152 - Lectures On Polytopes PDF
152 - Lectures On Polytopes PDF
152 - Lectures On Polytopes PDF
Ziegler
Lectures on
Polytopes
Revised First Edition
Springer
Gunter M. Ziegler
Technische Universta Berlin
Fachbereich Mathematik, MA 6-1
Berlin, D10623
Germany
Editorial Board
S Axler
Department of
Mathematics
San Francisco State University
San Francisco, CA 94132
USA
F.W Gehring
Department of
Mathematics
University of Michigan
Ann Arbor, MI 48109
USA
K.A. Ribet
Department of
Mathematics
University of California
at Berkeley
Berkeley, CA 94720-3840
USA
ISBN
ISBN
ISBN
ISBN
0-387-94329-3 Springer-Verlag
3-540-94329-3 Springer-Verlag
0-387-94365-X Springer-Verlag
3-540-94365-X Springer-Verlag
Preface
The aim of this book is to introduce the reader to the fascinating world
of convex polytopes. The book developed from a course that I taught at
the Technische Universitat Berlin, as a part of the Graduierten-Kolleg "Algorithmische Diskrete Mathematik." I have tried to preserve some of the
flavor of lecture notes, and I have made absolutely no effort to hide my
enthusiasm for the mathematics presented, hoping that this will be enough
of an excuse for being "informal" at times.
There is no P2C2E in this book.*
Each of the ten lectures (or chapters, if you wish) ends with extra notes
and historical comments, and with exercises of varying difficulty, among
them a number of open problems (marked with an asterisk*), which I hope
many people will find challenging. In addition, there are lots of pointers to
interesting recent work, research problems, and related material that may
sidetrack the reader or lecturer, and are intended to do so.
Although these are notes from a two-hour, one-semester course, they
have been expanded so much that they will easily support a four-hour
course. The lectures (after the basics in Lectures 0 to 3) are essentially
independent from each other. Thus, there is material for quite different twohour courses in this book, such as a course on "duality, oriented matroids,
and zonotopes" (Lectures 6 and 7), or one on "polytopes and polyhedral
complexes" (Lectures 4, 5 and 9), etc.
*P2C2E = "Process too complicated to explain" [434j
vi
Preface
Contents
Preface
Preface to the Second Printing
0
v
vii
1
22
23
27
27
32
37
39
43
45
47
49
51
51
55
59
64
65
Faces of Polytopes
2.1 Vertices, Faces, and Facets
2.2 The Face Lattice
2.3 Polarity
2.4 The Representation Theorem for Polytopes
2.5 Simplicial and Simple Polytopes
Contents
2.6 Appendix: Projective Transformations
Notes
Problems and Exercises
67
69
70
Graphs of Polytopes
3.1 Lines and Linear Functions in General Position
3.2 Directing the Edges ("Linear Programming
for Geometers")
3.3 The Hirsch Conjecture
3.4 Kalai's Simple Way to Tell a Simple Polytope
from Its Graph
3.5 Balinski's Theorem: The Graph is d-Connected
Notes
Problems and Exercises
77
77
80
83
93
95
96
97
103
104
107
109
113
115
119
127'
127
132
138
139
143
145
149
150
150
153
156
157
159
160
163
163
165
171
172
173
Contents
xi
175
177
179
183
184
191
191
195
198
208
217
224
225
231
232
239
246
254
258
268
275
281
291
292
299
310
319
320
321
References
325
Index
365
o
Introduction and Examples
part of the work (and fun) consists in seeing how intuition from life in
three dimensions can lead one (i.e., everyone, but not us) astray: there are
many theorems about 3-dimensional polytopes whose analogues in higher
dimensions fail badly. Thus, one of the main tasks for polytope theory is
to develop tools to analyze and, if possible, "visualize" the geometry of
higher-dimensional polytopes. Schlegel diagrams, Gale diagrams, and the
Lawrence construction are prominent tools in this direction tools for a
more solid analysis of what polytopes in d-space "really look like."
Notation 0.0. We stick to some special notational conventions. They are
designed in such a way that all the expressions we write down are "clearly"
invariant under change of coordinates.
In the following R d represents the vector space of all column vectors of
length d with real entries. Similarly, ( d)* denotes the dual vector space,
that is, the real vector space of all linear functions 1!`i R. These are
given by the real row vectors of length d.
The symbols x, x o ,
. , y, z always denote column vectors in IY (or
in R d1 ) and represent (affine) points. Matrices X, Y, Z,... represent sets
of column vectors; thus they are usually (d x m)- or (d x n)-matrices. The
order of the columns is not important for such a set of column vectors.
Also, we need the unit vectors ei in
which are column vectors, and
the column vectors 0 and 1 = Ei ei of all zeroes, respectively ail ones.
The symbols a, ao , al , . . . , b, c, . . . always denote row vectors in ( d)* ,
and represent linear forms. In fact, the row vector a E (Rd )* represents the
linear form t = fa :R d ---+ R, z
ax. Here ax is the scalar obtained as
the matrix product of a row vector (i.e., a (1 x d)-matrix) with a column
vector (a (d x 1)-matrix). Matrices like A, A', B, . . . represent a set of row
vectors; thus they are usually (n x d)- or (m x d)-matrices. Furthermore,
the order of the rows is not important.
We use 11 = (1, ... ,1) to denote the all-ones row vector in (Rd)*, or
in (Rdi)* Thus, liz is the sum of the coordinates of the column vector x.
Similarly, 0 = (0, . . . , 0) denotes the all-zeroes row vector.
Boldface type is reserved for vectors; scalars appear as italic symbols,
such as a, b, c, d,x, y.... Thus the coordinates of a column vector x will be
, xd E R, and the coordinates of a row vector a will be al,. - , adBasic objects for any discussion of geometry are points, lines, planes and
so forth, which are affine subspaces, also called flats. Among them, the
vector subspaces of r d (which contain the origin 0 E d) are referred to as
linear subspaces Thus the nonempty affine subspaces are the translates of
linear subspaces.
The dimension of an affine subspace is the dimension of the corresponding
linear vector space. Affine subspaces of dimensions 0, 1, 2, and d 1 in Rd
are called points, lines, planes, and hyperplanes, respectively.
For these lectures we need no special mathematical requirements: we just
assume that the listener/reader feels (at least a little bit) at home in the
real affine space Rd , with the construction of coordinates, and with affine
maps x 1----+ Ax + x o , which represent an affine change of coordinates if A
is a nonsingular square matrix, or an arbitrary affine map in the general
case.
Most of what we do will, in fact, be invariant under any affine change
of coordinates. In particular, the precise dimension of the ambient space is
usually not really important. If we usually consider "a d-polytope in
then the reason is that this feels more concrete than any description starting
with "Let V be a finite-dimensional affine space over an ordered field,
and ...."
We take for granted the fact that affine subspaces can be described by
affine equations, as the affine image of some real vector space k or as the
set of all affine combinations of a finite set of points,
n
F = {x E
: x = A0x0 + . - + Ax n for Ai G R,
E Ai = 1}.
conv(K) :=
fl { K'
Our sketch shows a subset K of the plane (in black), and its convex hull
conv(K), a convex 7-gon (including the shaded part).
AIX'
A1
Ak
4-1
xk-1) + Akxk
1 Ak
for Ak < 1. For example, the following sketch shows the lines spanned by
four points in the plane, and the convex hull (shaded).
Geometrically, this says that with any finite subset K0 C K the convex
hull conv(K) must also contain the projected simplex spanned by K0 . This
proves the inclusion "D" of
,Xk} C K, Ai
= 11.
>
i=1
But the right-hand side of this equation is easily seen to be convex, which
proves the equality.
Now if K
, xn } Rd is itself finite, then we see that its convex
fxl,
hull is
Our sketches try to illustrate the two concepts: the left figure shows a
pentagon constructed as a V-polytope as the convex hull of five points; the
right figure shows the same pentagon as an 7-1-polytope, constructed by
intersecting five lightly shaded halfspaces (bounded by the five fat lines).
Usually we assume (without loss of generality) that the polytopes we
study are full-dimensional, so that d denotes both the dimension of the
polytope we are studying, and the dimension of the ambient space Rd.
The emphasis of these lectures is on combinatorial properties of the faces
of polytopes: the intersections with hyperplanes for which the polytope is
entirely contained in one of the two halfspaces determined by the hyperplane. We will give precise definitions and characterizations of faces of
polytopes in the next two lectures. For the moment, we rely on intuition
from "life in low dimensions": using the fact that we know quite well what
a 2- or 3-polytope "looks like." We consider the polytope itself as a trivial
face; all other faces are called proper faces. Also the empty set is a face for
every polytope. Less trivially, one has as faces the vertices of the polytope,
which are single points, the edges, which are 1-dimensional line segments,
and the facets, i.e., the maximal proper faces, whose dimension is one less
than that of the polytope itself.
We define two polytopes P,Q to be corribinatorially equivalent (and denote this by Pi_-_, Q) if there is a bijection between their faces that preserves
the inclusion relation. This is the obvious, nonmetric concept of equivalence that only considers the combinatorial structure of a polytope, see
Section 2.2 for a thorough discussion.
Example
polytopes are line segments. Thus any two 0-polytopes are affinely isomorphic, as are any two 1-polytopes.
Two-dimensional polytopes are called polygons. A polygon with n vertices is called an n-gon. Convexity here requires that the interior angles (at
the vertices) are all smaller than 7r. The following drawing shows a convex
6-gon, or hexagon.
sin( 2"k
)) - 0 < k < n} C
P2(n) := cony {(cos()
n
n
,
The following drawing shows the regular hexagon P2 (6) in R2 . It is cornbinatorially equivalent, but not affinely isomorphic, to the hexagon drawn
above.
:=
IX E
Example 0.4. The three-dimensional cube C3 and the octahedron C3 4 ' are
familiar objects as well:
Cd := Ix
Ix E Rd :
We have chosen our "standard models" in such a way that they are
symmetric with respect to the origin. In this version there is a very close
connection between the two polytopes Cd and Cd: they satisfy
Cd''
Cd
E Cd 1
G Cd'},
that is, these two polytopes are polar to each other (see Section 2.3).
Now it is easy to see that the d-dimensional crosspolytope is a simplicial
polyt,ope, all of whose proper faces are simplices, that is, every facet has
the minimal number of d vertices. Similarly, the d-dimensional hypercube
is a simple polytope: every vertex is contained in the minimal number of
only d facets.
These two classes, simple and simplicial polytopes, are very important. In
fact, the convex hull of any set of points that are in general position in Rd
a simplicial polytope. Similarly, if we consider any set of inequalities is
in i d that are generic (i.e., they define hyperplanes in general position)
and whose intersection is bounded, then this defines a simple polytope.
Finally the two concepts are linked by polarity: if P and P are polar,
then one is simple if and only if the other one is simplicial.
(The terms "general position" and "generic" are best handled with some
amount of flexibility you supply a precise definition only when it becomes
clear how much "general position" or "genericity" is really needed. One can
even speak of "sufficiently general position"! For our purposes, it is usually
sufficient to require the following: a set of n> d points in Rd is in general
position if no d of them lie on a common affine hyperplane. Similarly, a set
of n > d inequalities is generic if no point satisfies more than d of them
with equality. More about this in Section 3.1.)
Here is one more aspect that makes the d-cubes and d-crosspolytopes
remarkable: they are regular polytopes polytopes with maximal symmetry. (We will not give a precise definition here.) There is an extensive and
very beautiful theory of regular polytopes, which includes a complete classification of all regular and semi-regular polytopes in all dimensions. A lot
can be learned from the combinatorics and the geometry of these highly
regular configurations ("wayside shrines at which one should worship on
the way to higher things," according to Peter McMullen).
At home (so to speak) in 3-space, the classification of regular polytopes
yields the well-known five platonic solids: the tetrahedron, cube and octahedron, dodecahedron and icosahedron. We do not include here a drawing of the icosahedron or the dodecahedron, but we refer the reader to
Example 0.5. There are a few simple but very useful recycling operations
that produce "new polytopes from old ones."
If P is a d-polytope and x 0 is a point outside the affine hull of P (for
this we embed P into n for some n> d), then the convex hull
pyr(P) := conv(P U {x 0 })
is a (d + 1)-dimensional polytope called the pyramid over P. Clearly the
affine and combinatorial type of pyr(P) does not depend on the particular
choice of x0 just change the coordinate system. The faces of pyr(P) are
the faces of P itself, and all the pyramids over faces of P.
Especially familiar examples of pyramids are the simplices (the pyramid over Ad is Ad+i ), and the Egyptian pyramid Pyr3 = pyr(P2(4)): the
pyramid over a square.
x+
x_
and the crosspolytopes, which are iterated bipyramids over a point,
bipyr(Cd ) =
Cd+1 A
10
and set
PxQ := { (1 : x E P, y E Q}.
Y
We get a polytope of dimension dim(P) + dim(Q), whose nonempty faces
are the products of nonempty faces of P and nonempty faces of Q.
in particular
The prism over a polytope P is the product of P with a segment,
prism(P) := Px A i .
This is polar to the bipyramid:
prism(P) = (bipyr(P A )) .
The smallest interesting prism is the one over a triangle,
also known as the triangular prism.
A2 X Ai ,
11
is defined by
/
t i--+ x(t) :=
t2
\ td
ER".
x(t 3 )
x(t4)
12
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
<=
0
0
0
0
0
0
0
0
0
0
24
40
60
84
120
180
252
360
504
840
END
N\N
Ii
E\T
Q \ S
13
S\
\I
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
#
*...*
*..**
*.**.
***..
****.
**.**
**..*
*4....
**...
.****
.**.*
.**.
.**..
..***
..**.
.**.
..**
..**
..*
***
*.*
..*
..*
*
*..
**,
.**
...
*.,
**,
.**
*..
**.
.**
**.
.**
***
:
:
:
:
:
:
:
:
:
:
:
:
:
:
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
4
11111 111
00000 000
From this pattern, one can derive that any two vertices of the polytope
are adjacent. We can also check this directly: every pair of vertices is contained in at least 3 facets. So, the edge 12 is contained in the facets (5) =
1238, (6) = 1234, (7) = 1245, (8) = 1256, (9) = 1267, and (10) = 1278.
Similarly, the edge [1,3} is contained in the facets (4) = 1348, (5) = 1238
and (6) = 1234.
Finally, we can note that there is a combinatorial symmetry that sends
vertex i to vertex 9 i; see Exercise 0 1 .
The following theorem and corollary contain a complete description of
the combinatorial structure of the cyclic polytopes as suggested by our
computation. Here we break our promise not to do any proofs in this introduction: mainly because the proofs are fun, and the results are a little
surprising (see Corollary 0.80.
14
0.
Cd(n) = convfx(t i ), .
is a simplicial d-polytope. A d-subset S C [n] forms a facet of Cd(n) if and
only if the following "evenness condition" is satisfied:
If i <j are not in S, then the number of k E S between i and j is even:
2 I Iffk : k S, < k < jj
for i, j V S.
1
1
x(t o) x( t i)
i
to
...
x(td)
1
td
1
ti
det
b0
tg
4c1 i
0<i<3<d
bi
td
This is easily proved by observing that both sides are polynomials and that
the determinants vanish whenever we have ti = ti for some i
j. From
the identity, one sees that no d +1 points on the moment curve are affinely
dependent. In particular, this shows that Cd (n) is a simplicial d-polytope.
Now let S = {i 1 ,. , id } C [n]. Then the hyperplane Hs through the
corresponding points x(t 8 ) is given by
HS
E Rd : Fs(x) = 01,
where
Fs (x) := det
xi"
1
x(t 1 )
...
x(t)
\
)
15
thus it has d different zeroes, and changes the sign at each of them. The
following sketch is supposed to illustrate this.
Now S forms a facet if and only if Fs(x(t i )) has the same sign for all the
points x(t i ) with i C [n]\S; that is, if Fs(x(t)) has an even number of sign
changes between t = ti and t = ti , for i < j and ti, j E [n]\S
LI
In particular, this criterion shows that the combinatorics of Cd(ti, - , tn)
do not depend on the specific choice of the parameters t i , so Cd (n) is well
defined as a combinatorial equivalence class of polytopes.
It is quite easy to extend the evenness condition to a characterization of
all the faces of Cd (n). This characterization then also shows the following
corollary (Exercise 0.8), for which we give an independent proof.
Corollary 0.8. The cyclic polytope Cd(n) is [ ] -neighborly, that is, any
subset S C [n] of IS! < vertices forms a face.
4.
det (x,x(t ii ), x
+e), . . , x(t), x
, x(M +d 2k)) .
,M+d-2k.
For d < 3 Corollary 0.8 just says that the points x(t i ) form vertices of
Cd(n): the points on the moment curve are in convex position. However,
16
Rd
17
d
we have taken the following drawing from his paper [445, Fig. 4].
the permutation
x,:t )
that maps x i) in such a way that two vertices are connected by an
edge if and only if the corresponding permutations differ by an adjacent
transposition. Check this in our drawing of 112:
132
123
312
213
321
231
18
2143
4312
2134
3412
3214
3241
*( * ( ** ) )
(*(**))*
* ((* *)* )
19
Whereas the first constructions of the associahedra were very much "ad
hoc," in Lecture 9 we will get an associahedron from a very natural construction due to Gel'fand, Zelevinsky Si Kapranov [213, 214]; as the "secondary polytope" of the n-gon [213, Rem. 7c)]. More generally, we will construct "fiber polytopes" there, a concept due to Billera Si Sturmfels [74, 75].
Recently, Mikhail M. Kapranov [286] constructed a new combinatorial
object KfI i , the perrnuto-associahedron, which combines the permutahedron and the associahedron. (Kapranov denotes it "KP,i ".) Its vertices
correspond to the different ways of multiplying n terms a l , a2, ... , an, in
arbitrary order, assuming that multiplication is neither commutative nor
associative and again there is a natural way to describe all the faces.
Our drawing shows KII 2 , a 12-gon.
1(3 2)
(1 3)2
2(3 1)
(2 3)1
Adi(k)
..
conv{v E 10,11 d :
E v, , k}
d
-= {X
.0<
xi
i=i
for 1 < k < d - 1.
20
Xf:7
The matrices X' are the 0/1-matrices with exactly one 1 per row and per
column. If we identify Rd2 with the set of all real (dxd)-matrices, then
the matrices X' are 0/1-vectors in Rd ", and their convex hull forms a
0/ 1-polytope
P(d)
This is an interesting polytope with many names: the Birkhoff polytope, the
perfect matching polytope of Kn the assignment polytope, the polytope of
doubly stochastic matrices, and so forth.
The polytope P(d) has d! vertices (by construction), d 2 facets, and dimension (d 1) 2 . In fact, a complete linear description is given by
P(d) = {X
dd.
d,
k=1
d
d.}
21
22
CfT (6); see Exercises 0.14 and 1.1(iv). However, it seems that the method
does not go beyond that: in general the algorithmic determination of all the
facets of Q is certainly much harder and more strenuous than examining
all the vertices of Q.
The problem of finding some of the facets, by using the combinatorial
properties of the traveling salesman problem, is a central problem for a
whole branch of mathematics, called "polyhedral combinatorics" see
Gr6tschel & Padberg [229] and Jnger, Reinelt & Rinaldi [269] for solid
introductions, including detailed information about the structure of the
polytopes QT(n) and Q(n).
Notes
The principal historical "classics" in the theory of polytopes are the 1852
treatment by Schlifli [437] published in 1901, the books by Bruckner [131]
(1900), Schoute [444] (1905), and Sommerville [470] (1929), and the volume by Steinitz & Rademacher [490] (1934) about 3-dimensional polytopes.
(A very helpful bibliography is Sommerville [471].) The modern theory of
polytopes was established by Griinbaum's 1967 book [234]. It should be
stressed that not only did Griinbaum present the major part of what was
known at the time, but his book also contains various pieces of progress
and substantial original contributions, and has been an inspiring source
of problems, ideas, and references to everyone working on polytopes since
then.
There are more recent books and surveys on polytopes. Many of them
concentrate on aspects related to the upper and lower bound theorems and
the g-theorem (among them McMullen & Shephard [374], Brondsted [126],
Stanley [478], and Hibi [252]) and on the various methods of f-vector theory; see Lecture 8. Other aspects are treated in Barnette's exposition on
3-polytopes [41], Schrijver's book on optimization [448], and the handbook
chapters by Kleinschmidt & Klee [301], and Bayer & Lee [591. Also, the
reader might find Pach's volume [401] inspiring.
In our lectures we avoid any larger discussion of general convex sets and
bodies, as well as of most of the convex-geometric aspects of polytopes.
We refer to Bonnesen & Fenchel [117], Schneider [440], and Ewald [189],
the point is that for a convex polytope, we can describe and discuss
everything in terms of vertices, edges, facets, etc. (i.e., a finite collection
of combinatorial data) and bypass the apparatus of support functionals,
nearest point maps, distances, volume, and integration, etc. Correspondingly, in this book we disregard all metric properties of polytopes, such as
volume, surface area, and width, which are part of a very interesting theory
of their own.
23
24
0.3 Show that the permutahedron Ild_ i C Rd (Example 0.10) has dimension d - 1, that it is a zonotope, and that it is simple.
Describe its 2d 2 facets, by constructing inequalities that determine
them.
0.4 Let a l > a2 > ... > ad be real numbers, not all equal. The generalized permutahedron (or orbit polytope) lid- i (ai, . . . , ad ) is the convex
hull of all the vectors given by all the permutations of the multiset
{a l , ... ,ad}.
Investigate the combinatorics of the generalized permutahedra. In
particular, show that their dimension is d - 1. Are they all simple?
(They are not.)
Under what conditions do all the edges of Hd_1 (a l , ... , an ) have the
same length? (Schoute [445, p. 5])
0.5 Let P = Cd C 1. d be the d-cube. Enumerate the 3d 1 faces of Cd,
and show that the nonempty faces are naturally associated with the
sign vectors in {+, -, O } d .
Given a linear function c E (R)*, how can one find a vertex that
maximizes c over P ("optimization problem")?
Given y E d , how do we tell whether y E P? If y V P, how can we
find an inequality that is valid for P but is violated by y ("separation
problem")?
For which other classes of polytopes discussed in Lecture 0 can you
easily solve these problems?
0.6 Describe Cd(d + 2), the cyclic d-polytopes with d + 2 vertices, combinatorially and explicitly.
Is the 2-neighborly polytope (A2 x A2) constructed in Example 0.5
combinatorially equivalent to C4 (6)?
0.7 Consider the cyclic polytope Cd (n) = conv{x(0), x(2), ... , x(n-1)}.
Show that there is an affine symmetry (an affine reflection) which
induces the symmetry i 4----+ n -1- 1 - i (that is, x(i - 1) 4 x(n -i)),
and thus the corresponding combinatorial symmetry of Cd(n).
0.8 From Gale's evenness condition, given in Theorem 0.7, derive a complete combinatorial description of all the faces of Cd(n).
From this, derive that the cyclic polytopes are LC-neighborly (Corollary 0.8).
25
0.9 Show (bijectively) that the number of ways in which 2k elements can
be chosen from [72] in "even blocks of adjacent elements" is (n-k k ).
Thus, derive from Gale's evenness condition that the formula for the
number of facets of Cd (n) is
41\
fd-i(Cd(n))
(n - 1 1- % 1 1
fd-i(Cd(n))
{ nn
(n-k
k -k)
2 (n- kk-1)
for d = 2k even,
for d= 2k + 1 odd.
(iti +
0.11* Is there a fast and simple way to decide whether a certain point
G Rd (with rational coordinates, say) is contained in the cyclic
polytope Cd(1, 2, .. . , n)?
(General theory namely the polynomial equivalence of optimization and separation according to Griitschel, Lovisz & Schrij ver [228]
implies that there is a polynomial algorithm for this task, since optimization over Cd(n) is easy, by comparing the vertices. However, we
ask for a simple combinatorial test, not using the ellipsoid method )
0.12 Prove the claims in Example 0.12 about the Birkhoff polytope P(d):
in particular, show that the dimension is (d- 1) 2 , and that the number
of facets is d2 .
The Birkhoff polytope P(d) and the permutahedron 11d_1 are closely
related: show that there is a canonical projection map P(d)
0.13 Draw the 3-dimensional associahedron K3. Justify the general for()
2n-2
mula '
,-2n n-1 for the number of vertices ofK_2.
0.14 Describe the combinatorial structure of the traveling salesman polytopes QT (3), QT(4), and QT(5). How many vertices and facets do
they have? Which vertices are adjacent? Are they simple, or simplicial? Similarly, try to describe Q'T (2), (3), and CA, (4).
26
0.15* What is the maximal number f(d) of facets of a d-dimensional 0/1polytope? How fast does f (d) grow asymptotically?
(It is not hard to see that
Nid '
for all large enough d. Here the upper bound is due to Rote [431],
while the lower bound is from explicit computation of "random 0/1polytopes" in low dimensions in combination with a "free sum" construction for 0/1-polytopes from [315]. The value 3.6 was achieved
in March 1997 by Thomas Christof for a random 0/1-polytope (of
dimension 13, with 254 vertices and at least 17,464,356 facets), using
his PORTA code and new ideas described in Christof igi Reinelt [148 ] .
For "current records" in the "Olympic race" for 0/1-polytopes with
many facets see [314] on the Web.)
0 1 0 0)
0 0 0 1
1 0 0 0
0 0 1 0
0.17 Show that the Birkhoff polytope P(d) C Rd2 contains the asymmetric
traveling salesman polytope QT (d) C d2 d C d2 . Does every facet
of P(d) yield a facet of QT(d)?
(For a detailed investigation, see Billera & Sarangarajan [72 ] .)
1
Polytopes, Polyhedra, and Cones
1.1
However, to make sure that the pain level does not go below zero, we
start with a few definitions. In the following, we work with two versions
of polyhedra in the course of this lecture we will see that they are
mathematically (but not algorithmically!) equivalent. The two concepts
28
P = P(A , z)
Ix E
d : Ax < z1
for some A E
mxd 7 Z
R.
(Here "Ax < z" is the usual shorthand for a system of inequalities, namely
ai x < z1 , . . , am x < zrn, where al . . am are the rows of A, and
Z1 ,... zrn, are the components of z.)
For the second version we need the notion of a cone: a nonempty set of
vectors C C Rd that with any finite set of vectors also contains all their
linear combinations with nonnegative coefficients. In particular, every cone
contains O. For an arbitrary subset Y. C j . d , we define its conical hull (or
positive hula cone(Y) as the intersection of all cones in I d that contain Y.
Clearly C := cone(Y) is a cone for every Y. Similar to the situation for
convex hulls (Lecture 0), one can easily see that
,
cone(Y) =
>
01.
Itiy i +
+ tny,,
29
Now we define a V-polyhedron to denote any finitely generated convexconical combination: a set P C Rd that is given in the form
P = conv(V) + cone(Y)
dxn'
)
as the Minkowski sum of a convex hull of a finite point set and the cone
generated by a finite set of vectors.
Thus, comparing this to Definition 0.1, we get that a V-polytope is a Vpolyhedron that is bounded, that is, contains no ray { u + tv : t > 0} with
y 0. For this we only need to observe that conv(V) is always bounded.
This follows from a trivial computation: if z E conv(V), then
min{vik : 1 <j < n} < xk < max{vik : 1 <j < n},
which encloses conv(V) in a bounded box. Similarly, an 7-1-polytope is the
same thing as a bounded fl-polyhedron.
Now we start with a basic version of the " representation theorem for
polytopes," which will be considerably strengthened and generalized in the
course of the proofs. See Section 2.4 for a definitive version.
Theorem 1.1 (Main theorem for polytopes).
A subset P c Rd is the convex hull of a finite point set (a V-polytope)
P = cony (V)
P = P (A , z)
The first two statements are trivial for a polytope presented in the form
P = P (A , z) (where the first is a special case of the second), but both are
nontrivial for the convex hull of a finite set of points. Similarly the last two
statements are easy to see for the convex hull of a finite point set, but axe
nontrivial for bounded intersections of halfspaces.
Theorem 1.1 is the version we really need, a very basic statement about
polytopes; however, it is not the most straightforward version to prove.
Therefore we generalize it to a theorem about polyhedra, due to Mot zkin
[384].
30
P = conv(V) + cone(Y)
for some V
Rd " Y G Rd'
7
P = P(A,z)
mxd
First note that Theorem 1.1 follows from Theorem 1_2 we have already
seen that polytopes are bounded polyhedra, in both the V- and the 7-tversions.
Theorem 1.2 can be proved directly, and the geometric idea for this is
sketched in Section 1.2. However, fighting one's way through the formulas is
quite strenuous, mainly because the points in conv(V) + cone(Y) are hard
to manipulate. It turns out that it is much easier to "homogenize": we
pass from affine d-space to linear (d+ *space; for this, we adjoin an extra
coordinate (which we will take as the zeroeth coordinate in the following),
mapping the point x E llid to the vector ( xl ) E Rd+ 1 .
This reduces Theorem 1.2 to the special case where P is a cone, which
can be proved more easily.
Theorem 1.3 (Main theorem for cones).
A cone C C I d is a finitely generated combination of vectors
C = cone(Y)
C = P(A,O)
We will prove Theorem 1.3 in Section 1.3. In the following we will usually refer to the polyhedral cones characterized by Theorem 1.3 simply as
"cones," because the objects we consider are clearly polyhedra. Note that
every cone C, by definition, contains the origin O.
Let us see here why Theorem 1.2 follows from Theorem 1.3 by homogenization. For this, we associate with every polyhedron P c I d a cone
C(P) C 110+ 1 , as follows.
If P = P(A, z) is an H-polyhedron, we define
C(P) :__
31
P = Ix R d :
(xl)
E C(P)}.
Ix
If P
d:
1 E Pl is an fl-polyhedron as well.
C(P) := cone ( 11 0
V Y)
Clearly, C(P) is again a V-polyhedron in
d+1 , and
P = fx ERd : ( xl ) E C(P)}.
Conversely, a simple computation shows that if C = cone(W) is any cone
in Rd+ I generated by vectors wi with wio > 0, then Ix c Rd
(1)
E C}
is a V-polyhedron.
Now, given any fl-polyhedron P, we can apply Theorem 1.3 to C(P),
to conclude that C(P) is a V-cone contained in f x E Etd-1-1 . x o > 01, so
P is a V-polyhedron as well. Conversely, if P is a V-polyhedron, then by
Theorem 1.3 the associated cone C(P) is an fl-polyhedron, and hence so
is P.
In both cases C(P) realizes the homogenization of P, which we will discuss in Section 1.4, once we have established the Farkas lemma. The geometric idea is depicted in the sketch above, which shows the cone in 3
associated with an affine polytope in R2 . If P is a polyhedron, then one
has to add the necessary "points at infinity" to P, to make sure that C(P)
is a (closed) polyhedron.
32
d.
itE
(X,
t, u) E
fx - xkek : x E PI
= fx E d : Xk = 0, 3yER:x+yekEPI.
for the projection of P in the direction of ek. The set proj k (P) is contained
in the hyperplane Ilk = fx C Rd : xk = 01. A closely related set is the
elimination
elimk (P)
fx - tek : x
C P, t C 1
{x Elie : 3yER:x+yekEP } .
Thus elimk (P) is the set of all points in Rd which project to proj k (P). In
particular, we get an isomorphism elimk (P) proj k (P) x R.
33
(1)
xi
2x 1
(2)
4 x2
x2
2x
(3)
(4)
+
+
(5)
(6)
+ 2 xi
2x i
x2
+ 6 x2
(7)
6x 1
x2
xi
xi
<
<
9
4
<
<
<
11
<
17
<
X2
.
.-'
'-'
xi
Now assume that we fix some x1, and ask for the possible values of x2.
Then we see that inequality (4) requires x 1 < 4. All other inequalities can
be rewritten to give either an upper bound on x2 (if the coefficient of x2 is
positive), or lower bound (if the coefficient of x2 is negative). Furthermore,
there is a solution for x2 if and only if every upper bound for x2 derived
this way is larger than every lower bound.
The sketch on the next page shows the projection of the 2-polytope P to
proj 2 (P), by eliminating the x2-variable. Here elim2(P) is the infinite strip
(shaded) of all points that lie above or below proj2 (P).
Observe how the points of P on any vertical line (i.e., with fixed x 1 ) are
bounded from above and below by inequalities with positive, respectively
negative, coefficient ai2 . If there is no solution, then some upper bound is
smaller than some lower bound: that is, the combination of two inequalities,
one with positive and one with negative ai2, leads to a restriction for the
possible values of xl.
Also note that there is one redundant inequality in the original system:
this leads to the effect that the same lower bound on x 1 arises from several
different pairs of inequalities.
34
12
dim )(I ) )
P'12(P)
..r
as
ax <
ax zi
and
ax <zj
( aik)xk ax k + ax zi.
Here the right-hand sides of the rewritten forms do not depend on xk, so the
first one yields an upper bound on xk, the second one a lower bound. The
combination of the two inequalities (multiplied by the positive coefficients
aik respectively ak) yields the condition
choose k < d.
Construct the matrix Alk G
"
35
, and
z) = {x E Rd : Ax < z}
can be written as the intersection of a polyhedron (in fact, a cone)
(z)
E Rd +m
1.7
Co (A) = cone({
ei
Ae i
by decomposing
(
X)
E ixi I
(sign(x i ))
ei
Aei
M
i=1
AXii
ei
36
From this we get a little geometric intuition, which suggests that we can
write down
P n Hk = conv(V/k),
where ink is the matrix (set) of column vectors constructed as
Vii (Vjk)Vi
: Vik
V ik
t vi . Vik =
>
O, V3k G 0 } .
Vik Vjk
For this it is quite clear that we get P n Hk D conv(V/k), but for the
converse we have to work a little. We omit this ugly little computation
here since you'll see it in the next section: it comes out a little nicer in the
homogeneous form (Lemma 1.6). Anyway, this way "in principle" one can
give an explicit representation of z as a convex combination of vectors in
17 1k .
LI
This finishes the proof of Theorems 1.1-1.3.
One can do a similar argument for P = conv(V) + cone(Y). However,
the corresponding computations become extremely tedious they are too
ugly even to leave them as an exercise. This is why we homogenize and
switch to cones, where all difficulties disappear.
37
t > 0}
d 3tER n :t>0, X= Ytl.
a k < 0}.
This can be satisfied. Namely, if aik > 0 and ail, < O, then we know that
< 1 ( a )x , which is equivalent to (aka i+ (aik)ai )x < 0,
aik
which holds because x E P(A/', 0).
Now we proceed to prove the "backward direction" of Theorem 1.3. For
this let C = P(A, 0) C Rd be an fl-cone. We can write it as
=
{X E
Rd : Ax < 0}
{( x ) E Rd+m : Ax < w}
n{
( x
E Rd-E rn : w = 0 } .
38
Here {(w
x ) E Rd+m : Ax < w} is a V-cone, as we have shown above.
x ) E Rdm : w = 0 } can be formed succesThe intersection with {(w
sively, by setting coordinates to zero one at a time, i.e., intersecting with
coordinate hyperplanes of the form Ilk = fy e yk = 01. Thus it
suffices to prove the following lemma.
Proof. First note that the vectors in Yrn all have xk-coordinate 0, so
clearly C n HkD cone(Y/k).
For the reverse inclusion, we consider some y = Yt E cone(Y) > 0)
with vk = O. Now either we have tiYa = 0 for all i, in which case we get
V E cone({y, : yik = 01), or we can expand vk = 0, to get
A:=
t (y 3 k) > O.
tiYik
j
Yik >0
yik<0
tiyi +
iYjk>O
Yik=o
ti yi +
1
A
E
yik=o
ti yl
tiyi +
3 yik<0
t( y3k)
J tiy,
i yik>0 ( j y3 k<0
Y3k<0
vik>0
tiyik) t yi
titi
tiyi +
Y3k)Y2 Yikl/j )A ((
2
j Y3k<C)
39
1.4
It was first pointed out by Kuhn [317] that with (termination of) Fourier-
Motzkin elimination we have also done all the work for the Farkas lemma.
This extremely important lemma appears in many different versions all
over the theory of polytopes and polyhedra. It is interesting to note that if
you look into different books and papers, you find quite different lemmas
all called "the Farkas lemma." All of these, however, are easily transformed
into each other.
Essentially, the Farkas lemma yields a characterization for the solvability
of a system of inequalities. There are variations for systems of inequalities
in various standard forms: Farkas lemmas for polyhedra and for cones,
for inequality systems with equalities, inequalities or strict inequalities, in
nonnegative, positive or unrestricted variables, and so on. There are also
quite different ways to formulate theorems "of Farkas type":
as theorems of the alternative (one inequality system has a solution
if and only if a second system has none),
as transposition theorems (because the second system can be derived
by transposing the matrix and vectors of the first),
as duality theorems (the duality theorem for linear programs is of
Farkas type),
as good characterizations (if a system has a solution, then any solution
vector proves this; if it has no solution, then the Farkas lemma yields
a dual vector that encodes this fact),
40
as certificates for validity (if an inequality is valid for the solution set
of a system, then it is a conical combination of the inequalities of the
system),
or, dual to this, as separation theorems (if a point does not lie in
a convex-conical hull, then it can be separated from it by a linear
functional).
We refer to Mangasarian [347] and to Stoer & Witzgall [492] for more
versions, extensions, and generalizations. Separation theorems "of Farkas
type" also hold for convex bodies. Infinite-dimensional versions are fundamental in functional analysis ("Hahn-Banach theorem").
Here is one basic version, characterizing the solvability of a general system of inequalities.
Proposition 1.7 (Farkas lemma I).
Let A E mxd and z c m .
Either there exists a point x E Rd with Ax < z,
or there exists a row vector c E ( m)* with c > 0, cA = 0 and cz < O,
but not both.
Proof. First observe that both conditions cannot hold at the same time:
otherwise there are a column vector x E Rd and a row vector c E
with
0 = Ox = (cA)x = c(Ax) < cz < 0,
which is a contradiction.
Now define P
P(A, z), and Q
P(( z A), o). We note that an
x E Rd exists with Ax < z if and only if Q contains a point with xo > 0.
Here Q is an 7-1-cone. Now we eliminate the variables xl,
xd from Q, to
get the 7-1-cone elim i elim2
elimd(Q)The key observation is that if we do Fourier-Motzkin elimination to get
elimi P(D, 0) = P(Dii, 0), then every inequality in the eliminated system
elimi (D) is a positive combination of at most two rows of D, so D/i can
be written as CID for a matrix Ci with only nonnegative entries, of which
at most two per row are nonzero.
Iterating this idea, we get
41
and thus the system C(z, A)x < 0 contains an inequality eyioxo <O with
'Yio > O. Let e be the row of C that yields this, then we have c(z, A) =
El
(7i0, 0), that is, ez --=
<O and cA = O.
Now comes another version of the Farkas lemma, for nonnegative solutions of systems of inequalities. Every such system can be rewritten as a
system of inequalities, and this is exactly what we do to prove it. In fact, the
proof nicely illustrates various simple-but-important techniques for rewriting systems of inequalities, such as the introduction of slack variables, and
rewriting unbounded variables as differences of nonnegative ones.
Proposition 1.8 (Farkas lemma II).
Let A E Rmxd and z E Rm.
Either there exists a point x E d with Ax = z, x > 0,
or there exists a row vector e E (Rm)* with cA > 0 and cz < O,
but not both.
Proof. We have the following equivalences:
< >
Id
FL I
< >
x: Ax = z, x > 0
z
zo
<O
Id
e, >
<
>
< >
(el
e2 )z <O
42
Proof. The "if" part is easy to see: the existence of z with Ax <z and
aox > zo contradicts both (ii) (as in Farkas lemma I) and (i) (with a similar
computation).
For the "only if" part, assume that neither (i) nor (ii) is satisfied. Then
we conclude that there is no b> 0 and 0 > 0 with bA = Oa and bz < Ozo:
otherwise, (i) would be satisfied for c := lo b if fi> 0, or (ii) for c := b, if
0 O.
Thus we can apply Farkas lemma I to compute
7 cz = zo,
ra
(1
c( A) = -ao
-
0 )
z -A
= (zo, -ao)
-0)'
A Yo > 0, y E Rd :
Ay
Yozo-
Now either we have yo > 0, then we put x : = --o- y, and this satisfies Ax < z
and aox > zo , or we have yo = 0, then we use the w constructed before
(remember?) and put x : = w + y. This x satisfies Ax = Aw + Ay <
z + 0 = z and aox = aow + aoy > zo + 0 = zo.
-Ei
-
The following, fourth and last version (but see the exercises) shows that
the Farkas lemma can also be used to separate a point from a V-polyhedron:
if x is not contained in P := conv(V)cone(Y), then there is an inequality
ax < a satisfied by P, but not by x.
Proposition 1.10 (Farkas lemma IV).
Let V E Rdxn , Y E isdxn' and x E d
Either there exist t, u > 0 with 1t= I and x = Vt + Yu,
or there exists a row vector (a, a) E (Rd+ 1 )* with av i < a for all i <
ayi < O for all j < re, while ax > a,
but not both.
,
n,
43
t) (lx )
y
O) ( u
ti (a,
a) E (1V+ 1 )* -.
(a ' a)
1.
(a, a) ( x ) <o
> A (a, a)
1.5
Using the Farkas lemma, we can give an invariant description of some very
important constructions (notably the recession cone and the homogenization of a convex set) and establish their basic properties. In Proposition 1.14
we will see that the homogenization homog(P) of a polyhedron coincides
with the "associated cone" C(P) that we used in Section 1.1.
Definition 1.11.
of P is defined as
P .
lineal(P) + (P n U)
u) fol.
44
Proposition 1.12.
cone:
rec(P) = cone(Y).
Proof. Both parts are "clear," aren't they? Not quite: on close inspection
we see that in part (ii) the direction rec(P) C cone(Y) is not entirely
obvious; it needs the Farkas lemma. Using version IV (for V = 0), we see
that if y tS cone(Y), then there exists a linear functional a with aY < 41)
and ay > O.
Now consider some z = Vt + Yu E conv(V) cone(Y), with t, u >
_ 0,
II t = 1. For this we get
E
1<i<n
tiavi <
max av i =-: K,
1<i<n
45
Again it is quite easy to see that for every convex set P, the homogenization homog(P) is a convex cone in Rd+ 1 . Furthermore, any P can be easily
recovered from its homogenization (if we don't mess with the coordinate
system) as
Proposition 1.14.
Let
PC
be a convex set.
((i zl
1l 0
r y
C(P).
46
Proof. For (i), without loss of generality we assume that X has full rank,
rank(X) d, by passing to the linear hull of X. Now let z E cone(X) and
Notes
47
x E conv(X)
(1
E cone( X
Notes
The material of this lecture is classical
our discussion is inspired by
Griitschel's treatment in [227].
We recommend Schrijver's book [448, Sect. 12.2] for more historical
comments, as a superb guide to the historical sources, with references
to the original papers by Fourier, Dines, and Motzkin, and also those by
Minkowski, Weyl, Farkas, Carathodory, and others.
The elimination method was developed by Motzkin in his 1936 doctoral
thesis [384]. We quote from Dantzig & Eaves [167]:
For years the method was referred to as the Motzkin Elimination
Method. However, because of the odd grave-digging custom of
looking for artifacts in long forgotten papers, it is now known
as the Fourier-Motzkin Elimination Method and perhaps will
eventually be known as the Fourier-Dines-Motzkin Elimination
Method.
The Fourier-Motzkin elimination method is not only a theoretical tool
with some care it can also be used for computations. "In practice,"
however, one has to deal with an effect known as combinatorial explosion:
every elimination step may transform m inequalities into up to [m2/4] new
ones, which means that after a few steps the number of inequalities in the
system can increase dramatically. However, many of the inequalities we get
by elimination are redundant: they can be deleted without changing the
polyhedron that is described by the system. Thus it is important to eliminate redundant inequalities from the system, and in fact to detect many of
them quickly, in order to keep the problem size and the computation times
down.
48
49
linear function, then the simplex algorithm (cf. Section 3.2) with "Bland's
rule" finds a path from every vertex of the polytope to the maximal vertex
of the polytope. Now if we have a simple polytope, these paths form a tree
that connects the vertices, and which can be searched easily. This yields a
very effective algorithm try Avis [24]. In the nonsimple case, one has to
search a tree on the (huge) set of all feasible bases; with some extra care
one can detect those bases that are lexicographically first at a vertex. (See
also Rote [430].) In the very nonsimple case, this does not seem to be very
efficient at first, though Avis [24] has reported the successful solution of
very large convex hull problems with reverse search; see also [137].
Avis, Bremner & Seidel [25, 26, 124], however, construct and analyze
classes of "bad" test examples for different types of convex hull algorithms.
In particular, products of cyclic polytopes of the form Cd(n) d seem to be
"universally" bad for all known types of convex hull algorithms.
50
1.2 Describe a Fourier-Motzkin elimination method to solve strict inequality systems Ix : Ax < zl. Use it to prove a representation
theorem for "open polyhedra."
1.3 Show that if C = cone(W) is any cone in d+1 generated by arbitrary
vectors wi (not necessarily with wio > 0), then Ix E Rd : (xl ) c CI
is a V-polyhedron.
1.4 State and prove a Farkas lemma for systems of the form Ax < z,
x>
O.
1.5 Prove the following Farkas lemma for equality and inequality constraints: for compatible matrices A, B C and vectors u, V w
either there exists a solution vector x for
,
Ax = u, Bx> v, Cx < w,
or there exist row vectors a, b, c with
<O.
1.7 State and prove a version of Carathodory's theorem for convexconical combinations.
1.8 Transportation polytopes have the form
P(d : a,b) =
for 1 <
k=1
k=1
xi,
j < d,
ELI
2
Faces of Polytopes
In this lecture we will discuss faces, and the face lattice. Here we restrict
our attention entirely to polytopes, although nearly everything can quite
easily be generalized to polyhedra; see the exercises.
I hope that the reader enjoys the ease with which we will get the results
in this lecture. In fact, nearly all results are "geometrically clear," and
as far as we need algebra to verify them, we get by with straightforward
computations and the Farkas lemmas.
2.1
F -= P riIx E : cx col
where ex < co is a valid inequality for P. The dimension of a face is the
dimension of its affine hull: dim(F) := dim(aff(F)).
For the valid inequality Ox <O, we get that P itself is a face of P. All
other faces of P, satisfying F C P, are called proper faces.
For the inequality Ox < 1, we see that 0 is always a face of P.
The faces of dimensions 0, 1, dim(P) 2, and dim(P) 1 are called vertices, edges, ridges, and facets, respectively. Thus, in particular, the vertices
are the minimal nonempty faces, and the facets are the maximal proper
faces. The set of all vertices of P, the vertex set, will be denoted by vert(P).
52
2. Faces of Polytopes
The following sketches show two valid inequalities for a 2-polytope; they
define a vertex and an edge, respectively.
In the following two propositions we collect some simple but basic facts
about faces.
Proposition 2.2.
Let P C Rd be a polytope.
Proof. Let P
t>O:v==V't, llt=1
<
3a : a (v
n ,) 0, a ( v1.2 ) <0
<
Proposition 2.3.
be a face of P.
53
(iv) F = Pn aff(F).
Proof. Let F be defined by the valid inequality cx < co.
For the first assertion of part (i), we see that F is a polytope from the
characterization of polytopes as bounded intersections of halfspaces: F is
the intersection of a polytope P with a polyhedron (hyperplane)
H := x E Rd : cx = col.
Furthermore, we find that F C aff(F) C H, which proves (iv).
For the second assertion of (i), note that vert(F) D Fn V =: Vo. For
the converse inclusion, let x E F, so that x can be represented as x = Vt,
with t > 0, 11 t 1. We compute
co
thus (cv, co)ti --= 0 for all i. This implies that ti = 0 for all j with
vi Vo , and thus x E conv(Vo ). From this we see F conv(170), and thus
vert(F) C Vo by Proposition 2 2(4 This completes the proof of (i).
For (ii), let
F= Pn { x e Rd : cx
co}
and
G=Pn{xERd :bx=b0}
for inequalities cx < co and bx < 1)0 that are valid for P. Then the inequality (c + b)x < co + bo is valid for P, and
Pn
d : (C b)X --= Co
b0 } --= F n G.
bo by
for all v
co cv
E V1. Then
we get that (b-FAc)x < bo -FAQ) is valid with strict inequality for all v c 14.
54
2. Faces of Polytopes
We conclude that G is a face of P. (The sketch below might tell you what
0
was "really going on" during this algebraic manipulation.)
CX =-- CI }.
Note that the construction of P/v depends on the choice of cl and of the
inequality ex < co ; however, the following result shows that the combinatorial type of P/v is independent of this.
55
F
u: P n aff({y} U F')
Frqx:cx---- c1},
F'.
7r :
bo bv
co c1
P/v.
Now v" is a convex combination of v and v', and from (b + Ao c)v" <
bo + Aoc i and (b + Aoc)v = bo + Aoci we get (b Aoc)v i <b0 + AociNow we check that the maps u and ir are inverses of each other: we
compute
2.2
56
2. Faces of Polytopes
[x,y] := {wES:x5.w<y}
the interval between x and y. An interval in S is boolean if it is isomorphic
to the poset Bk --= (2[k] , C) of all subsets of a k-element set, for some k.
A poset is bounded if it has a unique minimal element, denoted 6, and a
unique maximal element, denoted . The proper part of a bounded poset S
is 3 :-..-il.
A poset is graded if it is bounded, and every maximal chain has the same
length. In this case the length of a maximal chain in the interval [6, x] is the
rank of x, denoted by r(x). The rank r(S) := r(i) is also called the length
of S. For example, every chain is a graded poset, with r(C) = ICI 1, and
the boolean posets Bk are graded of length r(Bk) = k, for all k > 1.
A poset is a lattice if it is bounded, and every two elements x, y E S have
a unique minimal upper bound in S, called the Join x V y, and every two
elements x, y E S have a unique maximal lower bound in S, called the meet
x A y. (In fact, any two of these three conditions imply the third; also, if
every pair of elements has a join respectively meet, then also every finite
subset has a join respectively meet.)
If S is a graded lattice, then we call the minimal elements of S\ its
atoms, and the maximal elements of SVi its coatoms. Equivalently, the
atoms are the elements of rank 1, and the coatorns are the elements of rank
SVC),
r(S) 1.
A lattice is atomic if every element is a join x = al V ... V ak of k > 0
of atoms, where we obtain x = for k = 0, and an atom x = a l for k = 1.
Similarly, a lattice is coatomic if every element is a meet of coatoms.
We define the opposite poset Se" (or order dual) to have the same underlying set as S, with x < y in 50P if and only if y < x holds in S.
We use the graphical representation of posets by Hasse diagrams, that
is, graphs drawn in the plane so that the elements correspond to vertices,
where x < y holds if and only if there is an increasing path from x to y.
Here we only include the edges corresponding to cover relations, that is, if
x <y and [x, yl = {x, y}.
Of the posets in the following figure, the first one is not bounded, but
all others are. The second one is a lattice, but not graded. The third one
57
is graded (of length 3), but it is not a lattice. The fourth poset is a graded
lattice (of length 3), and the fifth one is even boolean (isomorphic to B3).
The fourth poset is neither atomic nor coatomic, but the fifth one is.
Why is all this interesting for us? Because we want to study the set of
faces of a convex polytope, ordered by inclusion.
Definition 2.6. The face lattice of a convex polytope P is the poset
L := L(P) of all faces of P, partially ordered by inclusion.
(i) For every polyt ope P the face poset L(P) is a graded lattice of length
dim(P) + 1, with rank function r(F) = dim(F) + 1.
(ii) Every interval [G, F] of L(P) is the face lattice of a convex polyt ope
of dimension r(F) r(G) 1.
(iii) ("Diamond property") Every interval of length 2 has exactly four
elements. That is, if G C F with r(F) r(G) = 2, then there are
exactly two faces H with GCHCF, and the interval [G, F] looks
like
(iv) The opposite poset L(P)P is also the face poset of a convex polyt ope.
(N) The face lattice L(P) is both atomic and coatomic.
58
2. Faces of Polytopes
Proof. To see that L(P) is a lattice it suffices to see that it has a unique
maximal element i = P and a unique minimal element = 0, and that
meets exist, with FAG=FnG; this is true because F n G is a face of F
and of G, and thus of P, by Proposition 2.3(h). And clearly every face of P
that is contained in F and in G must be contained in F n G.
We continue with part (ii). For this we can assume that F = P, by
Proposition 2.3(iii). Now if G = 0, then everything is clear. If G 0,
then it has a vertex y E G by Proposition 2.2(i), which is a vertex of P
by Proposition 2.3(iii). Now the face lattice of Pity is isomorphic to the
interval [M, PI of the face lattice L(P), by Proposition 2.4. Thus we are
done by induction on dim(G).
For part (i) it remains to see that the lattice L(P) is graded. If G c F
are faces of P, then from G = Pn aff(G) C Pn aff(F) = F, which holds by
Proposition 2.3(iv), we can conclude that aff(G) c aff(F), and thus that
dim(G) < dim(F). So it suffices to show that if dim(F) dim(G) > 2, then
there is a face H E L(P) with GCHC F. But by part (ii) the interval
[G, 1 ] is the face lattice of a polytope of dimension at least 1, so it has a
vertex, which yields the desired H.
Part (iii) is a special case of (ii): the "diamond" is the face lattice of a
1-dimensional polytope.
We don't prove part (iv) here -- but we will do so in the next section.
Finally, for part (v), the first part is immediate from Proposition 2.2(i),
where the atoms of L(P) correspond to the vertices of P, and the second part follows from this by taking the opposite poset, according to
part (iv).
0
This theorem contains quite restrictive information on the structure of
polytope face lattices (Exercise 2.3). We will get even more precise information later.
We note here that the face lattice is the proper framework to define combinatorial equivalence of polytopes. In fact, our previous definition (before
Example 0.2) can be restated as saying that P and Q are combinatorially
equivalent, P ,--' Q, if and only if L(P) r' L(Q): if their face lattices are
isomorphic.
By Proposition 2.2(i), this is equivalent to a bijection vert(P) 4-* vert(Q)
between the vertices of P and Q, in such a way that the vertex sets of faces
of P correspond (under this bijection) to the vertex sets of faces of Q. A
general observation is that in this context it is enough to deal with vertices
and facets, because the faces are exactly the intersections of facets, and the
vertex sets of faces are exactly the intersections of vertex sets of facets
see Exercise 2.7. (Abstractly, the key properties are that face lattices are
atomic and coatomic.)
Topologically, combinatorial equivalence corresponds to the existence of
a (piecewise linear) homeomorphism between the polytopes P-2'-- Q that
restricts to homeomorphisms between the facets (and hence all the faces)
2.3 Polarity
59
2.3
Polarity
(ii) if ay = ao and a
Eid=0= 1 ,
(iv) y can be represented as y =
points x 0 ,
xd c P.
d+1 i=0x
Proof. Part (i) implies that P is full-dimensional. From this we get that
part (ii) holds: if ax < a0 were valid for P, then y would be contained in
the face P n {x c d : ax = ao} of smaller dimension. Conversely, if part
(ii) holds, then no inequality can define a facet that contains y.
Part (iv) trivially implies part (iii), and from this we get part (ii) by an
easy calculation:
d
ao = ay
Ai axi
Ai ao = a.
i=0
60
2. Faces of Polytopes
1
1(Y a`
d +1
ed +
(ii) if ax < a0 is valid for P, with equality for y, then ax = ao holds for
all x P,
(iii) y can be represented in the form y =E ik_0 A i x i for k + 1 a.ffine/y
independent points x 0 ,... , xk E P and for parameters Ai > 0,
(iv) y can be represented as y = k _1_ 1
dent points x 0 , .. ,x k E P.
Eik_o xi
relint(P) 0. To see this, we can take the barycenter of the vertex set of P,
vi for vert(P) = {v 1 , ... ,vN}, or the barycenter of any
as y :=
set of dim(P) +1 affinely independent points (e.g., vertices) in P according
to Lemma 2.9(iv).
2.3 Polarity
61
P=
Lt j
relint(F).
FEL(P)
Clearly, the construction of the polar can be iterated, and thus we get
the polar of the polar, or double-polar, as
where we have identified Rd and (Rd)** in the natural way. Now let's examine the nearly obvious (?) basic properties of the polar and double-polar
constructions.
62
2. Faces of Polytopes
Theorem 2.11.
(0 P C Q implies P D QA and PA C QA
(ii) P C PA ,
(iii) P and PA are convex,
(iv) 0 E PA , and 0 E PA ,
fa : aV < 1 1,
P =
=
PI
where for the last equality "C" is trivial, while "D" follows from convexity
(or from a trivial computation).
For part (vii), we compute
PA =
2 3 Polarity
63
From this, by Farkas lemma I, there exists a c" > 0 with c"A = 0 and
el > O. With this we can put
i. - c/i_ t
e
c" l
1
F := lc
Theorem 2.12.
in Rd, and that
( )
For the other half, we use the description of P in parts (vi) and (vii) of
Theorem 2.11, and get
F<> =
64
2. Faces of Polytopes
where for the last equality, "D" is clear, while for "C" we have to work.
In fact, for this we can choose some x E relint(F), which satisfies Aix = 1
and A"x < 1, and rewrite cA ,----- c' A' + c"A" . Then by
1 = (cA)x --, (c` A i +c" Anx .---- ci (A`x)+c" (A" x) < ci l+c"1..----- cl ,--- 1
we have c"(A"x) = c"1, which by A"x < 1 implies c" = O.
(i) F is a face of P ,
(ii) F" = F, and
(iii) F C G holds if and only if F D G .
Corollary 2.14.
of P:
This, in particular, completes the proof of Theorem 2.7, the last two parts
of which we had deferred (remember?). It means that for every statement
about the combinatorial structure of polytopes, there is a "polar statement," where the translation reverses inclusion of faces, and interchanges
4-9. i = P
4---+ facets
vertices
edges ) ridges
etc.
Note that polarity also identifies the face lattices of facets with (the opposites of) the face lattices of vertex figures. Finally, it says that the "polar"
combinatorial descriptions of polytopes, as V-polytopes in terms of vertices,
and as 7-1-polytopes in terms of facets, are logically equivalent.
Nevertheless, the metric properties of the polarity construction depend
on the location of 0 in P, whereas the combinatorial ones do not. This
motivates to define that two polytopes P and Q are cornbinatorially polar
if L(P) '' L(Q)P. Thus the construction of P establishes the existence
of a combinatorially polar polytope for every polytope P.
2.4
This section has a (by now) simple task: to state and prove the general
representation theorem for polytopes. There is no real work left to do: we
have assembled all the ingredients, notably Fourier-Motzkin elimination,
the Farkas lemmas, Carathodory's theorem, and polarity. One new term
appears in its statement: the k-skeleton of a polytope is the union of its
k-dimensional faces.
65
66
Faces
of Polytopes
is a
This answers one question, and opens up two new ones: first, is this true
for all polytopes? We will see later that it holds for polytopes of dimension
d < 3 (the case d < 2 is trivial), but it fails in general (see Lecture 6).
Second, if integral coordinates exist, can we keep them reasonably small?
Again, the answer is yes if we are in low dimension, but in general we
have to cope with coordinates that grow doubly exponential in terms of
the number of vertices; see Goodman, Pollack Si Sturmfels [220]. For fixed
dimension d = 3, however, the problem seems to be open (Problem 4.16*).
We already saw numerous examples of simple and simplicial polytopes
in Lecture 0. The d-simplex, the d-cubes, and the dodecahedron are simple
polytopes. The d-simplex, the octahedron, the icosahedron, and all cyclic
67
2.6
Although linear transformations are our main tool to "put polytopes where
we need them," it is sometimes convenient to use more general transformations, which allow us even to "adjust the shape of a given polytope,"
known as projective transformations.
We can describe projective transformations in a very simple way with the
tools of this lecture (in particular, without construction of projective space
and use of projective geometry). For this, we proceed as follows. Given a
polytope P C j d, we embed it into an affine hyperplane H C d + 1 , and
construct homog(P), the homogenization of P. By construction, this is a
pointed cone. Now we cut this cone by a different hyperplane K C d+1,
which is then identified with Rd by an affine map.
-3111*
L(P)
L(hornog(P))
L(P`).
2. Faces of Polytopes
68
x
H = {( ) : x E Rd }
1
and
d+1
Xd+1
1(
--= {
Rd+1 :
ax + ad-4-1X d+1
Xd+1
11.
The hyperplane K is admissible if and only if (a, ad+1)(11') > 0 for all
vertices y E vert(P). We map K back to d via an affine map
( X )
Bx +
Xd+1
det(
if and only if
B z
a ad-Fi
O.
x )
xd+i
+ z'.
0, and thus
x.
E H
1
x)
EK
ax + ad+1 ( 1
Bx + z
+ E Rd.
ax + ad-Fi
Bx + z
+ : x E P}
ax + ad+1
under the conditions that
det
(B
a
z
ad+i j
Notes
69
.
K . ... ,
. ,
\
.....4110,
Notes
All the basic facts about polarity, the face lattice, and the various parts
of the representation theorem and their proofs are classical, due to Farkas,
Weyl, Minkowski, Carathodory, Motzkin, Kuhn, and others. Again we refer to Griinbaum [234] and Schrijver [448] for the history. Anyway, "history
will teach us nothing" (Sting).
(This was a message from our No Comment department.)
70
2 Faces of Polytopes
2.2 Show that every polytope is affinely isomorphic to a bounded intersection of an orthant with an affine subspace.
2.3 Construct a small poset that satisfies the conditions of Theorem 2.7
but does not correspond to a convex polytope. Does your example
correspond to some geometric object?
2.4 Prove directly (i.e., without using polarity) that every face of a polytope P is contained in a facet.
2.5 If two 0/1-polytopes are combinatorially equivalent, does it follow
that they are affinely isomorphic? (The answer is "no.")
2.6 Let f(d) be the number of combinatorial equivalence classes of d-dimensional 0/1-polytope,s. The first values are f(0) = f(1) , 1,
f(2) = 2, and f(3) = 8. Prove that 2 2d-2 < f(d) <22d for d> 5.
(This solves a problem of Billera & Sarangarajan [72, Sect. 3]. For the
lower bound A. Sarangarajan and I suggest that you consider all the
polytopes of the form P = cony(S) for sets S C {0, 1}d that satisfy
There are 22d-1-4 such polytopes P(S); show that their combinatorial
equivalence classes are "small.")
2.7 Assume that one is given the vertex-facet incidence matrix
Al(P)
71
(iv) Assume that of a polytope P you are given the dimension, the
vertex set, and a matrix M(P) e 10, 11'" such that every row
of M(P) represents a facet of P.
How can you tell whether this list of facets is complete?
(Remark: this is not too easy; one can use tools from Chapter 8,
or from homology theory.)
(Part (ii) is important: See Klee & Minty [302, p. 167], Amenta &
Ziegler [17], and elsewhere. (iii) points to an error in [302, p. 167].)
2.9 Define the face figure PI F for any face of P by Pf F (F*) , that
is, a polar of the face of P which corresponds to F. (The face figures
F are also known as the quotients of P. Thus a quotient of P is
the same thing as an iterated vertex figure.)
Show that this is a polytope of dimension
72
2. Faces of Polytopes
Show that this definition (with "0" instead of "1") is a special case
of our definition for arbitrary subsets.
Formulate and prove the analogs of our Theorems 2.11 and 2.12.
P = P(A,z).
An inequality in this system is called redundant if deleting it from the
inequality system Xx < z does not change the polyhedron; otherwise
the inequality is called irredundant.
(i) Derive from Farkas lemma III that an inequality is redundant if
and only if it can be written as a positive combination of other
inequalities in the system.
(ii) Derive from a Farkas lemma that if P = P(A,z) (6 and if none
of the inequalities of a system Ax < z is redundant, then each
of them defines a facet. Thus, every polytope is the intersection
of its facet-defining inequalities.
73
P =- P(A, z).
With every inequality ax < zi in this system, associate its vertex
set
Vi
E V : at v = zi l.
Show that the following criteria can be used to check whether an
inequality is redundant.
(i) The inequality ax < z is redundant if and only if 3/4 C Vi for
some j i.
(ii) If 14 = Vi, then either the inequalities are multiples of each
other, or they can both be deleted from the system.
(iii) An inequality is irredundant if and only if it defines a facet of P
and no multiple of it is contained in the system.
(iv) If 13/41 < d, then the inequality ax < z -
redundant.
P = P(A, z),
is known, such that the inequalities ax < zi describe all the distinct
facets of P, without duplication. For every inequality ax < zi in the
system, let 3/4 be the set of points in V which satisfy it with equality.
Show that from this, an irredundant description of proj d (P) C
can be obtained from the following criteria:
d-1
aid
aidzi
( aik)z i
74
2. Faces of Polytopes
Furthermore, show that if I Vi n vi < d 1, then the combined inequality in (ii) is redundant. In particular, this is the case if d > 2 and
n Vi = 0. (Give geometric proofs they are easier than algebraic
ones!)
Explain how, by using Fourier-Motzkin elimination (Theorem 1.4)
together with these redundancy criteria, one can obtain an complete
irredundant description of P
conv(V) C Rd, even if the set V
contains more points than just the vertices of P.
How can the criterion be adapted for the case of polyhedra, where
the input is a polyhedron given as P =-- conv(V) + cone(X)?
What happens in the situation where dim(P)
difficulty there be overcome?
75
2.19 Using a Farkas lemma, show that for every unbounded pointed polyhedron P there is an inequality ax < 1 such that
:= {x E P : ax <1}
is a polytope with a facet F' :=
E P : ax = 11, such that the
k-faces of P correspond to the unbounded (k+ *faces of P, and the
k-faces of P' that are not faces of F' are in bijection with the k-faces
of P.
= p.
1/cos( u)
R
It(
y(u) :=
sin( u)
cos(2u)
sin(2u)
cos (du)
sin(du) /
(0
,Y(ud)}
is combinatorially equivalent to the cyclic polytope C2d(n)(Hint: You will find a useful hint in Griinbaum [234, p. 67, Ex. 23].)
76
2. Faces of Polytopes
(n)
2y2 + y4
1
1-4
3 + 4y1 y3
y4
2y2 y4
3 4yi y3
:=
1 cos(u)
sin(u)
sin(u)
= 1 + cos(u)
2.22
sin(2p/r)
cos(2q71- -7 )
sin(2q7r)
3
Graphs of Polytopes
3.1
78
3. Graphs of Polytopes
Ix E
: ai x = 1}
HI :=
E Rd aix <1}
denotes the closed halfspace that contains P, and 11;1- denotes the other
closed halfspace. In particular,
n H2 n n H-
Yv
79
The following lemma shows that a direction vector for such a line can be
found arbitrarily close to any given vector.
Let P = P(11,1), and let u E Rd\O. If A > 0 is small
enough, then the line gu(A)) is in general position with respect to P, for
Lemma 3.2.
( A \
u
A:d
Proof. We use that si (A)
au") = Edk=1 at k(uk + Ac) is a nonvanishing polynomial in A of degree at most d, which has at most d positive
zeroes. The polynomials si (A) are distinct, since ai ai for i j.
From this we get that ai u( A) 0 0 for all, except at most d, positive values
of A, and that aiu(A ) ai u(A ) for all positive values with not more than
(3)d exceptions.
LI
80
Graphs of Polytopes
Lemma 3.4. Let P = P(A, 1), and let c E ( )*\ O. If A > 0 is small
enough, then the linear function c(A) x is in general position with respect
to P, for
(In this whole course we need very little graph theory, only some terminology. When in doubt, look it up in any graph theory book. For that
purpose, even a mediocre book would do. As for good ones, we recommend
Bondy Sz Murty [116 ] , Bollobils [11.4 ] , or Tutte [5141.)
We will consider orientations of G(P), which assign a direction to every
edge. An orientation is acyclic if there is no directed cycle in it. This implies
(because all our graphs are finite) that there is a sink: a vertex that does
not have an edge directed away from it. (Proof: Start at any vertex, and
keep on walking along directed edges until you close a directed cycle or get
stuck in a sink.)
Linear programming is (in a geometer's version) the task to find a point
xo E P that maximizes a linear function cx, that is, such that czo =
max{cx : x E PI =: co. Now we easily see that the maximum is achieved
in a vertex. In fact, Fo := Ix E P : ex = co l is a face of P, and thus every
vertex of Fo maximizes cx.
FO
81
I
Now if c is in general position, then this gives us a well-defined way to
direct the graph of P, by directing an edge convfyi , vi I from vi to vi if
cvi < m3 . (Because of the general position assumption, ties cannot occur.)
We call this the orientation of G(P) induced by e.
With this construction monotone paths on P (edge paths for which the
objective function increases strictly in each step) translate into directed
paths in the orientation of G(P) induced by c.
Lemma 3.6. Let y E vert(P) be a vertex, and let N(v) be the set of its
neighbors in G(P). Then the cone (based at y) spanned by the neighbors
of y contains P:
P C + cone{u v : u E N(v)}.
82
3 Graphs of Polytopes
Proof. This follows from our proof of Proposition 2.4: the neighbors of y
are in one-to-one correspondence with the vertices of the vertex figure P/v,
and thus it is equivalent to say that those vertices span a cone that contains
P. But we have also seen there that every ray emanating from ty to any
other point x E P contains a point of the vertex figure. This yields
y + cone{u - y : u E N(v)}.
III
83
Au (2, n) --..: n 2.
Is this plausible? Here are a few observations, most of them due to Klee
Sz Walkup [303].
84
3. Graphs of Polytopes
The Hirsch conjecture is true for d < 3 and all n (even in the monotone and unbounded versions discussed below, by Klee [2931), and for
n d < 5, by Klee & Walkup [3031.
One can show that for Conjecture 3.8 it is sufficient to consider simple
polytopes (see Exercise 3.5).
e If n < 2d, then any two vertices lie on a common facet. From this we
get d, n) < 6(d 1,n 1); iterating this, we get
A(d,
for
n < 2d.
Similarly, we get A u (d, n) < At, (n d, 2(n d)). In both cases these
inequalities hold with equality [303 ] : this is quite obvious in the unbounded case. Thus we restrict our attention to the case n> 2d.
More surprisingly [3031, the Hirsch conjecture for all dimensions would
follow if one could prove it for n = 2d for all dimensions. The special
case n =z 2d has become known as the d step conjecture.
Consider two vertices that do not lie on a common facet. Since each
of them lies on d facets, we see that the d-step conjecture concerns
a very special geometric situation: after a change of coordinates we
can assume that the first vertex y is given by y = 0, where the facets
it lies on are given by x, > 0, which describes the positive orthant
x > O. Then the other vertex u can be assumed to be u = 1, and its
facets describe an affine image of the positive orthant.
-
In this situation there are d edges leaving from u, whose other endpoints are on the hyperplanes {x : x, = 0 } . The claim is that we can
get from u to V in d steps.
From the special case of the d-cube we get that A(d, 2d) > d. Thus
the bound suggested by the d-step conjecture is certainly the best
possible, if it holds. Furthermore, Holt & Klee [2581 have shown that
A(d, n) > n d
that is, the Hirsch conjecture is also best possible for all n, if the
dimension is high enough.
85
If you look for counterexamples, a natural guess would be to consider the polars of cyclic polytopes C d(n) 4 ' , or more generally the
polars of neighborly polytopes since they have the largest numbers of vertices for given n and d (according to the upper bound
theorem; see Section 8.4). However, Klee [297] has shown that the
polars of cyclic polytopes satisfy the Hirsch conjecture. Beyond that,
Kalai [277] could prove that if P is the polar neighborly d-polytope
with n facets, then one has at least a polynomial diameter bound
(5(G) _<,_ & (n d) d log(n).
3. Graphs of Polytopes
86
Klee & Walkup [303] showed that the Hirsch conjecture is also false
for unbounded polyhedra although Hirsch's original conjecture
was asked for unbounded polyhedra. They proved that for n > 2d,
A u (d, n) > n d l_df5j. This is the best lower bound known for
.A.,i (d, n).
Hu (d, n) <n d,
respectively
H(d, n) <n d.
d min{ H
d]
i ,
In
dl
In particular, there is a 4-polytope with n = 8 facets for which every monotone path to the top needs at least five steps. However, in Todd's example
there is a two-step nonmonotone path, which first goes to the bottom, and
then directly to the top! This motivates the following, more restrictive,
version of the monotone Hirsch conjecture, which might as well be true
and which would imply the Hirsch conjecture (via a simple argument using
projective transformations; see Exercise 2.17).
Conjecture 3.9 (Strict monotone Hirsch conjecture).
Let P be a d-dimensional polyt ope with n facets, and let ex be a linear
function that is in general position with respect to P.
Then there is a strictly increasing path with respect to ex, from the
(unique) vertex v min that minimizes cx, to the (unique) vertex vmax that
maximizes ex, of length at most n d.
87
V max
CX
Vmin
To illustrate this for a trivial case, observe that for an n-gon the length
of a shortest monotone path "to the top" can be n 2 =-n), but if
we start "from the bottom," then we need at most 1 121-i
n) steps.
n)?
What about upper bounds on A(d, n) and
In 1967 Barnette [36, 234] proved that A u (d, n) < n3 3 . An improved
bound, Au (d, n) < n2d-3 , was proved in 1970 by Larman [321]. Barnette's
and Larman's bounds are linear in n but exponential in the dimension d.
After that, nothing happened for a long time. In short, we might summarize
the history by saying that the experts thought that the conjecture was
plausible until they tried to prove it and couldn't; therefore now they think
it is false, and can't prove that. However, in the long run Kalai might prove
to be right, when he writes about "the author's guess (which is as good as
the reader's)" [278]. The existence of a polynomial (or even linear) bound
for A(d,n) is still a major open problem...
However, recently Gil Kalai achieved a substantial breakthrough: in a
sequence of papers (each simpler and more striking than the preceding
one) he established the first subexponential bounds for the diameter of a
polytope. In November 1990 he proved flu (d, n) < n2V7L [278, Sect. 3].
In March 1991 he derived a "pseudopolynomial" bound for the diameter
problem [278]:
A v (d,n)
88
3. Graphs of Polytopes
Hu (d, n) <
Proof. The key to this is the notion of an active facet: given any vertex y
of a polyhedron P, and a linear function ex, a facet of P is active (for y)
if it contains a point that is higher than y (that is, either the facet is
unbounded with respect to ex, or it has a top vertex w with cv < cw).
For this proof, we also admit problems for which ex is not bounded
on P, and where the last step "to the top" takes a ray (unbounded 1face) on which ex has no upper bound. (You may think of the top as an
extra vertex u,, in this case, which is adjoined to the directed graph of the
problem.)
Let ft(d, n) be the number of steps that may be required to get to the
top vertex if we start from a vertex y for which the polyhedron has at most
n active facets (and an arbitrary number of nonactive ones!).
Since Hu (d, n) is monotone in n we immediately get
A (d, n) < Au (d, n) < Hu (d, n) < H (d, n).
Thus it suffices to prove the bounds of the theorem for il(d, n). In the
following we require d > 2 and n > O. In the "boundary cases" we get
1(2, n) = n
(all the edges on a monotone path to the top are active facets, and this
may be all of them if the problem is not bounded), and
= fl (d, d 2) = 0,
(if y is not the top vertex, then it has an increasing edge, which lies on
d 1 active facets).
To get a recursion for R(d, n), we verify a sequence of four simple facts:
Given any set T of k active facets of P, we can reach from y either the
top vertex, or a vertex in some facet of T, in at most H(d, nk) monotone
steps.
Let "Ax < z" be a minimal system that defines P (having one inequality
for each facet of P), and let P' := P(A', z') be the polyhedron obtained
by deleting the inactive constraints that don't contain y as well as all the
inequalities that correspond to facets in T. Then y is a vertex of P' (unless
y lies on a facet in T, in which case we have nothing to prove), and it has
at most n k active facets in P'.
Now consider a shortest path from y to the top in P'. This path makes
at most H(d, n k) steps, by definition. If it touches a facet in ,F after at
1.
89
most fi(d, n - k) steps on P, then we are done. If it doesn't, then the top
vertex of P, is also the top vertex of P, and the path to it in P' also yields
a path to the top vertex on P, of length at most 11(d, n - k).
2. If we cannot reach the top in _TI (d , n-k) monotone steps, then the collection g of all active facets that we can reach from y by at most fI(d,n - k)
monotone steps contains at least n - k + 1 active facets.
If there are k facets that cannot be reached, we can delete these facets
together with all the inactive facets, and get a problem where we can reach
the top in at most ft(d,n - k) steps; however, the path in this reduced
problem corresponds to the same path in the original problem, leading to
the same top vertex: Contradiction.
3. Starting at y, we can reach the highest vertex Iv () contained in any facet
F E g within at most H(d,n - k) + f1(d - 1, n - 1) monotone steps.
We need at most 17/(d, n - k) steps to reach any facet of g; this facet (of
dimension d - 1) has at most n - 1 facets, thus in it we can find a path to
its top of length at most 1-1(d - 1,n - 1).
4. From w o we can reach the top in at most H(d, k - 1) steps.
This is because none of the facets in g is active for inch and thus wo has
at most n - (n - k + 1) = k - 1 active facets.
Putting the monotone paths together, we get a bound
f (d, t) := 2- t 11(d, 2t )
f(d,t) <
-
d + t - 3)
d-2)
90
3. Graphs of Polytopes
for (d, t) (2,0), by induction on t > 0 and d > 2. From this we derive
H(d,21+Ll0g2 mi )
-.=
2 1+ Llog2 n -I f(d,
<
2n
<
2n (d - i) 1g2(n) =
1 Llog2 n j )
(d + Llog2 n j 2)
d 2
2 n i-Flog2 (di) ,
for n, d > 2, using the inequality ( t. ()) < (a -I- 1) b , which follows by induction over a > 0 and b> 0.
(In fact, there are various ways to derive bounds on fl(d,n) from the
recursion. This is a standard type of gymnastics for which you should get
training at your "analysis of algorithms" class. Here is another way to proceed, which obtains the original Kalai-Kleitman bound. We use the starting
values H(2, n) = n and ft(d, 0) = 0. Since H(d,n) grows monotonically
in n, we get a simple recursion
H(d,n)
<
i=3
<
<
91
best possible, that is, any proof for a substantially better upper bound
has to use more of the specific geometry of the problem. Not much of the
geometry was used in the preceding proof. (In fact, Kalai [278, Sect. 4]
indicates a very general abstract framework, of a "simplicial complex with
a fixed shelling order" (see Lecture 8), in which such upper bounds can be
proved.)
Finally, let us mention that the diameter bounds can indeed (not quite
directly) be used to construct algorithms for linear programming. In his
research, Kalai [279, Sect. 3] found randomized pivot rules for linear programming that roughly require an expected number of n4 `fd arithmetic
operations for every linear programming problem of dimension d with n
facets. See Exercise 3.9(ii) for a simple sketch, and [281] for the latest version.
Very similar results were reached independently and nearly simultaneously (on a completely different path) by Matouek, Sharir & Welz1 [355],
in the setting of a "dual simplex method."
/ /
d=2
d=2
d=2
dim(P) = 1
dim(P) = 2
dim(P) = 2
6(G(P)) =1
b(G(P)) = 1
S(G(P)) -- 2
d=3
d=3
d=3
dim(P) = 2
S(G(P)) = 2
dim(P) = 3
S(G(P)) = 2
dim(P) = 3
6(G(P)) -= 3
92
3. Graphs of Polytopes
y,
Id := [0,11d = conv({0,
d-1
1-0'
(Xxd)
Fw P n E
= 0 whenever wi 01
P(i) = P n fx E
xi
= 01
93
xi = 0}.
94
Graphs of Polytopes
The existence of good acyclic orientations of G(P) follows from Theorem 3.7: if cx is in general position for P, then it is also for all faces
of P. Our first goal is to distinguish intrinsically between good and bad
orientations of G(P).
95
respect to O. But vert(H) includes the set of all vertices that are < x with
respect to O. (Remember: vert(P) is an initial set with respect to O.) Thus,
vert(F) c vert(H). Since both H and G(F) are k-regular and connected,
vert(F) = vert(H) and G(F) = H. This completes the proof.
Remarks 3.13.
You could ask: do these parameters h? actually mean anything? They
do
we will come back to this when we study shellability of polytopes, in Lecture 8.
1. We do not have a practical way to distinguish between good and
bad orientations. The algorithm suggested by the preceding proof is
exponential in Ivert(P)I. We do not even know of an efficient way to
compute the number of facets of P from G(P).
2. It is false that general polytopes can be reconstructed from their
graphs this can be seen, for example, from the existence of neighborly (simplicial) polytopes.
Perles [405, 407] proved that simplicial d-polytopes are determined
by their [d/2]--skeleta; see Kalai [280]. General d-polytopes are determined by their (d-2)-skeleta, and this is best possible even for quasisimplicial polytopes (all of whose facets are simplicial polytopes); see
Grfinbaum [234, Ch. 12].
0.
[33]
The graph G(P) is d-connected for every d-polytope P.
96
3. Graphs of Polytopes
Notes
The graph of a polytope is treated with care in Griinbaum's book [234,
Chapters 11, 13 and 16].
As for linear programming, this is usually described in a much less geometric way, which is better suited for algorithmic treatment. Also, there is
of course much more to say than our simplified sketch in Section 3.2. We refer to the books by Dantzig [166], Chvg,tal [151], Schrijver [448], Grbtschel,
LovAsz Sz Schrijver [228], Padberg [404], and Borgwardt [118], and to [92,
Ch. 10] for various different aspects of the matter. As for "Dantzig's simplex algorithm" [166], let us just mention that it was already developed by
Kantorovich in the 1920s, but could not published in any reasonable form
for reasons that were equally ideological and stupid [285].
Klee & Kleinschmidt [299] is an inspiring survey on the Hirsch conjecture and its relatives; see also Kleinschmidt [307]. The material in Section
3.3 is derived from the papers by Kalai Sz Kleitman [278, 282, 279]. In
particular, the proof of Theorem 3.10 is from Kalai [279, Sect. 2]. Our discussion of 0/1-polytopes is based on ideas by Naddef [389] and by Kleinschmidt [308]. The observation in Theorem 3.11 that the extreme case here
is only achieved for d-cubes seems to be new (although not deep).
97
In a recent preprint, Lagarias, Prabhu Sz Reeds [319] discuss the configuration space of all d-step configurations for a fixed d, analyze its structure,
and relate the d-step problem to certain factorization problems for matrices.
They also suggested that there might in fact be at least 2c1-1 paths of length
d between the complementary vertices of any d-dimensional Dantzig figure.
However, Klee & Holt [257] have now shown that this is true for d < 4, but
false for all d> 4.
Our treatment of the reconstruction of polytopes from their graphs owes
heavy thanks (thefts) to the paper [272] of Gil Kalai, as indicated there.
We might repeat here that the graph of a polytope carries important information, but by far not all the relevant information about the structure of a
polytope. One aspect of this is the fact that the graph does not determine
the dimension of a general polytope; see also Exercise 3.4. Another one is
that there are far fewer different polytope graphs for various parameters
than there are different polytopes. In fact, Perles proved that the number
of nonisomorphic graphs of d-polytopes with d + k vertices is bounded by a
function of k (independent from d!). The proof for that see Kalai [280]
uses only some lemmas about finite set systems. In contrast to Perles'
result one can easily see (for example with the methods of Section 6.5) that
the number of different d-polytopes with d + 2 vertices (i.e., k = 2) is not
bounded.
1)
= kl.
98
3. Graphs of Polytopes
P` -= P(A,1 (A) ),
with 1 (A) , = 1+Ai as in Lemma 3.2, is a simple polytope whose facets
are in natural bijection with the facets of P.
Furthermore, show that then 45(G(11)> S(G(P)): thus it is sufficient
to prove the Hirsch conjecture for simple polytopes.
3.6 If P is a pointed polyhedron in R3 , show that the graph of all bounded
edges is connected. Show that it is not 2-connected in general. What
about higher dimensions?
3.7 Let P c Rd be a d-polytope with 2d facets, such that the facets
containing y = 0 determine the positive orthant u > O. Show that
the facets of P can have 2d 1 facets each, if d > 4. (This is why it
is hard to use inductive arguments for the d-step conjecture.)
99
<
E(d 1, n 1) + ji
E(d,n i)
}-
100
3.10
3. Graphs of Polytopes
z) C
(Basis version of linear programming). Let P
and let cx be a linear function on Rd . A subset of d of the inequalities
of P, say Aix < is a basis if A' has rank d (equivalently, A'x = z'
has a unique solution x E Rd ). A basis A'x < z' is feasible if the
unique solution of A' x = z' satisfies x E P, and dual feasible if it
maximizes cx over Ix E Rd : < (equivalently, if 0 maximizes
< OD.
cx over Ix E Rd :
(ii)
(iii)
(iv)
(xi)
101
3.15 A complex A is a collection of d-subsets of the set [n] := {1, 2, ... , n}.
(This defines an "abstract simplicial complex" in the sense of Section 8.5.) Two sets F, G E A are adjacent if they differ only in
one element. This defines a graph on the d-sets in A, and A is
called strongly connected if this graph is connected. Even stronger,
A is called ultraconnected if every nonempty subfamily of the form
AK := {FE A.K CFI is strongly connected.
(1) Show that the complex which corresponds to a simplicial dpolytope with vertex set (identified with) [n] is ultraconnected.
(ii) If P is a simple d-dimensional polyhedron whose set of facets is
labeled by [n], then there is a d-set associated with every vertex.
Show that the corresponding complex is ultraconnected.
(iii) Every shellable simplicial complex is ultraconnected. (Shellability is an important combinatorial concept: see Section 8.1 for
the definition.)
More generally, a pure simplicial complex is shellable if and only
if it has an ordering F1 , F2 ,. .. , F8 of its facets such that the subcomplex F1 UF2U...UFs is ultraconnected for all i.
(iv) Let A be the complex of the 4-sets 1234, 2345, 1346, 5678, 2678,
1578, and all the 4-sets that have two elements from 1234 and
two elements from 5678 but do not contain both 1 and 2 nor
both 5 and 6. Show that the distance between 1234 and 5678 in
A is 5. Show that A is ultraconnected.
(N) Describe an ultraconnected complex of triangles (i.e., d = 3) on
n vertices with diameter n 3.
(vi) Let A be an ultraconnected complex of triangles on n vertices.
Show that between any two triangles there is a path of triangles
which visit every vertex at most twice. Deduce that the diameter
is at most 2n.
(vii)* Can you improve 2n to 1.999n (or at least to 2n-1000, 2n-1)?
Can you find an ultraconnected collection of triangles on n vertices with diameter > 1.001n? (or at least n+100, or even n-2?)
(This combinatorial set-up for studying diameter questions is due
to Larman [321] and to Kalai [278]. In particular, in [321] Larman
showed that between every two vertices of an ultraconnected complex of d-sets on n vertices, there is a path that visits every vertex at
102
3. Graphs of Polytopes
most 2d-1 times, and this implies a bound of A u (d, n) <2d-mn. (See
also Klee & Kleinschmidt [299, Sect. 7].)
The unbounded 4-polyhedron with 8 facets by Klee & Walkup [303],
which fails the Hirsch bound, yields the complex of part (iv).
Subexponential diameter bounds for ultraconnected complexes were
found by Kalai, see [278, Sect. 4.1]. This exercise is also due to him.
Part (vil)* demonstrates the unbelievable gap between the known
lower and upper bounds. Note that by parts (i) and (ii), every upper bound that one can prove for the diameter of an ultraconnected
complex of d-sets is automatically also valid for Au (d, n).
3.16 Let P be a simple d-polytope such that every k-face of P has at most
2k facets.
(i) Show that the diameter of P is bounded above by d
(ii) Moreover show for such polytopes that for every objective function and any starting point one can "reach the top" in d steps.
(This is from Kalai [278, Thm. 3], where it is proved that for every
fixed r > 2, if every k-face of P has at most rk facets, then the
diameter and the hight of P are bounded by a polynomial in d.)
3.17 Given finite graphs G and H, we define that G is an induced subgraph
of H if we can obtain a graph isomorphic to G by deleting a set S of
vertices (and all edges adjacent to them) from H. We say that H is
a suspension of G if additionally we require that the vertices of S are
connected to all other vertices of H.
Show that every finite graph is an induced subgraph of the graph
of a 4-polytope.
(If G has n > 5 vertices, then start with C4(n), and introduce
an extra vertex beyond every edge that is missing in G.)
(ii) For every finite graph there is some suspension which is the
graph of a d-polytope, for some d.
(If G has n > 5 vertices, then start with C4(n), and introduce
an extra dimension and two new vertices for every edge in G
that is supposed to be missing.)
(ill)* Does every finite graph have a suspension that is the graph of a
4-polytope? What about the case where G is the graph with n
vertices but no edges?
(iv) Give an example of a 4-connected graph on n > 5 vertices which
is not the graph of a 4-polytope. Can you construct a 4-regular
graph with these properties?
(i)
(Perles [4081)
4
Steinitz' Theorem for 3-Polytopes
[487, 490]
G is the graph of a 3-dimensional polytope if and only if it is simple, planar,
and 3-connected.
Polytopal graphs are certainly simple: they have no loops or multiple
edges. The graph G(P) is planar for every 3-polytope (use radial projection to a sphere from an interior point, or a linear projection to a plane
from a point beyond a facet). Also, G(P) is 3-connected by Balinski's Theorem 3.14. Thus, the difficult part is the "if" part of the theorem: it requires
that we show how, given a 3-connected planar graph, one can construct a
3-polytope.
104
4.1
Again we need some basic graph theory. Let us, for a while, admit nonsimple
graphs as well, that is, graphs that can have loops and parallel edges.
Such a graph G is connected if there is a path between any two distinct
vertices of G. All the graphs we consider will be connected.
A graph G with at least 2 edges is 2-connected if it is connected, has
no loops, and cannot be disconnected by removing one vertex and all the
105
106
Two very basic "local" operations on graphs are the deletion of edges
-----' >
>
and the contraction of edges,
>
>
for which the two vertices of the edge are identified. Any graph that can
>
{ -->. >
or delete edges that are in parallel with others (this is the usual operation
for "making a graph simple").
f
We refer to the figure for the "natural" correspondence between the edges
of the triangle and the edges of the 3-star. Note that these operations,
replacing a K3 by a K1,3, preserve the number of edges in the graph.
107
This duality can be carried into our reduction for polytopes, because the
graph of a 3-polytope is exactly the dual graph of the polar polytope:
G(P)* = G(P).
4.2
108
Here the dotted lines in our sketch denote edges that may or may not be
present, and are not affected by the simple A-to-Y reduction.
Similarly, we get four types of simple Y-to-A transformations when we
consider a vertex y of degree 3 and distinguish how many of its neighbors
are already connected:
4-
AAAA
These four transformations are exactly the "polar operations" (operations in the dual graph) for the simple A-to-Y reductions.
Lemma 4.3. Let G be a 3-connected planar graph, and let the graph G'
be derived from G by a simple AY transformation.
If G' is the graph of a 3-polytope, then so is G.
109
ii
4.3
In this section we show that every 3-connected planar graph (with n > 4
edges) can be reduced to K4 by a sequence of simple AY transformations.
This is a special case of a much more powerful theorem (for 2-connected planar graphs plus a "return edge") that was first established by Epifanov [186]
and has a clever and simple proof by Truemper [509]. In this section, we
follow his expositions in [509] and [510, Sect. 4.3].
For a while, we will only require that the graphs considered are 2connected; in particular, we admit AY operations if they keep our graphs
2-connected.
In the simplicity of Truemper's approach to Epifanov's theorem, the
reader should appreciate the "power of a normal form theorem" (in this
case: the embedding of a planar graph as a minor of a grid graph, which is
also at the heart of Robertson Sz Seymour's work [428] on graph minors).
Here come three lemmas and a corollary, which together prove everything. [Working through this, you can practice "three levels of reading":
first read only the lemmas, and try to understand what they mean and
how much is trivial_ Second glance over the proofs, and try to see whether
you know how to do them yourself. If you know, just do it. If you don't,
try to find counterexamples. In the third step, work your way through the
proofs until you are confident that you found all the errors I made and the
shortcuts I missed. Tell me about them.]
110
G':
Let e, f, g be the three edges of G that are involved. One possibility is that
all three edges are contained in H: then they form a nonseparating triangle
in H as well, and we can perform the corresponding A-to-Y step H ---> H'.
Then by induction we get that H' is AY-reducible, and hence so is H.
In the other case, some of the three edges e, f, g do not appear in H.
What happened to them? Since H is simple, it is not possible that only
one or two of them were contracted. If all three were contracted, then we
can assume that first one edge was deleted, then the others were contracted.
Using that deletions and contractions commute, and possibly relabeling, we
can thus assume that the first edge that disappears when H is formed from
G, say e, is deleted.
But then we get the same minor H from G', by contracting the corresponding edge from G', because the deletion of e from G and the contraction
of e in G' result in the same graph. Again we are finished by induction. 0
Denote the grid graph with mn vertices and m(n-1) n(m-1) edges by
G(m, n). Clearly the grid graphs G(m, n) and G(n, m) are isomorphic.
111
Lemma 4.5.
Proof. For this, fix an embedding of G into the plane IR2 ; now split the
vertices of G in order to get a graph G' of which G is a minor, such that
all vertices of G have degree at most 3.
k- k
11
8
10
10
5
112
Lemma 4.6.
to K4Proof. We use two basic observations. First, if an edge connects two neighbors of a vertex of degree 3, then we can delete it, by performing first a
A-to-Y transformation and then a series reduction.
Using these two observations, we can reduce any grid graph G (m, n) to K4,
as follows.
First perform a series reduction in the upper-right corner. Then, assuming that m > 4, we take the two edges in the lower-left corner and perforn
a series reduction to get a single edge. We can move this single edge across
the whole grid in a sequence of degree-4 moves until we "hit the boundary."
There either the edge is parallel to the diagonal corner edge and we can
parallel-delete, or it can be deleted according to our degree-3 move.
113
This way we have deleted the first square in the last row; similarly, we can
delete the second square in this row, and so on; the last square is deleted
by two series reductions and one parallel reduction.
11.
1111111.11
EMILE
We can delete all the squares in the first column of our board symmetrically,
if n > 4.
II
<I>.
.'
A-
At the end, this leaves us with the grid G(3, 3) with one short corner, which
is easily reduced first to the "wheel" graph W4 of a square pyramid, and
then to IC4 .
0
Corollary 4.7. Every 3-connected planar graph G can be reduced to K4
by a sequence of simple AY transformations.
Proof. The three lemmas together show that G is AY-reducible. We follow
this reduction to the first point where parallel or series edges are created.
These can be reduced immediately, which also results in a 3-connected
planar graph G by Lemma 4.2. The graph we now have has fewer edges:
hence we are done by induction on size.
4.4
114
polygons.
Proof. This we get by representing G as the graph of a 3-polytope, and
then choosing a point x beyond one of the facets. "Viewing" the polytope
from this perspective (and projecting it to the facet along the "rays of
vision") gives us the required representation. 0
Notes
115
of P is the set of all matrices Y G R d " such that yk = xk for 1 < k < d+1,
and such that P is combinatorially equivalent to Q := cony{ y i , yn }
under the correspondence x i 4 yi .
It is easy to see that the realization space is an elementary semialgebraic set defined over Z, that is, a subset of a real vector space that can
be defined in terms of polynomial equations and strict inequalities with
integer coefficients. Such semialgebraic sets can be arbitrarily complicated
as topological spaces, in general (see Exercise 4.22). The following assumes
that you know what contractible means for a topological space: if not, just
accept that it says that the space has "no holes"; in particular, contractible
spaces are connected.
Theorem 4.11 (Steinitz' theorem).
[490, 4871
For every 3-polytope P, the realization space TZ(P) is contractible, and
thus connected.
All of these theorems on prescribed facets, shadow boundary, symmetry, realization space, and so on were proved with basically the same
technique, and some clever variations. Like Steinitz' original theorem, they
are far from trivial. One way to see this is that they all fail "one dimension
up," for d = 4. We will construct explicit examples in the next two lectures.
Notes
The reason for this version of the definitions for graph connectivity is that
they fit into a larger and very natural pattern. Namely, following Tutte [513,
5141 (see Truemper [510, p. 15]) one defines a k-separation (for k > 1) of
a graph G = (V, E) as a partition into two graphs, G 1 = (V1 , E1 ) and
G2 = (1/2 7 E2), which have exactly k vertices but no edges in common, and
which have at least k edges each. Thus, our sketch
shows a 4-separation, if each side has at least 4 edges (and thus at least
one cycle, or a nonseparating vertex, or both). For any k > 2, a connected
graph is k-connected if it has no t-separation for 1 < 1 < k.
116
Then one proves that a straight line drawing of a 3-connected planar graph
can be lifted to 3-space (to obtain a 3-polytope) if and only if it is correct in
this sense. Of this basic theorem, Maxwell proved one direction, the other
(harder) one was provided by Crapo Sz Whiteley [157, 158] [525, Sect. 1.31;
see also Hoperoft & Kahn [260, Sect. 3], and in particular Richter-Gebert
[424, Sect. 12.2].
One basic observation is that a drawing of a graph is correct if and
only if it has a drawing of the dual (with a vertex "at infinity") such that
117
\\A
With proper care this also extends to higher dimensions, in the setting
of d-diagrams and Schlegel diagrams (see Lecture 5), as was shown by
McMullen [370 ] , completing an earlier version of Aurenhammer [23 ] _
Perhaps the nicest version of the approach via correct drawings uses the
circle packing theorem.
118
(In that case, a good way to view the representation is that one (e.g.,
dual) vertex is represented by the complement of a disk. Then the whole
plane/sphere will be covered by the disks of the representation.)
Theorem 4.13.
119
We don't try to do details here: after all, this uses completely "nonlinear"
methods, whereas everything we did is a combination of linear algebra and
combinatorics.
Still a different polytope realization algorithm, for simplicial 3-polytopes,
was given by Das Si Goodrich [170]: it essentially works by doing "many
inverse Steinitz operations on independent vertices in one step," and thus
yields a linear-time realization algorithm that produces realizations with a
singly-exponential bound on the size of the integer coordinates (as does the
Onn-Sturmfels algorithm [398] [424, Sect. 13.2] for general 3-polytopes).
Elementary semialgebraic sets are basic objects. Exercise 4.22 indicates
that they can have very complicated structure. General semialgebraic sets
(for which also nonstrict inequalities are admitted in the defining system)
can be written as finite unions of elementary semialgebraic sets.
In this chapter we only met realization spaces that are quite trivial (contractible); this will change as soon as we have developed the theory to
analyze the realization spaces of some high-dimensional polytopes, in Lecture 6. The study of semialgebraic sets is called real algebraic geometry
a very active and fascinating field of research. We refer to Bochnak, Coste
& Roy [103] and Becker [61] for more information.
4.2
4.3 What is the problem with the construction of P from P', if G ---4 G'
is a simple Y-to-A transformation? For this, try to prove the Y-to-A
part of Lemma 4.3.
If you are comfortable with projective transformations (Section 2.7),
explain how to do this without using polarity (as we did).
4.4 Check how to do the reduction theorem for grid graphs entirely within
the framework of 3-connected graphs. Where are the problems? How
do you make grid graphs 3-connected? How many "basic operations"
do you need?
120
ve+f=
for the number of vertices v, the number of edges e, and the number
of facets f of a 3-polytope.
4.7 Deduce from Steinitz' theorem (from Corollary 4.9) that every planar
graph has a straight line drawing in the plane without intersections.
(In graph theory literature, this appears as a result of Wagner [517],
rediscovered by Fry [193]. It is also quite easy to prove this directly
by induction on the number of vertices; see, for example, Hartsfield
Sz Ringel [251, p. 167].)
4.8 Conclude from Exercise 4.6 that for every 3-polytope P, either the
polytope P itself or its polar P has a facet that is a simplex.
(This is not true for 4-polytopes for these read about the regular
24 cell, whose facets are octahedra and whose vertex figures are cubes;
see, for example, Coxeter [156, Sect. 8.2].)
-
4.9 The graph of the 3-cube, G(C3 ), has cycles that go through all the
vertices ("Hamiltonian cycles").
121
4.10 Show that we cannot prescribe the shape of two facets of a 3-polytope
(even if they have congruent intersection). In fact, in a triangular
prism
if one square face is prescribed such that the edges a and b are parallel,
then the other square faces have to have b and c respectively a and c
parallel.
4.11 One cannot prescribe two disjoint facets of a 3-polytope either. For
this, analyze the prism over an n-gon, and show the following. If
we prescribe the bottom n-gon, then the coordinatizations of the
completed prism are prescribed by n +3 linear parameters ("degrees
of freedom"), while we have 2n 3 choices for the shape of the top
facet. Thus, if we prescribe two generic n-gons, then they cannot be
built into a prism, if n > 7.
122
4.12 Show that we cannot prescribe the shape of the shadow boundary of
a 3-polytope.
(Jurgen Richter-Gebert, who noted that from a triangular prism, you
can get a hexagon as a projection, but the shape of the hexagon cannot be prescribed: see page 141! Earlier, Barnette [42 ] gave a proof for
the case where the 3-polytope is a tetrahedron with a stellar subdivision on every facet, and Q is a regular 8-gon. Also, you can use the
prisms of the previous exercise, and again count degrees of freedom.)
4.13 Show that not all 3-polytopes can be represented in such a way that
all facets touch the unit sphere, or that all vertices are on the unit
sphere. Which 3-polytopes can be represented this way?
(Remark: This is a classical problem that goes back to 1832; see
Steiner [486], Steinitz [488 ] , and Schulte [450 ] . The characterization
problem was solved very recently by Hodgson, Rivin & Smith [255].)
the figure).
o
(i) Show that f2(n) < c n3/2 for some c> O.
(In fact, Thiele [502, Satz 4.1.10] proved that the bound
f2(n 1).
123
(iii) Using part (i), show that the polar cyclic 3-polytope C3 (m) A
with n vertices (where n = 2m 4) can be represented on a
le x k' x le grid, where k' = f2( 71- + 1).
4.16* For n > 4 let f3 (n) be the smallest positive integer such that every
3-dimensional polytope with n vertices can be represented with integral vertices in {0, 1, ... , f3(n)} 3 . Let ,a(n) be the same function for
simplicia,1 3-polytopes.
Determine the asymptotic behavior of the functions f3(n) and Mn).
(Work by Goodman, Pollack & Sturmfels [220, Sect. 5 ] shows that
for d-dimensional simplicial polytopes with d 4 vertices one needs
vertex coordinates that grow doubly exponentially in d,
L9, (d + 4) > 22"
for some constant c> O. However, there is no such result that applies
to simplicial polytopes for any constant dimension d.
The case d =... 3 might be special: it is quite possible that there is a
quadratic upper bound on f3 (n); up to now only exponential upper
bounds f (n) < 4 3 and f3 (n) < 2 13122 are known, due to Onn &
Sturmfels [398] and Richter-Gebert [424, p. 143].)
4.17* For n > 4, let g(n) E N be the smallest number such that every 3
connected planar graph can be represented in the plane in such a way
that the vertices are on the g(n) x g(n) grid, the edges are straight,
and the bounded regions determined by the graph embedding, as well
as the union of all these regions, are strictly convex (i.e., all interior
angles are smaller than 7r).
-
124
Can you do this for the 3-polytope obtained by cutting off the vertices
of a tetrahedron?
That is, can you represent this polytope in such a way, with points on
its facets, that adjacent facets of P correspond to adjacent vertices
on the convex hull P of the points?
(For the special polytope P above this can be done, according to
Jrgen Richter-Gebert. The general problem is suggested by the misleading description of polar polytopes in the book by Bartels [51,
p. 74]. Grnbaum & Shephard posed the question as Problem 3
in [241]. Grnbaum sent me the following e-mail message:
As far as I know, the problem is still open. I am inclined
to believe that the answer is negative, and that once some
counterexamples are found, we will all be saying how obvious that is.
I agree.)
4.20 Which subsets of 118 are elementary semialgebraic?
4.21
Show that for every convex d-polytope with n-vertices, the realization
space is an elementary semialgebraic set in Rd ".
(For this, embed the polytope into 1" d+1 , consider the maximal determinants of a realization matrix, and note that the condition that d+1
points have to be on a common facet says that a certain subdeterminant is 0. Similarly, two points being on the same side of the facet
spanned by some basis means that the product of two determinants
is positive.)
4.22 Let S C
125
4.23 The following problems discuss different ways in which a planar graph
can be represented by objects in the plane.
(i)
Prove that the vertices of a planar bipartite graph can be represented by horizontal and vertical closed line segments in the
plane, such that the segments intersect if and only if the corresponding vertices are adjacent.
(Hartman, Newman & Ziv [250 1 )
(ii)
In fact, every planar bipartite graph can be represented by disjoint horizontal and vertical open line segments in the plane,
which touch if and only if the corresponding vertices are adjacent.
(de Fraysseix, Ossona de Mendez & Pach [399, Sect. 6.3] [195])
(iii)* Can every planar graph be represented by a family of line segments in the plane such that every vertex corresponds to a segment and adjacent vertices correspond to intersecting edges?
4.24 Describe a "fast" algorithm to test whether two 3-polytopes are cornbinatorially equivalent.
For this, assume that the combinatorial structure is given by the
vertex-facet incidence matrices of the polytopes; from these, construct
the graphs. Then use that graph isomorphism can be tested "in linear
time" for planar graphs, according to Hoperoft & Wong [259].
(As a theoretical concept of fast algorithms one has the theory of polynomial algorithms and NP-completeness; see Garey & Johnson [208].
126
4. Steinitz'
The question of whether a polynomial algorithm exists for isomorphism of general graphs is one of the prominent open problems in
this theory [208, pp. 155-1581.)
5
Schlegel Diagrams for 4-Polytopes
5.1
Polyhedral Complexes
Definition 5.1.
in Rd such that
128
Our drawing shows the poset L(Co) for the complex Co above. Note that
L(C) does not have a unique maximal element, unless C is the complex
of all faces of one single convex polytope. Thus, L(C) is not a lattice in
general (although it is a finite meet-semilattice: it has a minimal element,
and meets exist). If we adjoin an artificial maximal element i, then we get
a lattice L(C) := L(C) U 01.
The maps f : C ----* D between polyhedral complexes are all the maps
f: ICI -- IDI that are affine maps when restricted to poiytopes in C. Two
complexes are affinely isomorphic if there is such a map f : C p T, which
is a bijection between C and D = {f(F) : F e C} . Equivalently, f has to
129
(i) The complex C(P) of the polytope P is the complex of all faces of P.
The face poset of C(P) is the face lattice L(P).
(ii) The boundary complex c ( p) is the subcomplex of C(P) formed by
all proper faces of P. Thus its underlying set is IC(8P) I = ap =
P\relint(P). Its face poset is " (8 p) := L(P)\{P}.
(ill) A (polytopal) subdivision of a polytope P is a polytopal complex C
with the underlying space ICI = P. The subdivision is a triangulation
if all the polytopes in C are simplices. In particular, one is interested
in subdivisions and triangulations without new vertices, that is, where
the only zero-dimensional polytopes in the complex are the vertices
of P.
Our drawing shows (from left to right, everybody smile, please!) a poly
subdivision of a hexagon, a subdivision without new vertices, a tri--topai
angulation, and a triangulation without new vertices.
130
(ii) C is the set of all lower faces of P, projected down to Q, that is,
Q
It suggests a reformulation: regular subdivisions arise from piecewise linear
convex functions in the following way. Given the projection 7r : P -----> Q,
the function
f:Q *R,
f (x) = minty E R: (1 E P}
Y
P := cony{
x
f (x)
: x E Q}.
131
Example 5.4. For integers z 1 ,..., zd > 1, the pile of cubes Pd(zi,- - , zd)
is the polytopal complex formed by all unit cubes with integer vertices in
the d-box
:= { s E Rd :
that is, the polyhedral complex formed by the set of all cubes
--Axe
C(ki,...,kd)
for integers 0 <k < z, together with all their faces. So, our drawing
represents the piles of cubes P2 (6,4) and P3 (6,4,3).
P2(6. 4)
P3(6, 4, 3
In particular, the pile of cubes Pd(zi, .. , zd) has (zi-1-1)- .. -(zd-F1) vertices, which are given by
vert(Pd (zi ,
, zd)) = B(zi,
, zd) n zd.
:Rd + R,
f (x)
fi (x1)
+ - - + fd(xd)
:=
cony
: V E B(Zi,
Zd)
fl Zd}.
132
5.2
Schlegel Diagrams
H := aff(F) = ix G Rd : ax = z}
133
Y,(
The Schlegel diagram of P based at the facet F, denoted as D(P, F), is the
image under p of all proper faces of P other than F; that is, it is the set
system
Y,
is iso-
134
where the center vertex in the diagram represents the (unique) vertex of
the tetrahedron which is not on the facet we project to. Similarly, here are
135
And finally, here is an attempt to illustrate how you construct the Schlegel
diagram of a lifted pile of boxes, based on the big square facet on its top.
The drawing is not metrically correct in any sense just an attempt
to understand the geometric situation. Specifically, we go for the Schlegel
136
diagram of j53 (6, 4), from a point of view "above the polytope":
Nokui roil
fall11111114
#0111111 111111114k
where the center vertex in the diagram represents the (unique) vertex of
the 4-simplex which is not on the tetrahedron we project to; here is the
drawing for the 4-cube,
137
P = A2 X A2/
Although this sketch is quite sketchy, the picture might try to tell you about
the combinatorial structure of (the Schlegel diagram of) the 4-polytopes
P4(Z1/ Z2 1 Z3). In fact, from the Schlegel diagram (compare also to the 3dimensional case) we see that P4 (zi , z2 , z3) has exactly z1 z2 z3 + 7 facets.
The seven "big ones" are the big cubical facet on which the diagram is
based and the six 3-dimensional facets that are lifted piles of cubes: two
copies of P2 (Z1 1 Z2) at the bottom and the top of the diagram, two copies
of 'P2(zi, z3) at the front and the back of the diagram, and two copies
of P2 (Z2 ) z3 ) at the left and right sides of the Schlegel diagram. Between
these six big facets, we have the subcomplex formed by z1z2z3 little cubes,
which is isomorphic to the pile 'P3(zi, z2, z3) from which we started the
construction.
Make sure that you "see" this: it will be useful later (in Section 8.2)
when we study the lifted piles of cubes again.
138
5. Schlegel Diagrams
for
4-Polytopes
5.3 d-Diagrams
Schlegel diagrams were studied extensively around the turn of this century;
however, no one realized that there is a problem: not everything that "looks
like" a Schlegel diagram actually is one. (See the notes below.) Now we
define looks like.
Theorem 5.8.
Every 2-diagram is combinatorially equivalent to a Schlegel diagram.
139
5.4
Three Examples
140
Now we choose extra points: 7 in general position inside Q, but outside the
tetrahedron [2,3, 5,6], and 8 above the top of the original pyramid, so that
all facets of Q, except for the base [4, 5, 6], can be "seen" from 8.
The diagram D I , based on the tetrahedron G = [4, 5, 6, 8], consists of the
following ten 3-polytopes in R3 , and their faces:
A:
B, C:
D, E:
5 4 Three Examples
141
Our next example, D2 is a Schlegel diagram for which the total space
F = 17,2 1 cannot be prescribed. This shows that for a 4-polytope, we cannot
prescribe the shape of a facet in contrast to the situation for 3-polytopes,
as discussed in Section 4.4.
Example 5.11 (Barnette's first diagram). {43]
For this, we consider the Schlegel diagram for the prism over the square
pyramid I x Pyr3. This Schlegel diagram can be constructed from a regular
cube, and two vertices on the vertical axis of symmetry.
E4
From symmetric arguments for Ez , Ei+1 , and E, we see that for every
3-diagram that is combinatorially equivalent to the given one, the four
lines E1 , E2 1 E3 1 E4 are parallel, or else they intersect in a common point.
Thus if we start with a "skew" combinatorial cube that does not satisfy
this (such a cube is easy to get), then we cannot even complete it to a
142
The curved quadrangle partitions the interior of T3 into two "3-cell regions." Into each of them we place a new vertex and then perform a "stellar
subdivision": each vertex is joined to all the faces on the boundary of the
respective 3-cell, so each 3-cell is replaced by the (topological) pyramids
over four triangles and two quadrangles each, and their faces.
But we find that this "topological 3-diagram" is not realizable. In fact,
any two realizations of the tetrahedron are equivalent. Now we try to realize
the quadrangles F2 and F4 in Q2 by planar convex quadrangles. Then the
plane H2
aff(F2) contains E2 and a (unique) point of E4 Similarly,
.
Notes
143
the plane H4 := aff(F4) then contains E4 and a point of E2. Thus the
intersection 112 n H4 connects a point on E2 with a point on E4. This
means that the four points of F2 are not in convex position in T3.
Il
Notes
Schlegel diagrams are the most direct, and probably the most effective, tool
to visualize 4-dimensional objects. Of course, there is a certain amount of
training necessary to develop the geometric intuition. Don't let yourself get
discouraged, not even by statements like the following:
Here, however, a word of warning may be in order: do not try to
visualize n-dimensional objects for n > 4. Such an effort is not
only doomed to failureit may be dangerous to your mental
health. (If you do succeed, then you are in trouble.) To speak
of n-dimensional geometry with n > 4 simply means to speak
(Chvital [151, p. 252])
of a certain part of algebra.
This is wrong, and even Chvital acknowledges the fact that the correspondence between intuitive geometric terms and algebraic machinery can be
used in both ways [151, p. 250].
The technique of Schlegel diagrams was already used extensively in work
of Bruckner [130] early this century, where the distinction between Schlegel
diagrams and 3-diagrams was not made.
There are also simplicial examples known of 3-diagrams that are not
Schlegel diagrams, and not even combinatorially equivalent to such. The
first one was described by Griinbaum's abstract [233], which started the
subject. The first non-Schlegel 3-diagram with 8 vertices was found by
Griinbaum & Sreedharan [242], showing that one of Briickner's 3-diagrams
does not, as Bruckner thought, represent the combinatorial type of a 4polytope. It is now known as the Briickner sphere [234, p. 222]. A second
example of a simplicial 3-diagram that is not equivalent to a Schlegel diagram the Barnette sphere
was found by Barnette [35] a little later.
There are also simplicial 3-spheres that can be represented by topological
diagrams (as in our Example 5.12), but not by straight 3-diagrams. Both
kinds of examples are nicely presented in Ewald [189, Sect. IV.4 and IV.5].
However, there is something special happening in the case of simple diagrams. In fact, every simple d-diagram with d > 3 is the Schlegel diagram
of a (d + 1)-polytope. Thus, there are also things true in 3-space that are
false in 2 dimensions. This was proved by Whiteley [526], and in an even
stronger version by Rybnikov [435]. (They use a quite general setting for
"Mobility"; see Crapo & Whiteley [157, 158, 159].) Earlier results of Davis
[172] and Aurenhammer [23] did not include the "boundary" conditions
that pose extra constraints (as in Problem 5.7).
144
145
Show that the following 3-polytope (the capped prism) has a nonregular triangulation without new vertices.
(Kleinschmidt and Lee, see Lee [329, Sect. 6])
x,
5.3
(ii)
146
5.4 Define the product of two polyhedral complexes, in such a way that
the product of subdivisions of two polytopes P and Q is a subdivision
of the product P x Q.
Prove that the product subdivision is regular if and only if the original
subdivisions of P and Q were regular.
5.5 Compute the numbers fk of k-faces for the polytope P4(zi , z2, z3).
5.6 A "default" convex function is given by the paraboloid function
f:
Let V C d be a finite set of points (vertices). Show that the regular
subdivision of Q := conv(V) associated with this f has the following
property: for every facet of the subdivision there is a sphere that
contains all its vertices, but no other vertices from V.
(This subdivision is known as the Delaunay triangulation of the point
set V. It is of great importance for many aspects of computational
geometry; see for example Edelsbrunner [184)
5.7 Show that the following 2-diagram is regular, but not Schlegel:
5.8 Show that in general the polytopal complex C(aP)\IFI and the
Schlegel diagram D(P, F) are not affinely isomorphic. (For this, it
suffices to consider the 3-cube!)
However, show that p: G --) p(G) is a projective transformation,
for all proper faces G of P.
Problems
and Exercises
147
P , (A2 x A2) A Construct Schlegel diagrams for the cyclic polytopes C3 ( 6 ) and C4 ( 7 ) .
f (3) ---, 5
1(7) = 1493
f(4) =-- 16
f(5) ---, 67
5.11 For a 4-polytope, one cannot prescribe the shape of a 2-face! Namely,
Richter-Gebert [424, p. 91] [426] provides the following diagram of a
4-dimensional polytope X*:
148
5.12 Show that 2d simplices can be arranged in r d in such a way that any
two are adjacent (that is, the intersections are (d 1)-dimensional).
5.13 The smallest triangulation of the torus surface has 7 vertices, 21 edges
and 14 triangles. Construct it, and show that it can be embedded as a
subcomplex into C(C7(4)), and thus as a simplicial complex into r 3
p. 253]; see also Altshuler [13] and(CsAzdr[16],Ginbaum234
Bokowski Sz Eggert [106])
(This is due to Betke, Schulz & Wills [64]; see Barnette [44] for a
similar, but simplicial, Mbius strip that serves as an "impediment
for polyhedrality." It is not even true that every triangulated Mbius
strip has a straight embedding into 3 : see Brehm [122]. However,
the Schlegel diagram construction shows that no such Mbius strip
can appear in the boundary of a 4-polytope.)
6
Duality, Gale Diagrams,
and Applications
More about life in high dimensions: after "successfully" dealing with polytopes in four dimensions, we now study polytopes with few vertices, that is,
d-polytopes with only d-plus-a-few vertices. For this, we develop a duality
theory, which describes them in terms of structures in low dimensions.
This duality theory (developed by Perles in the 1960s, and recorded
by Griinbaum [234]) is classically known as Gale diagrams. Later it was
realized (apparently first by McMullen [366]) that Gale diagrams are a
manifestation of "oriented matroid duality."
Thus behind the construction of Gale diagrams one finds (barely hidden)
oriented matroids. Their theory was initiated in the 1970s by at least four
independent authors, Jon Folkman, Jim Lawrence, Robert G. Bland, and
Michel Las Vergnas; see [194] and [96].
By now oriented matroids form a theory with many facets, extensive
enough to fill thick books [92]. One aim of our lectures is to give a simple introduction to a few topics that help understand polytopes. Keys to
this include the description of the relative position of vertices in terms of
"circuits" and "cocircuits," as well as the simple duality between these two
descriptions, oriented mat roid duality, which in polytope theory manifests
itself in (linear and affine) Gale diagrams.
Thus this lecture includes a brief crash (?) course on oriented matroids.
(This will be continued in Lecture 7, when we discuss hyperplane arrangements and zonotopes.) Although oriented matroids need some amount of
new notation and terminology, there is little magic involved: just don't be
scared of names. For further reading, we refer to the "Orientation Session"
in Bjiirner et al. [92, Ch. 11.
150
6.1
E
ieP(Z)
_zixi =
A
E _zi
iEN(Z)
y,
151
so that
y
conv{x i : i E P(z)}
conv{z i : i C N(z)}
represents a point that lies both in the convex hull of the points with
positive coefficients and in the convex hull of the points with negative
coefficients.
( 0 3 5 5 2 0\
For example, let X =
be the vertex set of a
0
01221
hexagon
4
6
z ,-----4
1
\ 0 )
represents y =
4
1
n convfx 2 , x41.
152
change (Exercise 6.1). Together these three facts mean that the minimal
affine dependences completely determine the structure of the point configuration.
For the hexagon we considered before, the linear dependence can be
written as a sum
/ 1\
( 1\
l o\
5
_3
4
2
6
3
3
=
+
__
4
2
1
1
0
o)
\ o)
o)
\
which writes it as a sum of the two minimal dependences depicted here:
4
/
6
1
, Xz =- 0, 11 z = 0} = SIGN(a-Dep(X)),
\ 0 j
[ + \
,
and
0
0 i
153
(f
(xi),
,f
(zn)) = (ex i z,
, cx r, z) = cX zll ,
where our drawing lists the values of f in brackets, and illustrates the
hyperplane Hf := {x c Rd : f (x) = 0} as a dashed line.
We note that the set
a-Val(X) := {cX zl G(
:cG ORd r , z E R}
Mar.
154
(+, +, 0,0, +, +)
and
155
1 1 0 0 0 0 )
Xi = ( 0 0 1
1 0 0
0 0 0 0 1-1 /
= (1 1 0 0 0 0 )
X2
bl 0 1 1 0 0
0 0 0 0 1 -1
We are not patient enough to list all the vectors and covectors. Here,
however, are complete lists of all the cocircuits: the reader is not expected to
check their details, but just to convince him/herself how this is constructed
"from the picture," and that it "seems correct." Does it?
(0
(0
(0
(0
+(0
C*(X1): (0
(+
(+
(+
(+
+ (+
0
0
+
+
+
+
0
0
0
0
0 0 + ---.),
+
00),
0 + 0 +),
0 + + 0),
+ 0 0 +),
+ 0 + 0),
0 + 0 +),
0 + + 0),
+ 0 0 +),
+ 0 + 0),
0 0 0 0),
( 0 0 0 0 + ),
(+ 0 + 0 0),
+ ( 0 + + 0 0),
+(0 0 + 0),
+ ( 0 0 + 0 ),
+(0 + 0 + 0 + ),
C*(X2):
(0
(0
(0
(+
(+
(+
(+
(+
+ 0
+
+ +
0 0
0 0
0 +
0+
0
+ + 0 ),
0 0 + ), +
0 + 0),
+ 0 +),
+ 0),
0 0 + ),
0 + 0
0 0 0),
),
156
Similarly, the circuits are the columns of the matrices we describe next,
and their negatives:
+ + 0 \
++0
0 +
0+
0
0
/
6.2
C X2 )
(
+ + + 0 \
+++0
-0+
0++
0
\ 0
Vector Configurations
(x12 ) in Rd+'. To
vertices of a d-polytope), we associate the vectors vi
get our notation for dimensions straight, we introduce a new parameter,
called rank, as r := d + 1. As so often when dealing with a transition from
affine to linear, it is convenient to have an extra letter r for affine rank (i.e.,
linear dimension), which is one more than the affine dimension d. Thus we
have vectors v i E W.
In fact, what we get this way is an acyclic vector configuration V =
{v 1 ,. , v.,2 } C rr, characterized by the following two properties (which
are equivalent by a simple application of Farkas lemma II):
(i) There is no nonnegative dependence, i.e., no y > 0, y
Vy = 0.
0, such that
(ii) There is a linear function c E (W)* such that cV > O (i.e., cv, > 0
for all i).
157
Dep(V) := Iv E Rn : Vv = 0}
c W1.
V(V)
{sign(v) E {, , Or
:V E
W1, Vv = 0} = SIGN(Dep(V)),
and the signed circuits are the signed vectors of minimal nonempty support.
Dually, the space of value vectors on a vector configuration, which correspond to linear functions c E (W)*, is constructed as
).
6.3
Oriented Matroids
In this section, we meet oriented matroids for the first time. Just to keep
things more exciting (and to make sure that the timid reader isn't overwhelmed by the first sight), we won't lift all the veils on the first encounter.
So we won't even define oriented matroids this time, but we promise more
for later.
158
{ ()'(--)}'
vectors: V(V) =
f() , -
o0 } '
(--) , ( 0)
cocircuits: C*(V) =
{(0++),
(0---),
(- -0)}
covetrs:
V*(V) =
159
(a) Axiomaties
The collections of sign vectors that make up an oriented matroid are highly
structured, and not just random collections. In fact, they arise in the following way.
For any linear subspace U C R , we define the operator SIGN as
SIGN(U) := {sign(x) : x G U} c
Or.
There is a natural partial order on the set of signs {-1-, 0}: we set 0 <
and 0 < , while + and are incomparable:
0
This corresponds to the fact that a number that is slightly perturbed either
keeps its sign, unless it is zero, in which case its sign can change to -1-, or
to , or remain O.
On sets of sign vectors S C { +, ,0}n, we use componentwise partial
ordering: u < u' if and only if ui < u't holds for all positions i. Thus we
get, for example,
but
(0+ 0 +000--+)
because of the second position, where +
O.
We use the componentwise partial order to define the operator MIN,
which takes all the minimal nonzero sign vectors in S:
MIN(S) :=
We will apply the operator SIGN equally to sets of row vectors and of
column vectors. Similarly, we apply MIN both to sets of row sign vectors
and of column sign vectors.
With these conventions, we can describe the oriented matroid data for V
as follows:
160
were studying in Sections 6.1 and 6.2 were the vectors and circuits of two
realizable oriented matroids of rank r respectively ri r,
M = M(Val(X))
M* = M(Dep(X)).
and
It is easy to write down extensive lists of axioms that are satisfied by any
collection SIGN(U). So one gets to axiom systems for oriented matroids.
We'll get to that in Lecture 7, but without much detail. In fact, the proofs
relating various axiom systems for oriented matroids tend to involve hard
work, something we try to avoid on this show. (This is, however, at the
basis of oriented matroid theory; we refer to [92, Ch. 3]).
The oriented matroid of a point configuration is a delicate model for its
geometry. It provides a much finer model than what the matroid encodes
about a vector configuration. (If you want to know what a matroid is, see
Welsh [518], White [520], or Oxley [400].) One can prove that in fact the approximation of the combinatorial model to "geometric reality" is extremely
good this is made precise in Lawrence's "topological representation theorem." We'll get back to this in Lecture 7.
This means also that all the main features of the geometry of vector
configurations can be derived from formal properties (the axioms). In fact
there are very few geometric statements that would be true for vector
configurations but fail for oriented matroids so whatever we find in that
direction is even more exciting. (See Theorem 7.20 for an example.)
(13) Equivalence
Different sets of data "A" and "B" about a geometric situation are equivalent if any two configurations with the same data A also have the same
data B, and conversely. This means that (at least in principle) one can
construct the data A from the data B. Any set of data that is equivalent
to the set of circuits is referred to as the oriented matroid of V.
We now want to show that the four sets of data for an oriented matroid given by Definition 6.5 are equivalent. For that we need to define the
combinatorial analogue of the condition cx = 0.
Definition 6.6. Let x G {,,O} n and C E ({,
vectors. Then we define that "c-x = 0" if
for each i, we have
= 0 or x, = 0,
:= fc E ({+,
}
n
0 and ci = x i 0.
Or, we define
6 3 Oriented Matroids
161
In fact, the condition c-x = 0 holds if and only if there are real vectors
E Rn and c E (Rn)* such that sign(x) = x, sign(c) = c, and cx =0. For
example, we have
+)
(+00) ( = 0,
but
(+00+) )
0.
The reader might want to check the next two statements for the small
3-vector configuration after Definition 6.5. They are both solid theorems
in the setting of oriented matroids [92, Ch. 3]. For the realizable case, we
won't have a lot of problems with them.
Lemma 6.7. Let U C n be a vector subspace of dimension r, and let
U E U be a vector with sign(u) = u E SIGN(U) C {+, , O}.
The vector u can be written as a finite sum u = u1 + .. . + uk of k < r
vectors Ui G U whose sign vectors ui := sign(ui) are below u and minimal,
that is, such that ui < u, and Ui G MINSIGN(U)).
Proof. The vectors y E U whose sign vector sign(v) is componentwise
smaller than or equal to sign(u) form a polyhedral cone:
C(u) :=
U.
This cone is in fact pointed: all vectors X G C(u) satisfy Ein_ i uixi > 0,
with equality only for z = O.
Thus P(u) :=
E C(u) : Etn_ i uixi = 1} is a polytope of dimension
at most r 1. By the results in Lecture 1 (with Carathodory's Theorem
1.15) every point in P(u) can be written as a convex combination of at
most r vertices. By linearizing we get that u is the sum of k < r vectors on
extreme rays (1-faces) of C(u). Finally, we observe that the sign vectors on
the proper faces of C(u) are strictly smaller than u, and thus the minimal
El
nonzero sign vectors are precisely found on the extreme rays.
Proposition 6.8.
(MIN(SIGN(U))) -L
(SIGN(U))
SIGN(U ).
For the converse, let c c I+, , Oln \SIGN(U -L ). Then the conditions
cE U-1- ,
sign(c) = c
162
chik = 0
di > 0
di < 0
di = 0
for
for
for
for
1< k<r
iE P
icN
iE Z
dv k = 0
di > +1
< 1
= 0
for
for
for
for
1< k<r
iE P
N
i E Z.
Now we apply the Farkas lemma (Proposition 1.7 adapted for systems with
inequalities and equations, see Exercise 1.6) to get existence of
eix`i` = 0 and
for i E N
O for i C P <
EiEp Xj
EiEN <O.
c (MIN(SIGN(U))) .
El
Corollary 6.9. For any vector configuration V E Rrxn , the four sets of
data given by Definition 6.5 determine each other (denoted by it> ,7 j,1 as
follows:
Dep(V)
Val(V)
SIGN
SIGN
vectors V (V)
covectors V* (V)
MIN
MIN
C(V) circuits
C* (V) cocircuits
Thus any of the four sets of data determines the other three, and thus also
the oriented matroid M(V).
163
(c) Duality
A concept of duality is built into the whole structure of oriented matroids.
In fact, since the vectors and the covectors of an oriented matroid M(V)
arise in the same way as the sign vectors of a subspace, they also satisfy
the same axioms (ignoring a switch from row vectors to column vectors).
Definition 6.10. The dual of an oriented matroid M is the oriented
matroid M* with the following properties:
The vectors of M* are the covectors of M, and thus
the circuits of M* are the cocircuits of M.
The covectors of M* are the vectors of M, and thus
the cocircuits of M* are the circuits of M.
M* = M(Dep(V))
for
M := M(Val(V)).
164
V. We can
Consider a vector configuration V E r ", and let u
certainly delete ui from V, to get the new configuration V\ui .
-
Proposition 6.11.
follows:
V(V\ui) = {Y\i : y E V(V), = 0} V* (V\ui) ={ \TV : y E 1,* (V)}
C(V\ui) = fc\i : c E C(V), ci = 01 C*(V\ui ) = MIN{cV : C E C*(V)}.
for this we project V parallel
The dual operation is the contraction of
to ui to some hyperplane that does not contain ui . If u = 0, then we just
delete ui .
:=
CIL 3
cui
tLi
:=
, fti-vi, -
165
Proposition 6.12.
follows:
Let us mention two examples that show how deletion and contraction
appear in connection with polytopes.
Examples 6.13. Let P C Rd be a polytope, let X := vert(P) be its
vertex set, and let V E R"n be the corresponding vector configuration
in
(r = d+1).
If F is a face of P, then the vector configuration for F is obtained by
deleting from V all the vertices that do not lie on F.
If x E vert(P) C Rd is a vertex of P and vEVC Rr is the corresponding
vector, then the vector configuration of the vertex figure Plx is the contraction V/v. (In this case the projection hyperplane for the contraction
can be taken parallel to the hyperplane spanned by P.)
Note that by contracting we get a vector configuration that may also represent a lot of interior points of the vertex figure, corresponding to vertices
w E vert(P) such that [w, y] is not an edge of P.
6.4
166
that
Val(V) = lc C ( r ) : cG = 01
and
Dep(V) = {Gx : x E Ile -r }.
rank(G) = n r
and
VG = 0,
Dep(V) = Val(G)
and
M(G) = (M(V))*.
167
0
Comparison between these descriptions of "totally cyclic" and the corresponding ones that we have given for the dual concept "acyclic" on page 156
might give a feel for how the translation between dual concepts works on
the linear algebra level.
On the combinatorial side, we derive the following.
Corollary 6.16. A vector configuration V is acyclic if and only if the
following equivalent conditions hold:
+
(iii) Every i is contained in a nonnegative circuit.
In particular, a vector configuration V is acyclic if and only if its dual
configuration G is totally cyclic, and conversely.
Now we'll put the pieces together.
168
n vectors in
(polytope)
d+1 , vi := (xt )
n vectors g i
nr
-d-1
Gale diagram
Here the passage from Rd to Rd+1 is the usual embedding, used to linearize the situation. The dual configuration of this vector configuration is
a Gale diagram for P, determined uniquely up to a change of coordinates.
For the reduction of (Rn()* to (i n-d-2 ) * , we find a suitable vector
y E Rn-d-1 such that g i y 0 unless gi = 0, for all i. Then we associate
with gi the point
ai :=
92
fc c (Rn-d-1)*
cy 11
9iY
which we call a positive point in the affine space 11 -d-2 if gi y > 0, and a
negative point if gi y < O.
This yields the affine Gale diagram, a labeled point configuration
{a i ,... an } C Rn-d-2
where the point ai is labeled i if it is a positive point, labeled if it represents a negative point, and not represented by a point (or, represented by
a "special" point) in the case where g i = O.
-d-2 ) * does not lose combinatorial information: the
The reduction to (
circuits and cocircuits of this affine point configuration still represent the
cocircuits and circuits of P. This is extremely useful for polytopes with
"few vertices," where n d is small, as we will see in the following section.
Let us consider one example, to illustrate the technique.
169
Example 6.18. For the octahedra of Example 6.3, we have the matrices
vi =
1 1 1 1 1 1
1 1 0 0 0 0
0 0 1-1 0 0
00001-1)
1-1
I6 0
0 0
v2=
1 1 1 1
0 0 0 0
1-1 0 0
00 1-1
G2=
GI =
1 \
1
0
i
_ i_
1
_ 13
12
From this we can directly draw Gale diagrams (they are 2-dimensional),
and derive 1-dimensional affine Gale diagrams, for y = (1). Here they are,
linear and affine.
We use the convention for affine Gale diagrams that black dots denote
positive points, while white dots denote negative points.
It is really important that the reader figure out how to read off the circuits
and the cocircuits of the octahedra from their affine Gale diagrams: he or
170
\ - /
which are cocircuits for both Gale diagrams, and thus circuits of the octahedra. Similarly, from minimal affine dependences we read off sign vectors
like
(+ 00 + + 0),
which form circuits for the Gale diagram, and thus cocircuits for the octahedra. The only new feature is that the sign is reversed for any negative
point in the diagram.
Every spanning set G = fg 1 , . . . , gii } of n row vectors in (I
r can
be interpreted as the Gale diagram of a vector configuration of n vectors that span Rd . However, these vectors need not come from a (d - 1)polytope: the vector configuration might not be "affine" (acyclic), and even
if it is, the vectors need not be in convex position. However, there is a
simple combinatorial condition that characterizes Gale diagrams, see the
following theorem. It is important because it allows us to conclude the
existence of a (high-dimensional) polytope with specific properties from a
(low-dimensional) configuration of signed points.
) *
171
It yields a criterion that is very easy to check "by inspection" for the
interesting case n d = 4; see below.
Corollary 6.20 (Characterization of Gale diagrams of polytopes).
A configuration A = {a 1 ,.
an } points in (Rn-d-2 ) * , each of them declared to be either "positive" or "negative," that affinely spans ( n-d-2)* 1
is the Gale diagram of a d-polytope with n vertices if and only if the fol.
e
represents the bipyramid over a triangle (this is a simplicial polytope with
6 facets), and
172
173
\%11;W
irapi
69
mmimp-
10
174
V=
/1 1 1 1 1 1
1\
1 0-1 0 0 0
0
0 1 0 1 0 0 0
0
0 0 0 0 1 1
\ 0 0 0 0 0 1 1 /
/0 0 2 2
1 0-1 0
0 1 0 1
0 0 2 2
\O 0 2 2
0
0
0
1
0
0
0
0
0
1
1 \
0
0
1
0
0 /
175
1 1 1 1
1 2 2 2 0 0 0
2 0 0 6 0 0 6
2 0 6 0 0 6 0
1
1
0
3
1 )
3
3
3
0
176
Now we check the following facts, which together imply all we need to know.
1. This signed point configuration is the affine Gaie diagram of a 5
poiytope Q with 10 vertices.
For this we check that every cocircuit of this configuration has at least
two negative and two positive elements.
2. The triple (8, 9, 10) describes a triangle F --= [8, 9, 10] which is a 2 face
of the 5-polytope Q.
The point 1 is in the interior of the prism, so the points different from
8,9, 10 support a positive circuit (+++++++000).
3. The face figure Q / F is a hexagon whose diagonals cross.
In fact, we get the affine Gale diagram of Q / F by deleting the points
8, 9, 10 from the diagram for Q. But what is left then is the affine Gale
diagram of a hexagon [2, 6, 4, 5, 3, 7] with 1 as the intersection point of the
long diagonals:
7
286
23157
385
23167
2107
24156
4105
24167
397
34156
496
34157
5. Every Gale diagram G' with the same positive circuits contains the
diagram of the hexagon with crossing diagonals.
Consider any other (linear) Gale diagram G' on 10 points with the same
positive circuits. From the 3-point circuits we see that the sets 2'4'5'7'10',
2'3'5'6'8', and 3'4'6'7'9' have to be planar. However, they cannot collapse to
be on a line, because then the whole diagram would collapse to a plane and
couldn't have 5-point circuits. From this one can show (using projective
uniqueness of the triangular prism here we are skipping the detailed
arguments) that in suitable coordinates G1\1' coincides with G\1: the Gale
diagram G' consists of a triangular prism and its facet centers as well.
Now the 5-point circuits imply that the point l' has to be in the interior
of this triangular prism. Hence, if we consider the diagram G'\{8', 9', 10' } ,
then this has the 5-point circuits listed above (so it describes the right
hexagon), and it has 3-point cocircuits 1', 2', 5', 1', 3', 6', and 1', 4',7' (so
iii
the long diagonals of the hexagon cross in 1', as required).
177
178
Then we choose the point 5 on the diagonal through 3 and 4. Thus we have
five points which then successively determine the points 6, 7, ... , 13. The
fourteenth point 14 is then placed at the intersection of the lines 1,3 and
2, 4, but not on the line 12, 13: this is possible if the point 5 had been drawn
slightly Southeast of the center, as in our figure, or if it is taken slightly
Northwest of the center, which results in a figure that is a reflection of our
figure; however, it is impossible (it results in a different configuration, with
12, 13, 14 collinear) to get a realization of the same configuration which is
y diagonal: for that we'd have to
itself symmetric with respect to the z
choose 5 in the center, and would get 12, 13 and 14 collinear.
----
1'
*1
'41!illP
5 8*
13
10
.
3
14
12
A configuration with this effect is not too hard to construct, but this
example has a stronger property: If the point 5 is chosen close enough
to the midpoint of the segment [3, 41, then the "Southeast" and "Northwest" realizations not only have the same collinearities (and thus the same
unsigned cocircuits, the same matroid), but they have the same signed cocircuits
both realizations yield the same oriented matroid. We do not
know whether one really needs 14 points for this effect: can you do with
less (Problem 6.26*)?
From this, one can easily construct an affine planar Gale diagram that
has a disconnected realization space. However, we must also make sure
that every polytope that is (only!) combinatorially equivalent to the one we
construct has the same diagram. A very "aggressive" method is to replace
every point of the diagram by a pair of positive and negative points: this
yields the affine Gale diagram of a 24-polytope with only 28 vertices that
has two isotopy classes of realizations. Implicitly, this is what the "Lawrence
construction" does, which we will discuss below.
179
Another nonisotopic point configuration, of 21 points, is explicitly constructed in White [522]. A general construction method is due to Mnv [378].
In particular, Mnv's universality theorem (see below) shows that the "realization space" for planar point configuration can be arbitrarily complicated.
Suvorov [5001 and Jaggi et al. [266] furthermore construct non-isotopic
configurations whose points are in general position. From these, one can
using a technique of Sturmfels [110, Thm. 6.5] get simplicial polytopes
that violate the isotopy conjecture. This yields examples for a much more
general "universality theorem for polytopes," described in the next section.
6.6
C*(P) n
>
For example, triangular prisms are rigid, but octahedra are not. The concept of rigidity is only interesting because there are some rigid polytopes
around, although "most" polytopes are not rigid. Here is one construction
to get rigid polytopes.
180
GG)
of a polytope with 2n vertices in R2n-(n- r) -1 = 118.n+d ; this polytope is
denoted by A(V) C Rn+d and called the Lawrence polytope of V.
Equivalently, we get the Lawrence polytope A(V) by successive Lawrence
extensions V o -z (V): for this we replace each vector '"in V by two
new vectors
Vi+ := Vi er+i
and
:= v i
to get an acyclic vector configuration in I r+n, from which we pass to affine
space
If we start with an affine point configuration X, we can perform the
Lawrence liftings directly on the point configuration, without linearization.
Our figures illustrate both the "linear picture" of a single Lawrence lifting
applied to a vector configuration
ii+
0 i
181
Lawrence polytopes are rigid, that is, if P' is combinatorially equivalent to A(V), then the oriented matroid of P' is also isomorphic
to that of P. In particular, the Gale diagrams of P' and of A(V) are iso-
Theorem 6.27.
morphic.
Proof. In the Gale diagram G of the Lawrence polytope A(V), and in
every Gale diagram G' with the same positive circuits, the points come
in pairs of positive and negative points, since those pairs form positive
circuits.
Furthermore, if we take any other circuit, then it contains at most one
point from every such pair. Hence we get all the circuits from the positive
ones by replacing a positive point by the negative "other point" of its pair.
Thus all circuits of the diagram are determined by the positive ones, and
thus the configuration V is rigid.
0
The Lawrence construction has numerous applications. Perhaps the most
striking one is the "universality theorem" for polytopes, which we want to
describe now.
For this, one needs a suitable equivalence relation for semialgebraic sets;
we use Richter-Gebert's version from [424]. Two semialgebraic sets S and
T are stably equivalent if they can be related by a sequence of "rational
changes of coordinates" (such that f and f-1 are both rational functions
with Q-coefficients, and induce homeomorphisms of the sets that we consider) and "stable projections" (whose fibers are the relative interiors of
rational polyhedra) see Richter-Gebert [424, Sect. 2.5] for the precise
definitions and more details. Stable equivalence is a very "restrictive" concept. In fact, stably equivalent sets S and T
have the same homotopy type (in particular, S is connected if and only
if T is connected)
182
(Mnv [378])
Every elementary semialgebraic set defined over Z is stably equivalent to
the realization space of some polytope.
Every open elementary semialgebraic set defined over Z is stably equivalent to the realization space of some simplicial polytope.
Essentially, this means that the realization space of a polytope can be
"arbitrarily complicated" - it can be disconnected with many components,
it can consist of circles and spheres (can have "homology" in all dimensions), and can have all kinds of complicated singularities in general it
is certainly not a manifold, as claimed in [429, p. 18].
The theorem on which all of this is based is Mnar's universality theorem: the realization space for planar point configurations (i.e., for oriented
matroids of rank 3) can be any semialgebraic set, up to stable equivalence.
For a long time, there was no detailed proof available for this theorem.
Mnv's paper [378] only provides a sketch of the basic ideas for the "local"
version of the theorem; two further sketches are in Shor [463, Sect. 4] and
in Goodman & Pollack [219, Sect. 7]; see also Bj6rner et al. [92, Sect. 8.6].
Finally, a complete, detailed proof was provided by Giinzel [243]. His proof
also covers the far-reaching extension announced in Mn6v [379], the "universal partition theorem" for oriented matroids.
On the other hand, it is much easier to see (using the "van Staudt constructions" for addition and multiplication of points, of classical projective
geometry [254, Sect. VI.7] [111, Sect. 2.1] [521, Sect. 7] that the smallest
subfield of Ill over which all planar point configurations can be realized is
the field of all algebraic numbers A C 11'. This means with Theorem 6.28
that A is also the smallest field over which all polytopes can be realized.
There is a great new development: Richter-Gebert's Universality Theorem for 4-Polytopes, and the (even stronger) Universal Partition Theorem
for 4-Polytopes, with all their corollaries and extensions.
Universality Theorem for 4-Polytopes 6.29. (Richter-Gebert [4241)
Every elementary semialgebraic set defined over Z is stably equivalent to
the realization space of some 4-dimensional polytope.
In the course of this work done after the first version of this book
appeared Richter-Gebert solved quite a number of basic open problems
(see Problems 5.11*, 6.10*, and 6.11*). There is neither time nor space
to explain Richter-Gebert's work here: See Richter-Gebert [424], and the
announcement in [426]. See also Giinzel [244].
Not es
183
Notes
For all information about oriented matroids, we rely on the monograph by
Bjiirner et al. [92]. Other expositions that include surveys on oriented matroids are Bachem [30], Bachem Sz Kern [31], Bokowski & Sturmfels [111],
and Bokowski [105].
As you may have noticed, we have deliberately tried to keep linear algebra
concepts low-key. You may reformulate all of the basic constructions in
more advanced language. For that, the vector configuration V E rXn is
r r , the space Dep(V) is the kernel
considered as a linear map V :
of this map, Val(V) is the image of the dual map, and so forth.
It seems that deletion and contraction of an "element" are fundamental
operations in many areas: for graphs (see Section 4.1), for vector configurations and oriented matroids (see Section 6.3(d)), and for arrangements
and zonotopes (see the next lecture). In fact, there is a tremendous power
in proofs "by deletion and contraction," which proceed by induction on the
number of elements, and by putting a structure together from the information given by deletion and contraction of the same element. Zaslavsky's
work on hyperplane arrangements [535] is the classic source for that approach.
Gale diagrams are a tool that emerged from work of Gale [205] and
were developed to their full power and beauty by Perles, as documented in
Griinbaum's book [234]. Additional sources are the book by McMullen &
Shephard [374, Ch. 3], McMullen's survey [366], and the treatment (with
nice illustrations and examples) in Ewald's book [189]. It seems that the
close connection between the Gale diagram technique and oriented matroid
duality was first mentioned in [366], and the explicit identification was
worked out by Sturmfels [494].
The reduction to affine Gale diagrams is implicit in Perles' work (see
Griinbaum [234, p. 59]) and also used by Bokowski [109]; it appears as a tool
of its own standing in Sturmfels' work [495]. We mention for completeness
that any two affine Gale diagrams of the same polytope are connected by
a projective transformation and the corresponding reorientation.
Many interesting properties of polytopes can be profitably studied from
the oriented matroid point of view, not only via Gale diagrams. Surveys of
applications are in Griinbaum's book [234], in Bokowski & Sturmfels [111],
and in Bayer & Lee [59, Sect. 4].
Some authors distinguish between "Gale diagrams" and "Gale transforms." We did not make this distinction, but essentially what we constructed here were Gale transforms, while any configuration that represents
the dual oriented matroid is a Gale diagram. We also just note that there
are several useful reformulations and variations of the Gale diagram construction, among them a "coordinate-free" formulation [190] [366], which
was useful in the investigation of infinite-dimensional polytopes by Kleinschmidt & Wood [310, 532].
184
For issues related to the isotopy conjecture, we refer to [92, Sect. 8.6].
The Lawrence construction is due to Jim Lawrence (surprise), but he
never published it. It appears in Billera Sz Munson [73, Sect. 2] and is also
explained in detail (in oriented matroid terms) in [92, Sect. 9 31
A "cr-construction" to produce rigid 6-polytopes from planar configurations was given by Sturmfels in [493]. However, we found the arguments
in [493] to be incorrect. (Specifically, the claim that the orientation of all
the "outer simplices" of a polytope is determined by the combinatorics of
the face lattice is only true for simplicial polytopes: for this, consider the
cone over a nonrigid polytope, where all simplices are outer; however, the
polytopes produced by the a-construction are not simplicial, and they turn
out not to be rigid in general.)
6.2 List all the circuits and cocircuits for the hexagon discussed in Section 6.1. How many vectors and covectors are there? (Don't list them
all, there are many.)
6.3 Show that the definitions of vectors, circuits, and so on, for the affine
( 11 )
and linear ces
as are consistent: if V -then
a-Dep(X) -= Dep(V)
and
'
a-Val(X) = Val(V),
and thus we get the same oriented matroid (the same circuits, cocircuits, etc.) for X and for V.
6.4 Let D = (V, A) be a directed graph with arc set A = {1, 2, ... , n}.
Define the signed circuits of D to be the sign vectors u E 10, +1 In
that correspond to circuits in D together with a chosen orientation,
0;
as follows: if the arc i is not contained in the circuit, then u
if it is in the circuit and directed according to the orientation, then
ui ,--- +; and if it is directed opposite to the orientation, then ui = .
-
185
For example, for the digraph drawn here and the oriented circuit
marked in it we read off the signed circuit u next to the drawing.
Show that the signed circuits we get that way are from a realizable
oriented matroid (as in Definition 6.5), whose cocircuits correspond
to the minimal directed cuts in the graph. Interpret the vectors and
the covectors of this oriented matroid in terms of the graph.
(Hint: Associate a vector configuration with D. A canonical choice is
= ei e3 for an arc from the node j to the node i.)
6.5 Prove that our two characterizations of acyclic vector configurations
are equivalent. Prove that the dual of an acyclic configuration is totally cyclic (Corollary 6.16).
Describe a small vector configuration that is neither acyclic nor totally cyclic.
6.6 A d-polytope with n vertices is simplicial if and only if every nonempty coface has at least n d elements. Derive from this a characterization of the (affine) Gale diagrams that represent simplicial
polytopes.
6.7 Given a Gale diagram, how can one (computationally) enumerate the
facets of the corresponding polytope?
6.8 Show that the following diagrams represent the four different combinatorial types of 4-polytopes with 6 vertices.
188
Construct a simple polytope for which the shape of a facet cannot be prescribed. For example, you might examine the polar of
189
(ii)* Can you prescribe the shape of a facet for simple 4-polytopes?
6.20 Let a centrally symmetric polytope with 2n vertices in
as P conv(X) for
be given
X=
Show that the dependences and the value vectors of X can be reconstructed from those of the following set of only n + 1 points in Rd:
X0
= fu,u
v2, ,u +
Thus the combinatorics of P can be read off from the dual configuration Go C (Rn-d ) * to Vo, the central (Gale) diagram of P, due to
McMullen & Shephard [373].
6.22 Consider the prisms over simplices, prism(Ad), and construct their
Gale diagrams. Show that they all arise as Lawrence polytopes.
6.23 Show that the oriented matroid given as an example for a 5-polytope
with a nonprescribable 2-face is not rigid. (Use the fact that one can
perturb the point 1 without changing any positive circuit.)
Is Perles' example of a nonrational 8-polytope rigid?
186
Show that the oriented matroid of this is dual to the oriented matroid of the cyclic polytope Cd(rt): so it is the Gale diagram of a
polytope that is combinatorially equivalent to Cd(n). Is this really a
Gale transform of Cd(n)?
6.14 Show that the following conditions are equivalent for a 2d-dimensional
polytope P with n vertices:
(i) P is neighborly.
(ii) Every set of n - d points forms is a positive covector.
d+ 1
187
neg-
7:
___----.
1
_____-----c)
8
0-------_________46, .70________---
Using part (0, solve Per les' problem for simplicial d-polytopes
with n = d + 4 vertices: every simplicial d-polytope with d +
4 vertices is a quotient (i.e., combinatorially equivalent to an
iterated vertex figure) of a neighborly (2d+4)-polytope with 2d+
8 vertices.
(Hint: For part (i) you will need the characterization of Exercise 6.14.
The Gale diagram formulation of Perles' problem is due to Sturmfels. See Sturmfels [495, Sect. 7], where some partial results are also
derived. Part (ii) was done by Kortenkamp [313].)
190
7
Fans, Arrangements, Zonotopes,
and Tilings
Zonotopes are the images of n-cubes under affine projection maps. Since
for most aspects of polytope theory n-cubes are not very complicated, this
definition may hide the complexity and richness of this concept. Zonotopes
are interesting from various points of view. Their combinatorial structure
is closely linked to (and in a precise sense equivalent to) that of real linear
hyperplane arrangements.
The aim of this lecture is to provide basic geometric intuition and the
tools for a combinatorial description of zonotopes. We will see how zonotopes again are modeled by oriented matroids, and discuss the surprising
appearance of general ("nonrealizable") oriented matroids in the study of
zonotopal tilings, and of hyperplane arrangements.
7.1
Fans
Definition 7.1.
A fan in Rd is a family
.T. = {Ci , c2 , 7 CN}
192
.7- is pointed if {0} is a cone in .F (and thus is a face of every cone in .F).
It is sirnplicial if all its cones are simplicial cones, that is, cones spanned by
linearly independent vectors. Simplicial cones and fans are automatically
pointed.
The following figure shows three complete fans in R2 , with N = 13,
N = 11, and N = 3 cones, of which 6, 5, and 2, respectively, are fulldimensional. The first two fans are pointed (for d = 2 this implies they are
simplicial); the third one is not.
There are various equivalent or similar ways to define fans (see also the
notes to this lecture). What we have given here essentially just describes
a polyhedral complex of cones (as in Definition 5.1) whose union is d .
In particular, the definition implies that the relative interiors of the cones
in Y form a partition of space:
N
Ed- relint(Ci ) = Rd
Now comes the first reason why we look at fans in the theory of polytope,s.
Example 7.2. Let P be a polytope in Rd with 0 E relint(P). We define
the face fan of P as the set of all the cones spanned by proper faces of P:
7.1 Fans
193
Our figure indicates the construction of the face fan, for a 2-polytope with
a given origin in it. Note that the geometry of the face fan does depend on
the position of the origin in P.
Example 7.3. Let P be a nonempty polytope in Rd. For the normal fan
of P we take the cones of those linear functions which are maximal on a
fixed face of P. That is, for every nonempty face F of P we define
NF :=
(Rd) * : F C fx E P : cx = rnax cy : y E
411\0}.
The above figure illustrates the construction of the normal fan of a 2polytope: for this we have identified 2 with (R2 )* via the usual scalar
product, which accounts for the right angles in the figure.
The face fan and the normal fans are very natural objects associated
with a polytope. In particular, they come up in the theory of optimization.
For this note that the question "Which cone of Ar(P) does c lie in?" is
the linear programming problem max cx : E P. Similarly, the question
"Which cone of .F(P) does y lie in?" is the separation problem of finding
one single valid inequality that determines whether av E P, for various
a > O.
Example 7.4. Let A
{HI , . . . Hp } be a finite set of linear hyperplanes
in Rd, where each Hi is of the form Hi =
e Rd : cix = 01 for some
ci ( d)*.
Clearly, the arrangement A decomposes I d into a complete fan TA. The
cones of the fan are also referred to as the faces of the (linear) hyperplane
194
arrangement A. The combinatorics of the fan encodes a lot about the configuration of row vectors C = {c1 , - . . , cp} it is quite easy to see that the
cones in .7' are in natural correspondence with the covectors of C. We'll see
details about this in Theorem 7.16 and Corollary 7.18.
In the figure after Definition 7.1, the first and the third fan are given by
hyperplane arrangements, the second one is not.
Example 7.5. There are also complete simplicial fans .F0 in R3 that are
nonpolytopal, that is, not of the form ,T(P) for any polytope P.
For one possible construction, we start from a simplicial 2-diagram D
that is not a Schlegel diagram, for example the one we constructed before
Theorem 5.7. This we place into an affine plane, and take all the cones
over faces of the diagram. We complete .F0 by adding one extra ray and
the simplicial cones that are spanned by this ray together with the cones
generated by the boundary of the diagram D. Our figure illustrates the
construction:
If Y is a fan in
Y0 g
and
195
If .F and g are both fans in the same space Rd, then we define their
common refinement as
IC n c' : C E T, C`
0.
.FIV :---
IC nV:
CE .F}.
It is easy to check that all three constructions are again fans in the
sense of Definition 7.1. If you want, you can consider gv := {V} as a
(noncomplete) fan in Rd: then the restriction to V is the intersection with
gv, that is, .FIV =
F A gv-
Lemma 7.7. Let P C il,I P and Q C Rq be two polytopes. Then the normal
fan of the product PxQC P+41 is the direct sum
ED
Also, the cones in the fan are characterized by the sign function: each cone
in the fan can be identified with a vector in {, , O}d, and the orthants
correspond to the sign vectors in {, } d
7.2
Rd be an affine map,
z,
196
dxP and z E Rd .
with A E (Rd)P
If 7r is injective (that is, A has rank p), then we refer to it as an affine
transformation, and 7r(P) is affinely isomorphic to P.
If 7r is not required to be injective, then we refer to it as an affine projection or an affine map. In this case Q := 7r(P) is again a polytope, whose
dimension is dim(Q) = dim(7r(RP)) = rank(A), where we had assumed
dim(P) = p.
We usually assume that Ir is surjective we may do this, after restricting
the image of 7 to 7r( r'P) C R d
and soir maps the p-polytope P C RP to
a d-polytope Q c Rd. Also, if we are only interested in properties that are
invariant under translation, then we may assume that z = 0 and that the
map 7r is actually linear.
Definition 7.9.
A projection of polytopes
z, such that p C I
is an affine map 7r : RP ---> Rd, x i--- Ax
p-polytope, Q C Rd is a d-polytope, and 7r(P) = Q.
-
is a
a
The linear algebra behind this construction is quite simple. From the
surjective map 7r : RP --4 Rd we get a dual map 7r* : (R()* --- (RP)*,
mapping c 1---+ c o 7r, which is injective. This distinguishes a certain set of
functions on P (and thus on Rd ): those that are constant on the fibers
of 7r. The embedding 7r* is used in the following lemma.
197
The normal fan Ar(Q) of a projected polytope Q is isomorphic, via 7r*, to the restriction of Ar(p) to the image of 7r*, the linear
subspace 7r* (Rd )*:
Lemma 7.11.
Ar(Q)
-`--. Ar(P)I7r*(Rd )
P + Pi :, {z + x' : x E P, x' E P } .
P+7
198
CI
7r
(;)
c(x
g*
1V(13 + Pi ) =-- _N-(7r(P x PT ,
=
using Lemma 7.7 for the direct sums, Lemma 7.11 for the projection, and
the dual map 7r*(c) .--- (e, c).
0
7.3 Zonotopes
Zonotopes are special polytopes that can be viewed in various ways: for
example, as projections of cubes, as Minkowski sums of line segments, and
as sets of bounded linear combinations of vector configurations. Each of
these descriptions gives different insight into the combinatorics of zonotopes. The following includes several such descriptions, all of which lead us
to the same "associated" system of sign vectors that describes the combinatorics of a zonotope. The main goal will be to see in what sense zonotopes
and arrangements can be considered equivalent, and how the combinatorial
structure of a zonotope is given by an oriented matroid.
After the general discussion of projections in the last section we now
consider a very special (but interesting) case: projections of cubes, that is,
ir : P --+ Q, x i--> Vx + z is an arbitrary (surjective) affine map, but P
is the d-cube,
P = Cp = fx E RP : --1 < xi < 1 for all il.
7.3 Zonotopes
199
Definition 7.13. A zonotope is the image of a cube under an affine projection, that is, a d-polytope Z C d of the form
Z -= Z(V)
:=
VCp + z = {Vy + z : y C Cp }
P
i=1
dxp .
Z
----------
Z(V)
and thus Z(V) = Hui, vi] + ... + {vp ,vpl + z for an affine map given by
r(y) = V y + z.
In the following we will usually assume that Z -- Z is centrally symmetric with respect to the origin 0, corresponding to a linear map 7F: C --* Z.
Example 7.14. By definition, the cubes Cd are zonotopes, where the projection map can be taken to be the identity.
Also, every centrally symmetric, 2-dimensional 2p-gon P2(2p) arises as
the projection of a p-cube to the plane. In fact, if the vertices of P2(2p) are
x i,
Xp,Xp+17 7 X2p
x2 xi x2
I
2
'
2
r
I .-.[
'
XP 1
1-
200
One way to prove this is by induction on p, by taking any pair of opposite (parallel, of same length) edges, and showing that it corresponds to a
Minkowski summand of P2 (2p).
7
8
We invite the reader to provide his or her own proof.
11n- 1 =
1 + [-
e2 - e i e2 - e
2
'
e3 - e i e3 - e l
l1 + [
en
2
'
en_ 1 en
2
'
1+-
2
en 1 ,
l
2
Perhaps the easiest way to see that this Minkowski sum yields the right
polyhedron is first to observe that it is invariant under permutation of
coordinates, and then compute the points of the sum that maximize a
linear function C E (Rn)* with e 1 < e2 < . . . < Co : this is easily seen to be
the vertex
v
--=--
n+1
2
e2 - el
'+
n +1
=
2 1
n-1
el
2
en - ea,_ 1
2
e3 - e l
2
n-3
2
e2
--.
n + 1 - 2n
en
2
1)
=(2
;.1
7.3 Zonotopes
201
is a zonotope, by McMullen [360]. The same fails if only all the facets
(k = d -1) are centrally symmetric: in this case we have a counterexample in
the regular 24-cell (a sporadic 4-dimensional regular polytope, described for
example in Coxeter [156, Sect. 8.2]), whose facets are all regular octahedra;
counterexamples for all d > 4 can be found in McMullen's paper [365].
In particular, we see that being a zonotope is a geometric property, not
a combinatorial one. For example, the quadrilateral CQi is not a zonotope,
but the quadrilateral Q2 is.
Thus being a zonotope is preserved under affine equivalence (in fact, under
affine projections), but not in general under combinatorial equivalence.
It may require a second of thought to figure out that in general zonotopes
are not simple polytopes (though the permutahedra are). Our next picture
shows a zonotope generated by four line segments in R3 , no three of them
coplanar. The resulting zonotope (d = 3, p = 4) has 8 simple vertices of
degree 3, and 6 nonsimple vertices of degree 4. The figure indicates (by a
dotted line) that the vertex figure of the top vertex is a square.
202
F =
Aiei :
Ai = +1 for ui = +,
Ai = 1 for ori
Cp :
Recall from Section 6.3(a) the componentwise partial order on sign vectors,
which is induced by
We find that the smaller a face F of Cp , the larger its sign vector u(F)
will be in this partial order "<." Also, in order to obtain the whole face
lattice we have to add an extra minimal element, since the empty face does
not contain points, and it does not have an associated sign vector, either.
Thus, we get
(L(Cp), g)
{6} u
The following sketch shows the signs associated with the faces of C2, and
the face lattice L(C2) with the signs corresponding to its elements.
(00)
(0+)
(-
(00)
(0-)
7.3 Zonotopes
203
With this we get sign vectors not only for the faces of cubes, but also for
the faces of zonotopes: if ir : Cp + Z is the projection that defines the
zonotope Z, then for every nonempty face G E L(Z) of the zonotope we
get the nonempty face 71 --1 (G) E L(Cd), and thus we put
(L(Z), g)
i--,-t
({0-(c): G E L(Z)\0}, ? ),
where is defined to be smaller than any sign vector u(G). Caution: the
partial order on the face lattices is opposite to the order on sign vectors.
The larger a face of Z, the smaller a sign vector we associate with it, in
the partial order on sign vectors induced by 0 < and 0 < -.
This assignment of sign vectors to the faces of a zonotope may look a
little mysterious. In particular, the description that we have given (from
the projection of a cube) does not tell us much about the structure of
the collections of sign vectors that we get, and how they are related to the
matrix V that defines the projection. Thus, we take a "fresh start" here, and
obtain the same sign vectors from a different approach, via optimization.
The following theorem is so basic that we give two proofs. The first proof
shows how a sign vector erc, E {,,0}P is associated with every cone in
the normal fan Ar(Z) of the zonotope. Similarly, the second proof obtains
a sign vector u(G) E {, -, 0}P for every face G E L(Z) the same sign
vector that we have just "pulled down" from the p-cube, of course. Thus
after this we have three different constructions of the sign vector system
associated with a zonotope.
A = A v := {H1 , . . . , Hp }
in Rd given by
:=
lc E (Rd)* :
cv, = 01.
First Proof. For each of these hyperplanes we define the positive halfspace
by
Ilii-
1c E (Rd
) *
204
and the negative one similarly. Thus the normal fan of the single line segment [v i , /Ji], is the set {Hi, Ht, a hyperplane and the two halfspaces determined by it.
HT
Now we use Proposition 7.12 to see that the normal fan of the Minkowski
sum of the line segments [-vi, Ili] is the hyperplane arrangement Av , that
is, the common refinement of the fans of the individual hyperplanes.
The position of c with respect to the fan {Hi, Ht, HT} is determined by
the sign of cvi . In fact, if sign(evi ) = 0, then c lies in Hi ; if this sign is +,
then e is in the interior of Hz ; and if it is , then c lies in the interior
of HT.
Thus in the common refinement A v , the position of c is given by the
sign vector
sign(eV) E I+, MP,
whose first coordinate records the relative position with respect to HI , the
second coordinate refers to H2, and so on. In particular, we get distinct sign
vectors for the distinct cones in Ar(Z), and inclusion of cones corresponds
to the usual partial order on sign vectors.
Our sketch shows a zonotope, its normal cones (drawn into the same
figure with right angles), the normal fan assembled from them (which is a
hyperplane arrangement), and the signs that we associate with each of the
cones in the normal fan.
(-++)
(- 0+)
(0--
(0++)
(+0-)
(+--)
7.3 Zonotopes
205
cx
Zc =
{yEZ:Cy ISEWICZ}
Ein=1
A i = 1
1 < Ai < +1
if cvi <0,
Ai = +1
if cv i > 0 1.
if cvi = 0,
Thus the decision "which face of Z maximizes c" is equivalent to the decision, for each i, of whether c lies on the hyperplane Hi itself, on its negative
side, or on its positive side, that is, by the position of c in the fan of the
arrangement A.
The family of hyperplanes A thus gives a combinatorial interpretation for
the covectors of the configuration V = {v 1 ,. vn }. Here the interesting
case is the one where the configuration V spans Rd , such that the zonotope
Z(V) has dimension d, and the hyperplane arrangement Av is essential: the
intersection of all the hyperplanes is the origin, 1/1 n H2 n. n H = 101,
and thus the cones in Av are all pointed cones.
Corollary 7.17. Let V E Rdx P be a vector configuration in Rd . Then
there is a natural bijection between the following three families:
(the sign vectors of) the nonempty faces of the zonotope Z (V) C
(the sign vectors of) the faces of the hyperplane arrangement Ay,
the signed covectors of the configuration V.
L(Av)
V * (V) g (I+,
0 113 ) *
Similarly (assuming that V has full rank), there is a natural bijection between the following three families:
(the sign vectors of) the facets of the zonotope Z(V)
(the sign vectors of) the one-dimensional rays of the hyperplane arrangement A v , and
the signed cocircuits of the configuration V.
206
In notation:
facets(Z(V))
vert(Z(V)) -->
rays (Av)
C* (V).
The ith zone of Z(V) is the collection of all faces that have [vi , vi] as
a Minkowski summand. The zones geometrically form "belts" around the
surface of a zonotope, and in fact completely cover it.
These zones may also be held responsible for the name "zonotope." Under the bijection between faces of a zonotope and cones in its hyperplane
arrangement, the ith zone corresponds to the ith hyperplane Hi in the
arrangement.
Note that vi and Ili determine the same zone, and the same hyperplane,
exactly if they are parallel vectors. This is a degenerate case that one usually excludes from the discussion. In fact, there is an even more degenerate
case, if vi = 0 for some i. In this case v i does not contribute to the geometry, we can just ignore it when constructing the zonotope, H = (Rd)* is
the full dual space, and i is a one element circuit (known as a loop) in the
oriented matroid, which can safely be deleted.
The number of zones is the principal complexity measure for zonotopes:
it coincides with the number of different hyperplanes in the associated
arrangement, and with the number of equivalence classes of elements in
the oriented matroid. If we assume that V is simple, that is, there are no
zero or parallel vectors in V E Rdx P, then the number of zones is p. The
key observation is that this parameter can be read off directly from the
zonotope Z, and does not depend on the choice of V. However, for every
zonotope Z there is a simple vector configuration that defines it, and the
vector configuration is unique up to permutations and sign changes.
Under the translation from zonotopes to arrangements and back, simple zonotopes correspond to simplicial hyperplane arrangements a very
7.3 Zonotopes
207
In particular, if V spans Rd, then the arrangement is essential, the zonot ope
is full-dimensional, and its polar is a palytope.
Thus the fan of an essential hyperplane arrangement is always polytopal.
The following constructs an explicit example for this. Note that from our
set-up the proof is simple but the geometric fact that we can construct
a polytope that "fits" any given hyperplane arrangement is not obvious at
first sight.
Example 7.19. For the matrix
= (
-1 1
1
we construct
208
hyperplane arrangements
zonotopes
vector configurations
..
209
{ i
if u,
0,
y-
: u = --vi 74 0}.
> -U E V *
(V1) u E V*
("The negative of a covector is always a covector")
(V2) u,v E V*
> uovE V*
("The set of covectors is closed under composition")
(V3) u, V E V * , j E S(U,y) == NtNi E V*: w eliminates j between u and v
("The set of covectors admits elimination")
The rank of the oriented matroid V* is the largest number r such that V*
contains a chain of covectors of length r:
with X i E V * .
SIGN(U) = {sign(u) : u E U} C
Ip
of dimension r.
{, --, Or
210
40.
40.
Considering the individual components of the sign vectors, we see that the
5th coordinate of w is zero. Similarly, if the ith coordinates of u and y
don't have opposite sign, then the ith coordinate of any positive linear
combination has the sign sign(wi ) = (u o v) i , as required.
211
Remark 7.24:
212
Oriented matroids provide a unifying framework, consistent terminology, and widely applicable tools for several areas of geometry.
Within the last twenty years, an extensive theory for oriented matroids was developed, with many nontrivial results. Results about
oriented matroids are easily transported from one field of applications
(e.g., hyperplane arrangements) to another one (like polytopes). For
the existing body of theory, we refer to the book by Bjiirner et al. [92],
and the handbook chapter by Bokowski [105].
The theory of oriented matroids allows us to handle, in a precise
sense, the combinatorics of objects that are geometric (like certain
simplicial spheres) but that cannot be represented in real "Euclidean
space" (or at least not as polytopes).
So in our drawing, for the case d ----- 3, we obtain an affine arrangement Aaff
of n 1 = 4 lines in the affine (dotted) plane, from an original arrangement
213
Definition 7.25. A pseudoline is a polygonal curve without self-intersections, with finitely many break points in R2 , and whose ends "head off to
infinity" in opposite directions.
An arrangement of pseudolines is a finite set of pseudolines in the plane
such that
(i) any two pseudolines either are disjoint (then we call them parallel), or
they meet in a single point and cross in this point, and
(ii) being parallel is transitive (that is, if a pseudoline intersects one pseudoline of a parallel pair, then it also has to intersect the other one).
ti
214
For example, in our next drawing small arrows are used to indicate the
positive side for each pseudoline. The shaded region (without its boundary) is the face associated with X = (-1---), the bold edge (without
endpoints) is the face associated with (-0+--), whereas (---0--) does
not correspond to a face.
3
.---->.
, -e4
{ (X',+) : F is a face of P}
{ (VG, 0 ) : G is a face at infinity for P} U {(O, 0)}
g f+,,01n
{(--X F, ) : F is a face of P}
215
Proof. From an arrangement p of n 1 pseudolines, the previous procedure reconstructs a linear arrangement of n "pseudoplanes" through the
origin in R3 , where the nth pseudoplane is a straight plane, corresponding
to the line at infinity in P.
The oriented matroid v*(p) c {+, , 0}n arises from this arrangement
in the same way as in the case of a straight arrangement of planes in R3 .
Similarly, the proof that the system V* (P) is an oriented matroid is analogous to the realizable case in Proposition 7.22 except that the "linear
algebra arguments" of that proof have to be replaced by "combinatorial
arguments" that remain valid in the nonlinear case.
We leave the details to the reader, and refer to [92, Sect. 5.1], where the
111
proof is nicely done in even greater generality.
In fact, there is a surprisingly strong theorem available here: Lawrence's
topological representation theorem states that
every linear arrangement of n pseudohyperplanes in Rd yields an
oriented matroid V* C {,,O } of rank d.
Theorem 7.26 just presents the case d = 3 of this statement. (We do not
intend to give a precise definition of arrangements of pseudohyperplanes
here. Intuitively this should be clear; see [92, Ch. 5] for a careful explanation.) The topological representation theorem, however, also includes a
converse:
every oriented matroid on n elements of rank d can be represented
by an arrangement of n linear pseudohyperplanes in Rd , which is
essentially unique.
216
Pappus' theorem states that the three black dots are collinear in every line
arrangement that is combinatorially equivalent to this arrangement. This
implies that there is no straight representation of the following pseudoline
217
There are even nonrealizable pseudoline arrangements that are simple (no
three pseudolines cross in a point or are parallel) see Ringel [427] [237,
p. 42] and Exercise 7.16 for an example with only 9 pseudolines in the
plane In fact, all (simple or nonsimple) arrangements of pseudolines with
at most 8 pseudolines are realizable (for this count the line at infinity, if
it is there), and Ringers example is essentially the unique simple one with
9 pseudolines, according to Richter [420].
218
219
220
Thus, we work in a different direction: in view of the topological representation theorem it is sufficient to construct the oriented matroid associated
with a zonotopal tiling and this is just a system of sign vectors, without
any topology! Here we go.
The following provides the basic construction and shows how the faces in
a zonotopal tiling get sign vectors associated with them, almost canonically.
Construction (with Definitions) 7.31.
([104], [425, Sect. 11)
Let Z be a zonotopal tiling in d
Two edges e, e' E Z are defined to be equivalent if there is a sequence
e =----- eo , e l , ... , et = e' of edges in Z such that ei_ I and ei are opposite
edges in a 2-face of Z, for 1 < i < t.
If this divides the edges in Z into n equivalence classes, then n is the
number of zones of Z. Let El , 2 , ... , En enumerate the corresponding equivalence classes of edges. The ith zone of Z, denoted A, is the collection of
all those F E Z which have a face in E.
.
ei
221
xr
The set
0(2) := 1XF : F E
,
_ _negative side
of the zone Z1
In the next sketch, all zones have been labeled and directed. The associated sign vector is indicated for one vertex, one edge, and one 2-face of the
tiling.
222
V*
:----
{ (X, +) : X E 0(Z)}
U
(Y, 0) : E V*(V)}
{(X,---) : X E 0(Z)}.
{X e
{+, ,0}n :
,0}n 1 with
(X, 0) 1-1} = V*(V).
223
However, in the framework of pseudoarrangements one can construct oriented matroids (in terms of arrangements of pseudohyperplanes) that have
fewer than 2n simplicial regions. The first example of this kind was presented by Roudneff & Sturmfels [432].
The current "world records" about simple vertices are due to RichterGebert [421, Thm. 2.21, who constructed oriented matroids of rank 4 on 4n
elements that have only 6n simplicial regions. Furthermore, from RichterGebert's example [421, Thm. 2.3] we get an oriented matroid R(20) of rank
4 which has a pseudohyperplane "8" that is not adjacent to any simplicial
region. Furthermore, the restriction to the pseudohyperplane 8 (the contraction R(20)/8 of the oriented matroid) is realizable. Via the Bohne-Dress
theorem, these results translate into the following theorem.
224
Theorem 7.35.
(j ) [432]
(ii) [421, Thm. 2.2] For k > 2 there are zonotopal Wings of rank 4 (in 1R3 ),
with 4k 1 zones (n = 4k), which have only 3k + 1 = ln + 1 simple
vertices.
There is a 3-dimensional zonotopal tiling with 19
zones, which has no simple vertex on the boundary.
You should try to visualize these in view of the "geometric" description of the pseudoplane arrangements in Richter-Gebert's paper with many
drawings this is not out of reach. A photo of a geometric model for the oriented matroid of part (0, built by Bokowski and Richter-Gebert, can be
found in [105, p. 562].
Notes
The permutahedron was first written about by Schoute in 1911 [445], it
seems General zonotopes were known to Blaschke [97, p. 250]. The first
systematic investigation of zonotopes was in Bolker [112], followed immediately by Schneider [439], and then by McMullen [364], who developed
zonal diagrams a version of Gale diagrams suitable for "zonotopes with
few summands" (see Exercise 7.7). There is a revived interest now, due to
the connection to oriented matroids, hyperplane arrangements, aspects of
optimization, computational geometry and convexity, and so on. We refer
to the surveys by McMullen [366] and by Schneider & Weil [442], to [92,
Sect. 2.2], and to the paper by Gritzmann & Sturmfels [226] and the references therein. More on Minkowski sums can also be found in [226].
The subject of hyperplane arrangements has a lot of different aspects,
and we do not even try to give an introduction here. We refer to [92] for the
case of real hyperplane arrangements and their oriented matroids, and for
further references. Arrangements of lines and pseudolines (corresponding
to arrangements of rank 3) are beautifully discussed by Griinbaum in [237].
Fans, polytopal or not, are of great interest for algebraic geometry. In
particular they represent toric varieties. In this case, the interest is restricted to fans that are pointed (i.e., {0} e .r) and rational (every cone
is generated by rational vectors). We refer to books by Fulton [2011 and
Oda [396], and in particular to the combinatorial treatment in Ewald [189].
Simple zonotopes exist, but they are rare. As we have seen, they correspond to simplicial arrangements of hyperplanes. Examples of such arrangements arise naturally in the theory of Coxeter groups, root systems,
225
and Lie algebras [1201 [128] [265] [92, Sect. 2.3]. There is a conjecture that
except for a few "obvious" infinite families, most of which come from these
theories, there are only finitely many "sporadic" examples: but currently
no one seems to have the faintest idea how to prove this. We refer to work
by Granbaum [236] for the case r = 3, and to Griinbaum & Shephard [240]
for the case r = 4. The enumeration of all "known" arrangements of rank 3
attempted in [236] had only one addition and one correction up to now
(Griinbaum [237], and Barthel, Hirzebruch & Iltifer [52]) and might be essentially complete, while the enumeration of [2401 for r > 4 is probably far
from complete; see Alexanderson & Wetzel [7, 8, 9]. Up-to-date references
can be found in Wetzel [519].
The Bohne-Dress theorem was announced by Andreas Dress at the 1989
Symposium on Combinatorics and Geometry in Stockholm. It is a strikingly simple geometric observation that had previously eluded people. A
complete proof, however, is surprisingly difficult, and it took some time
until the complete written version by Bohne [104] was available. A simpler, more geometric proof is given by Richter-Gebert & Ziegler [425]. Our
sketch in Section 7.5 follows that paper.
The Bohne-Dress theorem relates the set of all zonotopal tilings on a
given zonotope with an extension space problem ("Is the space of all extensions of an oriented matroid homotopy equivalent to a sphere?," see
Sturmfels & Ziegler [499]). Thus zonotopal tilings allow one to study a special case of two very basic, general, and apparently very difficult problems,
the generalized Baues problem" of Billera, Kapranov & Sturtnfels [69 ], and
the problem of "Combinatorial Grassmannians" by MacPherson [346], see
also in Mnv & Ziegler [380]. We will discuss the setting of the Generalized Baues Problem in Lecture 9. For the problems themselves and their
ramifications we refer to the original sources.
226
Show that the projections to R2 are not good enough for this.
(Witsenhausen [531])
7.6 Interpret the deletion and the contraction of a vector in the configuration V in terms of zonotopes. That is, describe how the zonotopes
Z(V\v) and Z(V/v) can be constructed geometrically.
7.7 Let Z(V) be the zonotope generated by a configuration V E I. d x n
G E (R*( n-d / Xn be the dual vector configurawhich spans I.
tion.
).
(i) How can the combinatorics of the zonotope Z(V) be read off
from the configuration G?
(ii) Use this to describe the zonotopes with n < d + 2 zones.
(iii) Describe the relation between the zonotope Z(V) and its asso-
(Rn-d )* .
7.8 Assume that x, y E Rd are given. Give an explicit formula for some
small enough E > 0 such that sign(x Ey) = sign(x) o sign(y).
7.9 Let V* C {-1-, --,O }n be a system of sign vectors.
(j ) Assuming that (V0): 0 E V* holds, show that the axioms (V1)
and (V2) together are equivalent to the axiom
(V2'): u, V E V
==U 0 (V)
V.
227
o
y3 , v4 1)
\0 j
Show that the family V* := { 6(U): U C V} is an oriented matroid.
What is its rank?
What is the relation to the oriented matroid associated with such a
digraph according to Exercise 6.3?
7.11 Show directly from the axioms in Definition 7.21 that every oriented
matroid of rank r < 2 (that is, an oriented matroid V* C {, , Or
that does not contain a chain 0 < X < < X") is realizable.
C* =MIN(V*),
228
7.13 Show that for any vector subspace U C Rn , the set of minimal
nonempty supports in II = SIGN(U), given by MIN(SIGN(U)) =
fu E sign(U)\{0} : v E sign(U), supp(v) C supp(u) implies v = 01,
is the set of cocircuits of an oriented matroid, that is, it satisfies the
axioms of Exercise 7.12.
7.14 How can you test whether a given zonotopal tiling is the picture of an
actual zonotope? Show that, essentially, one has to decide whether a
certain polyhedron has nonempty interior, which can be solved as a
linear programming problem, as in Exercise 5.2 (i ).
So, is it true that the first figure in Section 7.5 represents the picture
of a 3-dimensional zonotope?
7.15 Show that every nontrivial zonotopal tiling in R2 (a tiling of a centrally symmetric polygon by centrally symmetric polygons) has a simple vertex on the boundary, and also a simple vertex in the interior.
7.16 Consider the following zonotopal tiling.
229
(The second configuration is closely related to Ringel's simple configuration of 9 pseudolines, one only has to delete the line at infinity
and perturb the arrangement there.
These drawings were produced by Jiirgen Richter-Gebert, using his
postscript program described and listed in [422], which produces exceptionally nice pictures of zonotopal Wings.)
230
7.18 For every d-zonotope, the numbers fk of k-faces satisfy the relations
>
dk+l
d
) fo .
k
(This is given in terms of hyperplane arrangements and oriented matroids in Fukuda, Tamura Sc Tokuyama [200, 1991.)
7.19 For any vector configuration V = {v 1 , , vn } c lY, prove the volume formula for its zonotope:
vol(Z(V)) = 2d -
E
1<ii<
Idet(v i
, vid )I .
<td<n
For this, decompose the zonotopes into parallelepipeds, whose volumes are given by determinants.
(McMullen, see Shephard [459, Sect. 51, from where we have also
taken the illustration.)
7.20* You can use the formula in the preceding exercise to compute the
volume of a zonotope, but that is not very effective: the formula
has
terms, which may all be nonzero.
Is there a fast (polynomial) way to compute the volume of a zonotope?
(Answer: most probably not this is #P-hard according to Dyer,
Gritzmann & Hufnagel [177 ] .)
(n
d)
Sheliability and
the Upper Bound Theorem
Perhaps the most famous result about convex polytopes is the EulerPoincar formula:
fi
fo
232
Then, we'll present McMullen's proof for the upper bound theorem, give
a glimpse of extremal set theory, and end with the famous 9-theorem, and
derive some of its surprising consequences. So, there's a lot to do: let's get
going.
... wait, here is one more remark. This lecture has a distinctive "topological" flavor. In fact, already the first correct and complete proof of the
Euler-Poincar formula, by Poincar [410, 411], was done using tools of algebraic topology that Poincar had developed himself. Here we will avoid
most topological subtleties, for example by restricting our attention to polyhedral subdivisions of polytopes and their boundaries, instead of subdivided
topological balls and spheres. Thus, for this lecture no knowledge is needed
of the wonderful subtleties of piecewise linear topology, nor of the powerful
machinery of algebraic topology. Nevertheless, it is helpful and desirable,
and good for your intuition, if you take, at least, an excursion into these
worlds. I recommend Stillwell [491], Munkres [388], Daverman [171], and
Bjeirner [85] as guides to different points of view.
(i) Every polytope P together with all its faces forms the polytopal complex C(P). The only maximal face ("facet") of this complex is P itself.
(ii) All the proper faces of P form the boundary complex C ( P), whose
facets are just the facets of P. This is a pure simplicial complex of
dimension dim(P) 1.
233
<
C(3F1 )
(ii) For 1 < j < s the intersection of the facet Fi with the previous
facets is nonempty and is a beginning segment of a shelling of the
(k 1)-dimensional boundary complex of Fi , that is,
j-1
Fi
FT)
G1
G2 U U
z=1
234
Examples 8.2.
(i) Every 0-dimensional complex is shellable, by definition. A 1-dimensional complex (a graph) is shellabIe if and only if it is connected. In
particular, this means that it is pure (i.e., has no isolated vertices). A
shelling order is an ordering of the edges e l , e2 , ... , es in such a way
that the set { e l , ... , ei } describes a connected subgraph for every j:
this comes from the condition that the intersection of the edge e3
with the earlier edges has to be 0-dimensional, and thus nonempty.
(ii) The following are three 2-complexes in the plane R2 .
The first two are not shellable, but the third one is. (Check this!) In
each of them, there is a beginning of a shelling indicated, that is, the
complex given by the numbered facets together with their faces is
shellable. However, if you try to add the last facet in any of the first
two examples, you violate condition (ii) of Definition 8.1.
(iii) Every simplex is shellable, and every ordering of its facets is a shelling
order. This immediately follows by induction on the dimension, since
the intersection of Fi with Fi (i < j) is always a facet of F3 in this
case.
(iv) The d-cubes are shellable: by induction on dimension one can show
that every ordering of the 2d facets F1 , F2 , ... , F2d such that the
first and the last facet are opposite, F1 = F2d, is a shelling order. (The condition F1 = F2d is sufficient, but not necessary, see
Exercise 8.1(i)!)
(N) The pile of cubes Pd(a l , ... , ad), see Example 5.4, is shellable for
arbitrary finite ai > 1. For this, we use part (iv) to see that the
lexicographic order on the little cubes in the pile is a shelling order.
Remarks 8.3.
(i) We will see in the next section that condition 8.1(i) is in fact redundant: the boundary complex C(8F1 ) of every polytope is shellable.
However, if one defines shellability more generally for cell complexes
rather than polytopes, as in Bjiirner [81], then this is necessary.
(ii) For simplicial complexes condition 8.1(i) is redundant because of Example 8.2(iii). In this situation, condition 8.1(ii) can also be simplified
considerably: it can be replaced by
235
8.1(W) For I < j < s the intersection of the facet Fi with the previous
facets is nonempty and pure (k 1)-dimensional.
In other words, for every i < j there exists some I <j such that
the intersection Fi n Fi is contained in Fi n Fi , and such that
Ft n Fi is a facet of Fi .
(iii)
8.1(ii11) For 1 < j < s the intersection of the facet Fi with the previous
facets
i--1
Fn (U Fi) = Gi u ... U Gr
i=1
Examples 8.4.
(i) Every polytopal subdivision of a 2-polytope is shellable see Exercise 8.0.
(ii) The boundary of every 3-polytope is shellable. This follows from (i):
for this we first shell a Schlegel diagram D(P,F) of P, which is a subdivision of a 2-polytope. This corresponds to a shelling of the whole
boundary OP except for the facet F. The shelling can be completed
by taking F as its last facet.
Subdivisions of 2-polytopes and boundaries of 3-polytopes are easy to
shell. One reason is that no matter how we start the shelling, we can't get
stuck. Let's introduce some fancy terminology for this.
236
Lemma 8.6.
ro
Two
ff
r
O
Af
mfAr.149 0
aria*
:1410
"Iir WIMP
lOr aI I I
41111/41ffilala
AIPerarawfa
10
kil150
PALP
ATIZI1
rillIAWAW0
41111IAIVAIWAWANIMIIII,AIMIA1171
idwarderAwderAwarAn
ArAiramorAmormrAlisimr,
ArrAmordwrArramorararraro
411/411101111P2111/4111FAININFAIIII
AVAIIPTAFAIVAIIIIIIPTAIINIVA/1
Adur,mrarr,mprdurdrPrar.arav
111111 111111111
111111
MM.
MINER
MEN
N.=
If we try to add any new cube of the pile P3(9, 9,4) to the subcomplex,
then the intersection with what is already there is not pure 2-dimensional,
or not connected, or both. Thus, this is a shellable part of P3(9, 9, 4), and
we are stuck: so P3 (9 7 9, 4) is not extendably shellable.
D
237
star(v, C)
Lemma 8.7. Let C be a shellable simplicial complex, with shelling order F1 , F2 ,. , Fs . Then the restriction of this order to star(?), C) yields a
shelling order for the star, and also for link(v, C).
Proof. We directly verify condition 8.1 (u i1 ): let Fi be a facet in the star
(so y G Fi ), and let Fi be an earlier facet that also lies in the star (with
< j and y E Fi ). Since we have a shelling of C, there is a facet F1 with
1 <j such that Fi n F3 C F1 n Fi . But this implies that y E F1, so F1 is in
the star of v.
The same proof also shows that we get a shelling of the link, since (in
the simplicial case) we have a bijection between k-faces G E link(?), C) and
LI
(k+1)-faces a = conv(G U y) E star(v, C).
Theorem 8.8. (Rudin [433 ] )
The 3-simplex 6.3 can be triangulated in a nonshellable way.
The first nonshellable triangulation of a tetrahedron (with 14 vertices,
41 facets, all vertices on the boundary) was constructed by Rudin [433]
in 1958. Her construction is subtle and hard to visualize, and it seems to
be the only one in the literature. So, instead of reproducing it, here is a
different construction that shows that the tetrahedron and the 3-cube have
nonshellable triangulations.
238
of the tetrahedron
T3 :=--
IX
E R3
+x+y+z<2
+x y z < 2
x+yz <2
x y + z < 2
2\
2 ) ( +2 ) ( +2 )
2 j , ( +2 ,
2 , +2
2
+2
+2
2
It turns out that the 12 edges we get (see the drawing) are grouped into
four disjoint triangles. The key property is that every edge is surrounded
by a triangle. By symmetry, it is in fact sufficient to verify this for one
single edge e E E12 .
The next step is to construct a triangulation of the cube C3, respectively
of the tetrahedron T3, which contains the edges in E12 as faces (not subdivided!). While this can be done explicitly, here we resort to a powerful
tool: Whitehead's completion lemma [523, Thm. 5], according to which every partial triangulation of R3 can be completed. (In fact, the same is true
in R d , according to Bing [76, Lemma 6]. See Bing [78, Sect. 1.2] for a textbook version.) Hence, we can take any triangulation of the boundary of C3,
respectively T3, that uses the vertices in V12 , plus the twelve edges in E12 ,
and complete this simplicial complex to a triangulation of C3, respectively
T3. There will be additional interior vertices necessary for this.
239
Since the circle C surrounds the edge e, we have to "pass over" a vertex
C e when we contract C in C3 . Also, the link of this vertex y in the
complex C3 , link(Y, C3), is a 2-dimensional shellable complex, by Lemma 8.7.
However, we can contract our circle C within IC3 I until it lies in link(v, C3 ),
but it cannot be contracted within the link, because then it would not pass
over v. This shows that the link is not shellable: contradiction.
El
8.2
Shelling Polytopes
240
Well, this is the essence. However, what one actually needs is that there
are shellings with very specific properties. These are obtained from the
Bruggesser-Mani construction, which yields the much more specific theorem
below (which includes later refinements by McMullen [3611, Danaraj & Klee
[163 ] , and by Bj6rner & Wachs [94 ] ). Also, this is the technical statement
that has an easy proof by induction on the dimension.
241
Here we can use our intuition to understand what visible means: a facet
F C P is visible from x if for every y E F the closed line segment [x, y]
intersects P only in the point y. Equivalently, F is visible from x if and
only if x and int(P) are on different sides of the hyperplane aff(F) spanned
by F. For example, if xG is beyond the face G (in the sense of Section 3.1),
then the facets that contain G are exactly those that are visible from xG.
aff(F2)
A
242
After a while, a new facet will appear on the horizon: the rocket passes
through a hyperplane aff(F2 ), and we label the corresponding facet F2Continuing this, we label the facets F3, F4, . . . in the order in which the
rocket passes through their hyperplanes, that is, in the order in which the
facets appear on the horizon, becoming visible from the rocket. Now we pass
through infinity, and imagine that the rocket comes back to the planet from
the opposite side. We continue the shelling by taking the facets in the order
in which we pass through the hyperplanes aff(Fi ), that is, in the order in
which the corresponding facets disappear on the horizon.
This "rocket flight" clearly gives us a well-defined ordering on the whole
set of facets. Also, the facets that are visible from x form a beginning
segment, since we see exactly those at the point where the rocket passes
through x.
To see that the ordering is a shelling, we consider the intersection OFi n
(F1 u u Fi _ i ). If Fi is added before we pass through infinity, then this
intersection is exactly the set of those facets of Fi that are visible from the
point t n aff(Fi ), at which F3 appears on the horizon. Thus, we know by
induction on the dimension that this collection of facets of Fi is shellable,
and can be continued to a shelling of the whole boundary aFi .
aff(F7 )
I
243
of the line i also reverses the line shelling. Thus the reverse of every line
shelling is not only a shelling (by Lemma 8.10), but a line shelling as well.
The Bruggesser-Mani construction has a lot of flexibility: we can get
shellings with special properties by suitable choice of the shelling line L.
Corollary 8.13.
Proof. For the first claim, choose any shelling line 1 which intersects the
boundary of P in the facets F and F'. (For example, choose x beyond F,
choose x' beyond F', and let f be the line determined by x and z'. Perturb
f to general position, if necessary.)
For the second claim, let x v be a point beyond the vertex y, and choose
the shelling line to contain this x v .
111
Corollary 8.14. Every Schlegel diagram is shellable.
More generally, every regular subdivision of a polytope is shellable.
Proof. For any Schlegel diagram D(P, F), choose a shelling of the polytope
P such that the facet F comes last. Thus the shelling of P also induces a
shelling of the Schlegel diagram D(P, F).
Every regular subdivision Ep(Q) of a d-polytope Q is, by Definition 5.3,
isomorphic to the complex of faces of a (d+1)-polytope P that are visible from a certain point z = Ted +i: and thus we can apply Theorem
8.12.
0
244
Proof. (1)
245
moiffo
Shellings of polytopes allow us to "build up" the boundary of a polytope step by step, adding one facet at a time. Thus one can do proofs by
induction on the number j of facets in the complex
C3 := C(111 U F2 U U F3 ),
f (C)
(fi,
fa,
- , fd)
c Nci+2 ,
, f_)
Note that all the f-vectors we consider start with the entry /I = 1,
corresponding to the empty face. The f-vectors of polytopes satisfy only
one nontrivial linear equation: the Euler-Poincar formula.
246
fo fi + - - + (- 1 )d ifd_ i
=1
X(D) :=
Now if V and TY are polytopal complexes such that the union is a polytopal
complex too (that is, if F n F' e D n V' for F E 1), F' E D'), then the
Euler characteristic is additive:
Fj))
fo
c(ap)
247
complex" on the finite set V. In particular, the facets of P are (d-1)simplices, and thus correspond to d-subsets of V. The complex C is pure
(d-1)-dimensional, so it is completely determined by the family of facets
C (V1
d ). All the other faces in C correspond to the subsets of facets in Y.
Now fix a shelling order F1, F2, . . . on the facets in T. We define the
restriction R of the face Fj as the set of all vertices y E Fj such that Fi \y
is contained in one of the earlier facets:
Ri
Fi : Fi\y
Rj c C C
In fact, a face G that is new is necessarily a subset of Fj : if it misses a vertex
y E R3, then it was already contained in a previous facet, by construction.
Finally, if G satisfies Rj C G C Fj but is not new, with G C Fi for some
< j, then by the definition of shellings G is contained in some F1 (1 < j)
such that Fi n Fi = FAw is a facet of Fi . From FM C F1 we get w E R3 ,
and from R3 C G C F1 n Fj = Fj \w we get w Rj: a contradiction.
Thus every shelling gives us a partition I W
/8 of the set of faces of
the simplicial complex into intervals of the form
:= {G : Ri C G c
For a partitionable simplicial complex the f-vector can be read off from
the partition. Namely, if 1R31 = i, then there are exactly (dk :ii ) (k-1)-faces
248
=
Let
hi
---
E
j= 1
IRA)
hi (C) denote the number of parts in the partition such that the
NZ
12
13
34 35
45
36 56
h(C) =
(1,4,2).
d= 2, n = 5,
249
12 34
45 35
d1
hk + (dk+l)hk_ i + + ( k i )h i + (d)ho.
However, the f-vector also determines the h-vector: from this formula we
can recursively compute hk from fk_ i together with (h0,... , hk _ i ). Here
is one way to do the bookkeeping. We consider the f polynomial
-
fox' + fix d =
xd -i
+ h 1 xd-1 + Itoxd
E hi
i=0
From the above derivation, we see that a shelling step with IRA = j contributes a summand of (z 1)d i to the f-polynomial. Thus, we get a
formula
d
f(x) =
h(x) = f(x
1).
250
Definition 8.18.
h-vector of C is
h(C) = (ho , h i ,
, hd)
fd
iN
E")"
i=o
hk
that is,
d- i
i.o
fk-1 -
- k + 2)
ne -3
2
k 1)fk-2 +
i)ife
In particular, we have ho = 1, h1
hd
fd- 1
Id-2
fd- 3
ifo(kd
1.),k(kti).
- d, and
( -1 ) d i f
( - 1) d
251
we have f =
1
1
1
h=(
12
5
4
7
3
1
12
1
1
1
15
11
10
20
16
15
9
6
1
6
8
15
22
7
7 23
23
37
30
60
60
30
6
1
1
1
= (
1
)
In particular, the h-vector has large negative components. This cannot happen for shellable complexes, but this one is not shellable: it
is not even pure, and it is disconnected.
Why do we study h-vectors? For various problems about simplicial polytopes, h-vectors are a much more convenient and concise way to encode the
information about the face numbers than f-vectors.
252
2 fd- 2 =
dfd
l.
This is the only new equation for d < 4, but in higher dimensions there are
more (and more complicated) ones. Dehn [173] did the case d = 5, and the
general case was done by Sommerville [472]. Here is the version in terms of
the h-vector.
Theorem 8.21 (Dehn-Sommerville equations).
The h-vector of the boundary of a simplicial d-polytope satisfies
hk = hd_ k
for k = 0,1, . , d.
Proof. (by McMullen [3611) We use Lemma 8.10 and the observations in
its proof. Namely, if F1 ,
, Fs is a shelling, then its reverse Fs ,
, F1 is
a shelling as well. Furthermore, if F, comes earlier than Fi (that is, i < j)
in the first shelling, then it comes later in the reversed shelling. From this
we see that the restriction set for Fi in the reversed shelling is exactly
F3 \R: the complement of the restriction set for the original shelling. Thus
if Fi contributes "1" to hk in the original shelling (where k = IR3 I), then
it contributes "1" to hd_ k in the reversed shelling (where dk
Thus the value of hk computed by the original shelling is the same as the
value of hd_k computed by the reversed shelling.
However, by Theorem 8.19 the h-vector is independent of the shelling
LI
chosen to compute it, and hence hk = hd_k.
Example 8.22. If the vertices of the octahedron are numbered as in the
sketch,
253
0,
h(C) = (1,3,3,1),
as we had computed in Example 8.20(H). Now we reverse the shelling, to
get the shelling order
456, 345, 246, 234, 156, 135, 126, 123,
where the corresponding minimal new faces are
ha,
E(-1)k ---t
E ( ir _k_i(d - i) Ji--i,
.
(I
)fi'
k
d- k
o
d-k
fk -1 =
E(-1)d-i
i=k
( k)fi_i.
254
fk-i(P) fk-i(Cd(n))Here equality for some k with [4] <k < d implies that P is neighborly.
The first fact to note is that we can restrict our attention to simplicial
polytopes.
Lemma 8.24. (Klee [296] and McMullen [359]) The vertices of a dpolytope P can be perturbed in such a way that the resulting polytope P`
(with the same number of vertices) is simplicial, and
fk-i(P)
fk-i(P)
Ad
simplicial.
This is not the hard part, so we avoid the distraction of a proof. Thus from
now on we only consider simplicial d-polytopes: this is essential, because it
allows us to use the Dehn-Sommerville equations! What do they get us?
First, we note that we always have
f k - 1 < (n
k)
with equality if and only if P is k-neighborly. This bound is achieved with
equality for k < Lti in the case of neighborly polytopes like the cyclic
To +
To
i=0
+ - - - + 714 1
pl c
if d is odd,
if d is even.
That is, the asterisk means that we take only half of the last term for i -=
if d is even, and take the whole last term for i =
=c-Y if d is odd.
Similarly, we will use
to denote a sum where only half of the first term
is taken if the starting index of the summation is integral.
For k > j, we have with this notation
E.
fk-1
E (d
i=o
i)
- (d i\
k i hi
i=0
i hi
di
i.0
_ i)
(8.25)
(k d +
where for the last equality we have substituted d i for i, and used the
Dehn-Sommerville equations.
Looking at that, we see that what we "really have to prove" is that for
k<
the neighborly simplicial polytopes not only maximize fk_ i (as we
know), but they also maximize hk. That is, the following lemma is "more
than enough."
Lemma 8.26.
[361, Lemma 2] Let P be a simplicial d-polytope on
fo = n vertices. Then for 0 < k < d,
hk(P)
(n d
1 +k ).
Equality holds for all k with 0 < k < t if and only if t < Lc4 j and P is
1-neighborly.
256
The statement and proof of this lemma are the key steps in McMullen's
(d) ifor
solution of the upper-bound conjecture [361] (with the notation gk_
what we now call hk)-
Proof. The proof is done by induction on k. The lemma is clearly true for
k = 0, since we have defined ho to be 1. Thus it suffices to verify
(nd+k\
k k+1
nd-1+k \ 7
k
that is,
(k + 1)hk+1 < (n
d + k)hk
(8.27)
for k > O.
We get this by putting together two parts. The first one is the formula
ha/v)
(8.27a)
vEvert(C)
Et,
257
If y E Ri , then we get a minimal new face of size IRj I-1 in the link of v.
So we get a contribution to hk(C/v) only if IRA = k+1, and in this case
we get a contribution of "1" to k+1 different terms on the left-hand side
of the equation. At the same time, we get that hk +1(C) increases by 1, so
the right-hand side increases by k+1, and we are even. (The right drawing
in our sketch depicts this case.)
This proves equation (8.27a).
The second part we need is an inequality,
hk (C/v)
5_
n hk (C).
(8.27b)
vEvert(C)
For this, we prove that hk(C/y) < hk(C) holds for all n vertices y E vert(C).
To see this, take a shelling that shells the star of y first. This means that
the minimal new face in C and in the link Civ coincide at every step while
we are shelling the star. Later, after the shelling has left the star, we may
get new contributions to hk(C), but not any more to hk(C/v). With this we
get the inequality (8.27b), and putting it together with equation (8.27a)
we derive the inequality (8.27).
What about the equality case? To get hk(C/v) = hk(C), it is necessary
that in a shelling that starts with the star of y, there is no "new" face of
size at most k outside the star of v. Thus we get that, for 1 > 1, the equality
hk(C/v) = hk(C) holds for all k <1 if and only if in a shelling that starts
with the star of y, there is no minimal new face of size at most 1 outside
the star of y. Equivalently, this says that every face G with at most IGI < 1
vertices is contained in the star of y, so that Cu {v} is a face, too. Equality
in (8.27b) holds only if we have equality for all vertices y. From this we get
that equality in (8.27b), and thus in (8.27), holds for all k < 1 if and only
if C is (1-1-1)-neighborly. 0
On the way, we have also computed the f-vector of the neighborly polytopes: for this we only have to put the equality case of Lemma 8.26 into
the formula 8.25.
Corollary 8.28.
vertices, then
di
ndl+i
i
)
i=aqk i) + (lc id + i)) (
for 0 < k < d. For every k this gives the maximal number of (k-1)-faces
for a d-polytope with n vertices.
For k = d, this reduces to a formula for the number of facets of the cyclic
polytope Cd (n):
fk-1 =
.1- 2 (n d 1 + i) .
E
i
i--=o
d
258
(Compare this to Exercise 0.9!) Note that in fixed dimension, fd_ i (Cd (n))
grows like a polynomial of degree VD in the number of vertices.
Here is a brief asymptotic argument, due to Seidel [454] (see also Mulmuley [387]), for this corollary to the upper bound theorem. Namely, consider
any shelling of OP. For every facet we get that either the restriction Ri or
its complement Fi \Ri has size at most
So, either in the shelling or in
its reverse we have that Fi has a restriction of size at most Ilk and the
restriction sets in a shelling are distinct by construction. Thus the number
From
of facets is at most twice the number of k-faces of P with k <
this we get
[g].
Lc.
Lti
in n.
8.5
We have used already that the simplicial complex C with n vertices can be
identified with a set system, the collection of subsets S(C) of an n-set,
S(C) := {vert(G) : G E L(C)}.
For the following we identify the vertex set of C with the set
[n] := {1, 2, . . .
and the k-faces of the complex with the (k+1)-subsets of the ground set [n],
for -1 < k < dim(C). Thus, if C is a pure (d-1)-dimensional simplicial
complex, then it is determined by its family of facets, which is a subset
of
the collection of d-subsets of [n].
The construction behind this identifies geometric simplicial complexes,
as we get them for example as boundary complexes of simplicial polytopes,
with abstract simplicial complexes, where we only retain the information
on the vertex set, and the information about "which vertex sets correspond
to faces, respectively facets." This approach is useful for all problems that
are not concerned with the geometry of a complex, but only with its combinatorial structure. The combinatorial structure of a complex, however, is
faithfully represented by the abstract set system: from the set system data,
it is easy to reconstruct the simplicial complex (this is a process known as
geometric realization of the abstract simplicial complex).
On the next page is a "generic picture" of a simplicial complex, viewed as
a set system. (Of course, this need not be your way of viewing set systems
supply your own sketch!) The left version just shows you the "shape"
of a simplicial complex within the lattice of all subsets of [n], while the
n
ci l ),
259
right side shows you the bipartite graph of all k-faces and (k-1)-faces,
that is, of all subsets of size k+1 and k within the complex. In both cases
the minimal element is 0, which is always supposed to be contained in the
complex, corresponding to fi = 1.
Lemma 8.29.
(Sperner [473]) If C is a simplicial complex of dimension
d on fo <n vertices, then for 1 < k < d one has
n-k
k 1)
k+1
h,.
< (n - d - 1 + k)
260
hk
-=
(cl-k+ 2
(d- k + 3)
J k-3
3
_ t -i\ f.d
lk j1 J 3 -1
-
k-4
J
for 1 < j <k - 1.
(If k is odd we'll have an extra term of (kd )f__1 at the end, but that's a
constant anyway.) By Sperner's lemma, we can bound this by
d - - 1
(d- j.-1\f3 J f
Vc -3 - 1)
1 f_ k-jj
(d- j. -1\ f3 in-j
- 3 - 1)
lj+1 k-jj
-1
for j
1, ... k - 1.
Since this lower bound is monotone in j, we only need to consider this for
j = k -1, and get the condition n > (k -1) + k(d - k + 1), which we had
required to hold.
LI
Note that this lemma is false without the assumption that n is large
enough: for the complex in Example 8.20(iii) we have d = 6, fa -= n -= 12,
and
(8) (n d 1 + 3)
h3 = 60 > 56 =
3)
From this elementary lemma we get McMullen's Lemma 8.26, and thus
a proof of the upper bound theorem, for polytopes with a sufficiently large
number of vertices. This simple proof not only works for polytopes: the
argument equally applies to all kinds of simplicial complexes that satisfy
the Dehn-Sommerville equations. This includes all spherical polytopes (corresponding to simplicial fans, see Kleinschmidt & Smilansky [309]), and
even more generally, all simplicial Eulerian pseudomanifolds (see Klee [295],
Bayer & Billera [58], Chan, Jungreis & Stong [139], and Stanley [480,
Sect. 3.14]).
Corollary 8.31 (The upper bound theorem for complexes with
many vertices).
Let C be a (d-1)-dimensional simplicial complex that
satisfies the Dehn-Sommerville equations hk = hd-k-
261
fk(e) 5_ fk(Cd(n)).
k
Proof. We get hk < ( n-d-1+10
) = hk(Cd(n)) for k < Lli from Lemma
8.30, and for k > Lg-j from the Dehn-Sommerville equations. The rest
follows from the fact that the fks are positive combinations of the his with
0
i < k + I.
In this proof, you can see some of the power of the translation of geometric simplicial complexes into finite set systems (abstract simplicial
complexes). In this setting, extremal problems about simplicial complexes
are a principal topic of "extremal set theory." We will review in the rest of
this section some basic concepts, constructions, and results from this field
some of them without proofs, to save time and space. You might want
to look at the wonderful survey by Greene & Kleitman [2231 for some of
the missing details.
A basic tool of extremal set theory is the use of various partial and
linear orderings on the k-subsets of an n-set (i.e., on the (k-1)-faces of a
complex). Since we assume that the vertex set is [n], that is, the vertices are
labeled 1, 2, ..., there is a natural linear ordering ("well-ordering") on the
vertex set. With this, we can in particular talk about the largest element
rnax(G), if G is nonempty.
Using this, we define the r-lex order (or reverse lexicographic ordering)
on the k-subsets of vertices. For this we write
G - H
if and only if G H and the largest element in which G and H differ is
in H, that is, if
max(G\H) < max(H\G).
Equivalently, this means that either max(G) < max(H), or max(H) =
max(G) =: p and G\p - H\p.
In the definition of the r-lex order, the number of elements n is not
specified. Thus, we can take "-<" as a linear order on the set of all the ksubsets of N. Furthermore, for every k-subset G c N, there is only a finite
number of k-subsets of N that are smaller than G, because G >-- H implies
that H C (ink]) for n :---,- max(G). This means that we can use the r-lex order
to enumerate all the k-subsets of N, as Fi (k), F2 (k),. . .. So, we define Fi (k)
to be the jth subset in this increasing listing according to r-lex order.
262
145 -
Fi (3) - F2(3)
F3(3)
F4(3) - F5(3)
F6(3) - F7(3)
F8(3)
(ak)
with ak > ak_i > - > a2 > a i > 0.
In fact, existence and uniqueness of this expansion are easy to verify, by
choosing ak first, ak_i after that, and so on. A more systematic explanation
may be the following. Define the integers ak > ak_i > - - - > a2 > al > 0
by setting
Fn+i (k)
(Here the subscript "<" indicates that the elements are listed in increasing
order.) Then there are exactly n different k-subsets GcN that are smaller
than fal +1, , a,+1} in r-lex order. Namely,
of them have a maximal
element smaller than ak + 1;
have maximal element ak -1- 1 but the
next smallest element smaller than ak_ i + 1; and so on.
One more thing is easy to see: the (k-1)--subsets of N that are contained
in some Fi (k) with j < n +1 also have maximal element smaller than ak +1,
or they have maximal element ak + 1 but the next element is smaller than
ak_i 1, etc. so there are exactly
(a
k
i
(a
kk -11 )
ak
k 1
()
k 2
(a2)
1
(al)
(7)
and from this we see that F8(3) = {5,4,1}, consistent with our listing
above. There are 7 smaller sets in r-lex order, where 4 = (34 ) have largest
element smaller than 5; 3 = (32 ) have largest element 5 but the next element
smaller than 4; and 0 = CD have the two largest elements 5 and 4 but the
smallest element smaller than 1 (impossible). Also, there are
(93(8)
()
263
2-subsets contained in the first eight 3-sets, namely 6 = (42 ) with largest
element smaller than 5; 3 = ( I ) with largest element 5 but the next one
smaller than 4; and 1 = (g) with the elements 5 and 4. Note that this last
one is contained in F8(3), but not in a smaller 3-set.
The r-lex order is very natural in various respects. For example, it yields
a shelling order for the (k-1)-skeleton of the simplex Ad 1 for k < n = d+ 1
(Exercise 8.24(0). In fact, many other linear orders work as well. However,
it is a fascinating open problem whether skeleta of simplices are extendably
shellable; see Problem 8.24(iii)*.
Perhaps the most basic result of extremal set theory, and a principal
application of r-lex order, is the characterization of the f-vectors of simplicial complexes. It is known as the Kruskal-Katona theorem [316] [291],
although Schiitzenberger [449 ] was earlier, and even before this Harper got
close: his paper [249] does not explicitly state the theorem, but the result
is easy to derive, and I was told that Harper was aware of it at the time.
Theorem 8.32 (Krusical-Katona theorem).
Let f = (fi, fo, fi , - - - , fdi) e 14+1 be a sequence of nonnegative integers. Then the following conditions are equivalent.
(ii) The family .9f) := {Fi(k) : 0 < k < d, 1 < j _< fk _ 1 } is a simplicial
complex (that is, with every set it contains all subsets).
(iii) f_ i = 1, and fk_ i > ak (fk ) for 0 < k < d-1.
Proof. The implication (ii) >.(i) is trivial, and the equivalence (ii) 4-4 (iii)
is clear with our construction of the "boundary operator" ak(n) above.
The remaining nontrivial part is (i) >-(ii): see Greene & Kleitman [223,
ii
p. 73] for a nice and simple proof by "compression."
A simplicial complex (on a vertex set V C N) is compressed if its k-faces
form an initial segment with respect to r-lex order, for all k, that is, if it is
a complex as given by Theorem 8.32(4
The compression technique mentioned for the last proof takes as an input a simplicial complex, and outputs a compressed simplicial complex
with the same f-vector. The technique stems from a paper by Lindstriim
& Zetterstriim [336]. It works quite the same way for multicomplexes (see
Macaulay's theorem 8.34 below), and also for a generalization of both theorems, due to Clements & Lindstr6m [153} [18, Sect. 9.1].
What we really need for the following is not this theorem for simplicial
complexes, but a version for "multicomplexes." For this, we introduce some
new terminology I guess you've seen some of this before, but perhaps
with different names.
A multiset is a finite sequence of elements that may contain repeated
elements. The order of the elements is irrelevant, but their multiplicities
264
OSA
fbi,b2,
where the subscript "<" indicates that we have arranged the elements in
weakly increasing order, bi < b2 < ... < bk. The size of a multiset is the
number of elements, counting multiplicities. So the multiset IT' above has
size IFI = k, and we would call it a k-multiset. Also, a submultiset G C F
is a multiset in which every element has smaller or equal multiplicity than
in F. Finally, a multicomplex is a finite collection of multisets that is closed
under taking submultisets.
Multicomplexes can be interpreted in a variety of different ways (see
Exercise 8.22). For example, they are equivalent to order ideals in Ng and
to systems of monomials that are closed under taking factors. Our sketch
shows the "generic" drawing of what a multicomplex might look like. Note
the small diamond shape at the bottom, which denotes all the sets in the
multiset system.
One can attempt a (quite technical) topological interpretation of multicomplexes, leading to the extensive apparatus of semisimplicial sets [358]
which we avoid. There are only a few pieces of topological terminology
we use. So, the dimension of a multiset is defined to be one less than its
size, dim(F) := 1 1.11 1; the dimension of a multicomplex is the greatest
dimension of a multiset it contains; and the f -vector of a multicomplex is
U-47 fo, fi, , f4, where fi is the number of multisets of dimension un
the multicomplex.
Here is a basic bijection, which takes the k-rnultisets with elements
from [n] to the k-subsets of [n+k-11:
4)
{b1 7
bk } <
1--
(n + k 1) ,
(8.33)
265
where the symbol on the left side denotes the number of k-multisets with
elements from [n] the multiset analogue of the binomial coefficient (z).
See Exercise 8.22 for three other proofs of this.
Many set concepts are easily generalized to multiset concepts, if we just
replace binomial coefficients by their multiset counterparts*. In particular,
we need the r lex order on k-multisets. For this we write
-
P - 6
if max(P) < max(d), or if max(P) = max(d) =: p and F\73 \p, where
"P\p" means that we remove exactly one copy of the largest element from
F. So, r-lex is a linear order on the set of all k-multisubsets of N. All the
nice properties of r-lex order on sets generalize to multisets. The reason
is that under the bijection 8.33, r-lex order on k-multisets is equivalent to
r-lex order on k-sets,
-
q5 ( P
--
Thus, for every k-multisubset, there is only a finite number of smaller ones,
and thus we can use r-lex order to enumerate and label the k-multisubsets
of N, as F1(k),F2(k), .... Thus we define F3 (k) to be the jth multiset in
the listing according to r-lex order, and find that in fact it is the 0-image
of the jth subset:
15(Pi(k)) = Fi(k)For example, the r-lex order on the 3-multisubsets of N begins
P1(3) - P2( 3) .-< P3( 3) -- P4( 3) - P5( 3) - P6( 3) - P7( 3) -- PO) - - - - ,
that is,
111 - 112
122
222
22)) ((b11))
((bkk)) ((kbk--11 )) ... r
with bk > bk_i > ... > b2 > bi > 0
Pn-Fl(k) =:
266
8.
There are exactly n different k-multisets that are smaller in r-lex order
than fbi + 1, ... , bk + 11. Namely, for ea of them the maximal element
is smaller than bk 1; for ekk -3 the maximal element is bk + 1 but the
next element is smaller than bk_i + 1; and so on.
One more thing is easy to see:* the (k-1)-multisets that are contained in
some Fi (n) with j <n 1 also have maximal element smaller than bk 1,
or they have maximal element bk +1 but the next smallest element smaller
than bk_i -I- 1, and so forth so there are exactly
k
))
of them.
For an example, again let k = 3 and n = 7. We can expand
7
((g))
0+M,
and from this we see that F8(3) = {1, 3, 3}, as in the listing above. There
are 7 smaller 3-multisets in r-lex order, where (() = 4 have largest element
smaller than 3 1 22))
( = 3 have largest element 3 but the next element smaller
than 3, and M = 0 have the two largest elements equal to 3 but the
smallest element smaller than 1. Also, there are
493 (8)
= (g) + ((i )) +
((c)) = g + () + ( -01) =
3+2+1 = 6
2-multisets contained in the first eight 3-sets, namely 3 = (( 2)) with largest
element smaller than 3, 2 = ((i )) with largest element 3 but the next one
smaller than 3, and 1----- ( -01) with the elements 3 and 3. This last one is
contained in F8(3), but not in a smaller 3-set.
Our main reason of doing multisets and their r-lex ordering is to get
some intuition for what multicomplexes are, how they behave to be able
to make sense out of the following theorem. It uses a "relative" (1)d of the
0-map 8.33, which takes a naultiset {bk, ... , b 1 }>, adds 1 to each of the
elements, adjoins d-k zeroes to the multiset, and then applies the 0-map
to get a set:
'Ild({bk, ,b11>)
dk
*Do you get a dj-vu feeling? Of course, what we are doing here for multisets is
exactly the same as we did for sets before!
267
(y) The family {(1)d(i 3 (k)) : 0 < k < d, 1 < j < hk } is the set of facets
of a shellable simplicial complex with h-vector h.
Proof. Again part (ii) >(i) is trivial, while (ii)-4=P.(iii) follows from our
previous discussion.
The part (i)(ii), from multicomplexes to compressed multicomplexes,
is originally due to Macaulay. It can be proved by the "compression" technique that we have mentioned in the proof of the Kruskal-Katona Theorem 8.32.
For (ii)(v), from multicomplexes to shellable complexes, this is the
special case "s = 1" of a construction in Bjeirner, Frank! & Stanley [89],
which takes a multicomplex and produces a pure complex from it:
3 c1
-4
If is the compressed multicomplex from (ii), then the pure complex i'd(- - -)"
is shellable. In fact, in this case r-lex order defines a shelling, and the restriction set is R(4)d(Fi(k))) = 4)d(Fi (k))\{dk,... , 2, 1 } , of cardinality k.
Thus every k-multiset in the multicomplex contributes "1" to hk in the hvector of the corresponding simplicial complex. For the proof with details
we refer to [89, Sect. 5].
The implication (v)(iv) is trivial, thus we are left with proving the
direction (iv)(i), from shellable complexes to multicomplexes. For this,
Stanley [475, 476] has given an algebraic argument: the multicomplex arises
in this case from a monomial basis for "the Stanley-Reisner ring modulo
a system of parameters." Is there a simple combinatorial argument? Note
that this innocent-looking claim in particular implies that
hk < ((h i )) = (hi + k _1)
for shellable complexes, and thus this reproves the upper bound theorem
(McMullen's Lemma 8.26)! In fact, this is the key to Stanley's proof of the
upper-bound conjecture for spheres [474] [478, Sect. 11.3].
268
{ 0 , 1,2,3 , 11,12,22 }
together with the first h3 sets from the list
8.6
From the last section, I hope we gathered some intuition for "what an
M-sequence is." All kinds of interpretations are useful: so, the best is to
alternate between various explanations, between
the f-vector of a multicomplex,
the h-vector of a shellable complex, and
a sequence of nonnegative integers satisfying ak(hk
Here comes one big reason why M-sequences are useful. It yields a complete characterization of the f-vectors of simplicial d-polytopes P. What
do we know about them so far? Forgetting about Id = 1, we know they can
be encoded by their h-vectors h(P) = (ho, h 1 , ... , hd), which satisfy the
Dehn-Sommerville equations hk = hd_ k for 0 < k < d. Also, we know that
h(P) is an M-sequence from Macaulay's Theorem 8.34(iv), which implies
the upper bound inequality
hk
<
(h i +kk 1) .
In quite a daring step, McMullen in 1970 combined all the then available
information (including the lower bound theorem by Barnette, see below)
269
g(P) :=
with go := ho = 1, and gk := hk
go , gil . - - )gL4J)
Lgi-
fk-1
d
E
(dk
: i) hi
i
i=o \
=
2.0
d+1 d
=
EE
(d i
j=0 i=j
d+1
(d+1j\
= V
2_,
.i=-0
d+1k)
(d + il k)) .
( (d + 1 j ) (j
Vl +1 k)
VI +1 k))o<i<LC,
0<k<d
270
[86]
g 1--* g- Md
g E NP
I +1 with gi -= nd-1, and
M1 = (1 2),
( 1 4 6 4
M3=0
0 1 3 2
M4 =
)
10
"
1 i )7
5
1
0
(
I 0
0
10
4
1
10
6
2
5
3
1
From this we get trivialities for d < 2. For d ,---- 3 we get the f-vectors of
simplicial 3-polytopes, which are also easy to get by elementary arguments
(Exercise 8.28). However, starting at d = 4, we get nontrivial characterizations: so the f-vectors of simplicial 4-polytopes are all the row vectors f of
the following form:
f(P4)
with gi, g2 _?_. 0, 492 (g2) ,...giThe matrices Md are given explicitly and are not hard to analyze, which
allows us to study the f-vectors of simplicial d-polytopes. In particular,
one can easily (using well-known recursions, monotonicity properties, and
so forth, of binomial coefficients) verify the following simple properties.
-,
( 1
0
0
0
0
0
0
0
8 0
1 0
0 0
0 0
28-0
70
10
00
56-0
21-0
60
1 0
70-0
35-0
15-0
56-0
35-0
20-0
28-0
21-0
15-1
50
10-1
10-3
80
71
6 2
53
271
and this means that the f-vectors of simplicial 7-polytopes are exactly the
vectors of the form
(go
1 8 28 56 70 56 28 8
0 1 7 21 35 35 21 6
0 0 1 6 15 20 14 4
000 1 5972
gl 92 93)
(UBT) The f-vector f (P) = gMd has its componentwise maximum if and
only if all the components of g are maximal, with
9k = ((91))
k
(n - d 2)
( Y i + k -1 )
1
<
fd .1,
fp fp+1?
for some p, that is, the f-vector has to be unimodal. It seems that this
question was first asked by Motzkin in the late 1950s; see [801.
For this, it is not hard (but a little tedious, perhaps) to check that the
rows of Md are unimodal: they first increase, until they reach a maximum, and then they decrease again. Furthermore, the maximum occurs in
columns with indices j between j = 1. 4j and j = 3
This type of
analysis yields the crucial part of the following theorem.
u-i-A.
Theorem 8.39.
and
272
Lc
The bounds
and [(31)j are best possible in the sense that for every
p and d with LI] < p <L (304--1) i, , there is a simplicial d-polytope whose
f-vector "peaks" at p:
f- 1 < fo <
Thus, the "shape" of the face lattice of simplicial convex polytopes looks
roughly as follows (taking into account also that they are "top heavy"; see
Exercise 8.34):
k=d
k = 3d/4
,*
k =--- d/2
k=0
Eiji5rner's theorem implies the unimodality conjecture for simplicial polytopes of dimension d < 10. With more work, one can get it up to dimension 15 (Bj8rner [80, 86]), and even to dimension 19 (Eckhoff [178]). Surprisingly enough, the unimodality conjecture for simplicial polytopes is false
in dimension 20, as was first discovered by Bjbrner [80] and Lee [324, 70].
Examples 8.40. The unimodality conjecture fails for a simplicial polytope
of dimension d --.-- 20 with the following f-vector, for which fii > fi2 <113.
1
f-i =
4203045807626
fo =
,-84060916163336
fi
798578704207074
4791472253296106
h = 20363758019368323
=
65164051780016980
15
= 162910744316489788
/6
= 325834059588060117
17
= 529707205213463823
18
19
= 709935971390166248
805494832051588614
ho
Ili
= 821976324224631043
/2 =
,_-13
115
116
117
118
119
=
=
,
=
=
546761228419958342
293715859557026466
106920718330384544
23458617733909980
2345861773390998
/
\
273
n d 1+ r,
(n d 2 + k)
gk :=
for k 1.
Now take d = 20, n = 169, and r = 4203045807457, and compute (i.e., let
MAPLE compute).
The existence of the corresponding polytope follows from the (necessity
part of the) g-theorem. However, the corresponding polytopes Cd (n)<r>
are also easy to construct "by hand": see Exercise 8.32.
If we go a little higher in dimension, then the same construction produces
nonunimodal f-vectors for simplicial polytopes with much fewer vertices:
so, for d = 30, n = 47, and r = 65555 one obtains a simplicial f-vector
with only fo = 65602 vertices. However, Eckhoff [178] observed that with a
more complicated f-vector one can do even better. The simplicial f-vector
with the smallest number of vertices he found is
1
f-i =
1320
fo =
869619
24650747
342491792
h =
3070918789
h =
19918328394
f5 =
99465082767
.f6 =
h = 397591643442
= 1306188319799
.f8
3593770140180
=
h
b. = 8397239870111
fi i = 16843753477928
f12 = 29259588507633
f13 = 44370698483306
f14 = 59263421467414
f15 = 70604148959649
f16 = 76609321169592
L7 ---= 78245589858777
fis = 78245589349944
fig = 78245589350797
ho = 76598891788386
f21 = 69592677861523
f22 = 55485099387534
f23 = 37137014371927
f24 = 20144065902012
f25 = 8558343705069
f26 .--- 2730558787586
613985498319
f27 =
86678396880
f28 =
f29--,--5778559792
fi =
f2 =
274
ho =
h1 =
h2 =
1
1289
830484
18) (i;fir)
1252344784
hi
h15 =
By the (sufficiency part of) the g-Theorem 8.32, the existence of the corresponding simplicial polytope follows, if we verify that these hi form an
M-sequence, and this is easy do it!
From these f-vectors of "large" polytopes, you can perhaps get a more
realistic feeling for what f-vectors look like "in practice." Observe how the
monotonicity statements of Bjetrner's Theorem 8.39 still hold. I guess the
more general moral is that you shouldn't rely too much on intuition from
3- and 4-polytopes when you want to get a feel for the behavior of "typical"
simplicial polytopes.
Furthermore, in contrast to all the detailed (essentially complete, by the
g-Theorem 8.35!) information known about f-vectors of simplicial polytopes, we do not know much about nonsimplicial polytopes. Our knowledge
is not even complete for 4-polytopes (Problem 8.29*). In Problems 8.33*
and 8.35* we ask basic questions about the f-vectors of general polytopes.
Here we end the chapter with a construction due to Eckhoff [178]
that "easily" produces nonunimodal f-vectors in low dimensions.
Example 8.41. (Eckhoff [178])
Let P be a simplicial polytope, and P' a simple polytope, both of dimension d. Now we "cut off" one vertex from P'; then, after a projective
transformation, we can "glue" the rest of P' onto a facet of P, to obtain
the connected sum P#P'. Instead of formal details for this construction,
we just provide a sketch of a 3-dimensional connected sum.
= (1 , f,
A-1)
f (P) =
Notes
then
275
f(P#11
- this is just the sum of the f-vectors, except that 1 has been subtracted
in the dimensions -1, 0 and d - 1, corresponding to the vertex of P' and
the facet of P that were deleted in the construction.
Now if P = Cd (n) is a cyclic d-polytope with many vertices, then its
f-vector peaks in dimension 1_ 3(d4-1) _], and the f-vector of its polar peaks
at in dimension VT]. This suggests that, if d and n are large enough,
then the f-vector of Cd(n)#Cd(n) A cannot be unimodal. For example,
straightforward computations, for d = 8 and n = 25, yield
f(C8(25)#C8(25) ) =
= (1, 7149, 28800, 46800, 46400, 46400, 46800, 28800, 7149).
Similarly, for d
f(C9 (18)#C9(18) )
= (1, 1447, 6588, 12984, 15618, 15552, 15618, 12984, 6588, 1447).
These polytopes, in dimension 8, and with less than 1500 vertices, you
might even consider as "small" (if you compare them to our previous,
simplicial examples).
Notes
Sections 8.1 and 8.2. Sch1511i [437] had made a shellability assumption for
his 1852 proof of the d-dimensional Euler-Poincar formula, but did not
specify the exact condition he needed. Thus the theory of shellability got
its basis with the paper by Bruggesser & Mani [132], published in 1971, in
which they first defined the concept. Bruggesser and Mani write in their
introduction: "We were surprised to find that Schlafli's assumption can be
justified in an almost trivial manner" [132, p. 197].
Since then, shelling has become a very basic and useful technique with
many (geometric and combinatorial) applications. We refer to Danaraj &
Klee [164], Bjtirner & Wachs [94], Bji5rner et al. [92, Sect. 4.7], and in
particular to Bji5rner [83] for further reading and references.
A nonshellable triangulation of a tetrahedron (with 14 vertices, 41 facets,
all vertices on the boundary) was constructed by Rudin [433] in 1958. This
discouraged geometers from trying to prove that the boundary complex of
every polytope is shellable. Rudin's ball can even be brought into convex
position [155, p. 305], so it can be considered as a nonshellable triangulation
of a 3-polytope without new vertices.
276
Notes
277
278
also presented in Fiiredi [197] and in Ewald [189, Sect. 111.7]; their proof
is based on "shifting," a linear algebra method by Kalai which you may
find explained in more detail in Bj8rner & Kalai [91]. Shifting also leads to
farreaching extensions of the upper bound theorem (for subcomplexes of
polytopes) by Kalai [277, Sect. 9], which in turn can be applied to the diameter problem [277]. A proof of the upper-bound conjecture that is valid for
general triangulated spheres, not necessarily shellable, was found by Stanley [474, 478] (see also Hibi [252]), using the commutative algebra methods
we mentioned before. Very nice surveys are Stanley [479] and Bjiirner [82].
See Clarkson [152] for a different, combinatorial proof. Considerable very
recent progress is provided by Novik [395].
Section 8.5. Extremal set theory is an extremely interesting and very widely applicable part of combinatorics, of which we have only "scratched the
surface." We recommend the paper by Greene & Kleitman [223] and the
book by Anderson [18] for more material. See also Fiiredi [197], Engel &
Gronau [184] and Engel [185]. A spectacular recent success of extremal set
theory methods applied to a polytope is due to Kahn & Kalai [271], with
a lovely one page version by Nilli [394]).
Section 8.6. Both parts of McMullen's g-conjecture were established in
1979. In that year Billera & Lee [324, 70, 71] proved the sufficiency of McMullen's conditions (they describe an ingenious combinatorial-geometric
construction of a simplicial polytope with any prescribed M-sequence as
its g-vector). The paper [71] is highly recommended for study: it has motivated some spectacular research, notably Kalai's construction of "many
nonpolytopal spheres" [274] [330].
The necessity part of McMullen's g-conjecture (i.e., that the g-vector has
to be an M-sequence in all cases) was in the same year proved by Stanley [477, 479]. This relied on heavy machinery from algebraic geometry: the
hard Lefschetz theorem for the cohomology of projective toric varieties. (It
may be noted that the algebraic geometry tools were not complete at the
time: the only available proof turned out to be faulty. A new and even
more technical one was eventually done by Saito [436], see Stanley [481,
p. 64], Fulton [201, Sect. 5.2], and also Oda [397].) A more elementary proof
of this half of the g-theorem was long searched for, and recently given
by McMullen [369]. The new proof also uses developments (McMullen's
polytope algebra, see McMullen [367, 368] and Morelli [382, 383]) outside
the scope of this book; however, it keeps getting simpler. McMullen's paper [371] explains that it is not necessary to study the polytope algebra for
this purpose: the (much simpler) "weight algebra" will do the job. McMullen's abstract concludes: "Thus a yet easier proof of the g-theorem is now
available." See [372] for the latest "state of the art."
Note that there is still no proof that would establish McMullen's conditions for simplicial spheres, like for simplicial fans (where fi counts the
(i + 1)-dimensional cones).
Notes
279
Ei
(1) In [49], Barnette, Kleinschmidt & Lee derive an upper bound theorem
for polytope pairs polytope pairs are important because they correspond
to the case of (unbounded) polyhedra, capturing also their "combinatorial
structure at infinity" (cf. Exercise 2.19). Similarly, there is a lower bound
theorem for polytope pairs by Lee [325].
(2) It is an open problem to formulate and prove an upper bound theorem
for centrally symmetric polytopes. Here one would call a polytope centrally k-neighborly if every set of k vertices, among them no two opposite
ones, form a face. Surprisingly enough, there seems to be no straightfor-
280
281
Show that one cannot, however, prescribe the last 2-face of a shelling
in general.
Show that a set of facets of the d-cube determines a shellable
subcomplex of 8Cd if and only if it contains no facets (is empty),
or all facets, or if it contains at least one facet such that the
opposite facet is not in the complex.
Deduce that the boundary complexes of the d-cubes are extendably shellable.
(ii) Describe a shelling of the d-dimensional crosspolytope. Use it
to compute the f-vector and the h-vector of the d-dimensional
crosspolytope.
(iii) Given the h-vector of a simplicial polytope P, how can one derive
the h-vector of the bipyramid bipyr(P)?
(iv)* Are the d-dimensional crosspolytopes Cd extendably shellable?
8.1 (i)
8.2
Show that an ordering F1, F2, . . . ,F8 of the facets of a pure simplicial
complex is a shelling order if and only if for every i > 1, the facet Fi
contains a unique minimal face which is not contained in an earlier
facet Fi with j < i.
Show that the permutation F1 , ... , Ft is a shelling of A if and only if
A is partitionable with a partition such that
Ri C F3
', .
i < j.
8.3* Is every d-diagram shellable? What can you say about the case d = 3?
(For d = 2 this is true, by Exercise 8.0. If you want a guess for d > 3,
I'd vote for "no," because of the rule of thumb, "if Bruggesser-Mani
doesn't shell it, then it isn't sheilable." )
282
(ii)
Show that the d-cubes Cd have perfect shellings, for all d > 1.
(iii)
(iv)
283
8.11 Prove the upper bound theorem for simple d-polytopes with n facets.
For this, consider a linear function C E (Rd ) in general position on a
simple d-polytope P C Rd with n facets. For t E R define hk (P, t)
to be the number of vertices in Ix E P: cx < t} that are the highest
point for k different edges (as suggested by the previous exercise).
By letting t increase, show that for all t E
(dk)hk (P,t) + (k+1)hk+1 (t) Ehk (F,t) < nhk+MP,t),
F
where the sum is over all the n facets of P, and the second inequality
follows from consideration of a linear function c for which the vertices
of F are smaller than all other vertices of P. Deduce from this that
hk (P)
(it k 1\
d k )'
for d > k > d [g], and from this the upper bound theorem.
(This is the "dual proof" of the upper bound theorem 8.23, from
McMullen [361, Note added in proof].)
284
285
fk
'
( i
i=k
for k= 0, 1, . .. , d
d( di)
i.,
0+1 j\
V1+1 kj .
= (n
d 1+j) .
i=o
For this, verify that the equation is true for d = j (by induction)
and for j = 0, and then use induction on d, where the left-hand
side and the right-hand side satisfy the same simple recursion.
(See [234, p. 149], and also [222, p. 169].)
Use this to compute the h-vector for simplicial neighborly polytopes, directly from Definition 8.18.
286
8.21 Use Exercise 2.20 to show that one need not consider unbounded
polyhedra for the upper bound theorem, as follows.
For every unbounded d-polyhedron P with at least two vertices, there
exists a d-polytope with the same number of facets, but with more
vertices than P.
What happens in the cases where P has at most one vertex?
8.22 Give natural bijections between
the k-multisets with elements from [n],
the monomials of degree k in the variables x 1 , x2,. x n , and
vectors z E Ng with liz = k.
Show that under these bijections, inclusion of multisets corresponds
to divisibility of monomials and to the componentwise ordering of
vectors.
_ (+k-1)
n k . For example,
Give three more proofs of ((z)
(i) Show that every k-multiset with elements in [n] corresponds to
a sequence like**I*11***1*1 with n stars and k-1 bars, where
the number of stars between the ith bar and the (il)st bar
is the multiplicity of i in the multiset. Then count the star-bar
strings.
(ii) Use induction on n, and a basic identity for binomial coefficients.
For inspiration, see also Stanley [480, Sect. 1.2].
8.23 On the vertex set [3n] = {1, 2, , 374, consider the pure complex of
dimension d = 3n 4 generated by the n facets [3n]\{3i-2, 3i-1, 3i}
for j = 1, ... n. There are 3n minimal nonfaces: the sets of cardinality
n that contain exactly one of 3i-2, 3i-1, 3i for each j.
By comparing this complex to an (n+1)-neighborly one, show that
we have
(3:1
h,
(nI4 ) + (n 4)3n
so that for n > 6 we have hn <0, and the upper bound condition
of Lemma 8.26 is violated for h n+1 (where n + 1 < [1 .1) for a pure
complex. (Wistuba & Ziegler [530])
8.24 Show that the k-skeleta of d-polytopes are shellable polyhedral complexes.
287
(i) Show that r-lex order defines a shelling order F1 (k), F2(k) - - for the (k-1)-skeleton of the d-simplex, by directly verifying
condition 8.1(W).
(ii) Show that in fact the k-skeleta of all shellable polyhedral complexes are shellable.
Pr Is the (k-1)-skeleton of every d-simplex extendably shellable?
(This is the shelling extension conjecture, due to Simon [464,
Ch. 5]; the conjecture is known to be true for k < 3, by Bjiirner
& Eriksson [88 ] , and for k > d 1, by Kalai.)
8.25 Show the following weaker version of Macaulay's theorem, which estimates the M-sequence h = (110 , h1, . . . hd) without using the subtle
operator 8k . If hk = (1) for some real x E IR and if k > 1, then
hk_i
<
(The result is known for shellable complexes, but only with algebraic
tools, like Kalai's "algebraic shifting." Is there a fully combinatorial
rule that with every shellable complex would associate a multicomplex whose f-vector is the h-vector of the complex? Can one prove it
for pure complexes satisfying the Dehn-Sommerville equations, like
Eulerian complexes [296 ] [58 ] ?)
L
11]
\ L4J
8.28 Show that the f-vectors of general 3-polytopes are exactly the vectors
of the form
f(P3)
with gi > O.
288
every d > 4. However, for d = 4 quite a number of necessary conditions is known, and the possible pairs (fi , fi) have been characterized
for all i < j. See Bayer [55] and Bayer & Lee [59, Sect. 3.8] for further
information.
8.30 Prove Byirner's Theorem 8.39.
For the first half, you need a lemma that verifies the monotonicity of
the rows of Md, and shows that the peak lies between j =
and
1 3(d-1) 1
Lc
For the second half, you can use g-vectors of the form
g
=
such that the kth row of Md peaks at mkp . Now let gl get "very
large."
f(8P) f(8P) +
MAO
+ f(aAd i)
-
2/(Ad 1).
-
8.34 For simplicial d-polytopes, show that fk < fd-2k and fk < Li-1-k,
for 0 < k < d; 3) ]. (Byirner [82, 86])
[
8.35* Does
289
n \
fi = 2/ 1) .
290
(v)* Does every cubical dpolytope with d> 4 have an even number
of vertices?
(For even d this was shown by Blind tS6 Blind ]100].)
8.39* Is every cubical polytope rational?
8.40 The twisted lexicographic ordering on the facets of Cd(n) is defined as
follows. Consider facets F = {i n . . . ,id} and G = {j 1 , ... , jd }, and let
k be the smallest index where they differ, ik ik- Setting io = jo = 0,
define that F e G holds either if ik <2k and ik-1 = ik-1 is even, or
if ik > ik and ik-1 = .1k -1 is odd. Otherwise we set F > ' G.
-
(i)
(ii)
(iii)
(This linear ordering is from Gartner 86 Ziegler [204, Sect. 4], motivated by Klee's construction [297, Thm. 1.1] of a Hamilton cycle
(a simple closed cycle in the graph which passes through all the vertices) in the graph of Cd (n)X. Part (iii) was observed by Robert
Hebble.)
8.41* Can the cyclic polytope Cd(n) be realized in such a way that it has
a Bruggesser-Mani shelling for which every facet is adjacent to the
previous one?
Equivalently, can the polars Cd(n) A of cyclic polytopes be realized in
such a way that they have a monotone path through all the vertices?
(For example, you could try to modify the ordering of the previous
exercise so that you obtain a shelling for all d, and maintain the
adjacency property (i). A positive answer to this problem would determine M (d, n) (see Problem 3.11*) to be the number of facets of
Cd(n), by the upper bound theorem.
It has been shown in Amenta Si Ziegler [16, 17] that a 2-dimensional
projection of a polar C4 (n) of a cyclic 4-polytope has no more than
3n vertices. This is equivalent to the conclusion that for large n a
monotone path through all vertices if it exists cannot be found
by a "shadow vertex algorithm," for any realization of C4(n) A . Equivalently, a Bruggesser-Mani shelling of C4(n) cannot in general be
generated by a 2-dimensional section of C4 (n) that would cut all the
facets.)
8.42* It seems to be an open problem to show that all f-vectors of cyclic
polytopes are unimodal. Are they?
9
Fiber Polytopes, and Beyond
292
9.1
(ii) 7r(F) C 7r(F1) implies F =F'n 77-1 (7r(F)), and thus, in particular,
F C Ft.
Every pol3rtope in a 7r-induced subdivision 7r(F) arises from a unique face
F G F, and the collection .T is part of the definition of 7r(F). Thus, we
usually abuse notation and call the family of polytopes F C L(P) itself the
7r-induced subdivision.
We define a partial order on these subdivisions by setting
F1 < F2
if and only if
Q).
293
For example, consider the projection sketched before, and label the vertices and edges of P:
V2
E2
173
P
E3
V5
'i
E4
V4
44,
/
Q
-7-0
w (P, (2)
/\
1-1
Y-2
294
Definition 9.2.
C
E (RP) * . Then wc :
x i ( r(x)
cx ) is a linear map from RP to Rq x so
c determines a polytope
71.0
f (
7r (x) )
.). ---.c
tre := i.
: x E P} C illq+1
cx
which projects down to Q via the map p that deletes the last coordinate.
Let Li (QC) C L(Qc) be the family of lower faces of Q c. Then
TC .=
(R .c) _i
L4.(Qc) .
C L(P)
wcoh(P1Q)
we will denote the subposet of 7-coherent subdivisions of Q, in the (usually
larger) poset of all 7-induced subdivisions.
Note that the subdivisions {ir(F) : F E .F} that arise this way are regular
by construction. Also 7r-coherent subdivisions are 7-induced, because 7r =
p o ire : p ---+ Q. It is not true that all the regular 7-induced subdivisions
are 7r-coherent, as we will see now. However, this is true if P is a simplex
(see Exercise 9.5).
295
P
V5
Q
and it is regular (all subdivisions of a 1-dimensional polytope are regular).
Q
However, this subdivision of Q is not it-coherent. In fact, if Q is any such
polytope whose lower faces yield the subdivision of Q, then there cannot
be a linear map rc : P 4 QC -f2---'
as required by Definition 9.2. In fact,
such a map would have to take yi to wi in the following sketch.
W5
Q
'
./
2 ---"*",.....-- W4
W3
However, this is impossible for an affine map, since the line [y 1 , /75] intersects the triangle [2,2 , v3, v4] in P, and there is no such intersection of
the triangle [w 2 , 10 3 ,104] with the line [w 1 , 105 ] in Q.
296
For a general projection of polytopes, the structure of the poset of all subdivisions is not clear. The "generalized Baues problem" of Billera, Kapranov
& Sturmfels [69] asked whether this poset always has the "homotopy type
of a (p q 1) sphere." For this Rambau & Ziegler [417] gave explicit counterexamples: For example, there is a projection of a 5-dimensional polytope
(simplicial, 2-neighborly, 10 vertices, 42 facets) to a hexagon such that the
poset of all subdivisions is disconnected. Nevertheless, Edelman & Reiner
[179] have proved that the generalized Baues conjecture is true in the case of
a (general position) projection of a simplex into the plane. The important
cases of general projections of d-simplices (related to spaces of triangulations) and the projections of d-cubes (related to zonotopal tilings and
oriented matroids, see Chapter 7) are still wide open and very tantalizing.
(See also Bj6rner [84], Sturmfels [497], and Mnv & Ziegler [384)
The figure on the next page shows the poset of all subdivisions for the
projection of Example 9.3.
We met a special case of this before: when P = Cp is a p-cube and Q is
thus a zonotope, then the set of all zonotopal tilings (those are the tilings of
Z = Q by faces of P = P) is the poset of all one-element oriented matroid
extensions, which appears in the Bohne-Dress Theorem 7.32 (Section 7.5).
In contrast to the set of all subdivisions, the poset of all 7r coherent
subdivisions is the face poset of a polytope: of the "fiber polytope" of the
projection . In the drawing on the next page, this is the part of the poset
drawn with solid lines the face poset of a hexagon (Exercise 9.1)!
-
f 1
'y(z)da : -y is a section of irl.
1 vol(Q) 14
R f (x) dx = vol(RH(ro),
00
297
298
The scaling factor 1/vol(Q) in the definition is only needed for this inclusion, but is irrelevant for the geometry of the fiber polytope.
proper subdivisions.
In particular, in the special case where P = Ap is a simplex we get that
the vertices of E(A p , Q) correspond to triangulations of Q this is the
case of secondary polytopes as considered in [213, 214]; see the next section.
Proof. (Sketch)
299
9.2
Some Examples
Some special cases of the fiber polytope construction are easily analyzed.
If Q = tql is a point (q = 0), then E(P, Q) = E(P,{q}) = P.
More generally, if P=RxQ is a product, and 71" : R x Q Q is the
canonical projection, then it is easy to see that E(P, Q) = E(RxQ,Q) R:
the fiber polytope is a translate of R. In fact, we get
Ir
:10
The lower end 51 of the interval arises as the integral of the collection
1
of "lower faces" of P, while st is the integral of the collection of "upper
faces," divided by vol(Q) in both cases.
300
These were trivial cases we will now do two more interesting ones.
We start with the permutahedron, and the (more general) "monotone path
polytopes." Then we will construct the associahedron, as the "secondary
polytope" of an n-gon. So both examples are special cases of important
constructions.
Let P C RP be a p-dimensional polytope, and consider
a nonzero linear function a E (RP) * on P. This defines a projection
Definition 9.7.
P --> Q := fax : x E P} C
to the 1-dimensional polytope Q [a min , am}, where
amin = min ax,
-= max ax.
XEP
XEP
II(P, a)
E(P, fax : x E
V1
Vn- 1
1.7n)
<
CV'
defines a section
P,
t-v i_ i +
for a: =
74' Q
Every such section 745 defines a point in the monotone path polytope
II(P, a), namely the integral
vol(Q)
70(x)dx =
(avi avo)
amax amin
vo
+ vi
2
(aVn aVn- 1 )
Vn- Vn
301
xi .
i=1
e l e2 e3 = 1
] n
cony ({
( z2
: j = 0, 1, 2,
, n} u { (
n2 (i n)2
) : j = 0, 1, 2, .
n}.
302
( 1
ecr(i)
()
(:::11) 2 )
2i 1)
which maps ea ( i) +
i2
(i
1)x k
:
n1
[0, 1[n that minimizes the integral
The (unique) section
fon ccry(x)dx describes a path in the 1-skeleton of [0, 1 ] :
"Ya : 0 --)
Ca
..
1-
el + e2
el
e2 + e3 = 1
303
Note that this is a special case of our discussion and computation after
Definition 9.7. Here the integral of 7' is given by the sum
ion
7' (x) dx
+ WM + 7' (2)) + - - -
1
1(2nl)e,( 1 ) + (2n-3)ecr(2) + ... + (1)e(,))
1 1.-.
/
= 2(2n 1 2i)e,( i)
2n + 1
1(
2
u ;(n)
11([0, l ] n , Il) =
&k
-
cony
n-k
n-k
ir (x)
( E in= 1 x i )
EiEil X i
=
(
7(x)
cAx).
304
el + e2 + e3 = 1
E([0, l] n , [ 0, nD = fx E
n
: Ilx =
2,
E x, <
iEA
IAI(2n IA1)
for 0 c A c [n]}.
2n
where we set e0 := O.
E(Q) := (d + 1) vol(Q)
v}.
Q),
305
Ain
Q
For every n-simplex P C
there is a projection map 7r :
that maps the vertices of Ain to the vertices of Q. Furthermore, P and
the map 7r are unique up to affine coordinate changes in n . So, for every
projection of an n-simplex P to Q, the fiber polytope E(P, Q) is affinely isomorphic to the secondary polytope E(Q). (Equivalently, one could also use
our "standard" n-simplex An c Rn+1.) Thus the secondary polytope is a
canonical object associated with any polytope Q. We refer to the papers by
Gel'fand, Zelevinsky & Kapranov [214], Billera, Gerfand & Sturmfels [68],
and Billera, Filliman & Sturmfels [67] for extensive discussions. The key
observation is that, by Theorem 9.6 together with Exercise 9.4, once the
map 7r : Ain Q is fixed, every regular subdivision of Q (without new
vertices) is 7r-coherent.
Corollary 9.10. The vertices of E(Q) are in bijection with the regular
triangulations of Q, via
(ei. +
+ eid ),
[v 0 ,... y id ] E T
where every regular triangulation of Q is represented by its collection of
d-dimensional simplices.
To get the associahedron from this construction, we use a "well-known"
bijective correspondence between the complete bracketings of a string of
n letters and the triangulations T without new vertices of the (n + 1)-gon
C2 (n + 1). So, to every complete bracketing a of the string 123 ... n of
length n we associate the corresponding triangulation of C2 (n 1), written
as a set of triples T(a) C ( 10' 3 1 ). We denote by En the set of all these
triangulations. For example, we get (using square brackets for the triples)
T2 =
=
306
T =-- 1[014],
[123],
[1341}
4(2.3)4)
C2(n + 1)
for vi
.
2) (z
vol[vi,vi ,vkl
-21- (j
(k
i)(k
j)
- (ei + ei
ek).
[i,j,kJ E T
Here the sum is over all triples i < j < k such that [71- (ei),Ir(ei),71 - (ek)1 is
a triangle in the triangulation T, of area 1(j (k i)(k j).
Taking this together with the above bijection, this defines a point in Rn
for every complete bracketing of 12 ... n. For example, with the bracketing
307
1
4 I.
3
0
\ 3
\ 9
( 66g
Our figure shows the corresponding section .1T, whose integral yields vT .
e1
V4
Vo
Vi
Vn := vol(C2 (n 1 ))
nEi(i 1)
2
i=2
(n + 1)
3 P
(dgi
cn) :,_ q 0 (C2(n-l- 1 )) (
3
)
(6n2 + 1)/
for n > 2.
308
From this we get that E(C2 (n + 1)) is contained in the affine subspace
of Rn given by the equations
n
ix,
3 vol(C2(n 1)) cn =
E i2xi .
3 vol(C2 (n + 1)) dn
n4
n2
4
6 n5 5n3 n
30
How do we get the facet-defining inequalities? We use the method outlined after Theorem 9.6. The facets correspond to the diagonals of C2(n+1),
which we interpret as going from yi to yi, for 0 < i < j < n, with
2 G j i < n. For every such "admissible" pair (i, j), we construct a
regular function
fi3
x
) := max{0, y + (i Ax ij}.
This formula can be derived from the condition that the linear function
and for (x) = ( ). The comey + (i + j)x ij vanishes for ()
Y
Y =()
/2
45 :,
ek 1--->
f i3
(kk2 ) .
*,
=
L-d fii (kk2) Xk
,1
kn
=E
max{0, k2 + (i + Ak ii} xk
k=1
n
=E
k=1
=E
(k i)(j k)Xle-
k=i
309
fj
min{eivT : T E 7-;,} =
i +1\3(j i) 2 -2
3
10
E(C2 (n + 1))
Kn_2 ,
xi
-2
i= 1
Xi
6n5 5n3 n
30
i 1) 3(j
3
i)2 2
10
xl
x2
x3
2x1+2x2
2x2 +2x3
>= 1
>= 1
>= 1
>= 10
>= 10
END
where the inequalities correspond to the diagonals [02], [13], [24], [031, and
[14] (in this order). PORTA requires knowing a valid point for this system,
so we give it the point that we had computed before.
Now the PORTA command traf v ass2. le q produces from this the
list of vertices and the incidence matrix for a pentagon, in a file named
ass2. ieq. poi:
-
310
DIM = 4
CONV_SECTION
( 1) 1 4 9 6
( 2) 4 1 10 6
( 3) 9 4 1 10
( 4) 10 1 4 9
( 5) 1 10 1 9
END
Il
N\
T\
S\
1
*..*. :
.*.* . :
..*.* :
.*..* :
*.*.. :
2
3
4
5
2
2
2
2
2
I 22222
9.3
...
for
fi := el
.- ei,
f 0 = 0,
311
Proposition 9.12.
7r : Aif, p C2(n+1),
fi i---4
(t),
which maps e a 1-4 (22 1 1 ) for i > 1. Its scaled fiber polytope
K4_ 2
:=3
(n-F 1)
3
f
.E(A n , C2(n 1 ))
v T :=
[i,j, ic]
(fi -I- fi
fk)
E Zn,
for all triangulations T of C2 (71 1) without new vertices. Here the sum
is over all triples i <j < k such that (71-(fi ),7r(fi ),r(f k )) is a triangle in
the triangulation T, of area 1(j i)(k j) (k j).
Furthermore, all vertices lie on a sphere around the origin:
n
E(VT) 2
(n 1) 30n4 33n2 2
3
i=1
70
n
E
xi
(2i 1) - xi
i=1
n
i=1
(n 1) n
\3)22
-,. =
n2 (n2 1)
4
3 (n 1) 6n2 + 1
6n5 5n3 n
1
=
30
3 ) 15
which describe the (n-2)-subspace of Rn that contains K_27 and facetdefining inequalities
((-2k + 1) i
:=
k=i+1
>
( j i + 1) 31j
3 )
02
10
for0<i<j<nand2 <ji<n-1.
This is just what we worked out in Example 9.11, after the linear transformation in Rn that sends ei 1-4 fi for 0 < i < n (that is, e0 = 0 to
La = 0)The magical little thing is that the vertices lie on a sphere. One can
prove this by analyzing the situation along an edge, corresponding to a
single rebracketingichange-of-diagonal, but there is no really good (that is,
geometric) explanation. Do you have any ideas? The strange thing is that
312
3 4.1 5 7.6.2 .
One can show quite easily that KlIn_i is in fact a graded, atomic, and
coatomic lattice of length n. Thus it "looks like" the face lattice of an
(n - 1)-polytope. The coatoms ("facets") of KII... 1 are the ordered partitions of {1, ... , n } , without brackets. The atoms ("vertices") correspond to
complete parenthesizations of permutations of the letters 1, 2, ... , n. The
edges are of two types: they correspond either to a single reparenthesization, or to a transposition of two adjacent letters that are grouped together.
For n = 3 we get the face lattice of a 12-gon, as follows:
1.2.3
1.23
1.3.2
13.2
K II2 :
3.1.2
3.12
313
2(3.1)
Now every permutation a = o- (1)o- (2) .. . a- (n) determines a simplex in
namely
An (a)
:=
n:1
X0 (1)
X012)
1.
[0, n]
Rn
L171:, = convff 1 ,
,f
The description of the simplices dn (u) in terms of their inequality systems also shows that they fit nicely together to form a triangulation of the
unit cube [0, 11n.
314
315
e2 + e3
el
=1
231
Each of these simplices has a natural map down to R2 , where we map
--{ ec (i )
- -
1--*
(! 7
z2
Since the simplices fit together so nicely to form a triangulation of the cube
[0,1r, and the projection maps are defined consistently on the vertices, we
obtain a continuous, but nonlinear "folding map"
H: [0, 1r --> C2 (n 1),
which is linear on the simplices
(o-). Furthermore, for every bracketed
permutation, there is an obvious section to this folding map! For this we
define the section on the vertices by
i
Vi = (.2 ) t--+ ecr(1) -F + ecr(i)
Z
associahedron.
316
Our figure illustrates the section -y" : C2(4) > [O, 11 3 that is associated
to the bracketed permutation a = 2(3.1) by this method.
e2 + e 3 + e l = 1
2(3.1)
Another way to view this construction is the following. The vertices of
the "special associahedron" IC4_2 in Proposition 9.12 satisfy y 1 > y2 >
... > un : in fact, the little associahedron
1
n-2
3 1)Kf
3 (n+
317
Ci
+ e2 e3 = 1
The only change we do for the formulas is that instead of the average
integral we take three times the integral, that is, we blow the polytope up
by a factor
) = 3vol(C2(n -I- 1)), in order to get integral coordinates.
WI 1
Theorem 9.15.
va
The formula
3-y'dx dy
Jc2(n+i)
1
--.
0 -0(k i)(k
:7) - (f
-()
f cr(i) f cr(k))
(z,j,k)ET(a)
Ex, =
i
n4 n2 ,
4
/.
318
= (n + 1) 30n4 33n2 + 2
k 3)
70
Proof. A proof with details is in [419], and we refer to the treatment there.
What is the idea? First the associahedron in Proposition 9.12 lies in the
n : 11 x ....= n 4 _4 n 2 , .
hyperplane H =-- Ix E
} Since the vertices of KlIn_ i can
be generated by permuting the coordinates of vertices of Krfi _ 2 , we get that
our polytope KlIn_ i is also contained in H.
Then we obtain the facet-defining inequalities. That is, to every ordered
partition 0 of [n] we associate a linear function c0 . For this let 0 have p
blocks, and write it as
(15 = a(ii)
- - (7 (ii) Gr(i2)
u(ip),
where the numbers ir and jr just tell us about the "block structure" of 0,
with
)ip-1
+1 = ip
ip =
n.
ct = ir + jr
9.4
319
fA,Z (p) c Q}
(0
QP+'.
(ii) In particular, take the (d- *simplex 2d1 = conv{e i ,..., ed} c Rd
on the hyperplane H := Ix E Rd lix = 11, as before. Then we
H with linear maps Rd 4 Rd
can identify affine maps with H
that fix O. Such maps are given by a matrix V = (v 1 , , vd) E Rdxd ,
where ei Thus the mapping polytope 4)(A d_ 1 , Ad- 1 ) turns
out to be
1) (Ad-1,Ad-1)
f (vii ) E Rctxd
< d,
320
(iii) Now we will consider the subset of all maps that preserve barycenters,
that is, map the barycenter 1 of 'd-i to itself. But the map takes
this to ti (vi + + vi), so we get extra conditions
d
= 1,
1 (Ad),
1 )
)
{( vii ) E Rd x d
v- >_o,
Ei vii =
1 for all
Notes
The original motivation for the constructions of secondary polytopes and of
fiber polytopes did not come from polytope theory, but from the theory of
A-hypergeometric functions [210, 212], and from state polytopes in commutative algebra [54] [498]. It turns out that also there are strong connections
to constructions in algebraic geometry [287] and elimination theory [288]. A
survey was given by Loeser [342] in the "Bourbaki seminar." We especially
321
recommend the recent book by Gelfand, Kapranov & Zelevinsky [212] for
study.
Our definitions (starting with the Definition 9.1 of 7r-induced subdivisions) are different in appearance from, but equivalent to, the original setup by Billera & Sturmfels [74], which works with vertex sets rather than
polytopes. Only the explicit version of 7r-coherent subdivisions in Definition 9.2 might be new here. We also refer to the survey of fiber polytopes
in [497], and to the alternative set-up (via normal fans) in [75].
Similarly, the original definition of secondary polytopes by GePfand,
Zelevinsky & Kapranov [213, 214] (see [212, Ch. 7]!) was quite different
from the one, due to Billera & Sturmfels [74], which we have presented in
Definition 9.9. Our presentation also reverses the historical order of things:
the ingenious construction of [213] now appears as a very special case of
the fiber polytope construction; let us just say that the secondary polytopes are in many respects the most fundamental case. In particular, every
fiber polytope can be written as a projection of a secondary polytope; see
Exercise 9.6.
The "permuto-associahedron" Kri n_ i is a combinatorial object introduced by Kapranov [286] (he denotes it as K/3,). The construction as
a polytope, and the generalization to "Coxeter-associahedra," by Reiner
& Ziegler [419], were born in December 1992, so to speak (and not baptized). The nonlinear effects appearing in this, like the sphericity in Proposition 9.12, suggest that there is much more to be discovered and that the
constructions are not yet well understood.
Compute the fiber polytope for the projection in Example 9.3. For
this, compute both the vertex coordinates, corresponding to the six
coherent tight sections.
(According to our discussion, you should get a hexagon!)
Also, compute the point in R3 which corresponds to the noncoherent
tight section of Example 9.3. Does it lie in the relative interior of the
fiber polytope?
Finally, use the methods described in this chapter to derive a description of the fiber polytope by equations and inequalities.
322
f
for a,
Xdl,
(ii) Describe Newton(det(xii)), for the determinant of an (n x n)matrix, with d = n2 different variables as entries.
9.5 If 7r : P
Q is a projection of polytopes, then the 7r-induced
subdivisions of Q only use vertices in the finite set 7r(vert(P)), which
need not all be vertices of Q.
Show that if P = Ap is a simplex, then all the regular 7r-induced
subdivisions of Q are 7r-coherent.
323
What about the volume and the barycenter of the general cyclic polytopes Cd(n+1) ?
9.8 Let P and Q be polygons in the plane. Show that the fiber polytope
of the projection
PxQ ---* PQ
from the product to the Minkowski sum has a fiber polytope that is
isomorphic to P + (--Q).
What goes wrong here if Q degenerates to a line segment?
324
is a zonotope as well.
(Billera & Sturmfels [74, Thm. 6.11)
Compute the fiber polytopes for the projections
and
C6
P12
2
'
+ Q2
) (P1Q 1)
show that
+ (PS22)-
References
7397-city traveling salesman instance solved, Optima No. 45, March 1995, 7 -8. (23)
Si
JOHN E. WETZEL:
A simphcial
326
References
simplicial
4-arrangement of 33 planes, Geometriae Dedicata 24 (1987), 245254. (225)
[10]
ALEXANDER
DANILOWITSCH ALEXANDROW:
Konvexe Po/yeder,
A simple proof of the upper bound theorem, European J. Math. 6 (1985), 211-214. (277)
M.
Shadows and slices of polytopes, in: Proceedings of the 12th Annual ACM Symposium on Computational Geometry, ACM Press 1996, 10-19. (290)
ZIEGLER:
Combinatorics of Finite Sets, Oxford Science Publications, Clarendon Press, Oxford University Press, Oxford 1987.
(263, 278)
Ein Algorithmusansatz fuir Realisierungsfragen im Ed getestet an kombinatorisehen 3-Sphdren, Staatsexamensarbeit, Universita Bochum 1982. (212)
References
[21]
327
[24] DAVID AVIs: A C implementation of the reverse search vertex enumeration algorithm, preprint, McGill University 1992; report and
code available by anonymous ftp from mutt .cs.mcgill ca, directory "pub/C". (49)
.
[25] DAVID AVIS & DAVID BREMNER: How good are convex hull algo-
[27]
Lows J.
&
WALTER KERN:
328
References
[34] IMRE BARANY 86 I.ASZIA5 LOVSZ: Borsuk's theorem and the number
Diagrams and Schlegel diagrams, in: "Combinatorial Structures and Their Applications" (Proc. Calgary Internat. Conference, Calgary 1969), Gordon and Breach, New York 1970,
pp. 1-4. (143)
A 2-manifold of genus 8 without the Wvproperty, Geometriae Dedicata 46 (1993), 211 214. (85)
References
329
An upper bound theorem for polytope pairs, Math. Operations Research 11 (1986), 451-464. (279)
Hamiltonian circuits
in certain prisms, Discrete Math. 5 (1973), 389-394. (100)
[51] HANS G. BARTELS: A priori Informationen zur Linearen Programmierung. Ober Ecken und Hyperfliichen auf Polyedern, Beitrage zur
Datenverarbeitung und Unternehmensforschung, Band 4, Verlag Anton Hain, Meisenheim am Glan 1973. (124)
[52] GOTTFRIED BARTHEL, FRIEDRICH HIRZEBRUCH & THOMAS HoFER, Geradenkonfigurationen und Algebraische Fleichen, Aspekte der
Mathematik D4, Vieweg, Wiesbaden 1987. (225)
[53] ALEXANDER
[54] DAVID BAYER & IAN MOFtRISON: Standard bases and geometric in-
variant theory, I. Initial ideals and state polytopes, J. Symbolic Computation 6 (1988), 209 217. (320)
-
[57] MARGARET
M. BAYER & LOUIS J. BILLERA: Generalized DehnSommerville relations for polytopes, spheres and Eulerian partially
ordered sets, Inventiones Math. 79 (1985), 143-157. (260, 280, 287)
[58] MARGARET
[59] MARGARET
[61] EBERHARD BECKER: On the real spectrum of a ring and its application to semialgebraic geometry, Bulletin Amer. Math. Soc. 15 (1986),
19-60. (119)
330
References
[62]
[63]
M.
Bander
und Maiusbander in konvexen Polytopen, Abh. Math. Sem. Univ.
Hamburg 44 (1975), 249-262. (148)
WILLS:
[65] Louis J. BILLERA: Homology of smooth splines: generic triangulations and a conjecture of Strang, Transactions Amer. Math. Soc. 310
(1988), 325-340. (129)
[66] Louis J. BILLERA, CLARA CHAN Az NIANDONG Liu: Flag complexes, labelled rooted trees, and star shellings, preprint 1996. (280)
[67] Louis J. BILLERA, PAUL FILLIMAN Sz BERND STURMFELS: Constructions and complexity of secondary polytopes, Advances in Math.
83 (1990), 155-179. (305)
[68] Louis J. BILLERA, IZRAIL M. GEL'FAND & BERND STURMFELS:
Duality and minors of secondary polyhedra, J. Combinatorial Theory,
Ser. B 57 (1993), 258-268. (305)
[69] Louis J. BILLERA, MIKHAIL M. KAPRANOV & BERND STURMFELS:
Cellular strings on polytopes, Proc. Amer. Math. Soc. 122 (1994),
549-555. (225, 296)
[70] Louis J. BILLERA & CARL W. LEE: Sufficiency of McMullen's conditions for f -vectors of simplicial polytopes, Bulletin Amer. Math.
Soc. 2 (1980), 181-185. (269, 272, 278)
[71] Louis J. BILLERA & CARL W. LEE: A proof of the sufficiency of
McMullen's conditions for f -vectors of simplicial polytopes, J. Combinatorial Theory, Ser. A 31 (1981), 237-255. (269, 278, 279)
[72] Louis J. BILLERA & A. SARANGARAJAN: All 0-1 polytopes are traveling salesman polytopes, Combinatorica 16 (1996), 175-188. (21, 26,
70)
[73] Louis J. BILLERA Si BETH SPELLMAN MUNSON: Polarity and inner
products in oriented matroids, European J. Combinatorics 5 (1984),
293-308. (184)
[74] Louis J. BILLERA Si BERND STURMFELS: Fiber polytopes, Annals
of Math. 135 (1992), 527-549. (19, 291, 298, 301, 321, 323, 324)
[75] Louis J. BILLERA Si BERND STURMFELS: Iterated fiber polytopes,
Mathematika 41 (1994), 348-363. (19, 291, 294, 321)
References
331
natorics" (R. Graham, M. Griitschel, and L. LovAsz, eds.), NorthHolland/Elsevier, Amsterdam 1995, pp. 1819-1872. (59, 232)
[86] ANDERS B.JaRNER: Partial unimodality for f -vectors of simplicial
number of faces of balanced Cohen-Macaulay complexes and a generalized Macaulay theorem, Combinatorica 7 (1987), 23-34. (267, 268,
287)
332
References
[90] ANDERS BJ45RNER, ADRIANO M. GARSIA Sc RICHARD P. STANLEY: An introduction to Cohen-Macaulay partially ordered sets, in:
"Ordered Sets" (I. Rival, ed.), D. Reidel, Dordrecht 1982, pp. 583615. (247)
[91] ANDERS B.16RNER Sc GIL KALAI:
clopedia of Mathematics, Vol. 46, Cambridge University Press, Cambridge 1993. (18, 59, 96, 142, Lectures 6 and 7, 233, 275, 277, 291)
[93] ANDERS BJ6RNER Si SVANTE LINUSSON: The number of
k-faces of
Sc
than 2d+1 faces, in: "The Litsz16 Fejes T6th Festschrift" (I. BdrAny,
J. Pach, eds.), Discrete Comput. Geometry 13 (1995), 321-345. (280)
[102] ROSWITHA BLIND Sc PETER MANI-LEVITSKA: On puzzles
and poly-
References
[104]
333
Eine kombinatorische Analyse zonotopaler Raumaufteilungen, Ph.D. dissertation, Bielefeld 1992; Preprint 92-041, Sonderforschungsbereich 343 "Diskrete Strukturen in der Mathematik,"
Universitat Bielefeld 1992, 100 pages. (220, 222, 225)
JOCHEN BOHNE:
Si
PETER KLEINSCHMIDT:
Simplicial convex 4-polytopes do not have the isotopy property, Port ugaliae
Math. 47 (1990), 309-318. (177)
JRGEN BOKOWSKI & ANTONIO GUEDES DE OLIVEIRA:
[110]
[111]
BERND STURMFELS:
Computational Syn-
thetic Geometry, Lecture Notes in Mathematics 1355, SpringerVerlag, Berlin Heidelberg 1989. (144, 179, 182, 183, 212)
[112] ETHAN D. BOLKER:
BLA BOLLOBS:
AAGE BONDESEN
102. (277)
[116] JOHN ADRIAN BONDY
& U. S. R.
MURTY:
334
References
Springer-Verlag, Berlin 1934; berichtigter Reprint, Springer-Verlag,
Berlin 1974; English Translation: Theory of Convex Bodies, BCS Associates Pub., Moscow, Idaho, 1987 (22)
[121]
NICOLAS BOURBAKI:
WALTER
BRAGGER:
[124]
[125]
R. BRIGHTWELL & EDWARD R. SCHEINERMAN: Representations of planar graphs, SIAM J. Discrete Math. 6 (1993), 214-229.
(118)
DAVID BREMNER:
GRAHAM
GEORGE MAXWELL:
(9, 225)
[129]
[130]
RICHARD
335
References
cells and spheres, Math. Scand. 29 (1971), 197-205. (233, 235, 240,
275)
[133] EWALD BURGER: Ober homogene lineare Ungleichungssysteme, Zeit-
metric convex polytopes having many vertices, J. Combinatorial Theory, Ser. A 58 (1991), 321 322. (280)
-
[139] CLARA CHAN, DOUGLAS JUNGREIS & RICHARD STONG: Buchsbaum and Eulerian complexes, J. Pure Appl. Algebra, to appear.
(260)
[140] BERNARD CHAZELLE: An optimal convex hull algorithm in any fixed
non-negative solutions of a system of linear equations, U.S.S.R. Computational Mathematics and Mathematical Physics 4 (1964), 151
158. (48)
336
References
THOMAS CHRISTOF:
A com-
plete description of the traveling salesman polytope on 8 nodes, Operations Research Letters 10 (1991), 497-500. (49)
[147]
MAREK CHROBAK, MICHAEL T. GOODRICH & ROBERTO TAMASSIA: Convex drawings of graphs in two and three dimensions, in: "12th
ACM Symp. Computational Geometry", ACM Press 1996, pp. 319328. (123)
[150]
Convex grid drawings of 3connected planar graphs, Preprint 1993; Int. J. Comput. Geometry
and Applications, to appear. (123)
[151]
pour les empilements de cercles, Inventiones Math. 104 (1991), 655-669. (118)
[155] ROBERT CONNELLY & DAVID W. HENDERSON: A convex 3-complex
not simplicially isomorphic to a strictly convex complex, Math. Proc.
Cambridge Phil. Soc. 88 (1980), 299-306. (275)
[154]
References
337
AKOS
[163] GOPAL DANARAJ Si VICTOR KLEE: Shellings of spheres and polytopes, Duke Math. J. 41 (1974), 443-451. (233, 240)
[164] GOPAL DANARAJ Si VICTOR KLEE: Which spheres are shellable?
in: "Algorithmic Aspects of Combinatorics" (B. Alspach et al., eds.),
Annals of Discrete Math. 2, 1978, pp. 33-52. (235, 275, 277)
[165] GOPAL DANARAJ Si VICTOR KLEE: A representation of 2-dimensional manifolds and its use in the design of a linear-time shelling
algorithm, Annals of Discrete Math. 2 (1978), 33-52. (281)
[166] GEORGE B. DANTZIG: Linear Programming and Extensions, Princeton University Press, Princeton 1963. (81, 83, 96)
[167] GEORGE B. DANTZIG Si B. CURTIS EAVES: Fourier-Motzkin elimination and its dual, J. Combinatorial Theory, Ser. A 14 (1973),
338
References
Academic
[172] CHANDLER DAVIS: The set of non-linearity of a convex piecewiselinear function, Scripta Math. 24 (1959), 219-228. (143)
[173] MAx DEHN: Die Eulersche Formel im Zusammenhang mit dem Inhalt in der nicht-Euklidischen Geometrie, Mathematische Annalen
61 (1905), 561-568. (252)
[174] A. K. DEWDNEY 86 JOHN K. VRANCH: A convex partition of
with application to Crum 's problem and Knuth 's post-office problem,
Utilitas Math. 12 (1972), 193-199. (144)
[175) MICHEL MARIE DEZA AL MONIQUE LAURENT: Geometry of Cuts and
Metrics, Algorithms and Combinatorics 15, Springer-Verlag, Berlin
Heidelberg 1997. (23)
[176] TOMASZ DUBEJKO & KENNETH STEPHENSON: Circle packing: Experiments in discrete analytic function theory, Experimental Math.
4 (1995), 307-348. (118)
[171
[181] JACK EDMONDS Sz ARNALDO MANDEL: Topology of oriented matroids, Ph.D. thesis of A. Mandel, University of Waterloo 1982, 333
pages. (216)
[182] FRITZ EHLERS: Eine Klasse komplexer Mannigfaltigkeiten und die
References
339
[184]
[185]
GRONAU:
KONRAD ENGEL:
[186] G. V. EPIFANOV: Reduction of a plane graph to an edge by a startriangle transformation, Soviet Math. Doklady 7 (1966), 13-17. (109)
[187]
Lattice Point
Problems, Pitman Monographs and Surveys in Pure and Applied
Mathematics 39, Longman, Essex, and John Wiley Si Sons, New
York 1989. (23)
[188]
HERVA LE VERGE: Complete linear descriptions of small asymmetric traveling salesman polytopes, Discrete Applied Math. 62 (1995), 193-208. (49)
[189]
GUNTER EWALD:
[190]
[191]
GUNTER
REINHARDT EULER Si
[192]
Stellar subdivisions
of boundary complexes of convex polytopes, Mathematische Annalen
210 (1974), 7-16. (97)
JIM LAWRENCE: Oriented matroids, J. Combinatorial Theory, Ser. B 25 (1978), 199-236. (149, 211, 216)
JON FOLKMAN Si
How
340
References
[198] KOMEI FUKUDA: CDD A C implementation of the double description method, available by anonymous ftp from ftp . epf 1 . ch, directory "incoming/dma". (48)
-
[200] KOMEI FUKUDA, AKIHISA TAMURA & TAKESHI TOKUYAMA: A theorem on the average number of subfaces in arrangements and oriented
matroids, Geometriae Dedicata 47 (1993), 129 142. (230)
-
[208] MICHAEL R. GAREY Si DAVID S. JOHNSON: Computers and Intractability. A Guide to the Theory of NP-Completeness, W. H. Freeman, San Francisco 1979. (125, 126)
[209] LYNN E. GARNER: An Outline of Projective Geometry, North Holland, New York 1981. (69)
References
341
Math. Doklady 40 (1990), 278-281. (19, 291, 298, 305, 306, 321)
[214] IZRAIL M. GELTAND, ANDREI V. ZELEVINSKIT Sz MIKHAIL M.
KAPRANOV: Discriminants of polynomials in several variables and
sion of the isotopy conjecture, in: Proc. Conf. "Discrete Geometry and
Convexity," New York 1982, (J.E. Goodman, E. Lutwak, J. Malkevitch, and R. Pollack, eds.), Annals of the New York Academy of
Sciences 440 (1985), 12-19. (177)
[217] JACOB E. GOODMAN SE RICHARD POLLACK:
&
BERND STURMFELS:
342
References
Concrete Mathematics. A Foundation for Computer Science, Addison-Wesley, Reading, MA 1989; second edition 1994, in preparation.
(277, 285)
J.
KLEITMAN:
addition
of polytopes: Computational complexity and applications to Griibner
bases, SIAM J. Discrete Math. 6 (1993), 246-269. (224)
[227]
[228]
[229]
Polyhedral Theory,
Chapter 8 in: "The Traveling Salesman Problem" (E.L. Lawler, J.K.
Lenstra, A.H.G. Rinnooy Kan, and D.B. Schmoys, eds.), Wiley-Interscience Series in Discrete Mathematics and Optimization, John Wiley
Si Sons, Chichester New York 1985, 251-360. (22)
[230]
Ulysses 2000: In
search of optimal solutions to hard combinatorial problems, Preprint
SC 93-34, ZIB Berlin 1993, 27 pages. (23)
MARTIN GR6TSCHEL:
References
343
hyperplanes, in: "Proc. Second Lousiana Conference on Combinatorics, Graph Theory and Computing" (R. C. Mullin et al., eds.), Louisiana State University, Baton
Rouge 1971, pp. 41-106. (225)
Polytopal graphs, in: "Studies in Graph Theory, Part II" (D. R. Fulkerson, ed.), MAA Studies in Math. 12, Mathematical Association of America, Washington DC 1975, pp. 201-224.
(116)
An enumeration of
Constructing the associahedron, unpublished manuscript, MIT 1984, 11 pages. (18, 304, 318)
[246]
MARK HAIMAN:
MARK HAIMAN:
344
References
[251] NORA HARTSFIELD & GERHARD RINGEL: Pearls in Graph Theory: A Comprehensive Introduction, Academic Press, San Diego 1990.
(120)
(122)
[256] KATHY HOKE: Extending shelling orders and a hierarchy of functions
of unimodal simple polytopes, Discrete Applied Math. 60 (1995), 211
217. (283)
& VICTOR
[258] FRED HOLT & VICTOR KLEE: Many polytopes meeting the conjectured Hirsch bound, Preprint 1996; Discrete Comput. Geometry, to
appear. (84)
[259] JOHN E. HOPCROFT &
& PETER J.
References
345
[262]
[263]
ROBERT
ROBERT
B.
HUGHES
&
MICHAEL
R.
ANDERSON: A triangulation of
the 6-cube with 308 simplices, Discrete Math. 117 (1993), 253-256.
(147)
[264]
ROBERT
[265]
&
STEFAN THIENEL:
346
[276]
References
GIL KALA1: On low-dimensional faces that high-dimensional
poly-
[278] GIL KALAI: Upper bounds for the diameter and height of graphs of
convex polyhedra, Discrete Comput. Geometry 8 (1992), 363-372.
(87, 91, 96, 101, 102)
[279]
[280]
GIL KALAI:
GIL KALAI:
[281] GIL KALAI: Linear programming, the simplex algorithm and simple
polytopes, Preprint 1997; in: "Proc. Int. Symp. Mathematical Programming" (Lausanne 1997), to appear. (91)
[282] GIL KALAI Sz DANIEL J. KLEITMAN: A quasi-polynomial bound for
the diameter of graphs of polyhedra, Bulletin Amer. Math. Soc. 26
(1992), 315-316. (87, 96)
[283] RAVI KANNAN: Algorithmic geometry of numbers, in: "Annual Review of Computer Science", Vol. 3, Annual Reviews Inc., New York
1987, pp. 231-267. (23)
[284] Goos
[285] LEONID V. KANTOROVICH: My journey in science (proposed report to the Moscow Mathematical Society), Russ. Math. Surveys 42
(1987), 233-270. (96)
[286] MIKHAIL M. KAPRANOV: Pei muto-associahedron, MacLane coherence theorem and asymptotic zones for the KZ equation, J. Pure and
Applied Algebra 85 (1993), 119-142. (19, 291, 310, 312, 313, 321)
[287] MIKHAIL M. KAPRANOV, BERND STURMFELS & ANDREI V.
ZELEVINSKY: Quotients of toric varieties, Mathematische Annalen
290 (1991), 643-655. (320)
[288] MIKHAIL M. KAPRANOV, BERND STURMFELS Sz ANDREI V.
ZELEVINSKY: Chow polytopes and general resultants, Duke Math. J.
67 (1992), 189-218. (320)
References
347
[289] JOHAN KARLANDER: A characterization of affine sign vector systems, preprint, KTH Stockholm 1992. (227)
[290] RICHARD M. KARP: Reducibility between combinatorial problems, in: "Complexity of Computer Computations" (R.E. Miller,
J.W. Thatcher, eds.), Plenum Press, New York 1972, pp. 85-103.
(21)
[291] GYULA KATONA: A theorem on finite sets, in: "Theory of Graphs,"
Proc. Colloquium at Tihany (Sept. 1966), Academic Press, and
Akadmiai Kiado, Budapest 1968, pp. 187-207. (263)
[292] YAN KE S.L JOSEPH O'ROURKE: Comment on Pach's animal problem, preprint 1988, 4 pages. (276)
[293] VICTOR KLEE: Paths on polyhedra I, J. Soc. Indust. Math. 13 (1965),
946-956. (84)
[294] VICTOR KLEE: A property of d-polyhedral graphs, J. Math. Mechanics 13 (1964), 1039-1042. (95)
[295] VICTOR KLEE: A combinatorial analogue of Poincares duality theorem, Canadian J. Math. 16 (1964), 517-531. (260)
[296] VICTOR KLEE: The number of vertices of a convex polytope, Canadian J. Math. 16 (1964), 701-720. (254, 277, 287)
[297] VICTOR KLEE: Paths on polyhedra II, Pacific J. Math. 17 (1966),
249-262. (85, 290)
[298] VICTOR KLEE: A d-pseudomanifold with fo vertices has at least dfo (d 1)(d + 2) d-simplices, Houston Math. J. 1 (1975), 81-86. (279)
[299] VICTOR KLEE & PETER KLEINSCHMIDT: The d-step conjecture and
its relatives, Math. Operations Research 12 (1987), 718-755. (83,
96, 102)
[300] VICTOR KLEE & PETER KLEINSCHMIDT: Geometry of the Gass-
348
References
[306]
[307]
PETER KLEINSCHMIDT:
[309]
New results for simplicial spherical polytopes, in: "Discrete and Computational Geometry: Papers from the DIMACS Special Year" (eds. J. E. Goodman,
R. Pollack, W. Steiger), DIMACS Series in Discrete Mathematics
and Theoretical Computer Science, Vol. 6, Amer. Math. Soc. 1991,
pp. 187-197. (260, 277)
[310]
[311]
[312]
GINA KOLATA:
PAUL KOEBE:
Math problem, long baffling, slowly yields. The traveling salesman problem still isn't solved, but computers can now get
most of the answers, New York Times, Tuesday March 12 (1991),
[314]
ULRICH H. KORTENKAMP:
ULRICH H. KORTENKAMP:
JfIRGEN RICHTER-GEBERT, A.
SARANGARAJAN & GNTER M. ZIEGLER: Extrernal properties of
0/1-polytopes, Discrete & Comput. Geometry 17 (1997), 439-448.
(26)
[316] JOSEPH B. KRUSKAL: The number of simplices in a complex, in:
"Mathematical Optimization Techniques" (R. Bellman, ed.), University of California Press, Berkeley CA, 1963, pp. 251-278. (263)
[315]
ULRICH
H.
KORTENKAMP,
References
349
[317] HAROLD W. KUHN: The solvability and consistency for linear equa-
matical Developments arising from Linear Programming" (J. C. Lagarias and M. J. Todd, eds.), Contemporary Mathematics 114 (1990),
pp. 3-19. (250, 279)
[329] CARL W. LEE: Regular triangulations of convex polytopes, in: "Applied Geometry and Discrete Mathematics The Victor Klee
Festschrift" (P. Gritzmann and Bernd Sturmfels, eds.), DIMACS Series in Discrete Mathematics and Theoretical Computer Science 4,
Amer. Math. Soc. 1991, pp. 443-456. (129, 145, 281)
350
References
(278)
[331] THOMAS LENGAUER: Combinatorial Algorithms
cuit Layout, John Wiley & Sons, Chichester New York, and Teubner,
Stuttgart 1990. (111)
A combinatorial
problem in the k-adic number system, Proc. Amer. Math. Soc. 18
[338]
NATHAN LINIAL,
RICHARD
[339] JESUS DE LOERA: Computing regular triangulations for point configurations, Discrete Comput. Geometry 15 (1996), 253 - 264. (145)
[340] JESUS DE LOERA, BERND STURMFELS & REKHA R. THOMAS:
Grabner bases and triangulations of the second hypersimplex, Cornbinatorica 15 (1995), 409-424. (145)
[341] LisszL6 Loviisz: An Algorithmic Theory of Numbers, Graphs and
Convexity, CMBS-NSF Regional Conference Series in Applied Mathematics 50, Society for Industrial and Applied Mathematics (SIAM),
Philadelphia 1986. (79)
Polytopes secondaires et discriminants, in:
"Sminaire Bourbaki", 43me anne, 1990-91, no. 742, Juin 1991,
Astrisque 201 202 203 (1991), 387-420. (320)
References
351
352
References
[358] J. PETER MAY: Simplicial Objects in Algebraic Topology, Van Nostrand Mathematical Studies, No. 11, D. Van Nostrand, Princeton
1967. (264)
[359] PETER MCMULLEN: On
of certain eu-
Comput. Geom-
polynomials, preprint 1995, 15 pages; in: Proc. "International Conference on Intuitive Geometry" (Budapest, Sept. 1995), to appear.
(278)
[373] PETER MCMULLEN & GEOFFREY C. SHEPHARD: Diagrams for centrally symmetric polytopes, Mathematika 15 (1968), 123-138. (189,
280)
References
[374] PETER MCMULLEN & GEOFFREY C. SHEPHARD:
353
Convex Polytopes
and the Upper Bound Conjecture, London Math. Soc. Lecture Notes
Series 3, Cambridge University Press, Cambridge 1971. (1, 22, 27,
183)
[375] PETER MCMULLEN & EGON SCHULTE:
generalized lower-
The universality theorem on the oriented mafroid stratification of the space of real matrices, in: "Discrete and
354
References
[386] THEODORE S. MOTZKIN, HOWARD RAIFFA, GERALD L. THOMPSON & ROBERT M. THRALL: The double description method, in:
"Contributions to the Theory of Games, Vol. II" (H. W. Kuhn and
A. W. Tucker, eds.), Annals of Math. Studies 28, Princeton University Press, Princeton 1953, pp. 81-103. (36, 48, 74)
[387] KETAN MULMULEY: Computational Geometry: An Introduction
Through Randomized Algorithms, Prentice Hall, Englewood Cliffs,
NJ 1994. (258, 277)
[388] JAMES R. MUNKRES: Elements of Algebraic Topology,
Addison-
References
[398]
355
JAMES G. OXLEY:
1992. (160)
[401] JiiNos PACH, ED.: New Trends in Discrete and Computational Geometry, Algorithms and Combinatorics 10, Springer-Verlag, Berlin
Heidelberg 1993. (22, 129)
[402] JNOS PACH Si PANKAJ K. AGARWAL: Combinatorial Geometry,
J. Wiley & Sons, New York 1995. (118)
[405] MICHA A. PERLES: Results and problems on reconstruction of polytopes, unpublished, Jerusalem 1970. (93, 95)
[406] MICHA A. PERLES: At most 2d+ 1 neighborly simplices in Ed, Annals
of Discrete Math. 20 (1984), 253-254. (144)
[407]
Reconstruction of polytopes A survey, in: "DIMACS workshop on polytopes and convex sets," abstracts, DIMACS,
Rutgers University, January 1990, 1 page. (95)
MICHA A. PERLES:
[408] MICHA A. PERLES: Problems 3.17 and 3.18, in: "DIMACS workshop
on polytopes and convex sets," list of problems, DIMACS, Rutgers
University, January 1990, 8 pages. (102, 289)
[409] MICHA A. PERLES Si NAGABHUSHANA PRABHU: A property of
graphs of convex polytopes, Note, J. Combinatorial Theory, Ser. A
62 (1993), 155-157. (95)
[410] HENRI POINCAR: Sur la gnralisation d'un theorem d'Euler relatif
aux polydres, Comptes Rend. Acad. Sel. Paris 117 (1893), 144-145.
(232)
[411] HENRI POINCAR: Complment a l'analysis situs, Rendiconti Circolo
Mat. Palermo 13 (1899), 285-343. (232)
356
References
[413] J. SCOTT PROVAN: Decompositions, shellings, and diameters of simplicial complexes and convex polyhedra, Ph.D. thesis, Cornell University 1977. (247)
& GNTER
polytopes and
M. ZIEGLER: Mapping
polytopes
Coxeter associahedra,
-
zonotopal W-
matroids, Doc-
Polytopes, Lecture
& GNTER
M. ZIEGLER:
Zonotopal
References
357
[427] GERHARD RINGEL: Teilungen der Ebene durch Geraden oder topolo-
&
PAUL
arrangements and mutations of oriented mat roids, Geometriae Dedicata 27 (1988), 153-170. (208, 223, 224)
358
References
G. SCHNEIDER: Neighbourliness of centrally symmetric convex polytopes in high dimensions, Mathernatika 22 (1975), 176-181.
(280)
[441]
ROLF
[442]
G. SCHNEIDER & WOLFGANG WEIL: Zonoids and related topics, in: "Convexity and Its Applications" (P. M. Gruber and J. Wills,
eds.), Birkhauser, Basel 1983, pp. 296-317. (224)
[443]
ROLF
WALTER SCHNYDER:
Analytic treatment of the polytopes regularly derived from the regular polytopes, Verhandelingen der Koninklijke Akademie van Wetenschappen te Amsterdam, Deel 11, No. 3,
Johannes Mller, Amsterdam 1911, 87 pages. (17, 24, 224)
PIETER HENDRIK SCHOUTE:
543-560. (118)
[448]
Theory of Linear and Integer Programming, Wiley-Interscience Series in Discrete Mathematics and Optimization, John Wiley & Sons, Chichester New York 1986. (19, 22,
23, 27, 47, 69, 96)
ALEXANDER SCHRIJVER:
tain polynomials of E. F. Moore and C. E. Shannon, in: "RLE Quarterly Progress Report No. 55," Research Lab. of Electronics, MIT
1959, pp. 117-131 (263)
Analogues of Steinitz's theorem about non-inscribable polytopes, in: "Intuitive Geometry" (Si6fok 1985), Colloquia
Soc. JAnos Bolyai 48, North-Holland, Amsterdam 1987, pp. 503-516.
(122)
References
359
[454] RAIMUND SEIDEL: The upper bound theorem for polytopes: an easy
proof of its asymptotic version, Comput. Geometry: Theory and Applications 5 (1995), 115 - 116. (258)
[455] MAJORIE SENECHAL: Crystalline Symmetries. An Informal Mathematical Introduction, Adam Hilger, Bristol, Philadelphia and New
York 1993. (9)
[456] MAJORIE SENECHAL & GEORGE FLECK, EDS.: Shaping Space. A
Polyhedral Approach, Design Science Collection, Birkhhuser Verlag,
Boston 1988. (103)
[457] ROBERT WILLIAM SHANNON: Simplicial cells in arrangements of hyperplanes, Geometriae Dedicata 8 (1979), 179-187. (207)
[458] IDO SHEMER: Neighborly polytopes, Israel J. Math. 43 (1982), 291314. (187)
[459] GEOFFREY C. SHEPHARD: Combinatorial properties of associated
zonotopes, Canadian J. Math. 26 (1974), 302 - 321. (230)
Three more animals, preprint, Dept. Computer Science, McGill University, Montral 1988, 5 pages. (276)
THOMAS C. SHERMER:
[465] NAVIN SINGHI & S. S. SHRIKHANDE: A reciprocity relation for tdesigns, European J. Combinatorics 8 (1987), 59 - 68. (284)
[466] ZEEV SMILANSKY: Convex hulls of generalized moment curves, Israel
J. Math. 52 (1985), 115 - 128. (76)
[467] ZEEv SMILANSKY: Bi cyclic 4 polytopes, Israel J. Math. 70 (1990),
82 -92. (76)
-
360
References
References
361
Studies in Pure Mathematics 11, Kinokuniya, Tokyo, and NorthHolland, Amsterdam/New York 1987, pp. 187-213. (280)
[483] RICHARD P. STANLEY: Subdivisions and local h-vectors, J. Amer.
362
References
[497] BERND STURMFELS: Fiber polytopes: A brief overview, in: "Special Differential Equations" (M. Yoshida, ed.), Kyushu University,
Fukuoka, 1991, pp. 117-124. (291, 296, 321)
[498] BERND STURMFELS: Asymptotic analysis of toric ideals, Memoirs of
the Faculty of Science, Kyushu University, Ser. A Mathematics 46
1965. (144)
The monotonic bounded Hirsch conjecture is
false for dimension at least 4, Math. Operations Research 5 (1980),
599-601. (86)
363
References
364
[530]
References
HARALD WISTUBA
&
[532]
R. WOOD, STEPHEN R. McDowELL Sz PETER KLEINSCHMIDT: Constructing infinite dimensional polytopes using Gale
transforms, preprint 1992; SEAMS Bull. Math., in print. (183)
[533]
[534]
HANS
GRAHAM
VLADIMIR
JOSEPH
[535] THOMAS ZASLAVSKY: Facing up to arrangements: Face count formulas for partitions of space by hyperplanes, Memoirs Amer. Math. Soc.
No. 154, 1 (1975). (183)
[536] GNTER M. ZIEGLER: Vier farbige Probleme Nachbargebiete und
Kartenfiirbung in drei und mehr Dimensionen, Jugend forscht (German National Science Fair) 1982. (144)
[537] GNTER M. ZIEGLER: Three problems about 4-polytopes, in: "Poly-
Index
active facet, 88
acyclic orientation, 80
acyclic vector configuration, 156
admissible hyperplane, 67
affine dependences, 150
affine Gale diagram, 168
affine hull, 3
affine map, 3, 195
affine subspaces, 2
affine transformation, 3, 196
affinely independent, 3
affinely isomorphic, 5, 128
animal problem, 276
arrangement, 193
affine, 212
of hyperplanes, 193
of pseudohyperplanes, 211
of pseudolines, 213
realizable, 216
assignment polytope, 20
associahedron, 18, 306
axiom systems, 160
Balinski's theorem, 95
Barnette sphere, 143
being explicit, 79
belt polytopes, 226
bicyclic polytope, 76
binomial expansion, 262
bipyramid, 9
Birkhoff polytope, 20, 320
Bohne-Dress theorem, 220, 225
boundary complex, 129, 232
bounded, 4, 29
Brfickner sphere, 143
Bruggesser-Mani shellings, 240
capped prism, 145
Carathodory curve, 75
Carathodory's theorem, 46
category of polytopes, 319
centrally symmetric, 17
characteristic cone, 43
circle packing theorem, 117
coface, 154
cofacet, 154
cofiber polytope, 319, 324
column vectors, 2
combinatorial explosion, 47
366
Index
invertible, 139
non-Schlegel, 139
simple, 138
simplicial, 138
topological, 142
Dehn-Sommerville equations, 252
Delaunay triangulation, 146
deletion, 106, 163
deletion and contraction, 183
Delta-Wye operation, 106
AY operation, 106
diagram, see d-diagramdiameter,
83
diamond property, 57
dimension, 2, 5, 51, 127, 232
double description method, 36, 48
d-polytope, 5
d-simplex, 7
d-step conjecture, 84
dual graph, 105
dual vector space, 2
duality, 163
duality theorems, 39
edges, 5, 51
Egyptian pyramid, 9
elimination, 32
equivalent, see
combinatorially equivalent
equivalent data, 160
Euler's formula, 120
Euler-Poincar formula, 231
extendably shellable, 235
extremal set theory, 261
compression, 263
face, 5, 51, 232
proper, 5
face fan, 192
face figure, 71, 187
face lattice, 57
face poset, 128
facet, 5, 51, 232
fan, 191
common refinement, 194
Index
direct sum, 194
face fan, 192
nonpolytopal, 194
normal fan, 193
restriction, 195
simplicial, 192
Farkas lemma, 39, 50
fast algorithms, 126
fiber polytope, 291, 296
flag vector, 280
flats, 2
Fourier-Motzkin elimination, 32,
35, 47
for cones, 37
4-polytopes, 127
f-vector, 245
nonunimodal, 272
Gale diagram, 149, 168
affine, 168
central, 189
zonal, 224
Gale's evenness condition, 12, 14
g-conjecture, 269
general position, 8, 79
generating functions, 277
generic, 8, 79
geometric realization, 258
geometry of numbers, 23
graph, 80
d-connected, 95
dimensionally ambiguous, 98
dual, 105
good orientation, 93
k-connected, 104, 116
k-regular, 94
layout, 111
of 3-polytopes, 103
of 4-polytopes, 102, 126
planar, 103
simple, 103
straight line drawing, 120
graph theory, 80
grid graph, 110
group action, 320
367
g-theorem, 269
g-vector, 269
Hamilton cycle, 290
hexagon, 6
Hirsch conjecture, 83
for 0/1-polytopes, 91
monotone, 86
strict monotone, 86
upper bounds, 87
history, 69
homogenization, 31, 44
7-1-polyhedron, 28
7-1-polytope, 29
h-vector, 248
generalized, 280
local, 281
hypercube, 7
hyperplane, 2, 212
linear, 2, 193
hyperplane arrangement, 193
affine, 212
essential, 205
linear, 193
hypersimplex, 19
induced subgraph, 80, 102
integer coordinates, 122
integral points, 23
interior, 60
interval, 56
isomorphic, see affinely isomorphic
isotopy conjecture, 177
join, 323
368
Index
length, 56
line shellings, 242
lineality space, 43
linear dependences, 157
linear forms, 2
linear programming, 80, 193
basis version, 100
linear subspaces, 2
lines, 2
link, 237
lower faces, 130, 294
Macaulay's theorem, 267
main theorem, 27
for cones, 30
for polyhedra, 30
for polytopes, 29
MAPLE, 279
mapping polytopes, 319
matching polytope, 320
matroid, 160
McMullen correspondence, 270
Minkowski sum, 28, 197
minor, 106
Mbius band, 148
moment curve, 11
monotone Hirsch conjecture, 86
monotone path polytope, 300
monotone paths, 81
M-sequences, 268
multicomplex, 264
multiset, 263
negative point, 168
neighborly polytopes, 16, 186, 187,
254
nets, 126
Newton polytope, 322
n-gon, 6
non-Pappus configuration, 217
nonrational 8-polytope, 172
nonrevisiting path conjecture, 85
normal fan, 193
notation, 2
octahedron, 7
orientation of a graph, 80
acyclic, 80, 93
oriented matroid, 149, 158, 160
axiom systems, 159
cocircuit axioms, 228
covector axioms, 209
deletion and contraction, 163
duality, 149, 163
equivalent data, 160
non-Pappus, 217
nonrealizable, 208, 216
rank, 159, 209
realizable, 158
what are they good for? 211
Pappus configuration, 216
partially ordered sets, 55
pencil of lines, 141
perfect matching polytope, 20
permutahedron, 17, 200, 301
generalized, 24
permuto-associahedron, 19, 310
piecewise linear, 130
pile of cubes, 131
planes, 2
point beyond F, 78
pointed polyhedra, 43
points, 2
polar polytope, 8, 59
combinatorially, 64
polar set, 61
polygons, 6
polyhedral complex, 127
polyhedral cones, 30
polyhedron
fl-polyhedron, 4, 28
pointed, 43
V-polyhedron, 29
polytopal complex, 127, 232
shellable, 233
polytope, 4, 5
bicyclic, 76
cd-index, 280
centrally symmetric, 17
Index
connected sum, 274
cubical, 23, 280
d-polytope, 5
edges, 51
equivalent, 5
face lattice, 57
faces, 51
facets, 51
flag vector, 280
4-polytopes, 127
f -vector, 245
graph, 80
g-vector, 269
7-i-polytope, 4, 29
h-vector, 246, 250
integer coordinates, 122
isomorphic, 5
join, 323
Minkowski sum, 28
neighborly, 16, 186, 187, 254
nonrational, 172, 186
polar, 59
product, 9
projection, 196, 292
proper face, 51
quotient, 71, 289
rational, 66
reconstruction problems, 95
representation theorem, 65
ridges, 51
rigid, 179
simple, 8, 66
simplicial, 8, 65
spherical, 260
3-polytopes, 103
vertices, 51
V-polytope, 4, 29
with "few vertices", 171
polytope algebra, 278
polytope pairs, 320
PORTA, 11, 48, 309
poset, 56
bounded, 56
graded, 56
interval, 56
369
rank, 56
positive halfspace, 203
positive hull, 28
positive point, 168
positive sign vector, 154
prescribing
shadow boundary, 114
shape of a facet, 114, 141, 174
2-face, 175
prism, 10
product, 9
projection, 16, 32, 292
projection of polytopes, 196
projective transformations, 67
applications, 74
projectively unique, 190
proper face, 5, 51
pseudoline, 213
pyramid, 9
quotients, 71, 289
Radon's theorem, 151, 184
rank, 56, 156
rational polytope, 66
realizable oriented matroid, 158
realization space, 115
recession cone, 43
redundancy criteria, 73
redundant inequality, 47, 72
criteria, 73
regular n-gon, 6
regular subdivision, 129
relative interior, 60
representation theorem, 65
reverse lexicographic ordering, 261
reverse search, 49
ridges, 51
row vectors, 2
Schlegel diagram, 133, 134
of cyclic polytope, 144
secondary polytope, 291, 305
section, 296
semialgebraic set, 115, 119
370
Index
tetrahedron, 6
3-polytopes, 103
topological representation
theorem, 211, 215
transportation polytope, 38, 50
traveling salesman polytope, 21
traveling salesman problem, 21
triangulation, 129
2-neighborly, 10
ultraconnected, 101
underlying set, 127, 232
unimodality conjecture, 271
universality theorem, 182
upper bound theorem, 16, 254
for centrally symmetric
polytopes? 279
for polytope pairs, 279
valid inequality, 51
value vector, 153, 157
vector configuration
acyclic, 156
dual, 165
simple, 206
totally cyclic, 167
vector space, 2
vector sum, 28, 197
vectors
column vectors, 2
row vectors, 2
vertex figure, 54
vertex set, 51
vertices, 5, 51
visible, 240
V-polyhedron, 29
V-polytope, 29
0/1-polytopes, 19, 25, 70
zonal diagrams, 224
zone, 206
zonotopal tiling, 218
zonotope, 17, 191, 199
associated, 226
generalized, 226
volume, 230