Download as pdf or txt
Download as pdf or txt
You are on page 1of 128

Number 10°

BrankoGaba
Anrangements a
THE LIBRARY

CARLETON COLLEGE
NORTHPIRLD. MN 85049

«R33 Griinbaum, Branko

BORROWER'S NAME
7] iY
SIN

GAYLORD 45
DATE DUE
'
\ iL

ae
_

° a = ie
Digitized by the Internet Archive
in 2022 with funding from
Kahle/Austin Foundation

https://archive.org/details/arrangementsspreOO000grun
ee ;
=a

oe 7 al) a ; :
- . :

el ea
7 ca

S er: a
0h

= Sees py ee i
aes ine NERY fates iN sta EAT es ae
=

sD

=
se SO
opaeeen ae rll:
Aone ons matin musty ‘sienna ani Ah
S-
7 : «A Siete mF ste, CHEM eresstipgl re
a
oe

benegaset extent Scot HEae


aoe Ot) atest yydcenteA mm: tule tee? het W gon) ae
~ = "TEE! ahve toe ethe wt an wie Gnu cae 46
ee ee rs ae |et ys ines F
7 — wade eledew
herrea
iah wt temo race eed 5.6
a? _ Pew pien wee! Beya rue ean inisteokt 1m
innaleaieciels silo
pAe rab ity 7

Ba : af i “—
= See righ ‘Phas ks an

~ f Titi lei: r
+ CNET ON eObLear Seu |
i, SAFIBL 9,
a MO tiie . _
Other Monographs in this Series

N Ome Irving Kaplansky: Algebraic and analytic aspects of operator algebras


2 . Gilbert Baumslag: Lecture notes on nilpotent groups
3 . Lawrence Markus: Lectures in differentiable dynamics
4 . H.S.M. Coxeter: Twisted honeycombs
Ds George W. Whitehead: Recent advances in homotopy theory
6 . Walter Rudin: Lectures on the edge-of-the-wedge theorem
7h. Yozo Matsushima: Holomorphic vector fields on compact Kahler manifolds
8 . Peter Hilton: Lectures in homologica!l algebra
9 . I. N. Herstein: Notes from a ring theory conference
Conference Board of the Mathematical Sciences
REGIONAL CONFERENCE SERIES IN MATHEMATICS

supported by the
National Science Foundation

sate
49 ~JNZTDA
/ Yb Of

Number 10

ARRANGEMENTS AND SPREADS

by

’ BRANKO GRUNBAUM
Se

, THE LIBRARY
CARLETON COLLEGE
NORTHFIELD, MINN. 55057

Published for the

Conference Board of the Mathematical Sciences

by the

American Mathematical Society

Providence, Rhode Island


Expository Lectures
from the CBMS Regional Conference
held at the University of Oklahoma, Norman, Oklahoma
June 21—25, 1971

International Standard Book Number 0—8218—1659—4


Library of Congress Catalog Card Number 71—38926
AMS 1970 Subject Classifications: 05—O0, 05—02,

05B99, 05C10, 50—00, 50A20, 50B20, 50B99, 50D20,


52—00, 52A10, 52A40, 55A20, 55A99.

Copyright © 1972 by the American Mathematical Society


Printed in the United States of America

All rights reserved except those granted to the United States Government.
May not be reproduced in any form without permission of the publishers.
CONTENTS
thee HLUERORC
TTS 1) jes oS SOR IR ers a ean lake a eee 1
hag TNTEAS
SST OSS) HE oa el oc eo n ele n e 4
Zee eSOlMOLDMSHI-LPES
Y OF ATCANGERICINUS io:cs seo 5c 401-eva eck oan seus coos coreuecacese ices loorestecoeest 4
2.2 Relations between the numbers of lines, vertices, edges and cells ........cccccseceseeseee 9
ASS, SIAN S SWE) BLAU TECELTU NZNO]1E0 | Ceo Ry nan te CE 16
Peo NUD CLSIOR CONUS OL Vat lOUS KANGS eoaecauses je cog b.5.8)s vapaurastassieeeS aula iad taredarspeean ner? 25
Ze MARIANI OTha ATE ATISCINOINDS covcotacerte es cere ce sheaves ahdchvagn (i ogsasdeentcoac cops eps os 33
POsVAThANSeMeNTS ASSOCIATE WITH SCTS'OL POLIS. cvcenncasscescneszeisescadsncdacsveavacouseealecos cans 36
Da BO tacte Ob ONS ANG CLASSI CAUONS 6 28: sans ce Beto es Sroc is 556 task oaningde tes Bc ok te tanec ca ero 37
3. Arrangements of pseudolines and arrangements of CUrVes ..............:cesceececeeeeeeeeeneees 40
SS SCUOOUNES ancdumonestr DIC Arran Pements, <n. 2.2 c<cicccccysevxesceues Levapensaneevncsaceese
Chaet 40
B.7 oome.- results on arrangements of pSCuUdOMNeS......-..2.¢: secacoev.neeeeunetansdactonsee
-ecectacee.sss 45
3.3) pAttangenrents.ol simple curves in the Euclidean plane. 2.0.0... sosncspesccessen-oaeeeroeesesees 25)
Epa ML aH ALLL IOIS eleanor wiih Sp aes otSoresfiery eae sees on as Coen ce saa aaetaoi chGey sae Coe meates 68
AN PYSPERCAVG ISINROL aS MORRO EC ECE a OE OPER, oR See RY Pee eR DET a
tal eeDe tTdICON ANG PIODETUICS OL, SPTCALS 2, a. wc <a oteaaaa sco isbstacew rte nossoepeounaadsupaasssaecvananas rei
BF MaisANANOLE SHOE SPE AOS peer awh ocoe oe an we Ns ete saan spac her neg cased oe scanas anaes aaeeesestge 81
AeA Thal CCCACItS AiG SpiedCSpada ec gh apoien “cde aacacasnd Suparleran onsevne Go apteats hea etches 85
AAs OPOLORICAl PLAN Su... sex vars cove needa seiiw eget este 16 Ere aimed vase Sag souks ophal aps seeveaecencdeenecpenes 86
TREUETENICESRE or ieee BE. AAI cbt, SE SHES bse cneadearierrreny Heeiae Sege SD
PNOVeS AUC AMINO TOOL coe, firicc 5, drersucc Ou ioe, fonts weesertaeevete tek doahuehacdin souatuac ne cgeaie toes tensteal 112
i

Onoeers® Tice
; }
7 a Bee
Ff A
7 7a : a
© ie) Oe

ATA
iy aa ail, 7 » eu 3 > \ie
+ wi Fri
b : 14, = Liters § aiows tt i

- jaye! a8 aL cae v¢ ue s@ Ge thes nil

© or « Ske Ujta tye uebenl)) 918 sae


et ri eh os Ta : (25 ty Ved i ee a

a ¥ ,
ole
aes ls ip etl elt sree ald
Pw aee! Ce
( prea)
ae >sol ’ iuupelil:
Pi 7 iP pe yt : irene

m. 5) rv Protea a Dae gy a i +e 1" “) éa


she ea by « j
Ss FY Be pe ty rity, u yy seed ai ‘ ae f

‘Of. eu ' a ak ee Bird)


2am rs :
2

; a at i as—
: 7 ‘ait as etbada) ‘ag ¥ mens

ii ts 7 7 ae es Ai ® aaa a

yo OCR: ans 80 OSS.) aseae ne


1

wees 77) + ij 7 .
ARRANGEMENTS AND SPREADS*

Chapter 1. Introduction
The present survey deals mostly with rather elementary mathematics,—so elementary in
fact that most of its results and problems are (or at least should be) understandable to under-
graduates. It was written out of the conviction that many neglected aspects of elementary
geometry deserve a wider dissemination, —because of their inherent beauty and interest, and
for the inspiration and understanding they can impart to our students and to ourselves.
During the last decade or two, it has become an article of faith with many of us that
research in mathematics and its teaching are worthwhile only if they deal with very general
and abstract topics. The elements of algebra, set theory, logic, topology, measure theory,
etc., — that is what advanced undergraduates and starting graduate students are being taught
in good schools; it would be foolish to say that this should be changed. What is unfortunate,
and what cripples most of the students for the rest of their professional lives, is that other
topics and points of view are only rarely presented. The students are made to understand
that research in, say, finite groups is “important”—no matter how esoteric the problem in-
vestigated,
— while topics in the Euclidean plane, for example, are certainly old-fashioned
since the Euclidean plane is a very special structure. Many aspects of geometry fare partic-
ularly badly from this attitude; they are even denied the right to exist unless they fit nice
algebraic patterns. In the most ridiculous and extreme aspects of such tendencies geometry
is equated with a subdiscipline of linear algebra.
As a counterthrust to that trend I would like to recall just a few of the many historical
instances in which the geometric patterns of thought or points of view led not only to
another theorem or two but were crucial in gaining new and revolutionary insights:
(i) The discovery of irrational numbers;
(ii) The invention of calculus;
(iii) The development of the axiomatic method;
(iv) The emergence of algebraic topology, and of functional analysis.

*) The first version of this survey was presented as an invited address at the Annual Spring Meeting
of the Michigan Section of the Mathematical Association of America held on April 4, 1970, at the Wayne
State University in Detroit, Michigan.
Research sponsored in part by the National Science Foundation through the Science Development
Program grant GU—2648, and by the Office of Naval Research through contract N00014—67—A—0103—
0003.
0) BRANKO GRUNBAUM

geometric
In most of these examples (and in many other cases) it was the spirit of
that the geomet-
thinking more than any specific fact that led to the breakthrough. I believe
ns, and
ric point of view will provide in the future similarly important impulses and inspiratio
lead to still unexplored disciplines and insights. Therefore, I am convinced that geometric
thinking ought to be cultivated among students much more than is the case at the present.
In the following pages I have surveyed a number of topics considered to be not-quite-
legitimate mathematics by some of our colleagues—though the bibliography shows that
others, of no lesser stature, were willing to think about them. One of my aims was to point
out the wealth of open problems in many areas in which some partial answers are already
known. The main thrust, however, was to show how this elementary and intuitively appeal-
ing material exhibits a far-reaching analogy to the subtle differences encountered in topology
between the piecewise linear and the topological situations. In our case, the main dichotomy
is between the rectilinear and the curvilinear arrangements, though in each type additional
subclasses are worth exploring.
An additional objective was the juxtaposition of the discrete and the continuous vari-
ants. This is again analogous to the situation in topology, but in the present context it
seems so far to have escaped attention.
For simplicity the whole exposition is restricted to the two-dimensional case. There is
no difficulty to extend many of the definitions and results to higher dimensions, but only in
a few places have I found it worthwhile to mention such material. (For a recent survey of
the results known about arrangements of hyperplanes in higher-dimensional spaces see the
first part of Griinbaum [1971].) Much of the beauty of the subject derives, in my opinion,
from the fact that already the two-dimensional case is interesting and non-trivial.
The exposition is organized as follows:
Arrangements of lines, that is the structures determined by finite families of straight
lines in the real projective plane, are examined in detail in Chapter 2. Isolated problems in
this area have been investigated for almost 150 years, but the only previous attempt at a
systematic exposition seems to be a short section in Griinbaum [1967]. Possibly part of the
reason for this neglect is to be found in the fact that quite a few of the problems and re-
sults extend naturally to various more general settings—some to the theory of geometric con-
figurations, to finite projective planes and to different kinds of combinatorial designs, some
to convex polyhedra in 3-space or to planar graphs, and so on. However, it is my firm con-
viction that there are great advantages to be derived from the consideration of all those
problems in a unified setting, although that implies some loss of generality for each individual
result. (This situation seems to be rather analogous to that in the theory of convex sets in
Euclidean spaces: There are very few results about such sets that could not be extended to
more general sets or spaces, the generalization varying from result to result. But the mutual
relationship among the various properties is best understood when they are considered in
their common core of validity.) Hence the present exposition is limited almost exclusively
to the real projective plane, with only a few hints at other possible settings. For all asser-
tions that are not selfevident or verifiable by exhaustive checking we provide either proofs or
references to the literature.
ARRANGEMENTS AND SPREADS 3

In Chapter 3 we deal mainly with arrangements of pseudolines in the real projective


plane, and arrangements of simple curves in the Euclidean plane. These are quite natural gen-
eralizations of the topic of Chapter 2, although they received very little attention in the litera-
ture. Many of the facts and conjectures mentioned are new, and it is hoped that they will
give impetus to additional research. The feeling that parallel consideration of vaguely related
topics may inspire meaningful advances led to the inclusion in Chapter 3 of the material on
selfintersection patterns, crossing patterns, etc.
The emphasis on the finite and the combinatorial which prevails in Chapters 2 and 3
changes to considerations of a more topological nature in Chapter 4, in which spreads of
curves are discussed. Spreads may be considered as continuous analogues of arrangements of
curves, and this point of view leads to a large number of open problems. However, most of
the facts known about spreads are related to various properties of convex sets, and an appre-
ciable part of the mainly expository Chapter 4 is devoted to this background.
By adding one more “degree of freedom” to the spreads of curves one obtains “topo-
logical planes”. This possibility is briefly discussed in the last section of Chapter 4, in which
an attempt is made to show the relation of arrangements and spreads to the recently very
popular theory of topological projective planes and related structures.
* * *

Throughout the survey, many conjectures are explicitly stated and many additional
problems are hinted at. It is rather obvious that the subject offers extremely varied oppor-
tunities for research. Due to the elementary nature of the topic, the solutions of many of 41

the open problems will probably require more inspiration than erudition—but is that not a
hallmark of beauty in mathematics?

University of Washington Branko Grunbaum


Seattle, September 1971

1. Symbols of this kind refer to notes added in proof starting on page 112.
4 BRANKO GRUNBAUM

Chapter 2. Arrangements of lines

2.1. Isomorphism-types of arrangements.


real
Arrangements of straight lines are among the simplest objects one may study in the
projective plane P. It is easily seen that similar investigations could be made in the Euclidean
plane E*; however, since the projective variant is somewhat simpler, and since most of the
information from the projective case may be utilized in the Euclidean by the simple ex-
pedient of adjoining the “line at infinity”, we shall concentrate our attention on arrange-
ments in the projective plane. Such arrangements (or their Euclidean counterparts) have
been studied, with varying emphasis, since Steiner [1826], von Staudt [1847], Schlafli
[1852], Wiener [1864], and Sylvester [1867]. Considering the large number of papers
written on different aspects of the subject it is rather surprising that no systematic exposition
is available. (A first attempt in that direction was made in Chapter 18 of Gruinbaum [1967].)
By an arrangement of lines A we mean a finite family of n= n(A) lines L,, °*+,L,,
in the real projective plane P. If there exists a point common to all lines L; we shall call
the arrangement trivial. Unless the opposite is explicitly stated we shall in the sequel assume
that all arrangements we are dealing with are non-trivial; therefore also n > 3. If no point
belongs to more than two of the lines L; the arrangement is called simple.
With an arrangement A there is associated the 2-dimensional cell complex into which
the lines of A decompose P (see, for example, Veblen-Young [1918, Chapter 9], Carver
[1941]). The vertices, edges, and cells (or polygons) of that complex are also said to belong
to the arrangement, and their numbers are denoted by f(A), f,(A), and f,(A). Two
arrangements are said to be isomorphic provided the associated cell complexes are isomorphic;
that is, if and only if there exists an incidence-preserving one-to-one correspondence between
the vertices, edges and cells of one arrangement and those of the other. The totality of ail
mutually isomorphic arrangements forms an isomorphism-type of arrangements. If all the
cells of an arrangement are triangles we shall say that the arrangement and its isomorphism
type are simplicial. Clearly the notions of simple and simplicial arrangements, and the num-
bers n(A) and f(A) are isomorphism invariants of A. (Some other equivalence classes of
arrangements shall be discussed later, but for most of the exposition the appropriate and
natural notion is that of isomorphism.) It is of some interest to note that although simple
arrangements have been studied frequently, simplicial arrangements appear only in Melchior
[1940]. This is rather surprising since the simplicial arrangements frequently occur as solu-
tions of extremal problems (see, for example, Theorem 2.5 below).
One of the first problems concerning arrangements is the determination of the numbers
c(n), c§(n) and c4(n) of different isomorphism-types of arrangements, simple arrangements,
or simplicial arrangements of n lines. Analogous enumeration problems for polytopes, com-
plexes, and graphs are notorious for thein difficulty, but it seems that even less is known con-
cerning the enumeration of types of arrangements. The few available results are as follows:
THEOREM 2.1. c(3)= 1; c(4) = 2; c(5)= 4; e(6)= 17.
ARRANGEMENTS AND SPREADS

Figure 2.1
The different isomorphism types of arrangements of 3, 4, 5, and 6 lines. (Vertices
belonging to 3 or more lines are indicated by small circles.)
6 BRANKO GRUNBAUM

In Figure 2.1 representatives of all types with n <6 are shown. (An erroneous value
for c(6) was quoted in Griinbaum [1967, p. 394].)
THEOREM 2.2. (3)=C4)=C6)=1; €@)=45 C-11-
For proofs of Theorem 2.2 and for representatives of the different types see Cummings

A A AL AS
N(3) N (4) N (5) N(6)

R(6) NC) A, (7)

N (8) R(8) N(Q)


R(Y)
Figure 2.2
The different isomorphism t ypes of simplic
i iciial arrangements wi i
(Line at infinity included in R (9).) ; oe
ARRANGEMENTS AND SPREADS 7

[1932a], [1932b], [1933], White [1932], R. Klee [1938], Griinbaum [1967, p. 394]. From
the work of Canham [1971] and Halsey [1971] (see also below, page 43) it may be conjec-
tured that c*(8) = 135.
THEOREM 2.3. ¢°(3) 234) = cA GS
cP (6) =c°(7) =e" (8) =.c*(9).=c4 (1) 52;
E> (10) =:e4 G2)=—"4;
C3) eee GAs;
CAS y="
The simplicial arrangements with at most 9 lines are illustrated in Figure 2.2; for addi-
tional illustrations see Griinbaum [1971].
Possibly of more interest than the enumeration results of Theorem 2.3 is the following
rather unexpected conjecture, a variant of which goes back to Melchior [1940].

is Nessa
i
iy 2s
SAE
f
cx <SI
ee

Figure 2.3
111
A simplicial arrangement of 16 lines (line at infinity included) with 38 vertices,
edges, and 74 cells (triangles).
Conjecture 2.1. For n> 38 we have
2 Tf rn=O0f1,2 (Gneds)
ENaS
1 ifin=3 (mod 4).

More precisely, the following three infinite families of simplicial arrangements are known:

SK
SSR
iS
OEE AIO
LLIN NOS.
NANA

Figure 2.4
The only, known type of simplicial arrangement with no non-trivial combinatorial
symmetries, (Line at infinity is included.)
ARRANGEMENTS AND SPREADS 9

(1) The near-pencil N(n),n > 3, consisting of n—1 lines through one point and one
line missing that point;
(2) The regular arrangement R(2k), k >3, consisting of the k lines determined by
the edges of a regular k-gon in the Euclidean plane and the k axes of symmetry of that
k-gon;
(3) The regular arrangement R(4m + 1),m > 2, consisting of the 4m lines of
R(4m) together with the “‘line at infinity”.
Besides these three infinite families of simplicial arrangements only 91 other types are
known (Grtinbaum [1971]), each one having at most 37 lines. (The “Catalogue of simpli-
cial arrangements” in Griinbaum [1971] contains 90 of those types; the additional type
was discovered meanwhile, and it is illustrated in Figure 2.3.)
Another remarkable feature of the known types of simplicial arrangements is revealed
in the following conjecture (compare Griinbaum [1971]), in which by “combinatorial sym-
metry” of an arrangement we mean any isomorphism of the arrangement with itself.

Conjecture 2.2. Except for a finite number of isomorphism types, each simplicial
arrangement of lines has a non-trivial group of combinatorial symmetries.

In fact only a single such type of arrangement is known: it contains 28 lines and is
shown in Figure 2.4.
* * *

Arrangements of lines have many relations to the theory of convex polyhedra (3-dimen-
sional convex polytopes). Some of those common properties are consequences of formal
connections; see, for example, Coxeter [1962], McMullen [1971], Griinbaum [1971]. How-
ever, one of the most fruitful aspects of those relations is the inspiration provided by analo-
gies between the two fields. For example, it is well known (see Steinitz-Rademacher [1934,
p. 347]) that if P’ and P” are isomorphic convex polyhedra then P’ is homotopic either
to P” or else to a polyhedron which is obtained from P” by symmetry in a plane. (Two
polyhedra are homotopic if one may be continuously transformed into the other through
isomorphic intermediate polyhedra.) Using the analogous notion of homotopic arrangements
we venture:
Conjecture 2.3. Every two isomorphic arrangements of lines are homotopic.

Conjecture 2.3 is open even for simple arrangements. In a related circle of ideas Ringel
[1957] proved that every two simple arrangements with the same number of lines are trans-
formable into each other by a finite sequence of steps each of which is either a homotopy,
or else the “switching” of a triangle determined by lines of the arrangement (see Figure 2.5).

2.2. Relations between the numbers of lines, vertices, edges and cells.
From various points of view it would be of interest to determine the set of all points
in the Euclidean 4-space representable in the form (n(A), fo(A), f,(A), f,(A)), where A
varies over all arrangements. Though nothing has been published on that set, very many re-
sults are known concerning its projections on the various coordinate planes and 3-spaces.
10 BRANKO GRUNBAUM

One of the simplest and best known such results is the Euler relation; though it holds
more generally for arbitrary cell decompositions of the projective plane, in the case of arrange-
ments it becomes particularly elementary. As is easily established by induction (see, for ex-
ample, Veblen-Young [1918, Section 185]), the numbers f; (i= 0, 1, 2) of vertices, edges,
and cells of each arrangement A satisfy Euler’s relation:
THEOREM 2.4. f,(A) — f,(A) + f(A) = 1.
It is not hard to prove (see Grunbaum [1967, pp. 401—402]) that the numbers of
faces of various dimensions satisfy also the linear inequalities

(*) 1 + f,(A) < f(A) S 2fofA) — 2.

Indeed we have:
THEOREM 2.5. The inequalities (*) determine the convex hull of the set of pairs
(fo(A), f,(A)) for all arrangements A. Moreover, equality on the left holds in (*) if and
only if A is a simple arrangement, while equality on the right is characteristic for simplicial
arrangements.

Figure 2.5
The “switching” of a triangle.

Unfortunately, as may be seen from a plot of the pairs (f9(A), f,(A)) (see Figure 2.6)
it seems that there is little hope of completely characterizing the set of those pairs. (In Fig-
ure 2.6 a numeral n is placed at each point which corresponds to an arrangement A with
n(A) =n; in case two different values of n occupy the same position, they are placed in a
circle. Figure 2.6 is complete for n <8 only.)
Theorem 2.5 is analogous to the result of Steinitz [1906] (see also Griinbaum [1967,
p. 190]) that the pairs (f,,f) which correspond to convex polyhedra are characterized by
the inequalities

4< (fy +4)/2<f,


<2f) -4.
Howeyet in that case the situation is simpler in as much as every pair (fo, f,) satisfying
the
inequalities indeed corresponds to some convex polyhedron.
ARRANGEMENTS AND SPREADS 11

Concerning the pairs (n(A), f,(A)) we have


THEOREM 2.6. For every arrangement A with n(A)=n we have

3657 60175 849 1011 1213141516 1718 19 20 21 22 23 24 25 26 27 28

Figure 2.6
The pairs (fo> fy) corresponding to arrangements of lines.
12 BRANKO GRUNBAUM
n
n<f,(A)< ie)
with equality at right if and only if A is a simple arrangement, and on the left if and only
if A is the near-pencil N(n).
While the assertions concerning the upper bound are completely trivial, those regarding
the lower bound are less immediate. Various proofs of that part were given by Erdos [1943],
Steinberg [1944], de Bruijn-Erdés [1948], Hanani [1951], [1954], Motzkin [1951], Bouten-
de Witte [1965], de Witte [1966a], [1966b], Ryser [1968], Buekenhout-Doyen [1970, p. 41].
The inequality fy =n deals only with the numbers of lines and of their points of in-
tersection. Hence it is meaningful in situations much more general than arrangements of
lines and may equivalently be formulated in the dual setting (“How many lines are deter-
mined by n points?’”’). As a matter of fact, most of the proofs listed above apply to such more
general incidence systems, or to the dual situation. The following proof, taken from Baster-
field-Kelly [1968], shows how few assumptions are needed for the validity of fy 27.
Let L,,°*:,L,, be the n lines and assume that they determine fy <n vertices
Vee ce, Vio: For each vertex V let t(V) be the number of lines incident with V, and
for each line L let r(L) be the number of vertices incident with L. Notice that if V; €
L, then r(l;) > ¢(V;), since each line through V, intersects L,. We have

wo are) 1 1
Le a aS —__—_—_—_——
2 fo~ (Lx) x fo - r(Lx) 2 fo — Vj)
VjOL K=O VOL k=o

fo n-1Vj)
=> mm
22 eee
n
j=1 To ty) it

and the contradiction reached shows that fo 2n.


The set (n(A), fo(A)) has not been completely determined, but some interesting re-
sults going beyond the inequalities of Theorem 2.6 are known about it. For example, it is
easily seen that the values fy = ( :)~ 1 and f, = is )-—3 are impossible for all n. This
is complemented by the following result of Erdos [1971a]:

THEOREM 2.7. There exist an integer n* and areal c>O such that for all n>n*
all pairs (n, fy) with Cn! * <= fos () — 4 correspond to arrangements.

On the other hand, the situation at the lower end of the range of values of fo is more
complicated. First, there is the result of Kelly-Moser [1958] (for n > 27, extended to the
values 10 <n < 26 by Elliott [1967]):
THEOREM 2.8. For n(A)=n> 10 either A= N(n) is the near-pencil with n lines
and f(A) =n, or else fy(A) >2n- 4.
An extension of this result was found by R. K. Guy [1971a] and Erdés [1971la]. It
refutes a conjecture of Griinbaum [1967, p. 404], and may be formulated as follows:
ARRANGEMENTS AND SPREADS 13

THEOREM 2.9. If k >3, if 1 <p <(25k* — 33k +.9)/2 (hence certainly if 1<p<
67) and if 2n > 27k* — 33k + 11, then there exists no arrangement A with n(A)=n and
We ee i ea Kgs.2)/2— p.
The result of Theorem 2.9 is probably not best possible in the sense that the conclusion

f, = 28
27
26
25
24
23
22

21
20

19
18
17
16
15
14
13
12
11
10

vy
aA
FmHA
Ww
«-
@

Figure 2.7
The pairs (n, fo) for n < 12 and fo < 28.
14 BRANKO GRUNBAUM

in which
may well hold even for values of n smaller than assumed. But already for k = 3,
precise
case pairs with fy = 3” — 12, fy = 3n — 13, etc., are shown to be impossible, the
range of validity is not known.
Figure 2.7 shows the existing pairs (n(A), fo(A)) for m<12 and fo < 28; each exist-
ing pair is indicated by a small cross. Most notable is the absence of the pair (12, 21).
Somewhat analogous is the situation regarding the pairs (n(A), f,(A)), although they
do not seem to have been considered in the literature. We shall discuss two such results.

THEOREM 2.10. Each arrangement A satisfies


n(A)
2n(A) - 2 <f,(A) < fits
a)

Equality on the left holds if and only if A is the near-pencil with n lines; equality on the
right holds if and only if A _ is simple.
PROOF. The upper bound follows from the observation that the number of cells does
not decrease if the lines of an arrangement are subjected to sufficiently small perturbations,
which change the given arrangement into a simple one. For simple arrangements (and only
for such arrangements) we have f, = O + 1, as is easily checked by induction. The lower
bound f, 2 2n — 2 is also easily established using induction. Indeed, it obviously holds for
small n (such as n= 3,4, 5,6). For n2>4, let L,, bea line of A such that the other
n—1 lines of A form a non-trivial arrangement A*. Then EA) & 2m—-b) —' 25 bet Jo
partitions at least two cells of A*, hence f,(A) 22 + ES = 2n-—2. The characteriza-
tion of the cases of equality also follows at once by induction.
In analogy to the situation described in Theorem 2.8 we have:
THEOREM 2.11. If n(A)=n and if A is not a near-pencil then

PRooF. The assertion is again easily checked for small n. Hence we assume n > 6
and the theorem established for arrangements of less than n lines. Since A is not a near-
pencil, the omission of any line of A yields a non-trivial arrangement. If the omission of
line L,, from A yields the near-pencil Ni —1) then L,, partitions at least n—2 cells of
N@—1) and thus f,(A) >n-2+f,(N@- 1))=n-2 + 2n-4=3n-6. If the omis-
sion of no line of A yields a near-pencil, let L,, be a line of A which is incident with at
least 3 vertices; such a line exists since n > 6. Then the arrangement A* obtained by omit-
ting L, from A satisfies, by the inductive assumption, Fete) = 3(ni= ll) 6. sincent,:
partitions at least 3 cells of A*, the desired result follows.
It is not hard to show that for n(A) >7 the equality in Theorem 2.11 characterizes
arrangements A consisting of n —2 lines through one point, the remaining two lines inter-
secting at a point on one of those n— 2 lines. By letting the last two lines intersect at a
point not on any of the first nm — 2 lines, an arrangement with f, =3n—5 is obtained. It
is also possible to show that, for sufficiently large n, if f, >3n—S5 then i, 2 4n=— 12:
ARRANGEMENTS AND SPREADS 1s)
As a matter of fact it may be shown that there exists no arrangement A with n(A) = 9 and
f, = 23, and we venture
Conjecture 2.4. If A is an arrangement with n(A)=n>9 and if (Ay = 3n 15
then f,(A) > 4n — 12.
(@) ie} O
to OO
Onan O10 ° fo=n-12
45 é - “OU OMAOne - O/-
OnigiOni.O °
Oy °
OMNI tol woO
Ol 1 °
ko bs = & = - = 0. - Om =O fe - 9
e) (@) o
OMiInol He
One Ome Ose O f,=3n-6
OnOnO 9
35 aos ey e O PO SOL Oy :
(e} (9) Oo fe
Onn OmO 9
° ° fe
° ° e
30 - - - - - - 0-0 f/f - - -
° @ ©
e) e) O D i
Qo © fe
fe) O° O fj =ene2

25 = - - - S46) = Claro =
Oo ° e e
°
oO 2) jf $@ O

oe 0 O
20 - - - - 0 - G/ - - 9 - - -
fe) fe '
[@) O O

re)
O) O

15 e

10

gi Hoty Yi215 VN a5)


Figure 2.8
The pairs (n, fo) known to exist. The diagram is complete for n < 10.
16 BRANKO GRUNBAUM

Results concerning the pairs (n, f,) analogous to Theorems 2.7 and 2.9 may be estab-
lished by easy variations of the proofs of Erdos [1971a].
The pairs (n(A), f,(A)) known to exist are shown in Figure 2.8. The diagram is com-
plete for n < 10.
It seems very likely that all the above results have analogues dealing with the pairs
(n(A), f,(A)); however, it seems that such questions have not been considered in the literature.
2.3. Vertices of given multiplicity.
In order to discuss additional results and problems concerning arrangements it is con-
venient to introduce the following notation. For a vertex V of an arrangement A let us
denote by 1(V) the multiplicity of V, that is the number of lines of A incident with V.
We shall denote by tA) the number of vertices of A which have multiplicity precisely /.
Also, we shall denote by r(A) the number of lines of A incident with precisely j vertices
of A, and by pA) the number of j-gons among the cells of (the complex associated with)
A. Then, by counting various incidences it is easily seen that

fo(A)= >, tA)


j22

n(A)= >> (A)


J22

f,(A) = as DAA)
S
jJ23

D> it(A)=2 j22


af, (A)=2 j>2 Dy 1A)
Do (A= j23

(= (2) 400.
j22

Many interesting facts and questions deal with the mutual relationships of the numbers
n(A), f(A), t)(A), 7(A), and pA). We shall next discuss some of the results and open
problems of this area.
Sylvester [1893] seems to have been the first to observe that if m points in the real
projective plane are not collinear then there exists a line containing precisely two of the
points. Using duality this may equivalently be formulated as:

THEOREM 2.12. Each non-trivial arrangement A satisfies

t,(A) > 0.
The proof of Theorem 2.12 hinted at in Sylvester [1893] was not convincing, but the
problem was forgotten and the whole question was quiescent till the early 1930’s. At that
time it was independently raised by P. Erdds and others. For various contributions to the
rather curious history of this now famous “‘Sylvester’s problem”, and for different proofs of
Theorem 2.12 (some in more general settings) see Erdés [1943], Steinberg [1944],
ARRANGEMENTS AND SPREADS e/

Coxeter [1948], [1949, p. 16], [1961, pp. 65, 181], [1962], de Bruijn-Erdés [1948],
Fejes-Toth [1948], Trevisan [1949], Dirac [1951], Motzkin [1951], Yaglom-Yaglom [1954],
Lang [1955], Hadwiger-Debrunner [1955], [1960, p. 58], Hadwiger-Debrunner-Klee [1964,
pp. 3, 42, 57] , Hadwiger-Debrunner-Yaglom [1965, pp. 11, 72], Williams [1968], Chakerian
[1970], Edelstein [1970], Buekenhout-Doyen [1970, p. 55]. As proposed by Anonymous 2
[1964] , it would seem to be of interest to clarify the axiomatic background of the “Sylvester
property” expressed by Theorem 2.12. For a related problem see Serre [1966]. For some
results on matroids which may be considered as generalizations of parts of Theorem 2.12 see
Murty [1969], [1970].
Contrary to widely held opinions, the first valid published proof of Theorem 2.12
appears in Melchior [1940]; Melchior was not aware of Sylvester’s problem, and he proved
the stronger result:

THEOREM 2.13. For every non-trivial arrangement

t,(A) 2 3.
PRooF. Using the above relations involving the t,’s and the D;, and applying Euler’s
relation we have

eB (3 =I; a3 >> (3 — k)p, = She =f, = 3f, at 2F; = 3% aa tis) = 3.


j>2 k>3
The inequality of Theorem 2.13 is obtained by dropping the non-positive terms on the left.
The conditions for equality in Theorem 2.13 (namely t-=p,=0 for 7 2 4; incorrectly
quoted in Coxeter [1942]) lead easily to the observation:
THEOREM 2.14. Each non-trivial arrangement A with n(A) 2 8 satisfies
t,(A) 2 4.
The above proof of Theorem 2.13 is due to Melchior [1940]; surprisingly, the fact that
t, 23 is an easy consequence of the Euler relation was rediscovered only in Kelly-Moser
[1958], although a related proof of Theorem 2.12 seems to have been supplied by N. Steenrod
(see Steinberg [1944]). Other proofs of Theorem 2.13 were given by Dirac [1951] and
Motzkin [1951]; the latter established also Theorem 2.14.
As a step beyond Theorems 2.13 and 2.14 the set of pairs ((A), t,(A)) may be inves-
tigated. Clearly t, < e ), with equality characterizing simple arrangements. As observed by
Canham [1971], it also follows from the last of the displayed equations on page 16 that f,
cannot assume any of the values Co )=juicl-7 = 1, 2, 4,5, 11..8,,141., 14, 17, Conceming
lower bounds for t,, Motzkin [1951] proved that te: i 2) >n; Kelly-Moser [1958] im-
proved this to
THEOREM 2.15. Each arrangement A satisfies

3
t,(A) our n(A).

This theorem gives the best possible estimate of the type ft, 2 cn, since the arrange-
ment A,(7) in Figure 2.2 shows that t, = 3 is possible for n= 7. However, it seems
18 BRANKO GRUNBAUM

problem
that this arrangement is the only one for which equality holds in Theorem 2.15. The
of determining the minimum of ¢, as a function of n has been mentioned very frequently

(see Dirac [1951], Grinbaum [1967, p. 404], Crowe-McKee [1968], Crowe [1969], Croft-
Guy [1971]); it has repeatedly been conjectured that t, [n/2]. More precisely, we ven-
ture

Conjecture 2.5. Except for n= 3, 4, 5, and 13, the minimum of t,(A) for arrange-
ments A with n(A)=n equals k if n= 2k, and it equals 3k if n=4k +1 or n=
Ak Bs
The exceptional arrangement with n= 13 (see Figure 2.9) was found in Crowe-McKee
[1968], where examples are also given showing that the bounds of Conjecture 2.5 cannot be
improved. (This resolves in the negative a question of Erdos [1961].) Those examples are:
the simplicial arrangements R(2k) and R(4k + 1), and the arrangement obtained from
R(4k + 4) by omitting one line.

<Ke<f

Figure 2.9
A simplicial arrangement with n= 13 and ty = 6. (Line at infinity is included.)

Recently Brakke [1971] established the validity of Conjecture 2.5 for n = 14, 16, 18,
and 22.
The information available on the pairs (n(A), t,(A)) is collected in Figure 2.10. The
data are complete for n <9. The known bounds for t, are indicated by the solid line,
while the thin line and the dashed line indicate the bounds of Theorem 2.15 and Conjecture 2.5.
For additional open problems concerning t, see Erdés [1961] and Croft-Guy [1971].
We turn now to the consideration of the relations between n(A) and EA) for Kk 2,
Some aspects of it were considered (mostly in the dual form) in the literature of mathemat-
ical puzzles and diversions; the known results are very fragmentary. Since t, = 0 is obvi-
ously possible for each k >3 and for all n, most of the interest centered on the question
ARRANGEMENTS AND SPREADS 19

how large can t, be compared to n. We shall use the notation ¢,(n) = max {f,(A)|n(A) =n}.
The first result is (H. Croft and P. Erdos, private communication):

20

10 = Welt pe OhNeer, ENS Tt


O07 OUMOWAA Z
oor A ;

10 15 25
Figure 2.10
The pairs (n, ty) known to exist. The diagram is complete for n < 9. Solid line
indicates the established minimum of f,; thin line denotes the bound ft, = [3n/7],
dashed lines the bound of conjecture 2.5. O indicates any arrangement, A a sim-
plicial arrangement.
20 BRANKO GRUNBAUM

THEOREM 2.16. For each k >3 there exists a constant c, >0 such that t,(n) >
c,n* forall n> k.
PRooF. We shall choose, in the lattice of integral points in E?, k directions deter-
mined by points of the lattice. For fixed m, let the arrangement consist of all the lines in
each of the chosen directions that can be drawn through lattice points ({ j) with 1 <i,
j<m. Then there are altogether n <b,m lines (for a suitable constant b,), k of which
pass through each of the m? chosen lattice points; hence t, > m > ER ee To complete
the proof of the theorem we only have to obtain the estimate for intermediate values of n,
a task that may be accomplished for example, by taking c, = bop:
This method yields, for example, the values c,= 1/32 and c, = 1/72, which seem to
be far from best possible. While some improvements are obviously possible using the same
idea, any essentially better estimate will probably have to use some other method. For
some results related to Theorem 2.16 see Karteszi [1963], [1964], Erdos [1971b].

Figure 2.11
The nine heavy lines show that t3(9) > 10; to obtain an arrangeme
nt with n= 10
11, 12, 13, 14, 15, or 16 lines having the maximal known
tz, the lines marked with
numbers up to n should be used together with the nine heavy lines.
ARRANGEMENTS AND SPREADS mil

In the special case k = 3 much better estimates are available. We have the following
result of A. H. Stone (private communication), which slightly improves (and greatly simpli-
fies the proof of) a result of Sylvester [1867] (see also Ball [1960, p. 105]):
THEOREM 2.17. t,(n) > [((n — 1)? + 4)/8] forall n.
PRooF. On the cubic curve y =x? in the Euclidean plane three points are collinear
if and only if the sum of their x-coordinates is 0. Therefore, if n= 2k + 1 is odd, the
points with x =0,+1,+2,++-,+k provide [(k* + 1)/2] lines with precisely 3° points;
if n= 2k the points with x =0,+1,+++,+(k-—1), k provide k(k —1)/2 such lines.
The dual configurations establish the theorem.
Sylvester [1886] (see also Ball [1960, p.105]) stated that t,(n) > [(n- 1) (1 — 2)/6] 43

=
for all n. However, no satisfactory proof seems to have been published for this assertion.
It is also not known whether for n > 13 we have t,(n)= [(n- 1) (1 - 2)/6].
The facts known about ¢,(n) for k= 3,4,5 and n < 30 are collected in Table 1.
Arrangements with large ¢, are shown in Figures 2.11 to 2.14.
As an analogue of Theorem 2.24 (below) and as a refinement of a relation in Griinbaum
[1967, p. 403] we have the following result, the proof of which may be patterned after the
proof of Theorem 2.24 by Canham [1969]. o4

Figure 2.12
Solid lines: n= 11 t= 6
All lines shown: n= 12 tg=7
With line at infinity: m= 13 ta = 9.
WD) BRANKO GRUNBAUM

Maximal value known to the author of


n \[((a-1P + 4/8) | [(@- 1) @- 2)/6] t3(A) t,(A) t.(A)

2 2 oY *] =
5
6 3 3 *4 ii! "||
*6 *D *]
4 5 5

6 7 7) eo) *]
8
9 8 9 —1Oae baie 5) +)

10 10 iD sap)
11 13 15 *2
WD 15 18 *3
13 18 DD *3
26 *4
14 Di

15 DS 30 28 &12 86
16 28 35 35 &15b, ¢c 86
87
ef,
18 $9b
19 &10b
&11
&lld
&12
&14
&16
&18
&21

#0 &40 &26
* The number is ¢,(7).
§ The number is ¢,(m) (K. A. Brakke, private communication).
& The number is probably ¢,(n).
a See Sylvester [1867].
b See Ball [1960], pp. 105—106, 127.
c See Dudeney [1907], pp. 19, 87, 103-104, 140, 172—173, 180-181.
d See Dudeney [1967], pp. 165—167, 374-376.
See Loyd [1914], pp. 38, 301, 344, 380.
f Dudeney [1967, pp. 166 and 374] gives an example with ft, = 20.

Table |
ARRANGEMENTS AND SPREADS

THEOREM 2.18. For any k distinct vertices Vas, *>, V, of an arrangement A we


have
k k
D> 0V,)
< n(A) + ()
j= »
* * *

For an arrangement A let r*(A) denote the largest value of 7 such that r(A) > 0.
Then we have the following simple result (Dirac [1951] ):

Zi
DN Le C\
\
>
\|
l
]
~

]
]
|

Figure 2.13
Subsets of the set of lines shown (and the line at infinity in some cases) yield the
arrangements with maximal known t, for n> 14 (see Table 1).
For example: Heavy lines: n=20 t,=20
All solid lines: St SY, Vee
All lines shown: n= 30 tq = 40.
24 BRANKO GRUNBAUM

THEOREM 2.19. For each arrangement A we have

r*(A)2vV nA) +1,

with strict inequality except for n(A) = 3.


Dirac [1951] also conjectured that r*(A) > [(1 + n(A))/2]. This conjecture is cer-
tainly false for small n since there exist arrangements (even simplicial ones) with the follow-
ing values:

x
ee A
INES

\
\
\

Figure 2.14
Subsets of the set of lines shown (and the line at infinity in some cases) yield the
arrangements with maximal known ts for n > 20 (see Table 1).
For example: Heavy
Ave lines: n= 20 ts = 11
Solid lines and line at infinity n=26 ts=21
All lines shown n= 30 ts = 26.
AL LANOCEMENTS AND SPREADS 25

Designation of the arrangement


mh) in Grinbaum [1971]

4,9)
15 4,5)
19 A,(09), A,09)
25 4,,(25)
31 A,(31)
37 A,(37)

No reasonable adjustment of Dirac’s conjecture has been proposed so far, and even the much
weaker conyectuse of Exdos [1961] is still open:
Conjecture 26. There exists
a real © > 0 such that r°(A)
2 cA) for all arrange-
wens b
* % z

Many othes directions of investigation into the relations among the numbers n, fy, by,
wa 1, ae possible, but concerning most of them only few results are known. Some of
these may be found in Sung-Mdchior [1936] , Jung [1937], Melchior [1937] , Koutsky-Poldk
[1969].
24. Numbers
of cells of various kinds.
We shal next investigate the numbers pfA) and we start by reporting on the relations
known to exist between the numbers n{A) and p,(A). We ase still very fas from any de-
tailed understanding of the situation, but some non-trivial results are known.
Tsxoves 2.20. For very nortriviol arrangement A we have
MAYS pfAy
Peoor. The flowing proof, due essentially to Levi [1926] (see also Moser [1964]),
is intexesting because of the convexity considesations involved. If A is a near-pencil then
pAA)=I6A)=
WA) -2> mA), therefore we may assume that A is not a neas-pencil
gh that nAy=n>3. Let L be any lineof A; we consider L as the Sine at infinity,
zak we Genteby V the set of all those vertices of A which belong to the resulting affine
hull of V is 2 polygon with at least 3 vertices v,,V2,V3,°°",
glene. Then the convex
ett, A which ts dso 2 vertex of A. The lines of A through a vertex v, determineat
least one angle the sides of which contain no other vertices of the set V. In the arrange-
ment A, these anges cossespond to triangular cells; hence L (and therefore each line of
A) contcins edges of atleast 3 triangular cells of A. Therefore the number of incidencesof
kinesA A and triangles (that is, triangulas cells) of A is at least 3n, and thus p, 70
Fos senghe assangements Theorem 2.20 has been proved already by Eberhard [1890],
whe dso Soserved that equality may hold for each n > 4, and that for each n 26 there
ots ever nontsomophic asangements with p,~n. We may, however, make the following
e If WA)= pA) then A is a simple arrangement
Conyectur2.7.
26 BRANKO GRUNBAUM

Another open question related to Theorem 2.20 is

Conjecture 2.8. If A is not a near-pencil then for each line L of A there are at
least n — 3 triangles of A which have no edge on L.

The validity of this conjecture in case of simple arrangements was asserted by Roberts
[1889], but his proof is not convincing.
Concerning the question of how large p, can be we have the following results (Grin-
baum [1967, p. 398]; R. J. Canham, private communication); here ¢(7) is the function de-
fined by $(2k) = [2k(2k — 1)/3] and (2k + 1)= [(2k + 1) (2k - 1)/3].
THEOREM 2.21. For each simple arrangement A with n(A)=n2>4 we have
p3(A) < o().
PRooF. Ina simple arrangement with n 24 it is impossible for two triangles to have a
common edge. Therefore each line of A can be incident with at most n — | triangles, and

' Figure 2.15


A simple arrangement of 12 lines
with P3 = 40.
ARRANGEMENTS AND SPREADS Dif

if n is odd with at most n—2 (this last fact was observed by R. J. Canham). Thus
3p3 <n(m— 1) forall n and 3p, <n(n—- 2) for odd n, completing the proof of Theo-
et) Doral
Probably it is possible to strengthen Theorem 2.21 to

Conjecture 2.9. Jf A is a simple arrangement with n(A)=n> 6 then


[n(n — 1)/3] for n=4 (mod 6)
p(A)<
[n(n — 2)/3] for n#4 (mod 6).
In Figures 2.15, 2.16, and 2.17 we show simple arrangements with 12, 15, 16, and 20
lines having the largest known values of p,.

Figure 2.16
Simple arrangements with maximal p3 (due to Simmons [1971]).
Lines shown n=15 p3= 65
With line at infinity: Opa 80.
28 BRANKO GRUNBAUM

Figure 2.17
A simple arrangement of 20 lines with
P3— 120.
ARRANGEMENTS AND SPREADS 29

No reasonable analogue of Theorem 2.21 is known for not necessarily simple arrange-
ments. We make the following
Conjecture 2.10. The estimate p,(A) < ¢(n) is valid for all arrangements A with
n(A)=n 216. Moreover, for n(A)=n 216 the maximum of p,(A) is attained for sim-
ple arrangements A.
For smaller values of n there are non-simple arrangements that violate both parts of
the conjecture. For example, the simplicial arrangement of 15 lines shown in Figure 2.18
satisfies p, = 66 > ¢(15) = 65. The known results on the pairs (n(A), p3(A)) are collected
in Figure 2.19. (See also Table 2 on page 56.)
Concerning p, it is not hard to establish

THEOREM 2.22. For each arrangement A with n(A)=n2>5 we have p,(A)<


n(n — 3)/2. For each n> 5 there exists a unique simple arrangement with n(A)=n and
p4(A) = nn — 3)/2.
We also have
Conjecture yh Equality in Theorem 2.22 is possible only for simple arrangements A.

The most interesting question about p, concerns the lower bound for simple arrange-
ments. It has been conjectured (see Griinbaum [1970b]) that p,(A) = 0 is possible for
simple A only in case n(A) is 3, 6, or 10. However, the recent example of Simmons

KE
AERA
ZINN

Figure 2.18
A simplicial arrangement of 15 lines with p3 = 66.
30 BRANKO GRUNBAUM

PR

120

110

100

90

80

70

60

9 90. TE Se dS ese oy, Ye i920


Figure 2.19
The pairs (n, P3) known to exist. The diagram is complete for n < 7.
Proved extrema for simple arrangements.
Known upper bounds for simple arrangements.
Conjectured upper bounds for simple arrangements.
Simple arrangement known to exist.
Simplicial arrangement known to exist.
Other arrangement known to exist.
ARRANGEMENTS AND SPREADS 31

[1971] (see Figure 2.16) shows that P4 = 9 is possible also for n= 16. We are tempted to
make

Conjecture 2.12. [f A is a simple arrangement with n(A)> 16 then p4(A) > 0.


It is unfortunate that no reasonable guess was ever made about the minimal possible
value of p,(A) for simple arrangements with arbitrary n(A)=n. The available informa-
tion on the pairs (n(A), p,(A)) is collected in Figure 2.20.

©
5: ¥ = si ; i: 7 ? -

20 = aS) 3 : ar = : =
O
O
O
O
15 = me A 7 J * ij i
(eras)
Oo" oO
ONES
Om ae)
10 Oe
QT CG)! 8@)
e oe
Ow'O.
oO
* Oreo
® oF
5 Olas OF A@s ines " er Ora: z ao
* * Epon tom Sad *
O * “eC, 7® O
* * * * %
*

O} @ *

eg re. DenyOw On 10 IZ, Se io. dO I7 ia, 1970


Figure 2.20
The known pairs (n, py). Complete for n <7 for simple arrangements, and for
n <6 for non-simple arrangements.
O indicates a simple arrangement
* indicates a non-simple arrangement.
32 BRANKO GRUNBAUM

Concerning the possible pairs (”(A),p,(A)) for k 25 there is no general information


available; we venture

Conjecture 2.13. For each k 25 the maximum of p,(A) for


and for each n >k,
all A satisfying n(A)= Nn is attained for some simple arrangement, and only simple arrange-
ments attain that maximum.

Another aspect of the relations among the p,’s and other quantities is treated in the
circle of ideas known as “Eberhard-type” theorems. One of the consequences of the equa-
tions on page 16 and Euler’s relation is

23 C= 0 eee
j>2 k>3

from which it follows that

p3 24+ Dk -4)py,
k2>5

with equality characterizing simple arrangements (compare Moser [1963]). Conversely, we


have the following result (Eberhard [1891, p. 221], Griinbaum [1967, p. 405] ):
THEOREM 2.23. For each sequence p3, Ds, De, ***» P m Of non-negative integers
such that

(**) p3=4 a Se (kK — 4)p,


k2>5

there exists a simple arrangement A (with n(A)=p3) such that


Dp; (A) = Px for k= om 5.0; ee ’ m.

The arrangement A constructed in the proof of this result in Griinbaum [1967, pp.
405—406] satisfies
P3
p4(A) = Le ee
2 k#4

In general there exist also arrangements having the same values of p,(A) for k #4 but
differing in p4(A) and n(A), However, it is not known what values of p,(A) are possible
for a given sequence p,,p.<,°** that satisfies (**). It is even not known whether there
are any cases in which p,(A)=0 beyond the four arrangements (with 3, 6, 10, and 16
lines) mentioned on pages 29 and 31.
Many Eberhard-type theorems are known for convex polyhedra, and for cell-decompo-
sitions of orientable 2-manifolds (see Griinbaum [1970a]. for details and for a list of refer-
ences; for additional results see Jendrol-Jucovié [1971a], [1971b], Jucovicé-Trenkler [1971]).
However, in the case of the projective plane nothing is known beyond Theorem 2.23, and
many questions remain to be explored both for arrangements, and for cell-decompositions
of the projective plane. Some readily constructed examples make it seem likely that the
situation is interesting and non-trivial in both variants.
* * *
ARRANGEMENTS AND SPREADS 33

In order to formulate the next result on relations between the p,;’s and n it is con-
venient to denote by p(F) the number of edges (or vertices) of the cell F of an arrange-
ment. Levi [1926] observed that p(F) <n(A)
for every arrangement A and also that
P,(A) <1. This was extended by N. Gunderson (see Carver [1941]) who proved that
P(F,) + p(F,) <n(A) + 4 whenever F, and F, are different cells of the arrangement A.
Extending these results and correcting Theorem 18.2.9 of Griinbaum [1967], Canham [1969]
proved:

THEOREM 2.24. If F,,°*+, F,, are different faces of an arrangement then


k
Do PF) < n(A) + 2k(k - 1).
j=1
This result is best possible whenever n > 2k(kK — 1); however, for small values of n
the problem of determining
k
a(k, n) = max (xP(F;) - na)
AMIN ted
is still open. Theorem 2.24 means that a(k, n)= 2k(k-—1) for n > 2k(kK — 1). It is not
hard to check that a(3, 6) = a(3, 7)=9 and a(3, 8)= 10; no other values are known.
2.5. Irrational arrangements.
In contrast to the great difficulties encountered in determining the numbers of non-
isomorphic types of arrangements already for: relatively small numbers of lines, it is worth
pointing out that the numbers c(n), c5(n) and c4(n) are “effectively computable” func-
tions of n; in other words, it is possible to devise an algorithm for computing, one after the
other, the values c(3), c(4), c(5), «++ etc. A crucially important subalgorithm of this algo-
rithm is based on Tarski’s decidability theorem for elementary algebra (Tarski [1951],
Seidenberg [1954], Robinson [1963, Theorem 4.2.28], Cohen [1967] ). (For a similar ap-
plication of the same technique to the number of isomorphism types of convex polytopes
see Griinbaum [1967, Section 5.5] .)
The significance of this remark is best understood by contrasting it with the following
consideration. Instead of the real projective plane P with which we are concerned, one may
investigate projective planes over other ordered fields. In particular, the projective plane
P(Q) over the field Q of rational numbers may be studied. In P(Q) arrangements of lines,
and isomorphism types of such arrangements, may be defined in complete analogy with the
definitions given above for P. The cell complex associated with an arrangement may be
properly defined, since its properties are based on the Jordan curve theorem for simple poly-
gonal curves which is valid in P(Q) and in the more general planes considered below; see,
for example, Hahn [1908], Lennes [1911]. Since P(Q) can obviously be interpreted as a
subplane of P, each “rational arrangement” (that is, arrangement in P(Q)) is isomorphic to
an arrangement in P. However, we have the following result (see Griinbaum [1967, Section
55).
THEOREM 2.25. There exist “essentially irrational” arrangements in P, that is arrange-
ments not isomorphic to any arrangement in P(Q).
34 BRANKO GRUNBAUM

An example of an essentially irrational arrangement is the arrangement C shown in


Figure 2.21 (a) and derived in an obvious fashion from the regular pentagon with vertices
AKBDC. C contains 9 lines and is the smallest essentially irrational arrangement. (A proof
of the fact that C is the smallest such arrangement was never published but was convinc-
ingly presented to me by J. H. Conway.) The proof of the essential irrationality of C fol-
lows immediately from the fact (which in turn, is a consequence of Theorem 2.28 below)
that in each arrangement isomorphic to C the points corresponding to A, B, E, F deter-
mine an irrational cross-ratio.
It follows from these remarks that the number of isomorphism types of rational
arrangements of 7 lines is less than c(n) for n 29. However, nothing more is known
about this number, and we even have

Conjecture 2.14. The number of different isomorphism types of rational arrangements


of n lines is not an effectively computable function of nN.

Support for this conjecture may be found in the recent negative solution (Matiyasevic
[1971]) of Hilbert’s [1900] problem about the existence of an algorithm for deciding the
solvability of Diophantine equations.
Also, the solutions of various existence or extremal problems concerning arrangements
in the rational projective plane will, in general, differ from the solutions of the analogous
problems in the real projective plane. For example, we have:

Conjecture 2.15. Except for the near-pencils, there exist only finitely many non-
isomorphic types of simplicial arrangements in the rational projective plane.
* * *

It is of some interest to note that there exist no arrangements “essentially unreal”


in the following sense:

Arra ngement iG
Arrangement C
“ (b)
Figure 2.21
ARRANGEMENTS AND SPREADS 35

THEOREM 2.26. Every arrangement of lines in a projective plane over some ordered
field is isomorphic to an arrangement in P.

A proof of this result may be given by a slight modification of the proof given by
Lindstrom [1971] to the analogous theorem about convex polytopes. As a matter of fact, a
stronger result is obtainable by the same method, since Theorem 2.26 remains valid if the
real projective plane P is replaced by the projective plane P(A) over the field A of real
algebraic numbers. This extension is, in a sense, best possible as shown by the following re-
sult due to H. A. Heilbronn (see Griinbaum [1967, p. 94]):
THEOREM 2.27. If F is an ordered field such that every arrangement in the real pro-
jective plane P is isomorphic to an arrangement in the projective plane P(F) over F, then
F contains a subfield isomorphic to the field A of the real algebraic numbers.

The proof of Theorem 2.27 consists in devising, for each non-rational algebraic number,
an arrangement which is a “projective construction” of this number (or of one of its real
conjugates). Such a “projective construction” for the number V3 is shown in Figure 2.22,
in which the triplets of numbers indicate the homogeneous coordinates of the vertices used
in the “construction’’.

100

010 on 0,3-J3.3 001


0,1,-1

Figure 2.22
A “projective construction” of /3
36 BRANKO GRUNBAUM

2.6. Arrangements associated with sets of points.


Let V={V,,°*+, V,,} bea set of m= m(V) > 3 points in the projective plane, not
all on one line. Then V determines the associated arrangement A(V) formed by all the
lines connecting pairs of points in V. It is easily checked that if m= 3 then the isomor-
phism type of A(V) is uniquely determined, but that for m= 4 two types of A(V) are
possible while for m= 5 the number of possible isomorphism types of A(V) is 4 (see
Figure 2.23). These numbers coincide with the numbers c(m), but for m= m(V)2> 6 it is

Figure 2.23
ARRANGEMENTS AND SPREADS Bi]

easily seen that the number of non-isomorphic types of A(V) is larger than c(m). Jordan
[1920] investigated the analogous problems in the Euclidean plane and 3-space. The num-
ber of possible types of A(V) obviously increases rapidly with m, but no results on its
rate of growth are known.
Due to the duality between points and lines in the projective plane, many problems
and results concerning arrangements of lines associated with sets V of m= m(V) points
are only reformulations of the corresponding questions and results about arrangements on
m=n(A) lines. For example, the possible pairs (n(A), fo(A)) obviously coincide with the
possible pairs (m(V), n(A(V))). However, in questions involving the cell complexes asso-
ciated with arrangements the two points of view lead frequently to completely different
problems. For example, almost nothing is known concerning the possible pairs #5
(m(V), fo(A(Y))). One of the fundamental constructions (see Mébius [1827, p. 205], Zindler
[1889], Levi [1929, p. 131]) of projective geometry may be mentioned here since it is re-
lated to these questions: Taking the vertices of A(V) as a new set V’, considering the ver-
tices V" of A(V’), and so on, the union of the sets V,7 = 1, 2, °**, is dense in the pro-
jective plane provided V is not the set of vertices of a near-pencil. Also fitting this frame-
work is the concept of “residence” introduced in Kelly-Moser [1958]; for each v,; € V its
residence is the convex polygon formed by all the cells of the cell complex associated with
A(V) that contain v;. This notion is very useful in the investigations of variants of Sylves-
ter’s problem, and will probably have additional applications. For related notions in higher
dimensions see Bonnice-Kelly [1971] and Rottenberg [1971].
All near-pencils are clearly obtainable in the form A(V) for suitable sets V. However,
the only other known simplicial arrangements obtainable in this form correspond to m(V) =
4 or 5 (see Figure 2.23), and 6, 7, or 9 (see Figure 2.24). We venture the following 6
Conjecture 2.16. Jf m(V) > 10 the only simplicial arrangements of the form A(V)
are the near-pencils.

2.7. Other problems and classifications .


Various properties of arrangements dealing with coloring, orientation and other aspects
have been considered in the literature (see Melchior [1940], Motzkin [1967a], Chakerian
[1970], Murty [1971], Motzkin-Rabin [1971]) and those investigations are only at their be-
ginning. However, in order not to lengthen this chapter excessively we shall refrain from a
discussion of the facts known about those properties, and conclude with a few remarks on
classifications of arrangements by criteria different from isomorphism.
A classification finer than isomorphism is provided by the “projective isomorphism”’;
an arrangement A is of a projectively unique type provided each arrangement isomorphic to
A is projectively isomorphic to it (that is, may be obtained as a projective image of A). 07
This notion is analogous to the concept of projectively unique types of convex polytopes
introduced in Griinbaum [1967, p. 68]; for related material see Griinbaum [1970a, Section
3.3], McMullen [1969], [1971]. It is easily checked that the near-pencil N(”) is of a pro-
jectively unique type if and only if n=3 or 4. However, we have (see Griinbaum [1971])
38 BRANKO GRUNBAUM

Conjecture 2.17. Except for the near-pencils N(n) with n 2 S, all simplicial arrange-
ments are of projectively unique types.

For investigations concerning projective classifications of sets of points (or, dually, of


sets of lines) see Coble [1915], Levi [1922], Musselman. [1931], [1936]. Coble’s paper is
of interest also since it implicitly develops a remarkable correspondence between arrange-
ments in spaces of various dimensions; this correspondence has been explicitly discovered and
investigated only much later by Motzkin [1957], [1967c] and especially by McMullen [1971].)
* * *

In certain instances it is of some interest to consider a partition of arrangements into


equivalence classes coarser than the partition into isomorphism types. We shall say that the
arrangements A and A’ are equiconfigurational (under a mapping ©) provided there ex-
ists a one-to-one correspondence © between the lines L,; and the vertices V, of A and
the lines and vertices of A’ such that O(L,) is incident with O(V;) if and only if L; is
incident with V;. Isomorphic arrangements are obviously equiconfigurational, and any two
simple arrangements with the same number of lines are equiconfigurational. However, it is
well possible for two arrangements to be isomorphic, and at the same time to be equicon-
figurational under a mapping © without being isomorphic under the same mapping @.

A BS AM

Figure 2.24
ARRANGEMENTS AND SPREADS 39

Easy examples for that possibility are provided by near-pencils with at least 5 lines; a
more interesting example is given by the arrangements C and (C’ in Figure 2.21, in which
vertices corresponding to each other under © are denoted by the same letter. Moreover,
those arrangements C and (’ have the following remarkable property:

THEOREM 2.28. If A is an arrangement equiconfigurational with C under a mapping


© then A is projectively equivalent to either C or to C’ under a mapping which extends @.
PROOF. Let us assume that an arrangement equiconfigurational with C is given in a
system of homogeneous coordinates in which K= (1, 0, 0), C= (0, 1, 0), D= (0, 0, 1), 7=(1, 1, 1),
hence £= (1, 1,0) and F=(1,0, 1). Assuming that J=(1, a, 8) it follows that A =(l,a, 1-9),
B=(1,1 mp; 6). For G, interpreted as intersection of the lines AD and CK we have
G = (1, a, 0), while the interpretation as intersection of BJ with CK yields G=
(1 — 6, 1 — 28, 0). Hence a= (1 — 28)/(1 — 8), and similar reasoning for H shows that
6B = (1 — 2a)/(1 — a). Therefore a=, hence a? — 3a +1=0 and thus either a= B=
(3 + ERY, or else a=B=(3 - /5 )/2. It is easy to verify that in the first case A is
projectively equivalent to C, in the second to (’. This completes the proof of Theorem
225,
Equiconfigurational properties and related aspects of the “Desargues’ configuration”
have been studied in detail (Levi [1929, p. 171], Maier [1939], Lauffer [1953a], [1953b] )
as have been those of the ‘‘Pappus configuration” (Levi [1929, p. 118], Kommerell [1941])
and some others (see, for example, Zacharias [1941], [1950], Novak [1959]).
* * *

The definition of equiconfigurational arrangements involves only the incidences of


points and lines, and it refers neither to the cell complexes associated with the arrangements,
nor to the order or separation properties of points on the lines. If applied to the various
combinatorial incidence systems (such as finite projective planes and similar structures; see,
for example, Ryser [1963], Hall [1967], Dembowski [1968]) it becomes the notion of iso-
morphism for those structures. Similarly, it is possible to consider in such more general
settings the concept of “‘configurations of lines”. Although “ ... there was a time when the
study of configurations was considered the most important branch of all geometry ...”
(Hilbert-Cohn-Vossen [1932, p. 85; p. 95 of the English translation] ), the interest in their
geometric aspects has waned in comparison to the purely combinatorial investigations (de-
signs of various types). The many papers dealing with configurations have never been sur-
veyed from the point of view of realizability in the real projective plane. Such an exposi-
tion would be very desirable since the field is inherently very interesting, but limitations of
space and time prevent us from attempting it here.
40 BRANKO GRUNBAUM

Chapter 3. Arrangements of pseudolines and arrangements of curves


3.1. Pseudolines and non-stretchable arrangements.
A careful consideration of the proofs of most of the results presented in the preceding
chapter reveals that the rectilinear nature of the lines which form the arrangements plays
only a very limited role. This leads naturally to the idea of investigating arrangements of
more general types; while different generalizations are possible (some will be considered
later), a very convenient setting is provided by the “arrangements of pseudolines” (or
“topological lines”) introduced by Levi [1926].
An arrangement of pseudolines P in the real projective plane P is any family of sim-
ple closed curves {P,,°°*,P,,} in P such that every two curves have precisely one point
in common, at which they cross each other. If all curves P; have a point in common we
call P trivial; except when the opposite is explicitly stated, in the sequel we shall assume
that nm >3 and that P is non-trivial.
It is clear that each arrangement of lines is also an arrangement of pseudolines. Since
it is well known (see, for example, Stenfors [1923], Kivikoski [1925]) that each simple
closed curve in P is the image of either a circle, or else of a straight line, under a self-
homeomorphism of P, it follows from the intersection condition imposed on pseudolines
that each P, is the image of a straight line under a self-homeomorphism of P. However,
the self-homeomorphism may depend on the P,; in question, and we cannot assert that the
arrangement P is the image of some arrangement of straight lines in P under a suitable
homeomorphism; as a matter of fact, we shall soon see examples in which no such homeo-
morphism exists.
Nevertheless, it follows from the Jordan curve theorem that each arrangement of
pseudolines P determines in P an associated cell complex (C(P). In analogy to the case of
arrangements of lines, the elements of C(P) of dimensions 0, 1, 2 shall be referred to as
the vertices, edges, and cells (or polygons) of C(P) and of P. We shall say that two arrange-
ments of pseudolines are isomorphic provided the associated cell complexes are isomorphic.
Notions of isomorphism types, simple or simplicial arrangements, and the numbers is ti,
r;, p;, etc. may be introduced in the same way for arrangements of pseudolines as we did
earlier for arrangements of lines.
We shall say that an arrangement of pseudolines P is stretchable provided P is iso-
morphic to an arrangement of (straight) lines. If all arrangements of pseudolines were
stretchable, there would be little point in studying them. However, as remarked already by
Levi [1926], we have:
THEOREM 3.1. There exist non-stretchable arrangements of pseudolines.

PRooFr. Following Levi [1926], we start by recalling the Pappus-Pascal theorem


(see Figure 3.1): If the points A,,A,, Aj are collinear, and the points B,, B,, B, are
collinear, and if C; is the intersection-point of the lines A,B, and A,B; for i#j #
k#i, then C,,C,, C3 are collinear. Now in order to obtain a non-stretchable arrange-
ment of pseudolines it is enough to take the arrangement of lines of the Pappus-Pascal
ARRANGEMENTS AND SPREADS 41

theorem and replace the line through C,, C, and C, by a pseudoline differing from it by
missing C3 (but passing through C, and C,, see Figure 312):
Similar constructions could be performed by suitable alterations of other incidence
theorems of projective geometry (such as the Desargues’ theorem, etc.).
The proof of Theorem 3.1 relies on certain multiple incidences among straight lines,
dictated by general theorems of projective geometry. One could believe that if attention is
restricted to simple arrangements, the phenomenon of non-stretchability may be eliminated.

Figure 3.1

Figure 3.2
42 BRANKO GRUNBAUM

(It seems that this attitude was implicitly taken by R. Klee [1938] in his attempts at enu-
merating the different isomorphism types of simple arrangements in the Euclidean plane and
in the projective plane.) However, it was shown by Ringel [1956] that regarding stretchabil-
ity the simple arrangements of pseudolines do not behave any better than arrangements in
general; Ringel established:
THEOREM 3.2. There exists a non-stretchable simple arrangement of 9 pseudolines.

Proor. Consider the arrangement P of 9 pseudolines shown in Figure 3.3; it is


patterned after arrangements in Ringel [1956] and Griinbaum [1970b]. Assume that we
found an arrangement A of straight lines isomorphic with P; we shall denote the vertices
of A by the same symbols as those of P, indicated in Figure 3.3. We start with the ver-
tices A,, B3, A,, B,, Az, B,; let C, be the intersection point of the lines A,B, and
A,B, of A (or the corresponding pseudolines of P), and similarly let C, be on the
straight lines A,B, and A,B, in the plane of A, while C, is on the lines A,B, and
A3B,. Since the six points A,, B, belong to two lines, it follows from the Pappus-Pascal
theorem that the intersection points C,, C,, C, of the “opposite sides” of the hexagon

Figure 3.3
ARRANGEMENTS AND SPREADS 43

A,B3A,B,A3B, must be collinear. On the other hand it is easily checked that the point
C3; belongs to the face of A that corresponds to the shaded quadrangle designated C, in
Figure 3.4, while C, is in the one corresponding to the quadrangle denoted C,. But then,
considering the heavily drawn lines in Figure 3.4, it follows that the points C,, C,, and C,
cannot be collinear. The contradiction reached completes the proof of Theorem 3.2.
It may be noted that the arrangements of Figures 3.2 and 3.3 contain 9 pseudolines
each; thus we are led to

Conjecture 3.1. Every arrangement of at most 8 pseudolines is stretchable.

The same examples show that the enumeration of the different isomorphism types of
arrangements (and of simple arrangements) of pseudolines differs from the corresponding
enumeration of arrangements of lines at least from n= 9 on. Independent enumerations by
Canham [1971] and Halsey [1971] show that all simple arrangements of 7 or fewer pseudo-
lines are stretchable; they also established that there exist 135 different isomorphism types
of simple arrangements of 8 pseudolines. It seems that non-stretchable simplicial arrange-
ments require larger values of n; the smallest known non-stretchable simplicial arrangement

Figure 3.4
BRANKO GRUNBAUM

Figure 3.5
A non-stretchable simplicial arrangement of 15 pseudolines
(line at infinity included).

Figure 3.6
A non-stretchable simplicial arrangement of 16 pseudolines
(line at infinity included).
ARRANGEMENTS AND SPREADS 45

has 15 pseudolines (Figure 3.5). Three non-isomorphic such arrangements are known for
n= 16; two are shown in Figures 3.6 and 3.7, while the third may be found in Gruinbaum
[1971].
Despite Theorems 3.1 and 3.2, there are various ways of representing arrangements
of pseudolines in specially convenient ways. Intuitively rather appealing is the following
result, which is easily proved by induction on the number of pseudolines:

THEOREM 3.3. Every arrangement of pseudolines is isomorphic to an arrangement


having polygonal pseudolines such that each cell is a convex polygon, the edges and vertices
of the arrangement being also edges and vertices of these polygons.

Ringel [1956] has given an abstract characterization of cell complexes associated with
simple arrangements of pseudolines. No such characterization is known in the general case.
* * *

The existence of non-stretchable simple arrangements: of pseudolines implies, among


others, a negative answer to the following problem of Motzkin [1967b]: Is every 4-valent
centrally symmetric 3-polytope with simple geodesics isomorphic to a centrally symmetric
polytope in which all the edges of each geodesic circuit are coplanar? For related material
see Zaks [1971] and Grinbaum [1971].
A remarkable non-stretchable arrangement of 10 pseudolines is shown in Figure 3.8.
Its rectilinear appearance is misleading, and it did mislead Kantor [1881] in his enumeration
of the configurations 103. (Through each of the 10 vertices marked in Figure 3.8 pass 3
of the pseudolines, and each of the 10 pseudolines contains 3 of the marked vertices.) Kan-
tor’s error was detected by Schréter [1889] who proved the non-stretchability of the arrange-
ment. $8
It would be rather interesting to characterize stretchable arrangements, or at least to
find non-trivial sufficient conditions for stretchability. However, even the easier of these
two problems seems to be very difficult.

3.2. Some results on arrangements of pseudolines.


The existence of non-stretchable arrangements implies that enumeration, existence, ex-
tremal and other problems considered in Chapter 2 may be reformulated, and need to be re-
considered, for the larger family of arrangements of pseudolines. As it happens, many of the
results discussed in Chapter 2 remain valid; without any change in the proofs we could have
established for arrangements of pseudolines the Theorems 2.4 to 2.14, 2.16 to 2.19, and
2.21 to 2.24. Indeed, the proofs of these results are either constructions (hence a fortiori
valid for pseudolines), or else they are based on incidence properties which are equally valid
for arrangements of lines and of pseudolines, and do not rely on rectilinearity.
However, there are non-trivial results concerning pseudolines, the analogues of which
for straight lines are either completely trivial, or else much easier to prove. We shall discuss
a few such results next.
We start by establishing an “enlargement lemma” due to Levi [1926]; it is completely
46 BRANKO GRUNBAUM

Figure 3.7
Another type of non-stretchable simplicial arrangement with 16 pseudolines
(line at infinity included).

Figure 3.8
A non-stretchable configuration 103.
ARRANGEMENTS AND SPREADS 47

trivial for arrangements of lines. The following proof is an adaptation of the one given by
Levi [1926].
THEOREM 3.4. Given an arrangement P of pseudolines P,, P,, °**, P n?
and points
V,, V. which do not both lie on any one of the pseudolines P,,+** , P,,, then there ex-
ists an arrangement of pseudolines Py, P,,°**,P, such that V, and V, are points of Po.

Before proving Theorem 3.4 we note the following lemma which, despite its ponderous
formulation, describes a simple geometric fact:
(*) Let P={P,,°*>,P,} be an arrangement of pseudolines, let H be (a subset of
the projective plane which is homeomorphic to) one of the closed halfplanes determined by
the pseudolines P, and P,, and let W,,W., © H. For each simple open arc CC H with
endpoints W, and W, let w(C) be the number (possibly infinite) of incidences of C
with the pseudolines P,,°**,P,, and let Cy be anarc of this type for which w(Cy)
attains the minimal possible value. Then Cy intersects (and crosses) just once those
pseudolines among P,, °°, P,, the traces of which on H separate (in H) W, from W,,
and Cy does not intersect the remaining pseudolines.

Indeed, from the assumed minimality of w(C,) it follows that Cg intersects each
pseudoline at most once: if 7,7, is a minimal arc of Cy such that 7, and 7, belong
to the same P,, the replacement of that arc by a suitable arc not crossing P; would reduce
w(C,) by 2 (see Figure 3.9 where the arc 7,7, and the replacement arc are dashed). On
the other hand, by the Jordan curve theorem C, must cross an odd number of times each
pseudoline which separates in H between W, and W,, and cross an even number of
times each non-separating pseudoline.

Figure 3.9
48 BRANKO GRUNBAUM

The Lemma (*) thus established, we turn to the proof of Theorem 3.4.
Consider first the case in which V,, V, belong to some of the pseudolines, say V, ©

P, and V,€P,. Then consider the two closed halfplanes H* and H** determined by
P, and P,, and in each of them apply to the points V, and V, the Lemma (*). Let
CS C H* and Ge C H** be the minimal arcs with endpoints V,, V, given by the lemma.
Since a pseudoline P, that separates V, and V, in H* (or H**) does not separate the
same points in H ** (respectively H*) and since each pseudoline P either contains one of
the points V, or V, or else separates them in one of the halfplanes H*, H**, it follows
that the simple closed curve {V,} U Bes U {V,} U Co* may be taken as the pseudoline
P,, completing the proof in this case.
If V, ۩P, but V, is on no pseudoline, let W, be a point on the boundary of the
cell which contains V,, chosen so that W, belongs to a pseudoline P, different from P,. Holt
is that closed halfspace determined by P, and P, which contains V,, and if H*™* is the
other closed halfspace, we apply Lemma (*) to the points V, and V, in H* obtaining
an arc G , and to the points V, and W, in H** obtaining an arc 6s * Denoting by
(V,, W,) an open arc with endpoints V, and W, belonging to the interior of the cell
that contains them, it follows that we can take for Py) the simple closed curve {V,} U
CHUL, LU VE, WUT CS”:
In order to visualize the above two cases (as well as in the following arguments) it is
convenient to use the model of the projective plane in which one of the pseudolines (say P,)
is taken as the boundary of the circular disc (hence each point of P, is represented by a
pair of diametral points). The two cases discussed above are illustrated in Figure 3.10.
If neither V, nor V, belong to any of the pseudolines, we choose points W, and
W, on different pseudolines P, and P, so that W, is on the boundary of the cell that
contains V, and W, on the boundary of the cell that contains V,. We then apply Lemma
(*) to the halfplanes H* and H** determined by P, and P,, and obtain the pseudoline
Py) by combining the arcs Gs and (oe with (V,, W,) and (V,, W,) as shown in Figure
3.11 for the two possible cases.
This completes the proof of Levi’s Theorem 3.4.
Theorem 3.4 is a useful tool in many questions about arrangements of pseudolines.
Recently it has been used by Kelly-Rottenberg [1971] in the proof of the following general-
ization of Theorem 2.15:

THEOREM 3.5. Every non-trivial arrangement P of pseudolines satisfies t,(P) =


3n(P)/7.
This result provides an affirmative solution to one of the problems raised in Griinbaum
[1970b]; other problems concerning pseudolines formulated in that paper are still open.
Earlier, Theorem 3.4 found application in Levi’s proof of the following result, which
generalizes Theorem 2.19:

THEOREM 3.6. Every non-trivial arrangement P of pseudolines satisfies p3(P) = n(P).


ARRANGEMENTS AND SPREADS 49

Figure 3.10

Figure 3.11

Figure 3.12 Figure 3.13


50 BRANKO GRUNBAUM

follow-
As in the case of arrangements of lines, this is an immediate consequence of the
ing result:

THEOREM 3.7. Let P be a non-trivial arrangement of pseudolines, and let P, be one


of the pseudolines. Then there exist at least 3 triangles in P each of which has one of its
edges in P,.

Our proof of Theorem 3.7 will be based on a lemma ((#*) below), which goes back
essentially to Steinitz [1922] (see also Steinitz-Rademacher [1934], and in particular the
discussion of “lenses” in Griinbaum [1967, pp. 239—241]). We begin by establishing the
lemma:

(«*) Let H be a closed halfplane determined by pseudolines P, and P, of the


arrangement P of pseudolines. Then H contains either a triangle T which has an edge on
P, but misses P,, or else H contains two triangles T, and T, both having edges on P,
and containing the intersection-point V of P, and P,.
Proor. We first observe that there is no generality lost by assuming that no pseudo-
line different from P, and P, passes through V and is contained in H; indeed, any
such pseudoline would determine, together with P,, a halfplane contained in H, and the
triangle or triangles contained in it would satisfy the lemma for the original pair of pseudo-
lines.
Next, we note that the trace Pe of each pseudoline P,,k 23, on H isa simple arc
which reaches from P, to P, but, by the above, misses V. We distinguish two cases:
(i) No two arcs Py intersect in the interior of H. Then the two faces of the
arrangement that are contained in H and contain V may obviously be taken as 7, and
T,. (Compare Figure 3.12.)
(ii) Some arcs les intersect in the interior of H; in other words, the interior of H
contains some vertices of the arrangement P. For each such vertex V, and for each arc
JP that passes through V; let P,(j) be that subare of Pe which has V; as one endpoint
and a point W, of P, as the other. We shall say that a vertex V; is a neighbor oP;
provided there exists a k such that P,(/) consists of the single edge (V ppl’) Cob. BP. Bor
a vertex V; that is a neighbor of P,, with (V;, W,,) an edge of P, let P,, be another
pseudoline through V; and let z(V;, k, m) be the number of faces of P contained in the
“patch” of H bounded by P,(j), P,,(/), and an arc of P,. We claim that if V;, P, and
P, are chosen so as to minimize 2(V;, k, m) then z(V;, k, m)= 1, and the single face of
P contained in the “patch” is a triangle 7 which has an edge on P, and misses P,. In-
deed, otherwise (see Figure 3.13) the arc P,,@) would contain vertices of P different from
V; and W,,. If V, is the nearest to W,, among those vertices, and if P, is a pseudoline
different from P,, passing through V,, then, by the Jordan curve theorem, the patch deter-
mined by P,,(r), P(r) and an are of P, satisfies z(V,, m, s) < z(V;, k, m), contradicting
the minimality assumption and thus completing the proof of Lemma (#).
Having established the lemma, we turn now to the proof of Theorem 3.7. Let Ti be
a triangle having an edge on P, (such a triangle exists by Lemma (**)), and let P, and
ARRANGEMENTS AND SPREADS 51

P, be the other two pseudolines containing the edges of T, (see Figure 3.14). Then, by
Lemma (**), the halfplane H;, determined by P, and P, and containing T,, contains
another triangle T, having an edge on P,, for j= 2,3. The triangles T,, T,, T, estab-
lish the assertion of Theorem 3.7.
We turn now to some instances in which arrangements of pseudolines behave differ-
ently from arrangements of lines.
In contrast to Conjecture 2.1, a number (seven at the present) of infinite families of
non-stretchable simplicial arrangements of pseudolines are known. They are derived in vari-
ous ways from regular polygons. For example, regular k-gons with k=O or 4 (mod 6)
lead to simplicial arrangements with 3k + 1 pseudolines and also, for k=O or 2 (mod 6)

Figure 3.14

to simplicial arrangements with 3k pseudolines, and to another family with 3k + 1 pseudo-


lines. The case k = 10 of the first type and the case k = 12 of the second type are shown
in Figures 3.15 and 3.16. Other infinite families of simplicial arrangements of pseudolines
are obtainable from regular polygons with an odd number of sides. A member (correspond-
ing to the undecagon) of one such family is shown in Figure 3.17. (In some of those fam-
ilies certain members with relatively few pseudolines are stretchable.)
Although there probably exist additional infinite families of simplicial arrangements
we believe that the following analogues of Conjectures 2.1 and 2.2 are valid.
Conjecture 3.2. There exists a constant c such that, for each n, the number of dif-
ferent isomorphism types of simplicial arrangements of n pseudolines is at most c.
Conjecture 3.3. Except for a finite number of types, each simplicial arrangement of
pseudolines has a non-trivial group of combinatorial symmetries.
Many instances are known (for n > 19) in which p,, the number of triangles, of a
simplicial arrangement of pseudolines exceeds the corresponding number for all known
simplicial arrangements of lines with the same 1. Moreover, two simplicial arrangements of
SP BRANKO GRUNBAUM

Figure 3.15
A non-stretchable simplicial arrangement of 31 pseudolines
(including line at infinity).
ARRANGEMENTS AND SPREADS

Figure 3.16
Non-stretchable simplicial arrangements of 36 and (with line at infinity)
37 pseudolines.
54 BRANKO GRUNBAUM

Wt
NS ZEIN
i
LVR
SS
W)

Figure 3.17
Non-stretchable simplicial arrangements of 33 and (with line at infinity)
34 pseudolines.
ARRANGEMENTS AND SPREADS 55

19 pseudolines each are known (see Figure 3.18) for which p3 = 108, thus contradicting the
extension of Conjecture 2.10 to arrangements of pseudolines.
Conjecture 2.7 cannot be extended to arrangements of pseudolines. As observed by
Canham [1971], the non-stretchable arrangement of 9 pseudolines obtained from the
arrangement in Figure 3.3 by “shrinking” the central triangle to a point contains 9 triangles
but is not simple. On the other hand, by conversely “splitting” suitable vertices of certain
simplicial arrangements of pseudolines it is possible to obtain arrangements of pseudolines
which have very large values of p,. The available data on the values of p, (compare also
Theorem 2.21 and Conjectures 2.9 and 2.10) are collected in Table 2.
3.3. Arrangements of simple curves in the Euclidean plane.
We turn now to a discussion of a number of variants and generalizations of the notion
of arrangements of pseudolines.
It is well known that the projective plane P may be represented by the set of all
pairs of antipodal points of a (unit) sphere S in Euclidean 3-space. In this way to each
arrangement of pseudolines in P there corresponds an arrangement of curves in S, each
curve being simple, closed and centrally symmetric, and every two curves having in common
precisely one pair of antipodal points. Considering a suitable stereographic projection of
such an arrangement into the Euclidean plane one is rather naturally led to consider the
“symmetric arrangements of curves” and more general “arrangements of curves”’.
An arrangement of curves C= {C,,°***,C,} in the Euclidean plane E? isa finite
family of simple closed curves C; with the property:
(i) Every two curves have precisely two points in common, at which they cross each
other.
An arrangement of curves C is trivial provided there exists a pair of points contained
in each curve of C. In the sequel we shall assume, unless the contrary is explicitly stated,
that all arrangements considered are non-trivial.

Figure 3.18
simplicial arrangements of 19 pseudolines with p3 = 108.
Non-stretchable
(Line at infinity included in both.)
56 BRANKO GRUNBAUM

Upper bound for p3 in Largest p3 known in arrangements of


simple arrangements
n lines pseudolines

3 4 4 4 4
4 4 4 6 6
5 5 > 8 8

6 10 10 12 \
7 11 11 16 16
8 16 16 20
) 21 2h 24
10 30 30 30

11 33 32 36
1 ct 40 40 42
13 47 52
14 60 56 58
15 65 65 66

16 80 80 74
17 85 84
18 102 90 92 96
19 107 102 108
20 126 120 120 110 120

21 133 126 130


22 154 138 140 146
23 161 148 150
24 184 176 158 162 174
25 191 180 186

26 216 208 190 200 208


a8) 234 200 204
28 fasye 210 224 240
29 261 224 238 254
30 290 280 240 250 276

31 299 252 280 290


S2 330 320 272 278 320
ie) 341 288 320 336
34 374 306 308 356
35 385 68 322

36 420 408 342 378 402


3) 431 360 396 420
38 468 456 380 450
39 481 76
40 520 420 424 500
41 533 440 456 520
42 574 560 462 462 550
43 587 84 476
Aq 630 616 506 Sd2 608
45 645 528 594 636
Table 2
ARRANGEMENTS AND SPREADS Si/

An arrangement of curves C is symmetric if it has the following property:


(ii) For every three curves C, C, and C” of C, the two points of CNC’ either
coincide with the two points of CMC”, or else separate them on C
Each arrangement of curves in the plane determines a cell complex decomposition of
the Euclidean plane; in general, digons (cells with only two edges and two vertices) may be
present. (The “outside” of the arrangement is considered as a cell.) In analogy to the situa-
tion concerning arrangements of lines or pseudolines it is therefore possible to define isomor-
phism of arrangements of curves, as well as notions such as fj tj, Pj, ete. It should be
noted that each arrangement of pseudolines P in the projective plane leads to an associated
symmetric arrangement of curves C in the Euclidean plane, such that FC) = 2f(P), t(C)=
2t;(P), ete.
On the other hand, for each symmetric arrangement of curves C condition (ii) im-
plies the existence of an involutory symmetry (without fixed elements) in the cell complex
determined by C in the plane. This may be used to show that each symmetric arrange-
ment of curves is isomorphic to an arrangement associated with an arrangement of pseudo-
lines in the projective plane. Though this shows that there is little independent interest in
symmetric arrangements of curves, there are some pedagogical benefits in the possibility of
substituting such arrangements in the Euclidean plane for arrangements of pseudolines in the
projective plane. In addition, this connection points to various directions in which meaning-
ful results about general arrangements of curves may be sought.
As we shall see presently, arrangements of curves differ in many properties from sym-
metric arrangements of curves. Many of those differences are related to the possibility of
digons being present; hence digon-free arrangements of curves form an interesting inter-
mediate class of arrangements.
Let an arrangement of curves C be called near-trivial provided there exists a point
common to all the curves in C. By a simple reduction it may be seen that Theorem 3.7 implies
THEOREM 3.8. Each near-trivial arrangement of curves C has at least 3 digons; that
is, op, (C)'= 3.
° This result is best possible in the sense that equality may hold for arbitrarily large
n(C) (compare Figure 3.19).

Figure 3.19
A typical representative of near-trivial arrangements of curves with three digons.
20
40
60
<@o0

Pairs

®
oa @0e@0®@Ood000

o ®00@080@04000d00

A simplicial
symmetric
|
D0000¢00 @00000000000408
BRANKO

Figure

A symmetric simpliciai
«oo e0e0cecececececececdo808nn0n000400

(n i fy) corresponding to
|

3.20
0e080000n00004a0 0©0000000000000000000000080<404
GRUNBAUM

400

O any
cn0000000
000008
0

10

arrangements
©200000000000000000000

© digon-free
oo0o000000

11
c000000
0% e0ec00
0% 0000000
00 0000000000

@o00

of curves,
@oo

12
aneceac00000G000
4

00000800
0 00000
04000808 ©000
(

400 0000000

13
ececann00 40@0000000
040008090
@0$ 00000

400

14
00000000
80800
000e0@00
®0®@0
00 0000
000000000000
|

15
0o0e0@DOo00000
00000000000000
5000@00
0 000
@o0o0
ARRANGEMENTS AND SPREADS ay)

In order to see an example for the changes in the results of Chapter 2 when the prob-
lems are considered in the more general setting of arrangements of curves, we shall first inves-
tigate the pairs (n(C), f,(C)) for arrangements of curves. We recall that Theorems 2.10 and
2.11, together with the remark on page 57 imply that for symmetric arrangements of curves
C either f,(C) = 4n(C)-—4 and C is associated with the near-pencil of n lines, or else
f,(C) 2 6n(C) — 12 (compare Figure 3.20). We have the following result, the proof of
which parallels that of Theorem 2.11.

THEOREM 3.9. Concerning the number f,(C) =f, of cells of an arrangement C of


n(C)=n curves we have either (i) f, = 2n, which happens if and only if C is a trivial
arrangement, or (ii) f, = 3n-—2, which happens if and only if either n= 4 and C is
the arrangement of Figure 3.21, or if C is the special near-trivial arrangement T(n) ob-
tained from a trivial arrangement of n—1 curves (see Figure 3.22); or (iii) f, 2 4n— 6.
Moreover, if f, =4n-6 then C is either derived from [(n— 1) in one of the
two ways indicated in Figure 3.23 (note that each of the two ways may lead to a number
of non-isomorphic types), or else n= 6 or 7 and C is one of the exceptional arrange-
ments shown in Figure 3.24.

Figure 3.21
An arrangement with (m, f,) = (4, 10)

Figure 3.22
A typical representative of near-trivial arrangements with f, = 3n — 2.
60 BRANKO GRUNBAUM

Through a straightforward but lengthy process it is similarly possible to characterize


all the arrangements of n curves with f, = 4n—5. Though the result is not very remark-
able in itself, it is of some interest in so far as all such arrangements are seen to contain
digons. Hence we have the following result, for which a direct proof would be desirable:
THEOREM 3.10. For each digon-free arrangement of curves we have

f, 24n—-4.
The information available concerning the pairs (, f,) is collected in Figure 3.20,
which is complete for n <10 (although possibly some © could be “upgraded” to ©). 09

Figure 3.20 suggests a number of conjectures, the simplest of which are:


Conjecture 3.4. No arrangement C of n(C)=n curves satisfies
Ate SGP (Cy< Ste 12
Conjecture 3.5. No digon-free arrangement C of n(C)=n curves satisfies
an —4-< f,(C)<on
—7
On Says) On — ee
Similar investigations may be made concerning the possible pairs (n(C), p,(C)). Here
the first possible value of k is k= 2, which is without analogue among arrangements of
lines.
Figure 3.25 shows the known pairs (n, p,). The values p, = 2n are obtained for
trivial arrangements of n curves, while p, = 2n ~ 2 holds, among others, for the (simple)
arrangements of the type indicated in Figure 3.26. It is also easily seen how to obtain sim-
ple arrangements with any value p, <2n—- 2. This leads to

Figure 3.23
Typical representatives of arrangements with 'D) = 4n — 6.
ARRANGEMENTS AND SPREADS

Figure 3.24
Some arrangements with Sf, = 4n—- 6.
62 BRANKO GRUNBAUM

10

Figure 3.25
The pairs (n, Py).

Figure 3.26
Typical representative of arrangements with Py = 2n—- a8
= = oe ie rie a
50-+- — Bas

A >
°
A
fe) 000f,000e0r0g0
>
@
°
ae

= = =a = eg ees 508
40 +
°
A a
° o oo
ae 0°@ Oo >
° Oo
Oo a
° Oo
& e oa
Oo oO
— yay — @ a o
oe — —
= 30
fe) Oo o
e A e
° o fa)
® a Oo
fe) Oo Oo
a oA a @
° a) Oo fs)
fe) a a a
oO o o oO
— a — B = o
== — == A
20
° is) ia) o
o @ a g Ss oooo.s

o Oo fa] ia)
A 0 ® rT] a a) (5)
(a)
femelle!
qepfal
fo)
tefishie)
feijeliel
(je)
fapfepqe
fa)

° o oO fa] o
° a r o Oo
fe) Oo Oo oO
s 0@ a fa] Oo
fe) o Oo
Oo — ra] = eae
— (e) —— 8 =
10
°
a a Oo o

5 6 id 8 9 10 11
ees 4

Figure 3.27
(n, P3) corresponding to digon-free arrangements of curves.
Pairs
A symmetric simplicial A simplicial
® symmetric simple O simple digon-free
® symmetric O digon-free.
64 BRANKO GRUNBAUM

Conjecture 3.6. A pair (n, p,) with n> 5S corresponds to an arrangement of curves
C if and only if either (i) p, = 2n, and C is trivial; or (ii) O<p, S 2n — 2.

The information available on the pairs (n, p,) is presented in Figure 3.27, which is
probably complete for n <10 (although possibly some of the symbols could be “upgraded”).
While Theorem 3.6 implies that p, > 2n for every symmetric arrangement of curves, the
examples of Figure 3.28 show that p, = 2n —4 is possible (for n > 6) for digon-free
arrangements. We have
Conjecture 3.7. For every digon-free arrangement of curves

ps = 2n— 4.

No details are known concerning the behaviour of p, in arrangements in which digons


are allowed. But even for digon-free, simple arrangements the situation is rather unclear. In
particular, while the method of proof of Theorem 2.21 implies (see page 26 for the definition
of ¢(n)) that p,(C) < 2¢(n) for each symmetric and simple arrangement C of n(C)=27
curves, the argumentation fails already for digon-free simple arrangements. As may be seen
from Figure 3.27 there indeed exist digon-free simple arrangements with p, > 2¢(n); exam-
ples of such arrangements for n= 7 and n= 8 are shown in Figure 3.29. No reasonable
guess has been made concerning the maximal possible value of p,(C) for all (or all simple
and/or digon-free) arrangements of n(C)=n curves.
Simplicial arrangements of curves exist in great profusion; already for m= 7 there
exist two non-symmetric types (see Figure 3.30). As shown in Figure 3.27, in many cases
the maximal number of triangles for arrangements with given n(C) seems to be attained
for (non-symmetric) simplicial arrangements (see Figures 3.31, 3.32 and 3.33 for arrangements

Figure 3.28
Heavy lines: A simple digon-free arrangement of 5 curves with P3 = 8. All lines:
A typical representative of digon-free arrangements of n curves with p3 = 2n—4.
ARRANGEMENTS AND SPREADS 65

Figure 3.29
Solid lines: A simple digon-free arrangement of 7 curves with P3 = 24 > 22 = 29(7).
All lines: A simple digon-free arrangement of 8 curves with P3 = 34 > 32 = 2¢(8).

‘@)@ Two types of non-symmetric


Figure 3.30
simplicial arrangements of 7 curves.
LN
AaaES
SESS
C e:
peat

on-symmetric simplicial arrangement of curves. Non-symmetric simplicial arrangement


of curves.

Ce
Lee

A non-symmetric simplicial arrangement of 12 curves with


ARRANGEMENTS AND SPREADS 67

with n= 9, 10, and 12 yielding the maximal known p,). Nothing seems to be known re-
garding the behaviour for large n; indeed, it seems even that the analogues of Conjectures
3.2 and 3.3 are not valid for simplicial arrangements of curves.
Similarly, no information seems to be available concerning the pairs (n(C), p,(C)) for
k 24 and C ranging over various types (or all) arrangements of curves.
Denoting (as on page 33) by p(F) the number of edges of the cell F of an arrange-
ment of curves, it is not hard to verify (W. Meyer, private communication):
THEOREM 3.11. Jf F is a cell of an arrangement of curves C then p(F) < 2n(0)- 2;
moreover, if C is digon-free then p(F) < 2n(C) - 4.
As shown by the examples in Figure 3.34, these estimates are best possible.
Extensions of Theorem 3.11 analogous to Canham’s Theorem 2.24 are still to be inves-
tigated, and so are various generalizations to arrangements of curves of Eberhard’s Theorem 2.23.
* * *

In analogy to the above discussion of the possible pairs (n, f,) and (n, p,), it is pos-
sible to investigate the pairs (”, fy) and (n, ¢,), thereby extending to arrangements of
curves the results of Chapter 2 concerning such pairs for arrangements of lines. The litera-
ture seems to contain no results on those questions, but some scattered observations may
be made that indicate that the differences in behaviour between arrangements of lines (or
pseudolines) and curves may be appreciable and interesting. For example t, = 6 and ft, =
6 are possible for digon-free arrangements with 8 and 11 curves, while it may be shown
(compare Table 1 on page 22) that for arrangements of curves associated with arrangements
of lines (or pseudolines) both these numbers are at most 4. If digons are permitted the dif-
ferences become even more pronounced; there exist arrangements of 6 and 7 curves which
have t, =3 or t; = 3. Unfortunately, no non-trivial general results seem to be known.
Possibly most challenging is the situation concerning the pairs (”, ¢,), in particular
for digon-free arrangements of curves. Denoting by ]x[ the smallest integer not less than
x, the result of Kelly-Rottenberg may be reformulated as

Figure 3.34
Typical representatives of arrangements of curves with maximal p (F).
68 BRANKO GRUNBAUM

THEOREM 3.12. For every symmetric arrangement of n curves


spe PAI UNE

The method of proof used in Kelly-Rottenberg [1971] does not work for (not neces-
sarily symmetric) digon-free arrangements of curves, and indeed the arrangement of 10
curves shown in Figure 3.32 has ¢, = 9 < 10 = 2)30/7[.
The information available on the pairs (n, t,) is presented in Figure 3.35, and we #10
venture

Conjecture 3.8. For all digon-free arrangements of curves C we have t,(C) 2


n(C) — 1; moreover, equality is possible only for n(C) = 1 (mod 3). #11

Another open and seemingly hard problem is to find the right analogue for (digon-
free) arrangements of curves of Levi’s extension lemma (Theorem 3.4). It is well possible
that an appropriate result in this direction would lead to solutions of some of the other
problems mentioned.

3.4. Generalizations.
A large number of concepts related to arrangements of curves may be devised and
studied, and it is only natural that some of them possess interesting features. We shall con-
clude this chapter by a brief discussion of several such possibilities, supplying in each case
all the bibliographic information available to us.
(I) A “weak arrangement of curves” is any family C={C,,°°*,C,} of n(C)= n
simple closed curves such that C;C; has at most two points whenever i #j, and if
C,; C; consists of two points the curves cross each other.
Thus “weak arrangements” differ from ‘“‘arrangements” in that some pairs of curves
may be disjoint, and some may have only an osculation vertex in common; we shall denote
the number of osculation vertices of C by w(C).
A weak arrangement of curves C is called Apollonian provided ail its vertices are
oscutation vertices; that is, w(C)=fo(C). The fact that a planar graph with n > 3 nodes
has at most 3n — 6 edges implies #12
THEOREM 3.13. Each Apollonian arrangement C of n>3 curves satisfies u(C)<
Si 6.

However, the determination of non-trivial upper bounds on w(C) in terms of n(C)


among all weak arrangements C seems to be rather hard; suitable families of examples
show that w(C)/n(C) > © for n—> © (Erdés—Griinbaum, in preparation).
Those arrangements of curves in which each curve is a circle have formed the subject
of a number of investigations, as have the analogously defined arrangements of curves,
circles, or great circles on the sphere. A very difficult question appears to be the determina-
tion of the (weak) arrangements of curves which are circularizable, that is isomorphic to
(weak) arrangements of circles. Another hard problem concerns the characterization of
arrangements isomorphic to arrangements of convex curves (boundaries of planar convex
sets). However, even the characterization of those arrangements of circles in the plane
ARRANGEMENTS AND SPREADS

oo

o °o

a 0@

e ie)

e
ie) ie)
@
(e)
50 ke e ° -

@
° °
° e
(eo)
e °
e) O°
@
1) ie)
- 0 - @ =
1)
@ O° a
° °
@ ® A A
e} ie} 4
[e) e a A
(e} (e}
e@ e@ 4 A a
° 10)
e@ e a A 4 a
(e} °
fo) ® a a a a
° O° 4
e@ ° a a a
° °
@ e A A A a
° O° 4
O° re) @ a A
° °
Ie. = "0 @ a a PA SA
ie) (eo) 4 aN
° e@ a a a a a a
° O° 4a 4
° a 4 4 a A
° 4 rN A 4
a A a A
° 4 4 4
A A A ® a A
a
a
4 4
a

n-=.3 4 5 6 7h 8 9 10 1 12 13 14 15 16 17 18

Figure 3.35
Pairs (n, ty) for digon-free arrangements of curves. Probably complete for n < 9.
A symmetric simplicial A simplicial
® symmetric O digon-free
70 BRANKO GRUNBAUM

to be
(or sphere) which are isomorphic to arrangements of great circles on a sphere seems
elusive. Possibly relevant in this context is the following observation of G. Ewald (private
communication):

THEOREM 3.14. An arrangement of circles C in the plane is the stereographic projec-


tion of an arrangement of great circles in a suitable sphere if and only if
(i) C is a symmetric arrangement of curves (i. e., satisfies conditions (i) and (ii) on
pages 55 and 57); and
(ii) For any four circles C,, Cy, C3, C4 of C the intersection points A, and A,
of C, and C,, and the intersection points A, and A, of C, and C4 are concyclic.

Questions about f,(C), t,(C) etc. for arrangements of circles were raised on many
occasions (Steiner [1826], Erdés [1957], [1961], Moser [1962], Croft-Guy [1971]); how-
ever, the results known are rather weak.
In the solution of a problem of Moser [1952], Bankoff [1970] asserted that if an
arrangement C of great circles on a sphere is not trivial then Lf = 2n. While the result
is clearly only a reformulation of Theorem 2.6, the proof in Bankoff [1970] is invalid.
To facilitate the formulation of some of the other results found in the literature it is
convenient to introduce the notions of the weak [generalized] arrangements of circles
associated with a set V of points (compare the similar notion considered in Section 2.6).
The weak arrangement C(V) is formed by all the circles determined by the points of V,
while the generalized arrangement G(V) is formed by all the circles and all the straight
lines that contain at least 3 points of V. Denoting by n(C(V)) or no(G(V)) the number
of circles in C(V) or G(V), and by n(G(V)) the total number of circles and lines in
G(V), we have the following results of Jucovic [1967] and Elliott [1967]:
THEOREM 3.15. Jf V consists of m(V)= m2 6 points then n LOS) | implies
n(G(V)) = [(Sm— 4)/3] while n(G(V)) > 1 implies n(G(V))= (Sm +5)/3]; fms
393 and n(G(V))>1 then n(G(V))> ("> ').
The last estimate is best possible for large m, as shown by sets V consisting of m—1
collinear points and a point not collinear with those. The requirement m > 393 is probably
too restrictive, but it was observed by B. Segre that for m= 8 the vertices V of a cube
(or a suitable projection of them into the plane) show the possibility of n(G(V)) = 20 <21=
(757 )e We'venture
Conjecture 3.9. For m(V)=m=>9 and n(G(V))>1 we have n(G(V))cd
> dese iy
for m 10 and n(G(V))
>1 we have even no(G(V))> (5 *).
The nine points of a 2-by-2 square in the integer lattice show that Ny(G(V))=
20.< 2) is possible for m(V)= 9.

THEOREM 3.16. Let V bea set of m(V)=m> 4 points in the plane such that
n(G(V)) > 1. Then each point of V belongs to a circle C of G(V) such that VOC
consists of precisely 3 points; the number r, of circles through precisely 3 points of V
satisfies
ARRANGEMENTS AND SPREADS 71

r,(C(V)) > max {4, 2n(n — 1)/63}.


The first part follows at once (by inversion) from Theorem 2.12. Various proofs for it
were given by Gupta [1953], Balasubramanian [1953], Smirnova [1956], Hadwiger-Debrunner
[1955], [1960] ,Hadwiger-Debrunner-Klee [1964], Hadwiger-Debrunner-Yaglom [1965],
Jucovic [1967]. The estimate of 73 is due to Elliott [1967], who also made the
Conjecture 3.10. Jf V is aset of m points such that n(G(V))>1 then r,(C(V))=
m*/6 + O(m).
* * *

(II) One possible generalization of arrangements of curves consists of considering such


families of simple closed curves in which each pair intersect in at most (or else in precisely)
2k points. The maximal possible numbers of cells for such arrangements involving n curves

EO) OP iia

2 See te

Figure 3.36
Selfintersection patterns with 1, 2, or 3 vertices.
72 BRANKO GRUNBAUM

were determined by Robinson [1945]. (The failure of Robinson’s [1945] results concerning
higher-dimensional arrangements was observed by Frame [1945]; it is caused by imprecision
in defining the conditions.) Arrangements of these types occur also in More’s [1959] dis-
cussion of “Venn diagrams”.
Considerably more attention has been given to the somewhat related topic of selfinter-
section patterns of closed curves. Given a closed curve C in the plane (or on the sphere),
with a finite number k of selfintersections (vertices) at each of which the curve just
crosses itself, a cell complex C= C(C) is determined in the plane (the unbounded region
being considered as a cell); we call it the selfintersection pattern of C. It is clearly possible
to define isomorphic selfintersection patterns, and to investigate the patterns in a manner
analogous to the way we investigated arrangements of curves. The various non-isomorphic
types with k <3 are shown in Figure 3.36. As an example of easily established results
we mention that p,(C(C)) is at most 27-1 whenever k is 27-1 or 2.
One aspect of selfintersection patterns has no analogues in the theory of arrangements
of curves although it is of interest for areas as disparate as knots in 3-space (both theoretical—
see, for example, Tait [1877], Reidemeister [1932], Crowell-Fox [1963], Treybig [1968] —
and practical—see, e. g., Ashley [1944, Chapter 20] ), functions of a complex variable (Titus
[1960], [1961] ), and continua in the plane (Whitney [1937]). This is the so-called Gauss
code of a selfintersection pattern, defined as follows. First the k vertices of the pattern
C(C) are assigned distinct symbols taken, for example, from the set {1, 2, ++, kK}. The
Gauss code of C is the (cyclically understood) sequence of length 2k of symbols formed
by proceeding along the curve and noting the symbol of each vertex as it is traversed. Thus 14
the k= 7 selfintersections of the curve C in Figure 3.37 yield the Gauss code 123451
26477365. The Gauss code of each curve C clearly contains each symbol precisely
twice. Less trivial is the following observation of Gauss [1823] (“parity condition”):
THEOREM 3.17. The two appearances of each symbol in the Gauss code of a selfinter-
section pattern are separated by an even number of places.

Re
Figure 3.37
ARRANGEMENTS AND SPREADS 73

Proofs of Theorem 3.17 were given by Wiener [1864] (for polygonal curves), Landsberg
[1911], Sz. Nagy [1927], Rademacher-Toeplitz [1930] ; in most of them (as well as in a 15
series of papers published by G. Sz. Nagy between 1927 and 1930) various generalizations
and strengthenings of Theorem 3.17 were considered. The more interesting question, however,
is what properties besides the parity conditionare needed to characterize sequences which
are Gauss codes. Gauss [1844] has already known that from k= 5 on the parity condition
is not sufficient; he gave as examples the sequences 1231245345 and 1231435
4 25 which satisfy the parity condition but are not Gauss codes. Solutions of the charac-
terization problem have been found recently (Treybig [1968], Marx [1969]); however, 16
they are of the same aesthetically rather unsatisfying character as Mac Lane’s [1937]
criterion for planarity of graphs. A characterization of Gauss codes in the spirit of the Kura-
towski criterion for planarity of graphs is still missing.
As additional open problems we should mention the investigation of selfintersection
patterns in which multiple-intersection vertices are allowed. For other problems see Guy
[1970], [1971a]. 417
* * *

(II) The last notion to be considered in the present chapter is somewhat dual to
arrangements of pseudolines: starting with a number of points in the Euclidean plane, some
arcs are drawn connecting them (subject to appropriate restrictions); a cell-complex-like
structure is thereby determined in the plane, with vertices (among them the original points),
edges (which are subsets of the arcs) and cells of various kinds. In order to simplify the ex-
position we shall use the following terminology. Let G denote an abstract graph with
nodes (vertices, “points”) V,,°*-, V,;, and arcs (edges, “lines”)"E,, *** , By @ cross-
ing pattern C= C(G) of G we mean the complex determined in the plane by a set of n
points conveniently denoted by V,,°°:, V,n and a set of m simple curves (arcs) denoted
by -£,,°°:, £,, with endpoints in the set V,, +, V,” such that:
(i) the curve £; contains the point V; if and only if in G the node V; is incident
with £;:
(ii) the intersection E; E,, of two of the curves is either
(a) empty; or
(b) a common endpoint of both; or
(c) a single point relatively interior to both, at which the curves cross each other;
(iii) except for V,,°**, V,, each point of the plane belongs to at most two of the
curves £,.
A weak crossing pattern is defined similarly, with (c) replaced by:
(c*) a single point relatively interior to both.
In a rectilinear crossing pattern all the curves E; are straight-line segments.
For any crossing pattern C(G) we shall denote by vu(C(G)) the number of vertices of
C(G) of type (c) (or (c*)), so that f,(C(G)) = 2 + v(CG)).
74 BRANKO GRUNBAUM

Crossing patterns of graphs were investigated in a number of different contexts. The


minimal crossing number v,(G) of a graph G is the minimal value of u(C(G)) among all
crossing patterns C(G) of G; the maximal crossing number u*(G) is defined analogously.
Mutatis mutandis for the definition of the minimal and maximal rectilinear crossing numbers
v,(G) and v*(G). (In the recent graph-theoretic literature v,(G) is frequently referred to as
the “crossing number” of G.)
Let K,, denote the complete graph with n nodes, and K,.,” the complete bipartite
graph with n’ and n” nodes. The recent interest in crossing patterns started with P. Turan’s
question about v,(K,',”) (see Guy [1969] for a detailed historical account); it led to
Conjecture 3.11. v,(K,' n”) = [n'/2] [(n’ — 1)/2] [n"/2][(n" — 1)/2].
Invalid proofs of Conjecture 3.11 have been published by Zarankiewicz [1954],
Busacker-Saaty [1965, p. 49], and claimed by Urbanik [1955]. For partial results see Guy
[1969], Harary [1969, p. 123], Kleitman [1970], BlaZek-Koman [1967] ; concerning multi-
partite graphs see Harborth [1971].
Another open problem is Guy’s [1960]
Conjecture 3.12. v,(K,,) = [n/2] [(n — 1)/2] [(n — 2)/2] [(@ — 3)/2]/4.
For partial results on v,(K,,) see Ringel [1964], Blazek-Koman [1964], Moon [1965],
Saaty [1967], [1969], [1971], Kainen [1968], Harary [1969, p. 123],Guy [1971b]. Concerning 18
results about v,(K,,) see Harary-Hill [1962], Guy [1970], Jensen [1971] (see also Leclere-
Monjardet [1969]). For analogous problems on the torus see Guy-Jenkyns [1969], Guy-
Jenkyns-Schaer [1968]. The problem of determining v,(G) in case G is the graph of the
d-dimensional cube was considered by Eggleton-Guy [1971]; for a class of graphs with
v,(G) = 1 see Guy-Harary [1967]. An algebraic approach to ‘crossing numbers” was in-
vestigated by Tutte [1970]. For a recent survey see Guy [197Ic].
Curiously, the numbers v*(K,,), v*(K," ,”), etc. seem not to have been investigated;
it is easy to see that v*(K,)= 1 but that there exist weak crossing patterns C(K,) with
ASUS) Neer
The implied assertion concerning v*(K,,) in Saaty [1969] is probably true, but unsub-
stantiated.
The numbers v'*(C,,), where C, is a simple circuit with
n nodes, were the subject
of many investigations; the earliest seem to be those of Balzer [1885], Brunel [1894], and
Brickner [1900, pp. 10—12].
However, all those are either incomplete or incorrect, and the
complete solution was found only by Steinitz [1923] (see also Steinitz [1922, §4]):

THEOREM 3.18. The values of v(C(C,,)) possible for rectilinear crossing patterns of
Care:
(i) for n= 2k +1, all integers from 0 to (2k + 1)(k-1) except (2k+ 1)(k-1)-1;
(ii) for n= 2k, all integers from 0 to 2K = 2) iz
An independent proof for the value of v*(C,,) resulting from Theorem 3.18 was given
by Bergmann [1969].
ARRANGEMENTS AND SPREADS 75

The range of values of u(C(C,,)) for C(C,,) varying over all crossing patterns of C,,
appears not to have been investigated. We have:

Conjecture 3.13. The range of values of v(C(C,,)) for all crossing patterns of C,, coin-
cides with that for all rectilinear crossing patterns of C.,,.
It was observed by L. M. Kelly and G. C. Shephard (private communications) that the
range of values of u(C(C,,)) for rectilinear crossing patterns C(C,,) in the projective plane
coincides with the set of integers from O to n(n — 3)/2 for all n> 3.
J. H. Conway raised the question (see Guy [1970], [1971a]) about the maximal pos-
sible number m of arcs £, in graphs G with n nodes which can be “thrackled”, that is
for which a crossing pattern C(G) exists in which every two curves have a common point.
(in other words, every pair of curves E, of C(G) has either a common endpoint, or else
cross each other.) Conway made the rather surprising
Conjecture 3.14. [fa graph with n nodes and m arcs can be thrackled then m <n.

Assuming the validity of Conjecture 3.14 Woodall [1971] characterized all graphs that
can be thrackled.
We shall say that a graph G may be weakly thrackled provided there exists a weak
crossing pattern C(G) in which the intersection of each pair of curves consists of precisely
one point. In analogy to Conjecture 3.14 we have

Conjecture 3.15. Jf a graph with n nodes and m arcs can be weakly thrackled then
We 2 = 3;
It is not hard to verify this conjecture for m <5; the examples in Figure 3.38 show
that equality may be attained at least for n <6.
ARRANGEMENTS AND SPREADS WH

Chapter 4. Spreads of curves


4.1. Definition and properties of spreads.
In the present chapter we shall discuss some of the properties of spreads of curves, ob-
jects which in many respects may be considered as continuous analogues of the arrangements
of lines and pseudolines we considered in Chapters 2 and 3.
Actually, the investigation of spreads of curves was motivated by results in the theory
of convex sets (see Grinbaum [1966]), and the analogy to arrangements seems to have es-
caped previous notice.
Properties of certain special families of chords (such as area-bisectors, diameters, etc.)
of planar convex sets have been investigated by many authors. In many instances it turned
out that the results do not depend very much on the particular type of chords considered,
or even on their being straight-line segments, but rather follow from the way members of
the family are related to each other. This led to the investigation of a number of such re-
sults in a more general setting, referred to in the literature by the terms “continuous families
of curves”. As this is only a vaguely descriptive term, we prefer now to use the more spe-
cific designation “spread of curves”. (According to Webster, one of the meanings of the
word “spread” is “continuous assemblage”. It should be noted, however, that “spread”’ is
used in a different meaning by some authors (see, for example, Dembowski [1968, p. 29])
to denote what others call “fibration” (see Segre [1964] ).)
Following Grtinbaum [1966] we shall now define the ‘“‘spreads of curves” and discuss
some of the results known about them; following that we shall give examples of families
of curves to which the theory of spreads applies. At the end of the chapter we shall discuss
the relation of spreads to arrangements.
Let K be a convex body (that is, a compact convex set with non-empty interior
int K) in the Euclidean plane £%, and let C= bd K be the boundary of K. A family
L ={L} of simple (open) arcs is a spread of curves in K provided:
(1) Each L © L is contained in int K, while its endpoints are different and belong
107 €
(2) Each p €C is an endpoint of a single curve L = L(p) in L.
(3) If L’ and L” are different curves in | then L' L" isa single point.
(4) The curve L(p) depends continuously (in the Hausdorff metric) on the point
DaG
Using the Jordan curve theorem and other continuity arguments it is not hard to verify
that each spread L of curves has also the following properties:
(5) For distinct L', L" € L the endpoints of L’ separate on C those of L’.
(6) A continuous involution * is defined on C by assigning to each p€C the
other endpoint p* of L(p).
(7) For distinct L', L" € L, the unique point L' L" depends continuously on
dy and
78 BRANKO GRUNBAUM

The above definition of spreads of curves is only a reasonable compromise between the
possibilities of far-reaching generalizations, and the wish to avoid technical complications.
Among the generalizations, we could mention the possibility of allowing C to be any simple
closed (Jordan) curve, or (see Zamfirescu [1969a] )the weakening of condition (3) to
(3') If L' and L” arein L then L’ OL" is connected.
A number of related notions may be treated in a completely analogous fashion. For ex-
ample, we may replace the set K in the definition by the whole plane E?, and modify appropri-
ately the conditions involving C; we shall refer to such objects as to “FE? spreads”. Or else, we
may insist that all the curves of the spread be straight-line segments (or straight lines in case of
E?-spreads); in those cases we shall speak about spreads of segments and spreads of lines.
Let k be acardinal number; a point x € K is called a k-tuple [k-fold] point of L
provided x belongs to precisely [at least] k different members of L. The set of all
k-tuple [all k-fold] points of L is denoted by 7,(L) [by F;,(L)]. Clearly F,(L)=
Ups T(L)-
Some of the simplest results on spreads of curves are the following:
THEOREM 4.1. T)(L)=2.
Proor. By the Jordan curve theorem the complement of each Z in int K consists
of two disjoint regions which, from an endpoint p of L =L(p), may be distinguished as
being “‘to the left” and “‘to the right” of L. The property “x € int K is in the “left” region
of L(p)’ depends continuously on p unless x € L(p); but “left” and “right” are inter-
changed for the two endpoints of L. Hence each x € int K belongs to at least one L € L.
The next result is slightly less trivial:

THEOREM 4.2. L 1 F,(L) = is possible for at most one L € L.


PRooF. Let Ly =L(pg) be a curve of the spread L, and let X(p)=Ly 1 L(p) be
the continuous mapping from the open arc A of C, with endpoints py and Po, into Lo-
Then the following three possibilities arise:
(a) X(p) is a strictly monotone function of p € A,
(b) X(p) is a monotone but not strictly monotone function of p € A,
(c) X(p) is not a monotone function of p € A.
In case (a) X establishes a homeomorphism between A and X(A) C Ly; through each
point of Ly there passes at most one curve L © L different from Ly. In case (b) there
exists a closed arc [p,, p,] C A such that K(p) is constant for all p € [P;, Py]; hence
the point X(p,) of Ly belongs to Ty(L) and thus to F;(L). In case (c) there exists in
Ly an open arc with endpoints X(p') and X(p") such that each point of that arc is the
image of at least two different points of A, and thus belongs to Ft).
Therefore the proof of Theorem 4.2 shall be completed if we show that L may con-
ily at most one curve of type (a). Assuming otherwise, let Ly = L(p,) and L, =L(p,) be
distinct curves of type (a). Let P, bea point of A not belonging to the set (Do Diop Pas Pik
thus L, = L(p,) is neither L, nor L,. Adopting the notation indicated in Figure 4.1 we
ARRANGEMENTS AND SPREADS 719

consider a point p belonging to that open arc (p%, p*) of A that does not contain py.
Since Ly is of type (a) the point Ly M L(@) belongs to the open arc (Po, Z)rotaLe:
similarly the point L, % L(p) belongs to the open arc (p,, z) of L,. But this is impos-
sible since (by the Jordan curve theorem) L(p) must intersect the closed subarc of Ly UL,
that contains p65, v, Zz, w, py, and thus either Ly N L(y) or L, VYL@) would consist of
at least two points. This completes the proof of Theorem 4.2.
It should be noted that L may indeed contain one exceptional curve Ly of type (a),
even if L is a spread of segments. For example, taking as C the circle of Figure 4.2, let
Ly =L@o) be a diameter of C, (x,, x.) an open subinterval of L,, and ¢ any homeo-
morphism of the open semicircle (pp, p%) onto (x,, 2). Then Lp» and the chords of C
determined by the segments [p, ¢(p)] for p © (Go, pd) yield a spread of segments in which
Lo is of type (a).
Theorem 4.2 may be found, together with the above proof, in Griinbaum [1966] ;
among its corollaries we mention:

THEOREM 4.3. For each spread [ either F,.(L)#@ or else card F3(L)=%.
THEOREM 4.4. If F3(L) isa single point x, then x, ©L for each cure LE L.
Other results about the sets F, and TJ, are:
THEOREM 4.5. For every spread L

F,(L) C cl F3(L).
(cl A denotes the closure of the set A.)
THEOREM 4.6. For every spread | and for every integer j 2 1, int T,(L) =o. If
OS# Fy AL) #F..(L) then int ORE PEACE) FO.
Theorem 4.5, and the first part of Theorem 4.6 in case j= 1, were established in
Griinbaum [1966]. Theorem 4.6 and additional results concerning relations among the sets
F,, and 7; are due to Zamfirescu [1969a]. As an unsolved problem dealing with this topic
we mention

Conjecture 4.1. If F,(L) #@ and 1<j<k<- then TL) #2.

Cc

Figure 4.1 Figure 4.2


80 BRANKO GRUNBAUM

Connectedness properties of the sets F’,(L), along with other topics concerning spreads
and some related, more general, objects were discussed in Grunbaum [1966] and Zamfirescu
[1967a], [1967b], [1967c], [1968a], [1968b], [1969a], [1969b]. As examples of such re-
sults we may mention the following (see Griinbaum [1966] for the assertions concerning Py
Zamfirescu [1969a] for F'):
THEOREM 4.7. The set F,(L) is L-convex, that is, for each L © L the set i
F;(L) is either empty or connected.
THEOREM 4.8. The sets F,(L) and F,(L) are L,(L)sets.
Here a set A is called an L,(L)-set provided every two points of A may be joined
within A by a curve composed of at most two arcs of elements of L.
Generalizing those results we have:
Conjecture 4.2. For each j > 1, the set FL) is L-convex, and it is an L,(L)-set.
Of particular relevance for the theory of convex sets is the following result of Zamfir-
escu [1969a] :
THEOREM 4.9. Jf g, and g, are continuous maps of L into itself, there exists
LEL such that LNg,€)8,(£) #2.
Proor. If g,(L)=g,(L) for some L € L, the assertion is obviously true. But if
PIL) = g,(L) 1 g,(L) is a single point for every L © L, then p(L)€L for some L by
the reasoning used in the proof of Theorem 4.1.
Similarly simple proofs may be given the following two results:

THEOREM 4.10. If © is a fixpoint-free continuous involution of L onto itself, and


if g is a continuous function from L to the reals, then there exists a curve L © L such
that g(L) = g(O(L)).
THEOREM 4.11. Let L and M be two spreads on the same set K, and let g bea
continuous map from L to M. Then there exists a curve L © L such that LQ g(L) #@.

Some generalizations of spreads of curves to higher dimensions have also been con-
sidered in the literature. Extending an earlier result of Forrester [1952], Stein [1954] has
proved certain theorems of that type. One of Stein’s results is the following generalization
of Theorem 4.1 formulated for spreads of lines (see Hadwiger [1961] for an elementary
proof in case n= 3):

THEOREM 4.12. [f to each direction u inE" a line L(u)=L(-u) parallel to u is


assigned in a continuous manner, then each point of E" belongs to some L(u).
For other results on spreads of lines in E” see Hammer [1954], Sobezyk [1956].
Generalizations of Theorem 4.2 to appropriately defined higher-dimensional spreads of
curves (in which case even the assertion F, # @ is non-trivial) were given by Kosiriski [1957],
[1958]. An easily stated special case of one of Kosiriski’s results is: #19
THEOREM 4.13. If | is a family of simple curves in the n-ball B” such that one
ARRANGEMENTS AND SPREADS 81

curve of the family has its endpoints in each pair of antipodal points of bd B", and the
curve depends continuously on its endpoints, then F3(L) #2.
4.2. Examples of spreads.
Zindler [1921], [1922] investigated systematically many types of “remarkable chords”
of planar convex bodies, thereby providing a number of examples of spreads of curves and,
in particular, spreads of segments. Conversely, the theorems on spreads quoted above spe-
cialize for those families of chords to results that have frequently been obtained in an inde-
pendent fashion.
The following types of “remarkable chords” were among those considered by Zindler
[1921]; the totality of chords of each type forms a spread of segments. For each type, the
“remarkable chord” is the unique chord L of the planar convex body K parallel to a given
direction u and having the property:
(i) L bisects the area of K;
(ii) L bisects the perimeter of K;
(iii) LZ divides K into parts having equal moments of inertia about aff L:
(iv) L divides bd K into parts having equal moments of inertia about aff L;
(v) L divides K into parts having centroids equidistant from L;
(vi) L is equidistant from the support lines of K parallel to L; we shall call such
chords “‘midparallels of K”’;
(vii) K is smooth and rotund and L is the longest chord parallel to u; such chords
are usually called “diameters of K’’.
Generalizations of the families of chords of type (vii) to all planar convex bodies, and
E*-spreads of lines (called “‘outwardly simple line families”) formed by “extended diameters”,
have been discussed by Hammer-Sobczyk [1953a], Smith [1961], Ceder [1964] and others
(see below; for a more detailed account see Hammer [1963]).
An additional type of “remarkable chords” of a smooth and rotund convex body K
is easily seen to be given by the set of all chords L of K such that
(viii) L is equidistant from the two chords of K parallel to L and having length
equal to half the length of L.
An important spread of curves is given by the “‘midcurves” of a rotund and smooth
planar convex body K. If wu is a direction in the plane, the midcurve L,, of K corre-
sponding to wu is the set of midpoints of all proper chords of K parallel to u. For dis-
cussions of the spread of midcurves see Brunn [1889], Emch [1913], Zindler [1921],
[1922], Steinhaus [1957]. It should be stressed that the above definition applies to all
planar convex bodies; however, in general the intersection of two different midcurves will
be either a point or a segment, so that the midcurves do not form a spread under the defin-
ition adopted here. It is also worth mentioning that the assertion concerning the intersec-
tion of two midcurves in the case of general K is not as trivial to prove as is often intimated.
However, concerning midcurves and many other spreads, in many instances it is enough to
82 BRANKO GRUNBAUM

establish an assertion for smooth and rotund K, since the general case may easily be deduced
by a limit process.
Part of the interest in midcurves is due to the following fact:
THEOREM 4.14. All midcurves of a planar convex body K are straight-line segments
(if and) only if K is an ellipse.
This result is essentially due to Bertrand [1842], where it is proved for not necessarily
convex curves, but with smoothness assumptions. The popularity of the theorem is reflected
in the number of published proofs, many of which require some degree of smoothness; see
Blaschke [1916, p. 158], [1923, p. 23], Nakajima [1928], Berger [1936], Kubota [1939],
Kneser [1949], Busemann-Kelly [1953, p. 140], Busemann [1955, p. 9] , Danzer-Laugwitz-
Lenz [1957], Stiss-Viet-Berger [1960].
Among the many known variants and generalizations of Theorem 4.12 the following
two seem worth mentioning in the present context:

THEOREM 4.15. Jf the planar convex body K has infinitely many straight midcurves
then K is an ellipse.

THEOREM 4.16. If € >0 and if K is a planar convex body with straight e-mid-
curves then K is an ellipse.

(Here we say that K has straight e-midcurves if for each x € bdK the midpoints of
all chords of K parallel to a support line of K at x and contained in an e-neighborhood
of x, are on a straight line.)
It may be noted in passing that Theorem 4.16 would not remain valid if its assump-
tions were weakened to read: For each x € bd K there is an € = e(x) such that the mid-
points of all chords of K, parallel to a support line of K at x and belonging to an

Figure 4,3
ARRANGEMENTS AND SPREADS 83

e(x)-neighborhood of x, are collinear. A non-elliptic example of K with this property may


be constructed (see Figure 4.3) by choosing two mutually perpendicular normals of an ellipse
which are not the axes of the ellipse, and taking the arc determined by these normals and its
mirror images in the normals.
One possible generalization of midcurves seems not to have been investigated in the
literature, although it appears to lead to interesting problems. To define such “bisection-
curves” let a mass-distribution M be given in the plane, or in a suitable part of it. Fora
given direction u, let T(u, 7) > Tu, j + 1) be a sequence of strips bounded by lines paral-
lel to u and of widths tending to 0. Let G(u, j) be the sequence of the centroids of the
part of the mass M that is contained in 7T(u, j). If G(u, /) is a convergent sequence, its
limit G(u) will be one of the points of the bisection-curve of M corresponding to direction
u. Under suitable restrictions on M the bisection-curves of M form a spread. The mid-
curves correspond to the case in which M is a uniform mass-distribution on a convex body
K. It seems likely that the investigation of bisection-curves and the analogues of Theorems
4.14, 4.15, and 4.16 will be of special interest in the case in which M is proportional to the
length of bd, or to the curvature of bd K, for sufficiently smooth convex bodies K.
The following are examples of theorems obtainable by specializing general results on
spreads of curves.
Theorem 4.11 clearly implies the result of Zindler [1921] that for each two types of
“remarkable chords” of the same body K there is a chord of K which is remarkable in
both senses. Brennan [1958] and Zitronenbaum [1959] observed this fact in relation to
spreads of bisectors of area and bisectors of perimeter (types (i) and (ii) above). Applied to
the E?-spreads of lines formed by the bisectors of each of two different (continuous) mass-
distributions in the plane, Theorem 4.11 yields the “bread-and-butter-sandwich theorem” 20

(the poor man’s version of the famous “ham-sandwich theorem’’—see, for exarnple, Stone-
Tukey [1942]) asserting the existence of a straight line bisecting both masses. In analogy to
Zindler’s result, Hadwiger [1961] proved that every two E%-spreads of straight lines have a
common line.
Theorem 4.4 easily implies the result of Zarankiewicz [1959] (see also Piegat [1963],
Menon [1966]) for bisectors, and of Viet [1956] for midcurves:
THEOREM 4.17. Jf a planar convex body has only one point through which pass 3 or
more area (or perimeter) bisectors, or midcurves, then this point is a center of symmetry of
the body.

For far-reaching generalizations of the result of Theorem 4.17 in case of the spread L
of midcurves see Chakerian-Stein [1966], where a measure of symmetry is defined using the
set T,(L). Many “remarkable points” of each convex body belong to F,(L), that is, bisect
at least three different chords: the (area) centroid (Bose [1935], Ehrhart [1955], Viet [1956]),
the perimeter centroid (Bose-Roy [1935]), the Steiner point (“curvature centroid”) (Bose
[1935]), the inellipse center and the circumellipse center (Behrend [1938] ), each sixpartite
point, etc. Concerning connectedness properties of F3(L) see Ceder [1965b].
84 BRANKO GRUNBAUM

The alternative of Theorem 4.3 was noted in different instances by Steinhaus [1955]. #22
Theorem 4.1 for the family L of midcurves of K implies that int K C F,(L); the
case of general K may be deduced by an approximation argument from that of smooth and
rotund K.
For strengthenings of Theorem 4.6 in the case of spreads of midcurves see Chakerian-
Stein [1966], and in the case of spreads of diameters (longest chords, type (vii) above) see
Hammer-Sobczyk [1953b], Ceder [1965a]. For the spreads of diameters, the validity of the
second part of Theorem 4.6 was established by Ceder [1964].
Theorem 4.7 may be used to derive the result of Buck-Buck [1949] concerning the ex-
istence of sixpartite points for each planar convex body K (that is, the existence of three
concurrent lines such that each of the six wedges determined by them contains 1/6 of the
area of K) and its generalizations (Eggleston [1953], Griinbaum [1963], Ceder [1965a]). 23
Applied to midcurves, Theorem 4.9 specializes to a result of Steinhaus [1957]: For every
planar convex body K and for positive reals a, 8B, y such that a + 8B + y= 77 there exist
mutually bisecting chords of K such that the angles between consecutive half-chords are
a, B, y, a, B, y. In case of the spread of chords of type (viii) Theorem 4.9 yields the existence
of an affinely regular hexagon inscribed into bd K (see Besicovitch [1948], Fary [1950],
Fejes Toth [1953, p. 102]).
Interpreting the involution © as perpendicularity it is easy-to derive from Theorem
4.10 the observation of Zindler [1921] that each planar convex body may be divided into
4 parts of equal area by a pair of perpendicular lines. Rather surprisingly, the following
problem due to B. J. Birch is still open: Given a real number A with O<A< 1/2, does
there exist for each convex body K (or for each continuous mass-distribution in the plane)
a pair of perpendicular lines such that the quadrants determined by them contain, clockwise,
the fractions A, A, 1/2 — A, 1/2 —2 of the area of K? Zindler’s remark provides an affirma-
tive answer for X= 1/4; the question is undecided for all other values of X.
An application of Theorem 4.10 to the spread of midcurves of K leads to

THEOREM 4.18. The boundary of every planar convex body contains the vertices of
a square.

This last result was the subject of many investigations. It seems that it was first estab-
lished by Toeplitz [1911] (though the proof appears not to have been published). Indepen-
dently it was discovered by Emch [1912], [1915]; for other proofs see Zindler [1921],
Christensen [1950], Kakeya [1916], Biernacki [1953].
It has been frequently conjectured (see, for example, Frink [1949]) that Theorem
4.18 may be generalized to the assertion that every Jordan curve in the plane contains the
vertices of a square. For sufficiently smooth (twice differentiable) Jordan curves this has
been proved by a number of authors; See Snirelman [1929], Ogilvy [1950], Jerrard [1961],
Guggenueimer [1965]. Related to this is a conjecture of Hadwiger [1971]: Every simple
closed curve in the Euclidean 3-space contains four distinct points which are the vertices
(perhaps collinear) of a parallelogram. Similarly, generalizing the fact mentioned above
ARRANGEMENTS AND SPREADS 85

concerning the existence of affine-regular hexagons inscribed in convex sets we make


Conjecture 4.3. Every Jordan curve in the plane contains the vertices of an affine-regu-
lar hexagon.

This conjecture is open even for smooth curves.Related to it is the result of Schaffer
[1968] that every simple closed curve in the Euclidean 3-space, which has the origin 0 asa
center of symmetry and which meets each ray issuing from 0 in at most one point, contains
the vertices of an affine-regular hexagon.
4.3. Arrangements and spreads.
One of the main reasons for presenting here the material just surveyed on spreads of
curves is the conviction that spreads are a natural, continuous counterpart of the discrete
arrangements of lines and of pseudolines considered in Chapters 2 and 3. Indeed, if a spread
of curves is given in the convex set K and if K is obtained from K by abstractly identify-
ing each pair of points of C= bd K which are the endpoints of a curve of the spread, then
K isa topological space homeomorphic to the projective plane. Every finite family of
curves of the spread determines a (possibly trivial) arrangement of pseudolines in the projec-
tive plane K. This remark leads at once to a wide variety of questions, most of which have
not been considered in the literature. We shall mention here only a few samples of very
natural questions of this type.
It is very easy to see that every arrangement of lines may be imbedded in (that is, en-
larged to) an E*-spread of curves (and even to an E?-spread of lines). We make
Conjecture 4.4. Every arrangement of pseudolines may be imbedded in an E* -spread
of curves.
Levi’s [1926] extension lemma (our Theorem 3.4) lends credibility to this conjecture,
but we were unable to derive from it a proof of the conjecture.
Unless all the curves of a spread pass through one point (‘‘trivial spread’’), the spread
contains four curves which determine a simple arrangement of 4 pseudolines, and other four
curves which determine a near-pencil of 4 pseudolines. It is not known what other types of
arrangements of pseudolines may be found in all non-trivial spreads.
Let two spreads L and L’ (in K and K’) be called isomorphic provided there is a
homeomorphism of K onto K’ that carries curves in [ to curves in L’. Isomorphic
spreads clearly have isomorphic collections of arrangements of finitely many of their curves;
it is not known to what extent the converse is valid. We believe
Conjecture 4.5. There exist non-isomorphic spreads L and L' such that every
arrangement contained in | is isomorphic to some arrangement contained in L', and vice
versa.
Conjecture 4.6. There exist uncountably many pairwise non-isomorphic spreads.

It seems not to be known in any case how to characterize (intrinsically) spreads isomor-
phic to some “natural” spreads, such as those formed by diameters, or area-bisectors, or
midcurves, etc.
86 BRANKO GRUNBAUM

It is easily seen that the spread of midcurves of a smooth and rotund convex body K
in the plane is isomorphic to the spread of midparallels (type (vi) above) of each body K*
polar to K (Heil [1971]). Bose [1935] and Heil [1971] show that for every smooth and
rotund convex body K in the plane each k-tuple point of the spread of midcurves of K is
also a k-tuple point of the spread of midparallels of K. This enables one to construct exam-
ples which establish:
THEOREM 4.19. Two spreads L and L’ on the same set K may satisfy T,,(L)=
T,(L) for all cardinals k without the spreads being isomorphic.
Finally, Sylvester’s [1893] problem concerning arrangements of lines or pseudolines
(see Theorems 2.12, 2.15, and 3.5) leads one to the following special case of Conjecture 4.1:
Conjecture 4.7. For every non-trivial spread of curves L we have T,(L)#2, that is,
there exist points belonging to precisely two curves of the spread.

Carrying the analogy to arrangements somewhat further we venture

Conjecture 4.8. For every non-trivial spread L

card TCL) = 3.
4.4. Topological planes.
Starting with Hilbert [1899] (or possibly even earlier) there have been very many
papers dealing with “topological planes” or higher-dimensional spaces. By a “topological
plane” we mean here roughly the following: Given is a set B, homeomorphic to the Euclid-
ean plane (or to the projective plane), and a family H = {4} of “lines”, that is simple (or
simple closed) curves in B, which has some of the properties of the family of all straight
lines in the Euclidean (or the projective) plane.
The properties assumed in the investigations of Cc 8
“topological planes” vary greatly from
one author to another, as do the authors’ aims and the terminology. We mention briefly
only a few examples:

(i) Hilbert [1899] and Moulton [1902] are interested in questions of mutual depen-
dence of axioms for Euclidean geometry; they assume as satisfied mainly the axioms of
incidence and order.
(ii) Artin [1940] starts with only very weak incidence assumptions on B and H
(B is not even assumed to be homeomorphic to the plane, nor the members of { to be
curves) and shows how to “coordinatize” B by certain algebraic objects. Algebraic proper-
ties of those objects are then found to be related to geometric properties of the “topological
planes”. In particular, the algebraic structure simplifies greatly if certain configurational
statements such as Desargues’ theorem are assumed to hold. This direction has led to a
great proliferation of literature;
for modern introductions to the subject the reader is referred
to Artzy [1965] or Bumcrot [1969].
(iii) Skornyakov [1954], Salzmann [1955], and others assume that the “line” deter-
mined by two points depends continuously on the points, and that the intersection point
ARRANGEMENTS AND SPREADS 87

of two “lines” depends continuously on the “lines”. Many of the investigations of this gen-
eral direction are devoted to characterizations of the Euclidean or the projective plane. A
survey of the literature may be found in Salzmann [1967]. The work of Gemignani leads
in a related direction (see Gemignani [1966] and later papers by the same author).
(iv) Related in spirit to (ii) and (iii) above, but also to the investigations on arrange-
ments of pseudolines and on spreads, is the “geometry of webs” (see especially Blaschke-Bol
[1928] ). The geometry of webs may be used as a very appealing introduction to projective
geometries and their coordinatization, and as a starting point for relating combinatorial con-
cepts with differential geometric ones. It is to be regretted that no suitable account of that
theory is available in English, and that the approach has become rather unfashionable.
(v) Busemann [1942], [1955] assumes B endowed with the structure of a metric
space, the members of fH being geodesics in that metric. In his own words, Busemann
gives “‘a geometric approach to qualitative problems in intrinsic geometry”’.
(vi) Haupt [1965] considers systems of curves in B which generalize, among others,
the family of all circles in the plane. Under very general conditions he proves various order
and continuity properties of his systems, which extend facts previously known about the be-
haviour of families of functions (see, for example, Tornheim [1950] ).
The number of examples could be greatly increased, but we shall not do so since the
only aim of the preceding lines (which obviously did not even attempt to do justice to the
topics mentioned) was to make the reader aware of the wide variety of investigations that
impinge on the notion of “topological plane”. In the following pages we shall also restrict
ourselves to hasty sketches, the purpose of which is to point to facts and problems that
seem relevant to the theories of arrangements and apreads we discussed earlier.
We shall call topological Euclidean plane T= T(B, H) any open and bounded con-
vex set B in the Euclidean plane E*, together with a family H = {H} of lines, that is
simple open arcs H in B, such that:
1. Each H has two distinct endpoints which belong to bd B.
2. For every two distinct points b,, b, of B there is a unique line H= H(b,,b,)EH
that contains them; H(b,, b,) depends continuously (in the Hausdorff metric for subsets
of £*) on the points b, and b,. Similar statements hold if b, © bd B provided
“b, H’”’ is replaced by oD; is an endpoint of H”’.
3. For each two different lines H,, H, © H the intersection H, H, is either a
single point which depends continuously on H, and H, which cross each other at that
point, or else H, © H, =g and the endpoints of H, coincide with those of H,. In the
latter case (or if they coincide) H, and H, are said to be parallel.
As in the case of spreads of curves considered earlier, the above definition is only a
convenient one; no attempt was made to remove from it superfluous requirements, or to
attain maximal generality. A number of related notions may be defined by minor changes;
for example, B could be the whole Euclidean plane E*, or else endpoints of lines could be
identified to yield topological projective planes, etc.
88 BRANKO GRUNBAUM

We shall say that topological Euclidean planes T(B, H) and T’(B’, H ’) are isomorphic
if there exists a homeomorphism of cl B onto cl B' which carries the lines in H onto
those in ’.
The question whether there exist non-denumerably many pairwise non-isomorphic topo-
logical Euclidean planes appears to be one of the many still unsolved problems.
As examples of topological Euclidean planes that were considered in the literature we
may mention:
(a) The model of the Euclidean plane discussed by Gans [1955], [1969] and Zeitler
[1970] (see also May [1954], Gans [1958]). In that model B @ a circular disc, and each
HEH is either a diameter of B, or else a semi-ellipse in B with a diameter of B as
major axis. As is easily seen, this model may be obtained by first radially projecting the
Euclidean plane E? onto the open lower hemisphere of a sphere tangent to E*, and then
orthogonally projecting that hemisphere into E*. A quite different but isomorphic model
results if the second step of this procedure is replaced by stereographic projection into E?
In that model (which seems not to have been discussed in the literature, and which was
brought to the author’s attention by G. C. Shephard) the lines are: the diameters of B, and
arcs of circles connecting antipodal points of bd B. Other models of the Euclidean plane
result by projecting the lower hemisphere from other points.
(b). Generalizing the notion of midcurves (see page 81), Sholander [1953] considered
“-curves” of a given convex body K C E*. Here a d-curve in direction u, for some
with O<A< 1, is the set of all points which divide proper chords of K in direction u
in the ratio A:(1-A). As noted by Sholander [1953], the family of all A-curves (in all
directions and for all values of A) of a strictly convex and smooth convex body K deter-
mines a topological Euclidean plane on K. In case K is a circular disc, the model of the
Euclidean plane mentioned at the beginning of (a) is obtained.
(c) Another generalization of midcurves to a topological Euclidean plane may be ob-
tained by taking the polars of the points outside a smooth and rotund planar convex body
K. If p€ bd K the polar L, of p with respect to. K is the set of all points q with
the property that p and q are harmonically separated by the two points in which the line
pq intersects bd K. Midcurves are, clearly, polars of points belonging to the line at infinity.
For investigations involving polars see Brunn [1889], Kojima [1919], Marchaud [1948],
[1959], John [1937], Locher-Ernst [1951], Vincze [1952], Gergely [1957], [1959].
A connection between topological planes and spreads of curves is based on the remark
that for each T and for each by) © B the members of H that contain by forma
(trivial) spread of curves in the closure cl B of B. Also, any continuous selection of one
line from each class of mutually parallel ones clearly yields a spread of curves. Similarly,
each choice of finitely many lines in H leads to a (possibly trivial) arrangement of pseudo-
lines in B, the projective plane obtained from cl B by identifying pairs of points of bdB
which are endpoints of lines in H.
Conversely, one may ask whether a spread of curves may be imbedded in (that is, ex-
tended to) a topological plane by suitably defining additional lines. We venture
ARRANGEMENTS AND SPREADS 89

Conjecture 4.9. For every spread of curves L there exists a topological Euclidean
plane (on the same set of points) that extends L.
We are able to establish only the following much weaker result:

THEOREM 4.20. If K is a strictly convex and smooth convex body in the Euclidean
plane E* and if L is a spread of segments in K, then L may be extended to a topolog-
ical Euclidean plane in K.

PROOF. The method of proof is motivated by the A-curves discussed in (b) above. For
each chord L © L with endpoints A, and A, we consider (see Figure 4.4) in the Euclid-
ean plane E* the tangents T, to K at points A;, and their intersection point S (which
may be at infinity). Then, for each point Qj) © K with Q, € L we construct a “parallel”
to L as the set of all points Q@© K such that the cross-ratios satisfy (QP; RS) = (Q5Py; Ro).
Here P and R are the points of intersection with Z, and with one of the arcs of bd K
determined by A, and A,, of the arbitrary line T through S, while Py and Ro are
defined similarly by the line 7) determined by S and Q). It is not hard to complete the
proof of the theorem by verifying that the family of all the curves constructed, together
with the curves (segments) of the spread L, satisfy the conditions 1, 2,3 of the definition
of a topological Euclidean plane. (Another proof of Theorem 4.20, patterned after the
Shephard model of the Euclidean plane discussed in (a) above, may be based on the well-
known fact that two distinct homothets of a strictly convex set in E* may have at most
two boundary points in common.)

Figure 4.4
Extending a spread of segments to a topological Euclidean plane.
90 BRANKO GRUNBAUM

In analogy to Conjecture 4.9 we also make

Conjecture 4.10. Every arrangement of pseudolines ts extendable to a topological pro-


jective plane.

In the special case of arrangements of lines this conjecture may be established very
easily.
Some other open problems are:
What arrangements and what spreads of curves are present in every topological plane, or
in each topological plane of some specified kind?
To what extent is a topological plane determined by the collection of isomorphism
types of the arrangements, and the spreads, it contains?
What collections of spreads (or arrangements) may be simultaneously imbedded in a
topological plane? In particular, does there exist a topological plane universal for all
spreads (or arrangements), that is, containing isomorphic images of all types. The existence
of non-stretchable arrangements of pseudolines (Theorems 3.1 and 3.2) shows that the real
projective plane is not universal for all arrangements (or even for all simple arrangements) of
pseudolines.
Do there exist reasonable criteria characterizing topological planes isomorphic to those
of some special kind? For example, what planes are isomorphic to those obtainable (as in
the proof of Theorem 4.20) by extending a spread of segments, or to those having A-curves
as lines? Some of the results of the investigations mentioned in (iii) above fit in this con-
text by giving characterizations of topological planes isomorphic to the Euclidean plane
(see Salzmann [1967] ).
* * *

Just as we saw in Chapter 3 how easily and naturally one may consider arrangements
of curves alongside those of pseudolines, it is possible and interesting to replace the “‘lines”’
in topological planes by suitable “‘circles”. This type of geometry is usually called a
“Mobius plane” or a “‘circle plane”. It is the subject of numerous investigations, and it
would lead us too far to list detailed references; the reader may become acquainted with the
field and its literature, for example, through the papers Ewald [1956a], [1956b], [1967],
Benz [1960], Strambach [1970], Heise [1970a], Buekenhout [1971]. For extensions to
more general structures see similarly Heise [1970b].
One of the first important results in this direction was the characterization (Hesselbach
[1933], van der Waerden—Smid [1934]) of “circle geometries” isomorphic to the geometry
determined by all the circles of the (inversive) plane. This characterization corrects lapses
and errors made in this context by Reidemeister [1924] and Blaschke [1924a], [1924b],
[1930]; the efforts of Reidemeister and Blaschke were directed towards the proof of the
famous conjecture that spheres are the only surfaces in 3-space with the property that all
geodesics emanating from one point converge at another. (This conjecture was established
by Green [1963] using a completely different approach.)
ARRANGEMENTS AND SPREADS 9)

In a different direction of investigations related to circle geometries we should mention


the work of Buckel [19534] which characterizes, among families of subsets of the Euclidean
plane, the family of all circles. In a companion paper, Buckel [1953b] characterizes in a
similar manner the family of all non-degenerate conics in the projective plane.
x % %

It seems certain that a more thorough search of the literature would produce many
more examples fitting here, and that the situation is similar for most of the topics we have
discussed. But J hope that the material presented will suffice to convince the reader of the
wealth of openings for research present in the field, and also to convey the spirit of unity
and interdependence that manifests itself in many of the results and problems concerning
arrangements, spreads, and topological planes.
92 BRANKO GRUNBAUM

REFERENCES

The references are arranged alphabetically by the author’s name. Each item is followed by a listing
(in square brackets [ ]) of those sections in which that item is mentioned. In as far as such information
is available, each item is provided with a reference to a reviewing journal; MR stands for ‘‘Mathematical
Reviews”, FM for “Jahrbuch tiber die Fortschritte der Mathematik”.

Anonymous
1964 Programma van Jaarlijkse Prijsvragen; Problem No. 6, Nieuw Arch. Wisk. (3)
12 (1964), 64. [2.3]
E. Artin
1940 Coordinates in affine geometry, Notre Dame Math. Collog. 1940, 15—20.
(= Collected Papers (1965), S05—510.) MR 3 p. 179. [4.4]
R. Artzy
1965 Linear geometry, Addison-Wesley, Reading 1965. MR 32 #6274. [4.4]
C. W. Ashley
1944 The Ashley Book of Knots, Doubleday and Co., Garden City, N. Y., 1944. [3.4]
N. Balasubramanian
1953 A theorem on sets of points, Proc. Nat. Inst. Sci. India 19 (1953), 839. MR 15
p. 551. [3.4]
W. W. R. Ball
1960 Mathematical recreations and essays, Revised by H. S. M. Coxeter,
MacMillan Co.,
New York 1960. MR 8 p. 440. [2.3]
®
R. Baltzer
1885 Eine Erinnerung an Mobius und seinen Freund Weiske, Ber. Sachs.
Ges. Wiss.
Leipzig (1885), 1-6. FM 17 p. 518. [3.4]
L. Bankoff
1970 Solution of Problem 130, Math. Mag. 43 (1970), 233. [3.4]
J. G. Basterfield and L. M. Kelly
1968 A characterization of sets of n points which determine n hyperpla
nes, Proc.
Cambridge Phil. Soc. 64 (1968), 585-588. MR 38 #2040. [2.2]
F. Behrend
1938 Ueber die kleinste umbeschriebene und die grosste einbeschrieben
e Ellipse eines
Konvexen Bereiches, Math. Ann. 115 (1938), 379-411.
FM 64 p. 731. [4.2]
W. Benz
1960 Uber Mobiusebenen, Jber. Deutsch. Math.-Verein. 63 (1960),
1-27. MR 22
#7012. [4.4]
K. H. Berger
1936 Eilinien mit perspektiv liegenden Tangenten-und
Seh nendreiecken, S.-B. Heidel-
berg. Akad. Wiss. 1936, pp. 1-11. FM 62 p. 700. [4.2]
H. Bergmann
1969 Die maximale Anzahl von Uberschneidungen bei einem
Polygon, Arch. Math. 20
(1969), 107-112. MR 40 #7939. [3.4]
ARRANGEMENTS AND SPREADS 93

J. Bertrand
1842 Démonstration d’un théoréme de géométrie, J. Math. Pures Appl. (1) 7 (1842),
215-216. [4.2]
A. S. Besicovitch
1948 Measure of asymmetry of convex curves, J. London Math. Soc. 23 (1948), 237—
240. MR 10 p. 320. [4.2]
M. Biernacki
1953 Sur quelques properiétés des ovales, Ann. Univ. M. Curie-Sktodowska (Lublin), 7
(1953), 103-112. MR 16 p. 950. [4.2]
W. Blaschke
1916 Kreis und Kugel, Veit, Leipzig 1916. (Reprint: Chelsea, N. Y. 1949). FM 64
p. 1109; MR 17 p. 887. [4.2]
1923 Vorlesungen uber Differentialgeometrie. 11, Grundlehren der Math. Wiss. Vol. 7.
Springer, Berlin 1923. FM 49 p. 499. [4.2]
1924a Line topologische Kennzeichnung der Kreise auf der Kugel, Hamburg Math. Abh.
3 (1924), 164-166. FM S50 p. 488. [4.4]
1924b Eine Kennzeichnung der Kreise auf der Kugel, Math. Z. 21 (1924), 209-210.
FM 50 p. 488. [4.4]
1930 1930 Vorlesungen uber Differentialgeometrie, 1, Third edition. Springer, Berlin 1930.
(Reprinted by Dover, N. Y. 1945.) FM 56 p. 588; MR 7 p. 391. [4.4]
W. Blaschke and G. Bol
1938 Geometrie der Gewebe, Grundlehren der Math. Wiss., Vol. 49. Springer, Berlin
1938. FM 64 p. 727. [4.4]
J. Blazek and M. Koman
1964 “A minimal problem concerning complete plane graphs,”’ Theory of Graphs and

its Applications. (Proc. Sympos. Smolenice, 1963), pp. 113—117. Publ. House
Czechoslovak Acad. Sci., Prague 1964. MR 30 #4249. [3.4]
1967 On an extremal problem concerning graphs, Comment. Math. Univ. Carolinae 8
(1967), 49-52. MR 35 #1506. [3.4]
W. E. Bonnice and L. M. Kelly
1971 On the number of ordinary planes, J. Combinat. Theory A 11 (1971), 45—S3.
[2.6]
R. C. Bose
1935 A note on the convex oval, Bull. Calcutta Math. Soc. 27 (1935), 55-60. FM
61 p. 1428. [4.2; 4.3]
R. C. Bose and S. N. Roy
1935 Some properties of the convex oval with reference to its perimeter centroid,
Bull. Calcutta Math. Soc. 27 (1935), 79-86. FM 61 p. 1428. [4.2]
M. Bouten and P. de Witte
1965 A new proof of an inequality of Szekeres, de Bruijn and Erdos, Bull. Soc. Math.
Belg. 17 (1965), 475-483. MR 33 #3170. [2.2]
94 BRANKO GRUNBAUM

K. A. Brakke
1971 Some new values of Sylvester’s function for n noncollinear points, J. Undergrad.
Math. (to appear). [2.3]
J. G. Brennan
1958 A property of a plane convex region, Note 2808. Math. Gazette 42 (1958), 301—
302.
M. Brickner
1900 Vielecke und Vielflache, Teubner, Leipzig 1900. FM 31 p. 479. [3.4]
N. G. deBruijn and P. Erdos
1948 Ona combinatorial problem, Proc. Nederl. Akad. Wet. 51 (1948), 1277-1279.
MR 10 p. 424. [2.2; 2.3]
G. Brunel
1894 Note sur le nombre de points doubles que peut presenter le périmetre d’un poly-
gone, Mem. soc. sci. phys. et nat. Bordeaux (4) 4 (1894), 273-276. FM 25
p. 873. [3.4]
H. Brunn
1889 Ueber Curven ohne Wendepunkte, Ackermann, Munchen 1889. FM 21 p. 815.
[4.2; 4.4]
R. C. Buck and E. F. Buck
1949 Equipartition of convex sets, Math. Mag. 22 (1949), 195—198. MR 10 p. 621.
[4.2]
W. Buckel
1953a Eine Kennzeichnung des Systems aller Kreise mit nichtverschwindendem Radius
der euklidischen Ebene, J. Reine Angew. Math. 191 (1953), 13—29. MR 15
p. 149. [4.4]
1953b Eine Kennzeichnung des Systems aller nichtzerfallenden Kegelschnitte der pro-
jektiven Ebene, J. Reine Angew. Math. 191 (1953), 165-178. MR 15 p. 149.
[4.4]
F. Buekenhout
1971 Inversions in locally affine circular spaces. 1, Math. Z. 119 (1971), 189—202.
[4.4]
F. Buekenhout and J. Doyen
1970 Linear Geometry, Lecture Notes, University of Brussels, 1970. [2.2; 2.3]
R. J. Bumcrot
1969 Modern projective geometry, Holt, Rinehart and Winston, New York 1969.
MR 38 #6450. [4.4]
R. G. Busacker and T. L. Saaty
1965 Finite graphs and networks, McGraw-Hill, New York 1965. MR 35 #79. [3.4]
H. Busemann
1942 Metric methods in Finsler spaces and in the foundations of geometry, Annals of
Math. Studies No. 8. Princeton University Press 1942. MR 4 p. 109. [4.4]
ARRANGEMENTS AND SPREADS 95

1955 The geometry of geodesics, Academic Press, New York 1955. MR 17 p. 779.
[4.2; 4.4]
H. Busemann and P. J. Kelly
1953 Projective geometry and projective metrics, Academic Press, New York 1953.
MR 14 p. 1008. [4.2]
R. J. Canham
1969 A theorem on arrangements of lines in the plane, Israel J. Math. 7 (1969), 393—
397. MR 40 #7938. [2.3; 2.4]
1971 Ph. D. Thesis, University of East Anglia, Norwich 1971. [2.1; 2.3; 3.1; 3.2]
W. B. Carver
1941 The polygonal regions into which a plane is divided by n straight lines. Amet.
Math. Monthly 48 (1941), 667-675. MR 3 p. 180. [2.1; 2.4]
J. Ceder
1964 On outwardly simple line families, Canad. J. Math. 16 (1964), 1-11. MR 28
#517. [4.2]
1965a Generalized sixpartite problems, Bol. Soc. Matem. Mexic. 9 (1965), 28-32. MR
32 476318. [4.2]
1965b On a problem of Griinbaum, Proc. Amer. Math. Soc. 16 (1965), 188-189. MR
29 #3413. [4.2]
G. D. Chakerian
1970 Sylvester’s problem on collinear points and a relative, Amer. Math. Monthly 77
(1970), 164-167. MR 41 #3305. [2.3; 2.7]
G. D. Chakerian and S. K. Stein
1966 Bisected chords ofa convex body, Arch. Math. 17 (1966), 561-565. MR 34
#6635. [4.2]
C. M. Christensen
1950 A square inscribed in a convex curve, (In Danish) Mat. Tidsskr. B. 1950, 22—26.
MR 12 p. 525. [4.2]
A. B. Coble
1915 Points sets and allied Cremona groups, Trans. Amer. Math. Soc. 16 (1915), 155—
198. FM 45 p. 234. [2.7]
P. Cohen
1967 Decision procedures for real and p-adic fields, Mimeographed notes, Stanford
University 1967. [2.5]
H. S. M. Coxeter
1942 Review of Melchior [1940], Mathematical Reviews, 3 (1942), 13. [2.3]
1948 A problem of collinear points, Amer. Math. Monthly 55 (1948), 26-28. MR
9 p. 458. [2.3]
1949 The real projective plane, McGraw-Hill, New York 1949. (Second edition:
Cambridge Univ. Press, London 1955). MR 10 p. 729. [2.3]
1961 Introduction to geometry, John Wiley and Sons, New York 1961. MR 23
#A1251. [2.3]
96 BRANKO GRUNBAUM

1962 The classification of zonohedra by means of projective diagrams, J. de Math.


pures appl. (9) 41 (1962), 137-156. MR 25 #4417. [2.1; 23]
H. Croft and R. K. Guy
1971 Research problems in intuitive mathematics, (In preparation.) [2.3; 3.4]
D. W. Crowe
1969 Sylvester’s problem on collinear points, Question 10, J. Undergrad. Math. 1
(1969), 133. [2.3]
D. W. Crowe and T. A. McKee
1968 Sylvester’s problem on collinear points, Math. Magaz. 41 (1968), 30-34. MR
38 #3761. [2.3]
R. H. Crowell and R. H. Fox
1963 Introduction to Knot Theory, Blaisdell, New York 1963. MR 26 #4348. [3.4]
L. D. Cummings
1932a Hexagonal systems of seven lines in a plane, Bull. Amer. Math. Soc. 38 (1932),
105-110. FM 58 p. 676. [2.1]
1932b Heptagonal systems of eight lines in a plane, Bull. Amer. Math. Soc. 38 (1932),
700-702. FM 58 p. 676. [2.1]
1933 On a method of comparison for straight-line nets, Bull. Amer. Math. Soc. 39
(1933), 411-416. FM 59 p. 610. [2.1]
L. Danzer, D. Laugwitz and H. Lenz
1957 Uber das Lownersche Ellipsoid und sein Analogon unter den einem Eikorper
einbeschriebenen Ellipsoiden, Arch. Math. 8 (1957), 214-219. MR 20 #1283.
[4.2]
P. Dembowski
1968 Finite Geometries, Ergebnisse der Math. u. Grenzgebiete, Vol. 44. Springer-Ver-
lag, New York 1968. MR 38 #1597. [2.7; 4.1]
G. A. Dirac
1951 Collinearity properties of sets of points, Quart. J. Math. 2 (1951), 221—227.
MR 13 p. 270. [2.3]
H. E. Dudeney
1907 The Canterbury puzzles and other curious problems, W. Heinemann, London
1907;,(2.3]
1967 536 puzzles and curious problems, Charles Scribner’s Sons, New York 1967.
1273)
V. Eberhard
1890 Eine Classification der allgemeinen Ebenensysteme, J. Reine Angew. Math. 106
(1890), 89-120. FM 22 p. 553. [2.4]
1891 Zur Morphologie der Polyeder, Teubner, Leipzig 1891. [2.4]
M. Edelstein
1970 Generalizations of the Sylvester problem, Math. Magazine 43 (1970), 250—254.
(2.3)
ARRANGEMENTS AND SPREADS 97

H. G. Eggleston
1953 Some properties of triangles as extremal convex curves, J. London Math. Soc.
28 (1953), 32-36. MR 14 p. 896. [4.2]
R. B. Eggleton and R. K. Guy
1971 The crossing-number of the n-cube, (to appear). [3.4]
E. Ehrhart
1955 Une généralisation du théoreme de Minkowski, C. R. Acad. Sci. Paris 240 (1955),
483—485. MR 16 p. 574. [4.2]
P. D. T. A. Elliott
1967 On the number of circles determined by n points, Acta. Math. Acad. Sci. Hun-
gar. 18 (1967), 181-188. MR 35 #4793. [2.2; 3.4]
A. Emch
1913 Some properties of closed convex curves in a plane, Amer. J. Math. 35 (1913),
407-412. FM 44 p. 561. [4.2]
1915 On the medians ofa closed convex polygon, Amer. J. Math. 37 (1915), 19—28.
FM 45 p. 732. [4.2]
P. Erdos
1943 Three point collinearity, Problem 4065. Amer. Math. Monthly 50 (1943), 65.
(2323.23)
1957 Néhany geometriai problémarol, Mat. Lapok 8 (1957), 86—92. MR 20 #6056.
[3.4]
1961 Some unsolved problems, Publ. Math. Inst. Hungar. Acad. Sci. 6 (1961), 221—
254. MR 31 #2106. [2.3; 3.4]
1971a On a problem of Griinbaum, Canad. Math. Bull. (to appear). [2.2]
1971b Topics in combinatorial analysis, (to appear). [2.3]
G. Ewald
1956a Axiomatischer Aufbau der Kreisgeometrie, Math. Ann. 131 (1956), 354-371.
MR 18 p. 502. [4.4]
1956b Uber den Begriff der Orthogonalitat in der Kreisgeometrie, Math. Ann. 131
(1956), 463-469. MR 21 #5159. [4.4]
1967 Aus konvexen Kurven bestehende Mobiusebenen, Abh. Math. Sem. Univ. Ham-
burg 30 (1967), 179-187. MR 35 #3536. [4.4]
I. Fary
1950 Sur la densité des réseaux de domaines convexes, Bull. Soc. Math. France 78
(1950), 152-161. MR 12 p. 526. [4.2]
L. Fejes Toth
1948 Review of Coxeter [1948], Math. Reviews 9 (1948), 458. [2.3]
1953 Lagerungen in der Ebene, auf der Kugel und im Raum, Springer, Berlin 1953.
MR 15 p. 248. [4.2]
A. Forrester
1952 A theorem on involutory transformations without fixed points, Proc. Amer. Math.
Soc. 3 (1952), 333-334. MR 14 p. 72. [4.1]
98 BRANKO GRUNBAUM

J. S. Frame
1945 Review of Robinson [1945], Math. Reviews 6 (1945), 215. [3.4]
G. K. Francis
1969 Null genus realizability criterion for abstract intersection sequences, J. Combinat.
Theory 7 (1969), 331-344. MR 40 #5475. [3.4]
O. Frink
1949 Problem 4325, Amer. Math. Monthly 56 (1949), 423—424. [4.2]

D. Gans
1954 A circular model of the euclidean plane, Amer. Math. Monthly 61 (1954), 23—
30. MR1S5p. 460. [4.4]
1958 Models of projective and euclidean space, Amer. Math. Monthly 65 (1958),
749-756. MR 20 #4805. [4.4]
1969 Transformations and geometries, Appleton-Century-Crofts, New York 1969. [4.4]
C2 Fe Gauss
1823 Zur Geometria Situs, Unpublished manuscript. Werke, Band 8, p. 272. Teubner,
Gottingen 1900. [3.4]
1844 Zur Geometrie der Lage, fiir Zwei Raumdimensionen, Unpublished manuscript.
Werke, Band 8, pp. 282—286. Teubner, Gottingen 1900. [3.4]
M. C. Gemignani
1966 Topological geometries and a new characterization of R™, Notre Dame J. For-
mal Logic 7 (1966), 57-100. MR 34 #8259. [4.4]
E. Gergely
1957 La généralisation de la theorie polaire sur les ovales et les ovaloides, (Romanian;
summaries in Russian and French) Acad. R. P. Romine Fil. Cluj. Stud. Cerc.
Mat. 8 (1957), 143-160. MR 20 #3501. [4.4]
1959 Eine Verallgemeinerung der polaren Theorie auf Eilinie und Eiflache, Mathemat-
ica (Cluj) 1 (24) (1959), 221-237. MR 23 #A3507. [4.4]
L. W. Green
1963 Auf Wiedersehensflachen, Ann. Math. (2) 78 (1963), 289-299. MR 27 #5206.
[4.4]
B. Griinbaum
1963 Measures of symmetry for convex sets, Proc. Sympos. Pure Math. Vol. 7 (Con-
vexity) (1963), 233-270. MR 27 #6187. [4.2]
1966 Continuous families of curves, Canad. J. Math. 18 (1966), 529-537. MR 33
#4783. [4.1]
1967 Convex polytopes, Interscience, London 1967. MR 37 #2085. [2.1; 2.2; 2.3;
DA Doe) wan
1970a Polytopes, graphs, and complexes, Bull. Amer. Math. Soc. 76 (1970), 1131—
1201. MR 40 #5480. [2.4; 2.7]
1970b The importance of being straight, Proc. 12th Bienn. Internat. Seminar of the
Canad. Math. Congress 1969 (1970), 243—254. [2.4; 3.1; 3.2]
ARRANGEMENTS AND SPREADS Sis)

OTs Arrangements of hyperplanes, Proc. Second Louisiana Conference on Combina-


torics and Graph Theory. Baton Rouge 1971 (to appear). [2.1; 2.7; 331]
H. Guggenheimer
1965 Finite sets on curves and surfaces, Israel J. Math. 3 (1965), 104-112. MR 32
#6326. [4.2]
H. Gupta
1953 Non-concyclic sets of points, Proc. Nat. Inst. Sci. India 19 (1953), 315—316.
MR 15 p. 139. [3.4]
R. K. Guy
1960 A combinatorial problem, Bull. Malayan Math. Soc. 7 (1960), 68—72. [3.4]
1969 “The decline and fall of Zarankiewicz’s theorem”, Proof Techniques in Graph
Theory, edited by F. Harary. Academic Press, New York 1969. MR 40 #7144.
[3.4]
1970 Twenty odd questions in combinatorics, Proc. Second Chapel Hill conference on
Combinatorial Mathematics and its Applications. Chapel Hill, 1970, pp. 209—237.
[3.4] j
1971a “Unsolved combinatorial problems’, Combinatorial Mathematics and its Applica-
tions, edited by D. J. A. Walsh, pp. 121-127. Academic Press, London 1971.
[2.2; 3.4]
1971b The crossing number of the complete graph, (to appear). [3.4]
197ic “Latest results on crossing numbers”, Recent Trends in Graph Theory, edited by
M. Capobianco, J. B. Frechen and M. Krolik. Springer-Verlag, New York 1971.
pp. 143-156. [3.4]
. Guy and F. Harary
1967 On the Mobius ladders, Canad. Math. Bull. 10 (1967), 493—496. MR 37 #98.
[3.4]
. Guy and T. A. Jenkyns
1969 The toroidal crossing number of K,,,, J. Combinat. Theory 6 (1969), 235—
250. MR 37 #5660. [3.4]
. Guy, T. A. Jenkyns and J. Schaer
1968 The toroidal crossing number of the complete graph, J. Combinat. Theory 4
(1968), 376-390. MR 36 #3682. [3.4]
H. Hadwiger
1961 Kleine Studie zur elementaren Stetigkeitsgeometrie, Joer. Deutsch. Math.-Verein.
64 (1951/62), 78-81. MR 25 #5421. [4.1; 4.2]
1971 Ungeloste Probleme Nr. 53, Elem. Math. 26 (1971), 58. [4.2]
H. Hadwiger and H. Debrunner
1955 Ausgewahlte Einzelprobleme der kombinatorischen Geometrie in der Ebene,
Enseign. Math. (2) 1 (1955), 56—89. (French translation by J. Chatelet, En-
seign. Math. (2) 3(1957), 35-70.) MR 17 p. 887. [2.3; 3.4]
1960 Kombinatorische Geometrie in der Ebene, Monogr. de |’Enseign. Math. No. 2,
Université, Genéve 1960. MR 22 #11310. [2.3; 3.4]
100 BRANKO GRUNBAUM

H. Hadwiger, H. Debrunner and V. Klee


1964 Combinatorial geometry in the plane, Holt, Rinehart and Winston, New York
1964. MR 29 #1577. [2.3; 3.4]
H. Hadwiger, H. Debrunner and I. M. Yaglom
1965 Combinatorial geometry of the plane, {In Russian.] Nauka, Moscow 1965.
MR 34 #3428. [2.3; 3.4]
H. Hahn
1908 Uber die Anordunungssdtze der Geometrie, Monatshefte Math. Phys. 19 (1908),
289-303. FM 39 p. 550. [2.5]
M. Hall, Jr.
1967 Combinatorial Theory, Blaisdell, Waltham 1967. MR 37 #80. [2.7]
E. Halsey
1971 Ph. D. Thesis, University of Washington, Seattle 1971. [2.1; 3.1]
P. C. Hammer
1954 Diameters of convex bodies, Proc. Amer. Math. Soc. 5 (1954), 304-306. MR
15 p. 819. [4.1]
1963 Convex curves of constant Minkowski breadth, Proc. Sympos. Pure Math. Vol.
7 (Convexity) (1963), 291-304. MR 25 #5421. [4.2]
P. C. Hammer and A. Sobczyk
1953a Planar line families. 1, Proc. Amer. Math. Soc. 4 (1953), 226—233. MR 14 p.
787. [4.2]
1953b Planar line families. II, Proc. Amer. Math. Soc. 4 (1953), 341—349. MR 15
p. 149. [4.2]
H. Hanani
1951 On the number of straight lines determined by n points, [In Hebrew]. Riveon
Lematematika 5 (1951), 10-11. MR 13 p. 5. [2.2]
1954 On the number of lines and planes determined by d_ points, Scientif. Publ.
Technion Haifa, 6 (1954), 58-63. MR 17 p. 294. [2.2]
F. Harary
1969 Graph Theory, Addison-Wesley, Reading, Mass. 1969. MR 41 #1566. [3.4]
F. Harary and A. Hill
1962 On the number of crossings in a complete graph, Proc. Edinburgh Math. Soc.
(2) 13 (1962-63), 333-338. MR 29 #602. [3.4]
H. Harborth

1971 Uber die Kreuzungszahl vollstandiger, n-geteilter Graphen, Math. Nachr. 48


(1971), 179-188. [3.4]
O. Haupt
1965 Verallgemeinerung zweier Satze tiber interpolatorische Funktionensysteme, Akad.
Wiss. Lit. Mainz Abh. Math. Natur. Kl. 1965, 239-255. MR 33 #7936. [4.4]
E. Heil
1971 Linienfamilien ebener konvexer Bereiche, (to appear). [4.3]
ARRANGEMENTS AND SPREADS 101

W. Heise
1970a Eine Definition des Mobiusraumes ,Manuscripta Math. 2 (1970), 39-47. MR
41 #7509. [4.4]
1970b Eine neue Klasse von Mobius m-Strukturen, Rend. Ist. di Matem. Univ. di
Trieste 2 (1970), 125—128. [4.4]
B. Hesselbach
1933 Uber zwei Vierecksdtze der Kreisgeometrie, Abh. Math. Sem. Univ. Hamburg 9
(1933), 265-271. FM 59 p. 1232. [4.4]
D. Hilbert
1899 Grundlagen der Geometrie, Teubner, Leipzig 1899. FM 30 p. 424. [4.4]
1900 Mathematische Probleme, Nachr. Ges. Wiss. Gottingen, Math.-Phys. K1 1900
(1900), 253—297. (Modified French translation: Sur les problémes futures des
matheématiques. C. R. 2. Congr. Internat. Math. Paris 1900 (1902), 58—114.)
FM 31 pp. 68, 905. [2.5]
D. Hilbert and S. Cohn-Vossen
1932 Anschauliche Geometrie, Grundlehren Math. Wiss. Vol. 37. Springer, Berlin
1932. (English translation: Geometry and the imagination, Chelsea, New York
1952.) FM 58 p. 597; MR 13 p. 76. [2.7]
S. Jendrol and E. Jucovic
1971a On the toroidal analogue of Eberhard’s theorem, (to appear). [2.4]
1971b On a conjecture by B. Griinbaum, (to appear). [2.4]
H. F. Jensen
1971 An upper bound for the rectilinear crossing number of the complete graph, J.
Combinatorial Theory (to appear). [3.4]
R. Jerrard
1961 Inscribed squares in plane curves, Trans. Amer. Math. Soc. 98 (1961), 234—241.
MR 22 #11354. [4.2]
F. John
1937 Polar correspondence with respect to a convex region, Duke Math. J. 3 (1937),
355—369. FM 63 p. 669. [4.4]
C. Jordan
1920 Sur la classification des constellations, C. R. Congr. Internat. Math. Strasbourg
1920, 410—436. FM 48 p. 652. [2.6]
E. Jucovic
1967 Beitrag zur kombinatorischen Inzidenzgeometrie, Acta Math. Acad. Sci. Hung.
18 (1967), 225-259. MR 36 #766. [3.4]
E. Jucovic and M. Trenkler
1971 On 4-valent graphs imbedded in orientable 2-manifolds, (to appear). [2.4]
W. Jung
1937 Untersuchungen tiber symmetrische Geradenkomplexe, Thesis, Univ. Berlin 1937.
47 pp. FM 63 p. 599. [2.3]
102 BRANKO GRUNBAUM

W. Jung and E. Melchior


1936 Symmetrische Geradenkomplexe. Ein Beitrag zur Theorie der Konfigurationen,
Deutsche Math. 1 (1936), 239-255. FM 62 p. 735. [2.3]
P. C. Kainen
1968 Ona problem of Erdos, J. Combinat. Theory 5 (1968), 374—377. MR 38 #72.
[3.4]
Nn. Kakeya
1916 On the inscribed rectangles of a closed convex curve, Tohoku Math. J. 9 (1916),
163-166. FM 46 p. 1116. [4.2]
S. Kantor
1881 Die Configurationen (3, 3), S.-Ber. Math.-Nat. Kl. Akad. Wiss. Wien 84 (1881),
1291-1314. FM 13 p. 460. [3.1]
F. Karteszi
1963 Alcuni problemi della geometria d’incidenza, Confer. Sem. Math. Univ. Bari No.
88 (1963), 14 pp. FM 31 #3926. [2.3]
1964 Intorno a punti allineati di certi reticoli circolari, Rend. Sem. Matem. Messina
9 (1964/65), 1-12. MR 40 #7923. [2.3]
= . M. Kelly and W. O. J. Moser
1958 On the number of ordinary lines determined by n points, Canad. J. Math 10
(1958), 210-219. MR 20 #3494. [2.2; 2.3; 2.6]
Ic. M. Kelly and R. Rottenberg
1971 Simple points in pseudoline arrangements, (to appear). [3.2; 3.3]
es). Kivikoski
1925 Ein einfacher Beweis des Jordan'schen Kurvensatzes nebst der projektiven Ver-
allgemeinerung des Satzes, Soc. Scient. Fenn., Comm. Phys.-Math. 2, 21 (1925),
12 pp. FM 51 p. 459. [3.1]
R. Klee
1938 Uber die einfachen Konfigurationen der euklidischen und der projektiven Ebene,
Focken und Oltmanns, Dresden 1938. FM 64 p. 1296. [2.1; 3.1]
D. J. Kleitman
1970 The crossing number of K, ,, J. Combinat. Theory 9 (1970), 315—323. [3.4]
M. Kneser
1949 Eibereiche mit geraden Schwerlinien, Math.-Phys. Semesterber. 1 (1949), 97—
98. MR 11 p. 386. [4.2]
fer Kojima
1919 On characteristic properties of the conic and quadric, Tohoku Sci. Reports 8
(1919), 67-78. FM 47 p. 682. [4.4]
K. Kommerell
1941 Die Pascalsche Konfiguration 9,, Deutsche Math. 6 (1941), 16-32. MR3 p.
1'79%[2.7]
ARRANGEMENTS AND SPREADS 103

A. Kosinski
1957 On involutions and families of compacta, Bull. Acad. Polon. Sci. Cl. II, 5 (1957),
1055—1060. MR 21 #2244. [4.1]
1958 On a problem of Steinhaus, Fund. Math. 46 (1958), 47-59. MR 24 #A2379.
[4.1]
K. Koutsky and V. Polak
1960 Note on the omittable points in complete systems of points and straight lines in
the plane, [In Czech. Russian and English summaries.] Casopis Pést. Mat. 85
(1960), 60-69. MR 24 #A448. [2.3]
T. Kubota
1939 Ein neuer Aufbau der euklidischen Geometrie in der affinen Ebene, Monatsch.
Math. Phys. 48 (1939), 96-102. FM 65 p. 638. [4.2]
G. Landsberg
1911 Beitrage zur Topologie geschlossener Kurven mit Knotenpunkten und zur Kron-
eckerchen Charakteristikentheorie, Math. Ann. 70 (1911), 563—579. FM 42
p. 209: [34]
D. W. Lang
1955 The dual of a well known theorem, Math. Gazette 39 (1955), 314. [2.3]
R. Lauffer
1953a Zur Topologie der Konfiguration von Desargues. 1, Math. Nachr. 9 (1953), 235—
240. MR 14 p. 1008. [2.7]
1953b Zur Topologie der Konfiguration von Desargues. 11, Math. Nachr. 10 (1953),
179-180. MR 15 p. 339. [2.7]
B. Leclerc and B. Monjardet
1969 Representation graphique d'un graphe, Math. et Sci. humaines 26 (1969), 51—
57. [3.4]
N. J. Lennes
1911 Theorems on the simple finite polygon and polyhedron, Amer. J. Math. 33
(1911), 37-62. FM 42 p. 511. [2.5]
F. Levi
1922 Die gegenseitige Lage von 5 und 6 Punkten in der projektiven Ebene, Ber. Ver.
Sachs. Ges. Wiss. Leipzig. Math. Phys. Kl. 74 (1922), 34-39. FM 48 p. 721.
227]
1926 Die Teilung der projektiven Ebene durch Gerade oder Pseudogerade, Ber. Math.-
Phys. KI. Sachs. Akad. Wiss. Leipzig 78 (1926), 256—267. FM 52 p. 575.
12.47-351; 322; 433]
1929 Geometrische Konfigurationen, Hirzel, Leipzig 1929. FM 55 p. 351. [2.6; 2.7]
B. Lindstrom
1971 On the realization of convex polytopes, Euler’s formula and Mobius functions,
Aequat. Math. (to appear). [2.5]
L. Locher-Ernst
1951 Polarentheorie der Eilinien, Elem. Math. 6 (1951), 1-7. MR 12 p. 436. [4.4]
104 BRANKO GRUNBAUM

S. Loyd
1914 Cyclopedia of puzzles, Lamb, New York 1914. [2.3]
S. Mac Lane
1937 A combinatorial condition for planar graphs, Fund. Math. 28 (1937), 22—32.
FM 63 p. 548. [3.4]
K. Maier
1939 Die Desarguessche Konfiguration, Deutsche Math. 4 (1939), 591-641. FM 65
p. 664. [2.7]
A. Marchaud
1948 Sur les ovales, Ann. Soc. Polon. Math. 21 (1948), 324-331. [4.4]
1959 Un théoréme sur les corps convexes, Ann. Sci. Ecole Norm. Sup. (3) 76 (1959),
283—304. [4.4]
M. L. Marx
1969 The Gauss realizability problem, Proc. Amer. Math. Soc. 22 (1969), 610—613.
MR 39 #6297. [3.4]
Yu. V. Matiyasevité
1971 Diophantine representation of enumerable predicates, Izvestia Akad. Nauk SSSR
35 (1971), 3-30. [2.5]
K. O. May
1954 The use of condensed graphs in analytic geometry, Amer. Math. Monthly 61
(1954), 31-32. MR 15 p. 460. [4.4]
P. McMullen
1969 Linearly stable polytopes, Canad. J. Math. 21 (1969), 1427-1431. MR 40
#6364. [2.7]
1971 On zonotopes, Trans. Amer. Math. Soc. 159 (1971). [2.1; 2.7]
E. Melchior
1937 Untersuchungen iiber ein Problem aus der Theorie der Konfigurationen, Schr.
math. Sem. Inst. angew. Math. Univ. Berlin 3 (1937), 181-206. FM 63 p.
1199. [2.3]
1940 Uber Vielseite der projektiven Ebene, Deutsche Math. 5 (1940), 461—475.
MRv3 peli [2cbe 2.37 247]
V. V. Menon
1966 A theorem on partitions of mass-distributions, Pacif. J. Math. 16 (1966), 133—
137. MR 32 #4602. [4.2]
A. F. Mobius
1827 Der barycentrischer Calcul, J. A. Barth, Leipzig 1827. (Ges. Werke, Vol. 1,
pp. 1—388. S. Hirzel, Leipzig 1885.) [2.6]
J. W. Moon
1965 On the distribution of crossings in random complete graphs, SIAM Journal 13
(1965), 506-510. MR 31 #3357. [3.4]
ARRANGEMENTS AND SPREADS 105

T. More, Jr.
1959 On the construction of Venn diagrams, J. Symb. Logic 24 (1959), 303—304.
MR 23 #A3683. [3.4]
L. Moser
1952 Problem 130, Mathematics Mag. (1952). [3.4]
1963 Problem P. 65, Canad. Math. Bull. 6 (1963), 113. Solution, ibid., 14 (1971),
129-130. [2.4]
1964 Problem 77, Canad. Math. Bull. 7 (1964), 137. Solution by B. Griinbaum, ibid.,
477-478. [2.4]
T. S. Motzkin
1951 The lines and planes connecting the points of a finite set, Trans. Amer. Math.
Soc. 70 (1951), 451-464. MR 12 p. 849. [2.3]
1957 Types of dissections, Abstract 112t. Bull. Amer. Math. Soc. 63 (1957), 35—36.
[2.7]
1967a Nonmixed connecting lines, Abstract 67T—605. Notices Amer. Math. Soc. 14
(1967), 837. [2.7]
1967b Combinatorial realization of centrally symmetric convex polyhedra, Research
Problem 3—8. J. Combinat. Theory 3 (1967), 411. [3.1]
1967c “‘Signs of minors’, Inequalities, edited by O. Shisha, pp. 225—240. Academic
Press, New York 1967. MR 36 #6432. [2.7]
T. S. Motzkin and M. O. Rabin
1971 Nonmixed connecting lines, J. Combinat. Theory (to appear). [2.7]
F.R. Moulton
1902 A simple non-Desarguesian plane geometry, Trans. Amer. Math. Soc. 3 (1902),
192-195. FM 33 p. 497. [4.4]
U.S. R. Murty
1969 “Sylvester matroids,”’ Recent progress in combinatorics, edited by W. T. Tutte,
283—286. MR 42 #1685. Academic Press, New York 1969. [2.3]
1970 Matroids with Sylvester property, Aequat. Math. 4 (1970), 44—S0. [2.3]
1971 How many magic configurations are there? Amer. Math. Monthly 78 (1971):
1000-1002. [2.7]
J. R. Musselman
1931 The planar imprimitive group of order 216, Amer. J. Math. 53 (1931), 333-342.
FM 57 p. 805. [2.7]
1936 A classification of planar sixpoints. Tohoku Math. J. 42 (1936), 114-117.
FM 62 p. 734. [2.7]
J. v. Sz. Nagy
1927 Uber ein topologisches Problem von Gauss, Math. Z. 26 (1927), 579-592. FM
53 p. 550. [3.4]
S. Nakajima
1928 Kilinien mit geraden Schwerlinien, Japan. J. Math. 5 (1928), 81-84. FM 54 p.
799. [4.2]
106 BRANKO GRUNBAUM

J. Novak
1959 Anwendungen der Kombinatorik auf das Studium ebener Konfigurationen
(12,, 163), (Czech., Russian and German Summaries). Casopis Pést. Mat. 84
(1959), 257—282. MR 24 #A447. [2.7]
C. S. Ogilvy
1950 Square inscribed in arbitrary simple closed curve, Solution of Problem 4325.
Amer. Math. Monthly 57 (1950), 423—424. [4.2]
E. Piegat
1963 O srednicach figur wypuktych plaskich, Roczn. Polsk. Towarz. Mat. Ser. 2, 7
(1963), 51-56. [4.2]
H. Rademacher and O. Toeplitz
1930 Von Zahlen und Figuren, Springer, Berlin 1930. FM 56 p. 62. [3.4]
K. Reidemeister
1924 Eine Kennzeichnung der Kugel nach W. Blaschke, J. Reine Angew. Math. 154
(1924), 8-14. FM SO p. 488. [4.4]
1932 Knotentheorie, Ergebn. Math. u. Grenzgeb. Vol. 1. Springer, Berlin 1932.
FM 58 p. 1202. [3.4]
G. Ringel
1956 Teilungen der Ebene durch Geraden oder topologische Geraden, Math. Z. 64
(1955), 79-102. MR 17 p. 651. [3.1]
1957 Uber Geraden in allgemeiner Lage, Elem. Math. 12 (1957), 75—82. MR 19 p.
1632 [254
1964 “Extremal problems in the theory of graphs,” Theory of Graphs and its Applica-
tions (Proc. Sympos. Smolenice, 1963), pp. 85—90. Publ. House Czechoslovak
Acad. Sci., Prague, 1964. MR 31 #4021. [3.4]
S. Roberts
1889 On the figures formed by the intercepts of a system of straight lines in a plane,
and on analogous relations in space of three dimensions, Proc. London Math.
Soc. 19 (1889), 405—422. FM 20 p. 592. [2.4]
A. Robinson
1963 Introduction to model theory and to the metamathematics of algebra, North-
Holland, Amsterdam 1963. MR 27 #3533. [2.5]
H. A. Robinson
1945 A problem of regions, Amer. Math. Monthly 52 (1945), 33-34. MR’‘6 p. 215.
[3.4]
R. R. Rottenberg
1971 On finite sets of points in P*, Israel J. Math.10 (1971), 160—171. [2.6]
H. J. Ryser
1963 Combinatorial Mathematics, Carus Math. Monograph No. 14. Wiley, New York
1963, MRi27 podk(2.7)
1968 An extension of a theorem of de Bruijn and Erdés on combinatorial designs,
J. Algebra 10 (1968), 246-261. MR 37 #5103. [2.2]
ARRANGEMENTS AND SPREADS 107

T. L. Saaty
1967 Two theorems on the minimum number of intersections for complete graphs, J.
Combinat. Theory 2 (1967), 571-584. MR 35 #2796. [3.4]
1969 Symmetry and the crossing number for complete graphs, J. of Research Nat. Bur.
Stand. 73B (1969), 177-186. MR 41 #6715. [3.4]
1971 On polynomials and crossing numbers of complete graphs, J. Combinat. Theory
10 (1971), 183-184. [3.4]
H. Salzmann
1955 Uber den Zusammenhang in topologischen projektiven Ebenen, Math. Z. 61
(1955), 489-494. MR 16 p. 845. [4.4]
1967 Topological planes, Advances in Math. 2 (1967), 1-60. MR 36 #3201. [4.4]
J. J. Schaffer
1968 Symmetric curves, hexagons, and the girth of spheres in dimension 3, Israel J.
Math. 6 (1968), 202—205. MR 38 #1610. [4.2]
L. Schlafli
1852 Theorie der vielfachen Kontinuitat, Posthumously published in: Neue Denschriften
der schweiz. Ges. Naturwiss. 38 (1901), iv + 239 pp. (= Ges. Math. Abh., Vol. 1,
167-387. Birkhauser, Basel 1950.) [2.1]
H. Schroter
1889 Ueber die Bildungsweise und geometrische Construction der Configurationen 10,,
Gottingen Nachr. 1889, 193-236. FM 21 p. 536. [3.1]
B. Segre
1964 Teoria di Galois, fibrazioni proiettive e geometrie non desarguesiane, Ann. Mat.
Pura Appl. 64 (1964), 1-76. MR 29 #6370. [4.1]
A. Seidenberg
1954 A new decision method for elementary algebra, Ann. Math. 60 (1954), 365-374.
MR 16 p. 209. [2.5]
J.-P. Serre
1966 Problem 5359, Amer. Math. Monthly 73 (1966), 89. [2.3]
M. Sholander
1953 Plane geometries from convex plates, Pacif. J. Math. 3 (1953), 667-671. MR
15 p. 246. [4.4]
G. J. Simmons
1971 A maximal 2-arrangement of sixteen lines in the projective plane, (to appear).
[2.4]
L. A. Skornyakov
1954 Topological projective planes, (in Russian), Trudy Moskov. Mat. Ob&8&. 3 (1954),
347-373. MR 16 p. 60. [4.4]
H. A. Smirnova
1956 The problem of g-circles, (in Russian), Mat. Sbornik, N. S. 38 (81) (1956),
397-399. MR 18 p. 227. [3.4]
108 BRANKO GRUNBAUM

T. J. Smith
1961 Line families in the plane, University of Wisconsin Thesis, 1961. [4.2]
L. G. Snirelman
1929 On certain geometric properties of closed curves, (in Russian), Sbornik rabot mat.
razd. sekc. estestv. tocnyh nauk Komakademii, Moscow 1929. Reprinted in
Uspehi Mat. Nauk 10 (1944), 34-44. MR 7 p. 35. [4.2]
A. Sobczyk
1956 Simple line families, Pacif. J. Math. 6 (1956), 541-552. MR 19 p. S6. [4.1]
G. K. C. von Staudt
1847 Geometrie der Lage, Nurnberg, 1847. [2.1]
S. Stein
1954 Families of curves, Proc. Amer. Math. Soc. 5 (1954), 745—747. MR 16 p. 157.
[4.1]
R. Steinberg
1944 Three point collinearity, Solution of Problem 4065, Amer. Math. Monthly 51
(1944), 169-171. [2.2; 2.3]
J. Steiner
1826 Einige Gesetze tiber die Theilung der Ebene und des Raumes, J. Reine Angew.
Math. 1 (1826), 349-364. [2.1; 3.4]
H. Steinhaus
1955 Quelques applications des principes topologiques a la géométrie des corps con-
vexes, Fund. Math. 41 (1955), 284—290. MR 16 p. 849. [4.2]
1957 On chords of convex curves, Bull. Acad. Polon. Sci. Cl. HI, 5 (1957), 595—597.
MR 19 p. 573. [4.2]
E. Steinitz
1906 Uber die Eulersche Polyederrelationen, Arch. Math. Phys. (3) 11 (1906), 86—88.
FM 37 p. 500. [2.2]
1922 Polyeder und Raumeinteilungen, Enzykl. math. Wiss. Vol. 3 (Geometrie), Part
3AB12, pp. 1-139. (1922). [3.2; 3.4]
1923 Uber die Maximalzahl der Doppelpunkte bei ebenen Polygonen von gerader
Seitenzahl, Math. Z. 17 (1923), 116-129. FM 49 p. 410. [3.4]
E. Steinitz and H. Rademacher
1934 Vorlesungen tiber die Theorie der Polyeder, Springer, Berlin 1934. FM 60 p.
497. [24332]
E. Stenfors
1923 Der Jordansche Kurvensatz in der projektiven Ebene und die v. Staudt'schen
Schnittpunktsdtze, Soc. Scient. Fenn., Comm. Phys.-Math. Vol. 2, No. 5 (1923),
6 pp. FM 49 p, 404. [3.1]
A. H. Stone and J. W. Tukey
1942 Generalized “sandwich” theorems, Duke Math. J. 9 (1942), 356—359. MR4
p. 75. [4.2]
ARRANGEMENTS AND SPREADS 109

K. Strambach
1970 Spharische Kreisebenen, Math. Z. 113 (1970), 266—292. MR 41 #7510. [4.4]
W. Siiss, U. Viet and K. H. Berger
1960 “Konvexe Figuren,” Grundziige der Mathematik, edited by H. Behnke, F. Bach-
mann, K. Fladt and W. Stiss, pp. 510-529. Vanderhoeck and Ruprecht, Gottin-
gen 1960. [4.2]
J. J. Sylvester
1867 Problem 2473, Mathematical questions and solutions from the Educational Times,
8 (1867), 104-107. [2.1; 2.3]
1886 Problem 2572, Mathematical questions and solutions from the Educational Times,
45 (1886), 127-128. (Reprinted in Vol. 59 (1893), 133-134.) [2.3]
1893 Problem 11851, Mathematical questions and solutions from the Educational
Times, 59 (1893), 98—99. [2.3; 4.3]
POG. Lait
1877 Some elementary properties of closed plane curves, Messenger Math. (2) 6
(1877), 132—133. (See also: Scientific Papers, Vol. 1., Cambridge Univ. Press
1898, pp. 273-347.) FM 9 p. 393. [3.4]
A. Tarski
1951 A decision method for elementary algebra and geometry, Univ. of California
Press, Berkeley 1951. MR 10 p. 499. [2.5]
C. J. Titus
1960 A theory of normal curves and some applications, Pacif. J. Math. 10 (1960),
1083-1096. MR 22 #5014. [3.4]
1961 The combinatorial topology of analytic functions on the boundary of a disc,
Acta Math. 106 (1961), 45—64. MR 23 #A339. [3.4]
O. Toeplitz
1911 Verh. schweiz. Naturforsch. Gesellschaft Solothurn, 1911, p. 197. [4.2]
L. Tornheim
1950 On n-parameter families of functions and associated convex functions, Trans.
Amer. Math. Soc. 69 (1950), 457-467. MR 12 p. 395. [4.4]
G. Trevisan
1949 Una condizione di allineamento per gli insiemi finiti di punti del piano euclideo,
Rend. Seminar. Mat. Univ. Padova 18 (1949), 258-261. MR 11 p. 383. [2.3]
L. R. Treybig
1968 A characterization of the double point structure of the projection of a poly-
gonal knot in regular position, Trans. Amer. Math. Soc. 130 (1968), 223-247.
MR 36 #878. [3.4]
W. T. Tutte
1970 Toward a theory of crossing numbers, J. Combinat. Theory 8 (1970), 45—S3.
MR 41 #6720. [3.4]
K. Urbanik
1955 Solution du probléme posé parP. Turan, Colloq. Math. 3 (1955), 200-201. [3.4]
110 BRANKO GRUNBAUM

O. Veblen and J. W. Young


1918 Projective geometry. II, Ginn, Boston 1918. (Reprint: Blaisdell, New York 1965).
FM 47 p. 582. [2.1; 2.2]
U. Viet
1956 Umkehrung eines Satzes von H. Brunn iiber Mittelpunktseibereiche, Math.-Phys.
Semesterber. 5 (1956), 141-142. MR 18 p. 667. [4.2]
I. Vineze
1952 Uber die Schwerlinie einer geschlossenen, konvexen Kurve, (In Hungarian; sum-
maries in Russian and German), C. R. Premier Congr. Math. Hungrois, 1950,
679—687. Akademiai Kiado, Budapest 1952. MR 15 p. 247. [4.4]
B. L. van der Waerden and L. J. Smid
1934 Eine Axiomatik der Kreisgeometrie und der Laguerregeometrie, Math. Ann. 110
(1934), 753-776. FM 61 p. 600. [4.4]
H. S. White
1932 The plane figure of seven real lines, Bull. Amer. Math. Soc. 38 (1932), 59—65.
FM 58 p. 675. [2.1]
H. Whitney
1937 On regular closed curves in the plane, Compositio Math. 4 (1937), 276—284.
FM 63 p. 647. [3.4]
C. Wiener
1864 Uber Vielecke und Vielflache, Leipzig 1864. [2.1; 3.4]
V. C. Williams
1968 A proof of Sylvester’s theorem on collinear points, Amer. Math. Monthly 75
(1968), 980—982. [2.3]
P. de Witte
1966a Combinatorial properties of finite linear spaces, Bull. Soc. Math. Belg. 18 (1966),
133-141. MR 34 #1913. [2.2]
1966b A new property of non-trivial finite linear spaces, Bull. Soc. Math. Belg. 18
(1966), 430-438. MR 35 #5345. [2.2]
D. R. Woodall
1971 “Thrackles and deadlock,’’ Combinatorial Mathematics and its Applications,
edited by D. J. A. Welsh, pp. 335—347. Academic Press, London 1971. [3.4]
A. M. Yaglom and I. M. Yaglom
1954 Non-elementary problems in elementary presentation, (in Russian), Biblioteka
Mat. Kru%ka, Moscow 1954. MR 17 p. 18. [2.3]
M. Zacharias
1941 Untersuchungen iiber ebene Konfigurationen (12,, 16), Deutsche Mathem. 6
(1941), 147-170. MR 8 p. 219. [2.7]
1950 Uber die harmonisch gekoppelten Hesseschen Konfigurationen (12,, 163) und
gewisse in ihnen enthaltene Konfigurationen (15,), Math. Nachr. 3 (1950),
243-256. MR12 p, 5230/27]
ARRANGEMENTS AND SPREADS 111

J. Zaks
1971 On realizing symmetric 3-polytopes, Israel J. Math. 10 (1971), 244—251. [3.1]
T. Zamfirescu
1967a Sur les familles continues de courbes. 1, Atti Accad. Naz. Lincei Rend. Cl. Sci.
Fis. Mat. Natur. (8) 42 (1967), 771-774. MR 38 #3843. [4.1]
1967b Sur les familles continues de courbes. II, Atti Accad. Naz. Lincei Rend. Cl. Sci.
Fis. Mat. Natur. (8) 43 (1967), 13-17. MR 38 #3843. [4.1]
1967c Théoréme dual concernant les familles continues de courbes, Bull. Acad. Roy.
Belg. (5) 53 (1967), 1385-1391. MR 39 #7505. [4.1]
1968a Sur les familles continues de courbes. III, Atti Accad. Naz. Lincei, Rend. Cl. Sci.
Fis. Mat. Natur. (8) 44 (1968), 639-642. MR 40 #2034a. [4.1]
1968b Sur les familles continues de courbes. IV, Atti Accad. Naz. Lincei Rend. Cl.
Sci. Fis. Mat. Natur. (8) 44 (1968), 753-758. MR 40 #2034b. [4.1]
1969a On planar continuous families of curves, Canad. J. Math. 21 (1969), 513—530.
MR 39 #7506. [4.1]
1969b Les courbes fermées doubles sans points triple associées a une famille continue,
Israel J. Math. 7 (1969), 69-89. MR 39 #2035. [4.1]
K. Zarankiewicz
1954 Ona problem of P. Turan concerning graphs, Fund. Math. 41 (1954), 137-145.
MR 16 p. 156. [3.4]
1959 Bisection of plane convex sets by lines, (in Polish), Wiadom. Mat. (2) 2 (1959),
228—234. MR 22 #7055. [4.2]
H. Zeitler
1970 Modelle der euklidischen und nichteuklidischen Geometrie, Praxis Math. 12
(1970), 33-38. [4.4]
K. Zindler
1889 Zur Theorie der Netze und Configurationen, S.-B. Math.-Nat. Cl. Akad. Wiss.
Wien 98 Part Ila (1889), 499-519. FM 21 p. 535. [2.6]
1921 Uber konvexe Gebilde. 11, Monatsh. Math. Phys. 31 (1921), 25-57. FM 48 p.
833. [4.2]
1922 Uber konvexe Gebilde. Il, Monatsh. Math. Phys. 32 (1922), 107-138. FM 48
p. 833. [4.2]
A. C. Zitronenbaum
1959 Bisecting an area and its boundary, Note 2845, Math. Gazette 43 (1959), 130—
131. [4.2]
I? BRANKO GRUNBAUM

Notes added in proof (March 1972)

1. (Page 3) Since the completion of the manuscript certain results have come to my
attention which deserve being mentioned together with those discussed in the text. The fol-
lowing notes deal with such results and references.

2. (Page 17) See also V. Klee (The use of research probiems in high school geometry,
Educational Studies in Math. 3 (1971), 482—489).
3. (Page 21) The problem of determining t,(n) was posed also by Alauda (Question
1664, Interméd. Math. 6 (1899), 245 and 18 (1911), 196—197); the values he indicates for
t;(9) and 1¢,(10) are too small.
4. (Page 21) For a relation between n, f, andthe numbers ¢, see E. B. Elliott,
Question 6362, Math. Questions from the Educ. Times, 34 (1881), p. 120 (solutions by W.B.
Grove, E. Rutter and others).
5. (Page 37) Certain related but easy questions were considered by E. Lucas (Récréa-
tions Mathématiques, vol. 4. Gauthier-Villars, Paris 1894, pp. 155—194; FM 25, p. 336) and
in a problem posed by E.-N. Barisien (Question 3901, Interméd. Math. 18 (1911), 171) and
solved by Welsch (ibid. 19 (1912), 18—21). Much harder are the problems concerning the
different ‘‘aspects” of finite sets of points; that is the question what permutations of the
given points correspond to the circular orders in which the points are visible from other
points of the (Euclidean) plane. There appear to have been no advances in this question
since the results and discussions by C.-A. Laisant (Régions du plan et de l’espace. Assoc.
Frang. Avanc. Sci. 1881, pp. 71-76, and Remarques sur la théorie des régions et des as-
pects, Bull. Soc. Math. France 10 (1881—82), 52—55; FM 14, p. 151), R. Perrin (Sur le
probléme des aspects, Bull Soc. Math. France 10 (1881—82), 103—127; FM 14, 152, and
Question 27, Interméd. Math. 1 (1894), 7—8) and A. Sainte-Lague (Géométrie de situation
et jeux, Mémorial Sci. Math. vol. 41. Gauthier-Villars, Paris 1929, pp. 3—6; FM 55, p. 974).

6. (Page 37) Two arrangements were inadvertently omitted from this list. They may
be obtained by adding to the marked points on the last arrangement of Figure 2.24 one of
the vertices, or two “‘neighboring”’ vertices, of the ‘inner pentagon”.
7. (Page 37) Let C, and C, be two cell complex decompositions of the projective
plane P. We shall say that C, and C, are piecewise projective provided there exists a
homeomorphism 6 of P onto itself, such that for each cell C of C, the set 0(C) isa
cell of C,, and the restriction of @ to C coincides with the restriction to C of a pro-
jective transformation. A remarkable result of E. Steinitz (Uber ein merkwuirdiges Polyeder
von einseitiger Gesamtoberflache, J. Reine Angew. Math. 130 (1905), 281-307; FM 37, p.
500) may be reformulated as follows: If C is a cell complex decomposition of P which
is piecewise projective with the simple arrangement of four lines, then C itself is a simple
arrangement of four lines. It may be conjectured that every cell complex decomposition of
P piecewise projective with a simple arrangement of lines is itself a simple arrangement of
lines.
ARRANGEMENTS AND SPREADS 113

8. (Page 45) The ten non-isomorphic configurations 10, of pseudolines form the
basis of the attractive game Configurations: Number Puzzles and Patterns for all Ages by H.L.
Dorwart (WFF ’N PROOF, New Haven, 1967).

9. (Page 60) There also exist simplicial arrangements of 12 curves with Tounocsane
of 13 curves with f, = 84.

10. (Page 68) In Figure 3.35 one could add the symbol for a simplicial arrangement
of 13 curves with t, = 21.

11. (Page 68) W. Meyer (On ordinary points in arrangements, to appear) has proved
that t, 2n/2 for every digon-free arrangement of n curves.
12. (Page 68) The question of finding upper bounds for w(C) in Apollonian
arrangements of n circles has been proposed by G. de Rocquigny (Question 1179, Interméd.
Math. 4 (1897), p. 267 and 15 (1908), p. 169).
13. (Page 71) F. Dumont (Question 1389, Interméd, Math. 5 (1898), p. 246) asked
about the maximal number of regions into which the plane may be decomposed by n co-
nics; this obviously corresponds to a special case of k = 2. The answer to Dumont’s ques-
tion given by P. Hendle (ibid. 6 (1899), p. 137) allows pairs of straight lines; if only ellipses
are allowed the answer becomes 2(n* —1n + 1), and is valid for all arrangements of simple
closed curves in the Euclidean plane, each pair of which intersect in at most 4 points.
14. (Page 72) A slightly different code, in some respects simpler than Gauss’ code,
was described by P. G. Tait (Listing’s Topologie, Philos. Mag. (5) 17 (1884), 30—46 =
Sci. Papers, vol. 2, pp. 85—98). Proceeding along C we assign the symbols a, b, c, d, etc.
to the first, third, fifth, seventh, etc. vertex, till all vertices have names; then we go around
C once more, reading off the symbols we find on the second, fourth, sixth, etc. vertex. For
example, in the selfintersection pattern of Figure 3.37, the letters a, b, c, d, e, f, g could
stand for the numbers 1, 3,5, 2, 4, 7, 6; the Tait code would be deagfobe.

15. (Page 73) In this context see also A. Sainte-Lague (Les réseaux (ou graphes),
Mémorial Sci. Math. vol. 18. Gauthier-Villars, Paris 1926, pp. 41-42; FM 52, p. 576)
and the works of E. Kronecker, H. Weber, and others quoted there.

16. (Page 73) A much earlier solution of the characterization problem was given,
together with many other results, in an apparently forgotten paper by M. Dehn (Uber kom-
binatorische Topologie, Acta Math. 67 (1936), 123—168; FM 62, p. 656-658).

17. (Page 73) Even if multiple-intersection vertices are allowed, the map determined
by each selfintersection pattern is 2-colorable. S. de la Campa (Question 1451, Interméd.
Math. 6 (1899), 29-30) posed the problem of determining the possible partitions of the
f, cells on the selfintersection pattern into the two color-classes for preassigned numbers
of vertices of various multiplicities (t,j): The problem is still open, as is also the question
what sequences (t,, ty, te, °°*, t>,) correspond to selfintersection patterns.

18. (Page 74) Another paper dealing with v,,(K,,) is R. K. Guy, Sequences associ-
ated with a problem of Turan and other problems. Proc. Balatonfured Combinatorics Conf.
114 BRANKO GRUNBAUM, 3 2

1969, Vol. 4 (1970), 553-569. For a variant of the rectilinear crossing numbers see M. E.
Watkins, A special crossing number for bipartite graphs: A research problem. Internat. Conf.
on Combinatorial Math. 1970, Ann. New York Acad. Sci. 175 (1970), 405—410; MR 42
#120.
19. (Page 80) H. Debrunner (Orthogonale Dreibeine in richtungsvollstandigen, stetigen
Geradenscharen des R*, Comment. Math. Helv. 37 (1962/63), 36—43; MR 26 #6833) has
proved the following conjecture of H. Hadwiger (Ungeloste Probleme Nr. 39, Elem. Math.
16 (1961), 30—31; Nachtrag, ibid. 18 (1963), 85): Every spread of lines in E>? contains
three mutually orthogonal lines with a common point. Debrunner also gives an example of
a spread of lines in E* which contains only one triple ‘of such lines.
20. (Page 83) The first mention of the planar “sandwich theorem” appears to be in
F. Levi’s paper Die Drittelungskurve (Math. Z. 31 (1930), 339-345; FM 55, p. 434) in
which the result is attributed to an oral communication from L. Neder. The main mass-
partition theorem of Levi’s paper was independently rediscovered and generalized by K. Kura-
towski and H. Steinhaus (Une application géométrique du théoréme de Brouwer sur les
points invariants, Bull. Acad. Polon. Sci., Cl. 3, 1 (1953), 83-86; MR 15, p. 336) and by
K. Borsuk (An application of the theorem on antipodes to the measure theory, Bull. Acad.
Polon. Sci., Cl. 3, 1 (1953), 87-90; MR 15, p. 204).
21. (Page 83) It is not known whether in any of the spreads of “‘remarkable chords ”

mentioned on page 81, or in the spread of midcurves, the exceptional curve allowed by
Theorem 4.2 actually exists.

22. (Page 84) Midcurves may be generalized to convex bodies K in E£”. For each
direction u, the set of centroids of the sets KH, where H varies over all hyperplanes
with normal u, forms the midcurve corresponding to u. (For some properties of such
curves see, for example, W. Blaschke, Uber affine Geometrie 1X: Verschiedene Bemerkungen
und Aufgaben. Ber. Verh. sachs. Ges. Wiss. Leipzig. Math.-Phys. KI. 69 (1917), 412—420;
T. Bonnesen and W. Fenchel, Theorie der konvexen Korper; Erbegnisse der Math. vol. 3,
Springer, Berlin 1934; reprint Chelsea, New York 1948; pp. 10—13, and the references
given there.) Zindler [1922] proved that F,(L)#@ for the spread L of midcurves of
each K C £3, and Steinhaus [1955] strengthened this to the assertion that either
Fy(L)#@ or card F; = 8. Theorem 4.13 implies F,(L)#@ for the spread L of mid-
curves of every convex body K C E”.
23. (Page 84) F. Levi (Die Drittelungskurve, Math. Z. 31 (1930), 339-345; FM 55,
p. 434) proves the existence of three area bisectors having a common point and enclosing
equal angles; his continuity proof applies also to any other preassigned angles between the
bisectors.
*06-DI
— I-130+

You might also like