Sei sulla pagina 1di 238

Foundations of Convex Geometry

AUSTRALIAN MATHEMATICAL SOCIETY LECTURE SERIES

Editor-in-chief: Professor J. H. Loxton, School of Mathematics, Physics,


Computing and Electronics, Macquarie University, NSW 2109, Australia

1 Introduction to Linear and Convex Programming, N. CAMERON


2 Manifolds and Mechanics, A. JONES, A. GRAY & R. HUTTON
3 Introduction to the Analysis of Metric Spaces, J. R. GILES
4 An Introduction to Mathematical Physiology and Biology, J. MAZUMDAR
5 2-Knots and their Groups, J. HILLMAN
6 The Mathematics of Projectiles in Sport, N. DE MESTRE
7 The Peterson Graph, D. A. HOLTON & J. SHEEHAN
8 Low Rank Representations and Graphs for Sporadic Groups,
C. PRAEGER & L. SOICHER
9 Algebraic Groups and Lie Groups, G. LEHRER (ed)
10 Modelling with Differential and Difference Equations,
G. FULFORD, P. FORRESTER & A. JONES
11 Geometric Analysis and Lie Theory in Mathematics and Physics,
A. CAREY & M. MURRAY (eds)
Foundations of Convex Geometry
W. A. COPPEL
Department ofTheoretical Physics,
Australian National University

CAMBRIDGE
UNIVERSITY PRESS
PUBLISHED BY THE PRESS SYNDICATE OF THE UNIVERSITY OF CAMBRIDGE
The Pitt Building, Trumpington Street, Cambridge CB2 lRP, United Kingdom
CAMBRIDGE UNIVERSITY PRESS
The Edinburgh Building, Cambridge CB2 2RU, UK http://www.cup.cam.ac.uk
40 West 20th Street, New York, NY 10011-4211, USA http://www.cup.org
10 Stamford Road, Oakleigh, Melbourne 3166, Australia

© Cambridge University Press 1998

This book is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without
the written permission of Cambridge University Press.

First published 1998

Printed in the United Kingdom at the University Press, Cambridge

Typeset by the author

A catalogue record for this book is available from the British Library

Library of Congress Cataloguing in Publication data

Coppel, W. A.
Foundations of convex geometry I W. A. Coppel.
p. cm. - (Australian Mathematical Society lecture series: 12)
Includes bibliographical references and index.
ISBN 0 521 63970 0
1. Convex geometry. I. Title. II. Series.
QA639.5C66 1998
516'.08-dc21 97-32149 CIP

ISBN 0 521 63970 0 paperback


For Maria Antonietta
Contents

Preface XI

Introduction 1

I Alignments 3

1 Partially ordered sets 3


2 Aligned spaces 6
3 Anti-exchange alignments 14
4 Antimatroids 16
5 Exchange alignments 20
6 Notes 24

II Convexity 27

1 Examples 27
2 Convex geometries 31
3 Additional axioms 36
4 Examples (continued) 43
5 Notes 47

III Linearity 49

1 Lines 49
2 Linear geometries 54
3 Helly's theorem and its relatives 66
vm Contents

IV Linearity (continued) 71

1 Faces 71
2 Polytopes 74
3 Factor geometries 81
4 Notes 84

v Density and Unendlngness 85

1 Dense linear geometries 85


2 Unending linear geometries 95
3 Cones 103
4 Polyhedra 109
5 Notes 114

VI Desargues 115

1 Introduction 115
2 Projective geometry 119
3 The Desargues property 124
4 Factor geometries 131
5 Notes 135

VII Vector Spaces 137

1 Coordinates 137
2 Isomorphisms 143
3 Vector sums 150
4 Notes 154

VIII Completeness 157

1 Introduction 157
2 Separation properties 164
3 Fundamental theorem of ordered geometry 172
4 Metric and norm 178
5 Notes 182
Contents IX

IX Spaces of Convex Sets 185


1 The Hausdorff metric 185
2 The space '€(X) 189
3 Embeddings of '€(X) 192
4 The Krein-Kakutani theorem 196
5 Notes 209

References 211

Notations 216

Axioms 217

Index 218
Preface

Euclid's Elements held sway in the mathematical world for more than two
thousand years. (For an English translation, see Euclid (1956).) In the nineteenth
century, however, the growing demand for rigour led to a re-examination of the
Euclidean edifice and the realization that there were cracks in it. The critical
reappraisal which followed was synthesized by Hilbert (1899), in his Grundlagen
der Geometrie, which has in tum held sway for almost a century.
In what ways may Hilbert's treatment be improved upon today? Apart from
technical improvements, a number of which were included in later editions of his
book, it may be argued that Hilbert followed Euclid too closely. The number of
undefined concepts is unnecessarily large and the axioms of congruence sit uneasily
with the other axioms. In addition, the restriction to three-dimensional space
conceals the generality of the results and seems artificial now.
For these reasons Hilbert's approach is often replaced today by a purely
algebraic one - the axioms for a vector space over an arbitrary field, followed by the
axioms for a vector space over the real field with a positive definite scalar product.
This permits a rapid development, but it simply begs the question of why it is
possible to introduce coordinates on a line. Only when one attempts to answer this
question does one realise that many important results in no way depend on the
introduction of coordinates. It is curious that some powerful advocates of
'coordinate-free' linear algebra have used real analysis to prove purely geometric
results, such as the Hahn-Banach theorem, without any apparent twinge of
conscience.
A system of axioms for Euclidean geometry in which the only undefined
concepts are point and segment was already given before Hilbert by Peano. We
follow his example in the present work, although some of our axioms differ. The
choice of a system of axioms is inherently arbitrary, since there will be many
equivalent systems. However, the purpose of an axiom system is not only to
provide a basis for rigorous proof, but also to reveal the structure of a subject.
XII Preface

From this point of view one axiom system may seem preferable to another which is
equivalent to it. The development should seem natural, almost inevitable.
Considerations of this nature have led us to isolate a basic structure, here called
a convex geometry, which is defined by two axioms only. Additional axioms are
chosen so that each individually, in conjunction with these, guarantees some
important property. Four such additional axioms define another basic structure,
here called a linear geometry, for which a rich theory may be developed.
Examples of linear geometries, in addition to Euclidean space, are hyperbolic space
and (hemi)spherical space, where a 'segment' is the geodesic arc joining two
points.
Apart from a dimensionality axiom, only seven axioms in all are needed to
characterize ordinary Euclidean or, more strictly, real affine space. However,
dimensionality plays a role in achieving this number, by eliminating some other
possibilities in one or two dimensions. Although the characterization of Euclidean
space may be regarded as our ultimate goal, it would be contrary to our purpose to
impose all the axioms from the outset. Instead we adjoin axioms successively, so
that results are proved under minimal hypotheses and may be applied in other
situations, which are of interest in their own right. For example, Proposition 111.17
establishes the theorems of Belly and Radon in any linear geometry, and Chapter IV
similarly extends the facial theory of polytopes. In this way each result appears, so
to speak, in its 'proper place'. When there is a branching of paths, we choose the
one which leads to our ultimate goal. This approach, I believe, has not previously
been pursued in such a systematic manner. I have found it illuminating myself and
hope that the illumination succeeds in shining through the present account.
We have spoken of the characterization of Euclidean space as our ultimate
goal, but we are actually concerned with characterizing only a convex subset of
Euclidean space. The greater freedom of the whole Euclidean space is
mathematically desirable, but it is reasonable to require only that our physical world
may be embedded in such a space. This type of non-Euclidean geometry was
already mentioned by Klein (1873) and was further considered by Schur (1909). In
view of the great historical importance of the parallel axiom, it is of interest that it
can play no role here, since it fails to hold in any proper convex subset of Euclidean
space (except a lower-dimensional Euclidean space).
The foundations of geometry have been studied for thousands of years, and
thousands of papers have been written on the subject. Since it is impossible to do
Preface xm

justice to all previous contributions in such a situation, we have deliberately


restricted the number of references. Thus the absence of a reference does not
necessarily imply ignorance of its existence and is not a judgement on its quality.
Some references are included simply to provide a time scale and others for their
useful bibliographies. Our task has been to select and organize, and sometimes
extend. Some open problems are mentioned at the ends of Chapters Ill, IV and
VITI.
The topic of axiomatic convexity has been omitted from the recent extremely
valuable Handbook of convex geometry (Vols. A and B, ed. P.M. Gruber and
J.M. Wills, North-Holland, Amsterdam, 1993), and ordered geometry receives
only passing reference (on p. 1311) in the equally valuable Handbook of incidence
geometry (ed. F. Buekenhout, North-Holland, Amsterdam, 1995). It is hoped that
the present work may in some measure repair these omissions. Our aim has been to
give a connected account of these subjects, which may be read without reference to
other sources. Since many results are established under weaker hypotheses than
usual, rather detailed proofs have been given in some cases.
The work is not arranged as a textbook, with starred sections and exercises,
and is perhaps more difficult than the usual final-year undergraduate or fir~t-year
graduate course. However, the mathematical prerequisites are no greater and I
believe that an interesting course could be constructed from the material here, which
would acquaint students with a cross-section of mathematics in contrast to the usual
compartmentalized course.
Familiarity is assumed with the usual language and notation of set theory, with
such algebraic concepts as group, field and vector space, and with Dedekind's
construction of the real numbers from the rationals. Concepts from other areas,
such as partially ordered sets and lattices, projective geometry, topology and metric
spaces, are defined in the text. A few elementary properties of metric spaces are
stated without proof. On the two or three occasions when appeal is made to some
other result the subsequent development is not at stake. Although the work is
essentially self-contained, the reader is still encouraged to consult other treatments,
both in the references cited in the notes at the end of each chapter and in the
introductory chapters of more specialized works on convexity theory, such as
Bonnesen and Fenchel (1934), Valentine (1964), Leichtweiss (1980) and Schneider
(1993). The remaining prerequisites for the present work are more substantial-the
ability and resolve to follow a detailed logical argument and a love of mathematics.
XIV Preface

For assistance in various ways I thank M. Albert, B. Davey, V. Klee, J. Reay,


J. Schiffer, V. Soltan and H. Tverberg. I am especially grateful to H. Tverberg for
the detection of errors, misprints and obscurities in the original manuscript. The
exposition and index have been improved by suggestions from the referees, and the
appearance by suggestions from the copy-editor.
As I write these lines on the eve of my retirement from paid employment in the
Institute of Advanced Studies at the Australian National University, I take the
opportunity to acknowledge that this book could not have been written without the
privileged working conditions which I have enjoyed.

Andrew Coppel
Introduction

In the present work convex subsets of a real vector space are characterized by a
small number of axioms involving points and segments. At the same time a
substantial part of the basic theory of convex sets is developed in a purely geometric
manner. This contrasts with traditional treatments, in which the origin has a
distinguished role and in which use is made of such analytic devices as duality and
the difference of two sets, not to mention metric properties.
We owe to Pasch (1882) the recognition of the role of order in Euclidean
geometry, both the order of points on a line and a more subtle order associated with
points in a plane. A basic undefined concept in Pasch's work was the segment
determined by two distinct points. By assuming various axioms involving
segments he was able to define a line and the order of points on a line. An
additional undefined concept was that of the plane determined by three points not
belonging to the same line.
Peano (1889) avoided this additional undefined concept by replacing Pasch's
planar ordering axiom by two further axioms involving segments. Thus he was
able to give a system of axioms for Euclidean geometry in which the only undefined
concepts were point and segment. Some of Peano's axioms were consequences of
the remaining axioms. An independent system of axioms, involving only points
and segments, was given by Veblen (1904).
Two axioms which were used from the outset by all these authors will be
mentioned here. The first, called density in the present work, states that the
segment [a,b] determined by two distinct points a,b contains a point c ¢ a,b. The
second, called unendingness in the present work, states that for any two distinct
points a,b there is a point c ¢ a,b such that b is contained in the segment [a,c]
determined by a,c. In recent years discrete mathematics has attained equal status
with continuous mathematics, and it seems desirable to develop the foundations of
geometry as far as possible without using these two axioms. This is what is done
here.
2 Introduction

One consequence is that our segments are closed (endpoints included), rather
than open (endpoints excluded) as in the work of the authors cited. However, this
is not a novelty; see, for example, Szczerba and Tarski (1979). What is new here is
the gradual introduction of axioms and an obstinate insistence on establishing
results without the use of axioms which are not required.
I
Alignments

This chapter is of a preparatory nature. For convenience of reference we begin by


defining some basic concepts related to partially ordered sets and lattices. We also
prove Hausdorffs maximality theorem, which will be used repeatedly. However,
the main topic of the chapter is alignments. An alignment, like a topology, is a
collection of subsets with certain properties. (The name is not very suggestive, but
has the merit that it is not used elsewhere in mathematics.) Alignments, or
equivalently algebraic hull operators, provide an abstract framework for such
concepts as extreme point, independent set, basis and/ace. Reily sets, Radon
sets and Caratheodory sets are also defined. Although they are of interest in the
present general context, they will acquire a more familiar form in Chapter III.
Two particular types of alignment are given special attention, those with the
exchange property and those with the anti-exchange property. Exchange
alignments provide an abstract framework for much of vector space theory,
including the concepts of hyperplane and dimension. Exchange alignments on a
finite set are well-known under the name of matroids. Anti-exchange alignments,
which are of more recent vintage, in the same way provide an abstract framework
for some aspects of convexity theory. We also study properties of anti-exchange
alignments on a finite set, which are here called antimatroids.

1 PARTIALLY ORDERED SETS

Although we will make no use of the theory of partially ordered sets and
lattices, we will at times use some of the concepts of these subjects and for
convenience of reference we state some definitions here.
4 I. Alignments

A set X is said to be partially ordered if a binary relation < is defined on X


with the properties

(01) x S: x (reflexivity),
(02) if x Sy and y S: x, then x = y (antisymmetry),
(03) if x S: y and y S: z, then x < z (transitivity).

Here a binary relation on X is just a subset R of the product set XxX and
x S: y denotes that (x,y) e R. It may be that neither x < y nor y S:x, in which case x
and y are said to be incomparable. A partially ordered set is totally ordered if no
two elements are incomparable.
If x S: y and x ~ y, we write x < y. Instead of x < y we may write y > x, and
instead of x < y we may write y > x.
We say that x e X is an upper bound for a subset Y of a partially ordered set X
if y < x for every y e Y, and a lower bound for Y if x < y for every y e Y. An
upper bound for Y is said to be a least upper bound, or supremum, for Y if it is a
lower bound for the set of all upper bounds. Similarly a lower bound for Y is said
to be a greatest lower bound, or bzfimum, for Y if it is an upper bound for the set of
all lower bounds. It follows from (02) that Y has at most one supremum and at
most one infimum.
A partially ordered set is a lattice if any two elements x,y have a supremum
x v y and an infimum x A y. It is a complete lattice if every subset Y has a
supremum and an infimum.
If X and Y are partially ordered sets, a map/: X -+ Y is said to be order-
preserving if x 1 < x 2 implies f(x 1) S:/(x2). It is said to be an order isomorphism
if, in addition, it is a bijection and /(x 1) < /(x2 ) implies x 1 < x 2 • An order
isomorphism between two lattices is a lattice isomorphism.
The preceding definitions are all rather basic. A few concepts of lesser
importance will be defined when they are encountered.
For us the most important example of a partially ordered set is the collection X
of all subsets of a given set C, where A S: B denotes that the subset A is contained in
the subset B. This partially ordered set is in fact a complete lattice, since the
supremum of any family {Aa} of subsets is their union UAa and the infimum is
their intersection n Aa.
One property of partially ordered sets will be proved here, since it will be used
repeatedly. This property may be formulated in several equivalent ways. We
1. Partially ordered sets 5

choose here a formulation due to Hausdorff (1927) and will always appeal to it.
Other formulations, including the well-ordering theorem and the popular Zorn's
lemma, may be found in Hewitt and Stromberg (1975).

HAUSDORFF'S MAXIMALITY THEOREM Every nonempty partially ordered


set X contains a maximal totally ordered subset.

Proof Let .srf. be the family of all totally ordered subsets of X. Evidently .srf. is not
empty, since it contains any singleton, i.e. any subset of X containing exactly one
element. If 5" is any subfamily of st which is totally ordered by inclusion, then the
union of all the (totally ordered) sets in 5" is again totally ordered and hence is in 91..
By the axiom of choice there exists a function f which associates to each
nonempty subset E of X an element/(£) of E. For any A e 91, let At be the set of
all x e X\A such that Aux e 91. We put

g(A) = {A uA/(At) if At¢ 0,


if At= 0.

The function g: .srf. ~st has the property that A


g(A) and that at most one element
5
of g(A) is not in A. We wish to show that g(A) =A for at least one A e 91, since
then At = 0 and A is a maximal element of 91.
Fix A 0 e st. We will call a subfamily ~ of st a 'tower' if it has the following
three properties:

(i) A 0 e ~'
(ii) if 5" is any subfamily of ~ which is totally ordered by inclusion, then the
union of all the sets in 5" is again in ~'
(iii) if B e ~' then g(B) e ~.

For example, the family of all B e .srf. such that A0 s B is a tower. If ~o is


the intersection of all towers, then ~o is itself a tower but no proper subfamily of
~o is a tower. Evidently A0 s B for every B e ~ 0 . It is enough to show that ~o
is totally ordered by inclusion. For the union B of all sets in ~o will be in ~ 0 , by
(ii), and g( B ) e ~ 0 , by (iii). Hence g( B ) s B , by the definition of B . Since
B 5 g( B ), by the definition of g, it follows that g( B ) = B .
Let ~ be the family of all C e ~o such that, for every B e ~ 0 , either B 5 C
or C 5 B. For example, A 0 e ~- For each Ce ~'let ~(C) be the family of all
Be ~o such that either B 5 C or g(C) s B. Again, Ao e ~(C). Thus (i) is
6 I. Alignments

satisfied by '<6 and by each ~(C). Evidently also (ii) is satisfied by '<6 and by each
~(C).
To show that ~(C) is a tower it remains to show that Be ~(C) implies
g(B) e ~(C), i.e. that B s C and g(C) s B each imply either g(B) s C or
g(C) s g(B). If g(C) s B then also g(C) s g(B), since B s g(B). If B = C, then
g( C) = g(B). If B c: C, then C cannot be a proper subset of g(B), since g(B) \ B
contains at most one element, and hence g(B) s C, since C e '<6 and g(B) e ~ 0 •
Thus ~(C) is a tower. Since ~o is minimal, it follows that ~(C) = ~o for
each C e '<6. Thus if B e ~o and C e '<6, then either B s C or g( C) s B. Thus
g(C) e '<6 and '<6 is a tower. Since ~o is minimal, it follows that '<6 = ~ 0. Hence,
by the definition of '<6, ~o is totally ordered. D

2 ALIGNED SPACES

Let X be a set and let '<6 be a collection of subsets of X. The collection '<6 will
be said to be an alignment on X if it has the following three properties:

(A 1) the set X is itself in '<6,


(A2) the illtersectioll of ally nonempty family of sets in '<6 is again a set in '<6,
(A3) the union of any nonempty family of sets in '<6 which is totally ordered
by inclusion is again a set in '<6.

For example, the collection of all subgroups of a group G is an alignment on


G, and the collection of all subspaces of a vector space V is an alignment on V.
The subsets of X which are in '<6 will be said to be convex. In subsequent
chapters we will study particular types of alignment, and the notion of convex set
will be correspondingly restricted.
For any set S s X, let [SJ denote the intersection of all convex sets which
contain S. Then, by (Al)-(A2), [SJ is itself a convex set, which we call the
convex hull of S.

PROPOSIDON 1 Convex hulls have the following properties:

(Hl) S s [SJ,
(H2) S s T implies [SJ s [T],
(H3) [[S]] = [S],
2. Aligned spaces 7

(H4) the convex hull of any set is the union of the convex hulls of all its
finite subsets.

Proof Properties (Hl) and (H2) follow immediately from the definition of
convex hull, and (H3) restates that [S] is convex. The proof of (H4) requires the
use of (A3) and is not so immediate.
A set S s X will be said to be 'good' if [S] = U [F], where F runs through all
finite subsets of S, and will be said to be 'excellent' if Su His good for every
finite set H s X. Thus every finite subset of X is excellent, and S u H is excellent
if S is excellent and H finite.
Let S be an arbitrary nonempty subset of X and let ~ be the family of all
excellent subsets of S. Then~ is not empty, since it contains every finite subset of
S. By Hausdorffs maximality theorem the family~, partially ordered by inclusion,
contains a maximal totally ordered subfamily ~. Put M = U Te'?J T. We are going
to show that Me~.
If His any finite subset of X, the family {[Tu H]: Te ~} is also totally
ordered by inclusion. Hence if we put CH= UTe'?J [Tu H], then CH is convex.
Since M u H s CH' it follows that [M u H] s CH· On the other hand, T s M
implies [Tu H] s [Mu H] and hence CH s [Mu HJ. Thus CH= [Mu H].
Consequently, if x e [M u H] then x e [Tu HJ for some Te ~. Since T is
excellent, it follows that x e [F u H] for some finite set F s T s M. Thus
M u H is good and M is excellent, as we wished to prove.
By the definition of~' no excellent subset of S properly contains M. Since
Mus is excellent ifs e S\M, it follows that M = S. Thus Sis indeed good. D

The properties (Hl)-(83) alone imply that, for arbitrary sets S,T s X,

[Su T] = [Su [T]J = [[S] u [T]J.

Let 2x be the collection of all subsets of the set X. A map h: 2x ~ 2x is said to


be a hull operator on X if it has the following three properties:

(i) ss h(S),
(ii) S s T implies h(S) s h(T),
(iii) h(h(S)) = h(S).

The hull operator is said to be algebraic if, in addition,


8 I. Alignments

(iv) x e h(S) implies x e h(F) for some finite set F s S.

We have shown that from any alignment on X we can derive an algebraic hull
operator. However, the process can also be reversed. If h: 2x-+ 2x is an algebraic
hull operator on X, then one sees immediately that the collection <(6 of all sets S s X
such that h(S) =Sis an alignment on X. In fact, there is a bijective correspondence
between alignments on X and algebraic hull operators on X. For our purposes it is
more convenient to take alignments as the starting-point.
An alignment <(6 on a set X may be said to be normed if it has the additional
property

(AO) the empty set 0 is in '<6.

There is no loss of generality in restricting attention to normed alignments. For let


'<6 be an arbitrary alignment on X, and put E = nCe<€ C.
Then the collection~ of
all sets {C \ E: C e '<6} is a normed alignment on X, since Ee '<6.
We will assume throughout the remainder of this section that we are given
a set X and a normed alignment '<6 on X. We will say that the pair (X,'<6) is an
aligned space, or simply that Xis an aligned space if the meaning is clear. The
qualification 'nonempty' in the statement of (A2) and (A3) may now be omitted,
and convex hulls now have not only the properties (Hl)-(H4) of Proposition 1
but also the property

(HO) [0] = 0.
The collection '<6 of all convex sets, partially ordered by inclusion, is a
complete lattice. For any family {Ca} of convex sets has an infimum, namely
nCa, and a supremum, namely the intersection of all convex sets which contain
UCa.
If S s X and e e S, then e is said to be an extreme point of the set S if
e ~ [S\e]. Clearly if e is an extreme point of S, then it is also an extreme point of
every subset of S which contains it. From the definition we obtain also

PROPOSITION 2 Let C be a convex set and e e C. Then e is an extreme point


of C if and only if C \ e is convex.

Proof If C \e is convex then [C \e] = C \e, and if C \e is not convex then


[C\e] = C. D
2. Aligned spaces 9

Since any intersection of convex sets is again a convex set, it follows from
Proposition 2 that if an arbitrary collection of extreme points is removed from a
convex set, then the remaining set is convex.
We will denote by E(S) the set of all extreme points of the set S. It can be
characterized in the following way:

PROPOSITION 3 If S s X, the11 E(S) is the i11tersection of all subsets of S


which have the same co11vex hull as S.

Proof If e is an extreme point of Sand if Tis a subset of S with [T] = [S], then
e e T, since [S\e] is a proper subset of [S]. On the other hand, if e e Sis not an
extreme point of S then e e [S \ e]. Hence S s [S \ e] and [S] = [S \ e]. But
e ~ S\e. Cl

A set S s X will be said to be indepe11de11t if every point of Sis an extreme


point of S.
It follows at once from this definition that, for an arbitrary set S s X, the set
E(S) of all extreme points of S is an independent set. Furthermore, any singleton is
an independent set and any subset of an independent set is again an independent set.
On the other hand, it follows from (H4) that an infinite set is independent if every
finite subset is independent. Consequently the union of a totally ordered collection
of independent sets is again an independent set. Hence, by Hausdorff s maximality
theorem, any independent subset of a set is contained in a maximal independent
subset.
A subset T of a set Swill be said to generate S if [T] = [S], and will be said to
be a basis of S if in addition it is independent.
It follows at once from the definitions that a subset T of a set S is a basis of S if
and only if T generates S, but no proper subset of T generates S. Hence a set has a
finite basis if it is generated by a finite subset. Any basis of a set S is a maximal
independent subset of S, but in general the converse is not true.
A subset A of a convex set C will be said to be a face of C if A is convex and
if, for every set S s C,
[S] nA s: [S nA].

It is actually sufficient to require that [F] n A s [F n A] for every finite set


F s C. For then, if S s C and x e [S] n A, there is a finite set F s: S such that
10 I. Alignments

xe [F] nA s [FnA] s [Sn A].

Obviously C is itself a face. The proper faces of a convex set C are the faces other
than c itself.

PROPOSITION 4 The collection of all faces of a convex set C is a normed


alignment on C.

Proof Evidently 0 and C are faces of C. If {A;: i e I} is a family of faces of C


and A = nie/ A;, then A is a convex subset of C. Suppose S s C and
x e [S] n A. Then x e [F] for some finite F s S. Moreover we may assume that
x ~ [F*] for every proper subset F* of F. Then, since x e [F] n A; s [F n A;],
we have F n A;= F for all i e I. Hence F s A; for all i e /,and so F s A.
Consequently x e [F] s [S n A], proving that A is a face of C.
If {A;: i e /} is a family of faces of C which is tot.ally ordered by inclusion and
A= U;eiA;, then A is a convex subset of C. Suppose S s C and x e [S] n A.
Then, for some i e I, x e [S] n A; s [Sn A;]. Hence [S] n A s [Sn A], and
A is a face of C. Cl

The collection of all faces of a convex set C, partially ordered by inclusion, is


also a complete lattice, since any family {Fal of faces has both an infimum, namely
n Fa' and a supremum, namely the intersection of all faces which contain U Fa·
PROPOSITION 5 If A is a/ace of the co11vex set C, then C\A is co11vex.
Proof Taking S = C\A in the definition of a face, we obtain [C\A] nA = 0.
Since [C\A] s C, it follows that [C\A] = C\A. D

PROPOSITION 6 If A,B,C are co11vex sets such that A is a face of B and B is a


face of C, then A is a/ace of C.

Proof Suppose S s C and x e [S] n A. Then x e [S] n B and hence


x e [Sn B], since Bis a face of C. From Sn B s Band x e [Sn B] n A it
follows that x e [S n A], since A is a face of B. D

PROPOSITION 7 If A is a face of the convex set C and B a face of the convex


set D, then A n B is a face of the convex set C n D.
In particular, if A is a face of the convex set C, then A is a face of any
convex set B such that A s B s C.
2. Aligned spaces 11

Proof Certainly E: =A n B is convex. Suppose S s: C n D and x e [S] n E.


From S s C and x e [S] n A we obtain x e [S n A]. From S n A s: D and
x E [S ()A] () B we obtain x E [S () E]. a
PROPOSITION 8 A singleton {e} is a face of a convex set C if and only if
[e] = { e} and e is an extreme point of C.

Proof If {e} is a face of C, then {e} is convex and, taking S = C \ e in the


definition of a face, e ~ [C\ e].
Suppose, on the other hand, that [e] = {e} and e is an extreme point of C. Let
S be any subset of C. If e ~ S then e ~ [S], since C \ e is convex. If e e S, then
e e [S]. In either case, [S] n {e} s: [Sn e]. a

We will say that a set S s: Xis a Radon set if for every partition

we also have [S 1] n [S2] = 0. This implies that for any pair of disjoint subsets
S',S" of S we have [S1 n [S"] = 0, since we can extend the pair to a partition
S 1 =S', S2 = S \ S'. Consequently any subset of a Radon set is again a Radon set.
On the other hand, it follows from (H4) that an infinite set is a Radon set if every
finite subset is a Radon set. Hence the union of a totally ordered collection of
Radon sets is again a Radon set, and any Radon set is contained in a maximal
Radon set. Moreover any singleton is a Radon set, by (HO).
We will say also that a set S s: Xis a Helly set if

flses [S \ s] = 0.

Any subset of a Belly set is again a Belly set. For if Sis a Belly set and Ts: S,
then

= r\es [S\s] = 0.
We now show that an infinite set S is a Belly set if every finite subset is a
Belly set. Indeed if S is not a Belly set, there exists a point s0 e [S \ s] for every
s e S. Taking s =s0 , it follows that there exists a finite set F0 s: S \ s0 such that
s0 e [F0]. If F0 = {s 1, ••• ,snl then, in the same way, for each i = l, ... ,n there
exists a finite set F; s: S \ s; such that s0 e [F;]. It follows that the finite subset
12 I. Alignments

F =F0 u F 1 u ... u Fn is not a Belly set. For if s ~ Fo then F0 =


S \ s and
s0 e [F0] Si [S\s], whereas ifs e F0 thens= s; for some i e { l, ... ,n} and hence

s0 e [F;] =[S \s].


It now follows, as for Radon sets, that the union of a totally ordered collection
of Belly sets is again a Belly set, and any Belly set is contained in a maximal Belly
set. Moreover, any singleton is a Helly set.

PROPOSffiON 9 Every Helly set is a Radon set, and every Radon set is an
independent set.

=
Proof Suppose that the set S X is not a Radon set. Then there exists a partition
*
S = S 1 u S2, S1 nS2=0 with [Si] n [S2] 0. Letx e [S 1] n [S2]. Ifs e S1,
then S 2 s S \s and hence x e [S \s]. Ifs e S 2, then S 1 =
S \s and again
x e [S\s]. Thus x e nses [S\s] and Sis not a Belly set.
On the other hand, by the definition of an extreme point, every point of a
Radon set R is an extreme point of R. Thus any Radon set is an independent set.
a
Let S be a Helly set with ISi = h + 1. (Here, and later, we denote the
cardinality of a set S by ISi.) If s 1, ••• ,sh+l are the elements of S, then the family of
convex sets C; = [S \s;] (i = l, ... ,h+l) has empty intersection, although every
subfamily of h of the sets has nonempty intersection. This statement admits a
converse:

PROPOSITION 10 Suppose that some finite family~ of at least h + 1 convex


sets has empty intersection, although every subfamily of h of the sets has
nonempty intersection. Then there exists a Helly set S s X with ISi = h + 1.

Proof If ~ s ~ is a minimal subfamily with empty intersection, then I~ ~ h + 1.


For each Fe ~ there exists a point ap which belongs to every convex set in ?I
*
except F. Moreover, if F' e ~and F' F, then ap· ap. *
If we put A = UFe'* ap then for every F e ~ we have A\ ap s F. Hence

Thus A is a Helly set containing at least h+ 1 points, and we can take S to be any
subset of A containing exactly h+ 1 points. D

We will say that a set S s Xis a Caratheodory set if


2. Aligned spaces 13

[S] ¢ UseS [S \s].


A CarathCodory set is nonempty and also finite, since the convex hull of any set is
the union of the convex hulls of all its finite subsets. Moreover any Caratheodory
set is an independent set, since ifs e [S\s] then [S] = [S\s]. Evidently also any
singleton is a Caratheodory set. However, a nonempty subset of a CarathCodory
set need not again be a Caratheodory set; an example is given at the end of Chapter
II.
For any set S s X and any x e [S], there exists a Caratheodory set F s S
such that x e [F]. For there exists a finite set F s S such that x e [F], and we can
choose F so that no proper subset has the same property.

Finally we consider some ways of manufacturing new aligned spaces from


given ones.

(i) Restriction: Let'€ be a normed alignment on the set X and let Y be an


arbitrary subset of X. Then the collection~ of all sets C n Y, where Ce '€,is a
normed alignment on Y.
If S s Y, then [S]~ = [S]ce n Y. Hence the extreme points of Sin the aligned
space (Y,~) are the same as the extreme points of Sin the aligned space (X,'€). It
follows that a subset of Y is independent in (Y,~) if and only if it is independent in
(X,'€).

(ii) Join: Let X 1, X 2 be disjoint sets and let '€k be a normed alignment on Xk
(k = 1,2). If X =X 1 u X2 and'€= {C: C = C 1 u C 2, where Ck e '€k (k = 1,2)},
then'€ is a normed alignment on X.
If S s X, then the extreme points of S in the aligned space (X,'€) are the
extreme points of S nX 1 in the aligned space (X 1,'€ 1), together with the extreme
points of Sn X2 in the aligned space (X2,'€ 2). Hence Sis independent in (X,'€) if
and only if Sn Xk is independent in (Xk,'€,J fork= 1,2.

(iii) Contraction: Let'€ be an alignment on the set X and let Y be an arbitrary


subset of X. Then the collection ~ of all subsets D of Y such that

[Du(X\Y)]nY =D
is an alignment on Y. The alignment ~ is normed if and only if X \ Y e "6.
14 I. Alignments

(iv) Intersection: Let~ and 9J be alignments on the same set X. It is easily


verified that h(S) =[S]'« n [S]2b is an algebraic hull operator on X. It follows that
the collection ~ of all sets C n D, where C e ~ and D e ~' is again an alignment
on X, which is normed if either~ or 9J is normed.

3 ANTI-EXCHANGE ALIGNMENTS

An aligned space X will be said to have the anti-exchange property, and its
alignment will be called an anti-exchange alignment, if the following additional
axiom is satisfied:

(.IE) for any distinct x,y e X and any S s X, if ye [S u x] but ye: [S], then
x e: [Su y].
This definition can be reformulated in several other ways:

PROPOSITION 11 An aligned space X has the anti-exchange property if and


only if one of the following equivalent properties holds:

(i) a subset C of a convex set D is a maximal proper convex subset of D if and


only if D \ C = {e}, where e is an extreme point of D;
(ii) any set S s X has the same extreme poillts as its convex hull [S];
(iii) for any convex set C c: X and any point x e X \ C, the set [Cu x] \x is
convex;
(iv) for any finite set S s X, [S] = [E(S)].

Proof Let C be a subset of the convex set D. If ID\ Cl = 1 then, by Proposition 2,


C is a maximal proper convex subset of D if and only if D \ C = { e}, where e is an
extreme point of D. Suppose ID\ Cl > 1 and let x,y be distinct points of D \ C. If
ye: [Cu x] then E: =[Cu x] is a convex set such that C c: E c D, and if
x e: [Cu y] then E': =[Cu y] is a convex set such that C c: E' c D. It follows
that ()E) => (i).
If S s X and e is an extreme point of [S] then e e S, by Proposition 3, and
hence e is an extreme point of S. Assume (i) holds, but (ii) does not. Then, for
some S s X and some e e S, e e: C: =[S \ e] and T: =[S] \ e is not convex. Hence
C c T and [S] = [71 =Tue. Consequently C is not a maximal proper convex
3. Anti-exchange alignments 15

subset of [S]. Let 9) be the family of all convex sets D such that C c: D c: [S].
Then 9), partially ordered by inclusion, contains a maximal totally ordered
subfamily ~. Evidently M = UDe'!f D is convex and M c: T, since e ~ M.
Consequently M is not a maximal proper convex subset of [S], which contradicts
the definition of ~. Thus (i) => (ii).
If C c: X is convex and x e X \ C, then x is an extreme point of the set C u x.
If (ii) holds, then xis also an extreme point of the set [Cu x] and hence [Cu x] \x
is convex, by Proposition 2. Thus (ii) => (iii).
Suppose S is a finite set such that [E(S)] c: [S]. Let T be a maximal subset of
S containing E(S) such that [T] c: [S]. Then there exists x e S\ [T]. If (iii) holds,
then C =[Tux] \xis convex. Since Tis maximal, it follows that [Tux]= [S]
and Sn C = T. Hence S =Tux and x ~ [S \x]. Since xis not an extreme point
of S, this is a contradiction. Thus (iii)=> (iv).
Suppose that for some x,y e X and some S s: X, ye [Su x] and x e
[Su y], but y ~ [S]. Then there exists a finite set F s: S such that ye [Fu x] and
x e [Fu y]. Put T =Fu {x,y}. If (iv) holds, then [T] = [E(nJ. Since x and y
are not extreme points of T, it follows that [T] = [F]. But this is a contradiction,
since [F] s: [S] and y ~ [S]. Thus (iv) => (lE). D

Throughout the remainder of this section we assume that the aligned space
X has the anti-exchange property and we derive some further simple properties.

PROPOSITION 12 If S s: X, then S has a basis if and only if [S] = [E(S)].


Moreover E(S) is then the unique basis of S.

Proof If T is a basis of S, then [T] = [S] and every point of T is an extreme point
of T. But, by Proposition 1 l(ii), T and S have the same extreme points. Hence
T = E(S) and [S] = [E(S)].
On the other hand if [S] = [E(S)] then E(S) is a basis of S, since it is
independent. a
It follows from Proposition 12 that a set has a basis if and only if its convex
hull has a basis, and the bases are then the same. Also, if S has a basis and
E(S) s: T, then [S] s: [T].
It will now be shown that if S is an independent set and T s: S, then

T = [T] n S.
16 I. Alignments

Indeed if x e [T] r. S , then x is an extreme point of S and hence also of [S].


Consequently x is an extreme point of the subset [T] and hence also of T. Thus
[T] r. S s T, which is all that requires proof.
Conversely, a set Sis independent if for each Ts S there exists a convex set
C such that T = C r. S . For suppose some x e S is not an extreme point of S.
Thenx e [S\x]. By hypothesis there exists a convex set C such thatS\x =Cr. S.
Then [S\x] s C. Hencex e Cr. S, which is a contradiction.
Again, if S and T are sets such that [SJ = [T] and if S is finite, then
[S] =[Sr. T]. For, by Proposition 1 l(ii), Sand T have the same set E of extreme
points. Thus E s S r. T. Since [E] = [S], by Proposition 1l(iv), the claim
follows.
It follows directly from the definition that the restriction of an aligned space X
with the anti-exchange property to an arbitrary subset Yagain has the anti-exchange
property. Furthermore the join of two disjoint aligned spaces with the anti-
exchange property again has the anti-exchange property. Finally, the intersection of
two anti-exchange alignments on the same set X is again an anti-exchange
alignment

4 ANTIMATROIDS

An aligned space X with the anti-exchange property is said to be an


antimatroid if it contains only finitely many points. Even if we were not initially
interested in this case, we could be led to it by restriction of an alignment on a given
set X to a finite subset Y.
We now consider some results which hold only in the finite case.

PROPOSITION 13 Let X be a finite set and'€ a collection of subsets of X.


Then (X,"6) is an aligned space if and only if

(o) 0, Xe "6,
(i) C,D e 'l6 implies Cr.De '€.

Moreover (X,'€) has the anti-exchange property if and only if, in addition,

(ii) Ce 'l6 and C :1:X imply C ux e '€/or some x e X\C.


4. Antimatroids 17

Proof Since Xis finite, (o) and (i) imply that the intersection of any family of sets
in ~ is again a set in ~' and it is trivially true that the union of any family of sets in
~ which is totally ordered by inclusion is again a set in~. Hence (X,~) is an
aligned space if and only if (o) and (i) hold
Suppose (X,~) is an aligned space with the anti-exchange property, Ce ~and
C -¢X. If C0: =X then, by Proposition ll(iv), C0 = [E(C0)]. Hence there exists a
point e 1 e E(C0) \C. Then, by Proposition 2, C 1 = C 0 \e 1 is also convex and
Cs C 1• If C -:1: C 1, let e2 e E(C 1)\C. Then in the same way C 2 = C 1 \e 2 is
convex and C s C2• If C '¢ C2 we can repeat the process. Since X is finite, there is
a least positive integer k such that C = Ck· Then C u ek is convex.
Suppose now that the aligned space (X,~) has the property (ii). If Ce ~and
x e X \ C, then there exists a chain of sets in~:

Hence x = Xs for some s with 1 < s < 11. Moreover, since [C u x] s


Cu {x 1, ••• ,xs}, we have [Cu x] =Cu {x;1, ••• ,x;), where 1 < i1 < ... < ir = s
Furthermore

[Cu x] \x =Cu {x; 1, ••• ,x;,_ 1 } =[Cu x] rt (Cu {x 1, ••. ,xs-t}) e ~­

Hence the aligned space (X,~) has the anti-exchange property, by Proposition
11(iii). a
In the statement of Proposition 13 we can evidently omit the requirement
X e ~ in (o) when (ii) holds. It is worth noting also that, although the anti-
exchange property is obviously preserved under restriction of an aligned space to an
arbitrary subset, it is not obvious that the property (ii) is preserved.
Antimatroids can also be characterized in the following way:

PROPOSITION 14 Let X be a fi11ite set and~ a collection of subsets of X.


Then (X,~) is a11 aligned space with the anti-excha11ge property if and only if

(o) 0,X e ~'


(i) if C,D e ~and C ~ D, then C\c e ~for some c e C\D.

Proof Suppose first that the aligned space (X,~) has the anti-exchange property,
and let C,D be convex sets with C ~ D. Since C is the convex hull of its extreme
18 I. Alignments

points, by Proposition ll(iv), some extreme point c of C is not contained in D.


Then C \ c is convex.
Suppose next that (X,'€) has the properties (o),(i). We show first that if
C ,D e '16, then C n D e '16. Evidently we may assume that C ~ D. Then
Ci =C \ c e '16 for some c e C \ D and Ci n D =C n D. If Ci s D, then
C n D = Ci e '€. If Ci <I: D, we can repeat the process. Since X is finite, the
process must eventually terminate, yielding C n D = Ci e '€.
By Proposition 13 it only remains to show that if Ce '16 and C '# X, then
C u x e '16 for some x e X \ C. By (i) there exists Xi e X \ C such that
X 1 =X\xi e '16. Evidently Cs X1• IfC =X 1, there is nothing more to do. If
C '# X 1, the argument can be repeated. The process must eventually terminate,
yielding c = xk and Xk-1 =cu Xk e '16. D

There is an interesting connection between this result and linguistics. Our


language has as its alphabet the elements of the finite set X. The admissible
words of our language are the finite sequences a = x 1 ••• xn,, where
X; e E(X \ {x1, ••• ,x;_ 1}) for 1 < i < m. This language obviously has the properties

(o) 0 is an admissible word;


(i) if a= x 1 ••• Xm is an admissible word, then x; '# Xj for 1 < i <j < m;
(ii) there exists an admissible word a= x 1 ••. Xn such that, for every x e X, x = X;
for some i.

Another property of our language is

(iii) if a= Xi and f3 = Yt ... Yn are admissible words, and if X; E {y 1, ... ,ynl


••• Xm

for some i, then 'Y =y 1 ... Y,,Xj is an admissible word for some xi e {x1, ••• ,xm}.

To see this, takej to be the least i for which X; E {y1,... ,ynl· Then there exist
distinct Y; 1, ••• ,y;i_1 e {y 1, ••• ,ynl such that x 1 = Y;1, ••• ,xi-I = Y;i-l . Since

X \ {y1, ... ,yn} s X \ {Y;1, ... ,y;i_1 },

we have xi e E(X \ {yi, ... ,ynD· Hence 'Y =y 1 ••• Y,,Xj is an admissible word.
A language with the properties (o)-(iii) is sometimes called a shelling
structure. It follows from Proposition 14 that, conversely, if X is the finite
alphabet of a shelling structure then an alignment with the anti-exchange property is
4. Antimatroids 19

obtained by taking " to be the collection of all sets C = X \ {x 1, ••. ,xm}, where
a =x 1 ••• Xm is an admissible word of the shelling structure.
Another result for finite aligned spaces with the anti-exchange property is the
following:

PROPOSITION 15 If Xis an antimatroid and if S s Xis a Helly set, then


there exists a Helly set C with ICI = ISi which is also convex.

Proof Obviously we may assume that S itself is not convex. For each x e [S] \ S
we have x 4! [S \ s] for some s e S. Choose such an x so that the number of s e S
for which x 4! [S \ s] is a minimum and let e be one such s. Then e is an extreme
point of S, since x e [S].
Let T =x u (S \e), so that Ts X and ITI = ISi- Since [S] \ e is convex, we
have [T] s [S] \e. We claim that Tis a Helly set. To establish this we will show
that for each ye [T] we have y 4! [T\t] for some t e T. If y 4! [S\e] we can take
t =x, since T\x =S\e.
Thus we now suppose that y e [S \ e]. Assume first that for some t e S we
have y 4! [S\t]. If x e [S\t], then t e T andx is the only element of T\t which is
not in S\t. Sincex e [S\t], it follows that [T\t] s [S\t] and hence y 4! [T\t].
Hence we may now assume that x 4! [S \ t] for every t e S for which
y 4! [S \ t]. Since x 4! [S \ e] and ye [S \ e] this implies, by the choice of x, that
y ES.
Since y e S and S is a Helly set, y 4! [S \ t] for some t e S. Evidently this
implies that y = t and that y is an extreme point of S. Moreover y -:1: e, since
ye [S \e], and hence ye T. If we put D = [S \y], then x 4! D but x e [S] =
[y u D]. Hence, by the anti-exchange property, y 4! [x u D] and, a fortiori,
y 4! [T\y]. Thus our claim is now established.
In this way we have replaced the Helly set S by another Helly set T with the
same cardinality but with [T] c: [S]. If Tis not convex we can repeat the argument
Since Xis finite, we must eventually arrive at a convex Helly set which has the
same cardinality as S. a
It should be noted that, in any aligned space, a convex set is a Helly set if and
only if it is independent, by the definitions and Proposition 9.
20 1. Alignments

S EXCHANGE ALIGNMENTS

Throughout this section the convex sets of the aligned space X will be called
affine sets. Furthermore the convex hull of a set S s X will be called its affine
hull and will be denoted by <S>, rather than by [S]. The reason for these changes
in terminology and notation is that in Chapter Ill, where the results of this section
will first be used, we will be concerned with a set X which is equipped with two
different alignments and it will be necessary to distinguish between them.
An aligned space X will be said to have the exchange property, and its
alignment will be said to be an exchange alignment, if the following additional
axiom is satisfied:

(E) for any x,y e X and any S Sii X, if ye <Su x> but y i <S>, then
x e <S uy>.

This definition can be reformulated in several other ways:

PROPOSIDON 16 An aligned space X has the exchange property if and only if


one of the following equivalent properties holds:

(i) for any S c X and any x e X \ <S>, if S is an independent set, then S u x


is also a11 independent set;
(ii) for any S s X, if T is a maximal independent subset of S, then T is a basis
ofS;
(iii) for any S s X, if T is an independent subset of S, then S has a basis B
such that T s B;
(iv) for any affine set A c X and any x e X \A, there is no affine set A' such
that Ac A' c <Aux>.

Proof Suppose first that Sis an independent set and xi <S>. If Su xis not
independent then, for some ye Su x, ye <(Su x) \y>. In fact ye S, since
x i <S>, and y i <S \y>, since S is independent. If X has the exchange property it
follows that x e <(S \y) u y> = <S>, which is a contradiction. This proves that
(E) => (i).
Suppose next that T is a maximal independent subset of the set S. If (i) holds,
5. Exchange alignments 21

then S s <I> and hence <S> = <T>. Thus T generates S and actually, since it is
independent, Tis a basis of S. Thus (i) =>(ii).
Suppose now that Tis an independent subset of the set S. The family ?"; of all
independent subsets of S which contain T, partially ordered by inclusion, contains a
maximal totally ordered subfamily ~. The union B of all the sets in ~ is an
independent set containing T. Moreover B is a maximal independent subset of S.
Hence (ii) => (iii).
Suppose A is an affine set, x ~ A, and there exists an affine set A' such that
A c A' c <A u x>. If (iii) holds, then A has a basis B and A' has a basis
B' ::> B. Furthermore B' u x has a basis B" containing B'. Since <B'> =A' and
<B "> = <A u x>, we must have B" = B' u x. If y e B' \ B, then
y i <(B' u x) \y>, since B' u xis independent. But Bux s (B' u x) \y and
<B u x> = <A u x>. Since ye <A u x>, this is a contradiction. This proves that
(iii) => (iv).
Finally suppose y e <Su x> but y i <S>. Then x i <.S>. If (iv) holds, then
<.S u y> = <.Su .t> and hence x e <Su y>. Thus (iv) => (E). C

The property (iv) is known as the covering property, since in any partially
ordered set (in our case, the lattice of affine sets) an element B is said to cover an
element A if A < B, but there is no element C such that A < C < B.
Throughout the remainder of this section we assume that the aligned space
X has the exchange property. It follows from Proposition 16(ii) that any subset of
X has a basis. Furthermore,

PROPOSmON 17 If T and T' are subsets of S such that T is independent and


T' generates S, then there is a subset T" ofT' such that Tu T" is a basis of S.

Proof Evidently <Tu T'> = <S>. Let ?"; be the collection of all independent sets
U such that T s U s T u T'. Then ~ is not empty, since it contains T. If we
partially order ~ by inclusion then, by Hausdorffs maximality theorem, ?";
contains a maximal totally ordered subcollection V. The union V of all sets in Vis
an independent set such that T s V s Tu T'. Thus V = T u T", where T" s T'
and T n T" = 0. Since V is a maximal independent subset of T u T', we have
<V> =<Tu T'>, by Proposition 16(ii). C
22 I. Alignments

PROPOSITION 18 If B 1 and B2 are bases of a set S, and if bi e B 1 \B2, then


there exists b2 e B 2 \Bi such that (B 1 \ b1) u b2 is also a basis of S.

Proof Since Bi is independent, b 1 i <.B 1 \b 1> and hence, since <.B 2> =<S>,
there exists b 2 e B 2 such that b 2 i <.B 1 \b 1>. Since b2 e <(Bi \bi) u b 1>, it
follows from the exchange property that b1 e <.B 3>, where B3 = (B 1 \b 1) u b2.
Since also B 1\b 1 s <8 3>, it follows that <.B3> = <S>.
It remains to show that B3 is independent. Assume on the contrary that, for
some b 3 e B 1 \b 1, we have b 3 e <(B 1 \ {b 1 ,b 3 }) u b 2>. Since b 3 i
<.Bi\ {b 1,b3 }>, it follows from the exchange property again that b2 e <Bi \bi>,
which is a contradiction. a

An affine set H c: Xis said to be a hyperplane (of X) if X covers H, i.e. if H


is a maximal proper affine subset of X. Equivalently, an affine set H c: X is a
hyperplane if <H u x> = X for every x e X \ H.
It follows that if H 1,H2 are distinct hyperplanes, then H 1 ~ H 2• Furthermore,
if B' is a basis of a hyperplane Hand x e X\H, then B' u xis a basis of X.
We now show that if A is an affine set and x e X \A, then there is a
hyperplane H such that A s H and x i H. Consider the collection ~ of all affine
subsets of X which contain A but do not contain x. If we suppose ~ partially
ordered by inclusion then, by Hausdorffs maximality theorem, there exists a
maximal totally ordered subcollection ~o- The union H of all affine sets in ~0 is an
affine set which contains A but not x. Thus if y i H, then x e <H u y> and
hence, by (E), y e <H u x>. Consequently X = <Hux> and X covers H.
It follows that if H 1,H2 are hyperplanes and if y e X \ (H 1 u H 2), x e
H1 \H2, then there is a hyperplane H 3 such that (H 1 n H 2) u ye H 3 and xi H 3.
For <(H 1 n H ~ u y> is an affine set which does not contain x.

PROPOSITION 19 If B is an independent set and S = {x1, ... ,xn} a finite set


such that x 1 i <.B> and xii <.B u {x 1, ..• ,xj-l }>for j = 2, ... ,n, then Bu Sis
an independent set.

Proof The result holds for n = l, by Proposition 16(i). We suppose n > 1 and use
induction on n. Then Bu S', where S' = {x1,... ,xn_iJ, is independent. Since
Xn i <.B u S'>, it follows that also B u s is independent. a
5. Exchange alignments 23

PROPOSITION 20 Let B be an independent set and let S,T be finite sets. If


B n S = 0 and Bu Sis an independent set such that <lJ u S> s <lJ u T>,
then ISi s ITI·
Proof Let S = {x1, ••• ,xml and T = {y 1, ••• ,ynl· Since x 1 e <Bu T> the set
Bu S 1, where Si = {xi,y 1, ••. ,yn}, is not independent. Consequently, by
Proposition 19, Y;1 e <B u {xi,y 1, ••• ,y;1_ 1 }> for some Y;1 e T. Thus if we put
Ti =x1 u (T\y;1), then Y;1 e <Bu Ti> and <Bu T1> =<Bu T>. Similarly the
set Bu S 2 , where S 2 = {x2 ,x 1,T\y;1 }, is not independent. Hence for some
Y;2 e T \y;1 we have Y;2 e <Bu T2>, where T2 = {x2,xi, T \ (y; 1 u Y;2) }, and
<Bu T2> =<Bu Ti>= <Bu T>. If n < m, then by proceeding in this way we
would obtain <lJ u {x1, •.. ,xn}> =<Bu T> and hence Xm e <B u {xi, ... ,xn}>,
which is a contradiction. a
PROPOSITION 21 Let A be a proper affine subset of X and let B u S be a
basis of X, where B is a basis of A and S is finite.
If B' u S' is any basis of X such that B' is a basis of A, then S' is finite
and IS1 =ISi.
Proof It follows at once from Proposition 20 that if B u T is a basis of X, then T
is finite and ITI = ISi. If B' is any basis for A then, by Proposition 17, X has a
basis of the form B' u T', where T's Bu S. Evidently T' n B =0 and hence
T' s S. But now, in the same way, X has a basis of the form Bu T, where
T s T'. Hence, by the first part of the proof, T = T' = S. Moreover, if B' u S' is a
basis of x, then S' is finite and 1s1 = ISi· a
If A is a proper affine subset of X, then X has a basis of the form B u S,
where B is a basis of A. We will say that A has finite codimension d in X if S is
finite and ISi =d, and that A has infinite codimension in X if S is infinite. By
Proposition 21, these definitions do not depend on the choice of B and S. Evidently
an affine set is a hyperplane if and only if it has codimension 1.
In particular we can take A = 0. We will say that X has dimension d if all its
bases arc finite and have cardinality d + l, and that it is infinite-dimensional if all
its bases are infinite.
In these definitions we may replace X by an affine subset. If A1 and A2 arc
finite-dimensional affine sets with A 1 c A2 then the difference in their dimensions,
24 I. Alignments

dim A2 - dim Alt is equal to the codimension of Ai in A2 • This holds also for
Ai = 0 if we assign to the empty set the dimension -1.

PROPOSmON 22 If A1 and A2 are affine sets such that Ai n A2 has finite


codimension n i11 A2, the11 Ai has fi11ite codimension m in <Aiu A 2>, where
m Sn.
· In particular, if Ai and A2 are fi11ite-dimensional affine sets, then Ai n A2
and <Ai u A2> are also finite-dimensional affine sets and

Proof Let B be a basis of Ai n A 2• Then A 1 has a basis of the form B u B 1 and


A 2 has a basis of the form Bu B 2 , where IB2 1 = n. Since <A 1 u A 2 > =
<Aiu B 2>, it follows that <Aiu A 2> has a basis of the form Bu B 1 u B 2',
where B 2' s B2• C

It follows directly from the definition that the restriction of an aligned space X
with the exchange property to an arbitrary subset Y again has the exchange
property. Furthermore the join of two disjoint aligned spaces with the exchange
property again has the exchange property.

6 NOTES

The term alignment originates with Jamison-Waldner (1982). The proof of


Proposition 1 given here is sketched in the dissertation of Jamison (1974).
There is also a bijective correspondence between hull operators on X and
collections of subsets of X with only the properties (Al)-(A2) in the definition of
an alignment, but the proof of this is almost trivial. Moreover, Sierksma (1984)
shows that in a sense no generality is lost by restricting attention to algebraic hull
operators and alignments.
An example of a set X and a collection '€ of subsets of X having the properties
(Al) and (A2), but not the property (A3), is the set X =R with '€ the collection
of all closed subsets of R. If C n = [1/n,l] (n e N ), then C n e '€ and
C1 S C2 S ••• , but UneN Cn = (0, 1] ~ '€.
For completeness it should be mentioned that property (A3) in the definition
of an alignment may be replaced by
6. Notes 25

(A3)' the union of any nonempty directed family of sets in ~ is again a set in
~-

Here a family ~ of sets is said to be directed if A,B e ~ implies A u B s C


for some C e ~. It is trivial that (A3)' => (A3), and it follows from Proposition 1
that (Al)-(A3) => (A3)'. Actually (A3) => (A3)'; a direct proof is given in
Erne (1984).
By omitting the convexity requirements in the definition of a face of a convex
set there is obtained the more general concept of an extreme subset of a set. The
properties of extreme subsets are discussed by Lassak (1986).
The concepts of Helly, Radon and Caratheodory sets were introduced in 1976
by Soltan, who called them h-, r- and c-independent sets; see Soltan (1984). In
Chapter III they will be related to the classical theorems of Helly, Radon and
Caratheodory for convex sets in Euclidean space. The concepts replace with
advantage the older concepts of Helly, Radon and Caratheodory numbers. In our
terminology the Helly (Caratheodory) number of an aligned space is the maximum
cardinality of any Helly (Caratheodory) set, whereas the Radon number is one more
than the maximum cardinality of any Radon set.
Some significant examples of aligned spaces with the anti-exchange property
will be given in Chapter II. Further results on finite aligned spaces with the anti-
exchange property are given in the surveys by Edelman and Jamison (1985) and
Duchet (1987). We have preferred the name 'antimatroid' to their 'convex
geometry', since the essential nature of convexity seems to us to be captured more
by the axiom (C) of the next chapter, than by the axiom (JE) of this chapter.
Shelling structures are studied by Korte and Lov,sz (1984), who attribute to
Bjorner the recognition of their equivalence with antimatroids. The term 'alternative
precedence structure' is preferred by Bjomer and Ziegler in White (1992), to avoid
confusion with the shellability of polyhedral complexes considered, for example, by
Ziegler (1995).
The exchange property (E) was first formulated, for real vector spaces, by
Grassmann (1844), §20. Matroids, i.e. finite aligned spaces with the exchange
property, have been extensively studied; see, for example, Welsh (1976), Oxley
(1992) or White (1992). They provide a common framework for independence in
algebra, incidence in geometry, partial transversals in families of subsets, and
26 I. Alignments

disjoint unions of trees in graphs. The resulting cross-fertilization has proved


remarkably fruitful.
II
Convexity

In this chapter we introduce the notion of a convex geometry, which is defined by


just two axioms involving points and sets of points called segments. A set is
convex if it contains, together with any two points, the segment which they
determine. The connection with the previous chapter is that the collection of all
convex sets is an alignment.
A real vector space is a convex geometry, and so is a vector space over any
ordered division ring. We give also other interesting examples of quite a different
nature. A basic property of convex geometries is that the convex hull of the union
of two convex sets C,D is the union of the convex hulls of all sets {c,d}, where
c e C and d e D. Already in this general setting it is possible to establish several
properties which were first observed for convex sets in a real vector space. In a
convex geometry also, extreme points and faces may be characterized more simply
than in an arbitrary alignment
Four more axioms are then introduced, each of which on its own ensures some
useful additional property of a convex geometry. These new axioms are also
satisfied in any vector space over an ordered division ring, which is the underlying
reason for our interest in them. Some further properties of Helly, Radon and
Caratheodory sets are established under the hypotheses of this chapter. Finally,
examples are given which show the independence of the six axioms and the
necessity of the axioms for some of the properties which have been established.

1 EXAMPLES

'Ordinary' convexity is defined in the following way:


28 II. Convexity

EXAMPLE 1 Let X be a vector space over the field R of real numbers. For any
x,y e X, define the segment [x,y] to be the set of all z e X which can be
represented in the form z =Ax+ (1 -A.)y, where A. e Rand 0 ~A.~ 1. Then a set
C =Xis said to be convex if x,y e C implies [x,y] s C.
For example, the subsets of R2 in Figure 1 are convex, but those in Figure 2
are not.

''''''''''''''''''''''
,,,,,,,,,,,,,,,,,,,,,,
''''''''''''''''''''''
,,,,,,,,,,,,,,,,,,,,,,
''''''''''''''''''''''
''''''''''''''''''''''
''''''''''''''''''''''
''''''''''''''''''''''
,,,,,,,,,,,,,,,,,,,,,,
''''''''''''''''''''''
''''''''''''''''''''''

Figure 1: Some convex sets in R 2

,,,,, ,,,,,,
'''''
,,,,,
''''' ''''''
''''''
'''''
,,,,, ''''''
'''''
''''' ''''''
''''''
'''''
''''' ''''''
''''''
'''''
,,,,,
''''' ''''''
''''''
,,,,,,
''''' ''''''
,,,,,,,,,,,,,,,,,,,
,,,,,,,,,,,,,,,,,,,
'''''''''''''''''''
,,,,,,,,,,,,,,,,,,,
'''''''''''''''''''
'''''''''''''''''''

Figure 2: Some non-convex sets in R 2

Many elementary properties of convex sets hold also in other structures, since
their proofs do not make full use of the preceding definition. Here are some
examples of the structures we have in mind. In each case we associate with any
two elements x,y of a set X a subset [x,y] of X containing them, and we then define
a set Cs X to be convex if x,y e C implies [x,y] s C.

EXAMPLE 2 In Example 1 the field R of real numbers can be replaced by the field
Q of rational numbers, leaving everything else unchanged. More generally, the
field R can be replaced by an arbitrary ordered division ring D.
A division ring differs from a field only in that multiplication need not be
commutative. A division ring D is ordered if it contains a subset P of positive
elements which is closed under addition and multiplication and is such that D is the
disjoint union of the sets {0}, P and - P = {- A.: A. e P}. If A.,µ e D we write
A. < µ when either µ - A. = 0 or µ - A. e P.
1. Examples 29

Any ordered field K can be embedded in a larger ordered field in the following
way. Let K(t) denote the field of all rational functions in one indeterminate t with
coefficients from K. An element of K(t) can be uniquely expressed in the form
f / g, where f and g are relatively prime polynomials in t with coefficients from K
and g has leading coefficient 1. If we define such an element to be 'positive' when
the leading coefficient off is positive, then K(t) also acquires the structure of an
ordered field. (This ordered field is 'non-archimedean ', since t > n for every
positive integer n.)
Any ordered field K can also be embedded in an ordered division ring which is
not a field. We sketch the construction, due to Hilbert, without giving detailed
proofs. The set L = L(K) of all formal Laurent series a = I, keZ aktk, where ak e K
and ak ~ 0 for at most finitely many k < 0, is a field if addition and multiplication
are defined by

where 'Yk = I,;+ j=k a;~j· Moreover, L is an ordered field if a = I, keZ aktk is
defined to be positive when, for some m e Z, am > 0 and ak = 0 if k < m. If
p e K, p > 0 and p ~ 1, then the map v: L ~ L defined by
'If (a) = I, keZ pkaktk if a = I, keZ aktk
is a nontrivial automorphism of the field L which preserves positivity.
Now let~= ~(K) be the set of all formal Laurent series A = I.veZ avsv,
where av e L and av ~ 0 for at most finitely many v < 0. If we define addition and
positivity as before, but define multiplication by

where Cv = I, A.+µ=v a1'lfA.(bµ), then ~ is a noncommutative ordered division ring


which contains K as a subfield.

EXAMPLE 3 In Example 1 or Example 2 we can take X to be not the whole vector


space, but a given nonempty convex subset. Thus we admit only those convex sets
in the vector space which are contained in X.
For example, we can take X to be the subset of Rm2 consisting of all mxm
positive definite symmetric matrices.
30 II. Convexity

EXAMPLE 4 Let X be a partially ordered set. For any x,y e X, define the
segment [x,y] to be the set {x,y} if x and y are incomparable, to be the set of all
ze X such thatx < z Sy if x < y, and to be the set of all ze X such thaty < z <x if
y Sx.

EXAMPLE 5 Let X be a Goin) semilattice, i.e., a partially ordered set in which


any two elementsx,y have a least upper boundx v y. For any x,y e X, define the
segment [x,y] to be the set {x,y,x v y}.
For example, we can take X to be the set of all subsets of a finite set E,
partially ordered by inclusion. In this case the least upper bound of two subsets x,y
of E is their union x u y.

EXAMPLE 6 Let X be a tree (i.e., a finite connected graph without circuits) and,
if x,y are vertices of X, define the segment [x,y] to be the set of all vertices z of X
which lie on the (unique) shortest path fromx toy.

EXAMPLE 7 Let X be a vector space over the field A = R(t) of all rational
functions in one variable with real coefficients. For any x,y e X define the segment
[x,y] to be the set of all z e X such that z = A.x + (1 - A.)y, where A. e A and
0 < A.(t) < 1 for all t e JR for which A.(t) is defined (i.e., for which the denominator
of A. does not vanish).

EXAMPLE 8 Let X be the set of all measurable functions x: T ~ JRn, where Tis a
given measure space. For any x,y e X, define the segment [x,y] to be the set of all
z e X such that z(t) e {x(t),y(t)} for all t e T, except a set of measure zero, i.e.
z ='XEX + (1 - 'XE)y, where 'XE is the characteristic function of some measurable set
Es T.

In order to make it apparent that a number of properties of convex sets hold


also in these structures we are going to adopt an axiomatic approach to convexity.
This approach has the merit of exhibiting the logical structure of the subject and of
establishing results under minimal hypotheses. It may also be argued that it forces
us to formulate definitions and construct proofs in the 'right' way. After studying
the general properties of 'convex geometries' we will return to the preceding
examples at the end of the chapter.
2. Convex geometries 31

2 CONVEX GEOMETRIES

Let X be a set and suppose that with any unordered pair {a,b} of elements of X
there is associated a subset [a,b] of X containing a and b. The elements of X will be
called points and the subsets [a,b] segments. We will say that a convex geometry
is defined on X if the following two axioms are satisfied:

(C) if c e [a,bi] and de [c,b2], then de [a,b]for some be [b1,b2];


(Ll) [a,a] ={a}.

Throughout this section and the next we will assume that a set Xis given
on which a convex geometry is defined. We are going to study the consequences
of this assumption.

Proof Let c e [c 1,c 2]. From c 1 e [a,b] and c e [c1'c 2 ] we obtain, by (C),
c e [a,b1 for some b' e [b,c2]. From c2 e [a,b] and b' e [b,c2] we obtain, by (C)
and (Ll), b' e [a,b]. From b' e [a,b] and c e [a,b1 we obtain, by (C) and (Ll)
again, c e [a,b ]. a
We show next that the axioms (C) and (Ll) together imply a generalization of
axiom (C):

PROPOSITION 2 If c 1 e [a,b 1], c2 e [a,b2] and c e [c 1,c2], thence [a,b]for


some b e [b 1,bi].

Proof Applying (C) twice, we obtain c e [a,b1 for some b' e [b 1,c 2 ] and
b' e [a,b] for some be [b 1,b2]. From c e [a,b1 and a,b' e [a,b] we now obtain
c e [a,b], by Proposition 1. a

We define a subset C of X to be convex if x,y e C implies [x,y] s C. For


example, the segment [x,y] is itself convex, by Proposition 1.
It follows at once from the definition that the collection ~ of all convex sets is
a normed alignment on X. Thus Xis an aligned space, and we may use the
definitions and results of Chapter I, Section 2. In particular, we define the convex
hull [S] of any set S s X to be the intersection of all convex sets which contain S.
Our notations are consistent, since the convex hull of the set {x,y} is the segment
32 II. Convexity

[x,y]. However, convex geometries possess many properties not possessed by all
aligned spaces, as we will now see.

PROPOSITION 3 For any nonempty convex set C and any point a i! C,

[a UC] = UceC [a,c].

Proof Suppose x 1 e [a,c 1] and x 2 e [a,c2], where c 1,c2 e C. If x e [x1,x2] then,


by Proposition 2, x e [a,c] for some c e [c 1,c2] s C. This proves that the set
D = Ucec [a,c] is convex. It is obvious that D is contained in every convex set
which contains both c and a. a

Since [x u [S]] = [x u S], it follows from Proposition 3 that for any


nonempty set S,
[x u S] = Uye [SJ [x,y].

In particular, if x e [a,b,c], then x e [a,d] for some de [b,c]. However, much


more is true:

PROPOSITION 4 For any nonempty sets S,T s X,

[SU T] = Uxe[S],ye[1l [x,y].

Proof Put
R = Uxe [SJ.ye [11 [x,y].

Since S u T s R s [S u T], we need only prove that R is convex. Thus we wish


to show that if z' e [x',y1 and z" e [x",y"], where x',x" e [S] and y',y" e [T], and if
z e [z',z"], then z e [x,y] for some x e [S] and ye [T].
Since z e [x',y',z"] s [x',x",y',y"], we have z e [x',w] for some
w e [x",y',y'']. Then we [x",y] for some ye [y',y"] s [T]. Thus z e [x',x",y],
and hence z e [x,y] for some x e [x',x"] s [S]. a
In convex geometries there is a simple characterization of extreme points:

PROPOSITION 5 If S s X and e e S, then e is an extreme point of S if and


only if e e [x,y], where x,y e S, implies e e {x,y}.

Proof If e e [x,y], where x,y e S \ e, then e e [S \ e] and hence e is not an


extreme point of S.
2. Convex geometries 33

Suppose, on the other hand, that e e [x,y], where x,y e S, implies x = e or


y = e. If e is not an extreme point of S, then e e [S\e]. Hence e e [F] for some
finite set F s S\e. We may assume that e E [F1 for every proper subset F' of F.
If u e F then, by Proposition 4, e e [u, v] for some v e [F \ u]. Since v '¢ e, it
follows that u =e. Since u E S\e, this is a contradiction. a
By Proposition I.8 and (Ll), a singleton {e} is a face of a convex set C if and
only if e is an extreme point of C. In convex geometries there is also a simple
characterization of arbitrary faces. As with intervals in R, for any a,b e X we
define 'half-open' and 'open' segments by

[a,b) = [a,b] \b, (a,b] = [a,b] \a, (a,b) = [a,b] \ {a,b} ..

PROPOSITION 6 A subset A of a convex set C is a face of C if and 011/y if A is


convex and, for any c,c' e C, (c,c') ("\A'¢ 0 implies c,c' e A.

Proof Let A be a convex subset of C. If F = {c }, where c e C, then [F] ("\A =


[F ("\A] for both c e A and c E A. Suppose next that F = {c,c'}, where c,c' are
distinct elements of C. Then [F] ("\A ~ [F ("\A] if and only if c,c' are not both in
A and (c,c') ("\A '¢ 0. Consequently, if A is a face of C then (c,c') fl A '¢ 0
implies c,c' e A.
Suppose on the other hand that (c,c') fl A '¢ 0, where c,c' e C, implies
c,c' e A. We wish to show that [F] ("\A s [F ("\A] for any finite set F s C.
From what has already been said, this is true if IFI < 2. Hence it is sufficient to
show that if [F] fl A s [F ("\A] for some finite set F s C with IFI ~ 2, and if
x e C\F, then
[x u F] fl A s [ (x u F) fl A].

By Proposition 4,

[x u F] ("\A = Uye[FJ [x,y] fl A,


and [x,y] fl A s [{x,y} fl A]. H x E A, it follows that

[x u F] fl A s [F] ("\A s [F fl A] = [(x u F) fl A].

On the other hand, if x e A then


34 II. Convexity

[x u F] f"'\ A s {x} u Uye [FJnA [x,y]

Si {x} u Uye[FnAJ [x,y]

- [x u (Ff"'\ A)]

- [(x u F) f"'\ A]. a


In an arbitrary aligned space, and even in an aligned space with the anti-
exchange property, a nonempty subset of a Carath6odory set need not again be a
Carath6odory set. However, convex geometries do have this desirable property:

PROPOSITION 7 Any nonempty subset of a Caratheodory set is again a


Caratheodory set.

Proof Let S be a Caratheodory set. Since Sis finite, it is enough to show that S\s
is also a Carath6odory set for every s e S. Moreover, since singletons are
Carath6odory sets, we may suppose that ISi > 2.
Assume, on the contrary, that for some s0 e S,

[S \so] = UseS\so [S \ (s u so)].

Then, by Proposition 4,

[S] = U, e (S \so] [so,t] = Uses\ s0 U, e (S \ (s0us)] [so,t] ·

But [s0,t] s [S\s], since s0 e S\s and t e [S\s], and hence

[S] 5 UseS\so [S \s] 5 Uses [S\s].

Since S is a Carath6odory set, this is a contradiction. a


In an arbitrary convex geometry the classes of Carath6odory, Radon and Belly
sets need not coincide:

EXAMPLE 9 Take X =R 3 and for any x ={x1,x2,x3), y =(y1,Y2,y3) in X define


[x,y] to be the set of all z = (z1tz2,z3) in X with min{x;,y;} S z; S max{x;,Y;}
(i =1,2,3). It is easily verified that with this definition Xis a convex geometry.
The convex subsets of X are the boxes with sides parallel to the coordinate axes. H
we take x =(0,1,2), y ={l,2,0) and z =(2,0,1), then [x,y,z] is the cube consisting
of all points w = (w 1,w2,w3) with 0 S w; S 2 (i = 1,2,3). The set {x,y,z} is a
Carath6odory set, since (1/2,1/2,1/2) E [x,y] u [y,z] u [x,z], and is also a Radon
2. Convex geometries 35

set, but it is not a Belly set, since (1,1,1) e [x,y] r. [y,z] r. [x,z]. It is easily
verified also that the four points (0,0,1), (1,2,3), (3,1,2), (2,3,0) form a Radon
set, but not a Carath6odory set, and that the six points (+ 1,0,0), (0, + 1,0),
(0,0, + 1) form an independent set, but not a Radon, Belly or Caratheodory set.

For a given set S X and a given point x e S, we say that a point y e S is


5
visible from x if [x,y] 5 S. We denote by Sx the set of all points of S which are
visible from x and we define the kernel of S to be the set

Obviously a set Sis convex if and only if K(S) =S. A set Sis said to be star-
shaped if K(S) ~ 0. An example in R2 of a set which is star-shaped, but not
convex, is the middle set in Figure 2 at the beginning of the chapter.

PROPOSITION 8 For any set S X and any x e S, the kernel K(Sx) of the set
5
of all points of S visible from x is the intersection of all maximal convex
subsets of S which contain x. Furthermore,

Proof Since singletons are convex, there exists a maximal convex subset of S
which contains x. Let y e K(Sx) and let M be any maximal convex subset of S
which contains x. Then M 5 Sx and hence, since y e K(Sx),

[y U M1 = UzeM [y,z] 5 Sx.

Since Sx 5 S and M is a maximal convex subset of S which contains x, it follows


thaty e M.
Conversely, let y belong to all maximal convex subsets of S which contain x
and let z e Sx· Since [x,z] 5 S, there exists a maximal convex subset M of S which
contains [x,z]. Then ye Mand hence [y,z] 5 M. If we [y,z], then [x,w] 5 M
and hence we Sx. Thus [y,z] 5 Sx. Since this holds for every z e Sx, it follows
that y e K(S.J.
Since K(S.J s Sx, we certainly have flxes K(S.J 5 K(S). It remains to show
that if y e K(S), then y e K(S.J for every x e S. By the definition of kernel, y e Sx
for every x e S. We wish to show that [y,z] 5 Sx for every z e Sx.
36 II. Convexity

If u e fy,z] then u e S, since ye K(S) and z e S. If v e [x,u], then v e


[x,y,z] and hence v e [y,w] for some we [x,z]. Then we S, since z e Sx, and
hence v e S, since ye K(S). Thus [x,u] s Sand u e Sr a
COROLLARY 9 For any set S s: X, the kernel K(S) is the intersection of all
maximal convex subsets of S. In particular, K(S) is itself a convex subset of S.
a
The first statement of Corollary 9 was proved (for X a real vector space) by
Toranzos (1967); the second statement was proved (for X = JR2) by Brunn (1913).

3 ADDITIONAL AXIOMS

We now examine the consequences of imposing various additional axioms on


a convex geometry. These axioms may seem arbitrary at first sight, but their natural
role will become apparent in the next chapter. We first consider the axiom

(L2) if be [a,c], c e [b,d] and b "* c, then be [a,d].


PROPOSITION 10 If the convex geometry X satisfies (L2}, then the
alignment of convex sets is an anti-exchange alignment.

Proof We will show that if e is an extreme point of S, then it is also an extreme


point of [S]. Assume on the contrary that e e [x,y], where x,y e [S] \ e. By
Proposition 4 we have x e [e,u] and y e [e, v] for some u, v e [S \ e]. From
"*
e e [x,y], ye [e,v] and y e we obtain, by (L2), e e [x,v]. From e e [x,v],
x e [e,u] and x '¢ e we obtain similarly e e [u,v]. Since e E [S \ e], this is a
contradiction. a
Thus under the hypotheses of Proposition 10 all the results of Chapter I,
Section 3 are valid. The axiom (L2) is actually essential for the convex sets of a
convex geometry to form an anti-exchange alignment. Indeed if be [a,c],
c e [b,d], b '¢ c and b E [a,d], then [a,b,d] \b contains a and c, but not b.

PROPOSITION 11 If the convex geometry X satisfies (L2) then, for any


a,b e X, (a,b] and (a,b) are convex sets.
3. Additional axioms 37

Proof We may assume a :I: b, by (Ll). Then (a,b] and [a,b) are convex, by
Proposition 10 and Proposition 1.11. Hence (a,b) = (a,b] f"\ [a,b) is also convex.
a
Another property in which we are interested is the no branchpoint property:

(L3) if c e [a,b] and be [a,c], then [a,b] f"\ [a,c] = {a}.

The following proposition was proved by Hammer (1977):

PROPOSITION 12 If the convex geometry X satisfies (L3) then, for any finite
Radon set Sand any subsets S1,S2 of S,

Proof The result holds by the definition of a Radon set if S 1 and S 2 are disjoint.
Suppose now that S 1 f"\ S 2 = {s} is a singleton. We may clearly assume that
IS1I > 1, IS2I > 1, and then

[S1] = Uxe[S 1 \s] [s,x]' [S2] = Uye[S2 \s] [s,y].

Suppose z e [s,x] [s,y] for somex e [S 1\s] andy e [S2 \s]. Thenx E [S2] and
f"\

y E [Si], because Sis a Radon set. In particular, x E [s,y] and y E [s,x]. Hence,
by (L3), z = s. Thus [Si] f"\ [Si]= {s} and the result holds in this case.
Suppose next that S1 f"\ S2 contains m > 2 elements and assume that the result
holds whenever S 1 f"\ S2 contains less than m elements. Write S 1 n S2 = s u R,
wheres E R, and put Ti= Si\ (s u R) (j = 1,2). Then

[S 1l = Uxe [Ru Ti] [s,x] ' [S2] = Uye [Ru T2] [s,y].

By the induction hypothesis,

Since [Ru Ti] 5 [S1], it follows that

[Ru Ti] f"\ [Ru T2] = [R].


Hence
[S 1] f"\ ([Ru T2] \[Ru T1]) = 0,
[S2] n ([Ru Ti]\ [Ru T2]) = 0.
38 II. Convexity

Suppose ze [s,x]n [s,y], wherexe [Ru Ti] and ye [RuT2 ]. If


x e [Ru T 2] or if ye [Ru T 1], then z e [s u R] = [S 1 n S21· Hence we may
suppose x E [R u T 2] and y E [R u Ti]. Then x E [S 2] and y E [Si]. In
particular, x E [s,y] and y E [s,x]. Hence, by (L3), z =s e [S1 n S2]. D

PROPOSITION 13 If the convex geometry X satisfies (L3), then the classes of


Reily sets and Radon sets coincide.

Proof By Proposition 1.9 we need only show that every Radon set is a Belly set.
Moreover we may restrict attention to finite sets, since an infinite set is a Radon
(Belly) set if and only if every finite subset is a Radon (Belly) set. But if Sis a
finite Radon set then, by Proposition 12,

nses [S\s] = [0] = 0. D

We are also interested in the additivity property:

(L4) if c e [a,b], then [a,b] = [a,c] u [c,b].

With its aid we can prove

PROPOSITION 14 If the convex geometry X satisfies (L4) then, for any set
S s X and any c e [S],
[S] = Uses [c U (S\s)].

Proof It is sufficient to prove the result for finite sets S, by (84). Suppose
S = {s 1, ••• ,snl· The result is trivial for n = 1 and it holds for n = 2, by (L4). We
use induction and assume that the result holds for all finite sets containing at most n
elements, where n > 2. Let T =s0 u Sand supposed e [T]. Then, by Proposition
4, de [s 0 ,c] for some c e [S]. Hence, by Proposition 4 and the induction
hypothesis,
[T] = Uze[SJ [so,z] = Ui=t [cu (T\s;)].
But, by Proposition 4, [c u (T \ s;)] is the union of all segments [x,y], with
x e [s 0 ,c] and ye [S \s;]. Since [s 0 ,c] = [s 0 ,d] u [d,c], it follows from
Proposition 4 again that

[cu (T\s;)] = [du (T\s;)] u [cud u (S\s;)] (i = l, ... ,n).


Hence
3. Additional axioms 39

[T] = U7=1 [du (T\s;)] u U7=1[cud u (S\s;)]

= Ui::1 [du (T\s;)] u [du S].

Thus the result holds also for all finite sets containing n+ 1 elements. D

PROPOSITION 15 If the convex geometry X satisfies (L4), then any


Caratheodory set is also a Helly set.

Proof If Sis finite and there exists a point x e nseS [S\s] then, by Proposition
14,
[S] = Uses [x U (S\s)] = Uses [S\s]. D
PROPOSITION 16 Suppose the convex geometry X satisfies both (L2) and
(L4). Let S = {s 1, .•• ,snl be a finite set and let d,e be distinct elements of [S].
Then [d,e] s [p,q], where p,q e U7= 1 [S\s;].
Proof By Proposition 14,

[S] = Ui::: 1 [du(S\s;)] = U7=1 [eu (S\s;)].


Thus e e [du (S\sj)] and de [e u (S\sk)] for somej,k e { l, ... ,n}. It follows
from Proposition 4 that e e [d,p] for some p e [S \sj] and de [e,q] for some
q e [S\sk]. Hence d,e e [p,q], by (L2), and [d,e] s [p,q], by Proposition 1. D

Finally we introduce one more axiom:

The axiom's label is chosen in honour of Peano (1889) who, developing the
work of Pasch, first used the axioms (C) and (P). Convex geometries satisfying
the axiom (P) possess a number of additional properties. Our first result is a
counterpart to Proposition 2:

PROPOSITION 17 Suppose the convex geometry X satisfies the axiom (P). If


c 1e [a,b1], c2 e [a,b2 ] and be [b1,b2], then there is a point c e [a,b] n [c 1,c2].

Proof By (P), there exist a point c' e [a,b] n [c 1 ,b 2 ] and a point


c e [a,c1 n [c 1,ci]. Since [a,c1 s [a,b], the result follows. D
40 II. Convexity

PROPOSITION 18 Suppose the convex geometry X satisfies the axiom (P).


Then the foil owing properties hold:

(i) if c e [a,b] and de [a,c], thence [b,d];


(ii) if c e [a,b] and be [a,c], then b = c;
(iii) ifce [a,b],then [a,c]rt[b,c]={c}.

Proof (i) There exists a point e e [b,d] n [c,c], by (P), and e = c, by (Ll).
(ii) Take d = b in (i).
(iii) If de [a,c] rt [b,c], thence [b,d], by (i), and hence d = c, by (ii). []

The preceding two results actually hold in any aligned space satisfying the
axiom (P). The following sand-glass property does require that the aligned space
be derived from a convex geometry - draw a picture in R.2 to see the reason for the
name!

PROPOSITION 19 Suppose the convex geometry X satisfies (P). If


x e [a,a1 n [b,b1 and ye [a,b], then x e [y,y1for some y' e [a',bl

Proof Since x e [a,a1 and ye [a,b] there exists a point z e [a',y] rt [b,x], by
(P). From x e [b,b1 and z e [b,x] we obtain, by Proposition 18(i), x e [b',z] s
[y,a',b1. Hence x e [y,y1 for some y' e [a',b1. D

The sand-glass property in tum implies the following general statement:

PROPOSITION 20 Suppose the convex geometry X satisfies (P). If c e [a,b]


then, for any set S s X,

[au cu S] rt [bu cu SJ = [cu SJ.

Proof Obviously the right side is contained in the left. Suppose, on the other
hand, that xis any element of the left side. Then x e [d,y] for some de [a,c] and
y e [S], and x e [e,z] for some e e [b,c] and z e [S]. By Proposition 18(i),
c e [b,d] and also c e [d,e]. Consequently, by Proposition 19, x e [c,w] for some
we [y,z] Si [S]. []

As an application we prove

PROPOSITION 21 Suppose the convex geometry X satisfies (P). If


p e [a 1, ••• ,an1 and b; e [p,a;] (i = l, ... ,n), then p e [b 1, ••• ,bn1·
3. Additional axioms 41

Proof Assume p e [b 1, ••• ,bk-1'ak,... ,an1 for some k. Then, by Proposition 20,

Since the assumption holds for k =1, the result follows by induction on k. a
A basic property of convex geometries with the property (P) is the following
separation theorem:

PROPOSITION 22 Suppose the convex geometry X satisfies (P). If C and D


are disjoint convex subsets of X, then there exist disjoint convex sets C' and D'
with C' u D' = X such that C s C', D s D'.

Proof Let ~ be the family of all convex sets C" which contain C but are disjoint
from D. Then ~ is nonempty, since it contains C. If we partially order ?:F by
inclusion then, by Hausdorffs maximality theorem,?; contains a maximal totally
ordered subfamily ?;0 • The union C' of all the sets in ?;0 is again a convex set
containing C but disjoint from D.
Since C' is maximal, for every x E C' we have

We will show that, for every x E C',

C' r'\ [x u D] = 0.

Assume on the contrary that, for some x E C ', there exists a point
c' e C' r'\ [x u D] and let d" e [x u C'] r'\ D . By Proposition 4 we have
c' e [x,d1 for some d' e D and d" e [x,c"] for some c" e C'. Hence, by (P),

[c",c1 ('\ [d',d"] * 0.


Since [c",c1 s C' and [d',d"] s D, this is a contradiction.
Consider now the family C§ of all convex sets D" which contain D but are
disjoint from C'. Then C§ is nonempty and contains a maximal totally ordered
subfamily <§0• The union D' of all the sets in <§0 is again a convex set containing D
but disjoint from C'. In the same way, for every y E D' we have

and
[y u D1 r'\ C' *0
D' r'\ [y u C1 = 0.
42 II. Convexity

Since C' is maximal, it follows that y e C'. That is, C' is the complement of D '.
a
The axiom (P) is essential for the validity of Proposition 22. For suppose
z1 e [x,yi], z2 e [x,y 2] and [y 1,z2] n [y 2,zi] = 0. If C' and D' are convex sets
such that [y 1,z2 ] s C' and [y2 ,zi] s D', then x e C' implies z 1 e C' n D' and
x e D' implies z2 e C' n D'.
A set H s X is said to be a hemispace if both H and X \ H are convex.
Clearly X itself is a hemispace, and the complement of a hemispace is again a
hemispace. Sometimes another terminology is more convenient: two sets C,D are
said to be a convex partition of X if they are convex, nonempty and

CuD=X, CnD=0.

From Proposition 22 we can obtain a separation theorem for any finite number
of sets. The proof will be based on the following preliminary result:

LEMMA 23 Suppose the convex geometry X satisfies (P). Let A = {a 1, ••• ,an}
and B = {b 1, ••• ,bnl be subsets of X such that b; e [p,a;] (i = l, ... ,n)for some
p e X. If A;= A\ a; and B; = B \ b;, then ni:: 1[A; u q] ¢ 0, ni::1[B; u a;] ;t 0.

Proof For n = 1 the result is trivial. For n = 2 both formulae say that
[a 1,b2] n [a 2,bi] ¢ 0, and thus the result holds by (P). Suppose now that the
result holds for sets of n > 2 points, and that the hypotheses are satisfied by
A,= Au an+l and B' =Bu bn+l· If x E ni:1CA; u q] then

Since A 'n+l =A, this establishes the first formula unless [x,an+il n [A u bn+tl =
0. In the latter case there exists, by Proposition 22, a convex partition C,D of X
such that [Au bn+il s C and [x,an+d s D. Since p e D implies bn+l e D, we
must have p e C. Then bn e C and x e [An u bnl s C, which is a contradiction.
The second formula is proved similarly by taking a pointy E nl!:1CB; u a;]
and showing that [y,bn+il n [Bu an+il = 0 leads to a contradiction. a

PROPOSITION 24 Suppose the convex geometry X satisfies (P). If C1, ••. ,Cn
are convex subsets of X such that ni:: 1C; = 0, then there exist hemispaces
H 1, ••• ,Hn such that C; s H; (i = 1,... ,n), ni::1H; = 0 and LJi=1H; = X.
3. Additional axioms 43

Proof Let H 1, ••• ,Hn be maximal convex sets such that C; s H; (i = 1,... ,n), and
n7=1H; = 0. Thus any convex set properly containing Hi intersects nii!jH;
(j = l, ... ,n). Since there exists a hemispace containing Hi and disjoint from
n;i! j H; , by Proposition 22, Hi must itself be a hemispace. Assume there exists a
point p e X\ Uj= 1Hi. Then [Hi up] intersects nii!jH; in some point bi, and
bi e fp,aj] for some aj e Hi. Applying Lemma 23 to the sets A= {a 1, ••• ,an} and
B = {b 1, ••. ,bn}, we see that the sets [(B \bi) u aj] (j = 1,... ,n) have a common
point. Since (B \bi) u ai s Hi, this contradicts n}=t Hi = 0. D

Finally we consider some ways of manufacturing new convex geometries from


.
given ones.

(i) Restriction: Let X be a convex geometry and Y a nonempty convex subset of


X. Since x,y e Y implies [x,y] s Y, by restriction to Y we again obtain a convex
geometry. Moreover this convex geometry on Y satisfies those of the axioms
(L2),(L3),(L4),(P) which are satisfied by the given convex geometry on X.

(ii) Join: LetX =X 1 u X 2, where X 1,X 2 are disjoint sets on each of which a
convex geometry is defined. For a,b e X we define [a,b] = {a,b} if a e X 1,
be X2 or if a e X2, be X1, and [a,b] = [a,bh if a,b e Xk (k = 1,2). It is easily
verified that this defines a convex geometry on X. Indeed the axiom (C) holds
trivially if c e {a1,bi} or if de { c,b2 } and otherwise holds in X because it holds in
X 1 and X2 • Moreover this convex geometry on X satisfies those of the axioms
(L2),(L3),(L4),(P) which are satisfied by both the given convex geometries on
x l,X2.

4 EXAMPLES (continued)

It will now be shown that a real vector space, or indeed a vector space over
any ordered division ring, is a convex geometry with the usual definition of
segments. Since (Ll) is obviously satisfied, we need only prove (C). The
geometric significance of the axiom (C) in this case is illustrated in Figure 3. We
have d = 0c + (1 - 0)b 2 and c = A.a + (1 - A.)b 1, where 0 < 0,A. <l. Evidently
44 II. Convexity

Figure 3

we may assume that 0 < 0,A. < 1. Hence if we put v = 1 - 0A., then 0 < v < 1.
Moreover d = (1 - v)a + vb, where

We are going to show that segments in a vector space over an ordered division
ring also satisfy the axioms (L2 )-(L4) and (P). It is easy to verify (L2); for
suppose
b = A.c + (1 - A.)a, c = µb + (1 - µ)d,

where 0 S A.,µ < 1. Substituting the second equation in the first, we obtain
b= ea + (1 - 0)d, where e = (1 - A.µ)-1(1 - A.) and thus O < e < 1. Thus a vector
space over an ordered division ring is actually an aligned space with the anti-
exchange property.
We can now give a concrete illustration of the results of Chapter I, which was
first proved by Doignon (1973):

PROPOSITION 25 Let zd denote the set of poillts in R.d whose coordinates


are all integers. Ifs is any finite subset of zd with ISi > 2d, then nseS [S \s]
contains a point of Zd.

Proof Put X = [S] (') zd. Then [X] = [S], since S s X s [S], and hence
X = [X] (') Zd. As we saw in Chapter I,~ becomes an aligned space with the anti-
cxchange property if we define a 'convex' set in X to be the intersection with X of a
convex set in Rd. We wish to show that S is not a Belly set in this aligned space.
Assume that Sis a Belly set Then, by Proposition 1.15, there exists a Helly
set T s X with ITI = ISi and T = [T] (') X. Since ITI > 2d, there exist points
4. Examples (continued) 45

y,z e T whose corresponding coordinates are all congruent (mod 2). Then
x =(y + z)/2 E zd. But x E [T] Si [X] = [S]. Hence x E x, and so x E T. Thus T
is not independent, which contradicts Proposition 1.9. a
Since the set S of all points in zd whose coordinates are either 0 or 1 is an
independent set such that S =[SJ fl zd, it follows from the remark after the proof of
Proposition 1.15 that the bound 2d in Proposition 25 is actually sharp.
It is obvious that the axioms (L3) and (L4) are satisfied in a vector space over
an ordered division ring with the usual definition of segments. Finally it will be
shown that the axiom (P) is also satisfied. The geometric significance of the axiom
(P) in this case is illustrated in Figure 4.
Suppose

for some A.,µ with 0 S A.,µ ~ 1. If A. = µ = 1 then (P) obviously holds, since
c 1 =b 1 and c2 =b2. Otherwise we have A.µ< 1 and µA.< 1. Thus if we put

a = (1 - A.)(l - µA.)-1, p= (1 - µ)(l _ A.µ)-1,

then 0 S a,p < 1. Moreover, since (1 - A.µ)-1 = A.(1 - µA.)-lA,-1, we have

Figure4

We leave it as an exercise for the interested reader to verify that Examples 4-8
at the beginning of this chapter are all convex geometries. Furthermore, Examples
4-7 satisfy the axiom (L2). The axioms (L3),(L4) and (P) are in general not
satisfied in Example 4, but they are all satisfied if the set X is totally ordered. The
46 II. Convexity

axioms (L4) and (P) are satisfied in Examples 5--6, but the axiom (L3) is in
general not satisfied. The axiom (P) is satisfied in Examples 7-8. However, the
axioms (L3)-(L4) are not satisfied in Example 7, and the axioms (L2)-(L4) are
not satisfied in Example 8.
Finally we show that the axioms studied in this chapter are mutually
independent.

EXAMPLE 10 We exhibit systems which satisfy all the axioms (C),(Ll)-


(L4),(P) with the exception of the one named:

(C) Take X = {a,b 1,b2 ,c,d} with [a,bi] = {a,b 1,c}, [c,b2] = {c,b2,d}, and
[x,y] = {x,y} otherwise. The collection of all sets C Si X such that c e C if
a,b 1 e C, and de C if c,b 2 e C, is an alignment on X and, furthermore, it has the
anti-exchange property. Proposition 7 no longer holds, since {a,b 1,b 2 } is a
Caratheodory set but {b 1,b2 } is not.

(Ll) Take any setX such that IXI > 1 with [x,y] =X for allx,y e X.

(L2) Take X = {a~b,c,d} with [a,c] = {a,b,c}, [b,d] = {b,c,d} and [x,y] = {x,y}
otherwise.
An example which is 'continuous' rather than 'discrete' may also be given.
Let X = {z = ei& e C: 0 ~ 0 < x/3 or 2x/3 < 0 < x or 4x/3 ~ 0 < Sx/3}. No two
points of X are antipodal and, for all x,y e X, we define [x,y] to be the set of all
points of X on the shorter arc of the unit circle with endpoints x,y.

(L3) Take X = {x0,x1,x2,x3 } to be the tree consisting of three branches joining x0


to x 1,x2,x3, with segments defined as in Example 6. Proposition 13 no longer
holds, since S = {x1,x2,x3 } is a Radon set but not a Belly set.
An example which is 'continuous' rather than 'discrete' may again be given.
Let X = {x = (~ 1 ,~ 2) e R2: ~2 = 0 or ~ 1 =0, ~ 2 > O} and for all x,y e X define
[x,y] to be the ordinary segment in R2 with endpointsx,y if this segment is entirely
contained in X, and otherwise to be the union of the ordinary segments [0,x] and
[0,y], where 0 =(0,0).

(IA) Take X = {a,b,c,d} with [a,b] = X and [x,y] ={x,y} otherwise. Proposition
14 no longer holds, since [a,b] :1: [a,c] u [b,c].
4. Examples (continued) 47

(P) Take X = {a,a',b,b',c} with [a,c] = {a,a',c }, [b,c] = {b,b',c }, and [x,y] =
{x,y} otherwise.

S NOTES

Examples 4-6 are studied as anti-exchange alignments, rather than as convex


geometries, by Jamison-Waldner (1982). Examples 7 and 8 appear in Prenowitz
(1961) and van de Vel (1993) respectively.
The earliest paper on axiomatic convexity is perhaps Levi (1951), who used
the conclusion of Proposition 14 as one of his axioms and also proved Proposition
1.9. Another early paper is Ellis (1952), where Proposition 22 is proved under
equivalent hypotheses to those used here. Our approach is closer to that of
Prenowitz (1961) and Calder (1971).
The books of Soltan (1984) and van de Vel (1993) are devoted to the axiomatic
theory of convexity and have extensive bibliographies. The proof of Proposition 24
is taken from van de Vel. In these works there is considerable freedom in the
choice of axioms. Our choice of axioms is directed by the wish to obtain ultimately
a set of axioms for Euclidean geometry.
m
Linearity

In this chapter we show how lines may be defined, and how the points of a line
may be totally ordered, using four of the axioms considered in Chapter II. The
notion of a linear geometry, which is defined by all six of the axioms of Chapter
II, is then introduced. A linear geometry possesses both an anti-exchange
alignment of convex sets and an exchange alignment of affine sets. This rich
structure makes it possible to establish many familiar geometrical properties,
including the existence of half-spaces associated with a hyperplane and the
property which Pasch showed to have been omitted by Euclid. Nevertheless a
linear geometry may in fact possess only finitely many points. However, it is
shown that all irreducible linear geometries which are not dense, in the sense of
Chapter V, have a rather simple form.
In conclusion we establish the theorems of Belly, Radon and Caratheodory in
any linear geometry. The extension to an arbitrary linear geometry of the theorems
of Bminy and Tverberg, which we prove for R.d, is left as an open problem.

1 LINES

Let X be a set and suppose that with any unordered pair {a,b} of elements of X
there is associated a subset [a,b] of X containing a and b. The elements of X will be
called points and the subsets [a,b] segments. Throughout this section we will
assume that the following four axioms are satisfied:

(Ll) [a,a] ={a};


(L2) if be[a,c], c e [b,d] and b ~ c, then be [a,d];
(L3) if c e
[a,b] and be [a,c], then [a,b] n [a,c] ={a};
(L4) if c e [a,b], then [a,b] = [a,c] u [c,b].
so Ill. Linearity

We propose to study the consequences of this assumption.

PROPOSmON 1 Fora/I a,b,c,de X,

(i) if c e [a,b] and de [a,c], thence [b,d];


(ii) if c e [a,b] and be [a,c], then b = c;
(iii) if c e [a,b], then [a,c] n [b,c] ={c}.

Proof (ii) Assume b -:I: c. Then be [a,a], by (L2), and hence b =a, by (Ll).
Thus c e [a,a] and hence c =a. Then b =c, which is a contradiction.
(iii) Supposed e [a,c] n [b,c]. Then de [a,b], by (L4). By (L4) also,
either c e [a,d] or c e [b,d]. In either event d = c, by (ii).
(i) Obviously we may assumed -:I: a,c. Then ct! [a,d], by (ii). By (L4),
de [a,b]. Since ct! [a,d], it follows from (L4) again that c e [b,d]. D

We recall that Proposition 1 was already proved, under different hypotheses,


in Proposition Il.18.
If a and b are distinct points, we define the line <a,b> to be the set of all
points c such that either c e [a,b] or a e [b,c] orb e [c,a]. If a = b we set
<a,b> ={a}.
It is easily verified that if Xis a vector space over any ordered division ring D,
with the standard definition of segments [a,b], then the line <a,b> is the set of all
c e X which can be represented in the fonn c = A.a + ( 1 - A.)b, where A. e D.
Clearly <a,b> = <6,a> and [a,b] s <a,b>. In particular, a,b e <a,b>.
Furthermore, if a,b,c are distinct points such that c e <a,b>, then also a e <b,c>
and be <c,a>.

PROPOSffiON 2 If c,d e <a,b> and c -:1: d, then <c,d> = <a,b>.

Proof Clearly we must have a -:1: b. It is sufficient to show that if c -:I: a,b then
<a,c> = <a,b>. In fact, by symmetry we need only show that <a,c> s <a,b>.
The following proof is arranged so as to appeal to Proposition 1(i) rather than
(L4), as far as possible.
Let x e <a,c>, so that either x e [a,c] or a e [c,.x] or c e [a,.x]. We wish to
show that either x e [a,b] or a e [b,.x] orb e [a,.x]. Evidently we may assume that
x -:I: a,b,c.
1. Lines 51

Suppose first that c e [a,b]. If a e [c,x] then a e [b,x], by (L2). If c e [a,x]


then x e [a,b] orb e [a,x], by (L3). If x e [a,c] then c e [b,x], by Proposition
l(i), and hence x e [a,b], by (L2).
Suppose next that a e [b,c]. If x e [a,c] then a e [b,x], by Proposition l(i).
If a e [c,x] then be [c,x] or x e [b,c], by (L3). Moreover, by Proposition l(i),
be [c,x] implies be [a,x] and x e [b,c] implies x e [a,b]. If c e [a,x] then
a e [b,x], by (L2).
Suppose finally that be [a,c]. If a e [c,x] then a e [b,x], by Proposition l(i).
If c e [a,x] then c e [b,x], by Proposition l(i), and hence be [a,x], by (L2). If
x e [a,c] then x e [a,b] or x e [b,c], by (L4). Moreover, by Proposition l(i),
x e [b,c] implies be [a,x]. D

COROLLARY 3 If there exist three disti11ct poi11ts a,b,c such that cf! <a,b>,
the11 for a11y disti11ct points x,y there exists a point z such that z f! <x,y>. D

Points which lie on the same line will be said to be co/li11ear. We show next
how the points of each line may be totally ordered:

PROPOSITION 4 Given two distinct points x,y e X, the points of the line
e = <x,y> may be totally ordered so that x < y and so that, for any points
a,b E .f with a < b, the segment (a,b) Consists Of al/ CE .f SUCh that a < C < b.
Moreover, this total ordering is unique.

Proof For any a,b e e, we write a < b if either of the following conditions is
satisfied:

(i) a e [x,b] and either ye [x,b] orb e [x,y],


(ii) x e [a,y] and either be [a,y] or ye [a,b].

It should be noted that (i) and (ii) say the same thing if a = x and that they cannot
both hold if a ~ x.
The definition obviously impliesx < y. Also, from the definition of a line it is
clear that for any a E .f we have a < a. Let a,b be points of .f such that both a < b
and b <a. We wish to show that this implies b =a.
Assume first that a e [x,b] and either ye [x,b] orb e [x,y]. Since x ~ y, this
implies x f! [b,y]. Since b <a, it follows that b e [x,a]. Hence b =a, by
Proposition l(ii). The case b e [x,a] and either ye [x,a] or a e [x,y] may be
discussed similarly. Suppose finally that x e [a,y] and x e [b,y]. Then either
52 III. Linearity

a e [b,y] orb e [a,y], by (L3). We cannot have also ye [a,b], since this would
imply y =a or y =b, respectively, and hence x = y. Consequently we must have
be [a,y] and a e [b,y], which implies b =a.
Suppose next that a,b,c are distinct points of esuch that a < b and b < c. We
wish to show that this implies a< c.
Assume first that be [x,c] and either ye [x,c] or c e [x,y]. If a e [x,b] then
a e [x,c] and hence a< c. If x e [a,y] then ye [x,c] implies ye [a,c], by (L2),
and c e [x,y] implies c e [a,y], by (L4). In both cases a< c. Assume next that
x e [b,y] and either c e [b,y] or ye [b,c]. If x e [a,y] and be [a,y] thence [a,y]
or ye [a,c], respectively. Thus again a< c. All remaining cases may be discussed
similarly.
e
It remains to show that if a,b are distinct points of then either a< b orb< a.
Assume that neither relation holds. Then if a e [x,b] we must have x e [b,y], and
if x e [b,y] we must have be [a,y]. Hence be [x,y], by (L2), and b = x, by
Proposition l(ii). But this implies b =a, which contradicts our assumption.
Therefore a E [x,b]. Similarly we can show that b E [a,x]. Consequently
x e [a,b].
If x e [a,y] we must have a e [b,y] and hence x e [b,y], and if x e [b,y] we
must have b e [a,y]. Therefore x E [a,y], and similarly x E [b,y]. Thus either
a e [x,y] or y e [a,x]. But a e [x,y] would imply x e [b,y], by (L2), and
ye [a,x] would imply x e [b,y], by Proposition l(i). Thus we cannot escape a
contradiction.
This proves that the relation< is a total ordering of the line t such that x s y.
Suppose now that a< band c e [a,b]. If (i) holds, thence [x,b] and hence c < b.
Moreover a e [x,c] and either ye [x,c] or c e [x,y], so that a< c. Similarly it may
be seen that a < c < b if (ii) holds.
Suppose on the other hand that c e t and a< c < b. We will show that
c e [a,b]. Indeed, if a e [c,b] then, by what we have just proved, c <a Sb and
hence c =a. Similarly if be [a,c], then c = b.
Finally, suppose that < is any total ordering of t with the properties in the
e
statement of the proposition. Let a,b be any distinct elements of and assume first
that (i) holds. If b < x, then b <a< x and x Sb< y, since y E [x,b]. Hence b =a,
which is a contradiction. Consequently x < b and x < a < b. Assume next that (ii)
holds. Then a <x < y and either a< b < y or a< y < b, since y :I: a. In any event
a< b, and so the total ordering is the one originally defined. D
1. Lines 53

To illustrate the application of Proposition 4 we prove

PROPOSITION 5 The union of two segments with more than one common
point is again a segment.

Proof Let [a,b] and [c,d] be two segments whose intersection contains the distinct
points x,y. Assume the line e =<x,y> totally ordered, as in Proposition 4, and
choose the notation so that a< b, c < d and a< c. Then y < b, since ye [a,b], and
c <x, since x e [c,d]. If d < b, then [c,d] s: [a,b]. On the other hand if b < d,
then a < c < x < y < b < d and hence

[a,b] u [c,d] = [a,c] u [c,d] = [a,d]. a


The following result shows that if all points of X are collinear, then the axioms
(C) and (P) of Chapter II both hold, and actually in a stronger form.

PROPOSITION 6 Suppose a,b 1,b 2 are collinear points and c 1 e [a,bi],


c 2 e [a,b2]. Then either c1 e [b 1,c2] or c2 e [b 2,ci]. Moreover, if de [c 1,c 2]
then either de [a,b1] or de [a,b2].

Proof Obviously we may assume c 1 ~ c2. If c2 e [a,ci] then c 1 e [b 1,c 2], by


Proposition l(i). Similarly if c 1 e [a,c 2] then c2 e [b 2 ,ci]. Hence we may
suppose a e (c 1,c2). But then both c 1 e [b1,c2] and c2 e [b2,c 1], by (L2).
If a e [b 1,b2], then c 1,c2 e [b 1,b2] and hence de [b 1,b2]. Thus de [a,bi] or
de [a,b 2], by (L4). If b 1 e [a,b2 ], then c 1 e [a,b 2] and hence de [a,b 2].
Similarly if b2 E [a,bi), then C2 E [a,b1] and hence d E [a,bi). C

The concept of collinearity makes it possible to obtain sharper forms (under


stronger hypotheses) for a number of results in Chapter II. These sharper forms are
based on the following simple results:

(a) Let a,b 1,b2 be non-collinear poi11ts and suppose c1 e (a,b 1), c2 e (a,b 2). If
de [b 1,c2] f"\ [b2,ci], then actually de (b 1,c2) f"\ (b2,c 1).

Proof It is sufficient to show that d ~


b 1,c 1. By Proposition 2 d = b 1 implies
b2 e <b 1,c 1> = <a,b 1>, which is a contradiction. Similarly d = c 1 implies
C2 E <a,b1> and again b2 E <a b1>. C1

Similarly we can prove


54 Ill. Linearity

(b) Let a,b 1,b2 be non-collinear points and suppose ci e (a,b 1}, c2 e (a,b 2],
c e (c 1,c 2). If c e [a,b]for some be [b 1,b 2], then actually c e (a,b) and
b e (b 1,b2). D

(c) Let a,b 1,b2 be non-collinear points and suppose c 1 e (a,b 1), c 2 e (a,b 2],
be (b 1 ,b 2 ). If there exists a point c e [a,b] r.. [c 1 ,c 2 ], then actually
c e (a,b) r.. (c 1,c2). D

(d) Let a,a',b,b' be non-collinear points and let x e (a,a') r.. (b,b'). If x e [y,y1
for some ye (a,b) and y' e [a',b1, the11 actually x e (y,y') and y' e (a',b'). D

By combining (a) with (P), and (b),(c),(d) with Propositions 11.2,17,19


respectively, we obtain, for co11vex geometries satisfyi11g the additional axioms
(L2)-(L4) and (P):

(P)' if a,bi ,b2 are non-collinear points and if c1 e (a,bi), c 2 e (a,b 2), the11
there exists a point d e (bi ,c2) r.. (b2,c1). D

PROPOSITION 11.2' Let a,bi ,b2 be 11on-collinear points and let c 1 e (a,b 1),
c2 e (a,b 2]. If c e (c 1,c2), then c e (a,b) for some be (b1,b2). D

PROPOSITION 11.17' Let a.bi ,b2 be non-collinear points and let c1 e (a,b1),
c2 e (a,b 2]. If be (b 1,b2), then there exists a point c e (a,b) r.. (c 1,c2). D

PROPOSITION 11.19' Let a,a',b,b' be non-colli11ear points. If ye (a,b) and


x e (a,a') r.. (b,b'), then x e (y,y'}for some y' e (a',b'). D

2 LINEAR GEOMETRIES

We define a linear geometry to be a convex geometry satisfying the additional


axioms (L2)-(L4) and (P). An example of a linear geometry with only finitely
many points is the setX = {a,a',b,b',c} with [a,a1 = {a,a',c}, [b,b1 = {b,b',c} and
[x,y] = {x,y} otherwise. In spite of such examples we are going to show that linear
geometries possess many of the properties of real vector spaces.
Throughout the remainder of this chapter we will assume that a set X is
given on which a linear geometry is defined. We define a subset A of X to be
2. Linear geometries 55

affine if x,y e A implies <x,y> s A. For example, the line <x,y> is itself affine,
by Proposition 2.
It follows at once from the definition that the collection SL of all affine sets is a
normed alignment on X. For any set S s X, we define its affine hull <S> to be the
intersection of all affine sets which contain S. Our notations are consistent, since
the affine hull of the set {x,y} is the line <x,y>. Evidently, for any set S s X, we
have S s [S] s <S>. The next result, inspired by Bennett and Birkhoff (1985), is
of fundamental importance:

PROPOSffiON 7 If C is a convex set, then

<C> = Ux,x'eC < x,x'>.

Proof It is sufficient to show that the set A = Ux,x'eC < x, x'> is affine, since it is
certainly contained in any affine set which contains C. Evidently we may assume
that C contains more than one point. Thus we wish to show that if b e <xi ,x2>,
c e <x3,x4 >, and a e <b,c>, where Xi , ••• ,x4 e C and Xi -:1: x 2, x 3 -:1: x4 , b -:1: c, then
there exist x 5 ,x6 e C with x 5 -:1: x 6 such that a e <x5 ,x6>. This is evident if
c e <xi,x2>, since then also a e <x(,x2>. Consequently we assume c e <xi,x2>,
and similarly b ~ <x3,X4>.
By symmetry, and the convexity of C, we need only consider the following
five cases:

(A) x 2 e [b,xi], x 4 e [c,x3], c e [b,a];


(B) x2 e [b,xi], X4 e [c,x3], a e [b,c];
(C) x2 e [b,x1], c e [x3,x4], a e [b,c];
(D) x2 e [b,xi], c e [x3,x4], c e [a,b];
(E) x2 e [b,xi], c e [x3,x4], be [a,c].

Consider first case (A). By (P) there exists a pointy e [a,x2] n [c,xi]. Since
c e <xi ,x2 >, we have y ~ x 2 and hence a e <x2,y>. Consequently we may
assume ye C. By (P) also there exists a point x 5 e [y,x 3 ] n [xi,x4]. Then
x5 e C and x5 e [a,x2,x3]. Hence x5 e [a,x6] for some x6 e [x2,x3]. Thus x6 e C
and a e <x5 ,x6> if x 5 -:1: x 6• We can now assume x5 = x 6 e [x2 ,x3]. If x 3 -:1: x 5
then x 2 e <x3,x5>, ye <x3,x5>, and hence a e <x3,x5>. Consequently we can
now assume x 3 = x 5 e [xi,x4]. Then, by (L2), x 4 e [c,xi]. Since ye C it
follows from (L4) that y e [c,x 4 ] and then from Proposition l(i) that
56 Ill. Linearity

x 4 e [x 1,y] s [a,x 1,x2]. Hence x 4 e [a,x7] for some x 1 e [x1,x2]. Then x 1 e C


and a e <x4 ,x1> if x 4 '¢ x 1• In fact we cannot have x 4 = X7, since this would imply
x 4 e [x1,x2], x3 e [x1,x2] and hence

Similarly in case (B) there exist points y e [b,x4 ] n [a,x3 ] and x 5 e


[y,xi] n [x2,x4]. The argument can now be completed as in the previous case.
In case (C) there exists a point x5 e [c,x2] n [a.xi]. Moreover x 5 e C, since
c e C, and x5 '¢ x 1, since c e <x 1,x2>. Hence a e <x 1,X5>.
Similarly in case (D) there exists a point x5 e [a.xi] n [c,xi] and the argument
can be completed as in the previous case.
Finally in case (E) we have c e C and x 2 e [a,c,xi]. Hence x 2 e [a,x5] for
some x 5 e [c,x 1]. Then x 5 e C, x 5 '¢ x 2 and a e <x2,x5>. D

The definitions and results of Chapter I, Section 2 may be applied to the


alignment of affine sets on X, as well as to the alignment of convex sets on X. To
avoid confusion, the qualification 'affine' will be made explicit. Thus we say that a
set S s X is affine independent if, for every x e S, x e <.S \x>. A subset T of a
set S is an affine gen~rator of S if <.T> = <.S> and an affine basis of S if, in
addition, T is affine independent.
According to the results of Chapter I, Section 2, an infinite set is affine
independent if every finite subset is affine independent, and any affine independent
set is contained in a maximal affine independent set. It will now be shown that the
alignment of affine sets is an exchange alignment:

PROPOSITION 8 For any set S s X, if ye <S u x> but ye <S>, then


x e <Su y>.

Proof Obviously we may assume that x e [Su y]. By Proposition 7,


ye <z 1,z2>, where z1,z2 e [Su x]. Therefore, by Proposition 11.4, z1 e [x,wi]
and z2 e [x,w2] for some w1,w2 e [S].
If ye [z 1 ,z 2 ], then ye [x,w 1 ,w 2 ] and hence ye [x,w] for some
we [w 1,w2] s [S]. Thus y '¢ w andx e <y,w> s <.Su y>.
By symmetry it only remains to consider the case z2 e (y,z1). We will assume
x e <y,w1,w2> and derive a contradiction.
2. Linear geometries 57

Since x E <y,w1,w2>, we must have z1 ~ x,w1• Thus z1 e (x,w 1) and hence,


by Proposition 11.2', z2 e (x,w) for some we (y,w 1). Hence we <x,z 2 > =
<x,w2>. Since x E <y,w 1,w2>, we must have w = w2• Since ye <w,w 1>, this is
a contradiction. D

It follows that all the results of Chapter I, Section 5 are valid for the alignment
of affine sets in a linear geometry. With the definition of dimension given there,
the empty set 0 has dimension -1, a point has dimension 0 and a line has
dimension 1. An affine set of dimension 2 will be called a plane. For finite-
dimensional X, an affine set H s X is a hyperplane if and only if
dim H =dim X - 1.
Pasch (1882) pointed out the incompleteness of Euclid's axioms for geometry
and introduced the following additional axiom: 'If a line in the plane of a triangle
does not pass through any of its vertices but intersects one of its sides, then it also
intersects another of its sides.' It will now be shown that Pasch 's axiom holds in
any linear geometry. (It is not difficult to show that, conversely, any convex
geometry is a linear geometry if it satisfies the axioms (L2 )-(L4) and Pasch 's
axiom.)

LEMMA 9 If a,b,c,d are points such that

[a,b] ("\ <c,d> =0, [a,c] ("\ <b,d> =0, [b,c] ("\ <a,d> = 0,

then d E <a,b,c>.

Proof The hypotheses evidently imply that a,b,c are not collinear and
d E [a,b] u [a,c] u [b,c]. We will assumed e <a,b,c> and derive a contradiction.
It follows from Proposition 11.16 that if de [a,b,c], then de [a,e] for some
e e [b,c], which contradicts [b,c] ("\ <a,d> = 0. Thus we now suppose
d E [a,b,c].
By Proposition 7, de <p,q>, where p,q e [a,b,c] and p ~ q. Moreover we
may choose the notation so that q e (p,d). Furthermore, by Proposition 11.16, we
may take p,q e [a,b] u [a,c] u [b,c].
If p,q e [a,b], then de <a,b> and either a e (b,d) orb e (a,d). In either case
<a,d> ("\ [b,c] ~ 0. It is obvious also that <a,d> ("\ [b,c] ~ 0 if q =a andp e [b,c]
orifp=aandqe [b,c].
58 Ill. Linearity

By symmetry it only remains to consider the case p e (a,b) and q e (b,c).


Theny e (b,r) for some re (a,dJ. Since q e (b,c) and r :I: c, either re (b,c) or
e
c (b,r). If re (b,c), then (a,d) ("\ (b,c) :I: 0. If c e (b,r), then c e (q,r) and hence
c e (d,e) for some e e (a,p) s (a,b). Thus <c,d> ("\ (a,b) :I: 0. D

PROPOSITION 10 Let a,b,c be non-collinear points and let .e be a line in the


plane <a,b,c> such that a,b,c E .e. If .e intersects (a,b), then .e also intersects
either (a,c) or (b,c), but not both.

Proof We show first that .e cannot intersect (a,b), (a,c) and (b,c). By symmetry it
is sufficient to show that if de (a,b), e e (a,c) and/ e (b,c), thenf E (d,e). But if
fe (d,e) then, by Proposition 11.2',/ e (a,g) for some g e (b,c). Hence a e <b,c>,
which is a contradiction.
It remains to show that .e intersects two 'sides' of the 'triangle' [a,b,c]. Since
a,b,c E .e, this follows at once from Proposition 11.16 if .e contains two distinct
points of [a,b,c]. Thus we now assume that .e contains a point e E [a,b,c]. Since
e e <a,b,c>, it follows from Lemma 9 that the points a,b,c may be named so that
there exists a point p e <a,e> ("\ [b,c]. Thus the hypothesis is now that .e contains a
point de (a,b) u (a,c) u (b,c).
Suppose first that de (a,b), which implies p :I: b. If a e (e,p), then de (eJ)
for some f e (b,p). If p e (a,e), then there exists a point/ e (d,e) ("\ (b,p). In both
cases/ e <d,e> ("\ (b,c).
Suppose next that de (b,c). We may assume that p :I: d and, without loss of
generality, that de (b,p). If a e (e,p), then there exists a point/ e (d,e) ("\ (a,b).
If p e (a,e), then de (el) for some/ e (a,b). In both cases/ e <d,e> ("\ (a,b).
The cased e (a,c) is reduced to the cased e (a,b) by interchanging band c.
D

Although Proposition 10 may seem rather special, it has some important


general consequences. The next lemma is proved by Lenz (1992) under stronger
hypotheses.

LEMMA 11 Let H be a hyperplane. If x 1,x2,x3,x4 are points o/X\H such that


2. Unear geometries 59

Proof Obviously we can assume X3 ~xi and x 4 ~ x 2 • The lines <xi,x 2>,
<x2 ,x3> and <x3,x4> intersect H in unique points h1th 2 and h 3. Moreover
hi E (Xi,X2), h2 E (X2,X3) and h3 E (X3,X4).
Suppose first that xi,x2,x3 are collinear. Then hi= h2 and either Xie (h 2 ,x3)
or x 3 e (h 2,xi). If Xi e (h 2,x 3) then, by (P), there exists a point h e
[h 2,h3] n (xi,x4). If x 3 e (h2,xi) then, by (C), there exists a point he [xi,x4]
such that h 3 e [h 2 ,h]. If h 3 = h2, then h2 ·e (x 1,x4). If h3 ~ h 2 , then h e
H n (X1,X4).
Thus we may assume that x 1,x2 ,x3 are not collinear, and hence h 1 ~ h 2 •
Suppose now that x 1 ,x 3,x4 are collinear, so that h 3 e <x 1 ,x 3>. Since
<h 1,h 2 ,h3> ~<x 1,x 2,x3>, we must have h3 e <h1th 2>. By Proposition 10,
h 3 ~ (x 1,x3). Since x 1,x3,x4 are collinear, it follows that h 3 e (x1,x4).
Thus we may assume that x 1,x3 ,x4 are not collinear and indeed, by similar
arguments, we may assume that no three of the points x 1,x2,x3,x4 are collinear.
Suppose next that <x2,x4> n H ~ 0. Then the line <x2 ,x4> intersects Hin a
unique point h. We show first that hi!: (x 2,x4 ). Assume on the contrary that
h e (x2,x4). Since x 2,x3,x4 are not collinear, it follows from Proposition 10 that
h ~ <h2,h3>. Hence <h2,h3,h> = <x2,x3,x4>, which is a contradiction.
Thus either x 4 e (x 2 ,h) or x 2 e (x4,h). If x 4 e (x 2 ,h) then, by (P)', there
exists a point h' e (h,h 1) n (x1,x4). If x2 e (x4,h) then, by Proposition 11.2', there
exists a point h' e (x1,x4) such that h 1 e (h,h').
Thus we may assume that <x 2 ,x4 > n H = 0, and similarly also that
<x 1,x3> n H = 0. By (P)', there exist a point x e (x2,h 3) n (x4,h 2) and a point
h e (x,x3 ) n (h 2,h3). Since <x2 ,x> n H ~ 0, it follows from the previous part of
the proof (with x4 replaced by x) that (x 1,x) n H ~ 0. Since <x,x4> n H ~ 0, it
further follows from the previous part of the proof (with x 2 replaced by x) that
(X1,X4) n H ~ 0. D
PROPOSITION 12 Let H be a hyperplane such that X \ H is not convex. Then
there exist unique nonempty convex sets H + and H _ such that
X\H = H+ u H_. Furthermore,

(i) H+ n H_ = 0,
(ii) ify e H+ and z e H_, then (y,z) n H ~ 0,
(iii) Hu H+ and Hu H_ are also convex.
60 Ill. Linearity

Proof Since X \ H is not convex, there exist points a,b e X \ H and c e H such
that c e (a,b). Let H+ denote the set of all points x e X such that (b,x) r'\ H 0, *
and let H _ denote the set of all points x e X such that (a,x) r'\ H 0. Then *
a e H +' b e H _ and H + r'\ H = 0, H_ r'\ H = 0.
We are going to show that also H + r'\ H _ = 0. Assume on the contrary that
there exists a point x e H + r'\ H _. If x e <a,b> then c e (a,x) r'\ (b,x), since
<a,b> r'\ H = {c}.
But this is impossible, since also c e (a,b). Thus a,b,x are not
collinear. If (a,x), (b,x) intersect H in h,h' respectively, then c E <h,h'> by
Proposition 10. Hence <c,h,h'> = <a,b,x>, which is a contradiction.
We show next that H_ is convex. Assume on the contrary that there exist
points x',x" e H _ and a point x e (x',x") such that x E H _. Then there exist points
h',h" e H such that h' e (a,x'), h" e (a,x"). Hence a,x',x" are not collinear and
h' * h". By Proposition 11.17' the segment (h',h") contains a point he (a,x). Then
he Hand x e H_, which is a contradiction. Thus H_ is convex, and similarly also
H+.
It follows at once from Lemma 11 that if y e H + and z e H _, then
*
(y,z) () H 0. It will now be shown that X =Hu H+ u H_. Since His a
hyperplane, it is sufficient to show that X': =Hu H + u H_ is affine. In fact it is
enough to show that if x' e H +' x" e X' and x e <X',x">, then x e X'. Since
x' e H+, the segment (b,x') contains a point h' e H.
Suppose first that x" = h" e H. If h" e (x,x') then x e H_, since x' e H +· If
x e (x',h") then, by (P), there exists a point h e [h',h"] () (b,x) and hence x e H +·
If x' e (x,h") and b,x,x' are collinear, then h' = h" e (b,x) and hence x e H +· If
x' e (x,h") and b,x,x' are not collinear then, by Proposition 11.2', there exists a
point he (b,x) such that h' e (h,h") and again x e H+. This proves that a line
containing a point of H and a point of H + is entirely contained in X', and similarly a
line containing a point of Hand a point of H_ is entirely contained in X'.
Suppose next that x" e H+ u H_ and the line <x',x"> contains no point of H.
Then x" e H+, since x" e H_ would imply (x',x") () H * 0. Hence the segment
(b,x") contains a point h" e H. If x e (x',x") then x e H+, since H+ is convex.
Now consider the case x' e (x,x"). By (P) ', there exists a point
ye (b,x') r'\ (h",x). Moreover y E H. If h' e (b,y), then the segment (b,x)
contains a point h such that h' e (h,h'') and hence x e H +· H h' e (x',y), then the
segment (x,x') contains a point h such that h' e (h,h'') and hence x e X' by what we
2. Linear geometries 61

have already proved. Since the same argument applies in the case x" e (x,x1, this
completes the proof that X = H u H+ u H_.
It now follows that Hu H +and Hu H_ are convex. For if (say) x 1 e H and
x 2 e H+, then (x 1,x2) rt H =0 and hence (x 1,x2) rt H_ =0.
Finally, suppose X\H is the union of two nonempty convex sets G+ and G_.
We may assume the notation chosen so that G+ rt H+ :1: 0. But then G+ s H+, by
(ii). Hence G _ n H _ :1: 0 and so, in the same way, G _ s H _. Since
G+ u G_ =H+ u H_, we must actually have G+ =H+ and G_ =H_. D

Proposition 12 says that a hyperplane has two 'sides', if its complement is not
convex. The convex sets H + and H _ in the statement of Proposition 12 will be
called the open half-spaces associated with the hyperplane H, and the convex sets
H+ u Hand H_ u H will be called the closed half-spaces associated with H.
(When X \ H is convex, we may call X \ H the open half-space and X the closed
half-space associated with the hyperplane H. This situation arises if, for example,
X is itself the closed half-space associated with a hyperplane H of a real vector
space.) Our topological terminology will be given more justification in Chapter V.
Lemma 9 can be generalized to 'higher dimensions in the following way:

PROPOSITION 13 Let H 1,H 2 ,H 3 be three distinct hyperplanes which


themselves have a common hyperplane Z. If xk e Hk \ Z (k = 1,2,3), and if

Proof Assume on the contrary that [x 1,x 2] rt H 3 = 0. Then x 1,x 2 ,x3 are not
collinear and hence Z :I: 0. Moreover, by Lemma 9,

Hence, by Proposition 8, for any z e Z we have

(*)

Since x 3 e X = it follows from Proposition 7 that x 3 e <y 1,y2>,


~ 1 ,x 2 ,Z>,

where Yj e [x1,x2 ,zj] for some zi e Z U= 1,2). Moreover(*) implies that z1 :I: z2,
Yj ~ [x 1,x 2] for j = 1,2, and x 3 ~ fy 1,y2]. Without loss of generality we may
suppose y 1 e (y2,x3).
62 111. Linearity

Put

Then dim X' = 3 and x 3 e X'. Furthermore Hk' = <xk,z 1,z 2 > (k = 1,2,3) are
distinct planes inX' with the common line Z' = <z 1,z2>. Moreover

Since y 2 e [x,z 2] for some x e [x 1,x 2] and y 2 t! [x 1,x2], we must have


[y2,z2l (') <x 1,x2> = 0. Since

it follows that [y2,z2] (') <x1,x2,z 1> = 0. Thus y 2 and z2 lie in the same open half-
space of X' associated with the plane <x1,x2,z1>. Since

it follows that x 3 and z2 lie in different open half-spaces of X' associated with the
plane <x1,x2,z 1>. Thus there exists a point

Similarly from [x 1,x3] (') H 2' = 0 we obtain

Consequently, by Lemma 9, z 1 t! <x 1,x2 ,u>. But u t! <x 1,x2>, since


z2 ~ <x 1,x2,x3 > and hence, by Proposition 8, z1 e <x1,x2 ,u>. Thus we have a
contradiction. D

The join of two convex geometries was defined in Chapter II and, in the
terminology of this chapter, it was shown there that the join of two linear
geometries is again a linear geometry. A linear geometry will be said to be
irreducible if it is not itself the join of two linear geometries. Since irreducible
2. Linear geometries 63

linear geometries are the building blocks of which all linear geometries are
composed, it is of interest to detennine their nature. To this task we now turn.

LEMMA 14 Let x,y 1,y2 be non-collinear points such that <x,y 1> :I:- {x,y 1} and
<x,y2> :I:- {x,y2}. If [y 1,y2] = {y 1,y 2 }, then there exist points z 1,z 2 such that
x e (y 1,z 1) () (y 2,z2) and the plane <x,y 1,y2> contains only the five points
x,yl ,y2,z1 ,z2.

Proof Let z" e <x,yk> with z" :I: x,yk (k = 1,2). Assume first that z2 e (x,y 2). If
y 1 e (x,z 1) then, by (P)', there exists a point we (y 1,y2) () (z 1,z2). If x e (y 1,z 1)
then, by Proposition 11.2', z2 e (y,z 1) for some ye (y 1,y2). If z 1 e (x,y 1) then, by
(P)', there exists a point u e (y 1,z2) n (y2 ,z 1) and, by Proposition 11.2', u e (x,v)
for some v e (y1,y2). Since (y 1,y2) =0 we conclude that (x,y2) =0, and likewise
(x,y 1) = 0.
Assume next that y 2 e (x,z 2). If x e (y 1,z 1), then y 2 e (zltu) for some
u e (y 1,z2) and there exists a point v e (x,u) n(y 1,y2). If y 1 e (x,z 1), then there
exist a point u' e (y1'z 2 ) () (y 2 ,z 1) and a point v' e (x,u) n (y 1,y 2 ). Since
(y 1,y2) = 0 we conclude that x e (y2,z2), and likewise x e (y 1,z 1).
Assume now that the line <x,y1> contains a point z '#: x ,y 1,z 1. Then in the
same way x e (y 1,z) and by interchanging z and z1, if necessary, we may suppose
that z e (x,z 1). But then z e (y 2,z') for some z' e (z 1 ,z 2) and hence, by
Proposition 11.19', x e (y',z') for some y' e (y 1 ,y 2 ). We conclude that
<x,y 1> = {x,y1,z 1}, and similarly <x,y2> = {x,y2,z2 }.
Assume next that the line <y 1,y2> contains a point z '#: y 1,y2• Without loss of
generality we may suppose y 2 e (y 1,z). Then x e (z,u) for some u e (y1,z2) and
there exists a point v e (u,y 2 ) n (x,y 1). Since (x,y 1) = 0, we conclude that
<y1,Y2> = {y1,Y2}.
Assume finally that the plane <x,y 1,y2> contains a point z '#: x,y1,Y2,z 1,z2.
Then

since (y 1,y2) =0 and z ~ <x,yy (k = 1,2). Furthermore

[x,yi] n <y2,z> = 0,

since (x,y1) = 0, z ~ <y 1,y2> and z ~ <x,y2>. Similarly

[x,y 2 ] () <y 1,z> =0.


64 Ill. Linearity

But this contradicts Lemma 9. D

COROLLARY 15 If not all points of the linear geometry X are collinear and if
every line contains at least three points then, for any distinct points a,b e X,
(a,b) -:1: 0. D

Irreducible linear geometries are completely characterized by the following


proposition:

PROPOSITION 16 If Xis an irreducible linear geometry, then exactly one of


the following alternatives holds:

(i) X contains only one point,


(ii) X contains at least three points and all points of X are collinear,
(iii) not all points of X are collinear and [a,b] -:!: {a,b} for any distinct points
a,b e X,
(iv) not all points of X are collinear and there exists z e X such that every line
which does not contt~in z has only two points, and every line which contains z
has exactly two other points with z in the segment determined by them.

Proof It is clear thatX is irreducible in each of the cases (i)-(iv), and that no two
of these cases can coexist. To prove the proposition we will assume that none of
(i)-(iv) holds and derive a contradiction.
Since X is irreducible, our assumptions imply that X contains at least three
points and not all points are collinear. If A is a nonempty proper affine subset of X
then, for some a e A and some x e X \ A, the line <a,x> contains more than two
points (since otherwise X \ A would be affine and X would be the join of A and
X\A).
In particular some line in X contains at least three points. It follows from
Hausdorffs maxima!ity theorem that there is an affine subset A of X, every line of
which contains at least three points, but which is not properly contained in another
affine subset with the same property. Since (iii) does not hold, it follows from
Corollary 15 that A-:!: X. Hence there exist points x e X \A and z e A such that
<z,x> contains at least three points.
Assume there exist distinct points x 1 ,x 2 e <A u x> \ A such that
2. Linear geometries 65

<xi,..t2> = {xi,x2 }. Then <A u x> =<A u Xi> and it follows from Proposition 7
that x2 e <x',x">, where x' e [a',xi] and x" e [a" ,Xi] for some a',a" e A. Since
Xi e= <a',a"> and x 2 e= <a',a"> u <a',xi> u <a",xi>, we may suppose that
x' e (a',xi) andx" e [a",xi). It follows from Lemma 14 thatx' e (x2,x") and that
the plane <a',a",xi> contains only five points. Hence x" =a" and the line <a',a">
contains only two points, which is a contradiction.
We conclude that if xi,x 2 are distinct points of <Aux> \A, then
<xi,x2> :I:- {xi,x2 }. But, since A is maximal, there exist distinct Yi,y2 e <Aux>
such that <yi,y2 > = {yi,y2 }. We may choose the notation so that Yi= a e A and
y 2 e= A. If <a,x> ¢ {a,x}, then y 2 :1:-x and <y2,x> :I:- {y2,x}. Hence, by Lemma
14, there exists a pointy such that x e (a,y) and <y,y2 > = {y,y2 }. Since ye= A,
this is a contradiction. Thus we may take y2 = x.
Since the lines <z,x> and <z,a> contain at least three points, it now follows
from Lemma 14 that there exist points a',x' such that z e (a,a') n (x,x') and the
plane <z,a,x> contains only the points z,a,x,a',x'. Hence, by Corollary 15,
A = { z,a,a'}.
Choose any point be= <z,a>. If <z,b> :I:- {z,b}, then any affine subset
properly containing <z,b> has a line containing only two points. In particular the
plane <z,a,b> has a line containing only two points. The preceding argument, with
A replaced by <z,a> and x by b, shows that there exists a point b' such that
z e (b,b1 and <z,a,b> = {z,a,b,a',b'}.
Let C be the union of z with the set of all points c e X such that <z,c> ¢ { z,c}.
It follows from what has been said that C is affine. Moreover C :1:-X, since (iv)
does not hold. Hence there exist points p e X \ C and q e C such that
<p,q> :I:- {p,q}. Since <z,p> = {z,p}, we must have q :I:- z and <z,q> ¢ {z,q}.
Hence <z,q> = {z,q,q'}, where z e (q,q'). But since <p,q> :I:- {p,q}, it follows
from Lemma 14 also that q e (z,q1. Thus we have a contradiction. D

An irreducible linear geometry for which case (iv) of Proposition 16 holds may
be called a multi-cross with (uniquely determined) centre z, since each plane
containing z has exactly four other points, ui,u 2 ,vi,v 2 say, with
z e (ui,Vi) n (u2,v2). If Vis a vector space over an ordered division ring D with
basis {e;} ie/t then the restriction of V to the set X = {O,e;,--e;} ie/ is a multi-cross
with centre 0.
66 111. Linearity

3 BELLY'S THEOREM AND ITS RELATIVES

It will now be shown that in a linear geometry Proposition 11.13 admits a


stronger formulation:

PROPOSITION 17 In a linear geometry the classes of Helly sets, Radon sets


and affine independent sets all coincide.

Proof By Proposition 11.13 we need only show that every affine independent set is
a Radon set, and that every Belly set is affine independent. Moreover we may
restrict attention to finite sets.
Assume first that S = {s 1, ••• ,snl is an affine independent set, but not a Radon
set. Then we may choose the notation so that [Si] n [S 2 ] 0, where
':/:
S 1 = {s1 , •.. ,Sm} and S2 = {sm+l , ... ,sn} for some m with 1 < m < ti. Let
x e [Si] n [S 2]. Then x ':/: s 1, since [S2] s: <S\s 1> and Sis affine independent.
On the other hand x e <s 1, ••• ,sm>. Hence we can choose k, with 1 < k < m, so
that x ~ <s 1 , ••• ,sk> but x e <s 1 , ••• ,sk+t>. Then sk+l e <x,slt···,sk>, by
Proposition 8. Since x e <sm+lt···,sn>, it follows that sk+l e <S\sk+ 1>, which is
a contradiction.
To show that a finite Belly set is affine independent we use induction on the
cardinality of the set. Since the result is true for singletons, it is sufficient to
establish the following assertion:

(#) If T = {x 1, ••• ,xn} is an affine independent set and if Xn+l e <T> \ T, then
S =Tu Xn+l is not a Belly set

This is certainly true for n = 2, since either x 3 e [x 1,x2] or x 1 e [x2,x3] or


x2 e [x1,x3]. We assume that n > 2 and the assertion holds for all smaller values of
n. We may assume also that Xn+l ~ [T], since Xn+l e [T] implies Xn+l e [S\x;]
for i = l, ... ,n+l.
By Proposition 7 we have Xn+l e <y,z>, where y,z e [T] and y ':/: z. By
Proposition 11.14, there exist xi,xk e T such that ye [z u (T \xj)] and
z e [y u (T \xk)]. Consequently ye [z,y1, where y' e [T\xj], and z e fy,z1,
where z' e [T\xk]. Then y,z e fy',z1, by (L2), and hence Xn+l e <y',z'>.
Since Xn+l ~ [T], we h~ve Xn+l ~
[y',z1. Without loss of generality, suppose
z' e (y',Xn+ 1). We may assume z' ~ [T \xi] since otherwise, by the induction
hypothesis, S\xi (and hence also S) is not a Belly set. Thenj ':/: k and y' e [xk,u],
3. Helly's theorem and its relatives 67

where u e [T \(xi u xk)]. Hence z' e [xk,xn+l•u], and so z' e [u, v], where
v e [xk,xn+il· Since u e [T\ (xj u xk)] and z' e [T\xk], it follows that

If v e T\xk, say v = X;, then Xn+l e <x;,X/r? and we are reduced to the case n =2.
If v t! T\xk then, by the induction hypothesis, the set

is not a Belly set. Thus there exists a point x e [T\xtJ such that x e [R \xi] for all
j e { 1, ... ,n} with j :I: k. Since v e [xk,xn+il' it follows that x e [S \x;] for
i = l, ... ,n+ 1. D

COROLLARY 18 For any set S s: X alld ally x e [S], there exists a finite
affine independent set F s: S such that x e [F] .

Proof Let F be a finite subset of S such that x e [F], but no proper subset of F
has the same property. Then F is a Caratheodory set. Hence F is a Helly set, by
Proposition 11.15, and thus also an affine independent set, by Proposition 17. D

Proposition 17 no longer holds if all the axioms for a linear geometry are
satisfied with the exception of (P). This is shown by the example illustrating the
independence of the axiom (P) in Example 11.10. In this example S = {a,a',b,b'} is
a Helly set, but it is not affine independent since a' e <a,c> s: <a,b,b'>.
If Sis an affine independent subset of Rd, then ISi S d + 1. Consequently, by
Proposition 17, any Belly set in Rd has at most d + 1 elements and hence, by
Proposition 1.10, a finite family of at least d + 1 convex sets in Rd has nonempty
intersection if every subfamily of d + 1 sets has nonempty intersection. This was
first proved by Helly (1923).
By Proposition 17 also, any Radon set in Rd has at most d + 1 elements.
Hence any subset S of R. d with ISi > d + 1 has disjoint subsets S 1,S 2 such that
[Si] n [S 2] :1: 0. This was first proved by Radon (1921), who used it to give the
first published proof of Helly's theorem.
Similarly it follows from Corollary 18 that if S s: Rd and x e [S], then there
exists a subset F of S with IFI < d + 1 such that x e [F]. This was first proved by
Caratheodory (1911).
68 Ill. Linearity

The formulation of Proposition 17 has several advantages over the classical


formulations: the condition given is both necessary and sufficient, the result holds
in vector spaces over any ordered division ring, and the vector space need not be
finite-dimensional. (For Corollary 18, see also Proposition V.2 below.)
There is an extensive literature dealing with generalizations, modifications and
analogues of the theorems of Helly, Radon and Caratheodory; see Danzer et al.
(1963), from which the present section takes its title, and Eckhoff (1979,1993).
We restrict ourselves to describing some interesting recent results in this area and
drawing attention to some open problems.
Barany ( 1982) has proved the following theorem, which contains
Caratheodory's theorem as a special case (S1 = ... =Sd+t =S). However, the proof
uses Caratheodory's theorem itself.

BARANY'S THEOREM If a point x belongs to the convex hulls of d+I sets


S 1, ••• ,Sd+t in Rd, then there exist points X; e S; (i = l, ... ,d+l) such that x
belongs to the convex hull of {x1, •.• ,xd+tl·

Proof We may assume, without loss of generality, that x = 0 and that JRd is
equipped with the Euclidean norm. Furthermore, by Caratheodory's theorem, we
may suppose that each set S; contains at most d+ 1 points. Then the collection 9' of
all sets S = {x 1, ••• ,xd+d withx; e S; (i = l, ... ,d+l) is finite. For any Se 9', put

d(S) = inf {llxll: x e [S]}.

Since [S] is compact, we have d(S) = llzll for some z e [S]. We wish to show that
d(S) =0 for some S e 9'. Since 9' is finite, it is sufficient to show that if d(S) > 0,
then there exists an S' e 9' such that d(S') < d(S).
If S is not affine independent then, by Caratheodory's theorem, z is a convex
combination of at most d points of S. If Sis affine independent, then the nearest
point of [S] to the origin lies on some proper face. In either case z e [S \xi] for
some j e { l, ... ,d+l }.
The origin cannot be separated from Si by any hyperplane, since 0 e [Sj].
Consequently the open half-space {x e Rd: (x - z,z) < 0}, which contains the
origin, must also contain a point x/ e Si. Let S' be the set which is obtained from S
by replacing xi by x/ and retaining the other x;. Then [S1 contains the segment
[x/.z], since it contains both x/ and z. But, for small A.> 0,
3. Helly's theorem and its relatives 69

llA.x/ + (1-A.)zll2 = llzll2 + 2A.(x/- z,z) + O(A.2)


< llzll 2•
Hence d(S1 < d(S). [J

Another proof of Barany's theorem is given by Kovijani~ (1994). It may be


asked if Barany's theorem admits the following extension to linear geometries:

Let S 1, ••• ,Sm be subsets of a linear geometry X such that dim <S 1, ••• ,Sm>
< m. If x e [S 1] r. ... r. [Sm], does there exist an affine independent set
{x 1, ••• ,xn}, with Xj e S ;ifor distinct illtegers i1 , ••• ,in e { l, ... ,m }, such that
XE [X1, ..• ,Xn]?

Tverberg (1966) established a strong generalization of Radon's theorem, and a


simpler proof has been given by Tverberg and Vretica (1993). The following
deduction of Tverberg's theorem from Barany's theorem, due to Onn, is contained
in Sarkaria (1992).

TVERBERG'S THEOREM If a set S s JRd colltains (r-l)(d+l) + 1 distinct


points, for some illteger r > l, thell S can be expressed as the union of r
pairwise disjoint subsets whose convex hulls have a common point.

Proof We can identify JRd with the hyperplane ~ 1 + ... + ~d+t = 1 in JRd+l. Let
v0 ,v 1, ••• ,vn, where n = (r-l)(d+l), be the points of Sin this identification. With
each point v; e S we associate a set S;' = {M;i, ... ,M;,} of (d+l)x(r-1) matrices
defined in the following way: if 1 < k < r the matrix M;k has v; as its k-th column
and zeros elsewhere, whereas every column of M;, is-v;. If we also identify the
set of all (d+l)x(r-1) matrices with JRn, we haven+ 1 sets S;' s JRn with IS;'I = r.
Moreover, the origin is contained in the convex hull of every set S; ', since
M ;1 + ... + M ;, = 0. Consequently, by Barany's theorem, there exist a matrix
M;k; e S;' and A;> 0 with I,~ 0 A;= 1 such that

L~o A.;M;k; = 0.

Let Ii denote the set of all i e {0, 1,... ,n} for which k; = j. Then we can rewrite the
last equation in the form

Equivalently,
70 Ill. Linearity

L iel1 A;V; = L ie/2 A;V; = ··· = L iel, A;V i·

Put µj = I, ieli A;. Since each v; lies on the hyperplane ~ 1 + ... + ~d+l = 1 and
L7=o A;= 1, it follows that µ1 =µ2 = ... =µ, = l/r. Hence, if we denote by sj
the set of all v; with i e Ji, the given set S has a partition into pairwise disjoint
subsets S1, ... ,S, whose convex hulls [S 1], ... ,[S,] have a common point. D

Tverberg's theorem is best possible, in the sense that the number


(r-l)(d+l) + 1 in its statement cannot be replaced by any smaller number. To
understand the significance of this number, define the r-th Tverberg number t, of
JRd to be the maximum cardinality of any set in JRd such that, for every partition of
the set into r pairwise disjoint subsets, the intersection of their convex hulls is
always empty. In particular, t 2 is the maximum cardinality of any Radon set in JRd.
Then Tverberg's theorem can be simply restated in the form: t, = (r-l)t2•
In the definition of Tverberg numbers, the space JRd can be replaced by an
arbitrary alignment. Eckhoff (1979) has conjectured that the inequality t, < (r-l)t2
holds in any alignment for which t 2 is finite. This sweeping generalization of
Tverberg's theorem (first formulated by Calder (1971) for Example 4 of Chapter IQ
has been neither proved nor disproved, although it is known that t 2 < oo implies
t, < oo for every r > 2. In the context of the present chapter it may be asked: does
Eckhoff's conjecture hold at least in any linear geometry?
There is another conjectured generalization of Tverberg's theorem within JRd
itself: if a set S s JRd contains r(d+ 1) distinct points, partitioned into d+ 1 subsets
A 1, ... ,Ad+t containing r points each, can S be expressed as the union of r pairwise
disjoint subsets S1, ... ,S, whose convex hulls have a common point in such a way
that every subset S; (i = l,... ,r) contains exactly one point from each subset Ak
(k = 1,... ,d+l)? A proof for r = 2 is contained in Barany and Larman (1992).
They also prove the conjecture for d =2.
IV
Linearity (continued)

We show here that faces may be given yet another characterization in a linear
geometry, and that polytopes possess most of their familiar properties in 1Rd. A
notable omission is the Euler-Poincare relation. It is shown also that from a given
linear geometry further factor geometries may be derived.

1 FACES

Throughout this chapter we assume that a set Xis given on which a linear
geometry is defined. The notion of 'face' of a convex set was defined in any
aligned space in Chapter I, and in Chapter II a simpler characterization was given in
any convex geometry. Yet another characterization will now be given in any linear
geometry:

PROPOSITION 1 A set A is a face of a convex set C if and only if


Cr. <A>= A and C\A is convex.

Proof Let A be a set such that Cr. <A>= A and C\A is convex. Then A s C
and A is convex, since it is the intersection of two convex sets. If a e (c,c1, where
a e A and c,c' e C, then at least one of c,c' is in A, since C \A is convex, and in
fact they both are, since C r. <A> =A. Thus A is a face of C, by Proposition 11.6.
On the other hand, if A is a face of C then C\A is convex, by Proposition 1.5.
Put B =Cr. <A>. Then A is a face of B, by Proposition 1.7, since B is convex
and A s B s C. By Proposition 111.7, if be B then be <a,a'> for some
a,a' e A. If be [a,a1 then be A, since A is convex. If a' e (a,b) or a e (a',b)
then be A, since A is a face of B. Thus B =A. D
72 IV. Linearity (continued)

COROLLARY 2 Let A and B be faces of a convex set C. Then A properly


contains B if and only if <A> properly contains <B>. D

PROPOSITION 3 If a convex set C is contained in a closed half-space of <C>


associated with a hyperplane H, then C n His a/ace of C.

Proof Put D = C n H and let H+ be the open half-space of <C> associated with H
such that Cs Hu H+. Then C n <D> =D, since D s C n <D> s C n H, and
C\D is convex, since C\D = C n H+. D

PROPOSITION 4 Let C be a convex set and H + an open half-space of <C>


associated with a hyperplane H. Also, let A be a subset of B = C n (Hu H+)
such that A n H is a face of C (possibly 0). Then A is a face of C if and only if
A is a/ace of B.

Proof By Proposition I.7 we need only show that if A is a face of B, then A is


also a face of C. Put H_ = <C> \(Hu H+)·
We show first that C n <A>= A. Since B n <A>= A, it is enough to show
that if ye C n H_, then y E <A>. Assume on the contrary that ye <A>. Then
ye <x,z> for some x,z e A and we may choose the notation so that z e (x,y).
Hence there exists a point we (y,z] n H. Moreover w ~ A, since An His a face
of C. Hence w '¢ z and z e (x,w). But, since A is a face of B, this contradicts
WE A.
We show next that C \A is convex. Assume on the contrary that, for some
points x,y e C \A, there exists a point z e (x,y) n A. Since B \A is convex, we
may assume that y e H _. Then x e H +' since H u H _ is convex. Hence there
exists a point we (x,y) n H, and z e [w,x), which leads to a contradiction in the
same way as before. D

PROPOSITION 5 Let C be a convex set and S any subset of C. Then the subset
As of C, consisting of the union of C n <.S> with the set of all points x,x' e C
such that (x,x') n <S> '¢ 0, is a face of C. Moreover As is contained in every
face of C which contains S.

Proof We show first that As is convex. Suppose z e (x,y), where x,y are distinct
elements of As with x E <.S> and z E <S>. Then z e C and there exist points
x' e As, a e <.S> such that a e (x,x'). If a,x,y are collinear, then a e (z,z') for
some z' e {x,y,x'} s C and so z e As. Hence we now suppose that a,x,y are not
1. Faces 73

collinear. Then, by (P)', there exists a point c e (a,y) n (x',z). If ye <S>, then
c e <S> and hence z e As. If y E <S>, there exist y' e As and b e <S> such that
b E (y,y1.
If b =a then, by Proposition II.19', a e (z,z1 for some z' e (x',y1 and hence
z e As. Thus we now suppose b ':/!a and hence y E <a,b>. By (P)', there exists a
point de (a,b) n (c,y1. Moreover de C n <S>, since a,b e C n <S>. Thus we
may suppose d ':/! z. We may also suppose that y' E <x',z>, since otherwise
de (z,x1 or de (z,y1. Then, by Proposition 11.2', there exists a point e e (x',y1
such that de (e,z). Thus z e As in every case.
We now show that As is a face of C. Suppose x e As and x e (c,c1, where
c,c' e C. If x e <S> then c,c' e As, by the definition of As. If x E <S>, then there
exist x' e As and a e <S> such that a e (x,x1. If a,c,c' are collinear, then c =a or
a e (c,x) or a e (c,x1. In any event c e As, and likewise c' e As. Thus we now
suppose that a,c,c' are not collinear. Then, by Proposition 11.2', a e (c,y) for some
y e (c',xl Hence c e As, and likewise c' e As.
It follows at once from Proposition 1 that any face of C which contains S must
also contain As. D

COROLLARY 6 Let C be a convex set and a e C. Then the set Aa consisting


of a and all points x,x' e C such that a e (x,x1 is a face of C. Moreover Aa is
contained in every face of C which contains a. D

A face F of a convex set C will be said to be a facet of C if <F> is a


hyperplane of <C>, and an edge of C if dim <F> = 1.

PROPOSITION 7 Three distinct face ts of a convex set cannot themselves have


a common nonempty facet.

Proof Assume on the contrary that three distinct facets F 1,F2,F3 of a convex set C
have a common nonempty facet F. Then <F 1>,<F 2 >,<F 3 > are distinct
hyperplanes of <C> with the common hyperplane <F>. Choose xk e Fk \F
(k = 1,2,3). Since C n <Ft> =Fk and C \ Fk is convex, we have

But this contradicts Proposition Ill.13. D


74 IV. Linearity (continued)

2 POLYTOPES

We define a polytope to be the convex hull of a finite set. Although this


definition makes sense in any convex geometry, or even in any aligned space, the
theory of polytopes is richer in a linear geometry. We show first that, for
polytopes, the notion of 'face' is especially significant.

PROPOSITION 8 Let P be a polytope and S the (finite) set of extreme points


of P. Then the faces of Pare the sets [11, where T s Sand [S \ 71r.<T>=0.

Proof By Proposition 1.11, P = [S]. We show first that if Fis a face of P, then
F = [71 for some T s S such that [S \ T] r. <T> =0. Evidently we may assume
that F ¢ 0, P. Let T denote the set of all extreme points of P which are contained
in F. Then [S\71 s P\F, since P\F is convex, and thus Tis a nonempty proper
subset of S. If x e F, then x e [y,z] for some y e [S \ 71 and z e [T]. Thus
ye P \F and z e F. Since x ¢ z would imply ye <.F>, which contradicts
Pr. <F> = F, we must have x = z e [T]. Hence F = [T] and [S\ T] r. <T> = 0.
We show next that [T] is a face of P if T is a subset of S such that
[S \ n r. <T> = 0. Let x E pr. <T>. Then x E [y,z], where y E [S \ n and
z e [71. In fact x = z, since x ¢ z would imply y e <T>. Thus P r. <T> = [T]. It
remains to show that P \ [T] is convex.
Suppose z e (x,y), where x e [S], ye [S \ T] and z e [T]. Then x e [y',z1,
where y' e [S \ 71 and z' e [T]. Since [S \ T] r. <T> = 0, we must actually have
x e (y',z') and x,y,z' are not collinear. Hence there exists a point z" e (y,y') such
that z e (z',z"). Since z" e [S \ T] r. <T>, this is a contradiction.
If P \ [T] is not convex then, for some points x' ,x" e P \ [T], there exists a
point z e (x',x") r. [T]. Then x' e [y',z'], where y' e [S \ T] and z' e [T].
Moreover, by what we have already proved, x' e (y',z') and x",y',z' are not
collinear. Hence there exists a point z" e (x",y') such that z e (z',z"). Since
z" e P r. <1'> = [T] this yields a contradiction, as we have already seen. a
COROLLARY 9 If Pis a po/ytope and Fa face of P, then Fis a polytope and
E(F) = E(P) r. F. D
COROLLARY 10 A polytope has only finitely many faces. More precisely, a
polytope with n extreme points has at most 2n faces. D
2. Polytopes 75

The bound 2" in Corollary 10 is actually attained in the case of a simplex:

PROPOSITION 11 Let P = [SJ, where S is a finite affine independent set.


Then a set F is a face of P if and only if F = [TJ for some T s S.

Proof To show that every set [TJ, where T s S, is a face of P it is sufficient to


show that [S\s] is a face of P, for every s e S. But this follows at once from
Proposition 8. a

We establish next some further properties, showing that the polytopes


considered here behave in the manner to which we are accustomed.

PROPOSITION 12 If P is a polytope and t a line such that P (') t ':I: 0, then


p (') eis a segment.

Proof Suppose P = [S], where S = {s 1, ••• ,sn}. Since the result is obvious if
n < 2, we assume that n > 2 and the result holds for all smaller values of n.
Obviously we may assume also that t contains two distinct points x,y of P. The
points of the line t may be totally ordered, as in Proposition IIl.4. By the induction
hypothesis, [S \s;] (') t either is empty or has the form [a;,b;], with a;< b;, for
each i e {l,... ,n}. If we put a= min; a; and b =max; b;, then [a,b] s P(') t.
Moreover, by Proposition 11.16, P cannot contain a point oft\ [a,b]. D

It may be noted that the axiom (P) is not required in the proof of Proposition
12.

PROPOSITION 13 If P is a polytope and L an affine set, then P (') L is a


polytope.

Proof Suppose P = [S], where S = {x 1, ••• ,xn}. Since the result is obvious if
n = 1, we assume that n > 1 and the result holds for all smaller values of n. Let
p e P (')L. By Proposition Il.14, we have P = Ui:1P;, where P; = [p u
(S\x;)].
It is enough to show that the sets P; (') L (i = 1,... ,n) are all polytopes. For if
P; (')Lis the convex hull of a finite set S; (i = l, ... ,n), then

Since S 1 u ... u Sn s P (') L and P (') L is convex, we must in fact have


[S 1 u ... u Sn]= P (') L.
76 IV. Linearity (continued)

Without loss of generality, we show. only that the set P 1 n L is a polytope.


Put P 1' = [x2, ••• ,.xnl and P' = P 1' n L. Since P' is the convex hull of a finite set,
by the induction hypothesis, we will complete the proof by showing that

The right side is certainly contained in the left, since p u P' = (p u P 11 n Land
P 1 = [jJ u P 11. On the other hand, the left side is contained in the right. For if
x e P 1 \ p then x e (p ,y] for some y e P 1 ', and if x e (P 1 n L) \ p then
ye P 1' n L. a

PROPOSITION 14 The intersection of a polytope P with a closed half-space of


X is again a polytope.

Proof Let H + be an open half-space of X associated with a hyperplane H and


H + u H the corresponding closed half-space. To show that P n (H + u H) is a
polytope we may obviously assume that Pis not contained in H+ u Hand, by
Proposition 13, we may assume that P n (H+ u H) is not contained in H. Then
X \ H is not convex and P contains points in both open half-spaces H +,H_ of X
associated with the hyperplane H. Hence the sets S+,S- of extreme points of Pin
H +,H_ are both nonempty. If S' is the set of extreme points of the polytope
P' = P n H, then the intersections of P with the closed half-spaces H+ u H,
H_ u H are the polytopes P + = [S' u S+], P_ = [S' u S_], since P + u P _ = P,
P+ n P_=P'. D

PROPOSITION 15 Any proper face F of a polytope P is contained in a facet of


P. Moreover, if dim F = dim P - 2, then F is contained in exactly two facets of
P and is their intersection.

Proof We begin by proving the following assertion:

(#) Let S be the set of extreme points of P, let S 1 be a subset of S such that
L 1: = <S 1> c <S> and let F 1 be a facet of the polytope P 1 =P n L 1• If
L 2: =<b,S1>, where be S\L 1, then the polytope P2 = P n L2 has a facet F2 such
that F 1 c F 2•

By hypothesis H 1 = <F 1> is a hyperplane of L 1 and, if a e S 1 \H 1, P 1 is


contained in the closed half-space of L 1 associated with H 1 which contains a. Let T
be the set of extreme points of P2 and let M be the open half-space of Li associated
2. Polytopes 77

with the hyperplane Li which contains b. We can choose bi ,... ,bm e T r. M so


that every element of Tr. M belongs to <b;,Hi> for some i and bi E <b;,Hi> if
j ¢ i. We will show that the set {bi,... ,bml is partially ordered by writing b; S bi
if [a,bj] rt <b;,Hi> ¢ 0.
It is evident that b; Sb; for every i. Suppose b; < bj and bi Sb;. Then there
exist points c; e [a,bj] r. <b;,H i > and cj e [a,b;] r. <bj,Hi >. Hence there exist a
point x e [b;,c;] r. [bj,cj] and a pointy e [b;,bj] such that x e [a,y]. Then
x e <b ;.H 1> r. <bj,H 1>. Moreover x ~ Hi' since y E Li· Hence
<b;,H1> =<.x,Hi> = <bj,H 1>, which implies b; =bi.
Suppose next that b; S bi and bi< bk. We wish to show that b; <bk. There
exist points c; e [a,bj] r. <b;,H 1> and ci e [a,bk] r. <bj,Hi>. If there exists a
point x e [bj,cj] r. <b;,H 1>, then x e [a,y] for some ye [bj,bk]. Moreover
x E H lt since y ~ L., and hence, as before, b; = bi S bk. Thus we may now
assume [bj,Cj] r. <b;,H1> = 0. Then bi and Cj lie in the same open half-space of L 2
associated with the hyperplane <b;,H 1>. Since a and bi lie in different open half-
spaces, it follows that a and Cj lie in different open half-spaces, and hence so also
do a and bk. Thus [a,bk.1 r. <b;,H 1> ¢ 0 and b; S bk.
This completes the proof that the set {bi, ... ,bm} is partially ordered. We now
choose the notation so that bm is a maximal element of this partially ordered set.
Thus, putting H 2 = <bm,Hi>, we have [a,b;] r. H 2 = 0 for all i < m. We are
going to show that F 2 = P r. H2 is a facet of the polytope P 2 =P r. L 2• Since
<F2> = H2 is a hyperplane of L 2, we need only show that P 2 is contained in the
closed half-space of Li associated with the hyperplane H2 which contains a.
Assume on the contrary that, for some a' e T, there exists a point
x e (a,a') r. H 2• Then, by construction, a' E M. Ha' e Li then x e Hi' since
bm E Li. But then a and a' lie in different open half-spaces of Li associated with
the hyperplane Hi, which is a contradiction because Fi is a facet of P 1• Thus we
now suppose a' E Li. Then a' e N, where N is an open half-space of L 2
associated with the hyperplane Li and N ¢ M. Thus there exists a point
ye (a',bm) r. Li. Hence a,a',bm are not collinear and there exists a point
z e (a,y) r. (bm,x). Since z e Li r. H 2, we must actually have z e Hi. Since
y e Pi' this again contradicts the fact that F 1 is a facet of Pi· This completes the
proof that F2 is a facet of P 2, and it is obvious that Fi c F 2• Thus we have now
proved(#).
78 IV. Unearity (continued)

If F 2 ' = P (')Li is a facet of P 2, then F 2 (') F 2' =Fi since H 2 (')Li= Hi. If
P (') Li is not a facet of P 2, then T (') N ~ 0 and we can in the same way choose
points d1t···,d,, e T (') N and show that F 2' = P (') H2', where H 2' = <d,,,Hi>, is
also a facet of P 2 containing Fi· Since the segment (bm,d,,) contains a point x e Li,
and H 2 ' = H 2 implies x e Hi' the facets F 2 and F 2' will be distinct if
(bm,d,,) (')Hi= 0. Thus they are certainly distinct if Fi is a face of P 2. Again,
F 2 (') F 2' =Fi since H2 (') H 2' =Hi. If we take Fi= F to be a face of P such that
dim F =dim P - 2, then P 2 = P and F is the intersection of the facets F 2,F2' of P.
Since Fis contained in at most two facets of P, by Proposition 7, this completes the
proof of the second statement of the proposition.
To prove the first statement of the proposition, take F 1 = F to be a face of P
such that dim F <dim P -2 and put Hi= <Fi>. Then Hi= <R>, where R is the
set of extreme points of P which are contained in F. If Li = <a,R>, where
a e S\Hi, then Fi is a facet of Pi= P (')Li. By repeatedly applying(#), we see
that Fis contained in a facet of P. D

From Proposition 15 we can deduce many other properties of polytopes:

COROLLARY 16 If F and Gare faces of a polytope P such that F c G, then


there exists a finite sequence F 0 ,F i , ... ,F, of faces of P, with F 0 = F and
F, = G, such that F;_i is a facet of F; (i = l, ... ,r). D

PROPOSITION 17 Any proper face F of a polytope P is the intersection of all


facets of P which contain it.

Proof Put d = dim P - dim F. If d = 1 the result is obvious and if d = 2 it holds


by Proposition 15. We assume that d > 2 and the result holds for all smaller values
of d. There exist faces G,G' of P such that G (') G' = F and Fis a facet of both G
and G'. Since, by the induction hypothesis, G and G' are the intersections of all
facets of P which contain them, so also is F. D

PROPOSITION 18 q F,G,H are faces of a po/ytope P such that F s: G s: H,


then there exists a face G' of P such that F s: G' s: H, F = G (') G', and every
face of P which contains both G and G' also contains H.

Proof We may suppose F c G c H, since if G = F we can take G' =Hand if


G = H we can take G' = F. Put n =dim H - dim F. ff n = 2 the result holds by
Proposition 15. We assume that n > 2 and the result holds for all smaller values of
2. Polytopes 79

n. If m =dim G - dim F then, since n > 2, either m > 1 or m < n - 1 (or both).
Without loss of generality suppose m > 1 and let Fi be a face of P such that
F c Fi c G and F is a facet of Fi· Then by the induction hypothesis there is a
face G" of P such that Fi s G" s H, Fi= G (') G" and every face which contains
both G and G" also contains H. Since G" ~ H, there also exists a face G' of P such
that F s G's G", F =Fir'\ G' and every face which contains both F 1 and G' also
contains G". Since G' s G",

G (') G' = (G (') G") (') G' = F 1 (') G' = F

and, since Fi s G, any face which contains both G and G' also contains G" and
hence also contains H. D

The next result shows, in particular, that polytopes are polyhedra:

PROPOSITION 19 Let P be a polytope and Fi, ... ,Fm its facets. If Q; is the
closed half-space of <P> associated with the hyperplane H; = <F;> which
contains P (i = l, ... ,m}, then

Proof Put Q = nf!1 Q;. Then obviously P s Q. We will assume that there exists
a point a e Q \P and derive a contradiction. Since a e <P>, we have a e <x,y>
for some distinct points x,y e P, by Proposition III.7. Moreover, by Proposition
12, we may choose x,y so that P (') <x,y> = [x,y] and we may choose the notation
so that x e (a,y). Then x and y belong to proper faces of P, by Corollary 6.
Hence, by Proposition 15, x and y belong to facets of P. Thus x e H k for some k.
Since a and y cannot lie in different open half-spaces of <P> associated with the
hyperplane Hk' it follows that a,y e Hk and hence y belongs to the same facets of P
as x. Hence, by Proposition 17, y belongs to every face of P which contains x. In
particular, by Corollary 6 again, x e (y,y') for some y' e P. But this is a
contradiction. a
From Propositions 14 and 19 we immediately obtain

PROPOSITION 20 The intersection of two polytopes is again a polytope. D

We can now show also that two-dimensional polytopes are polygons:


80 IV. Linearity (continued)

PROPOSITION 21 If P is a two-dimensional po/ytope, then the n ~ 3 extreme


points of P can be numbered e 1, ... ,en so that the facets of Pare precisely the
segments [e;,e;+ 11 (i = l, ... ,n-1) and [en,eiJ.

Proof Since the result is obvious if 11 = 3, we assume that n > 3 and the result
holds for two-dimensional polytopes with fewer than n extreme points. Let S
denote the set of extreme points of P. If e e S then, by Proposition 15, there exist
distinct e',e" e S such that [e,e1 and [e,e"] are facets of P. Suppose/e Sand
f-::1: e,e',e". Then <e,e'> n [e"J] = 0 = <e,e"> n [e'J]. Hence, by Lemma IlI.9,
<el> n [e',e"] -:I: 0. Since all points involved are extreme points of P, we must
actually have {e/) n (e',e") -:I: 0. It follows that [e',e"] is a facet of the polytope
[S\e]. The result now follows from the induction hypothesis. a
With any polytope P there is associated a graph, whose vertices are the
extreme points of P and whose edges are the edges of P. It will now be shown that
the graph of a d-dimensional polytope is d-co11nected, i.e. it is connected and
remains so whenever less than d vertices are removed. This was first proved, for
polytopes in Euclidean space, by Balinski (1961).

LEMMA 22 Let P be a polytope, Ga facet of P and Fa/ace of G. Then there


exists an extreme point e' of P such that e' ~ <G> and P n <e',F> is a face of
P.
Proof Put d =dim P. Since the result is trivial when d = 0 or 1, we assume that
d ~ 2 and the result holds for all smaller values of d.
We may suppose F c G, since if F = G we can take e' to be any extreme
point of P which is not in G. Let F' be a facet of G which contains F, and let G' be
a facet of P such that F' = G n G'. By the induction hypothesis there exists an
extreme point e' of G' such that e' ~ <F'> and G' n <e',F> is a face of G'. Then e'
is an extreme point of P, e' ~ <G> and P n <e',F> = G' n <e',F> is a face of P.
a
PROPOSITION 23 Let P be a d-dimensiona/ polytope, S the set of extreme
points of P and E a subset of S such that 0 < IS\ El < d. Then any two distinct
points e,e' e E can be connected in Eby edges of P, i.e. there exists a finite
sequence eo,e 1 , ... ,em of elements of E, with e0 = e and em = e', such that
[e;_1,e;] is an edge of P for i = l, ... ,m.
2. Polytopes 81

Proof Fix e e E. It follows from Lemma 22, with F = {e}, that there exists an
affine independent set Vof d+l extreme points of P, including e, such that [eJ] is
an edge of P for every f e V \ e. Hence we may assume that e' ~ V and that the
result holds for all d-dimensional polytopes with fewer extreme points than P.
Since <V> =<P>, it follows that e' e <S \ e'>.
By applying Proposition 19 to the polytope Q =[S \ e1, we see that there
exists a facet F of Q such that e' does not lie in the closed half-space of <P> = <Q>
associated with the hyperplane <.F> which contains Q. We have F = [T], where
Tc: S\e'. If/ e T then [e'J] is an edge of the polytope [e' u T], by Proposition
8, and hence also of P, by Proposition 4. Thus we may assume that e e U, where
U = S\ (e' u n. By the induction hypothesis any two distinct extreme points of Q
which are in E can be connected in E by edges of Q. Hence, by Proposition 4, e
can be connected in E to some extreme point f e T (') E by edges of P.
Consequently e can also be connected to e'. a

3 FACTOR GEOMETRIES

In this section we show how new linear geometries may be created from a
given one. This property, and the theory of polytopes constructed in the previous
section, are persuasive evidence that the concept of linear geometry has intrinsic
significance.
Let X be an arbitrary linear geometry, let H+ be an open half-space associated
with a hyperplane Hof X, and letL be any nonempty affine subset of H. For any
a e H +' we denote by Pa the closed half-space of <a u L> associated with the
hyperplane L which contains a. In symbols, Pa =<au L> (') (H + u H).
Consequently Pa \L =<au L> (') H+ and <Pa>= <au L>. Evidently Pa·= Pa
for any a' e Pa \L.
The collection H + : L of all such, closed half-spaces Pa with a e H + can be
given the structure of a linear geometry, the factor geometry of H + with respect to
L, in the following way. For any p 1,p 2 e H+: L, we define the sector [p 1,p2] to
be the set of all p e H+: L such that a e [a 1,a2] for some a e p \L, a1 e Pt \Land
a2 e p2 \L. As we now show, we may actually fix a1 and a2.
82 IV. Linearity (continued)

LEMMA 24 Suppose p e [p 1,p 2], where p 1,p 2 e H+: L. If a 1 e p 1 \Land


a2 e p 2 \L, then there exists a e p \L such that a e [a 1,a2].

Proof Evidently we may assume p 1 if: p2 and p if: p 1,p2. Then <P1> if: <P2>. By
hypothesis there exist points a' e p \L, a 1' e p 1 \L and a 2' e p2 \L such that
a' e [a 1 ',a 2 1. In fact a' e (a 1',a 2 '), since p if: p 1,p 2. If we put L; = <p ;> =
<a;' u L> (i = 1,2), then L 1 and L 2 are hyperplanes of L* = <a 1',a 2 ',L>.
Moreover L 1 n L 2 = L, by the exchange property. Since a' e (a1',a2 1, it follows
that L' = <p> =<a' u L> is also a hyperplane of L* and L' il:L 1,L2. Moreover a 1'
and a 2' lie in different open half-spaces of L* associated with the hyperplane L'.
From L 1 n L' =Land [a 1,a 11nL=0 we obtain [a 1,a 11nL'=0. In the same
way [a 2,a21nL'=0. Hence also a 1 and a 2 lie in different open half-spaces of
L* associated with the hyperplane L'. Consequently there exists a point a e
(a 1,a 2) n L'. Evidently a and a 2 lie in the same open half-space of L* associated
with the hyperplane L 1• But a and a' also lie in the same open half-space of L *
associated with the hyperplane L 1, since L 1 n L' = L. Therefore [a',a2] n L 1 = 0.
Assume a E p. Then ~here exist a point x e (a,a1 n L and a pointy e (a',a 2) such
that x e (a 1,y). Since ye [a',a2] n L 1, this is a contradiction. Thus a e p. 0

It will now be shown that, with sectors as 'segments', H+: L is a linear


geometry.

LEMMA 25 Suppose x e [a,b,c], where a,b,c are non-collinear points. If


x E [a,b],[a,c],[b,c], then x e (a,y)for some ye (b,c).

Proof By Proposition 11.16 there exist points y,z e [a,b] u [a,c] u [b,c] such that
a,x e [y,z]. Moreover a e {y,z}, since a is an extreme point of [a,b,c]. Without
loss of generality assume a= z. Then ye (b,c), since x E [a,b],[a,c], and x if: y,
since x E [b,c]. 0

PROPOSITION 26 If H + is an open half-space associated with a hyperplane H


of X and L any nonempty affine subset of H, then H+: Lis a linear geometry.

Proof We will prove that H+: L satisfies each of the axioms (C), (Ll)-(L4) and
(P). In only one case, namely (L3), does the proof present difficulty.

(Ll): It is obvious that [p,p] = {p} for any p e H+: L.


3. Factor geometries 83

(C): Suppose p' e [p,pi) and p" e [p',p2]. Choose a e p \L, at e Pt \L and
o 2 e p2 \L. Then, by Lemma 24, there exist a point o' e p'\L such that
o' e [o,oi) and a point o" e p"\L such that o" e [o',o2]. But o" e [o,d] for some
a E [Ot,a2]. If a e: L, it follows that p" E [p,p] for some p E Cpt,P21· On the
other hand, if a EL then o" E p \Land p" = p.

This already proves that H+ : L is a convex geometry.

(P): Suppose Pt' e [p,pi] and p 2' e [p,p2]. Choose o e p \L, o1 e Pt \Land
a 2 e p2 \L. Then there exist a point at' e Pt'\L such that Ot' e [o,oi] and a
point o 2 ' e p 2'\L such that o 2 ' e [o,o 2]. Hence there exists a point
o'e [01'021 (') Cot',02]. Moreover o'e H+, since Ot,02'e H+. Consequently
Pa·E [pt,P21 (') [Pt',p2].

(L4): Suppose p3 e Cpt,p 2] and p e [p 1,p2]. Choose o; e P; \L (i = 1,2). Then


there exist o3 e p3 \L such that o3 e [o 1,o2] and o e p \L such that o e [ot,o 2].
Hence o e [o t ,o 3] u [o 3,o 2], and p e [pt ,p 3] u [p 3,p21· Thus [pt ,P21 s
Cpt,p3] u [p3,p2l·
On the other hand, if p e [pt 'P 3] there exists o e p \ L such that
o e Cot,o 3] s Cat,o 2]. Hence Cpt,p 3] s Cpt,p 2], and similarly [p 3,p 2] s
Cpt,P2l· Thus Cpt,P21 = Cpt,p3] u [p3,p2].

(L2): Suppose p 2 e [pt,p 3], p 3 e [p 2,p 4] and p 2 ~ p 3. Choose o; e P; \L


(i = 1,3,4). Then there exist 02 e P2 \L such that 02 e Cot,03] and 03' e p3 \L
such that o3 ' e (02,04]. Moreover 02 ~ 03, since P2 ~ p3. If 03' = 03, it follows
that o3 e Cot,o4] and hence p3 e Cpt,p4]. If o3' ~ o3 , then o3' e [o3,o3"] for some
a3" e [Ot,04]. Since a 3" e <p3> (') H+, it follows that o3" e p3 \Land thus again
p3 E CP1tP4].

(L3): Suppose p 3 e: Cpt,p 2] and p 2 e: Cpt,p 3]. We wish to show that


Cpt,P2l (') Cpt,p3] ={pi}. We may suppose Pt e: [p2,p 3], since otherwise the
conclusion follows from Proposition II.18 (iii) (whose hypotheses we have already
shown to be satisfied). Choose o; e P; \L (i = 1,2,3). Then Ot ,a2,o3 are not
collinear and Cot,o2] (') <a3 ,L> = 0, Cot,o3] (') <o2,L> = 0, [02,03] (') <ot,L> =
0. Hence, by Lemma 111.9, <at,o2,o3> (') L = 0.
Assume, on the contrary, that there exists p e [pt ,p 2] (') [pt ,p3] with p ~ pt·
Then there exist o',a" e p \L such that o' e (ot,o 2), o" e (ot,o 3). Moreover
84 IV. Unearity (continued)

a' -:1: a", since a 1,a2 ,a3 are not collinear. Since a' e <a",L>, it follows from
Proposition 111.7 that a' e <a",/',/"> for some/',/" e L. Moreover <a',a"> n L =
0, since a',a" e <a 1,a2 ,a3>, and hence/' -:I:/".
Put Y = <l',l",a 1,a3>. Since a" e Y, we must have a' e Y and hence also
a2 e Y. Since Y contains the affine independent points a1,a2,a3,I', it follows that
dim Y =3. But, since Hk =<l',l",ak> (k = 1,2,3) are distinct planes in Y with the
common line Z = <1',I">, this contradicts Proposition Ill.13. D

4 NOTES

The faces of arbitrary convex sets do not share many of the properties
possessed by the faces of polytopes (or of polyhedra, which will be considered in
Chapter V). For example in R 3, with the usual definition of convexity, let
C =[p,q,S] be the convex hull of the set consisting of the circle S = {x =(~ 1 ,~ 2 ,~3):
(~ 1 - 1)2 + ~ 2 2 = l, ~ 3 = 0} and the two points p = (0,0,l), q = (0,0,-1). The
extreme points of C are p,q and all points of S except the origin. The edges of Care
the segments whose endpoints are distinct extreme points, not both of which are in
S. However, Chas no facets. (It may be noted also that, although C is compact,
the set of all extreme points is not closed.)
Polytopes in Euclidean space are extensively treated in the books of Br~ndsted
(1983), Grilnbaum (1967) and Ziegler (1995). The present treatment follows
Coppel (1995), but the key Proposition 15 has a stronger formulation here. It is
natural to ask if the Euler-Poincare relation, which connects the number of faces of
different dimensions of a polytope in JRd, can be generalized to any linear geometry.
Factor geometries are considered by Prenowitz (1961), Rubinstein (1964) and
Prenowitz and Jantosciak (1979), but they assume the axioms (D) and (U) which
will be considered in the next chapter.
v
Density and Unendingness

In this chapter two new axioms are introduced. A linear geometry is dense if no
segment contains exactly two points. The notions of intrinsic interior and convex
closure of a convex set are defined, and it is shown that in a dense linear geometry
they have many basic properties. In particular, any finite-dimensional convex set
has a nonempty intrinsic interior. A linear geometry is unending if any segment is
contained in another segment with different endpoints. An unending linear
geometry of dimension greater than one is necessarily also dense. In an unending
linear geometry the affine hull of the union of two affine sets A,B is the union of
the affine hulls of all sets {a,b}, where a e A and b e B, provided A (') B -:1: 0.
Furthermore a dense unending linear geometry may be given the structure of a
topological space, and actually of a Hausdorff space.
Following Prenowitz, we define products of arbitrary nonempty sets in a
dense linear geometry, and quotients also in a dense unending linear geometry.
Products and quotients possess many simple algebraic properties, and their
usefulness in elementary geometry has perhaps not been fully appreciated. We
further show that a substantial part of the usual theory of cones and polyhedra in
R.d carries over to any dense unending linear geometry.

1 DENSE LINEAR GEOMETRIES

A linear geometry will be said to be dense if it has the following property:

(D) if a,b e X and a -:1: b, then [a,b] -:1: {a,b}.

It is remarkable that this axiom actually implies the axiom (P), when all the other
axioms of a linear geometry are satisfied:
86 V. Density and unendingness

PROPOSITION 1 A convex geometry is a linear geometry if it satisfies the


axioms (L2)-(L4) and (D).

Proof It is sufficient to show that [y 1 ,z 2 ] n [y 2 ,z t1 ':/: 0 if z 1 e (x,y 1),


z2 e (x,y 2 ) and z1 ':/: z2 • Moreover, by Proposition III.6, we may assume that
x,y1,Y2 are not collinear.
By (D) we can choose z e (z 1,z2). Then z e (x,u) for some u e (y2,z 1), and
similarly z e (x,v) for some v e (y 1,z2). We may assume u ':/: v, since otherwise
there is nothing more to do. Then either u e (z,v) or v e (z,u). Without loss of
generality suppose u e (z,v). Then u e (z 1,w) for some we (y 1,z 2) and
w e (y 2,z 1 ') for some z1 ' e (x,y 1). Since the line <y 2 ,u> = <y 2 ,w> has at most
one point in common with the segment (x,y 1), we must have z1 ' = z1• Thus
we (y 1,z2) n (y 2 ,z 1). a

By Corollary III.15, a linear geometry is dense if not all points are collinear
and every line contains at least three points. Furthermore, a dense linear geometry
is irreducible. Throughout the remainder of this section we will assume that a
set X is given on which a dense linear geometry is defined.

PROPOSITION 2 The class of Caratheodory sets coincides with the class of


nonempty finite affine independent sets, and hence also with the class of
nonempty finite Helly sets and the class of nonempty finite Radon sets.

Proof We need only show that every nonempty finite affine independent set is a
Caratheodory set, by Propositions 11.15 and llI.17. Let S = {s 1, .•• ,snl be an affine
independent set. By (D) we can choose x 1 e (s 1,s2) and then, inductively,
xk e (xk-t,Sk+t) (k = 2, ... ,n - 1). Evidently Xn-l e [S] and, for each j < n,
Xj-l e Cstt···,sj] s <S\si+t>· However, xi~ <S\sj+t>, since Xj e <S\sj+t>
would imply Sj+t e <xi_ 1,xi> s <S\sj+t>, which contradicts the affine
independence of S. In particular,xn-1 ~ [S\sJ.
Assume, contrary to the proposition, that [S] = U7=1[S\s;]. Then Xn-1 e
[S \sm1 for some m < n. But if for some k with m < k < n we have xk e <S\sm>
then, since sk+l e S\sm, we also have xk-t e <xk,sk+t> s <S\sm>. It follows
that Xm e <S\sm> and, if m > l, even Xm-l e <S\sm>. Since the latter is a
contradiction, we conclude that m = 1 and x 1 e <S \ s 1>. But this is also a
contradiction, since St E <x1,S2> S <S\s1>. C
1. Dense linear geometries 87

The axiom (D) is essential for the validity of Proposition 2. For if x,y are
distinct points of a linear geometry such that [x,y] = {x,y}, then {x,y} is an affine
independent set but not a Caratheodory set.

PROPOSITION 3 If C is a convex set and Ai ,... ,Am affine sets such that

then C s A; for some i e { l, ... ,m}.

Proof The result is obviously true if m = 1. We assume that m > 1 and the result
holds for all smaller values of m. If C is not contained in Ai for some i then, by the
induction hypothesis, there exist a point a e C \ (U ~ 2 A;) and a point
be C\ (LJ~iiAi). Then a e Ai and be Am. If x e (a,b), then x e: Aiu Am.
Hence m > 2 and x e A 2 u ... u Am-i· Thus if we fix c e (a,b), then [a,c] s
Aiu ... u Am-i· Hence, by the induction hypothesis, [a,c] s Ai. Thus c e Ai,
which is a contradiction. D

We define the intrinsic interior Ci of a convex set C to be the subset of all


points of C which do not belong to a proper face of C. Thus Ci= C if C contains at
most one point.

PROPOSITION 4 Let C be a convex set and Ci its intrinsic interior. If x e Ci


and ye C, then (x,y) s Ci. In particular, Ci is also a convex set.

Proof Letze (x,y). If A is a face of C which contains z, then also x e A, by


Proposition 11.6. Hence A= C. D

The definition of 'intrinsic interior' may also be reformulated in the following


way:

PROPOSITION 5 The intrinsic interior Ci of a convex set C is the set of all


points a e C such that, for every b e C \a, there exists some c e C for which
a e (b,c).
Proof This follows at once from Corollary IV.6. D

We now introduce another concept of a topological nature and define the


convex closure C of a convex set C to be C itself if Ci =0, and otherwise to be the
88 V. Density and unendingness

set of ally e X such that [x,y) s Ci for some x e Ci. In either event C s C, by
Proposition 4.
In the preceding definition of 'convex closure' we can replace 'some' by
'every'. For suppose [x',y) s Ci and x e Ci. We wish to show that also
[x,y) s Ci. We may assume that x,x',y are not collinear, by Proposition 4. Since
x e Ci, there exists x" e C such that x e (x',x"). If z e (x,y), then z e (x",z~ for
some z' e (x',y). Since z' e Ci and x" e C, it follows that z e Ci.

PROPOSmON 6 If C is a convex set, then C is also a convex set.

Proof Obviously we may assume that Ci¢ 0. Suppose x,y e C, where x ¢ y.


We wish to show that if z e (x,y), then also z e C. That is, we wish to show that
c e Ci and we (c,z) imply we Ci. Evidently we may suppose that z e: [c,x] and
z e: [c,y]. Then c,x,y are non-collinear and hence, by Proposition 11.2', we (d,y)
for some de (c,x). Then de Ci, and hence we Ci. D

The definitions of 'intrinsic interior' and 'convex closure' make sense, but
have little significance, in an arbitrary convex geometry and the preceding properties
hold in an arbitrary linear geometry. However, the following results will make
essential use of the axiom (D).

PROPOSITION 7 If C is a convex set, then Ci = Ci and C = C.


Proof Obviously we may assume that C ¢ C, and then Ci¢ 0. Suppose x e Ci.
Then for every ye C \x there exists z e C such that x e (y,z). In particular this
holds for every ye Ci\x and then (y,z) s Ci, by the definition of C. Thus x e Ci
and Ci s Ci.
Suppose, on the other hand, thatx e Ci andy e C \x. Then (x,y) s Ci and,
by (D), there exists some w e (x,y). By the definition of Ci there exists z e C such
thatxe (w,z). Thenxe (y,z),by(L2). Thusxe CiandCis Ci.
This proves the first assertion of the proposition. To prove the second we
need only show that C s C, since the reverse inclusion is trivial. Suppose x e C.
Then (x,y] s Ci for every y e Ci. Since Ci= Ci, by what we have just proved,
this implies x e C. D

PROPOSITION 8 If C is a convex set, then (Ci)i =Ci. Moreover, if Ci¢ 0


then C; = C.
1. Dense linear geometries 89

Proof To prove the first assertion we may assume that Ci contains more than one
point. Suppose x,y e Ci, where x '¢ y. Then there exists z e C such that x e (y,z).
Moreover, by (D), there exists some w e (x,z). Then w e Ci and, by Proposition
III. l(i), x E (y,w]. Hence w '¢ y and x E (y,w). Thus x E Cii and Cii = c;.
Since C; '¢ 0, u e C if and only if (u,v] s: Ci for every v e Ci. Since
cu= Ci, u e C; if and only if the same condition is satisfied. D

PROPOSITION 9 If C is a convex set and c e C;, then C n (c,x) '¢ 0 for each
XE <C>\c.

Proof Evidently we may assume that xi! C. By Proposition 111.7 we have


x e <c 1,c2> for some c 1,c2 e C. Since xi! [c 1,c2], we may assume the notation
chosen so that c 1 e (c 2,x). Since c e Ci, there exists a point c3 e C such that
c e (c2,c3). If c,cltc2 are collinear and c 1 i! (c,x), then c e (c2 ,x) and c3 e (c,x).
Thus we now assume that c,c 1,c2 are not collinear. Then, by (P)', there exists a
point c4 e (cltc3) n (c,x). D

PROPOSITION 10 Let C and D be convex sets such that C s: D.


If C n Di '¢ 0, or if <C> = <D>, then Ci s: Di and C s: D.

Proof It is enough to show that Ci s: Di, since the relation C s: D then follows
from the definition of convex closure. Obviously we may assume that Ci '¢ 0.
Let x e Ci and suppose first that there exists a pointy e C n Di, where y '¢ x.
Since ye C, there exists a point z e C such that x e (y,z). Since z e D and y e Di,
it follows thatx e Di.
Suppose next that <C> = <D> and let ye D \C. Then there exists a point
x' e C n (x,y), by Proposition 9, and there exists a point y' e C such that
x e (x',y'), since x e Ci. Since y' e D and x e (y,y'), it follows that x e Di. D

PROPOSITION 11 For any convex sets C,D with Ci'¢ 0, Di'¢ 0, the following
statements are equivalent:

(i) Ci= Di,


(ii) C = D,
(iii) Ci s: D s: C.

Proof It follows at once from Propositions 7 and 8 that (i) and (ii) are equivalent.
Also it is obvious that Ci= Di implies Ci s: D, and that C = D implies D s: C.
90 V. Density and unendingness

Suppose finally that Ci s D s C. Applying Proposition 10 to the sets D and C


we obtain Di s Ci, and now applying it to the sets Ci and D we obtain Ci s Di.
D

PROPOSITION 12 If C =Ari B, where A and Bare convex sets such that


Ai ri Bi-¢ 0, then

Proof Letze Ai ri Bi and c e C\z. Since z e Ai and c e A, there exists a e A


such that z e (a,c). Similarly, since z e Bi and c e B, there exists be B such that
z e (b,c). By (L3) either a e [b,c] orb e [a,c]. Hence either a e C orb e C. In
any event, z e Ci. Thus Ai ri Bis Ci.
On the other hand, if we Ci and w '¢ z, then there exists c' e C such that
we (z,c'). It follows that we Ai ri Bi. Thus Ci s Ai ri Bi.
This proves the first statement of the proposition, and the second statement is
an immediate consequence. D

We can now derive a converse to Proposition 1.7:

PROPOSITION 13 Let C 1,C2 be convex sets and let F be a face of C 1 ri C 2


such that Fi'¢ 0. Then there exist faces F 1 of C 1 and F 2 of C2, with F 1i '¢ 0,
Fi'¢ 0, such that F = F 1 ri F 2 •
Proof Let x e Fi. Then F is the intersection of all faces of C 1 ri C2 which contain
x, since no face of C 1 ri C 2 containing x is properly contained in F and since the
intersection of two faces is again a face. Since faces form an alignment, there is a
face F k of Ck containing x which is contained in every face of Ck containing x
(k = 1,2). It follows that x e F 1i ri Fi, since any face of F k is also a face of Ck
(k = 1,2). Moreover F s F 1 ri F 2 , since F 1 riF 2 is a face of C 1 ri C 2 which
contains x. But x does not belong to any proper face of F 1 ri F 2 , since
x e (F 1 ri F 2)i by Proposition 12. Since F is a face of F 1 ri F 2 , we must have
F=F 1 riF2• D

We next derive some properties of polytopes:

PROPOSITION 14 If Pis a polytope, then P = P.


Proof We may assume that pi'¢ 0. Let x e pi and assume that there exists a
point y e P \ P. By Proposition IV.12, there exist a,b e P such that
1. Dense linear geometries 91

P n <x,y> = [a,bJ. Then ye= [a,bJ, x e (a,b) and we may choose the notation so
that be (x,y). Hence be pi, which is a contradiction. CJ

PROPOSITION 15 If P =[SJ, where Sis a finite affine independent set, then


pi'¢ 0 and
pi = [SJ\ Uses [S\s].

Proof The formula for pi follows from Proposition IV.11 and the definition of
intrinsic interior. Since S is a Caratheodory set, by Proposition 2, it now follows
that pi'¢ 0. CJ

Proposition 15 has the following important consequence:

PROPOSITION 16 If D is a nonempty finite-dimensional convex set, then


Di'* 0.

Proof Let S be a maximal affine independent subset of D and put C = [SJ. Then
Ci'¢ 0, by Proposition 15. But <C> = <D> and so, by Proposition 10, Ci s Di.
CJ

Thus if Pis a polytope, then pi'¢ 0. However, if Xis infinite-dimensional, it


necessarily contains a nonempty convex set C with Ci= 0. Indeed we can take
C = [S], where S = {x 1,x2, ••• } is any countable affine independent subset of X.
For assume x E Ci. Since c = u;:=lcn, where en= [X1, ••• ,xn], we have x E Cm
for some m. For any ye C \x, there exists z e C such that x e (y,z). Since
y ,z e Cn for some n > m and since Cm is a face of Cn' by Proposition IV .11, it
follows thaty E Cnr Hence c =cm, which is a contradiction.
It follows at once from Proposition 4 that <Ci> = <C> if Ci '¢ 0, and hence
<C> = <C> without exception. From the definitions and from Proposition IIl.12
we also obtain at once

PROPOSITION 17 Let H be a hyperplane such that X\H is not convex. If H+


and H _ are the open half-spaces associated with the hyperplane H, then

H +; =H+ , H_i =H_ ,


H+=H+uH, H_=H_uH. CJ

We can now also prove


92 V. Density and unendingness

PROPOSITION 18 If a and b are distinct points of X, then there exists a


hyperplane H such that a and b lie in different open half-spaces associated
with H.

Proof Choose c e (a,b) and consider the family ~ of all affine sets which contain
c but not a. If H is the union of a maximal totally ordered subfamily of ~, then
He ~. Moreover, if x e X\H then a e <Hux> and hence x e <Hu a>, by
Proposition III.8. Thus <H u a> = X and H is a hyperplane. D

For any a,b e X, put


ab = (a,b) if a'¢ b,
={a}ifa=b.

According to our definition, ab is not a segment if a '¢ b, since a,b ~ ab. However,
we will show that the analogues of the axioms (C) and (P) are satisfied.
We show first that if c e abi and de cb 2, then de ab for some b e bib2. If
a ,bi ,b2 are not collinear this follows from Proposition 11.2 ', and if a,bi ,b2 are not
all distinct it follows readily from (D). Suppose then that a,b 1,b2 are distinct but
collinear. If a e (bi,b2) then either d =a and we can take b =a, or de (a,c) and
we can take any be (c,d), or de (a,b 2) and we can take any b e (d,b 2). If
b 2 e (a,bi) we can take any be (bi,b 2) (') (c,b 1). If bi e (a,b 2) we can take any
b E(bi ,b2) (') (d,b2).
We show next that 1f Ci e abi and c2 e ab 2, then bic 2 (') b 2ci '¢ 0. If
a,b 1,b2 are not collinear this follows from (P)', and if a,b 1,b2 are not all distinct it
follows readily from (D). Suppose then that a,b 1,b2 are distinct but collinear. If
a e (bi,b 2), then a e (b 1,c2) (') (b 2,ci). Suppose b 2 e (a,bi). If ci e (bi,c 2) then
(b 2,ci) 5 (bi ,c2), and if c2 e [ci,b 1) then (b 2,c2) 5 (b 1,c2) (') (b 2,ci). The same
argument applies if b 1 e (a,b 2).
We will refer to these analogues of the axioms (C) and (P) as (C)" and (P)".
It is easily verified that the analogues of Propositions 11.2, 11.17 and 11.19 also
remain valid, i.e.

(i) if c1 e abi, c2 e ab2 and c e cic2, thence ab/or some be bib2;


(ii) if c1 e abi, c2 e ab 2 and be bib2, then there exists a point c e ab(') cic2;
(iii) if x e aa' (') bb' and y e ab, then x e yy' for some y' e a'b'.
1. Dense linear geometries 93

When required, these analogues will be referred to as Propositions 11.2", II.17" and
11.19".
If A and B are nonempty subsets of X, we define their product AB by

AB = U aeA,beB ab.

In other words, AB is the union of A n B with the set of all points c e X such that
c e (a,b) for some distinct points a e A and be B. The reason for the terminology
and notation is that products possess many of the properties of products of positive
numbers, as we now show.
First of all AB'#: 0, on account of (D), and it is obvious that, if A s C and
B s D, then AB s CD. Evidently also AB= BA. Furthermore the associative
law holds: for any nonempty sets A,B,C,

(AB)C = A(BC).

Indeed if x e (AB)C then, for some a e A, be B, c e C, we have x e cy and


ye ab. Hence, by (C)", x e az for some z e be. Thus (AB)C s A(BC) and, for
the same reason,
A(BC) = (CB)A s C(BA) =(AB)C.

For any nonempty set A s X we have A s AA. Moreover it follows at once


from the definitions that A is convex if and only if AA =A.
The proof of the next proposition illustrates how the algebraic properties of
products can replace involved geometric arguments.

PROPOSITION 19 If A and Bare nonempty convex sets, then their product AB


is also a nonempty convex set.

Proof (AB)(AB) = (AA)(BB) =AB. CJ

The following proposition essentially just summarises some earlier results:

PROPOSITION 20 Let C be a nonempty convex set. Then

(i) x e Ci if and only if x e cC for every c e C,


(ii) if x e Ci, then xC = Ci,
(iii) if Ci'#: 0, then Ci = Cic for every c e C.
94 V. Density and unendingness

Proof The statement (i) is just a refonnulation of Proposition 5, and (ii) follows at
once from (i) and Proposition 4.
If Ci:;: 0, then (ii) implies that Cic s Ci for any c e C. On the other hand, if
x e Ci and c e C, then x e cc' for some c' e C. If we choose x' e ex, then x' e Ci
and x e ex'. Hence Ci s: Cic. This proves (iii). CJ

In the statements of (i)-(iii) we may replace C by C (and keep Ci), by


Proposition 7.

PROPOSITION 21 Let A and B be nonempty convex sets. If x e AB,


ye [A u B] and y:;: x, then (x,y) s: AB.
Proof Letze (x,y). We have ye [a",b"] for some a" e A and b" e B. If
y e AB then also z e AB, since AB is convex. Hence we may assume without loss
of generality that ye A and ye B. Since x e ab for some a e A and be B, it
follows that z e xy s: yab and hence z e a'b for some a' e A. CJ

PROPOSITION 22 If A and Bare nonempty convex sets, then

[A u B]i = (AB)i.

Proof We may assume that [A u B] is not a singleton, and then also AB is not a
singleton by Proposition 21.
We show first that [A u B]i s: (AB)i. If x e [A u B]i, y e AB and y :;: x,
then x e (y,z) for some z e [A u B]. If z' e (x,z), then z' e (y,z) and hence
z' e AB by Proposition 21. Since x e (y,z'), it follows that x e AB and actually
x E (AB)i.
We show next that (AB)i s: [Au B]i. Suppose x e (AB);, ye [Au B] and
y:;: x. If y' e (x,y), then y' e AB by Proposition 21. Hence x e (y',z) for some
z e AB. Since AB s: [A u B] and x e (y,z), it follows that x e [A u B]i. D

PROPOSITION 23 If A and Bare convex sets such that Ai:;: 0 and Bi:;: 0,
then
[Au B]i = (AB); = AiBi :;: 0.

Proof By Proposition 22 we need only show that (AB); = AiBi. We divide the
proof into two parts.
1. Dense linear geometries 95

We show first that (AB)i s AiBi. Let x e (AB)i and let ye AiBi with y -:1= x.

Since AiBi s AB, we have x e (y,z) for some z e AB. Hence

We show next that AiBi s (AB)i. Letze AiBi and let c e AB. Then z e xy
for some x e Ai and ye Bi, and c e ab for some a e A and b e B. Hence x e aa'
for some a' e A, and y e bb' for some b' e B. Thus

z e (aa')(bb') = (ab)(a'b').

But abc =ab, since c e ab and ab= (ab)i. Hence z e c(AB)(AB) = c(AB). Thus
z e (AB)i, by Proposition 20(i). D

From Propositions 11 and 23 we immediately obtain

PROPOSITION 24 If A and Bare convex sets such that Ai ":I= 0 and Bi ":I= 0,
then
[AuB] = AB. CJ

From Proposition 23 we also obtain by induction a characterization of the


intrinsic interior of a polytope:

PROPOSITION 25 For any positive integer n and any a 1, ••• ,an e X,

2 UNENDING LINEAR GEOMETRIES

A linear geometry will be said to be unending if it has the following property:

(U) if a,b e X and a ":I= b, then a e [b,c]for some c ":I= a,b.

This axiom also implies the axiom (P), when all the other axioms of a linear
geometry are satisfied. Moreover, as we now show, an unending linear geometry
is 'generally' dense:
96 V. Density and unendingness

PROPOSmON 26 A convex geometry is a linear geometry if it satisfies the


axioms (L2)-(L4) and (U). Furthermore, it is also dense if not all points are
collinear.

Proof By Proposition III.6 and Proposition l, we need only show that the axiom
(D) is satisfied if not all points of X are collinear.
Let y 1,y2 be any two distinct points and choose z e: <y 1,y2>. By (U) we can
now choose z1 so that z e (y2,z 1) and then x so that z1 e (x,y 1). It follows from
Proposition 11.2' that z e (x,y) for some ye (y 1,y2). D

An unending linear geometry need not be dense if all points are collinear; for
example, take X = lR \ (0,1) and define the segment [x,y] to be the intersection with
X of the corresponding segment in JR.
It may be shown also that a convex geometry is a linear geometry if it
satisfies the axioms (L2),(L3),(P) and (U), i.e. these axioms imply the axiom
(L4). For suppose c,d e [a,b]. We wish to show that de [a,c] u [c,b].
Evidently we may assume a '¢ b, c '¢ a,b and d '¢ a,b,c. By (U) we may choose e
so that be (a,e). Since c,d e [a,b], it follows from Proposition II.18(i) that
be [c,e] fl [d,e]. Since b '¢ e, it now follows from (L3) that c e [d,e] or
de [c,e]. If de [c,e] then, since be [d,e], it follows from Proposition II.18(i)
that de [b,c]. Thus we now suppose c e [d,e]. From be [d,e] and de [a,b] we
obtain, by (L2), de [a,e]. From de [a,e] and c e [d,e] we obtain de [a,c], by
Proposition II. l 8(i) again.
Throughout the remainder of this section we assume that a set Xis given
on which a dense, unending linear geometry is defined.
An immediate consequence of this assumption is that, if H is a hyperplane of
X, then X \ H is not convex. Hence Proposition 111.12 and Proposition 17 now
apply to arbitrary hyperplanes. Another consequence of this assum~tion is that the
definition of intrinsic interior may be reformulated in the following way:

PROPOSffiON 27 Suppose C is a convex set and let c e C. Then c e C; if and


only if C fl (c,x) '¢ 0 for every x e <C> \ c.

Proof The necessity of the condition has already been proved in Proposition 9.
Suppose now that C fl (c ,x) '¢ 0 for every x e <C> \ c. For any d e C \ c, choose
x so that c e (d,x). If e e C fl (c,x), thence (d,e) and hence c e Ci. CJ
2. Unending linear geometries 97

A convex set C will be said to be basic if Ci = C and <C> = X. For example,


X itself is a basic convex set and so are the two open half-spaces associated with
any hyperplane. If C and D are basic convex sets with C n D '¢ 0 then, by
Propositions 12 and 9, C n D is also a basic convex set. Consequently a dense
unending linear geometry X can be given the structure of a topological space by
defining a set G s X to be open if G = 0 or if G is a union of basic convex sets.
Moreover, with this topology Xis a Hausdorff space, by Proposition 18. (The
various topological terms used here are defined in Section 4 of Chapter Vill.)
The notion of convex partition was introduced in Chapter II. Some further
properties of convex partitions can be derived under the hypotheses of the present
section:

PROPOSITION 28 Let C,D be a convex partition of X. If Ci'¢ 0, then also


Di'¢ 0 and both C ,Di and D ,Ci are convex partitions of X.

Proof Let x e Ci and ye D. If ye (x,z), then z e D. We will show that actually


z E Di.
Assume on the contrary that z ~ Di. Then there exists u e D such that
z e (u,v) implies v e C. Fix any such v. Evidently u ~ <x,z>. Since x e Ci, there
exists we C such thatx e (v,w). Then ye (t,w) for some t e (z,v). Since we C
and y e D, we must have t e D. Since z e (u,t) implies t e C, this is a
contradiction.
This proves that Di '¢ 0. It also proves that if y e C n D, then y ~ Di and
that if ye D \Di, then (x,y) s C. Thus C s X \Di and X \Di s C, which
shows that C ,Di is a convex partition of X. Moreover, C and D may now be
interchanged. 0

It is possible that under the hypotheses of Proposition 28 we have both Ci = C


and Di = D. For example, take X = Q to be the field of rational numbers, C to be
the set of all x e Q with x > 0 and x2 > 2, and D = X \ C.

For any a,b e X, put

alb = { c e X: a e (b,c)} if a'¢ b,


={a} ifa=b.

Thus c e alb if and only if a e be. More generally, for any nonempty subsets A,B
of X we define the quotient AIB by
98 V. Density and unendingness

AIB = UaeA,beB alb.


In other words, AIB is the union of A ri B with the set of all points c e X such that
a e (b,c) for some distinct points a e A and b e B. It follows from (U) that
A/B ~ 0. Evidently if A s C and B s D, then A!B s CID.
The reason for the terminology and notation is that quotients possess many of
the properties of ordinary fractions, provided equality is replaced by inclusion:

LEMMA 29 If A and B are nonempty subsets of X, then

(i) A s B(AIB),
(ii) A s (AB)IB,
(iii) A s Bl(BIA).

Proof Since these properties are really just reformulations of the definitions, we
only prove (i): if a e A, be B and c e ab, thence AB and a e (AB)/B. D

LEMMA 30 If A,B,C are nonempty subsets ofX, then

(i) (AIB)!C = Al(BC) = (AIC)IB,


(ii) Al(BIC) s (AC)IB,
(iii) A(BIC) s (AB)IC.

Proof (i) Suppose x e (alb)lc for some a e A, be B, c e C. Then there exists


ye X such that ye ex and a e by. Hence, by (C)", there exists z e be such that
a e xz. Thus x e al(bc).
On the other hand suppose x e al(bc) for some a e A, be B, c e C. Then
a e xy for some y e be. Hence, by (C) ", a e bz for some z e ex. Thus
x e (alb)lc.
This proves the first equality in (i) and, since BC= CB, the second follows.
(ii) Suppose x e al(blc) for some a e A, be B, c e C. Then there exists
y e X such that a e xy and b e cy. Hence, by (P) ", there exists z e ac ri bx.
Thus x e (ac)lb.
(iii) Suppose x e a(blc) for some a e A, be B, c e C. Then there exists
ye X such that x e ay and be cy. Hence, by (P)", there exists z e ab ri ex.
Thus x e (ab)lc. D
2. Unending linear geometries 99

LEMMA 31 If A,B,C,D are nonempty subsets of X, then

(i) (AIB)(CID) s: AC/BD,


(ii) (AIB)l(CID) s: AD/BC.

Proof (i) By Lemma 30(iii}, (AIB)(CID) s: ((AIB)C)ID and

(AIB)C = C(AIB) s: AC/B.


Hence, by Lemma 30(i),

(AIB)(CID) s: (ACIB)ID = AC/BD.

(ii) By Lemma 30(ii}, (AIB)l(CID) s: ((AIB)D)IC, and by Lemma 30(iii},

(AIB)D = D(AIB) s: ADIB.


Hence, by Lemma 30(i),

(AIB)l(CID) s: (ADIB)IC =AD/BC. D

PROPOSITION 32 If A and B are nonempty convex sets, then the quotient


AIB is also a nonempty convex set.

Proof Obviously AIB s: (AIB)(AIB). On the other hand, by Lemma 31(i},

(AIB)(AIB) s: (AA)l(BB) = A!B. D

For any nonempty set A s: X we have A s: A/A. It will now be shown that
affine sets are characterized by equality in this relation:

PROPOSITION 33 A nonempty set A is affine if and only if A/A =A.


Proof Suppose first that A is affine. If x e A/A, then there exist a,a' e A such
that a e xa' and hence x e A. Since A s: A/A, this proves that A= A/A.
Suppose next that A/A =A and let y,z be distinct elements of A. If y e (z,u),
then u e y/z s: A. If we (y,z), then ye (w,u) and we y/u s: A. If z e (y,v),
then v e zly s: A. Thus <y,z> s: A, which proves that A is affine. D

PROPOSITION 34 If A is a nonempty convex set, then <A> = A/A.

Proof If x e A/A, then there exist a,a' e A such that a e xa' and hence x e <A>.
100 V. Density and unendingness

On the other hand if x e <A> then, by Proposition III.7, x e <a,a'> for some
a,a' e A. Hence either x e A or a e (x,a~ or a' e (x,a). In every case x e A/A.
a
PROPOSITION 35 If A is a convex set and x e Ai, then <A> = Alx = x/A.

Proof It is obvious that Alx s <A> and x/A s <A>. Moreover x e Alx and
x e x/A. Suppose ye <A> and y :I: x. Then, by Proposition 9, there exists a point
z e A n (x,y). Hence y e zlx s Alx. Furthermore, since x e Ai, we have
xe (z,z~ for some z' e A. Hence x e (z',y) and ye xlz' s x/A. D

PROPOSITION 36 If A is a convex set and x e A, then the following


statements are equivalent:

(i) x E Ai,
(ii) A s x/A,
(iii) x/A is affine,
(iv) Alx is affine.

Proof It follows from Proposition 35 that (i) :::) (ii),(iii),(iv). To complete the
proof we show that (ii),(iii),(iv) :::) (i). We may assume that IAI > 1. Let y e A \x
and choose z,w e X so that x e (y,z), ye (x,w). Then z e x/A and we Alx.
If A s x/A, then ye x/A. Thus x e (a,y) for some a e A, and hence x e Ai.
If x/A is affine then ye x/A, since x e x/A. Hence A s x/A and x e Ai.
If Alx is affine then z e Alx, since x e Alx. Thus there exists a point
a' e A n (x,z). Then x e (a',y), and hence x e Ai. D

PROPOSITION 37 For any nonempty sets A,B, <A u B> = <AIB>.

Proof Since A s B(AIB), by Lemma 30(i) we have

A s A/A s Al(B(AIB)) = (AIB)l(AIB) s <AIB>.

Similarly, since B s Al(AIB), we have

B s BIB s (Al(AIB))IB = (AIB)l(AIB) s <.AIB>.

Since A/B s <A u B>, it follows that <A u B> = <.AIB>. D


2. Unending linear geometries 101

PROPOSITION 38 If A and Bare convex sets such that Ai n Bi'¢ 0, then

<A u B> = AIB = Ai/Bi.

Proof ff z e Ai n Bi then, by Propositions 35 and 20,

<A>l<B> = (Alz)l(Blz) s (zA)l(zB) =Ai/Bi s AIB s <A>l<B>.

Hence <A>l<B> = AIB =Ai/Bi. Consequently

(AIB)l(AIB) = (AIB)l(<A>l<B>) = (AIB)l((zlA)l(zlB))


s (AIB)l((zB)l(zA))
s (AIB)l(BIA)
s (AA)l(BB) = AIB.

Hence AIB is affine, by Proposition 33, and the result now follows from
Proposition 37. D

PROPOSITION 39 If A and B are convex sets such that Ai '¢ 0 and Bi '¢ 0,
then
(AIB)i =Ai/Bi-¢ 0.

Proof We show first that (AIB)i s Ai/Bi. Let x e (AIB)i and let y e Ai/Bi with
y '¢ x. Since Ai/Bi s AIB, we have x e (y,z) for some z e A/B. Hence, by
Lemma 31 and Proposition 20,

x E (Ai/Bi)(AIB) s (AiA)l(BiB) =Ai/Bi.

We show next that Ai/Bi s (AIB)i. If z e Ai/Bi, then z e xly for some
x e Ai, ye Bi. Hence x e Ai n (zBi) =Ai n (zB)i, by Proposition 23. It now
follows from Proposition 38 that

Ai/(zB)i = Al(zB) =<Au (zB)>.

Thus (AIB)lz = Al(zB) is affine and hence, by Proposition 36, z e (AIB)i. D

The next result may be regarded as an affine analogue of Proposition 11.4, but
it requires a supplementary hypothesis.
102 V. Density and unendingness

PROPOSITION 40 If A and B are affine sets such that A r. B :I: 0, then

<AU B> = AIB = UaeA,beB<a,b>.


Proof Since an affine set is its own intrinsic interior, and since it is obvious that

AIB s U aeA,beB <a,b> s <A u B>,

the result follows immediately from Proposition 38. D

The collection of all affine subsets of X, partially ordered by inclusion, is a


lattice, since any two affine sets A and B have an infimum A n B and a supremum
<A u B>. Our next result says that this lattice is weakly modular:

PROPOSITION 41 If A,B,C are affine sets such that Cs Band An B :1: 0,


then
<C u A> n B = <C u (A n B)>.
Proof The right side is certainly contained in the left, since Cu (A n B) is
contained in both B and <C u A>. To show that the left side is contained in the
right let p e A n B and x e <Cu A> n B. By Proposition 40 x e <a,b>, where
a e A and be <p u C>. If x e <p u C>, then x e <Cu (An B)>. If
x E <p u C> then, by Proposition 111.8, a e <x up u C>. Hence a e B and
againx e <Cu (A r.B)>. D

Proposition 41 has the following important consequence:

PROPOSITION 42 Suppose Ai and A 2 are affine sets such that Ai n A 2 -::/= 0.


Then Ai n A 2 has finite codimension in A 2 if and only if Ai has finite
codzmension in <Aiu A 2>, and the two codimensions are then equal.
In particular, if Ai and A 2 are finite-dimensional, then <Aiu A 2> is also
finite-dimensional and

Proof By Proposition 1.22, we need only show that if Ai has finite codimension
min <Aiu A 2>, then A 1 n A 2 has finite codimension n in A 2, where n < m. Let
B be a basis for Ai n A 2 • Then Ai has a basis of the form B u Bi and
<Aiu A 2> has a basis of the form Bu Bi u B 2', where B 2' s A 2 and IB2 1= m.
By Proposition 41, with A,B,C replaced by Ai,A 2,<B2'>, we have
2. Unending linear geometries 103

The hypothesis that X is unending cannot be omitted in Propositions 40-42.


For example, let X be the convex subset of JR..3 consisting of the origin (0,0,0) and
all points (x,y,z) with x,y,z > 0. If we take A to be the plane consisting of all points
(x,y,z) e X with y = x and B to be the plane consisting of all points (x,y,z) e X with
y = 2x, then A n B = { (0,0,0)} and hence

dim (A n B) + dim <A u B> < dim A + dim B.


Furthermore, if we take C to be the line consisting of all points (x,y,z) e B with
z =x, then <A u C> =X and A n B s C, but the point (1,2,2) e B does not lie on
any line <a,c> with a e A and c e C.
As an application of Proposition 42 we prove

PROPOSITION 43 Let H+,H_ be the open half-spaces of X associated with a


hyperplane H and let A be an affine subset of X. If A contains points of both
H + and H_, then L: =A n His a hyperplane of A and A n H +' A n H_ are the
open half-spaces of A associated with this hyperplane.

Proof The hypotheses imply that L ~A,0. Hence Lis a hyperplane of A, by


Proposition 42. Since A n H + is disjoint from L, it is contained in an open half-
space L+ of A associated with the hyperplane L. Let L_ be the other open half-space
of A associated with the hyperplane L. If x e An H+ andy e An H_, then (x,y)
contains a point z e L and hence y e L_. Since A n H+ s L+, A n H_ s L_ and
A is the union of L, A n H + and A n H_, we must actually have A n H + = L+,
A nH_=L_. a

3 CONES

In this section we assume, initially, that a set X is given on which a linear


geometry is defined.
For any z,x e X with z ~ x, we define the ray [z,x> through x from z to be the
set of all points y such that ye [z,x] or x e (z,y). We will sometimes find it
104 V. Density and unendingness

convenient to denote by (z,x> the 'open' ray [z,x> \ z. Evidently if x' e (z,x>, then
[z,x'> = [z,x> and (z,x'> = (z,x>.
A set K s X is said to be a cone with vertex z if x e K implies [z,x> s K.
The vertex need not be uniquely determined; the set of all vertices of a cone will be
called its vertex set.
For example, the empty set is a cone with X as its vertex set, and the singleton
{z} is a cone with vertex z. An arbitrary union of rays from z is a cone with vertex
z. A nonempty set Lis affine if and only if it is a cone with vertex set L.
The definition of a cone does not require it to be convex, but the convex case is
of greater interest. Thus if L is a nonempty affine set and x e X \ L then, by
Proposition 111.12, the closed half-space of <x u L>, which is associated with the
hyperplane Land which contains x, is a convex cone with every point of Las a
vertex.
It follows at once from the definition that if z is a vertex of a nonempty cone K,
then z e K. Moreover, if z and z' are vertices of K, then <z,z'> s K. In fact, as
we now show, every point of <z,z'> is a vertex.

PROPOSITION 44 The vertex set of a cone is an affine set.

Proof Let K be a cone and let u, v be vertices of K. We wish to show that if


we <u,v> and x e K, then ye [w,x] or x e (w,y) implies ye K.
Suppose first that we [u,v]. If ye [w,x], then there exists a pointy' e [v,x]
such that y e [u,y1. Moreover y' e K, since v is a vertex, and hence y e K, since u
is a vertex. On the other hand if x e (w,y), then there exists a point x' e [v,y] such
that x e [u,x1. Moreover x' e K, since u is a vertex, and hence y e K, since v is a
vertex.
Suppose next that u e [v,w]. If ye [w,x], then there exists a point
y' e [u,x]" [v,y]. Moreover y' e K, since u is a vertex, and hence ye K, since v
is a vertex. Similarly if x e (w,y), then there exists a point x' e [v,x] " [u,y].
Moreover x' e K, since v is a vertex, and hence y e K, since u is a vertex.
Since u and v can be interchanged, this completes the proof. D

For convex cones we can say more:

PROPOSITION 45 The vertex set of a convex cone K is a face of K.


3. Cones 105

Proof Let z be a vertex of K and suppose z e (z',z"), where z',z" e K. Then


<z',z"> s K, since z is a vertex. Let x e K \ <z',z">. If ye [z',x] then y e K,
since K is convex. On the other hand if x e (z',y), there exists a point
x' e (z,y) n (z",x). Then x' e K, since K is convex, and hence ye K, since z is a
vertex. Consequently z' is a vertex, and z" likewise. The result now follows from
Proposition 11.6. D

It follows that if K is a nonempty convex cone and L the affine set of all
vertices of K, then K\L is convex.
Let 'XL denote the collection of all convex cones whose vertex set contains L,
where L is a nonempty affine set. It is easily seen that 'XL is a normed alignment on
X. We define the L-conic hull [L,S> of any set S s X to be the intersection of all
convex cones in 'XL which contain S.
Evidently [L,S> e 'Xv Furthermore our notations are consistent, since for
L = {z} and S = {x} the L-conic hull of Sis the ray [z,x>. From the definition of
L-conic hull we immediately obtain the following properties:

[S] s [L,S> s <S u L>,


[L,S> = [L,[S]>,
[L',S> s [L,S>,

where L' is any nonempty affine subset of L.


When L is a singleton, L-conic hulls have a very simple structure:

PROPOSITION 46 If C is a convex set and z e X, then

[z,C> = U xeC [z,x>.


Proof Obviously the right side is contained in the left. On the other hand, the right
side contains C and is a cone with vertex z. Thus it merely remains to show that it
is convex.
Suppose y' e [z,x'> and y" e [z,x">, where x',x" e C. We need only show
that if y e (y',y"), then y e [z,x> for some x e [x',x"]. If y' e [x',z] and
y" e [x",z], then ye [x,z] for some x e [x',x"]. If x' e (y',z) and y" e [x",z],
then ye [z,z1 for some z' e [x",y1 and there exists a point x e [x',x"] n [z,z1.
Hence either y e [x,z] or x e (y,z). Finally, if x' e (y',z) and x" e (y",z), then
there exist a point z' e [y,z] n [x',y"] and also a point x e [x',x"] n [z,z1. Hence
x e (y,z). D
106 V. Density and unendingness

In order to progress further we assume, throughout the remainder of this


section, that the linear geometry X is both dense and unending.

PROPOSITION 47 Let L be a nonempty affine set. If K is a cone whose


vertex set contains L, then

K = LK = KIL = LKIL.

Conversely, if K = LKIL and L 5 K, then K is a cone whose vertex set


contains L.

Proof It follows directly from the definition of a cone and (U) that, if K is a cone
whose vertex set contains L, then K = LK and K =KIL. These two relations
obviously imply K =LKIL. Suppose on the other hand that K = LKIL and L s K.
Letze Land x e K\z, so thatx e y'lz'for some y' e LK and z' e L. We wish to
~how that [z,x> s K.

If x e (y,z), then y' e xz' =


yzz' and hence y' e yz" for some z" e zz'. Thus
z" e L, and ye LKIL = K. If ye (x,z) then, by (P)", there exists a point
w e yz' n y'z. Thus we (LK)L = LK and ye LKIL = K. Furthermore z e K, by
hypothesis. D

PROPOSITION 48 For any nonempty convex set C and any nonempty affine
set L,
[L,C> = [L u C]/L = LCIL u L.

Proof Put K = [Lu C]/L. Then K is a convex set containing L. From


C s L(CIL) s LCIL we obtain C s Kand

CIL s (LCIL)IL = LCIL.

Since [Lu C] =LC u Lu C, by Proposition 21, it follows that K = LCIL u L.


Hence
LKIL = (L(LCIL))IL u L s: (LCIL)IL u L = K.

On the other hand, since L s: LKIL and LC!L s: LKIL, we also have K s: LKIL.
Consequently K = LKIL and K is a cone whose vertex set contains L, by
Proposition 47. Furthermore, if K' is any convex cone containing C whose vertex
set contains L, then
K = LCIL u L s LK'IL = K'.
3. Cones 107

Consequently K = [L,C>. D

We consider next the structure of the intrinsic interior and convex closure of a
convex cone. If some vertex of a convex cone K is contained in its intrinsic interior
Ki then, by Proposition 45, every point of K is a vertex. Hence K is affine and
K =Ki = K. We now treat the general case:

PROPOSITION 49 If K is a convex cone whose vertex set contains the


nonempty affine set L, then Ki u L and K are also convex cones whose vertex
sets contain L.

Proof Obviously we may assume that Ki~ 0. Then H: =Ki u L is convex, since
Ki and L are convex and L s K. To show that H is a cone whose vertex set
contains L, we need only show that if z e L, x e Ki\L and x e (y,z), then ye Ki.
We certainly have ye K. Lety' e K\y and choose y" e X so thaty e y'y". Then
x e y'y"z and hence x e y'z' for some z' e y"z. Since x e Ki, it follows that
x e y'x' for some x' e xz' n K. Since x e yz and z' e y"z, it follows that x' e x"z
for some x" e yy". Moreover x" e K, since x' e Kand z is a vertex of K. Since
y e y'x", it follows that y e Ki.
To prove that K is also a convex cone whose vertex set contains L, we need
only show that K contains the ray [z,y> for any ye K \Kand z e L. Choose any
y' e [z,y>, x e Ki and x' e xy'. Suppose first that y' e (z,y). Then, by (C) ",
x' e zz' for some z' e xy. Consequently z' e Ki and hence x' e Ki. This proves
that y' e K. Suppose next that ye (z,y'). Then, by (P)", there exists a point
z" e zx' n xy. Consequently z" e Ki and hence x' e Ki. by the first part of the
proof. Thus again y' e K. D

PROPOSITION 50 Let L be a nonempty affine set and K a convex cone whose


vertex set is L. Then Kt= L/K is also a convex cone whose vertex set is L.
Moreover Kn Kt = L, (Kt) t = K, and K1 = z/K for any z e L.

Proof We may assume K ~L. Letze Land put Kz = z/K. Thus Kz is the union
of {z} with the set of all points x e X such that z e (x ,y) for some y e K.
Evidently Kz is a cone with vertex z, and Kz ~ L since K ~ L. Moreover Kz is
convex, by Proposition 11.19", and Ls: Kz since Ls: K.
108 V. Density and unendingness

In fact L =Kn Kz. For if z' e Kn Kz and z' -:I: z, then z e (z',z") for some
z" e Kand hence z' e L, since Lis a face of K by Proposition 45. Since it is clear
from the definition that (~)z = K, it follows that the vertex set of~ is ~ n K = L.
It only remains to show that Kz = LI K. In fact, since it is obvious that
Kz s L/K, we need only show that if x' e L/K, then x' e Kz. There exist y' e K
and z' e L such that z' e xy'. If we choose x" so that z e xx", then there exists a
pointy e zy' n z'x" by (P)". It follows that ye K, since K is convex, and hence
thatx" e K, since z' e L. Consequently x' e Kz. D

We will call Kt the convex cone symmetric to the convex cone K.


It is easily seen that, for any hyperplane H, the two closed half-spaces
associated with H are symmetric convex cones with H as their common vertex set.
Furthermore, if H+ and H_ are the two open half-spaces associated with H then, for
any nonempty affine set L s H, H+ u L and H_ u L are symmetric convex cones
with L as their common vertex set. This provides an example of a convex cone
with a different vertex set from that of its convex closure.

PROPOSITION 51 Let L be a nonempty affine set and let K,Kt be symmetric


convex cones with the common vertex set L. Then

(i) Ki -:I: 0 if and only if (Kt); -:I: 0,


(ii) Ki u L and (Kt); u L are symmetric convex cones with the common vertex
set L,
(iii) K and Kt are also symmetric convex cones.

Proof We show first that the vertex set of Ki u Lis exactly L. By Proposition 49
the vertex set of Ki u L contains L. Assume that Ki u L has a vertex w e Ki \L.
Then for some x e K the ray [w,x> is not wholly contained in K. However, the
line <x,w> contains points w',w" e Ki such that we (w',w"). Moreover w',w" are
vertices of Ki u L, since the set of all vertices is a face of Ki u L. Hence
<x, w> = <w', w"> s Ki u L, which is a contradiction.
If Ki -:1: 0 then, by Proposition 39, (Kt); = LI Ki -:1: 0. Since the relation
between Kand Kt is symmetric, this proves (i). Furthermore, since LIL= L,

This completes the proof of (ii).


3. Cones 100

To prove (iii) we may assume Ki'¢ 0. Fix z e L. Then, by Proposition 39,


we also have (Kt);= z/Ki. Suppose y' e Kt and x' e (Kt); with x' '¢ y'. Then
z e yy' for some ye K and z e xx' for some x' e Ki. If w' e x'y' then z e ww'
for some we xy, by Proposition 11.19". Thus we Ki and hence w' e (Kt);. This
proves that y' e Kt .
- -
Thus Kt =Kt. Replacing K by Kt, we obtain Kt t =
K. Since K 1 s; K 2
implies z/K1 s; z/K2 , it follows that also Kt s; Kt. D

4 POLYHEDRA

Throughout this section we will assume that a set X is given on which a


dense unending linear geometry is defined. A set P s; X is said to be a
polyhedron if it is affine or if it is the intersection of finitely many closed half-
spaces of its affine hull <I'>.
Examples of polyhedra in R2 are a strip bounded by two non-intersecting lines
and a sector bounded by two intersecting lines.
Obviously any polyhedron is convex and, by Proposition IV.19, any polytope
is a polyhedron. It is easily seen that a one-dimensional polyhedron is either a line
<a,b> or a ray [a,b> or a segment [a,b].
If Pis an affine set then, by Proposition 11.6, its only faces are P and 0. We
are going to show that the facial structure of polyhedra which are not affine
resembles that of polytopes.
If P is a polyhedron which is not affine, then there exist hyperplanes
H 1, ••• ,Hm of <P> with associated open half-spaces Hi+, Hi- (j = l,... ,m) such
that

(*)

The representation (*) will be said to be redundant if some term Hi+ can be
omitted and irredundant otherwise. Evidently if the representation (*) is
redundant, then an irredundant representation can be obtained from it by omitting
certain terms.
110 V. Density and unendingness

PROPOSITION 52 If a nonempty polyhedron P has the irredundant


n
representation (*), then P =P and pi = j:1 Hi+ -¢ 0. Furthermore,

(i) the distinct facets of P are Fi= P n Hi (j = 1,... ,m),


(ii) any nonempty proper face of P is a polyhedron, is contained in a facet of
P, and is the intersection of those facets of P which contain it,
(iii) a face of P is affine if and only if no nonempty face of P is properly
contained in it.

Proof Put
Mk = nb~k Hi+.

Then P =Mk n Hk + and Fk = P n Hk =Mk n Hk for each k e { l, ... ,m}. Since


P is not contained in Hkt Mk is not contained in Hk- . On the other hand, since the
representation (*) is irredundant, Mk is not contained in Hk+ . Thus Mk contains a
point of Hk+ and a point of Hk- , and hence also a point of Hk· ·
Thus Fk '¢ 0. We are going to show that actually <.Fi>= Hk. Assume on the
contrary that there exists a pointx e Hk such thatx E <.Fk>. By Proposition 111.7
we have x e <.x',x"> for some x',x" e P. Then x ¢X',x" and x',x" E Hk. Thus
x',x" e Mk n Hk +. Since x E [x',x"], we may choose the notation so that
x" e (x,x'). Picky e Mk n Hk - . Then (x',y) contains a pointy' e Mk n H k·
Since y' '¢ x, the points x,x',y are not collinear and there exists a point
y" e (x",y) n (x,y'). Then y" e Mk n Hk· Since x e <y',y"> and y',y" e Fk' this
is a contradiction.
Thus P n <.F k> = F k· Since also P \F k =Mk n Hk +, it follows from
Proposition IV .1 that F k is a face of P. In fact F k is a facet of P, since Hk is a
hyperplane of <P>. Moreover Fk is itself a polyhedron, since

Fk = nj~k Hi+ n Hk

and either H k s Hi+ or Hi+ n H k is a closed half-space of H k = <.F k>, by


Proposition 43.
We now show that any proper face F of P is contained in some facet Fk· It is
enough to show that F s H k and in fact, by Proposition 3, it is enough to show
that F s U j:1Hi. Assume on the contrary that there exists a point x e F such that
XE nj:1Hj+· Ifye P\xthenxe (y,z)forsomeze <P>,by(U). Moreover,
4. Polyhedra 111

sincey e nj:1 Hi+, we can choose z e nj:1 Hi+, by (D). Then z e P. Since F
is a face of P, it follows that y e F. Thus F =P, which is a contradiction.
This shows, in particular, that F 1, ••• ,Fm are the only facets of P. Since the
representation(*) is irredundant, they are certainly distinct.
It is obvious that nJ'= 1 Hi+ s Pi. Since no element of pi is contained in a
facet Fi, we must actually have pi = n j:1Hi+. Moreover Pi -¢ 0 since, by
Proposition 3, P is not contained in UJ'=1H j· If y e P and x e pi, then
(x,y) s Hi+ (j = 1,... ,m) and hence ye P. Thus P = P.
Let F be any nonempty proper face of P. As we have seen, F s F k for some
k. If F = Fk' then F is a polyhedron. If F '¢ Fk' then Fk is not affine. By what we
have already proved, the facets of Fk have the form

n · k.ff.+ nHknH· = FknF·


J-:1: ,I J I I

for some i '¢ k. Since F is a proper face of F k' it follows that F s F k n Fi for
some i '¢ k such that Fk n Fi is a facet of Fk. Moreover F = Fk n F; if and only if
F is a facet of F k and in this case F is a polyhedron. If F '¢ F k n F;, then Fk n F;
is not affine and the argument can be repeated. Since the process must eventually
terminate, this yields (ii) and (iii). Cl

COROLLARY 53 A polyhedron has only finitely many faces, and a nonempty


polyhedron has at least one nonempty affine face. Cl

COROLLARY 54 If F and G are faces of a polyhedron P such that F c G, then


there exists a finite sequence F 0 ,F 1, ••. ,F, of faces of P, with F 0 =F and
F, = G, such that F;_1 is a facet of Fi (i = 1,... ,r). Cl

It may be noted that, in Corollary 54, the length of the sequence is the
codimension of <F'> in <G> and consequently does not depend on the choice of
sequence.

COROLLARY 55 If F 1 is a facet of a polyhedron P and Fa nonempty facet of


F 1, then F is contained in exactly one other facet F 2 of P and F = F 1 n F 2•

Proof This follows at once from Proposition 52 and Proposition IV.7. D

PROPOSITION 56 If F,G,H are faces of a polyhedron P such that F s G s H,


then there exists a face G' of P such that F s G' s H, F = G n G', and every
face of P which contains both G and G' also contains H.
112 V. Density and unendingness

Proof The proof of the corresponding Proposition IV.18 for polytopes remains
valid if we replace differences of dimension by codimension. a
PROPOSITION 57 If P is a polyhedron and A an affine set, then P n A is a
polyhedron.

Proof Since P r'\ A =P n A r'\ <.P r'\ A>, we may assume that A = <P n A>.
Evidently we may assume also that A'¢ <P>, that P n A'¢ 0,A and that Pis not
affine. Let P have the irredundant representation (*). Then A is not contained in
Hi+ for at least one j, and we may choose the notation so that A s Hi+ if and
only if j > n. Then
Pr'\ A = nj= 1 Hi+ r'\ A

and, by Proposition 43, Hi+ r'\ A is a closed half-space of A associated with the
hyperplane Hin A for each j e { l, ... ,n}. Since A= <P n A>, it follows that
P r'\ A is a polyhedron. D

PROPOSITION 58 The intersection of two polyhedra is again a polyhedron.

Proof By Proposition 57 it is sufficient to show that if Pis a polyhedron which is


not affine, and if L+ is a closed half-space of an affine set A associated with a
hyperplane L of A, then P r'\ L+ is a polyhedron. Obviously we may assume that
A '¢ <.I'> and P r'\ L+ '¢ 0 ,P. Since P r'\ A is a polyhedron and P r'\ L+ =
P n A n L+ , we may further assume that P s A. Then H
- -
= <.I'> r'\ L is a
hyperplane of <.I'> and H+ = <P> r'\ L+ is a closed half-space of <.I'> associated
- --
with this hyperplane. Hence P r'\ L+ = P n H+ is a polyhedron. D

Finally we will detennine under what conditions a polyhedron is a polytope or


a cone.

LEMMA 59 If Pis a polyhedron, x e pi and ye <P> \P, then (x,y) contains a


point of P \Pi.

Proof Since P is not affine, it has an irredundant representation (*). Then


ye Hk- for some k. Since x e Hk+, there exists a pointy" e (x,y) n Hk. If
y" f! P this argument can be repeated with y" in place of y. Since the process must
terminate after at most m steps, (x,y) contains a pointy' e Pr'\ Hi for some j.
Since y' belongs to a facet of P, y' f! Pi. C
4. Polyhedra 113

PROPOSITION 60 A po/yhedro11 co11tai11i11g more tha11 011e poi11t is a polytope


if a11d 011/y if it does 11ot co11tai11 a11 e11tire ray of its affi11e hull.
Proof We need only show that if P is a polyhedron containing more than one
point, but not containing any entire ray, then Pis a polytope. Evidently Pis not
affine nor a closed half-space of its affine hull. Since any minimal nonempty face
of P is affine, it must be a singleton and hence an extreme point of P. It follows
that P has finite dimension d > 1. If d = l, then P is a segment and hence a
polytope. We assume that d > 1 and the result holds for polyhedra of dimension
less than d. Then any proper face of P is a polytope.
Letx e pi andy e <I'> \P. By Lemma 59, (x,y) contains a pointy' e P\Pi.
Since P does not contain a ray there exists also a point z e <I'> \ P such that
x e (y,z), and in the same way (x,z) contains a point z' e P \Pi. Since x e (y',z')
and y',z' are contained in one or other of the finitely many proper faces of P, it
follows from the induction hypothesis that p is a polytope. a
PROPOSITION 61 A 11011empty polyftedro11 P is a co11e if and only if it has a
unique no11empty affi11e face V. Moreover Vis the11 the vertex set of P.

Proof Let P be a nonempty polyhedron which is not affine. By Proposition 52,


the nonempty affine faces of P are the minimal nonempty faces.
Suppose first that P is a cone. Then the set V of all vertices of P is an affine
face of P, by Propositions 44 and 45. If v e V and x e P \ V, then [v,x> s P and
hence x e (v,x') for some x' e P. Consequently any face of P containing x must
contain V and thus cannot be affine.
Suppose next that P has a unique nonempty affine face V. Let d be the
codimension of <V> in <I'>. If d = l, then Pis a closed half-space associated with
the hyperplane <V> of <I'>. Thus Pis a cone with vertex set V. We now suppose
d > 1 and use induction on d. Assume that, for some v e V and some x e P \ v, the
ray [v ,x> is not entirely contained in P. Then x E V and x e (v ,y) for some
y e <I'>\ P. Hence, by Lemma 59, [x,y) contains a point x' e P \Pi. Thus x'
belongs to a proper face F of P. Hence Vs F and x e F. But then, by the
induction hypothesis, Fis a cone and [v,x> s F, which is a contradiction. a
114 V. Density and unendingness

S NOTES

The concepts of intrinsic interior and convex closure, for convex sets in a real
vector space, are discussed (among other things) by Bair and Fourneau (1975/80).
However, the definition of convex closure used here is slightly different from theirs
and has the advantage that in Proposition 7, for example, it need not be assumed
that Ci -:1: 0.
The properties of products and quotients in 'join spaces' are extensively treated
by Prenowitz (1961) and Prenowitz and Jantosciak (1979). They not only replace
geometry by algebra, but also obviate special consideration of degenerate cases.
Proposition 26 is already contained in Veblen (1904). Proposition 41 plays a
role in the lattice-theoretic characterization of affine geometry, which was initiated
by Menger (1936); see, for example, J6nsson (1959), where 'weakly modular' is
called 'special', and Maeda and Maeda (1970). Proposition 42, for the finite-
dimensional case, is already contained in Grassmann (1844), §126. Cones are
discussed for more general situations than Euclidean space by Prenowitz and
Jantosciak (1979) and Rubinstein (1964).
The idea of basing the theory of polyhedra on Proposition 3 is taken from
Griinbaum (1967). The usual analytic treatment replaces the theory of polyhedra by
the theory of linear inequalities; see, for example, Stoer and Witzgall (1970),
Tschemikow (1971) and Schrijver (1989). It should be mentioned, however, that
two significant results of this theory do not carry over to the more general situation
considered here. This is illustrated by the following simple example.
Let X be the open unit disc of R2, with the usual definition of segments, and
let P be the set of all points of X in the (closed) first quadrant. Then P is a
polyhedron with two facets. However, P is not the convex hull of its facets,
although it is neither affine nor a closed half-space. Furthermore, P is not the
convex hull of finitely many points and rays, although it is a polyhedral cone with
the origin as vertex.
VI
Desargues

There are many examples of the dense unending linear geometries studied in the
previous chapter, besides vector spaces over an ordered division ring. We give
several such examples here, including the projective plane over the field of rational
numbers. This example is not atypical, since in the next chapter we will show that
any dense linear geometry of dimension greater than 2 is isomorphic to a subset of a
projective space over an ordered division ring. In the present chapter we lay the
groundwork for this result by giving a brief introduction to projective geometry and
by showing that any dense linear geometry of dimension greater than 2 has the
Desargues property. Many dense two-dimensional linear geometries do not have
the Desargues property, but we establish several properties of those which do. We
show also that a factor geometry of any dense linear geometry is a dense linear
geometry with the Desargues property.

1 INTRODUCTION

The results which have already been established show that dense, unending
linear geometries closely resemble vector spaces over an ordered division ring,
which are indeed a special case. Nevertheless, there are several other examples:

EXAMPLE 1 Let X be any dense totally ordered set which has no least or greatest
element. For any x,y e X define the segment [x,y] to be the set of all z e X such
that x < z Sy or y < z S x, according as x Sy or y < x.
For example, we may take X to be the set of all ordered pairs (~ 1 ,~ 2) of real
numbers with the total ordering defined by (~ 1 ,~ 2) -< (11 1,11 2) if ~ 1 < 11 1, or if
~1 = 111 and ~2<112·
116 VI. Desargues

EXAMPLE 2 Let X be the open hemisphere consisting of all points


x = (l;o,l;lt···,l;n) e Rn+l with l;o 2 + l; 12 + ... + l;n 2 = 1 and l;o > 0. For any
x,y e X define the segment [x,y] to be {x} if x = y and otherwise to be the arc
joining x and y in X of the circle through x and y with centre at the origin.
It is easily seen that the spherical geometry thus defined is a dense unending
linear geometry. Indeed the only axioms which are not obviously satisfied are (C)
and (P). But a segment in Xis the projection from the origin onto the unit sphere
of the usual segment in the enveloping space Rn+l, and so (C) follows from the
corresponding property in Rn+ 1. Since X is obviously dense, (P) follows from
Proposition V .1.

EXAMPLE 3 Let X be the 'upper' half-space con~isting of all points


x = (l; 0,l; 1,... ,l;n) e Rn+l with l; 0 > 0, and let H be the bounding hyperplane
consisting of all points x = (l;0,l; 1,... ,l;n) with l;0 = 0. For any x,y e X with x :F: y
define the segment [x,y] to be the arc joining x and yin X of the circle through x
and y which is orthogonal to H (or the segment joining x and y of the line through
them, if this line is orthogonal to H).
We omit a detailed proof that the hyperbolic geometry thus defined is a dense
unending linear geometry.

EXAMPLE 4 Another example is the refracted plane, constructed by Moulton


(1902). LetX =R2 and define the segment [x,y] to be the ordinary segment joining
x and y, except in the case where the slope of the latter segment is positive and the

, ,•
,,
,,
, ,,' y
,,
0

Figure 1: Mou/ton's refracted plane


1. Introduction 117

second coordinates of x and y have opposite signs. In the excepted case, if


x = (~1,~2) n
and y = (Th,''12), where ~ 1 <Tl 1 and ~ 2 < 0 < 2, we define the
segment [x,y] = [y,x] to be the union of the segments [x,z] and [z,y], where
z = (~1,0) with ~1 = (T12~1-µ~2T1 1 )1(T12-µ~2), for some fixed real number
µ e (0,1). In other words, z is chosen on the 'horizontal' axis so that the slope of
the line <z,y> isµ times the slope of the line <X,z>, as shown in Figure 1. It is
readily verified that, with this definition, X is a dense unending linear geometry.

EXAMPLE 5 The preceding example is a special case of a vastly more general


construction due to Busemann (1955). Define a metric on R2 by putting (say)

Let Cf1 be a family of subsets of R2 such that

(i) for each subset !J e <§, there is a continuous injective map t ~ g(t) of R onto
e R}, and d(g(t),g(O)) ~ oo as t ~ +oo;
9, so that !J = {g(t): t
(ii) for any two distinct points a,b e R2, there is exactly one subset !Jab e <§which
contains both a and b.

The hypothesis (i) says that each !J e <§ is homeomorphic to R and is a closed
subset of R2. By adjoining to R2 a point at infinity and applying the Jordan curve
theorem (see, for example, Wall (1972)) to the resulting sphere s2, we see that
R2 \ !J has exactly two connected components and !J is their common boundary. We
will call these two connected components the 'sides' of 9.
If a= gab(a) and b = gab(p), we define the segment [a,b] to be the set of all
points c e !Jab with c =gab(t) for some t such that a St SP if a< p, or such that
p St Sa if p <a. We also put [a,a] ={a} for any a e R2. It may be
immediately verified that, with segments so defined, the axioms (Ll)-(L4) are all
satisfied. Thus we are free to use the results of Chapter 111, Section l, and !Jab is
what is there denoted by <a,b>. This notation will now also be used here.
We now show that the axiom (C) is also satisfied. Without loss of generality,
suppose c e (a,b 1) and de (c,b 2). By Proposition 111.6 we may assume that
b2 E <a,b 1>. Then b1 and b 2 are on opposite sides of <a,d>, since c and b2 are on
opposite sides of <a,d> but c and b 1 are on the same side. Thus there exists a point
b E (b1,b~ ~ <a,d>.
118 VI. Desargues

Assume first that be (a,d). Then c and dare on opposite sides of <b1,b2>,
since a and d are on opposite sides of <b 1,b2> but a and c are on the same side.
Since de (c,bi), this is a contradiction.
Assume next that a e (d,b). Then band b 1 are on opposite sides of <c,d>,
since a and b 1 are on opposite sides of <c,d> but a and b are on the same side.
Thus there exists a point b' e (b 1,b) n <c,d>. Since b' e (b 1,b2) and b2 e <c,d>,
this is a contradiction.
We conclude that de (a,b). Thus R2 is a convex geometry with the present
definition of segments. Since the axiom (U) is evidently satisfied, it follows from
Proposition V.26 that the axioms (P) and (D) are also satisfied.
As a specific example, consider the family CS consisting of all subsets

and all subsets

where a,(3;y e R. Each subset 9 aP is a translate of 9 = 9 00 and each subset /t"f is a


translate of It= .ftO. It follows that the family CB satisfies the condition (i) above.
We now show that CS also satisfies the condition (ii).
Let x = (~ 1 ,~ 2 ) and y = (11 1,11 2) be arbitrary distinct points of R 2. If
~ 1 = 11 1 =~'then x and y are both contained in .ftS, but are not both contained in
any other subset in CS. If ~ 1 '¢11 1 , then x and y are not both contained in any subset
/t'Y. We wish to show that they are both contained in a unique subset 9 aP e CS. We
require

which implies

Since ~ 1 '¢11 1 , this is a cubic equation for a with at least one real root. In fact it has
a unique real root, since the polynomial in a on the right side has a nonvanishing
derivative. Having determined a, we immediately obtain (3 = ~ 2 - (~ 1 -a)4 =
112 - (111 - a) 4 •
As another example, let 9 = {x = (~ 1 ,~ 2 ) e R2: ~ 2 > 0, ~ 1 = - ~ 2- 1 } be the
branch of the hyperbola ~ 1 ~ 2 = - 1 in the upper half-plane. Given any positive
numbers µ,p, there exist uniquely determined points x = (~ 1 ,~ 2), y = (11 1,112) on 9,
1. Introduction 119

with ~ 1 <11 1 , such that the straight line segment joining x and y has slopeµ and
length p. In fact, if we put a= p/2(1 + µ2)112, then

~1 = - a - (a2 + µ-1)112, 111 =a_ (a2 + µ-1)112.


Let

where a,p e R, be an arbitrary translate of II· Then it follows that, given any two
distinct pointsx =(~ 1 ,~ 2 ) andy =(11 1,11 2) ofR2 such that the straight line segment
joining x and y has slope µ > 0 and length p > 0, there is exactly one translate !lap
which contains both x and y; in fact we must take

a= (a2 + µ-1)112 + (~ 1 + 111 )/2, p = _ µ(a2 + µ-1)112+(~ 2 + 112 )/2.


Hence the family '§ consisting of all subsets !lap of R 2, together with all straight
lines in R2 which do not have positive slope,

where y,B,µ e R and µ < 0, satisfies the conditions (i) and (ii).

Further interesting examples of dense, unending linear geometries can be


given, but they require an acquaintance with projective geometry. For this reason,
and also for future reference, we now give a brief introduction to this subject.

2 PROJECTIVE GEOMETRY

A projective space P (in the terminology of some authors, an irreducible


projective space) is a set of elements, called 'points', and a nonempty collection of
nonempty proper subsets, called 'lines', such that

(i) any two distinct 'points' a,b are contained in exactly one 'line', denoted here
by <a,b>,
(ii) every 'line' contains at least three distinct 'points',
(iii) if <a,b> and <c,d> are distinct 'lines' with a common 'point' e '* a,b,c,d, then
<a,d> and <b,c> are also distinct 'lines' with a common 'point' f '* a,b,c,d (see
Figure 2).
120 VI. Desargues

Two 'lines' will be said to intersect if they have a common 'point' and 'points'
will be said to be collinear if they are all contained in a common 'line'. (Inverted
commas are used for several concepts in projective space to distinguish them from
the corresponding concepts in a vector space.) A projective plane is a projective
space in which any two 'lines' intersect.
For example, with any vector space V over a division ring D there is associated
a projective space P(V) whose 'points' are the one-dimensional vector subspaces of
V and whose 'lines' are the two-dimensional vector subspaces of V. If Wis a
vector subspace of V, then P(W) is said to be a 'subspace' of P(V), and if Wis a
maximal proper vector subspace of V, then P(W) is said to be a 'hyperplane' of
P(V).

Figure 2

Let V be an arbitrary vector space over a division ring D and let V* = D ~ V


be the vector space of all couples (A.,v), where A. e D and v e V, with the obvious
definitions of addition and multiplication by a scalar:

We can regard Vas a subset of the projective subspace P(V*) by identifying the
vector v e V with the one-dimensional subspace of V* containing the couple (l,v).
The remaining 'points' of P(V*), i.e. the one-dimensional subspaces of V*
containing a couple (O,v) for some v e V\ {0}, form a 'hyperplane' of P(V*) - the
2. Projective geometry 121

'hyperplane at infinity'. Thus P(V*) is the union of V and the 'hyperplane at


infinity'. The projective space P(V*) is called the projective completion of V and
will be denoted here by V.
More concretely, if Dis a division ring and n a positive integer, the 'points' of
the n-dimensional projective space pn = pn(D) may be represented by
(n + 1)-tuples of elements of D, not all zero, with two (n + 1)-tuples (~ 0,~ 1 , ••• ,~n)
and (~ 0 ',~ 1 ', ••• ,~n ') representing the same 'point' if ~;' = A.~; (i = 0,1, ... ,n) for
some A. e D. If x and y are distinct 'points', represented by the (n + 1)-tuples
(~o,~1'···,~n) and (110,11 1, ••• ,11n), then the 'points' of the 'line' <.x,y> are
represented by all (n + 1)-tuples (~ 0 ,~ 1 , .•• ,~n), where ~; = A.~; + µ 11;
(i = 0,1, ... ,n) for some A.,µ e D, not both zero. The projective space pn is the
projective completion of the usual vector space of all n-tuples (~ 1 , ••• ,~n) of
elements of D, the 'hyperplane at infinity' being the set of all 'points' of pn which
are represented by (n + 1)-tuples (~0 ,~ 1 , ••• ,~n) with ~o = 0.
Two distinct 'lines' of a projective space P(V) are 'coplanar' if they are
contained in a three-dimensional subspace of V. It follows at once from (i),(iii) that
two distinct 'coplanar lines' of P(V) intersect in a unique 'point'. An important
attribute of the projective space P(V) is Desargues' theorem:

If <a 1,bi>,<a 2 ,b2>,<a 3 ,b3> are three distinct 'lines' of P(V) with a
common 'point' p :I! a;,b; (i = 1,2,3), then the 'points' of intersection

are collinear.

We sketch the algebraic proof. Suppose the 'points' p,a;,b; of P(V) are
represented by the vectors x,cx;,P; (i = 1,2,3) of V. Since each is determined only
up to a non-zero scalar multiple, we may assume that

Then
cx.2 - cx.3 = P2 - p3, cx.3 - cx.1 = P3 - P1,
. cx.1 - a2 = P1 - P2·
Hence the 'points' c 1,c2,c 3 are represented by the vectors cx. 2 - cx 3,cx 3 - cx. 1,
cx. 1 - a 2. Since the three vectors have zero sum, they are contained in a two-
dimensional subspace of V and the three 'points' are collinear.
122 VI. Desargues

Desargues' theorem implies its own converse (cf. the proof of Proposition 3
below):

If a 1,a2 ,a3 and b 1,b2,b3 are non-collinear triples of 'points' in P(V) such
that

are distinct collinear 'points', then the 'lines' <a 1,b1>,<a2,b2>,<a 3,b3> have a
common 'point' p.

Suppose now that P(V) is the projective space associated with a vector space V
over an ordered division ring D. If a,b are distinct 'points' of P(V), represented by
the vectors a,p of V, then the 'line' <a,b> is the disjoint union of the 'points' a,b
and two open segments, represented by the vectors A.a+ µp, where A.,µ e D and
A.µ> 0 for one segment, A.µ< 0 for the other. The corresponding closed segments
are obtained by adjoining the two endpoints a,b.
It is readily seen that if <a;,b;> (i = 1,2,3,4) are four distinct 'lines' with a
common 'point' p '¢ a;,b; (i = 1,2,3,4), and if a3,a4 e <a 1,a2>, b3,b4 e <b 1,b2>,
then b3 ,b4 belong to the same open segment with endpoints b 1,b2 of the 'line'
<b1,b2> if a3,a4 belong to the same open segment with endpoints a 1,a2 of the 'line'
<a1,a2>.

EXAMPLE 6 The projective plane X over the field Q of rational numbers may be
given the structure of a dense, unending linear geometry in the following way.
Regard X as a subset of the real projective plane X' and choose a 'line' A.' in X'
which contains no point of X. (For example, if a 1,a2 e Rare such that l,a 1,a2
are linearly independent over Q, then a 1a 2 '¢: 0 and we can take A.' to be the 'line'
determined by the 'points' (l,a 1-1,0) and (l,O,a2-1).) Any two points x,y e X
with x '¢ y determine a unique 'line' l's X', and l' intersects A.' in a unique point
z' e: X. The points of I.' other than x and y form two disjoint open segments. We
define the segment [x,y] of X to be the intersection with X of the open segment of l'
which does not contain z', together with x and y.
Once again, the only axiom which needs to be verified is (C). One of the
configurations which can arise in the course of this verification is illustrated in
Figure 3.
2. Projective geometry 123

''
''
''
''
''
', A.'
''
''
''
''
b....... ....... .......
' ' ',
....... ....... '
....... ....... ....... ' ',

/
- - - - -- - - .......- ~
~-a--==:
''
/
Figure 3: Veblen's projective plane

This example of a dense, unending linear geometry, which was given by


Veblen (1904), may be readily generalized. Let D be an ordered division ring and
suppose D is a subring of an ordered division ring D' which has dimension> n,
considered as a vector space over D. It will be shown in Chapter VIII that such a
division ring D' exists, for any positive integer n, if D is not isomorphic to JR.
Then we can take X to be the n-dimensional projective space over D, and A.' to be a
'hyperplane' in the enclosing projective space X' over D' which contains no point of
X. (For example, if cx 1, ••• ,cxn e D' are such that l,cx 1, ••• ,cxn are linearly
independent over D, we can take A.' to be the 'hyperplane' determined by the
'points' (l ,cx 1-1,0, ... ,0) ,( l ,O,cx 2-I , •.• ,0), ... ,(l ,0,0, ... ,an-1).)

EXAMPLE 7 For any ordered division ring D and any positive integer n, the
n-dimensional projective space pn = pn(D) can be given the structure of a dense,
but not unending, linear geometry in the following way. We regard pn as the set of
(n + 1)-tuples of elements of D, not all zero, with proportional (n + 1)-tuples
identified, and Pi (0 < j < n) as the set of all (n + 1)-tuples (~ 0 ,~ 1 , .•• ,~n) with
~o = ~ 1 = ... = ~n-j-l = 0. In particular, po contains the unique 'point'
Po = (O, ... ,O,~n>·
For any x,y e pn with x :F- y put [x,x] = {x) and define [x,y] inductively in the
following way. If x,y e pl and x,y ':I: p 0 , take [x,y] to be the segment with
124 VI. Desargues

endpoints x,y of the 'line' <t,y> which does not contain p0; if x,y e pl and x =Po
or y =Po take [x,y] to be the segment with endpoints x,y of the 'line' <t,y> which
contains all (n + 1)-tuples A.x + µy with A.µ > 0. Now assume that [x,y] has been
defined for x,y e pk-1, where k > 1. If x,y e pk and x,y E pk-1, take [x,y] to be
the segment with endpoints x,y of the 'line' <t,y> which does not contain a point of
pk-1; if x,y e pk and exactly one of x,y is in pk-1, take [x,y] to be the segment with
endpoints x,y of the 'line' <t,y> which contains all (n + 1)-tuples A.x + µy with
A.µ> 0.
With these definitions X = pn is a linear geometry. The verification of the
axioms (L2)-(L4) reduces to their verification for n = l, and presents no
difficulty. To verify now the axiom (C) we may assume that a,b 1,b2 are not
collinear. If for any de [c,b2] we put d' = <a,d> n <b 1,b2>, then the map d ~ d'
is a bijection of [c,b2] onto a segment with endpoints b 1,b2 of the 'line' <b 1,b2>.
By considering separately the various possibilities, it may be seen that in all cases
this segment is [b 1,b2]. Since Xis obviously dense, the axiom (P) follows from
Proposition V.1. However, the dense linear geometry Xis not unending, since if
b :1: Po there is no c :1: Po such that Poe [b,c].

3 THE DESARGUES PROPERTY

A linear geometry will be said to have the Desargues property if the following
condition (illustrated in Figure 4) is satisfied:

(A) Let <a 1,b 1>, <a 2 ,b2>, <a 3,b3 > be three distinct lines with a common point
p :1: a;.b; (i = 1,2,3). If the lines <aj,ak> and <bj,bk> intersect for (j,k) = (1,2),
(1,3),(2,3), then the points of intersection c 3 ,c 2 ,c 1 (which are necessarily
uniquely determined and distinct) are collinear.

The Desargues property is vacuous in a one-dimensional linear geometry. It


will now be shown that it need not hold in a dense unending two-dimensional linear
geometry. (The qualification 'dense' is redundant, by Proposition V.26.) In
Example 4 take µ = 419 and a 1 = (0,4), a 2 = (6,2), a3 = (4,2), b 1 = (8,4),
b2 = (10,-2), b3 = (4,-1). Then the lines <a 1,b 1>, <a 2,b2 >, <a 3,b3 > have the
common point p = (4,4). The lines <a 1,a 2> and <b 1,b2> intersect in the point
c 3 = (9,1), the lines <a2 ,a3 > and <b2 ,b3 > intersect in the point c 1 = (-14,2), and
3. The Desargues property 125

the lines <a 1,a 3 > and <b 1,b 3 > intersect in the point c 2 = (160/29,36/29).
However, c2 does not lie on the line <c 1,c3>. (Only one of the lines mentioned,
namely <b 1,b3>, is a 'refracted' line. Since the ordinary affine plane has the
Desargues property, the refracted plane cannot.)

Figure 4: The Desargues property

In view of this example it is rather remarkable that the Desargues property


must hold in a dense linear geometry which is not two-dimensional:

PROPOSITION 1 If X is a dense linear geometry with dim X > 2, then X has


the Desargues property.

Proof Let p,a;,b;,c; (i = 1,2,3) be points satisfying the hypotheses of (A). We


suppose first that the affine set A = <p ,a 1,a 2 ,a 3> is three-dimensional. Then
<a 1,a 2 ,a 3> and <b 1,b 2,b 3> are distinct planes in A with the common points
c 1,c2,c3• Since the intersection of the two planes is affine, it must be a line. Thus
from now on we may suppose that A is a plane.
Choose x e X \A and then choose x e Pi, where P is the polytope
e:
<p,b3> = <a 3 ,b3 >, we can choose
[p,a 1,a 2,a 3 ,b 1,b2,b 3,c 1,c2,c3 ,x]. Since x
ye (p,x) so that <a 3,x> and <b3 ,y'> intersect within pi for any y' e (x,y]. Indeed
if a 3 e (p,b3) or p e (a 3,b3) we can choose any ye (p,x), whereas if b 3 e (p,a 3)
we can choose x' e pi so that x e (a 3 ,x') and then choose y e (p,x) n (b 3 ,x').
126 VI. Desargues

Similarly we can choose y e pi with x e (p,y) so that <a3,.x> and <b3 ,y'> intersect
within pi for any y' e (x,y].
Choose such a y e (p,x) if b 3 e (p,a 3}, and such a y with x e (p,y) if
a 3 e (p,b 3 ) or p e (a 3 ,b3). If z e <a3,x> n <b3 ,y>, then z e pi and in every case
x e (a 3 ,z). Since z e <a 3 ,a 1> = <c 2 ,a 1>, we can in the same way choose
a 3 ' e (x,z) so that <a 1 ,a 3 '> and <c 2 ,z> intersect in Pi. Similarly, since
z e <a 3,a 2> = <c 1,a2>, we can choose a 3 " e (x,z) so that <a2,a3 "> and <c 1,z>
intersect in Pi. Moreover, by choosing the one which is nearer to z, we may
suppose a 3 " = a 3 '. In the same way we can choose b 3 ' e (y,z} so that <b 1,b3 '>
intersects <c2,z> in pi and <b2,b3 '> intersects <c 1,z> in Pi. Furthermore, we may
now modify our choice of a 3 ' and b3 ' to ensure that p e <a3',b3 '>.
The line <c2,z> is not contained in the plane A'= <JJ,a 1,a3 '>. For otherwise
we would have x,a 3,b3 ,a 1,b 1 e A' and hence, since A is a plane, A =A', which
contradicts x e A. Since the line <c2,z> intersects both <a 1,a3 '> and <b 1,b3 '>, it
follows that the two points of intersection must be the same. Thus there exists a
point c 2 ' e <a 1 ,a 3 '> n <b 1 ,b 3 '>, and similarly there exists a point c 1 ' e
<a 2 ,a 3 '> n <b 2 ,b 3 '>. The points p,a;,bi,c;' (i = 1,2) and a 3 ',b3 ',c3 satisfy the
hypotheses of (A) and the pointsp,a 1,a 2 ,a 3 ' are not coplanar, since a2 e A'.
Hence, by the first part of the proof, c 1',c2 ',c3 are collinear. Thus c 3 e <c 1,c2,z>.
Since ze A and <c1tc 2 ,c3 > s A, it now follows from Proposition 111.8 that
C3 E <C1 ,C2>. 0

Although the Desargues property does not hold in all two-dimensional dense
linear geometries, Proposition 1 shows that it must hold in such a geometry if it can
be embedded in a dense linear geometry of higher dimension. Consequently we
will assume throughout the remainder of this section that X is a dense linear
geometry of dimension> 2 with the Desargues property.
We show first that a modified form of the Desargues property also holds:

PROPOSITION 2 Let <a 1,b 1>, <a 2,b2>, <a 3 ,b3 > be three distinct lines with a
*
common point p ai,b; (i = 1,2,3). If
3. The Desargues property 127

Proof The hypotheses imply that c1 -:1: a2,a3,b2,b3 .and c2 -:1: a 1,a3 ,b 1,b3• Moreover
c 1 -:1: c2 • Hence the lines <a 1,c2 >, <a 2 ,c 1>, <p,b 3> are distinct and have the
common point a3• It now follows from (A) that c3 e <b1,b2>. a

We establish next a converse of the Desargues property:

PROPOSITION 3 Let a 1 ,a 2 ,a 3 and b 1 ,b2 ,b3 be two non-collinear triples of


points such that

exist and are distinct collinear points. If p e <a 1,b 1> r. <a 2 ,b2>:, then also
p E <a3,b3>.

Proof Since the points c 1,c2 ,c3 are distinct, we have a; -:1: b; (i = 1,2,3). Since
a 1,a2,a3 and b 1,b2 ,b3 are not collinear, we also have c; -:I: a;,b; (i = 1,2,3). We
may further assume that c 1,c 2 -:I: a 1,a 2 ,b 1 ,b 2 • For suppose c 2 = a 1 • Then
a 2 e <c2,c3> = <c 1,c2>. Since a1,a2,a3 do not all lie on the line <c 1,c2>, it follows
that c 1 = a2. Thus a 1 e <b 1,b3>, a2 e <b2,b3>, and hence p =b3.
With this assumption the lines <a 1,a2>, <b 1,b2>, <c2,c 1> are distinct, and by
hypothesis they have the common point c 3 • We may assume in addition that
c3 ':/! a 1,a 2,b 1,b2 . For suppose c3 =a 1• Then a3 e <c2 ,c 3> = <c 1,c3> and hence
c 1 = a3, since a 1,a2,a3 are not collinear. Moreover <a 1,b 1> = <b 1,b2> and hence
p = b2• Since c 1 e <b2,b3> and c 1 = a 3 -:I: b3 , we obtain p e <a3 ,b3>.
With this additional assumption we have

The Desargues property and its converse may be loosely summarized by


saying that 'two triangles are perspective from a point if and only if they are
perspective from a line'.
In the next chapter we will make much use of the following result. Since
under our hypotheses two coplanar lines need not always intersect, the proof of the
result will be a good deal more complicated than that of its counterpart in projective
geometry.
128 VI. Desargues

PROPOSITION 4 Let a 1,a 2 ,a 3,a4 and a 1',a 2 ',a3',a 4 ' be two quadruples of
points, such that no three points of the same quadruple are collinear and none
of the points lies on a given line l. If <aj,ak> and <a/,ak'> intersect l in the
same point ajk,for (j,k) = (l,2},(l,3},(l,4},(2,3},(3,4}, and if <a 2 ,a4 > intersects
l in a 24 , then <a 2 ',a4 '> also intersects l in a24.

Proof Since a 1 ,a 2 ,a 3 are not collinear, we must have a 12 a 13 • Similarly


-:I:
a 12 ,a 13 -:1: a 14,a23 and a 34 -:1: a 13 ,a 14,a23 • The set-up is illustrated in Figure 5.

Figure 5

We consider first the case in which the two quadruples have a corresponding
point in common. Since ai =a/, ak = ak' for some j,k e {1,2,3,4} with j -:1: k
implies a;= a;' for all i e {1,2,3,4 }, we need only consider the case in which
a;=a;' for exactly one ie {l,2,3,4}. If a 1 =a 1 ', then the lines <a 2 ,a 2 '>,
<a4 ,a4 '>, <a 3 ,a 3 '> have the common point a 1 and hence, by Proposition 2,
a 24 e <a 2 ',a4 '>. If a 3 = a 3 ', then the lines <a4 ,a4 '>,<a 2,a2 '>,<a 1,a 1'> have the
common point a 3 and hence, in the same way, a 24 e <a2 ',a 4 '>. If a 2 =a 2 ' we
obtain a 24 e <a 2 ',a 4 '> by applying Proposition 3 to the triples a 3 ,a 1,a4 and
3. The Desargues property 129

a 3',a 1',a4'. If a 4 = a 4' we arrive at the same conclusion by applying Proposition 3


to the triples a 1,a3,a2 and a 1',a3 ',a2'· Thus we may now assume a;¢. a;'
(i = 1,2,3,4).
Assume first that <aj,ak> ¢. <a/,ak'> for (j,k) = (l,2),(1,3),(1,4),(2,3),(3,4).
Suppose also that there exists a point q e <a 1,a 1'> r'\ <a3,a3'>. Since a 1,a3,a2 and
a 1',a3',a 2 ' are non-collinear triples, and since a23,a 12,a 13 are distinct collinear
points, it follows from Proposition 3 that q e <a 2,a2'>. In the same way, since
a 1,a3,a4 and a 1',a3',a4' are non-collinear triples, and since a 34,a 14,a 13 are distinct
collinear points, q e <a4,a4'>. Thus q =<a2,a2'> r'\ <a4,a4'>. Since a2,a23,a2'
and a4,a34 ,a4' are non-collinear triples, and since a 3',q,a 3 are distinct collinear
points, it follows from Proposition 3 again that a24 e <a2 ',a4'>, as we wished to
show. Thus we may now suppose that <a 1,a 1'> r'\ <a 3,a 3'> = 0. Then
a 13 e (a 1',a31ifa13 e (a 1,a3),a 1'e (a3',a 13)ifa 1 e (a3,a 13 ),anda3'e (a 1',a 13)
if a 3 e (a 1,a13).
We now drop the assumption made at the beginning of the previous paragraph.
Our strategy will be to construct from the quadruple a 1,a2,a3,a4 another quadruple
b 1,b2,b3,b4, no three points of which are collinear, with <bj,bk> ¢. <a/,ak'> and
ajk e <bj,bk> for (j,k) = (1,2),(1,3),(1,4),(2,3),(3,4), with a 24 e <b 2,b4>, and
such that a 13 ,b 1,b3 are not in the same order as a 13 ,a 1',a3'.
By interchanging a 1 and a 3, if necessary, we may assume that
a 34 E (a 13 ,a 14). Choose any b 1 e (a 1,a 12). Put d = a 13 if a 12 e (a 13 ,a 23 ) or
a 13 e (a 12,a23 ), and put d = (a 1,a23 ) r'\ (b 1,a 13 ) if a 23 e (a 12 ,a 13 ). Then for any
b 3 e (b 1,d) there exists b 2 e (a 1,a 12 ) \bi such that a 23 e <b 2,b 3>. If
a 14 e (a 13 ,a 34 ) take b 4 = (b 1,a 14 ) r'\ (b 3,a 34 ), and if a 13 e (a 14 ,a 34 ) take
b4 e (bi,ai 4) so that b3 e (b4,a34).
By construction no three points of the quadruple b 1,b2,b3,b4 are collinear and
ajk e <bj,bk> for (j,k) = (l,2),(l,3),(1,4),(2,3),(3,4). Moreover bi e (ai,a 12),
b3 e (b 1,a13) and b2 e (a 1,a 12) \bi. We are going to deduce from these properties
that also a24 e <b2,b4>.
Put d' =a 13 if a 13 e (a 1,a3) or a 1 e (a3,a 13), and put d' = (a3,a 12) r'\ (b 1,a 13)
if a 3 e (a 1,a 13). Then every line through a 3 which intersects (b 1,d1 also intersects
(a 1,a12) \ b 1. Put d" = a 13 if a 12 e (a 13 ,a 23 ) or a 13 e (ai 2,a 23 ), and put
d" = (a 1,a23 ) r'\ (b 1,a 13 ) if a 23 e (a 12 ,a 13). Then every line through a 23 which
intersects (b 1,d") also intersects (ai,a 12 ) \bi. We may now choose
c 3 e (b 1,d1 r'\ (b 1,d") r'\ (b 1,b 3) so that c 3 E <a 2,a 3>, c3 E <a 3,a4 >, the line
130 VI. Desargues

<c3,a3> intersects (ai,a 12 ) \bi in a point p such that b 2 E [bi,p], and the line
<c3,a23 > intersects (ai,a 12 ) \bi in a point c 2 such that b 2 E [bi,c 2]. Evidently
bi,c2,c3 are not collinear. If ai 4 e (ai 3 ,a34) take c4 = (bi,ai 4) n (c3 ,a34), and if
ai 3 e (ai 4 ,a34) take c4 e (bi,ai 4) so that c3 e (c4 ,a34). In either case bi,c3,c4 are
not collinear and c4 = (bi,ai 4) n <c3,a34>. Indeed if we put ci =bi, then no three
points of the quadruple Ci ,c2,c3,c4 are collinear and ajk e <cj,ck> for (j,k) =
(1,2),(l,3),(1,4),(2,3),(3,4). Furthermore c; ¢ b;, not only for i = 2,3 but also for
i = 4, since

Since a 23 ,a 13 ,a 12 are distinct collinear points and p e <ci,ai> n <c3,a3>, by


applying Proposition 3 to the triples ci,c3,c2 and ai,a 3,a2 we obtain p e <c2,a2>.
Similarly by applying Proposition 3 to the triples ci,c3,c4 and ai,a 3,a4 we obtain
p e <c4 ,a4>. Thus
p e <c;,a;> (i = 1,2,3,4).
Moreover p -:t:. c;,a; (i = 1,2,3,4). For bi E <ai,a 3> implies p -:t:. bi,a1'c 3,a3.
Similarly c3 E <a2,a3> implies p -:t:. a 2,c2 and c3 E <a3,a4> implies p -:t:. a4,c4.
The triple a 2,a 23 ,c 2 is not collinear, since c 3 E <a 2,a 3>, and the triple
a4 ,a34 ,c4 is not collinear, since c3 E <a 3,a4>. Since c3,p,a3 are distinct collinear
points and a 24 e <a 2,a 4> n I. it follows from Proposition 3 again that
a 24 e <c2,c4>. Since c 2 e <bi,b2> and c 2 ¢ b 2, we have c2 E <b2 ,b3>. Thus the
points c 2,a 23 ,b 2 are not collinear, and similarly the points c 4 ,a 34 ,b4 are not
collinear. Applying Proposition 3 to these two triples we now obtain
a24 e <b2,b4>, since b3,bi,c3 are distinct collinear points and a 24 e <c2,c4> n I..
On account of the arbitrary nature of bi and b 3, we may suppose that
<bj,bk> ¢ <a/,ak'> for (j,k} = (l,3),(1,4),(2,3),(3,4).
However, if <ai,a2> =
<ai ',a 2'>, we will have <bi,b 2> = <ai ',a 2'>. We now apply to the quadruple
bi,b 2,b 3,b4 a similar procedure to that which we applied to the quadruple
a i ,a2,a3,a4.
Choose any bi' e (b 3 ,a 13 ). If a 12 e (a 13 ,a 23 ) or a 13 e (ai 2,a 23 ) put e =ai 3 ,
and if a 23 e (ai 2,a 13 ) put e = (bi ',a 13 ) n (g,a 23 ), with g =(ai,ai 3) n <bi ',ai 2>.
Then for any b 3 'e (bi',e), there exists b 2 'e (g,ai 2 )\bi' such that
a2 3 e <b2',b 3'>. Now define b 4' in the analogous way to b4. By construction no
three points of the quadruple bi',b2',b 3',b4' are collinear and ajke <b/,bk'> for
(j,k) = (l,2),(l,3},(l,4),(2,3),(3,4). In the same way as before we may show that
3. The Desargues property 131

a24 e <b2',b/>. On account of the arbitrary nature of b 1',b3 ' we may suppose that
<b/,bk'> :I: <a/,ak'> for (j,k) = (l,2},(1,4),(2,3},(3,4), but in addition
<b 1',b 3 '> = <b 1,b 3> :I: <a 1 ',a 3 '>. The quadruple b 1 ',b 2 ',b 3 ',b 4 ' has b 3 ' e
(b 1 ',a 13 ). If a 3' e (a 1 ',a 13 ), we are finished. To complete the proof we now
construct another quadruple b 1",b 2",b 3",b4 " with the same properties but with
b1" E (b3 ",a13).
We have b 3' e (b 1,a 13 ) and b 1 e (a 1,a 12). Choose any b 1" e (b 3',a 13). We
are going to choose ye (b 1",b3 ') so that, for each x e (b 1",y), the line <x,a 34>
intersects the line <b 1",a 14>. If a 13 e (a 14 ,a 34 ), choose any ye (b 1",b 3'). If
a 14 e (a 13 ,a 34 ) we choose y in the following way. If a 13 e (a 12 ,a 34 ), then
b 1" e (u,a 14 ) for some u e (b 1,a 12) and we take y' = (b 1,a 13 ) r'\ (u,a 34 ). If
a13 e (a 12 ,a34), then b 1" e (v,a 14) for some v e (a 1,a 13) and we take y' e (v,a34)
so that b 1" e (y',a 13 ). Then in either of the last two cases we choose any
Y E (b1 ",b3 ') ('\ (b{,y').
If a 12 e (a 13 ,a 23 ), take z e (a 1,a 13) so that b 1" e (z,a 12 ) and then choose
z" e (b 1",z) so that y" = (z",a 23 ) r'\ (b 1",b 1) satisfies y" e (b 1",y). Choose any
b 2" e (b 1",z") and then take b 3" = (b 1",y'') r'\ (b2",a23 ).
If a 13 e (a 12 ,a 23 }, take z e (y,a 23 ) so that b 1" e (z,a 12 ), and if
a23 e (a12,a13}, take z = (b 1 ",a 12 ) r'\ (y,a 23 ). In either case choose any
b 2" e (b 1",z) and then take b3" e (b{,y) so that b2" e (b3",a23).
In every case we now take b4 " = <b3",a34> r'\ <b1",a 14>. By construction no
three points of the quadruple b 1",b 2",b3",b 4 " are collinear, b 1" e (b3 ",a 13) and
ajk e <b/',bk"> for (j,k) = (l,2),(1,3),(1,4),(2,3),(3,4). In the same way as before
it follows that a24 e <b2",b4 "> and we may choose b 1",b2 " so that <b/',bk"> :I:
<a/,ak'> not only for (j,k) = (1,3), but also for (j,k) = (l,2},(l,4),(2,3),(3,4). D

4 FACTOR GEOMETRIES

In this section we study the way in which the Desargues property, and also the
properties of density and unendingness considered in Chapter V, are inherited by
the factor geometries of a linear geometry which were defined in Chapter IV.
132 VI. Desargues

PROPOSITION 5 Let X be a linear geometry, H+ an open half-space


associated with a hyperplane H of X, and L a nonempty affine subset of H. If
the linear geometry Xis dense (resp. unending), then the linear geometry H+: L
is also dense (resp. unending).

Proof Suppose p 1,p 2 e H + : Land p1 :I: p2. Choose a; e P; \L (i = 1,2). Then


a 1 :I: a 2 • If X is dense, there exists a e (a 1 ,a 2 ). Then a e H + and
p: =<au L> () (H+ u H) satisfies p e (p 1,pi).
If X is unending, there exists a 3' e X such that a 1 e (a2,a3'). If a 3' e H+' put
a 3 = a3 '. If a 3 ' E H + then, by Propositions III.12 and V.26, there exists a point
a 3 e (a 1,a3 ') n H +· Then a 1 e (a 2,a3) and p3: = <a3 u L> n (H + u H) satisfies
p 1 E (P2,p3). D

We next investigate the relationship between the dimensions of a linear


geometry and its factor geometries. We first observe that if p1,... ,pm e H+: L, and
if a; e P; \L (i = l, ... ,m), then [p 1,... ,pml is the set of all p e H+: L such that
p n [a 1, ••• ,am] :I: 0. This follows by induction from Lemma IV.24 (without any
extra hypotheses).
Now let X be a dense linear geometry, H+ an open half-space of X associated
with a hyperplane H, and La nonempty affine subset of H. Also let aO e H+, let
B = {bi: i e /} be an affine basis for L, and let C = {d: j e J} be an affine
independent set (with 0 E J and possibly J = 0) such thatB u C is an affine basis
for H. For each j e J, choose ai e (aO,ci) and put J' =Ju {0}, A = {ai: j e J'}.
Then Au Bis an affine basis for X.
For each j e J', put pi = <ai u L> n (H+ u H). We are going to show that
R = {pi: j e J'} is an affine basis for H + : L.
We show first that R affinely generates H + : L. Let p e H+ : L and choose
a e p \ L. Since A u B is an affine basis for X, there exist finite sets
{a 1, ••• ,am} s A and {b 1 , .•. ,bn} s B such that a e <a1'···,am,b 1, ••• ,bn>·
Hence, by Proposition III.7, a e <c',c"> for some c',c" e [a 1, ••• ,am,b1'···,bnl·
Moreover we may choose the notation so that c' E L. Then c' e [a',b') for some
a' e [a 1, ••• ,am] and b' e [blt···,bnl· Hence p' = <c' u L> () (H + u H) satisfies
p' e [p1, ... ,pml, where P; =<a; u L> n (H+ u H) (i = l, ... ,m). If c" e L, then
p = p'; if c" E L then, in the same way, p" = <c" u L> () (H+ u H) satisfies
p" e [p1, ... ,pml· Hence p e <p',p"> s <p1, ... ,pm>.
4. Factor geometries 133

It remains to show that R is affine independent. It is enough to show that if


{p 1, ••• ,pml is an affine independent subset of Rand if Pm+l e R is such that
Pm+l e <p1, ... ,pm>, then Pm+l = P; for some i Sm. We have Pm+l e <p',p">,
where p',p" e [p 1, ... ,pm1· Thus if P; =<a; u L> n (H+ u H}, where a; e A
(i = l, ... ,m+l), then there exist a' e p', a" e p" such that a',a" e [a 1, ••• ,am1· If
Pm+l e [p',p"], there exists dm+l e Pm+l such that dm+l e [a',a"] s [a 1, ••• ,am1
and hence am+l e <dm+l u L> s <alt···,am,B>. If p' e (Pm+ 1,p"}, there exists
a' E p' such that a' E (am+1,a") and again am+l E <a',a"> S <alt···,am,B>.
Similarly if p" e (Pm+ 1,p') we obtain am+l e <a 1, ••• ,am,B>. Since A u B is an
affine basis for X it follows that, for some i < m, am+l =a; and hence Pm+l = p;.
It follows that H + : L has finite dimension r if and only if L has codimension r
in H.
It will now be shown that, whether or not a dense linear geometry has the
Desargues property, its factor geometries always have the Desargues property.

PROPOSITION 6 Let X be a dense linear geometry, H+ an open half-space


associated with a hyperplane Hof X, and L a nonempty affine subset of H.
Then H+: Lis a dense linear geometry with the Desargues property.

Proof By Proposition IV.26 and Proposition 5, it only remains to prove that


H + : L has the Desargues property. H H + : L has four affine independent points, we
can apply Proposition 1. If H + : L has at most two affine independent points, the
Desargues property is vacuously satisfied. Thus we now assume that dim H + : L =
2. Then the Desargues property holds in X, by Proposition l, and H has an affine
basis of the form D u {c',c"}, where D is an affine basis for L. Choose d e D and
c e (d,c"). Then M = <D \ d,c,c'> is a hyperplane of H and d,c" lie in different
open half-spaces of H associated with the hyperplane M.
For any a e H+, let aa be the closed half-space of <au M> associated with
the hyperplane M which contains a. We are going to show that
H+: M = {aa: a e H+} is totally ordered by writing aa <ab if [d,b] n aa ¢. 0.
We show first that the ordering depends only on ab e H + : M and not on
be H+. Since it is evident that Ga= ab implies aa <ab, we may suppose that
Gan ab= M. Assume on the contrary that (d,b) n Ga¢. 0, but [d,b1 n aa = 0
for some b' e ab \M. Then d and b' lie in the same open half-space of X
associated with the hyperplane <a u M>, but d and b lie in different open half-
134 VI. Desargues

spaces. Hence band b' lie in different open half-spaces and there exists a point
x e (b,b1 n <au M>. Since x e a 0 nab, but x E M, this is a contradiction.
Suppose now that a 0 <ab and ab< a 0 • Then there exist points a' e a 0 \M
and b' e ab\ M such that a' e [d,b] and b' e [d,a]. Hence there exists a point
ye [a,a1 n [b,b1. Since ye a 0 nab and y E M, it follows that a 0 =ab.
Suppose next that a 0 S ab and ab < ac. Then there exists a point b' e ab\ M
such that b' e [d,c]. Also, since ab.= ab, there exists a point a' e a 0 \M such that
a' e [d,b1. Then a' e [d,c] and hence aa < ac.
It remains to show that, for any a,b e H +' either a 0 <ab or ab < a 0 •
Obviously we may assume b E a 0 and a E ab. Let C',C" be the closed half-spaces
of X associated with the hyperplanes H' = <a u M>, H" = <b u M> which
contain d. We intend to show that either a e C" orb e C'.
Assume on the contrary that a E C" and b E C'. Then there exist points
u e (d,a) n H" and v e (d,b) n H'. Hence d,a,b are not collinear and there exists
a point w e (a, v) n (b,u). Then w e H' n H" = M and there exists a point
t e (a,b) such that we (d,t). Since t e <d,w> s H and (a,b) s H +' this is a
contradiction.
From c" e H\ M it follows that c" E H',H". Let B ',B" be the closed half-
spaces of X associated with the hyperplanes H',H" which contain c". Then in the
same way we can show that either a e B" orb e B'. Since b E a 0 we cannot have
b e H' = C' n B', and since a E ab we cannot have a e H" = C" n B". Hence
a e C" implies b E C', and be C' implies a E C". But if b E C', then (d,b) n H' ~
0 and hence a 0 <ab. Similarly if a E C", then ab< a 0 •
We now proceed with the proof that H+: L has the Desargues property. Let
<p1,P1'>, <p2,P2'>, <p3,p3'> be three distinct 'lines' in H+: L with a common
'point' p ¢. p;,p;' (i = 1,2,3). Suppose also that the intersections

exist. Choose a; e p; \L, a;' e p;'\L (i = 1,2,3). We begin by showing that there
exists h' e H + such that the hyperplane H' = <h' u M> of X intersects (d,a;],
(d,a;1 (i = 1,2,3).
Let a;,a;' be the closed half-space of <a; u M>, resp. <a;' u M>, associated
with the hyperplane M which contains a;, resp. a;'. By what we have just proved
we may choose the notation so that a 1 ~ a;,a;' (i = 1,2,3). Thus if we take h' = a 1
and H' = <h' u M>, then [d,a;] n a 1 ¢. 0, [d,a;1 n a 1 ¢. 0 (i = 1,2,3). Since
4. Factor geometries 135

d E H', it follows that H' contains points of (d,a;], (d,a;1 (i = 1,2,3). By changing
notation we may now assume that a;,a;' e H' (i = 1,2,3).
Suppose first that L = {d}. Then any element of H + : L intersects H' in at
most one point. Since p e <p;,p;'> and p :I: p;,p;' (i = 1,2,3), it follows from
Lemma N.24 and the definition of a line that there exists a point P; e p \L such that
P; e <a;,a;'> and p; :I: a;,a;'. Moreover, by the preceding remark, P; = p is
independent of i. Furthermore the lines <a;,a;'> are distinct, since the 'lines'
<p;,p;'> are distinct. In the same way 'Yi contains both a point hi' e <a2 ,a3> and
a point hi" e <a2',a3 '>,and we must have hi'= hi". Thus the intersections

exist. Hence, by the Desargues property in X, hi ',h 2 ',h 3 ' are collinear.
Consequently y1,y2,y3 are collinear in H + : L.
Suppose next that L :I: { d} and put i = <D \ d>. Any a e H + is contained in a
unique p e H+: Land a unique p e H +: i. It is easily seen that if also a e H',
then p = p n H'. Hence if we put P; = p; n H', p;' = p;' n H' (i = 1,2,3), then
the lines <pi, Pi'>,<p2 , p 2 '>,<p3 , p3 '> in H + : i have a common point :I: f);, P;'
(i = 1,2,3). Furthermore the intersections

- -- -,-,,
"f3 =<pi, P2,> n <P1, P2 >

exist. But the Desargues property holds in H + : i, since dim H + : i = 3.


Consequently i 1, y2, y3 are collinear in H+: i. Since Y; = 'Y; n H' (i = 1,2,3), it
follows that 'Yi ,y2,y3 are collinear in H + : L. D

5 NOTES

Among the many texts on projective geometry, two which are particularly
concerned with its foundations are Heyting (1980) and the classic Veblen and
Young (1910/18). A basic theorem says that if Pis a projective space containing a
pair of disjoint 'lines', then there is a vector space Vover a division ring D such that
P = P(V). Moreover the same conclusion holds if no two 'lines' of P are disjoint
(i.e. if P is a projective plane), provided Desargues' theorem holds in P. In the
136 VI. Desargues

next chapter we will prove the analogues of these theorems for linear geometries, in
which case the division ring D is ordered.
Desargues' theorem dates from 1636. Smith (1959) gives a short account of
the life of Desargues, together with an English translation of the statement of his
theorem in the Oeuvres de Desargues. Desargues' theorem provides an interesting
combinatorial configuration of ten points and ten lines, each line containing three of
the points and each point contained in three of the lines. The ubiquitous Petersen
graph is obtained by taking the ten points as vertices, with two points joined by an
edge if they are not both contained in one of the ten lines.
Segre (1956) has given an 'algebraic' example of a projective plane which
does not have the Desargues property. Its 'points' are the 'points' (~0 ,~ 1 ,~ 2) of the
real projective plane, but its 'lines' are the two-parameter family of cubic curves

where e is a sufficiently small positive constant and p 0,p 1,p2 are homogeneous
parameters specifying the curve.
Many dense unending two-dimensional linear geometries which do not have
the Desargues property are provided by Example 5, in particular the two specific
cases discussed there. (Even Moulton planes with different refractive indices µ are
non-isomorphic.) Rather than attempt to enumerate all such geometries, we will
show in Chapter VID that they can still possess nice 'Euclidean' properties.
The proof of Proposition 4 is based on Sperner (1938), but some extra
argument is needed because we do not require that the linear geometry be unending.
VII
Vector Spaces

Here the results of the previous chapter are used to introduce coordinates, using a
splendid idea of von Staudt, as modified by Hilbert and then generalized by
Sperner. We define isomorphism of linear geometries and establish a result of
Doignon, that any dense linear geometry of dimension > 2 with the Desargues
property can be isomorphically embedded in a projective space over an ordered
division ring. In general such a geometry cannot be isomorphically embedded in a
vector space over an ordered division ring, but an open half-space of such a
geometry can be embedded in this way.
We also study properties involving vector sums of convex sets in a vector
space over an ordered division ring. In particular, we establish a useful algebraic
version of RAdstrom's cancellation law, which will be encountered in Chapter IX.

1 COORDINATES

We assume throughout this section that X is a dense linear geometry of


dimension > 2 with the Desargues property. We are going to show that it is
possible to define addition and multiplication of points of a segment so that all the
usual algebraic laws hold, with the exception of the commutative law for
multiplication. Actually the entire discussion will be based on Proposition VI.4.
Let z, w be any two distinct points. In order to define the sum of two points
a,b e (z, w}, choose any point w" E <z, w> and any point w' e (w, w"). There
exists a point a' e (z,w') such that the line <a,a'> intersects the segment (w,w").
Indeed we can take any a' e (a",w1, where a"= (a,w") ("\ (z,w1. Put

w0 =<a,a'> ("\ (w,w"), p =(b,w1 ("\ (a',w).


138 Vil. Vector spaces

Then actually w a e (w', w") and we can define the sum p = a + b by p =


<p,w0 > ("\ (z,w). (See Figure 1.)
This definition does not depend on the choice of w' and a'. For suppose
ro" E <z,w>, ro' e (w,ro") and a' e (z,ro') is such that <a,a'> intersects (w,ro") in
ro0 • Define it and 1t in the analogous way to p andp. Applying Proposition VI.4
to the quadruples w',p,a',w0 and ro', it,a',0>0 , we immediately obtain 1t = p.

w"

Figure 1: Addition

We show first that addition is commutative: a + b = b + a. Evidently we may


assume that a ;e: b. Put b' = (b,w0 ) n (z,w') and p = (a,w') n (b',w). Then (with
b' in the role of a') p = b +a is given by p = <p,w 0 > n (z,w). Applying
Proposition VI.4 to the quadruples a',p,w',w 0 and w',w0 ,b',p, we immediately
obtain p = p. (See Figure 2.)
We show next that addition is associative: (a+ b) + c =a+ (b + c). Besides
the notations already introduced, put w = <z , p> n (w, w ") and ii =
(c, w) n (a',w). Then (with w,p in the role of w',a') q = b + c is given by
q = <ii,w'> n (z,w), and similarly s = p + c is given bys= <ii,w0 > n (z,w).
Since (with w',a' as initially) r =a + q is also given by r =<ii ,wa> n (z, w), we
obtain r =s. (See Figure 3.)
It is easily seen that, for any points a,b e (z,w), there exists a point c e (z,w)
such that b = a + c if and only if a e (z ,b ). Moreover, c is then uniquely
determined and c e (z,b).
1. Coordinates 139

z w

Figure 2: Commutativity of addition

z a C Q r=S w

Figure 3: Associativity of addition

To define the product of two points a,b e (z,w), we fix a point e e (z,w).
There exists a point z' e (z,w1 such that the lines <e,z'> and <a,z'> both intersect
the segment (w,w"). Indeed we can take any z' e (z",w'), where
z" = (e,w") rt (z,w1 if a e (z,e] and z" = (a,w") rt (z,w1 if a e [e,w). Put
140 VII. Vector spaces

We= <e,z'> ("'\ (w,w"), W0 = <a,z'> ("'\ (w,w").


Then actually we,wa e (w',w"). Thus we can take p' = (b,we) n (z,w') and define
the product p = a·b by p = <p',w0 > n (z,w). Evidently p = b if a= e and p =a if
b = e. (See Figure 4.)
This definition does not depend on the choice of w' and z'. To see this, we
may assume a,b '#: e. Suppose ro" f! <z,w>, ro' e (w,ro") and~, e (z,ro') are such
that <e,~'> and <a,~'> intersect (w,ro") in roe and ro0 respectively. Define 1t' and
1t in the analogous way top' and p. Applying Proposition VI.4 to the quadruples
z',wa,we,p' and ~',roa,roe,1t', we immediately obtain 1t = p.

z e P a w

Figure 4: Multiplication

It is not true, in general, that multiplication is commutative. However, we


now show that multiplication is associative: (a·b)·c = a·(b·c). Choose z" e (z,w')
so that the lines <e,z">, <a,z"> and <b,z"> all intersect the segment (w',w") and
then choose Wee (w',w") so that all three points of intersection lie in (we,w"). If
we put z' = (e,we) n (z,w'), then actually z' e (z",w'). Hence there exist points
w0 = <a,z'> n (w',w") andp' = (b,we) n (z,w'). Furthermore, there exist points
wP = <e,p'> n (w',w") and q' = (c,wp) n (z,w'). Then p = a·b is given by
p = <p',w 0 > n (z,w) and (with p' in the role of z') q = b·c is given by
q = <q',we> n (z,w). Since q' = (c,wp) n (z,w'), s = p·c is given by
s = <q',w0 > n (z,w), and since q' = (q,we) n (z,w'), r = a·q is given by
r = <q',w0 > n (z,w). Thus r = s.
1. Coordinates 141

It will now be shown that the points of the segment (z,w) form a group under
multiplication. Since multiplication is associative and e acts as an identity element,
we need only show that for each point a e (z,w) there exists a point be (z,w) such
that a·b = e. But if e' = (e,w0 ) n (z,w') and b = <e',we> n (z,w), then a·b = e.
We prove next the distributive laws: (a+ b)·c =a·c + b·c and c·(a + b) =
c·a + c·b. Again put p =a + _b and choose z' e (z,w') so that the lines <e,z'>,
<a,z'>,<b,z'>,<p,z'> all intersect the segment (w,w"), in points we,wa,wb,wp say.
If q' = (c,we) n (z,w') then, by the definition of multiplication,

a·c = <q',w0 > n (z,w), b·c = <q',wb> n (z,w), p·c = <q',wp> n (z,w).

By the definition of additionp =<p,w0 > n (z,w), where p = (b,w') n (z',w). If


s' = (q',w) n (b.c,w') then, by applying Proposition VI.4 to the quadruples
b,p,z',p and b·c,p·c,q',s', we obtain w0 e <p·c,s'>. Since w0 e <a·c,q'>, this
proves that p·c =a·c + b·c.
To prove the second distributive law choose z' e (z,w') so that the lines <e,z'>
and <c,z'> intersect the segment (w',w''), in we and we. Then

C·a =<a',wc> (") (z,w), C·b =<b',wc> n (z,w),


where
a'= (a,we) n (z,w'), b' = (b,we) ri (z,w').

Moreover p =a + b and r =c·p are given by

p =<fi,we> ri (z,w), p =(a',w) n (b,w'),


and
r =<p',wc> ri (z,w), p' = (p,we) ri (z,w').

If we put f = (a',w) n (r,wc) then, by applying Proposition VI.4 to the quadruples


w,b,we,fi and w,c·b,wc,f, we obtain w'e <c·b,f>. This proves that r = c·a + c·b.
Put P = (z,w) and let D denote the collection of all ordered pairs (a 1,a2) with
a1'a 2 e P. Two elements (a 1,a2) and (a 1',a 2 ')of D will be said to be equal if
there exist x,x' e P such that

It is easily verified that this is indeed an equivalence relation on D. Moreover, for


any x e P, (a 1+x,a 2+x) = (a 1,a2). We also define the sum and product of two
elements of D by
142 VII. Vector spaces

It is easily verified that these definitions do not depend on the choice of elements
within an equivalence class, i.e., if (a 1',a21 = (a1'a2) and (b 1',b2 1 = (b 1,b2), then

Furthermore, the commutative and associative laws for addition, the associative law
for multiplication and the two distributive laws hold in D, since they hold in P.
The elements 0 =(a,a) and 1 =(e+a,a) act as identity elements for addition
and multiplication respectively. Moreover any element (a 1,a2) of D has an additive
inverse (a 2,a 1). We will show that any element (a 1,a2) :I= 0 has a multiplicative
inverse. Since a 1 :I= a2, either a 1 e (z,a 2) or a2 e (z,a1). If a 1 e (z,a2), there exists
x e P such that a 1 + x =a2 and hence (a 1,a2)·(a,x-1+a) = 1. If a 2 e (z,a 1), there
exists ye P such that a2 + y = a 1 and hence (a 1,a2)-(y-1+a,a) = 1.
Thus we have now shown that D is a division ring. Since the map
x--+ (x+a,a) of PintoD preserves sums and products, we may identify the point
x e P with the element (x+a,a) of D. In this way Pis embedded in D. Since Pis
closed under addition and multiplication, and Dis the disjoint union of the sets P,
{0} and -P, it follows that D is an ordered division ring with P as the subset of
positive elements.
We now show that if a and bare distinct points of P, then the 'open segment'
(a,b) is the set of all points ce P which can be represented in the form
c = A.a + (1 - A.)b, where 0< A. < 1. Without loss of generality assume a < b, i.e.,
a e (z,b), and suppose first that c admits such a representation. Since A.a< Ab we
have c < A.b + (1 - A.)b = b, and since (1 - A.)a < (1 - A.)b we have
a= A.a+ (1 -A.)a < c. Thus c e (z,b) and a e (z,c), which implies c e (a,b).
Suppose on the other hand that c e (a,b). Since a e (z,c) we have c =a+ a,
where a e P, and since c e (z,b) we have b = c + "(, where "( e P. Thus
b =a + p, where p =a + "( and hence "( < p. If we put A. =yp-1, then 0 < A. < 1
and
c = a + a = a + (1 - A.)p = A.a + (1 - A.)b.
2. Isomorphisms 143

2 ISOMORPHISMS

Two linear geometries X and Y are said to be isomorphic if there is a one-to-


one map f of X onto Y such that/([x 1,x2]) = [f(x 1)/(x2)] for all x 1,x2 e X. The
map f itself is said to be an isomorphism of X onto Y. It is easily seen that 1-1 is
an isomorphism of Y onto X. Moreover, if g is an isomorphism of Y onto Z, then
g 0 f is an isomorphism of X onto Z. If l is a line inX thenj{l) is a line in Y, and if
A is an affine independent subset of X thenj{A) is an affine independent subset of
Y.
With this terminology, we have shown in Section 1 that the segment (z,w) of a
dense linear geometry X of dimension ~2 with the Desargues property is
isomorphic to the set P of positive elements of an ordered division ring D. The
segment (z, w) of X is also isomorphic to the segment (0, 1) of D, since the map
f: x--+ x(l + x)-1 of Ponto (0,1) is an isomorphism. It follows that the segment
[z,w] of Xis isomorphic to the segment [0,l] of D.

PROPOSITION 1 Let X be a dense linear geometry of dimension ~ 2. If


a 1,a2,b1,b2 e X and a 1 '¢ a2, b 1 '¢ b 2, then the segments [a 1,a 2] and [b 1,b2] are
isomorphic.

Proof Suppose first that b 1 = a 1 and b 2 ~ <a 1,a 2>. Choose some c e (a 2 ,b2).
For any a e (a 1,a 2), there exist a unique a' e (b 2 ,a) ri (b 1,c) and a unique
be (b1,b2) such that a' e (a 2,b). Moreover every be (b 1,b2) may be obtained in
this way. If we put /(a) = b,f(a 1 ) = b 1 ,f(a 2 ) = b 2 , then it follows from
Propositions 11.17' and 11.2' that/ is an isomorphism of [a 1,a2] onto [b1,b2].
Suppose next that <a 1,a2> = <h 1,b2> and choose c2 ~ <a 1,a2>. By what we
have already proved, [a 1,a2] is isomorphic to [a 1,c2], [a 1,c2] is isomorphic to
[b 1,c2 ], and [b 1,c2 ] is isomorphic to [b 1,b 2 ]. Hence [a 1,a 2] is isomorphic to
[b1 ,b2].
Suppose finally that <a 1,a2> :I: <h 1,b2>. We may choose the notation so that
b 1 ~ <a 1,a2> and a 1 ~ <b1,b2>. Then, by what we have already proved, [a1'a 2]
is isomorphic to [a 1,bi] and [a1tbtl is isomorphic to [b1tb2]. Hence again [a1'a2]
is isomorphic to [b 1,b2]. Cl
144 VII. Vector spaces

That the hypothesis on the dimension cannot be omitted in Proposition 1 is


shown, for example, by the dense totally ordered set X consisting of all rational
numbers in the interval (0, 1] and all real numbers outside this interval.

PROPOSmON 2 Let X and X be dense linear geometries with the Desargues


property. Let Ao be an affine independent subset of X containing at least two
points and let fo be an isomorphism of C0 = [A 0 ] onto C0 s X. If a 1 e
X \ <A 0 > and a1 e X \ <Ao>, where Ao =/ 0 (A 0 ), then there exists an
isomorphism f of C =[Ao u ai] onto C = [Aou di] such that f(a 1) = a1 and
f(x) = f 0(x)for x e C0 •

Proof For convenience of writing put A = A0 u a 1 and A = Ao u a1• Also, put


C 1 = [A\ a0 ] where a 0 e A0 , andC1 = [A \ dol where a0 =f 0 (a 0 ). Further, put
Ao'= Ao \ao, Co'= [Ao1 and Ao'= Ao\ d0 , C0 ' =[Ao 1. Assume initially that
there exist be X such that a 1 e (a~b) and be X such that d1 e (d0 , b).
For any x 1 e C1 \C0 ' withx1 ¢ a1, there are a unique pointx1' e C0 ' such that
x 1 e (a 1,x 1') and a unique point x 0 e (a 0 ,x1 ') such that x 1 e (b,x0). If we put
7t(x 1) = x0 , 7t(a 1) = a0 and 7t(x) = x for x e C0 ', then 7t maps C 1 isomorphically
onto C0• In the same way, with b replaced by b, we can define an isomorphism it
of C1 onto Co with it(di) = do and it(y) = y for ye Co'. Then/1 = it- 1 lo 7t 0 0

is an isomorphism of C 1 onto C1 such that/1(a 1) = a1 and/1 (x) =f0 (x) for


xe C0 '=C0 riC1.
For any x e C \ [a 0 ,a i], there exist a unique x 0 e C 0 \ a 0 such that
x e (a 1,x0] and a unique x 1 e C 1 \a 1 such that x e (a0,xi]. Moreover, there exist
a unique x 0 ' e C 0 ' such that x 0 e (a 0 ,x0 1 and a unique x 1 ' e C 0 ' such that
x 1 e (a 1,x 11. Then x 0 ',x 1 ' e <a 0 ,a 1,x>, since x0 ,x1 e <a0 ,a 1,x>. Since also
x 0 ',x 1 ' e <Ao'>, x 0 ' ¢ x 1 ' would imply <a0 ,x0 ',x 1 '> = <a0 ,a 1,x> and hence
a 1 e <A 0>, which is a contradiction. Consequently x 0 ' =x1' =x', say.
It follows that/0 (x0) e (a0 ,i1 and/1(x 1) e (d1,i1, where .£' =fo(x') =
/ 1(x'). Moreover, if x ~ C0 u C 1, then / 0 (x0)/1(x 1) ¢ i' and hence there exists a
unique point i e (do/1(x 1)) ri (a1 / 0 (x0)). We define a map/: C\(ao,a 1) ~ C
by putting.f{x) = i if x ~ C0 u C 1,.f{x) = f 0 (x) if x e C0 and/(x) = f 1(x) if x e C 1•
To define j{z) for z e (a°'a 1) choose a e A 0 ' and x e (a,z). Then.f{x) is in the
intrinsic interior of [d0 , a1/(a)] and hencef(x) e <Jta),i) for some i e (d0 , a1).
We wish to show that i is independent of the choice of a and x. This is obvious if
we only change x. If a' e Ao'\ a and x' e (a',z), then f(a)/(x)/(x')/(a') are
2. lsonwrphisms 145

coplanar and hence i' = i. We now complete the definition off by putting
/(z) = i. It is not difficult to verify that/ is an isomorphism of C onto C with the
required properties.
It remains to show that our initial assumption may be removed. Choose
a 0 ' e Ao' and then c 0 e (a 0 ,a 0 ') and a 1 ' e (c 0 ,a 1). Also, let a0 ' = / 0 (a 0 '),
c0 = / 0(c0), and choose d1' e (c0 , d1). Then a 1' e (a 0,b) for some be (a0 ',a 1)
and d1 ' e (do, b) for some be (d0 ', d1). By what we have already proved, there
exists an isomorphism f' of C' =[Ao u a 11 onto C' = [Ao u a11 such that
f'(a 1 ') = d1 ' andf'(x) = f 0 (x) for x e C0 • Clearly it is enough now to show that
there exists an isomorphism cl> of C onto C' such that cj>(a 1) = a1' and cj>(x) = x for
x e C0. We will merely give the definition of cl> at other points.
For any x e C \ C0 with x ::1: a 1, there is a unique point x 0 e C 0 such that
x e (a 1,xo). If Xo ::/:co we put cj>(x) = x', where x' = (a 1 ',xo) ri (x,c 0 ). If
x e (a 1,c0 ), there is a unique x" e (a 1,a0) such that x e (x",a 0 ') and we define
cj>(x) = (cj>(x"),a0 ') ri (a 1',co). D

A subset of a linear geometry is said to be a simplex if it is the convex hull of


an affine independent set. Our labours so far in this chapter are essentially
summarised in the following characterization of simplices:

THEOREM 3 Let X be a dense linear geometry with dim X > 2, or with


dim X = 2 and the Desargues property. Then there is an ordered division ring D
such that any simplex C s X is isomorphic to a simplex C in a vector space i.
over D.

Proof We have already shown that there exists an ordered division ring D such
that, for any affine independent set A s X with IAI = 2, there is an affine
independent set A s D with IAI = 2 such that [A] is isomorphic to [A].
Suppose now that A is an affine independent set with IAI > 2. Let ?J be the
family of all triples (B, i. J), where B s A, i. is a vector space over D and/ is an
isomorphism of [B] onto a simplex in X. We partially order ?J by writing
(B, i. J) ~ (B', i. ',g) if B s B', i. is a subspace of X 'and g(x) = f(x) for x e [B].
Let ~ = { (B ;. X; /;)} be a maximal totally ordered subfamily of ?J and put
B =U B ;. X = U X;. If x e [B], then x e [B ;] for some i and we can define
/: [B] ~ i. by f(x) =f;(x). Moreover f(B) is an affine independent set, since any
finite subset is affine independent, and hence f is an isomorphism of [B] onto the
146 VII. Vector spaces

simplex [f(B )] = U [/(B ;)]. Thus (B. i. .f) is an element of ?J and


(B;. X; /;) =Ei (B. i. /) for all i. H B = A we are finished. If not, then there exists a
point a 1 e A \B. If i.' is a vector space over the division ring D which properly
contains i., then, by Proposition 2, there exists an element (Bu a 1,i. ',g) of ?J
such that (B. i. J) ~ (B u a 1, i. ',g). But this contradicts the maximality of 5". D

In Theorem 3 the whole space X need not be isomorphic to a convex subset of


a vector space over an ordered division ring (even if Xis also unending). This is
shown by Example Vl.6, since any two coplanar lines of a projectiv~ space
intersect. However, the whole space X may be characterized in the following way:

THEOREM 4 Let X be a dense linear geometry with dim X > 2, or with


dim X = 2 and the Desargues property. Then there exists a vector space V over
an ordered division ring D such that X is isomorphic to a set X' in the
projective completion V of V where, for any y',z' e X' with y' ~ z', (y',z1 is one
of the two closed segments with endpoints y',z' of the 'line' <y',z'>.

Proof Let E = {e;: i e /} be an affine basis for X. By Theorem 3 there exists an


isomorphism f of the simplex C = [E] onto a simplex C' = [E1 in a vector space V
over an ordered division ring D. Moreover we may assume that E' =
{e;' =/(e;): i e /} is an affine basis for V. Let V be the projective completion of V.
If x e X \ C, then there exist lines through x which contain more than one point
of C. For let {e 1 , ••• ,en} be a finite subset of E, with n ~ 3, such that
x e <e 1, ••• ,en> and put P = [e 1, ••• ,en1· For any p e pi the line <p,x> intersects P
in a nondegenerate segment, since x ~ P.
For any x e X let a,p be distinct lines through x, each containing more than
one point of C, and let a', P' be the images under f of an C, Pn C. The distinct
coplanar lines <a'>, <P'> intersect in a unique point x' of V. We are going to
show that x' does not depend on the choice of lines a,p. Clearly it is sufficient to
show that if y is a third line through x containing more than one point of C, and y'
the image under f of y n C, then x' e <y'>.
Let a 1,a2 be points of an C with a 2 e (x.a 1), let b 1,b2 be points of Pn C
with b 2 e (x.b 1), and let c 1,c 2 be points of y n C with c 1 e (x.c 2 ). Choose
d 1 e (a 1,b 1) and then d 2 e (x.d 1) so that d 2 e (a 2,w) for some we (d 1,b 1) and
d 2 e (b2, w) for some we (a 1,d1). Then d 1,d2,w, we C and
2. Isomorphisms 147

Let 8' be the image under/ of 8 n C, where 8 =<d1,d2>. Since the intersections

exist and are points of C, and since the lines <a 1,a2>,<c 1,c2>,<d1,d2> have the
common point x, it follows from the Desargues property that u, v, w are collinear
(see Figure 5). Under f the points a;, ... ,d; are mapped into points a/, ... ,d;' of C'

Figure 5

with preservation of segments and collinearity. Thus u',v',w' are collinear points of
V. Since V has the converse Desargues property (Proposition Vl.3), it follows
that the lines <a'>,<y'>,<8'> have a common point of intersection. Since the
intersection u = (b 1,c 1) n (b 2,c2) also exists and is a point of C, it follows in the
same way that the lines <P'>,<y'>,<8'> have a common point of intersection.
Hence the lines <a'>,<P'>,<y'> have a common point of intersection.
We now extend the definition off to X by putting/(x) = x'. We are going to
show next that distinct collinear points of X are mapped by f into distinct collinear
points of V. This implies, in particular, that/ is injective.
Let x,y,z be points of X such that x e (y,z) and let x',y',z' be their images
under/. There exists a finite subset {elt···,en} of E, with n ~ 3, such that
y,z e <e 1, ••• ,en>. Put P = [e 1, ••• ,en1 and choose p e pi with pi! <y,z>. There
exist points i, y, z e C such that p e (x, i), ye (p,y), ze (p,z). There exist also
148 Vil. Vector spaces

points x = (p,x) (y, z), y 1 e (i, z) such thatp e (y,y 1), and z1 e (i, y) such
f"\

thatp e (z ,z 1). Choose ye (y,z 1) and let x 1 e (i,y 1) be such that ye (y,x 1).
If we put a 1 = p, then there exist points

Let a2 e (z,c 1) be such that b 1 e (y,a2), let y2 e ( z ,y 1) be such that a 2 e (b 1,y 2),
and let z2 e (i,z 1) be such that c 1 e (a 2,z2). Finally, if we choose x 2 e (i,z 2),
there exist points

In this way (see Figure 6)

y x z

Figure 6

a; = (y,y;) f"\ (z,z;), b; = (z,z;) f"\ (x,x;), C; = (x,x;) f"\ (y,y;) (i = 1,2)
and
2. Isomorphisms 149

are distinct collinear points, it follows from Proposition VI.3 that the segments
(a 1,a2), (b1,b2), (c 1,c2) have a common point q. Since i, y, ze C, it follows that

x = <b t ,c1 > n <b2 ,c2 >, Y = <a1 ,c1 > n <a2 ,c2 i>, z = <a1 'bi
I I i I i I I i I I
' t > n <a2''b 2 > I

and <a 1',a2'>, <b 1',b2'>, <c 1',c2'> have the common point q' ~ a;',b;',c;' (i = 1,2).
Hence x',y',z' are distinct and collinear, by the Desargues property in V.
It will now be shown that/ maps the segment [y,z] of X onto the segment of
the line <y',z'> in V which has endpoints y',z' and contains x'. Suppose that
w e (y,z), where w ~ x, and put w = (p, w) n (y, z). Then the lines <y', y '>,
<z',z'>,<x',x'>,<w',w'> have the common point p' and x',w' e <y',z'>,
x',w' e <y',z'>. Since x ', w 'lie in the same open segment with endpoints y', z ',
it follows that x', w' lie in the same open segment with endpoints y',z'.
On the other hand, if w' is any point of the open segment with endpoints y',z'
which contains x', the point w' = <p',w'> n <y', z'> lies in the open segment
with endpoints y ', z' which contains x '. Hence w ' = /( w ) for some w e ( y, z)
and there exists a point we (y,z) such that w e (p,w). It follows that w' = /(w).
Thus f maps X isomorphically onto X' = f(X). D

Although in Theorem 4 the whole space X need not be isomorphic to a convex


set in a vector space over an ordered division ring, it will now be shown that any
open half-space of X is isomorphic to such a convex set.

THEOREM 5 Let X be a dense linear geometry with dim X > 2, or with


dim X = 2 and the Desargues property. If H+ is an open half-space of X
associated with a hyperplane H, then H + is isomorphic to a convex subset of a
vector space V over an ordered division ring D.

Proof Let E = {e;: i e /} be an affine basis for X such that H = <E\e 1> and
e 1 e H+· By Theorem 3 there exists an isomorphism/ of the simplex C = [E] onto
a simplex C' = [E1 in a vector space Vover an ordered division ring D. Moreover
we may assume that V = <E'>. By Theorem 4 and its proof we may extend f to an
isomorphism of X into the projective completion V of V. The hyperplane
<E' \ e 1 '> of V extends to a hyperplane H' of V. We will show that
f<H+) n H' = 0.
Assume on the contrary thatx': =f(x) e H' for somex e H+. Let {e 1, ••. ,en}
be a finite subset of E, with n > 3, such that x e <e 1, ••• ,en>. Put P = [e 1, ••• ,enl
150 Vil. Vector spaces

and Q = [e 2 , ••• ,en1· If h e Qi then <x,h> n P = [h,h+], where h+ e H +· Thus


there exist points y,z e (h,x) n C with y -:1: z. If h',y',z' are the images of h,y,z
under f, then h',y',z' are collinear and so are x',y',z'. Since y',z' E H', it follows
that h' = x'. Since this holds for any he Qi and/ is injective on C, we have a
contradiction.
By definition, V = P(V*), where V* is a vector space over the division ring
D. There exists a bijective linear map y: V* ~ V* inducing a bijective map
g: V ~ V such that g(H') is the hyperplane at infinity of V. Then g 0 f maps H +
isomorphically onto a convex subset of V. [J

In the statement of Theorem 5 we cannot replace 'open half-space' by 'closed


half-space'. This is shown by Example VI.7. In fact, since any two coplanar lines
of the projective space X = pn (11 > 2) intersect, X is not isomorphic to a convex
subset of a vector space over an ordered division ring. However, X is itself the
closed half-space associated with the hyperplane pn-1.

3 VECTOR SUMS

Throughout this sectio11 we assume that X is a vector space over an


ordered division ring D with the standard alignment of convex sets.
For any nonempty sets A,B ~ X and any a e D, we define

A+B = {a+b:ae A,be B},


a.A = {aa: a e A}.

The set A+ B will be called the vector (or Minkowski) sum of A and B, and aA
will be called a scalar multiple of A.
It is easily verified that the following properties hold for any nonempty sets
A,B ,C ~ X and any a,~ e D:

(i) A + B = B +A, (A + B) + C =A + (B + C),


(ii) a(A + B) =aA + a.B, (a~)A = a(~A), IA =A.

The singleton 0 = { 0} acts as a zero element:

(iii) A + 0 = A, OA = 0.
3. Vector sums 151

Furthermore,

(iv) if A~ B, then A+ C ~ B + C and aA c aB.

We consider now the case where the sets are convex.

LEMMA 6 If A,B are nonempty convex subsets of X and a e D, then A + B


and aA are also nonempty convex subsets of X. Moreover

(v) (A. + µ)A = A.A + µA for any A.,µ e D with A.,µ > 0.

Proof Since A,B are convex, for any A.,µ > 0 with A. + µ = 1 we have
A.A+ µA~ A and A.B + µB ~ B. Hence

A.(A + B) + µ(A + B) = (A.A + µA) + (AB + µB) ~ A + B


and
A.(aA)+ µ(aA) = a(a-lA.aA + a-lµaA) !;:; aA.

Thus A + B and aA are convex.


For any nonempty set A X and any A.,µ e D we have (A. + µ)A ~
!;:;

AA + µA. Hence to prove the last statement of the lemma we need only show that
if A is convex and if A.,µ > 0, A. + µ > 0, then A.A +µA !;:; (A. + µ)A. But if
x ="A.a.'+ µa", where a',a" e A, then

x =(A.+µ){ (A.+ µ)-lA.a' +(A.+ µ)-lµa"} e (A.+ µ)A. D

It may be noted that A is necessarily convex if (A.+ µ)A= A.A+ µA for all
A.,µ ~ 0, and that if IAI > 1, A. > 0 then

0 = (A.+ (-A.))A ~ AA + (-A.)A.

It follows by induction from Lemma 6 that, for any nonempty convex set A
and any positive integer n,
nA = A+ ... + A,

where there are n summands on the right

PROPOSITION 7 If S,T ~ X, then [S + 71 = [S] + [71.

Proof If x e [S + 71, then there exist Sj e Sand tie T (j = l, ... ,n) such that
152 VII. Vector spaces

x = l:-}=i 'Aj(sj + ti),


where 'Ai > 0 (j = l, ... ,n) and I,}=i 'Ai = 1. Hence x = y + z, where
y = l:-j=i Afj e [S] and z = l:-j=i 'A/j e [T]. Thus [S + 71 c [S] + [T].
On the other hand, if x e [S] + [71, then there exist s; e S (i = 1,... ,m) and
tie TU= l, ... ,n) such that
x = I-i'!i A;Si + I-j=i µli'
where A; ~ 0 (i = l ,... ,m ), µj > 0 (j = 1,... ,n) and I-~i A.i = 1 = I,}=i µj. Hence

x = I-i'!i I-'J=i 'Aiµj(s; + ti)


and x E [S + 71. Thus [S] + [71 ~ [S + T]. a
PROPOSITION 8 If A and Bare collvex sets such that Ai-¢ 0 and Bi-¢ 0, thell

(A + B)i = Ai + Bi.

Proof We show first that Ai+ Bis;: (A + B)i. Let z = x + y, where x e Ai and
ye Bi, and let c =a+ b, where a e A, be Band c '¢ z. There existsµ> 1 such
that, for 1 < 'A < µ,

x' =Ax+ (1 - 'A)a E A, y' = A.y + (1 - 'A)b E B.

Then, for 1 < 'A < µ,


'Az + (1 - 'A)c = x' + y' e A + B

and hence z e (A + B)i.


We show next that (A+ B)i s: Ai+ Bi. Let x e Ai and ye Bi. Then
z = x + y e (A + B)i by what we have just proved. Suppose w e (A + B)i and
w '¢ z. Since z e A + B, we have we (z,c) for some c e A + B. Thus c =a + b,
where a e A and be B, and, for some 0 such that 0 < 0 < l, w = 0z + (1 - 0)c.
Hence w =x' + y', where x' =ex+ (1 - 0)a E Ai and y' = 0y + (1 - 0)b E Bi.
Thus we Ai +Bi. [J

PROPOSITION 9 If Ci and C2 are nollempty convex sets, and if F is a


nonempty face of C =Ci + C2 , then there exist nonempty faces Fi of Ci a1Zd
F 2 of C2 such that F =Fi+ F 2.

Proof Put
3. Vector sums 153

F 1 = {xe C 1:x+ye Fforsomeye C 2 },


F 2 = {ye C 2:x+ye Fforsomexe Ci}.

Evidently F 1 and F 2 are nonempty convex sets.


Suppose X1 e F 1 and X1 = 0c1 + (1 - 0)c1 ',where C1,C1, e c 1 and 0 <a < 1.
Then x 1 +ye F for some ye C 2. Since x 1 + y = 0(c 1 + y) + (1 - 0)(c 1 ' + y) and
Fis a face of C, it follows that c 1,c1' e F 1. Hence F 1 is a face of C 1, and similarly
F 2 is a face of C2 •
Obviously F ~ F 1 + F 2 . On the other hand, if x 1 e F 1 and y 2 e F 2 , then
there exist y 1 e C2 and x2 e C 1 such that x 1 + y 1 e F and x2 + y2 e F. Hence

Since Fis a face of C, it follows thatx 1 + y 2 e F. Thus F 1 + F 2 ~ F. D

PROPOSITION 10 Let C =A+ B, where A and Bare nonempty convex sets.


If z is an extreme point of C, then it has a unique representation in the form
z = x + y, where x e A and ye B. Moreover, in this representation x is an
extreme point of A and y an extreme point of B.

Proof By Proposition 9 there exists a representation z = x + y, where x is an


extreme point of A and y an extreme point of B. Suppose also that z =x' + y',
where x' e A and y' e B. Since z = (l/2)(x + y') + (l/2)(x' + y) and z is an
extreme point of C, we must have x + y' = x' + y. Then 2z = 2x + y + y' =
2x' + y + y' and hence x =x', y =y'. [J

PROPOSITION 11 Let A,C,P be nonempty sets, with A arbitrary, C convex and


Pa po/ytope. If A+ P !::: C + P, then A~ C.

Proof We wish to show that if a+ P !::: C + P, then a e C. Let P = [p 1, ••• ,pn] be


the convex hull of n points. Since the result is obvious if n = 1, we assume that
n > 1 and the result holds for polytopes which are the convex hulls of fewer than n
points. Hence if we put Q = [p 2, ••• ,pn], it is sufficient to show that a + Q c C + Q.
If q e Q, then a + q = c + p for some c e C and p e P. Similarly
a+ p 1 = c' + p' for some c' e C andp' e P. But P = [p 1 u Q]. Hence if p e: Q,
then p = a.p 1 + (1 - a.)q', where q' e Q and 0 < a. s; 1. Similarly if a'¢ c', then
p' = ~p 1 + (1- ~)q", where q" e Q and 0 < ~ < 1. It follows that

P1 = q" + (1 - ~)- 1 (c' - a)


154 VII. Vector spaces

and hence
a+ q = c + aq" + a(l - p)-l(c' - a) + (1 - a)q'.
Consequently

(1 + a(l - p)-l](a + q) = c + a(l - p)-lc' + aq" + (1 - a)q' + a(l - p)-1q.

Since C and Q are convex, this shows that a + q e C + Q. Thus a + Q !::: C + Q. [J

PROPOSITION 12 If P 1 and P 2 are polytopes, then P = P 1 + P 2 is also a


polytope. Moreover any nonempty face F of P has a unique representation
F = F 1 + F 2, where Fk is a 11onempty face of Pk (k = 1,2).

Proof It follows at once from Proposition 7 that P is a polytope and from


Proposition 9 that any nonempty face F of P has a representation F = F 1 + F 2,
where F 1 is a nonempty face of P 1 and F 2 a nonempty face of P 2. We suppose F 1
and F 2 defined as in the proof of Proposition 9.
Suppose F = F 1' + F 2', where F 1' is a nonempty face of P 1 and F 2 ' a
nonempty face of P 2• Then Fk, ~ Fk (k = 1,2), by the definition of F 1,F2 in the
proof of Proposition 9. Since

it follows that F 1' + F 2 = F = F 1 + F2. Hence, by Proposition 11, F 1' = F 1 and


similarly F 2' = F 2• CJ

PROPOSITION 13 If Pis a polytope and A,B nonempty convex sets such that
A+ B = P, then A and Bare also polytopes.

Proof Let P = [p 1, ... ,pn1· Then Pk= ak +bk for some ak e A and bk e B
(k = l, ... ,n). If A'= [a 1, ••• ,an] and B' = [b 1, ••• ,bn], then A'!::: A, B' ~Band
hence A'+ B's:: P. On the other hand, Pk e A'+ B' (k = l,... ,n) and A'+ B' is
convex, by Lemma 6. Hence P s:;: A ' + B '. Thus P = A ' + B ', and so
A + B' ~A'+ B'. Hence A ~A', by Proposition 11. Consequently A =A', and
similarly B = B'. D

4 NOTES

It was first shown by von Staudt (1856/57) that it is possible to introduce


coordinates in space, independently of the notions of distance and congruence. His
4. Notes 155

discussion was based on the 'fundamental theorem of projective geometry', which


may be given the following formulation:

If a map of a line to itself is the product of finitely many perspectivities


and has at least three fixed points, then it is the identity map.

Some interesting remarks on the fundamental theorem of projective geometry are


contained in Pickert (1981) and Frank (1992).
Hilbert ( 1899) showed that the Desargues property could be used instead of
the fundamental theorem of projective geometry. In a noteworthy (but rather
inaccessible) paper, Spemer (1938) showed that Hilbert's argument could be
adapted to any unending two-dimensional linear geometry with the Desargues
property. Doignon (1976), using ideas of Busemann (1955), extended Spemer's
work to higher dimensions and proved Theorems 3-5 for any unending linear
geometry. The more general treatment here for dense, rather than unending, linear
geometries presents little difficulty after the corresponding extension in Proposition
VI.4.
The division ring in these theorems is a field, i.e. multiplication is
commutative, if and only if the fundamental theorem of projective geometry holds
or, equivalently, if and only if the linear geometry has the following 'Pappus
property':

Suppose a 1,b 1,c 1 are distinct collinear points, and a 2 ,b 2 ,c 2 are distinct
collinear points not on the same line; if <a 1,b 2> intersects <a 2 ,b 1>, <b 1,c 2 >
intersects <b 2 ,c 1>, and <c 1,a 2 > intersects <c 2 ,a 1>, then the three points of
intersection are collinear.

Furthermore the Pappus property actually implies the Desargues property. (It is
pointed out in Seidenberg (1976) that the original proof of this by Hessenberg
(1905) was incomplete.) Proofs of these results for ordinary projective space are
given in Heyting (1980), for example. We do not establish them here, since
commutativity of multiplication will be a consequence of the additional axiom (S) in
Chapter VIIl.
Proposition 10 appeared in Klee (1959). Proposition 11 is a topology-free
version of RAdstrom's cancellation law (Proposition IX.4). It may be noted that
Proposition 11 no longer holds if 'polytope' is replaced by 'polyhedron'. For
156 VII. Vector spaces

example, if X = JR.2, A = {(1,1)} and C = {(0,0)} are singletons, and P =


{ (x,y) e JR.2: x,y > O} is the first quadrant, then

A+PcP=C+P.

Bair and Fourneau (1975/80) contains additional results on vector sums, and
additional references.
VIII
Completeness

In this chapter we introduce a final axiom of completeness, which is an extension


to linear geometries of the Dedekind cut property for the real line. Indeed it is
shown that, in any complete dense linear geometry of dimension> 2, a segment
containing more than one point is isomorphic to the interval [0,1] of JR. It is further
shown that, in any complete dense unending linear geometry, two convex sets with
disjoint nonempty intrinsic interiors can be properly separated by a hyperplane.
This in turn implies the geometric form of the Hahn-Banach theorem and a
separation theorem for any finite number of convex sets. These results hold even if
the linear geometry is two-dimensional and does not have the Desargues property.
We prove next the fundamental theorem of ordered geometry that any
complete unending linear geometry of dimension > 2 with the Desargues property is
isomorphic to a convex set, which is its own intrinsic interior, in a real vector
space. The fundamental theorem may be established in other ways than that used
here, but they have the disadvantage of requiring dimension > 3. The necessity of
the various hypotheses of the theorem is illustrated by examples. It is somewhat
surprising that a finite-dimensional projective space over any ordered division ring
other than JR can be given the structure of a dense unending linear geometry. At the
end of the chapter we recall, for use in Chapter IX, some well-known properties of
metric spaces and nonned vector spaces.

1 INTRODUCTION

A linear geometry X will be said to be complete if the following axiom is


satisfied:
158 VIII. Completeness

(S) if a,b e X and if C is a convex subset of [a,b] such that a e C, then there
exists a point c e [a,b] such that either C = [a,c] or C = [a,c).

The axiom's label is chosen to indicate that c 'separates' the points of [a,b]
which are in C from those which are not. It is easily seen that the point c is
uniquely determined if the linear geometry X is dense. On the other hand, if
[a,b] = {a,b} for some distinct a,b e X and if C ={a}, then we can take either c =a
or c = b.
The axiom (S) may be regarded as an extension to linear geometries of the
Dedekind cut property for the real line JR. Any interval of JR is a complete dense
linear geometry. However, a segment in a complete dense linear geometry need not
be isomorphic to an interval of lR:

EXAMPLE 1 Let W be an uncountable set, e.g. the set of all sequences (an),
where an e {0, 1}. By the well-ordering principle, there is a total ordering of W
such that every nonempty subset of W has a least element. (The well-ordering
principle may be deduced from Hausdorffs maximality theorem; see, e.g., Hewitt
and Stromberg (1975).) Let Z be the set of all z e W for which the set
{w e W: w < z} is uncountable. If Z is nonempty, it has a least element ~- Put
Y = {y e W: y < ~} if Z ~ 0 and Y = W otherwise. Then Y is an uncountable
totally ordered set with the following properties:

(i) every nonempty subset of Y has a least element,


(ii) for each y e Y, the set {y' e Y: y' < y} is countable,
(iii) Y has no greatest element

The long line A is defined to be the set of all pairs (y ,p ), where y e Y and
p e (0,1), except for the pair (y0 ,0) where Yo is the least element of Y, with the
lexicographic ordering defined by (y 1,p 1) -< ()'2,P2) if Y1 < Y2, or if Y1 = Y2 and
p 1 < p 2. Since A is a dense totally ordered set with no least or greatest element, it
is a dense unending linear geometry with the definition of segments given in
Example VI.1. Furthermore the separation axiom (S) is satisfied. Since A has no
countable dense subset, it is not isomorphic to JR. However, any segment of A
containing more than one point is isomorphic to the interval (0,1] of JR.
Finally define a totally ordered set X = U ~=l X;, where X 1 and X 3 are copies
of the long line A, X2 = {p} and Xj < xk if xi e Xi, xk e Xk U < k). Then Xis a
1. Introduction 159

complete dense unending linear geometry, but the segment [a,b] is not isomorphic
to an interval of R. if a e X 1 and b e X \ X 1•

In Example 1 all points of X are collinear. It will now be shown that if X is a


complete dense linear geometry, not all points of which are collinear, then any
segment of Xis isomorphic to an interval ofR.. We first prove

LEMMA 1 Let S be a countable dense totally ordered set with no least or


greatest element. Then S is order isomorphic to the set Q of rational numbers.

Proof Let {s 1,s2, ••• } and {r1,r2 , ••• } be enumerations of Sand Q respectively.
We wish to construct a bijective map <p: S ~ Q such that <p(sj) ~ <p(sk) according
as sj ~ sk for all distinctj,k e N. The construction is inductive.
Put <p(s 1) = r 1 and write sl = s 1, rl = r 1• Suppose we have defined subsets
sn = {s•, ... ,sn}, Qn = {r•, ... ,rn} of S, Q respectively and a bijective map
cp: Sn ~ Q n such that <p (si) ~ cp (sk) according as si ~ sk for all disinct
j,k e { l ,... ,n}. If n is odd define rn+ 1 to be the rational number rk with least k
which is not in Qn. Then either rn+ 1 is less than all elements of Qn, or rn+ 1 is
greater than all elements of Qn, or there exist two elements ri,rk of Qn such that
ri < ,.n+I < rk, but ri <,.; < rk for no ri e Qn. We define sn+l to be the element sk of
S with least k which is correspondingly less than all elements of sn, or greater than
all elements of Sn, or satisfies si < sn+l < sk, and we put cp(sn+l) = rn+l. If n is
even we interchange the roles of Sand Q. Thus we define sn+l to be the element sk
of S with least k which is not in Sn and we define ,.n+ I = <p(sn+ I) to be the rational
number rk with least k which has the same order relation with respect to Qn as sn+l
has with respect to Sn. Since Sand Qare both dense, and have no least or greatest
element, this construction is always possible. Moreover each element of S and Q is
ultimately included. a
PROPOSITION 2 If X is a complete dense linear geometry with dim X > 2,
then every nondegenerate segment of X is isomorphic to the interval [0, 1] of R..

Proof Let a and b be distinct points of X and choose Yoe (a,b). Define a
sequence {Yml recurrently by Ym e (a,ym-l) (m > 1). If we suppose the segment
[a,b] totally ordered in accordance with Proposition III.4 so that a< b, then {Yml is
a decreasing sequence. Denote by C the set of all points x e [a,b] such that x < Ym
for all m. Then C is convex, a e C and be: C. Hence, by (S), there exists a point
160 VIII. Completeness

a' e [a,b) such that [a,a') s C and (a',b] s X \ C. By Proposition VII.I it is


enough to show that the segment [a',b] is isomorphic to an interval of the real line.
Thus we now take a =a'. Then for each x e (a,b) there is a positive integer m such
that y,,, e (a,x). By Proposition Ill.4 we need only show that there is a bijective
map of (a,b) onto R which preserves order.
We show first that, for any ye (a,b), there exists an lsomorphism.fy of [a,b]
onto [y,b] such that.fy(a) =y andfy(b) =b. Choose d ~ <a,b> and c e (b,d). For
any x e (a,b), there exists a unique x' e (a,c) r. (d,x). If we put / 1(x) = x',
/ 1(a) =a,f1(b) =c, then/1 is an isomorphism of [a,b] onto [a,c]. For any
x' e (a,c), there exists a unique y' e (c,y) r. (b,x'). If we put/2(x') = y',f2(a) = y,
/ 2 (c) = c, then/2 is an isomorphism of [a,c] onto [y,c]. For any y' e (c,y), there
exists a unique z e (y,b) such that y' e (d,z). If we put/3 (y') = z,/3 (y) = y,
then/3 is an isomorphism of [y,c] onto [y,b]. Evidently fy =/ 3 °/ 2 °/ 1 is
/ 3(c) = b,
an isomorphism of [a,b] onto [y,b] such that/y(a) = y andfy(b) = b. (In the
notation of Chapter VII, /y(x) =x + y.) Moreover /y(x) e (x,b) for any x e (a,b),
since y' e (x',b), and for any z e (x,b) there exists ye (a,z) such that.fy(x) = z (see
Figure 1). Furthennore.fy(x) </y(x) if x < x, and/y(x) < f-y(x) if y < y.

x y z

Figure 1: z = /y(x)

For any ye (a,b), the iterates y(O) = y, y(l) = !y(y), y(2) = f//y(y)), ... form an
increasing sequence. We now show that, for each x e (a,b), there exists a non-
negative integer n such that y(n) > x. Assume on the contrary that y(n) < x for all
n > 0. It follows from (S) that there exists z e (a,b) such that y<n> S z for all n > 0,
1. Introduction 161

but y<n> > z' for some n > 0 if z' < z. Since z e (y,b), there exists we (a,b) such
that fy(w) = z. Then w < z and hence y(n) > w for some n > 0. Consequently
y(n+l) > /y(w) = z, which is a contradiction.
We show next that the set E of all points Ym(n), where {Yml is the sequence
constructed at the outset and n is any non-negative integer, is dense in [a,b].
Suppose u,v e (a,b) and u < v. Then there exists t e (a, v) such thatf,(u) = v.
Choose m so that y m < t and y m < u. There is an integer p > 0 such that
Ym(p) < u < Ym(p+l). Then Ym(p+l) <fy(u) <f,(u) = v. Thus Ym(p+l) e (u,v), as we

wished to prove.
Thus (a,b) is a totally ordered set with no least or greatest element and with a
countable dense subset E. Consequently, by Lemma 1, there is a bijective order-
preseiving map cp of E onto the set Q of rational numbers. For any c e (a,b) \ E,
put
A = {x e E: x < c}, B = {x e E: x > c}.

Then A,B are disjoint nonempty subsets of E and E = A u B. Hence Q is the


union of the disjoint nonempty subsets cp(A),cp(B), and a < p for every a e cp(A),
Pe cp(B). By the Dedekind property for JR, there is a unique ye JR \ Q such that
a< y < P for every a e cp(A), Pe cp(B). It is easily seen that if we extend the
definition of cp to the whole of (a,b) by putting cp(c) = y, then cp is a bijective order-
preseiving map of (a,b) onto JR. D

Suppose X is the open unit disc in the Euclidean plane, with the usual
definition of segments. Through any point not on a given line there pass infinitely
many lines which do not intersect the given line. However, there are two 'special'
ones, namely the intersections with X of the lines in R 2 which pass through the two
endpoints on the unit circle of the given line. This example may be borne in mind in
the statement and proof of the following proposition.

PROPOSITION 3 Suppose X is a complete unending linear geometry. Let


a 1,a 2,b 1 be non-collinear points of X and let H+ be the open half-plane of the
plane TI= <a 1,a2,b 1> associated with the line H = <a 1,b1> which contains a 2 •
Then there exists b2 e H + such that, for any x e H +' [a 1,a 2> '1 [b 1,x> "# 0
if and only if x lies in the open half-plane ofll associated with the line <b 1,b2>
which contains a 1•
162 V/11. Completeness

Furthermore <a 1,a2> t\ <.b 1,b2> = 0 and, if the lines <a 1,a2>,<b 1,b2> are
ordered so that a 1 < a 2, b 1 < b2, then for any a',a" e <a 1,a 2> with a'< a" and
any b',b" e <b 1,b2> with b' < b", b" has the same properties with respect to
a',a",b' as b 2 has with respect to a 1,a2,b 1.

Proof Choose w so that b 1 e (a 1,w). Then, for any x e H +' the ray [b 1 ,x>
contains a unique point c e (a 1,a2) t\ [a 2,w). We wish to show that there exists
b 2 e (a 2,w) such that [b 1,c> intersects [a 1,a2> if and only if c e [a 2,b2).
Let C be the set of all c e [a 2,w) such that [b 1,c> intersects [a 1,a 2>.
Evidently a2 e C and C is convex. If we choose a 3 so that a2 e (a 1,a3), then there
exists a point u e (b 1 ,a 3 ) t\ (a 2 ,w) and u e C. If we choose a 0 so that
a 1 e (a 0 ,a2), then b 1 e (a 0 ,v) for some v e (a 2,w) and v e: C. It now follows
from (S) that there exists b 2 e (a 2,w) such that C = [a 2,b2] or C = [a2 ,b 2). It
remains to show that b2 e: C.
Assume on the contrary that there exists a pointp e [a 1,a2> t\ [b 1,b2>. Then
b 2 e (b 1,p) and a2 e (a 1,p). If we choose a 3 so thatp e (a2,a3), then there exist a
point q e (a3,b 1) t\ (w,p) and a point re (b 1,q) t\ (w,b 2). Hence re Ct\ (TI\ C),
which is a contradiction.
Suppose the lines t = <a 1,a2>, m = <.b 1,b2> ordered so that a 1 < a2, b 1 < b2•
Since b2 E H+, we have both [a1,a2> (\ [b1,b2> = 0 and e(') [b1,b2> = 0. We
will show that actually e(\
m = 0.
Assume on the contrary that there exists a point p e t (') m. Then p < b 1 on
m, and hence p <al on .f. If we choose ao E .f SOthat ao < p, then b1 E (ao,Q) for
some q E (a2,bi). Hence there exists a point r E e(') [b1 ,q>. Since b1 E (ao,r)
and b 1 e: t, this is a contradiction.
Let a',a" be points oft with a'< a" and let b',b" be points of m with b' < b".
It is evident that if a'= a 1 and b' = b 1, then b" has the same properties with respect
to the ordered triple a',a",b' as b 2 has with respect to the ordered triple a 1,a2,b 1.
The same conclusion holds if b' = b 1 and a 1 <a', since [a',a"> (') [b 1,x> = 0 for
any x e H_ in the open half-plane of TI associated with the line <a',b 1> which
contains a". The same conclusion holds also if a'= a 1 and b 1 < b', since
[a 1,a"> (') [?J',x> "# 0 for any x e H_ in the open half-plane of TI associated with
the line <a 1,b'> which contains a". Since t "m = 0, it follows from what has
already been proved and the total ordering of the lines t,m that the same conclusion
holds also in the general case. []
1. Introduction 163

The uniquely determined ray [b1tb2> will be called the asymptote through b 1
to the ray [a 1,a2>. The second part of Proposition 3 shows that it depends only on
the oriented line <a 1,a2>, and not on the choice of a1,a2• It further shows that, in a
sense, it depends only on the oriented line <b 1,b2>.
In the same way, reversing the orientation on <a1ta 2>, there exists an
asymptote [b1tb 2 t> through b 1 to the ray [a 2 ,a 1>. In Euclidean space
<b 1,b2> =<b 1,b2t>, but the restriction of the Euclidean plane to the open unit disc
already shows that this need not be true in general.
Proposition 3 shows that, for any line eand any point p ~ e' there is at least
one line m in the plane <p u .f> such that p e m and t r. m = 0. Thus we may
regard Proposition 3 as a replacement for Euclid's parallel axiom. However, unlike
parallelism in Euclidean space, the asymptote relationship may be neither symmetric
nor transitive. To see this, in the second specific case of Example VI.5 take
a 1 = (-1,1), a2 = (-1,2) and b 1 = (0,1), b2 = (0,2). Then <a 1,a2> r. <b 1,b2> = 0
since, in the notation of Chapter VI, <a 1,a 2 > = 11,- 1 and <b 1,b 2 > = IL 0 •
Furthermore, if a e (a 1,a2> then [a,bi] is the usual straight line segment, since
a = (-1,a) for some a > 1 and hence <a,b 1> = .fl-a,l. It follows that [b1tb 2> is
the asymptote through b 1 to the ray [a 1,a2>. However, if a 3 = (-1/2,2), it is easily
seen that the asymptote through a 1 to [b 1,b2> is the ray [a 1,a3> of 9.

PROPOSITION 4 If X is a complete unending linear geometry with dim X ~ 2,


then every line of X is isomo1phic to the real line R.

Proof It follows from Proposition 2 that, for any distinct points a,b e X, (a,b) is
order isomorphic to (0,1) and hence to R. We need only show that an arbitrary line
<a 1,a2> of X is order isomorphic to R.
Let b 1 be a point not on the line <a 1,a2>, choose w so that b 1 e (a 1,w), and
let b2 e (a 2,w) be defined as in the proof of Proposition 3. Also, choose a',a" so
that a 2 e (a 1 ,a ') and a 1 e (a 2 ,a"). By the proof of Proposition 3,
(a 2,a'> = [a 2,a'> \ a2 is order isomorphic to (a 2 ,b2), and hence to Rand to (1, 00) .
Similarly (a 1,a"> is order isomorphic to (a 1,b2t), and hence to Rand to (- oo,0).
Since [a 1,a2] is order isomorphic to [0,1], the result follows. a
164 VIII. Completeness

2 SEPARATION PROPERTIES

Throughout this section we will assume that a set X is given on which a


complete dense unending linear geometry is defined. We draw attention to the
fact that we do not assume the Desargues property. Although the Desargues
P.roperty necessarily holds if dim X > 2, the results which will be established are
also valid if dim X = 2 and the Desargues property does not hold.
The notion of 'convex partition' was introduced in Chapter II. It will now be
shown that, under the present hypotheses, convex partitions are 'generally'
determined by a hyperplane.

PROPOSITION 5 If C, Dis a co11vex partition of X such that Ci u Di"# 0,


then Ci, Di are both nonempty and H: = X \(Ci u Di) is a hyperp/a11e.
Moreover, the two open half-spaces associated with Hare Ci, Di a11d the two
closed half-spaces associated with H are C, D.

Proof By Proposition V .28, Ci "# 0 and Di "# 0. It follows from Proposition


V.28 also that C = X\Di, D =X\Ci and hence H = C n D. Thus His convex.
In fact His affine. For suppose a,b e Hand a e (b,c). Then c ~ Ci, since be C
and a ~ Ci. Similarly c ~ Di, and thus c e H.
Let c e Ci and de Di. By (S), the segment [c,d] contains a pointx such that
[c,x) s Ci and (x,d] s X\Ci. Evidently we must havex e H. It follows first that
Di s <.JI u c> and then that Ci s <.JI u c>. Hence <.JI u c> = X and H is a
hyperplane. Since Ci and Di are convex and partition X\H, they are the open half-
spaces associated with H. Hence, by Proposition V.28, C = Ci u H and
D=DiuH. D

Let C, D be a convex partition of X. If X is finite-dimensional then, by


Proposition V.16, Ci"# 0 and Di"# 0. However, when Xis infinite-dimensional it
is possible that Ci =Di =0, even though <C> =<D> =X. For example, take X to
be the set of all sequences (xn) of real numbers with at most finitely many non-zero
terms. We can give X the structure of a real vector space by putting
(Xn) + CYn> = (Xn + Yn) and A.(xn) = (Axn), and then define convexity in the usual
way. Let C, C' be the subsets consisting of all sequences (xJ whose last non-zero
term is positive, negative, and put D = C' u {0}. Then C, Dis a convex partition
of X. However, for each point x e X there exist points y,z e X, with x e (y,z),
2. Separation properties 165

such that (x,y] s C and (x,z] s C'. In fact, if x = (xn), where Xn =0 for n > m,
we can take y = (yn) and z = (zn), where Yn =Zn =Xn for n :I: m+ 1 and Ym+l = 1,
Zm+l = -1.
Uoder our present hypotheses there is a simple characteriution of affine sets:

PROPOSmON 6 A set C s X is affine if and only if C is convex and Ci = C.

Proof Suppose first that C is affine. Then C is certainly convex and, since Xis
unending, Ci= C. It follows at once that also C = C.
Suppose next that C is convex and Ci = C. If C is not affine, there exist
distinct points c,c' e C and x e X \ C such that c' e (c,x). By (S), there exists
ye (c,x] such that (c,y) s C, (y,x) s X \ C. Since Ci= C = C, it follows that
ye C, y :1:-x and (y,x) contains a point c" e C, which is a contradiction. D

Under our present hypotheses there is also a simple characterization of convex


sets whose 'boundary' is convex:

PROPOSITION 7 Let C be a convex set which is llot affine. If Ci :I: 0 alld if


H: = C \ Ci is convex, then H is a hyperplalle of <C>, Ci is an associated open
half-space alld C is the correspollding closed half-space.

Proof By Proposition 6, H :I: 0. Since Ci= Ci, it follows that C is not affine
and thus C c <C>. Let c e Ci and a e <C> \ C. Then, by Proposition V .9, there
exists a point c' e (c,a) (')Ci. Hence, by (S), there exists a point h e (c,a] such
that (c,h) s Ci and (h,a] s X \Ci. Then he H, since he C, h :I: a and hence
h E Ci.
We now show that D = <C> \ Ci is convex. Assume on the contrary that there
exists c e Ci such that c e (d,d') for some distinct d,d' e D. If de C choose a so
that de (c,a), and if d E C put a = d. Similarly if d' e C choose a' so that
d' e (c,a'), and if d' e= C put a'= d'. Then a,a' E C. By the argument of the
preceding paragraph, the segments (c,a) and (c,a') each contain a point of H. Since
H is convex, it follows that c e H, which is a contradiction.
Thus Ci ,D is a convex partition of <C>. By Proposition ~it now need only be
shown that H =D \Di. As in the proof of Proposition 5, Ci = <C> \Di. But
C =Ci, since Ci :1:-0, and H = C (')D. Hence H =D \Di. D
166 VIII. Completeness

Let H be a hyperplane of X and let S,T be nonempty subsets of X. The sets S


and T are said to be separated by the hyperplane H if S is contained in one of the
closed half-spaces associated with Hand Tis contained in the other. They are said
to be properly separated by H if, in addition, S u T H. ~
The case in which the sets S and T are convex is of particular interest. The
followingfirst separation theorem deals with this case:

PROPOSmON 8 Let A and B be convex sets such that Ai*' 0, <A>= X and
B ¢. 0. Then A and B can be separated by a hyperplane if and only if
Ai "' B = 0, and Ai is then contained in one of the open half-spaces associated
with this hyperplane.

Proof Suppose first that A and B can be separated by a hyperplane H. Then A is


not contained in H, since <A>= X, and hence Ai"' H = 0. Thus Ai is contained
in one of the open half-spaces associated with Hand Ai"' B = 0.
Suppose next that Ai "' B = 0. Then, by Proposition 11.22, there exists a
convex partition C,D of X with Ai s C, B s D. In fact Ai s Ci, by Proposition
V.10, since <Ai>= X = <C>. If we put H = X\ (Ci u Di) then, by Proposition 5,
H is a hyperplane and Ci, Di are the two open half-spaces associated with H. It
follows at once that A and B are separated by H. D

In Proposition 8 the hypothesis <A> = X is rather strong and the convex sets
A and B do not appear on an equal footing. These drawbacks are removed in the
following deeper-lying second separation theorem:

PROPOSITION 9 Let A and B be convex sets such that Ai¢. 0 and Bi¢. 0.
Then A and B can be properly separated by a hyperplane of X if and only if
Ai"' Bi= 0.

Proof Suppose first that A and Bare properly separated by a hyperplane H, and
assume that there exists a point x e Ai"' Bi. Then x e Hand there exists also a
point y e (A u B) \ H. Without loss of generality assume y e A \ H. Since
x e Ai, there exists a point z e A such that x e (y,z). Then y and z lie in different
open half-spaces associated with H, which is a contradiction.
Suppose next that Ai"' Bi= 0. We show first that if a hyperplane H properly
separates Ai and Bi, then it also properly separates A and B. Indeed Ai and Bi are
not both contained in H. If Ai, say, is not contained in H then, since Aii =Ai, Ai is
2. Separation properties 167

contained in an open half-space H + associated with H and Bi is contained in the


disjoint closed half-space H_ u H. Moreover, since Bii =Bi, either Bi s H_ or
Bi s H. It follows that A s H+ u Hand B s H_ u H. Thus, by replacing A by
Ai and B by Bi, we may now assume without loss of generality that A =A; and
B =Bi.
Suppose X': = <A u B> -:1: X and there exists a hyperplane H' of X' which
properly separates A and B. Letx' e X'\H'. By Hausdorffs maximality theorem
there exists an affine set H which contains H' but not x', and such that any affine set
properly containing H does contain x'. Thus if x e X\H, then x' e <x u H> and
so, since x' ~ H, x e <x' u H>. Hence <x' u H> = X and H is a hyperplane of
X. Evidently H flX' = H'. If H+ and H_ are the open half-spaces of X associated
with its hyperplane H, then H+ flX' and H_ n X' are the open half-spaces of X'
associated with its hyperplane H', since they are convex and partition X' \ H'. It
follows that A and Bare also properly separated by the hyperplane Hof X. Thus
we now assume, without loss of generality, that X =<A u B>.
Put C = AIB. Then C is convex, by Proposition V.32, and <C> =<A u B> =
X. Moreover B fl C = 0, since A fl B = 0, and Ci= C, by Proposition V.39.
Hence, by Proposition 8, B and C can be separated by a hyperplane H of X and C is
contained in an open half-space H+ associated with H. Moreover, since Bi= B, B
is contained in the other open half-space H_ or in H. In either case A fl H_ = 0,
and A s H + if B s H. Thus H properly separates A and B. D

COROLLARY 10 Two nonempty finite-dimensional convex sets A,B can be


properly separated by a hyperplane if and only if Ai fl Bi= 0. D

An important consequence of Proposition 9 is the following extension


theorem, which reduces to the geometric form of the Hahn-Banach theorem in the
case where X is a real vector space.

PROPOSffiON 11 If A is a convex set such that Ai -:1: 0 and if Bis a nonempty


affine set such that Ai fl B = 0, then there exists a hyperplane H such that
B s H and Ai fl H = 0.

Proof Since B is affine, Bi = B. Hence, by Proposition 9, there exists a


hyperplane H which properly separates A and B. H B s H, then Ai fl H = 0 and
there is nothing more to do. Otherwise, since B is affine, B is contained in an open
168 VIII. Completeness

half-space H_ of X associated with the hyperplane Hand A is contained in the


complementary closed half-space H + u H.
Put A'= H+ u H, and let C denote the set of all points c such that c e (a;b) for
some a e A' and be B. Then, since B is affine, B n C = 0 and A'= C.
Moreover, C is convex. For suppose c e (c 1,c2 ), where c 1 and c2 are distinct
elements of C. Then c; e (a;,bi) for some ai e A' and bi e B (i = 1,2). Hence
c e [a 1,a 2,b 1,b2] and c e [a,b] for some a e [a1'a 2] and be [b 1,b 2]. H c = b,
then c,c1'c 2 e <a 1,b1'b2> and a 1,a 2,c1'c2 all lie in the same open half-plane
associated with the line <b 1,b2>, which contradicts c e (c 1,c2). Hence c ¢band
c e C. Furthermore Ci ¢ 0, since H+ =Ci.
Consequently, by Proposition 8, there exists a hyperplane H' which separates
Band C. Moreover Ci s H+', where H+' is one of the open half-spaces associated
with H', and C s H +' u H'. It follows from the definition of C that B n H_' = 0,
where H_' is the other open half-space associated with H', and so B s H'.
Hence H ¢ H'. Furthermore H n H' = 0. For assume a' e H n H' and
choose any a e H such that a¢ a'. Then <a,a'> s H. Since H s C, it follows
that <a,a'> s H'. Thus H =H', which is a contradiction.
It follows that H s H+', and hence also A'= H+. u H s H+'· D

PROPOSITION 12 Let C be a convex set with Ci¢ 0 and let x e C \Ci. Then
there exists at least one hyperplane Hx of X, with x e H X' such that Ci is
contained in one of the open half-spaces associated with H x and C is
contained in the corresponding closed half-space.
Moreover, C is the intersection of all such closed half-spaces and Ci the
intersection of all such open half-spaces, for varying x e C \Ci.

Proof By Proposition 8, x and Ci can be separated by a hyperplane Hx' of <C> so


that Ci is contained in one of the open half-spaces associated with H/. Moreover,
since x e C, we must have x e Hx'· By Proposition 11 there exists a hyperplane
Hx of X containing Hx' such that Ci is contained in one of the open half-spaces
associated with Hx. Then C is contained in the corresponding closed half-space.
If Dis the intersection of all such closed half-spaces, for varying x e C \Ci,
then obviously C s D. Assume there exists a pointy e D \ C and choose z e Ci.
Then the segment [y,z] contains a point x such that [z,x) s Ci and (x,y] n Ci= 0.
Evidently x e C \Ci. Hence there exists a hyperplane containing x such that Ci is
contained in one of the associated open half-spaces. Since y is in the other
2. Separatio11 properties 169

associated open half-space, this is a contradiction. We conclude that C = D. The


corresponding characterization of Ci follows immediately. D

A hyperplane Hx with the properties described in Proposition 12 will be said to


be a supporting hyperplane of C at x, and the associated closed half-space which
contains C will be said to be a supporting half-space of C at x.
Evidently a hyperplane Hx of X which contains a pointx e C is a supporting
hyperplane of C at x if and only if C n H x is a proper face of C. The intersection
of all supporting half-spaces of C, at a fixed point x e C \Ci, is a convex cone with
vertex x, which will be called the supporting cone of C at x.
It follows at once from the definitions that if C is a convex cone with Ci~ 0,
then any supporting hyperplane of C contains the affine set of vertices of C.
Finally we consider separation properties for any finite number of convex sets.

PROPOSITION 13 Let C1, ... ,Cn (n > 1) be convex proper subsets of X such
that Ck;~ 0 (1 S k Sn) and

Then there exist hyperplanes Hk with associated open half-spaces Hk+ ,Hk-
such that Ck s Hk + (1 < k < n), Ck~ Hkfor at least one k, and

Proof By Proposition 9 the result is true for n = 2, with H 1 = H2 and H2+=He.


We assume that n > 2 and the result holds for all smaller values of n.
Suppose first that Ci n ... n Cni = 0. Then, by the induction hypothesis,
there exist hyperplanes Hk with associated open half-spaces Hk+,Hk- such that
Ck s Hk+ (2 < k < n), Ck ~ Hk for at least one k > 2, and

Choose x 1 e X \ C 1• By Proposition 11 there exists a hyperplane H 1 with


associated open half-spaces H 1+,Hc such thatx 1 e H 1 and C 1i s H 1+. Since
C 1 s H1+ , the result certainly holds in this case. Thus we may now suppose
nk=l,kil:jCi ~ 0 for eachj E { l,... ,n}.
If we put

then, by Proposition V.12,


170 VIII. Completeness

Hence, by Proposition 9, there exists a hyperplane !f 1 which properly separates B 1


and C 1• Thus we may name the open half-spaces H 1+,Hc associated with the
hyperplane H 1 so that C1 s H 1+ and B 1 s H 1-. Moreover, either C 1i s H 1+, or
C 1 s H 1 and B 1i s H 1-.
Thus the hypotheses of the proposition remain satisfied if C 1 is replaced by
D1 = H 1+. By the induction hypothesis, in the same way as before, we may
suppose that
Dii n nk=2,k;t:j Ci '¢ 0 for eachj E {2, ... ,n}.
If we put

then Bi -:i: 0 and Bin Ci= D 1i n B 1i = 0. Hence there exists a hyperplane H 2


which properly separates B 2 and C 2 • Thus the open half-spaces H 2+,H 2-
associated with the hyperplane H 2 may be named so that C 2 s H2+ and
B 2 s H 2- • Moreover, either Ci s H 2+ or C2 s H 2 and Bi s H 2-.
Thus the hypotheses of the proposition remain satisfied if also C2 is replaced
by D 2 = H2+. Proceeding in this way we define inductively for k = l, ... ,n
hyperplanes Hk with associated open half-spaces Hk+,Hk- such that Ck s Dk and
- --
Bk s Hk-, where Dk= Hk+ (1 <k<n),

and

Moreover, either Ck; s Dk;= Hk+ or Ck s Hk and Bk; s Hk-· Furthermore,


although
Bl n Ci = D 1; n ... n D k-l i n Ck; n ... n Cn; (1 s k < n)
is empty, all intersections obtained by omitting one term on the right are nonempty.
Since

the proof is complete unless we have Ck s H k for every k e {1,...,n}.


In this case all intersections obtained by omitting one term from
H 1 n ... n H n are nonempty and, for 1 < k < n, all intersections obtained by
omitting one term from
2. Separation properties 171

are nonempty. We are going to show that H 1 n ... n Hn = 0.


Assume on the contrary that there exists a point z e H 1 n ... n H n and let
ye B 1i. Then ye Hen H 2 n ... n Hn. If we choose x so that z e (x,y), then
x E H 1+ n H 2 n ... n H n· Let u E H 2 + n ... n H n-l + n H n• Then (u,x) =
H 2+ n ... n Hn_ 1+ n Hn. But, since x e H 1+, we can choose v e (u,x) so that
v e H 1+. Then v e H 1+ n ... n Hn_ 1+ n Hn, which contradicts Bni s Hn-·
It now follows from Proposition 11 that there exists a hyperplane Ln with
associated open half-spaces Ln + ,Ln- such that H 1 n ... n H n-l s Ln and
Hn S Ln+•
If z e H 1+ n H 2 n ... n H n-l and y e B 1i, then y e H 1- n H 2 n ... n H n
and there exists a point x e (y,z) such that x e H 1 n ... n H n-l · Since x e Ln and
ye Ln+, it follows that z e Ln-· Thus H 1+ n H 2 n ... n Hn-l s Ln-·
Suppose that, for some k with 1 < k < n -1, we have

H 1+ n ... n H k-l + n H k n ... n H n-l s Ln -.

We have just shown that this is true for k = 2 and we wish to show that if it holds
as written, then it also holds when k is replaced by k + 1. Let

and assume thatx e Ln. Lety e Bi, so that

and let

Since Hn s Ln+ and z e Ln-, by supposition, there exists a point we (y,z) such
that w e Ln. Then

Hence there exists z' e (x,w) such that z' e Hk. Then

z' e H 1+ n ... n Hk_1+ n Hk n ... n Hn-l n Ln,

which is contrary to our supposition. We conclude that


172 VIII. Completeness

Since (x,z) s H 1+ (') ... (') Hk+ (') Hk+l (') .•. (') Hn_ 1, we must actually have

Hence it follows by induction that

3 FUNDAMENTAL THEOREM OF ORDERED GEOMETRY

An ordered division ring D is a one-dimensional dense unending linear


geometry. It will now be shown that D is complete (if and) only if D is the field R
of real numbers.

PROPOSITION 14 Let D be an ordered division ring. If, for any convex set
C s [0,1] with 0 e C and 1 E C, there exists c e [0,1] such that [0,c) s C and
(c,l] s D \ C, then there is a bijective map of D onto the real field R which
preserves sums, products and order.

Proof The interval [0,1) may be linearly mapped onto the interval [a,b], for any
a,b e D with a ':I: b. Consequently the property in the statement of the proposition
continues to hold if 0 is replaced by a and 1 by b.
In any ordered division ring D, nl = 1 + ... + 1 (n times) is positive for any
positive integer n. It follows that the set of all elements ml·(nl)-1, where mis any
integer and n is any positive integer, is a field isomorphic to the field Q of rational
numbers. By identifying ml·(nl)-1 with mn-1, we may regard Q as embedded in
D.
We show first that for each a e D there exists a positive integer n such that
n > a. Assume on the contrary that n < a for every positive integer n. Then a > 0
and the set C of all be [O,a] such that n < b for every positive integer n is convex
and contains a, but not 0. Hence there exists c e [0,a] such that (c,a] s C and
[0,c) s D \ C. Moreover 1 < c, since 2 E C. Thus if we choose de D so that
1 + d = c, then 0 < d < c. Since d E C, we have m > d for some positive integer m.
Hence 1 + m > 1 + d =c, which is a contradiction.
3. Fundamental theorem of ordered geometry 173

Consequently, for any a,b e D with a> 0, there is a positive integer n such
that n > ba-l, and then na > b.
We show next that, for any a,b e D with b > a, there exists x e Q such that
b > x > a. Let n be a positive integer such that n(b - a) > 1 and let m be the least
integer such that m > na. Then nb > na + 1 > (m - 1) + 1 = m > na and hence
b > mn-l >a.
It will now be shown that D is a field. It is sufficient to show that ab = ba for
all a> 0, b > 0. Assume on the contrary that there exist a> 0, b > 0 such that
ab > ba. Then ab > x > ba for some x e Q. Thus b-lx > a and a > xb-l, hence
b-lx > xb-l. Then b-lx > y > xb-l for some ye Q. Thus b-1 > yx-l and
x-ly > frl. Sinceyx-1 =x-ly, this is a contradiction.
For any a e D, the set of all x e Q with x <a is a 'cut' in Q. Moreover, the
hypothesis of the proposition ensures that all cuts in Q may be obtained in this way.
It is readily verified that the mapping which makes correspond to a the real number
defined by this cut in Dedekind's construction of the reals from the rationals has all
the required properties. D

COROLLARY 15 In the statements of Theorems V/1.3 ,4 and 5 tJ,.e ordered


division ring D may be replaced by the real field R if the linear geometry X is
also assumed to be complete. D

Every convex subset C of a real vector space such that Ci = C is a complete


unending linear geometry. The following theorem essentially characterizes such
sets. Since in a sense it is the culmination of our work, we designate it the
fundamental theorem of ordered geometry.

THEOREM 16 If X is a complete unending linear geometry with dim X > 2, or


with dim X = 2 and the Desargues property, then X is isomorphic to a convex
subset C of a real vector space such that Ci = C.

Proof By Theorem VII.4 and Corollary 15, there is an isomorphism/ of X with a


set C in the projective completion V of a real vector space V. Moreover Ci = C,
since Xi =X.
Let a 0 ,b0 be distinct points of X. By Proposition V.18, there exists a
hyperplane H of X such that a0 ,b0 lie in different open half-spaces H +•H- of X
associated with the hyperplane H. The images Y,Y+,Y- of H,H+,H- under the
174 Vil/. Completeness

isomorphism/ are disjoint nonempty convex sets with union C. Moreover Y+; = Y+
and y_i = Y_, since H+i = H+ and H_i = H_.
The theorem will be proved if we show that there is a 'hyperplane' of V
disjoint from C. Since it is sufficient to show that there is a 'hyperplane' disjoint
from C in the projective subspace of V generated by C, we may assume that C itself
generates V. Then there is a unique 'hyperplane' K of V containing Y and Y+,Y_
are contained in the affine space A = V \ K. Since A is the affine hull of Y+ and of
Y_, it follows from Proposition 8 that there is a hyperplane L of A such that Y+ and
Y_ are contained in different open half-spaces of A associated with the hyperplane
L. The hyperplane L of the affine space A extends to a 'hyperplane' L of V.
Evidently Y+ and Y_ are disjoint from L, since Lis obtained from L by adjoining
points of K. It only remains to show that Y is also disjoint from L. Suppose
c' e Y, so that c' = f(c) for some c e H. Thence (a,b) for some a e H+, be H_.
If a'= f(a), b' =f(b), then the line <a',b'> of the affine space A contains a point of
L. Since c' e Land a',b' e L, it follows that c' e L. D

The axioms actually required for Theorem 16 are (C),(Ll)-(L4),(U) and


(S), since (P) and (D) are then implied by dim X > 1. We now consider the
necessity of the hypotheses in Theorem 16. Example VI. 7 shows that Theorem 16
no longer holds if in its statement we replace 'unending' by 'dense', even if we
omit the requirement that Ci= C. Examples Vl.4 and Vl.5 show that Theorem 16
no longer holds when dim X = 2 if we omit the requirement that X have the
Desargues property. Furthermore Example 1 of this chapter shows the necessity of
excluding the case dim X = 1. Finally we are going to show that in the statement of
Theorem 16 we cannot omit 'complete' and replace 'real vector space' by 'vector
space over an ordered division ring'.
We first prove

LEMMA 17 The real field JR has no proper subfield of which it is a finite


extension.

Proof It is sufficient to show that if K is a subfield of the complex field C such


that [C : K] = n, where 1 < n < oo, then n =2. Evidently C is a normal extension of
K, since any polynomial in K[x] is a product of linear factors in C[x] by the
fundamental theorem of algebra. Since the Galois group G of C over K has order
n, it has an element g of order p, where p is some prime divisor of n. By the
3. Fundamental theorem of ordered geometry 175

fundamental theorem of Galois theory, the elements of C fixed by the


automorphism g form a field F such that K s F s C and [C : F] = p. Since

xP - 1 = (x - 1)(1 + x + ... + xP-1 ),

and since an irreducible polynomial in F[x] has degree dividing [C: F] = p, the
irreducible factors of xP - 1 in F[x] are all linear. Thus F contains a primitive p-th
root of unity ~·
We are going to show that C = F(a), where aP e F. Let 'Ye C \ F, so that
C = F(y). If
CX = "( + ~g("() + ... + ~p-1 gP-l(y),
then
g(a) = g(y) + ~g2(y) + ... + ~P-lgP(y) = ~-la.

Thus a E F, since g(a) ':I: a. But g(a.P) = ~-Pcx,P = a.P. Hence a.I' e F and
C = F(a).
Choose p E c so that pP =a.
If b = n
'l=l gi-1(p), then g(b) = b and hence
be F. Moreover bP = Il ~=l gi-l(a) and gi-l(a) = ~-j+lcx, (j = l, ... ,p). If p -¢: 2,
then 1 + 2 + ... + (p - 1) = p(p - 1)/2 is divisible by p and hence bP = aP. Since
a E F, be F and F contains all p-th roots of unity, this is a contradiction. Hence
p = 2, ~ = -1 and b2 = - cx,2. Since (ib)2 = cx,2 and a E F, it follows that i E F. A
fortiori, i E K. If K(i) ':I: C, the same argument can be applied with K(i) in place of
K to obtain the contradiction i E K(i). Hence K(i) = C and n = 2. D

PROPOSITION 18 For any ordered division ring D which is not isomorphic to


the field JR of real numbers and any positive integer n, the projective space
pn(D) can be given the structure of a dense, unending linear geometry.

Proof The proof will be based on simple facts about filters. Let X be a nonempty
set. A nonempty collection~ of subsets of Xis said to be afilter on X when the
following properties hold:

(Fl) if A e ~and A s B s X, then Be ~;


(Fl) if A e ~ and B e ~, then A n B e ~;
(F3) 0 E ~.

Properties (Fl) and (F3) imply that the intersection of finitely many sets in ~
is nonempty and, since~ is nonempty, property (Fl) implies thatX e ~.
176 Vil/. Completeness

If ?J and ?J' are filters on X, we write '!Ji s '!Ji' if A e ?J implies A e '!Ji'.


Evidently this is a partial ordering of the set of all filters on X. An ultrajilter on X
is a filter which is a maximal element in this partial ordering. It follows· from
Hausdorfrs maximality theorem that, for any filter ?J on X, there exists an
ultrafilter Oll, on X such that ?J s oU..
If Oll, is an ultrafilter on X and A,B are subsets of X such that A u B e Oll,, then
either A e oU, or Be oU, (or both). For assume on the contrary that A,B E oU.. Then
the collection ?J of all sets F s X such that A u F e oU, is a filter on X. Since
oU, s ?J and Be ?J, this contradicts the hypothesis that Oll, is an ultrafilter.
It follows that if Oll, is an ultrafilter on X then, for any C s X, either C e Oll, or
X\ Ce oU.. We will show that, conversely, if ?J is a filter on X such that, for any
C s X, either Ce ?J or X\C e '!Ji, then '!Ji is an ultrafilter. Indeed otherwise there
exists E s X such that E E ?J and E () F -:I: 0 for every F e ?J. Since E E ?J
implies X\E e ?J and E () (X\E) = 0, this is a contradiction.
Now let D be an ordered division ring, Oll, an ultrafilter on a set X, and r the
set of all maps cl>: X ~ D . If cl> ,cl> ' e r, we write cl> -- cl> ' if the set
{x e X: cl>(x) = cl>'(x)} e Oll,. It is easily verified that this is an equivalence relation
on r. The set of equivalence classes will be denoted by t (X,Oll, ), or simply t,
since X and oU, will be fixed. If ~, ljf, ci> e t, we define ci> = ~ + '\ji if there exist
cl> e ~, 'I' e ljf, ro e ci> such that the set {x e X: m(x) = cl>(x) + 'lf(X)} e Oll,. This
definition does not depend on the choice of cl>,'lf,O> within their equivalence classes.
For if cl>' -- cl>, v' -- "'' m' - m, then the set {x E X: m'(x) = cl>'(x) + V'(x)}
contains the intersection of the sets {x e X: cl>'(x) = cl>(x)}, {x e X: V'(x) = 'lf(x)},
{x e X: m'(x) = m(x) }, {x e X: m(x) =cl>(x) + 'lf(X)} and consequently, since Oll, is
a filter, is in Oll,.
If ~, ljf , ci> e t, we may similarly define ci> = ~ · ljf if there exist cl> e ~,
'If e ljf, me ci> such that the set {x e X: m(x) = cl>(x)·'lf(x)} e Oll,. We may also
define ~ < 'if if there exist cl> e ~, 'If e ljf such that the set
{x e X: cl>(x) < 'lf(X)} e Oll,. Let 6, i be the equivalence classes in t containing the
constant maps cl>(x) =0, cl>(x) = 1 respectively, for every x e X. It is not difficult to
verify that with these definitions t is an ordered division ring. We give the
argument only when it depends on the fact that Oll, is actually an ultrafilter.
Suppose ~ e t and~ -:1: 6. We wish to show that there exists 'if e t such
that ~ · ljf = 'if·~ = i. If cl> e ~, then the set {x e X: cl>(x) = 0} E Oll,. Since Oll, is
3. Fundamental theorem of ordered geometry 177

an ultrafilter, it follows that the set {x e X: cp(x) '# 0} e oU. Define v: X--+ D by
'lf(X) = cp(x)-1 if cp(x) '# 0, = 1 otherwise. If 'If E \ji, then ~· \ji = \ji· ~ = i.
Suppose ~' \ji e t and ~ ~ \ji. We wish to show that '\fl < ~· If cp e ~'
'II e 'if, then the set {x e X: cp(x) < 'lf(X)} e oU. Since oU is an ultrafilter, it
follows that the set {x e X: 'lf(X) < cp(x)} e oU, and hence also the set
{x e X: 'lf(X) S cj>(x)} e oU. Thus \ji S ~.
For any a e D, define cl>a:X--+ D by ct>a(x) =a for every x e X. If
cl>a e {~a}, then the map a --+ ~a is an injection of D into t which preserves
addition, multiplication and order. Consequently we may regard D as embedded in
r.
Suppose now that the ordered division ring D has a gap; i.e., there exist
nonempty subsets A,B of D such that a< b for every a e A, be B, but there is no
c e D such that a< c < b for every a e A, be B. This implies that A(') B = 0.
We wish to construct an ordered division ring D' :::> D in which the gap has been
removed, i.e. there exists c' e D' such that a< c' < b for every a e A, be B.
For any a e A, let Fa = {x e A: a < x} and le~ '!Ji denote the collection of all
sets F s A which contain a set Fa for some a e A. Evidently '!Ji is a filter on A.
Let oU be an ultrafilter on A such that '!Ji s oU, and now take t = t (A, oU) to be the
ordered division ring constructed with X =A and with this ultrafilter. Define the
map ·Y: A --+ D by 'Y(x) = x for every x e A and let y be the corresponding element
oft. For each a e A, the set {x e A: a S y(x)} e oU, and for each be B, the set
{x e A: y(x) < b} =A e oU. Thus, regarding D as embedded in t, we have
a< y < b for every a e A, be B.
After these preparations we can proceed with the proof of Proposition 18.
Assume first that D is a proper subfield of R. Then, by Lemma 17, R is infinite-
dimensional, considered as a vector space over D. Hence, by the discussion of
Example VI.6, for any positive integer n the n-dimensional projective space over D
can be given the structure of a dense, unending linear geometry.
Thus we now assume that no extension of D is isomorphic to the real field R.
Since D itself is not isomorphic to R, it follows from Proposition 14 that the
interval [0,1] is the union of two nonempty subsets A,B such that a< b for every
a e A, b e B, but there is no c e D such that a < c < b for every a e A, b e B.
Hence, as we have seen, there exists an ordered division ring D' :::> D with an
element c' such that a Sc' Sb for every a e A, be B. Since no extension of Dis
isomorphic to R, this argument can be repeated any finite number of times.
178 VIII. Completeness

Consequently, for any positive integer n, there exists an ordered division ring
D(n) :::> D which has dimension > n as a vector space over D. The conclusion now

follows as in the previous case. D

4 METRIC AND NORM

For convenience of reference we repeat here some well-known definitions


concerning topological and metric spaces, and state some well-known results. The
proofs may be found, for example, in Hewitt and Stromberg (1975).
A topology is said to be defined on a setX, andX is said to be a topological
space, if a collection ~ of subsets of Xis given such that

(Tl) the whole set X and the empty set 0 are in~'
(Tl) the intersection of two sets in~ is again a set in~'
(T3) the union of any family of sets in ~ is again a set in ~.

A set A s Xis said to be open if A e ~and is said to be closed if X\A e ~. A


topological space is a Hausdorff space if any two distinct points are contained in
disjoint open sets.
An open cover of a topological space Xis a family {Gal of open sets such that
X = Ua Ga. A subcover is a subfamily which is also a cover. A topological space
X is said to be compact if each open cover has a finite subcover.
Given a subset A of a topological space X, the collection of all sets Gr. A,
where G is any open subset of X, defines a topology on A. If A is closed and X is
compact, then A is also a compact topological space with the topology induced by
that on X.
A metric is defined on a set X if with each ordered pair (x,y) of elements of X
there is associated a real number d(x,y) with the properties

(Dl) d(x,y) > 0, with equality if and only if x =y,


(D2) d(x,y) = d(y,x) for all x,y e X,
(D3) d(x,y) < d(x,z) + d(z,y) for all x,y,z e X.

The pair (X,d) is a metric space. A subset A of a metric space (X,d) is said to be
bounded if {d(x,y): x,y e A} is a bounded subset of R.
4. Metric and norm 179

A metric space (X ,d) has a natural topology in which the open sets are the
unions of sets of the form
N(x0,e) = {x e X: d(x0 ~) < e},

where x0 e X and e > 0. This topology will always be understood when speaking
of open or closed subsets of a metric space. For a metric space there is a simpler
characterization of compactness: a metric space Xis compact if and only if every
sequence {xn} of elements of X has a convergent subsequence {xntl, i.e. there
exists x e X such that d(xnt~)--+ 0 ask--+ co.
If a subset A of a metric space X is compact, with the topology induced by that
on X, then A is closed and bounded. In general it is not true that, conversely, a
closed and bounded set is compact, but it is true for X =Rd.
A sequence {xn} in a metric space Xis said to be a Cauchy sequence if for
each real e > 0 there is a corresponding integer p > 0 such that d(xm~n) < e for all
m,n > p. A metric space is said to be complete if every Cauchy sequence is
convergent Any compact metric space is complete.
The two uses of the word 'complete' in this chapter correspond to the two
ways, due to Dede.kind and Cantor, of constructing the real numbers from the
rationals. It is hoped that the meaning in each instance will be clear from the
context.
Suppose now that V is a real vector space. We say that a norm is defined on
V, and that Vis a normed vector space, if with each v e V there is associated a real
number llvll with the properties

(Nl) llvll > 0, with equality if and only if v = 0,


(N2) llv + wll < llvll + llwll for all v,w e V,
(N3) llavll = lal llvll for all v e V and all a e R.

It follows that Vis a metric space with the metric d(v,w) = llv-wll.
The following two propositions will be proved, since the proofs are less
accessible.

PROPOSITION 19 A finite-dimensional real vector space V can be normed.


Moreover, whatever the norm, any bounded closed subset is compact.

Proof Let d be the dimension of the vector space V and let e 1, ••• ,ed be a (vector
space) basis for V. Then each v e V can be uniquely represented in the form
180 VIII. Completeness

where aj e R (j = 1,... ,d). It may be immediately verified that

lvl = max 1sjsd lail


is a norm on V. Moreover, the unit ball U = {v e V: lvl < 1} is compact, since the
interval [-1,1] of R is compact. The unit sphere S = {v e V: lvl = 1} is also
compact, since it is a closed subset of U.
To show that, with respect to an arbitrary norm 1111, any bounded closed set is
compact, it is sufficient to show that this norm is equivalent to the norm 11, i.e.
there exist positive constants µ,p such that pjvl < llvll < µjvj for every v e V.
This is certainly true if d = l, since then llvll = la11 lle 1ll = lle1ll lvl. We
supposed> 1 and use induction on d. Write v = v' + v", where

By the induction hypothesis there exists a positive constantµ' such that, for every
v e V, llv'll < µ 'lv'I. Then

llvll < llv'll + llv"ll < µ'jv'I + ladl lledll < µjvj,

whereµ=µ'+ lle~j.
The real-valued function cp{v) = llvll is continuous on S, since

Let p =inf {cp(v): v e S} and let {vn} be a sequence of elements of S such that
cp(vn> ~ p. Since S is compact, by restricting attention to a subsequence we may
suppose that there exists v* e S such that lvn - v*I ~ 0. Then cp{v*) = p, since <p
is continuous, and p > 0, since v* :I: 0. Since llvll > p for lvl = 1, it follows from
(N3) that p lvl < llvll for every v e V. 0

PROPOSITTON 20 If V is a normed real vector space, then a metric d can be


defined on the corresponding projective space P(V). Moreover, the metric
space (P(V),d) is complete if V is complete and compact if V is finite-
dimensional.

Proof The elements of P(V) are the one-dimensional vector subspaces of V. If


x,y e P(V) and if u,v are non-zero vectors in x,y respectively, we put
4. Metric and norm 181

d(x,y) = min {ll<u~lull - v~lvll)ll, ll<u~lull + v~lvll>ll} ·


Evidently this definition does not depend on the choice of u e x and v. e y.
Moreover d(x,y) = d(y,x) and d(x,y) :c?: 0, with equality if and only if x = y.
Furthermore, the triangle inequality

d(x,y) < d(x,z) + d(z,y)

holds since, for any non-zero vectors u,v,w,

ll<u/llull - v~lvll)ll < ll<u~lull - w/llwll)ll + II<w/llwll - v~lvll)ll,


ll<u~lull - v~lvll)ll < ll<u~lull + w~lwll)ll + ll<w~lwll +v~lvll)ll,
ll<u~lull +v/llvll)ll < ll(u/llull-w/llwll)ll + ll<w~lwll +v~lvll)ll,
ll<u~lull + v~lvll)ll < ll<u~lull +w~lwll)ll +ll<w~lwll - v~lvll)ll.
The projective space P(V) is bounded with respect to this metric, since d(x,y) < 2
for all x,y e P(V).
Suppose now that Vis complete and let {xnl be a Cauchy sequence in P(V).
Then there exists a positive integer p such that d(xn,xp) < 1/2 for every n > p.
Choose a vector Un e Xn with llunll = 1. By replacing Un by- Un, if necessary, we
may assume that

Then, if m,n > p,

and hence

Consequently d(xm~,J =l um - unll and {Un} is a Cauchy sequence in V. Since Vis


complete, there exists a vector u e V with llull = 1 such that llun - ull -+ 0. If x is the
corresponding element of P(V), it follows directly that d(xn,x) -+ 0.
Suppose next that Vis finite-dimensional and let {xnl be any sequence in
P(V). Choose e Xn with llunll = 1. Since the unit sphere of V is compact, by
Un

Proposition 19, there exist a subsequence {vnl of {un} and a vector v e V with
llvll = 1 such that llvn - vii-+ 0. Let Yn'Y be the elements of P(V) corresponding to
vn,v. For n >p we have llvn-vll < 1 and hence llvn +vii> 1. Consequently
182 VIII. Completeness

In conjunction with Corollary 15, Propositions 19 and 20 show that any


complete dense linear geometry X with 2 < dim X < oo, or with dim X = 2 and the
Desargues property, can be given the structure of a metric space so that every closed
set is compact.

5 NOTES

The long line is a fruitful source of counterexamples in topology; see Steen and
Seebach (1978). For the roots of Proposition 3 in the investigations by Gauss on
non-Euclidean geometry, see Section 12.6 of Coxeter (1989). Proposition 4 is
proved in Doignon (1976).
Separation properties of two convex sets, and in particular the Hahn-Banach
theorem, play a fundamental role in functional analysis; see, for example, Bourbaki
( 1987) and Holmes (197 5). The separation of several convex sets, which is
important in optimization, is treated in Bair and Foumeau (1980) and Boltyanskii
(1975).
Theorem 16 is contained in Doignon (1976), but naturally it had many
forerunners. The theorem throws an interesting light on Examples 11.1, VI.2 and
VI.3. These examples can be given a common formulation in the following way: let
X be a connected finite-dimensional Riemannian manifold such that any two points
a,b of X are joined by a unique geodesic arc, and define the segment [a,b] to be the
set of all points on this arc. A theorem of Beltrami, in conjunction with Theorem
16, shows that the three cited examples are the only complete unending linear
geometries which can be obtained in this way if dim X > 2, or if dim X = 2 and X
has the Desargues property. Beltrami's theorem states that if Xis a connected
Riemannian n-manifold such that every point has a neighbourhood which can be
mapped homeomorphically into JRn, so that geodesic arcs are mapped onto straight
line segments, then X has constant curvature. The theorem is proved in Spivak
(1975), for example. A proof which assumes only continuity of the Riemannian
metric is sketched by Pogorelov (1991).
If dim X = 2 and X does not have the Desargues property, then the possibilities
are by no means so restricted. Busemann (1955), Section 11, shows that in
Example VI.5 a metric d can be defined on JR2 so that, for any points x,y,z e JR2,
5. Notes 183

(t) d(x,y) = d(x,z) + d(z,y) if and only if z e [x,y].

It follows that [x,y] is the unique geodesic arc joining x and y. Another proof,
using methods derived from integral geometry, is described in Salzmann (1967). It
may be asked: can a metric d with the property (t) be defined on any complete
unending two-dimensional linear geometry?
For an approach which starts from a metric space (X ,d) and defines the
segment [x,y] to be the set of all points z such that d(x,y) = d(x,z) + d(z,y), see
Rinow (1961) (and p. 420, in particular, for the connection with our approach).
Proposition 18 was first proved by Szczerba (1972) for two-dimensional V,
using model theory rather than filters. The main ingredients in the proof of Lemma
17, namely the fundamental theorem of algebra and the fundamental theorem of
Galois theory, are proved in Lang (1984), for example. Lemma 17 is actually a
special case of a more general result of Artin and Schreier: if an algebraically closed
field F has a proper subfield K of which it is a finite extension, then F has
characteristic zero and F = K(i), where i2 = - 1. See, for example, Priess-Crampe
(1983).
IX
Spaces of Convex Sets

This chapter is concerned with spaces of convex sets. We show that a metric may
be defined on the space ~(X) of all nonempty bounded closed subsets of a metric
space X. Furthermore this Hausdorff metric is complete if the metric of X is
complete. These results are then applied to the space '<6 (X) of all nonempty
bounded closed convex subsets of a normed real vector space X. The vector sum
of two elements of '<6(X) need not again be in '<6(X), if Xis infinite-dimensional.
However, the closure of the vector sum is in '<6 (X) and, with this modified
definition of addition, '<6 (X) has the structure of a commutative semigroup.
R4dstrom's cancellation law implies that this semigroup may be embedded in a
group. In fact '<6(X), with the Hausdorff metric and with the partial order induced
by inclusion, may be embedded isometrically and isomorphically in a vector lattice
with two further properties, which guarantee the existence of a norm. Vector
lattices with these two further properties are here called Kakutani spaces.
A basic theorem, the Krein-Kakutani theorem, says that an arbitrary Kakutani
space can be mapped isometrically and isomorphically onto a dense subset of the
Kakutani space C(K) of all continuous real-valued functions on some compact
Hausdorff space K. Thus, as a final result, for a given normed real vector space X
there exists a compact Hausdorff space K such that the space '<6(X) of all nonempty
bounded closed convex subsets of X may be embedded in the space C(K) of all
continuous real-valued functions on K. To make the discussion self-contained, we
also prove the Krein-Kakutani theorem.

1 THE HAUSDORFF METRIC

Let (X,d) be a metric space and let ~(X) denote the family of all nonempty
bounded closed subsets of X. We are going to show that a metric can be defined
also on ~(X).
186 IX. Spaces of convex sets

We first define the deviation of a pointy e X from a set A e ~(X) by

O(y,A) = infxeA d(x,y).


It is obvious that o(y,A) = 0 if ye A. On the other hand o(y,A) > 0 if y ~ A. For,
since the complement of A is open, there exists p > 0 such that no point of A is
contained in the open ball with centre y and radius p.
The Hausdorff distance between two sets A,B e ~(X) is now defined by

h(A,B) = max {SUPyeBO(y,A), SUPxeAO(x,B)}.

Evidently h{A,B) < oo, since A and B are bounded. Moreover h(A,B) > 0, with
equality if and only if A = B. It is obvious from the definition that
h{A,B) = h(B ,A}. It remains to prove the triangle inequality

h(A,B) < h(A,C) + h(C,B).

For all x e A, y e B, z e C we have

d(x,y) < d(x,z) + d(z,y).


Hence
o(x,B) < d(x,z) + o{z,B)
< d{x,z) + SUPweC o(w,B)
< d(x,z) + h(C,B).

Since this inequality holds for all z e C, it follows that

o(x,B) < o(x,C) + h(C,B)


< SUPueAO(u,C) + h{C,B)
< h{A,C) + h(C,B).
Thus
SUPxeAO(x,B) < h(A,C) + h{C,B).

Similarly we can show that

SUPyeBS(y,A) < h(A,C) + h(C,B).

Taking both together, we obtain the triangle inequality.


From now on we understand the space ~(X) to be equipped with the
Hausdorff metric. Thus we can talk about convergent sequences, and Cauchy
sequences, of nonempty bounded closed subsets of X.
1. The Hausdotff metric 187

PROPOSITION 1 If (X,d) is a complete metric space, then (~(X),h) is also a


complete metric space.

Proof We wish to show that if {A 11 } is any Cauchy sequence in ~(X), then there
exists A e ~(X) such that An-+ A as n-+ oo. For each n = 1,2, ... let Fn denote
the closure of the set Um~nAm , so that F 1 ~ F 2 ~ · · • • Then A = n;=t F n is a
bounded closed subset of X. We will show that A has the required properties.
Fix any e > 0. For each k = 1,2, · · · there exists a positive integer nk such that

Moreover we may assume that n 1 < n2 < · · · . Choose any point x1 e An 1 and
define inductively a sequence {xk} so that xk e An.t and d(xk,xk_ 1) < e/2k. Then
{xk} is a Cauchy sequence, since d(xj,xk) < e/2k for allj > k. Since Xis complete,
there exists a point y 1 e X such that xk -+ y 1 ask-+ oo. Clearly, d(x 1,y 1) Se.
Moreover y 1 e A, since Xj e F n,t for all j > k and F n,t is closed. Thus A is
nonempty and
S(x,A) < £ for all x e An 1•

On the other hand, let B denote the set of all points ye X such that S(y,An 1) < e.
Then An c B for all n > n 1• Since B is closed, it follows that F ni c B and hence
AcB. Thus
S(y,An 1) < £ for ally e A.

Therefore h(A,An 1) Se and, by the triangle inequality, h(A,An) < 2e for all n > n 1•
Thus we have shown that An -+ A as n -+ oo. 0

Let ':K(X) denote the family of all nonempty compact subsets of X. Then
'X(X) ~ ~(X), since any compact subset of Xis closed and bounded.

PROPOSITION 2 If (X,d) is a complete metric space, then ('j{(X),h) is also a


complete metric space.

Proof We need only show that ':K(X) is a closed subset of the complete metric
space GJ(X),h). Suppose An e ':K(X) and An -+ A in ~(X). We wish to show that
A e ':K(X). Thus we wish to show that any sequence {xn} of elements of A has a
subsequence converging to an element of A. In fact, since X is complete and A is
188 IX. Spaces of convex sets

closed, it is enough to show that the sequence {Xn} has a subsequence which is a
Cauchy sequence.
There exists a positive integer n 1 such that h(A,A,J S 1/2 for all n ~ n 1. Since
An1 is compact, it is covered by finitely many open balls with radius 1/2. It follows
that A is covered by finitely many open balls with the same centres and radius 1.
Thus there is a ball B 1 with radius 1 which contains Xn for infinitely many n. In the
same way there is a ball B 2 with radius 1/2 such that B 1 n B 2 contains Xn for
infinitely many n. In general, for each positive integer k there is a ball Bk with
radius 1/2k-l such that B 1 n ... n Bk contains Xn for infinitely many n. Hence
there exists an increasing sequence {n k} of positive integers such that
Xnk E B 1 n ... n Bk· Evidently the subsequence {Xnkl is a Cauchy sequence. a
The next result shows that the space of nonempty compact subsets of X has the
Bolzano-Weierstrass property if X itself has. The proof uses the diagonal process
which Cantor first used to show that the set of real numbers is uncountable.

PROPOSITION 3 /f(X,d) is a compact metric space, then (X(X),h) is also a


compact metric space.

Proof Since a compact metric space is necessarily complete, (JC (X),h) is a


complete metric space by Proposition 2. Thus we need only show that any
sequence {An} of elements of X(X) has a subsequence which is a Cauchy
sequence.
For a given £ 1 > 0 there is a positive integer m 1 such that the space Xis
covered by finitely many open balls B 11 ,... ,B lmi of radius £1. With each set An we
associate the collection of those balls B lj with which it has at least one point in
common. Since a set with m 1 elements has just 2m 1 - 1 nonempty subsets, there
exists a collection of balls B lj which is associated in this way with infinitely many
sets An. Let {An,l} be a subsequence of {An} such that with each set An,l there is
associated the same collection of balls B lj· Then S(x.An, 1) < 2£ 1 if x e Am,l and
hence h(Am,l ,An,l) < 2£ 1 for all m,n.
We now repeat this process with £ 1 replaced by e 2 = £ 1/2 and {An} replaced
by {An, 1 }. We obtain a subsequence {An, 2 } of {An,ll such that
h(Am,2,An,2) < 2£2 = £1 for all m,n. By repeating the process infinitely often and
constructing the diagonal sequence {An'}, where An'= An,n' we obtain a
subsequence of {An} which is a Cauchy sequence. D
2. The space C({,(X) 189

2 THE SPACE '€(X)

To apply the preceding results to convex sets we now impose more structure
on the metric space. Throughout this section we assume that Xis a normed real
vector space. Thus with each x e X there is associated areal number llxll with the
properties

(Nl) llxll > 0, with equality if and only if x = 0,


(N2) llx + Yll < llxll + llYll for all x,y e X,
(N3) Ila.xii = lcxl llxll for all x e X and all ex e R,
and Xis a metric space with the metric d(x,y) = llx-Yll· We will denote by Uthe
closed unit ball of X: U = {x e X: llxll < 1}. Evidently U is convex, by (Nl)-
(N3).
Let '€ = '€(X) denote the collection of all nonempty bounded closed convex
subsets of X. If A,B e '€, then A + B is a nonempty bounded convex subset of X,
but it need not be closed if Xis infinite-dimensional. However, if we define A EBB
to be the closure of A+ B, then A EBB e '€.
The properties in Chapter VII, Section 3, about addition of convex sets in a
vector space over an ordered division ring, remain valid in '€ ='€(X) with this new
definition of addition:

(i) A EB B = B EB A, (A EB B) EB C =A EB (B EB C),
(ii) cx(A EBB) =a.A EB cxB, (cx~)A = cx(~A), IA =A,
(iii) A EB 0 = A, OA = 0,
(iv) if A ~ B, then A EB C ~ B EB C and a.A c: cxB,
(v) if A.,µ~ 0, then (A.+ µ)A =A.A EB µA.
Since the proofs for the other properties are similar but simpler, we prove only the
associativity of addition:
(A EB B) EB C = A EB (B EB C).

Let D denote the closure of the set {a+ b + c: a e A, be B, c e C}. If


x e A EB B then, for each positive integer n, there exist an e A, bn e B and un e U
such that x = unln +an + bn. Hence if c e C, then x + c e D. Thus
(A EB B) + C ~ D and hence, since D is closed, (A EB B) EB C c D. Since the
190 IX. Spaces of convex sets

reverse inequality is obvious, this proves that (A E9 B) E9 C = D. But in the same


way we can prove that A E9 (B E9 C) = D.
We now derive some additional properties. If'€ is partially ordered by
inclusion, then

(vi) any A,B e '€ have a least upper bound A v B.

In fact A v B is the closure of [A u BJ. Furthermore,

(vii) for any A,B,C e '€, (A v B) E9 C = (A E9 C) v (B E9 C).

To prove this we need only show that (A v B) E9 C ~ (A E9 C) v (B E9 C), since


the reverse inequality is obvious. A similar argument to that used in proving the
associativity of addition shows that (A v B) E9 C is the closure of the set
[Au BJ+ C. But if x e [Au BJ+ C then, for some a e A, be B, c e C and
A. e [0,lJ,
x =A.a+ (1 -A.)b + c = A.(a + c) + (1 -A.)(b + c).

Thus x e (A E9 C) v (B E9 C), and hence (A v B) E9 C ~ (A E9 C) v (B E9 C).


Of particular importance is the following cancellation law for addition:

PROPOSmON 4 For any A,B,C e '€,

(viii) if A E9 Cf;; B E9 C, then A~ B.


Proof Choose a e A and ci e C. Since a+ ci e B E9 C, there exist bi e B,
c2 e C and Xi e 2-iu such that

Repeating this procedure, we define inductively bk e B, ck+i e C and xk e 2-kU


such that

Adding these relations fork= l, ...,n and mutiplying by 1/n, we obtain

where bn' =(bi + ... + bn)/n e B, since B is convex, and Xn' =(xi + ... + Xn)/n e
(1/n)U, by construction. Since C is bounded and Bis closed, it follows that a e B.
Thus A~ B. D
2. The space ~(X) 191

COROLLARY 5 If A,B,C e ~ andA E9 C =B E9 C, then A =B. D

To illustrate the role played by convexity in this result we now show ~at if
X = JR.d, Ac X and C =[A], then

tf-lA + C = tf-lC + C.

Since the left side is obviously contained in the right, and since d-lC + C =
d-l(d + l)C, we need only show that C ~ (d + 1)-lA + d(d + 1)-lC. If x e C then,
by Corollary 111.18, there exists a finite affine independent set F c A such that
x e [F]. Thus x = L ~ 1 A;X;, where X; e A and A; > 0 (i = l ,... ,m ), L ;!1 A; = 1
and m < d + 1. Moreover we may choose the notation so that A. 1 = max; A;. Then
A. 1 > m-1 > (d+l)-1 and

x = (d+l)-1x 1 + [A.1 -(d+l)-1Jx1 + L ~2A;X;

= (d+l)-1x 1 +d(d+l)-1 :E~ 1 µ;X;,

where µ; > 0 (i = 1,... ,m) and L ;!1µ; = 1. Thus


x E (d + 1)-lA + d(d + 1)-lC.

If we choose A bounded and closed, but not convex, then C is a bounded Closed
convex set, but tf-lA ~ d-lC.
For a normed vector space the definition of the Hausdorff metric can be
reformulated in the following way:

LEMMA 6 If X is a normed real vector space then, for any sets A,B e ?1(X),

h(A,B) = inf {A> 0: A~ B E9 A.U, B ~A E9 A.U}.

Proof If A ~ B E9 A.U then, for any x e A, S(x,B) < A. and consequently


SUPxeAB(x,B) <A. Hence if A~ B E9 AU andB ~A
E9 AU, then h(A,B) SA.
On the other hand if A ~ B E9 µU then, for some x e A, x E B E9 µU. Hence
d(x,y) >µfor ally e B, and so S(x,B) ~ µ. Consequently, if either A ~ B E9 µU
or B ~ A E9 µU, then h(A,B) ~ µ. The result follows. D
192 IX. Spaces of convex sets

LEMMA 7 If X is a normed real vector space, if {An} is a sequence of sets in


'€(X) and if An~ A in (?i(X),h), then A e '€(X).

Proof Since A is nonempty, bounded and closed, we need only show that it is
convex. Let z = A.x + (1 -A.)y, where x,y e A and 0 <A.< 1. By Lemma 6, for
any £ > 0 there is a positive integer m such that

An~ A+ eU, A~ An+ eU for every n > m.

Since An and U are convex, so also is An+ eU. Hence

z e An + eU ~ A + 2eU.

Since e is arbitrary and A is closed, it follows that z e A. D

From Lemma 7 and Proposition 1 we immediately obtain

PROPOSITION 8 If X is a complete normed real vector space then, with the


Hausdorff metric h derived from the norm, ('€(X),h) is also a complete metric
space. D

A finite-dimensional normed real vector space X is necessarily complete, and a


subset of X is compact if and only if it is both bounded and closed. Hence, by
applying Lemma 7 and Proposition 3 to a compact set X' ~ X we immediately
obtain the Blaschke selection principle:

PROPOSITION 9 If X is a finite-dimensional normed real vector space then,


with the Hausdorff metric h derived from the norm, any bounded sequence of
sets in ('€(X),h) has a convergent subsequence. D

3 EMBEDDINGS OF '€(X)

Since '€ has the structure of a commutative semigroup under addition, with a
zero element and with multiplication by non-negative real numbers as a semigroup
of operators, and since the cancellation law for addition holds, it is possible to
embed~ in a vector space. We now sketch the construction, which is similar to
that by which the semigroup of non-negative integers is embedded in the group of
all integers (and to that used at the end of Section 1 in Chapter Vill).
3. Embeddings of ~(X) 193

If (A,B) and (C ,D) are ordered pairs of elements of'€, we write (A,B) - (C ,D)
if A El1 D = B El1 C. It follows directly from Corollary 5 that - is an equivalence
relation. Let {A,B} denote the equivalence class containing the pair (A,B) and let
;;£ = fi(X) denote the set of all equivalence classes {A,B}.
If {A,B} and {C,D} are elements of;;£, we define their sum by

{A,B} + {C,D} = {A El1 C,B EBD}.

It is easily seen that the sum is uniquely defined, i.e. if {A',B'} = {A,B} and
{C',D'} = { C,D} then

{A',B'} + {C',D'} - {A,B} + {C,D}.

Similarly, for any a e R. we can define

a{A,B} = {aA,a.B} if a> 0,


= { (-cx.)B,(--a)A} if a< 0.

It is not difficult to verify that with these definitions;;£ is a vector space with {0,0}
as the zero vector. Evidently any {A,B} e ;;£has the form

{A,B} = {A,0} - {B,0}.

We can also define an order relation on ;;£ by

{A,B} < {C,D} ifAa1D~Ba1C.

It follows from Proposition 4 that this order relation is uniquely defined and is
indeed a partial order. Also, if {A,B} < {C,D} then, for any {E,F} e ff, and any
A.> 0,
{A,B} + {E,F} < {C,D} + {E,F},
A.{A,B} < A.{C,D}.

Furthermore any two elements {A,B} and {C,D} of;;£ have a least upper
bound, namely {G,B El1 D }, where G = (A El1 D) v (B El1 C). For we certainly
have
{ A,B} = {A e D' B e D}
< {G ,B e D}'
{C,D} = {B El1 C,B EBD} < {G,B EBD}.

On the other hand, if {A,B} < {E,F} and {C,D} < {E,F}, then
194 TX. Spaces of convex sets

A EB D EB F ~ B EB D EBE, B EB C EB F c B EB D EB E,
and hence
G EB F = (A Ea D EB F) v (B Ea C EB F) ~ B EB D Ea E.
Thus~ is a vector lattice.
We draw attention also to two other properties of ~. First, for any
{A,B} e ~there is a positive integer n such that

- n{ U,0} < {A,B} < n{ U,0}.

Indeed we can choose the positive integer m so large that A,B ~ mU and then, if
n > 2m, we will have A~ B + nU andB ~A+ nU =A+ (-n)U.
Secondly, if {A,B},{C,D} e ~and n{A,B}< {C,D} for every positive
integer n, then {A,B} < {0,0}. Equivalently, if nA Ea D ~ nB EB C for every
positive integer n, thenA ~ B. Indeed if a e A and de D, then there exist bn e B,
en e C and un e U such that

Since C and D are bounded, and B is closed, it follows that a e B.


In the terminology which will be introduced in the next section, we have
shown that ~ is a Kakutani space. It is shown there that a Kakutani space may
always be normed. Applying the definition of the norm to ~ we immediately verify
that the distance between {A,0} and {B,0}, with respect to this norm, is precisely
the Hausdorff distance

h(A,B) = inf {A.> 0: A~ B Ea A.U, B c A Ea A.U}.


If we define a map cp: '€ ~~by cp(A) = {A,0} for every A e '€,then the
preceding argument shows that

cp(A Ea B) = cp(A) + cp(B), cp(AA) = A.cp(A) for any A.> 0,


A ~ B if and only if cp(A) < cp(B), cp(A v B) = cp(A) v cp(B),
h(A,B) = llcp(A) - cp(B)ll.

Thus we have proved


3. Embeddings o/'€(X) 195

THEOREM 10 Let X be a normed real vector space and, with the Hausdorff
metric h derived from the norm, let ("6(X),h) be the metric space of nonempty
bounded closed convex subsets of X.
Then there exists an injective map <p o/"6 = "6(X) into a Kakutani space:£
with the following properties:

(i) <p preserves sums and non-negative real multiples,


(ii) <p preserves order and the supremum of any two elements,
(iii) <p preserves distance,
(iv) <(6' = <p("6) is a convex cone in:£ and'€' generates:£:

A.'€' + µ'€' ~ <(6 'for all A.,µ > O and:£ ='€' - '€ '. a
By identifying A with <p(A), we may regard'€ as embedded in:£. From
Theorem 10 we immediately obtain the translation invariance of the Hausdorff
metric (which can also be proved directly):

COROLLARY 11 Let X be a normed real vector space and, with the Hausdorff
metric h derived from the norm, let ("6(X),h) be the metric space of nonempty
bounded closed convex subsets of X. If A,B,C e <(6(X) and A.> 0, then

h(A E9 C,B E9 C) =h(A,B), h(AA,AB) =A.h(A,B). D

The prototype of a Kakutani space is the space C(K) of all continuous


functions /: K--+ JR, where K is a compact Hausdorff space and where, if
f,g e C(.K) and a e JR,/+ g and af are defined by

(f + g)(t) = f(t) + g(t), (af)(t) = af(.t) for every t e K,

and/< g is defined by
f(t) < g(t) for every t e K.

Then any f,g e C(.K) have a least upper bound/ v g defined by

(f v g)(t) = max lf(t),g(t)} for every t e K.

Moreover, for any f,g e C(K), if nf < g for every positive integer n, then/< 0.
Finally, for any f e C(.K) there is a positive integer n such that- ne <f < ne, where
e e C(.K) is the constant function defined by

e(t) = 1 for every t e K.


196 IX. Spaces of co11vex sets

The associated norm for the Kakutani space C(K) is the supremum norm:

llfll = sup{lf(t)l:te K}.

The Krein-Kakutani theorem, which will be proved in the next section, says
that an arbitrary Kakutani space can be mapped isomorphically and isometrically
onto a dense subset of the Kakutani space C(.K) for some compact Hausdorff space
K. By composing the injective map <p: <€(X) ~ fl of Theorem 10, with the linear
isometry and lattice isomorphism J of fl into C(K) guaranteed by the Krein-
Kakutani theorem, we immediately obtain

THEOREM 12 Let X be a normed real vector space and, with the Hausdorff
metric h derived from the norm, let (<€(X),h) be the metric space of nonempty
bounded closed convex subsets of X.
Then there exist a compact Hausdorff space K and an injective map 'I' of
<€ = <€(X) into the Kakutani space C(K) of all continuous real-valued/unctions
on K with the following properties:

(i) 'If preserves sums and non-negative real multiples,


(ii) 'If preserves order and the supremum of any two elements,
(iii) 'If preserves distance,
(iv) 'lf(U) is the constant function 1,
(v) <€'='If(<€) is a convex cone in C(K) and<€'-<€' is dense in C(K). CJ

By identifying A with 'lf(A), we may regard<€ as embedded in C(K).

4 THE KREIN-KAKUTANI THEOREM

In this section we prove the Krein-Kakutani theorem, not only to make our
account self-contained but also because the existence in our case of an order unit
makes possible some simplifications. At the same time the proof may serve as a
minicourse on convexity in functional analysis.
A linear functional on a normed real vector space Vis a map/: V ~ R such
that
J(ax + (3y) = aj(x) + (3/(y) for all x,y e Vandall a,(3 e R..
4. The Krein-Kakutani theorem 197

The linear functional f is said to be bounded if there exists a non-negative real


number µ such that
lf(x)I < µllxll for all x e V.

The set V' of all bounded linear functionals on Vis also a normed vector space if we
define
(a/+ pg)(x) = aft.x) + pg(x),
llf 11 = sup {lf(x)I: llxll = 1}.

Since/ is linear, we can also write

llf 11 = sup {lf(x)I: llxll < 1 }.

For any non-zero vector x0 e V, there is a linear functional f e V' such that
llf 11 = 1 and/(xo) = llxoll. This is a special case, with p(x) =llxll, w = {a.xo: a E R}
and g(ax0) = allxoll for a e R, of the analytic form of the Hahn-Banach theorem:

(HB) Let W be a vector subspace of the real vector space V and let p: V ~ R
be a sublinear functional on V, i.e.

p(x + y) < p(x) + p(y), p(/...x) = 'Ap(x) for all x,y e Vandall real A. > 0.

If g is a linear functional on W such that

g(y) < p(y) for every y e W,

then there exists a linear functional f on V such that f(y) = g(y) for every y e W
and
f(x) < p(x) for every x e V.

Proof Let

A= {(x,a) e V x R: p(x) <a}, B = {(x,a) e W x R: g(x) =a}.

Then A is a convex subset of V x R and B is a vector subspace of V x R. We will


show that Ai= {(x,a) e V x R: p(x) <a}, which implies Ai() B = 0.
Suppose (x,a) e Ai. If P > 0 then (O,p) e A and hence, for some A.> l,
A.(x,a) + (1 -A.)(O,p) e A. Thus

'Ap(x) = p(Ax) ~ A.a + (1 - A.)p,


198 IX. Spaces of convex sets

which implies p(x) <a. On the other hand, suppose p(x) <a and choose a' so
that p(x) <a'< a. For any (y,p) e V x JR, if 0 < 0 < 1 and 0 is sufficiently close
to 1, then
p(0x + (1 - 0)y) < 0a' + (1 - 0)p(y) < ea+ (1 - 0)p.

Hence (x,cx.) e Ai, by Proposition V.27.


It follows from Proposition VIII.11 that there exists a hyperplane H of V x R.
such that B ~Hand Air. H = 0. Since any element of V x JR can be uniquely
expressed as the vector sum of an element of H and a scalar multiple of a fixed
element of Ai, there exists a linear functional hon V x JR such that h(x,cx.) = 0 for
(x,a) e Hand h(x,cx.) > 0 for (x,cx.) e Ai. For ye W we have

h(y,g(y)) = 0 < h(y,1 + p(y))


and hence
0 < h(0,1 + p(y)-g(y)) = [l + p(y) -g(y)]h(0,1).

Since g(y) ~p(y), it follows that h(O,l) > 0. For any (x,cx.) e Vx JR,

h(x,cx.) = h(x,0) + cx.h(O,l).

Hence h(x,cx.) = 0 if and only if ex= /(x): = - h(x,0)/h(0,1). Evidently f is a linear


functional on V and/(y) = g(y) for y e W. Furthermore /(x) < p(x) for any x e V,
since
0 < h(x,p(x)) = [p(x) - f(x)]h(0,1). D

Since V' is also a normed vector space, we can in the same way define the
normed vector space V" =(V1' of all bounded linear functionals on V'. For any
x e V, the map lx: V' ~ R. defined by lif) = f(x) is a bounded linear functional on
V'. Moreover IVJI = l~I, since

IVJI =sup {Vx<f)I: IVll = 1} = sup {lf(x)I: llf 1 = 1}


and since, for any x ~ 0, there exists/ e V' with llfll = 1 and/(x) = 1~11. It follows
from (HB) also that the map x ~ lx is injective. Since lcu+py = aJx + fi!y, the map
x ~ lx isactually an isometric isomorphism of V into V".
Since V' is a normed vector space, it has a topology derived from the metric
d(f,g) = llf- gll.
However, there is a weaker topology on V' which is more
convenient for our purposes. The open sets of this weak* topology are the unions
of sets of the form
4. The Krein-Kakutani theorem 199

N(f0 ,e,A) = {/ e V': lf(x) -f0(x)I < e for all x e A},

where /o e V', e > 0 and A is a finite subset of V. The weak* topology is a


Hausdorff topology, since if f,g are distinct elements of V', then/(x) ¢ g(x) for
some x e V and hence there exist disjoint neighbourhoods N(f,e,x), N(g,e,x).
The convenience of the weak* topology stems from the Banach-Alaoglu
theorem:

(BA) The closed unit ball U' = {/ e V': llf 11 S 1} of V' is compact in the weak*
topology.

Proof lf/e U', then lf(x)I Sllxll and so/(x) lies in the compact interval
Ix= C- llxll,llxlll. If P = Ilxevlx then U' ~ P, since Pis the set of all functions
g: V -> R such that g(x) e Ix for every x e V. Moreover the topology U' inherits as
a subspace of Pis the weak* topology. ButP is a product of compact spaces and
hence itself compact, by Tychonoffs theorem. Hence, to show that U' is compact,
we need only show that it is a closed subset of P.
Let g e P be a point in the closure of U'. Then g maps V into R and
lg(x)I <II.xii for each x e V. Suppose z =ax+ py, where x,y e V and a,p e R.
For each e > 0, the set N = {h e P: lh(t) - g(t)I < e for all t e A}, where
A = {x,y,z}, is an open subset of P which contains g. Since g is in the closure of
U', there exists/ e U' r.N, and since/is linear,/(z) = af<.x) + pj(y). Hence

lg(z) - ag(x) - pg(y)I < (1 + !al + IPl)e.

Since this inequality holds for each e > 0, we inust actually have

g(z) = ag(x) + pg(y).

Thus g is a linear functional, and moreover g e U'. a


Another result which will be required is the Krein-Milman theorem:

(KM) If C is a nonempty weak* compact convex subset of V', then the set K of
all extreme points of C is nonempty and C is the weak* closure of the convex
hull of K.

Proof The family ~ of all nonempty weak* compact faces of C is nonempty, since
it contains C itself. If we regard ~ as partially ordered by inclusion then, for any
200 IX. Spaces of convex sets

F e '!J, there is a maximal totally ordered subfamily ~ containing F. The


intersection E of all members of ~ is a face of C. Moreover E is nonempty and
weak* compact, since each member of ~ is weak* compact. Thus E e '!J and no
proper subset of Eis in '!J. We will show that Eis necessarily a singleton.
Assume on the contrary that E contains distinct points f,g of V'. Then
f<.x) ¢. g(x) for some x e V. Put

µ = sup {h(x): he E},

and let D be the subset of all points of E at which the supremum is attained. Since
E is weak* compact, D is nonempty and weak* closed. Hence D is also weak*
compact. Moreover D is convex, since if h1th 2 e D and h = 9h 1 + (1 - 9)h 2 ,
where 0 < 9 < 1, then
h(x) = 9µ + (1 - 9)µ = µ.

On the other hand, if he D and h = 9h 1 + (1 - 9)h 2 , where h 1,h 2 e E and


0 < 9 < 1, then h 1,h 2 e D, sinceµ= 9h 1(x) + (1 - 9)h 2 (x) and h 1(x) < µ,
h2(x) S µ. Consequently D is a face of E, and therefore of C. But D is a proper
subset of E, since it does not contain both/ and g. Thus we have a contradiction.
This proves that any nonempty weak* compact face of C contains an extreme
point of C. H cl> is a bounded linear functional on V', a similar argument shows that
the set of all points of C at which cl> attains its supremum is a nonempty weak*
compact face of C. Consequently the supremum is attained, in particular, at an
extreme point of C.
Now let A be the weak* closure of the convex hull of the set K of all extreme
points of C. Then A ~C and hence A is weak* compact. Assume there exists a
point/ e C \A. By applying Proposition VIII.8 to the sets U1 + 2pU' and A for
some small p > 0, we see that there exists a hyperplane of V' such that U1 + pU' is
contained in one of the open half-spaces associated with this hyperplane and A in
the other. Thus, as in the proof of (HB), there are a linear functional cl> on V' and a
real number y such that cl><!+ pu') > y for every u' e U' and cl>(g) < y for every
g e A. Since this implies

[y- cl>(f)]/p < cl>(u') < - [y- cl>(/)]/p for all u' e U',
4. The Krein-Kakutani theorem 201

the linear functional ct> is actually bounded. Hence ct> attains its supremum on C at
an extreme point e of C. Since e e A, cl>(e) < y< ci><IJ. Hence/ e C, which is a
contradiction. a
A real vector space V is said to be a vector lattice if a partial order < is defined
on V with the properties

(i) if x,y e V and x < y, then x + z < y + z for every z e V and A.x < A.y for
every real A. > 0,
(ii) any two vectors x,y e V have a least upper bound x v y.

It follows that any two vectors x,y e V also have a greatest lower bound
x " y, namely
x "y = - ((-x) v (- y)).

We also have the following simple properties: if x,y,z e V and A.> 0, then

(x v y) + z =(x + z) v (y + z), (x "y) + z =(x + z) " (y + z),


A.(x v y) = (Ax) v (A.y), A.(x A y) =(Ax) A (A.y) if 'A > 0,
(x v y) + (x A y) = x + y.

[Proof Put
u = x v y, v = (x + z) v (y + z), w =(Ax) v (A.y).

From x,y < u we obtain x + z < u + z, y + z < u + z and hence v < u + z. On the
other hand from x + z < v, y + z < v we obtain x < v - z, y < v - z. Hence
u < v - z and u + z < v. This proves that v = u + z, and similarly we can prove that
w = AU if A > 0.
The corresponding relations with v replaced by " follow immediately. Also,
by what we first proved,

-(x "y) +x = ((-x) v (-y)) +x =0 v (x-y)


and hence
-(x "y) +x +y =0 v (x-y) +y =y v x. C]

These relations imply the lattice distributive laws: ifx,y,z e V, then

(x v y) " z = (x " z) v (y " z), (x "y) v z = (x v z) " (y v z).


202 IX. Spaces of convex sets

[Proof Put
u =(x " z) v (y " z), v =x v y.

Since x "z < v "z and y A z <v A z, we certainly have u < v A z. On the other
hand, since
x A z =x + z - (xv z), y A z =y + z -(y v z),
we have
x + z - (x v z) < u, y + z - (y v z) < u,

and hence x + z < u + (v v z), y + z < u + (v v z). It follows that v + z <


u + (v v z), i.e. v A z Su. This proves the first distributive law, and the second is
a consequence. 0]

In particular, if we put

x+ =x v 0, x- = - (x A 0) = (- x) v 0,
then
x+ ~ 0, x- ~ 0, x+ A x- = 0,
x+-x- =x, x+ +x- =xv (-x).

For convenience of writing we put

V+={xe V:x~O}.

We will say that a vector lattice V is a Kakutani space if it has the additional
properties

(iii) if x,y e V and nx < y for every positive integer n, then x < 0,
(iv) there exists e e V such that, for each x e V, there is a positive integer n
for which-ne <x < ne.

For any x e V, put


llxll = inf {A.> 0: -A.e Sx < A.e}.

It is easily verified that with this definition the Kakutani space Vis a normed vector
space. Moreover norm and order are connected by the properties

(N4) llxll = llx+ + x-11 for every x e V,


(NS) llx v Yll = max {llxll,llYll} if x,y e V+·
4. The Krein-Kakutani theorem 203

The real line R is itself a Kakutani space, with x v y = max {x ,y}, x A y =


min {x,y} and e = 1. It can be illuminating to use the lattice notations also in this
case, even though their use is easily avoided.
As mentioned in Section 3, the vector space V = C(.K) of all continuous real-
valued functions on a compact Hausdorff space K is a Kakutani space. In fact Vis
complete with respect to its norm, since R is complete and the limit of a uniformly
convergent sequence of continuous functions is again a continuous function. The
Krein-Kakutani theorem says, in effect, that every complete Kakutani space is of
this type.
Let V be a Kakutani space and let V' be the vector space of all bounded linear
functionals on V. We will show that V' can also be given the structure of a vector
lattice.
If f,g e V' we define f S g if/(x) < g(x) for every x e V +· This evidently
defines a partial order on V' with the property (i) in the definition of a vector lattice.
It remains to show that any two elements of V' have a least upper bound.
For any f,g e V' and any x e V+' put

h(x) = sup {.f(y) + g(x-y): 0 < y Sx}.


Then
f(x) < h(x), g(x) < h(x) for every x e V +·

Furthermore, if I e V' is a linear functional such that

f(x) < l(x), g(x) < l(x) for every x e V+'


then
f(y) + g(x - y) < l(y) + l(x - y) = l(x) for 0 < y < x,

and hence
h(x) < l(x) for every x e V+·

It is clear from the definition of h that h('A.x) = A.h(x) for every x e V + and
every A. e R+, and lh(x)I S <IYll +Ilg ll>llxll. We will show that also

If 0 < y 1 < x 1 and 0 < y 2 < x 2 , then y = y 1 + y 2 satisfies 0 < y S x 1 + x 2 •


Conversely, if 0 Sy Sx 1 +x2 and if we put y 1 =x 1 A y, y 2 = y -y 1, then
Y1 + Y2 = y, 0 < y 1 S x 1 and
204 IX. Spaces of convex sets

Hence

h(Xi +X2) = sup utY1) + /(y2) + g(x1-Y1) + g(x2-Y2): 0 <Yi < Xtt 0 < Y2 < X2}
= h(x1) + h(x2 ).

We show next that there exists a linear functional k: V ~ R such that


k(x) = h(x) for every x e V +· Every x e V admits the representation x =X+ - x_,
where x+,x- e V+. In general, if x = y- z, where y,z e V+, we define

k(x) = h(y) - h(z).

This definition does not depend on the particular representation of x as the


difference of two elements of V +· For if x = y' - z', where y',z' e V +' then from
y+ z' =y' + z we obtain

h(y) + h(z1 = h(y + z1 = h(y' + z) = h(y1 + h(z)


and hence
h(y) - h(z) = h(y1 - h(z') .

It may be verified similarly that k is linear,

k(x + x') =k(x) + k(x1, k(-x) =- k(x), k(A.x) =A.k(x) for A. e R+,

and thus is the required least upper bound of/and g.


Because a Kakutani space V has more structure than an arbitrary normed vector
space, the supremum norm on the dual space V',

IVll = sup {V<x)I: llxll < 1 },


also has additional properties. These properties will now be investigated.
First, let/: V ~ R be any linear functional such that/(y) > 0 for every y > 0.
If x e V and llxll = 1 then, for any£> 0, -(1 + e)e <x < (1 + e)e and hence

- (1 + e)/(e) S/(x) < (1 + e).f{e).

Consequently lftx)I <.f{e). Thus/ e V' and IVll =.f{e).


In agreement with the notation already introduced for V, put
,
V+ = {f e V':f(y) > 0 for every ye V+l·
4. The Krebi-Kakutani theorem 205

If/ e v; then llf 11 = f(e), by what we have just proved, and hence
(NS)' llf + gll = lltll + llgll iff.g e v;.
This property replaces the property (NS) in V. It will now be shown that the
property (N4) continues to hold in V', i.e.

(N4)' llfll = IJf+ + f-11 for every f e V'.

We have

llf+ + f-11 = f+(e) + f-(e)


= sup {fly): 0 Sy < e} + sup {-f(z): 0 S z < e}
= sup {fly- z): 0 Sy,z Se}.

If we put w = y - z, then
w+ - w- =w, w+ + w- =w v (- w).
Hence
IJf+ +f-11 = sup lf(w): w+ + w- Se}
or, since llw+ + w-11 = llwll,

IJr + f-11 = sup lf(w): llwll S 1} = llf 11.


Now put
C = {fe v;: llfll = 1}.
As we have seen, we can also write

c = {fe v;:f(e) = 1}.


It follows that C is a convex subset of the unit sphere in V' and that C is closed in
the weak* topology. Hence, by the Banach-Alaoglu theorem, C is weak* compact.
The extreme points of the convex set C may be characterized in the following way:

(LH) A linear functional f e C is an extreme point of C if and only if it is a


lattice homomorphism, i.e.

f(x v y) = f(x) v f(y), f(x A y) = f(x) A /(y) for all x,y e V.

Proof Suppose first that/is a lattice homomorphism. If/is not an extreme point
of C, then/= 8/1 + (1 - 8)/2 for some/1/ 2 e C with/1 ~ / 2 and some 8 e (0,1).
206 IX. Spaces of convex sets

Iff(x+) = 0 for some x e V, then / 1(x+) = / 2(x+) = 0, and similarly iff(r) = 0 then
/ 1 (x-) = / 2 (x-) = 0. If /(x) = 0 for some x e V, then /(x+) = f(x-) = 0, since
/(0) = 0 and/ is a lattice homomorphism. Since /;(x) = /;(x+) - /;(r) (i = 1;2), it
follows that/(x) =0 implies/1(x) =/ 2(x) =0.
The set H = {z e V:/(z) = 0} is a hyperplane of V which passes through the
origin. We have shown that / 1 (z) = / 2 (z) = 0 for every z e H. Since
/ 1(e) =/ 2(e) = 1 and every x e V has the form x = z + ae for some z e H, a e R,
it follows that/1(x) =a= / 2(x). Thus/1 = / 2, which is a contradiction.
Suppose next that f is an extreme point of C. Assume first that
f(x+) "/(r) = 0 for every x e V. Since x = x+ -r, we have f(x) = /(x+) - /(x-).
Since f(x+) > O,/(r) ~ 0 and min {f(x+)J(.r)} = 0, by hypothesis, it follows that
/(x+) = /(x)+,f(r) =f(x)-. From x v y = y + (x -y)+ we now obtain

f(x v y) = /(y) + f((x - y)+)


= /(y) + (f(x -y))+ = f(x) v /(y),

and similarly f(x A y) = /(x) A/(y). Thus/is a lattice homomorphism.


Hence to complete the proof we will assume thatft.z+) "/(z-) :'I: 0 for some
z e V and derive a contradiction. Evidently z+ ,z- are non-zero and are not scalar
multiples of one another. Since z+ " z- = 0, it follows that az+ + Pz- ~ 0 if and
only if a~ 0, p ~ 0. Hence (az+ + Pz-)+ = a+z+ + p+z- for all a,p e R.
Define p: V --> R by p(x) =ft.x+). Then p(x) > 0 for every x e V and p(x) = 0
if x < 0. Moreover p is sublinear. Let Il be the plane determined by O,z+,z- and
define a bounded linear functional/o: Il --> R by

fo(az+ + Pz-) = af(z+) for all a,p e R.

Since (az+ + Pz-)+ = a+z+ + p+z-, we have

p(az+ + Pz-) = f(a+z+ + ~+z-)


= a+/(z+) + ~+f(z-) > a/(z+).
Thus /o(X) S p(X) for all X E Il.
It follows from the Hahn-Banach theorem (HB) that there exists a bounded
linear functional/1: V--> R such that/1(x) S p(x) for all x e V and/1(x) =/ 0(x) for
x e Il. Put / 2 =f - / 1. If x > O then / 1(x) ~ 0, since - / 1(x) = / 1(-x) s p(-x) = 0,
and / 2 (x) ~ 0, since /(x) = p(x) ~/1 (x). Evidently also / 1(z+) = f(z+) ¢ 0 and
/ 2 (z-) = /(z-) ¢ 0. Thus if we put A.;= f;(e) then A.;> 0 (i = 1,2) and A. 1 + ~ = 1.
4. The Krein-Kakutani theorem 207

Since f = A. 1(A.cl/1) + A. 2(A.2-1/2), where A. 1-1/1,A. 2-1/2 e C, and since f is an


extreme point of C, this is the required contradiction. 0

Finally we require the lattice version of the Stone-Weierstrass theorem:

(SW) Let K be a compact Hausdorff space containing more than one point, let
C(K) be the Kakutani space of all continuous real-valued functions on K, and
let E be a vector sublattice of C(K), i.e. E is a subset of C(K) such that if
f,g e E and a,p e R then a/+ pg e E and f v g e E.
Suppose also that E contains the constant function 1 and that E separates
the points of K, i.e. if s,t e K and s '¢ t then there exists f e E such that
f(s) ~ f(t).
Then Eis dense in C(K), i.e. for any he C(K) and any e > 0, there exists
f e E such that llf- hll < e.

Proof We first observe that, for any distinct points s,t e Kand any a,p e JR, there
exists/ e E such that/(s) = a,f(t) = p. Indeed if g e Eis such that g(s) "¢ g(t), we
can t.ake
f = [g(s) -g(t)]-1( (a- ~)g + ~g(s)- a.g(t) }.

Lethe C(K) and, for any distinct points s,t e K, letfst e E be such that
fs 1(s) = h(s) andfs,(t) = h(t). For given e > 0, define open subsets Ost of K by

Ost = {u e K:fs1(u) > h(u)-e}.

Then, for each t e K, K =U seK 0 st' since s e 0 st· Since K is compact, there is a
finite subcover:

If ft= /s 11 v ... v fsmt' then ft e E,ft(t) = h(t) andft(u) > h(u) - e for all u e K.
Now define an open subset 0 1 of Kby

0 1 = {u e K:ft(u) < h(u) + e}.

Then K = U teK 0 1, since t e 0 1• Let

K = 0 11 u ... u 0 1n

be a finite subcover. If we put/= ft 1 A ••• A ftn, then/ e E and, for every u e K,

h(u)- £ <f(u) < h(u) + £. a


208 IX. Spaces of convex sets

We are now in a position to prove the Krein-Kakutani theorem:

(KK) If V is a Kakutani space, then there is a linear isometry and lattice


isomorphism of V onto a dense subset of the Kakutani space C (K) of all
continuous real-valued functions on some compact Hausdorff space K.

Proof Let
c = lf e v;: llfll = 1} = lf e v;:Jte) = 1 },

and let K be the set of all extreme points of C. The theorem will be established in a
very explicit manner by proving

(i) K is a compact Hausdorff space with the weak* topology on V',


(ii) for each x e V, the map lx: K---+ R defined by lif) = f(x) is a continuous
function on K,
(iii) the map J: V---+ C(K) defined by J(x) = lx has the properties

J(ax + py) = a.l(x) + (il(y) for all x,y e Vandall a,p e R,


IV°(x)ll = 1~11 for all x e V,
J(x v y) = J(x) v J(y), J(x " y) = J(x) " J(y) for all x,y e V,

(iv) J(V) is dense in C(K) andJ(e) is the constant function 1.

We have already seen that the weak* topology on V' is a Hausdorff topology,
that C is a compact convex subset of V', and that K is the set of all/ e C which are
lattice homomorphisms. It follows that K is closed, and hence compact. Since the
map J~: V' ---+ R defined by J~ (j) = /(x) is a bounded linear functional, its
restriction lx to K is certainly continuous and

Moreover, since every f e K is a lattice homomorphism,

Thus J(V) is a vector sublattice of C(K). Furthermore le is the constant function 1,


and iff,g are distinct elements of K thenlx(f) -:1:JxCg) for some x e V. Hence, by
(SW), J(V) is dense in C(K).
It remains to show that llJ(x)ll = 1~11 for every x e V. In fact we need only
verify this for every x e V+· For if x* = x+ + x- = x v (-x), then 1~*11=1~11 and,
4. The Krein-Kakutani theorem 200

since every f e K is a lattice homomorphism,/(x*) =max {f(x)/(-x)} = lf(x)j;


hence
l~(x)ll = sup {lftx)I: f e K}
= sup {f(x*): f e K} = l~(x*)ll.

Suppose now that x ~ 0. Since

llxll = sup f.ftx):/ e V', lltll < 1 },

it follows from (KM) that


llxll = sup {/(x):/e K'},

where K' is the set of extreme points of the unit ball U' = {/ e V', lltll ~ 1 }. It is
evident that if/ e K', then lltll = 1 and hence, by (N4)'-(N5)',

llf 11 + lltll = llr + t-11 = lltll = 1.

It follows that either j+ = 0 or/- = 0, since otherwise

I = IJf+ll <rtllfll) + lltll <-rtlltll)

would not be an extreme point of U'. Since x ~ 0, in evaluating llxll we can restrict
attention to those I e K' for which I= r' and these I are in K. Hence J is an
isometry:
llxll = sup f.ftx):/ e K} = l~(x)ll- D

5 NOTES

Hausdorff (1927) defined the 'Hausdorff' metric and proved Proposition 3.


Blaschke (1916) had already proved Proposition 9.
RAdstrom (1952) proved Proposition 4 and essentially showed that '€(X) could
be embedded in a nonned vector space. The lattice properties were added by
Pinsker (1966). Schmidt (1986) makes the connection with the Krein-Kakutani
theorem and gives applications of Theorem 10 to the theory of random sets and
inteival mathematics.
If Xis complete in Theorem 10, then'€' is also complete, by Proposition 8.
However, Debreu (1967) shows that~ is not complete even in the caseX = R.2. It
should be noted also that if'€ is regarded as actually embedded in ~, then one must
210 IX. Spaces of convex sets

distinguish between -A and (- l)A. In fact the two are distinct for any A e '€with
IAI > 1.
The Krein-K.akutani theorem is usually formulated for complete Kakutani
spaces, or 'AM-spaces with unit' in the official terminology. It was independently
proved by M.G. and S.G. Krein, and by K.akutani, in the years 1940-1941. The
theorem is discussed in a more general context in the books of Day (1973),
Schaefer (1974) and Meyer-Nieberg (1991). Tychonoffs theorem is proved in
Hewitt and Stromberg (1975), for example.
References

J. Bair and R. Fourneau (1975/80), Etude geometrique des espaces vectoriels


(2 vols.), Lecture Notes in Mathematics 489 and 802, Springer-Verlag,
Berlin.
M.L. Balinski (1961), On the graph structure of convex polyhedra in n-space,
Pacific J. Math. 11, 431-434.
I. Barany (1982), A generalization of Caratheodory's theorem, Discrete Math. 40,
141-152.
I. Barany and D.G. Larman (1992), A colored version of Tverberg's theorem,
J. London Math. Soc. 45, 314-320.
M.K. Bennett and G. Birkhoff (1985), Convexity lattices, Algebra Universalis
20, 1-26.
W. Blaschke (1916), Kreis und Kugel, Veit, Leipzig. [2nd ed., de Gruyter,
Berlin, 1956]
V.G. Boltyans.kii (1975), The method of tents in the theory of extremal problems,
Russian Math. Surveys 30, no. 3, 1-54.
T. Bonnesen und W. Fenchel (1934), Theorie der konvexen Korper, Springer-
Verlag, Berlin. [Reprinted, Chelsea, New York, 1948]
N. Bourbaki (1987), Topological vector spaces, Chapters 1-5, Springer-Verlag,
Berlin.
A. BrfZJndsted (1983), An introduction to convex polytopes, Springer-Verlag,
New York.
H. Brunn (1913), Ober Kemeigebiete, Math. Ann. 73, 436-440.
H. Busemann (1955), The geometry of geodesics, Academic Press, New York.
J.R. Calder (1971), Some elementary properties of interval convexities, J. London
Math. Soc. (2) 3, 422-428.
C. Caratheodory (1911), Ober den Variabilitatsbereich der Fourierschen
Konstanten von positiven harmonischen Funktionen, Rend. Circ. Mat.
Palermo 32, 193-217. [Reprinted in Gesammelte mathematische
Schriften, Band 3, pp. 78-110, Beck, Miinchen, 1955]
W.A. Coppel (1995), A theory of polytopes, Bull. Austral. Math. Soc. 52, 1-24.
H.S.M. Coxeter (1989), Introduction to geometry, 2nd ed., reprinted by Wiley,
New York.
212 References

L. Danzer, B. Grilnbaum and V. Klee (1963), Belly's theorem and its relatives,
Convexity (ed. V. Klee), pp. 101-180, Proc. Symp. Pure Math. 7, Amer.
Math. Soc., Providence, R.I.
M.M. Day (1973), Normed linear spaces, 3rd ed., Springer-Verlag, Berlin.
G. Debreu (1967), Integration of correspondences, Proc. Fifth Berkeley Symp.
Math. Statist. Probability (ed. L.M. Le Cam and J. Neyman), Vol. II,
Part l, pp. 351-372, University of California Press, Berkeley.
J.-P. Doignon (1973), Convexity in cristallographic lattices, J. Geom. 3, 71--85.
J.-P. Doignon (1976), Caracterisations d'espaces de Pasch-Peano, Acad. Roy.
Belg. Bull. Cl. Sci. (5) 62, 679-699.
P. Duchet (1987), Convexity in combinatorial structures, Rend. Circ. Mat.
Palermo (2) Suppl. No. 14, 261-293.
J. Eckhoff (1979), Radon's theorem revisited, Contributions to geometry (ed. J.
Tolke and J.M. Wills), pp. 164-185, Birkhauser, Basel.
J. Eckhoff (1993), Belly, Radon, and Caratheodory type theorems, Handbook of
convex geometry (ed. P.M. Gruber and J.M. Wills), Volume A, pp. 389-
448, North-Holland, Amsterdam.
P.H. Edelman and R.E. Jamison (1985), The theory of convex geometries, Geom.
Dedicata 19, 247-270.
J.W. Ellis (1952), A general set-separation theorem, Duke Math. J. 19, 417-421.
M. Erne (1984), Chains, directed sets and continuity, Institut fur Mathematik,
Universitit Hannover, Preprint Nr. 175.
Euclid (1956), The thirteen books of Euclid's elements, English translation by
T.L. Heath, 2nd ed., reprinted in 3 vols. by Dover, New York.
R. Frank (1992), Ein lokaler Fundamentalsatz fiir Projektionen, Geom. Dedicata
44, 53-66.
H. Grassmann (1844), Die Iineale Ausdehnungslehre, Leipzig. [Reprinted in
Gesammelte Werke 11, Chelsea, New York, 1969]
B. Griinbaum (1967), Convex polytopes, Interscience, New York.
R. Hammer (1977), Beziehungen zwischen den Satzen von Radon, Belly und
Caratheodory bei axiomatischen Konvexit!ten, Abh. Math. Sem. Univ.
Hamburg 46, 3-24.
F. Hausdorff (1927), Mengenlehre, 2nd ed., de Gruyter, Berlin. [English
translation of 3rd ed., Set theory, Chelsea, New York, 1962]
E. Belly (1923), Ober Mengen konvexer Korper mit gemeinschaftlichen Punkten,
Jahresber. Deutsch. Math.-Verein. 32, 175-176.
G. Hessenberg (1905), Beweis des Desarguesschen Satzes aus dem Pascalschen,
Math. Ann. 61, 161-172.
E. Hewitt and K. Stromberg (1975), Real and abstract analysis, 3rd printing,
Springer-Verlag, New York.
A. Heyting (1980), Axiomatic projective geometry, 2nd ed., North-Holland,
Amsterdam.
References 213

D. Hilbert (1899), Grundlagen der Geometrie, Berlin. [12th ed., Teubner,


Stuttgart, 1977; English translation of 10th ed., Foundations of geometry,
Open Court, LaSalle, Ill., 1971]
R.B. Holmes (1975), Geometric functional analysis and its applications,
Springer-Verlag, New York.
R.E. Jamison (1974), A general theory of convexity, Dissertation, University of
Washington, Seattle.
R.E. Jamison-Waldner (1982), A perspective on abstract convexity: classifying
alignments by varieties, Convexity and related combinatorial geometry
(ed. D.C. Kay and M. Breen), pp. 113-150, Dekker, New York.
B. Jonsson (1959), Lattice-theoretic approach to projective and affine geometry,
The axiomatic method (ed. L. Henkin et al.), pp. 188-203, North-
Holland, Amsterdam.
V. Klee (1959), Some characterizations of convex polyhedra, Acta Math. 102,
79-107.
F. Klein (1873), Ueber die sogenannte Nicht-Euklidische Geometric (Zweiter
Aufsatz), Math. Ann. 6, 112-145.
B. Korte and L. Lovasz (1984), Shelling structures, convexity and a happy end,
Graph theory and combinatorics (ed. B. Bollobas), pp. 219-232,
Academic Press, London.
Z. Kovijanic' (1994), A proof of Barany's theorem, Publ. Inst. Math. (Beograd)
(N.S.) 55 (69), 47-50.
S. Lang (1984), Algebra, 2nd ed., Addison-Wesley, Reading, Mass.
M. Lassak (1986), A general notion of extreme subset, Compositio Math. 57, 61-
72.
K. Leichtweiss (1980), Konvexe Mengen, Springer-Verlag, Berlin.
H. Lenz (1992), Konvexitat in Anordnungsraumen, Abh. Math. Sem. Univ.
Hamburg 62, 255-285.
F.W. Levi (1951), On Belly's theorem and the axioms of convexity, J. Indian
Math. Soc. 15, 65-76.
F. Maeda and S. Maeda (1970), Theory of symmetric lattices, Springer-Verlag,
Berlin.
K. Menger (1936), New foundations of projective and affine geometry, Ann. of
Math. 37, 456-482.
P. Meyer-Nieberg (1991), Banach lattices, Springer-Verlag, Berlin.
F.R. Moulton (1902), A simple non-desarguesian plane geometry, Trans. Amer.
Math. Soc. 3, 192-195.
J.G. Oxley (1992), Matroid theory, Oxford University Press, New York.
M. Pasch (1882), Vorlesungen ilber neuere Geometrie, Teubner, Leipzig. [2nd
ed. with appendix by M. Dehn, reprinted by Springer-Verlag, Berlin, 1976]
214 References

G. Peano (1889), I principii di geometria logicamente esposti, Fratelli Bocca,


Torino. [Reprinted in Opere scelte, Vol. II, pp. 56-91, Cremonese,
Roma, 1958.]
G. Pickert (1981), Projectivities in projective planes, Geometry - von Staudt's
point of view, (ed. P. Plaumann and K. Strambach), pp. 1-49, Reidel,
Dordrecht, Netherlands.
A.G. Pinsker (1966), The space of convex sets of a locally convex space
(Russian), Leningrad Inzh.-Ekonom. Inst. Trudy Vyp. 63, 13-17.
A.V. Pogorelov (1991), On a theorem of Beltrami, Soviet Math. Dok/. 43, 83--85.
W. Prenowitz (1961), A contemporary approach to classical geometry, Amer.
Math. Monthly 68, no. 1, part II, 67 pp.
W. Prenowitz and J. Jantosciak (1979), Join geometries, Springer-Verlag, New
York.
S. Priess-Crampe (1983), Angeordnete Strukturen: Gruppen, Korper, projektive
Ebenen, Springer-Verlag, Berlin.
J. Radon (1921), Mengen konvexer Korper, die einen gemeinsamen Punkt
enthalten, Math. Ann. 83, 113-115. [Reprinted in Collected works,
Vol. 1, pp. 367-369, Birkhauser, Basel, 1987]
H. RAdstrom (1952), An embedding theorem for spaces of convex sets, Proc.
Amer. Math. Soc. 3, 165-169.
W. Rinow (1961), Die innere Geometrie der metrischen Rii.ume, Springer-
Verlag, Berlin.
G.S. Rubinstein (1964), Theorems on the separation of convex sets (Russian),
Sibirsk. Mat. Zh. 5, 1098-1124.
H.R. Salzmann (1967), Topological planes, Adv. in Math. 2, 1-60.
K.S. Sarkaria (1992), Tverberg's theorem via number fields, Israel J. Math. 19,
317-320.
H.H. Schaefer (1974), Banach lattices and positive operators, Springer-Verlag,
Berlin.
K.D. Schmidt (1986), Embedding theorems for classes of convex sets, Acta Appl.
Math. 5, 209-237.
R. Schneider (1993), Convex bodies: the Brunn-Minkowski theory, Cambridge
University Press.
A. Schrijver (1989), Theory of linear and integer programming, Wiley,
Chichester.
F. Schur (1909), Grundlagen der Geometrie, Teubner, Leipzig.
B. Segre (1956), Plans graphiques algebriques reels non Desarguesiens et
correspondances cremoniennes topologiques, Rev. Roumaine Math. Pures
Appl. 1, no. 3, 35-50.
A. Seidenberg (1976), Pappus implies Desargues, Amer. Math. Monthly 83, 190-
192.
Re/ere11ces 215

G. Sierksma (1984), Extending a convexity space to an aligned space, Nederl.


Akad. Wetensch. lndag. Math. 46, 429-435.
D.E. Smith (1959), A source book in mathematics, Dover, New York.
V.P. Soltan (1984), Introduction to the axiomatic theory of convexity
(Russian), Shiintsa, Kishinev.
E. Spemer (1938), Zur Begriindung der Geometric im begrenzten Ebenenstiick,
Schriften der Konigsberg Gelehrten Gesellschaft 6, 121-143.
M. Spivak (1975), A comprehensive introduction to differential geometry,
Vol. IV, Publish or Perish, Boston, Mass.
L.A. Steen and J.A. Seebach Jr (1978), Counterexamples in topology, 2nd ed.,
Springer-Verlag, New York.
J. Stoer and C. Witzgall (1970), Convexity and optimization in finite
dimensions I, Springer-Verlag, Berlin.
L.W. Szczerba (1972), A paradoxical model of Euclidean affine geometry, Bull.
Acad. Po/on. Sci. Ser. Sci. Math. Astron. Phys. 20, 845-851.
L.W. Szczerba and A. Tarski (1979), Metamathematical discussion of some affine
geometries, Fund. Math. 104, 155-192.
F.A. Toranzos (1967), Radial functions of convex and star-shaped bodies, Amer.
Math. Monthly 14, 278-280.
S.N. Tschemikow (1971), Lineare Ungleichungen, VEB Deutscher Verlag der
Wissenschaften, Berlin. [Russian original, 1968]
H. Tverberg (1966), A generalization of Radon's theorem, J. London Math. Soc.
41, 123-128.
H. Tverberg and S. Vreeica (1993), On generalizations of Radon's theorem and the
ham sandwich theorem, European J. Combin. 14, 259-264.
F.A. Valentine (1964), Convex sets, McGraw-Hill, New York.
M.L.J. van de Vel (1993), Theory of convex structures, North-Holland,
Amsterdam.
0. Veblen (1904), A system of axioms for geometry, Trans. Amer. Math. Soc. 5,
343-384.
0. Veblen and J.W. Young (1910/18), Projective geometry (2 vols.), Ginn,
Boston, Mass. [Reprinted by Blaisdell, New York, 1965]
K.G.C. von Staudt (1856/7), Beitriige zur Geometrie der Lage (2 vols.),
Niirnberg.
C.T.C. Wall (1972), A geometric introduction to topology, Addison-Wesley,
Reading, Mass.
D.J.A. Welsh (1976), Matroid theory, Academic Press, London.
N. White (ed.) (1992), Matroid applications, Cambridge University Press.
G.M. Ziegler (1995), Lectures on polytopes, Springer-Verlag, New York.
Notations

< 4, 51, 195 [z,x> 103


<, >, > 4 (z,x> 104
v 4, 190, 195,201 [L,S> 105
/\ 4, 201 Kt 107
[S] 6, 31 P(V) 120
E(S) 9 V, pn(D) 121
ISi 12 a+ b 138
<S> 20, 55 a·b 140
dimX 24 A +B, aA, 0 150
[a,b] 28, 30, 31, 49 d(x,y), (X,d) 178
[a,b), (a,b], (a,b) 33 1111 179, 189, 196, 202

K(S) 35 ~(X) 185

<a,b> 50 S(y,A), h(A,B) 186


H+, H_ 59 CX(X) 187
'tr 70 U, <g(X), A ED B 189
As 72 ~(X) 193
A 0 73 C(K) 195
H+: L, [p1,P2] 81 V' 197
Ci, C 87 x+, x-, V+, e 202
,
AB 93 V+ 204
AIB 98
Axioms
(01)-(03) 4 (P) 39
(Al)-(A3) 6 (Ll)-(L4) 49
(Hl)-(H3) 6 (D) 85
(H4) 7 (U) 95
(AO),(HO) 8 (A) 124
(tE) 14 (S) 158
(E) 20 (Fl)-(F3) 175
(A3)' 25 (Tl)-(T3) 178
(C),(Ll) 31 (Dl)-(D3) 178
(L2) 36 (Nl)-(N3) 179, 189
(L3) 37 (N4),(N5) 202
(L4) 38 (N4) ',(NS)' 205

Propositions
(P)' 54 (C)",(P)" 92
11.2', 11.17', 11.19' 54 11.2", 11.17", 11.19" 93

Examples
11.1,2 28 Vl.2,3,4 116
Il.3 29 VI.5 117
11.4,5,6,7 ,8 30 Vl.6 122
11.9 34 VI.7 123
11.10 46 Vlll.1 158
Vl.1 115
Index

addition of points 138 bounded


affine linear functional 197
basis 56 set 178
generator 56
hull 20, 55 cancellation law 190, 153
independent 56 Caratheodory
set 20, 55, 99, 165 number 25
algebraic hull operator 7 set 12, 34, 86
aligned space 8 Caratheodory's theorem 67
alignment 6 Cauchy sequence 179
normed 8 closed
alternative precedence structure 25 half-space 61, 165
AM-space with unit 210 set 178
anti-exchange unit ball 189
alignment 14, 36 codimension 23
property 14, 16, 17 collinear points 51
antimatroid 16 compact topological space 178
asymptote 163 complete
·linear geometry 157, 173
Banach-Alaoglu theorem 199 metric space 179
Barany's theorem 68 cone 104, 113
basic convex set 97 conic hull 105
basis 9 contraction of an alignment 13
Beltrami's theorem 182 convex
binary relation 4 closure 87
Blaschke selection principle 192 geometry 31
Index 219

convex face
hull 6, 31 of polyhedron 110, 111
partition 42, 97, 164 of polytope 74, 76, 78
set 6, 28, 31 of sum 152, 154
cover 21 proper 10
open 178 facet
covering property 21 of convex set 73
of polyhedron 110
d-connected graph 80 of polytope 76, 79
dense linear geometry 85 factor geometry 81, 132
Desargues property 124, 127 filter 175
converse of 127 fundamental theorem
Desargues theorem 121, 136 of ordered geometry 173
converse of 122 of projective geometry 155
deviation 186
dimension 23, 57 generating set 9
directe.d family 25
division ring 28 Hahn-Banach theorem 167, 197
ordere.d 28, 142, 172 half-space 61
Hausdorff
Eckhoff's conjecture 70 distance 186
e.dge 73 metric 186
Euler-Poincare relation 84 space 178, 97
exchange Bausdorffs maximality theorem 5
alignment 20, 56 Belly
property 20 number 25
extension theorem 167 set 11, 38, 66
extreme Belly's theorem 67
point 8, 32, 153, 199 hemispace 42
subset 25 hull operator 7
algebraic 7
face hyperbolic geometry 116
of convex set 9, 33, 71, 84 hyperplane 22, 57
of intersection 10, 90 'hypeiplane' 120
220 Index

incomparable elements 4, 30 long line 158


independent set 9 lower bound 4
infimum 4
intersection matroid 25
of aligned spaces 14, 16 metric 178
of 'lines' 120 space 178
intrinsic interior 87, 96 Minkowski sum 150
irreducible linear geometry 62, 64 multi-cross 65
irredundant representation 109 centre of 65
isomorphism of linear geometries 143 multiplication of points 140

.. norm 179
JOin

of alignments 13, 16, 24 normed


of convex geometries 43 alignment 8
of linear geometries 62 vector space 179, 189
space 114
open
Kakutani space 202, 194 cover 178
kernel of a set 35, 36 half-space 61, 165
Krein-Kakutani theorem 208, 196 set 178, 97
Krein-Milman theorem 199 optimization 182
order isomorphism 4, 159
lattice 4 order-preserving map 4
complete 4 ordered division ring 28, 142, 172
homomorphism 205
isomorphism 4 Pappus property 155
vector 201, 194 partially ordered set 4, 30
L-conic hull 105 Pasch's axiom 57
line 50 Petersen graph 136
'line' 119 plane 57
linear Busemann's 117
functional 196 Moulton's 116
geometry 54 projective 120
linguistics 18 point 31, 49
Index 221

'point' 119 restriction


polygon 79 of an alignment 13, 16, 24
polyhedron 109, 79 of a convex geometry 43
polytope 74, 90, 113
positive element 28, 142 sand-glass property 40
product scalar multiple 150
of points 140 sector 81
of sets 93 segment 28, 30, 31, 49, 82
projective half-open 33
completion 121 open 33
geometry 119, 135, 155 semilattice 30
plane 120 separated sets 166
space 119, 135 properly 166
projective plane •120 separation
Segre's 136 properties 164
Veblen's 123 theorems 41, 166
shelling structure 18, 25
quotient simplex 145, 75, 91
of affine sets 102 singleton 5
of convex sets 99, 101 spherical geometry 116
of sets 97 star-shaped set 35
Stone-Weierstrass theorem 207
Radon subcover 178
number 25 'subspace' 120
set 11, 37, 66 sum
Radon's theorem 67 of convex sets 151, 153
Rldstrom's cancellation law 190, 209, 155 of points 138
ray 103 of polytopes 154
open 104 of sets 150
refracted plane 116 supporting
representation of polyhedron cone 169
irredundant 109 half-space 169
redundant 109 hypetplane 169
222 Index

supremum 4
symmetric convex cone 108

totally ordered set 4, 45, 51, 115, 159


topological space 178, 97
topology 178
of a metric space 179
tree 30
Tverberg number 70
Tverberg's theorem 69

ultrafilter 176
unending linear geometry 95
upper bound 4

vector
lattice 201, 194
sum 150
vertex
of a cone 104
ofagraph 80
vertex set
of a cone 104
of a convex cone 104
of a polyhedron 113
visible point 35

weakly modular lattice 102, 114


weak* topology 198

Potrebbero piacerti anche