

(k, s1, . . . , sm) �→ ks1 . . . sm : C×S1× . . .×Sm → G

is a homeomorphism. In particular, G ≈ C × Rm.

For a proof see Iwasawa 1949 Theorem 13, compare also Hofmann–

Terp 1994. The result has first been obtained for Lie groups. It canthen be extended to general locally compact, connected groups, becauseeach of these has arbitrarily small compact normal subgroups with Liefactor groups.

Exercises(1) If inversion in a topological group is uniformly continuous with respectto the right uniformity, then multiplication is uniformly continuous in twovariables.

(2) If H is a locally compact subgroup of the regular topological group G,then H is closed in G.

(3) A topological space is called 0-dimensional, if each point has arbitrarilysmall open and closed neighbourhoods. Show that every locally compact,0-dimensional group has arbitrarily small compact open subgroups.

Page 361: The Classical Fields

344 Appendix

63 Locally compact abelian groups and Pontryagin duality

Each locally compact abelian (or LCA) group A has a dual A∗, whichis again an LCA group. The Pontryagin–van Kampen Theorem assertsthat A∗∗ is canonically isomorphic to A. Therefore, each question con-cerning A is reflected in a question on A∗. Usually, one of the twoproblems is easier to deal with than the other. This makes Pontryaginduality the most important tool in the structure theory of LCA groups.For an introduction to duality theory see Pontryagin 1986 or Morris

1977. A full account is given in Hewitt–Ross 1963 Chapter VI; seealso the elegant exposition in Hofmann–Morris 1998 Chapter 7, orStroppel 2006.

In this section, all groups are commutative.

63.1 Characters Let A be an abelian topological group, and denote byT = R/Z the torus group; compare 8.7. A character of A is a continuoushomomorphism χ : A→ T. It is convenient to denote the values of χ in Tby 〈x, χ〉 rather than by χ(x). Under pointwise addition, the charactersof A form a group A∗.

63.2 The compact-open topology on A∗ is generated by all sets

W (C,U) = {χ ∈ A∗ | 〈C,χ〉 ⊆ U } ,

where C is compact and U is open in T; the sets W (C,U) with 0 ∈ U

form a neighbourhood base at 0. The group A∗ will always be equippedwith the compact-open topology. With this topology, A∗ is a topologicalgroup.

63.3 Adjoint Any continuous homomorphism γ : A → B induces

a continuous ‘adjoint’ homomorphism γ∗ : B∗ → A∗ defined by the

condition 〈a, γ∗β〉 = 〈aγ, β〉. Obviously, γ∗ is injective, if γ is surjective.

Continuity of γ∗ follows from 〈C, γ∗β〉 = 〈Cγ, β〉 and the fact that Cγ

is compact if C is.

63.4 Natural homomorphism For each a ∈ A, the evaluation map

δa = (χ �→ 〈a, χ〉) : A∗ → T is a character of A∗, and δA = (a �→ δa) :A → A∗∗ is a homomorphism. If δA is an isomorphism of topological

groups, then A is called reflexive.Note that the map δa is continuous: if Ω = W (a, U), then 〈a,Ω〉 ⊆ U

by definition.

63.5 If A is compact, then A∗ is discrete, and if A is discrete, then A∗

is compact.

Page 362: The Classical Fields

63 Locally compact abelian groups and Pontryagin duality 345

Proof If A is compact, then W (A,U) is open, and if U is small, thenW (A,U) = {0} because T has no small subgroups. If A is discrete, thenthe topology on A∗ is the topology of pointwise convergence, and A∗ isclosed in the compact space TA. �

63.6 Theorem Let A be an LCA group.

(a) Let C be a compact neighbourhood of 0 in A. If U is a sufficiently

small neighbourhood of 0 in T, then the closure of W (C,U) is

compact.

(b) The character group A∗ is locally compact.

The proof uses Ascoli’s theorem; for details see Morris 1977 Theo-rem 10, Hofmann–Morris 1998 Theorem 7.7(ii) or Stroppel 2006Theorem 20.5.

The duality theorem 63.27 asserts that A is reflexive, i.e., that δA isinjective and surjective, continuous and open, if A is locally compact.These four properties of δA will be discussed separately.

A topological space A is said to be a k-space if it has the followingproperty: O ⊆ A is open in A if, and only if, O ∩ C is open in C foreach compact subspace C of A.

In particular, A is a k-space, if each point of A has a countable neigh-bourhood base. (Use the fact that a convergent sequence together withits limit point is compact, or see Dugundji 1966 XI.9.3.)

63.7 Continuity If the underlying space of the topological group A is

a k-space, in particular, if A is locally compact, then δA : A → A∗∗ is

continuous.

A proof is given in Hofmann–Morris 1998 Theorem 7.7(iii). ForLCA groups, the proof is easier; see Hewitt–Ross 1963 24.2, Pon-

tryagin 1986 Definition 37 or Stroppel 2006 Lemma 20.10.

63.8 As topological groups, (A⊕B)∗ and A∗ ⊕B∗ are isomorphic.

63.9 Corollary If A and B are reflexive, so is A⊕B.

63.10 Vector groups The characters of R are given by x �→ ax + Zwith a ∈ R. Hence R is self-dual, moreover, Rn is self-dual for anyn ∈ N; see 8.31a (or Exercise 5 of Section 52).

63.11 Definition We say that A has enough characters, if for eacha ∈ A, a �= 0, there is a character χ ∈ A∗ such that 〈a, χ〉 �= 0, i.e. if thecharacters separate the elements of A. The following fact is obvious.

Page 363: The Classical Fields

346 Appendix

63.12 Injectivity The natural map δA : A → A∗∗ is injective if, and

only if, A has enough characters.

63.13 Theorem Each compact group has enough characters.

For a proof see Pontryagin 1986 Theorems 31 and 33 or Hofmann–

Morris 1998 Corollary 2.31. The case of general LCA groups will bereduced to the compact case; see 63.18.

63.14 Splitting Theorem Any LCA group A is isomorphic as a

topological group to Rn ⊕H, where H has a compact open subgroup.

Proof Hofmann–Morris 1998 Theorem 7.57; see also Pontryagin

1986 §39. �

63.15 Lemma If B is an open subgroup of A, then any character

β ∈ B∗ is a restriction α|B of some character α ∈ A∗. In other words,

if γ : B → A is injective and open, then the adjoint γ∗ : A∗ → B∗ is

surjective.

Proof The divisible group T is injective; see 1.23. This means that β

can be extended to a homomorphism α : A → T, and α is continuous,because B is open in A. �

63.16 Corollary Each discrete group D has enough characters.

Proof Obviously, each cyclic subgroup of D has enough characters andis open. �

63.17 Corollary If H has a compact open subgroup K, then H has

enough characters.

Proof There is a canonical projection κ : H → D onto the discrete groupD = H/K. If cκ �= 0, then 〈cκ, α〉 = 〈c, κ∗α〉 �= 0 for some α ∈ D∗ andκ∗α ∈ H∗. If 0 �= c ∈ K, then 〈c, γ〉 �= 0 for some γ in the restrictionH∗|K . �

63.18 If A is locally compact, then the natural map δA : A → A∗∗ is

injective.

Proof This is an immediate consequence of 63.9–17. See also Stroppel

2006 Proposition 20.12. �

63.19 Lemma Every finitely generated discrete group A is reflexive.

Page 364: The Classical Fields

63 Locally compact abelian groups and Pontryagin duality 347

Proof By the fundamental theorem on finitely generated abelian groups(Hofmann–Morris 1998 Theorem A1.11 or most books on abeliangroups), A is a direct sum of cyclic groups. Each finite cyclic group isself-dual, and Z and T are character groups of each other. Thus 63.9implies the assertion. �

63.20 Theorem If A is compact, then δA is surjective, and hence A is

reflexive.

Proof If A is compact, then D = A∗ is discrete, and C = D∗ is compactagain. Because of 63.13, the group A may be considered as a subgroupof C. By the very definition, A separates the points of D: for eachx ∈ D, x �= 0, there is some a ∈ A with 〈a, x〉 �= 0. If suffices to showthat A is dense in C, and this means that each non-empty open set O

in C of the form O = W (F,U) contains an element of A. Here, F iscompact in D, hence F is finite. According to the previous lemma, F

generates a reflexive subgroup G of D. The adjoint κ : D∗ → G∗ of theinclusion map G ↪→ D is surjective by 63.15. The restriction A|G = Aκ

separates the elements of G. Assume that Aκ �= G∗. Since G∗/Aκ iscompact, 63.13 implies that there is some x ∈ G∗∗ = G with x �= 0 and〈Aκ, x〉 = 〈A, x〉 = 0, which is a contradiction. Therefore, Aκ = G∗. Ifw ∈ O, then w|G ∈ Oκ, and there is some a ∈ A such that aκ = w|G.Then 〈a, F 〉 = 〈w,F 〉 ⊆ U and a ∈ O. As A is compact and δA is acontinuous bijection, δA is a homeomorphism. �

63.21 Corollary If D is discrete, then δD is surjective, hence D is

reflexive.

Proof The group C = D∗ is compact by 63.5. If the discrete groupS = D∗∗/D is not trivial, then some element c ∈ C∗∗ = C induces anon-trivial character on S. This means that 〈c,D∗∗〉 �= 0 and 〈c,D〉 = 0,but the last condition implies c = 0, which is a contradiction. �

63.22 Annihilators If L is a subgroup of the topological abelian groupA and if Λ is a subgroup of A∗, then

L⊥ = {χ ∈ A∗ | 〈L, χ〉 = 0} and Λ⊥ = {a ∈ A | 〈a,Λ〉 = 0}are called the annihilators of L and of Λ, respectively.

63.23 Lemma Assume that K is a compact open subgroup of H,

and let κ denote the canonical projection of H onto the discrete group

Page 365: The Classical Fields

348 Appendix

D = H/K. Then (ϕ �→ κϕ) : D∗ → K⊥ is an isomorphism of topological

groups, and K⊥ is a compact open subgroup of H∗.

This can be proved by direct verification; see Hofmann–Morris 19987.13.

63.24 Lemma Assume again that K is a compact open subgroup

of H. By 63.15, the inclusion ι : K → H has a surjective adjoint

ι∗ : H∗ → K∗ with kernel K⊥. Hence each x ∈ K⊥⊥ induces a character

on H∗/K⊥ ∼= K∗. Because of 63.20, there is an element zx ∈ K such

that 〈x, ξ〉 = 〈zx, ι∗ξ〉 = 〈zxδH , ξ〉 for each ξ ∈ H∗. Consequently,

KδH = K⊥⊥. �

63.25 Surjectivity If A is locally compact, then the natural map δA

is surjective.

Proof By the Splitting Theorem, it suffices to prove surjectivity for agroup H having a compact open subgroup K. Let δ = δH denote thenatural map H → H∗∗. The last two lemmas imply that K⊥ is compactand open in H∗ and that K⊥⊥ = Kδ ≤ Hδ = G ≤ H∗∗, moreover, K⊥⊥

is compact and open in H∗∗. Thus G is an open subgroup of H∗∗.If G < H∗∗, then there is a non-trivial character ϕ : H∗∗ → T with

〈G, ϕ〉 = 0. In particular, 〈K⊥⊥, ϕ〉 = 0. Lemma 63.24, applied toK⊥ ↪→ H∗ instead of K ↪→ H, gives an element ζ ∈ K⊥ such that〈x, ϕ〉 = 〈x, ζ〉 for all x ∈ H∗∗. It follows that 〈Hδ, ζ〉 = 〈H, ζ〉 = 0,hence ζ = 0 and ϕ = 0. This contradiction shows that G = H∗∗. �

63.26 If A is locally compact, then the natural map δA : A → A∗∗ is

open.

Proof Again, it suffices to consider a group H with a compact opensubgroup K. The map δ = δH is bijective and continuous, hence therestriction δ|K is a homeomorphism onto the image Kδ = K⊥⊥, andthis image is an open subgroup of H∗∗, as we have seen in the lastproof. Since δ is a group isomorphism, it follows that δ is open in aneighbourhood of each point. �

Combined, Theorems 63.7, 63.18, 63.25, and 63.26 can be stated as

63.27 Pontryagin duality If A is a locally compact abelian group,

then A is reflexive, i.e., the natural homomorphism δA : A → A∗∗ is a

homeomorphism and hence also a group isomorphism. Therefore, A∗∗

and A can be identified via δA; the groups A and A∗ are then duals of

each other.

Page 366: The Classical Fields

63 Locally compact abelian groups and Pontryagin duality 349

Note One should keep in mind that for the crucial steps 63.6, 63.7,63.13, and 63.14 no proofs have been given, only convenient references.See also Stroppel 2006 Theorem 22.6.

63.28 Evaluation If A is locally compact, then (x, ξ) �→ 〈x, ξ〉 is a

continuous map ω : A×A∗ → T.

Proof Only continuity of ω at (0, 0) has to be shown. Let V be a compactneighbourhood of 0 in A, and put Ω = W (V,U). Then 〈V, Ω〉 ⊆ U bythe very definition of Ω. �

Applications

63.29 Proposition If C is a connected subgroup of A and Γ is a

compact subgroup of A∗, then 〈C, Γ〉 = 0.

Proof Let ι : Γ ↪→ A∗ denote the inclusion. Its adjoint ι∗ : A → Γ∗ is acontinuous homomorphism into a discrete group, and Cι∗ is a connectedsubgroup of Γ∗. Hence Cι∗ = 0 and 〈C, ιΓ〉 = 〈Cι∗, Γ〉 = 0. �

63.30 Theorem Let A be an LCA group. Then A is connected if, and

only if, A∗ has no compact subgroups other than {0}.Proof If A is connected and Γ is a compact subgroup of A∗, then 〈A, Γ〉 =0 by the last proposition, and this means that Γ = 0.

For the converse, let C denote the connected component of A. By62.13 and Exercise 3 of Section 62, the factor group A/C is 0-dimensionaland has arbitrarily small compact open subgroups S/C, where C ≤ S

and S is open in A. If A is not connected, then A/C �= 0 and A/S �= 0for a suitable choice of S. Consider the canonical projection κ of A ontothe discrete group A/S = D. Its adjoint κ∗ : D∗ → A∗ is injective (63.3)and maps the compact group D∗ homeomorphically onto a non-trivialsubgroup of A∗. This proves the converse. �

An abelian group A is divisible (resp. torsion free) if each homomor-phism A→ A : x �→ n · x with n ∈ N is surjective (resp. injective).

63.31 Proposition If A is divisible, then A∗ is torsion free.

This is an immediate consequence of 63.3 and the identity 〈n · a, χ〉 =〈a, n·χ〉. The converse does not hold in general; see Hewitt–Ross 196324.44. The following is true, however.

Page 367: The Classical Fields

350 Appendix

63.32 Theorem If A is a torsion free abelian group, and if A is compact

or discrete, then A∗ is divisible.

Proof Put Λ = n · A∗. By 63.5, the group A∗ is discrete or compact.Hence Λ is closed in A∗. We have Λ⊥ = {a ∈ A | 〈n · a,A∗〉 = 0} = 0,since A is torsion free. The adjoint of the quotient map A∗ → A∗/Λ

yields an injection (A∗/Λ)∗ → Λ⊥ = 0, and n ·A∗ = A∗.If A is discrete, the proof follows also immediately from 63.15. �

Exercises(1) Let 0 → A → B → C → 0 be a proper exact sequence of LCA groups (i.e.,A → B is an embedding and B → C is a quotient morphism with kernel A).Then C∗ ∼= A⊥ ≤ B∗ and the adjoint sequence 0 → C∗ → B∗ → A∗ → 0 isalso exact and proper.

(2) Let D denote the group (R, +), taken with the discrete topology. Showthat the adjoint ε∗ of the identity mapping ε : D → R is not surjective. Infact, card R∗ < card D∗.

(3) Show that a compact abelian group A is connected if, and only if, it isdivisible.

64 Fields

This section deals with fields and field extensions. Most proofs are easilyaccessible in the literature, there are many algebra books with chapterson field theory. We mention Cohn 2003a and Jacobson 1985, 1989;see also Morandi 1996, Lang 1993, Bourbaki 1990.

64.1 Definition A field is a set F with two binary operations, addition+ and multiplication, such that the following axioms hold.(a) (F,+) is a commutative group with neutral element 0.(b) F× = F �{0} is a commutative group with respect to multiplication

(with neutral element 1).(c) Addition and multiplication are related by the distributive law

(a + b)c = ac + bc.

Let F be a subfield of a field E; then we speak of the field extensionE|F . Clearly E is a vector space over F ; we call the dimension dimF E

of this vector space the degree of E|F and denote it by [E : F ]. Anextension E|F is called finite, if its degree [E : F ] is finite. The followingsimple observation is very useful.

Page 368: The Classical Fields

64 Fields 351

64.2 Degree formula Let D|E and E|F be field extensions. Then

[D : F ] = [D : E][E : F ].

64.3 Quotients of rings Let M be an ideal of a commutative ring R.

Then the quotient ring R/M is a field if, and only if, M is a maximal

ideal in R.

Proof If R/M is a field, then the trivial ideal is maximal in R/M , henceM is maximal in R. Conversely, let M be maximal. If a ∈ R � M , thenthe ideal aR + M coincides with R. Thus there exists an element b ∈ R

with 1 ∈ ab + M , and this means that b + M is the inverse of a + M inR/M . �

64.4 Prime fields Each field F has a smallest subfield P , which isthe intersection of all subfields of F . One calls P the prime field of F .If P ∼= Q, then F is said to have characteristic zero, and one writeschar F = 0. Otherwise P ∼= Fp = Z/pZ for some prime number p, andthen F has the characteristic char F = p.

64.5 Definitions Let E|F be a field extension. An element a ∈ E

is called algebraic over F , if f(a) = 0 for some polynomial f �= 0 withcoefficients in F . By F [x] (or F [t]) we denote the polynomial ring over F

(compare 64.22), and we call a polynomial monic if its leading coefficientis 1. The monic polynomial f ∈ F [x] of minimal degree with f(a) = 0is uniquely determined; it is called the minimal polynomial of a over F .Such a minimal polynomial is irreducible in F [x].

The extension E|F is called algebraic (and the field E is called alge-braic over F ), if each element of E is algebraic over F .

Each algebraic extension E|F satisfies card E ≤ max{ℵ0, cardF}; thisis a consequence of 61.14: we have card F [x] = max{ℵ0, cardF}, andeach polynomial has only finitely many roots.

For a ∈ E we denote by F [a] the subring generated by F ∪ {a}, andF (a) denotes the subfield generated by F ∪ {a} (which is the field offractions of F [a]). For A ⊆ E, the ring F [A] and the field F (A) aredefined analogously.

64.6 Algebraic elements Let E|F be a field extension. An element

a ∈ E is algebraic over F if, and only if, F [a] has finite dimension n as

a vector space over F .

In this case, F [a] = F (a) is a field, the minimal polynomial g ∈ F [x]of a has degree n, the evaluation map F [x] → F (a) : f �→ f(a) is a

surjective ring homomorphism, and its kernel is the ideal (g) := gF [x].

Page 369: The Classical Fields

352 Appendix

Proof F [a] is the vector space over F generated by 1, a, a2, . . . . Thisvector space has finite dimension if, and only if, the elements 1, a, . . . , ak

are linearly dependent for some k ∈ N, and this means that a is algebraicover F .

If this holds, then g is the monic polynomial of minimal degree inthe ideal I = {f ∈ F [x] | f(a) = 0}. Since F [x] is a principal idealdomain, we infer that I = (g). The vector space F [x]/I has the basis{xi + I | 0 ≤ i < deg g }. Now F [x]/I ∼= F [a], hence deg g = n.

Since g is irreducible, I = (g) is a maximal ideal in the principalideal domain F [x]. By 64.3, the quotient F [x]/I ∼= F [a] is a field, henceF [a] = F (a). �

64.7 Transitivity If D|E and E|F are algebraic field extensions, then

D|F is algebraic.

Proof Let a ∈ D. Then an +∑n−1

i=0 ciai = 0 for suitable elements ci ∈ E

and n ∈ N. By 64.2, the field C = F (c0, c1, . . . , cn−1) is a finite extensionof F , hence [F (a) : F ] ≤ [C(a) : F ] = [C(a) : C][C : F ] ≤ n · [C : F ] isfinite. �

64.8 Splitting fields Let f ∈ F [x] be a non-zero polynomial. Choosean irreducible factor g of f . By 64.3, E = F [x]/gF [x] is a field, and weconsider F as a subfield of E by identifying c ∈ F with c + gF [x] ∈ E.Then f has a root in E, namely x + gF [x]. (For example, f = x2 + 1has the root x + fR[x] in C = R[x]/fR[x].)

Repeated application of this construction gives a finite extension E|Fsuch that f is a product of linear factors in E[x]. If E is minimal withthis property, i.e., if E is generated by F together with all roots of f ,then E is called a splitting field of f over F . The splitting field of f overF is uniquely determined up to isomorphism; see Cohn 2003a 7.2.3,Jacobson 1985 Section 4.3, Lang 1993 V.3.1. More generally, thereexists a splitting field for any set of non-zero polynomials over F .

Let q be a power of a prime p. A finite field F with cardF = q isa splitting field of the polynomial xq − x over Fp = Z/pZ (in fact, F

consists just of all roots of xq − x). Thus a finite field is determined upto isomorphism by its cardinality.

64.9 Definition An algebraic field extension E|F is called normal , ifeach irreducible polynomial f ∈ F [x] that has some root in E splits inE[x] into a product of linear factors.

Page 370: The Classical Fields

64 Fields 353

64.10 Theorem An algebraic extension E|F is normal precisely if E

is a splitting field of some set of non-zero polynomials over F .

Indeed, if E|F is normal, then E is a splitting field of the set of minimalpolynomials of elements in E. For the converse see Cohn 2003a 7.2.4,Jacobson 1989 Theorem 8.17 or Lang 1993 V.3.3.

64.11 Definition A polynomial is said to be separable if all its roots(in some splitting field) are simple. Let E|F be a field extension. Anelement a ∈ E is called separable over F , if a is algebraic over F andthe minimal polynomial of a over F is separable. The extension E|F iscalled separable if every element of E is separable over F .

Each multiple root of a polynomial f ∈ F [x] is also a root of thederivative f ′. If f is irreducible and charF = 0, then f ′ �= 0 cannotdivide f , hence f ′ and f have no common factor. This shows that eachalgebraic extension of fields with characteristic 0 is separable.

64.12 Primitive elements Let E|F be a finite separable extension.

Then E|F is simple, i.e., E = F (a) for a suitable ‘primitive’ element

a ∈ E.

This is proved in Cohn 2003a 7.9.2, Jacobson 1985 Section 4.14 andLang 1993 V.4.6.

64.13 Definition A field F is called algebraically closed , if F has noproper algebraic extension; this means that each irreducible polynomialin F [x] has degree 1, or equivalently, that each non-zero polynomialin F [x] has a root in F . An extension field F � of a field F is called analgebraic closure of F , if F �|F is algebraic and F � is algebraically closed.

64.14 Theorem Every field F has an algebraic closure F �.

For the proof, one uses Zorn’s Lemma (or some equivalent principle)to construct a maximal algebraic extension of F ; such an extension is al-gebraically closed by 64.7. See Jacobson 1989 Section 8.1 or Morandi

1996 3.14 for the details, and Cohn 2003a 7.3.4, 11.8.3 for other con-struction methods. These references show moreover that the algebraicclosure F � of a field F is uniquely determined up to isomorphism. Thisfact is also a consequence of the following result.

64.15 Theorem Let F � be an algebraic closure of the field F , let

α : F → F be a field automorphism of F , and let E|F be an algebraic

extension. Then α has an extension to a monomorphism E → F � of

fields. If E is algebraically closed, then any such extension is an isomor-

phism of E onto F �.

Page 371: The Classical Fields

354 Appendix

For a proof see Lang 1993 V.2.8, Bourbaki 1990 V §4 or Morandi

1996 3.20, 3.22. Using (an equivalent of) Zorn’s Lemma in the proofsof 64.14 and 64.15 is unavoidable, because some models of ZF set the-ory contain fields F without containing an algebraic closure of F ; seeLauchli 1962.

64.16 Cardinality of the algebraic closure We have cardF � =max{ℵ0, cardF}.

For the proof see the inequality in 64.5 and note that a finite field E

is not algebraically closed (as 1 +∏

a∈E(x− a) has no root in E).

64.17 Definitions Let E|F be a field extension and let Γ ≤ AutE bea group of automorphisms of E. Then

AutF E := {α ∈ AutE | α(a) = a for all a ∈ F }is a subgroup of AutE, and FixEΓ := {a ∈ E | α(a) = a for all α ∈ Γ}is a subfield of E.

An algebraic extension E|F is said to be a Galois extension if F =FixE AutF E, and then GalF E := AutF E is called the Galois group ofE|F . An algebraic extension E|F is a Galois extension if, and only if,E|F is normal and separable; see Cohn 2003a 7.6.1, 11.8.4.

64.18 Main Theorem of Galois Theory Let E|F be a finite Galois

extension and Γ = GalF E. Then |Γ| = [E : F ], and the following holds.

(i) The map Δ �→ FixEΔ is a lattice anti-isomorphism of the subgroup

lattice of Γ onto the lattice {D | F ≤ D ≤ E } of all fields between

F and E.

(ii) Let Δ ≤ Γ and D = FixEΔ. Then E|D is a Galois extension with

Galois group Δ, and the degree [D : F ] is the index of Δ in Γ.

The extension D|F is a Galois extension precisely if Δ is a normal

subgroup of Γ, and then GalF D ∼= Γ/Δ.

64.19 Transcendental elements Let E|F be a field extension. Anelement t ∈ E is said to be transcendental over F , if t is not algebraicover F . In this case, F [t] is isomorphic to the polynomial ring F [x](via evaluation as in 64.6), and F (t) is isomorphic to the field F (x) ofrational functions (which is the field of fractions of F [x]).

Such a field F (t) has many automorphisms: each of the substitutionst �→ t−1 and t �→ at + b with a ∈ F×, b ∈ F yields a unique auto-morphism of F (t) fixing all elements of F . Hence AutF F (t) containsthe group PGL2F (which is generated by these substitutions). In fact,

Page 372: The Classical Fields

64 Fields 355

AutF F (t) = PGL2F by a result related to Luroth’s Theorem; see Cohn

2003a 11.3.3 or Jacobson 1989 Section 8.14.We claim that [F (t) : F ] = max{ℵ0, cardF} = cardF (t) if t is trans-

cendental over F . Indeed, the elements (t−a)−1 with a ∈ F are linearlyindependent over F by Exercise 5 below; the other necessary estimatesare obvious.

64.20 Algebraic independence and transcendency bases LetE|F be a field extension. A subset T ⊆ E is called algebraically inde-pendent over F , if each t ∈ T is transcendental over F (T � {t}). Thenthe field F (T ) consists of all fractions of polynomials in finitely manyvariables taken from T ; such a field F (T ) is called a purely transcendentalextension of F .

If T �= ∅ is algebraically independent over F , then

[F (T ) : F ] = max{ℵ0, cardF, cardT} = cardF (T ) .

Indeed, if T is finite, then the results 64.2, 64.19 and 64.13b imply that[F (T ) : F ] = max{ℵ0, cardF}card T = max{ℵ0, cardF}; now the firstequation for [F (T ) : F ] follows from this in view of 61.13a, 61.14 andF (T ) =

⋃{F (X) | X ⊆ T,X is finite}. Again by 61.13b and 61.14 wehave cardF (T ) = max{cardF, [F (T ) : F ]} = [F (T ) : T ].

By Zorn’s Lemma, there exists a maximal set T ⊆ E which is alge-braically independent over F , a so-called transcendency basis of E|F ;then E is algebraic over F (T ). By a result due to Steinitz 1910, anytwo transcendency bases of E over F have the same cardinality; seeCohn 2003a 11.2.1, Jacobson 1989 Section 8.12, Morandi 1996 19.15or Bourbaki 1990 V §14 Theorem 3. This cardinality is called thetranscendency degree trdeg(E|F ) of E|F .

For every infinite field F , we have cardF = max{ℵ0, trdeg(F |P )},where P is the prime field of F .

64.21 Theorem (Steinitz) Two algebraically closed fields F, F ′ with

prime fields P, P ′ are isomorphic if, and only if, charF = charF ′ and

trdeg(F |P ) = trdeg(F ′|P ′).

Such a field F is an algebraic closure of P (T ) for each transcendencybasis T of F over P ; hence 64.21 follows from the fact that trdeg(F |P )is well-defined (64.20). Note that trdeg(F |P ) = [F : P ] = cardF if F isuncountable.

Each permutation of T induces an automorphism of P (T ), which ex-tends to F by 64.15. This shows that algebraically closed fields have

Page 373: The Classical Fields

356 Appendix

many automorphisms (see 52.10 for Q� and F �p). See also Exercise 4 and

Theorem 14.11.

Now we define some rings and fields which consist of various types offormal power series.

64.22 Power series and polynomials Let F be a field. We endowthe set F N0 of all maps (sequences) f : N0 → F with pointwise additionand with the convolution product fg defined by

(fg)(n) =∑n

i=0 f(i)g(n− i) .

Then F N0 is a commutative ring (and an F -algebra) with the unit ele-ment δ0 defined by δ0(0) = 1 and δ0(n) = 0 for n �= 0.

We claim that each element f ∈ F N0 with f(0) �= 0 is invertiblein this ring. Indeed, the map g defined recursively by g(0) = f(0)−1

and g(n) = −f(0)−1∑n

i=1 f(i)g(n − i) for n > 0 satisfies the equationfg = δ0.

Now we switch to the usual notation. For n ∈ N0 let δn ∈ F N0 bedefined by δn(n) = 1 and δn(m) = 0 for n �= m. The special elementt := δ1 is called an indeterminate, it satisfies tn = δn for n ∈ N0.The F -subalgebra of FN0 generated by t is the polynomial ring F [t] ={∑n

i=0 aiti | n ∈ N0, ai ∈ F }; it consists of those elements f ∈ F N0

which have finite support supp(f) := {n ∈ N0 | f(n) �= 0}, and deg F :=max supp(f) is the degree of a polynomial f �= 0.

Often an arbitrary element f ∈ F N0 is written formally as the powerseries f =

∑n≥0 f(n)tn, and the ring F N0 is written as the (formal)

power series ring

F [[t]] ={∑

n≥0 antn∣∣ an ∈ F

}.

The group of units of F [[t]] is the direct product of F× and of themultiplicative group 1 + tF [[t]].

The notation of these ring elements as series has a topological mean-ing: in the product topology of F [[t]] = F N0 obtained from the discretetopology on F , the series f is the limit of the polynomials

∑Nn=0 f(n)tn

with N ∈ N. Moreover, F [[t]] is the completion of the polynomial ringF [t] with respect to this topology; see 44.11. �

64.23 Laurent series Let F be a field. We denote by

L ={

f ∈ F Z∣∣ ∃k∈Z ∀n<k f(n) = 0

}the set of all mappings f : Z → F such that the support supp(f) =

Page 374: The Classical Fields

64 Fields 357

{n ∈ Z | f(n) �= 0} is bounded below. For 0 �= f ∈ L we define v(f) =min supp(f) ∈ Z, and v(0) :=∞. This set L is a commutative ring withpointwise addition and with the convolution product fg defined by thefinite sums

(fg)(n) =∑ {f(i)g(j) | i + j = n, i ≥ v(f), j ≥ v(g)} .

We claim that L is in fact a field. The mapping v : L → Z ∪ {∞}satisfies v(fg) = v(f)+ v(g) and v(f + g) ≥ min{v(f), v(g)} for f, g ∈ L

(compare 56.1). Therefore L0 := {f ∈ L | v(f) ≥ 0} is a subring ofL, and we identify L0 with the power series ring F [[t]] from 64.22 viathe isomorphism given by the restriction f �→ f |N0 . In particular, weconsider t = δ1 as an element of L with v(t) = 1. Each non-zero elementf ∈ L can be written as f = tng, where n = v(f) and g ∈ L withv(g) = 0. The element δ−1 ∈ L (defined by δ−1(−1) = 1 and δ−1(z) = 0for z �= −1) is the inverse of t = δ1, and g is invertible in L0 = F [[t]]by 64.22. Hence f is invertible in L, and L is a field. In fact, we haveshown that L is the field of fractions of the power series ring L0 = F [[t]].We remark that v is a valuation of L with valuation ring L0 and residuefield F ; see Section 56.

Now we switch to the usual notation. We have L =⋃

n≥0 t−nL0, hencewe write each f ∈ L formally as the Laurent series f =

∑n≥v(f) f(n)tn

(compare 64.22), and we write the field L as the (formal) Laurent seriesfield

F ((t)) := L ={∑

n≥z antn∣∣ z ∈ Z, an ∈ F

}.

In this notation, the multiplicative group of F ((t)) has the direct de-composition F ((t))× = F× × tZ × (1 + tF [[t]]).

Again the series notation has a topological meaning: with respect tothe valuation topology induced by v (see 13.2b), f is the limit of theLaurent polynomials

∑Nn=v(f) f(n)tn with v(f) ≤ N ∈ Z; note that F

is discrete in this topology. Moreover, F ((t)) is the completion withrespect to this valuation topology of the field F (t) of rational functions;see 44.11. A Laurent series f belongs to F (t) precisely if the coefficientsf(n) finally satisfy a linear recurrence relation; see Ribenboim 19993.1.N p. 92.

64.24 Puiseux series Let F be a field. We define the set P ⊆ F Q by

P =⋃

n∈N Pn, where Pn ={

f ∈ F Q∣∣ ∃k∈N supp(f) ⊆ n−1(N− k)

}.

Thus P consists of all maps f : Q → F with supports which are con-tained in a cyclic subgroup n−1Z of (Q, +) and bounded below. We

Page 375: The Classical Fields

358 Appendix

endow P with pointwise addition and with the convolution product fg,where (fg)(r) with r ∈ Q is defined as in 64.23, except that now the sum-mation extends over the finitely many pairs (i, j) ∈ supp(f) × supp(g)with i + j = r.

The bijection Pn → F ((t)) : f �→ f defined by f(z) = f(z/n) forz ∈ Z is a ring isomorphism, hence Pn is a field by 64.23. Thereforealso P =

⋃n Pn! is a field. We remark that the map v : P → Q ∪ {∞}

defined as in 64.23 is a valuation of P with value group Q and residuefield F .

Similarly as in 64.22, we define δq ∈ P for q ∈ Q by δq(q) = 1 andδq(x) = 0 for q �= x ∈ Q; then δqδr = δq+r for q, r ∈ Q. We writet := δ1 and tq := δq; then (t1/n)n = t for n ∈ N. The isomorphismPn → F ((t)) described above maps t1/n ∈ Pn to t ∈ F ((t)); we identifyvia this isomorphism and write F ((t1/n)) for Pn, and F [[t1/n]] for thevaluation ring {f ∈ Pn | v(f) ≥ 0} of Pn. Moreover we write

F ((t1/∞)) := P =⋃

n∈N F ((t1/n)) .

By 64.23 each element f ∈ Pn = F ((t1/n)) can be written as f =∑i≥N ait

i/n with N ∈ Z and ai ∈ F (these infinite sums converge inthe valuation topology defined by v). Such a series is called a Puiseuxseries, and F ((t1/∞)) is called the Puiseux series field over F .

The multiplicative group of F ((t1/∞)) has the direct decompositionF ((t1/∞))× = F× × tQ × {1 + f | v(f) > 0}, and {1 + f | v(f) > 0} =⋃

n>0(1 + t1/nF [[t1/n]]).Let F be an algebraically closed of characteristic 0. Then the Newton–

Puiseux theorem says that F ((t1/∞)) is an algebraic closure of the Lau-rent series field F ((t)); see, for example, Serre 1979 IV Proposition 8p. 68, Ribenboim 1999 7.1.A p. 186 or Volklein 1996 Theorem 2.4.

This implies that C((t1/∞)) is isomorphic to C; see 64.21 and ob-serve that trdeg(C((t1/∞))|Q) ≤ card C((t1/∞)) ≤ card CQ = card C =trdeg(C|Q); compare also 64.25 and 14.9.

64.25 Hahn power series Let F be a field and let Γ be an orderedabelian group. On the set

F ((Γ)) :={

f ∈ F Γ∣∣ supp(f) is a well-ordered subset of Γ

}one can define an addition and a multiplication as in 64.24, and this turnsF ((Γ)) into field, the field of Hahn power series over Γ. The elements ofF ((Γ)) with finite support form a subring, the group ring F [Γ], and the

Page 376: The Classical Fields

64 Fields 359

elements of F ((Γ)) with countable support form the subfield

F ((Γ))1 := {f ∈ F ((Γ)) | card supp(f) ≤ ℵ0 } ;

see Dales–Woodin 1996 2.7 and 2.15, Prieß-Crampe 1983 II §5,Shell 1990 Appendix B.7, Neumann 1949a or Engler–Prestel 2005Exercise 3.5.6 for details, and Ribenboim 1992 for generalizations.

As in 64.24, one has the valuation v : F ((Γ)) → Γ ∪ {∞} defined byv(f) = min supp(f) for f �= 0 and v(0) = ∞, with value group Γ andresidue field F .

Such a field F ((Γ)) or F ((Γ))1 is algebraically closed if, and only if,F is algebraically closed and Γ is divisible; see Dales–Woodin 19962.15, Ribenboim 1992 5.2 or Prieß-Crampe 1983 p. 52 Satz 6. Thisimplies that C((Q)) ∼= C ∼= C((Q))1; see 64.21.

Using 12.10 we infer that the fields R((Γ)) and R((Γ))1 are real closedfor each divisible ordered abelian group Γ. The unique ordering of thesefields is described by f > 0⇔ f(v(f)) > 0.

Exercises(1) Show that each finite subgroup G of the multiplicative group of a field iscyclic. In particular, the multiplicative group of every finite field is cyclic.

(2) Let E|F be a Galois extension with Galois group Γ. Then the minimalpolynomial of a ∈ E over F is the product

Q

b∈B(x − b), where B = {γ(a) |γ ∈ Γ} is the orbit of a under Γ.

(3) Let E be a splitting field of the polynomial f = x5 −16x+2 over Q. Showthat GalQE induces the full symmetric group on the five roots of f .

(4) Every algebraically closed field F has precisely 2card F field automorphisms.

(5) If t is transcendental over the field F , then the rational functions (t−a)−1

with a ∈ F are linearly independent over F .

(6) Let q = pn be a power of the prime p. Show that the automorphism groupof the finite field Fq of cardinality q is the cyclic group of order n which isgenerated by the Frobenius automorphism a �→ ap.

Page 377: The Classical Fields

Hints and solutions

1 The additive group of real numbers

(1) Every element x ∈ B′ is a linear combination of some finite set Bx ⊆ B.Note that B =

S

x∈B′ Bx and apply 61.13a.

(2) If a ∈ A has order pkr, where r is not divisible by the prime p, thenupk + vr = 1 for suitable integers u, v (see 1.5), and a = vr · a + upk · a. Thefirst summand has order pk, the order of the other summand is prime to p.The assertion follows by induction.

(3) We have R+×Zn < R+×Qn ∼= R+ by 1.17. The group R+ is divisible,but Zn and hence R+×Zn are not.

(4) If B is a basis of R over Q, then each map λ : B → Q defines a linear form.The kernel H of λ is a hyperplane, and λ is determined by H up to a scalarin Q×. Hence ℵ0 · cardH = card QB = ℵ ℵ

0 = 2ℵ by 1.12 and 61.15.

(5) 12(C+C) = {P∞

ν=1 cν3−ν | cν ∈ {0, 1, 2}} = [0, 1] generates R+ as a group.

(6) See 35.10.

(7) If M is a maximal subgroup of F+, then the quotient F/M is an abeliangroup without non-trivial proper subgroups, hence cyclic of prime order p.The divisibility of F+ implies that F = pF ⊆ M , a contradiction.

(8) Each such subgroup of R+ is isomorphic to Z2. Any automorphism of oddorder of Z2 is given by an integer matrix with determinant 1, since Aut Z2 =GL2Z. If X ∈ SL2R has trace t and finite order n > 1, then X2 = tX − Eand X3 = (t2 − 1)X − tE. Each complex eigenvalue λ of X satisfies |λ| = 1.This implies |t| ≤ 2. In the case t = 2, induction shows Xn = nX − (n− 1)E,and Xn = E implies X = E. If t = −2, then −X = E by the last argument.|t| = 1 yields X3 = ±E, and t = 0 gives X2 = −E. In particular, n = 5.

Alternative solution: Every automorphism of Z2 is described by an integer2× 2 matrix M . If M has order 5, then 1 is not an eigenvalue of M , hence Mis annihilated by the cyclotomic polynomial Φ5 = (x5 − 1)/(x− 1). Now Φ5 isirreducible over Q (apply Eisenstein’s criterion, which is due to Schonemann1846, to Φ5(x + 1)), hence Φ5 divides the characteristic polynomial of M ,which has degree 2, a contradiction.

(9) By 1.17, R/Q ∼= R.

360

Page 378: The Classical Fields

Hints and solutions 361

2 The multiplication of real numbers, with a digression on fields

(1) As Q(T ) =S{Q(S) | S ⊆ T ∧ card S < ℵ0 }, we may assume that T is

finite. Since Q(S∪{t}) = Q(S)(t), the claim can then be proved by induction.If F is any ordered field, then there is a unique ordering for the polynomialring F [t] such that F < t. This ordering yields the desired ordering of thefield of fractions F (t).

(2) Every involution of the rational vector space R yields such an extension.An isomorphism between two extensions maps the copies of R onto each other,since R+ has no subgroup of index 2. Involutions defining isomorphic exten-sions have eigenspaces of the same dimension; hence we obtain infinitely manyextensions. If an extension of R+ by C2 is commutative, then it is isomorphicto R ⊕ C2 by 1.23 and 1.22.

3 The real numbers as an ordered set

(1) The chain R is order complete, but neither strongly dense in itself norstrongly separable.

(2) If ϕ : R2lex → R< is injective and order-preserving, then each set {t} × R

is mapped onto a connected set, hence onto an interval, but any family ofpairwise disjoint open intervals in R is at most countable (Souslin’s condition).

4 Continued fractions

(1) Consider the continued fraction [ 1; 1, 1, 1, . . . ] for the golden ratio γ =

(√

5 + 1)/2. The approximating fractions have the form fν+1/fν with fν+1 =fν + fν−1; the fν = 1, 1, 2, 3, 5, 8, . . . are the Fibonacci numbers. The quo-tients fν+1/fν converge to γ > 8/5, and (8/5)2 > 5/2. Therefore, f6/f5 =13/8 > 8/5, f4+ν ≥ 5 · (8/5)ν and f2+νf3+ν ≥ 6 · (5/2)ν . By induction,qν = cνqν−1 + qν−2 ≥ fν , and the claim follows.

(2) Use pν = cνpν−1 + pν−2 and qν = cνqν−1 + qν−2.

(3) Let ν be even. Then pν/qν < ζ < pν+1/qν+1 and pν+1/qν+1 − pν/qν =(qνqν+1)

−1 < 12(q−2

ν + q −2ν+1), since the geometric mean is smaller than the

arithmetic mean.

5 The real numbers as a topological space

(1) A is open (closed) in X and in Y if there are open (closed) subsets U, Vin Z such that A = U ∩ X = V ∩ Y . Put U ∩ V = W . It follows thatA = W ∩ X = W ∩ Y = W ∩ (X ∪ Y ) = W ∩ Z = W is open (closed) in Z.

(2) The set Q is dense in R, and R is separable. Each interval of the form]a, c[= [a, c] is open and closed. Hence R is not connected and does not have acountable basis; in particular, R is not metrizable. Being orderable, the spaceR is normal, and order-completeness implies that R is locally compact.

(3) (a) Obviously X = Xiii is a Hausdorff space and Q is dense in X. If X isa disjoint union of two proper open subsets, then there is also a partition ofX into open subsets A and B such that A has a least upper bound s. As inthe case of the ordinary reals, s /∈ A and each interval ]u, s] contains elementsfrom A ∩ Q. This contradiction shows that X and the homeomorphic spaces] , x[ and ]x, [ are connected, but X is not locally connected because eachopen interval in Q is open in X.

Page 379: The Classical Fields

362 Hints and solutions

(b) The remaining open sets form a topology on X = Xii and Q is densein X. By definition, the complement of {0} is not open. Suppose that X ispartitioned into two open subsets A and B such that 1 ∈ B. Since [1, [ isconnected, one has ] ,−1] ⊆ A and [1, [ ⊆ B, but then 0 ∈ A∩B. Therefore,X is connected. Similarly, X is locally connected and each point separates Xinto two connected subsets.

(c) A non-empty intersection of two basic open sets is another one. HenceX = Xiv is a topological (Hausdorff) space. The countable set N× `

[0, 1[∩Q´

is dense in X and X is separable. Being a union of the connected space N andconnected ‘bristles’ each of which meets N the space X is connected. Since Nis locally connected, X has a basis of connected open sets. By 5.22, each point(n, 0) has a connected complement in N × {0}. Therefore, condition (iv′) isviolated.

(4) Any interval with rational endpoints is open and closed in I. Hence anypermutation of a finite number of pairwise disjoint intervals with rationalendpoints is induced by a homeomorphism of I, and n distinct points can bemapped by a homeomorphism to n other points whenever the two sets aredisjoint.

(5) T1000... and T0111... define the same vertex of T . A similar remark applieswhenever two binary expansions represent the same number. The diameterof Tc|ν converges to 0 as ν increases. Hence ϕ is well-defined and continuous.Obviously, ϕ is surjective. This curve ϕ is named after Polya 1913. Accordingto Lax 1973, the map ϕ is nowhere differentiable; for a triangle with sides oflength 9, 40, 41 however, the analogous construction yields a map ψ suchthat ψ′(t) = 0 almost everywhere; see also Prachar–Sagan 1996. More onspace-filling curves can be found in Sagan 1994.

(6) A typical example is S = {x ∈ R | x = 0 ∨ |x| > 1}.(7) By 4.11, the chain I of irrational numbers admits a complete metric.

(8) Assume that ϕ : L → L is continuous and increasing. Put c0 = 0 andcκ+1 = ϕ(cκ). The set { cκ | κ < ω } is bounded. Hence the cκ converge tosome c ∈ L and ϕ(c) = c.

(9) Each finitely generated Boolean algebra is finite and isomorphic to thepower set 2N of the finite set N of atoms (compare Cohn 2003a 3.4.5 orKoppelberg 1989 Corollary 2.8 p. 30).

List the elements of the given algebras A = {0, 1, a1, a2, . . . } and B ={0, 1, b1, b2, . . . }. Then there is an isomorphism ϕ of the algebra An generatedby a1, a2, . . . , an into B. Choose the first element b in B which is not in ϕ(An).Since the algebras are supposed to have no atoms, each interval of An containsfurther elements of A. Therefore there is a first element a ∈ A such ϕ has anextension ϕ′ : 〈An, a〉 → B with ϕ′(a) = b. Now interchange the roles of Aand B and proceed in the same way, so that both sequences will be exhausted.

Compare also Hodges 1993 p. 100.

(10) If q ∈ Q2, then each straight line through q with rational slope r = 0is contained in S. If u, v ∈ I, then there are increasing sequences of rationalnumbers aν converging to u and bν converging to v with decreasing slopes(bν+1 − bν)/(aν+1 − aν). Hence there is an increasing continuous map ϕ :[0, 1] → S (which is piecewise linear on each interval [0, �] with � < 1) suchthat ϕ(0) = (a1, b1) and ϕ(1) = (u, v).

Page 380: The Classical Fields

Hints and solutions 363

6 The real numbers as a field

(1) Well-order a maximal set T of algebraically independent elements of R.Note that any extension of a countable field by a countable set is itself count-able by 61.14. Therefore, T is uncountable. Choose tν as the first element inT such that tν /∈ Q({ tκ | κ < ν }) = Fν . Then the Fν form a chain of subfieldsof R of length at least Ω = ω1 (compare 61.7).

(2) Clearly Z[[x]]� is a subring of Z[[x]]. If f =P

n fnxn ∈ Z[[x]]� and |fn| ≤Cnk for all n ∈ N, then the series for f(1/2) converges absolutely in R. Thus

the evaluation f �→ f(1/2) is a surjective ring homomorphism Z[[x]]� → R.The ideal generated by 1 − 2x is contained in the kernel of this evaluation.Let f(1/2) = 0 and define g ∈ Z[[x]] by g = f · (1 − 2x)−1 = f · P

n 2nxn; weshow now that the coefficients gn of g are bounded by some polynomial in n.

We have gn =Pn

j=0 fj2n−j = 2n(f(1/2)−P

j>n fj2−j) = −2n P

j>n fj2−j ,

thus |gn| ≤ CP2n

j=n+1 jk2n−j +CP

j>2n jk2n−j ≤ C(2n)k +CP

j>2n jk2n−j .

Hence it suffices to show thatP

j>2n jk2n−j =P

j≥1(2n+j)k2−n−j is boundedabove by a constant which depends only on k and not on n.

The function h(x) := (2x + j)k2−x with fixed values j, k ∈ N is decreasingfor x ≥ x0 := k/(ln 2) − j/2 and increasing for 1 ≤ x ≤ x0, hence we have

(2n + j)k2−n ≤ h(x0) = (2k/(ln 2))k2−x0 ≤ (4k)k2j/2 for j, k, n ∈ N, in

view of ln 2 > 1/2. This impliesP

j≥1(2n + j)k2−n−j ≤ (4k)kP

j≥1 2−j/2 =

(4k)k/(√

2 − 1), which does not depend on n.

7 The real numbers as an ordered group

(1) Monotonicity being obvious, R2lex is an ordered group. It is not Archime-

dean, since n · (0, 1) = (0, n) < (1, 0).

(2) Assume that ϕ : A → Aut< B : a �→ ϕa is a homomorphism into thegroup of order-preserving automorphisms of B. Then the set A × B with thelexicographic ordering and the operation (a, b)⊕(c, d) =

`

a+c, b+ϕa(d)´

is an

ordered group, because d < d′ implies ϕa(d) < ϕa(d′). It is not Archimedeanfor the same reason as in Exercise 1.

(3) The additive group Z log 2 + Z log 3 is dense in R, since log 3/ log 2 /∈ Q;see 2.2, 1.6 and 35.10.

(4) By 7.12, the automorphisms α in question are the maps α = ϕr where r isa positive real number r such that rA = A. In particular, α(1) = r = m + ncand α(c) = rc = mc + nc2 = p + qc with m, n, p, q ∈ Z. If c is not quadraticover Q, then n = 0 and α = ϕr = ϕm, hence m = 1 and α = id.

Let c be quadratic over Q. Then ϕr is an endomorphism of the orderedgroup A precisely if 0 < r ∈ A and rc ∈ A, i.e., for 0 < r ∈ A ∩ Ac−1. Wehave A ∩ Ac−1 = (Z ⊕ Zc) ∩ (Z ⊕ Zc−1) = Z ⊕ (Zc ∩ Ac−1) = Z ⊕ Zkc wherek ∈ N is the smallest integer such that kc2 ∈ A, hence A∩Ac−1 = Z[kc] is thering generated by the real algebraic integer kc. By Dirichlet’s unit theorem,the group of units of Z[kc] is the direct product of {±1} with an infinitecyclic group (which is generated by the fundamental unit u > 1 obtained fromsolving some Pell equation); see Borevich–Shafarevich 1966 Chapter 2, 4.3

Theorem 5 (for example, Z[√

5] has the fundamental unit u = 2 +√

5). Thusthe positive units of Z[kc], hence also the automorphisms of the ordered groupA, form an infinite cyclic group.

Page 381: The Classical Fields

364 Hints and solutions

8 The real numbers as a topological group

(1) Cantor’s middle third set C is a meagre null set, and so is S = Z+C. FromC + C = [0, 2] (see Section 1, Exercise 5) it follows that S + C = R.

(2) This follows directly from the definitions.

(3) (a) Any proper subgroup of R is totally disconnected, because each con-nected subset of R is an interval and generates all of R. A proper vectorsubgroup A is dense in R; hence a compact neighbourhood contains an inter-val and generates R.

(b) By the definition of the Tychonoff topology (or the universal propertyof topological products) the direct product of any family of topological groupsis itself a topological group.

(4) By 8.26, the endomorphism ring P = Endc R consists of all maps ϕr =(x �→ rx) with r ∈ R. Clearly ϕ : R → P : r �→ ϕr is an isomorphism of theadditive groups, and ϕrs = ϕr ◦ ϕs, hence ϕ is also multiplicative. A typicalneighbourhood (C, U) in P is {ρ ∈ P | ρ(C) ⊆ U }, where C is compact andU is open in R. Continuity of ϕ−1 follows from the fact that ϕ−1 maps (1, U)to U . If rC ⊆ U , then V C ⊆ U for some neighbourhood V of r, and ϕ is alsocontinuous.

9 Multiplication and topology of the real numbers

(1) Each automorphism α of R× maps the connected component Rpos of R×

onto itself. By 8.24, α induces on Rpos a map x �→ xa with a = 0. Moreover,(−1)α = −1 since −1 is the only involution in R×, and (−x)α = −xα.

(2) Obviously, each motion in M preserves the distance d. Conversely, if a mapϕ of R satisfies |ϕ(x)−ϕ(y)| = |x−y| for all x, y ∈ R, then ϕa = x �→ ϕ(x)−ahas the same property. Putting a = ϕ(0), we get |ϕa(x)| = |x| = ±x. Bycontinuity, the sign is constant and ϕ ∈ M.

(3) The orientation preserving elements in the groups in question form a (nor-mal) subgroup A of index 2. Each element in the complement of A has order 2and generates a factor C2 of a semidirect decomposition.

10 The real numbers as a measure space

(1) Closed sets, and points in particular, are contained in Oδ. Each countableset is in Oδσ = O1 but not in O (because intervals are uncountable).

(2) Let Q = {rμ | μ ∈ N} and choose Oν|κ as a 2−κ-neighbourhood of rν1 .

(3) If k ∈ N and k > 1, then k · S ⊂ S and k · λ(S) = λ(k · S) ≤ λ(S). Hence

λ(S) is either 0 or ∞. A subspace of finite codimension has only countablymany cosets in R. The claim follows as in 10.9 and 10.10.

(4) If T ⊂ M and card T ≤ ℵ, then an argument analogous to 10.15 showsthat S ∪ T generates a σ-field of cardinality at most ℵ.

(5) Let H denote a hyperplane in the rational vector space R. Put C = C2∩H

and B =`

[0, 1] � C2

´∩H. Then λ(B) = 1/2 and λ(B ∪C) = 1 by the remark

following 10.11. Subadditivity of λ gives λ(C) ≥ 1/2, but λ(C) = 0 by 10.8.

(6) The first claim follows from 10.15. Consider now a non-measurable setC ⊂ C2 as in Exercise 5 and a homeomorphism ϕ of R which maps C2 onto

Page 382: The Classical Fields

Hints and solutions 365

Cantor’s middle third set C. Then ϕ(C) is a null set and therefore measurable,but its preimage C is not measurable.

(7) For a given sequence ν ∈ NN, choose μ ∈ A = A such that |ν1 − μ1| ≤|ν1 − ξ1| for all ξ = (ξι)ι ∈ A and, inductively, |νκ − μκ| ≤ |νκ − ξκ| for all(μ1, . . . , μκ−1, ξκ, ξκ+1 . . . ) ∈ A. Whenever there are two possibilities, selectthe smaller one as μκ. Put ρ(ν) = μ. Then ρ|A = idA, and ρ is continuousbecause μκ depends only on (ν0, . . . , νκ).

11 The real numbers as an ordered field

(1) Let F0 = Q and define inductively F2ν+1 = F2ν(exp F2ν) and F2ν+2 =F2ν+1(log(F2ν+1 ∩ Rpos)). Then each Fν is countable, and F =

S

ν∈NFν has

the required property.

(2) If a > 0, then F < at + b, but if c = 0 or a < 0, then F < (at + b)/(ct + d)does not hold.

(3) By 11.14, the field R with the given structure embeds as a subfield into Rwith the usual structure. The subfield is order complete because its orderingis isomorphic to the usual one, hence it contains the closure of Q.

(4) If F+ is Archimedean, then F embeds into R by 11.14, hence the mul-tiplicative group of positive elements is Archimedean. (Alternatively, a > 0implies that (1 + a)n > n · a is not bounded above.) Conversely, if N ≤ k inF , then 2n ≤ k and the multiplication is not Archimedean.

12 Formally real and real closed fields

(1) For f, g ∈ F [t] � {0}, we call the rational function f/g positive if theleading coefficient of the product fg is positive with respect to <. This definesan ordering on F (t); on R(t) this is the ordering P∞,+ introduced in Exercise 3below. For the uniqueness we observe that the condition F < t determinesthe extension uniquely on the polynomial ring F [t], and then also on the fieldF (t) of fractions of F [t].

(2) Let < be the unique ordering of R. The field Q(√

2) admits the field

automorphism α defined by α(a + b√

2) = a− b√

2 for a, b ∈ Q. Hence Q(√

2)

has the two domains of positivity P1 = {a+ b√

2 | a, b ∈ Q, a+ b√

2 > 0} and

P2 = α(P1) = {a + b√

2 | a, b ∈ Q, a − b√

2 > 0}.Now let P be any domain of positivity of Q(

√2). By applying α if necessary,

we may assume that√

2 ∈ P ; then we have to show that P = P1. Since Pand P1 are subgroups of index 2 in Q(

√2)×, it suffices to prove the inclusion

P ⊆ P1. We know that P ∩ Q is the set of positive rational numbers by 11.7.Proceeding indirectly, we assume that there are rational numbers a, b with

a + b√

2 ∈ P and a + b√

2 < 0. If b < 0, then −b√

2 ∈ P , hence a ∈ P , a > 0and a−b

√2 > 0. Thus the rational number a2−2b2 = (a+b

√2)(a−b

√2) ∈ P

is negative, a contradiction. The case a < 0 < b is treated analogously.

(3) P∞,+ is the domain of positivity of the ordering described in Exercise 1.Applying the field automorphisms of R(t) given by t �→ −t or t �→ ±t−1 + r(compare 64.19), we infer that also P∞,− and Pr,± are domains of positivity.

Let P be any domain of positivity of R(t). The corresponding ordering <induces on R the usual ordering (11.7). We can achieve that R < t by applyingthe field automorphisms just mentioned. Now Exercise 1 yields P = P∞,+.

Page 383: The Classical Fields

366 Hints and solutions

(4) The field C((t1/∞)) is algebraically closed by the Newton–Puiseux theorem;

see 64.24. Because C((t1/∞)) = R((t1/∞))(√−1), the field R((t1/∞)) is real

closed by 12.10. In the unique ordering < of R((t1/∞)), we have N < t−1,

since t = (t1/2)2 and 1−nt =`

P

k≥0

`

1/2k

´

(−n)ktk´2

are squares in R((t1/∞))for each n ∈ N.

(5) We show that each element s ∈ P is totally real: There exists a sequenceof quadratic field extensions Fν+1 = Fν(

√1 + c2

ν) such that F0 = Q and

s = a + b√

1 + c2 with a, b, c ∈ Fn. By induction we may assume that a, b, care totally real. If Fn(s) is contained in the Galois extension H of Q, and ifΓ = Aut H, then {sγ | γ ∈ Γ} is the set of all roots of the minimal polynomialof s over Q (see Section 64, Exercise 2). For every γ ∈ Γ the elements aγ , bγ

und cγ are real. Hence (1 + c2)γ = 1 + (cγ)2 > 0 and sγ is real.For the converse, let a be totally real. Then the minimal polynomial of a2

has only positive roots. Hence a2 is a sum of squares; see Jacobson 1989Theorem 11.7.

(6) Let M be a maximal subfield of R. Then R = M(a) for any a ∈ R � M ,and a is algebraic over M , otherwise M < M(a2) < R. Thus [R : M ] and[C : M ] = 2[R : M ] are finite, and 12.15 implies that 2 = [C : M ] = 2[R : M ],hence R = M .

(7) Proceed as in Section 6, Exercise 1, and note that the algebraic closure ofany countable field is countable by 64.16.

(8) Let p = char F . First we show that F is perfect, that is, either p = 0 orF = F p for p > 0. Indeed, if p > 0 and a ∈ F � F p, then each polynomial

f = xpk − a with k ∈ N is irreducible over F : if r is a root of f in some

splitting field, then f = (x − r)pk

; hence any non-trivial factor of f in F [x] isof the form c(x − r)d with c ∈ F× and 0 < d < pk. Then F contains rd and

rpk

= a, hence by Bezout’s theorem 1.5 also rg with g := gcd(d, pk). From

g < pk and (rg)pk/g = a we infer that a ∈ F p, a contradiction.As F is perfect, every finite extension of F is separable (see the proof of

12.15), hence simple (64.12). Thus the degrees of the finite extensions of Fare bounded, and there is a finite extension E of F of maximal degree. Bymaximality, E is algebraically closed. Now apply 12.15.

(9) If a field F is Euclidean, then F is Pythagorean with a unique ordering;see 12.8. Conversely, assume that F is Pythagorean with a unique ordering.By 12.5, the unique domain of positivity P of F consists of all non-zero sumsof squares of F , and P is the set of all non-zero squares of F since F isPythagorean. Hence F is Euclidean.

13 The real numbers as a topological field

(1) The inclusions a+nZ− (b+nZ) ⊆ a−b+nZ for a, b ∈ Q, n ∈ N show thatthe non-zero ideals of Z yield a topology τ on Q such that Q+ is a topologicalgroup. For p, s ∈ Z and q, r, n ∈ N we have (pr−1 + nrsZ)(qs−1 + nrsZ) ⊆pq(rs)−1 + nZ, hence τ is a ring topology of Q. However, τ is not a fieldtopology, because the set (1 + 2Z)−1 does not contain any coset 1 + nZ withn ∈ N.

Let τ be any ring topology of Q which is not a field topology, and let B beany Hamel B basis of R; compare 1.13. Endow QB with the product topology

Page 384: The Classical Fields

Hints and solutions 367

obtained from τ , and consider R as a subspace of QB . Then R is a topologicalring (since the multiplication is bilinear), but not a field (because Q retainsits topology τ).

(2) By 8.15 and 13.6 the topological additive group is isomorphic to R, andthen 13.7 gives the assertion.

(3) Since Q is dense in R, a continuous isomorphism Q(s) → Q(t) is theidentity, whence Q(s) = Q(t). This equation means that s = (at + b)/(ct + d)for some invertible matrix ( a b

c d ); see Cohn 2003a 11.3.3 or Jacobson 1989Section 8.14.

(4) If T is a transcendency basis, so is { t − rt | t ∈ T } for any choice ofnumbers rt ∈ Q.

(5) Use x2 = 1 + 2/((x − 1)−1 − (x + 1)−1) and 4xy = (x + y)2 − (x − y)2.

(6) Let c =

a −bb a

«

∈ H. Then c = 0 implies det c = aa + bb > 0, and c is

invertible. Hence H is a division algebra. Put i =`

1−1

´

and j = ( −11 ). If

ci = ic, then b = 0; if cj = jc, then a = a and b = b. Therefore, the centralizerof i and j consists of the (real) diagonal matrices in H and this is also thecentre of H. Moreover, H ∼= R4 and det c = |a|2 + |b|2 is the square of theEuclidean norm; see also 34.17.

(7) By induction, ccn < 2. For x < 2 we have x < cx because x−1 log xis strictly increasing on the open interval (1, e). Hence the sequence (cn)n

converges to a number s ≤ 2, and s = cs = 2.

(8) The contribution of finitely many elements an to Ca is negligible. If|an − s | < ε for all n, then |cn − s| < ε for all n. If an = (−1)n · (2n + 1),then Ca does not converge, but CCa does.

(9) IfQ

n(1 − an) > 0, then 1 +P

n an <Q

n(1 + an) <Q

n(1 − an)−1 < ∞.Conversely, assume that

P

n an < ∞. Discarding finitely many elements of

the sequence, we may assume thatP

n an < 1/2. ThenQ

n(1 − an)−1 <Q

n(1 + 2an) <Q

n exp(2an) = exp(2P

n an) < e.

(10) If the field F is arcwise connected, then there is also an arc C from 0to 1 (by homogeneity). Let W be an arbitrary neighbourhood of 0. For eachc ∈ C there are neighbourhoods Uc of c and Vc of 0 such that UcVc ⊆ W sincemultiplication is continuous. The compact arc C is covered by finitely manyof the sets Uc. The intersection V of the (finitely many) corresponding setsVc is a neighbourhood of 0 with V ⊆ CV ⊆ W , and CV is arcwise connectedbecause Cv is an arc from 0 to v.

(11) Let T ⊆ R be a transcendency basis over Q; then card T = card R (see64.20, 64.5). As mentioned in 13.10, there exists an arcwise connected Haus-dorff field topology on Q(T ) such that the subfield Q is discrete. Let B be abasis of the vector space R over Q(T ), and consider R as a subspace of theproduct space Q(T )B . This gives a Hausdorff ring topology on R (the multi-plication is bilinear), which is arcwise connected (as T is homeomorphic to R)and distinct from the usual topology of R (since Q is discrete).

Applying the result of Gelbaum, Kalisch and Olmsted mentioned in 13.10 tothe strange ring topology of R described in the previous paragraph, we obtaina field topology of R which is again arcwise connected. This field topology

Page 385: The Classical Fields

368 Hints and solutions

is distinct from the usual topology of R, because T is connected in the newtopology, but totally disconnected in the usually topology (since every intervalcontains infinitely many rational numbers).

14 The complex numbers

(1) A discrete subgroup D < C+ contains an element a of minimal absolutevalue |a| > 0. Let b ∈ D � Za such that the triangle with vertices 0, a, b hasminimal positive area; then D = Za + Zb. Compare also 8.6.

(2) If C is a subgroup of C× and c ∈ C � S1, we may assume that |c| > 1.Then the powers cn converge to ∞ and C is not compact.

(3) The mapping z �→ (z/|z|, log |z|) : C× → S1 × R is an isomorphism oftopological groups. The compact image ϕ(S1) is trivial by 8.6. The mapt �→ ϕ(et) is a continuous homomorphism of R+ into C+, hence R-linear,which gives the assertion.

(4) Each G-conjugacy class in N is connected and discrete, hence a singleton.

(5) Otherwise R or C would be of the form F (t) where t is transcendental overthe subfield F . Such a field F (t) has always automorphisms of order 3, forexample the F -linear automorphism determined by t �→ (1 − t)−1; see 64.19.This contradicts 6.4 and 14.13(i).

(6) If a is contained in some maximal subfield M , then M is real closed byExercise 5 and 12.15, hence Q(a) is formally real. Conversely, assume thatQ(a) is formally real. Let T be a transcendency basis (64.20) of C over Q(a).Then Q(a, T ) is formally real; see Exercise 1 of Section 12. Proposition 12.16yields a real closed field M with Q(a, T ) ⊆ M ⊂ C. From 12.10 we infer thatC = M(

√−1), hence M is a maximal subfield of C.

(7) The set Δ = { (a + b, a + ζb, a + ζ2b) | a ∈ C, b ∈ C× } is the orbit of theequilateral triangle (1, ζ, ζ2) under the group Aff C of similarity transforma-tions; see 14.17. Thus Δ is the set of all equilateral triangles.

By 14.11 and 14.12 there exist many discontinuous field automorphisms αof C. The set {ζ, ζ2} of all roots of unity of order 3 is invariant under α,hence Δ is invariant under α. Now R is not invariant under α (otherwise αwould be R-linear by 6.4, hence continuous), but R can be defined in termsof 0, 1 and the α-invariant set {i,−i} using any one of the mentioned geo-metric notions; for example, R = {x ∈ C | 0, 1, x are collinear} = {x ∈ C |(x, i) is congruent to (x,−i)}.(8) The extension E1E2|F is finite by 64.2. It is a Galois extension, since Ei

is the splitting field of some set Xi of separable polynomials over F (compare64.17), and E1E2 is the splitting field of X1∪X2. The Galois group of E1E2|Fis solvable, because it embeds into the direct product GalF E1 × GalF E2, viarestrictions to Ei.

(9) Every element g of a free group can be written uniquely in the form g =xe1

1 · · ·xenn where x1, . . . , xn are free generators with xi = xi+1 for 1 ≤ i < n

and 0 = ei ∈ Z. If g2 = 1, then g = g−1 = x−enn · · ·x−e1

1 , hence xi = xn+1−i

and ei = −en+1−i by uniqueness. If n = 2k − 1 is odd, then ek = −ek, acontradiction to ek = 0. If n = 2k > 0, then xk = xk+1, again a contradiction.Hence n = 0 and g = 1.

Similar arguments with normal forms show that free groups are torsion free;see Cohn 2003b 3.4.1, Jacobson 1989 2.13 p. 89 or Lang 1993 I.12.6 p. 74.

Page 386: The Classical Fields

Hints and solutions 369

21 Ultraproducts

(1) (a) We have char E = p; (b) every element of E is a sum of two squares;(c) if p is odd, then −1 is a square in E if, and only if, p ≡ 1 mod 4 or 2N ∈ Ψ;

(d) card E = ℵ; (e) F �pν = F �

p and E� < (F �p)Ψ for analogous reasons as in

21.8(f).

(2) If M ∈ Ψ and sν < tν for ν ∈ M , then also h(sν) < h(tν) for ν ∈ M .Hence hΨ is strictly increasing. Because h is surjective, so is hΨ.

(3) The image is [−1, 1]Ψ = {x ∈ RΨ | −1 ≤ x ≤ 1}.22 Non-standard rationals

(1) If x = (xν)ν/Ψ ∈ Q�, then there exists a polynomial f with rationalcoefficients such that f(xν) = 0 for almost all ν. This equation has onlyfinitely many solutions, and 21.2(c) shows that xν = c ∈ Q� for almost all ν.If x ∈ QΨ, then xν ∈ Q for all ν.

(2) If 2N ∈ Ψ, then sν2 < 2 for almost all ν ∈ N. If 2N /∈ Ψ, then the set of

all odd numbers belongs to Ψ and sν2 > 2 for almost all ν.

(3) By definition, π(pν) = ν. The prime number theorem implies ν · log pν <2pν for almost all ν ∈ N and hence k · ν < pν for each k ∈ N and allmost all ν.On the other hand, 2π(n) > n/ log n and 2k · log n <

√n, hence π(n)2 > k2 ·n

for each k ∈ N and almost all n.

23 A construction of the real numbers

(1) If a, b ∈ R and 0 < a � b < m ∈ N, then a ∈ N .

(2) Being a Euclidean field is a first-order property. Hence ∗E is a Euclideanfield. Obviously, ∗Q < ∗E, and K ≤ ∗E. Each element of K is algebraic overthe field ∗Q. Put c0 = 2 and cν+1 =

√cν . Then (cν)ν represents an element

c ∈ ∗E, and c is not of finite degree over ∗Q. Therefore, K < ∗E.

(3) Similarly, ∗P is a Pythagorean field. If x ∈ ∗P , then x4 = 2 since 4√

2 /∈ P ;see Exercise 5 of Section 12. Consequently, ∗P < ∗E.

24 Non-standard reals

(1) For x = (xν)ν ∈ RN, put Cx = {ν | xν = 0}. Then x ∈ M ⇔ Cx ∈ Ψ,and A ∈ Ψ if, and only if, Cx = A for some x ∈ M .

(2) Surjectivity is obvious from the definition ∗exp x = (exp xν)ν . This showsalso that the kernel of ∗exp is ∗Z2πi.

(3) A positive element b can be represented by a sequence (bν)ν , where bν > 0for all ν ∈ N. Then ba = (eaν log bν )ν/Ψ.

(4) The conjugation ι of ∗C fixes exactly the elements of ∗R. By 24.6 there isan isomorphism α : ∗C ∼= C, and F = ∗Rα is the fixed field of the involutionια = α−1ια.

(5) The map x �→ 1 − (1 + exp x)−1 is an order isomorphism of ∗R onto theinterval ]0, 1[ ⊂ ∗R. Each countable sequence in ∗R is bounded; in R this isnot true.

(6) If { bν | ν ∈ N} is a basis of ∗R, then, by 24.11, there is an element b ∈ ∗Rsuch that bν < b for all ν, moreover, N · b < c for some c ∈ ∗R. Each element

Page 387: The Classical Fields

370 Hints and solutions

in ∗R has a representation x =Pn

ν=1 rνbν with rν ∈ R. Choose m ∈ N withrν < m for ν = 1, 2, . . . , n. Then x < n · m · b < c, and c is not in the vectorspace spanned by the bν .

(7) Note that v(x) = v(1) = 0 ⇔ xR = R ⇔ x ∈ R×. Thus v(x) > 0 ⇔ x ∈ n,and V0 = n ∈ τ because 0 is in the interior of n. If γ = v(c), then Vγ = cn isalso open. On the other hand, each interval ]−c, c[ contains Vγ .

25 Continuity and convergence

(1) Suppose that s converges to t and that {ν ∈ N | s(ν) = 0} ∈ Ψ. Thent = 0 and ∗s(m) ≈ 0 for each strictly increasing sequence m. This means that{ν | s(ν) = 1} is finite or that s is finally constant.

(2) Put x = (ν)ν and y = (ν + ν−1)ν . Then x ≈ y, but y2 − x2 ≈ 2.

(3) It suffices to show that f is constant on some neighbourhood of eachpoint a ∈ R. Assume that there are elements sν ∈ (a − ν−1, a + ν−1) withf(sν) = f(a). Then s = (sν)ν/Ψ ∈ a + n and ∗f(s) = f(a).

26 Topology of the real numbers in non-standard terms

(1) The first part is easy, and the second part is a consequence of (A∩B)+n ⊆∗A ∩ ∗B = ∗(A ∩ B). Dually, ∗(A ∪ B) = ∗A ∪ ∗B by 21.2c, and the last claimfollows with 26.2.

(2) We have Q = R by 26.2, since R ⊆ ∗Q + n.

(3) The intersection of two dense open sets A, B is dense: if U is open andU = ∅, then (A ∩ B) ∩ U = A ∩ (B ∩ U) = ∅.

If the Cκ form a descending sequence of dense open sets, then D =T

κ Cκ

is dense: by 26.2 and translation invariance it suffices to show that ∗D∩n = ∅.Let cκ = (cκν)/Ψ ∈ ∗Cκ ∩ n. Then cκν ∈ Cι for ι ≤ κ and |cκν | < κ−1 foralmost all ν. The element cκκ may be chosen arbitrarily in Cκ, say such that|cκκ| < κ−1. Obviously, d = (cνν)ν/Ψ ∈ ∗D ∩ n.

Write Q = {rν | ν ∈ N} and put U =S

ν ]rν − ε2−ν , rν + ε2−ν [. Then Uhas Lebesgue measure < 4ε.

(4) A closed interval C is compact. There are xν ∈ C such that the f(xν)converge to t = sup f(C), and x = (xν)ν/Ψ ∈ ∗C. By 26.5, there is an elementc ∈ C with c ≈ x. Continuity implies f(c) ≈ ∗f(x) ≈ t, and f(c) = t.

(5) By Exercise 4, there is some c ∈ C such that f(x) ≤ f(c) for all x ∈ C. Ifx ∈ ∗C and x = (xν)ν/Ψ, then ∗f(x) = (f(xν))ν/Ψ ≤ f(c); see 21.6.

27 Differentiation

(1) There is some real r > 0 such that f(a) ≤ f(x) for x ∈ [a − r, a + r].Consequently, ∗f(x ± h) ≥ f(a) for all h ∈ n, and condition (∗) implies thatf ′(a) = 0.

(2) One has h−1`

g(a + h) − g(a)´ ≈ −h−1

`

f(a + h) − f(a)´

g(a)2.

31 The additive group of the rational numbers

(1) Every subgroup A ∼= Q of Q is divisible. By 31.10 we may assume that1 ∈ A. Then A contains all fractions 1/n with n ∈ N, hence A = Q.

(2) See Section 1, Exercise 7.

Page 388: The Classical Fields

Hints and solutions 371

(3) If p and q are distinct primes, then the map (x �→ px) induces an automor-phism of 〈pZ〉 = {apn | a, n ∈ Z}, but it maps 〈qZ〉 onto a proper subgroup.Hence 〈pZ〉 ∼= 〈qZ〉.(4) Let X be any set of primes and consider the subgroup AX of Q+ generatedby all sets qZ with q ∈ X. If p is a prime, then the map (a �→ pa) is surjectiveif, and only if, p ∈ X. Hence the groups AX are mutually non-isomorphic.

32 The multiplication of the rational numbers

(1) The subgroup of Q× generated by −1 together with any infinite set ofprimes is isomorphic to Q×.

(2) Exactly the infinite cyclic ones and {1}, because Q+ is locally cyclic.

(3) Consider the subgroups generated by an arbitrary set X ⊂ P such thatP � X is infinite.

(4) By 32.1, Q×pos

∼= Z(P). For q ∈ P, define (δpq)p ∈ Z(P) by δpp = 1 and

δpq = 0 for p = q. These elements generate the additive group Z(P). An

endomorphism ϕ ∈ End Z(P) is represented by the matrix (zpq)p,q ∈ Z(P)×P

whose q-th column is the image ϕ((δpq)p). Addition and composition of en-

domorphisms correspond to addition and multiplication of matrices in Z(P)×P.The additive group of Z(P)×P clearly is isomorphic to (Z(P))P ∼= (Q×

pos)P, in

agreement with 32.4.

(5) There is at least one prime q ≡ −1 mod 3 which divides (2 ·3 · · · p)−1 andat least one prime r ≡ −1 mod 4 which divides (2 · 3 · · · p) + 1.

(6) Let m be the product of all primes ≤ p. Note that (m− 1)(m2 +m+1) =m3 −1 and gcd(m−1, m2 +m+1) = gcd(m−1, 2m+1) = gcd(m−1, 3) = 1.If the prime r divides m2 + m + 1, then r > p and m3 ≡ 1 ≡ m mod r. Hence3 is the order of m in F×

r , whence 3 | r − 1.

(7) Each prime q ≡ −1 mod 4 in Z is also prime in J. In fact, if n is anodd natural number, then n2 ≡ 1 mod 4, hence c ∈ J implies cc ≡ −1 mod 4.Therefore cc = q and any factor c of q in J has norm 1 or norm q2.

33 Ordering and topology of the rational numbers

(1) The additive group S := {m ·2n | m, n ∈ Z} is strongly dense in itself andcountable, hence (S, <) ∼= (Q, <) by 3.4.

(2) Each interval ]n, n + 1[ is isomorphic to the chain (Q, <). Denote byτ : Q → Q the translation x �→ x + 1. There are 2ℵ0 automorphisms of thechain (Q, <) which fix each integer n and induce on ]n, n + 1[ a map which isequivalent to τ or to τ−1.

(3) With the lexicographic ordering, the chain Qn is strongly dense in itselfand countable, hence isomorphic to the chain (Q, <) by 3.4.

(4) There is an involution η ∈ H which inverts the order on the open interval

]−√2,√

2 [ (of length < 3) and induces the identity outside of this interval.If τ is the translation x �→ x + 1, then ητ 3η does not preserve the ordering;hence Γ is not normal in H.

Let HI consist of all elements of H which permute the set I of rationalintervals ]n +

√2, n + 1 +

√2[ with n ∈ N. The group HI induces on I the

full symmetric group Sym I (compare the proofs of 33.16 and 33.17), and

Page 389: The Classical Fields

372 Hints and solutions

ΓI := Γ ∩ HI acts trivially on I. Hence |H : Γ| ≥ |HI : ΓI | ≥ card Sym I = 2ℵ0

by 61.16, and 33.15 yields |H : Γ| = 2ℵ0 .

(5) A complete metric space X is a Baire space, i.e., any intersection of count-ably many dense open subsets is dense. However, the intersection of thecomplements of all singletons in X is empty.

(6) The chain (Q, <) is strongly dense in itself, hence each open interval is openin the Sorgenfrey topology σ. An interval ]a, b] is not open in the ordinarytopology τ . Therefore τ is a proper subset of σ. Obviously, σ has a countablebasis. An interval ]a, b] is also σ-closed. Consequently, (Q, σ) is a regulartopological space, hence metrizable.

34 The rational numbers as a field

(1) The two fields F1 := Q(√−2) and F2 := Q(

√−7) satisfy F+1

∼= F+2 ,

because [Fν : Q] = 2. We will show that F×ν

∼= Q×.Note that Fν is the field of fractions of the ring Rν of algebraic integers

in Fν . We have R1 = Z + Z√−2 and R2 = {x + y

√−7 | 2x, 2y ∈ Z ∧x − y ∈ Z}. The elements ±1 are the only units in Rν and Rν is a Euclideanring with respect to the norm N (for a, b ∈ R2 there is some q ∈ R2 suchthat N(ab−1 − q) ≤ 7/42 + 1/4 < 1). Hence both rings have unique primedecomposition. Moreover, there are infinitely many primes in Rν : if p ∈ P andp is not a prime in Rν , then p = uv with Nu = Nv = p, and u, v are primes inthe ring Rν . Obviously, −2 is not a square in F2. Therefore F1 ∼= F2.

See also Amer. Math. Monthly 93 1986 p. 744 Problem 6489. A moregeneral result due to Skolem is given in Fuchs 1973 Theorem 127.2.

(2) Assume that x2 + y2 = 3 with x, y ∈ Q. Then there exist relativelyprime integers a, b, c such that a2 + b2 = 3c2. Because m2 ≡ −1 mod 3 for allm ∈ N, it follows that a, b ≡ 0 mod 3 and then also c ≡ 0 mod 3, which is acontradiction.

(3) It suffices to show that the projection of C ∩ Q2 onto the x-axis is densein [−1, 1]. For u, v ∈ N, we have

`

(u2 − v2)/(u2 + v2), 2uv/(u2 + v2)´ ∈ C.

Hence our claim says that the numbers u2/(u2 + v2) are dense in the interval[0, 1] or, equivalently, that the numbers v2/u2 are dense in [0,∞[. The latteris obviously true.

(4) The group Δ consists of all rational matrices

a −bb a

«

with a2 + b2 = 1;

it is sharply transitive on C ∩ Q2.

(5) Let n be as in Corollary 34.9, and put T := {a2 + b2 | a, b ∈ N}.If v2 is odd or if n is divisible by some prime number p ≡ 1 mod 4, then n ∈

{2m2, km2, 2km2}, where k > 1 is a product of distinct primes p ≡ 1 mod 4.By 34.9 we have k = a2 + b2 with a, b ∈ N, as k is not a square; thus k ∈ T .Moreover, 2 ∈ T and 2k = (a + b)2 + (a − b)2 ∈ T , as a = b. We infer thatn ∈ T .

For the converse, let n ∈ T . Aiming for a contradiction, we assume thatv2 ≡ 0 mod 2 and vp = 0 for p ≡ 1 mod 4. Then n is a square that is notdivisible by any prime p ≡ 1 mod 4; see 34.9. By repeated application of 34.6,we infer from n ∈ T that 4e ∈ T with e ≥ 0. Let 4e = a2 + b2 with a, b ∈ N;then e > 0 and a2 + b2 ≡ 0 mod 4. Hence a and b are even. We conclude that4e−1 ∈ T , hence 40 = 1 ∈ T , which is absurd.

Page 390: The Classical Fields

Hints and solutions 373

(6) There are two essentially different representations: 30 = 52 + 22 + 12 + 02

and 30 = 42+32+22+12. Counting permutations and signs, the total numberis 8 · 3 · 24.

(7) Let a/b = s+ti ∈ Q(i) and choose z = x+yi ∈ J such that |x−s|, |y−t| ≤ 12.

Then a − bz = b(a/b − z) and N(a − bz) ≤ 12Nb.

36 Addition and topologies of the rational numbers

(1) According to 31.10, each group automorphism of Q+ is given by multi-plication with a rational number and therefore continuous. Let 0 = s ∈ S.If the rational subspace sQ is a proper subgroup of S, then S = sQ ⊕ H forsome non-trivial rational subspace H < S (here we use the axiom of choice).We obtain an automorphism α of S with α|H = id and α|sQ = id. Since H isdense in R and in S, such a map α is not continuous.

(2) Each neighbourhood of 0 generates the full group Q+, hence Q has noproper open subgroup. A cyclic subgroup is closed. If a subgroup A is notcyclic, then it is dense (1.4) and hence A = Q if A is closed.

37 Multiplication and topology of the rational numbers

(1) A typical neighbourhood inL

p pZ is given by specifying a finite numberof entries, while the remaining ones are arbitrary; hence μ is not continuous.Write r =

Q

p pvp(r). Continuity of μ−1 means that each map vp is continuous

on Q×pos; obviously, this is not the case.

(2) Via logarithms, Q×pos is isomorphic to a dense subgroup S of R+; see 2.2

and 1.4. Each character α : S → T is uniformly continuous and hence has aunique extension to a character α : R → T; compare 43.23. This proves thatS∗ ∼= R∗ ∼= R+.

51 The field of p-adic numbers

(1) The partial sumsPk

n=0 pn = (1 − pk+1)/(1 − p) converge for k → ∞ to

1/(1 − p), as pk+1 converges to 0.

(2) One hasP∞

n=0(p − 1)pn = (p − 1)/(1 − p) = −1 in Qp by Exercise 1.

In Q3, we want to solve (1+32)P

n cn3n = c0+c1·3+P

n≥2(cn+cn−2)3n = 1

with cn ∈ {0, 1, 2}. Reading this equation modulo 3kZ3 and applying induc-tion over k, we obtain c0 = 1, c1 = 0, c2 + c0 = 3, cn + cn−2 + 1 = 3 forn ≥ 3, hence cn = 2 for n ≡ 2, 3 mod 4 and cn = 0 for 0 < n ≡ 0, 1 mod 4.In passing, this yields the expansion 1 = 1 + 3 · 32 + 2 · 33 · P

j≥0 3j ; compareExercise 1.

WritingP

n≥0 cnpn = c0, c1c2c3 . . . and using Exercise 1, quotients in Qp

can also be determined by the familiar ‘long division’ procedure.

(3) Each seriesP

n≥k znpn with k ∈ Z and zn ∈ Z converges in Qp, in view of

|znpn|p ≤ p−n; compare Exercise 5. Hence by substituting p for x, we obtainan epimorphism Z((x)) → Qp of rings. This epimorphism maps Z[[x]] ontoZp; see 51.8 and 51.9.

The kernel of this epimorphism contains x−p, hence also the ideal generatedby x − p, and it remains to show that every f =

P

n≥k anxn ∈ Z((x)) with

f(p) = 0 is divisible by x − p in Z((x)).

Page 391: The Classical Fields

374 Hints and solutions

For this purpose we define integers bn with an = bn−1 − pbn for n ≥ k asfollows. Put bk−1 = 0. To define bn inductively, we infer from f(p) = 0 thatpn+1Zp contains

P

j≤n ajpj = anpn +

P

j<n(bj−1 − pbj)pj = (an − bn−1)p

n.

Hence p divides an − bn−1, and bn := (an − bn−1)/p is an integer. Thenf =

P

n≥k(bn−1 − pbn)xn = (x − p)P

n≥k bnxn.Compare Faltin et al. 1975.

(4) Let kn be the order of a in the group of units of the finite ring Z/pnZ.Then |akn − 1|p ≤ p−n, hence limn→∞ akn = 1 in Qp.

(5) The ultrametric inequality implies that the partial sums ofP

n an form aCauchy sequence whenever the sequence (an)n converges to 0.

Rearranging does not affect the property that (an)n converges to 0, hencethe rearranged series

P

n aπ(n) converges for every permutation π of N. For

n ∈ N and n′ := max π−1({1, 2, . . . , n}) we obtain |P

j≤n aj −P

j≤n′ aπ(j)|p ≤sup{ |aj |p | j > n}, which converges to 0 for n → ∞. Hence the rearrangedseries has the same value.

(6) If limn |an+1 − an|p = 0, then the sequence ak = a1 +Pk−1

n=1(an+1 − an)converges by Exercise 5.

(7) Since Zp � pZp is closed in Qp (see 51.6), it suffices to prove that P � {p}is dense in Zp � pZp. Because Z is dense in Zp by 51.9, every neighbourhoodof an element of Zp � pZp contains a subset a + pnZp with a ∈ Z � pZ andn ∈ N. The set a + pnZ contains infinitely many prime numbers accordingto a famous result of Dirichlet; for a proof see Shapiro 1950, Serre 1973VI §4, Borevich–Shafarevich 1966 Chapter 5 §3, Lang 1970 VIII §4 orNeukirch 1992 VII.5.14.

(8) This map f (which can be found in Witt 1975) is bijective for com-binatorial reasons. f is continuous, because the first n coordinates of f(x)depend only on the first n coordinates of x. Since both products are compactHausdorff spaces, f is a homeomorphism.

(9) This holds precisely if a ∈ 1+pZp. Indeed, if a ∈ 1+pZ, then |an−am|p =|an−m − 1|p = |a ± 1|p · |n − m|p by 37.5. As Z is dense in Zp (51.9), we havethe same equation for a ∈ 1 + pZp, hence f is continuous.

Conversely, if f is continuous, then f(1+pn) = a ·apn

converges to f(1) = a

for n → ∞, hence limn→∞ apn

= 1. Thus |a|p = 1. Moreover the map x �→ xp

acts as the identity on the quotient Zp/pZp∼= Fp, hence 1 belongs to the

compact set a + pZp.See also 37.6(ii).

(10) The set N is p-adically dense in Zp by 51.9 and f is continuous, hencef(Z) ⊆ f(Zp) ⊆ Zp for each prime p. Thus f(Z) ⊆ T

p Zp ∩ Q = Z.

(11) We claim that each power series fk =P

n≥0 nkxn represents on pZp a

rational function of x with rational coefficients (this implies thatP

n≥0 nkpn =

fk(p) ∈ Q). We proceed by induction on k and observe that f0 represents1/(1 − x). The power series

P

n≥1 nk+1xn−1 represents the derivative f ′k on

pZp (this is justified in Robert 2000 5.2.4). The derivative f ′k is a rational

function with rational coefficients (by induction). Hence also fk+1 = xf ′k

represents a rational function with rational coefficients.

Page 392: The Classical Fields

Hints and solutions 375

52 The additive group of p-adic numbers

(1) By 52.1, R and Qp are vector spaces over Q of the same infinite dimension,and the quotient modulo a one-dimensional subspace has the same dimension.

(2) Let U be a subgroup of index n in Z+p . Then nZp ⊆ U , hence 52.6 implies

that n is a power of p. Now use the proof of 52.3(i).

(3) By 52.2 we have to determine the sequences c = (cq)q ∈×q∈P�{p} Cq∞

which have finite order. This happens precisely if only finitely many compo-nents cq are non-trivial, and this means that c ∈ L

q∈P�{p} Cq∞ .

(4) The kernel of χ is Zp, and the image is isomorphic to the Prufer group Cp∞ .Moreover, χ is continuous, as Qp/Zp

∼= Cp∞ carries the discrete topology.

(5) The map χa is a character of F+, since x �→ ax is a continuous endo-morphism of F . Hence χ : F → F ∗ : a �→ χa is a monomorphism, which iscontinuous by the definition of the compact-open topology on F ∗.

In fact, χ is an embedding: let (aν)ν be a net in F such that χaν convergesto χa; we have to show that (aν)ν converges to a. The net (aν)ν has an accu-mulation point c in the one-point compactification F ∪ {∞} of F . If c = ∞,then (a−1

ν )ν accumulates at 0; see Grundhofer–Salzmann 1990 XI.2.11; al-ternatively, F is of type V (compare Warner 1989 Theorem 19.7(3)), thusone can use Exercise 3 of Section 57. Hence (a−1

ν b)ν accumulates at 0, andχ1(b) = χaν (a−1

ν b) accumulates at χa(0) = 0 for every b ∈ F , a contradictionto χ1 = 0. Therefore c ∈ F , and then χaν accumulates at χc, whence c = a bythe injectivity of χ. This shows that (aν)ν converges to a. (Another argumentuses the absolute value of F constructed in 58.5, as follows. It suffices to con-sider the case a = 1. A neighbourhood Ω of χ1 consists of all ϕ ∈ F ∗ mappinga ball B = {x ∈ F | |x| ≤ r } into an open subset J ⊆ R/Z with χ1(F ) ⊆ J .Choose c such that χ1(c) /∈ J . If χa ∈ Ω, then c /∈ aB and |a| < r−1|c|. Hence{a ∈ F | χa ∈ Ω} is contained in some compact ball and the claim follows.See also the proof of 8.31(a) and Hewitt–Ross 1963 25.1.)

Consequently, χ(F ) is closed in F ∗; see Section 62, Exercise 2. If χ(F ) = F ∗,then by duality (63.27), there is some x ∈ F× such that χa(x) = 0 for alla ∈ F , but then χ1(F ) = 0 and χ1 would be trivial.

(6) The quotient Sp is compact, since it is the image of the compact space[0, 1]×Zp. It is a Hausdorff space by 62.8, because { (z, z) | z ∈ Z} is discreteand closed in R × Zp. The sets R × {0} and R × Z have the same image inSp, hence this image is connected. Moreover R×Z is dense in R×Zp by 51.8,hence Sp is connected.

An element (t, x) ∈ R × Zp is mapped to a torsion element in Sp if, andonly if, n · t = n · x ∈ Z for some n ∈ N. By 52.2, the torsion subgroup of Sp

is isomorphic toL

q∈P�{p} Cq∞ .

(7) We have Aut Fpqn ∼= Cqn ; compare Exercise 6 of Section 64. Each element

of Cqn = Z/qnZ has the formP

j<n cjqj + qnZ with uniquely determined

numbers cj ∈ {0, 1, . . . , q − 1}. Any automorphism σ of Fp,q induces an auto-morphism σn on Fpqn , and σ is uniquely determined by the sequence (σn)n.

Thus σ corresponds toP

j cjqj ∈ Z+

q , and this correspondence is a groupisomorphism.

Page 393: The Classical Fields

376 Hints and solutions

53 The multiplicative group of p-adic numbers

(1) The first part is 53.1(iii). If n ≥ 2, then Un is an open subgroup of U1 andof U2; see 53.1(i). By 53.2, U1

∼= Z+p for p = 2 and U2

∼= Z+2 for p = 2. Result

52.3(i) implies that every open subgroup of Z+p is isomorphic to Z+

p .

(2) We have vp(n!) = max{e ∈ N0 | pe divides n!} =P

j≥1[n/pj ], where

[r] := max{z ∈ Z | z ≤ r }. By Exercise 5 of Section 51, the series for exp(x)

converges precisely if |xn/n!|p = |x|np · pvp(n!) converges to 0 for n → ∞.

The estimate vp(n!) ≤ nP

j≥1 p−j = n/(p − 1) gives |xn/n!|p ≤ |x|np ·pn/(p−1), hence exp(x) converges for p = 2 and |x|p ≤ p−1, and for p = 2 and|x|2 ≤ 2−2.

The estimates vp(n!) ≥ [n/p] > (n/p) − 1 and |xn/n!|p ≥ |x|np · p(n/p)−1

show that exp(x) does not converge for |x|p ≥ 1. If p = 2 and |x|2 = 2−1,

then exp(x) does not converge either, since then |xn/n!|2 = 2−n+v2(n!), andone has −2k + v2((2

k)!) = −2k +P

j≥1[2k−j ] = −1 for all k ∈ N.

(3) If x ∈ N, then`

xn

´ ∈ N0. We infer that`

xn

´ ∈ Zp for x ∈ Zp from the densityof N in Zp (51.9) and the continuity of polynomials. This implies that the seriesP

n≥0

`

xn

´

pn converges in Zp for every x ∈ Zp. The mapping β : Zp → U1 ⊆ Zp

is continuous, because β−1(a + pnZp) = {x ∈ Zp | Pn−1k=0

`

xk

´

pk ∈ a + pnZp }is the preimage of a + pnZp under a polynomial. Furthermore, β satisfiesβ(x + y) = β(x)β(y) for all x, y ∈ Zp, because this is true for x, y ∈ N bythe binomial theorem, and then in general by the density of N in Zp and thecontinuity of β.

(4) This can be proved by induction on n. Alternatively, one can use ideasfrom the proof of 37.5.

(5) If p = 2 and x ∈ Zp, then 1 + px2 ∈ 1 + pZp, which consists of squaresby 53.2 and 52.6. For the converse inclusion, consider x ∈ Qp � Zp. Then|px2|p = p−1|x|2p ≥ p−1+2 > 1, hence |1 + px2|p = |px2|p is a power of p with

odd exponent, and we infer that 1 + px2 is not a square.If x ∈ Z2, then 1 + 2x3 ∈ 1 + 2Z2; this group is 3-divisible by 53.2 and

52.6. For the converse let x ∈ Q2 � Z2. Then |2x3|2 = 2−1|x|32 > 1, hence|1+2x3|2 = |2x3|2 = 2n with n not divisible by 3. Thus 1+2x3 is not a cube.

These descriptions of Zp show that Zp is invariant under every ring endo-morphism of Qp, and this invariance is the crucial step for proving 53.5.

(6) Use the direct decomposition 53.3 of Q×p , the isomorphism 53.2 and the

divisibility property 52.6.

54 Squares of p-adic numbers and quadratic forms

(1) By 53.3 or 54.1, this happens if, and only if, p ≡ 1 mod 4.

(2) For p = 2 this is the special case n = 2 of Exercise 6 for Section 53; oruse 54.1. For p = 2 we infer from 54.1 that Q�

2 = 〈4〉 × (1 + 8Z2) consistsof all 2-adic numbers 2r(c0 + 2c1 + 4c2 + 8c3 + · · · ) such that r is even andc0 + 2c1 + 4c2 ≡ 1 mod 8.

(3) One has to characterize when d is a square in Qp. By 54.1, this holds forp = 2 precisely if d is a non-zero square modulo p, and for p = 2 precisely ifd ≡ 1 mod 8.

Page 394: The Classical Fields

Hints and solutions 377

(4) Suppose that x2 + y2 + z2 + 1 = 0 with x, y, z ∈ Q2. We may assume that|x|2 = max{|x|2, |y|2, |z|2}. Then division by x2 gives a relation of the sametype with x, y, z ∈ Z2. Since Z2/8Z2

∼= Z/8Z (see 51.7), we obtain the samerelation in the finite ring Z/8Z. This is a contradiction, since 0, 1, 4 are theonly squares of Z/8Z.

The quadratic form x2 + y2 + z2 takes the seven values 1, 1 + 1 = 2,1+1+1 = 3, 4+1+1 = 6, 9+4+1 = 14 = (−2)(−7), 16+4+1 = 21 = (−3)(−7)and 25 + 16 + 1 = 42 = (−6)(−7). Since −7 = 1 − 8 is a square in Q2, thisshows that x2 + y2 + z2 represents all square classes of Q2 except the squareclass of −1; see 54.1.

(5) By the Chinese remainder theorem we have Z/mZ ∼= L

i Z/peii Z as rings,

with prime numbers pi. Hence it suffices to consider the special case wherem = pe is a power of a prime p. By 54.7 the quadratic form f is isotropicover Qp, and by homogeneity there exists a solution x ∈ Zn

p � (pZp)n of

f(x) = 0. Since Zp/peZp∼= Z/mZ we obtain a non-trivial solution of f(x) = 0

in (Z/mZ)n by reducing each coordinate of x modulo peZp.

55 Absolute values

(1) The condition ϕ1(a) < 1 is equivalent to limn an = 0 with respect to thetopology defined by ϕ1, hence equivalent to ϕ2(a) < 1 if the two absolutevalues define the same topology.

Assume that ϕ1(a) < 1 ⇔ ϕ2(a) < 1. For x, y ∈ F×, m, n ∈ Z and a :=xmyn we obtain m log ϕ1(x)+n log ϕ1(y) < 0 ⇔ m log ϕ2(x)+n log ϕ2(y) < 0,hence log ϕ1(x)/ log ϕ2(x) = log ϕ1(y)/ log ϕ2(y) is a constant s > 0.

(2) By Exercise 1 we find a, a′ ∈ F with ϕ1(a) < 1 ≤ ϕ2(a) and ϕ2(a′) <

1 ≤ ϕ1(a′). Then b := a/a′ satisfies ϕ1(b) < 1 < ϕ2(b). The elements

1− (1+ bn)−1 = 1− b−n(1+ b−n)−1 converge to 0 in F1 and to 1 in F2. Hencethe closure C of the diagonal contains the pair (0, 1). Replacing b by b−1

shows that (1, 0) ∈ C. Since C is an F -subspace of F 2, this yields C = F 2.

(3) Let a ∈ bF . Since F is dense in bF , also F − a is dense in bF , hence we findb ∈ F with bϕ(b−a) < bϕ(a). Then bϕ(a) = max{bϕ(b−a), bϕ(a)} = bϕ(b−a+a) =ϕ(b); see 55.3.

(4) For the subadditivity of ϕs it suffices to show that a ≤ b + c impliesas ≤ bs + cs, where 0 < a, b, c ∈ R. We may assume that a = 1; then theconclusion is true if b ≥ 1 or c ≥ 1. For b, c ≤ 1 we have 1 ≤ b + c ≤ bs + cs.

56 Valuations

(1) If F is algebraic over a finite field, then F× consists of roots of unity, henceevery homomorphism of F× into an ordered group is trivial. If F is not analgebraic extension of a finite field, then F contains a subfield F0 isomorphicto Q or Fp(x). By 56.3 we find valuations on F0, which extend to F by 56.9.In order to obtain an absolute value on F , use Zorn’s Lemma to enlarge anyvaluation ring of F to a maximal subring R of F ; then R is a valuation ring,and by 56.8 the corresponding value group is Archimedean, which leads to anabsolute value by 56.2.

(2) As in the proof of 56.7, one uses the fact that k(x) is a principal idealdomain; for details see Cohn 2003a 9.1 p. 312, Ribenboim 1999 3.1.K p. 89,Engler–Prestel 2005 Theorem 2.1.4b p. 30 or Bourbaki 1972 VI.1.4 p. 380.

Page 395: The Classical Fields

378 Hints and solutions

(3) If v is principal and π is a prime element, then every element of F× hasthe form aπn with n ∈ Z and v(a) = 0, i.e. a ∈ R×. Hence F× = πZ × R×,and the non-trivial ideals of R are the principal ideals πnR with n ≥ 0.

Conversely, if R is a principal ideal domain, then its maximal ideal is of theform πR. One shows that

T

n πnR = {0} and that v(π) generates the valuegroup; see Cohn 2003a 9.1.3, Warner 1989 Theorem 21.3 or Bourbaki 1972VI.3.6 Proposition 9 p. 392 for details.

If F is complete and Fv is finite, then one finds in R a root of unity ζ oforder card F×

v as in 53.1(ii). As in 53.1(iii) we obtain R× = 〈ζ〉 × (1 + M).

(4) Choose e ∈ E with v(e) > 0. Then en =Pn−1

i=0 aiei for some n ∈ N

and suitable elements ai ∈ F , a0 = 0. The assumption that v is trivial on Fleads to v(

P

i aiei) = min{v(aie

i) | ai = 0} = v(a0) = 0, a contradiction tov(en) = nv(e) = 0.

(5) In order to prove that w(fg) ≤ w(f) + w(g) for polynomials f =P

aixi

and g =P

bixi, consider i0 = min{ i | v(ai) + iδ = w(f)} and j0 = min{ j |

v(bj) + jδ = w(g)}. For details see Bourbaki 1972 VI.10.1 Lemma 1 p. 434,or Engler–Prestel 2005 Theorem 2.2.1.

(6) If R is a valuation ring of F , then F = R ∪ (R � {0})−1, hence F is thefield of fractions of R. Conversely, let F be the field of fractions of a maximalsubring R. Assume that x ∈ F �R with x−1 /∈ R. By maximality, the subringR[x] generated by R ∪ {x} is F , and similarly R[x−1] = F . Hence 1 ∈ R[x]and 1 ∈ R[x−1], and this leads to a contradiction, as in the proof of 56.9.

(7) One example is the field F = k((Γ)) of Hahn power series; see 64.25. Asmaller example is the field of fractions of the group ring kΓ ⊆ k((Γ)); compareRibenboim 1999 13.1.C p. 368.

(8) Pure transcendency is ruled out by the fact that Qp has trivial automor-phism group; see 53.5 and 64.19. If Qp|F is a finite extension, then Qp = F (a)for some a by 64.12. The splitting field E of the minimal polynomial of a overF is a Galois extension of F of finite degree; see 64.10. We may assume thatE ⊆ Q�

p; then Qp = F (a) ⊆ E. Result 56.15 gives a (unique) extension w ofthe p-adic valuation of Qp to a principal valuation of E, and 56.13 shows thatall field automorphisms of E are isometries and therefore continuous with re-spect to w. Hence GalF E fixes each element of Q and of its topological closureQp, whence Qp ⊆ F . (See also Ribenboim 1999 6.2.K p. 167)

(9) The quaternion skew field H contains elements i, j with i2 = j2 = −1 andij = −ji. Let R be a valuation ring of H. Then R contains all elements ofH× of order 4 (like i), hence all a ∈ H with a + a = 0 and aa = 1. Theseelements form a 2-sphere in A := {a ∈ H | a + a = 0} ∼= R3, hence theygenerate A additively. We conclude that the ring R contains A and iA = A,hence R ⊇ A + iA = H, a contradiction.

If v is an extension of vp to the rational quaternions, then i, j are units ofthe valuation ring R of v. Since p belongs to the maximal ideal M of R, theimage of the ring Z + Zi + Zj + Zij in the residue skew field R/M is a finitesubring of R/M , hence a finite skew field. By Wedderburn’s theorem (compareCohn 2003a 7.8.6 or Jacobson 1985 7.7), that image is commutative. Thusthe difference of 1 and −1 belongs to M , hence 2 ∈ M ∩ Z = pZ and p = 2.

It remains to show that the 2-adic valuation of Q extends to the rationalquaternions. In fact, an extension w of the 2-adic valuation v2 of Q2 to the 2-

Page 396: The Classical Fields

Hints and solutions 379

adic quaternions H2 := Q2 +Q2i+Q2j +Q2ij is given by w(x) = v2(xx)/2, as

we show now. The algebra H2 is a skew field, because its norm xx =P4

i=1 x2i

is anisotropic by Exercise 4 of Section 54. The mapping w : H×2 → Q+ is

a group homomorphism. For every a ∈ H2, the restriction of w to the fieldQ2(a) is a valuation of Q2(a) by 56.15; in particular, w(1+a) ≥ min{w(a), 0}.This implies that w is a valuation (compare the proof of 56.5).

57 Topologies of valuation type

(1) One has field topologies by 13.2. If U = {a ∈ F | v(a) > γ }, then(F � U)−1 = {a ∈ F | v(a) ≥ −γ } is bounded; similar arguments apply toabsolute values and ordered fields.

(2) As F is a field, each set aU with a ∈ F× is a neighbourhood of 0. For anyneighbourhood V of 0, there exists a ∈ F× with aU ⊆ V by 57.2(iii).

(3) The Hausdorff property holds for τ∞ precisely if the points 0 and ∞ areseparated by τ∞; this is equivalent to the existence of a neighbourhood of 0which is bounded and closed. By 13.4, τ is regular, hence each neighbourhoodcontains a closed neighbourhood.

Let U ∈ τ be an open neighbourhood of 0. The preimage {∞}∪(U �{0})−1

under inversion belongs to τ∞ precisely if (F � U)−1 is contained in somebounded (closed) subset of F . This means that the closed set (F � U)−1 isbounded; hence inversion is continuous at ∞ precisely if τ is of type V .

(4) Let τp be the p-adic topology of Q, and let Bp = {a ∈ Q | |a|p < 1}.Then τX is the supremum of the topologies τp with p ∈ X; this means thatS{ τp | p ∈ X } is a subbasis of τX . Hence each τX is a field topology, and τp

is of type V by Exercise 1. If X is finite, thenT

p∈X Bp belongs to τX and is

bounded (by 57.2(iii)), hence τX is locally bounded.Now let X be infinite. The sets Un :=

T

p∈X,p≤n pn−1Bp with n ∈ N form aneighbourhood base at 0 of τX . We show that none of the sets Un is bounded(hence τX is not locally bounded), using 57.2(iii). Choose q ∈ X with q > n.If a ∈ Q×, then x = |a|q

Q

p∈X,p≤n pn ∈ Un and |ax|q = 1, which shows thataUn ⊆ Bq.

Let X contain distinct primes p, q. The balls Bp and Bq belong to τX andcontain 0, and (Q � Bp)

−1 = {x ∈ Q | |x|p ≤ 1}. If a ∈ Q×, then |a|q hasp-adic absolute value 1, and a|a|q ∈ a(Q � Bp)

−1 has q-adic absolute value 1,hence a(Q � Bp)

−1 ⊆ Bq. This shows that τX is not of type V ; see 57.2(iii).Note that τ{2,3} is the topology of the diagonal D in 44.12.

(5) If F is of type V and if W is a neighbourhood of 0, then we find aneighbourhood U of 0 with U(F � W )−1 ⊆ W . Now xy ∈ U and y ∈ F � Wimply that x = xyy−1 ∈ U(F � W )−1 ⊆ W .

Conversely, assume that for every neighbourhood W of 0 there exists aneighbourhood U of 0 with xy ∈ U ⇒ x ∈ W ∨ y ∈ W . Then U(F � W )−1 ⊆W , because y ∈ F � W and x = (xy−1)y ∈ U imply xy−1 ∈ W . Hence(F � W )−1 is bounded.

(6) Let F be a topological field of type V . We verify the condition in 44.8for concentrated filterbases C on F which do not converge to 0. There existneighbourhoods W , W ′ of 0 with W /∈ C and W ′ + W ′ ⊆ W , and an elementB ∈ C with B−B ⊆ W ′. If x ∈ B∩W ′, then B = B−x+x ⊆ B−B+W ′ ⊆ W ,a contradiction to W /∈ C; thus B ∩ W ′ = ∅. Since F is of type V , the sets

Page 397: The Classical Fields

380 Hints and solutions

B−1 ⊆ (F � W ′)−1 are bounded, hence B−1 · B−1 is bounded by 57.2(ii).Hence for every neighbourhood U of 0 there exists a neighbourhood U ′ of0 with U ′ · B−1 · B−1 ⊆ U , and then an element C ∈ C with C ⊆ B andC−C ⊆ U ′. This yields C−1−C−1 ⊆ (C−C)·C−1 ·C−1 ⊆ U ′ ·B−1 ·B−1 ⊆ U ;hence C−1 is concentrated.

We remark that the completion bF is again of type V ; compare Warner1989 Theorem 19.12.

(7) One checks that Ra is a subring of F ; for example, if x, y ∈ Ra and n ∈ N,then nv(xy) = nv(x) + nv(y) ≥ 2n min{v(x), v(y)} = min{2nv(x), 2nv(y)} >−v(a). Moreover, Ra contains the valuation ring R of v, and a−1 /∈ Ra. HenceRa is a valuation ring of F , and R and Ra induce the same valuation topologyon F by 56.10.

It remains to show that Ra is a maximal subring of F ; then 56.8 impliesthat the value group determined by Ra is Archimedean. Let b ∈ F �Ra. Thenmv(b) + v(a) ≤ 0 for some m ∈ N, hence v(b−ma−1) ≥ 0 and b−ma−1 ∈ R.We infer that a−n = (bm(b−ma−1))n belongs to the ring R[b] generated by Rand b for every n ∈ N, hence R[b] ⊇ S

n∈Na−nR = F and therefore Ra[b] = F .

58 Local fields and locally compact fields

(1) The polynomial f(x + 1) =Pp

i=1

`

pi

´

xi−1 is an Eisenstein polynomial (seeSchonemann 1846). Reducing the coefficients modulo p gives the polynomialf = xp−1 ∈ Fp[x]. Hence each non-trivial factor of f in Zp[x] has its constantterm in pZp. Since f(0+1) = p, this implies that f is irreducible in Zp[x]. Theirreducibility in Qp[x] is then a consequence of the Gauss Lemma; compareCohn 2003a 7.7.2, Jacobson 1985 2.16 Lemma 2 or Lang 1993 IV §2.

(2) Let α be the generator of the Galois group of F |Qp. By 56.15 the p-adic valuation vp of Qp has a unique extension w to F which is given byw(x) = vp(xxα)/2 for all x ∈ F . Thus we have w(F ) = vp(Qp) = Z preciselyif vp(xxα) is even for all x ∈ F . Writing x = a+ b

√c, we obtain the condition

that vp(a2 − b2c) is even for a, b ∈ Qp. This is true for b = 0, and for b = 0 wehave vp(a2 − b2c) = vp((a/b)2 − c) + 2vp(b).

If p = 2 and c = 3, then v2(1 − 3) = 1, hence Q2(√

3)|Q2 is ramified.

Let p = 2 and c = 5. We claim that Q2(√

5)|Q2 is unramified. Otherwisev2(a

2 − 5) would be odd for some a ∈ Q2. If v2(a) = 0, then v2(a2 − 5) =

2 min{v2(a), 0} is even (see 56.1). Hence v2(a) = 0, which means that a ∈1 + 2Z2. Therefore a2 − 5 ∈ 1 − 5 + 8Z2 and v2(a

2 − 5) = 2, a contradiction.Alternatively, we observe that −3 · 5 = −15 ∈ 1 + 8Z2 is a square in Q2 by

54.1, hence Q2(√

5) = Q2(√−3) = Q2((−1+

√−3)/2). Since (−1+√−3)/2 is

a root of unity of order 3 = 22 − 1, the field Q2(√

5) is the unique unramifiedquadratic extension of Q2; see 58.2.

(3) Each polynomial x2 + tnx + t is separable over F = F2((t)), with rootsan, bn = ta−1

n in a fixed algebraic closure of F . By exchanging an and bn wecan achieve that an belongs to the natural compact valuation ring of F (an) =F (an, bn); see 56.15. If there were only finitely many extension types F (an)|F ,then for some k ∈ N the field F (ak) would contain an for infinitely many n ∈ N.Then some subsequence of (an)n has a limit a ∈ F (ak). Since a2

n+tnan+t = 0for n ∈ N and limn→∞ tn = 0, we infer that a2 + t = 0. Thus F (a) is aninseparable quadratic extension of F contained in F (ak). Hence F (a) = F (ak),a contradiction, as F (ak)|F is separable.

Page 398: The Classical Fields

Hints and solutions 381

(4) From 54.1 it follows that 17 = 1 + 8 · 2 is a square in Q2 (and in F2), and2 = 62(1 − 17 · 2 · 6−2) is a square in Q17 (and in F17). If p = 2, 17, then atleast one of the numbers 2, 17, 34 = 2 · 17 is a square in Fp (since the squaresof F×

p form a subgroup of index 2; see Exercise 1 of Section 64) and hence alsoin Qp (by Exercise 2 of Section 54).

(5) The mapping x + yb �→„

x cyγ

y xγ

«

, where x, y ∈ E, is a Z-algebra isomor-

phism as required. Direct computation shows that the matrices of the shape

A =

x cyγ

y xγ

«

form a Z-subalgebra of E2×2. Since det A = xxγ − cyyγ , the

non-zero matrices of this shape are invertible precisely if c = xxγ for all x ∈ E.The Cayley–Hamilton theorem implies that A−1 det A = (x+xγ)I −A, hencethe inverses have the same shape, as x + xγ , det A ∈ Z.

(6) In the notation of 58.10, one has Zn = Fqn and aγr

= aqr

for a ∈ Fqn ;compare 58.2. Substituting t for b in the definition of [Zn|Z; γr, x] gives anisomorphism as required.

61 Ordinals and cardinals

(1) Consider the set of elements with the smallest first coordinate.

(2) Countable subsets of R can be described by maps N → R, and card RN =ℵℵ0 = ℵ by 61.13b.

(3) One may assume that Aι ⊂ Bι and Bι ∩ Bκ = ∅ for ι = κ. Consider any

map ϕ :S

ι Aι →×ιBι : s �→ ϕs and choose ψ(ι) ∈ Bι � {ϕs(ι) | s ∈ Aι }.

Then ψ = ϕs for each s, and ϕ is not surjective.

(4) We have ℵω =P

ν∈ω ℵν <Q

ν∈ω ℵν ≤ ℵℵ0ω . With ℵω = 2ℵ0 this leads to

a contradiction.

62 Topological groups

(1) If inversion is uniformly continuous, then for each neighbourhood U of 1there is a symmetric neighbourhood S such that yx−1 ∈ S implies x−1y ∈ U .Hence S ⊆ xUx−1, and this true for each x. Therefore, V :=

T

x xUx−1 is aninvariant neighbourhood of 1.

If V is invariant and V 2 ⊆ U , then V aV b ⊆ Uab for all a and b, and this isequivalent with the assertion.

(2) Consider a symmetric (compact) neighbourhood V of 1 such that H ∩ V 2

is compact. If x ∈ H, then H ∩ xV = ∅ and hence ∅ = Hx ∩ V ⊆ H ∩ V ⊆H ∩ V 2 ⊆ H.

(3) If U is a compact open neighbourhood of 1, then by the usual compactnessarguments, there is a neighbourhood V of 1 such that UV ⊆ U . This impliesthat V generates an open (and hence compact) subgroup of U .

63 Locally compact abelian groups and Pontryagin duality

(1) Since B → C is surjective, the adjoint C∗ → B∗ is injective. In fact,C∗ ∼= A⊥ ≤ B∗. Hence there is an exact sequence 0 → A⊥ → B∗ → X → 0.The first argument shows that X∗ ∼= A⊥ = A, and reflexivity implies thatX ∼= A∗.

Page 399: The Classical Fields

382 Hints and solutions

(2) As a rational vector space, D has a basis b of cardinality ℵ. The values of acharacter of D can be chosen arbitrarily on b. Hence card D∗ = ℵℵ = 2ℵ > ℵ.On the other hand, R∗ ∼= R has cardinality ℵ.

(3) The dual A∗ is discrete by 63.5. Hence A∗ is torsion free if, and only if, eachcompact subgroup of A∗ is trivial. The claim follows now from 63.30–63.32.

64 Fields

(1) Let d be a divisor of |G| = n. The polynomial xd − 1 has at most d roots.If G contains an element of order d, then the set {g ∈ G | gd = 1} has size d,hence the number of elements in G of order d is given by the Euler functionϕ(d). Thus ψ(d) := card{g ∈ G | g has order d} ∈ {0, ϕ(d)} for all divisors dof n. Now n =

P

d ψ(d) ≤ P

d ϕ(d), and consideration of a cyclic group showsthat the last sum has the value n. We conclude that ψ = ϕ, in particularψ(n) = ϕ(n) = 0, whence G is cyclic. (This proof is due to Gauß.)

(2) The group Γ acts on E[t] by acting on the coefficients of polynomials, fixingprecisely the polynomials in F [t]. Therefore f =

Q

b∈B(x − b) has coefficientsin F . Clearly f(a) = 0. If g ∈ F [t] satisfies g(a) = 0, then x − a divides g inE[t]. Applying the elements of Γ we see that x − b divides g for each b ∈ B,whence f divides g. This shows that f is the minimal polynomial of a over F .

(3) By Eisenstein’s criterion (compare Schonemann 1846), f is irreduciblein Q[x]. For n = −3, 0, 1, 2, the signs of f(n) alternate, hence f has at leastthree real roots. The sum of the squares of the five roots xk of f is 0, becauseP

k xk = 0 =P

h�=k xhxk and thereforeP

k x2k = 0. Hence f has two roots

z, z ∈ C which are not real. Complex conjugation gives an element τ in theGalois group Γ = GalQE which acts as the transposition (z, z) on the set ofroots of f . By Exercise 2, Γ is transitive on the five roots. Hence Γ containsan element σ of order 5, which acts as a 5-cycle on the roots. Together, σ andτ generate the symmetric group of degree 5.

(4) By 61.16, the field F has at most 2card F automorphisms. On the otherhand, the groups Aut Q� and Aut F �

p have cardinality 2ℵ0 ; see 52.10 and 64.15.Any field is an algebraic extension of some purely transcendental field Q(T ) orFp(T ); see 64.20. If F is countable, then the extension result 64.15 implies thatF has at least 2ℵ0 field automorphisms. If F is uncountable, then card F =card T ; see 64.20. There are 2card T permutations of T by 61.16, and by 64.15each of these permutations extends to a field automorphism of F .

(5) Assume that (t − a)−1 is an F -linear combination of rational functions(t − bi)

−1 with bi ∈ F � {a}. Then (t − a)−1 = f/Q

i(t − bi) with f ∈ F [t],hence t− a divides

Q

i(t− bi) in F [t], which is a contradiction to the fact thatF [t] is a unique factorization domain.

(6) The field of fixed elements of the Frobenius automorphism ϕ is the primefield Fp, hence Aut Fq = GalFpFq has order n; compare 64.18. By Exercise 1,ϕ has order n, hence Aut Fq = 〈ϕ〉.

Page 400: The Classical Fields

References

Numbers at the end of an entry indicate the pages where the entry is quoted.

U. Abel 1980. On the group of piecewise linear monotone bijections of anarc. Math. Z. 171, 155–161. 64

U. Abel 1982. A homeomorphism of Q with Q as an orbit.Elem. Math. 37, 108–109. 205

U. Abel and J. Misfeld 1990. Congruences in decompositions of the realand rational numbers. Rend. Mat. Appl. (7 ) 10, 279–285. 64

N. A’Campo 2003. A natural construction for the real numbers.Preprint arXiv math GN/0301015. 2

J. Aczel 1966. Lectures on Functional Equations and their Applications.New York: Academic Press. 6

J. F. Adams 1969. Lectures on Lie Groups.New York: Benjamin. 67, 90

P. Alexandroff 1924. Uber die Metrisation der im Kleinen kompaktentopologischen Raume. Math. Ann. 92, 294–301. 48

N.L. Alling 1962. On exponentially closed fields.Proc. Amer. Math. Soc. 13, 706–711. 117, 315

N.L. Alling 1987. Foundations of Analysis over Surreal Number Fields.Amsterdam: North-Holland. 1

N.L. Alling and S. Kuhlmann 1994. On ηα-groups and fields.Order 11, 85–92. 78, 129

C. Alvarez 1999. On the history of Souslin’s problem.Arch. Hist. Exact Sci. 54, 181–242. 27

R.D. Anderson 1958. The algebraic simplicity of certain groups of homeo-morphisms. Amer. J. Math. 80, 955–963. 54, 206

G.E. Andrews, S.B. Ekhad, and D. Zeilberger 1993. A short proof ofJacobi’s formula for the number of representations of an integer as a sumof four squares. Amer. Math. Monthly 100, 274–276. 213

R.D. Arthan 2004. The Eudoxus real numbers.Preprint arXiv math HO/0405454. 2

H. Bachmann 1967. Transfinite Zahlen.Berlin: Springer. Second edition. 337, 338

R. Baer 1970a. Dichte, Archimedizitat und Starrheit geordneter Korper.Math. Ann. 188, 165–205. 119

383

Page 401: The Classical Fields

384 References

R. Baer 1970b. Die Automorphismengruppe eines algebraisch abgeschlossen-en Korpers der Charakteristik 0. Math. Z. 117, 7–17. 150

R. Baer and H. Hasse 1932. Zusammenhang und Dimension topologischerKorperraume. J. Reine Angew. Math. 167, 40–45. 146

J.C. Baez 2002. The octonions.Bull. Amer. Math. Soc. (N.S.) 39, 145–205. xii

A. Baker 1975. Transcendental Number Theory.Cambridge: Cambridge University Press. (Reprinted 1990) 201, 221

A. Baker 1984. A Concise Introduction to the Theory of Numbers.Cambridge: Cambridge University Press. 201

J.A. Baker, C.T. Ng, J. Lawrence, and F. Zorzitto 1978. Sequencetopologies on the real line. Amer. Math. Monthly 85, 667–668. 69

B. Banaschewski 1957. Uber die Vervollstandigung geordneter Gruppen.Math. Nachr. 16, 51–71. 264

B. Banaschewski 1998. On proving the existence of complete ordered fields.Amer. Math. Monthly 105, 548–551. 248

H. Bauer 1968. Wahrscheinlichkeitstheorie und Grundzuge der Masstheorie.Berlin: de Gruyter. (Translation: Probability Theory and Elements ofMeasure Theory. New York: Academic Press 1981) 202

J. E. Baumgartner and K. Prikry 1977. Singular cardinals and the gener-alized continuum hypothesis. Amer. Math. Monthly 84, 108–113. 339

A.F. Beardon 2001. The geometry of Pringsheim’s continued fractions.Geom. Dedicata 84, 125–134. 31

E. Becker 1974. Euklidische Korper und euklidische Hullen von Korpern.J. Reine Angew. Math. 268/269, 41–52. 125

M. Benito and J. J. Escribano 2002. An easy proof of Hurwitz’s theorem.Amer. Math. Monthly 109, 916–918. 30

C.D. Bennett 1997. Explicit free subgroups of Aut(R,≤). Proc. Amer.Math. Soc. 125, 1305–1308. 64

E. Beth and A. Tarski 1956. Equilaterality as the only primitive notionof Euclidean geometry. Indag. Math. 18 (Nederl. Akad. Wetensch. Proc.Ser. A 59), 462–467. 153

M. Bhargava 2000. On the Conway–Schneeberger fifteen theorem.In: Quadratic Forms and their Applications (Dublin, 1999), vol. 272 ofContemporary Mathematics, pages 27–37. Providence, RI: AmericanMathematical Society. 213

G. Birkhoff 1948. Lattice Theory. Colloquium Publication 25. New York:American Mathematical Society. Second revised edition. 25, 26, 27, 164

G.D. Birkhoff and H. S. Vandiver 1904. On the integral divisors of an−bn.Ann. Math. (2 ) 5, 243–252. 231

A. Blass and J.M. Kister 1986. Free subgroups of the homeomorphismgroup of the reals. Topology Appl. 24, 243–252. 64

T. S. Blyth 2005. Lattices and Ordered Algebraic Structures.London: Springer. 77, 246

J. Bochnak, M. Coste, and M.-F. Roy 1998. Real Algebraic Geometry.Berlin: Springer. 122, 133, 134

Z. I. Borevich and I. R. Shafarevich 1966. Number Theory. New York:Academic Press. (Third Russian edition: Moscow: Nauka 1985. Germantranslation: Basel: Birkhauser 1966) 284, 297, 323, 325, 363, 374

N. Bourbaki 1966. General Topology. Paris: Hermann.(Reprints: Berlin: Springer 1989, 1998) 67, 164, 340, 341, 342

Page 402: The Classical Fields

References 385

N. Bourbaki 1972. Commutative Algebra. Paris: Hermann. (Second edition:Berlin: Springer 1989) 146, 304, 306, 326, 330, 377, 378

N. Bourbaki 1990. Algebra II.Chapters 4–7. Berlin: Springer. (Reprinted 2003) 290, 350, 354, 355

G.E. Bredon 1963. A new treatment of the Haar integral.Michigan Math. J. 10, 365–373. 326

H. Brenner 1992. Ein uberabzahlbares, uber Q linear unabhangiges Systemreeller Zahlen. Math. Semesterber. 39, 89–93. 7

N.G. de Bruijn 1976. Defining reals without the use of rationals. Indag.Math. 38 (Nederl. Akad. Wetensch. Proc. Ser. A 79), 100–108. 1

E.B. Burger and T. Struppeck 1996. DoesP∞

n=0 1/n! really converge?Infinite Series and p-adic Analysis.Amer. Math. Monthly 103, 565–577. 284

G. Cantor 1895. Beitrage zur Begrundung der transfiniten Mengenlehre.Erster Artikel. Math. Ann. 46, 481–512. (Zweiter Artikel: Math. Ann.49 1897, 207–246) 25

J.W. S. Cassels 1957. An Introduction to Diophantine Approximation.Cambridge: Cambridge University Press. 30

J.W. S. Cassels 1978. Rational Quadratic Forms.London: Academic Press. x, 297, 299, 325

J.W. S. Cassels 1986. Local Fields.Cambridge: Cambridge University Press. 285, 289, 304, 315, 323, 325

C.C. Chang and H. J. Keisler 1990. Model Theory. Amsterdam: North-Holland. Third edition. 156, 170

M.D. Choi and T.Y. Lam 1977. Extremal positive semidefinite forms.Math. Ann. 231, 1–18. 134

C.O. Christenson and W.L. Voxman 1977. Aspects of Topology.New York: Dekker. (Revised second edition: Moscow, ID: BCS Associates1998) 40, 42, 43, 50, 51, 53, 55, 283

J.C. Ch’uan and L. Liu 1981. Group topologies on the real line. J. Math.Anal. Appl. 81, 391–398. 69, 89

K. Ciesielski 1997. Set Theory for the Working Mathematician. Cambridge:Cambridge University Press. 337

L.W. Cohen and G. Ehrlich 1963. The Structure of the Real Number Sys-tem. Princeton, NJ: Van Nostrand. xii

P. J. Cohen 1966. Set Theory and the Continuum Hypothesis. New York:Benjamin. 339

S.D. Cohen and A.M.W. Glass 1997. Free groups from fields. J. LondonMath. Soc. (2 ) 55, 309–319. 64

D.L. Cohn 1993. Measure Theory.Boston, MA: Birkhauser. 104

P.M. Cohn 1995. Skew Fields. Theory of General Division Rings.Cambridge: Cambridge University Press. 330, 332

P.M. Cohn 2003a. Basic Algebra. London: Springer.16, 20, 131, 133, 145, 285, 289, 293, 306, 315, 323, 333, 350, 352, 353,354, 355, 362, 367, 377, 378, 380

P.M. Cohn 2003b. Further Algebra and Applications.London: Springer. 332, 333, 368

W.W. Comfort and S. Negrepontis 1974. The Theory of Ultrafilters.New York: Springer. 155

Page 403: The Classical Fields

386 References

A. Connes 1998. A new proof of Morley’s theorem. In: Les relations entre lesmathematiques et la physique theorique, pages 43–46. Bures-sur-Yvette:

Institut des Hautes Etudes Scientifiques. 151M. Contessa, J. L. Mott, and W. Nichols 1999. Multiplicative groups of

fields. In: Advances in Commutative Ring Theory (Fez, 1997), vol. 205of Lecture Notes in Pure and Applied Mathematics, pages 197–216. NewYork: Dekker. 17, 19, 72

J.H. Conway 1976. On Numbers and Games. London: Academic Press.(Second edition: Natick, MA: A K Peters 2001) 1

J.H. Conway 1997. The Sensual (Quadratic) Form. Washington, DC: Math-ematical Association of America. 212, 279, 297

J.H. Conway 2000. Universal quadratic forms and the fifteen theorem.In: Quadratic Forms and their Applications (Dublin, 1999), vol. 272 ofContemporary Mathematics, pages 23–26. Providence, RI: AmericanMathematical Society. 213

J.H. Conway and D.A. Smith 2003. On Quaternions and Octonions: theirGeometry, Arithmetic, and Symmetry. Natick, MA: A K Peters. xii, 211

L. Corwin 1976. Uniqueness of topology for the p-adic integers.Proc. Amer. Math. Soc. 55, 432–434. 290

H. Dales and W. Woodin 1996. Super-real Fields. Totally Ordered Fieldswith Additional Structure.Oxford: Oxford University Press. 1, 166, 169, 246, 359

G. Darboux 1880. Sur le theoreme fondamental de la geometrie projective.Math. Ann. 17, 55–61. 74

R. Dedekind 1872. Stetigkeit und irrationale Zahlen. Braunschweig: Vieweg.(Sixth edition 1960. Translation: Essays on the Theory of Numbers. NewYork: Dover Publications 1963) xii

O. Deiser 2005. Der Multiplikationssatz der Mengenlehre.Jahresber. Deutsch. Math.-Verein. 107, 88–109. 338

B. Deschamps 2001. A propos d’un theoreme de Frobenius.Ann. Math. Blaise Pascal 8, 61–66. 129

K. J. Devlin and H. Johnsbraten 1974. The Souslin Problem, vol. 405 ofLecture Notes in Mathematics. Berlin: Springer. 27

J. Dieudonne 1944. Sur la completion des groupes topologiques.C. R. Acad. Sci. Paris 218, 774–776. 263

J. Dieudonne 1945. Sur les corps topologiques connexes.C. R. Acad. Sci. Paris 221, 396–398. 146

D. Dikranjan 1979. Topological characterization of p-adic numbers and anapplication to minimal Galois extensions.Ann. Univ. Sofia Fac. Math. Mec. 73, 103–110. 287

J.D. Dixon and B. Mortimer 1996. Permutation Groups.New York: Springer. 72

E.K. van Douwen and W.F. Pfeffer 1979. Some properties of the Sor-genfrey line and related spaces. Pacific J. Math. 81, 371–377. 69

E.K. van Douwen and H.H. Wicke 1977. A real, weird topology on thereals. Houston J. Math. 3, 141–152. 69

M. Droste and R. Gobel 2002. On the homeomorphism groups of Cantor’sdiscontinuum and the spaces of rational and irrational numbers.Bull. London Math. Soc. 34, 474–478. 54

M. Dugas and R. Gobel 1987. All infinite groups are Galois groups over anyfield. Trans. Amer. Math. Soc. 304, 355–384. 75

Page 404: The Classical Fields

References 387

J. Dugundji 1966. Topology. Boston, MA: Allyn and Bacon.32, 57, 58, 90, 164, 200, 202, 335, 336, 337, 338, 342, 343, 345

A. Durand 1975. Un systeme de nombres algebriquement independents.C. R. Acad. Sci. Paris Ser. A 280, 309–311. 7

H. Ebbinghaus, H. Hermes, F. Hirzebruch, M. Koecher, K. Mainzer,J. Neukirch, A. Prestel, and R. Remmert 1991. Numbers.New York: Springer. (Translation of: Zahlen. Berlin: Springer 1988)

xii, 128, 138, 144, 151, 211, 216, 278, 304, 315, 330P. Ehrlich (ed.) 1994. Real Numbers, Generalizations of the Reals, and

Theories of Continua, vol. 242 of Synthese Library. Dordrecht: Kluwer.xii, 1

P. Ehrlich 2001. Number systems with simplicity hierarchies: a general-ization of Conway’s theory of surreal numbers. J. Symbolic Logic 66,1231–1258. Corrigendum: J. Symbolic Logic 70 2005, 1022. 1, 76, 124

R. Ellis 1957. Locally compact transformation groups.Duke Math. J. 24, 119–125. 340

C. Elsholtz 2003. Kombinatorische Beweise des Zweiquadratesatzes undVerallgemeinerungen. Math. Semesterber. 50, 77–93. 209

C. Elsner 2000. Uber eine effektive Konstruktion großer Mengen algebraischunabhangiger Zahlen. Math. Semesterber. 47, 243–256. 7

R. Engelking 1969. On closed images of the space of irrationals.Proc. Amer. Math. Soc. 21, 583–586. 111

R. Engelking 1978. Dimension Theory.Amsterdam: North-Holland. 197

A. J. Engler and A. Prestel 2005. Valued Fields.Berlin: Springer. 184, 279, 304, 306, 359, 377, 378

D.B.A. Epstein 1970. The simplicity of certain groups of homeomorphisms.Compositio Math. 22, 165–173. 64

P. Erdos 1940. The dimension of the rational points in Hilbert space.Ann. Math. (2 ) 41, 734–736. 197

P. Erdos, L. Gillman, and M. Henriksen 1955. An isomorphism theoremfor real-closed fields. Ann. Math. (2 ) 61, 542–554. 167

D.M. Evans and D. Lascar 1997. The automorphism group of the field ofcomplex numbers is complete. In: Model Theory of Groups and Auto-morphism Groups (Blaubeuren, 1995), vol. 244 of London MathematicalSociety Lecture Notes, pages 115–125. Cambridge: Cambridge UniversityPress. 150

F. Faltin, N. Metropolis, B. Ross, and G.-C. Rota 1975. The real num-bers as a wreath product. Adv. Math. 16, 278–304. 1, 374

A. Fedeli and A. Le Donne 2001. The Sorgenfrey line has a locally pathwiseconnected connectification. Proc. Amer. Math. Soc. 129, 311–314. 69

S. Feferman 1964. The Number Systems. Foundations of Algebra and Anal-ysis. Reading, MA: Addison-Wesley. xii

U. Felgner 1976. Das Problem von Souslin fur geordnete algebraische Struk-turen. In: Set Theory and Hierarchy Theory (Bierutowice, 1975), vol. 537of Lecture Notes in Mathematics, pages 83–107. Berlin: Springer. 27

W. Felscher 1978/79. Naive Mengen und abstrakte Zahlen I.Naive Mengen und abstrakte Zahlen II: Algebraische und reelle Zahlen.Naive Mengen und abstrakte Zahlen III: Transfinite Methoden.Mannheim: Bibliographisches Institut. xii

Page 405: The Classical Fields

388 References

B. Fine and G. Rosenberger 1997. The Fundamental Theorem of Algebra.New York: Springer. 128, 144

N. J. Fine and G.E. Schweigert 1955. On the group of homeomorphismsof an arc. Ann. Math. (2 ) 62, 237–253. 58, 63, 64

G. Flegg 1983. Numbers. Their History and Meaning.New York: Schocken Books. xii, 1

L. Fuchs 1970. Infinite Abelian Groups, vol. I.New York: Academic Press. 184, 186

L. Fuchs 1973. Infinite Abelian Groups, vol. II.New York: Academic Press. 96, 207, 372

H. Geiges 2001. Beweis des Satzes von Morley nach A. Connes.Elem. Math. 56, 137–142. 151

M. Gerstenhaber and C.T. Yang 1960. Division rings containing a realclosed field. Duke Math. J. 27, 461–465. 129

L. Gillman 2002. Two classical surprises concerning the axiom of choice andthe continuum hypothesis. Amer. Math. Monthly 109, 544–553. 339

L. Gillman and M. Jerison 1960. Rings of Continuous Functions.Princeton, NJ: Van Nostrand. (Reprint: New York: Springer 1976) 118

A.M.W. Glass 1981. Ordered Permutation Groups.Cambridge: Cambridge University Press. 27, 205

A.M.W. Glass and P. Ribenboim 1994. Automorphisms of the orderedmultiplicative group of positive rational numbers. Proc. Amer. Math.Soc. 122, 15–18. 221

W. Glatthaar 1971. Unterebenen und Kollineationen der komplexen Ebene.Zulassungsarbeit, University of Tubingen. 152

A.M. Gleason and R. S. Palais 1957. On a class of transformation groups.Amer. J. Math. 79, 631–648. 66

K. Godel 1940. The Consistency of the Axiom of Choice and of the Gener-alized Continuum Hypothesis, vol. 3 of Annals of Mathematics Studies.Princeton, NJ: Princeton University Press. 339

R. Goldblatt 1998. Lectures on the Hyperreals.New York: Springer. 154

S.W. Golomb 1959. A connected topology for the integers.Amer. Math. Monthly 66, 663–665. 45

H. Gonshor 1986. An Introduction to the Theory of Surreal Numbers. Cam-bridge: Cambridge University Press. 1

R.A. Gordon 1994. The Integrals of Lebesgue, Denjoy, Perron, and Hen-stock. Providence, RI: American Mathematical Society. 202

C. Gourion 1992. A propos du groupe des automorphismes de (Q;≤).C. R. Acad. Sci. Paris Ser. I Math. 315, 1329–1331. 204

F.Q. Gouvea 1997. p-Adic Numbers: an Introduction. Berlin: Springer.Second edition. 278, 283, 284

M. J. Greenberg 1967. Lectures on Algebraic Topology.New York: Benjamin. 86, 87

M. J. Greenberg 1969. Lectures on Forms in Many Variables.New York: Benjamin. 279

G.R. Greenfield 1983. Sums of three and four integer squares.Rocky Mountain J. Math. 13, 169–175. 212

E. Grosswald 1985. Representations of Integers as Sums of Squares.New York: Springer. 212, 213, 215

Page 406: The Classical Fields

References 389

T. Grundhofer 2005. Describing the real numbers in terms of integers.Arch. Math. (Basel) 85, 79–81. 2

T. Grundhofer and H. Salzmann 1990. Locally compact double loops andternary fields. In: Quasigroups and Loops: Theory and Applications (edsO. Chein, H. O. Pflugfelder and J. D. H. Smith), Chapter XI, pages 313–355. Berlin: Heldermann. 139, 375

V. S. Guba 1986. A finitely generated complete group (Russian).Izv. Akad. Nauk SSSR Ser. Mat. 50, 883–924. Translated in Math. USSRIzv. 29, 233–277. 5

Y. Gurevich and W.C. Holland 1981. Recognizing the real line.Trans. Amer. Math. Soc. 265, 527–534. 27

J.A. Guthrie, H.E. Stone, and M.L. Wage 1978. Maximal connectedexpansions of the reals. Proc. Amer. Math. Soc. 69, 159–165. 69

L. Hahn 1994. Complex Numbers and Geometry. Washington, DC: Mathe-matical Association of America. 151

H. Halberstam 1974. Transcendental numbers.Math. Gaz. 58, 276–284. 221

P.R. Halmos 1944. Comment on the real line.Bull. Amer. Math. Soc. 50, 877–878. 99

P.R. Halmos 1950. Measure Theory. New York: Van Nostrand.(Reprint: New York: Springer 1974) 104, 202, 326

F. Halter-Koch 1982. Darstellung naturlicher Zahlen als Summe vonQuadraten. Acta Arith. 42, 11–20. 212

G. Hamel 1905. Eine Basis aller Zahlen und die unstetigen Losungen derFunktionalgleichung: f(x + y) = f(x) + f(y).Math. Ann. 60, 459–462. 6

G.H. Hardy and E.M. Wright 1971. An Introduction to the Theory ofNumbers. Oxford: Oxford University Press. Fourth edition. (Fifth edition1979) 28, 159, 192, 201, 213

W.S. Hatcher and C. Laflamme 1983. On the order structure of the hy-perreal line. Z. Math. Logik Grundlag. Math. 29, 197–202. 169

F. Hausdorff 1914. Grundzuge der Mengenlehre. Leipzig: Veit & Comp.(Reprint: New York: Chelsea 1949) 25, 166, 167

F. Hausdorff 1935. Mengenlehre. Berlin: de Gruyter. Third edition. (Trans-lation: Set Theory. New York: Chelsea 1957) 109, 110

U. Heckmanns 1991. Beispiel eines topologischen Korpers, dessen Vervoll-standigung ein Integritatsring, aber kein Korper ist.Arch. Math. (Basel) 57, 144–148. 268

I. N. Herstein 1975. Topics in Algebra.Lexington, MA: Xerox. Second edition. 213

E. Hewitt and K.A. Ross 1963. Abstract Harmonic Analysis, vol. I.Springer: Berlin. 90, 99, 283, 284, 326, 340, 341, 342, 344, 345, 349, 375

D. Hilbert 1903. Grundlagen der Geometrie. Leipzig: Teubner. Second edi-tion. (14th edition 1999. Translation: Foundations of Geometry. Chicago:Open Court 1988) 125

D. Hilbert and S. Cohn-Vossen 1932. Anschauliche Geometrie. Berlin:Springer. (Reprinted 1996. Translation: Geometry and the Imagination.New York: Chelsea 1952) 215

M.D. Hirschhorn 1987. A simple proof of Jacobi’s four-square theorem.Proc. Amer. Math. Soc. 101, 436–438. 213

Page 407: The Classical Fields

390 References

G. Hjorth and M. Molberg 2006. Free continuous actions on zero-dimen-sional spaces. Topology Appl. 153, 1116–1131. 54

E. Hlawka, J. Schoissengeier, and R. Taschner 1991. Geometric andAnalytic Number Theory. Berlin: Springer. 68

W. Hodges 1993. Model Theory.Cambridge: Cambridge University Press. 362

J. E. Hofmann 1958. Zur elementaren Dreiecksgeometrie in der komplexenEbene. Enseignement Math. (2 ) 4, 178–211. 151

K.H. Hofmann and S.A. Morris 1998. The Structure of Compact Groups.Berlin: de Gruyter. 66, 90, 99, 343, 344, 345, 346, 347, 348

K.H. Hofmann and C. Terp 1994. Compact subgroups of Lie groups andlocally compact groups. Proc. Amer. Math. Soc. 120, 623–634. 343

O. Holder 1901. Die Axiome der Quantitat und die Lehre vom Maß.Ber. Verh. Kgl. sachs. Ges. Wiss. Leipzig, math.-phys. Kl. 53, 1–64.English translation (in two parts): J. Mitchell and C. Ernst, J. Math.Psychology 40 1996, 235–252 and 41 1997, 345–356. 77

W.C. Holland 1992. Partial orders of the group of automorphisms of the realline. In: Proceedings of the International Conference on Algebra, Part 1(Novosibirsk, 1989), vol. 131 of Contemporary Mathematics, pages 197–207. Providence, RI: American Mathematical Society. 59

M. Holz, K. Steffens, and E. Weitz 1999. Introduction to Cardinal Arith-metic. Basel: Birkhauser. 337

A.E. Hurd and P.A. Loeb 1985. An Introduction to Nonstandard RealAnalysis. Orlando, FL: Academic Press. 154

W. Hurewicz and H. Wallman 1948. Dimension Theory. Princeton, NJ:Princeton University Press. Second revised edition. 200

A. Hurwitz 1898. Uber die Komposition der quadratischen Formen vonbeliebig vielen Variablen. Nachr. Ges. Wiss. Gottingen, 309–316. 216

T. Husain 1966. Introduction to Topological Groups.Philadelphia, PA: Saunders. 342

K. Iwasawa 1949. On some types of topological groups.Ann. Math. (2 ) 50, 507–558. 343

K. Jacobs 1978. Measure and Integral.New York: Academic Press. 110

N. Jacobson 1985. Basic Algebra. I. New York: Freeman. Second edition.16, 127, 130, 138, 145, 190, 330, 350, 352, 353, 378, 380

N. Jacobson 1989. Basic Algebra. II. New York: Freeman. Second edition.122, 133, 134, 167, 184, 303, 304, 306, 315, 322, 323, 326, 331, 332, 333,350, 353, 355, 366, 367, 368

F.B. Jones 1939. Concerning certain linear abstract spaces and simple con-tinuous curves. Bull. Amer. Math. Soc. 45, 623–628. 42

R.R. Kallman 1976. A uniqueness result for topological groups.Proc. Amer. Math. Soc. 54, 439–440. 290

R.R. Kallman and F.W. Simmons 1985. A theorem on planar continuaand an application to automorphisms of the field of complex numbers.Topology Appl. 20, 251–255. 147

I. Kapuano 1946. Sur les corps de nombres a une dimension distincts ducorps reel. Rev. Fac. Sci. Univ. Istanbul (A) 11, 30–39. 146

H. J. Keisler 1994. The hyperreal line.In: Ehrlich 1994, pages 207–237. 154

Page 408: The Classical Fields

References 391

B. von Kerekjarto 1931. Geometrische Theorie der zweigliedrigen kon-tinuierlichen Gruppen. Abh. Math. Sem. Hamburg 8, 107–114. 141

H. Kestelman 1951. Automorphisms of the field of complex numbers.Proc. London Math. Soc. (2 ) 53, 1–12. 148

A.Ya. Khinchin 1964. Continued Fractions. Chicago, IL: University ofChicago Press. 28, 30

J.O. Kiltinen 1973. On the number of field topologies on an infinite field.Proc. Amer. Math. Soc. 40, 30–36. 138, 166

A.M. Kirch 1969. A countable, connected, locally connected Hausdorff space.Amer. Math. Monthly 76, 169–171. 45

G. Klaas, C.R. Leedham-Green, and W. Plesken 1997. Linear Pro-p-Groups of Finite Width, vol. 1674 of Lecture Notes in Mathematics.Berlin: Springer. 324

H. Kneser 1960. Eine kontinuumsmachtige, algebraisch unabhangige Mengereeller Zahlen. Bull. Soc. Math. Belg. 12, 23–27. 7

H. Kneser and M. Kneser 1960. Reell-analytische Strukturen der Alexan-droff-Geraden und der Alexandroff-Halbgeraden.Arch. Math. (Basel) 11, 104–106. 48

N. Koblitz 1977. p-Adic Numbers, p-adic Analysis, and Zeta-Functions.New York: Springer. (Second edition 1984) 284

S. Koppelberg 1989. Handbook of Boolean Algebras, Vol. 1 (ed. J. D. Monk).Amsterdam: North-Holland. 362

H.-J. Kowalsky 1958. Kennzeichnung von Bogen.Fund. Math. 46, 103–107. 36, 42

L. Kramer 2000. Splitting off the real line and plane.Results Math. 37, 119. 142

F.-V. Kuhlmann, S. Kuhlmann, and S. Shelah 1997. Exponentiation inpower series fields. Proc. Amer. Math. Soc. 125, 3177–3183. 117

S. Kuhlmann 2000. Ordered Exponential Fields. Providence, RI: AmericanMathematical Society. 117

L. Kuipers and H. Niederreiter 1974. Uniform Distribution of Sequences.New York: Wiley. 68

M. Laczkovich 1998. Analytic subgroups of the reals. Proc. Amer. Math.Soc. 126, 1783–1790. 7, 112

T.Y. Lam 2005. Introduction to Quadratic Forms over Fields. Providence,RI: American Mathematical Society.

122, 125, 130, 133, 134, 216, 297, 298, 299, 325E. Landau 1930. Grundlagen der Analysis. Leipzig: Akademische Ver-

lagsgesellschaft. (Reprints: New York: Chelsea 1951, Darmstadt: Wis-senschaftliche Buchgesellschaft 1970, and Lemgo: Heldermann 2004.Translation: Foundations of Analysis. New York: Chelsea 1966) xii

S. Lang 1966. Introduction to Transcendental Numbers. Reading, MA:Addison-Wesley. 221

S. Lang 1970. Algebraic Number Theory. Reading, MA: Addison-Wesley.(Second edition: Berlin: Springer 1994) 279, 289, 323, 324, 374

S. Lang 1971. Transcendental numbers and diophantine approximations.Bull. Amer. Math. Soc. 77, 635–677. 201, 221

S. Lang 1993. Algebra. Reading, MA: Addison Wesley. Third edition.(Revised third edition: Berlin: Springer 2002) 16, 20, 122, 131, 133,288, 289, 292, 303, 304, 350, 352, 353, 354, 368, 380

Page 409: The Classical Fields

392 References

D. Lascar 1992. Les automorphismes d’un ensemble fortement minimal.J. Symbolic Logic 57, 238–251. 150

D. Lascar 1997. The group of automorphisms of the field of complex numbersleaving fixed the algebraic numbers is simple. In: Model Theory of Groupsand Automorphism Groups (Blaubeuren, 1995), vol. 244 of London Math-ematical Society Lecture Notes, pages 110–114. Cambridge: CambridgeUniversity Press. 150

H. Lauchli 1962. Auswahlaxiom in der Algebra.Comment. Math. Helvet. 37, 1–18. 134, 354

P. Lax 1973. The differentiability of Polya’s function.Adv. Math. 10, 456–464. 362

C.R. Leedham-Green and S. McKay 2002. The Structure of Groups ofPrime Power Order. Oxford: Oxford University Press. 324

J. B. Leicht 1966. Zur Charakterisierung reell abgeschlossener Korper.Monatsh. Math. 70, 452–453. 130

H.W. Lenstra, Jr. and P. Stevenhagen 1989. Uber das Fortsetzen vonBewertungen in vollstandigen Korpern.Arch. Math. (Basel) 53, 547–552. 315

A. Levy 1979. Basic Set Theory.Berlin: Springer. 337

T. Lindstrøm 1988. An invitation to nonstandard analysis.In: Nonstandard Analysis and its Applications (Hull, 1986), pages 1–105.Cambridge: Cambridge University Press. 154

E. Livenson 1937. An example of a non-closed connected subgroup of thetwo-dimensional vector space. Ann. Math. (2 ) 38, 920–922. 93

H. Lombardi and M.-F. Roy 1991. Elementary constructive theory of or-dered fields. In: Effective Methods in Algebraic Geometry (Castiglioncello,1990), vol. 94 of Progress in Mathematics, pages 249–262. Boston, MA:Birkhauser. 133

L.H. Loomis 1945. Abstract congruence and the uniqueness of Haar measure.Ann. Math. (2 ) 46, 348–355. 326

M. Lopez Pellicer 1994. Las construcciones de los numeros reales.In: History of Mathematics in the XIXth Century, Part 2 (Madrid, 1993),pages 11–33. Madrid: Real Academia de Ciencias Exactas, Fısicas y Nat-urales. xii, 1

R. Lowen 1985. Lacunary free actions on the real line.Topology Appl. 20, 135–141. 60

H. Luneburg 1973. Einfuhrung in die Algebra.Berlin: Springer. 284

D. J. Lutzer 1980. Ordered topological spaces. In: G.M. Reed (ed.), Surveysin General Topology, pages 247–295. New York: Academic Press. 164

G. Malle and B. Matzat 1999. Inverse Galois Theory.Berlin: Springer. 145, 150

A. Markov 1945. On free topological groups (Russian).Izvestia Akad. Nauk SSSR 9, 3–64. 341

D.A. Martin 1976. Hilbert’s first problem: the continuum hypothesis.In: F.E. Browder (ed.), Mathematical Developments arising fromHilbert Problems (Proc. Sympos. Pure Math. 28, 1974), pages 81–92.Providence, RI: American Mathematical Society. 339

G.E. Martin 1998. Geometric Constructions.New York: Springer. 124, 125

Page 410: The Classical Fields

References 393

B. Mazur 1993. On the passage from local to global in number theory.Bull. Amer. Math. Soc. (N.S.) 29, 14–50. 325

G.H. Meisters and J.D. Monk 1973. Construction of the reals via ultra-powers. Rocky Mountain J. Math. 3, 141–158. 160

A.H. Mekler 1986. Groups embeddable in the autohomeomorphisms of Q.J. London Math. Soc. (2 ) 33, 49–58. 205

J. van Mill 1982a. Homogeneous subsets of the real line.Compositio Math. 46, 3–13. 57

J. van Mill 1982b. Homogeneous subsets of the real line which do not admitthe structure of a topological group. Indag. Math. 44 (Nederl. Akad.Wetensch. Proc. Ser. A 85), 37–43. 57

J. van Mill 1992. Sierpinski’s technique and subsets of R.Topology Appl. 44, 241–261. 26

A.W. Miller 1979. On the length of Borel hierarchies.Ann. Math. Logic 16, 233–267. 109

A.W. Miller 1984. Special subsets of the real line. In: Handbook of Set-theoretic Topology, pages 201–233. Amsterdam: North-Holland. 51

A.W. Miller 1993. Special sets of reals. In: Set Theory of the Reals (RamatGan, 1991), vol. 6 of Israel Mathematical Conference Proceedings, pages415–431. Ramat Gan: Bar-Ilan University. 51

D. Montgomery 1948. Connected one-dimensional groups.Ann. Math. (2 ) 49, 110–117. 91

D. Montgomery and L. Zippin 1955. Topological Transformation Groups.New York: Interscience. 88, 141

R.L. Moore 1920. Concerning simple continuous curves.Trans. Amer. Math. Soc. 21, 333–347. 42

P. Morandi 1996. Field and Galois Theory.New York: Springer. 145, 350, 353, 354, 355

S.A. Morris 1977. Pontryagin Duality and the Structure of Locally CompactAbelian Groups. Cambridge: Cambridge University Press. 344, 345

S.A. Morris 1986. A characterization of the topological group of real num-bers. Bull. Austral. Math. Soc. 34, 473–475. 90

S.A. Morris and S. Oates-Williams 1987. A characterization of the topo-logical group of p-adic integers. Bull. London Math. Soc. 19, 57–59. 287

S.A. Morris, S. Oates-Williams, and H.B. Thompson 1990. Locallycompact groups with every closed subgroup of finite index.Bull. London Math. Soc. 22, 359–361. 287

A.F. Mutylin 1966. Imbedding of discrete fields into connected ones (Rus-sian). Dokl. Akad. Nauk SSSR 168, 1005–1008. Translated in SovietMath. Dokl. 7, 772–775. 138

A.F. Mutylin 1968. Connected complete locally bounded fields. Completenot locally bounded fields (Russian). Mat. Sbornik (N.S.) 76 (118), 454–472. Translated in Math. USSR Sbornik 5, 433–449. 138, 139, 321

K. Nagami 1970. Dimension Theory.New York: Academic Press. 91, 92

M. Di Nasso and M. Forti 2002. On the ordering of the nonstandard realline. In: Logic and Algebra, vol. 302 of Contemporary Mathematics, pages259–273. Providence, RI: American Mathematical Society. 164

J. Neukirch 1992. Algebraische Zahlentheorie. Berlin: Springer. (Translation:Algebraic Number Theory. Berlin: Springer 1999)

284, 289, 290, 304, 322, 374

Page 411: The Classical Fields

394 References

B.H. Neumann 1949a. On ordered division rings.Trans. Amer. Math. Soc. 66, 202–252. 359

B.H. Neumann 1949b. On ordered groups.Amer. J. Math. 71, 1–18. 76

J. von Neumann 1928. Ein System algebraisch unabhangiger Zahlen.Math. Ann. 99, 134–141. 7

P.M. Neumann 1985. Automorphisms of the rational world.J. London Math. Soc. (2 ) 32, 439–448. 200, 205

I. Niven 1956. Irrational Numbers.New York: Mathematical Association of America. 28

I. Niven 1961. Numbers: Rational and Irrational.New York: Random House. 201

P. J. Nyikos 1992. Various smoothings of the long line and their tangentbundles. Adv. Math. 93, 129–213. 48

J. E. Nymann 1993. The sum of the Cantor set with itself.Enseign. Math. (2 ) 39, 177–178. 106

J.C. Oxtoby 1971. Measure and Category.New York: Springer. 108

R. S. Palais 1968. The classification of real division algebras.Amer. Math. Monthly 75, 366–368. 138, 331

V. Pambuccian 1990. On the Pythagorean hull of Q.Extracta Math. 5, 29–31. 125

S. Pauli and X.-F. Roblot 2001. On the computation of all extensions of ap-adic field of a given degree. Math. Comput. 70, 1641–1659. 324

A.R. Pears 1975. Dimension Theory of General Spaces.Cambridge: Cambridge University Press. 197

J. Petro 1987. Real division algebras of dimension > 1 contain C.Amer. Math. Monthly 94, 445–449. 138

A. Pfister 1995. Quadratic Forms with Applications to Algebraic Geometryand Topology. Cambridge: Cambridge University Press. 122, 134, 216

G. Polya 1913. Uber eine Peanosche Kurve.Bull. Acad. Sci. Cracovie Ser. A, 305–313. 362

L. S. Pontryagin 1932. Uber stetige algebraische Korper.Ann. Math. (2 ) 33, 163–174. 138

L. S. Pontryagin 1986. Topological Groups. Selected works vol. 2 (ed. R. V.Gamkrelidze). New York: Gordon and Breach. Third edition. (Germantranslation: Leipzig: Teubner 1957/1958) 90, 139, 340, 344, 345, 346

K. Potthoff 1981. Einfuhrung in die Modelltheorie und ihre Anwendungen.Darmstadt: Wissenschaftliche Buchgesellschaft. 170

V. Powers 1996. Hilbert’s 17th problem and the champagne problem.Amer. Math. Monthly 103, 879–887. 134

K. Prachar and H. Sagan 1996. On the differentiability of the coordinatefunctions of Polya’s space-filling curve.Monatsh. Math. 121, 125–138. 362

A. Prestel 1984. Lectures on Formally Real Fields, vol. 1093 of LectureNotes in Mathematics. Berlin: Springer. 122, 133, 134

A. Prestel and C.N. Delzell 2001. Positive Polynomials. From Hilbert’s17th Problem to Real Algebra. Berlin: Springer. 122, 133, 134

S. Prieß-Crampe 1983. Angeordnete Strukturen: Gruppen, Korper, projek-tive Ebenen. Berlin: Springer. 76, 77, 166, 167, 246, 359

Page 412: The Classical Fields

References 395

S. Prieß-Crampe and P. Ribenboim 2000. A general Hensel’s lemma.J. Algebra 232, 269–281. 279

W. Rautenberg 1987. Uber den Cantor-Bernsteinschen Aquivalenzsatz.Math. Semesterber. 34, 71–88. 337

I. Reiner 1975. Maximal Orders. London: Academic Press. (Correctedreprint: Oxford: Oxford University Press 2003) 323, 324, 331, 332, 333

G. Ren 1992. A note on (Q × Q) ∪ (I × I).Questions Answers Gen. Topology 10, 157–158. 70

P. Ribenboim 1985. Equivalent forms of Hensel’s lemma.Expo. Math. 3, 3–24. 279

P. Ribenboim 1992. Fields: algebraically closed and others.Manuscripta Math. 75, 115–150. 124, 125, 133, 359

P. Ribenboim 1993. Some examples of lattice-orders in real closed fields.Arch. Math. (Basel) 61, 59–63. 124, 133

P. Ribenboim 1999. The Theory of Classical Valuations. New York: Springer.289, 290, 304, 306, 313, 315, 323, 324, 357, 358, 377, 378

A.M. Robert 2000. A Course in p-adic Analysis.New York: Springer. 278, 284, 285, 323, 324, 326, 374

L. Robertson and B.M. Schreiber 1968. The additive structure of integergroups and p-adic number fields.Proc. Amer. Math. Soc. 19, 1453–1456. 287

A.M. Rockett and P. Szusz 1992. Continued Fractions. River Edge, NJ:World Scientific. 28, 30

W. Roelcke and S. Dierolf 1981. Uniform Structures on Topological Groupsand their Quotients. New York: McGraw-Hill. 341

C.A. Rogers and T.E. Jayne 1980. Analytic Sets.London: Academic Press. 110, 112

K.F. Roth 1955. Rational approximations to algebraic numbers.Mathematika 2, 1–20. Corrigendum page 168. 30

K.F. Roth 1960. Rational approximations to algebraic numbers. In: Proceed-ings of the International Congress of Mathematicians (Edinburgh, 1958),pages 203–210. Cambridge: Cambridge University Press. 30

G. Rousseau 1987. On a construction for the representation of a positiveinteger as the sum of four squares. Enseign. Math. (2 ) 33, 301–306. 213

J. E. Rubin 1967. Set Theory for the Mathematician.San Francisco, CA: Holden-Day. 335, 337

H. Rubin and J. E. Rubin 1985. Equivalents of the Axiom of Choice II.Amsterdam: North-Holland. 336

M.E. Rudin 1969. Souslin’s conjecture.Amer. Math. Monthly 76, 1113–1119. 27

H. Sagan 1994. Space-filling Curves.New York: Springer. 362

S. Saks 1937. Theory of the Integral. Warszawa–New York: Stechert.(Second revised edition: New York: Dover Publications 1964) 112

H. Salzmann 1958. Kompakte zweidimensionale projektive Ebenen.Arch. Math. (Basel) 9, 447–454. 59

H. Salzmann 1969. Homomorphismen komplexer Ternarkorper.Math. Z. 112, 23–25. 147

H. Salzmann 1971. Zahlbereiche. I: Die reellen Zahlen. University of Tubin-gen. Mimeographed lecture notes prepared by R. Lowen. xii, 39

Page 413: The Classical Fields

396 References

H. Salzmann 1973. Zahlbereiche. II: Die rationalen Zahlen. III: Die kom-plexen Zahlen. University of Tubingen. Mimeographed lecture notes pre-pared by H. Hahl. xii, 221

H. Salzmann, D. Betten, T. Grundhofer, H. Hahl, R. Lowen, andM. Stroppel 1995. Compact Projective Planes.Berlin: de Gruyter. xii, 72, 91, 92, 141, 211

T. Sander 1991. Existence and uniqueness of the real closure of an orderedfield without Zorn’s lemma. J. Pure Appl. Algebra 73, 165–180. 133

W. Scharlau 1985. Quadratic and Hermitian Forms.Springer, Berlin. 122, 125, 133, 134, 216, 297, 299, 325, 331, 332, 333

B. Schnor 1992. Involutions in the group of automorphisms of an alge-braically closed field. J. Algebra 152, 520–524. 150

T. Schonemann 1846. Von denjenigen Moduln, welche Potenzen von Prim-zahlen sind. J. Reine Angew. Math. 32, 93–105. 360, 380, 382

E. Schonhardt 1963. Die Satze von Pappus, Pascal und Desargues als Iden-titaten in komplexen Zahlen. Tensor (N.S.) 13, 223–231. 151

H. Schwerdtfeger 1979. Geometry of Complex Numbers. New York: DoverPublications. Corrected reprint of the 1962 edition. 151

J.-P. Serre 1973. A Course in Arithmetic. New York: Springer.(Corrected second printing 1978) 284, 293, 297, 325, 374

J.-P. Serre 1979. Local Fields. New York: Springer.(Corrected second printing 1995) 315, 323, 324, 331, 332, 333, 358

D.B. Shakhmatov 1983. Imbeddings in topological fields and the construc-tion of a field whose space is not normal (Russian).Comment. Math. Univ. Carolin. 24, 525–540. 136

D.B. Shakhmatov 1987. The structure of topological fields, and cardinalinvariants (Russian). Trudy Moskov. Mat. Obshch. 50, 249–259, 262.Translated in Trans. Moscow Math. Soc. 1988, 251–261. 136

H.N. Shapiro 1950. On primes in arithmetic progressions.Ann. Math. (2 ) 52, 217–230, 231–243. 374

S. Shelah 1983. Models with second order properties. IV. A general methodand eliminating diamonds. Ann. Pure Appl. Logic 25, 183–212. 125

N. Shell 1990. Topological Fields and Near Valuations.New York: Dekker. 135, 136, 139, 146, 321, 359

W. Sierpinski 1920. Sur une propriete topologique des ensembles denombra-bles denses en soi. Fund. Math. 1, 11–16. 197

R. Sikorski 1964. Boolean Algebras.Berlin: Springer. Second edition. 54

C. Small 1982. A simple proof of the four-squares theorem.Amer. Math. Monthly 89, 59–61. 213

C. Small 1986. Sums of three squares and levels of quadratic number fields.Amer. Math. Monthly 93, 276–279. 212

H. J. Smith 1855. De compositione numerorum primorum formae 4λ + 1 exduobus quadratis. J. reine angew. Math. 50, 91–92. 209

R.M. Smullyan and M. Fitting 1996. Set Theory and the Continuum Prob-lem. Oxford: Oxford University Press. 335, 338, 339

S. Solecki and S.M. Srivastava 1997. Automatic continuity of group op-erations. Topology Appl. 77, 65–75. 340

R.H. Sorgenfrey 1947. On the topological product of paracompact spaces.Bull. Amer. Math. Soc. 53, 631–632. 69

Page 414: The Classical Fields

References 397

T. Soundararajan 1969. The topological group of the p-adic integers.Publ. Math. Debrecen 16, 75–78. 290

T. Soundararajan 1991. On the cardinality of the group of automorphismsof algebraically closed extension fields.Acta Math. Vietnam. 16, 147–153. 150

M. Souslin 1920. Probleme 3.Fund. Math. 1, 223. 27

E. Specker 1954. Verallgemeinerte Kontinuumshypothese und Auswahlax-iom. Arch. Math. (Basel) 5, 332–337. 339

L.A. Steen and J.A. Seebach, Jr. 1978. Counterexamples in Topology.New York: Springer. Second edition. (Reprint: New York: Dover Publi-cations 1995) 45

E. Steinitz 1910. Algebraische Theorie der Korper.J. reine angew. Math. 137, 167–309. (Re-edited by R. Baer and H. Hasse,Berlin: de Gruyter 1930 and New York: Chelsea 1950) 355

J. Stillwell 2002. The continuum problem.Amer. Math. Monthly 109, 286–297. 339

M. Stroppel 2006. Locally Compact Groups.Zurich: EMS Publishing House. 66, 90, 249, 340, 344, 345, 346, 349

T. Szele 1949. Die Abelschen Gruppen ohne eigentliche Endomorphismen.Acta Univ. Szeged. Sect. Sci. Math. 13, 54–56. 96

S. Tennenbaum 1968. Souslin’s problem.Proc. Nat. Acad. Sci. U.S.A. 59, 60–63. 27

J. Tits 1974. Buildings of Spherical Type and Finite BN-Pairs, vol. 386 ofLecture Notes in Mathematics. Berlin: Springer. 152

J.K. Truss 1997. Conjugate homeomorphisms of the rational world.Forum Math. 9, 211–227. 205

P. Ullrich 1998. The genesis of Hensel’s p-adic numbers. In: Charlemagneand his Heritage. 1200 Years of Civilization and Science in Europe, vol. 2(Aachen, 1995), pages 163–178. Turnhout: Brepols. 278

H. Volklein 1996. Groups as Galois Groups.Cambridge: Cambridge University Press. 145, 150, 358

S. Wagon 1990. The Euclidean algorithm strikes again.Amer. Math. Monthly 97, 125–129. 209

M. Waldschmidt 1992. Linear Independence of Logarithms of AlgebraicNumbers, vol. 116 of IMSc Reports. Madras: Institute of Mathemati-cal Sciences. 7

J.A. Ward 1936. The topological characterization of an open linear interval.Proc. London Math. Soc. 41, 191–198. 42

J. P. Ward 1997. Quaternions and Cayley Numbers.Kluwer, Dordrecht. xii

S. Warner 1989. Topological Fields. Amsterdam: North-Holland.135, 139, 146, 249, 263, 268, 276, 277, 278, 293, 303, 304, 306, 313, 315,321, 330, 375, 378, 380

W.C. Waterhouse 1985. When one equation solves them all.Amer. Math. Monthly 92, 270–273. 130

A.G. Waterman and G.M. Bergman 1966. Connected fields of arbitrarycharacteristic. J. Math. Kyoto Univ. 5, 177–184. 138

A. Weil 1967. Basic Number Theory. New York: Springer. (Second edition1973) 166, 289, 293, 314, 323, 324, 326, 330, 332, 333

Page 415: The Classical Fields

398 References

W. Wi‘es�law 1988. Topological Fields. New York: Dekker. (Earlier version:

Acta Univ. Wratislav. 1982) 138, 139, 146, 321, 329E. Witt 1975. Homoomorphie einiger Tychonoffprodukte. In: Topology

and its Applications (St. John’s, Newfoundland, 1973), vol. 12 of Lec-ture Notes in Pure and Applied Mathematics, pages 199–200. New York:Dekker. 283, 374

W.H. Woodin 2001. The continuum hypothesis I, II. Notices Amer. Math.Soc. 48, 567–576, 681–690. 339

T.-S. Wu 1962. Continuity in topological groups.Proc. Amer. Math. Soc. 13, 452–453. 340

S.W. Young 1994. The representation of homeomorphisms on the intervalas finite compositions of involutions.Proc. Amer. Math. Soc. 121, 605–610. 61

D. Zagier 1990. A one-sentence proof that every prime p ≡ 1 (mod 4) is asum of two squares. Amer. Math. Monthly 97, 144. 209

D. Zagier 1997. Newman’s short proof of the prime number theorem.Amer. Math. Monthly 104, 705–708. 159, 192

H. Zassenhaus 1970. A real root calculus. In: Computational Problems inAbstract Algebra (Oxford, 1967), pages 383–392. Oxford: Pergamon. 133

Page 416: The Classical Fields

Index

η1-field, 166η1-set, 166σ-algebra, 104σ-field, 104

absolute value, 113, 135, 223, 270, 271,279, 281, 300, 306

affine group, 101affine plane, 177Alexandroff’s long line, 47algebraic, 351algebraic closure, 353algebraic number, 200algebraically closed, 353algebraically independent, 7, 355analytic set, 110anti-automorphism, 211arc, 34, 40arc component, 43Archimedean, 76, 119, 122, 240, 300Archimedean ordered field, 242arcwise connected, 43Artin’s trick, 305Artin–Schreier, Theorem of, 123, 126,

130, 133axiom of choice, 335axiom of foundation, 335

Bezout’s Theorem, 4Bolzano–Weierstraß, Theorem of, 175Borel field, 104, 109Borel set, 104, 109, 110bounded (in a topological ring), 316bounded chain, 32Brauer group, 333

Cantor set, 14, 52, 54–56, 106, 112, 283cardinal, cardinality, 337Cauchy sequence, 250chain, 22, 25

character group, 97, 98, 288, 291, 344characteristic of field, 351circle, 34, 41, 42, 84cofinal, 119compact, 32compact-open topology, 344complete, 22, 250, 251, 265, 311complete measure space, 104completely ordered field, 243completely ordered group, 76, 240completion, 236, 243, 253, 257, 265, 270concentrated filterbase, 250, 254, 255,

263, 265, 266, 268connected component, 343continued fraction, 28, 29, 195, 229continuum hypothesis, 170, 339continuum problem, 6, 339convergence of filterbases, 249convex, 118coterminal, 22, 236countable basis, 32countable chain condition, 27countably compact, 50covering dimension, 91covering map, 86cyclic algebra, 331, 334cyclic group, 2

Darboux, Theorem of, 74Dedekind cut, 77, 161, 167, 237degree of field extension, 350degree valuation, 275, 307dense, 22, 119dimension, 91Diophant’s identity, 208direct product, 8direct sum, 8divisible, 5, 10domain invariance, 43domain of positivity, 113

399

Page 417: The Classical Fields

400 Index

embedding of topological groups, 253endomorphism field, 96endomorphism ring, 95, 97, 182, 288equivalent valuations, 309Euclidean field, 124, 126, 134, 162Euclidean plane, 151, 178exponential field, 117exponential function, 15exponentially closed, 315

Fermat number, 189Fibonacci number, 190, 361field, 71, 350field extension, 350field of fractions, 207field topology, 135filter, 154, 249

neighbourhood, 250filterbase, 249flow, 59formally real, 122fundamental theorem of algebra, 128,

144

Galois extension, 354Galois group, 289, 354generalized continuum hypothesis, 27,

339

Haar measure, 326Hahn power series, 124, 169, 358Hahn–Mazurkiewicz, Theorem of, 53Hamel basis, 6, 107Hausdorff space, 32Hausdorff’s maximal chain principle,

336Hensel’s Lemma, 278, 325Hilbert’s 17th problem, 134Holder, Theorem of, 77homeomorphism, 57homeomorphism group, 57homogeneous, 26, 49, 204

ideal in a chain, 236identification map, 42infinitely small, 158infinitesimal, 158, 160, 162injective group, 10interval, 22involution, 58irrational numbers, 56, 70isotropic, 298

Kowalsky–Durbaum and Fleischer,Theorem of, 319

Kronecker’s Theorem, 67, 68, 226

Laurent series, 273, 285, 356LCA group, 344least upper bound, 22Lebesgue measure, 105Leibniz series, 215lexicographic ordering, 78Liouville, Theorem of, 201local field, 322local homeomorphism, 65locally arcwise connected, 43locally bounded, 316locally compact, 33, 251locally connected, 33locally cyclic, 4, 179long line, 48, 70long ray, 47

Mal’cev–Iwasawa, Theorem of, 343manifold, 42meagre set, 108measurable, 105, 112measure (space), 104Mersenne prime, 16, 188metric group, 222minimal polynomial, 351monad, 162monic polynomial, 351monotonicity, 75, 113

natural valuation, 322Newton and Puiseux, Theorem of, 358non-Archimedean, 300normal field extension, 352

one-parameter group, 59order isomorphism, 24order of group element, 3order relation, 22order topology, 23order-preserving, 24order-rigid, 119ordered field, 113, 135ordered group, 75, 82ordering, 22ordering valuation, 315ordinal, 335Ostrowski, Theorem of, 302, 303

p-adic absolute value, 223, 279, 281p-adic integer, 281p-adic metric, 222, 223p-adic number, 273p-adic topology, 223, 231, 281p-divisible, 287perfect space, 55polynomial ring, 356Pontryagin duality, 348

Page 418: The Classical Fields

Index 401

Pontryagin, Theorem of, 137, 332power series, 17, 75, 273, 356p-primary component, 11primary component, 180prime element, 306prime field, 74, 207, 351prime number, 188prime number theorem, 159, 192principal valuation, 306projective plane, 177Prufer group, 12, 13, 180, 284, 285pseudo-compact, 50, 165Puiseux series, 18, 134, 357purely transcendental extension, 355Pythagorean field, 124, 134, 162

quadratic form, 213, 215, 297, 325quaternion, xii, 129, 139, 211, 316quotient topology, 42

ramification index, 323rank of free abelian group, 186real algebraic number, 30real closed field, 124reflexive, 344regular measure, 104, 327residue degree, 323residue field, 308retraction, 10rigid, 74ring completion, 265ring topology, 135Roth, Theorem of, 30

semi-topological group, 82, 340semicontinuous, 68semidirect product, 101, 102separable (field extension), 353separable chain, 23separable space, 23separating set or point, 35sequentially compact, 50Shafarevich and Kaplansky, Theorem

of, 317simple group, 206simply connected, 86skew field, 71solenoid, 89, 97, 99, 100, 291Sorgenfrey topology, 69, 82, 100, 166,

340Souslin set, 110–112Souslin tree, 27Souslin’s condition, hypothesis, 27splitting field, 352square class, 295square free, 45standard part, 162

Steinitz, Theorem of, 355strongly dense, 23subadditive, 300support, 61supremum, 22surreal numbers, 1, 76, 124

T1-space, 32topological field, 135topological group, 81, 340topological ring, 135topological vector space, 328topologically nilpotent, 317torsion free, 3torsion group, 11, 180torus, 9, 13, 65, 84, 88torus topology, 65totally disconnected, 55totally ordered set, 22totally ramified, 323transcendency basis, 355transcendency degree, 355transcendental, 220, 354triangle inequality, 113, 223type V , 319

ultrafilter, 154free, 155principal, 155

ultrametric, 223, 301ultrapower, ultraproduct, 155unconditionally complete, 32uniformly continuous, 253uniquely divisible, 5unramified, 323

valuation, 135, 306valuation ring, 160, 308value group, 306

weakly dense, 22, 236well-ordered, 335well-ordering principle, 335

Zorn’s lemma, 335