0% found this document useful (0 votes)
267 views80 pages

Lectures On Discrete Geometry

This document is an excerpt from a textbook on discrete geometry. It begins by reviewing basic geometric concepts like linear and affine subspaces. An affine subspace is defined as a translation of a linear subspace by a vector. The affine hull of a set is the intersection of all affine subspaces containing the set. An affine combination of points is a weighted average where the weights sum to 1. The document then discusses general position and convex sets.

Uploaded by

Tanuj04
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
267 views80 pages

Lectures On Discrete Geometry

This document is an excerpt from a textbook on discrete geometry. It begins by reviewing basic geometric concepts like linear and affine subspaces. An affine subspace is defined as a translation of a linear subspace by a vector. The affine hull of a set is the intersection of all affine subspaces containing the set. An affine combination of points is a weighted average where the weights sum to 1. The document then discusses general position and convex sets.

Uploaded by

Tanuj04
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 80

Jir Matousek

Lectures on Discrete Geometry


Excerpt

Springer-Verlag
Berlin Heidelberg NewYork
London Paris Tokyo
Hong Kong Barcelona
Budapest
Contents

Notation and Terminology iii

1 Convexity 1
1.1 Linear and Affine Subspaces, General Position . . . . . . . . . . . . . 1
1.2 Convex Sets, Convex Combinations, Separation . . . . . . . . . . . . 5
1.3 Radons Lemma and Hellys Theorem . . . . . . . . . . . . . . . . . . . . . 9
1.4 Centerpoint and Ham Sandwich . . . . . . . . . . . . . . . . . . . . . . . . . . 14

5 Convex Polytopes 17
5.1 Geometric Duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.2 H-Polytopes and V -Polytopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.3 Faces of a Convex Polytope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.4 Many Faces: The Cyclic Polytopes . . . . . . . . . . . . . . . . . . . . . . . . 32
5.5 The Upper Bound Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.7 Voronoi Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

6 Number of Faces in Arrangements 51


6.1 Arrangements of Hyperplanes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.2 Arrangements of Other Geometric Objects . . . . . . . . . . . . . . . . . 55

Bibliography 67
Notation and Terminology

This section summarizes rather standard things, and it is mainly for reference.
More special notions are introduced gradually throughout the book. In order
to facilitate independent reading of various parts, some of the definitions are
even repeated several times.
If X is a set, |X| denotes the number of elements (cardinality) of X. If X
is a multiset, in which some elements may be repeated, then |X| counts each
element with its multiplicity.
The very slowly growing function log x is defined by log x = 0 for x 1
and log x = 1 + log (log2 x) for x > 1.
For a real number x, bxc denotes the largest integer less than or equal
to x, and dxe means the smallest integer greater than or equal to x. The
boldface letters R and Z stand for the real numbers and for the integers,
respectively, while Rd denotes thep d-dimensional Euclidean space. For a point
x = (x1 , x2 , . . . , xd ) Rd , kxk = x21 + x22 + + x2d is the Euclidean norm
of x, and for x, y Rd , hx, yi = x1 y1 + x2 y2 + + xd yd is the scalar product.
Points of Rd are usually considered as column vectors.
The symbol B(x, r) denotes the closed ball of radius r centered at x in
some metric space (usually in Rd with the Euclidean distance), i.e., the set
of all points with distance at most r from x. We write B n for the unit ball
B(0, 1) in Rn . The symbol A denotes the boundary of a set A Rd , that
is, the set of points at zero distance from both A and its complement.
For a measurable set A Rd , vol(A) is the d-dimensional Lebesgue mea-
sure of A (in most cases the usual volume).
Let f and g be real functions (of one or several variables). The notation
f = O(g) means that there exists a number C such that |f | C|g| for all
values of the variables. Normally, C should be an absolute constant, but if
f and g depend on some parameter(s) that we explicitly declare to be fixed
(such as the space dimension d), then C may depend on these parameters
as well. The notation f = (g) is equivalent to g = O(f ), f (n) = o(g(n))
to limn (f (n)/g(n)) = 0, and f = (g) means that both f = O(g) and
f = (g).
For a random variable X, the symbol E[X] denotes the expectation of X,
and Prob[A] stands for the probability of an event A.
iv Notation and Terminology

Graphs are considered simple and undirected in this book unless stated
otherwise,
 so a graph G is a pair (V,
 E), where V is a set (the vertex set ) and
E V2 is the edge set. Here Vk denotes the set of all k-element subsets
of V . For a multigraph, the edges form a multiset, so two vertices can be
connected by several edges. For a given (multi)graph G, we write V (G) for
the vertex set and E(G) for theedge set. A complete graph has all possible
V

edges; that is, it is of the form V, 2 . A complete graph on n vertices is
denoted by Kn . A graph G is bipartite if the vertex set can be partitioned
into two subsets V1 and V2 , the (color) classes, in such a way that each edge
connects a vertex of V1 to a vertex of V2 . A graph G0 = (V 0 , E 0 ) is a subgraph
of a graph G = (V, E) if V 0 V and E 0 E. We also say that G contains
a copy of H if there is a subgraph G0 of G isomorphic to H, where G0 and
H are isomorphic if there is a bijective map : V (G0 ) V (H) such that
{u, v} E(G0 ) if and only if {(u), (v)} E(H) for all u, v V (G0 ). The
degree of a vertex v in a graph G is the number of edges of G containing v.
An r-regular graph has all degrees equal to r. Paths and cycles are graphs as
in the following picture,
... ...
paths cycles

and a path or cycle in G is a subgraph isomorphic to a path or cycle, respec-


tively. A graph G is connected if every two vertices can be connected by a
path in G.
We recall that a set X Rd is compact if and only if it is closed and
bounded, and that a continuous function f : X R defined on a compact X
attains its minimum (there exists x0 X with f (x0 ) f (x) for all x X).
The CauchySchwarz inequality is perhaps best remembered in the form
hx, yi kxk kyk for all x, y Rn .
A real function f defined on an interval A R (or, more generally, on a
convex set A Rd ) is convex if f (tx + (1t)y) tf (x) + (1t)f (y) for all
x, y A and t [0, 1]. Geometrically, the graph of f on [x, y] lies below the
segment connecting the points (x, f (x)) and (y, f (y)). If the second derivative
satisfies f 00 (x) 0 for all x in an (open) interval A R, then f is convex
on A. Jensens inequality is a straightforward generalization of the definition
of convexity: f (t1 x1 + t2 x2 + + tn xn ) t1 f (x1 ) + t2 f (x2 ) + + tn f (xn )
for all choices of nonnegative ti summing to 1 and all x1 , . . . , xn A. Or in
integral
R form,if Ris a probability measure on A and f is convex on A, we have
f A x d(x) A f (x) d(x). In the language of probability theory, if X
is a real random variable and  f : R R is convex, then f (E[X]) E[f (X)];
for example, (E[X])2 E X 2 .
1

Convexity

We begin with a review of basic geometric notions such as hyperplanes and


affine subspaces in Rd , and we spend some time by discussing the notion
of general position. Then we consider fundamental properties of convex sets
in Rd , such as a theorem about the separation of disjoint convex sets by a
hyperplane and Hellys theorem.

1.1 Linear and Affine Subspaces, General Position


Linear subspaces. Let Rd denote the d-dimensional Euclidean space. The
points are d-tuples of real numbers, x = (x1 , x2 , . . . , xd ).
The space Rd is a vector space, and so we may speak of linear subspaces,
linear dependence of points, linear span of a set, and so on. A linear subspace
of Rd is a subset closed under addition of vectors and under multiplication
by real numbers. What is the geometric meaning? For instance, the linear
subspaces of R2 are the origin itself, all lines passing through the origin,
and the whole of R2 . In R3 , we have the origin, all lines and planes passing
through the origin, and R3 .
Affine notions. An arbitrary line in R2 , say, is not a linear subspace unless
it passes through 0. General lines are what are called affine subspaces. An
affine subspace of Rd has the form x + L, where x Rd is some vector and L
is a linear subspace of Rd . Having defined affine subspaces, the other affine
notions can be constructed by imitating the linear notions.
What is the affine hull of a set X Rd ? It is the intersection of all affine
subspaces of Rd containing X. As is well known, the linear span of a set X
can be described as the set of all linear combinations of points of X. What
is an affine combination of points a1 , a2 , . . . , an Rd that would play an
analogous role? To see this, we translate the whole set by an , so that an
becomes the origin, we make a linear combination, and we translate back by
2 Chapter 1: Convexity

+an . This yields an expression of the form 1 (a1 an ) + 2 (a2 an ) + +


n (an an )+an = 1 a1 +2 a2 + +n1 an1 +(11 2 n1 )an ,
where 1 , . . . , n are arbitrary real numbers. Thus, an affine combination of
points a1 , . . . , an Rd is an expression of the form

1 a1 + + n an , where 1 , . . . , n R and 1 + + n = 1.

Then indeed, it is not hard to check that the affine hull of X is the set of all
affine combinations of points of X.
The affine dependence of points a1 , . . . , an means that one of them can
be written as an affine combination of the others. This is the same as the
existence of real numbers 1 , 2 , . . . n , at least one of them nonzero, such
that both

1 a1 + 2 a2 + + n an = 0 and 1 + 2 + + n = 0.

(Note the difference: In an affine combination, the i sum to 1, while in an


affine dependence, they sum to 0.)
Affine dependence of a1 , . . . , an is equivalent to linear dependence of the
n1 vectors a1 an , a2 an , . . . , an1 an . Therefore, the maximum possible
number of affinely independent points in Rd is d+1.
Another way of expressing affine dependence uses lifting one dimension
higher. Let bi = (ai , 1) be the vector in Rd+1 obtained by appending a new
coordinate equal to 1 to ai ; then a1 , . . . , an are affinely dependent if and only
if b1 , . . . , bn are linearly dependent. This correspondence of affine notions in
Rd with linear notions in Rd+1 is quite general. For example, if we identify
R2 with the plane x3 = 1 in R3 as in the picture,

x3 = 1

x3 = 0
0

then we obtain a bijective correspondence of the k-dimensional linear sub-


spaces of R3 that do not lie in the plane x3 = 0 with (k1)-dimensional affine
subspaces of R2 . The drawing shows a 2-dimensional linear subspace of R3
and the corresponding line in the plane x3 = 1. (The same works for affine
subspaces of Rd and linear subspaces of Rd+1 not contained in the subspace
xd+1 = 0.)
This correspondence also leads directly to extending the affine plane R2
into the projective plane: To the points of R2 corresponding to nonhorizontal
1.1 Linear and Affine Subspaces, General Position 3

lines through 0 in R3 we add points at infinity, that correspond to hori-


zontal lines through 0 in R3 . But in this book we remain in the affine space
most of the time, and we do not use the projective notions.
Let a1 , a2 , . . . , ad+1 be points in Rd , and let A be the d d matrix with
ai ad+1 as the ith column, i = 1, 2, . . . , d. Then a1 , . . . , ad+1 are affinely
independent if and only if A has d linearly independent columns, and this is
equivalent to det(A) 6= 0. We have a useful criterion of affine independence
using a determinant.
Affine subspaces of Rd of certain dimensions have special names. A (d1)-
dimensional affine subspace of Rd is called a hyperplane (while the word plane
usually means a 2-dimensional subspace of Rd for any d). One-dimensional
subspaces are lines, and a k-dimensional affine subspace is often called a
k-flat.
A hyperplane is usually specified by a single linear equation of the form
a1 x1 + a2 x2 + + ad xd = b. We usually write the left-hand side as the scalar
product ha, xi. So a hyperplane can be expressed as the set {x Rd : ha, xi =
b} where a Rd \ {0} and b R. A (closed) half-space in Rd is a set of
the form {x Rd : ha, xi b} for some a Rd \ {0} and some b R; the
hyperplane {x Rd : ha, xi = b} is its boundary.
General k-flats can be given either as intersections of hyperplanes or as
affine images of Rk (parametric expression). In the first case, an intersection
of k hyperplanes can also be viewed as a solution to a system Ax = b of linear
equations, where x Rd is regarded as a column vector, A is a k d matrix,
and b Rk . (As a rule, in formulas involving matrices, we interpret points
of Rd as column vectors.)
An affine mapping f : Rk Rd has the form f : y 7 By + c for some d k
matrix B and some c Rd , so it is a composition of a linear map with a
translation. The image of f is a k 0 -flat for some k 0 min(k, d). This k 0 equals
the rank of the matrix B.
General position. We assume that the points (lines, hyperplanes,. . . ) are
in general position. This magical phrase appears in many proofs. Intuitively,
general position means that no unlikely coincidences happen in the consid-
ered configuration. For example, if 3 points are chosen in the plane without
any special intention, randomly, they are unlikely to lie on a common line.
For a planar point set in general position, we always require that no three
of its points be collinear. For points in Rd in general position, we assume
similarly that no unnecessary affine dependencies exist: No k d+1 points
lie in a common (k2)-flat. For lines in the plane in general position, we
postulate that no 3 lines have a common point and no 2 are parallel.
The precise meaning of general position is not fully standard: It may
depend on the particular context, and to the usual conditions mentioned
above we sometimes add others where convenient. For example, for a planar
point set in general position we can also suppose that no two points have the
same x-coordinate.
4 Chapter 1: Convexity

What conditions are suitable for including into a general position as-
sumption? In other words, what can be considered as an unlikely coincidence?
For example, let X be an n-point set in the plane, and let the coordinates of
the ith point be (xi , yi ). Then the vector v(X) = (x1 , x2 , . . . , xn , y1 , y2 , . . . , yn )
can be regarded as a point of R2n . For a configuration X in which x1 = x2 ,
i.e., the first and second points have the same x-coordinate, the point v(X)
lies on the hyperplane {x1 = x2 } in R2n . The configurations X where some
two points share the x-coordinate thus correspond to the union of n2 hy-
perplanes in R2n . Since a hyperplane in R2n has (2n-dimensional) measure
zero, almost all points of R2n correspond to planar configurations X with all
the points having distinct x-coordinates. In particular, if X is any n-point
planar configuration and > 0 is any given real number, then there is a con-
figuration X 0 , obtained from X by moving each point by distance at most ,
such that all points of X 0 have distinct x-coordinates. Not only that: Almost
all small movements (perturbations) of X result in X 0 with this property.
This is the key property of general position: Configurations in general
position lie arbitrarily close to any given configuration (and they abound
in any small neighborhood of any given configuration). Here is a fairly gen-
eral type of condition with this property. Suppose that a configuration X
is specified by a vector t = (t1 , t2 , . . . , tm ) of m real numbers (coordinates).
The objects of X can be points in Rd , in which case m = dn and the tj
are the coordinates of the points, but they can also be circles in the plane,
with m = 3n and the tj expressing the center and the radius of each circle,
and so on. The general position condition we can put on the configuration
X is p(t) = p(t1 , t2 , . . . , tm ) 6= 0, where p is some nonzero polynomial in m
variables. Here we use the following well-known fact (a consequence of Sards
theorem; see, e.g., Bredon [Bre93], Appendix C): For any nonzero m-variate
polynomial p(t1 , . . . , tm ), the zero set {t Rm : p(t) = 0} has measure 0 in
Rm .
Therefore, almost all configurations X satisfy p(t) 6= 0. So any condition
that can be expressed as p(t) 6= 0 for a certain polynomial p in m real
variables, or, more generally, as p1 (t) 6= 0 or p2 (t) 6= 0 or . . . , for finitely or
countably many polynomials p1 , p2 , . . ., can be included in a general position
assumption.
For example, let X be an n-point set in Rd , and let us consider the con-
dition no d+1 points of X lie in a common hyperplane. In other words, no
d+1 points should be affinely dependent. As we know, the affine dependence
of d+1 points means that a suitable d d determinant equals 0. This deter-
minant is a polynomial (of degree d) in the coordinates of these d+1 points.
Introducing one polynomial for every (d+1)-tuple of the points, we obtain
n

d+1 polynomials such that at least one of them is 0 for any configuration X
with d+1 points in a common hyperplane. Other usual conditions for general
position can be expressed similarly.
1.2 Convex Sets, Convex Combinations, Separation 5

In many proofs, assuming general position simplifies matters consider-


ably. But what do we do with configurations X0 that are not in general
position? We have to argue, somehow, that if the statement being proved is
valid for configurations X arbitrarily close to our X0 , then it must be valid
for X0 itself, too. Such proofs, usually called perturbation arguments, are of-
ten rather simple, and almost always somewhat boring. But sometimes they
can be tricky, and one should not underestimate them, no matter how tempt-
ing this may be. A nontrivial example will be demonstrated in Section 5.5
(Lemma 5.5.4).

Exercises
1. Verify that the affine hull of a set X Rd equals the set of all affine
combinations of points of X. 2
2. Let A be a 2 3 matrix and let b R2 . Interpret the solution of the
system Ax = b geometrically (in most cases, as an intersection of two
planes) and discuss the possible cases in algebraic and geometric terms.
2

3. (a) What are the possible intersections of two (2-dimensional) planes


in R4 ? What is the typical case (general position)? What about two
hyperplanes in R4 ? 3
(b) Objects in R4 can sometimes be visualized as objects in R3 moving
in time (so time is interpreted as the fourth coordinate). Try to visualize
the intersection of two planes in R4 discussed (a) in this way.

1.2 Convex Sets, Convex Combinations, Separation


Intuitively, a set is convex if its surface has no dips:
not allowed in a convex set
x y

1.2.1 Definition (Convex set). A set C Rd is convex if for every two


points x, y C the whole segment xy is also contained in C. In other words,
for every t [0, 1], the point tx + (1 t)y belongs to C.
The intersection of an arbitrary family of convex sets is obviously convex.
So we can define the convex hull of a set X Rd , denoted by conv(X), as the
intersection of all convex sets in Rd containing X. Here is a planar example
with a finite X:

X conv(X)
6 Chapter 1: Convexity

An alternative description of the convex hull can be given using convex


combinations.
1.2.2 Claim. A point x belongs to conv(X) if and only if there exist Pn points
x1 , x2 , . . . xn XPand nonnegative real numbers t1 , t2 , . . . , tn with i=1 ti =
n
1 such that x = i=1 ti xi .
Pn
The expression i=1 ti xi as in the claim is called a convex combination
of the points x1 , x2 , . . . , xn . (Compare this with the definitions of linear and
affine combinations.)
Sketch of proof. Each convex combination of points of X must lie in
conv(X): For n = 2 this is by definition, and for larger n by induction.
Conversely, the set of all convex combinations obviously contains X, and it
is convex. 2
In Rd , it is sufficient to consider convex combinations involving at most
d+1 points:
1.2.3 Theorem (Caratheodorys theorem). Let X Rd . Then each
point of conv(X) is a convex combination of at most d+1 points of X.
For example, in the plane, conv(X) is the union of all triangles with
vertices at points of X. The proof of the theorem is left as an exercise to the
subsequent section.
A basic result about convex sets is the separability of disjoint convex sets
by a hyperplane.
1.2.4 Theorem (Separation theorem). Let C, D Rd be convex sets
with C D = . Then there exists a hyperplane h such that C lies in one
of the closed half-spaces determined by h, and D lies in the opposite closed
half-space. In other words, there exist a unit vector a Rd and a number
b R such that for all x C we have ha, xi b, and for all x D we have
ha, xi b.
If C and D are closed and at least one of them is bounded, they can be
separated strictly; in such a way that C h = D h = .
In particular, a closed convex set can be strictly separated from a point.
This implies that the convex hull of a closed set X equals the intersection of
all closed half-spaces containing X.
Sketch of proof. First assume that C and D are compact (i.e., closed and
bounded). Then the Cartesian product C D is a compact space, too, and
the distance function (x, y) 7 kx yk attains its minimum on C D. That
is, there exist points p C and q D such that the distance of C and D
equals the distance of p and q.
The desired separating hyperplane h can be taken as the one perpendic-
ular to the segment pq and passing through its midpoint:
1.2 Convex Sets, Convex Combinations, Separation 7

h
C
p
q
D

It is easy to check that h indeed avoids both C and D.


If D is compact and C closed, we can intersect C with a large ball and
get a compact set C 0 . If the ball is sufficiently large, then C and C 0 have the
same distance to D. So the distance of C and D is attained at some p C 0
and q D, and we can use the previous argument.
For arbitrary disjoint convex sets C and D, we choose Sa sequence C1
C2 C3 of compact convex subsets of C with n=1 Cn = C. For
example, assuming that 0 C, we can let Cn be the intersection of the
closure of (1 n1 )C with the ball of radius n centered at 0. A similar sequence
D1 D2 is chosen for D, and we let hn = {x Rd : han , xi = bn } be a
hyperplane separating Cn from Dn , where an is a unit vector and bn R. The
sequence (bn )n=1 is bounded, and by compactness, the sequence of (d+1)-
component vectors (an , bn ) Rd+1 has a cluster point (a, b). One can verify,
by contradiction, that the hyperplane h = {x Rd : ha, xi = b} separates C
and D (nonstrictly). 2

The importance of the separation theorem is documented by its presence


in several branches of mathematics in various disguises. Its home territory is
probably functional analysis, where it is formulated and proved for infinite-
dimensional spaces; essentially it is the so-called HahnBanach theorem. The
usual functional-analytic proof is different from the one we gave, and in a
way it is more elegant and conceptual. The proof sketched above uses more
special properties of Rd , but it is quite short and intuitive in the case of
compact C and D.
Connection to linear programming. A basic result in the theory of
linear programming is the Farkas lemma. It is a special case of the duality
of linear programming (discussed in Section ??) as well as the key step in its
proof.

1.2.5 Lemma (Farkas lemma, one of many versions). For every d n


real matrix A, exactly one of the following cases occurs:
(i) The system of linear equations Ax = 0 has a nontrivial nonnegative
solution x Rn (all components of x are nonnegative and at least one
of them is strictly positive).
8 Chapter 1: Convexity

(ii) There exists a y Rd such that y T A is a vector with all entries strictly
negative. Thus, if we multiply the jth equation in the system Ax = 0 by
yj and add these equations together, we obtain an equation that obviously
has no nontrivial nonnegative solution, since all the coefficients on the
left-hand sides are strictly negative, while the right-hand side is 0.

Proof. Let us see why this is yet another version of the separation theorem.
Let V Rd be the set of n points given by the column vectors of the
matrix A. We distinguish two cases: Either 0 conv(V ) or 0 6 conv(V ).
In the former case, we know that 0 is a convex combination of the points
of V , and the coefficients of this convex combination determine a nontrivial
nonnegative solution to Ax = 0.
In the latter case, there exists a hyperplane strictly separating V from 0,
i.e., a unit vector y Rd such that hy, vi < hy, 0i = 0 for each v V . This is
just the y from the second alternative in the Farkas lemma. 2

Bibliography and remarks. Most of the material in this chapter is


quite old and can be found in many surveys and textbooks. Providing
historical accounts of such well-covered areas is not among the goals
of this book, and so we mention only a few references for the specific
results discussed in the text and add some remarks concerning related
results.
The concept of convexity and the rudiments of convex geometry
have been around since antiquity. The initial chapter of the Handbook
of Convex Geometry [GW93] succinctly describes the history, and the
handbook can be recommended as the basic source on questions re-
lated to convexity, although knowledge has progressed significantly
since its publication.
For an introduction to functional analysis, including the Hahn
Banach theorem, see Rudin [Rud91], for example. The Farkas lemma
originated in [Far94] (nineteenth century!). More on the history of the
duality of linear programming can be found, e.g., in Schrijvers book
[Sch86].
As for the origins, generalizations, and applications of Caratheo-
dorys theorem, as well as of Radons lemma and Hellys theorem dis-
cussed in the subsequent sections, a recommendable survey is Eckhoff
[Eck93], and an older well-known source is Danzer, Grunbaum, and
Klee [DGK63].
Caratheodorys theorem comes from the paper [Car07], concerning
power series and harmonic analysis. A somewhat similar theorem, due
to Steinitz [Ste16], asserts that if x lies in the interior of conv(X) for
an X Rd , then it also lies in the interior of conv(Y ) for some Y X
with |Y | 2d. Bonnice and Klee [BK63] proved a common general-
ization of both these theorems: Any k-interior point of conv(X) is a
k-interior point of conv(Y ) for some Y X with at most max(2k, d+1)
1.3 Radons Lemma and Hellys Theorem 9

points, where x is called a k-interior point of a set C if it lies in the


relative interior of the convex hull of some k+1 affinely independent
points of C.

Exercises
1. Give a detailed proof of Claim 1.2.2. 2
2. Write down a detailed proof of the separation theorem. 3
3. Find an example of two disjoint closed convex sets in the plane that are
not strictly separable. 1
4. Let f : Rd Rk be an affine map.
(a) Prove that if C Rd is convex, then f (C) is convex as well. Is the
preimage of a convex set always convex? 2
(b) For X Rd arbitrary, prove that f (conv(X)) = conv(f (X)). 1
5. Let X Rd . Prove that diam(conv(X)) = diam(X), where the diameter
diam(Y ) of a set Y is sup{kx yk: x, y Y }. 3
6. A set C Rd is a convex cone if it is convex and for each x C, the ray


0x is fully contained in C.
(a) Analogously to the convex and affine hulls, define the appropriate
conic hull and the corresponding notion of combination (analogous
to the convex and affine combinations). 3
(b) Let C be a convex cone in Rd and b 6 C a point. Prove that there
exists a vector a with ha, xi 0 for all x C and ha, bi < 0. 2
7. (Variations on the Farkas lemma) Let A be a dn matrix and let b Rd .
(a) Prove that the system Ax = b has a nonnegative solution x Rn if
and only if every y Rd satisfying y T A 0 also satisfies y T b 0. 3
(b) Prove that the system of inequalities Ax b has a nonnegative
solution x if and only if every nonnegative y Rd with y T A 0 also
satisfies y T b 0. 3
8. (a) Let C Rd be a compact convex set with a nonempty interior, and
let p C be an interior point. Show that there exists a line ` passing
through p such that the segment ` C is at least as long as any segment
parallel to ` and contained in C. 4
(b) Show that (a) may fail for C compact but not convex. 1

1.3 Radons Lemma and Hellys Theorem


Caratheodorys theorem from the previous section, together with Radons
lemma and Hellys theorem presented here, are three basic properties of con-
vexity in Rd involving the dimension. We begin with Radons lemma.
1.3.1 Theorem (Radons lemma). Let A be a set of d+2 points in Rd .
Then there exist two disjoint subsets A1 , A2 A such that

conv(A1 ) conv(A2 ) 6= .
10 Chapter 1: Convexity

A point x conv(A1 )conv(A2 ), where A1 and A2 are as in the theorem,


is called a Radon point of A, and the pair (A1 , A2 ) is called a Radon partition
of A (it is easily seen that we can require A1 A2 = A).
Here are two possible cases in the plane:

Proof. Let A = {a1 , a2 , . . . , ad+2 }. These d+2 points are necessarily affinely
dependent. That is, there exist real numbers 1 , . . . , d+2 , not all of them 0,
Pd+2 Pd+2
such that i=1 i = 0 and i=1 i ai = 0.
Set P = {i: i > 0} and N = {i: i < 0}. Both P and N are nonempty.
We claim that P and N determine the desired subsets. Let us put A1 =
{ai : i P } and A2 = {ai : i N }. We are going to exhibit a point x that is
contained in P the convex hulls of both thesePsets.
Put S = iP i ; we also have S = iN i . Then we define
X i
x= ai . (1.1)
S
iP
Pd+2 P P
Since i=1 i ai = 0 = iP i ai + iN i ai , we also have
X i
x= ai . (1.2)
S
iN

The coefficients of the ai in (1.1) are nonnegative and sum to 1, so x is a


convex combination of points of A1 . Similarly, (1.2) expresses x as a convex
combination of points of A2 . 2
Hellys theorem is one of the most famous results of a combinatorial nature
about convex sets.
1.3.2 Theorem (Hellys theorem). Let C1 , C2 , . . . , Cn be convex sets in
Rd , n d+1. Suppose that the intersection of every d+1 of these sets is
nonempty. Then the intersection of all the Ci is nonempty.
The first nontrivial case states that if every 3 among 4 convex sets in
the plane intersect, then there is a point common to all 4 sets. This can be
proved by an elementary geometric argument, perhaps distinguishing a few
cases, and the reader may want to try to find a proof before reading further.
In a contrapositive form, Hellys Tntheorem guarantees that whenever
C1 , C2 , . . . , Cn are convex sets with i=1 Ci = , then this is witnessed by
some at most d+1 sets with empty intersection among the Ci . In this way,
many proofs are greatly simplified, since in planar problems, say, one can deal
with 3 convex sets instead of an arbitrary number, as is amply illustrated in
the exercises below.
1.3 Radons Lemma and Hellys Theorem 11

It is very tempting and quite usual to formulate Hellys theorem as fol-


lows: If every d+1 among n convex sets in Rd intersect, then all the sets
intersect. But, strictly speaking, this is false, for a trivial reason: For d 2,
the assumption as stated here is met by n = 2 disjoint convex sets.
Proof of Hellys theorem. (Using Radons lemma.) For a fixed d, we
proceed by induction on n. The case n = d+1 is clear, so we suppose that
n d+2 and that the statement of Hellys theorem holds for smaller n.
Actually, n = d+2 is the crucial case; the result for larger n follows at once
by a simple induction.
Consider sets C1 , C2 , . . . , Cn satisfying the assumptions. If we leave out
any one of these sets, the remaining sets have a nonempty T intersection by
the inductive assumption. Let us fix a point ai j6=i Cj and consider the
points a1 , a2 , . . . , ad+2 . By Radons lemma, there exist disjoint index sets
I1 , I2 {1, 2, . . . , d+2} such that

conv({ai : i I1 }) conv({ai : i I2 }) 6= .

We pick a point x in this intersection. The following picture illustrates the


case d = 2 and n = 4:

a1
C2
a4 x a3
C1
a2
C3 C4

We claim that x lies in the intersection of all the Ci . Consider some i


{1, 2, . . . , n}; then i 6 I1 or i 6 I2 . In the former case, each aj with j I1 lies
in Ci , and so x conv({aj : j I1 }) Ci . For i T 6 I2 we similarly conclude
n
that x conv({aj : j I2 }) Ci . Therefore, x i=1 Ci . 2
An infinite version of Hellys theorem. If we have an infinite collection
of convex sets in Rd such that any d+1 of them have a common point, the
entire collection still need not have a common point. Two examples in R1 are
the families of intervals {(0, 1/n): n = 1, 2, . . .} and {[n, ): n = 1, 2, . . .}.
The sets in the first example are not closed, and the second example uses
unbounded sets. For compact (i.e., closed and bounded) sets, the theorem
holds:
1.3.3 Theorem (Infinite version of Hellys theorem). Let C be an ar-
bitrary infinite family of compact convex sets in Rd such that any d+1 of the
sets have a nonempty intersection. Then all the sets of C have a nonempty
intersection.
12 Chapter 1: Convexity

Proof. By Hellys theorem, any finite subfamily of C has a nonempty inter-


section. By a basic property of compactness, if we have an arbitrary family
of compact sets such that each of its finite subfamilies has a nonempty inter-
section, then the entire family has a nonempty intersection. 2

Several nice applications of Hellys theorem are indicated in the exercises


below, and we will meet a few more later in this book.

Bibliography and remarks. Helly proved Theorem 1.3.2 in 1913


and communicated it to Radon, who published a proof in [Rad21]. This
proof uses Radons lemma, although the statement wasnt explicitly
formulated in Radons paper. References to many other proofs and
generalizations can be found in the already mentioned surveys [Eck93]
and [DGK63].
Hellys theorem inspired a whole industry of Helly-type theorems.
A family B of sets is said to have Helly number h if the following holds:
Whenever a finite subfamily F BTis such that every h or fewer sets
of F have a common point, then F = 6 . So Hellys theorem says
that the family of all convex sets in Rd has Helly number d+1. More
generally, let P be some property of families of sets that is hereditary,
meaning that if F has property P and F 0 F, then F 0 has P as well.
A family B is said to have Helly number h with respect to P if for
every finite F B, all subfamilies of F of size at most h having P
implies F having P . That is, the absence of P is always witnessed by
some at most h sets, so it is a local property.

Exercises
1. Prove Caratheodorys theorem (you may use Radons lemma). 4
2. Let K Rd be a convex set and let C1 , . . . , Cn Rd , n d+1, be
convex sets such that the intersection of every d+1 of them contains a
translated copy of K. Prove that then the intersection of all the sets Ci
also contains a translated copy of K. 2
This result was noted by Vincensini [Vin39] and by Klee [Kle53].
3. Find an example of 4 convex sets in the plane such that the intersection
of each 3 of them contains a segment of length 1, but the intersection of
all 4 contains no segment of length 1. 1
4. A strip of width w is a part of the plane bounded by two parallel lines at
distance w. The width of a set X R2 is the smallest width of a strip
containing X.
(a) Prove that a compact convex set of width 1 contains a segment of
length 1 of every direction. 3
(b) Let {C1 , C2 , . . . , Cn } be closed convex sets in the plane, n 3, such
that
Tn the intersection of every 3 of them has width at least 1. Prove that
i=1 Ci has width at least 1.
2
1.3 Radons Lemma and Hellys Theorem 13

The result as in (b), for arbitrary dimension d, was proved by Sallee


[Sal75], and a simple argument using Hellys theorem was noted by Buch-
man and Valentine [BV82].
5. Statement: Each set X R2 of diameter at most 1 (i.e., any2 points
have distance at most 1) is contained in some disc of radius 1/ 3.
(a) Prove the statement for 3-element sets X. 2
(b) Prove the statement for all finite sets X. 2
(c) Generalize the statement to Rd : determine the smallest r = r(d) such
that every set of diameter 1 in Rd is contained in a ball of radius r (prove
your claim). 4
The result as in (c) is due to Jung; see [DGK63].
6. Let C Rd be a compact convex set. Prove that the mirror image of C
can be covered by a suitable translate of C blown up by the factor of d;
that is, there is an x Rd with C x + dC. 4
7. (a) Prove that if the intersection of each
Tn 4 or fewer among convex sets
C1 , . . . , Cn R2 contains a ray then i=1 Ci also contains a ray. 4
(b) Show that the number 4 in (a) cannot be replaced by 3. 2
This result, and an analogous one in Rd with the Helly number 2d, are
due to Katchalski [Kat78].
8. For a set X R2 and a point x X, let us denote by V (x) the set of all
points y X that can see x, i.e., points such that the segment xy is
contained in X. The kernel of X is defined as the set of all points x X
such that V (x) = X. A set with a nonempty kernel is called star-shaped.
(a) Prove that the kernel of any set is convex. 1
(b) Prove that if V (x) V (y) V (z) 6= for every x, y, z X and X is
compact, then X is star-shaped. That is, if every 3 paintings in a (planar)
art gallery can be seen at the same time from some location (possibly
different for different triples of paintings), then all paintings can be seen
simultaneously from somewhere. If it helps, assume that X is a polygon.
5

(c) Construct a nonempty set X R2 such that each of its finite subsets
can be seen from some point of X but X is not star-shaped. 2
The result in (b), as well as the d-dimensional generalization (with ev-
ery d+1 regions V (x) intersecting), is called Krasnoselskis theorem; see
[Eck93] for references and related results.
9. In the situation of Radons lemma (A is a (d+2)-point set in Rd ), call
a point x Rd a Radon point of A if it is contained in convex hulls of
two disjoint subsets of A. Prove that if A is in general position (no d+1
points affinely dependent), then its Radon point is unique. 3
10. (a) Let X, Y R2 be finite point sets, and suppose that for every subset
S X Y of at most 4 points, S X can be separated (strictly) by a
line from S Y . Prove that X and Y are line-separable. 3
(b) Extend (a) to sets X, Y Rd , with |S| d+2. 5
The result (b) is called Kirchbergers theorem [Kir03].
14 Chapter 1: Convexity

1.4 Centerpoint and Ham Sandwich


We prove an interesting result as an application of Hellys theorem.
1.4.1 Definition (Centerpoint). Let X be an n-point set in Rd . A point
x Rd is called a centerpoint of X if each closed half-space containing x
n
contains at least d+1 points of X.
Let us stress that one set may generally have many centerpoints, and a
centerpoint need not belong to X.
The notion of centerpoint can be viewed as a generalization of the me-
dian of one-dimensional data. Suppose that x1 , . . . , xn R are results of
measurements of an unknown real parameter x. How do we estimate x from
the xi ? We can use the arithmetic mean, but if one of the measurements is
completely wrong (say, 100 times larger than the others), we may get quite
a bad estimate. A more robust estimate is a median, i.e., a point x such
that at least n2 of the xi lie in the interval (, x] and at least n2 of them lie
in [x, ). The centerpoint can be regarded as a generalization of the median
for higher-dimensional data.
1
In the definition of centerpoint we could replace the fraction d+1 by some
1
other parameter (0, 1). For > d+1 , such an -centerpoint need not
1
always exist: Take d+1 points in general position for X. With = d+1 as in
the definition above, a centerpoint always exists, as we prove next.
Centerpoints are important, for example, in some algorithms of divide-
and-conquer type, where they help divide the considered problem into smaller
subproblems. Since no really efficient algorithms are known for finding
exact centerpoints, the algorithms often use -centerpoints with a suit-
1
able < d+1 , which are easier to find.

1.4.2 Theorem (Centerpoint theorem). Each finite point set in Rd has


at least one centerpoint.

Proof. First we note an equivalent definition of a centerpoint: x is a cen-


terpoint of X if and only if it lies in each open half-space such that
d
|X | > d+1 n.
We would like to apply Hellys theorem to conclude that all these open
half-spaces intersect. But we cannot proceed directly, since we have infinitely
many half-spaces and they are open and unbounded. Instead of such an open
half-space , we thus consider the compact convex set conv(X ) .


conv( X)
1.4 Centerpoint and Ham Sandwich 15

d
Letting run through all open half-spaces with |X | > d+1 n, we obtain
d
a family C of compact convex sets. Each of them contains more than d+1 n
points of X, and so the intersection of any d+1 of them contains at least
one point of X. The family C consists of finitely
T many distinct sets (since X
has finitely many distinct subsets), and so C = 6 by Hellys theorem. Each
point in this intersection is a centerpoint. 2

In the definition of a centerpoint we can regard the finite set X as defin-


ing a distribution of mass in Rd . The centerpoint theorem asserts that for
1
some point x, any half-space containing x encloses at least d+1 of the to-
tal mass. It is not difficult to show that this remains valid for continuous
mass distributions, or even for arbitrary Borel probability measures on Rd
(Exercise 1).
Ham-sandwich theorem and its relatives. Here is another important
result, not much related to convexity but with a flavor resembling the cen-
terpoint theorem.
1.4.3 Theorem (Ham-sandwich theorem). Every d finite sets in Rd can
be simultaneously bisected by a hyperplane. A hyperplane h bisects a finite
set A if each of the open half-spaces defined by h contains at most b|A|/2c
points of A.
This theorem is usually proved via continuous mass distributions using
a tool from algebraic topology: the BorsukUlam theorem. Here we omit a
proof.
Note that if Ai has an odd number of points, then every h bisecting Ai
passes through a point of Ai . Thus if A1 , . . . , Ad all have odd sizes and their
union is in general position, then every hyperplane simultaneously bisecting
them is determined by d points, one of each Ai . In particular, there are only
finitely many such hyperplanes.
Again, an analogous ham-sandwich theorem holds for arbitrary d Borel
probability measures in Rd .
Center transversal theorem. There can be beautiful new things to dis-
cover even in well-studied areas of mathematics. A good example is the fol-
lowing recent result, which interpolates between the centerpoint theorem
and the ham-sandwich theorem.
1.4.4 Theorem (Center transversal theorem). Let 1 k d and let
A1 , A2 , . . . , Ak be finite point sets in Rd . Then there exists a (k1)-flat f
such that for every hyperplane h containing f , both the closed half-spaces
1
defined by h contain at least dk+2 |Ai | points of Ai , i = 1, 2, . . . , k.
The ham-sandwich theorem is obtained for k = d and the centerpoint
theorem for k = 1. The proof, which we again have to omit, is based on a
result of algebraic topology, too, but it uses a considerably more advanced
16 Chapter 1: Convexity

machinery than the ham-sandwich theorem. However, the weaker result with
1 1
d+1 instead of dk+2 is easy to prove; see Exercise 2.

Bibliography and remarks. The centerpoint theorem was es-


tablished by Rado [Rad47]. According to Steinleins survey [Ste85],
the ham-sandwich theorem was conjectured by Steinhaus (who also
invented the popular 3-dimensional interpretation, namely, that the
ham, the cheese, and the bread in any ham sandwich can be simulta-
neously bisected by a single straight motion of the knife) and proved
by Banach. The center transversal theorem was found by Dolnikov
[Dol92] and, independently, by Zivaljevic and Vrecica [ZV90].
Significant effort has been devoted to efficient algorithms for find-
ing (approximate) centerpoints and ham-sandwich cuts (i.e., hyper-
planes as in the ham-sandwich theorem). In the plane, a ham-sandwich
cut for two n-point sets can be computed in linear time (Lo, Matousek,
and Steiger [LMS94]). In a higher but fixed dimension, the complexity
of the best exact algorithms is currently slightly better than O(nd1 ).
A centerpoint in the plane, too, can be found in linear time (Jadhav
and Mukhopadhyay [JM94]). Both approximate ham-sandwich cuts
(in the ratio 1 : 1+ for a fixed > 0) and approximate centerpoints
1
(( d+1 )-centerpoints) can be computed in time O(n) for every fixed
dimension d and every fixed > 0, but the constant depends expo-
nentially on d, and the algorithms are impractical if the dimension is
not quite small. A practically efficient randomized algorithm for com-
puting approximate centerpoints in high dimensions (-centerpoints
with 1/d2 ) was given by Clarkson, Eppstein, Miller, Sturtivant,
and Teng [CEM+ 96].

Exercises
1. (Centerpoints for general mass distributions)
(a) Let be a Borel probability measure on Rd ; that is, (Rd ) = 1 and
each open set is measurable. Show that for each open half-space with
() > t there exists a compact set C with (C) > t. 2
(b) Prove that each Borel probability measure in Rd has a centerpoint
(use (a) and the infinite Hellys theorem). 2
2. Prove that for any k finite sets A1 , . . . , Ak Rd , where 1 k d, there
exists a (k1)-flat such that every hyperplane containing it has at least
1
d+1 |Ai | points of Ai in both of its closed half-spaces for all i = 1, 2, . . . , k.
1
5

Convex Polytopes

Convex polytopes are convex hulls of finite point sets in Rd . They consti-
tute the most important class of convex sets with an enormous number of
applications and connections.
Three-dimensional convex polytopes, especially the regular ones, have
been fascinating people since the antiquity. Their investigation was one of
the main sources of the theory of planar graphs, and thanks to this well-
developed theory they are quite well understood. But convex polytopes in
dimension 4 and higher are considerably more challenging, and a surpris-
ingly deep theory, mainly of algebraic nature, was developed in attempts to
understand their structure.
A strong motivation for the study of convex polytopes comes from prac-
tically significant areas such as combinatorial optimization, linear program-
ming, and computational geometry. Let us look at a simple example illus-
trating how polytopes can be associated with combinatorial objects. The
3-dimensional polytope in the picture
2341 1342

2431 1243
1432
2143
3421 1423 1234

1324 2134
3412 2413
2314 3124
4312
4213 3214
18 Chapter 5: Convex Polytopes

is called the permutahedron. Although it is 3-dimensional, it is most natu-


rally defined as a subset of R4 , namely, the convex hull of the 24 vectors
obtained by permuting the coordinates of the vector (1, 2, 3, 4) in all possible
ways. In the picture, the (visible) vertices are labeled by the correspond-
ing permutations. Similarly, the d-dimensional permutahedron is the con-
vex hull of the (d+1)! vectors in Rd+1 arising by permuting the coordinates
of (1, 2, . . . , d+1). One can observe that the edges of the polytope connect
exactly pairs of permutations differing by a transposition of two adjacent
numbers, and a closer examination reveals other connections between the
structure of the permutahedron and properties of permutations.
There are many other, more sophisticated, examples of convex polytopes
assigned to combinatorial and geometric objects such as graphs, partially or-
dered sets, classes of metric spaces, or triangulations of a given point set. In
many cases, such convex polytopes are a key tool for proving hard theorems
about the original objects or for obtaining efficient algorithms. Two impres-
sive examples are discussed in Chapter ??, and several others are scattered
in other chapters.
The present chapter should convey some initial working knowledge of
convex polytopes for a nonpolytopist. It is just a small sample of an extensive
theory. A more comprehensive modern introduction is the book by Ziegler
[Zie94].

5.1 Geometric Duality


First we discuss geometric duality, a simple technical tool indispensable in
the study of convex polytopes and handy in many other situations. We begin
with a simple motivating question.
How can we visualize the set of all lines intersecting a convex pentagon
as in the picture?
a5
a1 a4

a2
a3

A suitable way is provided by linepoint duality.


5.1.1 Definition (Duality transform). The (geometric) duality transform
is a mapping denoted by D0 . To a point a Rd \{0} it assigns the hyperplane
D0 (a) = {x Rd : ha, xi = 1},
and to a hyperplane h not passing through the origin, which can be uniquely
written in the form h = {x Rd : ha, xi = 1}, it assigns the point D0 (h) =
a Rd \ {0}.
5.1 Geometric Duality 19

Here is the geometric meaning of the duality transform. If a is a point


at distance from 0, then D0 (a) is the hyperplane perpendicular to the line
0a and intersecting that line at distance 1 from 0, in the direction from 0
towards a.
a
1

0 D0 (a)

A nice interpretation of duality is obtained by working in Rd+1 and iden-


tifying the primal Rd with the hyperplane = {x Rd+1 : xd+1 = 1}
and the dual Rd with the hyperplane = {x Rd+1 : xd+1 = 1}. The
hyperplane dual to a point a is produced as follows: We construct the
hyperplane in Rd+1 perpendicular to 0a and containing 0, and we intersect
it with . Here is an illustration for d = 2:


a
0

D0 (a)

In this way, the duality D0 can be naturally extended to k-flats in Rd , whose


duals are (dk1)-flats. Namely, given a k-flat f , we consider the (k+1)-
flat F through 0 and f , we construct the orthogonal complement of F , and
we intersect it with , obtaining D0 (f ).
Let us consider the pentagon drawn above and place it so that the origin
lies in the interior. Let vi = D0 (`i ), where `i is the line containing the side
ai ai+1 . Then the points dual to the lines intersecting the pentagon a1 a2 . . . a5
fill exactly the exterior of the convex pentagon v1 v2 . . . v5 :

v4
`1
v5
a5
a4 v3
a1 0
v1 = D0 (`1 ) a2
a3
v2
20 Chapter 5: Convex Polytopes

This follows easily from the properties of duality listed below (of course, there
is nothing special about a pentagon here). Thus, the considered set of lines
can be nicely described in the dual plane. A similar passage from lines to
points or back is useful in many geometric or computational problems.
Properties of the duality transform. Let p be a point of Rd distinct
from the origin and let h be a hyperplane in Rd not containing the origin.
Let h stand for the closed half-space bounded by h and containing the
origin, while h+ denotes the other closed half-space bounded by h. That is,
if h = {x Rd : ha, xi = 1}, then h = {x Rd : ha, xi 1}.
5.1.2 Lemma (Duality preserves incidences).
(i) p h if and only if D0 (h) D0 (p).
(ii) p h if and only if D0 (h) D0 (p) .

Proof. (i) Let h = {x Rd : ha, xi = 1}. Then p h means ha, pi = 1.


Now, D0 (h) is the point a, and D0 (p) is the hyperplane {y Rd : hy, pi = 1},
and hence D0 (h) = a D0 (p) also means just ha, pi = 1. Part (ii) is proved
similarly. 2

5.1.3 Definition (Dual set). For a set X Rd , we define the set dual to
X, denoted by X , as follows:

X = {y Rd : hx, yi 1 for all x X} .

Another common name used for the duality is polarity; the dual set would
then be called the polar set. Sometimes it is denoted by X .
Geometrically, X is the intersection of all half-spaces of the form D0 (x)
with x X. Or in other words, X consists of the origin plus all points y
such that X D0 (y) . For example, if X is the pentagon a1 a2 . . . a5 drawn
above, then X is the pentagon v1 v2 . . . v5 .
For any set X, the set X is obviously closed and convex and contains
the origin. Using the separation theorem (Theorem 1.2.4), it is easily shown
that for any set X Rd , the set (X ) is the closure of conv(X {0}). In
particular, for a closed convex set containing the origin we have (X ) = X
(Exercise 3).
For a hyperplane h, the dual set h is different from the point D0 (h).1
For readers familiar with the duality of planar graphs, let us remark that
it is closely related to the geometric duality applied to convex polytopes in
R3 . For example, the next drawing illustrates a planar graph and its dual
graph (dashed):

1
In the literature, however, the star notation is sometimes also used for the dual
point or hyperplane, so for a point p, the hyperplane D0 (p) would be denoted by
p , and similarly, h may stand for D0 (h).
5.1 Geometric Duality 21

Later we will see that these are graphs of the 3-dimensional cube and of
the regular octahedron, which are polytopes dual to each other in the sense
defined above. A similar relation holds for all 3-dimensional polytopes and
their graphs.
Other variants of duality. The duality transform D0 defined above is just
one of a class of geometric transforms with similar properties. For some pur-
poses, other such transforms (dualities) are more convenient. A particularly
important duality, denoted by D, corresponds to moving the origin to the
minus infinity of the xd -axis (the xd -axis is considered vertical). A formal
definition is as follows.
5.1.4 Definition (Another duality). A nonvertical hyperplane h can be
uniquely written in the form h = {x Rd : xd = a1 x1 + + ad1 xd1 ad }.
We set D(h) = (a1 , . . . , ad1 , ad ). Conversely, the point a = (a1 , . . . , ad1 , ad )
maps back to h.
The property (i) of Lemma 5.1.2 holds for this D, and an analogue of (ii)
is:
(ii0 ) A point p lies above a hyperplane h if and only if the point D(h) lies
above the hyperplane D(p).

Exercises
1. Let C = {x Rd : |x1 | + + |xd | 1}. Show that C is the d-di-
mensional cube {x Rd : max |xi | 1}. Picture both bodies for d = 3.
2

2. Prove the assertion made in the text about the lines intersecting a convex
pentagon. 2
3. Show that for any X Rd , (X ) equals the closure of conv(X {0}),
where X stands for the dual set to X. 3
4. Let C Rd be a convex set. Prove that C is bounded if and only if 0
lies in the interior of C. 2
5. Show that C = C if and only if C is the unit ball centered at the origin.
2

6. (a) Let C = conv(X)T Rd . Prove that C = xX D0 (x) . 2


T
(b) Show that if C = hH h , where H is a collection of hyperplanes not
passing through 0, and if C is bounded, then C = conv{D0 (h): h H}.
2

(c) What is the right analogue of (b) if C is unbounded? 2


22 Chapter 5: Convex Polytopes

7. What is the dual set h for a hyperplane h, and what about h ? 3


8. Verify the geometric interpretation of the duality D0 outlined in the text
(using the embeddings of Rd into Rd+1 ). 2
9. (a) Let s be a segment in the plane. Describe the set of all points dual
to lines intersecting s. 1
(b) Consider n 3 segments in the plane, such that none of them contains
0 but they all lie on lines passing through 0. Show that if every 3 among
such segments can be intersected by a single line, then all the segments
can be simultaneously intersected by a line. 3
(c) Show that the assumption in (b) that the extensions of the segments
pass through 0 is essential: For each n 1, construct n+1 pairwise
disjoint segments in the plane that cannot be simultaneously intersected
by a line but every n of them can (such an example was first found by
Hadwiger and Debrunner). 4

5.2 H-Polytopes and V -Polytopes


A convex polytope in the plane is a convex polygon. Famous examples of
convex polytopes in R3 are the Platonic solids: regular tetrahedron, cube,
regular octahedron, regular dodecahedron, and regular icosahedron. A convex
polytope in R3 is a convex set bounded by finitely many convex polygons.
Such a set can be regarded as a convex hull of a finite point set, or as an
intersection of finitely many half-spaces. We thus define two types of convex
polytopes, based on these two views.
5.2.1 Definition (H-polytope and V -polytope). An H-polyhedron is
an intersection of finitely many closed half-spaces in some Rd . An H-poly-
tope is a bounded H-polyhedron.
A V -polytope is the convex hull of a finite point set in Rd .
A basic theorem about convex polytopes claims that from the mathemat-
ical point of view, H-polytopes and V -polytopes are equivalent.
5.2.2 Theorem. Each V -polytope is an H-polytope. Each H-polytope is a
V -polytope.
This is one of the theorems that may look obvious and whose proof
needs no particularly clever idea but does require some work. In the present
case, we do not intend to avoid it. Actually, we have quite a neat proof in
store, but we postpone it to the end of this section.
Although H-polytopes and V -polytopes are mathematically equivalent,
there is an enormous difference between them from the computational point
of view. That is, it matters a lot whether a convex polytope is given to
us as a convex hull of a finite set or as an intersection of half-spaces. For
example, given a set of n points specifying a V -polytope, how do we find
its representation as an H-polytope? It is not hard to come up with some
5.2 H-Polytopes and V -Polytopes 23

algorithm, but the problem is to find an efficient algorithm that would allow
one to handle large real-world problems. This algorithmic question is not yet
satisfactorily solved. Moreover, in some cases the number of required half-
spaces may be astronomically large compared to the number n of points, as
we will see later in this chapter.
As another illustration of the computational difference between V -po-
lytopes and H-polytopes, we consider the maximization of a given linear
function over a given polytope. For V -polytopes it is a trivial problem, since
it suffices to substitute all points of V into the given linear function and select
the maximum of the resulting values. But maximizing a linear function over
the intersection of a collection of half-spaces is the basic problem of linear
programming, and it is certainly nontrivial.
Terminology. The usual terminology does not distinguish V -polytopes and
H-polytopes. A convex polytope means a point set in Rd that is a V -polytope
(and thus also an H-polytope). An arbitrary, possibly unbounded, H-poly-
hedron is called a convex polyhedron. All polytopes and polyhedra considered
in this chapter are convex, and so the adjective convex is often omitted.
The dimension of a convex polyhedron is the dimension of its affine hull.
It is the smallest dimension of a Euclidean space containing a congruent
copy of P . The phrase d-dimensional polytope is often abbreviated to d-
polytope.
Basic examples. One of the easiest classes of polytopes is that of cubes.
The d-dimensional cube as a point set is the Cartesian product [1, 1]d.

d=1 d=2 d=3


As a V -polytope, the d-dimensional cube is the convex hull of the set {1, 1}d
(2d points), and as an H-polytope, it can be described by the inequalities
1 xi 1, i = 1, 2, . . . , d, i.e., by 2d half-spaces. We note that it is also
the unit ball of the maximum norm kxk = maxi |xi |.
Another important example is the class of crosspolytopes (or generalized
octahedra). The d-dimensional crosspolytope is the convex hull of the co-
ordinate cross, i.e., conv{e1 , e1 , e2 , e2 , . . . , ed , ed }, where e1 , . . . , ed are
the vectors of the standard orthonormal basis.

d=1 d=2 d=3


24 Chapter 5: Convex Polytopes

Pd
It is also the unit ball of the `1 -norm kxk1 = i=1 |xi | . As an H-polytope,
it can be expressed by the 2d half-spaces of the form h, xi 1, where runs
through all vectors in {1, 1}d.
The polytopes with the smallest possible number of vertices (for a given
dimension) are called simplices.
5.2.3 Definition (Simplex). A simplex is the convex hull of an affinely
independent point set in some Rd .
A d-dimensional simplex in Rd can also be represented as an intersection
of d+1 half-spaces, as is not difficult to check.
A regular d-dimensional simplex is the convex hull of d+1 points with all
pairs of points having equal distances.

d=0 d=1 d=2 d=3

Unlike cubes and crosspolytopes, d-dimensional regular simplices do not have


a very nice coordinate representation in Rd . The simplest and most useful
representation lives one dimension higher: The convex hull of the d+1 vectors
e1 , . . . , ed+1 of the standard orthonormal
basis in Rd+1 is a d-dimensional
regular simplex with side length 2.
(0, 0, 1)

(1, 0, 0)

(0, 1, 0)

Proof of Theorem 5.2.2 (equivalence of H-polytopes and V -poly-


topes). We first show that any H-polytope is also a V -polytope. We proceed
by induction on d. The case d = 1 being trivial, we suppose that d 2. T
So let be a finite collection of closed half-spaces in Rd such that P =
is nonempty and bounded. For each , let F = P be the intersection
of P with the bounding hyperplane of . Each nonempty F is an H-polytope
of dimension at most d1 (correct?), and so it is the convex hull of a finite
set V F by the inductive hypothesis. S
We claim that P = conv(V ), where V = V . Let x P and let `
be a line passing through x. The intersection ` P is a segment; let y and z
be its endpoints. There are , such that y F and z F (if y were
5.2 H-Polytopes and V -Polytopes 25

not on the boundary of any , we could continue along ` a little further


within P ).
y F

`
P x
z
F

We have y conv(V ) and z conv(V ), and thus x conv(V V )


conv(V ).
We have proved that any H-polytope is a V -polytope, and it remains to
show that a V -polytope can be expressed as the intersection of finitely many
half-spaces. This follows easily by duality (and implicitly uses the separation
theorem).
Let P = conv(V ) with V finite, and assume that T 0 is an interior point
of P . By Exercise 5.1.6(a), the dual body P equals vV D0 (v) , and by
Exercise 5.1.4 it is bounded. By what we have already proved, P is a V -
polytope, and by Exercise 5.1.6(a) again, P = (P ) is the intersection of
finitely many half-spaces. 2

Bibliography and remarks. The theory of convex polytopes is


a well-developed area covered in numerous books and surveys, such
as the already recommended recent monograph [Zie94] (with addenda
and updates on the web page of its author), the very influential book
by Grunbaum whose new edition [Gru03] is accompanied by surveys
of recent developments, the chapters on polytopes in the handbooks
of discrete and computational geometry [GO97], of convex geometry
[GW93], and of combinatorics [GGL95], or the books McMullen and
Shephard [MS71] and Brnsted [Br83], concentrating on questions
about the numbers of faces. Recent progress in combinatorial and com-
putational polytope theory is reflected in the collection [KZ00]. For
analyzing examples, one should be aware of (free) software systems
for manipulating convex polytopes, such as polymake by Gawrilow
and Joswig [GJ00].
Interesting discoveries about 3-dimensional convex polytopes were
already made in ancient Greece. The treatise by Schlafli [Sch01] writ-
ten in 185052 is usually mentioned as the beginning of modern theory,
and several books were published around the turn of the century. We
refer to Grunbaum [Gru67], Schrijver [Sch86], and to the other sources
mentioned above for historical accounts.
The permutahedron mentioned in the introduction to this chapter
was considered by Schoute [Sch11], and it arises by at least two other
quite different and natural constructions (see [Zie94]).
26 Chapter 5: Convex Polytopes

There are several ways of proving the equivalence of H-polytopes


and V -polytopes. Ours is inspired by a proof by Edmonds, as presented
in Fukudas lecture notes (ETH Zurich). A classical algorithmic proof
is provided by the FourierMotzkin elimination procedure, which pro-
ceeds by projections on coordinate hyperplanes; see [Zie94] for a de-
tailed exposition. The double-description method is a similar algorithm
formulated in the dual setting, and it is still one of the most efficient
known computational methods. We will say a little more about the
algorithmic problem of expressing the convex hull of a finite set as the
intersection of half-spaces in the notes to Section 5.5.
One may ask, What is a vertex description of an unbounded H-
polyhedron? Of course, it is not the convex hull of a finite set, but it
can be expressed as the Minkowski sum P + C, where P is a V -poly-
tope and C is a convex cone described as the convex hull of finitely
many rays emanating from 0.

Exercises
1. Verify that a d-dimensional simplex in Rd can be expressed as the inter-
section of d+1 half-spaces. 2
2. (a) Show that every convex polytope in Rd is an orthogonal projection
of a simplex of a sufficiently large dimension onto the space Rd (which
we consider embedded as a d-flat in some Rn ). 3
(b) Prove that every convex polytope P symmetric about 0 (i.e., with
P = P ) is the affine image of a crosspolytope of a sufficiently large
dimension. 3

5.3 Faces of a Convex Polytope


The surface of the 3-dimensional cube consists of 8 corner points called
vertices, 12 edges, and 6 squares called facets. According to the perhaps more
usual terminology in 3-dimensional geometry, the facets would be called faces.
But in the theory of convex polytopes, the word face has a slightly different
meaning, defined below. For the cube, not only the squares but also the
vertices and the edges are all called faces of the cube.
5.3.1 Definition (Face). A face of a convex polytope P is defined as
either P itself, or
a subset of P of the form P h, where h is a hyperplane such that P is
fully contained in one of the closed half-spaces determined by h.
5.3 Faces of a Convex Polytope 27

P
h

We observe that each face of P is a convex polytope. This is because P is


the intersection of finitely many half-spaces and h is the intersection of two
half-spaces, so the face is an H-polyhedron, and moreover, it is bounded.
If P is a polytope of dimension d, then its faces have dimensions 1, 0,
1, . . ., d, where 1 is, by definition, the dimension of the empty set. A face
of dimension j is also called a j-face.
Names of faces. The 0-faces are called vertices, the 1-faces are called
edges, and the (d1)-faces of a d-dimensional polytope are called facets. The
(d2)-faces of a d-dimensional polytope are ridges; in the familiar 3-dimen-
sional situation, edges = ridges. For example, the 3-dimensional cube has 28
faces in total: the empty face, 8 vertices, 12 edges, 6 facets, and the whole
cube.
The following proposition shows that each V -polytope is the convex hull
of its vertices, and that the faces can be described combinatorially: They are
the convex hulls of certain subsets of vertices. This includes some intuitive
facts such as that each edge connects two vertices.
A helpful notion is that of an extremal point of a set: For a set X Rd ,
a point x X is extremal if x 6 conv(X \ {x}).
5.3.2 Proposition. Let P Rd be a (bounded) convex polytope.
(i) (Vertices are extremal) The extremal points of P are exactly its ver-
tices, and P is the convex hull of its vertices.
(ii) (Face of a face is a face) Let F be a face of P . The vertices of F are
exactly those vertices of P that lie in F . More generally, the faces of F
are exactly those faces of P that are contained in F .
The proof is not essential for our further considerations, and it is given at
the end of this section (but Exercise 9 below illustrates that things are not
quite as simple as it might perhaps seem). The proposition has an appropriate
analogue for polyhedra, but in order to avoid technicalities, we treat the
bounded case only.
Examples. A d-dimensional simplex has been defined as the convex hull of
a (d+1)-point affinely independent set V . It is easy to see that each subset of
d+1
V determines a face of the simplex. Thus, there are k+1 faces of dimension
k, k = 1, 0, . . . , d, and 2d+1 faces in total.
The d-dimensional crosspolytope has V = {e1 , e1 , . . . , ed , ed } as the
vertex set. A proper subset F V determines a face if and only if there is
28 Chapter 5: Convex Polytopes

no i such that both ei F and ei F (Exercise 2). It follows that there


are 3d +1 faces, including the empty one and the whole crosspolytope.
The nonempty faces of the d-dimensional cube [1, 1]d correspond to
vectors v {1, 1, 0}d. The face corresponding to such v has the vertex
set {u {1, 1}d: ui = vi for all i with vi 6= 0}. Geometrically, the vector v
is the center of gravity of its face.
The face lattice. Let F (P ) be the set of all faces of a (bounded) convex
polytope P (including the empty face of dimension 1). We consider the
partial ordering of F (P ) by inclusion.
5.3.3 Definition (Combinatorial equivalence). Two convex polytopes
P and Q are called combinatorially equivalent if F (P ) and F (Q) are isomor-
phic as partially ordered sets.
We are going to state some properties of the partially ordered set F (P )
without proofs. These are not difficult and can be found in [Zie94].
It turns out that F (P ) is a lattice (a partially ordered set satisfying
additional axioms). We recall that this means the following two conditions:
Meets condition: For any two faces F, G F(P ), there exists a face
M F(P ), called the meet of F and G, that is contained in both F and
G and contains all other faces contained in both F and G.
Joins condition: For any two faces F, G F(P ), there exists a face
J F(P ), called the join of F and G, that contains both F and G and
is contained in all other faces containing both F and G.
The meet of two faces is their geometric intersection F G.
For verifying the joins and meets conditions, it may be helpful to know
that for a finite partially ordered set possessing the minimum element and the
maximum element, the meets condition is equivalent to the joins condition,
and so it is enough to check only one of the conditions.
Here is the face lattice of a 3-dimensional pyramid:
P

5
1234 125 235 345 145

3 12 23 34 14 15 25 35 45
4
1 2
P 1 2 3 4 5

The vertices are numbered 15, and the faces are labeled by the vertex sets.
5.3 Faces of a Convex Polytope 29

The face lattice is graded, meaning that every maximal chain has the same
length (the rank of a face F is dim(F )+1). Quite obviously, it is atomic: Every
face is the join of its vertices. A little less obviously, it is coatomic; that is,
every face is the meet (intersection) of the facets containing it. An important
consequence is that combinatorial type of a polytope is determined by the
vertexfacet incidences. More precisely, if we know the dimension and all
subsets of vertices that are vertex sets of facets (but without knowing the
coordinates of the vertices, of course), we can uniquely reconstruct the whole
face lattice in a simple and purely combinatorial way.
Face lattices of convex polytopes have several other nice properties, but no
full algebraic characterization is known, and the problem of deciding whether
a given lattice is a face lattice is algorithmically difficult (even for 4-dimen-
sional polytopes).
The face lattice can be a suitable representation of a convex polytope in
a computer. Each j-face is connected by pointers to its (j1)-faces and to
the (j+1)-faces containing it. On the other hand, it is a somewhat redundant
representation: Recall that the vertexfacet incidences already contain the
full information, and for some applications, even less data may be sufficient,
say the graph of the polytope.
The dual polytope. Let P be a convex polytope containing the origin in
its interior. Then the dual set P is also a polytope; we have verified this in
the proof of Theorem 5.2.2.
5.3.4 Proposition. For each j = 1, 0, . . . , d, the j-faces of P are in a
bijective correspondence with the (dj1)-faces of P . This correspondence
also reverses inclusion; in particular, the face lattice of P arises by turning
the face lattice of P upside down.
Again we refer to the readers diligence or to [Zie94] for a proof. Let us
examine a few examples instead.
Among the five regular Platonic solids, the cube and the octahedron are
dual to each other, the dodecahedron and the icosahedron are also dual, and
the tetrahedron is dual to itself. More generally, if we have a 3-dimensional
convex polytope and G is its graph, then the graph of the dual polytope
is the dual graph to G, in the usual graph-theoretic sense. The dual of a
d-simplex is a d-simplex, and the d-dimensional cube and the d-dimensional
crosspolytope are dual to each other.

We conclude with two notions of polytopes in general position.


5.3.5 Definition (Simple and simplicial polytopes). A polytope P is
called simplicial if each of its facets is a simplex (this happens, in particular, if
the vertices of P are in general position, but general position is not necessary).
A d-dimensional polytope P is called simple if each of its vertices is contained
in exactly d facets.
30 Chapter 5: Convex Polytopes

The faces of a simplex are again simplices, and so each proper face of a sim-
plicial polytope is a simplex. Among the five Platonic solids, the tetrahedron,
the octahedron, and the icosahedron are simplicial; and the tetrahedron, the
cube, and the dodecahedron are simple. Crosspolytopes are simplicial, and
cubes are simple. An example of a polytope that is neither simplicial nor
simple is the 4-sided pyramid used in the illustration of the face lattice.
The dual of a simple polytope is simplicial, and vice versa. For a simple
d-dimensional polytope, a small neighborhood of each vertex looks combina-
torially like a neighborhood of a vertex of the d-dimensional cube. Thus, for
each vertex v of a d-dimensional simple polytope, there are d edges emanat-
ing from v, and each k-tuple of these edgesuniquely determines one k-face
incident to v. Consequently, v belongs to kd k-faces, k = 0, 1, . . . , d.
Proof of Proposition 5.3.2. In (i) (vertices are extremal), we assume
that P is the convex hull of a finite point set. Among all such sets, we fix one
that is inclusion-minimal and call it V0 . Let Vv be the vertex set of P , and
let Ve be the set of all extremal points of P . We prove that V0 = Vv = Ve ,
which gives (i). We have Ve V0 by the definition of an extremal point.
Next, we show that Vv Ve . If v Vv is a vertex of P , then there is a
hyperplane h with P h = {v}, and all of P \ {v} lies in one of the open
half-spaces defined by h. Hence P \ {v} is convex, which means that v is an
extremal point of P , and so Vv Ve .
Finally we verify V0 Vv . Let v V0 ; by the inclusion-minimality of V0 ,
we get that v 6 C = conv(V0 \ {v}). Since C and {v} are disjoint compact
convex sets, they can be strictly separated by a hyperplane h. Let hv be the
hyperplane parallel to h and containing v; this hv has all points of V0 \ {v}
on one side.

C
v

h hv

We want to show that P hv = {v} (then v is a vertex of P , and we are


done). The set P \ hv = conv(V0 ) \ hv , being the intersection of a convex set
with an open half-space, is convex. Any segment vx, where x P \ hv , shares
only the point v with the hyperplane hv , and so (P \ hv ) {v} is convex as
well. Since this set contains V0 and is convex, it contains P = conv(V0 ), and
so P hv = {v} indeed.
As for (ii) (face of a face is a face), it is clear that a face G of P contained
in F is a face of F too (use the same witnessing hyperplane). For the reverse
direction, we begin with the case of vertices. By a consideration similar to
that at the end of the proof of (i), we see that F = conv(V )h = conv(V h),
5.3 Faces of a Convex Polytope 31

where h is a hyperplane defining F . Hence all the extremal points of F , which


by (i) are exactly the vertices of F , are in V .
Finally, let F be a face of P defined by a hyperplane h, and let G F be
a face of F defined by a hyperplane g within h; that is, g is a (d2)-dimen-
sional affine subspace of h with G = g F and with all of F on one side. Let
be the closed half-space bounded by h with P . We start rotating the
boundary h of around g in the direction such that the rotated half-space
0 still contains F .

G F
h g

h0 P

If we rotate by a sufficiently small amount, then all the vertices of P not


lying in F are still in the interior of 0 . At the same time, the interior of 0
contains all the vertices of F not lying in G, while all the vertices of G remain
on the boundary h0 of 0 . So h0 defines a face of P (since all of P is on one
side), and this face has the same vertex set as G, and so it equals G by the
first part of (ii) proved above. 2

Exercises
1. Verify that if V Rd is affinely independent, then each subset F V
determines a face of the simplex conv(V ). 2
2. Verify the description of the faces of the cube and of the crosspolytope
given in the text. 3
3. Consider the (n1)-dimensional permutahedron as defined in the intro-
duction to this chapter.
(a) Verify that it really has n! vertices corresponding to the permutations
of {1, 2, . . . , n}. 2
(b) Describe all faces of the permutahedron combinatorially (what sets
of permutations are vertex sets of faces?). 3
(c) Determine the dimensions of the faces found in (b). In particular, show
that the facets correspond to ordered partitions (A, B) of {1, 2, . . . , n},
A, B 6= , and count them. 3
4. Let P R4 = conv{ei ej : i, j = 1, 2, 3, 4, i 6= j}, where e1 , . . . , e4 is
the standard basis (this P is called the 24-cell). Describe the face lattice
of P and prove that P is combinatorially equivalent to P (in fact, P can
be obtained from P by an isometry and scaling). 3
5. Using Proposition 5.3.2, prove the following:
(a) If F is a face of a convex polytope P , then F is the intersection of P
with the affine hull of F . 2
(b) If F and G are faces of a convex polytope P , then F G is a face,
too. 1
32 Chapter 5: Convex Polytopes

6. Let P be a convex polytope in R3 containing the origin as an interior


point, and let F be a j-face of P , j = 0, 1, 2.
(a) Give a precise definition of the face F 0 of the dual polytope P cor-
responding to F (i.e., describe F 0 as a subset of R3 ). 2
(b) Verify that F 0 is indeed a face of P . 2
7. Let V Rd be the vertex set of a convex polytope and let U V . Prove
that U is the vertex set of a face of conv(V ) if and only if the affine hull
of U is disjoint from conv(V \ U ). 3
8. Prove that the graph of any 3-dimensional convex polytope is 3-connected;
i.e., removing any 2 vertices leaves the graph connected. 5
9. Let C be a convex set. Call a point x C exposed if there is a hyperplane
h with C h = {x} and all the rest of C on one side. For convex polytopes,
exposed points are exactly the vertices, and we have shown that any
extremal point is also exposed. Find an example of a compact convex set
C R2 with an extremal point that is not exposed. 3
10. (On extremal points) For a set X Rd , let ex(X) = {x X: x 6
conv(X \ {x})} denote the set of extremal points of X.
(a) Find a convex set C Rd with C 6= conv(ex(C)). 1
(b) Find a compact convex C R3 for which ex(C) is not closed. 4
(c) By modifying the proof of Proposition 5.3.2, prove that C = conv(ex(C))
for every compact convex C Rd (this is a finite-dimensional version of
the well known KreinMilman theorem). 4

5.4 Many Faces: The Cyclic Polytopes


A convex polytope P can be given to us by the list of vertices. How difficult
is it to recover the full face lattice, or, more modestly, a representation of P
as an intersection of half-spaces? The first question to ask is how large the
face lattice or the collection of half-spaces can be, compared to the number
of vertices. That is, what is the maximum total number of faces, or the
maximum number of facets, of a convex polytope in Rd with n vertices? The
dual question is, of course, the maximum number of faces or vertices of a
bounded intersection of n half-spaces in Rd .
Let fj = fj (P ) denote the number of j-faces of a polytope P . The vector
(f0 , f1 , . . . , fd ) is called the f -vector of P . We thus assume f0 = n and we
Pd
are interested in estimating the maximum value of fd1 and of k=0 fk .
In dimensions 2 and 3, the situation is simple and favorable. For d = 2, our
polytope is a convex polygon with n vertices and n edges, and so f0 = f1 = n,
f2 = 1. The f -vector is even determined uniquely.
A 3-dimensional polytope can be regarded as a drawing of a planar graph,
in our case with n vertices. By well-known results for planar graphs, we have
f1 3n6 and f2 2n4. Equalities hold if and only if the polytope is
simplicial (all facets are triangles).
5.4 Many Faces: The Cyclic Polytopes 33

In both cases the total number of faces is linear in n. But as the dimension
grows, polytopes become much more complicated. First of all, even the total
number of faces of the most innocent convex polytope, the d-dimensional
simplex, is exponential in d. But here we consider d fixed and relatively
small, and we investigate the dependence on the number of vertices n.
Still, as we will see, for every n 5 there is a 4-dimensional convex
polytope with  n vertices and with every two vertices connected by an edge,
i.e., with n2 edges! This looks counterintuitive, but our intuition is based
on the 3-dimensional case. In any fixed dimension d, the number of facets
can be of order nbd/2c , which is rather disappointing for someone wishing to
handle convex polytopes efficiently. On the other hand, complete desperation
is perhaps not appropriate: Certainly not all polytopes exhibit this very bad
behavior. For example, it is known that if we choose n points uniformly at
random in the unit ball B d , then the expected number of faces of their convex
hull is only o(n), for every fixed d.
It turns out that the number of faces for a given dimension and number of
vertices is the largest possible for so-called cyclic polytopes, to be introduced
next. First we define a very useful curve in Rd .
5.4.1 Definition (Moment curve). The curve = {(t, t2 , . . . , td ): t R}
in Rd is called the moment curve.
5.4.2 Lemma. Any hyperplane h intersects the moment curve in at most
d points. If there are d intersections, then h cannot be tangent to , and thus
at each intersection, passes from one side of h to the other.

Proof. A hyperplane h can be expressed by the equation ha, xi = b, or


in coordinates a1 x1 + a2 x2 + + ad xd = b. A point of has the form
(t, t2 , . . . , td ), and if it lies in h, we obtain a1 t + a2 t2 + + ad td b = 0. This
means that t is a root of a nonzero polynomial ph (t) of degree at most d,
and hence the number of intersections of h with is at most d. If there are d
distinct roots, then they must be all simple. At a simple root, the polynomial
ph (t) changes sign, and this means that the curve passes from one side of
h to the other. 2

As a corollary, we see that every d points of the moment curve are affinely
independent, for otherwise, we could pass a hyperplane through them plus
one more point of . So the moment curve readily supplies explicit examples
of point sets in general position.
5.4.3 Definition (Cyclic polytope). The convex hull of finitely many
points on the moment curve is called a cyclic polytope.
How many facets does a cyclic polytope have? Each facet is determined
by a d-tuple of vertices, and distinct d-tuples determine distinct facets. Here
is a criterion telling us exactly which d-tuples determine facets.
34 Chapter 5: Convex Polytopes

5.4.4 Proposition (Gales evenness criterion). Let V be the vertex set


of a cyclic polytope P considered with the linear ordering along the mo-
ment curve (larger vertices have larger values of the parameter t). Let F =
{v1 , v2 , . . . , vd } V be a d-tuple of vertices of P , where v1 < v2 < < vd .
Then F determines a facet of P if and only if for any two vertices u, v V \F ,
the number of vertices vi F with u < vi < v is even.

Proof. Let hF be the hyperplane affinely spanned by F . Then F determines


a facet if and only if all the points of V \ F lie on the same side of hF .
Since the moment curve intersects hF in exactly d points, namely at
the points of F , it is partitioned into d+1 pieces, say 0 , . . . , d , each lying
completely in one of the half-spaces, as is indicated in the drawing:
3 5
1
hF
0 2 4

Hence, if the vertices of V \ F are all contained in the odd-numbered pieces


1 , 3 , . . ., as in the picture, or if they are all contained in the even-numbered
pieces 0 , 2 , . . ., then F determines a facet. This condition is equivalent to
Gales criterion. 2
Now we can count the facets.
5.4.5 Theorem. The number of facets of a d-dimensional cyclic polytope
with n vertices (n d+1) is
   
n bd/2c n bd/2c 1
+ for d even, and
bd/2c bd/2c 1
 
n bd/2c 1
2 for d odd.
bd/2c
For fixed d, this has the order of magnitude nbd/2c .

Proof. The number of facets equals the number of ways of placing d black
circles and n d white circles in a row in such a way that we have an even
number of black circles between each two white circles.
Let us say that an arrangement of black and white circles is paired if any
contiguous segment of black circles has an even length (the arrangements
permitted by Gales criterion need not be paired because of the initial and
final segments). The number of paired arrangements of 2k black circles and
n 2k white circles is nk
k , since by deleting every second black circle we
get a one-to-one correspondence with selections of the positions of k black
circles among n k possible positions.
5.5 The Upper Bound Theorem 35

Let us return to the original problem, and first consider an odd d = 2k+1.
In a valid arrangement of circles, we must have an odd number of consecutive
black circles at the beginning or at the end (but not both). In the former case,
we delete the initial black circle, and we get a paired arrangement of 2k black
and n12k white circles. In the latter case, we similarly delete the black
circle at the end and again get a paired arrangement as in the first case. This
establishes the formula in the theorem for odd d.
For even d = 2k, the number of initial consecutive black circles is ei-
ther odd or even.  In the even case, we have a paired arrangement, which
contributes nk k possibilities. In the odd case, we also have an odd num-
ber of consecutive black circles at the end, and so by deleting the first and
last black circles we obtain a paired arrangement of 2(k1) black circles and
n2k white circles. This contributes nk2

k1 possibilities. 2

Bibliography and remarks. The convex hull of the moment curve


was studied by by Caratheodory [Car07]. In the 1950s, Gale con-
structed neighborly polytopes by induction. Cyclic polytopes and the
evenness criterion appear in Gale [Gal63]. The moment curve is an
important object in many other branches besides the theory of convex
polytopes. For example, in elementary algebraic topology it is used
for proving that every (at most countable) d-dimensional simplicial
complex has a geometric realization in R2d+1 .

Exercises
1. (a) Show that if V is a finite subset of the moment curve, then all the
points of V are extreme in conv(V ); that is, they are vertices of the
corresponding cyclic polytope. 2
(b) Show that any two cyclic polytopes in Rd with n vertices are com-
binatorially the same: They have isomorphic face lattices. Thus, we can
speak of the cyclic polytope. 3
2. Show that for cyclic polytopes in dimensions 4 and higher, every pair
of vertices is connected by an edge. For dimension 4 and two arbitrary
vertices, write out explicitly the equation of a hyperplane intersecting the
cyclic polytope exactly in this edge. 3
3. Determine the f -vector of a cyclic polytope with n vertices in dimensions
4, 5, and 6. 3

5.5 The Upper Bound Theorem


The upper bound theorem, one of the earlier major achievements of the theory
of convex polytopes, claims that the cyclic polytope has the largest possible
number of faces.
36 Chapter 5: Convex Polytopes

5.5.1 Theorem (Upper bound theorem). Among all d-dimensional con-


vex polytopes with n vertices, the cyclic polytope maximizes the number of
faces of each dimension.
In this section we prove only an approximate result, which gives the cor-
rect order of magnitude for the maximum number of facets.
5.5.2 Proposition (Asymptotic upper bound theorem).  A d-dimen-
n
sional convex polytope with n vertices has at most 2 bd/2c facets and no
d+1 n

more than 2 bd/2c faces in total. For d fixed, both quantities thus have
the order of magnitude nbd/2c .
First we establish this proposition for simplicial polytopes, in the following
form.
5.5.3 Proposition. Let P be a d-dimensional simplicial polytope. Then
(a) f0 (P ) + f1 (P ) + + fd (P ) 2d fd1 (P ), and
(b) fd1 (P ) 2fbd/2c1 (P ).
This implies Proposition 5.5.2 for simplicial polytopes, since the number
n

of (bd/2c1)-faces is certainly no bigger than bd/2c , the number of all bd/2c-
tuples of vertices.
Proof of Proposition 5.5.3. We Pdpass to the dual polytope P , which
is simple. Now we need to prove k=0 fk (P ) 2 f0 (P ) and f0 (P )
d

2fdd/2e (P ).
Each face of P has at least one vertex, and every vertex of a simple
d-polytope is incident to 2d faces, which gives the first inequality.
We now bound the number of vertices in terms of the number of dd/2e-
faces. This is the heart of the proof, and it shows where the mysterious
exponent bd/2c comes from.
Let us rotate the polytope P so that no two vertices share the xd -co-
ordinate (i.e., no two vertices have the same vertical level).
Consider a vertex v with the d edges emanating from it. By the pigeonhole
principle, there are at least dd/2e edges directed upwards or at least dd/2e
edges directed downwards. In the former case, every dd/2e-tuple of edges
going up determines a dd/2e-face for which v is the lowest vertex. In the
latter case, every dd/2e-tuple of edges going down determines a dd/2e-face
for which v is the highest vertex. Here is an illustration, unfortunately for
the not too interesting 3-dimensional case, showing a situation with 2 edges
going up and the corresponding 2-dimensional face having v as the lowest
vertex:

v
5.5 The Upper Bound Theorem 37

We have exhibited at least one dd/2e-face for which v is the lowest vertex or
the highest vertex. Since the lowest vertex and the highest vertex are unique
for each face, the number of vertices is no more than twice the number of
dd/2e-faces. 2
Warning. For simple polytopes, the total combinatorial complexity is pro-
portional to the number of vertices, and for simplicial polytopes it is pro-
portional to the number of facets (considering the dimension fixed, that is).
For polytopes that are neither simple nor simplicial, the number of faces of
intermediate dimensions can have larger order of magnitude than both the
number of facets and the number of vertices; see Exercise 1.
Nonsimplicial polytopes. To prove the asymptotic upper bound theorem,
it remains to deal with nonsimplicial polytopes. This is done by a perturba-
tion argument, similar to numerous other results where general position is
convenient for the proof but where we want to show that the result holds
in degenerate cases as well. In most instances in this book, the details of
perturbation arguments are omitted, but here we make an exception, since
the proof seems somewhat nontrivial.
5.5.4 Lemma. For any d-dimensional convex polytope P there exists a d-
dimensional simplicial polytope Q with f0 (P ) = f0 (Q) and fk (Q) fk (P )
for all k = 1, 2, . . . , d.

Proof. The basic idea is very simple: Move (perturb) every vertex of P by a
very small amount, in such a way that the vertices are in general position, and
show that each k-face of P gives rise to at least one k-face of the perturbed
polytope. There are several ways of doing this proof.
We process the vertices one by one. Let V be the vertex set of P and
let v V . The operation of -pushing v is as follows: We choose a point v 0
lying in the interior of P , at distance at most from v, and on no hyperplane
determined by the points of V , and we set V 0 = (V \ {v}) {v 0 }. If we
successively v -push each vertex v of the polytope, the resulting vertex set is
in general position and we have a simple polytope.
It remains to show that for any polytope P with vertex set V and any
v V , there is an > 0 such that -pushing v does not decrease the number
of faces.
Let U V be the vertex set of a k-face of P , 0 k d1, and let V 0
arise from V by -pushing v. If v 6 U , then no doubt, U determines a face of
conv(V 0 ), and so we assume that v U . First suppose that v lies in the affine
hull of U \ {v}; we claim that then U \ {v} determines a k-face of conv(V 0 ).
This follows easily from the criterion in Exercise 5.3.7: A subset U V is the
vertex set of a face of conv(V ) if and only if the affine hull of U is disjoint
from conv(V \ U ). We leave a detailed argument to the reader (one must use
the fact that v is pushed inside).
38 Chapter 5: Convex Polytopes

If v lies outside of the affine hull of U \ {v}, then we want to show that
U 0 = (U \ {v}) {v 0 } determines a k-face of conv(V 0 ). The affine hull of U
is disjoint from the compact set conv(V \ U ). If we move v continuously by
a sufficiently small amount, the affine hull of U moves continuously, and so
there is an > 0 such that if we move v within from its original position,
the considered affine hull and conv(V \ U ) remain disjoint. 2

Bibliography and remarks. The upper bound theorem was con-


jectured by Motzkin in 1957 and proved by McMullen [McM70]. Many
partial results have been obtained in the meantime. Perhaps most no-
tably, Klee [Kle64] found a simple proof for polytopes with not too few
vertices (at least about d2 vertices in dimension d). That proof applies
to simplicial complexes much more general than the boundary com-
plexes of simplicial polytopes: It works for Eulerian pseudomanifolds
and, in particular, for all simplicial spheres, i.e., simplicial complexes
homeomorphic to S d1 . Presentations of McMullens proof and Klees
proof can be found in Zieglers book [Zie94]. A nice variation was
described by Alon and Kalai [AK85].
The proof of Lemma 5.5.4 by pushing vertices inside is similar to
an argument in Klee [Kle64], but he proves more and presents the
proof in more detail.
Convex hull computation. What does it mean to compute the convex
hull of a given n-point set V Rd ? One possible answer, briefly
touched upon in the notes to Section 5.2, is to express conv(V ) as
the intersection of half-spaces and to compute the vertex sets of all
facets. (As we know, the face lattice can be reconstructed from this
information purely combinatorially; see Kaibel and Pfetsch [KP01]
for an efficient algorithm.) Of course, for some applications it may
be sufficient to know much less about the convex hull, say only the
graph of the polytope or only the list of its vertices, but here we will
discuss only algorithms for computing all the vertexfacet incidences
or the whole face lattice. For a more detailed overview of convex hull
algorithms see, e.g., Seidel [Sei97].
For the dimension d considered fixed, there is a quite simple and
practical randomized algorithm. It computes the convex hull of n
points in Rd in expected time O(nbd/2c + n log n) (Seidel [Sei91],
simplifying Clarkson and Shor [CS89]), and also a very complicated
but deterministic algorithm with the same asymptotic running time
(Chazelle [Cha93b]; somewhat simplified in Bronnimann, Chazelle,
and Matousek [BCM99]). This is worst-case optimal, since an n-vertex
polytope may have about nbd/2c facets. There are also output-sensitive
algorithms, whose running time depends on the total number f of faces
of the resulting polytope. Recent results in this direction, including an
algorithm that computes the convex hull of n points in general posi-
5.5 The Upper Bound Theorem 39

tion in Rd (d fixed) in time O(n log f + (nf )11/(bd/2c+1) (log n)c(d) ),


can be found in Chan [Cha00b].
Still, none of the known algorithms is theoretically fully satisfac-
tory, and practical computation of convex hulls even in moderate di-
mensions, say 10 or 20, can be quite challenging. Some of the algo-
rithms are too complicated and with too large constants hidden in the
asymptotic notation to be of practical value. Algorithms requiring gen-
eral position of the points are problematic for highly degenerate point
configurations (which appear in many applications), since small per-
turbations used to achieve general position often increase the number
of faces tremendously. Some of the randomized algorithms compute
intermediate polytopes that can have many more faces than the fi-
nal result. Often we are interested just in the vertexfacet incidences,
but many algorithms compute all faces, whose number can be much
larger, or even a triangulation of every face, which may again increase
the complexity. Such problems of existing algorithms are discussed in
Avis, Bremner, and Seidel [ABS97].
For actual computations, simple and theoretically suboptimal al-
gorithms are often preferable. One of them is the double-description
method mentioned earlier, and another algorithm that seems to be-
have well in many difficult instances is the reverse search of Avis and
Fukuda [AF92]. It enumerates the vertices of the intersection of a given
set H of half-spaces one by one, using quite small storage. Conceptu-
T
ally, one thinks of optimizing a generic linear function over H by a
simplex algorithm with Blands rule. This defines a spanning tree in
the graph of the polytope, and this tree is searched depth-first starting
from the optimum vertex, essentially by running the simplex algorithm
backwards. The main problem of this algorithm is with degenerate
vertices of high degree, which may correspond to an enormous number
of bases in the simplex algorithm.

Exercises
1. (a) Let P be a k-dimensional convex polytope in Rk , and Q an `-dimen-
sional convex polytope in R` . Show that the Cartesian product P Q
Rk+` is a convex polytope of dimension k + `. 2
(b) If F is an i-face of P , and G is a j-face of Q, i, j 0, then F G is
an (i + j)-face of P Q. Moreover, this yields all the nonempty faces of
P Q. 3
(c) Using the product of suitable polytopes, find an example of a fat-
lattice polytope, i.e., a polytope for which the total number of faces has
a larger order of magnitude than the number of vertices plus the number
of facets together (the dimension should be a constant). 3
40 Chapter 5: Convex Polytopes

(d) Show that the following yields a 5-dimensional fat-lattice polytope:


The convex hull of two regular n-gons whose affine hulls are skew 2-flats
in R5 . 3
For recent results on fat-lattice polytopes see Eppstein, Kuperberg, and
Ziegler [EKZ03].

5.7 Voronoi Diagrams


Consider a finite set P Rd . For each point p P , we define a region reg(p),
which is the sphere of influence of the point p: It consists of the points
x Rd for which p is the closest point among the points of P . Formally,

reg(p) = {x Rd : dist(x, p) dist(x, q) for all q P },

where dist(x, y) denotes the Euclidean distance of the points x and y. The
Voronoi diagram of P is the set of all regions reg(p) for p P . (More precisely,
it is the cell complex induced by these regions; that is, every intersection of
a subset of the regions is a face of the Voronoi diagram.) Here an example of
the Voronoi diagram of a point set in the plane:

(Of course, the Voronoi diagram is clipped by a rectangle so that it fits into a
finite page.) The points of P are traditionally called the sites in the context
of Voronoi diagrams.
5.7.1 Observation. Each region reg(p) is a convex polyhedron with at most
|P |1 facets.
Indeed, \
reg(p) = {x: dist(x, p) dist(x, q)}
qP \{p}

is an intersection of |P | 1 half-spaces. 2
5.7 Voronoi Diagrams 41

For d = 2, a Voronoi diagram of n points is a subdivision of the plane


into n convex polygons (some of them are unbounded). It can be regarded as
a drawing of a planar graph (with one vertex at the infinity, say), and hence
it has a linear combinatorial complexity: n regions, O(n) vertices, and O(n)
edges.
In the literature the Voronoi diagram also appears under various other
names, such as the Dirichlet tessellation.
Examples of applications. Voronoi diagrams have been reinvented and
used in various branches of science. Sometimes the connections are surprising.
For instance, in archaeology, Voronoi diagrams help study cultural influences.
Here we mention a few applications, mostly algorithmic.
(Post office problem or nearest neighbor searching) Given a point set
P in the plane, we want to construct a data structure that finds the point
of P nearest to a given query point x as quickly as possible. This prob-
lem arises directly in some practical situations or, more significantly, as
a subroutine in more complicated problems. The query can be answered
by determining the region of the Voronoi diagram of P containing x. For
this problem (point location in a subdivision of the plane), efficient data
structures are known; see, e.g., the book [dBvKOS97] or other introduc-
tory texts on computational geometry.
(Robot motion planning) Consider a disk-shaped robot in the plane. It
should pass among a set P of point obstacles, getting from a given start
position to a given target position and touching none of the obstacles.

If such a passage is possible at all, the robot can always walk along
the edges of the Voronoi diagram of P , except for the initial and final
segments of the tour. This allows one to reduce the robot motion problem
to a graph search problem: We define a subgraph of the Voronoi diagram
consisting of the edges that are passable for the robot.
(A nice triangulation: the Delaunay triangulation) Let P R2 be a finite
point set. In many applications one needs to construct a triangulation of
P (that is, to subdivide conv(P ) into triangles with vertices at the points
of P ) in such a way that the triangles are not too skinny. Of course, for
42 Chapter 5: Convex Polytopes

some sets, some skinny triangles are necessary, but we want to avoid
them as much as possible. One particular triangulation that is usually
very good, and provably optimal with respect to several natural criteria,
is obtained as the dual graph to the Voronoi diagram of P . Two points
of P are connected by an edge if and only if their Voronoi regions share
an edge.

If no 4 points of P lie on a common circle then this indeed defines a


triangulation, called the Delaunay triangulation2 of P ; see Exercise 5.
The definition extends to points sets in Rd in a straightforward manner.
(Interpolation) Suppose that f : R2 R is some smooth function whose
values are known to us only at the points of a finite set P R2 . We
would like to interpolate f over the whole polygon conv(P ). Of course,
we cannot really tell what f looks like outside P , but still we want a
reasonable interpolation rule that provides a nice smooth function with
the given values at P . Multidimensional interpolation is an extensive
semiempirical discipline, which we do not seriously consider here; we
explain only one elegant method based on Voronoi diagrams. To compute
the interpolated value at a point x conv(P ), we construct the Voronoi
diagram of P , and we overlay it with the Voronoi diagram of P {x}.

d
x
a
b c

The region of the new point x cuts off portions of the regions of some of
the old points. Let wp be the area of the part of reg(p) in the Voronoi
diagram of P that belongs to reg(x) after inserting x. The interpolated
value f (x) is
w
P p
X
f (x) = f (p).
pPqP wq

An analogous method can be used in higher dimensions, too.


2
Being a transcription from Russian, the spelling of Delaunays name varies in
the literature. For example, in crystallography literature he is usually spelled
Delone.
5.7 Voronoi Diagrams 43

Relation of Voronoi diagrams to convex polyhedra. We now show that


Voronoi diagrams in Rd correspond to certain convex polyhedra in Rd+1 .
First we define the unit paraboloid in Rd+1 :

U = {x Rd+1 : xd+1 = x21 + x22 + + x2d }.

For d = 1, U is a parabola in the plane.


In the sequel, let us imagine the space Rd as the hyperplane xd+1 = 0 in
R . For a point p = (p1 , . . . , pd ) Rd , let e(p) denote the hyperplane in
d+1

Rd+1 with equation

xd+1 = 2p1 x1 + 2p2 x2 + + 2pd xd p21 p22 p2d .

Geometrically, e(p) is the hyperplane tangent to the paraboloid U at the point


u(p) = (p1 , p2 , . . . , pd , p21 + + p2d ) lying vertically above p. It is perhaps
easier to remember this geometric definition of e(p) and derive its equation
by differentiation when needed. On the other hand, in the forthcoming proof
we start out from the equation of e(p), and as a by-product, we will see that
e(p) is the tangent to U at u(p) as claimed.
5.7.2 Proposition. Let p, x Rd be points and let u(x) be the point of U
vertically above x. Then u(x) lies above the hyperplane e(p) or on it, and the
vertical distance of u(x) to e(p) is 2 , where = dist(x, p).

U
u(x)

2
e(p)

u(p)
xd+1 = 0
p x

Proof. We just substitute into the equations of U and of e(p). The xd+1 -
coordinate of u(x) is x21 + + x2d , while the xd+1 -coordinate of the point
of e(p) above x is 2p1 x1 + + 2pd xd p21 p2d . The difference is
(x1 p1 )2 + + (xd pd )2 = 2 . 2

Let E(p) denote the half-space lying above the hyperplane e(p). Consider
an n-point set P Rd . By Proposition 5.7.2, x reg(p) holds if and only
if e(p) is vertically closest to U at x among all e(q), q P . Here is what we
have derived:
5.7.3 Corollary. The Voronoi
T diagram of P is the vertical projection of the
facets of the polyhedron pP E(p) onto the hyperplane xd+1 = 0. 2
44 Chapter 5: Convex Polytopes

Here is an illustration for a planar Voronoi diagram:

5.7.4 Corollary. The maximum total number of faces of all regions of the
Voronoi diagram of an n-point set in Rd is O(ndd/2e ).

Proof. We know that the combinatorial complexity of the Voronoi diagram


equals the combinatorial complexity of an H-polyhedron with at most n
facets in Rd+1 . By intersecting this H-polyhedron with a large simplex we
can obtain a bounded polytope with at most n+d+2 facets, and we have not
decreased the number of faces compared to the original H-polyhedron. Then
the dual version of the asymptotic upper bound theorem (Theorem 5.5.2)
implies that the total number of faces is O(ndd/2e ), since b(d+1)/2c = dd/2e.
2

The convex polyhedra in Rd+1 obtained from Voronoi diagrams in Rd


by the above construction are rather special, and so a lower bound for the
combinatorial complexity of convex polytopes cannot be automatically trans-
ferred to Voronoi diagrams. But it turns out that the number of vertices of a
Voronoi diagram on n points in Rd can really be of order ndd/2e (Exercise 2).
Let us remark that the trick used for transforming Voronoi diagrams to
convex polyhedra is an example of a more general technique, called lineariza-
tion or Veronese mapping, which will be discussed a little more in Section ??.
This method sometimes allows us to convert a problem about algebraic curves
or surfaces of bounded degree to a problem about k-flats in a suitable higher-
dimensional space.
The farthest-point Voronoi diagram. The projection of the H-poly-
hedron pP E(p)op , where op denotes the half-space opposite to , forms
T
the farthest-point Voronoi diagram, in which each point p P is assigned
the regions of points for which it is the farthest point. It can be shown that
all nonempty regions of this diagram are unbounded and they correspond
precisely to the points appearing on the surface of conv(P ).
5.7 Voronoi Diagrams 45

Bibliography and remarks. The concept of Voronoi diagrams in-


dependently emerged in various fields of science, for example as the
medial axis transform in biology and physiology, the WignerSeitz
zones in chemistry and physics, the domains of action in crystallo-
graphy, and the Thiessen polygons in meteorology and geography. Ap-
parently, the earliest documented reference to Voronoi diagrams is a
picture in the famous Principia Philosopiae by Descartes from 1644
(that picture actually seems to show a power diagram, a generalization
of the Voronoi diagram to sites with different strengths of influence).
Mathematically, Voronoi diagrams were first introduced by Dirichlet
[Dir50] and by Voronoi [Vor08] for the investigation of quadratic forms.
For more information on the interesting history and a surprising va-
riety of applications we refer to several surveys: Aurenhammer and
Klein [AK00], Aurenhammer [Aur91], and the book Okabe, Boots,
and Sugihara [OBS92]. Every computational geometry textbook also
has at least a chapter devoted to Voronoi diagrams, and most papers
on this subject appear in computational geometry.
The Delaunay triangulation (or, more correctly, the Delaunay tes-
sellation, since it need not be a triangulation in general) was first
considered by Voronoi as the dual to the Voronoi diagram, and later
by Delaunay [Del34] with the definition given in Exercise 5(b) below.
The Delaunay triangulation of a planar point set P optimizes sev-
eral quality measures among all triangulations of P : It maximizes the
minimum angle occurring in any triangle, minimizes the maximum
circumradius of the triangles, maximizes the sum of inradii, and so
on (see [AK00] for references). Such optimality properties can usually
be proved by local flipping. We consider an arbitrary triangulation T
of a given finite P R2 (say with no 4 cocircular points). If there
is a 4-point Q P such that conv(Q) is a quadrilateral triangulated
by two triangles of T but in such a way that these two triangles are
not the Delaunay triangulation of Q, then the diagonal of Q can be
flipped:

not locally locally


Delaunay
; Delaunay

It can be shown that every sequence of such local flips is finite and
finishes with the Delaunay triangulation of P (Exercise 7). This pro-
cedure has an analogue in higher dimensions, where it gives a simple
and practically successful algorithm for computing Delaunay trian-
gulations (and Voronoi diagrams); see, e.g., Edelsbrunner and Shah
[ES96].
46 Chapter 5: Convex Polytopes

Generalizations of Voronoi diagrams. The example in the text with


robot motion planning, as well as other applications, motivates var-
ious notions of generalized Voronoi diagrams. First, instead of the
Euclidean distance, one can take various other distance functions, say
the `p -metrics. Second, instead of the spheres of influence of points,
we can consider the spheres of influence of other sites, such as dis-
joint polygons (this is what we get if we have a circular robot moving
amidst polygonal obstacles). We do not attempt to survey the numer-
ous results concerning such generalizations, again referring to [AK00].
Results on the combinatorial complexity of Voronoi diagrams under
non-Euclidean metrics and/or for nonpoint sites will be mentioned in
the notes to Section ??.
In another, very general, approach to Voronoi diagrams, one takes
the Voronoi diagram induced by two objects as a primitive notion. So
for every two objects we are given a partition of space into two regions
separated by a bisector, and Voronoi diagrams for more than two ob-
jects are built using the 2-partitions for all pairs. If one postulates a
few geometric properties of the bisectors, one gets a reasonable theory
of Voronoi diagrams (the so-called abstract Voronoi diagrams), includ-
ing efficient algorithms. So, for example, we do not even need a notion
of distance at this level of generality. Abstract Voronoi diagrams (in
the plane) were suggested by Klein [Kle89].
A geometrically significant generalization of the Euclidean Voronoi
diagram is the power diagram: Each point p P is assigned a real
weight w(p), and reg(P ) = {x Rd : kx pk2 w(p) kx qk2
w(q) for all q P }. While Voronoi diagrams in Rd are projections
of certain convex polyhedra in Rd+1 , the projection into Rd of every
intersection of finitely many nonvertical upper half-spaces in Rd+1 is
a power diagram. Moreover, a hyperplane section of a power diagram
is again a power diagram. Several other generalized Voronoi diagrams
in Rd (for example, with multiplicative weights of the sites) can be
obtained by intersecting a suitable power diagram in Rd+1 with a
simple surface and projecting into Rd , which yields fast algorithms;
see Aurenhammer and Imai [AI88].
Another generalization are higher-order Voronoi diagrams. The
kth-order Voronoi diagram of a finite point set P assigns to each k-
point T P the region reg(T ) consisting of all x Rd for which the
points of T are the k nearest neighbors of x in P . The usual Voronoi
diagram arises for k = 1, and the farthest-point Voronoi diagram for
k = |P | 1. The kth-order Voronoi diagram of P Rd is the projec-
tion of the kth level facets in the arrangement of the hyperplanes e(p),
p P (see Chapter 6 for these notions). Lee [Lee82] proved that the
kth-order Voronoi diagram of n points in the plane has combinato-
5.7 Voronoi Diagrams 47

rial complexity O(k(nk)); this is better than the maximum possible


complexity of level k in an arrangement of n arbitrary planes in R3 .
Applications of Voronoi diagrams are too numerous to be listed here,
and we add only a few remarks to those already mentioned in the
text. Using point location in Voronoi diagrams as in the post office
problem, several basic computational problems in the plane can be
solved efficiently, such as finding the closest pair in a point set or the
largest disk contained in a given polygon and not containing any of
the given points.
Besides providing good triangulations, the Delaunay triangulation
contains several other interesting graphs as subgraphs, such as a min-
imum spanning tree of a given point set (Exercise 6). In the plane,
this leads to an O(n log n) algorithm for the minimum spanning tree.
In R3 , subcomplexes of the Delaunay triangulation, the so-called -
complexes, have been successfully used in molecular modeling (see,
e.g., Edelsbrunner [Ede98]); they allow one to quickly answer ques-
tions such as, how many tunnels and voids are there in the given
molecule?
Robot motion planning using Voronoi diagrams (or, more gener-
ally, the retraction approach, where the whole free space for the robot
is replaced by some suitable low-dimensional skeleton) was first con-
sidered by ODunlaig and Yap [OY85]. Algorithmic motion planning
is an extensive discipline with innumerable variants of the problem.
For a brief introduction from the computational-geometric point of
view see, e.g., [dBvKOS97]; among several monographs we mention
Laumond and Overmars [LO96] and Latombe [Lat91].
The spatial interpolation of functions using Voronoi diagrams was
considered by Sibson [Sib81].

Exercises
1. Prove that the region reg(p) of a point p in the Voronoi diagram of a
finite point set P Rd is unbounded if and only if p lies on the surface
of conv(P ). 2
2. (a) Show that the Voronoi diagram of the 2n-point set {( ni , 0, 0): i =
1, 2, . . . , n} {(0, 1, nj ): j = 1, 2, . . . , n} in R3 has (n2 ) vertices. 3
(b) Let d = 2k+1 be odd, let e1 , . . . , ed be vectors of the standard
orthonormal basis in Rd , and let e0 stand for the zero vector. For
i = 0, 1, . . . , k and j = 1, 2, . . . , n, let pi,j = e2i + nj e2i+1 . Prove that
for every choice of j0 , j1 , . . . , jk {1, 2, . . . , n}, there is a point in Rd for
which the nearest points among the pi,j are exactly p0,j0 , p1,j1 , . . . , pk,jk .
Conclude that the Voronoi diagram of the pi,j has combinatorial com-
plexity (nk ) = (ndd/2e ). 3
48 Chapter 5: Convex Polytopes

3. (Voronoi diagram of flats) Let 1 , . . . , d1 be small distinct positive


numbers and for i = 1, 2, . . . , d1 and j = 1, 2, . . . , n, let Fi,j be the
(d2)-flat {x Rd : xi = j, xd = i }. For every choice of j1 , j2 , . . . , jd1
{1, 2, . . . , n}, find a point in Rd for which the nearest sites (under the
Euclidean distance) among the Fi,j are exactly F1,j1 , F2,j2 , . . . , Fd1,jd1 .
Conclude that the Voronoi diagram of the Fi,j has combinatorial com-
plexity (nd1 ). 3
This example is from Aronov [Aro02].
4. For a finite point set in the plane, define the farthest-point Voronoi di-
agram as indicated in the text, verify the claimed correspondence with
a convex polyhedron in R3 , and prove that all nonempty regions are
unbounded. 3
5. (Delaunay triangulation) Let P be a finite point set in the plane with no
3 points collinear and no 4 points cocircular.
(a) Prove that the dual graph of the Voronoi diagram of P , where two
points p, q P are connected by a straight edge if and only if the bound-
aries of reg(p) and reg(q) share a segment, is a plane graph where the
outer face is the complement of conv(P ) and every inner face is a triangle.
3

(b) Define a graph on P as follows: Two points p and q are connected


by an edge if and only if there exists a circular disk with both p and q
on the boundary and with no point of P in its interior. Prove that this
graph is the same as in (a), and so we have an alternative definition of
the Delaunay triangulation. 3
6. (Delaunay triangulation and minimum spanning tree) Let P R2 be a
finite point set with no 3 points collinear and no 4 cocircular. Let T be a
spanning tree of minimum total edge length in the complete graph with
the vertex set P , where the length of an edge is just its Euclidean length.
Prove that all edges of T are also edges of the Delaunay triangulation of
P. 3
7. (Delaunay triangulation by local flipping) Let P R2 be an n-point set
with no 3 points collinear and no 4 cocircular. Let T be an arbitrary
triangulation of conv(P ). Suppose that triangulations T1 , T2 , . . . are ob-
tained from T by successive local flips as described in the notes above (in
each step, we select a convex quadrilateral in the current triangulation
partitioned into two triangles in a way that is not the Delaunay triangu-
lation of the four vertices and we flip the diagonal of the quadrilateral).
(a) Prove that the sequence of triangulations is always finite (and give
as good an estimate for its maximum length as you can). 3
(b) Show that if no local flipping is possible, then the current triangula-
tion is the Delaunay triangulation of P . 4
8. Consider a finite set of disjoint segments in the plane. What types of
curves may bound the regions in their Voronoi diagram? The region of a
5.7 Voronoi Diagrams 49

given segment is the set of points for which this segment is a closest one.
2

9. Let A and B be two finite point sets in the plane. Choose a0 A arbi-
trarily. Having defined a0 , . . . , ai and b1 , . . . , bi1 , define bi+1 as a point
of B \ {b1 , . . . , bi } nearest to ai , and ai+1 as a point of A \ {a0 , . . . , ai }
nearest to bi+1 . Continue until one of the sets becomes empty. Prove that
at least one of the pairs (ai , bi+1 ), (bi+1 , ai+1 ), i = 0, 1, 2, . . ., realizes the
shortest distance between a point of A and a point of B. (This was used
by Eppstein [Epp95] in some dynamical geometric algorithms.) 3
10. (a) Let C be any circle in the plane x3 = 0 (in R3 ). Show that there exists
a half-space h such that C is the vertical projection of the set h U onto
x3 = 0, where U = {x R3 : x3 = x21 + x22 } is the unit paraboloid. 1
(b) Consider n arbitrary circular disks K1 , . . . , Kn in the plane. Show that
there exist only O(n) intersections of their boundaries that lie inside no
other Ki (this means that the boundary of the union of the Ki consists
of O(n) circular arcs). 3
11. Define a spherical polytope as an intersection of n balls in R3 (such
an object has facets, edges, and vertices similar to an ordinary convex
polytope).
(a) Show that any such spherical polytope in R3 has O(n2 ) faces. You
may assume that the balls are in general position. 4
(b) Find an example of an intersection of n balls having quadratically
many vertices. 3
(c) Show that the intersection of n unit balls has O(n) complexity only.
4
6

Number of Faces in
Arrangements

Arrangements of lines in the plane and their higher-dimensional generaliza-


tion, arrangements of hyperplanes in Rd , are a basic geometric structure
whose significance is comparable to that of convex polytopes. In fact, ar-
rangements and convex polytopes are quite closely related: A cell in a hyper-
plane arrangement is a convex polyhedron, and conversely, each hyperplane
arrangement in Rd corresponds canonically to a convex polytope in Rd+1
of a special type, the so-called zonotope. But as is often the case with dif-
ferent representations of the same mathematical structure, convex polytopes
and arrangements of hyperplanes emphasize different aspects of the structure
and lead to different questions.
Whenever we have a problem involving a finite point set in Rd and parti-
tions of the set by hyperplanes, we can use geometric duality, and we obtain
a problem concerning a hyperplane arrangement. Arrangements appear in
many other contexts as well; for example, some models of molecules give rise
to arrangements of spheres in R3 , and automatic planning of the motion of
a robot among obstacles involves, implicitly or explicitly, arrangements of
surfaces in higher-dimensional spaces.
Arrangements of hyperplanes have been investigated for a long time from
various points of view. In several classical areas of mathematics one is mainly
interested in topological and algebraic properties of the whole arrangement.
Hyperplane arrangements are related to such marvelous objects as Lie alge-
bras, root systems, and Coxeter groups. In the theory of oriented matroids
one studies the systems of sign vectors associated to hyperplane arrangements
in an abstract axiomatic setting.
We are going to concentrate on estimating the combinatorial complexity
(number of faces) in arrangements and neglect all the other directions.
52 Chapter 6: Number of Faces in Arrangements

6.1 Arrangements of Hyperplanes


We recall from Section 4.1 that for a finite set H of lines in the plane, the
arrangement of H is a partition of the plane into relatively open convex
subsets, the faces of the arrangement. In this particular case, the faces are
the vertices (0-faces), the edges (1-faces), and the cells (2-faces).1
An arrangement of a finite set H of hyperplanes in Rd is again a partition
of Rd into relatively open convex faces. Their dimensions are 0 through d. As
in the plane, the 0-faces are called vertices, the 1-faces edges, and the d-faces
cells. Sometimes the (d1)-faces are referred to asSfacets.
The cells are the connected components of Rd \ H. To obtain the facets,
we consider the (d1)-dimensional arrangements induced in the hyperplanes
of H by their intersections with the other hyperplanes. That is, for each
h H we take the connected components of h \ h0 H: h0 6=h h0 . To obtain
S
k-faces, we consider every possible k-flat L defined as the intersection of some
dk hyperplanes of H. The k-faces S of the arrangement lying within L are
the connected components of L \ (H \ HL ), where HL = {h H: L h}.
Remark on sign vectors. A face of the arrangement of H can be described
by its sign vector. First we need to fix the orientation of each hyperplane
h H. Each h H partitions Rd into three regions: h itself and the two
open half-spaces determined by it. We choose one of these open half-spaces as
positive and denote it by h , and we let the other one be negative, denoted
by h .
Let F be a face of the arrangement of H. We define the sign vector of
F (with respect to the chosen orientations of the hyperplanes) as (F ) =
(h : h H), where
+1 if F h ,

h = 0 if F h,
1 if F h .

The sign vector determines the face F , since we have F = hH hh , where


T
h0 = h, h1 = h , and h1 = h . The following drawing shows the sign
vectors of the marked faces in a line arrangement. Only the signs are shown,
and the positive half-planes lie above their lines.

h1 0+00
++ 00+

0
h2 h4 h3

1
This terminology is not unified in the literature. What we call faces are sometimes
referred to as cells (0-cells, 1-cells, and 2-cells).
6.1 Arrangements of Hyperplanes 53

Of course, not all possible sign vectors correspond to nonempty faces. For n
lines, there are 3n sign vectors but only O(n2 ) faces, as we will derive below.
Counting the cells in a hyperplane arrangement. We want to count
the maximum number of faces in an arrangement of n hyperplanes in Rd . As
we will see, this is much simpler than the similar task for convex polytopes!
If a set H of hyperplanes is in general position, which means that the
intersection of every k hyperplanes is (dk)-dimensional, k = 2, 3, . . . , d+1,
the arrangement of H is called simple. For |H| d+1 it suffices to require that
every d hyperplanes intersect at a single point and no d+1 have a common
point.
Every d-tuple of hyperplanes in a simple arrangement determines exactly
one vertex, and so a simple arrangement of n hyperplanes has exactly nd
vertices. We now calculate the number of cells; it turns out that the order of
magnitude is also nd for d fixed.
6.1.1 Proposition. The number of cells (d-faces) in a simple arrangement
of n hyperplanes in Rd equals
     
n n n
d (n) = + + + . (6.1)
0 1 d

First proof. We proceed by induction on the dimension d and the number


of hyperplanes n. For d = 1 we have a line and n points in it. These divide the
line into n+1 one-dimensional pieces, and formula (6.1) holds. (The formula
is also correct for n = 0 and all d 1, since the whole space, with no
hyperplanes, is a single cell.)
Now suppose that we are in dimension d, we have n1 hyperplanes, and
we insert another one. Since we assume general position, the n1 previous
hyperplanes divide the newly inserted hyperplane h into d1 (n1) cells by
the inductive hypothesis. Each such (d1)-dimensional cell within h parti-
tions one d-dimensional cell into exactly two new cells. The total increase in
the number of cells caused by inserting h is thus d1 (n1), and so

d (n) = d (n 1) + d1 (n 1).

Together with the initial conditions (for d = 1 and for n = 0), this recurrence
determines all values of , and so it remains to check that formula (6.1)
satisfies the recurrence. We have

d (n 1) + d1 (n 1) = n1
  n1
+ n1

0 + 1 0
h i
+ n1 n1 n1 n1
   
2 + 1 + + d + d1

= n1 n n n
   
0 + 1 + 2 + + d = d (n).

2
54 Chapter 6: Number of Faces in Arrangements

Second proof. This proof looks simpler, but a complete rigorous presenta-
tion is perhaps somewhat more demanding.
We proceed by induction on d, the case d = 0 being trivial. Let H be a set
of n hyperplanes in Rd in general position; in particular, we assume that no
hyperplane of H is horizontal and no two vertices of the arrangement have
the same vertical level (xd -coordinate).
Let g be an auxiliary horizontal hyperplane lying below all the vertices.
A cell of the arrangement of H either is bounded from below, and in this
case it has a unique lowest vertex, or is not bounded from below, and then it
intersects g. The number of cells of the former type is the same as the number
of vertices, which is nd . The cells of the latter type correspond to the cells
in the (d1)-dimensional arrangement induced within g by the hyperplanes
of H, and their number is thus d1 (n). 2

What is the number of faces of the intermediate dimensions 1, 2, . . . , d1


in a simple arrangement of n hyperplanes? This is not difficult to calculate
using Proposition 6.1.1 (Exercise 1); the main conclusion is that the total
number of faces is O(nd ) for a fixed d.
What about nonsimple arrangements? It turns out that a simple arrange-
ment of n hyperplanes maximizes the number of faces of each dimension
among arrangements of n hyperplanes. This can be verified by a perturbation
argument, which is considerably simpler than the one for convex polytopes
(Lemma 5.5.4), and which we omit.

Bibliography and remarks. The paper of Steiner [Ste26] from 1826


gives formulas for the number of faces in arrangements of lines, circles,
planes, and spheres. Of course, his results have been extended in many
ways since then (see, e.g., Zaslavsky [Zas75]). An early monograph on
arrangements is Grunbaum [Gru72].
The questions considered in the subsequent sections, such as the
combinatorial complexity of certain parts of arrangements, have been
studied mainly in the last twenty years or so. A recent survey dis-
cussing a large part of the material of this chapter and providing many
more facts and references is Agarwal and Sharir [AS00].
The algebraic and topological investigation of hyperplane arrange-
ments (both in real and complex spaces) is reflected in the book Orlik
and Terao [OT91]. Let us remark that in these areas, one usually
considers central arrangements of hyperplanes, where all the hyper-
planes pass through the origin (and so they are linear subspaces of
the underlying vector space). If such a central arrangement in Rd is
intersected with a generic hyperplane not passing through the origin,
one obtains a (d1)-dimensional affine arrangement such as those
considered by us. The correspondence is bijective, and so these two
views of arrangements are not very different, but for many results, the
formulation with central arrangements is more elegant.
6.2 Arrangements of Other Geometric Objects 55

The correspondence of arrangements to zonotopes is thoroughly


explained in Ziegler [Zie94].

Exercises
1. (a) Count the number of faces of dimensions 1 and 2 for a simple ar-
rangement of n planes in R3 . 2
(b) Express the number of k-faces in a simple arrangement of n hyper-
planes in Rd . 2
2. Prove that the number of unbounded cells in an arrangement of n hyper-
planes in Rd is O(nd1 ) (for a fixed d). 2
3. (a) Check that an arrangement of d or fewer hyperplanes in Rd has no
bounded cell. 2
(b) Prove that an arrangement of d+1 hyperplanes in general position in
Rd has exactly one bounded cell. 3
4. How many d-dimensional cells are there in the arrangement of the d2


hyperplanes in Rd with equations {xi = xj }, where 1 i < j d? 3


5. How many d-dimensional cells are there in the arrangement of the hy-
perplanes in Rd with the equations {xi xj = 0}, {xi xj = 1}, and
{xi xj = 1}, where 1 i < j d? 5
6. (Flags in arrangements)
(a) Let H be a set of n lines in the plane, and let V be the set of vertices
of their arrangement. Prove that the number of pairs (v, h) with v V ,
h H, and v h, i.e., the number of incidences I(V, H), is bounded by
O(n2 ). (Note that this is trivially true for simple arrangements.) 2
(b) Prove that the maximum number of d-tuples (F0 , F1 , . . . , Fd ) in an
arrangement of n hyperplanes in Rd , where Fi is an i-dimensional face
and Fi1 is contained in the closure of Fi , is O(nd ) (d fixed). Such d-tuples
are sometimes called flags of the arrangement. 3
7. Let P = {p1 , . . . , pn } be a point set in the plane. Let us say that points
x, y have the same view of P if the points of P are visible in the same
cyclic order from them. If rotating light rays emanate from x and from y,
the points of P are lit in the same order by these rays. We assume that
neither x nor y is in P and that neither of them can see two points of P
in occlusion.
(a) Show that the maximum possible number of points with mutually
distinct views of P is O(n4 ). 2
(b) Show that the bound in (a) cannot be improved in general. 4

6.2 Arrangements of Other Geometric Objects


Arrangements can be defined not only for hyperplanes but also for other
geometric objects. For example, what is the arrangement of a finite set H of
segments in the plane? As in the case of lines, it is a decomposition of the
56 Chapter 6: Number of Faces in Arrangements

plane into faces of dimension 0, 1, 2: the vertices, the edges, and the cells,
respectively. The vertices are the intersections of the segments, the edges are
the portions of the segments after removing S the vertices, and the cells (2-
faces) are the connected components of R2 \ H. (Note that the endpoints
of the segments are not included among the vertices.) While the cells of line
arrangements are convex polygons, those in arrangements of segments can be
complicated regions, even with holes:

It is almost obvious that the total number of faces of the arrangement of n


segments is at most O(n2 ). What is the maximum number of edges on the
boundary of a single cell in such an arrangements? This seemingly innocuous
question is surprisingly difficult, and most of Chapter ?? revolves around it.
Let us now present the definition of the arrangement for arbitrary sets
A1 , A2 , . . . , An Rd . The arrangement is a subdivision of space into con-
nected pieces again called the faces. Each face is an inclusion-maximal con-
nected set that crosses no boundary. More precisely, first we define an
equivalence relation on Rd : We put x y whenever x and y lie in the
same subcollection of the Ai , that is, whenever {i: x Ai } = {i: y Ai }.
So for each I {1, 2, . . . , n}, we have one possible equivalence class, namely
{x Rd : x Ai i I} (this is like a field in the Venn diagram of the Ai ).
But in typical geometric situations, most of the classes are empty. The faces
of the arrangement of the Ai are the connected components of the equivalence
classes. The reader is invited to check that for both hyperplane arrangements
and arrangements of segments this definition coincides with the earlier ones.
Arrangements of algebraic surfaces. Quite often one needs to con-
sider arrangements of the zero sets of polynomials. Let p1 (x1 , x2 , . . . , xd ),. . . ,
pn (x1 , x2 , . . . , xd ) be polynomials with real coefficients in d variables, and let
Zi = {x Rd : pi (x) = 0} be the zero set of pi . Let D denote the maximum
of the degrees of the pi ; when speaking of the arrangement of Z1 , . . . , Zn ,
one usually assumes that D is bounded by some (small) constant. Without
a bound on D, even a single Zi can have arbitrarily many connected compo-
nents.
In many cases, the Zi are algebraic surfaces, such as ellipsoids, paraboloids,
etc., but since we are in the real domain, sometimes they need not look like
surfaces at all. For example, the zero set of the polynomial p(x1 , x2 ) = x21 +x22
consists of the single point (0, 0). Although it is sometimes convenient to think
of the Zi as surfaces, the results stated below apply to zero sets of arbitrary
polynomials of bounded degree.
It is known that if both d and D are considered as constants, the maximum
number of faces in the arrangement of Z1 , Z2 , . . . , Zn as above is at most
6.2 Arrangements of Other Geometric Objects 57

O(nd ). This is one of the most useful results about arrangements, with many
surprising applications (a few are outlined below and in the exercises). In
the literature one often finds a (formally weaker) version dealing with sign
patterns of the polynomials pi . A vector {1, 0, +1}n is called a sign
pattern of p1 , p2 , . . . , pn if there exists an x Rd such that the sign of pi (x)
is i , for all i = 1, 2, . . . , n. Trivially, the number of sign patterns for any n
polynomials is at most 3n . For d = 1, it is easy to see that the actual number
of sign patterns is much smaller, namely at most 2nD + 1 (Exercise 1). It is
not so easy to prove, but still true, that there are at most C(d, D) nd sign
patterns in dimension d. This kind of result is generally called the Milnor
Thom theorem (with some more justice, it should also bear the names of
Oleinik and Petrovski, who proved slightly weaker results much earlier).
Here is a more precise (and more recent) version of this statement, where the
dependence on D and d is specified quite precisely.
6.2.1 Theorem (Number of sign patterns). Let p1 , p2 , . . . , pn be d-
variate real polynomials of degree at most D. The number of faces in the
arrangement of their zero sets Z1 , Z2 , . . . , Zn Rd , and consequently
Pd the
number of sign patterns of p1 , . . . , pn as well is at most 2(2D)d i=0 2i 4n+1

i .
For n d 2, this expression is bounded by
 d
50Dn
.
d

Proofs of these results are not included here because they would require
at least one more chapter. They belong to the field of real algebraic geometry.
The classical, deep, and extremely extensive field of algebraic geometry mostly
studies algebraic varieties over algebraically closed fields, such as the complex
numbers (and the questions of combinatorial complexity in our sense are
not among its main interests). Real algebraic geometry investigates algebraic
varieties and related concepts over the real numbers or other real-closed fields;
the presence of ordering and the missing roots of polynomials makes its flavor
distinctly different.
Arrangements of pseudolines. An arrangement of pseudolines is a nat-
ural generalization of an arrangement of lines. Lines are replaced by curves,
but we insist that these curves behave, in a suitable sense, like lines: For ex-
ample, no two of them intersect more than once. This kind of generalization
is quite different from, say, arrangements of planar algebraic curves, and so it
perhaps does not quite belong to the present section. But besides mentioning
pseudoline arrangements as a useful and interesting concept, we also need
them for a (typical) example of application of Theorem 6.2.1, and so we kill
two birds with one stone by discussing them here.
An (affine) arrangement of pseudolines can be defined as the arrangement
of a finite collection of curves in the plane that satisfy the following conditions:
58 Chapter 6: Number of Faces in Arrangements

(i) Each curve is x-monotone and unbounded in both directions; in other


words, it intersects each vertical line in exactly one point.
(ii) Every two of the curves intersect in exactly one point and they cross
at the intersection. (We do not permit parallel pseudolines, for they
would complicate the definition unnecessarily.)2
The curves are called pseudolines, but while being a line is an absolute no-
tion, being a pseudoline makes sense only with respect to a given collection
of curves.
Here is an example of a (simple) arrangement of 5 pseudolines:
5
4
3
2
1

Much of what we have proved for arrangements of lines is true for arrange-
ments of pseudolines as well. This holds for the maximum number of vertices,
edges, and cells, but also for more sophisticated results like the Szemeredi
Trotter theorem on the maximum number of incidences of m points and n
lines; these results have proofs that do not use any properties of straight lines
not shared by pseudolines.
One might be tempted to say that pseudolines are curves that behave
topologically like lines, but as we will see below, in at least one sense this is
profoundly wrong. The correct statement is that every two of them behave
topologically like two lines, but arrangements of pseudolines are more general
than arrangements of lines.
We should first point out that there is no problem with the local struc-
ture of the pseudolines, since each pseudoline arrangement can be redrawn
equivalently (in a sense defined precisely below) by polygonal lines, as a wiring
diagram:

2
This affine definition is a little artificial, and we use it only because we do
not want to assume the readers familiarity with the topology of the projective
plane. In the literature one usually considers arrangements of pseudolines in
the projective plane, where the definition is very natural: Each pseudoline is a
closed curve whose removal does not disconnect the projective plane, and every
two pseudolines intersect exactly once (which already implies that they cross at
the intersection point). Moreover, one often adds the condition that the curves
do not form a single pencil; i.e., not all of them have a common point, since
otherwise, one would have to exclude the case of a pencil in the formulation of
many theorems. But here we are not going to study pseudoline arrangements in
any depth.
6.2 Arrangements of Other Geometric Objects 59

5
4
3
2
1

The difference between pseudoline arrangements and line arrangements is of


a more global nature.
The arrangement of 5 pseudolines drawn above can be realized by straight
lines:
5
4

3
2
1

What is the meaning of realization by straight lines? To this end, we need


a suitable notion of equivalence of two arrangements of pseudolines. There
are several technically different possibilities; we again use an affine notion,
one that is very simple to state but not the most common. Let H be a col-
lection of n pseudolines. We number the pseudolines 1, 2, . . . , n in the order
in which they appear on the left of the arrangement, say from the bottom
to the top. For each i, we write down the numbers of the other pseudolines
in the order they are encountered along the pseudoline i from left to right.
For a simple arrangement we obtain a permutation i of {1, 2, . . . , n} \ {i}
for each i. For the arrangement in the pictures, we have 1 = (2, 3, 5, 4),
2 = (1, 5, 4, 3), 3 = (1, 5, 4, 2), 4 = (5, 1, 3, 2), and 5 = (4, 1, 3, 2). For
a nonsimple arrangement, some of the i are linear quasiorderings, meaning
that several consecutive numbers can be chunked together. We call two ar-
rangements affinely isomorphic if they yield the same 1 , . . . , n , i.e., if each
pseudoline meets the others in the same (quasi)order as the corresponding
pseudoline in the other arrangement. Two affinely isomorphic pseudoline ar-
rangements can be converted one to another by a suitable homeomorphism
of the plane.3

3
The more usual notion of isomorphism of pseudoline arrangements is defined for
arrangements in the projective plane. The arrangement of H is isomorphic to the
arrangement of H 0 if there exists a homeomorphism of the projective plane
onto itself such that each pseudoline h H is mapped to a pseudoline (h)
H 0 . For affinely isomorphic arrangements in the affine plane, the corresponding
arrangements in the projective plane are isomorphic, but the isomorphism in the
projective plane also allows for mirror reflection and for relocating the infinity.
Combinatorially, the isomorphism in the projective plane can be described using
the (quasi)orderings 1 , . . . , n as well. Here the i have to agree only up to
60 Chapter 6: Number of Faces in Arrangements

An arrangement of pseudolines is stretchable if it is affinely isomorphic to


an arrangement of straight lines.4 It turns out that all arrangements of 8 or
fewer pseudolines are stretchable, but there exists a nonstretchable arrange-
ment of 9 pseudolines:

p r
q

The proof of nonstretchability is based on the Pappus theorem in projective


geometry, which states that if 8 straight lines intersect as in the drawing, then
the points p, q, and r are collinear. By modifying this arrangement suitably,
one can obtain a simple nonstretchable arrangement of 9 pseudolines as well.
Next, we show that most of the simple pseudoline arrangements are non-
stretchable. The following construction shows that the number of isomor-
2
phism classes of simple arrangements of n pseudolines is at least 2(n ) :
gm
..
.
g2
g1
p2
p1
hm
..
.
h2
h1

We have m n3 , and the lines h1 , . . . , hm and g1 , . . . , gm form a regular grid.


Each of the about n3 pseudolines pi in the middle passes near (n) vertices of
this grid, and for each such vertex it has a choice of going below it or above.
2
This gives 2(n ) possibilities in total.
Now we use Theorem 6.2.1 to estimate the number of nonisomorphic sim-
ple arrangements of n straight lines. Let the lines be `1 , . . . , `n , where `i
has the equation y = ai x + bi and a1 > a2 > > an . The x-coordinate

a possible reversal and cyclic shift for each i, and also the numbering of the
pseudolines by 1, 2, . . . , n is not canonical.
We also remark that two arrangements of lines are isomorphic if and only if
the dual point configurations have the same order type, up to a mirror reflection
of the whole configuration (order types are discussed in Section ??).
4
For isomorphism in the projective plane, one gets an equivalent notion of stretch-
ability.
6.2 Arrangements of Other Geometric Objects 61

b b
of the intersection `i `j is aij aji . To determine the ordering i of the in-
tersections along `i , it suffices to know the ordering of the x-coordinates of
these intersections, and this can be inferred from the signs of the polynomials
pijk (ai , bi , aj , bj , ak , bk ) = (bi bj )(ak ai ) (bi bk )(aj ai ). So the num-
ber of nonisomorphic arrangements of n lines is no larger than the number
of possible sign patterns of the O(n3 ) polynomials pijk in the 2n variables
a1 , b1 , . . . , an , bn , and Theorem 6.2.1 yields the upper bound of 2O(n log n) . For
large n, this is a negligible fraction of the total number of simple pseudoline
arrangements. (Similar considerations apply to nonsimple arrangements as
well.)
The problem of deciding the stretchability of a given pseudoline arrange-
ment has been shown to be algorithmically difficult (at least NP-hard). One
can easily encounter this problem when thinking about line arrangements and
drawing pictures: What we draw by hand are really pseudolines, not lines,
and even with the help of a ruler it may be almost impossible to decide ex-
perimentally whether a given arrangement can really be drawn with straight
lines. But there are computational methods that can decide stretchability in
reasonable time at least for moderate numbers of lines.

Bibliography and remarks. A comprehensive account of real al-


gebraic geometry is Bochnak, Coste, and Roy [BCR98], and a more
recent treatment stressing an algoritmic approach is Basu, Pollack,
and Roy [BPR03]. Among the many available introductions to the
classical algebraic geometry we mention the lively book Cox, Little,
and OShea [CLO92].
The first bounds in the spirit of Theorem 6.2.1 were obtained by
Oleinik and Petrovski; in particular, Oleinik in her 1951 paper [Ole51]
showed that the maximum number of connected components, or more
generally the sum of the Betti numbers, of the zero set V Rd of a
polynomial of degree D is at most D(D 1)d1 , assuming that V is a
smooth hypersurface. Later Milnor [Mil64] proved the slightly worse
bound of D(2D 1)d1 with V an arbitrary real algebraic set, that
is, the set of common zeros in Rd of polynomials of degree at most D.
Similar results were obtained independently by Thom [Tho65]. Milnor
also established a bound for a basic semialgebraic set, that is, a set
defined by a conjunction of polynomial inequalities. Bounds on the
number of sign patterns and cells in arrangements, which were con-
sidered only later, can be derived from the MilnorThom results by
elementary reductions (see, e.g., the seminal paper of Ben-Or [BO83],
where lower bounds on the complexity of a number of basic algorith-
mic problems were derived, or Alon [Alo86]). Warren [War68] showed
that the number of d-dimensionalPd cells in the arrangement as in The-
orem 6.2.1 is at most 2(2D)d i=0 2i ni . The extension of Warrens
bounds to faces of all dimensions was obtained by Pollack and Roy
[PR93].
62 Chapter 6: Number of Faces in Arrangements

Sometimes we have polynomials in many variables, but we are in-


terested only in sign patterns attained at points that satisfy some
additional algebraic conditions. Such a situation is covered by a re-
sult of Basu, Pollack, and Roy [BPR96]: The number of sign patterns
attained by n polynomials of degree at most D on a k-dimensional
algebraic variety V Rd , where V can be defined by polynomials of
degree at most D, is at most nk O(D)d .
While bounding the number of sign patterns of multivariate poly-
nomials appears complicated, there is a beautiful short proof of an
almost tight bound on the number of zero patterns, due to Ronyai,
Babai, and Ganapathy [RBG01], which we now sketch (in the sim-
plest form, giving a slightly suboptimal result). A vector {0, 1}n is
a zero pattern of d-variate polynomials p1 , . . . , pn with coefficients in a
field F if there exists an x = x() F d with pi (x) = 0 exactly for the i
with i = 0. We show that if all the pi have degree  at most D, then the
Dn+d
number of zero patterns cannot Q exceed d . For each zero pattern
, let q be the polynomial i: i 6=0 pi . We have deg q Dn. Let us
consider the q as elements of the vector space L of all d-variate poly-
nomials over F of degree at most Dn. Using the basis of L consisting
of all monomials of degree at most Dn, we obtain dim L Dn+d

d . It
remains to verify that the q are linearly
P independent (assuming that
no pi is identically 0). Suppose that q = 0 with F not all
0. Choose a zero pattern with 6= 0 P and with the largest possible
number of 0s, and substitute x() into q . This yields = 0,
a contradiction.
Pseudoline arrangements. The founding paper is Levi [Lev26], where,
among others, the nonstretchable arrangement of 9 lines drawn above
was presented. A concise survey was written by Goodman [Goo97].
Pseudoline arrangements, besides being very natural, have also
turned out to be a fruitful generalization of line arrangements. Some
problems concerning line arrangements or point configurations were
first solved only in the more general setting of pseudoline arrange-
ments, and certain algorithms for line arrangements, the so-called
topological sweep methods, use an auxiliary pseudoline to speed up
the computation; see [Goo97].
Infinite families of pseudolines have been considered as well, and
even topological planes, which are analogues of the projective plane
but made of pseudolines. It is known that every finite configuration
of pseudolines can be extended to a topological plane, and there are
uncountably many distinct topological planes; see Goodman, Pollack,
Wenger, and Zamfirescu [GPWZ94].
Oriented matroids. The possibility of representing each pseudoline
arrangement by a wiring diagram makes it clear that a pseudoline ar-
rangement can also be considered as a purely combinatorial object.
6.2 Arrangements of Other Geometric Objects 63

The appropriate combinatorial counterpart of a pseudoline arrange-


ment is called an oriented matroid of rank 3. More generally, similar to
arrangements of pseudolines, one can define arrangements of pseudo-
hyperplanes in Rd , and these are combinatorially captured by oriented
matroids of rank d+1. Here the rank is one higher than the space di-
mension, because an oriented matroid of rank d is usually viewed as a
combinatorial abstraction of a central arrangement of hyperplanes in
Rd (with all hyperplanes passing through 0).
There are several different but equivalent definitions of an oriented
matroid. We present a definition in the so-called covector form. An
oriented matroid is a set V {1, 0, 1}n that is symmetric (v V
implies v V ), contains the zero vector, and satisfies the following
two more complicated conditions:
(Closed under composition) If u, v V , then u v V , where
(u v)i = ui if ui 6= 0 and (u v)i = vi if ui = 0.
(Admits elimination) If u, v V and j S(u, v) = {i: ui = vi 6=
0}, then there exists w V such that wj = 0 and wi = (u v)i for
all i 6 S(u, v).
The rank of an oriented matroid V is the largest r such that there is
an increasing chain v1 v2 vr , vi V , where u  v means
ui  vi for all i and where 0 1 and 0 1. At first sight, all this
may look quite mysterious, but it becomes much clearer if one thinks
of a basic example, where V is the set of sign vectors of all faces of a
central arrangement of hyperplanes in Rd .
It turns out that every oriented matroid of rank 3 corresponds to
an arrangement of pseudolines. More generally, Lawrences represen-
tation theorem asserts that every oriented matroid of rank d comes
from some central arrangement of pseudohyperplanes in Rd , and so
the purely combinatorial notion of oriented matroid corresponds, es-
sentially uniquely, to the topological notion of a (central) arrangement
of pseudohyperplanes.5
Oriented matroids are also naturally obtained from configurations
of points or vectors. In the notation of Section ?? (Gale transform), if
a is a sequence of n vectors in Rr , then both the sets sgn(LinVal(a))
and sgn(LinDep(a)) are oriented matroids in the sense of the above
definition. The first one has rank r, and the second, rank nr.
We are not going to say much more about oriented matroids, re-
ferring to Ziegler [Zie94] for a quick introduction and to Bjorner, Las
Vergnas, Sturmfels, White, and Ziegler [BVS+ 99] for a comprehensive
account.
5
The correspondence need not really be one-to-one. For example, the oriented
matroids of two projectively isomorphic pseudoline arrangements agree only up
to reorientation.
64 Chapter 6: Number of Faces in Arrangements

Stretchability. The following results illustrate the surprising difficulty


of the stretchability problem for pseudoline arrangements. They are
analogous to the statements about realizability of 4-dimensional con-
vex polytopes mentioned in Section 5.3, and they were actually found
much earlier.
Certain (simple) stretchable arrangements of n pseudolines require
coefficients with 2(n) digits in the equations of the lines, in every
straight-line realization (Goodman, Pollack, and Sturmfels [GPS90]).
Deciding the stretchability of a given pseudoline arrangement is NP-
hard (Shor [Sho91] has a relatively simple proof), and in fact, it is
polynomially equivalent to the problem of solvability of a system of
polynomial inequalities with integer coefficients. This follows from re-
sults of Mnev, published in Russian in 1985 (proofs were only sketched;
see [Mne89] for an English version). This work went unnoticed in the
West for some time, and so some of the results were rediscovered by
other authors.
Although detailed proofs of such theorems are technically demand-
ing, the principle is rather simple. Given two real numbers, suitably
represented by geometric quantities, one can produce their sum and
their product by classical geometric constructions by ruler. (Since ruler
constructions are invariant under projective transformations, the num-
bers are represented as cross-ratios.) By composing such constructions,
one can express the solvability of p(x1 , . . . , xn ) = 0, for a given n-
variate polynomial p with integer coefficients, by the stretchability of a
suitable arrangement in the projective plane. Dealing with inequalities
and passing to simple arrangements is somewhat more complicated,
but the idea is similar.
Practical algorithms for deciding stretchability have been studied
extensively by Bokowski and Sturmfels [BS89] and by Richter-Gebert
(see, e.g., [RG99]).
Mnev [Mne89] was mainly interested in the realization spaces of ar-
rangements. Let H be a fixed stretchable arrangement. Each straight-
line arrangement H 0 affinely isomorphic to H can be represented by
a point in R2n , with the 2n coordinates specifying the coefficients in
the equations of the lines of H 0 . Considering all possible H 0 for a given
H, we obtain a subset of R2n . For some time it was conjectured that
this set, the realization space of H, has to be path-connected, which
would mean that one straight-line realization could be converted to
any other by a continuous motion while retaining the affine isomor-
phism type.6 Not only is this false, but the realization space can have
arbitrarily many components. In a suitable sense, it can even have
6
In fact, these questions have been studied mainly for the isomorphism of arrange-
ments in the projective plane. There one has to be a little careful, since a mirror
reflection can easily make the realization space disconnected, and so the mirror
reflection (or the whole action of the general linear group) is factored out first.
6.2 Arrangements of Other Geometric Objects 65

arbitrary topological type. Whenever A Rn is a set definable by


a formula involving finitely many polynomial inequalities with inte-
ger coefficients, Boolean connectives, and quantifiers, there is a line
arrangement whose realization space S is homotopy equivalent to A
(Mnevs main result actually talks about the stronger notion of sta-
ble equivalence of S and A; see, e.g., [Goo97] or [BVS+ 99]). Similar
theorems were proved by Richter-Gebert for the realization spaces of
4-dimensional polytopes [RG99], [RG97].
These results for arrangements and polytopes can be regarded as
instances of a vague but probably quite general principle: Almost
none of the combinatorially imaginable geometric configurations are
geometrically realizable, and it is difficult to decide which ones are.
Of course, there are exceptions, such as the graphs of 3-dimensional
convex polytopes.
2
Encoding pseudoline arrangements. The lower bound 2(n ) for the
number of isomorphism classes of pseudoline arrangements is asymp-
totically tight. Felsner [Fel97] found a nice encoding of such an arrange-
ment by an n n matrix of 0s and 1s, from which the isomorphism
type can be reconstructed: The entry (i, j) of the matrix is 1 iff the jth
leftmost crossing along the pseudoline number i is with a pseudoline
whose number k is larger than i.

Exercises
1. Let p1 (x), . . . , pn (x) be univariate real polynomials of degree at most D.
Check that the number of sign patterns of the pi is at most 2nD+1. 2
2. (Intersection graphs) Let S be a set of n line segments in the plane. The
intersection graph of S is the graph on n vertices, which correspond to
the segments of S, with two vertices connected by an edge if and only if
the corresponding two segments intersect.
(a) Prove that the graph obtained from K5 by subdividing each edge
exactly once is not the intersection graph of segments in the plane (and
not even the intersection graph of any arcwise connected sets in the
plane). 4
(b) Use Theorem 6.2.1 to prove that most graphs are not intersection
graphs of segments: While the total number of graphs on n given vertices
is 2( 2 ) = 2n /2+O(n) , only 2O(n log n) of them are intersection graphs of
n 2

segments (be careful about collinear segments!). 3


(c) Show that the number of (isomorphism classes of) intersection graphs
of planar arcwise connected sets, and even of planar convex sets, on n
2
vertices is much larger: 2(n ) . (The right order of magnitude in both of
2
these cases is 2(3/8+o(1))n , as was determined by Pach and Toth [PT03].)
4
66 Chapter 6: Number of Faces in Arrangements

3. (Number of combinatorially distinct simplicial convex polytopes) Use


Theorem 6.2.1 to prove that for every dimension d 3 there exists Cd > 0
such that the number of combinatorial types of simplicial polytopes in
Rd with n vertices is at most 2Cd n log n . (The combinatorial equivalence
means isomorphic face lattices; see Definition 5.3.3.) 4
Such a result was proved by Alon [Alo86] and by Goodman and Pollack
[GP86].
4. (Sign patterns of matrices and rank) Let A be a real n n matrix. The
sign matrix (A) is the n n matrix with entries in {1, 0, +1} given
by the signs of the corresponding entries in A.
(a) Check that A has rank at most q if and only if there exist n q
matrices U and V with A = U V T . 3
(b) Estimate the number of distinct sign matrices of rank q using Theo-
rem 6.2.1, and conclude that there exists an n n matrix S containing
only entries +1 and 1 such that any real matrix A with (A) = S has
rank at least cn, with a suitable constant c > 0. 3
The result in (b) is from Alon, Frankl, and Rodl [AFR85] (for another
application see [Mat96]).
Bibliography

The references are sorted alphabetically by the abbreviations (rather than by


the authors names).
[ABS97] D. Avis, D. Bremner, and R. Seidel. How good are convex hull
algorithms? Comput. Geom. Theory Appl., 7:265302, 1997.
[AF92] D. Avis and K. Fukuda. A pivoting algorithm for convex hulls
and vertex enumeration of arrangements and polyhedra. Dis-
crete Comput. Geom., 8:295313, 1992.
[AFR85] N. Alon, P. Frankl, and V. Rodl. Geometrical realization of set
systems and probabilistic communication complexity. In Proc.
26th IEEE Symposium on Foundations of Computer Science,
pages 277280, 1985.
[AG86] N. Alon and E. Gyori. The number of small semispaces of a
finite set of points in the plane. J. Combin. Theory Ser. A,
41:154157, 1986.
[AI88] F. Aurenhammer and H. Imai. Geometric relations among
Voronoi diagrams. Geom. Dedicata, 27:6575, 1988.
[AK85] N. Alon and G. Kalai. A simple proof of the upper bound
theorem. European J. Combin., 6:211214, 1985.
[AK00] F. Aurenhammer and R. Klein. Voronoi diagrams. In J.-R. Sack
and J. Urrutia, editors, Handbook of Computational Geometry,
pages 201290. Elsevier Science Publishers B.V. North-Holland,
Amsterdam, 2000.
[Alo86] N. Alon. The number of polytopes, configurations, and real
matroids. Mathematika, 33:6271, 1986.
[AMS94] B. Aronov, J. Matousek, and M. Sharir. On the sum of squares
of cell complexities in hyperplane arrangements. J. Combin.
Theory Ser. A, 65:311321, 1994.
[AMS98] P. K. Agarwal, J. Matousek, and O. Schwarzkopf. Computing
many faces in arrangements of lines and segments. SIAM J.
Comput., 27(2):491505, 1998.
[APS93] B. Aronov, M. Pellegrini, and M. Sharir. On the zone of a
surface in a hyperplane arrangement. Discrete Comput. Geom.,
9(2):177186, 1993.
68 Bibliography

[Aro02] B. Aronov. A lower bound on Voronoi diagram complexity. Inf.


Process. Lett., 83:183185, 2002.
[AS00] P. K. Agarwal and M. Sharir. Arrangements and their appli-
cations. In J.-R. Sack and J. Urrutia, editors, Handbook of
Computational Geometry, pages 49119. North-Holland, Ams-
terdam, 2000.
[Aur91] F. Aurenhammer. Voronoi diagrams: A survey of a fundamental
geometric data structure. ACM Comput. Surv., 23(3):345405,
September 1991.
[BCM99] H. Bronnimann, B. Chazelle, and J. Matousek. Product range
spaces, sensitive sampling, and derandomization. SIAM J.
Comput., 28:15521575, 1999.
[BCR98] J. Bochnak, M. Coste, and M.-F. Roy. Real Algebraic Geometry.
Springer, Berlin etc., 1998. Transl. from the French, revised and
updated edition.
[BEPY91] M. Bern, D. Eppstein, P. Plassman, and F. Yao. Horizon the-
orems for lines and polygons. In J. Goodman, R. Pollack, and
W. Steiger, editors, Discrete and Computational Geometry: Pa-
pers from the DIMACS Special Year, volume 6 of DIMACS
Series in Discrete Mathematics and Theoretical Computer Sci-
ence, pages 4566. American Mathematical Society, Association
for Computing Machinery, Providence, RI, 1991.
[BK63] W. Bonnice and V. L. Klee. The generation of convex hulls.
Math. Ann., 152:129, 1963.
[BO83] M. Ben-Or. Lower bounds for algebraic computation trees. In
Proc. 15th Annu. ACM Sympos. Theory Comput., pages 8086,
1983.
[BPR96] S. Basu, R. Pollack, and M.-F. Roy. On the number of cells
defined by a family of polynomials on a variety. Mathematika,
43:120126, 1996.
[BPR03] S. Basu, R. Pollack, and M.-F. Roy. Algorithms in real algebraic
geometry. Algorithms and Computation in Mathematics 10.
Springer, Berlin, 2003.
[Bre93] G. Bredon. Topology and Geometry (Graduate Texts in Math-
ematics 139). Springer-Verlag, Berlin etc., 1993.
[Br83] A. Brnsted. An Introduction to Convex Polytopes. Springer-
Verlag, New York, NY, 1983.
[BS89] J. Bokowski and B. Sturmfels. Computational Synthetic Geom-
etry. Lect. Notes in Math. 1355. Springer-Verlag, Heidelberg,
1989.
[BV82] E. O. Buchman and F. A. Valentine. Any new Helly numbers?
Amer. Math. Mon., 89:370375, 1982.
[BVS+ 99] A. Bjorner, M. Las Vergnas, B. Sturmfels, N. White, and
G. M. Ziegler. Oriented Matroids (2nd edition). Encyclopedia
Bibliography 69

of Mathematics 46. Cambridge University Press, Cambridge,


1999.
[Car07] C. Caratheodory. Uber den Variabilitatsbereich der Koeffizien-
ten von Potenzreihen, die gegebene Werte nicht annehmen.
Math. Ann., 64:95115, 1907.
[CEG+ 90] K. Clarkson, H. Edelsbrunner, L. Guibas, M. Sharir, and
E. Welzl. Combinatorial complexity bounds for arrangements
of curves and spheres. Discrete Comput. Geom., 5:99160, 1990.
[CEG+ 93] B. Chazelle, H. Edelsbrunner, L. Guibas, M. Sharir, and
J. Snoeyink. Computing a face in an arrangement of line seg-
ments and related problems. SIAM J. Comput., 22:12861302,
1993.
[CEGS89] B. Chazelle, H. Edelsbrunner, L. Guibas, and M. Sharir. A
singly-exponential stratification scheme for real semi-algebraic
varieties and its applications. In Proc. 16th Internat. Colloq.
Automata Lang. Program., volume 372 of Lecture Notes Com-
put. Sci., pages 179192. Springer-Verlag, Berlin etc., 1989.
[CEM+ 96] K. L. Clarkson, D. Eppstein, G. L. Miller, C. Sturtivant, and
S.-H. Teng. Approximating center points with iterative Radon
points. Internat. J. Comput. Geom. Appl., 6:357377, 1996.
[CF90] B. Chazelle and J. Friedman. A deterministic view of random
sampling and its use in geometry. Combinatorica, 10(3):229
249, 1990.
[CGL85] B. Chazelle, L. J. Guibas, and D. T. Lee. The power of geo-
metric duality. BIT, 25:7690, 1985.
[Cha93a] B. Chazelle. Cutting hyperplanes for divide-and-conquer. Dis-
crete Comput. Geom., 9(2):145158, 1993.
[Cha93b] B. Chazelle. An optimal convex hull algorithm in any fixed
dimension. Discrete Comput. Geom., 10:377409, 1993.
[Cha00a] T. M. Chan. On levels in arrangements of curves. In Proc. 41st
IEEE Symposium on Foundations of Computer Science, pages
219227, 2000.
[Cha00b] T. M. Chan. Random sampling, halfspace range reporting, and
construction of ( k)-levels in three dimensions. SIAM J. Com-
put., 30(2):561575, 2000.
[Cha00c] B. Chazelle. The Discrepancy Method. Cambridge University
Press, Cambridge, 2000.
[Cla88a] K. L. Clarkson. Applications of random sampling in computa-
tional geometry, II. In Proc. 4th Annu. ACM Sympos. Comput.
Geom., pages 111, 1988.
[Cla88b] K. L. Clarkson. A randomized algorithm for closest-point
queries. SIAM J. Comput., 17:830847, 1988.
[CLO92] D. Cox, J. Little, and D. OShea. Ideals, Varieties, and Algo-
rithms. Springer-Verlag, New York, NY, 1992.
70 Bibliography

[CS89] K. L. Clarkson and P. W. Shor. Applications of random sam-


pling in computational geometry, II. Discrete Comput. Geom.,
4:387421, 1989.
[dBvKOS97] M. de Berg, M. van Kreveld, M. Overmars, and O. Schwarzkopf.
Computational Geometry: Algorithms and Applications.
Springer-Verlag, Berlin, 1997.
[Del34] B. Delaunay. Sur la sphere vide. A la memoire de Georges
Voronoi. Izv. Akad. Nauk SSSR, Otdelenie Matematicheskih i
Estestvennyh Nauk, 7:793800, 1934.
[DGK63] L. Danzer, B. Grunbaum, and V. Klee. Hellys theorem and its
relatives. In Convexity, volume 7 of Proc. Symp. Pure Math.,
pages 101180. American Mathematical Society, Providence,
1963.
[Dir50] G. L. Dirichlet. Uber die Reduktion der positiven quadratischen
Formen mit drei unbestimmten ganzen Zahlen. J. Reine Angew.
Math., 40:209227, 1850.
[Dol92] V. L. Dolnikov. A generalization of the ham sandwich the-
orem. Mat. Zametki, 52(2):2737, 1992. In Russian; English
translation in Math. Notes 52,2:771779, 1992.
[Eck93] J. Eckhoff. Helly, Radon and Caratheodory type theorems. In
P. M. Gruber and J. M. Wills, editors, Handbook of Convex
Geometry. North-Holland, Amsterdam, 1993.
[Ede98] H. Edelsbrunner. Geometry of modeling biomolecules. In P. K.
Agarwal, L. E. Kavraki, and M. Mason, editors, Proc. Workshop
Algorithmic Found. Robot. A. K. Peters, Natick, MA, 1998.
[EKZ03] D. Eppstein, G. Kuperberg, and G. M. Ziegler. Fat 4-polytopes
and fatter 3-spheres. In Discrete Geometry: In honor of W.
Kuperbergs 60th birthday (A. Bezdek, ed.), Pure and Applied
Mathematics Vol. 253, pages 239265. Marcel Dekker Inc., New
York, NY, 2003.
[EOS86] H. Edelsbrunner, J. ORourke, and R. Seidel. Constructing
arrangements of lines and hyperplanes with applications. SIAM
J. Comput., 15:341363, 1986.
[Epp95] D. Eppstein. Dynamic Euclidean minimum spanning trees and
extrema of binary functions. Discrete Comput. Geom., 13:111
122, 1995.
[ES96] H. Edelsbrunner and N. R. Shah. Incremental topological flip-
ping works for regular triangulations. Algorithmica, 15:223241,
1996.
[ESS93] H. Edelsbrunner, R. Seidel, and M. Sharir. On the zone theorem
for hyperplane arrangements. SIAM J. Comput., 22(2):418429,
1993.
Bibliography 71

[Far94] G. Farkas. Applications of Fouriers mechanical principle (in


Hungarian). Math. Termes. Ertesito, 12:457472, 1893/94. Ger-
man translation in Math. Nachr. Ungarn 12:127, 1895.
[Fel97] S. Felsner. On the number of arrangements of pseudolines.
Discrete Comput. Geom., 18:257267, 1997.
[Gal63] D. Gale. Neighborly and cyclic polytopes. In V. Klee, editor,
Convexity, volume 7 of Proc. Symp. Pure Math., pages 225232.
American Mathematical Society, 1963.
[GGL95] R. L. Graham, M. Grotschel, and L. Lovasz, editors. Handbook
of Combinatorics. North-Holland, Amsterdam, 1995.
[GJ00] E. Gawrilow and M. Joswig. polymake: a framework
for analyzing convex polytopes. In G. Kalai and G. M.
Ziegler, editors, PolytopesCombinatorics and Computation,
pages 4374. Birkhauser, Basel, 2000. Software available at
http://www.math.tu-berlin.de/diskregeom/polymake/.
[GO97] J. E. Goodman and J. ORourke, editors. Handbook of Discrete
and Computational Geometry. CRC Press LLC, Boca Raton,
FL, 1997.
[Goo97] J. E. Goodman. Pseudoline arrangements. In J. E. Goodman
and J. ORourke, editors, Handbook of Discrete and Computa-
tional Geometry, pages 83110. CRC Press LLC, Boca Raton,
FL, 1997.
[GP84] J. E. Goodman and R. Pollack. On the number of k-subsets
of a set of n points in the plane. J. Combin. Theory Ser. A,
36:101104, 1984.
[GP86] J. E. Goodman and R. Pollack. Upper bounds for configurations
and polytopes in <d . Discrete Comput. Geom., 1:219227, 1986.
[GPS90] J. E. Goodman, R. Pollack, and B. Sturmfels. The intrinsic
spread of a configuration in <d . J. Amer. Math. Soc., 3:639
651, 1990.
[GPWZ94] J. E. Goodman, R. Pollack, R. Wenger, and T. Zamfirescu.
Arrangements and topological planes. Amer. Math. Monthly,
101(10):866878, 1994.
[Gru67] B. Grunbaum. Convex Polytopes. John Wiley & Sons, New
York, NY, 1967.
[Gru72] B. Grunbaum. Arrangements and Spreads. Regional Conf. Ser.
Math. American Mathematical Society, Providence, RI, 1972.
[Gru03] B. Grunbaum. Convex Polytopes. Second edition prepared by
V. Kaibel, V. Klee, and G. M. Ziegler. Springer, Berlin, 2003.
[GW93] P. M. Gruber and J. M. Wills, editors. Handbook of Convex Ge-
ometry (volumes A and B). North-Holland, Amsterdam, 1993.
[JM94] S. Jadhav and A. Mukhopadhyay. Computing a centerpoint of
a finite planar set of points in linear time. Discrete Comput.
Geom., 12:291312, 1994.
72 Bibliography

[Kat78] M. Katchalski. A Helly type theorem for convex sets. Can.


Math. Bull., 21:121123, 1978.
[Kir03] P. Kirchberger. Uber Tschebyschefsche Annaherungsmethoden.
Math. Ann., 57:509540, 1903.
[Kle53] V. Klee. The critical set of a convex body. Amer. J. Math.,
75:178188, 1953.
[Kle64] V. Klee. On the number of vertices of a convex polytope. Cana-
dian J. Math, 16:701720, 1964.
[Kle89] R. Klein. Concrete and Abstract Voronoi Diagrams, volume
400 of Lecture Notes Comput. Sci. Springer-Verlag, Berlin etc.,
1989.
[Kol01] V. Koltun. Almost tight upper bounds for vertical decomposi-
tions in four dimensions. In Proc. 42nd IEEE Symposium on
Foundations of Computer Science, 2001.
[KP01] V. Kaibel and M. E. Pfetsch. Computing the face lattice of
a polytope from its vertex-facet incidences. Technical Report,
Inst. fur Mathematik, TU Berlin, 2001.
[KZ00] G. Kalai and G. M. Ziegler, editors. PolytopesCombinatorics
and Computation. DMV-seminar Oberwolfach, Germany,
November 1997. Birkhauser, Basel, 2000.
[Lat91] J.-C. Latombe. Robot Motion Planning. Kluwer Academic Pub-
lishers, Boston, 1991.
[Lee82] D. T. Lee. On k-nearest neighbor Voronoi diagrams in the
plane. IEEE Trans. Comput., C-31:478487, 1982.
[Lev26] F. Levi. Die Teilung der projektiven Ebene durch Gerade
oder Pseudogerade. Ber. Math.-Phys. Kl. sachs. Akad. Wiss.
Leipzig, 78:256267, 1926.
[LMS94] C.-Y. Lo, J. Matousek, and W. L. Steiger. Algorithms for ham-
sandwich cuts. Discrete Comput. Geom., 11:433, 1994.
[LO96] J.-P. Laumond and M. H. Overmars, editors. Algorithms for
Robotic Motion and Manipulation. A. K. Peters, Wellesley, MA,
1996.
[Mat96] J. Matousek. On the distortion required for embedding finite
metric spaces into normed spaces. Israel J. Math., 93:333344,
1996.
[McM70] P. McMullen. The maximal number of faces of a convex poly-
tope. Mathematika, 17:179184, 1970.
[Mil64] J. W. Milnor. On the Betti numbers of real algebraic varieties.
Proc. Amer. Math. Soc., 15:275280, 1964.
[Mne89] M. E. Mnev. The universality theorems on the classification
problem of configuration varieties and convex polytopes vari-
eties. In O. Y. Viro, editor, Topology and GeometryRohlin
Seminar, volume 1346 of Lecture Notes Math., pages 527544.
Springer, Berlin etc., 1989.
Bibliography 73

[MS71] P. McMullen and G. C. Shephard. Convex Polytopes and the


Upper Bound Conjecture, volume 3 of Lecture Notes. Cambridge
University Press, Cambridge, England, 1971.
[Mul93] K. Mulmuley. Computational Geometry: An Introduction
Through Randomized Algorithms. Prentice Hall, Englewood
Cliffs, NJ, 1993.
[OBS92] A. Okabe, B. Boots, and K. Sugihara. Spatial Tessellations:
Concepts and Applications of Voronoi Diagrams. John Wiley
& Sons, Chichester, UK, 1992.
[Ole51] O. A. Oleinik. Estimates of the Betti numbers of real algebraic
hypersurfaces (in Russian). Mat. Sbornik (N. S.), 28(70):635
640, 1951.
[OT91] P. Orlik and H. Terao. Arrangements of Hyperplanes. Springer-
Verlag, Berlin etc., 1991.
[OY85] C. ODunlaing and C. K. Yap. A retraction method for plan-
ning the motion of a disk. J. Algorithms, 6:104111, 1985.
[PR93] R. Pollack and M.-F. Roy. On the number of cells defined by a
set of polynomials. C. R. Acad. Sci. Paris, 316:573577, 1993.
[PT03] J. Pach and G. Toth. How many ways can one draw a graph?,
2003. Manuscript, accepted for publication in Combinatorica.
[Rad21] J. Radon. Mengen konvexer Korper, die einen gemeinsamen
Punkt enthalten. Math. Ann., 83:113115, 1921.
[Rad47] R. Rado. A theorem on general measure. J. London Math. Soc.,
21:291300, 1947.
[RBG01] L. Ronyai, L. Babai, and M. K. Ganapathy. On the number
of zero-patterns of a sequence of polynomials. J. Amer. Math.
Soc., 14(3):717735 (electronic), 2001.
[RG97] J. Richter-Gebert. Realization Spaces of Polytopes. Lecture
Notes in Mathematics 1643. Springer, Berlin, 1997.
[RG99] J. Richter-Gebert. The universality theorems for oriented ma-
troids and polytopes. In B. Chazelle et al., editors, Advances
in Discrete and Computational Geometry, Contemp. Math. 223,
pages 269292. Amer. Math. Soc., Providence, RI, 1999.
[Rud91] W. Rudin. Functional Analysis (2nd edition). McGraw-Hill,
New York, 1991.
[Sal75] G.T. Sallee. A Helly-type theorem for widths. In Geom. Metric
Lin. Spaces, Proc. Conf. East Lansing 1974, Lect. Notes Math.
490, pages 227232. Springer, Berlin etc., 1975.
[Sch01] L. Schlafli. Theorie der vielfachen Kontinuitat. Denkschriften
der Schweizerichen naturforschender Gesellschaft, 38:1237,
1901. Written in 185051. Reprinted in Ludwig Schlafli, 1814
1895, Gesammelte mathematische Abhandlungen, Birkhauser,
Basel 1950.
74 Bibliography

[Sch11] P. H. Schoute. Analytic treatment of the polytopes regularly de-


rived from the regular polytopes. Verhandelingen der Koning-
lijke Akademie van Wetenschappen te Amsterdam, 11(3), 1911.
[Sch86] A. Schrijver. Theory of Linear and Integer Programming.
Wiley-Interscience, New York, NY, 1986.
[Sei91] R. Seidel. Small-dimensional linear programming and convex
hulls made easy. Discrete Comput. Geom., 6:423434, 1991.
[Sei97] R. Seidel. Convex hull computations. In J. E. Goodman and
J. ORourke, editors, Handbook of Discrete and Computational
Geometry, chapter 19, pages 361376. CRC Press LLC, Boca
Raton, FL, 1997.
[Sha01] M. Sharir. The ClarksonShor technique revisited and ex-
tended. In Proc. 17th Annu. ACM Sympos. Comput. Geom.,
pages 252256, 2001.
[Sho91] P. W. Shor. Stretchability of pseudolines is NP-hard. In
P. Gritzman and B. Sturmfels, editors, Applied Geometry and
Discrete Mathematics: The Victor Klee Festschrift, volume 4 of
DIMACS Series in Discrete Mathematics and Theoretical Com-
puter Science, pages 531554. AMS Press, 1991.
[Sib81] R. Sibson. A brief description of natural neighbour interpola-
tion. In V. Barnet, editor, Interpreting Multivariate Data, pages
2136. John Wiley & Sons, Chichester, 1981.
[Ste26] J. Steiner. Einige Gesetze uber die Theilung der Ebene und des
Raumes. J. Reine Angew. Math., 1:349364, 1826.
[Ste16] E. Steinitz. Bedingt konvergente Reihen und konvexe Systeme
I; II; III. J. Reine Angew. Math, 143; 144; 146:128175; 140;
152, 1913; 1914; 1916.
[Ste85] H. Steinlein. Borsuks antipodal theorem and its generaliza-
tions and applications: a survey. In A. Granas, editor, Methodes
topologiques en analyse nonlineaire, pages 166235. Colloq.
Semin. Math. Super., Semin. Sci. OTAN (NATO Advanced
Study Institute) 95, Univ. de Montreal Press, Montreal, 1985.
[SU00] J.-R. Sack and J. Urrutia, editors. Handbook of Computational
Geometry. North-Holland, Amsterdam, 2000.
[Tho65] R. Thom. On the homology of real algebraic varieties (in
French). In S.S. Cairns, editor, Differential and Combinato-
rial Topology. Princeton Univ. Press, 1965.
[Vin39] P. Vincensini. Sur une extension dun theoreme de M. J. Radon
sur les ensembles de corps convexes. Bull. Soc. Math. France,
67:115119, 1939.
[Vor08] G. M. Voronoi. Nouvelles applications des parametres conti-
nus a la theorie des formes quadratiques. deuxieme Memoire:
Recherches sur les parallelloedres primitifs. J. Reine Angew.
Math., 134:198287, 1908.
Bibliography 75

[War68] H. E. Warren. Lower bound for approximation by nonlinear


manifolds. Trans. Amer. Math. Soc., 133:167178, 1968.
[Wel01] E. Welzl. Entering and leaving j-facets. Discrete Comput.
Geom., 25:351364, 2001.
[Zas75] T. Zaslavsky. Facing up to Arrangements: Face-Count Formulas
for Partitions of Space by Hyperplanes, volume 154 of Memoirs
Amer. Math. Soc. American Mathematical Society, Providence,
RI, 1975.
[Zie94] G. M. Ziegler. Lectures on Polytopes, volume 152 of Gradu-
ate Texts in Mathematics. Springer-Verlag, Heidelberg, 1994.
Corrected and revised printing 1998.
[ZV90] R.T. Zivaljevic and S.T. Vrecica. An extension of the ham
sandwich theorem. Bull. London Math. Soc., 22:183186, 1990.

You might also like