You are on page 1of 250

A friendly invitation

to Fourier analysis on polytopes

Sinai Robins

August 25, 2021


Contents

1 Introduction 1

2 A motivating problem:
tiling a rectangle with rectangles 7

2.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.2 Nice rectangles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.3 Conventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3 Examples that nourish the theory 16

3.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3.2 Dimension 1 - the classical sinc function . . . . . . . . . . . . . . . . . . . . 17

3.3 The inverse Fourier transform . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3.4 Bernoulli polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.5 The cube, and its Fourier transform . . . . . . . . . . . . . . . . . . . . . . . 23

3.6 The simplex, and its Fourier transform . . . . . . . . . . . . . . . . . . . . . 24

3.7 Stretching and translating . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.8 The parallelepiped, and its Fourier transform . . . . . . . . . . . . . . . . . . 29

3.9 The cross-polytope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.10 Observations and questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

1
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4 Tools of the trade:


basic Fourier analysis 43

4.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.3 Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.4 Fourier series: fast and loose . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.5 Fourier series: rigorous statements . . . . . . . . . . . . . . . . . . . . . . . . 49

4.6 The Schwartz space, and other goodies . . . . . . . . . . . . . . . . . . . . . 50

4.7 Poisson Summation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.8 The convolution operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.9 Plancherel’s Theorem, and the Parseval identity . . . . . . . . . . . . . . . . 59

4.10 More crucial properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.11 Approximate identity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.12 A practical Poisson summation formula . . . . . . . . . . . . . . . . . . . . . 65

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5 The geometry of numbers -


Minkowski’s Fundamental Theorem, and Siegel’s extension 73

5.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

5.2 Minkowski’s convex body Theorem . . . . . . . . . . . . . . . . . . . . . . . 74

5.3 Siegel’s extension of Minkowski:


a Fourier transform identity for convex bodies . . . . . . . . . . . . . . . . . 75

5.4 Tiling and multi-tiling Euclidean space by translations of polytopes . . . . . 79

5.5 Extremal bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5.6 More about centrally symmetric polytopes . . . . . . . . . . . . . . . . . . . 84

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

2
6 An introduction to Euclidean lattices 95

6.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

6.2 Introduction to lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

6.3 Discrete subgroups -


an alternate definition of a lattice . . . . . . . . . . . . . . . . . . . . . . . . 99

6.4 Lattices defined by congruences . . . . . . . . . . . . . . . . . . . . . . . . . 101

6.5 The Gram matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

6.6 Dual lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

6.7 The successive minima of a lattice, and Hermite’s constant . . . . . . . . . . 110

6.8 Hermite normal form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

6.9 The Voronoi cell of a lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

6.10 Quadratic forms and lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

7 The Fourier transform of a polytope via its vertex description:


The Brion theorems 127

7.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

7.2 Cones, simple polytopes, and simplicial polytopes . . . . . . . . . . . . . . . 129

7.3 Tangent cones, and the Fourier transform of a simple polytope . . . . . . . . 131

7.4 Fourier-Laplace transforms of cones . . . . . . . . . . . . . . . . . . . . . . . 134

7.5 The Brianchon-Gram identity . . . . . . . . . . . . . . . . . . . . . . . . . . 141

7.6 Proof of Theorem 7.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

7.7 An application of transforms to the volume of a simple polytope, and for its
moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

7.8 Brion’s theorem - the discrete form . . . . . . . . . . . . . . . . . . . . . . . 147

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

8 The angle polynomial of a polytope 156

3
8.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

8.2 What is an angle in higher dimensions? . . . . . . . . . . . . . . . . . . . . . 157

8.3 Local solid angles for a polytope, and Gaussian smoothing . . . . . . . . . . 158

8.4 The Gram relations for solid angles . . . . . . . . . . . . . . . . . . . . . . . 164

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166

9 Counting integer points in polytopes - the Ehrhart theory 168

9.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

9.2 Computing integer points in polytopes via the discrete Brion Theorem . . . 169

9.3 Examples, examples, examples.... . . . . . . . . . . . . . . . . . . . . . . . . 173

9.4 The Ehrhart polynomial of an integer polytope . . . . . . . . . . . . . . . . 177

9.5 Unimodular polytopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

9.6 Rational polytopes and quasi-polynomials . . . . . . . . . . . . . . . . . . . 181

9.7 Ehrhart reciprocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

9.8 The Möbius inversion formula for the face poset . . . . . . . . . . . . . . . . 185

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

10 Sphere packings 194

10.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

10.2 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195

10.3 Upper bounds for sphere packings via Poisson summation . . . . . . . . . . . 196

10.4 The Fourier transform of the ball . . . . . . . . . . . . . . . . . . . . . . . . 199

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

11 The Fourier transform of a polytope via its hyperplane description:


the divergence Theorem 203

11.1 Intuition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204

4
11.2 The divergence Theorem, and a combinatorial
divergence theorem for polytopes . . . . . . . . . . . . . . . . . . . . . . . . 204

11.3 Generic frequencies versus special frequencies . . . . . . . . . . . . . . . . . . 212

Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216

5
Acknowledgements

The famous saying “no man is an island” is doubly-true in Mathematics, and indeed I’ve had
the good fortune to know and learn from many interesting people, concerning the contents
of this book. Special thanks goes to Ricardo Diaz, my first collaborator along these topics.
I would like to thank the following people, from the bottom of my heart, for their valuable
input and interesting discussions about some of these topics over the years:

Ian Alevy, Artur André, Christine Bachoc, Tamar Bar, Imre Bárány, Alexander Barvinok,
Matthias Beck, Dori Bejleri, Luca Brandolini, Michel Brion, Henry Cohn, Leonardo Colzani,
Amalia Culiuc, Pierre Deligne, Jesús A. De Loera, Michel Faleiros, Lenny Fukshansky,
Nick Gravin, Martin Henk, Didier Henrion, Roberto Hirata Junior, Jeffrey Hoffstein,
Alex Iosevich, Marvin Knopp, Mihalis Kolountzakis, Matthias Köppe, Jean Bernard Lasserre,
Nhat Le Quang, Rafael Zuolo Coppini Lima, Sameer Iyer Fabrı́cio Caluza Machado,
Romanos Malikiosis, Máté Matolci, Tyrrell McAllister, Victor Moll, Mel Nathanson,
James Pommersheim, Jim Propp, Thales Paiva, Jill Pipher, Jorge Luis Ramı́rez Alfonsı́n,
Ethan Reiner, Bruce Reznick, Tiago Royer, Nicolas Salter, Gervásio Santos, Richard Schwartz,
Dima Shiryaev, Joseph Silverman, Richard Stanley, Christophe Vignat, Sergei Tabachnikov,
Giancarlo Travaglini, Kevin Woods, Ren Yi, Günter Ziegler, Chuanming Zong.

6
Chapter 1

Introduction

What is a Fourier transform? Why is it so useful?


How can we apply Fourier transforms and Fourier se-
ries - which were originally used by Fourier to study
heat diffusion - in order to better understand topics in
discrete and combinatorial geometry, number theory,
and sampling theory?

To begin, there are some useful analogies: imagine


that you are drinking a milk-shake (lactose-free), and
you want to know the ingredients of your tasty drink.
You would need to filter out the shake into some of
its most basic components. This decomposition into
its basic ingredients may be thought of as a sort of
“Fourier transform of the milk-shake”. Once we un-
derstand each of the ingredients, we will also be able
to restructure these ingredients in new ways, to form
many other types of tasty goodies. To move the anal- Figure 1.1: Joseph Fourier
ogy back into mathematical language, the milkshake
represents a function, and each of its basic ingredi-
ents represents for us the basis of sines and cosines; we may also think of a basic ingredient
more compactly as a complex exponential e2πinx , for some n ∈ Z. Composing these basic
ingredients together in a new way represents a Fourier series.

Mathematically, one of the most basic kinds of milk-shakes is the indicator function of the
unit interval, and to break it down into its basic components, mathematicians, Engineers,
Computer scientists, and Physicists have used the sinc function (since the 1800’s):

sin(πz)
sinc(z) :=
πz

1
with great success, because it happens to be the Fourier transform of the unit interval [− 12 , 21 ]:
Z 1
2
e−2πizx dx = sinc(z),
− 12

as we will compute shortly in identity (3.5). Somewhat surprisingly, comparatively little


energy has been given to some of its higher dimensional extensions, namely those extensions
that arise naturally as Fourier transforms of polytopes.

One motivation for this book is to better understand how this 1-dimensional function –
which has proved to be extremely powerful in applications – extends to higher dimensions.
Namely, we will build various mathematical structures that are motivated by the question:

What is the Fourier transform of a polytope?

Of course, we will ask “how can we apply it”? An alternate title for this book might have
been:
We’re taking Poisson summation and Fourier transforms of polytopes
for a very long ride....

Historically, sinc functions were used by Shannon (as well as Hardy, Kotelnikov, and Whit-
taker) when he published his seminal work on sampling theory and information theory.

In the first part of this book, we will learn how to use the technology of Fourier transforms of
polytopes in order to build the (Ehrhart) theory of integer point enumeration in polytopes,
to prove some of Minkowski’s theorems in the geometry of numbers, and to understand when
a polytope tiles Euclidean space by translations.

In the second portion of this book, we give some applications to active research areas which
are sometimes considered more applied, including the sphere-packing problem, and the sam-
pling of signals in higher dimensions.

There are also current research developments of the material developed here, to the learning
of deep neural networks. In many applied scientific areas, in particular radio astronomy,
computational tomography, and magnetic resonance imaging, a frequent theme is the recon-
struction of a function from knowledge of its Fourier transform. Somewhat surprisingly, in
various applications we only require very partial/sparse knowledge of its Fourier transform
in order to reconstruct the required function, which may represent an image or a signal.

There is a rapidly increasing amount of research focused in these directions in recent years,
and it is therefore time to put many of these new findings in one place, making them accessible
to a general scientific reader. The fact that the sinc function is indeed the Fourier transform
of the 1-dimensional line segment [− 21 , 21 ], which is a 1-dimensional polytope, gives us a first
hint that there is a deeper link between the geometry of a polytope and the analysis of its
Fourier transform.

2
Indeed one reason that sampling and information theory, as initiated by Claude Shannon,
works so well is precisely because the Fourier transform of the unit interval has this nice
form, and even more-so because of the existence of the Poisson summation formula.

The approach we take here is to gain insight into how the Fourier transform of a polytope can
be used to solve various specific problems in discrete geometry, combinatorics, optimization,
approximation theory, and the Shannon-Whittaker sampling theory in higher dimensions:

(a) Analyze tilings of Euclidean space by translations of a polytope

(b) Give wonderful formulas for volumes of polytopes

(c) Compute discrete volumes of polytopes, which are combinatorial approximations to


the continuous volume

(d) Introduce the geometry of numbers, via Poisson summation

(e) Optimize sphere packings, and get bounds on their optimal densities

(f) Study the Shannon-Whittaker sampling theorem and its higher-dimensional siblings

(g) Recover a polytope by the inverse problem of knowing enough of its moments

Let’s see at least one direction that quickly motivates the study of Fourier transforms. In par-
ticular, we often begin with simple-sounding problems that arise naturally in combinatorial
enumeration, discrete and computational geometry, and number theory.

Throughout, an integer point is any vector v := (v1 , . . . , vd ) ∈ Rd , all of whose coordinates


vj are integers. In other words, v belongs to the integer lattice Zd . A rational point is
a point m whose coordinates are rational numbers, in other words m ∈ Qd . We define the
Fourier transform of a function f (x):
Z
ˆ
f (ξ) := f (x)e−2πihξ,xi dx, (1.1)
Rd

defined for all ξ ∈ Rd for which the latter integral converges, and where we use the standard
inner product ha, bi := a1 b1 + · · · + ad bd . We will also use the notation F(f ) for the Fourier
transform of f , which is useful in some typographical contexts, for example when considering
F −1 (f ).

We introduce one of the main objects of study in this book, the Fourier transform of a
polytope P, defined by:
Z Z
−2πihξ,xi
1̂P (ξ) := 1P (x)e dx = e−2πihξ,xi dx, (1.2)
Rd P

3
where the function 1P (x) is the indicator function of P, defined by
(
1 if x ∈ P
1P (x) :=
0 if not.
Thus, the words “Fourier transform of a polytope P” will always mean the Fourier transform
of the indicator function of P.

The Poisson summation formula, named after


Siméon Denis Poisson, tells us that for any “suffi-
ciently nice” function f : Rd → C we have:
X X
f (n) = fˆ(ξ).
n∈Zd ξ∈Zd

In particular, if we were to naively set f (n) := 1P (n),


the indicator function of a polytope P, then we would
get:
X X
1P (n) = 1̂P (ξ), (1.3)
n∈Zd ξ∈Zd

which is technically false in general due to the fact


that the indicator function 1P is a discontinuous func-
tion on Rd .

However, this technically false statement is very use-


ful! We make this claim because it helps us build Figure 1.2: Siméon Denis Poisson
intuition for the more rigorous statements that are
true, and which we study in later chapters. For ap-
plications to discrete geometry, we are interested in the number of integer points in a closed
convex polytope P, namely |P ∩ Zd |. The combinatorial-geometric quantity |P ∩ Zd | may
be regarded as a discrete volume for P. From the definition of the indicator function of
a polytope, the left-hand-side of (1.3) counts the number of integer points in P, namely we
have by definition

X
1P (n) = |P ∩ Zd |. (1.4)
n∈Zd

On the other hand, the right-hand-side of (1.3) allows us to compute this discrete volume of
P in a new way. This is great, because it opens a wonderful window of computation for us
in the following sense:

X
|P ∩ Zd | = 1̂P (ξ). (1.5)
ξ∈Zd

4
We notice that for the ξ = 0 term, we have
Z Z
−2πih0,xi
1̂P (0) := 1P (x)e dx = dx = vol(P), (1.6)
Rd P

and therefore the discrepancy between the continuous volume of P and the discrete volume
of P is
X
|P ∩ Zd | − vol(P) = 1̂P (ξ), (1.7)
ξ∈Zd −{0}

showing us very quickly that indeed |P ∩ Zd | is a discrete approximation to the classical


Lebesgue volume vol(P), and pointing us to the task of finding ways to evaluate the transform
1̂P (ξ). From the trivial but often very useful identity

1̂P (0) = vol(P),

we see another important motivation for this book: the Fourier transform of a polytope is
a very natural extension of volume. Computing the volume of a polytope P captures a
bit of information about P, but we also lose a lot of information.

On the other hand, computing the Fourier transform of a polytope 1̂P (ξ) uniquely determines
P, so we do not lose any information at all. Another way of saying this is that the Fourier
transform of a polytope is a complete invariant. In other words, it is a fact of life that

1̂P (ξ) = 1̂Q (ξ) for all ξ ∈ Rd ⇐⇒ P = Q.

Combinatorially, there are brilliant identities (notably the Brion identities) that emerge
between the Fourier and Laplace transforms of a given polytope, and its facets and vertex
tangent cones.

In Statistics, the moment generating function of any probability distribution is given by a


Fourier transform of the indicator function of the distribution, hence Fourier transforms arise
very naturally in Statistical applications. At this point, a natural glaring question naturally
comes up:

How do we compute the Fourier transform of a polytope 1̂P (ξ)? (1.8)

And how do we use such computations to help us understand the important “error” term
X
1̂P (ξ)
ξ∈Zd −{0}

that came up naturally in (1.7) above?

5
There are many applications of the theory that we will build-up. Often, we find it instruc-
tive to sometimes give an informal proof first, because it brings the intuitive ideas to the
foreground, allowing the reader to gain an overview of the steps. Then, later on, we revisit
the same intuitive proof again, making it rigorous.

The Poisson summation formula is one of our main stars, and has a relatively easy proof. But
it constitutes a very first step for many of our explorations. It may even be said that, from
this perspective, the Poisson summation formula is to combinatorial analysis as a microscope
is to our vision. It enhances our ability to see mathematical facts, and often in a surprisingly
simple way. So it’s a question of what we do with these tools - where do we point them?

A word about prerequisites for this book: Linear Algebra is always very useful! A couple
of calculus courses are helpful as well, with perhaps a touch of real analysis. In particular,
familiarity with infinite series is assumed.

We will assume some familiarity with the basic definitions of polytopes and their faces,
although at places we will remind the reader of some of these definitions. There are many
excellent texts that introduce the student to the classical language of polytopes, in particular
the two classics: Günter Ziegler’s “Lectures on Polytopes” [163], and Branko Grünbaum’s
“Convex Polytopes” [68].

For an easy introduction to the interactions between polytopes and lattice point enumera-
tion, the reader is invited to consult “Computing the continuous discretely: integer point
enumeration in polytopes”, by Beck and Robins [17]. The level of the current book is aimed
at advanced undergraduates and beginning graduate students in various fields, and in par-
ticular Mathematics, Computer Science, Electrical Engineering, and Physics.

This book proceeds by developing an intuitive understanding first, using many examples and
analogies, and this intuition then points us to a rigorous path for the details of the ensuing
proofs.

[PLEASE NOTE: THIS BOOK IS IN DRAFT FORM, and some proofs and citations
are to be inserted. Many additional solutions to the exercises will be inserted as well.]

Sinai Robins June 2021

IME, University of São Paulo

6
Chapter 2

A motivating problem:
tiling a rectangle with rectangles

Ripping up carpet is easy – tiling is the issue.


– Douglas Wilson

2.1 Intuition

To warm up, we begin with a simple tiling


problem in the plane. A rectangle will be
called nice if at least one of its sides is
an integer. We prove a now-classical fact
about tiling a rectangle with nice rectangles,
namely Theorem 2.1, and we focus on the
method of the straightforward proof.

This proof brings to the foreground an im-


portant idea: by simply taking a Fourier
transform of a body B, we immediately get
interesting geometric consequences for B. In
particular, we will see throughout this book
various ways in which the Fourier transform
of a geometric body is a natural extension of Figure 2.1: A rectangle tiled by nice rectangles
its volume, sometimes in a continuous way,
and sometimes in a discrete way. So in order
to study relationships between volumes of bodies, it is very natural and useful to play with
their Fourier transforms.

7
2.2 Nice rectangles

The tilings that we focus on, in this small chapter, are tilings that are composed of smaller
rectangles, all of which have their sides parallel to the axes, and all of which are nice. There
are at least 14 different known proofs [159] of Theorem 2.1. Here we give the proof that uses
very basic Fourier tools, from first principles, motivating the chapters that follow. The idea
for this proof goes back to Nicolaas Govert De Bruijn [42].

Theorem 2.1 (De Bruijn). Suppose we tile a fixed rectangle R with smaller, nice rectangles.
Then R is a nice rectangle.

Proof. Suppose that the rectangle R is tiled with smaller rectangles R1 , . . . , RN , as in Fig-
ure 2.1. Due to our tiling hypothesis, we have
N
X X
1R (x) = 1Rk (x) + (± indicator functions of lower-dimensional polytopes), (2.1)
k=1

where the notation 1S (x) always means we are using indicator functions. To ease the reader
into the computations, we recall that the Fourier transform of the indicator function of any
rectangle R := [a, b] × [c, d] is defined by:
Z Z bZ d
−2πihξ,xi
1̂R (ξ) := 1R (x)e dx = e−2πi(ξ1 x1 +ξ2 x2 ) dx1 dx2 . (2.2)
R2 a c

Now we may formally take the Fourier transform of both sides of (2.1). In other words we
simply multiply both sides of (2.1) by the exponential function e−2πihξ,xi and then integrate
both sides over R2 , to get:
N
X
1̂R (ξ) = 1̂Rk (ξ). (2.3)
k=1

In (2.3), we have used the fact that a 2-dimensional integral over a 1-dimensional line seg-
ment always vanishes, due to the fact that a line segment has measure 0 relative to the
2-dimensional measure of the 2-dimensional transform. Let’s compute one of these integrals,

8
over a generic rectangle Rk := [a1 , a2 ] × [b1 , b2 ]:
Z Z
−2πihx,ξi
1̂Rk (ξ) := 1Rk (x)e dx = e−2πihx,ξi dx (2.4)
R2 Rk
Z b2 Z a2
= e−2πihx,ξi dx (2.5)
b1 a1
Z a2 Z b2
−2πiξ1 x1
= e dx1 e−2πiξ2 x2 dx2 (2.6)
a1 b1
−2πiξ1 a2 −2πiξ1 a1
e −e e−2πiξ2 b2 − e−2πiξ2 b1
= · (2.7)
−2πiξ1 −2πiξ2
−2πi(ξ1 a1 +ξ2 b1 )
1 e
= 2
(e−2πiξ1 (a2 −a1 ) − 1)(e−2πiξ2 (b2 −b1 ) − 1), (2.8)
(−2πi) ξ1 ξ2
valid for all (ξ1 , ξ2 ) ∈ R2 except for the union of the two lines ξ1 = 0 and ξ2 = 0. Considering
the latter formula for the Fourier transform of a rectangle, we make the following leap of
faith:

Claim. Suppose that R is a rectangle whose sides are parallel to the axes. Then
 
R is a nice rectangle ⇐⇒ 1̂R ( 11 ) = 0. (2.9)

Proof of the claim. We consider the last equality (2.8). We see that

1̂Rk (ξ) = 0 ⇐⇒ (e−2πiξ1 (a2 −a1 ) − 1)(e−2πiξ2 (b2 −b1 ) − 1) = 0, (2.10)

which is equivalent to having either e−2πiξ1 (a2 −a1 ) = 1, or e−2πiξ2 (b2 −b1 ) = 1. But we know
that due to Euler, e2πiθ = 1 if and only if θ ∈ Z (Exercise 2.1), so we have

1̂R (ξ) = 0 ⇐⇒ ξ1 (a2 − a1 ) ∈ Z or ξ2 (b2 − b1 ) ∈ Z. (2.11)

Now, if R is a nice rectangle, then one of its sides is an integer, say a1 − a2 ∈ Z without
  loss
1
of generality. Therefore ξ1 (a2 − a1 ) ∈ Z for ξ1 = 1, and by (2.11), we see that 1̂R ( 1 ) = 0.
 
Conversely, if we assume that 1̂R ( 11 ) = 0, then by (2.11) either
1 · (a2 − a1 ) ∈ Z or 1 · (b2 − b1 ) ∈ Z, proving the claim.

To finish the proof of the theorem, byhypothesis


 each little rectangle Rk is a nice rectangle,
so by the claim above it satisfies 1̂Rk ( 11 ) = 0. Returning to (2.3), we see that therefore
1̂R (ξ) = N
P 1
k=1 1̂Rk (ξ) = 0, for ξ = ( 1 ), and using the claim again (the converse part of it
this time), we see that R must be nice.

The proof of Theorem 2.1 was simple and elegant, motivating the use of Fourier transforms
of polytopes in the ensuing chapters. The claim, namely equation (2.9), offers an intriguing

9
springboard for deeper investigations - it tells us that we can convert a geometric statement
about tiling into a purely analytic statement about the vanishing of a certain integral trans-
form. Later, when we learn about Theorem 5.5, we will see that this small initial success of
(2.9) is part of a larger theory. This is the beginning of a beautiful friendship...

2.3 Conventions

We mention some conventions that we use throughout the book. First, we note that whenever
we are given a complex-valued function f : Rd → C, we may write f in terms of its real and
imaginary parts: f (x) := u(x) + iv(x). Now its integral is defined by
Z Z Z
f (x)dx := u(x)dx + i v(x)dx, (2.12)
Rd Rd Rd
so that all of our Fourier transforms are really reduced to the usual integration of real-valued
functions on Euclidean space (see Exercise 2.4). This is good news for the reader, because
even though we see complex functions in the integrand, elementary calculus suffices.

Let S ⊂ Rd be a set. Then we say that S is measurable if the integral


Z
dx exists.
S
Equivalently, we may call S measurable if the indicator function 1S is an integrable function,
by definition of the integral. A set S is said to have measure zero if
Z
dx = 0.
S
In R, for example, we may also define a set S of measure 0 by saying that, given any  > 0,
there exists a countable collection of open intervals In that cover all of S, and whose total
length satisfies ∞
P
|I |
n=1 n < . But we will assume the reader knows the definition(s) of an
integral (either the Riemann integral or the Lebesgue integral), circumventing discussions
about σ-algebras of sets, so that the background required of the reader is kept to a minimum.

The point we want to make here is that most things are in fact easier than the reader may
have previously thought.

We say that a statement A(x) concerning points x ∈ Rd holds for almost every x ∈ Rd if
the set of x ∈ Rd for which A(x) is false is a set of measure 0.

Finally, a general word about definitions. Whenever we want to define a new object called
N , in terms of some combination of previously known mathematical objects called K, we
will use the notation
N := K.

10
Notes
(a) This little chapter was motivated by the lovely article written by Stan Wagon [159],
which gives 14 different proofs of Theorem 2.1. The article [159] is important because
it shows how tools from one field can leak into another field, and thus may lead to
important discoveries in the future.

(b) In a related direction, we might wonder which polygons, and more generally which
polytopes, tile Euclidean space by translations with a lattice. It turns out (Theo-
rem 5.5) that this question is equivalent to the statement that the Fourier transform
of P vanishes on a (dual) lattice.

(c) In the context of the Hilbert space of functions L2 ([0, 1]), Exercise 2.3 is one step to-
wards showing that the set of exponentials {en (x)}n∈Z is a basis for L2 ([0, 1]). Namely,
the identity above shows that
R 1 these basis elements are orthogonal to each other - their
inner product hea , eb i := 0 ea (x)eb (x)dx vanishes for integers a 6= b. Thus, the iden-
tity of Exercise 2.3 is often called the orthogonality relations for exponentials, over
L2 ([0, 1]). To show that they span the space of functions in L2 ([0, 1]) is a bit harder,
but see [155] for details.

(d) The question in Exercise 2.16 for Z was originally asked by Paul Erdös in 1951, and
has an affirmative answer. This question also has higher-dimensional analogues:

Suppose we give a partition of the integer lattice Zd into a finite, disjoint


union of translated sublattices. Is it always true that at least two of these
sublattices are translates of each other?

The answer is known to be false for d ≥ 3, but is still unsolved for d = 2 (see [55],[24]).

Exercises

2.1. ♣ Show that if x ∈ C, then e2πix = 1 if and only if x ∈ Z.

2.2. Show that |ez | ≤ e|z| , for all complex numbers z ∈ C.

2.3. ♣ Here we prove the orthogonality relations for the exponential functions
defined by en (x) := e2πinx , for each integer n. Recall that the complex conjugate of any
complex number x + iy is defined by

x + iy := x − iy,

11
so that eiθ := e−iθ for real θ. Prove that for all integers a, b:
Z 1 (
1 if a = b
ea (x)eb (x)dx = (2.13)
0 0 if not.

2.4. Here the reader may gain some practice with the definitions of integrals that use complex-
valued integrands f (x) := u(x) + iv(x). We recall for the reader the following definition:
Z Z Z Z
f (x)dx := (u(x) + iv(x)) dx := u(x)dx + i v(x)dx, (2.14)
Rd Rd Rd Rd

a linear combination of two real-valued integrals. Recalling that by definition,


Z
1̂[0,1] (ξ) := e−2πiξx dx,
[0,1]

show directly from definition 2.14 and from Euler’s identity eiθ = cos θ + i sin θ, that for any
nonzero ξ ∈ R, we have
e−2πiξ − 1
Z
e−2πiξx dx = .
[0,1] −2πiξ

Notes. Another way of thinking about this exercise is that it extends the ‘Fundamental
theorem of calculus’ to complex-valued functions in a rather easy way. The anti-derivative
−2πiξx
of the integrand f (x) := e−2πiξx is F (x) := e−2πiξ , and we are saying that it is ok to use it in
place of the usual anti-derivative in Calculus 1 - it is consistent with definition 2.14. In the
future, we generally do not have to break up complex integrals into their real and imaginary
parts, because we can make use of the fact that antiderivatives of complex-valued functions
are often simple, such as the one in this example.

We also note that this is not calculus with a complex variable, because the domains of
our integrands, as well as the measures we are using throughout this book, in order to
integrate, are always defined over real Euclidean space Rd . This means we are still using
basic Calculus.

2.5. ♣ We recall that the N ’th roots of unity are by definition the set of N complex solutions
to z N = 1, and are given by the set {e2πik/N | k = 0, 1, 2, . . . , N − 1} of points on the unit
circle. Prove that the sum of all of the N ’th roots of unity vanishes. Precisely, fix any
positive integer N ≥ 2, and show that
N −1
2πik
X
e N = 0.
k=0

12
2πi
Figure 2.2: The 6’th roots of unity, with ζ := e 6 . Geometrically, Exercise 2.5 tells us that
their center of mass is the origin.

2.6. Prove that, given positive integers M, N , we have


N −1
(
1 X 2πikM 1 if N | M
e N =
N k=0 0 if not.

Notes. This result is sometimes referred to as “the harmonic detector” for detecting when a
rational number M N
is an integer; that is, it assigns a value of 1 to the sum if M
N
∈ Z, and it
M
assigns a value of 0 to the sum if N 6∈ Z.

2.7. ♣ Here we prove the orthogonality relations for roots of unity. Namely, fix any
two nonnegative integers a, b, and prove that
N −1
(
1 X 2πika − 2πikb 1 if a ≡ b mod N
e N e N = (2.15)
N k=0 0 if not.

Notes. In a later chapter on Euclidean lattices (Chapter 6), we will see that the identity 2.15
is a special case of the more general orthogonality relations for characters on lattices. From
this perspective, this exercise is the orthogonality relations on the finite cyclic group Z/N Z.
There are more general orthogonality relations for characters of group representations, which
play an important role in Number Theory.

2.8. Show that for any positive integer n, we have


n−1
Y
n= (1 − ζ k ),
k=1

where ζ := e2πi/n .

13
2.9. An N ’th root of unity is called a primitive root of unity if it is not a k’th root of
unity for some smaller positive integer k < N . Show that the primitive N ’th roots of unity
are precisely the numbers e2πik/N for which gcd(k, N ) = 1.

2.10. The Möbius µ-function is defined by:


(
(−1) number of prime factors of n if n > 1
µ(n) :=
1 if n = 1.

Prove that the sum of all of the primitive N ’th roots of unity is equal to the Möbius µ-function,
evaluated at N :
2πik
X
e N = µ(N ). (2.16)
1≤k<N
gcd(k,N )=1

2.11. ♣ We follow the Weierstrassian approach to defining the complex exponential ez for
all complex z ∈ C:

z
X 1 n
e := z , (2.17)
n=0
n!
which converges absolutely for all z ∈ C. We also have the (Weierstrassian) definitions of
cos z and sin z:
∞ ∞
X 1 X 1
cos z := (−1)n z 2n , sin z := (−1)n−1 z 2n−1 ,
n=0
(2n)! n=1
(2n − 1)!

both converging absolutely again for all z ∈ C. Prove that Euler’s formula has the extension:

eiz = cos z + i sin z,

valid for all z ∈ C.

Notes. Karl Weierstrass developed a rigorous and beautiful theory of real and complex
functions, beginning with such a power series approach.

2.12. Here the reader needs to know a little bit about the quotient of two groups (this is one of
the few exercises that assumes group theory). We prove that the group of ‘real numbers mod
1’ under addition, is isomorphic to the unit circle, under multiplication of complex numbers.
Precisely, we can define h : R → S 1 by h(x) := e2πix .

(a) We recall the definition of the kernel of a map, namely ker(h) := {x ∈ R | h(x) = 1}.
Show that ker(h) = Z.

14
(b) Using the first isomorphism Theorem for groups, show that R/Z is isomorphic to the
unit circle S 1 .

2.13. Using gymnastics with roots of unity, we recall here a very classical solution to the
problem of finding the roots of a cubic polynomial.

(a) Let ω := e2πi/3 , and show that we have the polynomial identity:

(x + a + b)(x + ωa + ω 2 b)(x + ω 2 a + ωb) = x3 − 3abx + a3 + b3 .

(b) Using the latter identity, solve the cubic polynomial: x3 − px + q = 0 by substituting
p = 3ab and q = a3 + b3 .

2.14. Thinking of the function sin(πz) as a function of a complex variable z ∈ C, show that
its zeros are precisely the set of integers Z.

2.15. Here we give another equivalent condition for a rectangle in Theorem 2.1 to be a nice
rectangle, using the same definitions as before.

Let’s call ξ ∈ Z2 a generic integer point if ξ is not orthogonal to any of the edges of R. In
other words, a generic integer vector satisfies hξ, pi 6= 0, for all p ∈ R, and in particular
p = 0 is not generic, nor is any point p on the x-axis or the y-axis. Then

R is a nice rectangle ⇐⇒ 1̂R (ξ) = 0, for all generic points ξ ∈ Z2 . (2.18)

2.16 (Erdös, 1951). Erdös asked: “Can the set Z>0 of all positive integers be partitioned
(that is, written as a disjoint union) into a finite number of arithmetic progressions, such
that no two of the arithmetic progressions will have the same common difference?”

Suppose that we have a list of disjoint arithmetic progressions, each with its common differ-
ence ak :
{a1 n + b1 | n ∈ Z}, . . . , {aN n + bN | n ∈ Z},
where a1 ≤ a2 ≤ · · · ≤ aN , and N ≥ 2. Prove that in any such partitioning of the integers,
there are at least two arithmetic progressions that have the same maximal aN .

Notes. For example, if we write Z = {4n + 1 | n ∈ Z} ∪ {2n | n ∈ Z} ∪ {4n + 3 | n ∈ Z},


a disjoint union of 3 arithmetic progressions, then we see that the common difference of 4
appears twice. Erdös noticed that such a phenomenon must always occur. (See also Exercise
6.26 for an extension to lattices in Rd ).

15
Chapter 3

Examples that nourish the theory

A pint of example is worth a gallon of advice


– Anonymous

Figure 3.1: The first periodic Bernoulli polynomial P1 (x), sometimes called the sawtooth
function, which turns out to be one of the building blocks of integer point enumeration in
polytopes

3.1 Intuition

One way to think about the Fourier transform of a polytope P ⊂ Rd is that it simultaneously
captures all of the moments of P, thereby uniquely defining P. Here we begin concretely by

16
computing some Fourier transforms of various polytopes in dimensions 1 and 2, as well as
the Fourier transforms of some simple families of polytopes in dimension d as well.

The 2-dimensional computations will get the reader more comfortable with the basics. In
later chapters, once we learn a little more theory, we will return to these families of polytopes
and compute some of their Fourier transforms in general.

We also see, from small examples, that the Bernoulli polynomials immediately enter into
the picture, forming natural building blocks. In this chapter we compute Fourier transforms
without thinking too much about convergence issues, to let the reader run with the ideas. But
commencing with the next chapter, we will be more rigorous when using Poisson summation,
and with convergence issues.

3.2 Dimension 1 - the classical sinc function

We begin by computing the classical 1-dimensional example of the Fourier transform of the
symmetrized unit interval P := [− 12 , 21 ]:

Z
1̂P (ξ) := 1P (x) e−2πixξ dx (3.1)
ZR
= e−2πixξ dx (3.2)
[− 12 , 12 ]

e ( )ξ − e−2πi( −1
−2πi 21 2 )
ξ
= (3.3)
−2πiξ
cos(−πξ) + i sin(−πξ) − (cos(πξ) + i sin(πξ))
= (3.4)
−2πiξ
sin(πξ)
= , (3.5)
πξ
valid for all ξ 6= 0. The latter function is also known as the sinc function. We notice that
ξ = 0 is a removable singularity, so that we may define the continuous sinc-function by
(
sin(πx)
πx
, if x 6= 0
sinc(x) := (3.6)
1 if x = 0,
which is in fact infinitely smooth, via Lemma 3.1 below. Next, we introduce the inverse
Fourier transform, or as it is often called, the Fourier inversion formula: property:
(F ◦ F)(f )(ξ) = f (−ξ), (3.7)
an extremely useful tool (see [155]).

17
Figure 3.2: The function sinc(x), which is Fourier transform of the 1-dimensional polytope
P = [− 12 , 21 ].

Example 3.1. A famous and historically somewhat tricky integral formula for the sinc
function is the following fact:
Z ∞ Z ∞
sin(πx)
sinc(x)dx := dx = 1, (3.8)
−∞ −∞ πx
also known as ‘the Dirichlet integral’. The careful reader notices that the latter integrand
R ∞ sin(πx)
is not absolutely convergent, which means that −∞ πx dx = ∞ (Exercise 3.22). So we
have to specify what we really mean by the identity (3.8). The rigorous claim is:
Z N
sin(πx)
lim dx = 1.
N →∞ −N πx
Let’s see an intuitive derivation of (3.8), using (3.1). We saw above that the Fourier transform
of the indicator function of the interval P := [− 12 , 12 ] is:
sin(πξ)
F(1P )(ξ) = , (3.9)
πξ
so that  
sin(πξ)
F = (F ◦ F)(1P )(ξ) = 1P (−ξ). (3.10)
πξ
Using the definition of the Fourier transform, the latter identity is:
Z
sin(πx) −2πiξx
e dx = 1P (ξ), (3.11)
R πx
and now evaluating both sides at ξ = 0 gives us (3.8).

Why did we say this is intuitive and not yet rigorous? It is because we applied the Fourier
inversion formula to a function that was not in L1 (R), namely the sinc function.

18
3.3 The inverse Fourier transform

We want to use the fundamental fact that the Fourier transform is invertible, and more
precisely thatR its inverse has the following form. We call a function f : Rd → C absolutely
integrable if Rd |f (x)|dx < ∞, and we write this as f ∈ L1 (Rd ).

Theorem 3.1. Given a function f such that both f ∈ L1 (Rd ) and fˆ ∈ L1 (Rd ), we have

(F ◦ F)f (x) = f (−x), (3.12)

for all x ∈ Rd .

Identity 3.12 tells us that the inverse Fourier transform F −1 exists, and is almost F itself. A
moment’s thought reveals that we may rewrite (3.12) in the somewhat more familiar form:
Z
f (x) = fˆ(ξ)e2πihξ,xi dξ. (3.13)
Rd

We can extend Example 3.1 in a natural way to all Fourier pairs of functions, {f (x), fˆ(ξ)},
as follows. Simply let x = 0 in (3.13), to get:
Z
f (0) = fˆ(x)dx. (3.14)
Rd

To summarize, Example 3.1 is simply identity (3.14) with f (x) := 1[− 1 , 1 ] (x).
2 2

Another nice - and very useful - fact about the Fourier transform of a polytope is that it
is an entire function, meaning that it is differentiable everywhere. This differentiability is
already observable in the sinc function above, with its removable singularity at the origin.

Lemma 3.1. Let P ⊂ Rd be a d-dimensional polytope. Then 1̂P (ξ) is an entire function of
ξ ∈ Cd .

Proof. Because P is compact, we can safely differentiate under the integral sign (this is a
special case of Lebesgue’sR Dominated Convergence Theorem). Namely, for any coordinate
variable ξk , we have: dξdk P e−2πihξ,xi dx = P dξdk e−2πihξ,xi dx = 2πi P xk e−2πihξ,xi dx, and it is
R R

clear that all possible derivatives exist in this manner, because the integrand is infinitely
smooth.

We also have the very fortuitous fact that the Fourier transform of any polytope P ⊂ Rd
is a complete invariant, in the following sense. We recall that by definition a polytope is in
particular a closed set.

19
Lemma 3.2. Let P ⊂ Rd be a polytope. Then 1̂P (ξ) uniquely determines P. In other words,
given any two closed d-dimensional polytopes P, Q ⊂ Rd , we have

1̂P (ξ) = 1̂Q (ξ) for all ξ ∈ Rd ⇐⇒ P = Q.

In other words, for any polytope P, its Fourier transform 1̂P uniquely determines the polytope.

Proof. If P = Q, it is clear that 1̂P (ξ) = 1̂Q (ξ) for all ξ ∈ Rd . Conversely, suppose that
1̂P (ξ) = 1̂Q (ξ) for all ξ ∈ Rd . Using Fourier inversion, we may take the Fourier transform of
both sides of the latter equation to get 1P (−ξ) = 1Q (−ξ), for all ξ ∈ Rd .

A fascinating consequence of Lemma 3.2 is that when we take the Fourier transform of a
polytope, then all of the combinatorial and geometric information of P is contained
in the formula of its transform, somewhere in there...... So we may begin to create a complete
dictionary between the geometry and combinatorics of a polytope (in the space domain), and
its Fourier transform (in the frequency domain).

3.4 Bernoulli polynomials

We introduce the Bernoulli polynomials, which turn out to be a sort of “glue” between
discrete geometry and number theory, as we will see throughout the book. The Bernoulli
polynomials are defined via the following generating function:

text X tk
= Bk (x) . (3.15)
et − 1 k=0 k!

It’s fruitful to sometimes restrict the Bernoulli polynomials to the unit interval [0, 1], and
then periodize them. In other words, using

{x} := x − bxc,

the fractional part of x, we may define the n’th periodic Bernoulli polynomial:

Pn (x) := Bn ({x}), (3.16)

for n ≥ 2. Since Pn (x) is periodic on R with period 1, it has a Fourier series, and in fact:

n! X e2πikx
Pn (x) = − , (3.17)
(2πi)n kn
k∈Z−{0}

20
valid for x ∈ R (Exercise 3.9).

When n = 1, we have the first Bernoulli polynomial


1
P1 (x) := x − bxc − ,
2
which is very special (see Figure 3.1). For one thing, it is the only periodic Bernoulli
polynomial that is not continuous everywhere, and we note that its Fourier series does not
converge absolutely, although it is quite appealing:
1 X e2πikx
P1 (x) = − , (3.18)
2πi k
k∈Z−{0}

valid for all x ∈


/ Z. Hence special care must be taken with P1 (x). Exercises 3.4 through 3.17
illustrate some of the important properties of these polynomials. Exercise 3.33 provides a
rigorous proof of the convergence of (3.18).
Example 3.2. The first few Bernoulli polynomials are:
B0 (x) = 1 (3.19)
1
B1 (x) = x − (3.20)
2
1
B2 (x) = x2 − x + (3.21)
6
3 1
B3 (x) = x3 − x2 + x (3.22)
2 2
1
B4 (x) = x4 − 2x3 + x2 − (3.23)
30
5 5 1
B5 (x) = x5 − x4 + x3 − x (3.24)
2 3 6
5 1 1
B6 (x) = x6 − 3x5 + x4 − x2 + (3.25)
2 2 42
The Bernoulli numbers are defined to be the constant terms of the Bernoulli polynomials:
Bk := Bk (0).
The first few Bernoulli numbers are:
1 1 1 1
B0 = 1, B1 = − , B2 = , B3 = 0, B4 = − , B5 = 0, B6 = .
2 6 30 42

It follows quickly from the definition 3.15 above that for odd k ≥ 3, Bk = 0 (Exercise 3.15).
From the generating function 3.15 the Bernoulli numbers are defined via

X tk
t
= Bk . (3.26)
et − 1 k=0 k!


21
Historically, the first appearance of the Bernoulli polynomials occurred while Jakob Bernoulli
tried to compute sums of powers of integers. In particular, Bernoulli showed that:
n−1
X Bd (n) − Bd
k d−1 = ,
k=1
d

for all integers d ≥ 1 and n ≥ 2 (Exercise 3.8). An interesting identity that allows us to
compute the Bernoulli numbers recursively rather quickly is:
n  
X n+1
Bk = 0,
k=0
k

valid for all n ≥ 1 (Exercise 3.17).

Some of the most natural, and beautiful, Fourier series arise naturally from the periodized
Bernoulli polynomials. The following intuitive application of the Poisson summation formula
already suggests an initial connection between periodized Bernoulli polynomials and Fourier
transforms of polytopes - even in dimension 1.
Example 3.3 (Intuitive Poisson summation). In this example we allow ourselves to be
completely intuitive, and unrigorous at this moment, but often such arguments are useful in
pointing us to their rigorous counterparts. Consider the 1-dimensional polytope P := [a, b],
and restrict attention to the case of a, b 6∈ Z. If we could use the Poisson summation formula
X X
f (n) = fˆ(ξ),
n∈Zd ξ∈Zd

applied to the function f (x) := 1P (x), then we would get:


X X
1P (n)“ = ” 1̂P (ξ)
n∈Z ξ∈Z
X e−2πiξb − e−2πiξa
“ = ” 1̂P (0) +
−2πiξ
ξ∈Z−{0}

1 X e−2πiξb 1 X e−2πiξa
“ = ” (b − a) − +
2πi ξ 2πi ξ
ξ∈Z−{0} ξ∈Z−{0}
2πiξb
1 X e 1 X e2πiξa
“ = ” (b − a) + −
2πi ξ 2πi ξ
ξ∈Z−{0} ξ∈Z−{0}
   
1 1
“ = ” (b − a) − {b} − + {a} −
2 2
“ = ” b − {b} − (a − {a}) = bbc − bac.
Since
P we already know how to evaluate the LHS of Poisson summation above, namely that
n∈Z 1P (n) = # {Z ∩ P} = bbc − bac, we have confirmed that Poisson summation has given

22
us here the correct formula, in spite of the lack of rigor here. Why is the intuitive argument
above not rigorous yet? In order to plug a function f into Poisson summation, and consider
convergence at each point of the domain, f and its Fourier transform fˆ must both satisfy
some growth conditions at infinity, at least ensuring proper convergence of both sides of the
Poisson summation formula. We will see such conditions later, in Chapter 4, section 4.23.
Once we learn how to use Poisson summation, we will return to this example (see Example
8.3). 
P P
We recall that a series n∈Z an is said to converge absolutely if n∈Z |an | converges. It’s
easy to see that the series in (3.18) for P1 (x) does not converge absolutely. Such convergent
series that do not converge absolutely are called conditionally convergent.

To prove rigorously that the conditionally convergent series (3.18) does in fact converge, see
Exercises 3.29, 3.30, 3.32, and 3.33, which include the Abel summation formula, and the
Dirichlet convergence test (although extremely useful, we will not use them that much in
the ensuing chapters).

3.5 The cube, and its Fourier transform

Perhaps the easiest way to extend the Fourier transform of the unit interval is to consider
the d-dimensional unit cube  d
1 1
 := − , .
2 2
What is its Fourier transform? When we compute a Fourier n transform of a functiono f , we
will say that {f, fˆ} is a Fourier pair. We have seen that 1[− 1 , 1 ] (x), sinc(ξ) is a Fourier
2 2
pair in dimension 1.

Example 3.4. Due to the fact that the cube is the direct product of line segments, it follows
that the ensuing integral can be separated into a product of integrals, and so it is the product
of 1-dimensional transforms:
Z
1̂ (ξ) = 1 (x)e−2πihx,ξi dx (3.27)
R d
Z
= e−2πi(x1 ξ1 +···+xd ξd ) dx (3.28)

d Z 1
Y 2
= e−2πixk ξk dxk (3.29)
k=1 − 12
d
Y sin(πξk )
= , (3.30)
k=1
πξk

23
valid for all ξ ∈ Rd such that none of their coordinates vanishes. So here we have the Fourier
pair ( )
d
Y sin(πξk )
1 (x), .
k=1
πξk

In general, though, polytopes are not a direct product of lower-dimensional polytopes, so we


will need to develop more tools to compute their Fourier transforms. 

3.6 The simplex, and its Fourier transform

Another basic building block for polytopes is the standard simplex, defined by

:= x ∈ Rd x1 + · · · + xd ≤ 1, and all xk ≥ 0 .

(3.31)

x
0

Figure 3.3: The standard simplex in R2

Example 3.5. Just for fun, let’s compute the Fourier transform of 4 for d = 2, via brute-
force. We may use the following parametrization (called a hyperplane description) for this
standard triangle: 
= (x, y) x + y ≤ 1, and x ≥ 0, y ≥ 0 .

24
Hence, we have:
Z 
1̂ (ξ1 , ξ2 ) := e−2πi xξ1 +yξ2 dxdy
Z 1 Z y=1−x 
−2πi xξ1 +yξ2
= e dydx
0 y=0
Z 1  −2πiyξ2 
−2πixξ1 e y=1−x
= e dx
−2πiξ2 y=0

0
Z 1
1
e−2πixξ1 e−2πi(1−x)ξ2 − 1 dx

=
−2πiξ2 0
Z 1
1
e−2πix(ξ1 −ξ2 ) e−2πiξ2 − e−2πixξ1 dx

=
−2πiξ2 0
1 e−2πiξ2 −2πi(ξ1 −ξ2 ) 1 e−2πiξ1 − 1
= (e − 1) −
(−2πi)2 ξ2 (ξ1 − ξ2 ) (−2πi)2 ξ1 ξ2
 −2πiξ1 −2πiξ2 −2πiξ1

1 e −e e −1
= − .
(−2πi)2 ξ2 (ξ1 − ξ2 ) ξ1 ξ2

We may simplify further by noticing the rational function identity

e−2πiξ1 e−2πiξ1 e−2πiξ1


− = ,
ξ2 (ξ1 − ξ2 ) ξ1 ξ2 ξ1 (ξ1 − ξ2 )

giving us the symmetric function of (ξ1 , ξ2 ):


 −2πiξ1
e−2πiξ2

1 e 1
1̂ (ξ1 , ξ2 ) = + + . (3.32)
(−2πi)2 ξ1 (ξ1 − ξ2 ) ξ2 (ξ2 − ξ1 ) ξ1 ξ2

We need the concept of a convex set X ⊂ Rd , defined by the property that for any two
points x,y ∈ X, the line
segment joining them also lies in X. In other words, the line
segment λx + (1 − λ)y 0 ≤ λ ≤ 1 ⊂ X, ∀x, y ∈ X.

Given any finite set of points S := {v1 , v2 , . . . , vN } ⊂ Rd , we can also form the set of all
convex linear combinations of S by defining
( N
)
X
conv(S) := λ1 v1 + λ2 v2 + · · · + λN vN λk = 1, where all λk ≥ 0 . (3.33)
k=1

Given any set U ⊂ Rd (which is not restricted to be finite), we define the convex hull of U
as the set of convex linear combinations, taken over all finite subsets of U , and denoted by
conv(U ).

25
We define a polytope as the convex hull of any finite set of points in Rd . This definition of
a polytope is called its vertex description. We define a k-simplex ∆ as the convex hull
of a finite set of vectors {v1 , v2 , . . . , vk+1 }:

∆ := conv{v1 , v2 , . . . , vk+1 },

where 0 ≤ k ≤ d, and v2 − v1 , v3 − v1 , . . . , vk+1 − v1 are linearly independent vectors in Rd .


The points v1 , v2 , . . . , vk+1 are called the vertices of ∆, and this object is one of the basic
building-blocks of polytopes, especially when triangulating a polytope.

The simplex ∆ is a k-dimensional polytope, sitting in Rd . When k = d, the dimension of ∆


equals the dimension of the ambient space Rd - see Figure 3.4. We have already computed

Figure 3.4: A 3-simplex and its faces, which are lower-dimensional simplices as well

the Fourier transform of a particular 2-simplex, in (3.32).

More generally, let’s compute the Fourier transform of any 2-simplex in R2 . In order to
handle a general triangle, let ∆ be any triangle in the plane, with vertices
a a a
v1 := ( b11 ) , v2 := ( b22 ) , v3 := ( b33 ) .

Can we reduce the computation of 1̂∆ to our already known formula for 1̂ , given by (3.32)?
We first notice (after a brief cup of coffee) that we can map any triangle in the plane to the
standard triangle, by using a linear transformation followed by a translation:

∆ = M ( ) + v3 , (3.34)

26
where M is the 2 × 2 matrix whose columns are v1 − v3 and v2 − v3 . We are now ready to
compute the Fourier transform of a general triangle ∆:
Z Z
−2πihξ,xi
1̂∆ (ξ) = e dx = e−2πihξ,xi dx.
∆ M ( )+v3

Making the substitution x := M y + v3 , with y ∈ , we have dx = | det M |dy, and so


Z Z
−2πihξ,xi
e dx = | det M | e−2πihξ,M y+v3 i dy
M ( )+v3
Z
−2πihξ,v3 i T
= | det M |e e−2πihM ξ,yi dy

= | det M |e−2πihξ,v3 i 1̂ (M T ξ)
= | det M |e−2πihξ,v3 i 1̂

hv1 − v3 , ξi, hv2 − v3 , ξi
 −2πiz1
e−2πiz2

−2πihξ,v3 i 1 e 1
= | det M |e + + ,
(−2πi)2 z1 (z1 − z2 ) z2 (z2 − z1 ) z1 z2
where we’ve used our formula (3.32) for the FT of the standard triangle (thereby bootstrap-
ping out way to the general case) with z1 := hv1 − v3 , ξi, and z2 := hv2 − v3 , ξi. Substituting
these values into the latter expression, we finally arrive at the FT of our general triangle ∆:
e−2πihv1 ,ξi e−2πihv2 ,ξi e−2πihξ,v3 i
 
| det M |
1̂∆ (ξ) = (−2πi)2 + + .
hv1 − v3 , ξihv1 − v2 , ξi hv2 − v3 , ξihv2 − v1 , ξi hv3 − v1 , ξihv3 − v2 , ξi
(3.35)

We can notice in equation (3.35) many of the same patterns that had already occurred in
Example 3.8. Namely, the Fourier transform of a triangle has denominators that are products
of linear forms in ξ, and it is a finite linear combination of rational functions multiplied by
complex exponentials.

Also, in the particular case of equation (3.35), 1̂∆ (ξ) is a symmetric function of v1 , v2 , v3 , as
we might have expected.

Using exactly the same ideas that were used in equation (3.35), it is possible to prove (by
induction on the dimension) that the Fourier transform of a general d-dimensional simplex
∆ ⊂ Rd is:
N
X e−2πihvj ,ξi
1̂∆ (ξ) = (vol ∆)d! Qd [k 6= j], (3.36)
j=1 k=1 hvj − v k , ξi
where the vertex set of P is {v1 , . . . , vN } (Exercise 3.27), and in fact the same formula
persists for all complex ξ ∈ Cd such that the products of linear forms in the denominators
do not vanish.

However, looking back at the computation leading to (3.35), and the corresponding compu-
tation which would give (3.36), the curious reader might be thinking:

27
“There must be an easier way!”

But never fear - indeed there is. So even though at this point the computation of 1̂∆ (ξ)
may be a bit laborious (but still interesting), computing the Fourier transform of a general
simplex will become quite easy once we will revisit it in a later chapter (see Theorem 7.1).

3.7 Stretching and translating

The perspicacious reader may have noticed that in order to arrive at the formula (3.35) above
for the FT of a general triangle, we exploited the fact that the Fourier transform interacted
peacefully with the linear transformation M , and with the translation by the vector v. Is
this true in general?

Indeed it is, and we record these thoughts in the following two lemmas, which will become our
bread and butter for future computations. In general, given any invertible linear transfor-
mation M : Rd → Rd , and any function f : Rd → C whose FT (Fourier transform) exists, we
have the following useful interaction between Fourier transforms and linear transformations.

Lemma 3.3 (Stretch).


1
fˆ M −T ξ

◦ M )(ξ) =
(f\ (3.37)
| det M |

◦ M )(ξ) := Rd f (M x)e−2πihξ,xi dx. We perform the change


R
Proof. By definition, we have (f\
of variable y := M x, implying that dy = | det M |dx, so that:
Z
1 −1
(f ◦ M )(ξ) =
\ f (y)e−2πihξ,M yi dy
| det M | Rd
Z
1 −T
= f (y)e−2πihM ξ,yi dy
| det M | Rd
1
fˆ M −T ξ .

=
| det M |

What about translations? They are even simpler.

Lemma 3.4 (Translate). For any translation T (x) := x + v, where v ∈ Rd is a fixed vector,
we have
(f[ ◦ T )(ξ) = e2πihξ,vi fˆ(ξ). (3.38)

28
◦ T )(ξ) := Rd f (T x)e−2πihξ,xi dx, so that performing
R
Proof. Again, by definition we have (f[
the simple change of variable y = T x := x + v, we have dy = dx. The latter integral becomes
Z
(f ◦ T )(ξ) =
[ f (y)e−2πihξ,y−vi dy
R d
Z
=e 2πihξ,vi
f (y)e−2πihξ,yi dy := e2πihξ,vi fˆ(ξ).
Rd

In general, any function φ : Rd → C of the form


φ(x) = M x + v, (3.39)
where M is a fixed linear transformation and v ∈ Rd is a fixed vector, is called an affine
transformation. For example, we’ve already seen in (3.34) that the right triangle was
mapped to the more general triangle ∆ by an affine transformation. So the latter two lemmas
allow us to compose Fourier transforms very easily with affine transformations.
Example 3.6. Considering any measurable set B ⊂ Rd , let’s translate B by a fixed vector
v ∈ Rd , and compute 1̂B+v (ξ).

We note that because 1B+v (ξ) = 1B (ξ − v), the translate lemma applies, but with a minus
sign. That is, we can use T (x) := x − v and f := 1B to get:
−2πihξ,vi
1̂B+v (ξ) = (1\
B ◦ T )(ξ) = e 1̂B (ξ). (3.40)


3.8 The parallelepiped, and its Fourier transform

Now that we know how to compose the FT with affine transformations (translations and
linear transformations), we can easily find the FT of any parallelepiped in Rd by using our
d
formula for the Fourier transform of the unit cube  := − 21 , 12 , which we derived in


Example 3.4:
d
Y sin(πξk )
1̂ (ξ) = , (3.41)
k=1
πξk
d
for all ξ ∈ R such that all the coordinates of ξ do not vanish. First, we translate the cube
 by the vector ( 12 , · · · , 12 ), to obtain
 
1 1
C :=  + , ··· , = [0, 1]d .
2 2

29
It’s straightforward to compute its FT as well (Exercise 3.2), by using Lemma 3.4, the
‘translate’ lemma:
d
1 Y 1 − e−2πiξk
1̂C (ξ) = . (3.42)
(2πi)d k=1 ξk

Next, we define a d-dimensional parallelepiped P ⊂ Rd as an affine image of the unit cube.


In other words, any parallelepiped has the description

P = M (C) + v,

for some linear transformation M , and some translation vector v. Geometrically, the cube
is stretched and translated into a parallelepiped.

Figure 3.5: Mapping the unit cube to a parallelepiped

For the sake of concreteness, will will first set v := 0 and compute the Fourier transform
of P := M (C), where we now give M as a d × d invertible matrix whose columns are
w1 , w2 , . . . , wd . Because the cube C may be written as a convex linear combination of the
basis vectors ej , we see that P may be written as a convex linear combination of M ej = wj .
In other words, we see that the parallelepiped P has the equivalent vertex description:
( d )
X
P= λk wk all λk ∈ [0, 1] .
k=1

30
To review the basics, let’s compute the FT of our parallelepiped P from first principles:
Z Z
−2πihξ,xi
1̂P (ξ) := e dx = e−2πihξ,xi dx (3.43)
P M (C)
Z
= | det M | e−2πihξ,M yi dy (3.44)
ZC
T
= | det M | e−2πihM ξ,yi dy := | det M | 1̂C M T ξ

(3.45)
C
d
| det M | 1 − e−2πihwk ,ξi
Y
= . (3.46)
(2πi)d k=1 hwk , ξi

where in the third equality we used the substitution x := M y, with y ∈ C, yielding dx =


| det M |dy. In the last equality, we used our known formula (3.42) for the FT of the cube
C, together with the elementary linear algebra fact that the k’th coordinate of M T ξ is given
by hwk , ξi.

Finally, for a general parallelepiped, we have Q := P + v, so that by definition


( d
)
X
Q= v+ λk wk all λk ∈ [0, 1] .
k=1

Noting that 1P+v (ξ) = 1P (ξ − v), we compute the Fourier transform of Q by using the
‘translate lemma’ (Lemma 3.4), together with formula (3.46) for the Fourier transform of P:
d
| det M | Y 1 − e−2πihwk ,ξi
1̂Q (ξ) = e−2πihξ,vi , (3.47)
(2πi)d k=1
hwk , ξi

for all ξ ∈ Rd , except for those ξ that are orthogonal to one of the wk (which are edge vectors
for Q).

Example 3.7. A straightforward computation shows that if we let v := − w1 +···+w 2


d
, then
Pd
Q := {v + k=1 λk wk | all λk ∈ [0, 1]} is symmetric about the origin, in the sense that
x ∈ Q ⇐⇒ −x ∈ Q (Exercise 3.24). In other words, the center of mass of this new Q is
now the origin. Geometrically, we’ve translated the previous parallelepiped by using half its
‘body diagonal’. For such a parallelepiped Q, centered at the origin, formula (3.47) above

31
gives
d
2πihξ,
w1 +···+wd
i | det M | Y 1 − e−2πihwk ,ξi
1̂Q (ξ) = e 2 (3.48)
(2πi)d k=1 hwk , ξi
d
| det M | Y eπihwk ,ξi − e−πihwk ,ξi
= (3.49)
(2πi)d k=1 hwk , ξi
d
| det M | Y (2i) sin(πhwk , ξi)
= (3.50)
(2πi)d k=1 hwk , ξi
d
Y sin(πhwk , ξi)
= | det M | . (3.51)
k=1
πhwk , ξi
To summarize, for a parallelepiped that is symmetric about the origin, we have the Fourier
pair ( )
d
Y sin(πhwk , ξi)
1Q (x), | det M | .
k=1
πhw k , ξi

We could have also computed the latter FT by beginning with our known Fourier transform
(3.41) of the cube , composing the FT with the same linear transformation M of (3.43),
and using the ‘stretch’ lemma, so everything is consistent. 

3.9 The cross-polytope

Another natural convex body in R2 is the cross-polytope



♦2 := (x1 , x2 ) ∈ R2 |x1 | + |x2 | ≤ 1 .

(3.52)

In dimension d, the cross-polytope ♦d can be defined similarly by its hyperplane de-


scription
♦d := (x1 , x2 , . . . , xd ) ∈ Rd |x1 | + |x2 | + · · · + |xd | ≤ 1 .

(3.53)
The cross-polytope is also, by definition, the unit ball in the L1 -norm on Euclidean space,
and from this perspective a very natural object. In R3 , the cross-polytope ♦3 is often called
an octahedron.

In this section we only work out the 2-dimensional case of the Fourier transfrom of the
crosspolytope, In Chapter 7, we will work out the Fourier transform of any d-dimensional
cross-polytope, 1̂♦d , because we will have more tools at our disposal.

Nevertheless, it’s instructive to compute 1̂♦2 via brute-force for d = 2 here, in order to gain
some practice.

32
x2

1 x1

1
x3

Figure 3.6: The cross-polytope ♦ in R3 (courtesy of David Austin)

Example 3.8. Using the definition of the Fourier transform, we first compute the FT of the
2-dimensional cross polytope:
Z
1̂♦2 (ξ) := e−2πihξ,xi dx. (3.54)
♦2

In R2 , we may write ♦2 as a union of the following 4 triangles:

∆1 := conv(( 00 ) , ( 10 ) , ( 01 ))
∆2 := conv(( 00 ) , ( −1 0
0 ) , ( 1 ))
∆3 := conv(( 00 ) , ( −1 0
0 ) , ( −1 ))
0
∆4 := conv(( 00 ) , ( 10 ) , ( −1 )).

Since these four triangles only intersect in lower-dimensional subsets of R2 , the 2-dimensional
integral vanishes on such lower dimensional subsets, and we have:

1̂♦2 (ξ) = 1̂∆1 (ξ) + 1̂∆2 (ξ) + 1̂∆3 (ξ) + 1̂∆4 (ξ). (3.55)

Recalling from equation (3.32) of example 3.5 that the Fourier transform of the standard
simplex ∆1 is
2 
e−2πiξ1 e−2πiξ2
 
1 1
1̂∆1 (ξ) = + + , (3.56)
2πi ξ1 ξ2 (−ξ1 + ξ2 )ξ1 (ξ1 − ξ2 )ξ2

we can compute 1̂∆2 (ξ), by reflecting ∆2 about the x2 − axis (the Jacobian of this transfor-

33
mation is 1), and using the already-computed transform (3.56) of ∆1 :
Z
1̂∆2 (ξ1 , ξ2 ) := e−2πi(x1 ξ1 +x2 ξ2 ) dx
Z ∆2
= e−2πi(−x1 ξ1 +x2 ξ2 ) dx
Z∆1
= e−2πi(x1 (−ξ1 )+x2 ξ2 ) dx
∆1

= 1̂∆1 (−ξ1 , ξ2 )).

Similarly, we have 1̂∆3 (ξ1 , ξ2 ) = 1̂∆1 (−ξ1 , −ξ2 ), and 1̂∆4 (ξ1 , ξ2 ) = 1̂∆1 (ξ1 , −ξ2 ).

Hence we may continue the computation from equation 3.55 above, putting all the pieces
back together:

1̂♦2 (ξ) = 1̂∆1 (ξ1 , ξ2 ) + 1̂∆1 (−ξ1 , ξ2 ) + 1̂∆1 (−ξ1 , −ξ2 ) + 1̂∆1 (ξ1 , −ξ2 ) (3.57)
2 
− e2πiξ1 − e2πiξ2
 
1 1
= + + (3.58)
2πi ξ1 ξ2 (−ξ1 + ξ2 )ξ1 (ξ1 − ξ2 )ξ2
2 
e−2πiξ1 e2πiξ2
 
1 −1
+ + + (3.59)
2πi ξ1 ξ2 (ξ1 + ξ2 )ξ1 (ξ1 + ξ2 )ξ2
2 
e−2πiξ1 e−2πiξ2
 
1 1
+ + + (3.60)
2πi ξ1 ξ2 (ξ1 − ξ2 )ξ1 (−ξ1 + ξ2 )ξ2
2 
e2πiξ1 e−2πiξ2
 
1 −1
+ + + (3.61)
2πi ξ1 ξ2 (ξ1 + ξ2 )ξ1 (ξ1 + ξ2 )ξ2
 
1 cos(2πξ1 ) cos(2πξ2 ) cos(2πξ1 ) cos(2πξ2 )
=− 2 + + + (3.62)
4π (ξ1 − ξ2 )ξ1 (−ξ1 + ξ2 )ξ2 (ξ1 + ξ2 )ξ1 (ξ1 + ξ2 )ξ2
 
1 cos(2πξ1 ) − cos(2πξ2 )
=− 2 . (3.63)
2π (ξ1 + ξ2 )(ξ1 − ξ2 )

There is another fundamental relationship between the cross-polytope and the cube P :=
[−1, 1]d . To see this relationship, we define, for any polytope P ⊂ Rd , its dual polytope:

P ∗ := x ∈ Rd hx, yi ≤ 1, for all y ∈ P .



(3.64)

It is an easy fact (Exercise 3.25) that the cross-polytope and the cube P := [−1, 1]d are dual
to each other, as in the figure below.

34
Figure 3.7: The cube and the cross-polytope are duals of each other

3.10 Observations and questions

Now we can make several observations about all of the formulas that we found so far, for
the Fourier transforms of various polytopes. For the 2-dimensional cross-polytope, we found
that  
1 cos(2πξ1 ) − cos(2πξ2 )
1̂♦2 (ξ) = − 2 . (3.65)
2π (ξ1 + ξ2 )(ξ1 − ξ2 )

(a) It is real-valued for all ξ ∈ R2 , and this is due to the fact that ♦2 is symmetric about
the origin (see section 5.6).
Question 1. Is it true that any symmetric property of a polytope P is somehow mir-
rored by a corresponding symmetric property of its Fourier transform?

Although this question is not well-defined at the moment (it depends on how we define
‘symmetric property’), it does sound exciting, and we can morph it into a few well-
defined questions later.
(b) The only apparent singularities of the FT in (3.65) (though they are in fact removable
singularities) are the two lines ξ1 − ξ2 = 0 and ξ1 + ξ2 = 0, and these two lines are
perpendicular to the facets of ♦2 , which is not a coincidence (see Chapter 11).
(c) It is always true that the Fourier transform of a polytope is an entire function, by
Lemma 3.1, so that the singularities in the denominator (ξ1 + ξ2 )(ξ1 − ξ2 ) of (3.65)
must be removable singularities!
(d) The denominators of all of the FT’s so far are always products of linear forms in ξ.
Question 2. Is it true that the Fourier transform of any polytope is always a finite
sum of rational functions times an exponential, where the denominators of the rational
functions are always products of linear forms?

35
Answer: (spoiler alert) Yes! It’s too early to prove this here, but we will do so in
Theorem 7.2.

(e) We may retrieve the volume of ♦2 by letting ξ1 and ξ2 tend to zero (Exercise 3.21),
as always. Doing so, we obtain limξ→0 1̂♦2 (ξ) = 2 = Area(♦2 ).

Notes
(a) Another way to compute 1♦ (ξ) for the 2-dimensional cross-polytope ♦ is by starting
with the square [− 21 , 12 ]2 and applying a rotation of the plane by π/4, followed by a
simple dilation. Because we know that linear transformations interact in a very elegant
way with the FT, this method gives an alternate approach for the Example 3.8 in R2 .
However, this method no longer works for the cross-polytope in dimensions d ≥ 3,
where it is not (yet) known if there is a simple way to go from the FT of the cube to
the FT of the cross-polytope.
More generally, one may ask:
Question 3. is there a nice relationship between the FT of a polytope P and the FT
of its dual?

(b) We note that P1 (x) is defined to be equal to 0 at the integers, because its Fourier series
naturally converges to the mean of the discontinuity of the function, at each integer.

(c) It has been known since the work of Riemann that the Bernoulli numbers occur as
special values of the Riemann zeta function (see Exercise 4.6). Similarly, the Hurwitz
zeta function, defined for each fixed x > 0 by

X 1
ζ(s, x) := ,
n=0
(n + x)s

has a meromorphic continuation to all of C, and its special values at the negative
integers are the Bernoulli polynomials Bn (x) (up to a multiplicative constant).

(d) There are sometimes very unusual (yet useful) formulations for the Fourier transform
of certain functions. Ramanujan [120] discovered the following remarkable formula for
the Fourier transform of the Gamma function:

π Γ(a)Γ(a + 21 )
Z
−2πiξy
|Γ(a + iy)|e dy = , (3.66)
R cosh(πξ)2a

valid for a > 0. For example with a := 12 , in the language of this chapter we have the
Fourier pair {|Γ( 21 + iy)|, cosh(πξ)
π
}.

36
Exercises

Problems worthy of attack prove their worth by fighting back.


– Paul Erdös

3.1. ♣ Show that the Fourier transform of the closed interval [a, b] is:

e−2πiξa − e−2πiξb
1̂[a,b] (ξ) = ,
2πiξ
for ξ 6= 0.

3.2. Show that the Fourier transform of the unit cube C := [0, 1]d ⊂ Rd is:
d
1 Y 1 − e−2πiξk
1̂C (ξ) = , (3.67)
(2πi)d k=1 ξk

d
 all ξd ∈ R , except for the union of hyperplanes
valid for defined by
H := x ∈ R ξ1 = 0 or ξ2 = 0 . . . or ξd = 0 .

3.3. Suppose we are given two polynomials p(x) and q(x), of degree d. If there are d + 1
distinct points {z1 , . . . , zd+1 } in the complex plane such that p(zk ) = q(zk ) for k = 1, . . . , d+1,
show that the two polynomials are identical. (Hint: consider (p − q)(zk ))

3.4. To gain some facility with generating functions, show by a brute-force computation with
Taylor series that the coefficients on the right-hand-side of equation (3.15), which are called
Bn (x) by definition, must in fact be polynomials in x.

In fact, your direct computations will show that for all n ≥ 1, we have
n  
X n
Bn (x) = Bn−k xk ,
k=0
k

where Bj is the j’th Bernoulli number.

3.5. ♣ Show that for all n ≥ 1, we have

Bn (1 − x) = (−1)n Bn (x).

37
3.6. ♣ Show that for all n ≥ 1, we have
Bn (x + 1) − Bn (x) = nxn−1 .

3.7. ♣ Show that for all n ≥ 1, we have


d
Bn (x) = nBn−1 (x).
dx

3.8. ♣ Prove that:


n−1
X Bd (n) − Bd
k d−1 = ,
k=1
d
for all integers d ≥ 1 and n ≥ 2.

3.9. ♣ Show that the periodic Bernoulli polynomials Pn (x) := Bn ({x}), for all n ≥ 2, have
the following Fourier series:
n! X e2πikx
Pn (x) = − , (3.68)
(2πi)n k6=0 k n

valid for all x ∈ R. For n ≥ 2, these series are absolutely convergent. We note that from
the definition above, Bn (x) = Pn (x) when x ∈ (0, 1).

3.10. Show that the greatest integer function bxc (often called the ‘floor function’) enjoys
the property:
N −1  
X k
x+ = bN xc ,
k=0
N
for all x ∈ R, and all positive integers N , and that in the same range we also have
N −1  
X k
x+ = {N x} .
k=0
N

3.11. Show that the Bernoulli polynomials enjoy the following identity, proved by Joseph
Ludwig Raabe in 1851:
N −1  
n−1
X k
Bn (N x) = N Bn x + ,
k=0
N
for all x ∈ R, all positive integers N , and for each n ≥ 1.

Notes. Such formulas, in these last two exercises, are also called “multiplication Theorems”,
and they hold for many other functions, including the Gamma function, the dilogarithm, the
Hurwitz zeta function, and many more.

38
3.12. ♣ Here we give a different method for defining the Bernoulli polynomials, based on the
following three properties that they enjoy:

1. B0 (x) = 1.
d
2. For all n ≥ 1, dx Bn (x) = nBn−1 (x).
R1
3. For all n ≥ 1, we have 0 Bn (x)dx = 0.

Show that the latter three properties imply the original defining property of the Bernoulli
polynomials (3.15).

3.13. Here is a more explicit, useful recursion for computing the Bernoulli polynomials.
Show that
n−1  
X n
Bk (x) = nxn−1 ,
k=0
k
for all n ≥ 2.

3.14. Use the previous exercise, together with the known list the first 6 Bernoulli polynomials
that appear in equation 3.25, to compute B7 (x).

3.15. Show that for odd k ≥ 3, we have Bk = 0.

3.16. Show that the even Bernoulli numbers alternate in sign. More precisely, show that

(−1)n+1 B2n ≥ 0,

for each positive integer n.

3.17. Show that the Bernoulli numbers enjoy the recursive property:
n  
X n+1
Bk = 0,
k=0
k

for all n ≥ 1.

39
3.18. Show that the Bernoulli numbers enjoy the following asymptotics:

(2n)!
B2n ∼ 2
(2π)2n

as n → ∞. Here we are using the usual notation for asymptotic functions, namely that
f (n) ∼ g(n) as n → ∞ if limn→∞ fg(n)
(n)
→ 1.

3.19. ♣ Show that the following integrals converge and have the closed forms:
Z ∞ r
π
cos(x2 )dx = , (3.69)
−∞ 2
Z ∞ r
2 π
sin(x )dx = . (3.70)
−∞ 2

Notes. These integrals are called Fresnel integrals, and they are related to the Cornu spiral,
which was created by Marie Alfred Cornu. Marie used the spiral as a tool for computing
diffraction patterns that arise naturally in optics.

3.20. Prove the following Gamma function identity, using the sinc function:

sin(πx) 1
= ,
πx Γ(1 + x)Γ(1 − x)

for all x 6∈ Z.

Notes. This identity is often called Euler’s reflection formula.

3.21. ♣ Using the formula for the Fourier transform of the 2-dimensional cross-polytope ♦,
derived in the text, namely
 
1 cos(2πξ1 ) − cos(2πξ2 )
1̂♦ (ξ) = − 2 ,
2π ξ12 − ξ22

find the area of ♦ by letting ξ → 0 in the latter formula.

R ∞ sin(πx)
3.22. ♣ Show that −∞ πx dx = ∞.

40
3.23. There are (at least) two different ways of periodizing a given function f : R → C with
respect to Z. First, we can define F1 (x)
P := f ({x}), so that F1 is periodic on R with period 1.
Second, we may also define F2 (x) := n∈Z f (x + n), which is also a periodic function on R
with period 1.
R
Find an integrable (meaning that R f (x)dx converges) function f for which F1 6= F2 , as
functions.
P
Notes. In Chapter 4, we will see that the latter function F2 (x) := n∈Z f (x + n) captures
a lot more information about f , and often captures all of f as well.

3.24. Given linearly independent vectors w1 , . . . , wd ∈ Rd , let v := − w1 +···+w


2
d
, and define
Pd
Q := {v + k=1 λk wk | all λk ∈ [0, 1]}, a parallelepiped. Show that Q is symmetric about
the origin, in the sense that x ∈ Q ⇐⇒ −x ∈ Q.

3.25. ♣ Show that the d-dimensional cross-polytope ♦ and the cube  := [−1, 1]d are dual
to each other.

3.26. Prove the following 2-dimensional integral formula:

BeA − AeB
Z
1
eAλ1 eBλ2 dλ1 dλ2 = + , (3.71)
λ1 ,λ2 ≥0
λ +λ ≤1
AB(A − B) AB
1 2

valid for all A, B ∈ C such that AB(A − B) 6= 0.

3.27. Using the ideas of Example 3.35, prove (by induction on the dimension) that the
Fourier transform of a general d-dimensional simplex ∆ ⊂ Rd is given by:
N
X e−2πihvj ,ξi
1̂∆ (ξ) = (vol ∆)d! Q [k 6= j], (3.72)
j=1 1≤k≤d hvj − vk , ξi

for all ξ ∈ Rd , where the vertex set of P is {v1 , . . . , vN }.

3.28. Let’s answer the natural question: Does Euler’s formula eix = cos x+i sin x extend from
a real variable x, to a complex variable z? We assume here that ez , cos z, and sin P
z are defined
by the following absolutely convergent infinite series: ez := ∞ 1 n ∞ 1 2n
P
n=0 n! z , cos z := n=0 2n! z ,
and

41
3.29 (Abel summation by parts). ♣ Here we prove the straightforward but very useful tech-
nique of Niels Abel, called Abel summation by parts. Suppose Pwe are given two sequences
∞ ∞ n
{an }n=1 , and {bn }n=1 . We define the finite partial sums Bn := k=1 bk . Then we have
n
X n−1
X
ak bk = an Bn + Bk (ak − ak+1 ), (3.73)
k=1 k=1

for all n ≥ 2.

Notes. Using the forward difference operator, it’s easy to recognize identity (3.73) as a
discrete version of integration by parts.

3.30 (Dirichlet’s convergence test). ♣ Suppose we are given a real sequence {an }∞
n=1 ,

and a complex sequence {bn }n=1 , such that

(a) {an } is monotonically decreasing to 0, and


(b) | nk=1 bk | ≤ M , for some positive constant M , and all n ≥ 1.
P

P∞
Then k=1 ak bk converges.

3.31. Prove that for all x ∈ R − Z, we have the following important identity, called the
“Dirichlet kernel”, after Peter Gustav Lejeune Dirichlet:
n 1

X sin 2πx(n + )
e2πikx = 2
. (3.74)
k=−n
sin(πx)

3.32. For any fixed x ∈ R − Z, show that we have the bound on the following exponential
sum: n
X
2πikx 1
e ≤ . (3.75)

| sin(πx)|


k=1

P∞ e2πima
3.33. ♣ Prove that m=1 converges, given any fixed a ∈ R − Z.
m

Notes. We see that, although ∞ e2πima


P
m=1 m
does not converge absolutely, Abel’s summation
formula (3.73) gives us
n n n−1 k
X e2πika 1 X 2πira X  X 2πira  1
= e + e ,
k=1
k n r=1 k=1 r=1
k(k + 1)

and the latter series does converge absolutely. So we see that Abel summation transforms
one series (that barely converges at all) into another series that converges more rapidly.

42
Chapter 4

Tools of the trade:


basic Fourier analysis

“. . . Fourier’s great mathematical poem.”


[Referring to Fourier’s mathematical theory of the conduction of heat]

– William Thomson Kelvin

Figure 4.1: The unit cube  := [0, 1]3 , in R3 , which tiles the space by translations. Which
other polytopes tile by translations? How can we make mathematical use of such tilings? In
particular, can we give an explicit basis of exponentials for functions defined on ?

4.1 Intuition

Because we will use tools from Fourier analysis throughout, we introduce them here as an
outline of the field, with the goal of applying them to the discrete geometry of polytopes,
lattices, and their interactions.

43
In this chapter we develop, and sometimes quote, the basic and necessary tools of Fourier
analysis, so that we may tackle problems in the enumerative combinatorics of polytopes, in
number theory, and in some other fields. We emphasize that the Poisson summation formula
allows us to discretize integrals, in a sense that will be made precise in later chapters.

One pattern that the reader may have already noticed, among all of the examples of Fourier
transforms of polytopes computed thus far, is that each of them is a linear combination of a
very special kind of rational function of ξ, multiplied by a complex exponential that involves
a vertex of the polytope:
M
X 1
1̂P (ξ) = Qd e2πihvk ,ξi , (4.1)
k=1 j=1 hωj,k (vk ), ξi

where the vertices of P are v1 , . . . , vN , and where M ≥ N . We observed that in all of our
examples thus far, the denominators are in fact products of linear forms, as in (4.1). We
will be able to see some of the more precise geometric structure for these products of linear
forms, which come from the edges of the polytope, once we learn more about Fourier-Laplace
transforms of cones.

It is rather astounding that every single fact about a given polytope P is somehow hiding
inside these rational-exponential functions given by (4.1), due to the fact that the Fourier
transform 1̂P is a complete invariant (Lemma 3.3).

4.2 Introduction

In the spirit of bringing the reader very quickly up to speed, regarding the applications
of Fourier analytic tools, we outline the basics of the field, and only prove some of them.
Nowadays, there are many good texts on Fourier analysis, see Note (a).

One of the most useful tools for us is the Poisson summation formula, so we will indeed
prove it in some generality. As we will see, the Fourier transform is a very friendly creature,
allowing us to travel back and forth between the “space domain” and the “frequency domain”
to obtain many useful results. The readers who are already familiar with basics of Fourier
analysis may easily skip this chapter without impeding their understanding of the rest of the
book. For the rest of this chapter, we sometimes recall some standard facts, and the reader
is invited to look at the books mentioned above for their relevant proofs. We also note that
this chapter is only meant as an intuitive introduction to Fourier analysis, so although we
think about the Hilbert space L2 ([0, 1]d ), we will often not dwell on the more subtle questions
of whether a function belongs to L2 ∩ L1 , or L1 ∩ C 1 , etc. because we want to get the reader
up and running.

44
First, we introduce the space of square-integrable functions on the cube  := [0, 1]d :
Z
2 d
L () := {f : [0, 1] → C | |f (x)|2 dx < ∞}.
[0,1]d

The function space L2 () turns out to be a Hilbert space, because it is endowed with the
natural inner product Z
hf, gi := f (x)g(x)dx,
[0,1]d

relative to which it is a complete metric space (see [57] for more on Hilbert spaces). The
norm of a function f ∈ L2 () is defined by:
sZ
p
kf k := hf, f i = |f (x)|2 dx.
[0,1]d

Because L2 () is a Hilbert space, we have the Cauchy-Schwarz inequality here:


Z Z  21 Z  12
2 2
f (x)g(x)dx ≤ |f (x)| dx |g(x)| dx , (4.2)
  

for all f, g ∈ L2 (), with equality if and only if f (x) = Cg(x) for some constant C (Exercise
4.2). It is in this space, L2 (), that we build Fourier series.

The fact that L2 () is a Hilbert space means that it is a very friendly space, due to the
extra structure that exists here, namely the ability to use inner products to measure distance
between functions. This extra structure allows us to “do linear algebra” in an analogous
manner to finite-dimensional vector spaces, and in particular talk about bases of exponentials
(as in the following section, and definition (4.7) below).

In any Hilbert space, one of the most useful inequalities is the Cauchy-Schwarz inequality:

hu, vi ≤ kukkvk, (4.3)

with equality if and only if u, v are linearly independent. At this point the curious reader
d
might wonder
Pd “are there any other inner products on R , besides the usual inner product
d
hx, yi := x y
k=1 k k ?” A classification of all inner products that exist on R is given in
Exercise 4.15.

There are (at least) two other function spaces that come up very naturally as well. First,
the space of square-integrable functions on Rd is defined by:
 Z 
2 d d
2
L (R ) := f : R → C |f (x)| dx < ∞ .
Rd

45
Second, the space of absolutely integrable functions on Euclidean space is often used,
and is defined by:  Z 
1 d d

L (R ) := f : R → C |f (x)|dx < ∞ .
Rd

The usual theory progresses by first considering identities between functions in the Hilbert
space L2 (Rd ), and then later imposing other growth conditions to make various series con-
verge pointwise. One of the most common growth conditions is for a function to belong
to L1 (Rd ), which is also quite natural, and is the continuous analogue of absolute conver-
gence for a series. There are many fascinating facts about all of these functions spaces,
including the fact that L1 (Rd ) is not a Hilbert space, as we can easily show by exhibiting a
counter-example to the Cauchy-Schwarz inequality, as follows.

Example 4.1. We claim that the Cauchy-Schwarz inequality (which is one of the most
1
useful inequalities)
R is in fact falseRin L (R). If the Cauchy-Schwarz inequality was true here,
1 R 1
we would have R f (x)g(x)dx ≤ ( R |f (x)| dx) ( R |g(x)|2 dx) 2 for all functions f, g ∈ L1 (R).
2 2

As a counterexample, let
1
f (x) := 1(0,1) (x) √ .
x
It’s easy to see that f ∈ L1 (R):
Z Z 1
1 1 1
1(0,1) (x) √ dx = √ dx = .
R x 0 x 2
R1
But R f (x) · f (x)dx = 0 x1 dx diverges, so that we do not have a Cauchy-Schwarz inequality
R

in L1 (R), because here both the left-hand-side and the right-hand-side of such an inequality
do not even converge.

However, if the functions f, g above are bounded on R, then we do have a Cauchy-Schwartz


inequality in L2 (R), and we let the reader enjoy its verification. 

It is sometimes interesting to derive analogues between norms in Rd and norms in an infi-


nite dimensional function space. Among the many norm relations in Rd , we mention one
elementary but interesting relation:

kxk1 ≤ n kxk2 ,

for all vectors x ∈ Rd (Exercise 4.1).

For any k := (k1 , . . . , kd ) ∈ Zd≥0 , we can define the multivariable differential operator


Dk := .
∂xk11 · · · ∂xkdd

46
d
Example 4.2. In R1 , this is the usual k’th derivative: Dk f (x) := dxk
f (x). In R2 , for
example, we have D(1,7) f (x) := ∂x1∂∂x7 f (x). 
2

The order of the differential operator Dk is by definition |k| := k1 + · · · + kd . To define


spaces of differentiable functions, we call a function f : Rd → C a C m -function if all partial
derivatives Dk f of order |k| ≤ m exists and are continuous. We denote the collection
of all such C m -functions on Euclidean space by C m (Rd ). When considering infinitely-
differentiable functions on Euclidean space, we denote this space by C ∞ (Rd ).

Finally, we mention another basic ingredient, which holds for all of these functions spaces,
the triangle inequality for integrals:
Z Z
f (x)dx ≤ |f (x)|dx, (4.4)

S S

for any measurable subset S ⊂ Rd . It follows that if f is bounded on Rd , say f (x) ≤ M for
all x ∈ Rd , then Z Z
f (x)dx ≤ |f (x)|dx ≤ measure(S) · M, (4.5)

S S
a simple but very useful bound for many applications.

4.3 Orthogonality

For the unit cube  := [0, 1]d , there is a natural inner product for the space of square-
integrable functions V := L2 (), as we mentioned above, and this inner product is defined
by: Z
hf, gi := f (x)g(x)dx. (4.6)
[0,1]d

We focus on a particularly useful sequence of exponential functions, which turn out to be a


basis for V . For each n ∈ Z, we define en : Rd → C by:

en (x) := e2πihn,xi . (4.7)

This countable collection of exponentials turns out to form a complete basis for V . The
orthogonality is the first step, which we prove next, and we simply quote the more difficult
fact that they also span all of V .
Theorem 4.1 (Orthogonality relations for the exponentials en (x)).
Z (
1 if n = m
en (x)em (x)dx = (4.8)
[0,1]d 0 if not.

47
Proof. Because of the geometry of the cube, we can proceed in this case by separating the
variables, so that for n 6= m we compute:
Z Z
en (x)em (x)dx = e2πihn−m,xi dx
[0,1]d [0,1]d
d
Y e2πi(nk −mk ) − 1
= = 0,
k=1
2πi(nk − mk )

because each nk − mk ∈ Z.

4.4 Fourier series: fast and loose

Let’s see how we can expand (certain) functions in a Fourier series, as well as find a formula
for their series coefficients, in a foot-loose and carefree way - i.e. abandoning all rigor for the
moment.

Given that the sequence of exponential functions {en (x)}n∈Zd forms a basis of V := L2 (),
we know from Linear Algebra that any function f ∈ V may be written in terms of this basis:
X
f (x) = a(n)en (x). (4.9)
n∈Zd

How do we compute the Fourier coefficients an ? Let’s go through the intuitive process here,
ignoring convergence issues. Well, again by Linear Algebra, we take the inner product of
both sides with a fixed basis element ek (x):
X
hf (x), ek (x)i = h a(n)en (x), ek (x)i
n∈Zd
X
= a(n)hen (x), ek (x)i
n∈Zd
X
= a(n)δ(n, k)
n∈Zd
= a(k),

where we’ve used the orthogonality relations, Theorem 4.1 above, in the third equality.

48
Therefore, it must be the case that

a(k) = hf (x), ek (x)i


Z
:= f (x)e2πihk,xi dx
[0,1] d
Z
= f (x)e−2πihk,xi dx,
[0,1]d

also called the Fourier coefficients of f .

4.5 Fourier series: rigorous statements

Let’s record the intuitive arguments made in the previous section, which constitute the basis
for the classical result that a periodic function on Rd , belonging to L2 (), has a pointwise
convergent Fourier series.

Theorem 4.2 (Fourier series for periodic functions). Let F : Rd → C be a periodic


function, continuous on [0, 1)d , and furthermore let F ∈ L2 (). Then F has the Fourier
series: X
F (x) = a(n)e2πihn,xi , (4.10)
n∈Zd

for all x ∈ (0, 1)d , and the Fourier coefficients of F have the integral formula:
Z
a(n) = F (u)e−2πihn,ui du, (4.11)
[0,1]d

for all n ∈ Zd . 

We note that the Fourier coefficients above are integrals over the unit cube [0, 1]d . We will
encounter below, in the Poisson summation formula, a similar integral, but over all of Rd .
Such an integral comes up naturally during the typical “unfolding” procedure of the proof
of Poisson summation, and will be called a Fourier transform (more on this in section 4.7).
We therefore define the Fourier transform of f , for all x ∈ Rd , by the integral:
Z
fˆ(x) = f (u)e−2πihx,ui du. (4.12)
Rd

For the real line, we have the following refined version of Theorem 4.2. We use the standard

notation f (x+0 ) := lim→0 f (x0 + ), and f (x0 ) := lim→0 f (x0 − ), where  is always chosen
to be positive. We call f piecewise smooth if f ∈ C 1 (R).

49
Theorem 4.3. Let f : R → C be a periodic function, with fundamental domain [0, 1), and
piecewise smooth on R. Then, for each t ∈ R, we have
N
X f (t+ ) + f (t− )
lim fˆ(n)e2πint = . (4.13)
N →∞
n=−N
2

(For a proof of Theorem 4.3 see [155]). 

4.6 The Schwartz space, and other goodies

We recall that our definition of a ‘nice function’ was any function f : Rd → C for which the
Poisson summation formula holds. Here we give various sufficient conditions for a function
f to be nice. A Schwartz function f : R → C is defined as any function f ∈ C ∞ (R) that
satisfies the following growth condition: for all integers a, k ≥ 0,
d
|xa f (x)| is bounded on R. (4.14)
dxk

For Rd , we can define Schwartz functions similarly: they are infinitely differentiable func-
tions f : Rd → C such that for all vectors a, k ∈ Zd≥0 we have:
|xa Dk f (x)| is bounded on Rd , (4.15)
where xa := xa11 · · · xadd is the standard multi-index notation. Intuitively, a Schwartz function
decreases ‘at infinity’ faster than any polynomial function. We also define the Schwartz
space S(Rd ) to be set of all Schwartz functions f : Rd → C.
Theorem 4.4. The Fourier transform maps the Schwartz space S(Rd ) one-to-one, onto
itself. (See Exercise 4.14)

2
Example 4.3. The Gaussian function Gt (x) := e−t||x|| is a Schwartz function, for each fixed
t > 0.

To see this, we first consider R1 , where we note that the 1-dimensional Gaussian is a Schwartz
function, as follows. We observe that xa | dxdk Gt (x)| = Hn (x)Gt (x), where Hn (x) is a univariate
polynomial. Since limx→∞ eHt||x||
n (x)
2 = 0, Gt (x) is a Schwartz function. Now we note that
the product of Schwartz functions is again a Schwartz function; hence the d-dimensional
Gaussian, Gt (x) := e−t||x|| = dk=1 e−txk , a product of 1-dimensional Gaussians, is a Schwartz
2 Q 2

function. 

Another essential fact is Fubini’s theorem, which allows us to interchange integrals with
integrals, and series with integrals, for product spaces. If we write Rd = Rm × Rn , and we
denote a point z ∈ Rd by z := (x, y), then we may also write f (z) := f (x, y).

50
Theorem 4.5 (Fubini). Let f ∈ L1 (Rd ). Then:
Z Z Z 
f (z)dz = f (x, y)dx dy, (4.16)
Rd Rn Rm

and Z Z Z 
f (z)dz = f (x, y)dy dx. (4.17)
Rd Rm Rn


There is also a version of Fubini’s theorem that uses the counting measure in one of the
factors of Rm × Rn , giving us:
!
X Z  Z X
f (x, ξ)dx = f (x, ξ) dx.
ξ∈Zn Rm Rm ξ∈Zn

(See [133], p. 220, for a proof of Theorem 4.5)

Theorem 4.6 (Lebesgue dominated convergence). Suppose that we have a sequence of


(measurable) functions fn (x) : Rd → C, for n = 1, 2, 3, . . . , and suppose we have a limit
function f (x) = limn→∞ fn (x), for each x ∈ Rd .

If there exists a function g ∈ L1 (Rd ) such that


|fn (x)| ≤ g(x), n = 1, 2, 3, . . .
then f ∈ L1 (Rd ), and Z Z
lim fn (x)dx = f (x)dx.
n→∞ Rd Rd


There is an extremely useful application of Lebesgue’s dominated convergence theorem,


which allows us to interchange summations with integrals as follows.

Theorem 4.7. Suppose that we have a sequence of (measurable) functions fn (x) : Rd → C,


for n = 1, 2, 3, . . . , such that
∞ Z
X
|fn (x)|dx < ∞.
n=1 Rd
P∞
Then the series n=1 fn (x) converges for all x ∈ Rd , and we have:
X∞ Z Z X ∞
fn (x)dx = fn (x)dx.
n=1 Rd Rd n=1

(See [129], p. 26 for a proof of Theorem 4.6, and [129], p. 29 for a proof of Theorem 4.7)

51
4.7 Poisson Summation

We formally introduce the Poisson summation formula, one of the most useful tools in
analytic number theory, and in discrete / combinatorial geometry. There are many different
families of sufficient conditions that a function f can satisfy, in order for Poisson summation
to be applicable to f .

Figure 4.2: Spaces of functions for Poisson summation

Theorem 4.8 (Poisson summation formula, I). Given a Schwartz function f : Rd → C,


we have X X
f (n + x) = fˆ(ξ)e2πihξ,xi , (4.18)
n∈Zd ξ∈Zd

valid for all x ∈ Rd . In particular, we have:


X X
f (n) = fˆ(ξ). (4.19)
n∈Zd ξ∈Zd

Both sides of (4.18) converge absolutely, and are continuous functions on Rd .

d
P
Proof. If we let F (x) := n∈Zd f (n + x), then we notice that F is periodic on R , with
d d
the cube [0, 1)
P as a fundamental domain. The argument is easy: fix any m ∈ Z . Then
F (x + m) = n∈Zd f (n + x + m) = k∈Zd f (x + k), because Zd + m = Zd . By Theorem 4.2,
P
F has a fourier series, so let’s compute it:
X
F (x) := a(k)e2πihk,xi ,
k∈Zd

52
R
where a(k) = [0,1)d F (u)e2πihk,ui du for each fixed k ∈ Zd . Let’s see what happens if we
massage a(k) a bit:
Z
a(k) := F (u)e−2πihk,ui du (4.20)
[0,1)d
Z X
= f (n + u)e−2πihk,ui du (4.21)
[0,1)d
n∈Zd
XZ
= f (n + u)e−2πihk,ui du. (4.22)
[0,1)d
n∈Zd

The interchange of summation and integral in the latter step is allowed by Theorem 4.7
above, because the integrand satisfies |f (n + u)e−2πihk,ui | = |f (n + u)| ∈ L1 (Rd ), the latter
due to the fact that f is a Schwartz function.

Now we fix an n ∈ Zd in the outer sum of (4.22), and make the change of variable in the
integral: n + u := w, so that du = dw.

A critical step in this proof is the fact that as u varies over the cube [0, 1)d , n + u varies over
all of Rd because we have a tiling of Euclidean space by the unit cube: [0, 1)d + Zd = Rd .
We note that under this change of variable, e−2πihk,ui = e−2πihk,w−ni = e−2πihk,wi , because
k, n ∈ Zd and hence e2πihk,ni = 1. Thus, we finally have:
Z
a(k) = f (w)e−2πihk,wi dw := fˆ(k),
Rd

a(k)e2πihk,xi = fˆ(k)e2πihk,xi .
P P
so that F (x) = k∈Zd k∈Zd

We define a function f : Rd → C to be a nice function if the Poisson summation formula


X X
f (n + x) = fˆ(ξ)e2πihξ,xi (4.23)
n∈Zd ξ∈Zd

holds for f pointwise, for each x ∈ Rd . We will give various different sets of sufficient
conditions for a function f to be nice. Figure 4.2 suggests a simple containment relation
between some of these function spaces, as we can (and will) easily prove.

There are a few things to notice about the classical, and pretty proof of Theorem 4.18. The
first is that we began with any square-integrable function f defined on all of Rd , and forced
a periodization of it, which was by definition F . This is known as the “folding” part of the
proof. Then, at the end of the proof, there was the “unfolding” process, where we summed
an integral over a lattice, and because the cube tiles Rd , the sum of the integrals transformed
into a single integral over Rd .

The second thing we notice is that the integral over Rd , which by definition is called the
“Fourier transform of f ” (4.12), appears quite naturally, due to the tiling of Rd by the

53
unit cube [0, 1)d . Hopefully there will now be no confusion as to the difference between the
integral over the cube, and the integral over Rd , both appearing together in this proof.

There are various different families of functions for which the adjective ‘nice’ applies, and
one of the simplest to understand is the Schwartz class of functions. But there is another
family of nice functions that is extremely useful.

Theorem 4.9 (Poisson summation formula, II). [151] Let f ∈ L2 (Rd ), and suppose there
exist positive constants δ, C such that for all x ∈ Rd :
C C
|f (x)| < and |fˆ(x)| < . (4.24)
(1 + |x|)d+δ (1 + |x|)d+δ

Then we have the pointwise identity:


X X
f (n + x) = fˆ(ξ)e2πihξ,xi , (4.25)
n∈Zd ξ∈Zd

In addition, both sides of (4.25) converge absolutely, and are continuous functions on Rd .
We call a function that enjoys property (4.24) a Poisson function.

In other words, the growth conditions (4.24) guarantee that f is a nice function. We call
the space of functions that satisfy the hypotheses of Theorem 4.9, the Poisson space of
functions. A proof of Theorem 4.9 is given in Stein and weiss [151].

We observe that if a function f is a Schwartz function, then using the fact that the Fourier
transform maps S(Rd ) bijectively onto itself, we see that fˆ also satisfies the same growth
conditions, and is again a Schwartz function. Hence both f and fˆ decay faster than any
polynomials, and in particular they are both in the Poisson space, as Figure 4.2 suggested.
So we’ve shown that the Poisson space is contained in the Schwartz space.

There are other families of nice functions in the literature, including ‘functions of bounded
variation’, and ‘absolutely continuous’ functions, which we will not delve into here.

We will use a slightly more general version of the Poisson summation formula, which holds
for any lattice, and which follows rather quickly from the Poisson summation formula above.
We define a (full-rank) lattice L := M (Zd ) ⊂ Rd , the image of the integer lattice under an
invertible linear transformation M . The dual lattice of L is defined by L∗ := M −T (Zd ),
where M −T is the inverse transpose matrix of the real matrix M (see Section 6.6 for more
on dual lattices).

As we’ve seen in Lemma 3.3, Fourier Transforms behave beautifully under compositions with
any linear transformation. We will use this fact again in the proof of the following extension
of Poisson summation, which holds for all lattices and is quite standard.

54
Theorem 4.10 (Poisson summation formula, III). Given a Poisson function f : Rd → C,
we have X 1 X ˆ
f (n + x) = f (m)e2πihx,mi , (4.26)
n∈L
det L m∈L ∗

valid for all x ∈ Rd . In particular, we have


X 1 X ˆ
f (n) = f (ξ). (4.27)
n∈L
det L ξ∈L∗

Both sides of (4.26) converge absolutely and are continuous functions on Rd .

Proof. Any lattice (full-rank) may be written as L := M (Zd ), so that det L := | det M |.
Using the Poisson summation formula (4.18), with the change of variable n = M k, with
k ∈ Zd , we have:
X X
f (n) = (f ◦ M )(k)
n∈L k∈Zd
X
= (f\
◦ M )(ξ)
ξ∈Zd
1 X
fˆ M −T ξ

=
| det M | d ξ∈Z
1 X
= fˆ(m).
det L m∈L∗

where in the third equality we used the elementary ‘Stretch’ Lemma 3.3, and in the fourth
equality we used the definition of the dual lattice L∗ := M −T Zd .

As an afterthought, it turns out that the special case (4.27) also easily implies the general
case, namely (4.26) (Exercise 4.18).

A traditional application of the Poisson summation formula is the quick derivation of the
functional equation of the theta function. We first define the Gaussian function by:
d π 2
Gt (x) := t− 2 e− t ||x|| , (4.28)
for each fixed t > 0, and for all x ∈ Rd , as depicted in Figure 4.3.

Two immediately interesting properties of the Gaussian are:


Z
Gt (x)dx = 1, (4.29)
Rd

for each t > 0, and


2
Ĝt (m) = e−πt||m|| , (4.30)

55
Figure 4.3: The Gaussian family of functions Gt (x) with t = 1, t = .5, t = .3, and t = .1
respectively.

properties which are important in Statistics as well (Exercises 4.19 and 4.20). Each fixed
 gives us one Gaussian function and intuitively, as  → 0, this sequence of Gaussians
approaches the “Dirac delta function” at the origin, which is really known as a “generalized
function”, or “distribution” (Note (c)).

Example 4.4. The classical theta function (for the integer lattice) is defined by:
2
X
θ(t) = e−πt||n|| . (4.31)
n∈Zd

This function plays a major role in analytic number theory. One of its first historical appli-
cations was carried out by Riemann himself, who proved its functional equation (eq. (4.32)
below) and then applied a “Mellin
P∞transform” to it, to prove the functional equation of the
1
Riemann zeta function ζ(s) := n=1 ns .

We claim that it has the functional equation


 
1 d
θ = t 2 θ(t), (4.32)
t

for all t > 0. This will follow immediately from the Poisson summation formula 4.19 by

56
using f (x) := Gt (x). Using our knowledge of its FT, from 4.30, we have:
X X
Gt (n) = Ĝt (ξ)
n∈Zd ξ∈Zd
2
X
= e−πt||ξ|| := θ(t).
ξ∈Zd

d π 2 d
Gt (n) := t− 2 e− t ||n|| := t− 2 θ 1
P P 
Since by definition n∈Zd n∈Zd t
, (4.32) is proved. 

4.8 The convolution operation

For f, g ∈ L1 (Rd ), we define their convolution by


Z
(f ∗ g)(x) = f (x − y)g(y)dy,
Rd

for all x ∈ Rd for which the integral makes sense. It is possible to think intuitively of this
analogue of multiplication as: “this is how waves like to multiply”, via Lemma 4.2. We have
the following essential and standard facts.
Lemma 4.1. For f, g ∈ L1 (Rd ), their convolution (f ∗ g)(x) exists for almost all x ∈ Rd . 
Lemma 4.2.
\
(f ∗ g)(ξ) = fˆ(ξ)ĝ(ξ),
for all f, g ∈ L2 (Rd ). 

Lemma 4.2 means that convolution of functions in the space domain corresponds to the
usual multiplication of functions in the frequency domain (and vice-versa). It is easy to
check that the convolution product is commutative, distributive and associative, meaning
that f ∗ g = g ∗ f , f ∗ (g + h) = f ∗ g + f ∗ h, and f ∗ (g ∗ h) = (f ∗ g) ∗ h, respectively, for
all f, g, h ∈ L1 (Rd ).
Example 4.5. When P := [− 12 , 12 ], the convolution of 1P with itself is drawn in Figure 4.4.
We can already see that this convolution is a continuous function, hence a little smoother
than the discontinuous function 1P . Using Lemma 4.2 we have
 2
sin(πξ)
(1\
P ∗ 1P )(ξ) = 1̂P (ξ)1̂P (ξ) = .
πξ
We’ve used equation 3.1 in the last equality, for the Fourier transform of our interval P here.
Considering the graph in Figure 4.5, for the Fourier transform of the convolution (1P ∗ 1P ),

57
Figure 4.4: The function (1P ∗ 1P ) (x), with P := − 12 , 21
 

Figure 4.5: The Fourier transform (1\


P ∗ 1P )(ξ), which is equal to the smooth, nonnegative
 2
function sin(πξ)
πξ
:= sinc2 (ξ).

we see that this positive function is already much more tightly concentrated near the origin,
as compared with sinc(x) := 1̂P (ξ).

Another useful intuitive idea is that convolution is a kind of averaging process, and that the
convolution of two functions becomes smoother than either one of them. For our applications,
when we consider the indicator function 1P (x) for a polytope P, then this function is not
continuous on Rd , so that the Poisson summation formula does not necessarily hold for it.
But if we consider the convolution of 1P (x) with a Gaussian, for example, then we arrive at
the C ∞ function
(1P ∗ Gt )(x),
for which the Poisson summation does hold. In the sequel, we will use the latter convolved
function in tandem with Poisson summation to study “solid angles”.

58
It’s worth mentioning here that for any bounded measurable set K ⊂ Rd , we have
Z
(1K ∗ 1K )(y) := 1K (x)1K (y − x)dx (4.33)
Rd
Z
= 1K∩(−K+y) (x)dx (4.34)
Rd
Z
= dx (4.35)
K∩(−K+y)

= vol (K ∩ (−K + y)) , (4.36)

so that the convolution of indicator functions gives volumes.

4.9 Plancherel’s Theorem, and the Parseval identity

One of the main results in Fourier analysis is the Plancherel Theorem, which tells us that
the Fourier transform is an isometry of metric spaces. In other words, the Fourier transform
preserves norms of functions: kfˆkL2 = kf kL2 .
Theorem 4.11 (Plancherel I). Let f ∈ L2 (Rd ). Then
Z Z
|fˆ(ξ)| dξ =
2
|f (x)|2 dx.
Rd Rd

Proof. We let g(x) := f (−x), so that


Z
ĝ(ξ) = f (−x)e−2πihx,ξi dx
Rd
Z
= f (−x)e2πihx,ξi dx
Rd

= fˆ(ξ).

We define h := f ∗ g, and by Lemma 4.2 we have ĥ(ξ) = fˆ(ξ)ĝ(ξ), so that ĥ(ξ) = kfˆ(ξ)k2 .
R R R
Now, h(0) := Rd f (0 − x)g(x)dx = Rd f (−x)f (−x)dx = Rd |f (x)|2 .

On the other hand, h(0) = Rd ĥ(ξ)dξ = Rd |fˆ(ξ)|2 dξ. We therefore have


R R

Z Z
ˆ 2
|f (ξ)| dξ = |f (x)|2 dx.
Rd Rd

59
More generally, we have the following extended version of Plancherel’s Theorem, which has
an essentially identical proof (Exercise 4.21).

Theorem 4.12 (Plancherel II). Let f, g ∈ L2 (Rd ). Then hf, gi = hfˆ, ĝi. In other words,
Z Z
f (x)g(x)dx = fˆ(x)ĝ(x)dx. (4.37)
Rd Rd

Example 4.6. The sinc function, which we recall is defined by


(
sin(πx)
πx
, if x 6= 0
sinc(x) :=
1 if x = 0,

plays an important role in many fields, and here we will glimpse another aspect of its im-
portance, as an application of Plancherel’s identity (4.37) above. Let’s show that
Z (
1 if n = m
sinc(x − n)sinc(x − m)dx = (4.38)
R 0 if not.

Using Plancherel, with P := [− 21 , 21 ], we have


Z Z
sinc(x − n)sinc(x − m)dx = F(sinc(x − n))(ξ)F(sinc(x − m))(ξ)dξ
R ZR

= 1P (ξ)e2πiξn 1P (ξ)e2πiξm dξ
ZR

= e2πiξ(n−m) dξ
P
= δ(n, m),

by the orthogonality of the exponentials over P := [− 12 , 21 ] (Exercise 2.3).

So we see that the collection of functions



sinc(x − n) n ∈ Z

forms an orthogonal collection of functions in the Hilbert space L2 (R), relative to its norm. It
turns out that when one studies Shannon’s sampling theorem, these translated sinc functions
are in fact a complete basis for the Hilbert subspace of L2 (R) that consists of ‘bandlimited
functions’. 

Returning now to the Hilbert space L2 (), we have the following useful result for periodic
functions, known as Parseval’s identity .

60
Theorem 4.13 (Parseval’s
P identity). Suppose that f, g ∈ L2 (). We expand both in their
Fourier series: f (x) = n∈Zd an e2πihn,xi , and g(x) = n∈Zd bn e2πihn,xi . Then
P
Z X
f (u)g(u)du = an b n . (4.39)

n∈Zd

In particular, when f = g, we obtain


Z X
|f (u)|2 du = |an |2 . (4.40)

n∈Zd

Proof. The proof is a straightforward application of the orthogonality relations. We compute:


Z Z X X
f (u)g(u)du = an e2πihn,ui bn e−2πihm,ui du
 
n∈Zd m∈Zd
X Z
= an bm e2πihn,ui e−2πihm,ui du

n,m∈Zd
X
= an bm [n = m]
n,m∈Zd
X
= an b n ,
n∈Zd

using the orthogonality relations, Theorem 4.1, in the penultimate equality.

Parseval’s identity (4.40) is used very frequently in almost every branch of science. The
following is a simple Number-theoretic application.
Example 4.7. Let f (x) := P1 (x) := {x} − 12 , the first periodic Bernoulli polynomial. Then
as we have seen, f has the (conditionally convergent) Fourier series
1 X e2πikx
f (x) = − ,
2πi k
k∈Z−{0}

−1 1
for all x ∈
/ Z. So here we have ak = 2πi k
, by definition of f . Using Parseval’s identity (4.40)
R1 2
P 2
0
|f (u)| du = n∈Z |an | , we compute:
X 1 X 1 1 X 1
|an |2 = = ,
n∈Z
4π 2 n2 2π 2 n≥1 n2
n∈Z−{0}

while
Z 1 Z 1  2 Z 1 2
2 1 1 1
|f (u)| du = {x} − dx = x− dx = .
0 0 2 0 2 12

61
Therefore
X 1 π2
2
= .
n≥1
n 6
In a very similar manner one can evaluate the Riemann zeta function at all positive even
integers 2k, beginning with the Fourier series for the periodic Bernoulli polynomial Pk (x)
(Exercise 4.6) . 

4.10 More crucial properties

It is natural to wonder about the asymptotic decay rate of the Fourier coefficients of a
function f . In this direction we have the Riemann–Lebesgue lemma, as follows.
Lemma 4.3. Let f ∈ L1 (Rd ). Then:
lim fˆ(ξ) = 0.
|ξ|→∞

We collect here a few standard and useful properties of the Fourier transform.
Lemma 4.4. Suppose we are given any functions f, g ∈ L1 (Rd ). Then the following proper-
ties hold:

(a) Let T : R → R be the ‘translation by h’ function, defined by T (x) := x + h, for all


x ∈ R, and a fixed h ∈ R. Then we have:
F(f ◦ T )(m) = fˆ(m)e2πihm .

(b) Conversely, we also have:


F f (x)e−2πixh (m) = fˆ(m + h).


(c) Suppose in addition that f 0 ∈ L1 (R). Then


F (f 0 ) (m) = 2πimfˆ(m).

(d) Conversely, if we also have xf (x) ∈ L1 (R), then


 
d ˆ
F (−2πixf (x)) (m) = f (m).
dx

(See exercises 4.10, 4.11, 4.12).

62
4.11 Approximate identity

It is a sad fact of life that there is no identity in L1 (Rd ) for the convolution product - in
other words, there is no function h ∈ L1 (Rd ) such that
f ∗h=f (4.41)
for all f ∈ L1 (Rd ).

Why is that? Suppose there was such a function - then taking the Fourier transform of both
sides of (4.41), we would also have fˆ ĥ = fˆ for all f ∈ L1 (Rd ). Picking an f whose transform
is nowhere zero, we can divide both sides by fˆ now, to conclude that ĥ = 1. Now taking
inverse Fourier transforms, we solve for h, getting
Z
h(x) = e2πihx,ξi dx, (4.42)
Rd

an extremely interesting integral that unfortunately diverges. So there is no function that


plays the role of an identity for the convolution product (see note (c)). But it turns out
that we can get close! Here is how we may do it, and as a consequence we will be able
to rigorously apply the Poisson summation formula to a wider class of functions, including
smoothed versions of the indicator function of a polytope.

An approximate identity is any sequence of functions


φn (x) := nd φ(nx), (4.43)
which is defined for any (absolutely) integrable function φ : Rd → C with the additional
property that Z
φ(x)dx = 1.
Rd
R
It’s easy to show that the latter two equalities imply that Rd φn (x)dx = 1, for all n ≥ 1
(Exercise 4.23). So scaling φ by these n’s has the effect of squeezing φ so that it is becomes
concentrated near the origin, while maintaining a total mass of 1. Then intuitively a sequence
of such φn functions approach the “Dirac delta-function” at the origin (which is a distribution,
not a function).

There are many families of functions that give an approximate identity. In practice, we
will seldom have to specify exactly which sequence φn we pick, because we will merely use
the existence of such a sequence to facilitate the use of Poisson summation. Returning now
to the motivation of this section, we can recover the next-best-thing to an identity for the
convolution product, as follows.
Theorem 4.14. Suppose we are given a function f ∈ L1 (Rd ), which is continuous at a point
p ∈ Rd . Suppose that f ∗ φ exists. Then for any approximate identity φn we have
lim (f ∗ φn ) (p) = f (p). (4.44)
n→∞

63
Proof. We begin by massaging the convolution product:
Z
(φn ∗ f )(p) := φn (x)f (p − x)dx
Rd
Z  
= φn (x) f (p − x) − f (p) + f (p) dx
d
ZR   Z
= φn (x) f (p − x) − f (p) dx + f (p) φn (x)dx
Rd Z Rd
 
= f (p) + φn (x) f (p − x) − f (p) dx,
Rd
R
using the assumption that Rd φn (x)dx = 1. Using the definition of φn (x) := nd φ(nx), and
making a change of variable u = nx in the latter integral, we have:
Z   1
 
(φn ∗ f )(p) := f (p) + φ(u) f p − u − f (p) du.
Rd n

In the second part of the proof, we will show that as n → ∞, the latter integral tends to
zero. We will do this in two steps, first bounding the tails of the integral in a neighborhood
of infinity, and then bounding the integral in a neighborhood of the origin.

Step 1. Given any  > 0, we note that the latter integral converges, so the ‘tails are
arbitrarily small’. In other words, there exists an r > 0 such that
Z    
1
φ(u) f p − u − f (p) du < .

kuk>r n

Step 2. Now we want to bound kuk<r φ(u) f p − n1 u − f (p) du. We will use the fact that
R  
R
Rd
|φ(u)|du = M , a constant. Also, by continuity of f at p, we can pick an n sufficiently
large, such that:  
f p − u − f (p) <  ,
1

n M

when k n1 uk < r. Putting all of this together, and using the triangle inequality for integrals,
we have the bound
Z     Z  
1 1
φ(u) f p − u − f (p) du ≤ |φ(u)| f p − u − f (p) du < .

n n

kuk<r kuk<r

Therefore, as n → ∞, we have (φn ∗ f )(p) −→ f (p).

We note that a point of discontinuity of f , Theorem 4.14 may be false even in dimension 1,
as the next example shows.

64
Example 4.8. Let f (x) := 1[0,1] (x), which is discontinuous at x = 0 and x = 1. We claim
that for p = 1, for example, we have
1
lim (f ∗ φn )(p) = f (p),
n→∞ 2
so that the result of Theorem 4.14 does not hold at this particular p, because p lies on the
boundary of the 1-dimensional polytope [0, 1]. When p ∈ int([0, 1]), however, Theorem 4.14
does hold. 

4.12 A practical Poisson summation formula

In practice, we want to apply Poisson summation to indicator functions 1P of polytopes and


convex bodies. With this in mind, it’s useful for us to have our own, home-cooked version
of Poisson summation that is made for this culinary purpose.
2 d
R
Throughout, we fix any compactly supported function ϕ ∈ L (R ), with Rd
ϕ(x)dx = 1,
and we set ϕ (x) := 1d ϕ x , for each  > 0.


Theorem 4.15 (Poisson summation formula IV). Let f (x) ∈ L2 (Rd ) be a compactly sup-
ported function, and suppose that for each x ∈ Rd , we have:
f (x) = lim+ {ϕε ∗ f (x)} . (4.45)
ε→0

Then, for each ε > 0, we have absolute convergence:


X
ϕ
b (εm) fb(m) < +∞,

m∈Zd

and for each x ∈ Rd , we have


( )
X X
f (n + x) = lim+ b (εm) fb(m) e2πihm,xi
ϕ . (4.46)
ε→0
n∈Zd m∈Zd

We note that by our assumption on f , the left-hand-side of equation (4.46) is a finite sum.
For a detailed proof of Theorem 4.15, see [26].

Notes
(a) There are some wonderful introductory books that develop Fourier analysis thoroughly
from first principles, such as the books by Stein and Shakarchi [152] and Giancarlo

65
Travaglini [155]. The reader is also encouraged to read more advanced but fundamental
introductions to Fourier analysis, in particular the books by Mark Pinsky [114], Edward
Charles Titchmarsh [154], and Stein and Weiss [151]. In addition, the book [156]
by Audrey Terras is a good introduction to Fourier analysis on finite groups, with
applications. A more informal introduction to Fourier analysis, focusing on various
applications, is given by Brad Osgood [111].

(b) There are some “elementary” techniques that we will use, from the calculus of a com-
plex variable, but which require essentially no previous knowledge in this field. In
particular, suppose we have two analytic functions f : C → C and g : C → C, such
that f (zk ) = g(zk ) for a convergent sequence of complex numbers zk → L, where L is
any fixed complex number. Then f (z) = g(z) for all z ∈ C.
The same conclusion is true even if the hypothesis is relaxed to the assumption that
both f and g are meromorphic functions, as long as the sequence and its limit stay
away from the poles of f and g.

(c) The “Dirac delta function” is part of the theory of “generalized functions” and may be
d π 2
intuitively defined by the full sequence of Gaussians Gt (x) := t− 2 e− t ||x|| , taken over
all t > 0. The observation that there is no identity for the convolution product on Rd
is a clear motivation for a theory of generalized functions, beginning with the Dirac
delta function. Another intuitive way of “defining” the Dirac delta function is:
(
∞ if x = 0
δ0 (x) :=
0 if not,

even though this is not a function. But in the sense of distributions (i.e. generalized
functions), we have lim→0 Gt (x) = δ0 (x).
More rigorously, the δ-function belongs to a theory of distributions that was developed
by Laurent Schwartz in the 1950’s and by S.L. Sobolev in 1936, where we can think
of generalized functions as linear functionals on the space of all bump functions on Rd
(see the book by Lighthill [94] for a nice introduction to generalized functions).
Such generalized functions were originally used by the Physicist Paul Dirac in 1920,
before the rigorous mathematical theory was even created for it, in order to better
understand quantum mechanics.

(d) Of great practical importance, and historical significance, a bump function is defined
as any infinitely smooth function on Rd , which is compactly supported. In other words,
a bump function enjoys the following properties:

• φ has compact support on Rd .


• φ ∈ C ∞ (Rd ).

66
Bump functions are also called test functions, and if we consider the set of all bump
functions on Rd , under addition, we get a vector space V , whose dual vector space is
called the space of distributions on Rd .

(e) The cotangent function, appearing in some of the exercises below, is the unique mero-
morphic function that has a simple pole at every integer, with residue 1 (up to multi-
plication by an entire function with the same residues). The cotangent function also
forms an entry point for Eisenstein series in number theory, through the corresponding
partial fraction expansion of its derivatives.

(f) A deeper exploration into projections and sections of the unit cube in Rd can be found
in “The cube - a window to convex and discrete geometry”, by Chuangming Zong
[164]. In [86], Alexander Koldobsky gives a thorough introduction to sections of convex
bodies, intersection bodies, and the Busemann-Petty problem.

(g) There are numerous other identities throughout mathematics that are equivalent to
special cases of Poisson summation, such as the Euler-MacLaurin summation formula,
the Abel-Plana formula, and the Approximate sampling formula of signal analysis (see
[25] for a nice treatment of such equivalences for functions of 1 real variable, and
functions of 1 complex variable).

(h) What are some of the limitations of Poisson summation? One may wonder if ‘nice’
functions might include all continuous functions f such that fˆ is also continuous, and
such that f ∈ L1 (Rd ) and fˆ ∈ L1 (Rd ). But there is an important counterexample, by
Yitzhak Katznelson ([81], Ch. VI, exercise 15).

Exercises

In theory, there is no difference between theory and practice;


but in practice, there is! – Walter J. Savitch

p
4.1. ♣ Recalling that the L2 -norm is defined by kxk2 := x21 + · · · + x2d , and the L1 -norm
is defined by kxk1 := |x1 | + · · · + |xd |, we have the following elementary norm relations.

(a) Show that kxk2 ≤ kxk1 , for all x ∈ Rd .



(b) On the other hand, show that we also have kxk1 ≤ d kxk2 , for all x ∈ Rd .

67
4.2. ♣ Show that the Cauchy-Schwarz inequality holds in the Hilbert space L2 ():
Z Z  21 Z  12
f (x)g(x)dx ≤ |f (x)|2 dx |g(x)|2 dx , (4.47)
  

for all f, g ∈ L2 (), with equality if and only if f (x) = Cg(x) for some constant C

it −it it −it
4.3. We know that the functions u(t) := cos t = e +e 2
and v(t) := sin t = e −e 2i
are
natural, partly because they parametrize the unit circle: u2 + v 2 = 1. Here we see that there
are other similarly natural functions, parametrizing the hyperbola.

(a) Show that the following functions parametrize the hyperbola u2 − v 2 = 1:

et + e−t et − e−t
u(t) := , v(t) := .
2 2
t −t
(This is the reason that the function cosh t := e +e
2
is called the hyperbolic cosine, and
et −e−t
the function sinh t := 2 is called the hyperbolic sine)
cosh t et +e−t
(b) The hyperbolic cotangent is defined as coth t := sinh t
= et −e−t
. Using Bernoulli num-
bers, show that t coth t has the Taylor series:

X 22n
t coth t = B2n t2n .
n=0
(2n)!

4.4. Fix t > 0, and let f (x) := e−2πt|x| , for all x ∈ R. Show that f is not a Schwarz function.

4.5. ♣ We continue with the same function as in the previous exercise, f (x) := e−2πt|x| .

(a) Show that fˆ(ξ) = t 1


π ξ 2 +t2
, for all ξ ∈ R.

(b) Using Poisson summation, show that:


tX 1 X
= e−2πt|m| .
π n∈Z n2 + t2 m∈Z

4.6. ♣ Here we evaluate the Riemann zeta function at the positive even integers.

68
(a) Show that
X 1 + e−2πt
e−2πt|n| = := coth(πt),
n∈Z
1 − e−2πt
for all t > 0.

(b) Show that the cotangent function has the following (well-known) partial fraction expan-
sion: ∞
1 X 1
π cot(πx) = + 2x 2 − n2
,
x n=1
x
valid for any x ∈ R − Z.

(c) Let 0 < t < 1. Show that



tX 1 1 2X
2 2
= + (−1)m+1 ζ(2m) t2m−1 ,
π n∈Z n + t πt π m=1
P∞ 1
where ζ(s) := n=1 ns is the Riemann zeta function, initially defined by the latter
series, which is valid for all s ∈ C with Re(s) > 1.

(d) Here we show that we may quickly evaluate the Riemann zeta function at all even
integers, as follows. We recall the definition of the Bernoulli numbers, namely:
z z X B2m 2m
= 1 − + z .
ez − 1 2 m≥1 2m!

Prove that for all m ≥ 1,

(−1)m+1 (2π)2m
ζ(2m) = B2m .
2 (2m)!
π2 π4
Thus, for example, using the first 3 Bernoulli numbers, we have: ζ(2) = 6
, ζ(4) = 90
,
π6
and ζ(6) = 945 .

4.7. For each n ≥ 1, let Tn (x) = cos(nx). For example, T2 (x) = cos(2x) = 2 cos2 (x) − 1, so
T2 (x) = 2u2 − 1, a polynomial in u := cos x.

(a) Show that for all n ≥ 1, Tn (x) is a polynomial in cos x.

(b) Can you write xn + x1n as a polynomial in the variable x + x1 ? Would your answer
be related to the polynomial Tn (x)? What’s the relationship in general? For example,
 2
2 1 1
x + x2 = x + x − 2.

69
Notes. The polynomials Tn (x) are very important in applied fields such as approximation
theory, and optimization, because they have many useful extremal properties. They are
called Chebyshev polynomials.

4.8. The hyperbolic secant is defined by


2
sech(πx) := , for x ∈ R.
eπx + e−πx
Show that sech(πx) is its own Fourier transform:

F(sech)(ξ) = sech(ξ),

for all ξ ∈ R.

Notes. The reader may need some background in complex analysis for this exercise.

4.9. Using the previous exercise, conclude that


Z
1 1
πx −πx
dx = .
R e +e 2

The following exercises give more practice in computing/handling general Fourier transforms
and their important properties. Throughout, we assume that the Fourier transform of f
exists, where f : R → C is any integrable function.

4.10. ♣ For any f ∈ L1 (R), show that


 
F f (x)e−2πixh (m) = fˆ(m + h),

for all m, h ∈ R.

4.11. ♣ For any f ∈ L1 (R), such that f 0 ∈ L1 (R) as well, show that

F (f 0 ) (m) = 2πimfˆ(m),

for all m ∈ R.

4.12. ♣ Conversely, for any f ∈ L1 (R) such that xf (x) ∈ L1 (R) as well, show that:
 
  d ˆ
F − 2πixf (x) (m) = f (m),
dx
for all m ∈ R.

70
4.13. ♣ Prove that: Z 1
1
P1 (ax)P1 (bx)dx = .
0 12 gcd2 (a, b)
1
for all positive integers a, b. Here P1 (x) := x − {x} − 2
is the first periodic Bernoulli
polynomial.

Notes. This integral is called a Franel integral, and there is a substantial literature about
related integrals. In 1924, Jérôme Franel related this integral to the Riemann hypothesis,
and to Farey fractions.

4.14. ♣ Let f : R → C belong to the Schwarz class of functions on R, denoted by S(R).


Show that fˆ ∈ S(R) as well.

4.15. ♣ Here we answer the very natural question “What are the other inner products on
Rd , besides the usual inner product hx, yi := dk=1 xk yk ?”
P

The fact is that all inner products are related to each other via positive definite matrices, as
follows. We recall from Linear Algebra that a symmetric matrix is called positive definite if
all of its eigenvalues are positive. Prove that the following two conditions are equivalent:

1. hx, yi is an inner product on Rd .

2. hx, yi := xT M y, for some positive definite matrix M .

4.16. For any positive real numbers a < b < c < d, define

f (x) := 1[a,b] (x) + 1[c,d] (x).

Can you find a, b, c, d such that fˆ(ξ) is nonzero for all ξ ∈ R?

4.17. ♣ Show that the only eigenvalues of the linear operator f → fˆ are {1, −1, i, −i}, and
show that each of these eigenvalues is achieved by some function. You can assume that f
belongs to L2 (Rd ).

4.18. ♣ Show that the special case of Poisson summation, 4.27, implies the general case,
Theorem 4.10.

71
4.19. ♣ We define the Gaussian, for each fixed  > 0, and for all x ∈ Rd , by
1 π 2
G (x) := d e−  ||x|| . (4.48)
 2

Show that: Z
G (x)dx = 1.
Rd

4.20. ♣ Show that, for all m ∈ Rd , the Fourier transform of the Gaussian G (x) is:
2
Ĝ (m) = e−π||m|| .

4.21. ♣ For all f, g ∈ S(Rd ), show that hf, gi = hfˆ, ĝi.

4.22. Show that for any convex body B ⊂ Rd , (1B ∗ 1B ) (x) is a continuous function of
x ∈ Rd .

4.23. ♣ Given any approximate identity sequence φ , as defined in (4.43), show that for
each  > 0, Z
φ (x)dx = 1.
Rd

4.24. We define the ramp function x+ := max{x, 0}, which may also be written as
(
x if x ≥ 0
x+ :=
0 if x < 0.

Show that the ramp also has the representation


x + |x|
x+ = , (4.49)
2
for all x ∈ R.

Notes. This ramp function is quite common nowadays in machine learning, and has a strong
presence in approximation theory.

4.25. (hard) ♣ Here we show that there exist compactly supported functions, (which are also
radial) whose Fourier transform is strictly positive on all of Rd . Let φ : Rd → R be defined
by

Z ∞
kxk2 √
φ(x) := 1− (1 − t)+ dt
0 t +

where the constant λ ≥ d−1


2
if d ≥ 2, and for dimension d = 1, we set λ = 12 . Using properties
of J-Bessel functions, show that φ̂(ξ) > 0 for all ξ ∈ Rd . (see [32] for more details)

72
Chapter 5

The geometry of numbers -


Minkowski’s Fundamental Theorem,
and Siegel’s extension

“Henceforth space by itself, and time by itself, are doomed to fade away into
mere shadows, and only a kind of union of the two will preserve an independent
reality.” – Hermann Minkowski

5.1 Intuition

To see a wonderful and fun application


of Poisson summation, we give a rela-
tively easy proof of Minkowski’s fundamen-
tal theorem, in the Geometry of Numbers.
Minkowski’s theorem gives the existence of
an integer point inside symmetric bodies
in Rd , once we know their volume is suf-
ficiently large.

In fact we first prove a more powerful iden-


tity which is a classical result of Carl Lud-
wig Siegel (Theorem 5.3), yielding an iden-
tity between Fourier transforms of convex
bodies and their volume. Our proof of this
identity of Siegel uses Poisson summation, Figure 5.1: Hermann Minkowski
applied to the convolution of an indicator

73
function with itself. The geometry of num-
bers is an incredibly beautiful field, and too vast to encompass in just one chapter (see note
(c)). We hope this chapter, a small bite of a giant fruit, gives the reader motivation to pursue
the interactions between convex bodies and lattices even further.

5.2 Minkowski’s convex body Theorem

Minkowski initiated the field that we


call today ‘the geometry of numbers’,
around 1890. To begin, we define a
body in Rd as a bounded, closed set.
Most of the time, it is useful to work
with convex bodies that enjoy the fol-
lowing symmetry. We call a body
B centrally symmetric, also called
symmetric about the origin, if for
all x ∈ Rd we have

Figure 5.2: A convex, symmetric body in R2 , with x ∈ P ⇐⇒ −x ∈ P. (5.1)


area bigger than 4, containing two nonzero integer a body Q is called symmetric if some
points. translation of Q is centrally symmetric.

For example, the ball {x ∈ Rd | kxk ≤ 1} is centrally symmetric, and the translated ball
{x ∈ Rd | kx − wk ≤ 1} is symmetric, but not centrally symmetric.

Theorem 5.1 (Minkowski’s convex body Theorem for Zd ). Let B be a d-dimensional convex
body in Rd , symmetric about the origin.
If vol B > 2d , then B must contain a nonzero integer point in its interior. (5.2)

Sometimes this classical and very useful result of Minkowski is stated in its contrapositive
form: Let B ⊂ Rd be any convex body, symmetric about the origin.
If the only integer point in the interior of B is the origin, then vol B ≤ 2d . (5.3)

It is natural, and straightforward, to extend this result to any lattice L := M (Zd ), by simply
applying the linear transformation M to both the integer lattice, and to the convex body B.
The conclusion is the following, which is the version that we will prove as a consequence of
Siegel’s Theorem 5.3.

74
Theorem 5.2 (Minkowski’s convex body Theorem for a lattice L). Let B be a d-dimensional
convex body in Rd , symmetric about the origin, and let L be a (full rank) lattice in Rd .

If vol B > 2d (det L), then B must contain a nonzero point of L in its interior. (5.4)

Proof. The proof appears below - see (5.27).

These very important initial results of Minkowski [105] have found applications in algebraic
number theory, diophantine analysis, combinatorial optimization, and other fields. In the
next section we show that Minkowski’s result (5.4) follows as a special case of Siegel’s formula.

Figure 5.3: The Rhombic dodecahedron, a 3-dimensional symmetric polytope that tiles R3
by translations, and is another extremal body for Minkowski’s convex body Theorem.

5.3 Siegel’s extension of Minkowski:


a Fourier transform identity for convex bodies

An important construction in the geometry of number is the Minkowski sum of convex


bodies. Given two convex bodies K, L ⊂ Rd , their Minkowski sum is defined by

K + L := {x + y | x ∈ K, y ∈ L}.

Another related construction, appearing in some of the results below, is

K − L := {x − y | x ∈ K, y ∈ L},

the Minkowski difference of K and L. A very useful special case is the gadget known as the
Minkowski symmetrized body of K, defined by
1 1
K − K. (5.5)
2 2
75
Given any set K ⊂ Rd , the difference K − K is centrally symmetric. To see this, suppose
x ∈ K − K, so we may write x = y − z, with y, z ∈ K. Then −x = z − y ∈ K − K.

In addition, we have the fortuitous and easy fact that any set K ⊂ Rd is centrally symmetric
if and only if we have the equality
1 1
K − K = K. (5.6)
2 2
(Exercise 5.3). Now suppose we are given two convex bodies K, L ⊂ Rd . Then the resulting
bodies K + L, K − L turn out to also be convex (Exercise 5.2). Another important geometric
notion is the dilation of a convex body by a positive real number t:
tB := {tx | x ∈ B},
The most basic version of Siegel’s theorem is the following identity, which assumes that a
convex body K is symmetric about the origin.

Theorem 5.3 (Siegel). Let B be any d-dimensional convex body in Rd , symmetric about the
origin, and suppose that the only integer point in the interior of B is the origin. Then
4d X 2
2d = vol B + 1̂ 21 B (ξ) . (5.7)

vol B d
ξ∈Z −{0}

We now prove the following extension of the latter Siegel formula (5.7), namely (5.8) below,
which applies to bodies that are not necessarily convex, nor necessarily symmetric about
the origin. Our proof of Theorem 5.4 below consists of yet another application of Poisson
summation. It turns out that if K is any convex body, then f := 1 1 K ∗ 1− 1 K is a nice
2 2
function (Exercise 5.9), in the sense that Poisson summation (4.23) holds for f . So Theorem
5.3 is a consequence of the following extension.

Theorem 5.4 (Siegel’s formula, for a general body K, and a lattice L). Let K ⊂ Rd be
any measurable set (so K is not necessarily symmetric or convex), for which the convolution
1 1 K ∗ 1− 1 K is a nice function. Let B := 12 int(K) − 21 int(K) be its symmetrized body. If the
2 2
only integer point in the interior of B is the origin, then

d 4d X 2
2 = vol K + 1̂ 12 K (ξ) . (5.8)

vol K d
ξ∈Z −{0}

More generally, if we replace the lattice Zd by any full-rank lattice L, and assume that the
only lattice point of L in the interior of B is the origin, then we have:
4d X 2
2d det L = vol K + 1̂ 12 K (ξ) . (5.9)

vol K ∗
ξ∈L −{0}

76
Proof. We start with the function
 
f (x) := 1 1 K ∗ 1− 1 K (x), (5.10)
2 2

which is continuous on Rd , and we plug f into Poisson summation (4.19):


X X
f (n) = fˆ(ξ). (5.11)
n∈Zd ξ∈Zd

We first compute the left-hand-side of Poisson summation, using the definition of f :


X XZ
f (n) = 1 1 K (y)1− 1 K (n − y)dy (5.12)
2 2
Rd
n∈Zd n∈Zd
XZ
= 1 1 intK (y)1− 1 intK (n − y)dy, (5.13)
2 2
Rd
n∈Zd

where the last step follows from the fact that the integral does not distinguish between a
closed convex set or its closure. Now we follows the definitions of containment: y ∈ 21 K and
n − y ∈ − 12 K imply that the integer point n ∈ 12 K − 12 K. But by hypothesis 12 K − 12 K
contains the origin as its only interior integer point, so the left-hand-side of the Poisson
summation formula contains only one term, namely the n = 0 term:
X XZ
f (n) = 1 1 K (y)1− 1 K (n − y)dy (5.14)
2 2
Rd
n∈Zd n∈Zd
Z
= 1 1 K (y)1− 1 K (−y)dy (5.15)
2 2
Rd
Z
= 1 1 K (y)dy (5.16)
2
Rd
 
1 vol K
= vol K = . (5.17)
2 2d

77
On the other hand, the right-hand-side of Poisson summation gives us:
X X
fˆ(ξ) = 1̂ 1 K (ξ)1̂− 1 K (ξ) (5.18)
2 2
ξ∈Zd ξ∈Zd
XZ Z
2πihξ,xi
= e dx e2πihξ,xi dx (5.19)
1
K − 12 K
ξ∈Zd 2

XZ Z
2πihξ,xi
= e dx e2πih−ξ,xi dx (5.20)
1 1
K K
ξ∈Zd 2 2

XZ Z
2πihξ,xi
= e dx e2πihξ,xi dx (5.21)
1 1
K K
ξ∈Zd 2 2

X 2
= 1̂ 12 K (ξ) (5.22)

ξ∈Zd
2 X 2
= 1̂ 1 K (0) + 1̂ (ξ) (5.23)

1
2K
2
ξ∈Zd −{0}
2
vol K X 2
= + 1̂ 21 K (ξ) , (5.24)

4d
ξ∈Zd −{0}

where we have pulled out the ξ = 0 term from the series (5.22). So we’ve arrived at
vol K vol2 K X 2
= + 1̂ 21 K (ξ) ,

2d 4d
d ξ∈Z −{0}

yielding the required identity:


4d X 2
2d = vol K + 1̂ 12 K (ξ) .

vol K
ξ∈Zd −{0}

Finally, to prove the stated extension to all lattices L, we use the slightly more general form
of Poisson summation, Theorem 4.10, valid for any lattice L:
X 1 X ˆ
f (n) = f (ξ). (5.25)
n∈L
det L ξ∈L ∗

1
All the steps of the proof above are identical, except for the factor of det L
, so that we arrive
at the required identity of Siegel for arbitrary lattices:
vol K vol2 K 1 X 2
= d + 1̂ 12 K (ξ) . (5.26)

2d 4 det L det L ∗ ξ∈L −{0}

78
The proof of Minkowski’s convex body Theorem for lattices, namely Theorem 5.2 above,
now follows immediately.

Proof of Theorem 5.2. [Minkowski’s convex body Theorem for a lattice L]

Applying Siegel’s Theorem 5.4 to the centrally symmetric body B := K, in the statement of
Siegel’s theorem, we see that the lattice sum on the right-hand-side of identity (5.8) contains
only non-negative terms. It follows that we immediately get the analogue of Minkowski’s
result for a given cenetrally symmetric body B and a lattice L, in its contrapositive form.
Namely, if the only lattice point of L in the interior of B is the origin, then

2d det L ≥ vol B. (5.27)

In fact, we can easily extend Minkowski’s Theorem 5.2, using the same ideas of the latter
proof, by using Siegel’s Theorem 5.4 so that it applies to non-symmetric bodies as well (but
there’s a small ‘catch’ - see Exercise 5.14).

5.4 Tiling and multi-tiling Euclidean space by transla-


tions of polytopes

First, we give a ‘spectral’ equivalence for being able to tile Euclidean space by a single
polytope, using only translations by a lattice. It will turn out that the case of equality in
Minkowski’s convex body Theorem is characterized precisely by the polytopes that tile Rd
by translations. These bodies are called extremal bodies.

More generally, we would like to also consider the notion of multi-tiling, as follows. We say
that a polytope P k-tiles Rd by using a set of translations L if
X
1P+n (x) = k, (5.28)
n∈L

for all x ∈ Rd , except those points x that lie on the boundary of P or its translates under
L (and of course these exceptions form a set of measure 0 in Rd ). In other words, P is a
k-tiling body if almost every x ∈ Rd is covered by exactly k translates of P.

Other synonyms for k-tilings in the literature are multi-tilings of Rd , or tiling at level k.
When L is a lattice, we will say that such a k-tiling is periodic. A common research theme
is to search for tilings which are not necessarily periodic, but this is a difficult problem in
general. The classical notion of tiling, such that there are no overlaps between the interiors of

79
any two tiles, corresponds here to the case k = 1. We have the following dictionary between
multi-tiling or Euclidean space by translations of a convex body P, and a property of the
zero set of the Fourier transform of P, due to Kolountzakis [87].

Theorem 5.5. Fix any integer k ≥ 1. A body P k-tiles Rd by translations with a lattice L
if and only if the following two conditions hold:

1. 1̂P (ξ) = 0, for all nonzero ξ ∈ L∗ , the dual lattice.


vol P
2. k = det L
.

Proof. We begin with the definition of multi-tiling, so that by assumption


X
1P+n (x) = k, (5.29)
n∈L

for all x ∈ Rd except those points x that lie on the boundary of P or its translates under
L (and of course these exceptions form a set of measure 0 in Rd ). A trivial but useful
observation is that
1P+n (x) = 1 ⇐⇒ 1P (x − n) = 1,
P
so we can rewrite the defining identity (5.29) as n∈L 1P (x − n) = k. Now we notice that
the left-hand-side is a periodic function of x, namely
X
F (x) := 1P (x − n)
n∈L

in x with L as its set


is periodic P Pof periods. This is easy to see: if we let l ∈ L, then
F (x + l) = n∈L 1P (x + l − n) = m∈L 1P (x + m) = F (x), because the lattice L is invariant
under a translation by any vector that belongs to it.

The following ‘intuitive proof’ would in fact be rigorous if we were allowed to use ‘generalized
functions’, but since we do not use them in this book, we label this part of the proof
as ‘intuitive’, and we then give a rigorous proof, using functions rather than generalized
functions.

[Intuitive proof] By Theorem 4.2, we may expand F into its Fourier series, because it is a
periodic function on Rd . Now by Poisson summation, namely Theorem 4.10, we know that
its Fourier coefficients are the following:
X 1 X
1P (x + m) = 1̂P (ξ)e2πihξ,xi , (5.30)
m∈L
det L ξ∈L∗
P
By our multi-tiling assumption (5.29), we have n∈L 1P+n (x) = k. So comparing the latter
assumption with (5.30), we have two Fourier series that represent the same function, so by

80
uniqueness of Fourier series, the corresponding coefficients on both sides must be equal to
each other. Equating the constant terms, we have
1 vol P
k= 1̂P (0) = ,
det L det L
while equating all of the other terms we see that 1̂P (ξ) = 0 for all nonzero ξ ∈ L∗ .

[Rigorous proof] In order to apply Poisson summation, it is technically necessary to replace


1P (x) by a smoothed version of it, in (5.30). Because this process is so common and useful
in applications, this proof is instructive. So we let φn be any approximate identity, as in
(4.44), and we will impose the extra assumption that φn is supported on P. Now we apply
our ‘practical Poisson summation’ (eq. 4.46) to 1P ∗ φn :
X 1 X
(1P ∗ φn ) (x + m) = 1̂P (ξ)φ̂n (ξ)e2πihξ,xi , (5.31)
m∈L
det L ξ∈L∗

Using Theorem 4.14, and the fact that the summands on the left-hand-side of (5.31) are
compactly supported,Pwe may take the limit as n → ∞ inside the finite sum on the LHS
of (5.31), to obtain m∈L 1P (x + m), which is equal to the constant k, by assumption.
Summarizing, we have:
X 1 X
k= 1P (x + m) = lim 1̂P (ξ)φ̂n (ξ) e2πihξ,xi . (5.32)
n→∞ det L
m∈L ξ∈L∗

Due to the fact that 1P and φn are both supported on P, we may drop the limit, as long as
we evaluate the RHS of (5.32) for all sufficiently large values of n. Using the uniqueness of
Fourier series, we may compare coefficients on both sides of
1 X
k= 1̂P (ξ)φ̂n (ξ) e2πihξ,xi . (5.33)
det L ξ∈L∗

1 vol P
Comparing the constant terms, and using φ̂n (0) = 1, we obtain k = 1̂ (0)φ̂n (0)
det L P
= det L
,
as claimed. Comparing all other terms, we obtain

0 = 1̂P (ξ)φ̂n (ξ) e2πihξ,xi ,

for all nonzero ξ ∈ L∗ . If we knew that φ̂n (ξ) 6= 0 for any nonzero ξ ∈ L∗ , we would be
done, for then we would have 1̂P (ξ) = 0 for all nonzero ξ ∈ L∗ . In fact, using Exercise 4.25,
we can allow our approximate identity φ to enjoy an additional constraint: φ̂(ξ) > 0 for all
ξ ∈ Rd .

In 1905, Minkowski gave necessary conditions for a polytope P to tile Rd by translations.


Later, Venkov and independently McMullen found sufficient conditions as well, culminating
in the following fundamental result.

81
Figure 5.4: An extremal body in R2 , relative to the integer lattice, which is a hexagon. It
has area 4, and no integer points in its interior. We also get a 2-parameter family of such
extremal bodies, parametrized by the point p ∈ R2 in the figure. It is clear from the picture
that this family of extremal bodies consists of either symmetric hexagons, or symmetric
quadrilaterals.

Theorem 5.6 (Minkowski-Venkov-McMullen). A polytope P tiles Rd by translations if and


only if the following 3 conditions hold:

1. P is a symmetric polytope.
2. The facets of P are symmetric polytopes.
3. Fix any face F ⊂ P of codimension 2, and project P onto the 2-dimensional plane that
is orthogonal to the (d − 2)-dimensional affine span of F . Then this projection is either
a parallelogram, or a centrally symmetric hexagon.

5.5 Extremal bodies

An extremal body is a convex, symmetric body K for which we have equality in Minkowski’s
convex body Theorem:
vol K = 2d (det L).
If we just look at equation 5.8 a bit more closely, we quickly get a nice corollary that arises by
combining Theorem 5.5 and Siegel’s Theorem 5.3. Namely, equality occurs in Minkowski’s
convex body theorem if and only if K tiles Rd by translations. Let’s prove this.

82
Theorem 5.7 (Extremal bodies). Let K be any convex, centrally symmetric subset of Rd ,
and fix a full-rank lattice L ⊂ Rd . Suppose that the only point of L in the interior of K is
the origin. Then:

1
2d det L = vol K ⇐⇒ 2
K tiles Rd by translations with the lattice L.

Proof. By Siegel’s formula (5.9), we have

4d X 2
2d det L = vol K + 1̂ 12 K (ξ) . (5.34)

vol K
ξ∈L∗ −{0}

Therefore, the assumption 2d det L = vol K holds ⇐⇒

4d X 2
0= 1̂ 21 K (ξ) , (5.35)

vol K
ξ∈L∗ −{0}

⇐⇒ all of the non-negative summands 1̂ 1 K (ξ) = 0, for all nonzero ξ ∈ L∗ . Now we


2
would like to use Theorem 5.5 to show the required tiling equivalence, namely that 12 K tiles
Rd by translations with the lattice L. We have already verified condition (a) of Theorem
5.5, applied to the body 12 K, namely that 1̂ 1 K (ξ) = 0, for all nonzero ξ ∈ L∗ . To verify
2
condition (b) of Theorem 5.5, we notice that because vol 21 K = 21d vol K, it follows that

vol( 1 K )
2d det L = vol K is equivalent to 1 = det2L , so that we may apply Theorem 5.5 with
P := 12 K, and with the multiplicity k := 1.

There is an extension of theorem 5.6, the Minkowski-Venkov-McMullen result, to multi-


tilings.

Theorem 5.8. [67] If a polytope P multi-tiles Rd by translations with a discrete set of


vectors, then

1. P is a symmetric polytope.

2. The facets of P are symmetric polytopes.

In the case that P ⊂ Rd is a rational polytope, meaning that all the vertices of P have
rational coordinates, the latter two necessary conditions for multi-tiling become sufficient
conditions as well [67].

83
Figure 5.5: The truncated Octahedron, one of the 3-dimensional polytopes that tiles R3 by
translations.

5.6 More about centrally symmetric polytopes

It’s both fun and instructive to begin by seeing how very simple Fourier methods can give
us deeper insight into the geometry of symmetric polytopes. The reader may glance at the
definitions above, in (5.1).

Example 5.1. Consider the cross-polytope ♦ ⊂ R3 , defined in Chapter 3. This is a centrally


symmetric polytope, but each of its facets is not a symmetric polytope, because its facets
are triangles. 

If all of the k-dimensional faces of a polytope P are symmetric, for each 1 ≤ k ≤ d, then P
is called a zonotope. Zonotopes form an extremely important class of polytopes, and have
various equivalent formulations.

Lemma 5.1. A polytope P ⊂ Rd is a zonotope ⇐⇒ P has one of the following properties.

(a) P is a projection of some n-dimensional cube.

(b) P is the Minkowski sum of a finite number of line segments.

A projection here means any affine transformation of P, where the rank of the associated
matrix may be less than d.

Zonotopes have been very useful in the study of tilings ([163], [17]). For instance, in dimen-
sion 3, the only polytopes that tile R3 by translations with a lattice are zonotopes, and there
is a list of 5 of them (up to an isomorphism of their face posets), called the Fedorov solids,
and drawn in Figure 5.8 (also see our Note (e) below).

84
By definition, any zonotope is a symmetric polytope, but the converse is not true; for exam-
ple, the cross-polytope is symmetric, but it has triangular faces, which are not symmetric,
so the crosspolytope is not a zonotope.

Figure 5.6: A 3-dimensional zonotope, called the rhombic dodecahedron, showing in bold its
4 line segments whose Minkowski sum generate the object.

Example 5.2. Consider the following 3 line segments in R2 : conv{( 00 ) , ( 10 )}, conv{( 00 ) , ( 21 )},
and conv{( 00 ) , ( 13 )}. The Minkowski sum of these three line segments, by definition a zono-
tope in R2 , is the symmetric hexagon whose vertices are ( 00 ) , ( 10 ) , ( 21 ) , ( 33 ) , ( 31 ) , ( 43 ). Notice
that once we graph it, in Figure 5.7, the graph is hinting to us that this body is a projection
of a 3-dimensional cube, and indeed this turns out to be always true for Minkowski sums of
line segments. 

Example 5.3. A particular embedding of the truncated octahedron P, drawn in Figure 5.5,
is given by the convex hull of the set of 24 vertices defined by all permutations of (0, ±1, ±2).
We note that this set of vertices can also be thought of as the orbit of just the one point
(0, 1, 2) ∈ R3 under the hyperoctahedral group (see [36] for more on the hyperoctahedral
group). It turns out that this truncated octahedron P tiles R3 by translations with a lattice
(Exercise 5.10). 

As the following Lemma shows, it is easy to detect/prove whether or not S is centrally


symmetric by just observing whether or not its Fourier transform is real-valued.

Lemma 5.2. A bounded, measurable set S ⊂ Rd is symmetric about the origin ⇐⇒

1̂S (ξ) ∈ R,

for all ξ ∈ Rd .

85
Figure 5.7: The Minkowski sum of 3 line segments in the plane, forming a 2-dimensional
zonotope.

Proof. Suppose that the set S is centrally symmetric. Then we have


Z Z
1̂S (ξ) := e 2πihξ,xi dx = e−2πihξ,xi dx (5.36)
S ZS
= e2πihξ,xi dx (5.37)
Z−S
= e2πihξ,xi dx := 1̂S (ξ), (5.38)
S
(5.39)

showing that the complex conjugate of 1̂S is itself, hence that it is real-valued.

Conversely, suppose that 1̂S (ξ) ∈ R, for all ξ ∈ Rd . We use the fact that the Fourier
transform is invertible:
(F ◦ F)(1S )(x) = 1S (−x), (5.40)
for all x ∈ Rd . To show that S is centrally symmetric, we need to show that 1−S (x) = 1S (x),
for all x ∈ Rd . Further, by 5.40, it now suffices to show that 1̂−S (ξ) = 1̂S (ξ), for all ξ ∈ Rd .
We therefore compute:
Z Z
1̂−S (ξ) := e 2πihξ,xi
dx = e−2πihξ,xi dx (5.41)
−S S
Z
= e2πihξ,yi dy (5.42)
S

:= 1̂S (ξ) (5.43)


= 1̂S (ξ), (5.44)

86
Figure 5.8: The Fedorov solids, the only 3-dimensional polytopes that tile R3 by translations.
All 5 of them are zonotopes, and they are also extreme bodies for Minkowski’s convex body
theorem. The top three, from left to right, are: the Truncated octahedron, the Rhombic
dodecahedron, and the Hexarhombic dodecahedron. The bottom two are the cube and the
hexagonal prism.

for all ξ ∈ Rd , where we have used the assumption that 1̂S (ξ) is real-valued in the last
equality.

Example 5.4. The interval P := [− 12 , 21 ] is a symmetric polytope, and indeed we can see
that its Fourier transform 1̂P (ξ) is real-valued, namely we have 1̂P (ξ) = sinc(ξ), as we saw
in equation (3.6). 

Example 5.5. The cross-polytope ♦2 is a symmetric polytope, and as we verified in dimen-


sion 2, equation (3.57), its Fourier transform 1♦2 (ξ) is real-valued. 

Alexandrov [1], and independently Shephard [143], proved the following remarkable fact.

87
Theorem 5.9 (Alexandrov and Shephard). Let P be any real, d-dimensional polytope, with
d ≥ 3. If all of the facets of P are centrally symmetric, then P is centrally symmetric. 
Example 5.6. The converse to the latter result is clearly false, as demonstrated by the
cross-polytope in dimension d > 2: it is centrally symmetric, but its facets are not symmetric
because they are simplices and we know that no simplex (of dimension ≥ 2) is symmetric
(Exercise 11.9). 

Peter McMullen discovered the following


wonderful generalization of Theorem 5.9.
Theorem 5.10 (McMullen). Let P be any
real, d-dimensional polytope, with d ≥ 3, and
fix any positive integer k with 2 ≤ k ≤ d − 1.

If all the k-dimensional faces of P are sym-


metric, then all (k + j)-dimensional faces of
P are symmetric as well (including P itself ),
for all 1 ≤ j ≤ d − k.  Figure 5.9: A3 3-dimensional zonotope that
does not tile R by translations.
So the Alexandrov-Shephard Theorem 5.9
above follows as a special case of McMullen’s
theorem by setting k = d − 1.

One might wonder what happens if we ‘discretize the volume’ of a symmetric body K, by
counting integer points, and then ask for an analogue of Minkowski Theorem 5.1. In fact,
Minkowski already had a result about this too (and he had so many beautiful ideas that it’s
hard to put them all in one place!). We give Minkowski’s own elegant and short proof.
Theorem 5.11 (Minkowski, 1910). Let K ⊂ Rd be any d-dimensional, convex, symmetric
set. If the only integer point in the interior of K is the origin, then

K ∩ Zd ≤ 3d . (5.45)

Proof. We define the map φ : Zd → (Z/3Z)d , by reducing each coordinate modulo 3. Now
we claim that when restricted to the set K ∩ Zd , our map φ is 1 − 1. The statement of the
theorem follows directly from this claim. So let x, y ∈ K ∩ Zd , and suppose φ(x) = φ(y).
Then, by definition of the map φ, we have
1
n := (x − y) ∈ Zd , (5.46)
3
Now we define C to be the interior of the convex hull of x, −y, and 0. Because K is
symmetric, and x, y ∈ K, we know that −y ∈ K as well, so that C ⊂ int(K). Now using the

88
convexity of C, we also see that n ∈ C, because n is a non-trivial convex linear combination
of 0, x, −y.

Therefore n ∈ int(K) as well. Altogether, n ∈ int(K) ∩ Zd = {0}, which forces n = 0. Hence


x − y = 0.

Theorem 5.11 is often called Minkowski’s 3d theorem. An immediate and natural question
is: which bodies account for the ‘equality case’ ? One direction is easy to see: if K is the
integer cube whose vertices are ±e1 ± e2 ± · · · ± ed , then it is clear that K is symmetric about
the origin, and the only integer point in its interior is the origin. In addition, vol K = 2d ,
and K contains 3d integer points. It is a bit surprising, perhaps, that only in 2012 was it
proved that this integer cube is the only case of equality in Minkowski’s 3d theorem [106].

In a different direction, it turns out that the volume of the Minkowski symmetrized body
1
2
K − 12 K, which appeared quite naturally in some of the proofs above, can be related in a
rather precise manner to the volume of K itself. The consequence is the following inequality,
known as the Rogers-Shephard inequality [128],
   
1 1 2d
vol K ≤ vol K− K ≤ vol K, (5.47)
2 2 d

where equality on the left holds ⇐⇒ K is a symmetric body, and equality on the right holds
⇐⇒ K is a simplex (see Cassels [35]). There is also an extension of the Rogers-Shephard
inequality to two distinct convex bodies K, L ⊂ Rd :
 
2d
vol (K − L) vol (K ∩ L) ≤ vol K vol L. (5.48)
d

([128] and [71]). A quick way of proving (5.47) is by using the ubiquitous Brunn-Minkowski
inequality ([138], section 7.1) which tells us the following.

Theorem 5.12. If K and L are convex subsets of Rd , then


1 1 1
vol(K + L) d ≥ vol(K) d + vol(L) d , (5.49)

with equality if and only if K and L lie in parallel hyperplanes or are homothetic to each
other.

Two sets A, B ⊂ Rd are called homothetic if A = λB + v, for some fixed v ∈ Rd , and some
λ > 0 (or either A or B consist of just one point).

89
Notes
(a) Siegel’s original proof of Theorem 5.3 used Parseval’s identity, but the spirit of the two
proofs is similar.

(b) In Exercise 5.5 below, we see three equivalent conditions for a 2-simplex to be uni-
modular. In higher dimensions, a d-simplex will not satisfy all three conditions, and
hence this exercise shows one important ‘breaking point’ between 2-dimensional and
3-dimensional discrete geometry.

(c) There are a growing number of interesting books on the geometry of numbers. One
encyclopedic text that contains many other connections to the geometry of numbers
is Peter Gruber’s book [66]. Two other excellent and classic introductions are Siegel’s
book [145], and Cassels’ book [35]. An expository introduction to some of the elements
of the Geometry of numbers, at a level that is even appropriate for high school students,
is given by Olds, Lax, and Davidoff [110]. For upcoming books, the reader may also
consult Martin Henk’s lecture notes ‘Introduction to geometry of numbers’ [75], and
the book by Lenny Fukshansky and Stephan Ramon Garcia, ‘Geometry of Numbers’
[58].

(d) The Brunn-Minkowski inequality is fundamental to many branches of mathematics,


including the geometry of numbers. A wonderful and encyclopedic treatment of the
Brunn-Minkowski inequality, with its many interconnections, appears in [138].

(e) The Fedorov solids are depicted, and explained via the modern ideas of Conway and
Sloan, in an excellent expository article by David Austin [2]. For a view into the
life and work of Evgraf Stepanovich Fedorov, as well as a fascinating account of how
Fedorov himself thought about the 5 parallelohedra, the reader may consult the article
by Marjorie Senechal and R. V. Galiulin [141]. The authors of [141] also discuss the
original book of Fedorov, called An Introduction to the Theory of Figures, published
in 1885, which is now considered a pinnacle of modern crystallography. Fedorov later
became one of the great crystallographers of his time.
In R4 , it is known that there are 52 different combinatorial types of 4-dimensional
parallelohedra. In R5 , the complete classification of all the combinatorial types of
5-dimensional paralellohedra was completed in 2016 [50], where the authors found
110, 244 of them.

(f) The field of multi-tiling is still growing. One of the first important papers in this field
was by Mihalis Koloutzakis [87], who related the multi-tiling problem to a famous
technique known as the idempotent theorem, and thereby proved that if we have a
multi-tiling in R2 with any discrete set of translations, then we also have a multi-tiling
with a finite union of lattices. A recent advance is an equivalence between multi-tiling
and certain Hadwiger-type invariants, given by Nir Lev and Bochen Liu [92]. Here the
authors show as well that for a generalized polytope P ⊂ Rd (not necessarily convex

90
or connected), if P is spectral, then P is equidecomposable by translations to a cube
of equal volume.
Another natural question in multi-tiling, which is still open, is the following:
Question 4. Suppose that P multi-tiles with a discrete set of translations D. Do we
really need the set D of translates of P to be a very complicated discrete set, or is it
true that just a finite union of lattices suffices? Even better, perhaps one lattice always
suffices?

In this direction, Liu proved recently that if we assume that P multi-tiles with a finite
union of lattice, then P also multi-tiles with a single lattice [96]. This is big step in the
direction of answering Question 4 in general. An earlier, and smaller step, was taken
in [64], where the authors answered part of Question 4 in R3 , reducing the search from
an arbitrary discrete set of translations, to translations by a finite union of lattices.
Taken together, the latter two steps imply that in R3 (and in R2 ), any multi-tiling with
a discrete set of translations also occurs with just a one lattice.
In a different direction, the work of Gennadiy Averkov [3] analyzes the equality cases for
an extension of Minkowski’s theorem, relating those extremal bodies to multi-tilers. In
[161], Qi Yang and Chuanming Zong show that the smallest k for which we can obtain
a nontrivial k-tiling in R2 is k = 5, and the authors characterize those 5-tiling bodies,
showing in particular that if a convex polygon is a 5-tiler, then it must be either an
octagon, or a decagon.
Question 5. In Rd , what is the smallest integer k such that there exists a d-dimensional
polytope P that k-tiles Rd by translations?

(g) We say that a body P (any compact subset of Rd ) is ‘spectral’ if the function space
L2 (P) possesses an orthonormal, complete basis of exponentials. There is a fascinating
and vast literature about such spectral bodies, relating them to tiling, and multi-tiling
problems. One of the most interesting and natural questions in this direction is the
following conjecture, by Bent Fuglede.
The Fuglede conjecture asks whether the following is true.
Question 6. Does P tile Rd by translations ⇐⇒ P is spectral?

In the case that P is convex, one might hope that more is true. Indeed, in 2003 Alex
Iosevich, Nets Katz, and Terry Tao [76] proved that the Fuglede conjecture is true for
all convex domains in R2 . In 2021, this conjecture was proved for all convex domains
(which must necessarily be polytopes by an additional simple argument), in the work of
Nir Lev and Máté Matolcsi [93]. For non-convex bodies, Tao found a counter-example
to the Fuglede conjecture (Question 6).
In a related direction, Sigrid Grepstad and Nir Lev [62] showed that for any bounded,
measurable subset S ⊂ Rd , if S multi-tiles by translations with a discrete set, then S
has a Riesz basis of exponentials.

91
(h) We have seen in Theorem 5.5 that the zero set of the Fourier transform of a polytope is
very important, in that it gave us a necessary and sufficient condition for multi-tiling.
But the zero set of the FT also gives more information, and an interesting application
of the information content in the zero set is the Pompeiu problem. The Pompeiu
problem is an ancient problem (defined in 1929 by Pompeiu) that asks the following:
which bodies P ∈ Rd are uniquely characterized by the collection of their integrals over
P, and over all rigid motions of P? An equivalent formulation is the following.

Question 7. Does the vanishing of all of the integrals


Z
f (x)dx = 0, (5.50)
σ(P)

taken over all rigid motions σ that include translations, imply that f = 0?

A body P ⊂ Rd , for which the answer to the question above is affirmative, is said to
have the Pompeiu property. It is still an open problem, in general dimension, whether
all convex bodies have the Pompeiu property. It is known, by the work of Brown,
Schreiber, and Taylor [30] that P has the Pompeiu property ⇐⇒ the collection of
Fourier transforms 1̂σ(P) (z), taken over all rigid motions σ of Rd , have a common zero
z. It was also known that all polytopes have the Pompeiu property. Recently, in [99],
Fabricio Machado and SR showed that the zero set of the FT does not contain (almost
all) circles, and as a consequence we get a simple new proof that all polytopes have
the ‘Pompeiu property’.

Exercises

5.1. Suppose that in R2 , we are given a symmetric, convex body K of area 4, which contains
only the origin. Prove that B must tile R2 by translations.

5.2. ♣ Given convex d-dimensional bodies K, L ⊂ Rd , prove that K + L is convex, and that
K − L is convex.

5.3. ♣ Given any set C ⊂ Rd , prove that:


1 1
C is centrally symmetric ⇐⇒ C − C = C.
2 2

92
5.4. ♣ The support of a function f is defined here as

supp(f) := {x ∈ Rd | f(x) 6= 0}. (5.51)

Suppose that we are given two convex bodies A, B ⊂ Rd . Show that

supp(1A ∗ 1B ) = A + B,

where the addition is the Minkowski addition of sets. Note. Many books often use the closure
of the set (5.51) as the definition of the support of f , but for our purposes the definition above
is sufficient.

5.5. ♣ Suppose we have a triangle ∆ whose vertices v1 , v2 , v3 are integer points. Prove that
the following properties are equivalent:

(a) ∆ has no other integer points inside or on its boundary (besides its vertices).

(b) Area(∆) = 21 .

(c) ∆ is a unimodular triangle, which in this case means that v3 − v1 and v2 − v1 form a
basis for Z2 .

(Hint: You might begin by “doubling” the triangle to form a parallelogram.)

1
5.6. Show that in Rd , an integer simplex ∆ is unimodular if and only if vol(∆) = d!
.

5.7. In R3 , find an integer simplex ∆ that has no other integer points inside or on its
boundary (other than its vertices of course), but such that ∆ is not a unimodular simplex.

5.8. Prove that for any polytope P, 1̂P is not a Schwartz function.

5.9. ♣ (hard-ish) Show that if K is any convex body, then 1K ∗ 1K is a nice function, in
the sense of (4.23). In other words, show that the Poisson summation formula holds for the
function f (x) := (1K ∗ 1K ) (x).

Hint. Use the Parseval identity, valid for functions f ∈ L2 (Rd ). For this particular exercise,
feel free to use the results of all of the later sections (though in general we refrain from such
a ‘look ahead’).

93
5.10. Show that the truncated octahedron, defined in Example 5.3, tiles R3 by using only
translations with a lattice. Which lattice can you use for this tiling?

5.11. Define
√ f (x) := a sin x + b cos x, for constants a, b ∈ R. Show that the maximum value
of f is a + b2 , and occurs when tan x = ab .
2

5.12. Find an example of a symmetric polygon P ⊂ R2 that multi-tiles (nontrivially) with


multiplicity k = 5.

Notes. A trivial multi-tiling for P is by definition a multi-tiling that uses P, with some
multiplicity k > 1, but such that there also exists a 1-tiling (classical) using the same P.

5.13. Let K ⊂ Rd be centrally symmetric. Show that


 
1 1
K∩ K + n 6= φ ⇐⇒ n ∈ K.
2 2

5.14. ♣ Here we use Siegel’s theorem 5.4 to give the following extension of Minkowski’s
classical Theorem 5.2, but for bodies K that are not necessarily symmetric, nor necessarily
convex.

Namely, let K be any bounded, measurable subset of Rd , with positive d-dimensional volume.
Let B := 21 K − 12 K be the symmetrized body of K (hence B is a centrally symmetric set
containing the origin). Let L be a (full rank) lattice in Rd . Prove the following statement:

If vol K > 2d (det L), then B must contain a nonzero point of L in its interior.

Notes. We note that the positive conclusion of the existence of a nonzero integer point holds
only for the symmetrized body B, with no guarantees for any integer points in K.

94
Chapter 6

An introduction to Euclidean lattices

Lattices quantify the idea of periodic structures.


– Anonymous

Less is more.....more or less.


– Ludwig Mies van der Rohe

Figure 6.1: A fundamental parallelepiped (half-open), for a lattice L, generated by the


vectors v1 and v2 .

95
6.1 Intuition

We introduce Euclidean lattices here, which may be thought of intuitively as regularly-spaced


points in Rd , with some hidden number-theoretic structure. Another intuitive way to think
of lattices is that they are one of the most natural ways to discretize Euclidean space.
A lattice in Rd is also the most natural extension of an infinite set of equally-spaced points
on the real line. In the real-world, lattices come up very naturally when we study crystals,
for example.

It is perhaps not surprising that number theory comes in through study of the integer lattice
Zd , as it is the d-dimensional extension of the integers Z. Moreover, whenever we study
almost any periodic behavior, lattices naturally come up, essentially from the definition of
periodicity in Euclidean space. And of course, where there are lattices, there are Fourier
series, as we also saw in Chapter 4.

6.2 Introduction to lattices

Definition 6.1. A lattice is defined by the integer linear span of a fixed set of linearly
independent vectors {v1 , . . . , vm } ⊂ Rd :

L := n1 v1 + · · · + nm vm ∈ Rd all nj ∈ Z .

(6.1)

The most common lattice is the integer lattice



Zd := (x1 , . . . , xd ) ∈ Rd all xj ∈ Z .


However, we often encounter different types of lattices, occurring very naturally in practice,
and it is natural to ask how they are related to each other. The first thing we might notice
is that, by Definition 6.1, a lattice may also be written as follows:
     
 | | ... | n 1 n
 .1 
 

.
L := v1 v2 ... vm   ..   ..  ∈ Z
  m
:= M (Zm ), (6.2)
 | | ... |

nm nm

where by definition, M is the d × m matrix whose columns are the vectors v1 , . . . , vm . This
set of basis vectors {v1 , . . . , vm } is called a basis for the lattice L, and m is called the rank
of the lattice L. In this context, we also use the notation rank(L) = m.

We will call M a basis matrix for the lattice L. But there are always infinitely many other
bases for L as well, and Lemma 6.6 below shows how they are related to each other.

Most of the time, we will be interested in full-rank lattices, which means that m = d;
however, sometimes we will also be interested in lattices that have lower rank, and it is

96
important to understand them. The determinant of a full-rank lattice L := M (Zd ) is
defined by
det L := | det M |.
It is easy to prove that this definition is independent of the choice of basis matrix M , which
is the content of Lemma 6.6 below.
Example 6.1. In R1 , we have the integer lattice Z, but we also have lattices of the form rZ,
1
for any real number r. It’s√easy to show that any lattice in R
√ is of this latter type (Exercise
6.5). For example, if r = 2, then all integer multiples of 2 form a 1-dimensional lattice.

Example 6.2. In R2 , consider the lattice L generated by the two integer vectors v1 := ( −1
3 )
and v2 := (−4
1 ), drawn in Figure 6.1. A different choice of basis for the same lattice L is
−3 −8
{ −2 , −9 }, drawn in Figure 6.2. We note that det L = 11, and indeed the areas of both
half-open parallelepipeds equals 11. 

A fundamental parallelepiped for a lattice L with basis {v1 , . . . , vm } is:



D := λ1 v1 + · · · + λm vm all 0 ≤ λk < 1 , (6.3)
also known as a half-open parallelepiped.

Figure 6.2: A second fundamental parallelepiped for the same lattice L as in Figure 6.1

Given any lattice L ⊂ Rd , and any fixed fundamental parallelepiped D of L, we have the
pleasant property that D tiles Rd by translations with vectors from L. Put a little differently,
this useful fact tells us that any x ∈ Rd may be written uniquely as
x = n + f, (6.4)

97
where n ∈ L, and f ∈ D (Exercise 6.1).

How do we define the determinant of a general lattice L of rank r? We can start by observing
how the (squared) lengths of vectors in L behave w.r.t. a given basis of L:
* r r
+
X X X
kxk2 = cj vj , ck vk = cj ck hvj , vk i := cT M T M c, (6.5)
j=1 k=1 1≤j,k≤r

where M T M is an r×r matrix whose columns are basis vectors of L. With this as motivation,
we define: √
det L := M T M , (6.6)
called the determinant of the lattice L. This definition coincides, as it turns out, with
the Lebesgue measure of any fundamental parallelepiped of L (Exercise 6.15).

Given two lattices L ⊂ Rd , and M ⊂ Rd , such that


L ⊆ M,
we say that L is a sublattice of M. For example, Figure 6.3 shows a rank 1 sublattice of
the integer lattice Z2 , together with its determinant.

On the other hand, sublattices that have the same rank are very interesting, and quite useful
in applications. Given a sublattice L of M, both of the same rank, a crucial idea is to think
of all of the translates of L by an element of the coarser lattice M, which we call:

M/L := L + m m ∈ M . (6.7)
Each such translate L + m is called a coset of L in M, and the collection of all of these
cosets, namely M/L, is called the quotient lattice (or quotient group) .
Theorem 6.1. Let L ⊆ M be any two lattices of the same rank. Then

det L
1. det M
is an integer.
det L
2. The positive integer det M
is equal to the number of cosets of L in M. In other words,
det L
|M/L| = det M
.

For a proof of Theorem 6.1, see [58].


Example 6.3. Let M := Zd , and L := 2Zd , the sublattice consisting of vectors all of whose
coordinates
 d are even
d
integers. So L ⊂ M, and the quotient lattice M/L consists of the sets
2Z + n n ∈ Z . It is (almost) apparent that the number of elements of the latter set is

exactly 2d , so in our new notation we have Zd /2Zd = 2d .

We may also think of this quotient lattice Zd /2Zd as the discrete unit cube, namely {0, 1}d ,
a common object in theoretical computer science. 

98
Figure 6.3: A sublattice L ⊂ Z2 of rank 1, which has just one basis √ Here L has a
√ vector.
T
1-dimensional fundamental parallelepiped, showing that det L = v v = 5.

6.3 Discrete subgroups -


an alternate definition of a lattice

The goal here is to give another useful way to define a lattice. The reader does not need any
background in group theory, because the ideas here are self-contained, given some background
in basic linear algebra.
Definition 6.2. We define a discrete subgroup as a set S ⊂ Rd , together with the
operation of vector addition between all of its elements, which enjoys the following two
properties.

(a) [The subgroup property] If x, y ∈ S, then x − y ∈ S.

(b) [The discrete property] There exists a positive real number δ > 0, such that
the distance between any two distinct points of S is at least δ.

In particular, it follows from Definition 6.2 (a) that the zero vector must be in S, because
for any x ∈ S, it must be the case that x − x ∈ S. The distance function that we alluded to
in Definition 6.2 (b) is the usual Euclidean distance function, which we denote here by
v
u d
uX
kx − yk2 := t (xk − yk )2 .
k=1

99
Example 6.4. The lattice Zd is a discrete subgroup of Rd . In dimension 1, the lattice rZ is
a discrete subgroup of R, for any fixed r > 0. Can we think of discrete subgroups that are
not lattices? The answer is given by Lemma 6.1 below. 

The magic here is the following very useful way of going back and forth between this new
notion of a discrete subgroup of Rd , and our Definition 6.1 of a lattice. The idea of using
this alternate Definition 6.2, as opposed to our previous Definition 6.1 of a lattice, is that it
gives us a basis-free way of proving and discovering facts about lattices.

Lemma 6.1. L ⊂ Rd is a lattice ⇐⇒ L is a discrete subgroup of Rd . 

(For a proof see [66]).

Example 6.5. Given any two lattices L1 , L2 ⊂ Rd , let’s show that S := L1 ∩ L2 is also
a lattice. First, any lattice contains the zero vector, and it may be the case that their
intersection consists of only the zero vector. For any vectors x, y ∈ S, we also have x, y ∈ L1 ,
and x, y ∈ L2 , hence by the subgroup property of L1 and of L2 , we know that both x−y ∈ L1 ,
and x − y ∈ L2 . In other words, x − y ∈ L1 ∩ L2 := S. To see why the discrete property
of Definition 6.2 holds here, we just notice that since x − y ∈ L1 , we already know that
|x − y| > δ1 , for some δ1 > 0; similarly, because x − y ∈ L2 , we know that |x − y| > δ2 for
some δ2 > 0. So we let δ := min(δ1 , δ2 }, and we have shown that S is a discrete subgroup of
Rd . By Lemma 6.1, we see that S is a lattice.

If we had used Definition 6.1 of a lattice to show that S is indeed a lattice, it would require
us to work with bases, and the proof would be longer and less transparent, for this problem.

Example 6.6. Consider the following discrete set of points in Rd :
( d
)
X
Ad−1 := x ∈ Zd xk = 0 ,
k=1

for any d ≥ 2, as depicted in Figure 6.4. Is Ad a lattice? Using the definition 6.1 of a lattice,
it is not obvious that Ad is a lattice, because we would have to exhibit a basis, but it turns out
that the following set of vectors may be shown to be a basis: {e2 − e1 , e3 − e1 , · · · ed − e1 },
and hence Ad is a sublattice of Zd , of rank d − 1 (Exercise 6.10).

Just for fun, we will use Lemma 6.1 to show that Ad is indeed a lattice. To verify the
subgroup
Pd propertyPof Definition 6.2P(a) suppose that x, y ∈ Ad . Then by definition we have
d d
k=1 xk = 0 and k=1 yk = 0. So k=1 (xk − yk ) = 0, implying that x − y ∈ Ad .

To verify the discrete property of Definition 6.2 (b) suppose we are given two distinct points
x, y ∈ Ad . We can first compute their “cab metric” distance function, in other words the
L1 -norm defined by
kx − yk1 := |x1 − y1 | + · · · + |xd − yd |,

100
Figure 6.4: The lattice A1 , and the lattice A2 , with basis {v1 , v2 }

By assumption, there is at least one coordinate where x and y differ, say the k’th coordinate.
Then kx − yk1 := |x1 − y1 | + · · · + |xd − yd | ≥ 1, because all of the coordinates are integers,

and xk 6= yk by assumption. Since the L1 -norm and the L2 -norm are only off by d (by
Exercise 4.1), we have: √
dkx − yk2 ≥ kx − yk1 ≥ 1,
so the property 6.2 (b) is satisfied with δ := √1 , and we’ve shown that Ad is a lattice. 
d

We note that the lattices Ad defined in Example 6.6 are very important in many fields of
Mathematics, including Lie algebras (root lattices), Combinatorial geometry, and Number
theory.

6.4 Lattices defined by congruences

In this section we develop some of the theory in a concrete manner. A classic example of a
lattice defined by an auxiliary algebraic construction is the following. Suppose we are given
a constant integer vector (c1 , . . . , cd ) ∈ Zd , where we further assume that gcd(c1 , . . . , cd ) = 1.

101
Let
C := x ∈ Zd c1 x1 + · · · + cd xd ≡ 0

mod N , (6.8)
where N is a fixed positive integer.

Is C a lattice? Indeed, we can see that C is a lattice by first checking Definition 6.2 (a).
For any x, y ∈ C, we have c1 x1 + · · · + cd xd ≡ 0 mod N and c1 y1 + · · · + cd yd ≡ 0 mod N .
Subtracting these two congruences gives us c1 (x1 − y1 ) + · · · + cd (xd − yd ) ≡ 0 mod N , so
that x − y ∈ C. The verification of Definition 6.2 (b) if left to the reader, and its logic is
similar to Example 6.6.

There is even a simple formula for the volume of a fundamental parallelepiped for C:
det C = N, (6.9)
as we prove below, in (6.20). Perhaps we can solve an easier problem first. Consider the
discrete hyperplane defined by:

H := x ∈ Zd c1 x1 + · · · + cd xd = 0 ,


Is H a lattice? We claim that H itself is indeed a lattice (also known as a rank-(d − 1)


sublattice of Zd ), and since its verification is quite similar to the arguments above, we leave
this as Exercise 6.19.

The fundamental parallelepiped (which is (d − 1)-dimensional) of H also has a wonderful


formula, as follows. First, we recall a general fact (from Calculus/analytic geometry) about
hyperplanes, namely that the distance δ between any two parallel hyperplanes c1 x1 + · · · +
cd xd = k1 and c1 x1 + · · · + cd xd = k2 is given by
|k1 − k2 |
δ=p 2 . (6.10)
c1 + · · · + c2d
(see Exercise 6.3)

Lemma 6.2. For any discrete hyperplane H := x ∈ Zd c1 x1 + · · · + cd xd = 0 , we have:
q
det H = c21 + · · · + c2d . (6.11)

Proof.We first fix a basis {v1 , . . . , v


d−1 } for the (d − 1)-dimensional sublattice defined by
d
H := x ∈ Z | c1 x1 + · · · + cd xd = 0 . We adjoin to this basis one new vector, namely any
integer vector w that translates H to its nearest discrete hyperplane companion H + w,
where we define
H + w := x ∈ Zd c1 x1 + · · · + cd xd = 1 .


It’s easy to see that there are no integer points strictly between these two hyperplanes H
and H + w, and so the parallelepiped P formed by the edge vectors v1 , . . . , vd−1 , w is a
fundamental domain for Zd , hence has volume 1.

102
On the other hand, we may also calculate the volume of P by multiplying the volume of its
base times its height, using (6.10):

1 = vol P = (volume of the base of P)(height of P) (6.12)


= (det H) · δ (6.13)
1
= (det H) p 2 , (6.14)
c1 + · · · + c2d
p
and so det H = c21 + · · · + c2d .

To get started, it follows directly from the definition 6.8 of C that we may write the lattice
C as a countable, disjoint union of translates of H:

C := x ∈ Zd c1 x1 + · · · + cd xd = kN, where k = 1, 2, 3, . . . .

(6.15)

To be concrete, let’s work out some examples.

Example 6.7. Using Lemma 6.2, we can easily compute the determinant of the Ad lattice
from Example 6.6: √ √
det Ad = 1 + 1 + · · · + 1 = d.

Example 6.8. As in Figure 6.5, consider the set of all integer points (m, n) ∈ R2 that satisfy

2m + 3n ≡ 0 mod 4.

In this case the related hyperplane is the line 2x + 3y = 0, and the solutions to the latter
congruence may be thought of as a union of discrete lines:
  
x
2
C= ∈ Z 2x + 3y = 4k, and k ∈ Z .
y

In other words, our lattice C, a special case of (6.8), can in this case be visualized in Figure
6.5 as a disjoint union of discrete lines. If we denote the distance between any two of these
adjacent discrete lines by δ, then using (6.10) we have:
4
δ=√ . (6.16)
32 + 22
Finally, the determinant of our lattice C here is the area of the shaded parallelepiped:

det C = δ 32 + 22 = 4. (6.17)

103
Figure 6.5: The lattice of Example 6.8

Eager to prove the volume relation det C = N , we can use the ideas of Example 6.8 as a
springboard for this generalization. Indeed, Example 6.8 and the proof of Lemma 6.2 both
suggest that we should compute the volume of a fundamental parallelepiped P, for the lattice
C (as opposed to the lattice Zd ), by using a fundamental domain for its base, and then by
multiplying its volume by the height of P.
Lemma 6.3. Given a constant integer vector (c1 , . . . , cd ) ∈ Zd , with gcd(c1 , . . . , cd ) = 1, let

C := x ∈ Zd c1 x1 + · · · + cd xd ≡ 0 mod N ,

(6.18)

where N is a fixed positive integer. Then C is a lattice, and

det C = N.

Proof.
 We fix a basis {v1 , . . . , v d−1 } for the (d − 1)-dimensional sublattice defined by H :=
x ∈ Zd | c1 x1 + · · · + cd xd = 0 , and we adjoin to this basis one new vector, namely any
integer vector w that translates H to its nearest discrete hyperplane companion

H + w := x ∈ Zd c1 x1 + · · · + cd xd = N .


Together, the set of vectors {v1 , . . . , vd−1 , w} form the edge vectors of a fundamental par-
allelepiped P for the lattice C, whose hight δ is the distance between these two parallel

104
hyperplanes H and H + w. Using (6.10), we can may calculate the volume of P (which is
by definition equal to det C) by multiplying the volume of its ‘base’ times its ‘height’:

det C = (volume of the base of P)(height of P) = (det H)δ (6.19)


N
= (det H) p 2 = N, (6.20)
c1 + · · · + c2d
p
using the fact that det H = c21 + · · · + c2d from Lemma 6.2.

6.5 The Gram matrix

There is another very natural matrix that we may use to study lattices, which we can motivate
as follows. Suppose we are given any basis for a lattice L ⊂ Rd , say β := {v1 , . . . , vr }, where
1 ≤ r ≤ d. By definition L = M (Zd ), and rank(L) = r, where the columns of M are defined
by the basis vectors from β, and so M is a d × r matrix. We can therefore represent any
x ∈ Rd uniquely in terms of the basis β like this:

x = c1 v1 + · · · + cr vr , (6.21)

and the squared length of x is:


* r r
+
X X X
kxk2 = cj vj , ck vk = cj ck hvj , vk i := cT M T M c, (6.22)
j=1 k=1 1≤j,k≤r

where c := (c1 , . . . , cr )T is the coefficient vector defined by (6.21).

It’s therefore very natural to focus on the matrix M T M , whose entries are the inner products
hvj , vk i of all the basis vectors of the lattice L, so we define

G := M T M,

a Gram matrix for L. It’s clear from the computation above in (6.22) that G is positive
definite. Although G does depend on which basis of L we choose, it is an elementary fact
that det G is independent of the basis of L.

Because we are always feeling the urge to learn more Linear Algebra, we would like to see why
any real symmetric matrix B is the Gram matrix of some set of vectors. To see this, we apply
the Spectral Theorem: B = P DP T , for some orthogonal matrix P√ and a√diagonal matrix
D with nonnegative diagonal elements. √ So we can write B = (P D)(P D)T := M T M ,
T
where we defined the matrix M := (P D) , so that the columns of M are the vectors whose
corresponding dot products form the symmetric matrix B, and now B is a Gram matrix.

105
To review some more linear algebra, suppose we are given a real symmetric matrix A. We
recall that such a matrix is called positive definite if in addition we have the positivity
condition
xT Ax > 0,
for all x ∈ Rd . Equivalently, all of the eigenvalues of A are positive. The reason is easy:
Ax = λx for a non-zero vector x ∈ Rd implies that

xT Ax := hx, Axi = hx, λxi = λkxk2 ,

so that xT Ax > 0 if and only if λ > 0. In the sequel, if we only require a symmetric matrix
A that enjoys the property xT Ax ≥ 0 for all x ∈ Rd , then we call such a matrix positive
semidefinite.

Also, for a full-rank lattice, we see that B := M T M will be positive definite if and only if
M is invertible, so that the columns of M are a basis. Since a positive definite matrix is
symmetric by definition, we’ve proved:
Lemma 6.4. Suppose we are given a real symmetric matrix B. Then:

(a) B is positive definite if and only if it is the Gram matrix of a basis.


(b) B is positive semidefinite if and only if it is the Gram matrix of some set of vectors.

What about reconstructing a lattice, knowing only one of its Gram matrices? This is almost
possible to accomplish, up to an orthogonal transformation, as follows.
Lemma 6.5. Suppose that G is an invertible matrix, whose spectral decomposition is G =
P DP T . Then
G = X T X, (6.23)
if and only if √
X = Q( DP T ),
where Q is an orthogonal matrix.

Proof. The assumption G = X T X guarantees that G is symmetric and has positive eigen-
values, so by the Spectral Theorem we have:

G = P DP T ,

where D is a diagonal matrix consisting of the positive eigenvalues of G, and P is an orthog-


onal matrix consisting of eigenvectors of G. Setting X T X = P DP T , we must have
√ √
I = X −T P DP T X −1 = (X −T P D)(X −T P D)T , (6.24)

106

where we define D to be the diagonal matrix whose diagonal elements √ are the positive
−T
square roots of the eigenvalues of G. From
√ 6.24, −T
it follows that X P √D is an orthogonal
−T −T
matrix, let’s call it Q . Finally, X P D = Q implies that X = Q DP T .

So Lemma 6.5 allows us to reconstruct a lattice L, up to an orthogonal transformation,


by only knowing one of its Gram matrices. To better understand lattices, we need the
unimodular group, which we write as SLd (Z), under matrix multiplication:

SLd (Z) := M M is a d × d integer matrix, with | det M | = 1 . (6.25)
The elements of SLd (Z) are called unimodular matrices.
Example 6.9. Some typical elements of SL2 (Z) are
−1 0
  
S = −10 10 , T := 11 10 , and − I := 0 −1 ,
so we include matrices with determinant equal to −1 as well as 1. 

Any lattice L has infinitely many fundamental parallelepipeds and (Exercise 6.14) it is a
nice fact that they are all images of one another by the unimodular group. Now, suppose a
lattice L is defined by two different basis matrices: L = M1 (Zd ) and L = M2 (Zd ). Is there
a nice relationship between M1 and M2 ?
Lemma 6.6. If a full-rank lattice L ⊂ Rd is defined by two different basis matrices M1 , and
M2 , then
M1 = M2 U,
where U ∈ SLd (Z), a unimodular matrix.

In particular, det L is independent of the choice of basis matrix M .

Proof. By hypothesis, we know that the columns of M1 , say v1 , . . . , vd , form a basis of L,


and that the columns of M2 , say w1 , . . . , wd , also form a basis of L. So we can begin by
writing each fixed basis vector vj in terms of all the basis vectors wk :
d
X
vj = cj,k wk ,
k=1

for each j = 1, . . . , d, and for some cj,k ∈ Z. We may collect all d of these identities into
matrix form:
M1 = M2 C,
where C is the integer matrix whose entries are defined by the integer coefficients cj,k above.
Conversely, we may also write each basis vector wj in terms of the basis vectors vk : wj =
Pd
k=1 dj,k vk , for some dj,k ∈ Z, getting another matrix identity:

M2 = M1 D.

107
Altogether we have
M1 = M2 C = (M1 D)C,
and since M1−1 exists by assumption, we get DC = I, the identity matrix. Taking determi-
nants, we see that
| det D|| det C| = 1,
and since both C and D are integer matrices, they must belong to SLd (Z), by definition.
Finally, because, because a unimodular matrix U has | det U | = 1, we see that any two basis
M1 , M2 matrices satisfy | det M1 | = | det M2 |.

Using similar techniques, it is not hard to show the following fact (Exercise 6.13).

Theorem 6.2. Fix a full-rank lattice L ⊂ Rd . The group of one-to-one, onto, linear trans-
formations from L to itself is equal to the unimodular group SLd (Z).

6.6 Dual lattices

Every lattice L := M (Zd ) has a dual lattice, which we have already encountered in the
Poisson summation formula for arbitrary lattices. The dual lattice of L was defined by:

L∗ = M −T (Zd ). (6.26)

But there is another way to define the dual lattice, which is coordinate-free:

L∗ := x ∈ Rd | hx, ni ∈ Z, for all n ∈ L .



(6.27)

Lemma 6.7. The two definitions above, (6.26) and (6.27), are equivalent.

Proof. We let A := L∗ := M −T (Zd ), and B := x ∈ Rd hx, ni ∈ Z, for all n ∈ L . We first


fix any x ∈ A. To show x ∈ B, we fix any n ∈ L, and we now have to verify that hx, ni ∈ Z.
By assumption, x = M −T m for some m ∈ Zd , and n = M k, for some k ∈ Zd . Therefore

hx, ni = hM −T m, ni = hm, M −1 ni = hm, ki ∈ Z,

because both m, k ∈ Zd . So we have A ⊂ B. For the other direction, suppose that y ∈ B,


so by definition
hy, ni ∈ Z, for all n ∈ L.
We need to show that y = M −T k for some k ∈ Zd , which is equivalent to M T y ∈ Zd .
Noticing that the k’th element of M T y is hn, yi with n belonging to a basis of L, we are
done, by (6.7). Therefore A = B.

108
Example 6.10. Let L := rZd , the integer lattice dilated by a positive real number r. It’s
dual lattice is L∗ = 1r L, because a basis for L is M := rI, implying that a basis matrix for
L∗ is M −T = 1r I. We also notice that det L = rd , while det L∗ = r1d . 

A fundamental relation between a full-rank lattice and its dual follows immediately from
Definition 6.26: det(L∗ ) := det(M −T ) = det1M = det1 L , which we record as:

(det L)(det L∗ ) = 1. (6.28)

If we consider any integer sublattice of Zd , say L ⊂ Zd , together with its dual lattice L∗ in
the same space, some interesting relations unfold between them. Let’s consider an example.

Example 6.11. In R2 , let L := m ( 11 ) + n ( 14 ) m, n ∈ Z , a lattice with det L = 3 that
is depicted in Figure 6.6 by the larger green balls. Its dual lattice is
 
∗ 1 4 −1

L := (a ( −1 ) + b ( 1 )) a, b ∈ Z ,

3
whose determinant equals 31 , and is depicted in Figure 6.6 by the smaller orange balls. So L
is a coarser lattice than L∗ .

We can verify that the relation (6.28) works for this example: det L∗ = 13 = det1 L . We also
notice that L is a sublattice of L∗ . We may notice here that L∗ /L forms a finite group of
order 9 = (det L)2 , which is equal to the number of cosets of the coarser lattice L in the
finer lattice L∗ . 

The dual lattice also appears as the kernel of a certain map, as follows. Suppose that for
each point n ∈ L, we define a function called a character of L:

χn (y) := e2πihn,yi , (6.29)

whose domain is the whole space Rd . To see a connection with the dual lattice L∗ , we may
consider the simultaneous kernel of all of these functions taken together:

Ker := {x ∈ Rd | χn (x) = 1, for all n ∈ L}.

It’s clear that Ker = L∗ , because e2πiz = 1 if and only if z ∈ Z.

Now let’s consider the collection of all of these characters:

GL := {χn | n ∈ L}. (6.30)

If we multiply these character together by defining χn χm := χn+m , then GL forms a group,


called the group of characters of L. To see that this multiplication makes sense, we can
compute:
(χn χm )(x) = e2πihn,xi e2πihm,xi = e2πihn+m,xi = χn+m (x).

109
Figure 6.6: The lattice of Example 6.11, together with its dual lattice

Even more is true: GL is isomorphic, as a group, to the lattice L (Exercise 6.11) via the map
n → χn . Intuitively, one of the great benefits of group characters is that by using the magic
of just two-dimensional complex numbers, we can study high-dimensional lattices.
Example 6.12. For the integer lattice Zd , its group of characters is composed of the following
functions, by definition:
χn (x) := e2πihn,xi ,
for each n ∈ Zd . 

6.7 The successive minima of a lattice, and Hermite’s


constant

To warm up, we recall a very classical inequality of Hadamard, giving a bound on determi-
nants. Intuitively, Hadamard’s inequality tells us that if we keep all the lengths of the sides

110
of a parallelepiped constant, and consider all possible parallelepipeds P with these fixed side
lengths, then the volume of P is maximized exactly when P is rectangular.
Theorem 6.3 (Hadamard’s inequality). Given a non-singular matrix M , over the reals,
whose column vectors are v1 , . . . , vd , we have:

| det M | ≤ kv1 kkv2 k · · · kvd k,

with equality if and only if all of the vk ’s are pairwise orthogonal.

Proof. We use the following matrix decomposition from Linear Algebra: M = QR, where
Q is an orthogonal matrix, R := [ri,j ] is an upper-triangular matrix, and rkk > 0 (this
decomposition is a well-known consequence of the Gram-Schmidt Qd process applied to the
columns of M). So now we know that | det Q| = 1, and det R = k=1 rkk , and it follows that
d
Y
| det M | = | det Q det R| = det R = rkk .
k=1

Let’s label the columns of Q by Qk , and the columns of R by Rk . We now consider the
matrix M T M = RT QT QR = RT R. Comparing the diagonal elements on both sides of
M T M = RT R, we see that kvk k2 = kRK k2 . But we also have kRK k2 ≥ rkk 2
, so that
kvk k ≥ rkk . Altogether we have
d
Y d
Y
| det M | = rkk ≤ kvk k. (6.31)
k=1 k=1

The case of equality occurs if and only if kRK k2 = rkk


2
for all 1 ≤ k ≤ d, which means that R
is a diagonal matrix. Thus, we have equality in inequality (6.31) if and only if M T M = RT R
is a diagonal matrix, which means that the columns of M are mutually orthogonal.

A very important characteristic of a lattice L is the length of its shortest nonzero vector:
 

λ1 (L) := min kvk v ∈ L − {0} .

Every lattice has at least two shortest nonzero vectors, because if v ∈ L, then −v ∈ L.
Thus, when we use the words ‘its shortest vector’, we always mean that we are free to make
a choice between any of its vectors that have the same shortest, nonzero length. A natural
question, which has many applications, is “how short is the shortest nonzero vector in L, as
we somehow vary over all normalized lattices L?”
Example 6.13. We define the following lattice in R2 :

L := m ( 10211 ) + n ( 200 ) m, n ∈ Z .
16

111
What is the shortest nonzero vector in this lattice L? Without using any fancy Theorems,
we might still try simple subtraction, sort of mimicking the Euclidean algorithm. So for
−4
example, we might try ( 200 102
16 ) − 2 ( 11 ) = −6 , which is pretty short. So we seem to have
gotten lucky - we found a relatively short vector. But here comes the impending question:

Question 8. How do we know whether or not this is really the shortest nonzero vector in
our lattice L? Can we find an even shorter vector in L?

This is not easy to answer in general, and we need to learn a bit more theory even to approach
it in R2 . Moreover, in dimensions d ≥ 3, the corresponding problem of finding a shortest
nonzero vector in any given lattice is terribly difficult - it is considered to be one of the most
difficult problems in computational number theory. 

To capture the notion of the second-smallest vector in a lattice, and third-smallest vector,
etc, we begin by imagining balls of increasing radii, centered at the origin, and we can (at
least theoretically) keep track of how they intersect L.

Let Br be the ball of radius r, centered at the origin. For each fixed j, with 1 ≤ j ≤ d, let
r be the smallest positive real number such that Br contains at least j linearly independent
lattice points of L. This value of r is called λj (L), the j’th successive minima of the
lattice.

Another way of saying the same thing is:


 
λj (L) := min r > 0 dim span (L ∩ Br ) ≥ j . (6.32)

By definition, we have |λ1 (L)| ≤ |λ2 (L)| ≤ · · · ≤ |λd (L)|.

Example 6.14. For L := Zd , the shortest nonzero vector has length λ1 (Zd ) = 1, and the
successive minima for Zd all have the same value. One choice for their corresponding vectors
is v1 := e1 , . . . , vd := ed , the standard basis vectors. 

Example 6.15. In R2 , there is a very special lattice, sometimes called the hexagonal
lattice (also known as the Eisenstein lattice):
( √  )
3  
1
L := m 21 + n | m, n ∈ Z .
2
0

This lattice has det L = 23 and is generated by the 6’th roots of unity (Exercise 6.6). Given
the basis above, we see that here we have λ1 (L) = λ2 (L) = 1. It also turns out to be an
extremal lattice in the sense that it (more precisely a dilate of it) is the lattice that achieves
Hermite’s constant γ2 , below, over all lattices in R2 . (Exercise 6.7). 

112
Figure 6.7: The hexagonal lattice, also known as the Eisenstein lattice

Example 6.16. Let’s define the following family of 2-dimensional lattices. For each t > 0,
we let  t 
e 0
M := , and we let Lt := M (Zd ),
e−t
so that we get a parametrized family of lattices. While all of the lattices in this family have
det L = 1, their shortest nonzero vectors approach 0 as t → ∞, since λ1 (Lt ) = e−t . So we
see that it does not necessarily make sense to talk about the shortest nonzero vector among
a collection of lattices, but it will make sense to consider a “max-min problem” of this type
(Hermite’s constant (6.33) below). 

For each dimension d, we define Hermite’s constant as follows:



γd := max λ1 (L)2 L is a full-rank lattice in Rd , with det L = 1 .

(6.33)
In words, Hermite’s constant is retrieved by varying over all normalized lattices in Rd , which
have determinant 1, picking out the smallest squared norm of any nonzero vector in each
lattice, and then taking the maximum of these smallest norms. In a later chapter, on sphere
packings, we will see an interesting interpretation of Hermite’s constant in terms of the
densest lattice packing of spheres.

We next give a simple bound, in Theorem 6.4 below, for the shortest nonzero vector in a
lattice and hence for Hermite’s constant. But first we need to give a simple lower bound for
the volume of the unit ball, in Lemma 6.8. Curiously, Hermite’s constant γd is only known
precisely for 1 ≤ d ≤ 8, and d = 24, as of this writing.

113
Lemma 6.8.  d
2r
vol(Br ) ≥ √ .
d
n o
Proof. The cube C := x ∈ Rd all |xk | ≤ √rd is contained in the ball Br : if x ∈ C then
Pd  2
x 2
≤ d √r = r2 . So the volume of the ball Br is greater than the volume of the
k=1 k d
 d
cube, which is equal to √2rd .

The following result of Minkowski give a bound for the shortest nonzero vector in a lattice.

Theorem 6.4 (Minkowski). Suppose that L ⊂ Rd is a full-rank lattice. Then the shortest
nonzero vector v ∈ L satisfies √ 1
kvk ≤ d(det L) d . (6.34)
√ 1
In other words, λ1 (L) ≤ d(det L) d .

Proof. The idea is to apply Minkowski’s convex body Theorem 5.2 to a ball of sufficiently
large radius. Let r := λ1 (L) be the length of the shortest nonzero vector in L, and consider
the open ball Br of radius r. By definition, Br does not contain any lattice points of L. So
by Minkowski’s convex body Theorem, and Lemma 6.8,
 d
2λ1 (L)
√ ≤ vol(Br ) ≤ 2d det L.
d
√ 1
It follows that λ1 (L) ≤ d (det L) d , proving the claim.

Despite the latter bound (6.34) on the shortest nonzero vector in a lattice, there are currently
no known efficient algorithms to find such a vector, and it is thought to be one of the most
difficult problems we face today. In practice, researchers often use the LLL algorithm to find
a ‘relatively short’ vector in a given lattice, and the same algorithm even finds a relatively
short basis for L.

While we may not know explicitly all of the short vectors in a given lattice, it is often
still useful to construct an ellipsoid that is based on the successive minima of a lattice, as
we do below. In the spirit of reviewing basic concepts from Linear Algebra, an ellipsoid
boundary centered at the origin is defined by the (d − 1)-dimensional body
( d
)
2
X hx, b j i
x ∈ Rd 2
=1 , (6.35)
j=1
c j

114
for some fixed orthonormal basis {b1 , . . . , bd } of Rd . Here the vectors bj are called the
principal axes of the ellipsoid, and the cj ’s are the lengths along the principal axes of the
ellipsoid. A more geometric way of defining an ellipsoid (which turns out to be equivalent to
our definition above) is attained by applying a linear transformation M to the unit sphere
S d−1 ⊂ Rd (Exercise 6.20).

Figure 6.8: An ellipsoid in R3 .

Lemma 6.9. Corresponding to the successive minima of a full-rank lattice L, we have


d linearly independent vectors v1 , . . . , vd , so that by definition kvk k := λk (L). We apply
the Gram-Schmidt algorithm to this set of vectors {v1 , . . . , vd }, obtaining a corresponding
orthonormal basis {b1 , . . . , bd } for Rd .

Now, we define the following open ellipsoid by:


( d
)
2
X hx, b k i
E := x ∈ Rd 2 <1 , (6.36)
k=1
λ k

whose axes are the bk ’s, and whose radii are the λk := λk (L). We claim that E does not
contain any lattice points of L.

Pd v ∈ L. Let 1 ≤ k ≤ d
Proof. We fix any vector the maximal index such that λk (L) ≤ kvk.
beP
We may write v = j=1 hv, bj ibj , so that kvk = dj=1 hv, bj i2 .
2

Now v must , . . . vk } = span{b1 , . . . bk }, forP


Pdlie in span{v1P some 1 ≤ k ≤ d. Hence we may
write v = j=1 hv, bj ibj = j=1 hv, bj ibj , so that kvk = kj=1 |hv, bj i|2 . We now check if v
k 2

is contained in E:
d k k
X hv, bj i2 X hv, bj i2 1 X 2 kvk2
= ≥ hv, bj i = ≥ 1,
j=1
λj 2 j=1
λj 2 λk 2 j=1 λk 2

so that v 6∈ E.

115
More generally, we have the following refinement of Theorem 6.4, which gives us a bound on
the first d shortest (nonzero) vectors in a lattice.
Theorem 6.5 (Minkowski). The successive minima of a full-rank lattice L enjoy the prop-
erty:
  d1 √   d1
kλ1 (L)k · · · kλd (L)k ≤ d det L .

Proof. Using Lemma 6.9, the ellipsoid E contains no lattice points belonging to L, so that
by Minkowski’s convex body Theorem, we have vol E ≤ 2d det L. We also know that
d
! d
! d
Y Y 2
vol E = λj vol B1 ≥ λj √ .
j=1 j=1
d

Altogether, we have !
d d
d
Y 2
2 det L ≥ vol E ≥ λj √ ,
j=1
d
arriving at the desired inequality.

  d1
We notice that kλ1 (L)k · · · kλd (L)k ≥ kλ1 (L)k, because kλ1 (L)k ≤ kλk (L)k for all in-
dices 1 < k ≤ d. We therefore see that Theorem 6.5 is indeed a refinement of Theorem
6.4.

Example 6.17. The E8 lattice is defined by


( 8
)
 1 8 X
E8 := (x1 , x2 , · · · x8 ) ∈ Z8 ∪ Z + xk ≡ 0 mod 2 . (6.37)
2 k=1

It turns out that the E8 lattice gives the optimal solution to the sphere packing problem, as
well as the optimal solution for the kissing number problem in R8 . 

6.8 Hermite normal form

We call a lattice L an integral lattice if L ⊂ Zd . Further, we may recall that any lattice
L ⊂ Rd has infinitely many bases, so it may seem impossible at first to associate a single
matrix with a given lattice. However, there is an elegant way to do this, as follows.
Example 6.18. Suppose we are given a lattice L as the integral span of the vectors

v1 := ( 31 ) , v2 := ( −22 ) ,

116
which clearly has determinant 8. Then any integer linear combinations of v1 and v2 is still
in L. In particular, mimicking Gaussian elimination, we place v1 and v2 as rows of a matrix,
and row-reduce over the integers:
         
3 1 3 1 0 −8 1 3 1 3
→ → → → ,
−2 2 1 3 1 3 0 −8 0 8

where at each step we performed row operations (over Z) that did not change the lattice.
Hence we have a reduced basis for L, consisting of ( 13 ) and ( 08 ).

Figure 6.9: The lattice L of Example 6.18, depicted by the bold green points, and showing
the original basis {v1 , v2 } of L, and the Hermite-reduced basis of L

We notice that the resulting matrix is upper-triangular, with positive integers on the diago-
nal, nonnegative integers elsewhere, and in each column the diagonal element is the largest
element in that column.

There is another way to interpret the matrix reductions above, by using unimodular matrices,
as follows. The first reduction step can be accomplished by the multiplication on the left by
a unimodular matrix:     
1 0 3 1 3 1
=
1 1 −2 2 1 3
Similarly, each step in the reduction process can be interpreted by multiplying on the left
by some new unimodular matrix, so that at the  end of the process we have a product of
3 1
unimodular matrices times our original matrix . Because a product of unimodular
−2 2
matrices is yet another unimodular matrix, we can see that we arrived at a reduction of the
form:    
3 1 1 3
U = ,
−2 2 0 8
where U is a unimodular matrix. 

117
The point of Example 6.18 is that a similar matrix reduction persists for all integer lattices,
culminating in the following result, which just hinges on the fact that Z has a division
algorithm.
Theorem 6.6. Given an invertible integer d × d matrix M , there exists a unimodular matrix
U with U M = H, such that H satisfies the following conditions:

1. [H]i,j = 0 if i > j.
2. [H]i,i > 0, for each 1 ≤ i ≤ d.
3. 0 ≤ [H]i,j < [H]i,i , for each i > j.

Property 3 tells us that each diagonal element [H]i,i in the i’th column of H is the largest
element in the i’th column.

Moreover, the matrix H is the only integer matrix that satisfies the above conditions. 

The matrix H in Theorem 6.6 is called the Hermite normal form of M . To associate a
unique matrix to a given integral full-rank lattice L ⊂ Rd , we first choose any basis of L,
and we then construct a d × d integer matrix M whose rows are the basis vectors that we
chose. We then apply Theorem 6.6 to M , arriving at an integer matrix H whose rows are
another basis of L, called the Hermite-reduced basis.
Corollary 6.1. There is a one-to-one correspondence between full-rank integral lattices in
Rd and integer d × d matrices in their Hermite Normal Form. 

6.9 The Voronoi cell of a lattice

The Voronoi cell of a lattice L is defined by



V(L) := x ∈ Rd kxk ≤ kx − vk, for all v ∈ L .

(6.38)
In other words, the Voronoi cell of a lattice L is the set of all point in space that are closer
to the origin than to any other lattice point in L. Because the origin wins the battle of
minimizing this particular distance function, it is also possible to construct the Voronoi cell
by using half-spaces. Namely, for each v ∈ L, we define the half-space

Hv := x ∈ Rd hx, vi ≤ kvk ,


and observe that the Voronoi cell is also given by


\
V(L) = Hv .
v∈L−{0}

118
Example 6.19. The Dn lattice is defined by
( n
)
X
n
Dn := x ∈ Z xk ≡ 0 mod 2 ,
k=1

and is often called the “checkerboard” lattice. In particular, in R4 , the D4 lattice turns out
to be a fascinating object of study.

The Voronoi cell of D4 is called the 24-cell, as depicted in Figure 6.10, a 4-dimensional
polytope with some wonderful properties. It is one of the few polytopes that is self-dual. It
is also an example of a polytope P in the lowest possible dimension d (namely d = 4) such
that P tiles Rd by translations, and yet P is not a zonotope.

By (6.20) above, we see that det D4 = 2. 

Figure 6.10: The Voronoi cell of the D4 lattice in R4 , known as the 24-cell.

A fascinating open problem is the Voronoi conjecture, named after the Ukrainian mathe-
matician George Voronoi, who formulated it in 1908:

Conjecture 1 (Voronoi). A polytope P tiles Rd by translations if and only if P is the Voronoi


cell of some lattice L, or P is affinely equivalent to to such a Voronoi cell.

Example 6.20. For the lattice An ⊂ Rn+1 defined in Example 6.6, its Voronoi cell turns out
to have beautiful and important properties: A2 ⊂ R3 is a hexagon, A3 ⊂ R4 is a truncated
octahedron (one of the Fedorov solids), and so on (Conway and Sloane [38]).

119
6.10 Quadratic forms and lattices

The study of lattices is in a strong sense equivalent to the study of positive definite quadratic
forms, over integer point inputs, for the following simple reason. Any positive definite
quadratic form f : Rd → R is defined by f (x) := xT Ax, where A is a positive definite
matrix, so the image of the integer lattice under f is {xT Ax | x ∈ Zd }. On the other hand,
any full-rank lattice in Rd is given by L := M (Zd ), for some real (non-singular) matrix M .
By definition, this implies that the square of the norm of any vector in L has the following
shape: kvk2 = v T v = xT M T M x, for some x ∈ Zd . We notice that M T M in the last identity
is positive definite.

We may summarize this discussion as follows. Given any lattice L := M (Zd ), we have

kvk2 v ∈ L} = {xT Ax x ∈ Zd ,


where A := M T M is positive definite.

So the distribution of the (squared) norms of all vectors in a given lattice is equivalent to
the image of Zd under a positive definite quadratic form.

Interestingly, despite this equivalence, for an arbitrary given lattice L it is not known in
general whether the knowledge of the norms of all vectors in L uniquely determines the
lattice L. In very small dimensions it is true, but for dimensions ≥ 4 there are some
counterexamples due to Alexander Schiemann ([134], [135]).

The above equivalence between lattices in Rd and quadratic forms is straightforward but
often useful, because it allows both algebraic and analytic methods to come to bear on
important problems involving lattices.

Notes
(a) Kurt Mahler was one of the main contributors to the development of the Geometry
of Numbers. We mention here one of his more advanced results, involving limits
of lattices, called Mahler’s compactness theorem (also known as Mahler’s selection
theorem). So far we worked with one lattice at a time, but it turns out to be fruitful
to work with infinite sets of lattices.
Theorem 6.7 (Mahler). Fix ρ > 0, C > 0. Then any infinite sequence of lattices
L ⊂ Rd such that

min kxk x ∈ L − {0} ≥ ρ, and det L ≤ C,

has an infinite convergent subsequence of lattices.

120
In other words, Mahler realized that among all lattices of volume 1, if a sequence of
lattices diverges, then it must be true that the lengths of the shortest nonzero vectors
of these lattice tend to zero. To complete the story, we should define what it means
for a sequence of lattices {Ln }∞
n=1 to converge to a fixed lattice L. One way to define
this convergence is to say that there exists a sequence of bases βn of the lattices Ln
that converge to a basis β of L, in the sense that the j’th basis vector of βn converges
to the j’th basis vector of β.

(b) There is a well-known meme in Mathematics: “Can one hear the shape of a drum?”,
which is the title of Mark Kac’s famous paper regarding the desire to discern the shape
of a drum from its ‘frequencies’. An analogous question for lattices, studied by John
Conway, is “which properties of quadratic forms are determined by their representation
numbers?”. For further reading, there is the lovely little book by Conway called “The
sensual quadratic form”, which draws connections between quadratic forms and many
different fields of Mathematics [39].
Of course, no library is complete without the important and biblical “Sphere Packings,
Lattices and Groups”, by John H. Conway and Neil Sloane [38].

(c) The idea of periodicity, as embodied by any lattice in Rd , also occurs on other manifolds,
besides Euclidean space. If we consider a closed geodesic on a manifold, then it’s
intuitively clear that as we flow along that geodesic, we have a periodic orbit along
that geodesic. One important family of manifolds where this type of periodicity occurs
naturally is the family of Hyperbolic manifolds. Following the philosophy that ‘if we
have periodicity, then we have Fourier-like series’, we discover that there is also an
hyperbolic analogue of the Poisson summation formula, known as the Selberg trace
formula, and this type of number theory has proved extremely fruitful.

(d) A strong bound for Hermite’s constant in dimension d was given by Blichfeldt [22]:
   2
2 d d
γd ≤ Γ 2+ .
π 2

(e) The family of diagonal matrices in Example 6.16 is very important in the study of
homogeneous dynamics, because it acts by multiplication on the left, on the space
of all lattices that have det L = 1. This fascinating action is sometimes called the
“modular flow”, and was studied intensively by Etienne Ghys. A beautiful result in
this direction is that the periodic orbits of the modular flow are in bijection with the
conjugacy classes of hyperbolic elements in the modular group SL2 (Z), and furthermore
that these periodic orbits produce incredible knots in the complement of the trefoil
knot.

(f) Gauss initiated the systematic study of finding the minimum value of positive definite,
binary quadratic forms f (x, y) := ax2 + 2bxy + y 2 , over all integer inputs (x, y) ∈ Z2 .
Gauss’ theory is also known as a reduction theory for positive definite binary quadratic

121
forms, and is now a popular topic that can be found in many standard Number Theory
books. By the discussion above, in section 6.10, it is clear that minimizing positive
definite quadratic forms is closely related to finding a vector of smallest length in a
lattice, and to finding a “fat” fundamental parallelepiped for the lattice.
(g) It is clear that because lattices offer a very natural way to discretize Rd , they continue to
be of paramount importance to modern research. In particular, the theory of modular
forms, with linear (Hecke) operators that are defined using lattices and their fixed finite
index sublattices, is crucial for modern number theory. Euclidean lattices are also the
bread-and-butter of crystallographers.

Exercises

6.1. ♣ Given a lattice L ⊂ Rd , and any fundamental parallelepiped D of L, show that any
x ∈ Rd may be written uniquely as x = p + n, where n ∈ L, and p ∈ D.

Notes. This exercise gives a rigorous version of the statement “A fundamental parallelepiped
of a lattice L tiles Euclidean space by translations with L”.

6.2. ♣

1. Prove that (Zd )∗ = Zd .


2. More generally, prove that for any lattice L ⊂ Rd , we have (L∗ )∗ = L.

6.3. ♣ Show that the distance δ between any two parallel hyperplanes c1 x1 + · · · + cd xd = k1
and c1 x1 + · · · + cd xd = k2 is given by
|k1 − k2 |
δ=p 2 .
c1 + · · · + c2d

6.4. Given an integer point n ∈ Zd , we call the set of all integer multiples of n a lattice
line (also known as a rank-1 sublattice). Suppose we are given a full-rank, rational lattice
L ⊂ Rd (so that it has a rational basis matrix). Suppose, in addition, that we are also given
a fixed lattice line l1 .
Prove or disprove: the lattice L and the lattice line l1 always intersect in another lattice line:
L ∩ l1 = l2 ,
where l2 is another lattice line in Zd .

122
6.5. ♣ Let L be a lattice in R1 . Show that L = rZ for some real number r.

6.6. The hexagonal lattice is the 2-dimensional lattice defined by

L := {m + nω | m, n ∈ Z}, where ω := e2πi/3 .



3
Prove that det L = 2
, and give a description of the dual lattice to the hexagonal lattice.

6.7 (hard). Show that the hexagonal lattice attains the minimal value for Hermite’s constant
in R2 , namely γ22 = √23 .

6.8. Let L ⊂ R2 be any rank 2 lattice. Show that there exists a basis β := {v, w} of L such
that the angle θβ between v and w satisfies
π π
≤ θβ ≤ .
3 2

6.9. Suppose that M is a d × d matrix, all of whose d2 elements are bounded by B. Show
d
that | det M | ≤ B d d 2 .

(Hint: consider Hadamard’s inequality 6.3)


d
Notes. It follows from this exercise that if all of the elements of M are ±1, then | det M | ≤ d 2 .
Such matrices are important in combinatorics and are called Hadamard matrices. It is known
that if d > 2, then Hadamard matrices only exists when 4 | d. But for each d = 4m, it is
not known whether a d × d Hadamard matrix exists, except for very small cases.

6.10. ♣ Show that the following set of vectors is a basis for Ad :

{e2 − e1 , e3 − e1 , · · · , ed − e1 } ,

where the ej are the standard basis vectors. Hence Ad is a rank-(d − 1) sublattice of Zd , by
definition.

6.11. ♣ Recall that GL is the group of characters of the lattice L, under the usual multi-
plication of complex numbers, and that the lattice L is a group under the usual operation of
vector addition. Show that they are isomoprhic as groups: GL ' L.

123
6.12. ♣ Here we prove the orthogonality relations for characters of a lattice, for
sublattices of Zd . Fix a sublattice L of Zd , and let D be a fundamental parallelepiped for L.
Using the notation in Exercise 6.11, prove that for any two characters χa , χb ∈ GL , we have:
(
1 X 1 if χa = χb
χa (n)χb (n) = (6.39)
det L 0 if not.
n∈D∩Zd

6.13. ♣ Prove Theorem 6.2.

6.14. ♣ Prove that any two fundamental parallelepipeds (as defined in the text) of L, say
D1 and D2 , must be related to each other by an element of the unimodular group:

D1 = M (D2 ),

for some M ∈ SLd (Z).

6.15. ♣ Given a √sublattice L ⊂ Rd of rank r, show that our definition of its determinant,
namely det L := M T M , conincides with the Lebesgue measure of any of its fundamental
parallelepipeds.

(Here M is a d × r matrix whose columns are basis vectors of L)

6.16. Show that a set of vectors v1 , . . . , vm ∈ Rd , where 1 ≤ m ≤ d, are linearly independent


⇐⇒ their Gram matrix is nonsingular.

6.17. Prove that for any given lattice L ⊂ R2 , any set of shortest nonzero vectors for L
generate the lattice L.

Note. As a reminder, the first two shortest nonzero vectors of L may have equal length.

6.18. Find a lattice L ⊂ R5 such that any set of five shortest nonzero vectors of L do not
generate the lattice L.

6.19. ♣ Consider the discrete hyperplane defined by:



H := x ∈ Zd c1 x1 + · · · + cd xd = 0 ,


Show that H is a lattice (also known as a rank-(d − 1) sublattice of Zd ).

124
6.20. Here we give the details for (6.35), the definition of an ellipsoid in Rd . Starting over
again, we fix an orthonormal basis {b1 , . . . , bd } for Rd , and we define the following matrix:
 
| | ... |
M := c1 b1 c2 b2 ... cd bd  ,
| | ... |
where the ck ’s are positive scalars. We now apply the linear transformation M to the unit
sphere S d−1 := {x ∈ Rd | kxk2 = 1} in Rd , and we recall what this means. Now we define
the EllipsoidM := M (S d−1 ), a (d − 1)-dimensional object. In the spirit of review, we recall
the definition M (S d−1 ) := {u ∈ Rd | u = M x, x ∈ S d−1 }.

(a) Show that ( )


d 2
X hx, b j i
EllipsoidM = x ∈ Rd 2
=1 . (6.40)
j=1
c j


(b) We recall that the unit ball in Rd is defined by B d := x ∈ Rd kxk2 ≤ 1 . Show that
for the open ellipsoid body E (a d-dimensional object), as defined in (6.36), we have
the d-dimensional volume formula:
d
Y
d
vol(E) = vol(B ) cj .
j=1

6.21. We will use the equation (6.40) definition of an ellipsoid, from above. We can extend
the previous exercise in the following way. Let A be any d × d real matrix, and look at the
action of A on the unit sphere S d−1 ⊂ Rd . Suppose that rank(A) = r. Show:

(a) If r = d, then A(S d−1 ) is a d-dimensional ellipsoid, defined by an equation of the form
(6.40).

(b) If r < d, then A(S d−1 ) is an r-dimensional ellipsoid.

6.22. Suppose that A is a positive definite, real matrix. Solve for (i.e. characterize) all
matrices X that are the ‘square roots’ of A:

A = X 2.

6.23. Suppose that a certain 2-dimensional lattice L has a Gram matrix


 
2 −1
G := .
−1 2

Reconstruct L (i.e. find a basis for L), up to an orthogonal transformation.

125
6.24. Find a 2 by 2 matrix M that enjoys one of the properties of a positive semidefinite
matrix, namely that xT M x ≥ 0, for all x ∈ R2 , but such that M is not symmetric.

6.25. Show that any real 2 by 2 matrix A is positive definite if and only if both trace(A) > 0
and det A > 0.

6.26. (hard) Erdös’ question, given in Exercise 2.16, possesses a natural extension to dimen-
sion d.

Question 9. Suppose that the integer lattice Zd is partitioned into a disjoint union of a
finite number of translates of integer sublattices, say:

Zd = {L1 + v1 } ∪ {L2 + v2 } ∪ · · · ∪ {LN + vN }.

Is it true that there are at least two integer sublattices, say Lj , Lk , that enjoy the property
that Lk = Lj + w, for some integer vector w?

Here we prove that in R3 , Question 9 has a negative answer. In particular, find a partition of
Z3 into 4 integer sublattices, such that no two of them are integer translates of one another.
Using an easy extension to d > 3, also show that the answer to the question above is ‘no’, if
d ≥ 3.

Notes. Question 9 remains unsolved in dimension d = 2 [55].

126
Chapter 7

The Fourier transform of a polytope


via its vertex description:
The Brion theorems

See in nature the cylinder, the sphere, the cone.


– Paul Cézanne

Figure 7.1: The Dodecahedron in R3 , an example of a simple polytope. In Exercise 7.7, we


compute its Fourier-Laplace transform by using Theorem 7.1 below.

127
7.1 Intuition

Here we introduce the basic tools for computing precise expressions for the Fourier transform
of a polytope. To compute transforms here, we assume that we are given the vertices of a
polytope P , together with the local geometric information at each vertex of P, namely its
neighboring vertices in P ⊂ Rd . It turns out that computing the Fourier-Laplace transform
of the tangent cone at each vertex of P completely characterizes the Fourier transform of P.

One of the basic results here, called the discrete version of Brion’s Theorem (7.5), may
be viewed as an extension of the finite geometric sum in dimension 1, to sums in integer
cones, in dimension d. Some basic families of polytopes are introduced, including simple
polytopes and their duals, which are simplicial polytopes. These families of polytopes play
an important role in the development of Fourier analysis on polytopes.

Figure 7.2: The C60 Carbon molecule, also known as a buckeyball, is another example of a
simple polytope. The nickname “buckeyball’ came from Buckminster Fuller, who used this
molecule as a model for many other tensegrity structures. (the graphic is used with permis-
sion from Nanografi, at https://phys.org/news/2015-07-scientists-advance-tunable-carbon-
capture-materials.html)

128
7.2 Cones, simple polytopes, and simplicial polytopes

One of the most important concepts in combinatorial geometry is the definition of a cone
K ⊂ Rd , with an apex v, defined by;
( N
)
X
K := v + λk wk | λk ≥ 0 . (7.1)
k=1

The edge vectors of K are among the vectors w1 , . . . , wN . A fun exercise is to show that
the following two conditions are equivalent:

(a) A cone K has apex at the origin.

(b) K is a cone that enjoys the property λK = K, for all λ > 0.

(Exercise 7.10).

If the vectors w1 , . . . , wN span a k-dimensional subspace of Rd , we say that the cone K has
dimension k. When a k-dimensional cone K ⊂ Rd has exactly k linearly independent edge
vectors w1 , . . . wk ∈ Rd , we call such a cone a simplicial cone.

Figure 7.3: The cone on the left is pointed, and has edges w1 , w2 . The cone on the right,
with edges w1 , w2 , is also a half-space and it is not pointed.

A pointed cone is a cone K ⊂ Rd with apex v, such that its edge vectors w1 , . . . wN
are linearly independent. The following 4 conditions give equivalent characterizations of a
pointed cone K:

129
(a) There exists a hyperplane H such that H ∩ K = v.
(b) The translated cone C := K − v, with apex at the origin, enjoys C ∩ (−C) = {0}.
(c) K has a unique apex.
(d) K does not contain an entire line.

(Exercise 7.13). Part (d) is equivalent to the statement that for a non-pointed cone K, there
exists a vector u ∈ Rd such that K + u = K. We note that every cone has an apex, it’s
just that the apex may not be unique, for example when K is a half-space. All cones are
unbounded regions, by definition, so some care will have to be taken when integrating over
them. On the other hand, they are ‘almost linear’, because for a cone with apex at the
origin, we have
x, y ∈ K =⇒ x + y ∈ K.
This property makes them extremely helpful in the analysis of polytopes (for example, Sec-
tion 7.5).

An n-dimensional polytope P ⊂ Rd is called a simplicial polytope if every facet of P is a


simplex. Equivalently:

(a) Each facet of P has exactly n vertices.


(b) Each k-dimensional face of P has exactly k + 1 vertices, for 0 ≤ k ≤ n − 1.

It is a fun exercise to show that any simplicial cone is always a pointed cone (Exercise 7.11),
but the converse is clearly false.

By contrast with the notion of a simplicial polytope, we have the following ‘dual’ family of
polytopes.

An n-dimensional polytope P ⊂ Rd is called a simple polytope if every vertex is contained


in exactly n edges of P. Equivalently:

(a) Each vertex of P is contained in exactly n of its facets.


(b) Each k-dimensional face of P is contained in exactly d − k facets, for all k ≥ 0.

Example 7.1. Any d-dimensional simplex ∆ is a simple polytope. In fact, any k-dimensional
face of the simplex ∆ is also a simplex, and hence a simple polytope of lower dimension.

The 3-dimensional dodecahedron, in Figure 7.7, is also a simple polytope. Its edge graph,
which is always a planar graph for a convex polytope, consists of 20 vertices, 30 edges, and
12 faces. 

130
Example 7.2. A d-dimensional simplex also happens to be a simplicial polytope. The
3-dimensional icosahedron is a simplicial polytope. 

It is a nice exercise to show that the only polytopes which are both simple and simplicial
are either simplices, or 2-dimensional polygons (Exercise 7.14).

Example 7.3. The d-dimensional cube [0, 1]d is a simple polytope. Its dual polytope, which
is the cross-polytope ♦ (see (3.9)), is a simplicial polytope. 

One might ask: are the facets of a simple polytope necessarily simplicial polytopes? Again,
an example helps here.

Example 7.4. The 120-cell is a 4-dimensional polytope whose 3-dimensional boundary is


composed of 120 dodecahedra [137]. The 120-cell is a simple polytope, but because all of its
facets are dodecahedra, it does not have any simplicial facets. 

As becomes apparent after comparing the notion of a simple polytope with that of a simplicial
polytope, these two types of polytopes are indeed dual to each other, in the sense of duality
that we’ve already encountered in definition (3.64)

Lemma 7.1. P ⊂ Rd is a simple polytope ⇐⇒ P ∗ is a simplicial polytope.

(see Grünbaum [68] for a thorough study of this duality). This duality between simple
and simplicial polytopes suggests a stronger connection between our geometric structures
thus far, and the combinatorics inherent in the partially ordered set of faces of P. Indeed,
Grünbaum put it elegantly:

“In my opinion, the most satisfying way to approach the definition of polyhedra
is to distinguish between the combinatorial structure of a polyhedron, and the
geometric realizations of this combinatorial structure.” [69]

7.3 Tangent cones, and the Fourier transform of a sim-


ple polytope

An important step for us is to work with the Fourier-Laplace transform of a cone, and then
build some theorems that allow us to simplify many geometric computations, by using the
frequency domain on the Fourier transform side.

131
We may define the tangent cone of each face F ⊂ P as follows:

KF = {q + λ(p − q) | q ∈ F, p ∈ P, λ ∈ R≥0 } . (7.2)

We note that in general KF does not necessarily contain the origin. The tangent cone is
also known as the cone of feasible directions. Intuitively, we can imagine standing at the
point q ∈ F, and looking in the direction of all points that belong to P . Then we take the
union of all of these directions.

Figure 7.4: The triangle P has three vertex tangent cones: Kv1 , Kv2 , Kv3 . The picture is
meant to signify that these cones are, of course, unbounded.

In the case that the face F is a vertex of P, we call this tangent cone a vertex tangent
cone. The vertex tangent cone Kv , which is a cone with apex v, may also be generated by
the edge vectors vk − v, where [vk , v] is an edge of P:
N
X
Kv = {v + λk (vk − v) | all λk ≥ 0, and the vk are the neighboring vertices of v}, (7.3)
k=1

a construction we will often use in practice.

The tangent cone of an edge of a 3-dimensional convex polytope is an infinite wedge contain-
ing the whole line passing through that edge, while the tangent cone of a vertex (for a convex
polytope) never contains a whole line (Exercise 7.12). For non-convex polytopes, there are
many competing definition for the vertices, and not all of them agree. One definition for the
vertices of non-convex polytopes appears in [9], using Fourier transforms of cones. But in
this book we focus on convex polytopes.

132
Example 7.5. For the unit cube  := [0, 1]d , the tangent cone at the vertex v = 0 is

K0 = {λ1 e1 + λ2 e2 + λ3 e3 + · · · + λd ed | λk ≥ 0} ,

which also happens to be the positive orthant Rd≥0 . On the other hand, the tangent cone
of  at the vertex v = (1, 0, . . . , 0) is:

Kv = v + {λ1 (−e1 ) + λ2 e2 + λ3 e3 + · · · + λd ed | λk ≥ 0} ,

where ej is the standard unit vector along the j’th axis. 

Example 7.6. To relate some of these definitions, consider a d-dimensional simplex ∆ ⊂ Rd .


Located at each of its vertices v ∈ ∆, we have a tangent cone Kv , as in (7.3), and here Kv
is a simplicial cone. The simplex ∆ is both a simple polytope and a simplicial polytope. 

Brion proved the following extremely useful result, Theorem 7.1, concerning the Fourier-
Laplace transform of a simple polytope P. To describe the result, we consider each vertex v
of P, and we fix the d edges w1 , . . . , wd that emanate from v (not necessarily unit vectors).
Placing these edge vectors as columns of a matrix Mv , we define

det Kv := | det Mv |,

the absolute value of the determinant of the ensuing matrix.

Theorem 7.1 (Brion’s theorem - the continuous form, 1988). Let P ⊂ Rd be a


simple, d-dimensional polytope. For each vertex tangent cone Kv of P, we fix a set of edges
of Kv , say w1 (v), w2 (v), . . . , wd (v) ∈ Zd . Then
d
e−2πihv,zi det Kv
Z 
−2πihu,zi 1 X
e du = Qd (7.4)
P 2πi v a vertex of P k=1 hwk (v), zi

for all z ∈ Cd such that the denominators on the right-hand side do not vanish. 

The proof of Brion’s Theorem 7.1, which is one of the cornerstones of Fourier transforms
of polytopes, appears in section 7.6 below. We note that the determinant det Kv clearly
depends on our choice of edge vectors w1 , . . . , wd for the cone Kv , but it is straightforward
Kv
that the quotient Qd det
hw (v),zi
does not depend on the choice of edge vectors (Exercise 7.1).
k=1 k

133
7.4 Fourier-Laplace transforms of cones

First, let’s understand what it means to consider the complex vector z := x + iy ∈ Cd in the
transform Z
1P (z) := e−2πihu,zi du.
P

Our inner product hu, zi := u1 z1 + · · · + ud zd is always the usual inner product on Rd , defined
without using the Hermitian inner product here. In other words, we simply use the usual
inner product on Rd , and then formally substitute complex numbers zk into it. This means,
by definition, that
Z Z
−2πihu,zi
e du = e−2πihu,x+iyi (7.5)
P Z P

= e−2πihu,xi e2πhu,yi du, (7.6)


P

so that we have an extra useful real factor of e2πhu,yi that makes the integral converge quite
rapidly over unbounded domains, provided that hu, yi < 0. If we set y = 0, then it’s clear
that we retrieve the usual Fourier transform of P, while if we set x = 0, we get a new integral,
which we call the Laplace transform of P. Finally, the Fourier-Laplace transform of
P is defined by: Z
1̂P (z) := e−2πihu,zi du
P
d
valid for any z ∈ C for which the integral converges.

One clear reason for the use and flexibility of the full Fourier-Laplace transform is the fact
that for a cone K, its usual Fourier transform diverges. But if we allow a complex variable
z ∈ Cd , then the integral does converge on a restricted domain. Namely, the Fourier-Laplace
transform of a cone K is defined by:
Z
1̂K (z) := e−2πihu,zi du,
K

for a certain set of z ∈ Cd , but we can easily understand its precise domain of convergence.
For an arbitrary cone K ⊂ Rd , we define its polar cone by:

Ko := {y ∈ Rd | hy, ui < 0 for all u ∈ K},

which is an open cone. As one might expect, there is the following duality. If K1 ⊂ K2 , then
K2o ⊂ K1o (Exercise 7.15).

134
Example 7.7. Given the 1-dimensional cone K0 := R≥0 , we compute its Fourier-Laplace
transform:
Z Z ∞
−2πiuz 1
e du = e−2πiuz du = = e−2πiu(x+iy) |u=∞
u=0
K0 0 −2πiz
1
= e−2πiux e2πuy |u=∞
u=0
−2πiz
1 1 1
= (0 − 1) = ,
−2πiz 2πi z
valid for all z := x + iy ∈ C such that y < 0. We note that for such a fixed complex z,
|e−2πiuz | = e2πuy is a rapidly decreasing function of u ∈ R. 

Figure 7.5: A simplicial, pointed cone in R3 , with apex v and edge vectors w1 , w2 , w3

Now let’s work out the Fourier-Laplace transform of a d-dimensional cone whose apex is the
origin.

Lemma 7.2. Let K ⊂ Rd be a simplicial, d-dimensional cone, with apex at the origin. If
the edges of K are labelled w1 , . . . , wd , then
det K
Z
1
1̂K (z) := e−2πihu,zi du = d Qd .
K (2πi) k=1 hwk , zi

Furthermore, the domain of convergence for the latter integral is naturally associated with
the polar cone, and given by:

{z := x + iy ∈ Cd | y ∈ Ko }.

135
Proof. We first compute the Fourier-Laplace transform of the positive orthant K0 := Rd≥0 ,
with a complex vector z = x + iy ∈ Cd :
Z
1̂K0 (z) := e−2πihz,ui du (7.7)
K0
Z Z
−2πiz1 u1
= e du1 · · · e−2πizd ud dud (7.8)
R≥0 R≥0
d  d
Y 0−1 1 1
= = . (7.9)
k=1
−2πizk 2πi z1 z2 · · · zd

Next, the positive orthant K0 may be mapped to the cone K by a linear transformation.
Namely, we may use the matrix M whose columns are defined to be the edges of K, so that
by definition K = M (K0 ). Using this mapping, we have:
Z
1̂K (z) := e−2πihz,ui du
K Z
= | det M | e−2πihz,M ti dt
ZK0
T
= | det M | e−2πihM z,ti dt
K0
 d
1 | det M |
= Qd .
2πi k=1 hwk , zi

where in the second equality we’ve made the substitution u = M t, with t ∈ K0 , u ∈ K, and
du = | det M |dt. In the final equality, we used equation (7.9) above, noting that the k’th
element of the vector M T z is hwk , zi, and we note that by definition | det M | = det K.

For the domain of convergence of the integral, we observe that

e−2πihu,zi = e−2πihu,x+iyi = e−2πihu,xi e2πhu,yi ,



and because e−2πihu,xi = 1, the integral K e−2πihu,zi du converges ⇐⇒ hu, yi < 0 for all
R

u ∈ K. But by definition of the polar cone, this means that y ∈ Ko .

Example 7.8. Given the 2-dimensional cone K := {λ1 15 + λ2 −32 | λ1 , λ2 ∈ R≥0 }, we


 

compute its Fourier-Laplace transform, and find its domain of convergence. By Lemma 7.2,
Z
1 17
1̂K (z) := e−2πihu,zi du = 2
,
K (2πi) (z1 + 5z2 )(−3z1 + 2z2 )

valid for all z= ( zz12 ) :=x + iy such that y ∈ Ko . Here the polar cone is given here by
Ko = {λ1 −15 + λ1 −2 −3 | λ1 , λ2 ∈ R≥0 }. 

136
To compute the Fourier-Laplace transform of a simplicial cone K whose apex is v ∈ Rd ,
we may first compute the transform of the translated cone K0 := K − v, whose apex is at
the origin, using the previous lemma. We can then use the fact that the Fourier transform
behaves in a simple way under translations, namely

1̂K+v (z) = e2πihz,vi 1̂K (z),

to obtain the following result (Exercise 7.4).

Corollary 7.1. Let Kv ⊂ Rd be a simplicial d-dimensional cone, whose apex is v ∈ Rd .


Then
1 e−2πihv,zi det Kv
Z
1̂Kv (z) := e−2πihu,zi du = Qd , (7.10)
Kv (2πi)d k=1 hwk , zi

a rational-exponential function. More generally, for any d-dimensional cone Kv ⊂ Rd with


apex v, we can always triangulate Kv into M (v) simplicial subcones Kj (v) [45], and apply
the previous result to each simplicial subcone, obtaining:
M (v)
e−2πihv,zi X det Kj (v)
Z
−2πihu,zi
1̂Kv (z) := e du = d Qd , (7.11)
Kv (2πi) j=1 k=1 hwj,k (v), zi

a rational-exponential function.

There is an extension of Brion’s Theorem 7.1, due to Barvinok, for the Fourier transform
of a polytope, which allows us to drop the assumption that P is a simple polytope. This
extension has the merit that it is structurally the same as Theorem 7.1, for all real polytopes.
However, for a non-simple polytope, the question of computing efficiently the Fourier-Laplace
transforms of all of its tangent cones becomes unwieldy, as far as we know (this problem is
related to the P 6= N P problem, because computing the volume of a polytope is already
known to be NP-hard in general, and the volume is just the Fourier transform evaluated at
one point: vol P = 1P (0)).

The following fundamental result is the main Theorem in this chapter. When we write 1̂Kv (z)
below, by definition we mean its expression as a rational-exponential function, which is the
meromorphic continuation of the initial integral definition of the same object.

Theorem 7.2 (Fourier-Laplace transform of any real polytope). Let P ⊂ Rd be any


d-dimensional polytope. For each vertex v of P, we consider the tangent cone Kv of P. Then:
Z X
e−2πihu,zi du = 1̂Kv (z)
P v∈V

137
for almost all z ∈ Cd . We may also write
Z X e−2πihv,zi M
X (v)
det Kj (v)
−2πihu,zi
e du = Qd , (7.12)
P v∈V
(2πi)d j=1 k=1 hwj,k (v), zi

a rational-exponential function, using the notation of Corollary 7.1.

Proof. (See Section 7.6 below)

Example 7.9. Let’s work out a 2-dim’l example of Brion’s Theorem 7.1, using Fourier-
Laplace transforms of tangent cones. We will find the rational-exponential function for
the Fourier-Laplace transform of the triangle ∆, whose vertices are defined by v1 := ( 00 ),
v2 := ( a0 ), and v3 := ( 0b ), with a > 0, b > 0.

First, the tangent cone at the vertex v1 := ( 00 ) is simply the nonnegative orthant in this case,
with edge vectors w1 = ( 10 ) and w2 = ( 01 ). Its determinant, given these two edge vectors, is
equal to 1. Its Fourier-Laplace transform is
Z
1 1
e−2πihx,zi dx = , (7.13)
Kv1 (2πi)2 z1 z2

and note that here we must have both =(z1 ) > 0 and =(z2 ) > 0 in order to make the integral
converge. Here we use the standard notation =(z) is the imaginary part of z.

The second tangent cone at vertex v2 has edges w1 = ( −ab ) and w2 = ( −b0 ) (recall that we
don’t have to normalize the edge vectors at all). Its determinant has absolute value equal to
ab, and its Fourier-Laplace transform is

2
(ab)e−2πiaz1
Z 
−2πihx,zi 1
e dx = , (7.14)
Kv2 2πi (−az1 + bz2 )(−az1 )
and here the integral converges only for those z for which =(−az1 +bz2 ) > 0 and =(−az1 ) > 0.

Finally, the third tangent cone at vertex v3 has edges w1 = ( −ba ) and w2 = ( −b0 ). Its
determinant has absolute value equal to ab, and its Fourier-Laplace transform is

2
(ab)e−2πibz2
Z 
−2πihx,zi 1
e dx = . (7.15)
Kv3 2πi (az1 − bz2 )(−bz2 )
and here the integral converges only for those z for which =(az1 − bz2 ) > 0 and =(−bz2 ) > 0.

We can again see quite explicitly the disjoint domains of convergence in this example, so
that there is not even one value of z ∈ C2 for which all three Fourier-Laplace transforms of

138
all the tangent cones converge simultaneously. Despite this apparent shortcoming, Brion’s
identity (7.1) still tells us that we may somehow still add these local contributions of the
integrals at the vertices combine to give us a formula for the Fourier-Laplace transform of
the triangle:
2 
−b e−2πiaz1 −a e−2πibz2
Z  
−2πihx,zi 1 1
1̂∆ (z) := e dx = + + , (7.16)
∆ 2πi z1 z2 (−az1 + bz2 )z1 (az1 − bz2 )z2

which is now magically valid for all generic (z1 , z2 ) ∈ C2 ; in other words, it is now valid for
all (z1 , z2 ) ∈ C2 except those values which make the denominators vanish. 

Corollary 7.2. The Fourier-Laplace transform of any real polytope P ⊂ Rd , with vertex set
V , is a rational-exponential function of the form:
Z X e−2πihv,zi M
X (v)
det Kj (v)
−2πihu,zi
e du = Qd ,
P v∈V
(2πi)d j=1 k=1 hwj,k (v), zi

for all z ∈ Cd , except for those values of z for which any of the denominators vanish.

Proof. We recall (see [45]) that we may triangulate any cone into simplicial cones. Using
this algorithm, we can triangulate any tangent cone Kv into smaller simplicial cones Kj (v),
noting that the identity 1Kv = 1K1 (v) + · · · + 1KM (v) (v) implies the Fourier identity

1̂Kv = 1̂K1 (v) + · · · + 1̂KM (v) (v).

Hence Theorem 7.2 gives us:


Z X
e−2πihu,zi du = 1̂Kv (z)
P v∈V

XM
X (v)
= 1̂Kj (v)
v∈V j=1

XM
X (v) −2πihv,zi
e det Kj (v)
= Qd ,
v∈V j=1
(2πi)d k=1 hwj,k (v), zi

where the last step follows from Corollary 7.1.


Example 7.10. What is the Fourier transform of a hexagon?

Suppose we have a hexagon H that is symmetric about the origin; then we know that
its Fourier transform is real-valued, by Lemma 5.2. In this case it makes sense to form a
3-dimensional graph of the points (x, y, 1̂H (x, y)), as in Figure 7.6.

139
To be concrete, let’s define a (parametrized) hexagon H with the following vertices:
 2c   c   −c 
v1 = √ , 0 , v2 = √ , c , v3 = √ , c , v4 = −v1 , v5 = −v2 , v6 = −v3 ,
3 3 3
for each fixed parameter c > 0. Just for fun, our hexagon is scaled so that it has an inscribed
circle of radius c, which may be useful in future applications.

To use Brion’s theorem, we compute the Fourier Transforms of the 6 vertex tangent cones
of H. For v1 , the two rays defining Kv1 are w1 := v2 − v1 = (− √c3 , c) and w2 := v6 − v1 =
(− √c3 , −c), so the Fourier Transform of Kv1 is:
2c
−2πi √ z1 2c

2 √ − 4πic
√ z
e 3
3 2 3 e 3 1
1̂Kv1 (z) = = √ √ .
(−2πi)2 (− √c3 z1 + cz2 )(− √c3 z1 − cz2 ) (2π)2 (−z1 + 3z2 )(z1 + 3z2 )

For v2 , the two rays are w1 := v3 − v2 = (− √2c3 , 0) and w2 := v1 − v2 = ( √c3 , −c), giving us:

−2πi( √c z1 +cz2 ) 2c

2 √ −2πci( √1 z +z )
1 2
e 3
3 3 e 3
1̂Kv2 (z) = −2c = √ .
(−2πi)2 √ z1 ( √c z1 − cz2 ) (2π)2 z1 (z1 − 3z2 )
3 3

For v3 , the two rays are w1 := v4 − v3 = (− √c3 , −c) and w2 := v2 − v3 = ( √2c3 , 0), giving us:

−2πi(− √c z1 +cz2 ) √c
√ −2πci(− √1 z1 +z2 )
e 3
3 3 e 3
1̂Kv3 (z) = c 2c = √ .
(−2πi)2 (− √3 z1 − cz2 ) √3 z1 2
(2π) z1 (z1 + 3z2 )

By the inherent symmetry of our hexagon H, the computations for the other tangent cones
are just 1̂K−v (z) = 1Kv (−z), so we have:
Z
1̂H (z1 , z2 ) := e−2πihξ,zi dξ
H
= 1̂Kv1 (z) + 1̂Kv1 (−z) + 1̂Kv2 (z) + 1̂Kv2 (−z) + 1̂Kv3 (z) + 1̂Kv3 (−z)
√ 4πc 2πc
 2πc
 ! (7.17)
2 cos( √ z1 ) cos √ z1 + 2πcz2 cos √ z1 − 2πcz2
3 3 3 3
= 2 √ √ + √ + √ .
2π (−z1 + 3z2 )(z1 + 3z2 ) z1 (z1 − 3z2 ) z1 (z1 + 3z2 )

140
Figure 7.6: A graph of the Fourier transform 1̂H (x, y) of the symmetric hexagon H in
Example 7.10

7.5 The Brianchon-Gram identity

Brion’s theorem is particularly useful because whenever we are given a polytope in terms of
its local data at the vertices - including the edge vectors for each vertex tangent cone - we can
easily write down the Fourier transform of a simple polytope, by Theorem 7.1. Consequently,
we will be able to give a clean formulation for the volume of a simple polytope as well, as
we will soon see in Theorem 7.4.

The following combinatorial identity may be thought of as a geometric inclusion-exclusion


principle, and is quite general, holding true for any convex polytope, simple or not. For a
proof of the following result see, for example, [12] or [17].

Theorem 7.3 (Brianchon-Gram identity). Let P be any convex polytope. Then


X
1P = (−1)dimF 1KF , (7.18)
F ⊆P

where the sum takes place over all faces of P, including P itself. 

It turns out that the Brianchon-Gram relations (7.18) can be shown to be equivalent (in the
sense that one easily implies the other) to the Euler-Poincare relation (Exercise 7.17) for

141
the face-numbers of a convex polytope, which says that
f0 − f1 + f2 − · · · + (−1)d−1 fd−1 + (−1)d fd = 1. (7.19)
Here fk is the number of faces of P of dimension k.
Example 7.11. If we let P be a 2-dimensional polygon (including its interior of course)
with V vertices, then if must also have V edges, and exactly 1 face, so that (7.19) tells us
that V − V + 1 = 1, which is not very enlightening, but true. 
Example 7.12. If we let P be a 3-dimensional polytope with V vertices, E edge, and F
facets, then (7.19) tells us that V −E +F −1 = 1, which retrieves Euler’s well known formula
V −E+F =2
for the Euler characteristic of 3-dimensional polytopes. 

7.6 Proof of Theorem 7.1

In this section we provide a proof of Brion’s identity, namely Theorem 7.1, from first princi-
ples. This new proof uses some of the Fourier techniques that we’ve developed so far. Since
this book offers a friendly approach, we first give a short outline of the relatively simple ideas
of the proof:

Step 1. We begin with the Brianchon-Gram identity (a standard first step) involving the
indicator functions of all of the tangent cones of P.

Step 2. We now multiply both sides of the Brianchon-Gram identity (7.18) with the function
2
e2πihx,ξi−kxk , where we fix an  > 0, and then we will integrate over all x ∈ Rd . Using these
integrals, due to the damped Gaussians for each fixed  > 0, we are able to keep the same
domain of convergence for all of our ensuing functions.

Usually, the relevant Laplace integrals over the vertex tangent cones have disjoint domains
of convergence, lending the feeling that something magical is going on with the disjoint
domains of convergence. Getting around this problem by defining functions that have the
same domain of convergence was exactly the motivation for this proof.

Step 3. Now we let  → 0 and prove that the limits of each integral gives us something
meaningful. Using integration by parts, we prove that for any vertex tangent cone K the
R 2
corresponding integral K e2πihx,ξi−kxk dx converges, as  → 0, to the desired exponential-
rational function. In an analogous but easier manner, we will also prove that the corre-
sponding integral over a non-pointed cone (which includes all faces of positive dimension)
converges to zero.

142
Onto the rigorous details of the proof. We will make use of the following technical, but
crucial Lemma. We favor a slightly longer but clearer expositional proof over a shorter,
more obscure proof.

Lemma 7.3. Let Kv be a d-dim’l simplicial pointed cone, with apex v, and edge vectors
w1 , . . . , wd ∈ Rd . Then
d
e2πihv,zi |det Kv |

−1
Z
2πihx,ξi−kxk2
lim+ e dx = Qd , (7.20)
→0 Kv 2πi k=1 hw k (v), ξi
Qd
for all ξ ∈ Cd such that k=1 hwk (v), ξi 6 0.
=

Proof. We begin by noticing that we may prove the conclusion in the case that v = 0, the
origin, and for simplicity write Kv := K in this case. First we make a change of variables,
mapping the simplicial cone K to the nonnegative orthant Rd≥0 by the matrix M −1 , where
M is the d by d matrix whose columns are precisely the vectors wk . Thus, in the integral of
(7.20), we let x := M y, with y ∈ Rd≥0 , so that dx = |det M | dy. Recalling that by definition
det K = | det M |, We have
Z Z
2πihx,ξi−||x||2 2
e dx = |det K| e2πihM y,ξi−||M y|| dy. (7.21)
K Rd≥0

It is sufficient to therefore show the following limiting identity:


 d
−1
Z
2πihM y,ξi−||M y||2 1
lim+ e dy = Qd . (7.22)
→0 Rd≥0 2πi k=1 hwk (v), ξi

To see things very clearly, we first prove the d = 1 case. Here we must show that
Z ∞
2 −1
lim+ e2πixξ−x dx = , (7.23)
→0 0 2πiξ

for all ξ ∈ C − {0}, and we see that even this 1-dimensional case is interesting. We proceed
2
with integration by parts by letting dv := e2πixξ dx and u := e−x , to get
Z ∞ 2πixξ x=+∞ Z ∞ 2πixξ
2πixξ−x2 −x2 e e 2
e dx = e − (−2x)e−x dx (7.24)

2πiξ x=0 2πiξ

0 0
Z ∞
−1  2
= + xe2πixξ−x dx (7.25)
2πiξ πiξ 0
√ Z ∞
−1  2πi √u ξ 2
= + e ue−u du (7.26)
2πiξ πiξ 0

143

where we’ve used the substitution u := x in the last equality (7.26). We now notice that
Z ∞
2πi √u ξ 2
lim e ue−u du = lim ĝ(w),
→0 0 w→∞

2
where g(u) := ue−u 1[0,+∞] (u), and where w := √1 ξ. Luckily, as  → 0, our w → ∞. By
the Riemann–Lebesgue lemma 4.3, we know that limw→∞ ĝ(w) = 0, and so we arrive at the
desired limit (7.23).

We now proceed with the general case, which just uses the 1-dimensional idea above several
times. To prove (7.22), we first fix the variables y2 , . . . , yd and perform integration by parts
on y1 first. Thus, we let
t ξi
dv1 := e2πihM y,ξi dy1 = e2πihy,M dy1 = e2πi(y1 hw1 ,ξi+···+yd hwd ,ξi) dy1 , (7.27)

thought of as a function of only y1 . Carrying out the integration in the variable y1 , we have
t 2
v1 = e2πihy,M ξi / (2πihw1 , ξi). We let u1 := e−||M y|| , also thought of as a function of y1 alone.
2
We have du1 = −L(y)e−||M y|| dy1 , where L(y) is a real polynomial in y, with coefficients
from M . Integrating by parts in the variable y1 now gives us
Z Z  ∞ Z ∞ 
2πihM y,ξi−||M y||2
e dy = dy2 · · · dyd u1 v1 − v1 du1 (7.28)

Rd≥0 Rd−1 0 0
≥0
" t 2 Z ∞ #
e2πihy,M ξi−||M y|| y1 =∞
Z
 t 2
= dy2 · · · dyd + L(y)e2πihy,M ξi−||M y|| dy1
2πihw , ξi 2πihw , ξi

d−1
R≥0 1 y 1 =0 1 0
(7.29)
t 2 √
−e2πiht,M ξi−||M t||
Z Z
 t 2
= dt + L(y)e2πihy,M ξi−||M y|| dy (7.30)
Rd−1
≥0
2πihw1 , ξi 2πihw1 , ξi Rd≥0

−1
Z Z
2πiht,M t ξi−||M t||2  t 2
= e dt + L(y)e2πihy,M ξi−||M y|| dy, (7.31)
2πihw1 , ξi Rd−1 ≥0
2πihw1 , ξi Rd≥0

where we’ve used t := (y2 , . . . , yd ). Repeating exactly the same process of integration by
parts, we see that we get a sum of d terms, where the first term does not contain any
 factors, and all the other terms do contain  factors in the numerators, with positive
exponents. Therefore, when we complete the d-many integration by parts iteratively, and
−1 d 1

finally let  tend to zero, only the leading term remains, namely 2πi Qd . We’ve
k=1 hwk ,ξi
shown that (7.22) is true.

Proof. (of Theorem 7.1) We begin with the Brianchon Gram identity:
X
1P = (−1)dimF 1KF . (7.32)
F ⊆P

144
2
We fix any ξ ∈ Rd , and any  > 0. Multiplying both sides of (7.32) by e2πihx,ξi−kxk , and
integrate over all x ∈ Rd , we have:
Z Z
2πihx,ξi−kxk2 2
X
dimF
1P (x)e dx = (−1) 1KF (x)e2πihx,ξi−kxk dx. (7.33)
Rd F ⊆P Rd

Equivalently, Z Z
2πihx,ξi−kxk2 2
X
dimF
e dx = (−1) e2πihx,ξi−kxk dx. (7.34)
P F ⊆P KF

For each fixed  > 0, all integrands in (7.34) are Schwartz functions, and so all of the integrals
in the latter identity now converge absolutely (and rapidly). We identify two types of tangent
cones that may occur on the right-hand side of (7.34), for each face F ⊆ P.

Case 1. When F = v, a vertex, we have the vertex tangent cone Kv : these are the tangent
cones that exist for each vertex of P. It is a standard fact that all of these vertex tangent
cones areRpointed cones. Letting  → 0, and calling on Lemma 7.3, we obtain the required
2
limit for Kv e2πihx,ξi−kxk dx.

Case 2. When F is not a vertex, we have the tangent cone KF , and it is a standard fact
that in this case KF always contains a line. Another standard fact in the land of polytopes
is that each tangent cone in this case may be written as KF = Rk ⊕ Kp , the direct sum of a
copy of Euclidean space with a pointed cone Kp for any point p ∈ F. (as a side-note, it is
also true that dim F = k + dim(Kp )).

We would like to show that for all faces F that are not vertices of P, the associated integrals
tend to 0: Z
2
e2πihx,ξi−kxk dx → 0,
KF

as  → 0. Indeed,
Z Z
2πihx,ξi−kxk2 2
e dx = e2πihx,ξi−kxk dx (7.35)
KF k
R ⊕Kp
Z Z
2πihx,ξi−kxk2 2
= e dx e2πihx,ξi−kxk dx. (7.36)
Rk Kp

R 2
The integral Rk e2πihx,ξi−kxk dx is precisely the usual Fourier transform of a Gaussian, which
π 2
is known to be the Gaussian G (x) := −d/2 e−  kxk by Exercise 4.20. It is apparent that for
any fixed nonzero value of x ∈ Rd , we have lim→0 G (x) = 0. Finally, by Lemma 7.3 again,
R 2
the integral Kp e2πihx,ξi−kxk dx is finite, because Kp is another pointed cone. Therefore the
product of the integrals in (7.36) tends to zero, completing the proof.

145
7.7 An application of transforms to the volume of a
simple polytope, and for its moments

The following somewhat surprising formula for the volume of a simple polytope gives us
a very rapid algorithm for computing volumes of simple polytopes. We note that it is an
NP-hard problem [8] to compute volumes of general polytopes, without fixing the dimension.
Nevertheless, there are various other families of polytopes whose volumes possess tractable
algorithms.

Theorem 7.4 (Lawrence [91]). Suppose P ⊂ Rd is a simple, d-dimensional polytope. For a


vertex tangent cone Kv of P, fix a set of edges of the cone, say w1 (v), w2 (v), . . . , wd (v) ∈ Rd .
Then
(−1)d X hv, zid det Kv
vol P = (7.37)
d! v a vertex of P dk=1 hwk (v), zi
Q

for all z ∈ Cd such that the denominators on the right-hand side do not vanish. More
generally, for any integer k ≥ 0, we have the moment formulas:

(−1)d k! hv, zik+d det Kv


Z X
k
hx, zi dx = . (7.38)
(k + d)! v a vertex of P dm=1 hwm (v), zi
Q
P

Proof. We begin with Brion’s identity (7.7), and we substitute z := tz0 for a fixed complex
vector z0 ∈ Cd , and any positive real value of t:
d
e−2πihv,z0 it det Kv
Z 
−2πihu,z0 it 1 X
e du = .
2πi v a vertex of P td dm=1 hwm (v), z0 i
Q
P

Now we expand both sides in their Taylor series about t = 0. The left-hand-side becomes:
∞ d P∞ 1 j
j=0 j! (−2πihv, z0 it) det Kv
Z X 
1 k 1 X
(−2πihu, z0 it) du =
P k=0 k! 2πi v a vertex of P td dm=1 hwm (v), z0 i
Q

Integrating term-by-term on the left-hand-side, we get:


∞ d ∞ j−d
tk

det Kv
Z
X 1 X X t
(−2πi)k k
hu, z0 i du = Qd (−2πi)j hv, z0 ij .
k=0
k! P 2πi v a vertex of P m=1 hwm (v), z0 i j=0 j!

Comparing the coefficients of tk on both sides, we have:


d
(−2πi)k

det Kv
Z
1 X 1
k
hu, z0 i du = Qd (−2πi)k+d hv, z0 ik+d ,
k! P 2πi v a vertex of P m=1 hwm (v), z0 i (k + d)!

146
and simplifying, we arrive at the moment formulas:

hv, z0 ik+d det Kv


Z
k! X
hu, z0 ik du = (−1)d Qd .
P (k + d)! v a vertex of P m=1 hwm (v), z0 i

In particular, when k = 0, we get the volume formula (7.37).

7.8 Brion’s theorem - the discrete form

Example 7.13 (Finite geometric sums). Consider the 1-dimensional polytope P := [a, b],
where a, b ∈ Z. The problem is to compute the finite geometric series:
X X
e2πinz = qn,
n∈P∩Z a≤n≤b

where we’ve set q := e2πiz . Of course, we already know that it possesses a ‘closed form’ of
the type:
X q b+1 − q a
n
q = (7.39)
a≤n≤b
q−1
q b+1 qa
= − , (7.40)
q−1 q−1
because we already recognize this formula for a finite geometric sum. On the other hand,
anticipating the discrete form of Brion’s theorem below, we first compute the discrete sum
corresponding to the vertex tangent cone at the vertex a, namely a≤n q n :
P

qa
q a + q a+1 + · · · = . (7.41)
1−q
Now we compute the the sum corresponding to the vertex tangent cone at vertex b, namely
n
P
n≤b q :
qb q b+1
q b + q b−1 + · · · = = . (7.42)
1 − q −1 q−1
Summing these two contributions, one from each vertex tangent cone, we get:

qa q b+1 X
+ = qn,
1 − q q − 1 a≤n≤b

147
by the finite geometric sum identity, thereby verifying Theorem 7.5 for this example. This
example shows that Brion’s Theorem 7.5 (the discrete version) may be thought of as a
d-dimensional extension of the finite geometric sum.

But something is still very wrong here - namely, identity (7.41) converges for |q| < 1, while
identity (7.42) converges only for |q| > 1, so there is not even one value of q for which the
required identity (7.40) is true. So how can we make sense of these completely disjoint
domains of convergence ?! 

To resolve these conundrums, there is another very useful result of Michel Brion [28] that
comes to the rescue. We will discretize the continuous form of Brion’s Theorem 7.1, using
the Poisson summation formula, to arrive at a very useful, discrete form of Brion’s Theorem
(also due to Brion) using very different methods.

To this discrete end, we define the integer point transform of a rational polytope P by
X
σP (z) := ehn,zi .
n∈P∩Zd

We similarly define the integer point transform of a rational cone Kv by the series
X
σKv (z) := ehn,zi . (7.43)
n∈Kv ∩Zd

First, we need a slightly technical but easy Lemma.

Lemma 7.4. Let Kv be a rational cone, with apex at v. We pick any compactly supported
approximate identity φ , and we define:
X
1int Kv (x)e2πihx,zi ∗ φ (n).

RK (z) := lim
→0
n∈Zd

Then RK (z) is a rational-exponential function of z, valid for almost all z ∈ Cd , and is the
meromorphic continuation of the series defined by
X
σint Kv (z) := e2πihn,zi .
n∈Zd ∩int Kv

It turns out that the continuous form of Brion’s theorem, namely Theorem 7.1, can be used
to prove the discrete form of Brion’s theorem, namely Theorem 7.5 below.

148
Theorem 7.5 (Brion’s theorem - the discrete form, 1988). Let P ⊂ Rd be a rational,
d-dimensional polytope, and let N be the number of vertices of P. For each vertex v of P,
we consider the open vertex tangent cone int Kv of int P, the interior of P. Then
σint P (z) = σint Kv1 (z) + · · · + σint KvN (z). (7.44)
for all z ∈ Cd − S, where S is the hyperplane arrangement defined by the (removable)
singularities of all of the transforms 1̂Kvj (z).

Proof. We will use the continuous version of Brion, namely Theorem 7.1, together with the
Poisson summation formula, to deduce the discrete version here. In a sense, the Poisson
summation formula allows us to discretize the integrals.

Step 1. [Intuition] To begin, in order to motivate the rigorous proof that follows, we will
use Poisson summation on a function 1P (n)e2πihn,zi that “doesn’t have the right” to be used
in Poisson summation, because it is discontinuous. But this first step brings the intuition to
the foreground. Then, in Step 2, we will literally “smooth” out the lack of rigor in Step 1,
making everything rigorous.
X X
e2πihn,zi := 1P (n)e2πihn,zi
n∈P∩Zd n∈Zd
X
= 1̂P (z + ξ)
ξ∈Zd
X 
= 1̂Kv1 (z + ξ) + · · · + 1̂Kv1 (z + ξ)
ξ∈Zd
X X
= 1̂Kv1 (z + ξ) + · · · + 1̂KvN (z + ξ)
ξ∈Zd ξ∈Zd
X X
= 1Kv1 (n)e2πihn,zi + · · · + 1KvN (n)e2πihn,zi
n∈Zd n∈Zd
X X
:= e2πihn,zi + · · · + e2πihn,zi ,
n∈Zd ∩Kv1 n∈Zd ∩KvN

where we have used the Poisson summation formula in the second and fifth equalities. The
third equality used Brion’s Theorem 7.1.

Step 2 [Rigorous proof]. To make Step 1 rigorous, we pick any compactly supported ap-
proximate identity φ , and form a smoothed version of the function in step 1. Namely we
let
f (x) := (1P (x)e2πihx,zi ) ∗ φ (x),
so that now we are allowed to apply Poisson summation to f , because it is a Schwartz func-
tion. Recalling Theorem 4.14, we know that at a point x ∈ Rd of continuity of 1P (x)e2πihx,zi ,
we have
lim f (x) = 1P (x)e2πihx,zi .
→0

149
To proceed further, it is therefore natural to consider points x ∈ int P, the interior of P,
because 1P is continuous there, while it is not continuous on the boundary of P. To recap,
we have so far the equalities
X X
e2πihn,zi := 1int P (x)e2πihx,zi
n∈int P∩Zd n∈Zd
X
= lim f (n)
→0
n∈int P∩Zd
X
= lim f (n),
→0
n∈int P∩Zd

where we’ve used the fact that f is compactly supported, because it is the convolution of
two compactly supported functions. So the exchange above, of the sum with the limit, is
trivial because the sum is finite. With this in mind, the Poisson summation formula, applied
to the Schwarz function f , gives us:
X X X
e2πihn,zi = lim 1int P e2πihx,zi ) ∗ φ (n)

f (n) = lim
→0 →0
n∈int P∩Zd n∈int P∩Zd n∈Zd
X
e2πihx,zi ) ∗ φ (ξ)

= lim F (1int P
→0
n∈Zd
X
= lim 1̂int P (z + ξ)φ̂ (ξ)
→0
ξ∈Zd
X 
= lim 1̂int Kv1 (z + ξ) + · · · + 1̂int Kv1 (z + ξ) φ̂ (ξ)
→0
ξ∈Zd
X X
F (1int Kv1 e2πihx,zi ) ∗ φ (ξ) + · · · + lim F (1int KvN e2πihx,zi ) ∗ φ (ξ)
 
= lim
→0 →0
ξ∈Zd ξ∈Zd
X X
= lim (1int Kv1 e2πihx,zi ) ∗ φ (n) + · · · + lim (1int KvN e2πihx,zi ) ∗ φ (n)
→0 →0
n∈Zd ξ∈Zd

= σint Kv1 (z) + · · · + σint KvN (z),

using Lemma 7.4 in the last step.

Notes

(a) There is a large literature devoted to triangulations of cones, polytopes, and general
point-sets, and the reader is invited to consult the excellent and encyclopedic book on
triangulations, by Jesús de Loera, Jörg Rambau, and Francisco Santos [45].

150
(b) The notion of a random polytope has a large literature as well, and although we do
not go into this topic here, one classic survey paper is by Imre Bárány [8].
(c) The attempt to extend Ehrhart theory to non-rational polytopes, whose vertices have
some irrational coordinates, is ongoing. The pioneering papers of Burton Randol [108]
[124] extended integer point counting to algebraic polytopes, meaning that their ver-
tices are allowed to have coordinates that are algebraic numbers. Recently, a growing
number of papers are considering all real dilates of a rational polytope, which is still
rather close to the Ehrhart theory of rational polytopes.
In this direction, it is natural to ask how much more of the geometry of a given polytope
P can be captured by counting integer points in all of its positive real dilates. Suppose
we translate a d-dimensional integer polytope P ⊂ Rd by an integer vector n ∈ Zd .
The standard Ehrhart theory gives us an invariance principle, namely the equality of
the Ehrhart polynomials for P and P + n:
LP+n (t) = LP (t),
for all integer dilates t > 0.
However, when we allow t to be a positive real number, then it is in general false that
LP+n (t) = LP (t) for all t > 0.
In fact, these two Ehrhart functions are so different in general, that by the very re-
cent breakthrough of Tiago Royer [130], it’s even possible to uniquely reconstruct
the polytope P if we know all the counting quasi-polynomials LP+n (t), for all inte-
ger translates n ∈ Zd . In other words, the work of [130] shows that for two rational
polytopes P, Q ⊂ Rd , the equality LP+n (t) = LQ+n (t) holds for all integer translates
n ∈ Zd ⇐⇒ P = Q. It is rather astounding that just by counting integer points in
sufficiently many translates of P, we may completely reconstruct the whole polytope P
uniquely. Royer further demonstrated [131] that such an idea also works if we replace a
polytope by any symmetric convex body. It is now natural to try to prove the following
extended question.
Question 10. Suppose we are given polytopes P, Q ⊂ Rd . Can we always find a finite
subset S ⊂ Zd (which may depend on P and Q) such that
LP+n (t) = LQ+n (t) for all n ∈ S, and all t > 0 ⇐⇒ P = Q?

Exercises

7.1. ♣ Although det Kv depends on the choice of the length of each edge of Kv , show that
the ratio Qd | det
hw
Kv |
remains invariant if we replace each edge wk (v) of a simplicial cone by
k=1 k (v),zi
a constant positive multiple of it, say αk wk (v)

151
(Here z is any generic complex vector, meaning that hwk (v), zi =
6 0).

7.2. Consider the regular hexagon P ⊂ R2 , whose vertices are the 6’th roots of unity.

(a) Compute the area of P using Theorem 7.4.


(b) Compute all of the moments of P, as in Theorem 7.4.

7.3. Compute the Fourier transform of the triangle ∆ whose vertices are given by
(1, 0), (0, 1), (−c, −c),
where c > 0.

7.4. ♣ Prove Corollary 7.1 for a simplicial cone Kv , whose apex is v, by translating a cone
whose vertex is at the origin, to get:
1 e−2πihv,zi det Kv
Z
1̂Kv (z) := e−2πihu,zi du = Qd .
Kv (2πi)d k=1 hwk , zi

7.5. Consider the following 3-dimensional polytope P, whose vertices are as follows:
{(0, 0, 0), (1, 0, 0), (0, 1, 0), (1, 1, 0), (0, 0, 1)}.
“a pyramid over a square”. Compute its Fourier-Laplace transform 1̂P (z).

7.6. Werecall that the 3-dimensional cross-polytope (also called an octahedron) was defined
d
by ♦ := (x1 , x2 , x3 ) ∈ R | |x1 | + |x2 | + |x3 | ≤ 1 . Compute the Fourier-Laplace transform
of ♦ by using Theorem 7.2.

(Here not all of the tangent cones are simplicial cones, but we may triangulate each vertex
tangent cones into simplicial cones).

152
7.7 (hard-ish). Here we will find the Fourier transform of a dodecahedron P, centered at the
origin. Suppose we fix the following 20 vertices of P:
1 1 1
{(±1, ±1, ±1), (0, ±φ, ± ), (± , 0, ±φ), (±φ, ± , 0)},
φ φ φ

where φ := 1+2 5 . It turns out that P is a simple polytope. Compute its Fourier-Laplace
transform using Theorem 7.1.

Notes. All of the vertices of P given here can easily be seen to lie on a sphere S of radius 3,
and this is a regular embedding of the dodecahedron. It is also true (though a more difficult
fact) that these 20 points maximize the volume of any polytope whose 20 vertices lie on the
surface of this sphere S.

Figure 7.7: A climbing wall in Sweden, made up of Dodecahedrons, showing one of their
real-life applications

7.8. Define the 3-dimensional polytope P := conv{(0, 0, 0), (1, 0, 0), (0, 1, 0), (0, 0, 1), (a, b, c)},
where we fix real the positive real numbers a, b, c. Compute 1̂P (z), by computing the Fourier-
Laplace transforms of its tangent cones.

(Note. Here, not all of the tangent cones are simplicial cones).

7.9. This exercise extends Exercise 7.5 to Rd , as follows. Consider the d-dimensional
polytope P, called a “pyramid over a cube”, defined by the convex hull of the unit cube
[0, 1]d−1 ⊂ Rd−1 , with the point (0, 0, . . . , 0, 1) ∈ Rd . Compute its Fourier-Laplace transform
1̂P (z).

7.10. ♣ Show the following two conditions are equivalent:

153
(a) A cone K has apex at the origin.

(b) K is a cone that enjoys the property λK = K, for all λ > 0.

7.11. ♣ Suppose we are given a d-dimensional simplicial cone K ⊂ Rd (so be definition K


has exactly d edges). Show that K must be pointed.

7.12. ♣ Show that for any polytope P ⊂ Rd , a vertex tangent cone Kv never contains a
whole line.

7.13. ♣ Show that if K is a cone with an apex v (not necessarily a unique apex), the
following conditions are equivalent:

(a) K is a pointed cone.

(b) There exists a hyperplane H such that H ∩ K = v.

(c) The translated cone C := K − v, with apex at the origin, enjoys C ∩ (−C) = {0}.

(d) K has a unique apex.

(e) K does not contain an entire line.

7.14. ♣ Show that the only polytopes that are both simple and simplicial are either simplices,
or 2-dimensional polygons.

7.15. ♣ Show that if we have reverse inclusions for polar cones. Namely, if we have cones
K1 ⊂ K2 , then K2o ⊂ K1o .

7.16. Show that if we take the Minkowski sum K1 + K2 of two cones K1 , K2 ⊂ Rd , then

(K1 + K2 )o = K1o ∩ K2o .

7.17. ♣

(a) Show that the Brianchon-Gram relations (7.18) imply the Euler-Poincare relation for
the face-numbers of a convex polytope P:

f0 − f1 + f2 − · · · + (−1)d−1 fd−1 + (−1)d fd = 1, (7.45)

where fk is the number of faces of P of dimension k.

154
(b) (hard) Conversely, given a d-dimensional polytope P ⊂ Rd , show that the Euler-
Poincare relation above implies the Brianchon-Gram relations:
X
1P (x) = (−1)dimF 1KF (x),
F ⊂P

for all x ∈ Rd .

Notes. Interestingly, even though the above two conditions are equivalent, condition (b) is
often more useful because we may sum or integrate w.r.t. the variable x.

155
Chapter 8

The angle polynomial of a polytope

x
0

Figure 8.1: A discrete volume of the triange P, called the angle polynomial of P. Here we
sum local angle weights, relative to P, at all integer points.

8.1 Intuition

There are infinitely many ways to discretize the classical notion of volume, and here we offer
a second path, using ‘local solid angles’. Given a rational polytope P, we will place small
spheres at all integer points in Zd , and compute the proportion of the local intersection of
each small sphere with P. This discrete, finite sum, gives us a new method of discretizing
the volume of a polytope, and it turns out to be a more symmetric way of doing so. To go
forward, we first discuss how to extend the usual notion of ‘angle’ to higher dimensions, and
then use Poisson summation again to pursue the fine detail of this new discrete volume.

156
8.2 What is an angle in higher dimensions?

The question of how an angle in two dimensions extends to higher dimensions is a basic one
in discrete geometry. A natural way to extend the notion of an angle is to consider a cone
K ⊂ Rd , place a sphere centered at the apex of K, and then compute the proportion of the
sphere that intersects K. This intuition is captured more rigorously by the following integral:
Z
2
ωK = e−πkxk dx. (8.1)
K

called the solid angle of the cone K.

The literature has other synonyms for solid angles, arising in different fields, including the
volumetric moduli [65], and the volume of a spherical polytope [17], [46], [49].

Figure 8.2: A solid angle in R3 - note the equivalence with the area of the geodesic triangle
on the sphere.

We can easily show that the latter definition of a solid angle is equivalent to the volume of
a spherical polytope, using polar coordinates in Rd , as follows. We denote the unit sphere
by S d−1 := {x ∈ Rd |Rkxk = 1}. Then using the fact that the Gaussians give a probability
2
distribution, namely Rd e−π||x|| dx = 1 (which we know by Exercise 4.19), we have
R −πkxk2 R ∞ −πr2 d−1 R
K
e dx 0R
e r dr S d−1 ∩K dθ
ωK = R 2
−πkxk dx
= ∞ −πr2 r d−1 dr
R (8.2)
d e 0
e S d−1

RR
d−1 dθ
= SR ∩K (8.3)
S d−1


vold−1 K ∩ S d−1
= , (8.4)
vold−1 (S d−1 )

157
where vold−1 denotes the volume measure on the surface of the (d − 1)-dimensional sphere
S d−1 . We may think of (8.4) as the normalized volume of a spherical polytope defined by
the intersection of the cone K with the unit sphere. Thus for any cone K ⊂ Rd , we have

0 ≤ ωK ≤ 1.

We used polar coordinates in the second equality (8.2) above: x = (r, θ), with r ≥ 0, θ ∈
S d−1 . The Jacobian in the change of variables is dx = rd−1 drdθ.

We note that when K = Rd , so that the cone is all of Euclidean space, the integral (8.1)
becomes Z
2
e−π||x|| dx = 1,
Rd
by Exercise 4.19. This computation confirms that we do indeed have the proper normaliza-
tion with ωK = 1 if and only if K = Rd .

Example 8.1. If K ⊂ Rd is a half-space, then ωK = 12 . If K := Rd≥0 , the positive orthant,


then
Z Z !d
2 2 1
ωK = e−π||x|| dx = e−πu du = d .
Rd≥0 R≥0 2

8.3 Local solid angles for a polytope, and Gaussian


smoothing

Here we want to define solid angles relative to a fixed polytope. So given any polytope
P ⊂ Rd , we fix any point x ∈ Rd and define a local solid angle relative to P as follows.
The normalized solid angle fraction that a d-dimensional polytope P subtends at any point
x ∈ Rd is defined by
vol(S d−1 (x, ) ∩ P)
ωP (x) = lim . (8.5)
→0 vol (S d−1 (x, ))

Here, ωP (x) measures the fraction of a small (d − 1)-dimensional sphere S d−1 (x, ) centered
at x, that intersects the polytope P. We will use the standard notation for the interior of
a convex body, namely int(P), and for the boundary of a convex body, namely ∂P. As a
side-note, we mention that balls and spheres can be used interchangeably in this definition,
meaning that the fractional weight given by (8.5) is the same using either method (see
Exercise 9.13).

158
It follows from the definition of a solid angle that 0 ≤ ωP (x) ≤ 1, for all x ∈ Rd , and that
(
1 if x ∈ int(P)
ωP (x) =
0 if x ∈/ P.
But when x ∈ ∂P, we have ωP (x) > 0. For example, if x lies on a codimension-two face of
P, then ωP (x) is the fractional dihedral angle subtended by P at x.

Returning to discrete volumes, Ehrhart and Macdonald analyzed a different discrete volume
for any polytope P. Namely, for each positive integer t, define the finite sum
X
AP (t) := ωtP (n), (8.6)
n∈Zd

where tP is the t’th dilation of the polytope P. for P. In other words, AP (1) is a new
discrete volume for P, obtained by placing at each integer point n ∈ Zd the weight ωtP (x),
and summing all of the weights.
Example 8.2. In Figure 8.1, the solid angle sum of the polygon P is
AP (1) = θ1 + θ2 + θ3 + 3 21 + 4 = 6.


Here the θj ’s are the three angles at the vertices of P. 

Using purely combinatorial methods, Macdonald showed that for any integer polytope P,
and for positive integer values of t,
(
a1 t if d is odd,
AP (t) = (vol P)td + ad−2 td−2 + ad−4 td−4 + · · · + (8.7)
a2 t2 if d is even.
We will call AP (t) the angle-polynomial of P, for integer polytopes P and positive inte-
ger dilations t. However, when these restrictions are lifted, the sum still captures crucial
geometric information of P, and we will simply call it the (solid) angle-sum of P.

We define the heat kernel, for each fixed positive , by


d π 2
G (x) := − 2 e−  kxk , (8.8)
R
for all x ∈ Rd . By Exercises 4.19 and 4.20, we know that Rd
G (x)dx = 1 for each fixed ,
and that
2
Ĝ (ξ) = e−πkξk . (8.9)

The convolution of the indicator function 1P by the heat kernel G will be called the Gaus-
sian smoothing of 1P :
Z Z
(1P ∗ G )(x) := 1P (y)G (x − y)dy = G (y − x)dy (8.10)
Rd Z P
d π 2
= − 2 e−  ky−xk dy, (8.11)
P

159
a C ∞ function of x ∈ Rd , and in fact a Schwartz function (Exercise 9.9). The following
Lemma provides a first crucial link between the discrete geometry of a local solid angle and
the convolution of 1P with a Gaussian-based approximate identity.

Lemma 8.1. Let P be a full-dimensional polytope in Rd . Then for each point x ∈ Rd , we


have
lim(1P ∗ G )(x) = ωP (x). (8.12)
→0

Proof. We have
Z
(1P ∗ G )(x) = G (y − x)dy
ZP Z
= G (u)du = G1 (v)dv.
1
u∈P −x √

(P −x)

In the calculation above, we make use


√ of the evenness of G in the second equality. The
substitutions u = y − x and v = u/  are also used. Following those substitutions, we
change the domain of integration from P to the translation P − x, and to the dilation of
P − x by the factor √1 .

Now, when  approaches 0, √1 (P − x) tends to the cone K at the origin, subtended by P − x.
The cone K is in fact a translation of the tangent cone of P at x. Thus, we arrive at
Z
lim(1P ∗ G )(x) = G1 (v)dv = ωK (0) = ωP (x).
→0 K

Putting things together, the definition 8.6 and Lemma 8.1 above tell us that
X X
AP (t) = ωtP (x) = lim(1tP ∗ G )(n). (8.13)
→0
n∈Zd n∈Zd

We would like to interchange a limit with an infinite sum over a lattice, so that we may use
Poisson summation, and although this is subtle in general, it’s possible to carry out here,
because the summands are rapidly decreasing.

Lemma 8.2. Let P be a full-dimensional polytope in Rd . Then


X
AP (t) = lim (1tP ∗ G )(n). (8.14)
→0
n∈Zd

160
(For a proof of Lemma 8.2 see [49]).

Next, we apply the Poisson summation formula to the Schwartz function


f (x) := (1P ∗ G )(x):
X
AP (t) = lim (1tP ∗ G )(n) (8.15)
→0
n∈Zd
X
= lim 1̂tP (ξ)Ĝ (ξ) (8.16)
→0
ξ∈Zd
2
X
= lim 1̂tP (ξ) e−πkξk (8.17)
→0
ξ∈Zd
2
X
= td lim 1̂P (tξ) e−πkξk (8.18)
→0
ξ∈Zd
2
X
= td 1̂P (0) + lim td 1̂P (tξ) e−πkξk (8.19)
→0
ξ∈Zd −{0}
2
X
= td (vol P) + lim td 1̂P (tξ) e−πkξk , (8.20)
→0
ξ∈Zd −{0}

where we used the fact that Fourier transforms interact nicely with linear transformations
of the domain:
Z Z Z
−2πihξ,xi −2πihξ,tyi
1̂tP (ξ) = e dx = td
e dy = td
e−2πihtξ,yi dy = td 1̂P (tξ).
tP P P

We also used the simple change of variable x = ty, with y ∈ P, implying that dx = td dy, as
well as the Fourier transform formula for the heat kernel (8.9).

Altogether, we now have:


X
AP (t) = td (vol P) + td lim (1̂P (tξ) ∗ G )(n), (8.21)
→0
n∈Zd −{0}

suggesting a polynomial-like behavior for the angle polynomial AP (t).

The next step will be to use our knowledge of the Fourier transform of the polytope P, in
the right-hand-side of (8.21), for which even a 1-dimensional example is interesting.

Example 8.3. Let’s compute the angle polynomial of the 1-dimensional polytope P := [a, b],
with a, b ∈ R. We will use our knowledge of the 1-dimensional Fourier transform of an

161
interval, from Exercise 3.1, to compute:
2
X
AP (t) = (b − a)t + lim 1̂P (tξ) e−πξ (8.22)
→0
ξ∈Z−{0}
X  e−2πitξb − e−2πitξa  2
= (b − a)t + lim e−πξ (8.23)
→0 −2πiξ
ξ∈Z−{0}
X e−2πitbξ−πξ2 X e−2πitaξ−πξ2
= (b − a)t + lim − lim (8.24)
→0 −2πiξ →0 −2πiξ
ξ∈Z−{0} ξ∈Z−{0}

(8.25)

Throughout this example, all series converge absolutely (and quite rapidly) due to the exis-
2
tence of the Gaussian damping factor e−πξ . Let’s see what happens when we specialize the
vertices a or b - perhaps we can solve for these new limits?

case 1. a, b ∈ Z. This is the case of an integer polytope, which in this case is an interval
in R1 . Because we are restricting attention to integer dilates t, and since a, b, ξ ∈ Z, we have
e−2πitξb = e−2πitξa = 1. Therefore
X  e−2πitξb − e−2πitξa  2
AP (t) = (b − a)t + lim e−πξ = (b − a)t + 0.
→0 −2πiξ
ξ∈Z−{0}

We arrive at
AP (t) = (b − a)t,
so that the solid angle sum AP (1) is exactly the length of the interval we considered. We may
compare this discrete volume with the other discrete volume, namely the Ehrhart polynomial
of this interval: L[a,b] (t) = (b − a)t + 1.

case 2. a = 0, b ∈
/ Z. Here one of the two series in (8.24) is:
X e−2πitaξ−πξ2 X e−πξ2
= = 0,
−2πiξ −2πiξ
ξ∈Z−{0} ξ∈Z−{0}

because the summand is an odd function of ξ. But we already know by direct computation
that in this case A[0,b] (t) = 21 + bbtc, we can solve for the other limit:

1 X  e−2πitξb  2
+ bbtc = bt + lim e−πξ
2 →0 −2πiξ
ξ∈Z−{0}

So this simple example has given us a nice theoretical result. We record this rigorous proof
above as Lemma 8.3 below, after relabelling bt := x ∈ R. 

162
Lemma 8.3. For any x ∈ R, we have

1 X e−2πixξ−πξ2 1
lim = x − bxc − .
2πi →0 ξ 2
ξ∈Z−{0}

Theorem 8.1. Let P be an integer polygon. Then the angle polynomial of P is:

AP (t) = (areaP)t2 ,

for all positive integer dilations t.

It turns out that this result, for AP (1), is easily equivalent to the well-known Pick’s formula
for an integer polygon.
Theorem 8.2 (Pick’s formula, 1899). Let P be an integer polygon. Then
1
AreaP = I + B − 1,
2
where I is the number of interior integer points in P, and B is the number of boundary
integer points in P.

y y y

∪ =
x x x
0 0 0
P1 P2 P1 ∪ P2

Figure 8.3: Additive property of the angle polynomial

There is also a way to characterize the polytopes that k-tile Rd by translations, using solid
angle sums. Gravin, Robins, and Shiryaev [67, Theorem 6.1] gave the following characteri-
zation.
Theorem 8.3. A polytope P k-tiles Rd by integer translations if and only if
X
ωP +v (λ) = k,
λ∈Zd

for every v ∈ Rd .

163
8.4 The Gram relations for solid angles

How does our elementary school identity, giving us the sum of the angles of a triangle, extend
to higher dimensions? We describe the extension here, mainly due to Gram. First, for each
face F of a polytope P ⊂ Rd , we define the solid angle of F , as follows. Fix any x0 ∈ int F ,
and let
ωF := ωP (x0 ).
We notice that this definition is independent of x0 , as long as we restrict x0 to the relative
interior of F .
Example 8.4. If P is the d-dimensional cube [0, 1]d , then each of its facets F has ωF = 12 .
1
However, it is a fact that for the cube, a face of dimension k has a solid angle of 2d−k
(Exercise 8.9). In particular a vertex v of this cube, having dimension 0, has solid angle
ωv = 21d . 
Theorem 8.4 (Gram relations). Given any d-dimensional polytope P ⊂ Rd , we have
X
(−1)dim F ωF = 0.
F ⊂P

(For a proof of Lemma 8.4 see [17]).


Example 8.5. Let’s see what the Gram relations tell us in the case of a triangle ∆. For
each edge E of ∆, placing a small sphere at a point in the interior of E means half of it is
inside ∆ and half of it is outside of ∆, so that ωE = 12 . Next, each vertex of ∆ has a solid
angle equal to the usual (normalized) angle θ(v) at that vertex. Finally ∆ itself has a solid
angle of 1, because picking a point p in the interior of ∆, and placing a small sphere centered
at p, the whole sphere will be contained in ∆. Putting it all together, the Gram relations
read:
X
0= (−1)dim F ωF
F ⊂∆
 
0 1 1 1 1
= (−1) (θ(v1 ) + θ(v2 ) + θ(v3 )) + (−1) + + + (−1)2 · 1
2 2 2
1
= θ(v1 ) + θ(v2 ) + θ(v3 ) − ,
2
which looks familiar! We’ve retrieved our elementary-school knowledge, namely that the
three angles of a triangle sum to π radians. So the Gram relations really are an extension of
this fact. 

What about R3 ?

164
Example 8.6. Let’s see what hidden secrets lie behind the Gram relations for the standard
simplex ∆ ⊂ R3 . At the origin v0 = 0, the tangent cone is the positive orthant, so that
ω(v0 ) = 18 . The other 3 vertices all “look alike”, in the sense that their tangent cones are
all isometric, and hence have the same solid angle ωv . What about the edges? In general,
it’s a fact that the solid angle of an edge equals the dihedral angle between the planes of
its two bounding facets (Exercise 8.10). There are two types of edges here, as in the figure.
For an edge
D E which
 E lies on the boundary of the skew facet, we have the dihedral angle
1 0
1
cos φ = √3 1 , 1 = √13 , so that ωE = φ = cos−1 √13 . It’s straightforward that for the
1 0
other type of edge, each of those 3 edges has a solid angle of 14 . Putting it all together, we
see that
X
0= (−1)dim F ωF
F ⊂∆
   
1 1 1 1
= (−1) 0
+ 3ωv 1
+ (−1) 3 + 3 cos−1 √ + (−1)2 · 4 + (−1)3 · 1.
8 4 3 2

Solving for ωv , we get ωv = cos−1 √13 − 81 . So we were able to compute the solid angle of at
a vertex of ∆ in R3 , using the Gram relations, together with a bit of symmetry. 

Related to the topics above is the fact that the angle polynomial possesses the following
fascinating functional equation (For a proof of Theorem 8.5, and an extension of it, see [46]).
Theorem 8.5 (Functional equation for the angle polynomial). Given a d-dimensional
rational polytope P ⊂ Rd , we have

AP (−t) = AP (t),

for all t ∈ Z. 

Notes

(a) Let’s compare and contrast the two notions of discrete volumes that we have encoun-
tered so far. For a given rational polytope P, we notice that the Ehrhart quasi-
polynomial LP (t) is invariant when we map P to any of its unimodular images. That
is, any rational polytope in the whole orbit of the unimodular group SLd (Z)(P) has
the same discrete volume LP (t). This is false for the second discrete volume AP (t) - it
is not invariant under the modular group (Exercise 8.8). But AP (t) is invariant under
the large finite group of the isometries of Rd that preserve the integer lattice (known
as the hyperoctahedral group).

165
So we see that AP (t) is more sensitive to the particular embedding of P in space,
because it is dependent upon a metric. It is reasonable to expect that it can distinguish
between “more” rational polytopes, but such a question remains to be formalized.
The angle polynomial also has the advantage of being a much more symmetric poly-
nomial, with half as many coefficients that occur in the Ehrhart polynomial of integer
polytopes.
However, LP (t) has its advantages as well - to compute a local summand for AP (t) :=
P
n∈Zd ωtP (x) requires finding the volume
P of a local spherical polytope, while to com-
pute a local summand for LP (t) := n∈Zd 1 is quite easy: it is equal to 1.
But as we have seen, computing the full global sum for AP (t) turns out to have its
simplifications.

(b) The interesting undergraduate dissertation of Nhat Le Quang [118], from 2010, gives
a thorough analysis of solid angle sums in R2 , for rational polygons.

(c) The recent work of Gervásio [132] gives an online implementation for the calculation
of solid angles in any dimension, with open source code.

(d) In [49], there is an explicit description for some of the coefficients of the solid angle
polynomial AP (t) of a d-dimensional polytope, for all positive real dilations t > 0.
Indeed, the approach in [49] uses the Fourier analytic landscape.

Exercises
     
1 1 1
8.1. Let K = {λ1 0 + λ2 1 + λ3 1 | λ1 , λ2 , λ3 ≥ 0}, a simplicial cone. Show that the
0 0 1
1
solid angle of K is ωK = 48
.

8.2. We recall the 2-dimensional cross-polytope ♦ := {(x1 , x2 ) ∈ R2 | |x1 | + |x2 | ≤ 1} . Find,


from first principles, the angle quasi-polynomial for the rational polygon P := 31 ♦, for all
integer dilations of P.

8.3. We recall that the 3-dimensional cross-polytope was defined by

♦ := (x1 , x2 , x3 ) ∈ R3 | |x1 | + |x2 | + |x3 | ≤ 1 .




Compute the angle polynomial of A♦ (t).

166
8.4. We recall that the d-dimensional cross-polytope was defined by

♦ := (x1 , x2 , . . . , xd ) ∈ Rd | |x1 | + |x2 | + · · · + |xd | ≤ 1 .




Compute the angle polynomial of A♦ (t).

8.5. Let P be an integer zonotope. Prove that the angle polynomial of P is

AP (t) = (vol P)td ,

valid for all positive integers t.

8.6. Let P be a rational interval [ ac , db ]. Compute the angle quasi-polynomial AP (t) here.

8.7. Define the rational triangle ∆ whose vertices are (0, 0), (1, NN−1 ), (N, 0), where N ≥ 2 is
a fixed integer. Find the angle quasi-polynomial A∆ (t).

8.8. ♣ For each dimension d, find an example of an integer polytope P ⊂ Rd and a unimod-
ular matrix U ∈ SLd (Z), such that the angle quasi-polynomials AP (t) and AU (P) (t) are not
equal to each other for all t ∈ Z>0 .

8.9. ♣ For the cube  := [0, 1]d , show that any face F ⊂  that has dimension k has the
1
solid angle ωF = 2d−k .

8.10. ♣ Show that the solid angle ωE of an edge E (1-dimensional face) of a polytope equals
the dihedral angle between the hyperplanes defined by its two bounding facets. (Hint: use the
unit normal vectors for both facets)

8.11. Using the Gram relations, namely Theorem 8.4, compute the solid angle at any vertex
of the following regular tetrahedron:
n      o
1 0 0 1
T := conv 0 1 , 0 , 1 .
0 0 1 1

167
Chapter 9

Counting integer points in polytopes -


the Ehrhart theory

How wonderful that we have met with a paradox. Now we have some hope of
making progress.
– Niels Bohr

9.1 Intuition

A basic question in discrete geometry


is “how do we discretize volume?”

One method of discretizing the vol-


ume of P is to count the number of
integer points in P. Even in R2 , this
question may be highly non-trivial,
depending on the arithmetic proper-
ties of the vertices of P. Ehrhart
first considered integer dilations of
a fixed, integer polytope P, and de-
fined:

LP (t) := |Zd ∩ tP|, (9.1)

where tP is the t’th dilate of P, and


t is a positive integer. Ehrhart showed that LP (t) is a polynomial in the positive integer
parameter t, known as the Ehrhart polynomial of P.

168
From the lens of Fourier analysis, Ehrhart polynomials may be computed by ‘averaging’ the
Fourier transform of a polytope over the full integer lattice:
X X
LP (t) := |Zd ∩ tP| = 1tP (n) = 1̂tP (ξ), (9.2)
n∈Zd ξ∈Zd

where the latter identity uses Poisson summation, but because we may not use indicator
functions directly in Poisson summation, some care is required and the process of smoothing
may be applied to 1P .

We can also compare this combinatorial way to discretize volume, namely equation (9.1),
and the discrete volumes of the previous chapter which used solid angles.

More generally, given a function f : Rd → C, we may sum the values of f at all integer
points and observe how close this sum gets to the integral of f over P. This approach is
known as Euler-Maclaurin summation over polytopes, and is a current and exciting topic of
a growing literature (see Note (g) below).

9.2 Computing integer points in polytopes via the dis-


crete Brion Theorem

Here we present a proof of Ehrhart’s result, using Brion’s Theorem 7.5, the discrete form,
which we now recall. When all the vertices of a polytope P have rational coordinates, we
call P a rational polytope .

Let P ⊂ Rd be a rational, d-dimensional polytope, and let N be the number of its vertices.
For each vertex v of P, we consider the vertex tangent cone Kv of P. Once we dilate P by
t, each vertex v of P gets dilated to become tv, and so each of the vertex tangent cone Kv
of P simply get shifted to the corresponding vertex tangent cone Ktv of tP. Thus, we have
X X X
e2πihn,zi = e2πihn,zi + · · · + e2πihn,zi , (9.3)
n∈tP∩Zd n∈Ktv1 ∩Zd n∈KtvN ∩Zd

for all z ∈ Cd − S, where S is the hyperplane arrangement defined by the (removable)


singularities of all of the transforms 1̂Kvj (z). To simplify notation, we may absorb the
1
constant 2πi into the complex vector z by replacing z by 2πi z, so that we may assume
without loss of generality that Brion’s Theorem 7.5 has the form

X X X
ehn,zi = ehn,zi + · · · + ehn,zi . (9.4)
n∈tP∩Zd n∈Ktv1 ∩Zd n∈KtvN ∩Zd

169
We recall that the discrete Brion Theorem tells us that:

σP (z) = σKv1 (z) + · · · + σKvN (z). (9.5)

And now we notice that when z = 0, the left-hand-side gives us precisely


X
1 := |Zd ∩ tP|,
n∈tP∩Zd

which is good news - it is the Ehrhart polynomial LP (t), by definition. The bad news is that
z = 0 is a singularity of the right-hand-side of (9.3). But then again, there is still more good
news - we already saw in the previous chapter that it is a removable singularity. So we may
let z → 0 to see what happens.

Example 9.1. We compute the integer point transform of the standard triangle in the
plane, using Brion’s Theoem 9.4. Namely, for the standard triangle

∆ := conv(( 00 ) , ( 10 ) , ( 01 )),

as depicted in Figure 9.1, we find σ∆ (z). By definition, the integer point transform of its

Figure 9.1: The standard triangle, with its vertex tangent cones

170
vertex tangent cone Kv1 is
1 0
ehn1 ( 0 )+n2 ( 1 ),zi
X X
σKv1 (z) := ehn,zi =
n∈Kv1 ∩Zd n1 ≥0,n2 ≥0
X X
= en1 z1 en2 z2
n1 ≥0 n2 ≥0
1
= .
(1 − ez1 )(1 − ez2 )

For the vertex tangent cone Kv2 , we have


1 −1 +n −1 ,zi
eh( 0 )+n1 ( 0 ) 2( 1 )
X X
σKv2 (z) := ehn,zi =
n∈Kv2 ∩Zd n1 ≥0,n2 ≥0
X X
= e z1 en1 (−z1 ) en2 (−z1 +z2 )
n1 ≥0 n2 ≥0
z1
e
= .
(1 − e−z1 )(1 − e−z1 +z2 )

Finally, for the vertex tangent cone Kv3 , we have


0 0 1
eh( 1 )+n1 ( −1 )+n2 ( −1 ),zi
X
σKv3 (z) :=
n1 ≥0,n2 ≥0
X X
= ez2 en1 (−z2 ) en2 (z1 −z2 )
n1 ≥0 n2 ≥0
e z2
= .
(1 − e−z2 )(1 − ez1 −z2 )

Altogether, using 9.4 we have

σP (z) = σKv1 (z) + σKv2 (z) + σKv3 (z) (9.6)


1 ez1 ez2
= + + . (9.7)
(1 − ez1 )(1 − ez2 ) (1 − e−z1 )(1 − e−z1 +z2 ) (1 − e−z2 )(1 − ez1 −z2 )

Example 9.2. We find a formula for the Ehrhart polynomial LP (t) := |Z2 ∩ tP| of the
standard triangle, continuing the computation of the previous example. It turns out, as we
show in the section that follows, that the method we use here is universal - it can always be
used to find the Ehrhart polynomial of any rational polytope.

171
In this example we are lucky in that we may use brute-force to compute it, since the number
of integer points in the t-dilate of P may be computed along the diagonals:
(t + 1)(t + 2) 1 3
LP (t) = 1 + 2 + 3 + · · · + (t + 1) = = t2 + t + 1.
2 2 2
Now we can confirm this lucky answer with our brand new machine, as follows. Using (9.4),
and the formulation (9.7) from the previous example, we have the integer point transform
for the dilates of P:
X X X X
ehn,zi = ehn,zi + ehn,zi + ehn,zi (9.8)
n∈t∆∩Zd n∈Ktv1 ∩Zd n∈Ktv1 ∩Zd n∈Ktv3 ∩Zd

1 etz1 etz2
= + + (9.9)
(1 − ez1 )(1 − ez2 ) (e−z1 − 1)(e−z1 +z2 − 1) (e−z2 − 1)(ez1 −z2 − 1)
:= F1 (z) + F2 (z) + F3 (z), (9.10)
where we have defined F1 , F2 , F3 by the last equality. We can let z → 0 along almost any
direction, but it turns out that we can simplify our computations by taking advantage of the
x
symmetry of this polytope, so we will pick z = ( −x ), which will simplify our computations
(see Note (f)). Here is our plan:

x
(a) We pick z := ( −x ).
(b) We expand all three meromorphic functions F1 , F2 , F3 in terms of their Laurent series
in x, giving us Bernoulli numbers.
(c) Finally, we let x → 0, to retrieve the constant term (which will be a polynomial function
of t) of the resulting Laurent series.

To expand F1 (z), F2 (z), F3 (z) in their Laurent series, we recall the definition 3.26 of the
t
= ∞ tk
P
Bernoulli numbers in terms of their generating function, namely et −1 k=0 Bk k! :

−1 X xm X (−x)n
F1 (x, −x) = Bm Bn
x2 m≥0 m! n≥0 n!
x x2 x x2
  
−1 3 3
= 2 1− + + O(x ) 1+ + + O(x )
x 2 12 2 12
−1 1
= 2 − + O(x)
x 3

Similarly, we have
t2 2
x x2+ O(x3 ) (2x) (2x)2
  
1 + tx + 2!
x 3 3
F2 (x, −x) = 1+ + + O(x ) 1+ + + O(x )
2x2 2 12 2 12
1 3 2 t 3t t2
= 2+ + + + + + O(x)
2x 4x 3 2x 4 4

172
Now, by symmetry we see that F3 (x, −x) = F2 (−x, x), so that by (9.10) and the latter
expansions, we finally have:
x 3 1
ehn,( −x )i = F1 (x, −x) + F2 (x, −x) + F2 (−x, x) = 1 + t + t2 + O(x).
X

d
2 2
n∈t∆∩Z

x
Letting z := ( −x ) → 0 in the latter computation, we retrieve the (Ehrhart) polynomial
answer: X 3 1
1 = L∆ (t) = 1 + t + t2 ,
d
2 2
n∈t∆∩Z

as desired. 

9.3 Examples, examples, examples....

Example 9.3. We work out the integer point transform σK (z) of the cone
 
K := {λ1 31 + λ2 12 | λ1 , λ2 ∈ R≥0 },

Drawn in the figures below. We note that here det K = 5, and that there are indeed 5 integer
points in D, its half-open fundamental parallelepiped.

We may ‘divide and conquer’ the integer point transform σK (z), by breaking it up into 5
infinite series, one for each integer point in D, as follows:
X X X X X X
σK (z) := ehn,zi := + + + + ,
n∈K∩Zd 0 1 2 2
(0) (1) (1) (2) (2) 3

where
  
X X 1 +n1 3 +n2 1 ,zi
:= eh 1 1 2

( 11 ) n1 ≥0,n2 ≥0
  
1 X 3 1
= eh 1 ,zi ehn1 1 +n2 2 ,zi
n1 ≥0,n2 ≥0
  
1 X 3 X 1
= eh 1 ,zi en1 h 1 ,zi en2 h 2 ,zi
n1 ≥0 n2 ≥0
z1 +z2
e
= ,
(1 − e3z1 +z2 )(1 − ez1 +2z2 )

173
Figure 9.2: The 5 integer points in a fundamental parallelepiped D, of K.

and similarly we have


X e2z1 +z2
= ,
(1 − e3z1 +z2 )(1 − ez1 +2z2 )
( 21 )
X e2z1 +2z2
= ,
(1 − e3z1 +z2 )(1 − ez1 +2z2 )
( 22 )
X e3z1 +2z2
= ,
(1 − e3z1 +z2 )(1 − ez1 +2z2 )
( 32 )
and finally
X 1
= .
(1 − e3z1 +z2 )(1 − ez1 +2z2 )
( 00 )
To summarize, we have the following expression:
X 1 + ez1 +z2 + e2z1 +z2 ) + e2z1 +2z2 + e3z1 +2z2
ehn,zi = .
(1 − e3z1 +z2 )(1 − ez1 +2z2 )
n∈K∩Zd

Equivalently, we may use multinomial notation: let qj := ezj , so that by definition

q1 n1 · · · qd nd := ez1 n1 · · · ezd nd = ehz,ni .

174
1

Figure 9.3: The point 1
in D, with its images in K under translations by the edge vectors
of K.

It is common to use the following shorthand multinomial definition: q n := q1 n1 · · · qd nd . With


this multinomial notation, we have
X 1 + q1 q2 + q1 2 q2 + q1 2 q2 2 + q13 q22
qn = .
(1 − q13 q2 )(1 − q1 q22 )
n∈K∩Zd

Example 9.4. Here we will compute the integer point transform of the triangle ∆ defined by
the convex hull of the points ( 00 ) , ( 31 ) , ( 36 ), as shown in Figure 9.4 below. We first compute
the integer point transforms of all of its tangent cones. For the vertex v1 , we already computed
the integer point transform of its tangent cone in the previous example.

For the vertex


 v2 , we notice that its vertex tangent cone is a unimodular cone, because
0 −1
| det −1 −2 | = 1. Its integer point transform is:
0 +n −1 ,zi
 
h( 3 )+n1 ( −1 )
X X
hn, zi 2 −2
σKv2 (z) := e = e 6

n∈Kv2 ∩Zd n1 ≥0,n2 ≥0


X
= e3z1 +6z2 ) en1 (−z2 ) en2 (−z1 −2z2 )
n1 ≥0,n2 ≥0
3z1 +6z2
e
= .
(1 − e−z2 )(1 − e−z1 −2z2 )

175
Figure 9.4: A triangle with vertices v1 , v2 , v3 , and its vertex tangent cones

Equivalently, using the notation from Example 9.3 above,


X
n q13 q26
σKv2 (z) := q =
d
(1 − q2−1 )(1 − q1−1 q2−2 ).
n∈Kv2 ∩Z

For vertex v3 , the computation is similar to vertex tangent cone Kv1 , and we have:

+n1 −3
 
h( 3 ) +n2 ( 0
1 ), zi
X X
hn, zi −1
σKv3 (z) := e = e 1

n∈Kv3 ∩Zd n1 ≥0,n2 ≥0


X
= e3z1 +z2 e(−3z1 −z2 )n1 e2πi(z2 )n2
n1 ≥0,n2 ≥0

1 + e + e−2z1
−z1
= e3z1 +z2
(1 − e3z1 +z2 )(1 − ez2 )
e3z1 +z2 + e2z1 +z2 + ez1 +z2
=
(1 − e3z1 +z2 )(1 − ez2 )
q 3 q 2 + q 2 q 2 + q1 q 2
= 1 −3 1−1 .
(1 − q1 q2 )(1 − q2 )

Finally, putting all of the three vertex tangent cone contributions together, using (9.3), we

176
get:

σ∆ (z) = σKv1 (z) + σKv2 (z) + σKv3 (z)


1 + q1 q2 + q1 2 q2 + q1 2 q2 2 + q13 q22 q13 q26 q13 q2 + q12 q2 + q1 q2
= + + .
(1 − q13 q2 )(1 − q1 q22 ) (1 − q2−1 )(1 − q1−1 q2−2 ) (1 − q1−3 q2−1 )(1 − q2 )


As these examples suggest, there is a thread that they all share, namely that their numerators
are polynomials that encode the integer points inside a fundamental parallelepiped Π which
sits at the vertex of each vertex tangent cone. The proof is fairly easy - we only need to put
several geometric series together. Let’s formalize this.

First, given any d-dimensional simplicial cone K ⊂ Rd , with edge vectors ω1 , . . . , ωd , we


define the fundamental parallelepiped of K by:

Π := {λ1 ω1 + · · · + λd ωd | all 0 ≤ λj < 1}, (9.11)

a half-open parallelepiped. In the same way that we’ve encoded integer points in polytopes
using σP (z), we can encode the integer points in Π by defining
X
σΠ (z) := ehz,ni .
n∈Zd ∩Π

For rational simplicial cones Kv , it turns out that their integer point transforms, defined in
(7.43), have a pretty structure theorem (for a proof see [17]), and are rational functions of
the variables ez1 , . . . , ezd , as follows.
Theorem 9.1. Given a d-dimensional simplicial cone Kv ⊂ Rd , with apex v ∈ Rd , and with
d linearly independent integer edge vectors ω1 , . . . , ωd ∈ Zd , we have:
X σΠ+v (z)
ehn,zi = Qd . (9.12)
n∈Kv ∩Zd k=1 (1 − ehωk ,z i)

9.4 The Ehrhart polynomial of an integer polytope

Eugene Ehrhart initiated a systematic study of the integer point enumerator



LP (t) := tP ∩ Zd ,

for an integer polytope P, which Ehrhart proved was always a polynomial function of the pos-
itive integer dilation parameter t. Ehrhart also proved that for a rational polytope P ⊂ Rd ,

177
the integer point enumerator LP (t) is a quasi-polynomial in the positive integer parameter
t, which means by definition that
LP (t) = cd td + cd−1 (t)td−1 + · · · + c1 (t)t + c0 (t), (9.13)
where each cj (t) is a periodic function of t ∈ Z>0 .

The study of Ehrhart polynomials and Ehrhart quasi-


polynomials has enjoyed a renaissance in recent years
([12], [17]), and has some suprising connections to
many branches of science, and even to voting theory,
for example.

Theorem 9.2 (Ehrhart). For an integer polytope P ⊂


Rd , its discrete volume LP (t) is a polynomial functions
of t, for all positive integer values of the dilation pa-
rameter t. Moreover, we have
LP (t) = (vol P)td + cd−1 td−1 + · · · + c1 t + 1. (9.14)

As a first application, we show that the discrete vol-


ume of a (half-open) parallelepiped has a particularly
elegant and useful form.
Figure 9.5: Eugene Ehrhart
Lemma 9.1. Let D be any half-open integer paral-
lelepiped in Rd , defined by
D := {λ1 w1 + · · · + λd wd | 0 ≤ λ1 , . . . , λd < 1} ,
where w1 , · · · wd ∈ Zd are linearly independent. Then:
#{Zd ∩ D} = vol D,
and for each positive integer t, we also have
#{Zd ∩ tD} = (vol D) td .

Proof. We can tile tD by using td translates of D, because D is half-open. Therefore


#{Zd ∩ tD} = #{Zd ∩ D}td ,
and by definition #{Zd ∩ tD} = LD (t). On the other hand, we also know by Ehrhart’s
Theorem 9.2 that LD (t) is a polynomial for integer values of t, whose leading coefficient is
vol D. Since LD (t) = #{Zd ∩ D}td for all positive integer values of t, we conclude that
#{Zd ∩ D} = vol D.

178
Theorem 9.3 (Ehrhart). For a rational polytope P ⊂ Rd , its discrete volume LP (t) is a
quasi-polynomial function of t, for all positive integer values of the dilation parameter t. In
particular, we have
LP (t) = (vol P)td + cd−1 (t)td−1 + · · · + c1 (t)t + c0 (t), (9.15)
where each quasi-coefficient ck (t) is a periodic function of t ∈ Z>0 .

9.5 Unimodular polytopes

A d-dimensional integer simplex ∆ is called a unimodular simplex if ∆ is the modular


image of the standard simplex ∆standard , the convex hull of the points {0, e1 , . . . , ed } ⊂ Rd ,
where ek := (0, . . . , 0, 1, 0, . . . , 0) is the standard unit vector pointing in the direction of the
positive axis xk .

Figure 9.6: A unimodular polygon - each vertex tangent cone is a unimodular cone. It
is clear from the construction in the Figure that we can form arbitrarily large unimodular
polygons.
       
0 1 1 1
Example 9.5. Let ∆ := conv 0 , 0 , 1 , 1 , their convex hull. Then ∆ is a
0 0 0 1  
1 1 1
unimodular simplex, because the unimodular matrix 0 1 1 maps the standard simplex
0 0 1
∆standard to ∆. 

179
Figure 9.7: A unimodular cone at v, appearing as one of the vertex tangent cones in Figure
9.6. We notice that its half-open fundamental parallelepiped, with vertex at v, does not
contain any integer points other than v.

It is not difficult to show that the tangent cone of a unimodular simplex possesses edge
vectors that form a lattice basis for Zd . Thus, it is natural to define a unimodular cone
K ⊂ Rd as a simplicial cone, possessing the additional property that its d edge vectors form
a lattice basis for Zd .
Example 9.6. We consider the polygon P in Figure 9.6. An easy verification shows that
each of its vertex tangent cones is unimodular. For example, focusing on the vertex v, we see
from Figure 9.7, that its vertex tangent cone is Kv := v + {λ1 ( −21 ) + λ2 ( −11 ) | λ1 , λ2 ≥ 0}.
Kv is a unimodular cone, because the matrix formed by the its two edges ( −21 ) and ( −11 ) is
a unimodular matrix. 

More generally, a simple, integer polytope is called a unimodular polytope if each of


its vertex tangent cones is a unimodular cone. Unimodular polytopes are the first testing
ground for many conjectures in discrete geometry and number theoery. Indeed, we will see
later that the number of integer points in a unimodular polytope, namely |Zd ∩ P|, admits a
simple and computable formula, if we are given the local tangent cone information at each
vertex. By contrast, it is in general thought to be quite difficult to compute the number of
integer points |Zd ∩ P|, even for (general) simple polytopes, a problem that belongs to the
NP-hard class of problems (if the dimension d is not fixed).
Lemma 9.2. Suppose we have two integer polytopes P, Q ⊂ Rd , which are unimodular
images of each other:
P = U Q,
for some unimodular matrix U . Then LP (t) = LQ (t), for all t ∈ Z≥0 .

180
Lemma 9.3. Suppose that ∆ ⊂ Rd is a d-dimensional integer simplex. Then ∆ is a uni-
modular simplex ⇐⇒ (d − 1)∆ does not contain any interior integer points.

9.6 Rational polytopes and quasi-polynomials

The following properties for the floor function, the ceiling function, and the fractional part
function are often useful. It’s convenient to include the following indicator function, for the
full set of integers, as well: (
1 if x ∈ Z
1Z (x) := ,
0 if x ∈/Z
the indicator function for Z. For all x ∈ R, we have:

(a) dxe = − b−xc

(b) 1Z (x) = bxc − dxe + 1

(c) {x} + {−x} = 1 − 1Z (x)

(d) dxe = x + 1 − {x} − 1Z (x)


 n−1  n
(e) Let m ∈ Z>0 , n ∈ Z. Then m
+1= m
.

(Exercise 9.14)

Example 9.7. Let’s find the integer point enumerator LP (t) := |Z ∩ tP| of the rational line
segment P := [ 31 , 1]. Proceeding by brute-force, for t ∈ Z>0 we have
   
t t
LP (t) = , t ∩ Z = btc − +1 (9.16)
3 3
 
t
=t+ − +1 (9.17)
3
 
t t
=t+− − − +1 (9.18)
3 3
 
2 t
= t− − + 1, (9.19)
3 3

a periodic function on Z with period 3. Here we used property (a) in the third equality. In
fact, here we may let t be any positive real number, and we still obtain the same answer, in
this 1-dimensional case.

181
Now we will compare this to a new computation, but this time from the perspective of the
t
vertex tangent cones. For the  t Ktv1 := [ 3 , +∞), we can parametrize the integer points
 t cone
in this cone by Ktv1 ∩ Z = { 3 , 3 + 1, . . . }, so that
1
σKtv1 (z) = ed 3 ez enz = ed 3 ez
t X t
.
n≥0
1 − ez

For the cone Ktv2 := (−∞, t], we can parametrize the integer points in this cone by Ktv2 ∩Z =
{t, t − 1, . . . }, so that
X 1
σKtv2 (z) = et·z enz = etz .
n≤0
1 − e−z

So by the discrete Brion Theorem (which is here essentially a finite geometric sum), we get:
1 1
enz = ed 3 ez
X t
+ etz
1−ez 1 − e−z
n∈[ 3t ,t]
   2 2 ! 
t t z 1 1 1
=− 1+ z+ + ··· − + z + ···
3 3 2! z 2 12
2
  
2z 1 1 1
+ 1 + (t + 1)z + (t + 1) + ··· − + z + ···
2! z 2 12
   
1 t 1 t
= − + (t + 1) − + O(z) −→ t − + 1,
2 3 2 3
as z → 0, recovering the same answer 9.16 above. 

Example 9.8. Let’s find the integer point enumerator LP (t) := |Z2 ∩ tP| of the rational
triangle   1   
0
P := conv ( 00 ) , 02 , 21 .
First we will proceed by brute-force (which does not always work well), and then we will use
the machinery of (9.4).

For the brute-force method, we need to consider separately the even integer dilates and the
odd integer dilates. Letting t = 2n be a positive even integer, it’s clear geometrically that

LP (t) := |Z2 ∩ 2nP| = 1 + 2 + · · · + n


t t
n(n + 1) ( + 1)
= = 2 2
2 2
1 2 1
= t + t.
8 4

182
On the other hand, if t = 2n − 1, then we notice that we never have an integer point on the
diagonal face of P, so that in this case we get:
t+1 t+1
2 ( 2 + 1) 1 1 3
LP (t) := |Z ∩ (2n − 1)P| = 1 + 2 + · · · + n = 2
= t2 + t + .
2 8 2 8

Alternatively, we may also rederive the same answer by using the Brion identity (9.4). We
can proceed as in Example 9.2. The only difference now is that the vertex tangent cones
have rational apices. So although we may still use the same edge vectors to parametrize   the
d 0
integer points in Ktv3 ∩ Z , we now have a new problem: the rational vertex v3 = 12 . But
 
d 0
in any case, we get: {n ∈ Ktv3 ∩ Z } = { 2t + n1 ( −10 ) + n2 ( −11 ) | n1 , n2 ∈ Z≥0 }. We invite
the reader to complete this alternate derivation of the Ehrhart quasi-polynomial LP (t) in
this case. 

Figure 9.8: A rational triangle, which happens to be a rational dilate of the standard
simplex.

9.7 Ehrhart reciprocity

There is a wonderful, and somewhat mysterious, relation between the Ehrhart polynomial of
the (closed) polytope P, and the Ehrhart polynomial of its interior, called int P. We recall
our convention that all polytopes are, by definition, closed polytopes. We first compute
LP (t), for positive integers t, and once we have this polynomial in t, we formally replace t

183
by −t. So by definition, we form LP (−t) algebraically, and then embark on a search for its
new combinatorial meaning.

Theorem 9.4 (Ehrhart reciprocity). Given a d-dimensional rational polytope P ⊂ Rd , let


Lint P (t) := |Zd ∩ int P|, the integer point enumerator of its interior. Then

LP (−t) = (−1)d Lint P (t), (9.20)

for all t ∈ Z.

Offhand, it seems like ‘a kind of magic’, and indeed Ehrhart reciprocity is one of the most
elegant geometric inclusion-exclusion principles we have. Some examples are in order.

Example 9.9. ForP the unit cube  := [0, 1]d , we can easily compute from first principles
d
L (t) = (t + 1)d = k=0 kd tk . For the open cube int , we can also easily compute


d  
d
X d
Lint  (t) = (t − 1) = tk (−1)d−k
k=0
k
d  
d
X d
= (−1) (−t)k
k=0
k
d
= (−1) L (−t),

using our known polynomial L (t) = (t + 1)d . 

Example 9.10. For the standard simplex ∆, we consider its t-dilate, given by
d
X
d
t∆ := {(x1 , . . . , xd ) ∈ R | xi ≤ t, and all xk ≥ 0}.
k=1

We can easily compute its Ehrhart polynomial, by using combinatorics. We need to find the
number of nonnegative integer solutions to

x1 + · · · + xd ≤ t,

which is equal to L∆ (t), for a fixed positive integer t. We can introduce a ‘slack variable’ z,
to transform the latter inequality to an equality: x1 + · · · + xd + z = t, where 0 ≤ z ≤ t. By
a very classical and pretty argument, (involving placing t balls into urns that are separated
t+d

by d walls) this number is equal to d (Exercise 9.15). So we found that
 
t+d (t + d)(t + d − 1) · · · (t + 1)
L∆ = = , (9.21)
d d!

184
a degree d polynomial, valid for all positive integers t.

What about the interior of ∆? Here we need to find the number of positive integer solutions
to x1 + · · · + xd < t, for each positive integer t. It turns out that by a very similar argument
as above (Exercise 9.16), the number of positive integer solutions is t−1 d
= Lint ∆ (t). So is
it really true that    
d d−t t−1
(−1) = ?
d d
Let’s compute, substituting −t for t in (9.21) to get:
 
−t + d (−t + d)(−t + d − 1) · · · (−t + 1)
L∆ (−t) = =
d d!
(t − d)(t − d + 1) · · · (t − 1)
= (−1)d
  d!
t − 1
= (−1)d = (−1)d Lint ∆ (t),
d
confirming that Ehrhart reciprocity works here as well. 

9.8 The Möbius inversion formula for the face poset

Given a polytope P ⊂ Rd , the collection of all faces F of P - including the empty set and P
itself - is ordered by inclusion. This ordering forms a partially ordered set, and is called the
face poset. There is a particularly useful inversion formula on this face poset.

Theorem 9.5 (Möbius inversion formula for the face poset). Given any function g : P → C,
we may define a sum over the face poset of P:
X
h(P) := g(F ). (9.22)
F ⊆P

We then have the following inversion formula:


X
g(P) := (−1)dim F h(F ). (9.23)
F ⊆P

To prove (again) that for positive integer values of t, the angle polynomial AP (t) is indeed a
polynomial in t, we may use the following useful little relation between solid angle sums and

185
integer point sums. We recall that for any polytope F, the integer point enumerator for the
relative interior of F was defined by Lint F (t) := |Zd ∩ int F|.

For each face F ⊆ P, we define the d-dimensional solid angle of the face F by picking
any point x inside the relative interior of F and denoting

ωP (F) := ωP (x).

Theorem 9.6. Let P be a d-dimensional polytope in Rd . Then we have


X
AP (t) = ωP (F)Lint F (t). (9.24)
F ⊆P

Proof. The polytope P is the disjoint union of its relatively open faces F ⊆ P, and similarly
the dilated polytope tP is the disjoint union of its relatively open faces tF ⊆ tP. We
therefore have: X X X
AP (t) = ωtP (n) = ωtP (n)1int(tF ) (n).
n∈Zd F ⊆P n∈Zd

But by definition each ωtP (n) is constant on the relatively open face int(tF) of tP, and we
denoted this constant by ωP (F). Altogether, we have:
X X X
AP (t) = ωP (F) 1int(tF ) (n) := ωP (F)Lint F (t).
F ⊆P n∈Zd F ⊆P

Theorem 9.7. Given an integer polytope P ⊂ Rd , the discrete volume AP (t) is a polynomial
in t, for integer values of the dilation parameter t.

Proof. By Ehrhart’s Theorem 9.2, we know that for each face F ⊆ P, Lint F (t) is a
polynomial function of t, for positive integers t. By Theorem 9.6, we see that AP (t) is
a finite linear combination of polynomials, with constant coefficients, and is therefore a
polynomial in t.

We may apply Theorem 9.5 to invert the relationship in Theorem 9.6 between solid angle
sums and local Ehrhart polynomials, to get the following consequence of the Möbius inversion
formula.

Corollary 9.1. Let P ⊂ Rd be a d-dimensional polytope. Then we have


X
Lint P (t) = (−1)dim F AF (t). (9.25)
F ⊆P

186
Proof. We begin with the identity of Theorem 9.6:
X
AP (t) = ωP (F) Lint F (t), (9.26)
F ⊆P

and we use the Möbius inversion formula (9.23) to get:


X
ωP (P)Lint P (t) = (−1)dim F AF (t). (9.27)
F ⊆P

But ωP (P) = 1, by definition, and so we are done.

Figure 9.9: An integer triangle, for which we compute AF (1) for each face F ⊂ P in Example
9.11, and use Möbius inversion to find LintP (1)

Example 9.11. Let’s work out a special case of Corollary 9.1, in R2 , for the triangle P
appearing in Figure 9.9, with t = 1. P has vertices v1 := ( −13 ) , v2 := ( −12 ) , v3 := ( 41 ), and
edges E1 , E2 , E3 .

We have to compute AF (1) for each face F ⊂ P. At the vertices, we have Av1 (1) = θ1 ,
Av2 (1) = θ2 , and Av3 (1) = θ3 . For the edges of P, we have:

AE1 (1) = θv2 + 21 + θv3 ,

AE2 (1) = θv3 + θv1 ,


AE3 (1) = θv1 + θv2 .
Finally, for P itself, we have

AP (1) = 6 + 21 + θv1 + θv2 + θv3 = 7.

187
Putting everything together, we have:
X    
dim F
(−1) AF (1) = Av1 (1) + Av2 (1) + Av3 (1) − AE1 (1) + AE2 (1) + AE3 (1) + AP (1)
F ⊆P
   
= θv1 + θv2 + θv3 − θv2 + 12 + θv3 + θv3 + θv1 + θv1 + θv2 + 7
1 3
= − + 7 = 6 = Lint P (1),
2 2
the number of interior integer points in P. 

Finally, we mention a fascinating open problem by Ehrhart.

Question 11 (Ehrhart, 1964). Let B ⊂ Rd be a d-dimensional convex body with the origin
as its barycenter. If the origin is the only interior integer point in B, then

(d + 1)d
vol B ≤ ,
d!
and futhermore the equality holds if and only if B is unimodularly equivalent to (d + 1)∆,
where ∆ is the d-dimensional standard simplex.

Ehrhart proved the upper bound for all d-dimensional simplices, and also for all convex
bodies in dimension 2. But Question 11 remains open in general (see [109] for more details).

Notes

(a) Ehrhart theory has a fascinating history, commencing with the fundamental work of
Ehrhart [51], [52], [53], in the 1960’s. Danilov [41] made a strong contribution to
the field, but after that the field of Ehrhart theory lay more-or-less dormant, until it
was rekindled by Jamie Pommersheim in 1993 [115], giving it strong connections to
Toric varieties. Using the Todd operators to discretize certain volume deformations of
polytopes, Khovanskii and Pukhlikov discovered a wonderful result that helped develop
the theory further (see Theorem 12.6 of [17]). In 1993, Alexander Barvinok [11] gave
the first polynomial-time algorithm for counting integer points in polytopes in fixed
dimension.
In recent years, Ehrhart theory has enjoyed an enthusiastic renaissance (for example,
the books [12], [17], [60]). For more relations with combinatorics, the reader may enjoy
reading Chapter 4 of the classic book “Enumerative Combinatorics”, [150] by Richard
Stanley.

188
(b) Regarding the computational complexity of counting integer points in polytopes, Alexan-
der Barvinok settled the problem in [11] by showing that for a fixed dimension d, there
is a polynomial-time algorithm, as a function of the ‘bit capacity’ of any given rational
polytope P ⊂ Rd , for counting the number of integer points in P.
(c) It is also true that for integer polytopes which are not necessarily convex (for example
simplicial complexes), the integer point enumerator makes sense as well. In this more
general context, the constant term of the corresponding integer point enumerator equals
the (reduced) Euler characteristic of the simplicial complex.
(d) For more information about the rapidly expanding field of Euler-MacLaurin summation
over polytopes, a brief (and by no means complete) list of paper in this direction consists
of the work by Berligne and Vergne [15], Baldoni, Berline, and Vergne [5], Garoufalidis
and Pommersheim [61], Brandolini, Colzani, Travaglini, and Robins [27], Karshon,
Sternberg, and Weitsman ([79], [80]), and very recently Fischer and Pommersheim
[56].
(e) There are some fascinating relations between an integer polytope P and its dual poly-
tope P ∗ . In particular, let P ⊂ R2 be an integer polygon (convex) whose only interior
integer point is the origin. Such polygons are called reflexive polygons, and up to uni-
modular transformations there are only a finite number of them in each dimension. If
we let B(P) be the number of integer points on the boundary of P, then Bjorn Poonen
and Fernando Villegas proved [116] that
B(P) + B(P ∗ ) = 12.
One way to see why we get the number “12” is to consider Bernoulli numbers and
Dedekind sums, but in [116] the authors give 4 different proofs, including Toric varieties
and modular forms.
(f) The trick used in Example 9.2 of picking the particular vector z := (x, −x), which
turns out to simplify the computations a lot, is due to Michel Faleiros.
(g) In a future version of this book, we will also delve into Dedekind sums, which arise very
naturally when considering the Fourier series of certain rational-exponential functions.
To define a general version of these sums, let L be a d-dimensional lattice in Rd ,
let w1 , . . . , wd be linearly independent vectors from L∗ , and let W be a matrix with
the wj ’s as columns. For any d-tuple e = (e1 , . . . , ed ) of positive integers ej , define
|e| := kj=1 ej . Then, for all x ∈ Rd , a lattice Dedekind sum is defined by
P

1 X e−2πihx,ξi 2
LL (W, e; x) := lim Qk e−πkξk . (9.28)
→0 (2πi)|e| e
j=1 hwj , ξi
j
ξ∈L
hwj ,ξi6=0,∀j

Gunnells and Sczech [70] have an interesting reduction theorem for these sums, giving
a polynomial-time complexity algorithm for them, for fixed dimension d.

189
Exercises

9.1. Consider the 1-dimensional polytope P := [a, b], for any a, b ∈ Z.

(a) Show that the Ehrhart polynomial of P is LP (t) = (b − a)t + 1.


(b) Find the Ehrhart quasi-polynomial LP (t) for the rational segment Q := [ 31 , 12 ].

9.2. We recall that the d-dimensional cross-polytope was defined by


♦ := (x1 , x2 , . . . , xd ) ∈ Rd | |x1 | + |x2 | + · · · + |xd | ≤ 1 .


(a) For d = 2, find the Ehrhart polynomial L♦ (t).


(b) For d = 3, find the Ehrhart polynomial L♦ (t).

9.3. Using the same notation for the d-dimensional cross-polytope ♦ as above, show that its
Ehrhart polynomial is
d   
X d t−k+d
L♦ (t) = ,
k=0
k d
for all t ∈ Z>0 .

9.4. Let d = 2, and consider the cross-polytope ♦ ⊂ R2 . Find the Ehrhart quasi-polynomial
LP (t) for the rational polygon P := 21 ♦.

9.5. Suppose ∆ is the standard simplex in Rd . Show that the first d dilations of ∆ do not
contain any integer points in their interior:
t(int ∆) ∩ Zd = φ,
for t = 1, 2, . . . , d. In other words, show that Lint P (1) = Lint P (2) = · · · = Lint P (d) = 0.
Conclude that the same statement is true for any unimodular simplex.

9.6. Here we show that the Bernoulli polynomial Bd (t), is essentially equal to the Ehrhart
polynomial LP (t) for the “Pyramid over a cube” (as defined in Exercise 7.5). We recall the
definition: let C := [0, 1]d−1 be the d − 1-dimensional cube, considered as a subset of Rd , and
let ed be the unit vector pointing in the xd -direction. Now we define P := conv{C, ed }, a
pyramid over the unit cube. Show that its Ehrhart polynomial is
1
LP (t) = (Bd (t + 2) − Bd ),
d
for t ∈ Z>0 .

190
9.7. For any integer d-dimensional (convex) polytope P ⊂ Rd , show that
d  
!
(−1)d X d
vol P = 1+ (−1)k LP (k) , (9.29)
d! k=1
k

which can be thought of as a generalization of Pick’s formula to Rd .

Note. Using iterations of the forward difference operator

∆f (n) := f (n + 1) − f (n),

the latter identity may be thought of a discrete analogue of the d’th derivative of the
Ehrhart polynomial. This idea in fact gives another method of proving (9.29).

9.8. Show that Pick’s formula is the special case of Exercise 9.7 when the dimension d = 2.
That is, given an integer polygon P ⊂ R2 , we have
1
AreaP = I + B − 1,
2
where I is the number of interior integer points in P, and B is the number of boundary
integer points in P.

9.9. Show that the convolution of the indicator function 1P with the heat kernel G , as in
equation (8.10), is a Schwartz function.

9.10. Show that a unimodular triangle always has area equal to 12 .

9.11. Show that the Ehrhart polynomial of the standard simplex ∆ ⊂ Rd is


 
t+d
L∆ (t) = .
d

9.12. Consulting Figure 9.6:

(a) Find the integer point transform of the unimodular polygon in the Figure.

(b) Find the Ehrhart polynomial LP (t) of the integer polygon P from part (a).

191
9.13. ♣ Show that (8.5) is equivalent to the following definition, using balls instead of
spheres. Recall that the unit ball in Rd is define by B d := {x ∈ Rd | kxk ≤ 1}, and
similarly the ball of radius , centered at x ∈ Rd , is denoted by B d (x, ). Show that for all
sufficiently small , we have
vol(S d−1 (x, ) ∩ P) vol(B d (x, ) ∩ P)
= .
vol(S d−1 (x, )) vol(B d (x, ))

9.14. Here we gain some practice with ‘floors’, ‘ceilings’, and ‘fractional parts’. First, we
recall that by definition, the fractional part of any real number
( x is {x} := x − bxc. Next,
1 if x ∈ Z
we recall the indicator function of Z, defined by: 1Z (x) := .
0 if x ∈/Z

Show that:

(a) dxe = − b−xc


(b) 1Z (x) = bxc − dxe + 1
(c) {x} + {−x} = 1 − 1Z (x)
(d) dxe = x + 1 − {x} − 1Z (x)
 n−1  n
(e) Let m ∈ Z>0 , n ∈ Z. Then m
+1= m
.

9.15. ♣ Show that the number of nonnegative integer solutions x1 , . . . , xd , z ∈ Z≥0 to

x1 + · · · + xd + z = t,
t+d

with 0 ≤ z ≤ t, equals d
.

9.16. ♣ Show that for each positive integer t, the number of positive integer solutions to
x1 + · · · + xd < t is equal to t−1

d
.

9.17. We define the rational triangle whose vertices are (0, 0), (1, NN−1 ), (N, 0), where N ≥ 2
is a fixed integer. Prove that the Ehrhart quasi-polynomial is in this case
p−1 2 p+1
LP (t) = t + t + 1,
2 2
for all t ∈ Z> 0.

Notes. So we see here a phenomenon known as ‘period collapse’, where we expect a quasi-
polynomial behavior, with some nontrivial period, but in fact we observe a strict polynomial.

192
9.18. Here we show that the Ehrhart polymomial LP (t) remains invariant under the full
unimodular group SLd (Z). In particular, recalling definition 6.25, of a unimodular matrix,
show that:

(a) Every element of SLd (Z) acts on the integer lattice Zd bijectively.

(b) Let P be an integral polytope, and let Q := A(P), where A ∈ SLd (Z). Thus, by
definition P and Q are unimodular images of each other. Prove that

LP (t) = LQ (t),

for all t ∈ Z>0 .

(c) Is the converse of part (b) true? In other words, suppose that LP (t) = LQ (t), for all
positive integers t. Does it necessarily follow that Q := A(P), for some unimodular
matrix A ∈ SLd (Z)?

193
Chapter 10

Sphere packings

The problem of packing, as densely as possible, an unlimited number of equal


nonoverlapping circles in a plane was solved millions of years ago by the bees,
who found that the best arrangement consists of circles inscribed in the hexagons
of the regular tessellation. – H. S. M. Coxeter

There is geometry in the humming of the strings. There is music in the spacing
of the spheres. – Pythagoras

Figure 10.1: A lattice sphere packing, using the hexagonal lattice, which gives the densest
packing in 2 dimensions.

10.1 Intuition

The sphere packing problem traces its roots back to Kepler, and it asks for a packing of
solid spheres in Euclidean space that achieves the maximum possible density. In all of the

194
known cases, such optimal configurations - for the centers of the spheres - form a lattice.
It’s natural, therefore, that Fourier analysis comes into the picture. We prove here a result
of Cohn and Elkies, from 2003, which is a beautiful application of Poisson summation, and
gives upper bounds for the maximum densities of sphere packings in Rd .

At this point it may be wise to define carefully all of the terms - what is a packing? what is
density? Who was Kepler?

10.2 Definitions

A sphere packing in Rd is any arrangement of spheres of fixed radius r > 0 such that no
two interiors overlap, so we do not preclude the possibility that the spheres may touch one
another at some points on their boundary.

A lattice packing is a sphere packing with the property that the centers of the spheres form
a lattice L ⊂ Rd . Relaxing this restriction a little, a periodic packing is a sphere packing
with a lattice L, together with a finite collection of its translates, say L + v1 , . . . , L + vN ,
such that the differences vi − vj ∈/ L. This means that the centers of the spheres may be
placed at any points belonging to the disjoint union of L, together with its N translates.

Figure 10.2: A periodic packing, which is not a lattice packing, with two translates of the
same lattice.

The density of any sphere packing is intuitively the proportion of Euclidean space covered
by the spheres, in an asymptotic sense, but rather than go into these technical asymptotic
details, we will simply define a density function for lattice packings and for general periodic
packings, as follows. Given a lattice packing, with the lattice L ⊂ Rd , and with spheres of
radius r, we define its lattice packing density by

vol B d (r)
∆(L) := , (10.1)
det L
195
where B d (r) is a ball of radius r. This lattice packing corresponds to placing a sphere of
radius r at each lattice point of L, guaranteeing that the spheres do not overlap.

More generally, given a period packing with a lattice L and a set of translates v1 , . . . , vN , we
define its periodic packing density by

N vol B d (r)
∆periodic (L) := , (10.2)
det L
corresponding to placing a sphere of radius r at each point of L, and also at each point of its
translates L + v1 , . . . , L + vN . It’s not hard to prove that the latter definition 10.2 matches
our intuition that any fixed fundamental parallelepiped of L intersects this configuration of
spheres in a set whose measure is exactly N vol B d (r) (Exercise 10.1).

Henceforth, we use the words ‘packing density’ to mean ‘periodic packing density’, and we
always restrict attention to periodic packings - see the Notes for technical remarks involving
any sphere arrangement, and why periodic packings are sufficient.

We define the sphere packing problem as follows:

What is the maximum possible packing density, in any periodic packing of


spheres?

In other words, the problem asks us to find the maximum density ∆periodic L, among all
lattices (and their finite collections of translates). The sphere packing problem also asks us
to find, if possible, the lattice that achieves this optimal density.

Many other questions naturally arise: is such a lattice unique in each dimension? Are there
examples of dimensions d for which a single lattice does not suffice, in order to obtain the
maximum sphere packing, but for which we do in fact need to use some lattice translates
of a fixed lattice? The sphere packing problem is a very important problem in Geometry,
Number theory, Coding theory, and information theory.

10.3 Upper bounds for sphere packings via Poisson


summation

Here we give an exposition of the ground-breaking result of Henry Cohn and Noam Elkies on
the sphere packing problem. This result sets up the machinery for finding certain magical
functions f , as defined in Theorem 10.1 below, that allow us to give precise upper bounds
on ∆periodic L. The main tool is Poisson summation again, for arbitrary lattices.

196
Theorem 10.1 (Cohn-Elkies). Let f : Rd → R be a nice function, not identically zero,
which enjoys the following three conditions:

1. f (x) ≤ 0, for all kxk ≥ r.

2. fˆ(ξ) ≥ 0, for all ξ ∈ Rd .

3. f (0) > 0, and fˆ(0) > 0.

Then the periodic packing density of any d-dimensional sphere packing has the upper bound

f (0)
∆periodic (L) ≤ vol B d (r).
ˆ
f (0)

Proof. Suppose we have a periodic packing with spheres of radius r, a lattice L, and trans-
N vol(B d (r))
lation vectors v1 , . . . , vN , so that by definition the packing density is δ := det L
.

By Poisson summation, we have


X 1 X ˆ
f (n + v) = f (ξ)e2πihv,ξi , (10.3)
n∈L
det L ξ∈L∗

converging absolutely for all v ∈ Rd . Now we form the following finite sum and rearrange
the right-hand-side of Poisson summation:
X X 1 X ˆ X
f (n + vi − vj ) = f (ξ) e2πihvi −vj ,ξi (10.4)
1≤i≤j≤N n∈L
det L ξ∈L∗ 1≤i≤j≤N
1 X X 2πihvk ,ξi 2

= fˆ(ξ) e . (10.5)
det L ξ∈L∗ 1≤k≤N

Now, every summand on the right-hand-side of (10.5) is nonnegative, because by the second
assumption of the Theorem, we have fˆ ≥ 0, so that the whole series can be bounded from
ˆ(0)N 2
below by its constant term, which for ξ = 0 gives us the bound fdet L
.

On the other hand, considering the vectors n + vi − vj on the left-hand-side of (10.4),


suppose we have kn + vi − vj k ≥ r. Then the first hypothesis of the Theorem guarantees that
f (n + vi − vj ) ≤ 0. If we have kn + vi − vj k < r, then the vector n + vi − vj is contained in
the sphere of radius r, centered at the origin, but this means that it must be the zero vector:
n + vi − vj = 0. By assumption, the difference between any two translations vi − vj is never
a nonzero element of L, so now we have vi = vj , which implies that n = 0. We conclude
that the only positive contribution from the left-hand-side of (10.4) is the n = 0 term, and
so the left-hand-side of (10.4) has an upper bound of N f (0).

197
Altogether, Poisson summation gave us the bound:
X X 1 X ˆ X 2πihvk ,ξi 2 fˆ(0)N 2
N f (0) ≥ | f (n + vi − vj )| = f (ξ) e ≥ .
1≤i≤j≤N n∈L
det L ξ∈L∗ 1≤k≤N
det L

Simplifying, we have
f (0) N ∆periodic (L)
≥ := ,
fˆ(0) det L vol B d (r)
where the last equality follows from the definition of ∆periodic (L).

Example 10.1 (The trivial bound). Let L be a full-rank lattice in Rd , whose shortest
nonzero vector has length r > 0. We define the function

f (x) := 1K (x) ∗ 1K (x),

where K is the ball of radius r, centered at the origin. We claim that f satisfies all of the
conditions of Theorem 10.1. Indeed, by the convolution Theorem,
 2
ˆ \
f (ξ) = (1K ∗ 1K )(ξ) = 1̂K (ξ) ≥ 0,

for all ξ ∈ Rd , verifying condition 2. Condition 1 is also easy to verify, because the support
of f is equal to the Minkowski sum (by Exercise 5.4) K + K = 2K, a sphere of radius 2r.
It follows that f is identically zero outside
R a sphere of radius 2r. R For condition 3, by the
definition of convolution we have f (0) = Rd 1K (0 − x)1K (x)dx = Rd 1K (x)dx = vol(K) > 0.
 2
Finally, fˆ(0) = 1̂K (0) = vol2 (K) > 0.

By the Cohn-Elkies Theorem 10.1, we know that the packing density of such a lattice is
therefore bounded above by

f (0) vol(K)
vol B d (r) = vol(K) = 1,
fˆ(0) vol2 (K)

the trivial bound. So we don’t get anything interesting, but all this tells us is that our par-
ticular choice of function f above was a poor choice, as far as density bounds are concerned.
We need to be more clever in picking our magical f . 

Although it is far from trivial to find magical functions f that satisfy the hypothesis of the
Cohn-Elkies Theorem, and simultaneously give a strong upper bound, there has been huge
success recently in finding exactly such functions - in dimensions 8 and 24. These recent
magical functions gave the densest sphere packings in these dimensions, knocking off the
whole sphere packing problem in dimensions 8 and 24.

198
This exciting story continues today, and we mention some of the recent spectacular appli-
cations of the Cohn-Elkies Theorem, initiated recently by Maryna Viazovska for R8 , and
then extended by a large joint effort from Henry Cohn, Abhinav Kumar, Stephen D. Miller,
Danylo Radchenko, and Maryna Viazovska, for R24 [37]. Here is a synopsis of some of their
results.
Theorem 10.2. The lattice E8 is the densest periodic packing in R8 . The Leech lattice is
the densest periodic packing in R24 . In addition, these lattices are unique, in the sense that
there do not exist any other periodic packings that achieve the same density.

At the moment, the provably densest packings are known only in dimensions 1, 2, 3, 8, and 24.
Each dimension seems to require slightly different methods, and sometimes wildly different
methods, such as R3 . For R3 , the sphere packing problem was solved by Hales, and before
Hales’ proof, it was an open problem since the time of Kepler. Somewhat surprisingly, the
sphere packing problem is still open in all other dimensions.

In R4 , it is very tempting and natural to think of the lattice D4 as a possible candidate for
the densest lattice sphere packing in R4 , but this is still unknown.

10.4 The Fourier transform of the ball

Whenever considering packing or tiling by a convex body B, we have repeatedly seen that
taking the Fourier transform of the body, namely 1̂B , is very natural, especially from the
perspective of Poisson summation. Thus, it is also natural here to consider  the FT of a ball
in
Rd . While we are at it, let’s dilate the unit ball by c > 0, defining Bc := x ∈ Rd | kxk ≤ c .

To compute the Fourier transform of 1Bc , a very classical computation, we first define the
Bessel function Jp of order p ([54], page 147), which comes up naturally here:
Z π
(x/2)p
Jp (x) := 1 √
 eix cos ϕ sin2p (ϕ) dϕ. (10.6)
Γ p+ 2 π 0

Lemma 10.1. The Fourier transform of Bc , the ball of radius c in Rd centered at the origin,
is Z  d/2
−2πihξ,xi c 
1̂Bc (ξ) := e dx = Jd/2 2πckξk .
Bc kξk

Proof. Taking advantage of the inherent rotational symmetry of the ball, and also using the
fact that the Fourier transform of a radial function is again radial (Exercise 10.4), we have:
1̂Bc (ξ) = 1̂Bc (0, . . . , 0, kξk),

199
for all ξ ∈ Rd . With c = 1 for the moment, we therefore have:
Z
1̂B1 (ξ) = e−2πixd kξk dx1 . . . dxd ,
kxk≤1

Now we note that for each fixed xd , the function being integrated is constant and the integra-
tion domain for the variables x1 , . . . , xd−1 is a (d − 1)-dimensional ball of radius (1 − x2d )1/2 .
d−1
d−1
π 2
Using the classical fact that the volume of this ball is (1 − x2d ) 2
Γ( d+1
(see Exercise 10.5),
2 )
we have
d−1 Z 1 d Z π
π 2 −2πixd kξk d−1 π2
1̂B1 (ξ) = d+1 e (1 − x2d ) 2 dxd = √ e2πikξk cos ϕ sind ϕ dϕ.
πΓ d+1

Γ( 2 ) −1 2 0

Using the definition (10.6) of the J-Bessel function, we get


d
1̂B1 (ξ) = kξk− 2 J d 2πkξk ,

2

and consequently
  d2
c 
1̂Bc (ξ) = J d 2πckξk .
kξk 2

We call a function f : Rd → C radial if it is invariant under all rotations of Rd . In other


words, we have the definition

f is radial ⇐⇒ f ◦ σ = f,

for all σ ∈ SOd (R), the orthogonal group. Another way of describing a radial function is
to say that the function f is constant on each sphere that is centered at the origin, so that
intuitively a radial function only depends on the norm of its input.

A very useful fact in various applications of Fourier analysis (in particular medical imaging)
is that the Fourier transform of a radial function is again a radial function, and in fact we’ve
already used this fact in the proof of Lemma 10.1.

Now we may notice that if we have a magical function f that enjoys all three hypotheses
of the Cohn-Elkies Theorem 10.1, then it is easy to see that f ◦ σ also satisfies the same
hypotheses, for any σ ∈ SOd (R) (Exercise 10.6), and therefore we make take radial functions
as candidates for magical functions as well.

200
Notes

(a) Each dimension d appears to have a separate theory for sphere packings. This intuition
is sometimes tricky to conceptualize, but there are facts that help us do so. For
example, it is a fact that the Gram matrix (see 6.22) of a lattice L ⊂ Rd consists
entirely of integers, with even diagonal elements ⇐⇒ d is divisible by 8. For
this reason, it turns out that the theta series of a lattice possesses certain functional
equations (making it a modular form) if and only if 8 | d, which in turn allows us to
build some very nice related ‘magical’ functions f that are sought-after in Theorem
10.1, at least for d = 8 and d = 24 so far.
In dimension 2, it is an open problem to find such magical functions, even though we
have an independent proof that the hexagonal lattice is the optimal sphere packing
lattice.

(b) Johannes Kepler (1571 –1630) was a German astronomer and mathematician. Kepler’s
laws of planetary motion motivated Sir Isaac Newton to develop further the theory of
gravitational attraction and planetary motion. Kepler conjectured that the densest
packing of sphere is given by the “face-centered cubic” packing. It was Gauss (1831)
who first proved that, if we assume the packing to be a lattice packing, then Kepler’s
conjecture is true. In 1998 Thomas Hales (using an approach initiated by L. Fejes
Tóth (1953)), gave an unconditional proof of the Kepler conjecture.

(c) It is also possible, of course, to pack other convex bodies. One such variation is to pack
regular tetrahedra in R3 . The interesting article by Jeffrey Lagarias and Chuanming
Zong [89] gives a nice account of this story.

(d) Regarding lower bounds for the optimal density of sphere packings, Keith Ball [6]
discovered the following lower bound in all dimensions:
(n − 1)
∆periodic (L) ≥ ζ(n),
2n−1
where ζ(s) is the Riemann zeta function. Recently, Akshay Venkatesh [157] has given
an improvement over the known lower bounds by a multiplicative constant. For all
sufficiently large dimensions, this improvement is by a factor of at least 10, 000.

Exercises

10.1. Given a periodic lattice packing, by N translates of a lattice L ⊂ Rd , show that


any fixed fundamental parallelepiped of L intersects the union of all the spheres in a set of

201
measure N vol B d (r) , where r := 21 λ1 (L). Thus, we may compute the density of a periodic


sphere packing by just considering the portions of the spheres that lie in one fundamental
parallelepiped.

10.2. Suppose we pack equilateral triangles in the plane, by using only translations of a fixed
equilateral triangle. What is the maximum packing density of such a packing? Do you think
it may be the worst possible density among translational packings of any convex body in R2 ?

10.3. Here we show that the integer lattice is in fact a very poor choice for sphere packing.

(a) Compute the packing density of the integer lattice Z2 .

(b) Compute the packing density of the integer lattice Zd .

10.4. Show that the Fourier transform of a radial function f : Rd → C is always another
radial function.


10.5. Show that the volume of the unit ball B1 := x ∈ Rd | kxk ≤ 1 is:
d−1
π 2
vol B1 = .
Γ d+1
2

10.6. Show that if we have a magical function f that enjoys all 3 hypotheses of Theorem 10.1,
then f ◦ σ also satisfies the same hypotheses, for any orthogonal transformation σ ∈ SOd (R).

202
Chapter 11

The Fourier transform of a polytope


via its hyperplane description:
the divergence Theorem

Like a zen koan, Stokes’ Theorem tells us that in the end, what happens on the
outside is purely a function of the change within.
–Keenan Crane

Figure 11.1: A real vector field in R2

203
11.1 Intuition

The divergence Theorem is a multi-dimensional version of “integration by parts”, a very


useful tool in 1-dimensional calculus. When we apply the divergence Theorem, described
below, to a polytope, we obtain a kind of combinatorial version of the divergence Theorem,
allowing us to transfer some of the complexity of computing the Fourier transform of a
polytope to the complexity of computing corresponding Fourier transforms of its facets.
This kind of game can be iterated, yielding interesting geometric identities and results for
polytopes, as well as for discrete volumes of polytopes.

In the process, we also obtain another useful way to compute the Fourier transform of a
polytope in its own right.

11.2 The divergence Theorem, and a combinatorial


divergence theorem for polytopes

To warm up, we recall the Divergence Theorem, with some initial examples. A vector field
on Euclidean space is a function F : Rd → Cd that assigns to each point in Rd another vector
in Cd , which we will denote by

F (x) := (F1 (x), F2 (x), . . . , Fd (x)) ∈ Cd .

If F is a continuous (respectively, smooth) function, we say that F is a continuous vector


field (respectively, smooth vector field). If all of the coordinate functions Fj are real-
valued functions, we say that we have a real vector field.

We define the divergence of F at each x := (x1 , . . . , xd ) ∈ Rd by

∂F1 ∂Fd
divF(x) := + ··· + ,
∂x1 ∂xd
assuming that F is a smooth (or at least once-differentiable) vector field. This divergence
of F is a measure of the local change (sink versus source) of the vector field at each point
x ∈ Rd . Given a surface S ⊂ Rd , and an outward pointing unit normal vector n, defined at
each point x ∈ S, we also define the flux of the vector field F across the surface S by
Z
F · n dS,
S

where dS denotes the Lebesgue measure Pdof the surface S, and where the dot product F · n
is the usual inner product hF, ni := k=1 Fk nk . We will apply the divergence Theorem
(which is technically a special case of Stokes’ Theorem) to a polytope P ⊂ Rd , and its

204
(d − 1)-dimensional bounding surface ∂P. Intuitively, the divergence Theorem tells us that
the total divergence of a vector field F inside a manifold is equal to the total flux of F across
its boundary.
Theorem 11.1 (The Divergence Theorem). Let M ⊂ Rd be a piecewise smooth manifold,
and let F be a smooth vector field. Then
Z Z
divF(x)dx = F · n dS. (11.1)
M S

Example 11.1. Let P ⊂ Rd be a d-dimensional polytope, containing the origin, with


defining facets G1 , . . . , GN . Define the real vector field
F (x) := x,
for all x ∈ Rd . First, we can easily compute here the divergence of F , which turns out to be
constant:
∂F1 ∂Fd ∂x1 ∂xd
divF(x) = + ··· + = + ··· + = d.
∂x1 ∂xd ∂x1 ∂xd
If we fix any facet G of P then, due to the piecewise linear structure of the polytope, every
point x ∈ G has the same constant outward pointing normal vector to F , which we call nG .
Computing first the left-hand-side of the divergence Theorem, we see that
Z Z
divF(x)dx = d dx = (vol P)d. (11.2)
P P

Computing now the right-hand-side of the divergence Theorem, we get


Z Z XN Z
F · n dS = hx, ni dS = hx, nG i dS.
S ∂P k=1 Gk

Now it’s easy to see that the inner product hx, nG i is constant on each facet G ⊂ P, namely
it is the distance from the origin to G (Exercise 11.3), denoted by dist(G). So we now have
Z XN Z
F · n dS = hx, nGk idS
∂P k=1 Gk

XN Z N
X
= dist(Gk ) dS = dist(Gk ) vol Gk ,
k=1 Gk k=1

so that altogether we the following conclusion from the divergence Theorem:


N
1X
vol P = dist(Gk ) vol Gk . (11.3)
d k=1

known as “the pyramid formula” for a polytope, a classical result in Geometry, which also
has a very easy geometrical proof (Exercise 11.1). 

205
Example 11.2. Let P ⊂ Rd be a d-dimensional polytope with defining facets G1 , . . . , GN ,
and outward pointing unit vectors nG1 , . . . , nGN . We fix any constant vector λ ∈ Cd , and we
consider the constant vector field

F (x) := λ,

defined for all x ∈ Rd . Here the divergence of F is divF(x) = 0, because F is constant, and
so the left-hand-side of Theorem 11.1 gives us
Z
divF(x)dx = 0.
P

Altogether, the divergence Theorem gives us:


Z N Z
X
0= F · n dS = hλ, nGk idS
∂P k=1 Gk

XN Z
= hλ, nGk i dS
k=1 Gk
N
X
= hλ, vol(Gk )nGk i,
k=1

and because this holds for any constant vector λ, we can conclude that
N
X
vol(Gk )nGk = 0. (11.4)
k=1

Identity (11.4) is widely known as the Minkowski relation for polytopes. There is a mar-
velous converse to the latter relation, given by Minkowski as well, for any convex polytope.
[See Theorem 11.7] 

Now we fix ξ ∈ Rd and apply the divergence Theorem to the vector-field

F (x) := e−2πihx,ξi ξ. (11.5)

First,

∂ e−2πihx,ξi ξ1

∂(e−2πihx,ξi ξd )
divF(x) = + ··· +
∂x1 ∂xd
2 −2πihx,ξi
= (−2πiξ1 )e + · · · + (−2πiξd2 )e−2πihx,ξi
= −2πikξk2 e−2πihx,ξi .

206
So here the divergence Theorem gives us
Z Z
2 −2πihx,ξi
−2πi||ξ|| e dx = e−2πihx,ξi hξ, ni dS,
x∈P ∂P

which gives us another version of the Fourier transform of a polytope:


Z
1̂P (ξ) := e−2πihx,ξi dx
x∈P
Z
1
= hξ, nie−2πihx,ξi dS
−2πikξk2 ∂P
Z Z
1 −2πihx,ξi 1
= hξ, nG1 ie dS + · · · + hξ, nGN ie−2πihx,ξi dS
−2πikξk2 G1 −2πikξk2 GN
hξ, nG1 i hξ, nGN i
= 1̂G 1 (ξ) + · · · + 1̂G (ξ),
−2πikξk2 −2πikξk2 N
where we used the fact that the boundary R∂P of Ra polytope Ris a finite union of (d − 1)-
dimensional polytopes (its facets), to write ∂P = G1 + · · · + GN , a sum of integrals over
the N facets of P. We’ve arrived at the following formula for the Fourier transform of P:
Lemma 11.1. Given any d-dimensional polytope P ⊂ Rd , with outward pointing normal
vector nG to each facet G of P, its Fourier transform has the form
1 X hξ, nG i
1̂P (ξ) = 1̂G (ξ) (11.6)
−2πi G⊂∂P ||ξ||2

for all nonzero ξ ∈ Rd . Here the integral that defines each 1̂G is taken with respect to Lebesgue
measure that matches the dimension of the facet G ⊂ ∂P .

To simplify the notation that will follow, we can also the Iverson bracket notation, defined
as follows. Suppose we have any boolean property P (n), where n ∈ Zd ; that is, P (n) is
either true or false. Then the Iverson bracket [P ] is defined by:
(
1 if P is true
[P ] = (11.7)
0 if P is false
Now we may refine (11.6) above as follows:
1 X hξ, nG i
1̂P (ξ) = vol P [ξ = 0] + 1̂G (ξ) [ξ 6= 0]. (11.8)
−2πi G⊂∂P ||ξ||2

Later, after Theorem 11.2 below, we will return to the Iverson bracket, and be able to use
it efficiently. To proceed further, we need to define the affine span of a face F of P:
( k k
)
X X
aff(F) := λj vj | k > 0, vj ∈ F, λj ∈ R, and λj = 1 . (11.9)
j=1 j=1

207
Figure 11.2: The affine span of a face F , its linear span
√ , and the projection of ξ onto F .
Here we note that the distance from the origin to F is 20.

In other words, we may think of the affine span of a face F of P as follows. We first
translate F so that this translate, call if F0 , contains the origin. Then we take all real linear
combinations of points of F0 , obtaining a vector subspace of Rd , which we call the linear
span of F . Another way to describe the linear span of a face F of P is:
lin(F ) := {x − y | x, y ∈ F } .
Finally, we translate this subspace lin(F) back using the same translation vector, to obtain
aff(F) (see Figure 11.2).
Example 11.3. The affine span of two distinct points in Rd is the unique line in Rd passing
through them. The affine span of three points in Rd is the unique 2-dimensional plane
passing through them. The affine span of a k-dimensional polytope F ⊂ Rd is a translate
of a k-dimensional vector subspace of Rd . Finally, the affine span of a whole d-dimensional
polytope P ⊂ Rd is all of Rd . 

In formalizing (11.6) further, we will require the notion of the projection of any point ξ ∈ Rd
onto the linear span of any face F ⊆ P, which we abbreviate by ProjF ξ:
ProjF ξ := Projlin(F) (ξ). (11.10)
(see Figure 11.2) We will also need the following elementary fact. Let F be any k-dimensional
polytope in Rd , and fix the outward-pointing unit normal to F , calling it nF . It is straight-
forward to show that if we take any point xF ∈ F , then hxF , nF i is the distance from

208
the origin to F . Therefore, if ProjF ξ = 0, then a straightforward computation shows that
hξ, xF i = kξkdist(F) (Exercise 11.3).

We now record the reasoning above, leading to (11.6), more formally but with one slight
extension: we replace P by any k-dimensional face F of P, as follows.

Theorem 11.2 (Combinatorial Divergence Theorem). Let F be a polytope in Rd , where


1 ≤ dim F ≤ d. For each facet G ⊆ F , we let n(G, F ) be the unit normal vector to G, with
respect to lin(F). Then for each ξ ∈ Rd , we have:

(a) If ProjF ξ = 0, then


1̂F (ξ) = (vol F )e−2πikξkdist(F) . (11.11)

(b) If ProjF ξ 6= 0, then

1 X hProjF ξ, n(G, F)i


1̂F (ξ) = 1̂G (ξ). (11.12)
−2πi G⊂∂F ||ProjF ξ||2

We notice that, as before, we are getting rational-exponential functions for the Fourier
transform of a polytope, but now we are using the facets of P when applying Theorem 11.2
to P itself.

We are now set up to iterate this process, defined by Theorem 11.2, reapplying it to each
facet G ⊂ ∂P. Let’s use the Iverson bracket, defined in (11.7), and apply the combinatorial
divergence Theorem 11.2 to P twice:

1 X hξ, nF i
1
1̂P (ξ) = vol P [ξ = 0] + 2
[ξ 6= 0] 1̂F1 (ξ)
−2πi F ⊂∂P ||ξ||
1

1 X hξ, nF i
1
= vol P [ξ = 0] + [ξ 6= 0]
−2πi F ⊂∂P ||ξ||2
1
 1 X hProjF ξ, n(F2 , F1 )i 
· (vol F1 )e−2πihξ,xi [ProjF1 ξ = 0] + 2
1̂F2 (ξ)[Proj F1 ξ 6
= 0]
−2πi F ⊂∂F ||ProjF2 ξ||2
2 1

1 X hξ, nF i(vol F1 )e−2πihξ,xi


1
= vol P [ξ = 0] + [ξ 6= 0][ProjF1 ξ = 0]
−2πi F ⊂∂P ||ξ||2
1

1 X X hξ, nF i hProjF ξ, n(F2 , F1 )i


1
+ 2
1̂F2 (ξ) [ξ 6= 0][ProjF1 ξ 6= 0]
(−2πi)2 F1 ⊂∂P F ⊂∂F
||ξ||2 ||ProjF2 ξ||2
2 1

209
Figure 11.3: The face poset of a triangle

It is an easy fact that the product of two Iverson brackets is the Iverson bracket of their
intersection: [P ][Q] = [P and Q] (Exercise 11.11). Hence, if we define

F ⊥ := {x ∈ Rd | hx, yi = 0 for all y ∈ linF},

Then we see that P ⊥ = {0}, and we can rewrite the latter identity as

1 X hξ, nF i(vol F1 )e−2πihξ,xi


1̂P (ξ) = vol P [ξ ∈ P ⊥ ] + 1
[ξ ∈ F1⊥ − P ⊥ ]
−2πi F ⊂∂P ||ξ||2
1

1 X X hξ, nF i hProjF ξ, n(F2 , F1 )i


+ 1 2
1̂F2 (ξ) [ξ 6∈ F1⊥ ].
(−2πi)2 F1 ⊂∂P F ⊂∂F
||ξ||2 ||ProjF2 ξ||2
2 1

In order to keep track of the iteration process, we will introduce another book-keeping device.
The face poset of a polytope P is defined to be the partially ordered set (poset) of all faces
of P, ordered by inclusion, including P and the empty set.

Example 11.4. Consider a 2-dimensional polytope P that is a triangle. We have the


following picture for the face poset FP of P, as in Figure 11.3. It turns out that if we
consider a d-simplex P, then its face poset FP has the structure of a “Boolean poset”, which
is isomorphic to the edge graph of a (d + 1)-dimensional cube.

We only have to consider rooted chains in the face poset FP , which means chains whose root
is P . The only appearance of non-rooted chains are in the following definition. If G is a
facet of F , we attach the following weight to any (local) chain (F, G), of length 1, in the face
poset of P :

−1 hProjF (ξ), n(G, F )i


W(F,G) (ξ) := . (11.13)
2πi k ProjF (ξ)k2

210
Figure 11.4: A symbolic depiction of the face poset FP , where P is a 3-dimensional tetra-
hedron. Here the points and arrows are drawn suggestively, as a directed graph. We can
see all the rooted chains, beginning from a symbolic vertex in the center, marked with the
color purple. The rooted chains that terminate with the yellow vertices have length 1, those
that terminate with the green vertices have length 2, and those that terminate with the blue
vertices have length 3.

Note that these weights are functions of ξ rather than constants. Moreover, they are all
homogeneous of degree −1. Let T be any rooted chain in FP , given by
T := (P → F1 → F2 , . . . , → Fk−1 → Fk ),
so that by definition dim(Fj ) = d − j. We define the admissible set S(T) of the rooted
chain T to be the set of all vectors ξ ∈ Rd that are orthogonal to the linear span of Fk but
not orthogonal to the linear span of Fk−1 . In other words,
S(T) := {ξ ∈ Rd | ξ ⊥ lin(Fk ), but ξ 6⊥ lin(Fk−1 )}
= {ξ ∈ Rd | ξ ∈ Fk⊥ − Fk−1

}.

Finally, we define the following weights associated to any such rooted chain T:

(a) The rational weight RT (ξ) = R(P →...→Fk−1 →Fk ) (ξ) is defined to be the product of
weights associated to all the rooted chains T of length 1, times the Hausdorff volume
of Fk (the last node of the chain T). It is clear from this definition that RT (ξ) is a
homogenous rational function of ξ.

(b) The exponential weight ET (ξ) = E(P →...→Fk−1 →Fk ) (ξ) is defined to be the evaluation of
e−2πihξ,xi at any point x on the face Fk :
ET (ξ) := e−2πihξ,x0 i , (11.14)

211
for any x0 ∈ Fk . We note that the inner product hξ, x0 i does not depend on the position
of x0 ∈ Fk .

(c) The total weight of a rooted chain T is defined to be the rational-exponential function

WT (ξ) = W(P →...→Fk−1 →Fk ) (ξ) := RT (ξ)ET (ξ)1S(T) (ξ), (11.15)

where 1S(T) (ξ) is the indicator function of the admissible set S(T) of T.

By repeated applications of the combinatorial divergence Theorem 11.2, we arrive at a de-


scription of the Fourier transform of P as the sum of weights of all the rooted chains of the
face poset FP , as follows.

Theorem 11.3.
X X
1̂P (ξ) = WT (ξ) = RT (ξ)ET (ξ)1S(T) (ξ), (11.16)
T T

valid for any fixed ξ ∈ Rd .

For a detailed proof of Theorem 11.3, see [49]. Using this explicit description of the Fourier
transform of a polytope, we will see an application of it in the following section, for the
coefficients of Macdonald’s angle quasi-polynomial. In the process, equation (11.16), which
gives an explicit description of the Fourier transform of a polytope, using the facets of P as
well as lower-dimensional faces of P, will become even more explicit with some examples.

11.3 Generic frequencies versus special frequencies

Given a polytope P ⊂ Rd , we call a vector ξ ∈ Rd a generic frequency (relative to P) if ξ


is not orthogonal to any face of P. All other ξ ∈ Rd are orthogonal to some face F of P, and
are called special frequencies. Let’s define the following hyperplane arrangement, given
by the finite collection of hyperplanes orthogonal to any edge of P:

H := {x ∈ Rd | hx, F1 i = 0, for any 1-dimensional edge F1 of P }.

Then it is clear that the special frequencies are exactly those vectors that lie in the hyperplane
arrangement H. So we see from Theorem 11.3 that for a generic frequency ξ, we have
X
1̂P (ξ) = RT (ξ)e−2πihξ,F0 i , (11.17)
T:P →...→F1 →F0

212
where the F0 faces are the vertices of P. In other words, for generic frequencies, all of our
rooted chains in the face poset of P go all the way to the vertices. The special frequencies,
however, are more complex. But we can collect the special frequencies in ‘packets’, giving
us the following result.

Theorem 11.4 (Coefficients for Macdonald’s angle quasi-polynomial). [49] Let P be a d-


dimensional rational polytope in Rd , and let t be a positive real number. Then we have the
quasi-polynomial
Xd
AP (t) = ai (t)ti ,
i=0

where, for 0 ≤ i ≤ d,
2
X X
ai (t) := lim+ RT (ξ)ET (tξ) e−πkξk , (11.18)
→0
ξ∈Zd ∩S(T) l(T)=d−i

where l(T) is the length of the rooted chain T in the face poset of P , RT (ξ) is the rational
function of ξ defined above, ET (tξ) is the complex exponential defined in (11.14) above, and
Zd ∩ S(T) is the set of all integer points that are orthogonal to the last node in the chain T ,
but not to any of its previous nodes.

See [49] for the detailed proof of Theorem 11.4.

We call the coefficients ai (t) the quasi-coefficients of the solid angle sum AP (t). As a
consequence of Theorem 11.4, it turns out that there is a closed form for the codimension-1
quasi-coefficient, which extends previous special cases of this coefficient.

We recall our first periodic Bernoulli polynomial, from (3.16):


(
x − bxc − 21 if x ∈/Z
P1 (x) := (11.19)
0 if x ∈ Z,

where bxc is the integer part of x.

Theorem 11.5. [49] Let P be any real polytope. Then the codimension-1 quasi-coefficient
of the solid angle sum AP (t) has the following closed form:
X vol(F )
ad−1 (t) = − P1 (hvF , xF it), (11.20)
F a facet of P
kvF k
with vF 6=0

where vF is the primitive integer vector which is an outward-pointing normal vector to F ,


xF is any point lying in the affine span of F , and t is any positive real number.

213


We note that, rather surprisingly, the latter formula shows in particular that for any real
polytope P, the quasi-coefficient ad−1 (t) is always a periodic function of t > 0, with a period
of 1. Although it is not necessarily true that for any real polytope the rest of the quasi-
coefficients ak (t) are periodic functions of t, it is true that in the case of rational polytopes,
the quasi-coefficients are periodic functions of all real dilations t, as we show below.

We recall that zonotopes are projections of cubes or, equivalently, polytopes whose faces
(of all dimensions) are symmetric. We also recall the result of Alexandrov and Shephard
(Theorem 5.9) from chapter 5: If all the facets of P are symmetric, then P must be symmetric
as well. The following result appeared in [13], and here we give a different proof, using the
methods of this chapter.

Theorem 11.6. Suppose P is a d-dimensional integer polytope in Rd all of whose facets are
centrally symmetric. Then
AP (t) = (vol P)td ,
for all positive integers t.

Proof. We recall the formula for the solid angle polynomial AP (t).
2
X
AP (t) = lim+ 1̂tP (ξ)e−πkξk . (11.21)
→0
ξ∈Zd

The Fourier transform of the indicator function of a polytope may be written as follows,
after one application of the combinatorial divergence formula.
 
d −1 d−1 X hξ, nF i
1̂tP (ξ) = t vol(P)[ξ = 0] + t 1̂F (tξ)[ξ 6= 0], (11.22)
2πi F ⊆P
kξk2
dim F =d−1

where we sum over all facets F of P. Plugging this into (11.21) we get
2
e−πkξk
 
d −1 d−1
X X
AP (t) − t vol(P) = t lim+ hξ, nF i1̂F (tξ) (11.23)
2πi →0 kξk2 F ⊆P
ξ∈Zd \{0}
dim F =d−1

Thus, if we show that the latter sum over the facets vanishes, then we are done.

The assumption that all facets of P are centrally symmetric implies that P itself is also
centrally symmetric, by Theorem 5.9. We may therefore combine the facets of P in pairs of
opposite facets F and F 0 . We know that F 0 = F + c, where c is an integer vector, using the
fact that the facets are centrally symmetric.

214
Therefore, since n0F = −nF , we have
hξ, nF i1̂F (tξ) + hξ, −nF i1̂F +c (tξ)
= hξ, nF i1̂F (tξ) − hξ, nF i1̂F (tξ)e−2πihtξ,ci
= hξ, nF 1̂F (tξ) 1 − e−2πihtξ,ci = 0,


because htξ, ci ∈ Z when both ξ ∈ Zd and t ∈ Z. We conclude that the entire right-hand
side of (11.23) vanishes, and we are done.

Fourier analysis can also be used to give yet more general classes of polytopes that satisfy
the formula AP (t) = (vol P)td , for positive integer values of t (See also [100], [43]).

There is a wonderful result of Minkowski that gives a converse to the relation (11.4), as
follows.
Theorem 11.7 (The Minkowski problem for polytopes). Suppose that u1 , . . . , uk ∈ Rd are
unit vectors that do not lie in a hyperplane. Suppose further that we are given positive
numbers α1 , α2 , . . . , αk > 0 that satisfy the relation
α1 u1 + · · · + αk uk = 0.
Then there exists a polytope P ⊂ Rd , with facet normals u1 , . . . , uk ∈ Rd , and facet areas
α1 , α2 , . . . , αk . Moreover, this polytope P is unique, up to translations. 

There is a large body of work, since the time of Minkowski, that is devoted to extensions of
Minkowski’s Theorem 11.7, to other convex bodies, as well as to other manifolds.

Notes

(a) We could also define another useful vector field, for our combinatorial divergence Theo-
rem, besides our vector field in equation (11.5). Namely, if we define F (x) := e2πihx,ξi λ,
for a fixed λ ∈ Cd , then we would get the analogous combinatorial divergence formula
as shown below in (Exercise 11.4), and such vector fields have been used, for example,
by Alexander Barvinok [10] in an effective way. To the best of our knowledge, the first
researcher to use iterations of Stokes’ formula to obtain lattice point asymptotics was
Burton Randol [122], [123].
(b) The Minkowski problem for polytopes can also be related directly to generalized isoperi-
metric inequalities for mixed volumes, as well as the Brunn-Minkowski inequality for
polytopes, as done by Daniel Klain in [84].

215
Exercises

11.1. ♣ We define the distance from the origin to F , denoted by dist(F), as the length of the
shortest vector of translation between aff(F) and lin(F) (resp. the affine span of F and the
linear span of F , defined in (11.9)). Figure 11.2 shows what can happen in such a scenario.

(a) Suppose that we consider a facet F of a given polytope P ⊂ Rd , and we let nF be the
unit normal vector to F . Show that the function
xF → hxF , nF i
is constant for xF ∈ F , and is in fact equal to the distance from the origin to F . In
other words, show that
hx, nF i = dist(F).
(b) Show that if ProjF ξ = 0, then hξ, xF i = kξkdistF.

11.2. Here we prove the elementary geometric formula for a pyramid over a polytope. Namely,
suppose we are given a (d − 1)-dimensional polytope P, lying in the vector space defined by
the first d−1 coordinates. We define a pyramid over P, of height h > 0, as the d-dimensional
polytope defined by
Pyr(P) := conv{P, h · ed },
where ed := (0, 0, . . . , 0, 1) ∈ Rd . Show that
h
vol Pyr(P) = vol P.
d

11.3. ♣ Prove the Pyramid formula, (11.3) in Example 11.1, for a d-dimensional polytope
P which contains the origin, but now using just elementary geometry:
N
1X
vol P = dist(Gk ) vol Gk , (11.24)
d k=1

where the Gk ’s are the facets of P, and dist(Gk ) is the distance from the origin to Gk .

11.4. ♣ Show that if we use the alternative vector field F (x) := e−2πihx,ξi λ in equation (11.5),
with a constant nonzero vector λ ∈ Cd , then we get:
1 X hλ, nG i
1̂P (ξ) = 1̂G (ξ), (11.25)
−2πi G⊂∂P hλ, ξi

valid for all nonzero ξ ∈ Rd . Note that one advantage of this formulation of the Fourier
transform of P is that each summand in the right-hand-side of (11.25) is free of singularities,
assuming the vector λ has a nonzero imaginary part.

216
Figure 11.5: The meaning of Minkowski’s relation in dimension 2 - see Exercise 11.8

11.5. Show that the identity (11.25) of Exercise 11.4 is equivalent to the vector identity:
1 X
ξ 1̂P (ξ) = nG 1̂G (ξ),
−2πi G⊂∂P

valid for all ξ ∈ Rd .

11.6. Show that the result of Exercise 11.5 quickly gives us the Minkowski relation (11.4):
X
vol(G)nG = 0.
facets G of P

11.7. Continuing Exercise 11.4, show that by iterating this particular version of the Fourier
transform of a polytope P, k times, we get:
k
1 X Y hλ, nGj ,Gj−1 i
1̂P (ξ) = 1̂Gk (ξ), (11.26)
(−2πi)k Gk ⊂Gk−1 ⊂···G1 ⊂∂P j=1
hλ, Proj G j−1
ξi

valid for all nonzero ξ ∈ Rd , and where we sum over all chains Gk ⊂ Gk−1 ⊂ · · · G1 of length
k in the face poset of P, with codim(Gj ) = j.

11.8. Show that in the case of polygons in R2 , the Minkowski relation (11.4) has the meaning
that the sum of the pink vectors in Figure 11.5 sum to zero. In other words, the geometric
interpretation of the Minkowski relation in dimension 2 is that the sum of the boundary
(pink) vectors wind around the boundary and close up perfectly.

11.9. ♣ Let’s consider a simplex ∆ ⊂ Rd whose dimension satisfies 2 ≤ dim ∆ ≤ d. Show


that ∆ is not a symmetric body.

217
11.10. Let F ⊂ Rd be a centrally symmetric, integer polytope of dimension k. Show that the
distance from the origin to F is always a half-integer or an integer. In other words, show
that
1
dist(F) ∈ Z.
2
(See Exercise 11.1 above for the definition of distance of F to the origin)

11.11. ♣ To get more practice with the Iverson bracket, defined in (11.7), show that:

(a) [P and Q] = [P][Q].

(b) [P or Q] = [P] + [Q] − [P][Q].

11.12. Show that for any polytope P ⊂ Rd , its Fourier transform 1̂P (ξ) is not in L1 (Rd ).

218
Solutions and hints
There are no problems, just pauses between ideas.
– David Morrell, Brotherhood of the Rose

Chapter 2

Exercise 2.1 By Euler, we have 1 = eiθ = cos θ + i sin θ, which holds if and only if cos θ = 1,
and sin θ = 0. The latter two conditions hold simultaneously if and only if θ ∈ 2πk, with
k ∈ Z.

a2 +b2
Exercise 2.2 Let z := a + bi, so that |ez | = |ea+bi | = |ea ||ebi | = ea · 1 ≤ e = e|z| .

Exercise 2.3 In case a 6= b, we have


Z 1 1
e2πi(a−b)
Z
ea (x)eb (x)dx = e2πi(a−b)x dx = − 1 = 0,
0 0 2πi(a − b)

because we know that a − b ∈ Z. In case a = b, we have


Z 1 Z 1
ea (x)ea (x)dx = dx = 1.
0 0

PN −1 2πik
Exercise 2.5 Let S := k=0 e N , and note that we may write
X 2πik
S= e N .
k mod N

2πim
Now, pick any m such that e N 6= 1. Consider
2πim
X 2πi(k+m)
e N S= e N
k mod N
2πin
X
= e N = S,
n mod N

2πim 2πim
so that 0 = (e N − 1)S, and since by assumption e N 6= 1, we have S = 0.

219
−1 N
Exercise 2.6 We use the finite geometric series: 1 + x + x2 + · · · + xN −1 = xx−1 . Now, if
2πiM
N 6 |M , then x := e N 6= 1, so we may substitute this value of x into the finite geometric
series to get:
N −1 2πiM N
1 X 2πikM e N −1
e N = 2πiM
N k=0 e N −1
0
= 2πiM = 0.
e N −1
P −1 2πikM P −1
On the other hand, if N | M , then N1 N k=0 e
N = N1 Nk=0 1 = 1.

Exercise 2.8 We begin with the factorization of the polynomial xn −1 = nk=1Q (x−ζ k ), with
Q
n−1
ζ := e2πi/n . Dividing both sides by x −Q1, we obtain 1 + x + x2 + · · · + xn−1 = k=1 (x − ζ k ).
n−1 k
Now substituting x = 1, we have n = k=1 (1 − ζ ).

Exercise 2.7
N −1 N −1
1 X 2πika − 2πikb 1 X 2πik(a−b)
e N e N = e N .
N k=0 N k=0

Therefore, using Exercise 2.6, we see that the latter sum equals 1 exactly when N | a − b,
and vanishes otherwise.

Exercise 2.4 By definition,


Z Z Z
−2πiξx
e dx := cos(2πξx)dx + i sin(2πξx)dx
[0,1] [0,1] [0,1]
sin(2πξ) − cos(2πξ) + 1
= +i
2πξ 2πξ
i sin(2πξ) cos(2πξ) − 1
= +
2πiξ 2πiξ
2πiξ
e −1
= .
2πiξ

Exercise 2.9 Suppose to the contrary, that a primitive N ’th root of unity is of the form
2πim/N m N
e , where gcd(m, N ) > 1. Let m1 := gcd(m,N )
, and k := gcd(m,N )
, so that by assumption
2πim/N 2πim1 /k
both m1 and k are integers. Thus e =e , a k’th root of unity, with k < N , a
contradiction.

220
Exercise 2.14 We recall Euler’s identity:

eiw = cos w + i sin w,

which is valid for all w ∈ C. Using Euler’s identity first with w := πz, and then with
w := −πz, we have the two identities eπiz = cos πz + i sin πz, and e−πiz = cos πz − i sin πz.
Subtracting the second identity from the first, we have
1 πiz
e − e−πiz .

sin(πz) =
2i
Now it’s clear that sin(πz) = 0 ⇐⇒ eπiz = e−πiz ⇐⇒ e2πiz = 1 ⇐⇒ z ∈ Z, by
Exercise 2.1.

Exercise 2.16 We will assume, to the contrary, that we only have one arithmetic progression
with a common difference of aN , the largest of the common differences. We hope to obtain
a contradiction.

To each arithmetic progression {ak n + bk | n ∈ Z}, we associate the generating function


X
fk (q) := q ak n+bk ,
ak n+bk ≥0, n∈Z

where |q| < 1, in order to make the series converge. The hypothesis that we have a tiling of
the integers by these N arithmetic progressions translates directly into an identity among
these generating functions:
X X ∞
X
a1 n+b1 aN n+bN
q + ··· + q = qn.
a1 n+b1 ≥0, n∈Z aN n+bN ≥0, n∈Z n=0

Next, we use the fact that we may rewrite each generating function in a ‘closed form’ of the
q bk
following kind, because they are geometric series: fk (q) := ak n+bk ≥0, n∈Z q ak n+bk = 1−q
P
ak .

Thus, we have:
q b1 q bN 1
+ · · · + = .
1 − q a1 1 − q aN 1−q
Now we make a ‘pole-analysis’ by observing that each rational function fk (q) has poles at
2πi
precisely all of the k’th roots of unity. The final idea is that the ‘deepest’ pole, namely e N ,
cannot cancel with any of the other poles. To make this idea precise, we isolate the only
rational function that has this pole (by assumption):
q bN
 b1
q bN −1

1 q
= − + ··· + .
1 − q aN 1−q 1 − q a1 1 − q aN −1
2πi
Finally, we let q → e N , to get a finite number on the right-hand-side, and infinity on the
left-hand-side of the latter identity, a contradiction.

221
Chapter 3
Rb
Exercise 3.1 If ξ = 0, we have 1̂[a,b] (0) := a
e0 dx = b − a. If ξ 6= 0, we can compute the
integral:
Z b
1̂[a,b] (ξ) := e−2πiξx dx
a
e−2πiξb − e−2πiξa
= .
−2πiξ

Exercise 3.2 Beginning with the definition of the Fourier transform of the unit cube [0, 1]d ,
we have:
Z
1̂ (ξ) = e2πihx,ξi dx
Z1 Z 1 Z 1
2πiξ1 x1 2πiξ2 x2
= e dx1 e dx2 · · · e2πiξd xd dxd
0 0 0
d
1 Y e−2πiξk − 1
= ,
(−2πi)d k=1
ξk

valid for all ξ ∈ Rd , except for the finite union of hyperplanes defined by
H := {x ∈ Rd | ξ1 = 0 or ξ2 = 0 . . . or ξd = 0}.

Exercise 3.4 To see that the generating-function definition of the Bernoulli polynomials in
fact gives polynomials, we first write the Taylor series of the following two analytic functions:

t X Bk
t
= tk
e − 1 k=0 k!

xt
X xj tj
e = .
j=0
j!
Multiplying these series together by brute-force gives us:

! ∞ !
j
t X B k k
X x
ext = t tj (11.27)
et − 1 k=0
k! j=0
j!

!
X X Bk xj
= tn (11.28)
n=0 j+k=n
k! j!
∞ n
!
X X Bk xn−k
= tn . (11.29)
n=0 k=0
k! (n − k)!

222
The coefficient of tn on the LHS is by definition n!1 Bn (x), and by uniqueness of Taylor series,
this must also be the coefficient on the RHS, which is seen here to be a polynomial in x. In
fact, we see more, namely that
n
X Bk xn−k
1
Bn (x) = ,
n! k=0
k! (n − k)!
Pn n

which can be written more cleanly as Bn (x) = k=0 k
Bk xn−k .

Exercise 3.5 Commencing with the generating-function definition of the Bernoulli poly-
nomials, equation 3.15, we replace x with 1 − x in order to observe the coefficients Bk (1 − x):

X Bk (1 − x) tet(1−x)
tk =
k=0
k! et − 1
tet e−tx
=
et − 1
te−tx
=
1 − e−t
−te−tx
= −t
e −1

X Bk (x)
= (−t)k ,
k=0
k!

where the last equality follows from the definition of the same generating function, namely
equation 3.15, but with the variable t replaced by −t. Comparing the coefficient of tk on
both sides, we have Bk (1 − x) = (−1)k Bk (x).

223
Exercise 3.6 To show that Bn (x + 1) − Bn (x) = nxn−1 , we play with:
∞ 
tet(x+1) tet(x)

X Bk (x + 1) k Bk (x) k
t − t = t − t
k=0
k! k! e −1 e −1
tetx tet(x)
= et −
et − 1 et − 1
t tetx
= (e − 1) t
e −1
= tetx

X xk k+1
= t
k=0
k!

X xk−1 k
= t
k=1
(k − 1)!

X kxk−1
= tk .
k=1
k!

Therefore, again comparing the coefficients of tk on both sides, we arrive at the required
identity.

d
Exercise 3.7 We need to show that B (x)
dx n
= nBn−1 (x). Well,

X d Bk (x) k d tetx
t =
k=0
dx k! dx et − 1

X Bk (x)
=t tk
k=0
k!

X Bk (x)
= tk+1
k=0
k!

X Bk−1 (x)
= tk
k=1
(k − 1)!

X Bk−1 (x) k
= k t ,
k=1
k!

so that comparing the coefficient of tk on both sides, the proof is complete.


Pn
Exercise 3.30 Considering the partial sum Sn := k=1 ak bk , we know by Abel summation
that
n−1
X
Sn = an Bn + Bk (ak − ak+1 ),
k=1

224
for each n ≥ 2, where Bn := nk=1 bk . By assumption, |Bn | := | nk=1 bk | ≤ M , and the ak ’s
P P
are going to 0, P so we see that the first part of the right-hand-side approaches zero, namely:
|an Bn | := |an || nk+1 bk | → 0, as n → ∞.

Next, we have
n−1
X n−1
X n−1
X n−1
X
| Bk (ak − ak+1 )| ≤ |Bk ||ak − ak+1 | ≤ M |ak − ak+1 | = M (ak − ak+1 ),
k=1 k=1 k=1 k=1

where the last equality holds because by assumption the ak ’s are decreasing. But the last
equals −M an + M a1 , and we have limn→∞ (−M an + M a1 ) = M a1 , a finite limit.
finite sum P
Therefore n−1k=1 Bk (ak − ak+1 ) converges absolutely, and so Sn converges, as desired.

Exercise 3.32 We fix x ∈ R − Z, and let z := e2πix , which lies on the unit circle, and by
assumption z 6= 1. Then

n n z n+1 − 1
≤ 2 ,
X X
2πikx k
e = z = (11.30)

z−1 z−1


k=1 k=1

because |z n+1 − 1| ≤ |z n+1 | + 1 = 2. We also have

|z − 1|2 = |e2πix − 1||e−2πix − 1| = |2 − 2 cos(2πx)| = 4 sin2 (πx),


2 1
so that we have the equality z−1 = sin(πx) . Altogether, we see that

n
X
2πikx 1
e ≤ . (11.31)

| sin(πx)|


k=1

P∞ e2πima
Exercise 3.33 We fix a ∈ R − Z and need to prove that m=1 m
converges. Abel’s
summation formula (3.73) gives us
n n n−1 k
X e2πika 1 X 2πira X  X 2πira  1
= e + e ,
k=1
k n r=1 k=1 r=1
k(k + 1)

so that
∞ ∞ X
k
X e2πika X
2πira
 1
= e .
k=1
k k=1 r=1
k(k + 1)
and the latter series in fact converges absolutely.

Chapter 4
225

Exercise 4.1 The first part follows from the fact that a2 + b2 ≤ |a| + |b|, which is clear by
squaring both sides. For the second part, we use the Cauchy-Schwarz inequality, with the
two vectors x := (a1 , . . . , ad ) and (1, 1, . . . , 1):
q √ √
kxk1 := |a1 | · 1 + · · · + |ad | · 1 ≤ a21 + · · · + a2d 1 + · · · + 1 = d kxk2 ,

which also shows that we obtain equality if and only if (a1 , . . . , ad ) is a scalar multiple of
(1, 1, . . . , 1).

Exercise 4.5 With f (x) := e−2πt|x| , t > 0, we have:


Z
ˆ
f (ξ) := e−2πt|x|−2πixξ dx
Z R0 Z +∞
= e2πtx−2πixξ
dx + e−2πtx−2πixξ dx
−∞ 0
Z 0 Z +∞
= e2πx(t−iξ) dx + e−2πx(t+iξ) dx
−∞ 0
2πx(t−iξ) −2πx(t+iξ)

e 0 e
= +

2π(t − iξ) −∞ −2π(t + iξ) 0

1 1
= +
2π(t − iξ) 2π(t + iξ)
t
= ,
π(t2 + ξ 2 )
valid for all ξ ∈ R.

For the second part, we may use Poisson summation:


X tX 1
e−2πt|n| = .
n∈Z
π ξ∈Z ξ 2 + t2

Chapter 5
Exercise 5.3 This is purely set-theoretic, in the sense that no extra structure is needed for
the set C. If C is symmetric about the origin, then C = −C, and this implies 12 C − 21 C =
1
2
C + 21 (−C) = 12 C + 12 (C) = C. Conversely, suppose that
1 1
C− C=C (11.32)
2 2
Let’s pick any x ∈ C, and now we need to show that −x ∈ C. Since x belongs to the RHS
of (11.32), we know that there must exist y, z ∈ C such that 12 y − 12 z = x. This implies that
−x = 12 z − 12 y ∈ 12 C − 12 C = C.

226
Exercise 5.4 The support of f is supp(f) := {x ∈ Rd | f(x) 6= 0}. So given two convex
bodies A, B ⊂ Rd ,
 Z 

d
supp(1A ∗ 1B ) := y ∈ R 1A (x)1B (y − x)dx 6= 0 .
Rd

But the 1A (x) and 1B (y − x) are always nonnegative. Moreover they are positive ⇐⇒ both
1A (x) = 1 and 1B (y − x) = 1 ⇐⇒ both x ∈ A and y − x ∈ B ⇐⇒ y ∈ B + A, the
Minkowski sum of A and B.

Exercise 5.11 We use the Cauchy-Schwartz inequality:


D E2
( ab ) , ( cos := (a sin x + b cos x)2 ≤ a2 + b2 sin2 x + cos2 x = a2 + b2 .
 
sin x )
x

By the equality condition of Cauchy-Schwartz, we see that the maximum is obtained when
the two vectors are linearly dependent, which gives tan x = ab .

Exercise 5.7 Define ∆ := conv{(0, 0, 0), (1, 1, 0), (1, 0, 1), (0, 1, 1)}, an integer 3-simplex.
It’s clear that ∆ is subset of the unit cube [0, 1]3 , and therefore ∆ has no integer points
in its interior. To see that ∆ is not a unimodular simplex, we can consider its tangent K0
cone at the origin, which has primitive
 integer
 vectors (1, 1, 0), (1, 0, 1), (0, 1, 1), so that the

1 1 0
determinant of K0 is equal to det 1 0 1 = 2 > 1.
0 1 1

Exercise 5.8 Suppose to the contrary, that for some polytope P we have 1̂P (ξ) = g(ξ), a
Schwartz function. Taking the Fourier transform of both sides of the latter equality, and
using the fact that the Fourier transform takes Schwartz functions to Schwartz functions, we
conclude that 1P (−x) = ĝ(−x) is a Schwartz function. But this is a contradiction, because
the indicator function of any polytope is not even continuous.

Chapter 6
Exercise 6.20 It’s easy to see that the inverse matrix for M is
 T
| | ... |
M −1 :=  c11 b1 c12 b2 ... c1d bd  .
| | ... |
The image of the unit sphere under the matrix M is, by definition:
M (S d−1 ) := {u ∈ Rd | u = M x, x ∈ S d−1 }
= {u ∈ Rd | M −1 u ∈ S d−1 }
1 1
= {u ∈ Rd | 2 hb1 , ui2 + · · · + 2 hbd , ui2 = 1},
c1 cd

227
using our description of M −1 above.

For part (b), we begin with the definition of volume, and we understand that we want to
compute the volume of the region M (B d ) := {u ∈ Rd | u = M y, with kyk ≤ 1}.
Z
vol(EllipsoidM ) := du
M (B d )
Z
= | det M | dy
Bd
= | det M | vol(B d ).

using the change of variable u = M y, with y ∈ B d . We also used the Jacobian, which gives
du = | det M |dy.

Finally, we note that the matrix M T M is a diagonal matrix, with diagonal entries c2k , due
to the fact that the bk ’s form an orthonormal basis. Thus we use: | detQM |2 = | det M T M | =
Q d 2 d
k=1 ck , so taking the positive square root, we arrive at | det M | = k=1 ck , because all of
the ck ’s are positive by assumption.

Exercise 6.9 We consider each k’th


√ row of M as a√vector, call it vk . By assumption, the
norm of vk is bounded by kvk ≤ B 2 + · · · B 2 = B d. Using Hadamard’s inequality 6.3,
we have:
 √ d
| det M | ≤ kv1 k · · · kvd k ≤ B d .

Chapter 9
Exercise 9.6 Here P := conv{C, ed }, where C is the (d−1)-dimensional unit cube [0, 1]d−1 .
To compute the Ehrhart polynomial LP (t) here, we use the fact that a ‘horizontal’ slice of
P, meaning a slice parallel to C, and orthogonal to ed , is a dilation of C. Thus, each of these
slices counts the number of points in a k-dilate of C, as k varies from 0 to t + 1. Summing
over these integer dilations of C, we have
t+1 t+1
X
d−1
X 1
LP (t) = (t + 1 − k) = k d−1 = (Bd (t + 2) − Bd ),
k=0 k=0
d

where the last step holds thanks to Exercise 3.8.

228
Bibliography

[1] A. D. Alexandrov, A theorem on convex polyhedra, Trudy Mat. Int. Steklov, Sect.
Math, 4:87, (193 3).
[2] David Austin, Fedorov’s five parallelohedra, Notices of the American Math. Society,
Feature column, 2013.
http://www.ams.org/publicoutreach/feature-column/fc-2013-11
[3] Gennadiy Averkov, Equality Case in Van der Corput’s Inequality and Collisions in
Multiple Lattice Tilings, Discrete & Computational Geometry 65, (2021), 212–226.
[4] László Babai, On Lovfisz’ lattice reduction and the nearest lattice point problem, Com-
binatorica 6, (1986), 1–13.
[5] Velleda Baldoni, Nicole Berline, and Michèle Vergne, Local Euler–Maclaurin expansion
of Barvinok valuations and Ehrhart coefficients of a rational polytope, Integer points
in polyhedra—geometry, number theory, representation theory, algebra, optimization,
statistics, Contemp. Math., vol. 452, Amer. Math. Soc., Providence, RI, 2008, pp. 15–
33.
[6] Keith Ball, A lower bound for the optimal density of lattice packings, International
Mathematics Research Notices, Vol 1992, Issue 10, 1992), 217–221.
[7] W. Banaszczyk, New bounds in some transference theorems in the geometry of numbers,
Math. Ann. 296 (1993), 625–635.
[8] Imre Bárány, Random points and lattice points in convex bodies, Bull. Amer. Math.
Soc. (N.S.) 45 (2008), no. 3, 339–365.
[9] Imre Bárány, Arseniy Akopyan, and Sinai Robins, Algebraic vertices of non-convex
polyhedra, Advances in Math, 308, (2017), 627-644.
[10] Alexander Barvinok, Exponential integrals and sums over convex polyhedra, Funkt-
sional. Anal. i Prilozhen. 26 (1992), no. 2, 64–66.
[11] Alexander Barvinok, A polynomial time algorithm for counting integral points in poly-
hedra when the dimension is fixed, Math. Oper. Res. 19 (1994), no. 4, 769–779.

229
[12] Alexander Barvinok, Integer points in polyhedra, Zurich Lectures in Advanced Math-
ematics, European Mathematical Society (EMS), Zurich, 2008.

[13] Alexander Barvinok and James E. Pommersheim, An algorithmic theory of lattice


points in polyhedra, New Perspectives in Algebraic Combinatorics (Berkeley, CA, 1996–
97), Math. Sci. Res. Inst. Publ., vol. 38, Cambridge Univ. Press, Cambridge, 1999,
pp. 91–147.

[14] Victor V. Batyrev, Dual polyhedra and mirror symmetry for Calabi–Yau hy-
persurfaces in toric varieties, J. Algebraic Geom. 3 (1994), no. 3, 493–535,
arXiv:alg-geom/9310003.

[15] Nicole Berline and Michéle Vergne, Local Euler-Maclaurin formula for polytopes, Mosc.
Math. J., 7 (3) (2007), 355–386.

[16] József Beck, Probabilistic Diophantine approximation, Randomness in lattice point


counting, Springer Monographs in Mathematics, Springer, Cham, (2014), 1–487.

[17] Matthias Beck and Sinai Robins, Computing the continuous discretely: integer-point
enumeration in polyhedra, 2’nd edition, Springer, New York, (2015), 1–285.

[18] Matthias Beck and Sinai Robins, Explicit and efficient formulas for the lattice point
count in rational polygons using Dedekind–Rademacher sums, Discrete Comput. Geom.
27 (2002), no. 4, 443–459, arXiv:math.CO/0111329.

[19] Matthias Beck, Ricardo Diaz, and Sinai Robins, The Frobenius problem, rational
polytopes, and Fourier–Dedekind sums, J. Number Theory 96 (2002), no. 1, 1–21.
arXiv:math.NT/0204035.

[20] Matthias Beck, Sinai Robins, and Steven V Sam, Positivity theorems for solid-angle
polynomials, Beiträge Algebra Geom. 51 (2010), no. 2, 493–507, arXiv:0906.4031.

[21] Christian Bey, Martin Henk, and Jörg M. Wills, Notes on the roots of Ehrhart polyno-
mials, Discrete Comput. Geom. 38 (2007), no. 1, 81–98, arXiv:math.MG/0606089.

[22] H. F. Blichfeldt, The minimum value of quadratic forms, and the closest packing of
spheres, Math. Annalen 101 (1929), 605–608.

[23] Sebastian Böcker and Zsuzsanna Lipták, The money changing problem revisited: com-
puting the Frobenius number in time O(ka1 ), Computing and combinatorics, Lecture
Notes in Comput. Sci., vol. 3595, Springer, Berlin, 2005, 965–974.

[24] Maciej Borodzik, Danny Nguyen, and Sinai Robins, Tiling the integer lattice with
translated sublattices, Moscow Journal of Combinatorics and Number Theory, Vol 6,
issue 4, (2016), 3–26.

230
[25] P. L. Butzer, P. J. S. G. Ferreira, G. Schmeisser and R. L. Stens, The Summation
Formulae of Euler?Maclaurin, Abel?Plana, Poisson, and their Interconnections with
the Approximate Sampling Formula of Signal Analysis, Results. Math. 59 (2011), 359–
400.

[26] Luca Brandolini, Leonardo Colzani, Sinai Robins, and Giancarlo Travaglini, Pick’s
Theorem and Convergence of multiple Fourier series, The American Mathematical
Monthly, 128, Issue 1, (2021), 41–49.

[27] Luca Brandolini, Leonardo Colzani, Sinai Robins, and Giancarlo Travaglini, An Euler-
MacLaurin formula for polygonal sums, to appear in Transactions of the AMS, 2021.

[28] Michel Brion, Points entiers dans les polyèdres convexes, Ann. Sci. Ecole Norm. Sup.
4 21 (1988), no. 4, 653–663.

[29] Michel Brion and Michèle Vergne, Residue formulae, vector partition functions and
lattice points in rational polytopes, J. Amer. Math. Soc. 10 (1997), no. 4, 797–833.

[30] L. Brown, B. Schreiber, and B.A. Taylor, Spectral synthesis and the Pompeiu problem,
Ann. Inst. Fourier, 23 (3) (1973), 125–154.

[31] Richard A. Brualdi and Peter M. Gibson, Convex polyhedra of doubly stochastic ma-
trices. I. Applications of the permanent function, J. Combinatorial Theory Ser. A 22
(1977), no. 2, 194–230.

[32] M. D. Buhmann, Radial functions on compact support, Proceedings of


the Edinburgh Mathematical Society. Series II, 41, (1998), no. 1, 33–46.
https://doi.org/10.1017/S0013091500019416

[33] Kristin A. Camenga, Vector spaces spanned by the angle sums of polytopes, Beiträge
Algebra Geom. 47 (2006), no. 2, 447–462, arXiv:math.MG/0508629.

[34] Leonard Carlitz, The reciprocity theorem for Dedekind-Rademacher sums, Acta Math.
XXIX, (1976), 309–313.

[35] J. W. S. Cassels, An introduction to the geometry of numbers, Corrected reprint of the


1971 edition, Classics in Mathematics, Springer-Verlag, Berlin, (1997), 1–344.

[36] William Y. C. Chen and Peter L. Guo, Equivalence classes of full-dimensional 0/1-
polytopes with many vertices, Discrete and Computational Geometry 52 (2) (2014),
630–662.

[37] Henry Cohn, Abhinav Kumar, Stephen D. Miller, Danylo Radchenko and Maryna Via-
zovska, The sphere packing problem in dimension 24, Annals of Mathematics SECOND
SERIES, Vol. 185, No. 3 (2017), 1017–1033.

231
[38] John H. Conway and Neil J. A. Sloane, Sphere Packings, Lattices and Groups, Third
Edition, with additional contributions by E. Bannai, R. E. Borcherds, J. Leech, S. P.
Norton, A. M. Odlyzko, R. A. Parker, L. Queen and B. B. Venkov Grundlehren der
Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences],
290, Springer-Verlag, New York, (1999), 1–703.

[39] John H. Conway, The sensual quadratic form, The Carus Mathematical Monographs,
published by the MAA, (1997), 1–152.

[40] Dan Cristofaro-Gardiner, Teresa Xueshan Li, and Richard Stanley, New examples of
period collapse, (2015). arXiv:1509.01887v1

[41] Vladimir I. Danilov, The geometry of toric varieties, Uspekhi Mat. Nauk 33 (1978),
85–134, 247.

[42] N. G. De Bruijn, Filling boxes with bricks, Amer. Math. Monthly 76 (1969), 37–40.

[43] Pierre Deligne, Sergei Tabachnikov, and Sinai Robins, The Ice Cube Proof, The Math-
ematical Intelligencer, Vol 36, no. 4, 2014, 1–3.

[44] Jesús A. De Loera, Raymond Hemmecke, and Matthias Köppe, Algebraic and Geomet-
ric Ideas in the Theory of Discrete Optimization, MOS-SIAM Series on Optimization,
vol. 14, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA;
Mathematical Optimization Society, Philadelphia, PA, 2013.

[45] Jesús de Loera, Jörg Rambau, and Francisco Santos, Triangulations. Structures for al-
gorithms and applications, Algorithms and Computation in Mathematics, 25. Springer-
Verlag, Berlin, (2010), 1–535.

[46] David Desario and Sinai Robins , Generalized solid-angle theory for real polytopes, The
Quarterly Journal of Mathematics, 62 (2011), no. 4, 1003–1015, arXiv:0708.0042.

[47] Persi Diaconis and Anil Gangolli, Rectangular arrays with fixed margins, Discrete Prob-
ability and Algorithms (Minneapolis, MN, 1993), Springer, New York, 1995, pp. 15–41.

[48] Ricardo Diaz and Sinai Robins , The Ehrhart polynomial of a lattice polytope, Annals
of Math. (2) 145 (1997), no. 3, 503–518.

[49] Ricardo Diaz, Quang-Nhat Le and Sinai Robins, Fourier transforms of polytopes,
solid angle sums, and discrete volumes, preprint. https://drive.google.com/file/
d/0B223XJaVpyE_MU16UER2VnFQRHc/view

[50] Dutour Sikirić M, Garber A, Schürmann A, Waldmann C, The complete classification


of five-dimensional Dirichlet-Voronoi polyhedra of translational lattices, Acta Crystal-
logr A, Found Adv. Nov 1, (2016) 72 (Pt 6), 673–683.

232
[51] Eugène Ehrhart, Sur les polyèdres rationnels homothétiques à n dimensions, C. R.
Acad. Sci. Paris 254 (1962), 616–618.

[52] Eugène Ehrhart, Sur un problème de géométrie diophantienne linéaire II, J. reine.
angew. Math. 227, (1967), 25–49.

[53] Eugène Ehrhart, Polynômes arithmétiques et méthode des polyèdres en combinatoire,


Birkhäuser Verlag, Basel, 1977, International Series of Numerical Mathematics, Vol.
35.

[54] Charles L. Epstein, Introduction to the mathematics of medical imaging, Society for
Industrial and Applied Mathematics (SIAM), Philadelphia, PA, second edition, 2008.

[55] David Feldman, Jim Propp, and Sinai Robins, Tiling lattices with sublattices I, Discrete
& Computational Geometry, Vol. 46, No. 1, (2011), 184–186.

[56] Benjamin Fischer and Jamie Pommersheim, An algebraic construction of sum-integral


interpolators, preprint, 2021. https://arxiv.org/abs/2101.04845v1

[57] Gerald Folland, Fourier analysis and its applications, Wadsworth & Brooks/Cole Ad-
vanced Books & Software, (1992), 1–433.

[58] Lenny Fukshansky and Stephan Ramon Garcia, Geometric Number Theory, Cambridge
University Press, to appear in 2022.

[59] Lenny Fukshansky and Sinai Robins, Bounds for solid angles of lattices of rank three,
J. Combin. Theory Ser. A 118 (2011), no. 2, 690–701, arXiv:1006.0743.

[60] William Fulton, Introduction to Toric Varieties, Annals of Mathematics Studies, vol.
131, Princeton University Press, Princeton, NJ, 1993.

[61] Stavros Garoufalidis and James Pommersheim, Sum-Integral interpolators and the
Euler-MacLaurin formula for polytopes, Transactions of the AMS, 364, Number 6,
June 2012, 2933–2958

[62] Sigrid Grepstad and Nir Lev, Multi-tiling and Riesz bases, Advances in Mathematics,
252(2014), 1–6.

[63] Ewgenij Gawrilow and Michael Joswig, polymake: a framework for analyzing con-
vex polytopes, Polytopes—combinatorics and computation (Oberwolfach, 1997), DMV
Sem., vol. 29, Birkhäuser, Basel, 2000, pp. 43–73, Software polymake available at
www.polymake.org/.

[64] Nick Gravin, Mihail Kolountzakis, Sinai Robins, and Dmitry Shiryaev, Structure results
for multiple tilings in 3D, Discrete and Computational Geometry, (2013), Vol. 50,
1033–1050.

233
[65] D. Gourion, D. and A. Seeger, Deterministic and stochastic methods for computing
volumetric moduli of convex cones, Comput. Appl. Math. 29, (2010), 215–246.

[66] Peter M. Gruber, Convex and Discrete Geometry, SpringerVerlag, Berlin, 2007, 1–590.

[67] Nick Gravin, Sinai Robins, and Dmitry Shiryaev, Translational tilings by a polytope,
with multiplicity, Combinatorica 32 (2012), no. 6, 629–649, arXiv:1103.3163.

[68] Branko Grünbaum, Convex Polytopes, Graduate Texts in Mathematics, vol. 221,
Springer-Verlag, New York, 2003. Second edition, prepared by V. Kaibel, V. Klee,
and G. M. Ziegler (original edition: Interscience, London, 1967).

[69] Branko Grünbaum, Are your polyhedra the same as my polyhedra?, in Aronov, Boris;
Basu, Saugata; Pach, János; Sharir, Micha (eds.), Discrete and Computational Geom-
etry: The Goodman-Pollack Festschrift, Algorithms and Combinatorics, (2003).

[70] Paul Gunnells and Robert Sczech, Evaluation of Dedekind sums, Eisenstein cocycles,
and special values of L-functions, Duke Math. J. 118 (2003), no. 2, 229–260.

[71] David Alonso-Gutiérrez, C. Hugo Jiménez, Rafael Villa, Brunn–Minkowski and Zhang
inequalities for convolution bodies, Advances in Mathematics, 238 (2013), 50–69.

[72] Godfrey H. Hardy and John E. Littlewood, Some problems of Diophantine approxima-
tion: The lattice-points of a right-angled triangle (Second memoir), Abh. Math. Sem.
Univ. Hamburg, no. 1 (1922), no. 1, 211–248.

[73] Martin Henk, Achill Schürmann, and Jörg M. Wills, Ehrhart polynomials and succes-
sive minima, Mathematika 52 (2005), no. 1–2, 1–16 (2006), arXiv:math.MG/0507528.

[74] Martin Henk, Inequalities between successive minima and intrinsic volumes of a convex
body, Monatsh. Math. 110 (1990) 279–282.

[75] Martin Henk, An introduction to geometry of numbers, lecture notes, preprint.

[76] Alex Iosevich, Nets Katz, and Terence Tao, The Fuglede spectral conjecture holds for
convex planar domains, Mathematical Research Letters 10, (2003) 559–569.

[77] Ravi Kannan, Lattice translates of a polytope and the Frobenius problem, Combinator-
ica 12 (1992), no. 2, 161–177.

[78] Jean-Michel Kantor and Askold G. Khovanskiı̆, Une application du théorème de


Riemann–Roch combinatoire au polynôme d’Ehrhart des polytopes entiers de Rd , C.
R. Acad. Sci. Paris Sér. I Math. 317 (1993), no. 5, 501–507.

[79] Yael Karshon, Shlomo Sternberg, and Jonathan Weitsman, The Euler–Maclaurin for-
mula for simple integral polytopes, Proc. Natl. Acad. Sci. USA 100 (2003), no. 2,
426–433.

234
[80] Yael Karshon, Shlomo Sternberg, and Jonathan Weitsman, Exact Euler-Maclaurin
formulas for simple lattice polytopes, Advances in Applied Mathematics, (2007), Vol
39 (1), 1–50.

[81] Yitzhak Katznelson, An introduction to harmonic analysis, Third edition, Cambridge


Mathematical Library, Cambridge University Press, Cambridge, (2004), 1–314.

[82] Eric Katz and Alan Stapledon, Local h-polynomials, invariants of subdivisions, and
mixed Ehrhart theory, Adv. Math., 286 (2016), 181–239.

[83] Donald Knuth, Notes on generalized Dedekind sums, Acta Arith. (1977), 297–325.

[84] Daniel A. Klain, The Minkowski problem for polytopes, Advances in Mathematics, 185
(2004), no. 2, 270–288.

[85] David V. Feldman and Daniel A. Klain, Angles as probabilities, American Mathemat-
ical Monthly, 116 (2009), no. 8, 732–735.

[86] Alexander Koldobsky, Fourier analysis in convex geometry, Mathematical Surveys and
Monographs, American Mathematical Society, Providence, RI, 1–170.

[87] Mihalis N. Kolountzakis. On the structure of multiple translational tilings by polygonal


regions. Discrete and Computational Geometry, 23 (4), (2000), 537–553.

[88] Jeffrey C. Lagarias and Günter M. Ziegler, Bounds for lattice polytopes containing a
fixed number of interior points in a sublattice, Canadian J. Math. 43, (1991), no. 5,
1022–1035.

[89] Jeffrey C. Lagarias and Chuanming Zong, Mysteries in packing regular tetrahedra,
Notices Amer. Math. Soc. 59 (2012), no. 11, 1540–1549.

[90] Jean B. Lasserre and Eduardo S. Zeron, On counting integral points in a convex rational
polytope, Math. Oper. Res., 28(4) (2003), 853–870.

[91] Jim Lawrence, Polytope volume computation, Math. Comp. 57 (1991), no. 195, 259–
271.

[92] Nir Lev and Bochen Liu, Multi-tiling and equidecomposability of polytopes by lattice
translates, Bulletin of the London Math. Society, Vol 51, issue 6, (2019), 1079–1098.

[93] Nir Lev and Máté Matolcsi, The Fuglede conjecture for convex domains is true in all
dimensions, preprint, 2021.

[94] Lighthill, M. J., Introduction to Fourier analysis and generalised functions, Cambridge
University Press, New York (1960), 1–79.

[95] Eva Linke, Rational Ehrhart quasi-polynomials, J. Combin. Theory Ser. A 118 (2011),
no. 7, 1966–1978, arXiv:1006.5612.

235
[96] Bochen Liu, Periodic structure of translational multi-tilings in the plane, preprint,
2019. https://arxiv.org/abs/1809.03440

[97] Ian G. Macdonald, The volume of a lattice polyhedron, Proc. Cambridge Philos. Soc.,
59 (1963), 719–726.

[98] Ian G. Macdonald, Polynomials associated with finite cell-complexes, J. London Math.
Soc. (2) 4 (1971), 181–192.

[99] Fabricio Caluza Machado and Sinai Robins, The null set of a polytope and the Pompeiu
property for polytopes, preprint, 2021.

[100] Fabricio Caluza Machado and Sinai Robins, Coefficients of the solid angle and Ehrhart
quasi-polynomials, preprint, 2019.

[101] Tyrrell B. McAllister and Kevin M. Woods, The minimum period of the Ehrhart quasi-
polynomial of a rational polytope, Journal of Combinatorial Theory, Series A 109 (2005)
345–352.

[102] Peter McMullen, Lattice invariant valuations on rational polytopes, Arch. Math., 31,
(1978), 509–516.

[103] Peter McMullen, Non-linear angle-sum relations for polyhedral cones and polytopes,
Math. Proc. Cambridge Phil. Soc., 78, (1975), 247–261.

[104] Peter McMullen, Angle-sum relations for polyhedral sets, Mathematika 33 (1986), no.
2, 173–188.

[105] Hermann Minkowski, Geometrie der Zahlen, Teubner, Leipzig, 1896.

[106] Jan Draisma, Tyrrell B. McAllister, and Benjamin Nill, Lattice-Width Directions and
Minkowski’s 3d -Theorem, SIAM Journal on Discrete Mathematics, Vol. 26, No. 3
(2012), 1104–1107.

[107] Robert Morelli, Pick’s theorem and the Todd class of a toric variety, Adv. Math. 100
(1993), no. 2, 183–231.

[108] Marina Nechayeva and Burton Randol, Asymptotics of weighted lattice point counts
inside dilating polygons, Additive number theory, Springer, New York, (2010), 287–
301.

[109] Benjamin Nill and Andreas Paffenholz, On the equality case in Ehrhart’s volume con-
jecture, Adv. Geom. 14 (2014), no. 4, 579–586.

[110] C. D. Olds, Anneli Lax, Giuliana P. Davidoff, Mathematical Association of America,


(2000), 1–193.

236
[111] Osgood, Brad G, Lectures on the Fourier transform and its applications, Pure and
Applied Undergraduate Texts, 33, American Mathematical Society, Providence, RI,
2019. 1–693.

[112] Sam Payne, Ehrhart series and lattice triangulations, Discrete Comput. Geom. 40
(2008), no. 3, 365–376, arXiv:math/0702052.

[113] Micha A. Perles and Geoffrey C. Shephard, Angle sums of convex polytopes, Math.
Scand. 21 (1967), 199–218.

[114] Mark A. Pinsky, Introduction to Fourier Analysis and Wavelets, Brooks/Cole, Pacific
Grove, California, (2002), 1–376.

[115] James E. Pommersheim, Toric varieties, lattice points and Dedekind sums, Math. Ann.
295 (1993), no. 1, 1–24.

[116] Bjorn Poonen and Fernando Rodriguez-Villegas, Lattice polygons and the number 12,
Amer. Math. Monthly 107 (2000), no. 3, 238–250.

[117] Alexander Postnikov, Permutohedra, associahedra, and beyond, Int. Math. Res. Not.
(2009), no. 6, 1026–1106, arXiv:math/0507163.

[118] Quang-Nhat Le, A discrete Stokes formula and the solid-angle sum of polytopes, un-
dergraduate dissertation, (2009).

[119] Quang-Nhat Le and Sinai Robins, Macdonald’s solid-angle sum for real dilations of
rational polygons, preprint.

[120] Srinivasa Ramanujan, Some definite integrals, Messenger of Mathematics 44 (1915),


10–18.

[121] Jorge Luis Ramirez Alfonsin, Complexity of the Frobenius problem, Combinatorica, 16
(1), (1996), 143–147/

[122] Burton Randol, On the Fourier transform of the indicator function of a planar set,
Trans. Amer. Math. Soc., 139 (1969), 271–276.

[123] Burton Randol, On the asymptotic behavior of the Fourier transform of a convex set,
Trans. Amer. Math. Soc., 139 (1969), 279–285.

[124] Burton Randol, On the number of integral lattice-points in dilations of algebraic poly-
hedra, Internat. Math. Res. Notices (1997) no. 6, 259–270.

[125] Bruce Reznick, Lattice point simplices, Discrete Math. 60 (1986), 219–242.

[126] Jason M. Ribando, Measuring solid angles beyond dimension three, Discrete Comput.
Geom. 36 (2006), no. 3, 479–487.

237
[127] C.A. Rogers, Packing and covering, Cambridge Tracts in Mathematics and Mathemat-
ical Physics, No. 54, Cambridge University Press, New York, 1964.

[128] C. A. Rogers and G. C. Shephard, Convex bodies associated with a given convex body,
Journal of the London Math. Soc., 1 (1958), no. 3, 270–281.

[129] Walter Rudin, Real and complex analysis, Third edition, McGraw-Hill Book Co., New
York, (1987), 1–416.

[130] Tiago Royer, Reconstruction of rational polytopes from the real-parameter Ehrhart
function of its translates, preprint, 2017. https://arxiv.org/abs/1712.01973

[131] Tiago Royer, Reconstruction of symmetric convex bodies from Ehrhart-like data,
preprint, 2017. https://arxiv.org/abs/1712.03937

[132] Gervásio Protásio dos Santo Neto, The theory and computation of solid angles, Master’s
thesis, IME, Universidade de São Paulo, (2021), 1–85.

[133] Paul Sally, Fundamentals of Mathematical Analysis, AMS (The Sally series), Pure and
Applied Undergraduate texts (20), (2013), 1–384.

[134] Alexander Schiemann, Ein Beispiel positiv definiter quadratischer Formen der Dimen-
sion 4 mit gleichen Darstellungszahlen, Arch. Math. 54 (1990), 372–375.

[135] Alexander Schiemann, Temare positiv defInite quadratische Fonnen mit gleichen
Darstellungszahlen, Dissertation, Bonn, 1993.

[136] Ludwig Schläfli, Theorie der vielfachen Kontinuität, Ludwig Schläfli, 1814–1895,
Gesammelte Mathematische Abhandlungen, Vol. I, Birkhäuser, Basel, 1950, pp. 167–
387.

[137] Saul Schleimer and Henry Segerman, Puzzling the 120-cell, Notices Amer. Math. Soc.
62 (2015), no. 11, 1309–1316.

[138] Rolf Schneider, Convex Bodies: The Brunn–Minkowski Theory, 2nd edition, Encyclo-
pedia of Mathematics and its Applications, Cambridge University Press, 2013.

[139] Rolf Schneider and Wolfgang Weil, Stochastic and integral geometry, Springer Science
& Business Media, 2008.

[140] Alexander Schrijver, Combinatorial Optimization. Polyhedra and Efficiency. Vol. A–C,
Algorithms and Combinatorics, vol. 24, Springer-Verlag, Berlin, 2003.

[141] Marjorie Senechal and R.V. Galiulin, An Introduction to the Theory of Figures: the
Geometry of E.S. Fedorov, Structural Topology, 10, (1984), 5–22.

238
[142] Jeffrey Shallit, The Frobenius problem and its generalizations, Developments in lan-
guage theory, Lecture Notes in Comput. Sci., vol. 5257, Springer, Berlin, 2008, pp. 72–
83.

[143] Geoffrey C. Shephard, Polytopes with centrally symmetric faces, Canadian J. Math.,
19 (1967), 1206–1213.

[144] Geoffrey C. Shephard, An elementary proof of Gram’s theorem for convex polytopes,
Canad. J. Math. 19 (1967), 1214–1217.

[145] Carl Ludwig Siegel, Lectures on the Geometry of Numbers, Springer-Verlag, Berlin,
1989, Notes by B. Friedman, rewritten by Komaravolu Chandrasekharan with the
assistance of Rudolf Suter, with a preface by Chandrasekharan.

[146] Maxim M. Skriganov, Ergodic theory on homogeneous spaces and the enumeration of
lattice points in polyhedra (Russian), Dokl. Akad. Nauk 355 (1997), no. 5, 609–611.

[147] Maxim M. Skriganov, Ergodic theory on SL(n), Diophantine approximations and


anomalies in the lattice point problem, Invent. Math. 132 (1998), no. 1, 1–72.

[148] Duncan M. Y. Sommerville, The relation connecting the angle-sums and volume of a
polytope in space of n dimensions, Proc. Roy. Soc. London, Ser. A 115 (1927), 103–119.

[149] Richard P. Stanley, Combinatorial reciprocity theorems, Advances in Math. 14 (1974),


194–253.

[150] Richard P. Stanley, Enumerative Combinatorics, Volume 1, Second edition, Cambridge


Studies in Advanced Mathematics, vol. 49, Cambridge University Press, Cambridge,
2012.

[151] Elias Stein and Guido Weiss, Introduction to Fourier analysis on Euclidean spaces,
Princeton University Press, Princeton Mathematical Series, No. 32, Princeton, N.J.,
1971.

[152] Elias Stein and Rami Shakarchi, Fourier analysis, an introduction, Princeton Lectures
in Analysis, 1. Princeton University Press, Princeton, NJ, 2003, 1–311.

[153] Berndt Sturmfels, (1995). On vector partition functions, Journal of Combinatorial The-
ory Series A. 72 (2), 302–309.

[154] E. C. Titchmarsh, Introduction to the Theory of Fourier Integrals, Oxford University


Press, Oxford, 1937.

[155] Giancarlo Travaglini, Number theory, Fourier analysis and geometric discrepancy, Lon-
don Mathematical Society Student Texts, 81. Cambridge University Press, Cambridge
(2014), 1–240.

239
[156] Audrey Terras, Fourier Analysis on Finite Groups and Applications, London Math-
ematical Society, Student Texts, vol. 43, Cambridge University Press, Cambridge,
(1999).

[157] Akshay Venkatesh, A note on sphere packings in high dimension, Int. Math. Res. Not.
IMRN (2013), 1628–1642.

[158] Sven Verdoolaege, Software package barvinok, (2004), electronically available at


http://freshmeat.net/projects/barvinok/.

[159] Stan Wagon, Fourteen Proofs of a result about tiling a rectangle, The American Math-
ematical Monthly 94 (1987), 601–617.

[160] Kevin Woods, The unreasonable ubiquitousness of quasi-polynomials, Electronic Jour-


nal of Combinatorics 21 (1), Paper 1.44, (2014), 1–23.

[161] Qi Yang and Chuanming Zong, Multiple lattice tilings in Euclidean spaces, Canad.
Math. Bull. 62, (2019), no. 4, 923–929.

[162] Stephen T. Yau and Letian Zhang, An upper estimate of integral points in real simplices
with an application to singularity theory, Math. Res. Lett. 13 (2006), no. 6, 911–921.

[163] Günter M. Ziegler, Lectures on polytopes, Graduate Texts in Mathematics, Volume


152, Springer-Verlag, New York, 1995.

[164] Chuanming Zong, The cube - a window to convex and discrete geometry, Cambridge
University Press, (2006), 1–174.

240
Index

24-cell, 120 Dirac, Paul, 67


Dirichlet’s convergence test, 43
Abel summation by parts, 43 discrepancy, 5
Abel, Niels, 43 discrete subgroup, 100
affine span, 208 discrete volume, 5
affine transformation, 30 divergence Theorem, 206
Alexandrov, A. D. , 88 dual lattice, 55, 109
angle polynomial, 157, 160
angle polynomial: functional equation, 167 Ehrhart conjecture, 189
Ehrhart reciprocity, 187
Barvinok, Alexander, 138, 190, 215, 216 Ehrhart theory, 172, 182
basis matrix, 97 ellipsoid, 115
basis-free, 101 Erdös, Paul, 12
Bernoulli number, 38 Euler-Maclaurin summation, 173
Bernoulli polynomial, 21, 39 Euler-Poincare relation, 142
Bessel function, 200 extremal body, 83
Blichfeldt, 122
Brianchon-Gram identity, 142 face poset, 204, 211
Brion, 134 face-numbers, 143
Brion’s theorem - the continuous form, 134 Fedorov solids, 85, 88
Brion’s theorem - the discrete form, 150 Fourier analysis, 44
Brunn-Minkowski inequality, 90 Fourier coefficients, 50
Fourier series, 50
centrally symmetric polytope, 85
Fourier series for periodic functions, 50
Chebyshev polynomials, 71
Fourier transform, 3, 18, 45
combinatorial divergence theorem, 205
Fourier transform of a polytope, 4
complete invariant, 6
fractional part, 21
cone, pointed, 130, 155
Fubini’s theorem, 51
cones, 130
full rank lattice, 97
convex hull, 26
fundamental parallelepiped, 98
convolution, 58
coset, 99 Gauss, 202
Coxeter, 195 Gaussian, 56, 73
cross-polytope, 33, 35, 37, 85, 153, 170 Gaussian smoothing, 160
generic frequencies, 213
De Bruijn, Nicolaas Govert, 9
Gram relations, 165, 166
determinant of a general lattice, 99

241
Hadamard’s inequality, 112 Plancherel Theorem, 60
heat kernel, 160 Poisson summation formula, 4, 6, 23, 52, 53,
55, 56, 58, 66, 69, 72, 75, 77, 78, 81,
indicator function, 4 150, 173, 198, 227
integer lattice, 96, 97 Poisson summation formula for lattices, 56
integer point, 3 polytope, 3
integer point transform, 149 Pompeiu problem, 92
inverse Fourier transform, 18 pyramid formula, 206
Kelvin, William Thomson, 44 Pythagoras, 195
Kepler, Johannes, 202 quasi-coefficients, 214
lattice, 96, 97 quotient lattice, 99
lattice basis, 97 radial function, 201
Lebesgue dominated convergence theorem, 52 ramp function, 73
linear span, 209 rational point, 3
Möbius µ-function, 15 rational polytope, 173
machine learning, 73 Riemann-Lebesgue lemma, 63
Mahler, Kurt, 121 root of unity, primitive, 15
Minkowski, 74 Schwartz function, 51
Minkowski problem for polytopes, 216 Schwartz space, 50
Minkowski sum, 76, 85, 93, 155, 199, 228 Schwartz, Laurent, 67
Minkowski’s 3d theorem, 89 Shannon, Claude, 3
Minkowski, Hermann, 74 Shephard, 88
moment formulas, 147 shortest nonzero vector in a lattice, 112
moments, 147 Siegel’s formula, 74, 77
nice function, 50 simple polytope, 129, 147
Niels Bohr, 172 simplex, 27
Sinc function, 18
octahedron, 33 sinc function, 19
2
orthogonality of exponentials in L ([0, 1]), 12 slack variable, 189
orthogonality relations for lattices, 125 solid angle, 157–159
orthogonality relations for the exponentials solid angle of a face, 165
en (x), 48 sphere packings, 195
orthogonality, roots of unity, 14 standard simplex, 25, 34, 166, 183, 188, 189,
193, 194
parallelepiped, 30, 42 Stoke’s formula, 204
Parseval’s identity, 61 stretch lemma, 29
Parseval’s Theorem, 60 sublattice, 99
periodic Bernoulli polynomial, 21 symmetrized body, 76, 77, 90
periodicity, 97
Pick’s formula, 164 tangent cone, 133
Pick’s formula, generalization, 192 tangent cones, 133

242
theta function, 57
tiling, 9, 10, 54, 80, 83, 85, 92, 220, 222
Tiling a rectangle, 8
translate lemma, 29
triangle inequality for integrals, 48

unimodular group, 108, 109


unimodular matrix, 108

vertex description of a polytope, 27


volume, 5, 147
volume of a simple polytope, 147
volume of a spherical polytope, 158
volumetric moduli, 158
Voronoi conjecture, 120

zonotope, 85

243

You might also like