Ergodic Theory Number Theory
Ergodic Theory Number Theory
Abstract. The main goal of this survey is the description of the fruitful
interaction between Ergodic Theory and Number Theory via the study of
two beautiful results: the first one by Ben Green and Terence Tao (about
long arithmetic progressions of primes) and the second one by √ Noam Elkies
and Curtis McMullen (about the distribution of the sequence { n} mod 1).
More precisely, during the first part, we will see how the ergodic-theoretical
ideas of Furstenberg about the famous Szemerédi theorem were greatly
generalized by Green and Tao in order to solve the classical problem of
finding arbitrarily long arithmetical progression of prime numbers, while
the second part will focus on how Elkies and McMullen used the ideas
of Ratner’s theory (about the classification of ergodic measures related to
unipotent
√ dynamics) to compute explicitly the distribution of the sequence
{ n} on the unit circle.
The initial plan of this book was to cover the Green-Tao, Elkies-McMul-
len and Einsiedler-Katok-Lindenstrauss theorems. However, due to the
usual problem of limitation of space and time, we were forced to make a
choice between these three beautiful results. Because Einsiedler-Katok-
Lindenstrauss theorem is a little bit more difficult to explain from the
technical point of view (in the author’s opinion), we have chosen to discuss
Green-Tao and Elkies-McMullen theorems.
More precisely, this book has two parts: the first part, by A. Arbieto, C.
G. Moreira and C. Matheus, consists of the first two chapters and concerns
Green and Tao theorem, while the last one, by C. Matheus, consists of the
third chapter and concerns Elkies and McMullen theorem. The resume of
the contents of these chapters is:
• in the first chapter (part I), we will make a historical review of the
basic questions and theorems about the additive and multiplicative
properties of integer numbers. In particular, we are going to see that
there are several problems about the additive properties of prime
numbers which are very easy to state but very difficult to solve: e.g.,
it is not known whether there are infinitely many pairs of twin prime
numbers, i.e., pairs of prime numbers whose difference is 2 (this is
known as the Twin Prime Conjecture) and it is still open the so-
called Goldbach conjecture saying that every even natural number
≥ 4 is the sum of two prime numbers. Also, we will see that another
classical conjecture (solved by Ben Green and Terence Tao) says
that there are arbitrarily large arithmetical progressions of prime
numbers. The biggest known progression of primes (to the best of
the authors’ knowledge) is
Ben Green. The basic plan of the first chapter is to discuss some
“preparatory” results in the direction of Green and Tao theorem such
as Szemerédi theorem and its ergodic-theoretical proof by Fursten-
berg.
• finally, in the last chapter (part II), we will completely change the
subject from the additive properties of prime numbers
√ to the Elkies-
McMullen calculation of the distribution law of n (mod 1). In
particular, we subdivide this chapter into three sections: the first
two concerns√the translation of the problem of computing the dis-
tribution of n (mod 1) into an ergodic-theoretical problem and
the last section concerns the solution of the corresponding ergodic
problem via Ratner theory of homogenous flows.
Evidently, as the reader can infer from this summary, the parts I and
II are completely independent, so that the reader can chose where he/she
wants to start reading the book.
Finally, we would like to apologize for the omission of Einsiedler-Katok-
Lindestrauss theorem: as a form of compensation, C. Matheus would
like to say that he’s planning to include some notes about Einsiedler-
Katok-Lindenstrauss theorem in his mathematical blog “Disquisitiones
Mathematicae” (http://matheuscmss.wordpress.com/) in a near future.
Acknowledgments
The first two chapters are strongly based on our book [1] “Aspectos
Ergódicos da Teoria dos Números” (in Portuguese) for a mini-course of
the 26th Brazilian Mathematical Colloquium, which was, in turn, based
on our discussions with Jimmy Santamaria in an Ergodic-Theory seminar
at IMPA around Green-Tao theorem. In particular, we would like to thank
Jimmy Santamaria for his decisive participation in this seminar. Also, we
would like to thank Yuri Lima and Jimmy Santamaria for a careful revision
of previous versions of this work.
Furthermore, this book was written during the pos-doctoral stay of
Carlos Matheus at Collège de France (under the supervision of Jean-
Christophe Yoccoz). In particular, he is thankful to Jean-Christophe for
the several interesting mathematical (and non-mathematical) discussions,
advices and friendship, and the two secretaries of the Mathematics section
of College de France, Dominique Bidois and Veronique Sainz, for their
invaluable support and friendship. Moreover, Carlos Matheus is grateful
to his beloved wife Aline Gomes Cerqueira for her patience and support
(both mathematical and non-mathematical).
Last, but not least, it is a great pleasure to acknowledge Étienne Ghys,
Maria Eulália Vares and Vladas Sidoravicius, current editors of “Ensaios
Matemáticos”, for their invitation to start this project (after our mini-
course during the 26th Brazilian Math. Colloquium). In particular, we
are grateful to Maria Eulália Vares whose strong encouragement and kind
words kept us working on this project.
2 Green-Tao-Szemerédi theorem 38
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
9
10 CONTENTS
3 Elkies-McMullen theorem 75
3.1 Distribution of sequences on the circle . . . . . . . . . . . . 75
3.2 Ergodic version of Elkies-McMullen theorem . . . . . . . . . 78
3.2.1 Scheme of the proof of Elkies and McMullen theorem 78
3.2.2 Some preliminary reductions . . . . . . . . . . . . . 80
3.2.3 Geometrical interpretation of Le N . . . . . . . . . . . 83
3.3 Study of L via Ratner theorems . . . . . . . . . . . . . . . . 85
3.3.1 Computation of F assuming Theorem 3.3.1 . . . . . 85
3.3.2 Theorem 3.3.1 and homogenous flows . . . . . . . . 91
3.4 Equidistribution of non-linear horocycles . . . . . . . . . . . 92
3.4.1 The distribution law of a loop . . . . . . . . . . . . . 95
3.4.2 Ratner theorem and classification of µ . . . . . . . . 97
3.4.3 Non-linearity and torsion points . . . . . . . . . . . 100
3.4.4 End of the proof of Theorem 3.4.3 . . . . . . . . . . 102
Bibliography 103
Part I
Green-Tao theorem
Chapter 1
Additive properties of
prime numbers
1.1 Introduction
One of the oldest concepts in Mathematics is the the notion of prime
number. By definition, an integer number p is prime if it is divisible only
by 1 and by itself.
Prime numbers are important objects in Number Theory due the unique
factorization theorem (saying that any integer number can be written as
a product of prime numbers in an essentially unique way). Another ele-
mentary property of prime numbers is the fact that they are precisely the
integer numbers p such that Z/pZ is a field.
Obviously, due to the multiplicative character of the definition of prime
numbers, it is fairly easy to extract its multiplicative properties. For in-
stance, the product of two primes is certainly not a prime number and
there are no geometrical progressions of primes of length ≥ 3 formed only
by prime numbers.
On the other hand, the situation changes considerably when one poses
some questions about the additive character of the prime numbers. For
example, one can ask whether the sum of two prime numbers is still a
prime number. Of course, the answer is: it depends. In fact, 2+3=5 is
prime and 2+5=7 is prime, but neither 3+5=8 isn’t prime nor 7+2=9.
However, the Bertrand’s postulate says that, for every natural number N ,
there exists a prime number between N and 2N = N + N . In particular,
this shows that the following question deserves a little bit of attention:
Are there arithmetical progressions of length ≥ 3 formed only by prime
numbers? In the case of an affirmative answer to this question, how many
of them exist once the length of the arithmetical progression is fixed?
13
14 Chapter 1. Additive properties of prime numbers
We are going to see (in the first two chapters of this book) that this prob-
lem was solved by Ben Green and Terence Tao. However, before entering
this issue, let us take a little trip around the world of the prime numbers in
order to see some related questions about the additive properties of prime
numbers and its partial solutions.
An important result due to Brun [2] gives a flavor of the difficulty of this
conjecture: namely, Brun proved that, even in the case of the existence of
infinitely many pairs of twin primes, it is a very hard task to locate them
because they are very rare. More precisely, Brun’s theorem says that the
series of the inverse of the twin primes converges (to a certain number
called Brun’s constant):
1 1 1 1 1 1 1 1
( + ) + ( + ) + ( + ) + ( + ) + ... < +∞.
3 5 5 7 11 13 17 19
Later on, we will reformulate this conjecture in a more analytical language.
In particular, if we take two colors and we give one color to the set of
prime numbers P and the other color to the composite (i.e., non-prime)
numbers, we obtain:
Later in this chapter, we will see two proofs of Van der Waerden theorem
(one of them is combinatorial and the other is ergodic-theoretical).
|[1, N ] ∩ A|
d(A) = lim sup .
N →∞ N
|P ∩ [1, N ]| 1
= (1 + o(1)).
N log N
Although the prime numbers form a subset of zero density, the existence
of infinitely many arithmetical progressions of length 3 formed only by
prime numbers was showed in 1939 by Van der Corput (before Roth’s
theorem):
Finally, in 2004, Ben Green and Terence Tao [9] proved the general result
(about arbitrarily large arithmetic progressions formed only by primes).
This theorem is the main object of the first two chapters of this book:
Theorem 1.3.3 (Green-Tao). The primes contains arbitrarily large arith-
metical progressions.
This means that the prime numbers are less sparse than the perfect squares.
The Erdös-Turán conjecture claims that any set with this property con-
tains arbitrarily large arithmetical progressions (so that Green-Tao theo-
rem solves a particular case of this conjecture):
Conjecture 1 (Erdös-Turán). Let A ⊂ N be a subset such that
X 1
= +∞.
n
n∈A
Thus, one can suppose that we find a polychromatic fan inside {bkN1 +
1, . . . , bkN1 + N1 } for every b = 1, . . . , N2 . In other words, for each b =
1, . . . , N2 , we have a(b), r1 (b), . . . , rd−1 (b) ∈ {1, . . . , N1 } and distinct colors
c0 (b), c1 (b), . . . , cd−1 (b) ∈ {1, . . . m} such that c(bkN1 + a(b)) = c0 (b) and
c(bkN1 +a(b)+jri (b)) = ci (b) for every j = 1, . . . , k−1 and i = 1, . . . , d−1.
We say that these are the first and second properties of the fan associated
to b. In particular, the map
is a coloring with md N1d colors of the set {1, . . . , N2 }. Using again our in-
ductive hypothesis on k, if N2 is sufficiently large, there exists some arith-
metic progression b+[0, k−1)s which is monochromatic with respect to this
new coloring, say that its color has the form (a, r1 , . . . , rd−1 , c1 , . . . , cd−1 ).
Up to reversing the position of the progression, we can suppose that s is
negative.
At this point, the idea is to convert this huge progression of identical
polychromatic fans of degree d − 1 (in the sense that their combinatorial
type is fixed by the coloring (a, r1 , . . . , rd−1 , c1 , . . . , cd−1 )) in a new poly-
chromatic fan with degree d in order to close the inductive argument. Let
b0 = (b − s)kN1 + a ∈ {1, . . . , N } and consider:
If the base point b0 has the same color of a spoke, we found a monochro-
matic arithmetic progression of length k. Otherwise, the base point has
a distinct color from the spokes, so that we found a polychromatic fan of
radius k and degree d. This ends the inductive step and, a fortiori, the
proof of the theorem.
will introduce the precise definition of the shift map and we will see how
this important tool was applied by Furstenberg and Weiss to give a proof
of van der Waerden theorem.
Let A = {a1 , . . . , ak } be a finite alphabet. Consider the set Ω of all
infinite words obtained from the letters of this alphabet:
Ω = {(x1 , x2 , . . . , xn , . . . ) ; xi ∈ A, ∀ i}.
This set has a natural structure of metric space with respect to the follow-
ing distance: given x = (x1 , x2 , . . . ) and y = (y1 , y2 , . . . ), define
1
d(x, y) := if l is the smallest integer such that xl 6= yl .
l
The shift map T : Ω → Ω is:
T (x1 , x2 , x3 , . . . ) = (x2 , x3 , x4 , . . . ).
It is a simple exercise to show that the shift map is continuous with respect
to the distance d.
From these concepts, Furstenberg proved the van de Waerden theorem
via the following topological dynamical theorem (whose complete proof is
presented in the appendix to this chapter):
Theorem 1.4.3 (Topological Multiple Recurrence - Furstenberg and Weiss).
Let T : X → X be a continuous dynamical system on a compact metric
space X. For all k ∈ N and ε > 0, there exist x ∈ X and n ∈ N such that
d(T in (x), x) < ε for every i = 1, . . . , k. Moreover, given any dense subset
Z ⊂ X, we can take x ∈ Z.
Assuming this result, let us see how one can prove van der Waerden
theorem. Let A = {c1 , . . . , cs } be the set of colors and z = (z1 , z2 , z3 , . . . ) ∈
AN a given coloring of N, where zi ∈ A is the color of the integer i. Consider
T : AN → AN the shift map. From the definition of the distance d, we have
that, for x, y ∈ AN and m, l ∈ N, it holds d(T m (x), T l (y)) < 1 if and only
if xm+1 = yl+1 .
In particular, for a given coloring z ∈ AN , an arithmetic progression
m, m + n, . . . , m + kn is monochromatic if and only if zm = zm+n = · · · =
zm+kn , that is, if and only if:
d(T m−1 (z), T m−1+in (z)) = d(T m−1 (z), T in (T m−1 (z)))
< 1, for i = 1, . . . , k.
Z Z
f dµk → f dµ.
X X
k−1
1X i ∗
µk = (T ) η.
k i=0
T ∗ µ = T ∗ (lim µnk )
= lim(T ∗ (µnk ))
nk −1
1 X
= lim( (T i+1 )∗ (η))
nk i=0
nk
1 X
= lim( ( (T i )∗ (η) − η + (T nk )∗ η))
nk i=0
nk
1 X
= lim (T i )∗ (η)
nk i=0
= µ.
In the second equality, we used the fact that the push-forward opera-
tor T ∗ is continuous in the weak-* topology. This is true because T is
continuous: indeed, suppose that µk → µ weakly-* and fix a continuous
f : X → R, so that we have that f ◦ T is also continuous and, a fortiori,
Z Z Z Z
∗
f d(T µk ) = f ◦ T dµk → f ◦ T dµ = f d(T ∗ µ).
X X X X
such that, if |y| < δ, then µ(A ∩ (A − y)) > µ(A) − ε. Hence
Pk−1
µk = k1 j=0 δT j (x) . Then, as we already know, up to passing to a sub-
sequence, we can assume that µ = lim µk is a T -invariant probability
measure.
Define Y = {(yn ); y1 = 1}. Since Y is a compact subset, we have
that µ(Y ) = lim µk (Y ) = lim k1 |A ∩ [1, k]| > 0 (by hypothesis). Thus, by
Furstenberg multiple recurrence theorem, it follows that there exists N
such that µ(Y ∩ T −N (Y ) ∩ · · · ∩ T −(k−1)N (Y )) > 0. In particular, there
is z ∈ Y ∩ T −N (Y ) ∩ · · · ∩ T −(k−1)N (Y ). I.e., there exists some integer x
such that x, x + N, . . . , x + (k − 1)N ∈ A. This ends the proof of Szemerédi
theorem.
For any k ≥ 1 and 0 < δ ≤ 1, there exists a large integer NSZ (k, δ) ≥ 1
such that, for every N ≥ NSZ , any subset A ⊂ {1, . . . , N } of cardinality
|A| ≥ δN contains some arithmetic progression of length k.
Proof. Let m0 = NSZ (k, δ/2). Then, for every m ≥ m0 , any subset of
{1, 2, . . . , m} with cardinality ≥ δm/2 contains some arithmetic progres-
sion of length k. Let N be a large integer. For each 1 ≤ r ≤ ⌊N/m0 ⌋,
we divide {1, 2, . . . , N } into r arithmetic progressions of ratio r, e.g.,
{1 ≤ n ≤ N | n ≡ a(mod r)}, for each a with 0 ≤ a ≤ r − 1. Each
of these APs has ⌊N/r⌋ elements (at least), and, therefore, they can be
decomposed into a union of ⌊⌊N/r⌋/m0 ⌋ disjoint arithmetic progressions of
ratios r, lengths ≥ m0 (and almost equal), so that their diameters belong
to the interval [r(m0 − 1), r(2m0 − 1)].
Now, if A ⊂ {1, 2, . . . , N } satisfies |A| ≥ δN , we have that, for each r,
3δ
#{0 ≤ a ≤ r − 1 | #A ∩ {1 ≤ n ≤ N | n ≡ a(mod r)} ≥ ⌊N/r⌋} ≥
4
δr δ 3δ
(since t < ⇒ t + (1 − t) < δ).
4 − 3δ 4 − 3δ 4
δ δ 3δ
On the other hand, because t < ⇒ t+ (1−t) < , if #A∩{1 ≤
4 − 2δ 2 4
3δ
n ≤ N | A ≡ a(mod r)} ≥ ⌊N/r⌋ for a certain r, then, it follows
4
δ
that · ⌊⌊N/r⌋/m0 ⌋ of the created arithmetic progressions of length
4 − 2δ
≥ m0 (at least) intersects A with a relative proportion of δ/2 (at least),
⌊N/m
P 0 ⌋ δr
so that a fortiori, it should contain a k-AP. This gives us ·
r=1 4 − 3δ
δ
⌊⌊N/r⌋/m0 ⌋ > β(δ, m0 )N 2 k-APs (at least) contained in A (for a
4 − 2δ
large N ), where β(δ, m0 ) = δ 2 /64m20 . Of course, some of these APs can
be double-counted sometimes (for different choices of r), but once we fix
d
the diameter d of the AP, r must be a divisor of d between and
2m0 − 1
d d
, i.e., r = ′ , where k − 1 ≤ r′ ≤ 2m0 − 1. Consequently, there are
k−1 r
2m0 − k + 1 possibilities for r′ (at most) and a fortiori for r, so that each
AP is counted 2m0 − k + 1 times (at most). Hence, A contains at least
δ2
α(k, δ)N 2 k-APs, where α(k, δ) = ·
64m20 (2m0 − k + 1)
Remark 1.6.2. The basic difference between the proof of the quantitative
Szemerédi theorem and its previous versions is the finitary nature of the
arguments (allowing to explicit bounds on NSZ ). The proofs of the other
versions are infinitary arguments (they use to some extend the Axiom of
Choice) only permits us to show the existence of NSZ without any bound
on its magnitude. The strategy of the proof of the quantitative Szemerédi
is used during the proof of Green-Tao theorem as we are going to see in
the next chapter (we also recommend [16]).
A. Arbieto, C. Matheus and C. G. Moreira 31
1 X
E(f (n)|n ∈ A) = E(f |A) = f (n).
|A|
n∈A
E(Λ|[1, N ]) = 1 + o(1).
Dividing by N , we get that the prime number theorem implies the state-
ment about the expectations of Λ.
32 Chapter 1. Additive properties of prime numbers
Thereom 1.7.1.
3 This famous conjecture, one of the seven Millennium prize problems of Clay Mathe-
matical Institute (who offers 1 million dollars for its solution), says that the (non-trivial)
zeros of Riemann zeta function ζ(s) (a complex-analytic function related to the prime
∞
P
numbers obtained by analytic continuation of 1/ns , ℜ(s) > 1) are located in the
n=1
line ℜ(s) = 1/2.
A. Arbieto, C. Matheus and C. G. Moreira 33
Let (vj )j∈J be an arbitrary shape of Gaussian integers. Then, the set of
Gaussian primes contains infinitely many constellations with this
prescribed shape.
Remark 1.7.1. Just to stress the difficulty of the twin prime conjecture,
let us observe that this conjecture is much stronger than the result ∆ = 0
of Goldston, Pintz and Yıldırım (whose proof is highly non-trivial!).
At this stage, the proof proceeds by induction (on k). Suppose that the
theorem holds for some k ≥ 1, i.e., for all ε > 0 there exists x ∈ X and
n ∈ N such that d(T in (x), x) < ε for each i = 1, . . . , k. We claim that the
set of such points x is actually dense in X.
Indeed, let U ⊂ X be an arbitrary open subset and pick B ⊂ U a small
ball of radius strictly less than ε. Define Bm = (T m )−1 (B), so that these
subsets form an open cover of X (by the minimality assumption). Using
the compactness of X, we can extract a finite subcover {Bm1 , . . . , Bmr }.
Let δ > 0 be the Lebesgue number of this open subcover, that is, a number
such that any ball of radius δ is contained inside some element of the
subcover. Take x and n such that d(T in (x), x) < δ for i = 1, . . . , k (whose
existence is assured by the inductive hypothesis) and denote by D the ball
of center x and radius δ. Then, by our choice of δ, there exists j such that
D ⊂ Bmj . In particular, T mj (D) ⊂ B, that is, the elements T mj (T in (x))
belong to the ball of radius ε centered on T mj (x) ∈ U . This proves our
denseness claim.
Now, let’s go back to the proof of the theorem. Fix ε > 0. By the induc-
tive hypothesis, there are x0 and n0 such that d(T in0 x0 , x0 ) < ε/2 for i =
1, . . . , k. Taking x1 such that T n0 (x1 ) = x0 , we have d(T (i+1)n0 x1 , x0 ) <
ε/2 for i = 1, . . . , k. Hence, d(T in0 (x1 ), x0 ) < ε/2 for i = 1, . . . , k + 1.
By continuity, there exists ε1 < ε such that d(y, x1 ) < ε1 implies
d(T in0 (y), x0 ) < ε/2 for i = 1, . . . , k + 1. By our denseness claim, there
are y1 and n1 such that d(y1 , x1 ) < ε1 /2 and d(T in1 (y1 ), y1 ) < ε1 /2 for
i = 1, . . . , k. By the triangular inequality, we have:
By compactness, there are l > m such that d(xl , xm ) < ε/2. By the
triangular inequality, we have:
and √
kxk + kx′ k = 2 k.
Thus, since the equality in the triangular inequality kx + x′ k ≤ kxk +
kx′ k can only occur when the vectors (a1 , . . . , an ) and (a′1 , . . . , a′n ) are
proportional, we see that x = x′ = x′′ (because the norms of these vectors
are the same by hypothesis).
On the other hand, there are dn vectors (a1 , . . . , an ) satisfying the con-
straint 0 ≤ ai < d and there are n(d − 1)2 + 1 possible values of k. Con-
sequently, for some k = K0 , the subset Sk (n, d) must have cardinality (at
least)
dn dn−2
> ·
n(d − 1)2 + 1 n
Since every element of Sk (n, d) has modulus < (2d − 1)n , if we define
it holds
ν((2d − 1)n ) > dn−2 /n.
so that
dn−2 (N 1/n − 1)n−2
ν(N ) ≥ ν((2d − 1)n ) > >
n 2n−2 n
N 1−(2/n)
= (1 − N −1/n )n−2
2n−2 n
N 1−(2/n) log n (n−1) log 2
> n−1
= N 1−(2/n)− log N − log N
2 √n
1− 2 2 log 2+ε
√
>N log N .
and define Ak (n, d) := φ−1 (Sk (n, d)). As we already know, Sk (n, d) is
free from 3-APs (except for the trivial 3-APs {x, x, x}). This implies that
Ak (n, d) can only contain arithmetic progressions (n, n + r, n + 2r) when r
is a multiple of (2d − 1)n . In particular, the maximal number of 3-APs in
Ak (n, d) is N 2 /(2d − 1)n . On the other hand, when φ(x) ∈ {0, . . . , d − 1}n ,
1
the probability of x to belong to Ak (n, d) is n(d−1) 2 +1 . Therefore, we get
For a brief exposition of M. Elkin estimate see the paper [10] of B. Green
and J. Wolf.
Chapter 2
Green-Tao-Szemerédi
theorem
2.1 Introduction
The main goal of this chapter is the discussion of some of the ideas behind
the proof of Green-Tao theorem.
Roughly speaking, the proof has two main steps:
Once these two facts are established, we will see that the Green-Tao
theorem follows immediately (see the section 2.2).
However, before entering (in section 2.4) into the details of the previous
outline, we will try to motivate the concepts and tactics of the proof of
the Green-Tao-Szemerédi theorem via a complete proof of Roth theorem
(which corresponds to the particular case k = 3 of Szemerédi theorem) in
the section 2.3, while (unfortunately) we will skip completely the discussion
of the results related to the work of Goldston-Yıldırım.
The organization of this chapter is:
38
A. Arbieto, C. Matheus and C. G. Moreira 39
Closing this brief introduction, let us remark that when we reach the
end of this chapter, the Green-Tao theorem will be proved except for the
results of Goldston and Yıldırım whose beautiful proof will be omitted
because it is purely number-theoretical (and, thus, outside the scope of
this “ergodic-theoretical” book).
• this reduces our task to show that the set of prime numbers has a
pseudorandom behaviour; this fact follows more or less directly from
the works of Goldston and Yıldırım.
In the sequel, we give the details for these items. To do so, we start
with the definition of pseudorandomness:
Now we have a function still seeing (some) prime numbers such that it
doesn’t see neither the powers of primes nor the annoying non-uniformity
2 Basically this occurs because the prime numbers (and the von Mangoldt function)
are concentrated, for any q > 1 integer, on the φ(q) residual classes a(mod q) with
(a, q) = 1 (where φ(q) is the Euler totient function), while any pseudorandom measure
must be equidistributed along all congruence classes modulo q; since the quocient φ(q)/q
can be made arbitrarily small, we see that the von Mangoldt function doesn’t have
pseudorandom majorants.
42 Chapter 2. Green-Tao-Szemerédi theorem
the proof of Green-Tao theorem) that it suffices to take w(n) a very large constant
depending only on k (but not on N ).
A. Arbieto, C. Matheus and C. G. Moreira 43
it holds
Λ3 (f, f, f ) ≥ c(3, δ) − oδ (1).
In other words, we want some lower bounds on Λ3 (f, f, f ). We begin
with the simple remark that it is fairly easy to get upper bounds: e.g., by
Young inequality,
if 1 ≤ p, q, r ≤ ∞ and p1 + 1q + 1r ≤ 2.
On the other hand, we are only interested in lower bounds for Λ3 and,
a priori, upper bounds are not useful for our interests. However, we can
decompose f into a “good” part g = E(f ) and a“bad” part b = f − E(f ).
Using the multilinearity of Λ3 , we can use this decomposition to split up
Λ3 (f, f, f ) into eight pieces:
(a non-realistic case). But, we can refine this upper bound argument via
Harmonic Analysis, or, more precisely the Fourier transform:
we obtain that
X
Λ3 (f, g, h) = fb(ξ1 )b
g (ξ2 )b
h(ξ3 )×
ξ1 ,ξ2 ,ξ3
From Plancherel formula kf kL2 (ZN ) = kfbkl2 (ZN ) and Hölder inequality,
it follows that
g kl4 (ZN ) kb
|Λ3 (f, g, h)| ≤ kf kL2 (ZN ) kb hkl4 (ZN ) . (2.1)
and
kgkL1 (ZN ) , kbkL1 (ZN ) = O(δ).
Then
Λ3 (f, f, f ) = Λ3 (g, g, g) + O(δ 5/4 kbbkl4 (ZN ) )
and
Λ3 (f, f, f ) = Λ3 (g, g, g) + O(δkbbkl∞ (ZN ) ).
Proof. By hypothesis, kgkL2 (ZN ) , kbkL2 (ZN ) = O(δ 1/2 ), so that Plancherel
theorem says that
Logically, one should work in details this idea in order to see that it leads to
a proof of Roth theorem. In this direction, let us introduce the definition:
Here, it is fundamental to stress out that the constant c(ε, K) > 0 doesn’t
depend on N . On the other hand, by considering the shift dynamics
T (x) := x + 1 and by fixing r such that krξj k ≤ ε (where 1 ≤ j ≤ K), we
get:
kfqp ◦ T r − fqp kL2 (ZN ) ≤ C(K)ε.
Combining this information with the triangular inequality, it follows:
E(f · (f ◦ T r ) · (f ◦ T 2r )) ≥ δ 3 /4.
A. Arbieto, C. Matheus and C. G. Moreira 47
Lemma 2.3.3. Let b be a bounded function with kbbkl∞ (ZN ) ≥ σ > 0 and
0 < ε ≪ σ. Then, there exists a function of the form χ(x) = exp(2πixξ/N )
such that the associated sigma-algebra Bε,χ satisfies
This shows that kE(b|Bε,χ )kL2 (ZN ) ≥ σ − Cε ≥ σ/2, so that the proof of
the lemma is complete.
The last ingredient in the proof of Roth theorem is the following struc-
ture proposition:
A. Arbieto, C. Matheus and C. G. Moreira 49
For the proof of this proposition, we will employ the following algorithm:
• Stage 0 : We start with B and Be equal to the trivial sigma-algebra
{0, ZN }. Note that the inequality (2.4) is automatically satisfied at
this stage.
• Stage 1 : Consider B a sigma-algebra of the form Bε1 ,χ1 ∨ · · · ∨
Bεn ,χn , where χj (x) = exp(2πixξj /N ). Since the function E(f |B)
is bounded and B-measurable, the corollary 2.3.1 says that one can
find K depending on δ, n, ε1 , . . . , εn such that E(f |B) is (σ/2, K)-
quasi-periodic.
e If kbbkl∞ (Z ) ≤ F (δ, K),
e and b = f −E(f |B).
• Stage 2 : Put g = E(f |B) N
we end the algorithm. Otherwise, we go to the third stage.
• Stage 3 : Since we didn’t end the algorithm at the stage 2, we have
kbbkl∞ > F (δ, K). By Lemma 2.3.3, we can find ε ≪ F (δ, K) and
a function χ of the form χ(x) = exp(2πixξ/N ) whose associated
sigma-algebra Bε,χ verifies
Combining this inequality with (2.3) and the proposition 2.3.1, we have
Taking F “sufficiently small”, we can absorb the second term of the right-
hand side via the first term, so that
• 1st step: to define a class of norms (Gowers norms k.kU k−1 ) in order
to control the expectation of a k-AP to reside in the support of f ;
observe that, by proposition 2.3.1, in the particular case k = 3, the
l4 norm of the Fourier transform is a good candidate;5
• 2nd step: to make an energy increment argument.
kf + gkU d ≤ kf kU d + kgkU d .
kf kU d−1 ≤ kf kU d ,
Remark 2.4.1. Since the norms k.kU d are homogenous, this shows that
k.kU d are semi-norms. However, k.kU 1 isn’t a norm because kf kU 1 = E(f ).
However, one can prove (by direct calculation):
X 1
kf kU 2 = ( fb(ξ)4 ) 4 ,
and
c0 = 0.
This reduces our problem to prove that
k−1
Y
E fj (x + cj r)|x, r ∈ ZN = O(kf0 kU k−1 ) + o(1).
j=0
A. Arbieto, C. Matheus and C. G. Moreira 53
We divide the proof of this inequality into two parts: in the first part
we prove a Cauchy-Schwarz type inequality and we apply this inequality
k − 1 times to
Qthe left-hand side of the previous
identity in order to get a
k−1
control of E j=0 fj (x + cj r)|x, r ∈ ZN via a weighted sum of f0 over
(k −1)-dimensional cubes; in the second part we show that the linear forms
conditions implies that these weights are equal to 1 in average (so that the
theorem follows).
In order to state the Cauchy-Schwarz type inequality in a reasonable
way, we introduce a little bit more of notation. Given 0 ≤ d ≤ k − 1,
two vectors y = (y1 , . . . , yk−1 ) ∈ (ZN )k−1 and y = (yk−d
′ ′
, . . . , yk−1 ) ∈
d (S)
(ZN ) , and a subset S ⊂ {k − d, . . . , k − 1}, we define the vector y =
(S) (S)
(y1 , . . . , yk−1 ) ∈ (ZN )k−1 by
(S) yi if i ∈/S
yi :=
yi′ if i ∈ S.
In other words, S indicates the components of y (S) coming from y ′ instead
of y.
Lemma 2.4.1. Let ν : ZN → R+ be a measure and φ0 , . . . , φk−1 :
(ZN )k−1 → ZN some functions of the (k − 1) variables yi such that φi
doesn’t depend on yi for each 1 ≤ i ≤ k − 1. Suppose that f0 , . . . , fk−1 ∈
L1 (ZN ) are some functions satisfying |fi (x)| ≤ ν(x) for every x ∈ ZN and
0 ≤ i ≤ k − 1. For each 0 ≤ d ≤ k − 1 and 1 ≤ i ≤ k − 1, define
Y k−d−1
Y
Jd :=E ( fi (φi (y (S) ))×
S⊂{k−d,...,k−1} i=0
!
k−1
Y
× ν 1/2 (φi (y (S) ))y ∈ (ZN )k−1 , y ′ ∈ (ZN )d ,
i=k−d
and
Y
Pd := E ν(φk−d−1 (y (S) ))|y ∈ (ZN )k−1 , y ′ ∈ (ZN )d .
S⊂{k−d,...,k−1}
where
Y
G(y, y ′ ) := fk−d−1 (φk−d−1 (y (S) ))ν −1/2 (φk−d−1 (y (S) ))
S⊂{k−d,...,k−1}
and
Y k−d−2
Y
′
H(y, y ) := E( fi (φi (y (S) ))
S⊂{k−d,...,k−1} i=0
k−1
Y
× ν 1/2 (φi (y (S) ))|yk−d−1 ∈ ZN ).
i=k−d−1
k−2
Y
k−1 k−2−d
|J0 |2 ≤ Jk−1 Pd2 .
d=0
w∈{0,1}k−1
Because |fj (x)| ≤ ν(x), we see that our task is reduced to prove
Y
E |W (x, h) − 1| ν(x + w · h)|x ∈ ZN , h ∈ (ZN )k−1 = o(1).
w∈{0,1}k−1
Proof. By expanding the square, we see that it suffices to prove that, for
q = 0, 1, 2, it holds
Y
E W (x, h)q ν(x + w · h)|x ∈ ZN , h ∈ (ZN )k−1 = o(1).
w∈{0,1}k−1
w∈A
58 Chapter 2. Green-Tao-Szemerédi theorem
for some fixed A ⊂ {0, 1}k−1 , we see that it has the form
E ν(φ1 (x)) . . . ν(φ|A| (x))x ∈ ZkN ,
where x := (x, h1 , . . . , hk−1 ) and φ1 , . . . , φ|A| are an ordering of the |A|
linear forms x 7→ x + w · h with w ∈ A. Obviously, each of these linear
forms isn’t a rational multiple of any other, so that the (2k−1 , k, 1)-linear
forms condition can be used to conclude the proof of the lemma.
Let us resume the discussion of this subsection.
Resume of the subsection “Gowers norms”:
In this subsection we identified a class of norms (namely, Gowers norms)
naturally associated to the problem of counting arithmetic progressions
whose elements belong to the support of a given family of functions and
we proved Theorem 2.4.1 saying that the Gowers norms can effectively
give upper bounds on the number of such progressions up to a negligible
error. As we saw during the proof of Roth theorem, this good upper bound
is useful during the task of getting lower bounds of certain expectations
(our primary goal). The next step will be to introduce the concept of
anti-uniformity, which plays a fundamental role during the decomposition
of arbitrary functions into good and bad parts.
2.4.2 Anti-Uniformity
Since the Gowers norms (for d ≥ 2) are genuine norms, we can consider
the dual norms:
kgk(U k−1 )∗ := sup |hf, gi|,
kf kU k−1 ≤1
k−1
• kDF k(U k−1 )∗ = kF k2U k−1−1 and
k−1
• if |F | ≤ 1 + ν, then kDF kL∞ ≤ 22 −1
+ o(1).
k−1
2
Proof. The identity hF, DF i = kF kU k−1 follows directly from the defini-
function f , we have
k−1
2 −1
|hf, DF i| ≤ kf kU k−1 kF kU k−1 .
uniformly for every x ∈ ZN . On the other hand, the definition of the dual
function says that Dν1/2 can be expanded as
Y
E k−1
ν1/2 (x + w · h) h ∈ ZN .
w∈{0,1}k−1 −{0}
Remark 2.4.4. The proof of this lemma is the unique place where we
use the linear forms condition with inhomogeneuos terms bi 6= 0; indeed,
during the previous argument, all of the bi ’s are equal to x.
Making the change of variables h(j) = h + H (j) for any h ∈ (ZN )k−1 ,
taking the averages on h, we expand the products on j and interchanging
the expectations, we can rewrite this expression in terms of Gowers inner
product:
E(h(fw,H )w∈{0,1}k−1 iU k−1 |H ∈ ((ZN )k−1 )K ),
where H = (H (1) , . . . , H (K) ), f0,H := f , fw,H := gw·H for w 6= 0 with
w · H = (w · H (1) , . . . , w · H (K) ) and
K
Y
gu(1) ,...,u(K) (x) := Fj (x + u(j) ) for all u(1) , . . . , u(K) ∈ ZN .
j=1
Using the definitions of the Gowers norms and gu(1) ,...,u(K) , we can expand
this term as
Y K
Y
(j) (1) (K)
E Fj (x + u + h · w)
e x, u , . . . , u k−1
∈ ZN , h ∈ ZN .
k−1 j=1
e
w∈{0,1}
By factorizing, we obtain
K
Y
E E(Fj (x + u(j) + h · w) k−1
e | u(j) ∈ ZN ) x ∈ ZN , h ∈ ZN .
j=1
Using the assumption |Fj (x)| ≤ ν(x), our task is reduced to the verification
of the estimate
!
k−1
e | u ∈ ZN )K x ∈ ZN , h ∈ ZN
E E(ν(x + u + h · w) .
At this point, we are ready to use the correlation condition, which says
that
!
X
E ν(y + h · w)
e y ∈ ZN ≤ τ (h · (w e′ )),
e−w
k−1
′ w,
e we ∈{0,1}
′ ,w6
e =w
e
62 Chapter 2. Green-Tao-Szemerédi theorem
for all w,
e we′ ∈ {0, 1}k−1 distinct. But, since h 7→ h · (w e′ ) is a covering
e−w
K
map, the left-hand side is E(τ ) = OK (1).
Now let’s go back to the proof of the proposition. Recall that the
lemma 2.4.4 says that a basic anti-uniform function takes its values in
k−1 k−1
the interval I = [−22 , 22 ]. By Weierstrass approximation theorem,
given ε > 0, there exists a polynomial P close to the continuous function
Φ in the sense that
Remark 2.4.5. The unique place in this book where we used the correla-
tion condition was during the final part of the proof of Lemma 2.4.5.
Proof. The starting point of the argument is the following lemma ensuring
that each function generates a sigma-algebra:
Proof of lemma 2.4.6. Putting together Fubini’s theorem with the fact
that ν is a measure, we have
Z 1 X
E(1G(x)∈[ǫ(n−σ+α),ǫ(n+σ+α)] (ν(x) + 1)|x ∈ ZN )dα
0 n∈Z
Define Bǫ,σ (G) as the σ-algebra whose atoms are G−1 ([ǫ(n + α), ǫ(n + α +
1)]) for n ∈ Z. Note that Bǫ,σ (G) is well-defined because the intervals
[ǫ(n + α), ǫ(n + α + 1)] are a partition of the real line.
Clearly, if B is a σ-algebra, then the function G restricted to a atom of
B ∨ Bǫ,σ (G) takes its values on an interval of diameter ǫ, which gives the
first item of the lemma (G belongs to its own σ-algebra). Now, let A :=
G−1 ([ǫ(n + α), ǫ(n + α + 1)]) be a atom of Bǫ,σ (G). Since G takes its values
on I, we have n = O(1/ǫ) (otherwise, A = ∅). This proves the second item
of the lemma (bounded complexity). Finally, let ψσ : R → [0, 1] be a fixed
bump function such that ψσ = 1 on [σ, 1 − σ] and ψσ = 0 on [−σ, 1 + σ],
and define ΨA (x) := ψσ ( xǫ − n − α). Obviously, ΨA varies on a compact
subset Eǫ,σ of C 0 (I) (since n and α are bounded) and the identity (2.5)
implies the third item of the lemma (nice approximation by continuous
functions of G).
At this point, we come back to the proof of Proposition 2.4.2. We take
B := Bǫ,σ (DF1 ) ∨ · · · ∨ Bǫ,σ (DFK ), where Bǫ,σ (DFj ) is the sigma-algebra
provided by Lemma 2.4.6. Clearly the first item of Proposition 2.4.2 fol-
lows from the first item of Lemma 2.4.6. On the other hand, since each
Bǫ,σ (DFj ) has O(1/ǫ) atoms, B is generated by OK,ǫ (1) atoms. We say
that an atom A of B is small if E((ν + 1)1A ) ≤ σ 1/2 . Denote by Ω the
union of all small atoms. Then, Ω ∈ B and the second item of Propo-
sition 2.4.2 holds. In order to prove the last item of this proposition, it
suffices to show
E((ν − 1)1A )
= E(ν − 1|A) = oK,ǫ,σ (1) + OK,ǫ (σ 1/2 )
E(1A )
for all non-small atom A. From the smallness definition, we have
for a non-small A. Thus, since σ is small and N is large, our task is reduced
to the verification of
so that
Proof. The basic strategy is the same of the structure theorem of Fursten-
berg7 : we start with the trivial sigma-algebra B = {∅, ZN } and we look
at the function f − E(f |B). If it is uniform (i.e., the third item above
holds), we are done. Otherwise, we use the results about anti-uniformity
to find an anti-uniform function G1 with non-trivial correlation with f
and we add the level sets of G1 to the sigma-algebra B. The non-trivial
correlation property will tell us that the L2 norm of E(f |B) increases by a
non-trivial amount8 , while the pseudorandomness of ν shows that E(f |B)
stays uniformly bounded. At this point, we repeat this procedure until
f − E(f |B) becomes sufficiently uniform; note that the algorithm stops
k
in a finite number of steps (of order 22 /ǫ) due to the definite energy
increment at each step.
Now let us write this strategy in a more organized manner. Fix ǫ and let
k
K0 be the smallest integer greater than 1+22 /ǫ. We will need a parameter
0 < σ ≪ ǫ and we will take N large depending on ǫ and σ. We construct
B and Ω via a sequence of basic anti-uniform functions DF1 , . . . , DFK
on ZN , exceptional subsets Ω0 ⊂ · · · ⊂ ΩK ⊂ ZN and sigma-algebras
B0 ⊂ · · · ⊂ BK for some 0 ≤ K ≤ K0 as follows:
• Stage 2: If
k
kFK+1 kU k−1 > ǫ1/2
we define BK+1 := BK ∨ Bǫ,σ (DFK+1 ) and we go to the stage 3.
so
E(ν|Bj−1 )(x) = 1 + Oj−1,ǫ (σ 1/2 ),
for all x ∈
/ Ωj−1 . Since 0 ≤ f (x) ≤ ν(x), we conclude the pointwise
estimates
and
k(1 − 1Ω )E(ν − 1|BK1 +1 )kL∞ = OK1 ,ǫ (σ 1/2 ).
Define ΩK1 +1 := Ω ∪ ΩK1 . Obviously, ΩK1 +1 has the required properties
to execute the stage 3 without errors, so that we can go to the stage 2 of
the K1 + 1-th interaction (or we end the algorithm without any errors), so
that the inductive argument is complete.
In other words, at this moment, we proved that there are only two
possibilities for the algorithm: either it ends without errors or it goes all
the way to the K0 -th interaction. In order to conclude the proof of the
theorem, we affirm that, if the algorithm reaches the stage 3 of the K0 -
th iterate, then the following key-property (namely, the energy increment
estimate) holds:
On the other hand, the pointwise estimates (2.7), (2.9), (2.6) above show
that
h(1Ωj − 1Ωj−1 )(f − E(f |Bj−1 ), DFj )i = Oj,ǫ (σ 1/2 ),
while Lemma 2.4.6 and the estimate (2.9) tell us
h(1 − 1Ωj )(f − E(f |Bj−1 ), DFj − E(DFj |Bj ))i = O(ǫ).
A. Arbieto, C. Matheus and C. G. Moreira 69
|h(1 − 1Ωj )(f − E(f |Bj−1 )), E(DFj |Bj )i| ≥ ǫ1/2 − Oj,ǫ (σ 1/2 ) − O(ǫ).
Since the functions (1−1Ωj ), E(DFj |Bj ) and E(f |Bj−1 ) are Bj -measurable,
we can replace f by E(f |Bj ), so that we get
|h(1 − 1Ωj )(E(f |Bj ) − E(f |Bj−1 )), E(DFj |Bj )i| ≥ ǫ1/2 − Oj,ǫ (σ 1/2 ) − O(ǫ).
Thus, the triangular inequality and (2.11) show that the energy increment
estimate (2.10) follows from the following estimate
However, the left-hand side above can be expanded (by the cosine law) as
k(1 − 1Ωj )E(f |Bj−1 )k2L2 + k(1 − 1Ωj )(E(f |Bj ) − E(f |Bj−1 ))k2L2
+ 2h(1 − 1Ωj )E(f |Bj−1 ), (1 − 1Ωj )(E(f |Bj ) − E(f |Bj−1 ))i.
h(1 − 1Ωj )E(f |Bj−1 ), (1 − 1Ωj )(E(f |Bj ) − E(f |Bj−1 ))i = Oj,ǫ (σ 1/2 ).
h(1 − 1Ωj )E(f |Bj−1 ), E(f |Bj ) − E(f |Bj−1 )i = Oj,ǫ (σ 1/2 ).
Now we observe that, since the function (1−1Ωj−1 )E(f |Bj−1 ) is measurable
with respect to Bj−1 , it is orthogonal to the function E(f |Bj ) − E(f |Bj−1 )
because Bj−1 is a sub-sigma-algebra of Bj (by construction). In particular,
we can again rewrite the previous expression as
h(1Ωj − 1Ωj−1 )E(f |Bj−1 ), E(f |Bj ) − E(f |Bj−1 )i = Oj,ǫ (σ 1/2 ).
70 Chapter 2. Green-Tao-Szemerédi theorem
Using again the fact that (1Ωj − 1Ωj−1 )E(f |Bj−1 ) is a Bj -measurable (so
that it must be orthogonal to f −E(f |Bj )), we see that the previous identity
is equivalent to
Moreover, we have fAU ≤ 1+oǫ (1), so that we can use Szemerédi theorem9 .
In particular:
Remark 2.4.6. The careful reader noticed the strong analogy between the
previous estimates and the estimates of the proposition 2.3.1. In fact, as
it is natural to expect, we separated, in both arguments, a “good” term (it
was Λ3 (g, g, g) in Roth case and E(fAU (n) . . . fAU (n + (k − 1)r)|n, r ∈ ZN )
in Green-Tao case) which is relatively big (its order was δ 3 in Roth case
and c(k, δ) − oǫ (1) in Green-Tao case) and our task was to control the
“bad” terms (Λ3 (., ., .) where the bad function b must appear in some entry
in Roth case and E(f∗1 (n) . . . f∗k (n + (k − 1)r)|n, r ∈ ZN ) where the bad
function fU must appear in some entry in Green-Tao case). In order
to accomplish this goal, we used Hölder inequality in Roth case and the
generalized von Neumann theorem in Green-Tao case to reduce the problem
to the (non-trivial) “fact” that b in Roth case and fU in Green-Tao can
be taken uniform. Logically, this fact was obtained from the generalized
Koopman-von Neumann theorem in Green-Tao case, which uses (in its
proof ) the energy increment argument, as we annouced in the beginning to
the section.
Elkies-McMullen theorem
Chapter 3
Elkies-McMullen theorem
75
76 Chapter 3. Elkies-McMullen theorem
In other words, the “natural scale” for the study of the lengths |Ji | of
the gaps is 1/N . Taking this “natural scale” into account, we introduce
the following definition:
when N → ∞.
Coming back to the sequence {nα } with 0 < α < 1, some numerical
experiments suggest that it is exponentially distributed for every α 6= 1/2.
Furthermore, M. Boshernitzan observed (numerically) in 1993 a special
distribution for the specific case α = 1/2. However, a rigorous confirmation
of this numerical observation of Boshernitzan was only recently obtained
by N. Elkies and C. McMullen.
More precisely, N. Elkies and C. McMullen [6] showed √ the following
theorem about the distribution of the gaps of the sequence { n} (mod 1):
where F2 (t) and F3 (t) are explicit real-analytic functions. Moreover, F (t)
isn’t analytic (and it isn’t even C 3 ) at the points t = 1/2 and t = 2. In
other words, F (t) exhibits a genuine phase transition at the points t = 1/2
and t = 2.
C. Matheus 77
6 2 3 1
F2 (x) = ( (4r − 1) 2 ψ(r) + (1 − 6r) log r + 2r − 1) if ≤ x ≤ 2,
π2 3 2
6
F3 (x) = (f (α) − g(α) − h(α)) if x ≥ 2.
π2
√ √ √
Here ψ(r) = tan−1 [(2r−1)/ 4r − 1]−tan−1 [1/ 4r − 1], 2α = 1− 1 − 4r,
f (α) = 4(1 − 4α)(1 − α)2 log(1 − α), g(α) = 2(1 − 2α)3 log(1 − 2α) and
h(α) = 2α2 .
In particular, we see that F (t) is continuous at t = 1/2 and t = 2 (with
values F (1/2) = 6/π 2 and F (2) = 6(log 2 − 1/2)/π 2 ). Also, F (t) is C 1
but F (t) is not C 2 at t = 2 and F (t) is not C 3 at t = 1/2 (as a simple
argument with Taylor series around these points shows). Another direct
consequence of this√explicit formula of F (t) is the fact that the “tail” of
the distribution of n (mod 1) is not exponential: F (t) ∼ 3t−3 /π 2 when
t → ∞. Comparing this fact with the example 3.1.1, we √ see that the
appearance of large gaps is more frequent for the sequence n (mod 1)
than it is for a random sequence of points.
This being said, we will concentrate our efforts on the discussion of the
beautiful proof of Elkies and McMullen theorem. Since the proof of this
result involves a number of technical steps, we will describe in the next
paragraph a very rough sketch of the main arguments, leaving the details
to the subsequent sections.
Roughly speaking, Elkies and McMullen idea consists into a translation
of √the problem of the computation of the distribution F (t) of the gaps
of n (mod 1) to the calculation of the probability of a random affine
lattice of R2 to intersect a certain fixed triangle. The advantage of this
apparently artificial approach resides in the fact that the Ergodic Theory
of random (affine) lattices is well-understood due to the so-called Ratner
theory, which allows us to compute precisely the desired probability (of a
random lattice to meet a fixed triangle).
At this point, we are ready to discuss this scheme with a little bit more
of details.
78 Chapter 3. Elkies-McMullen theorem
4. also, one can show that we can assume that N is a square (without
loss of generality);
5. under these assumptions, we look at the linear approximation a2 −
x2 + 2(a + x)I of (I + a)2 and note that
n ∈ a2 − x2 + 2(a + x)I (for 0 ≤ n ≤ N )
m
2
√
(Z + x ) ∩ 2(a + x)I 6= ∅ (for 0 ≤ a + x ≤ N );
6. this last condition can be rewritten as T ∩ Z2 6= ∅ where T ⊂ R2 is
the triangle of area t given by
√
T := {(a, b) : b + x2 ∈ 2(a + x)I and a + x ∈ [0, N ]};
10. using this uniform distribution result, we get that PN (t) → p(t) when
N → ∞, where p(t) is the probability of a random affine lattice to
intersect the standard triangle St ;
11. finally, by reversing this translation, one can√show that p′′ (t) =
−F (t), where F (t) is the gap distribution of{ n}, which ends the
proof of Elkies and McMullen theorem because p′′ (t) can be explicitly
calculated from natural expressions for the Haar probability µE .
After this informal description of Elkies and McMullen arguments, we
will start a more or less detailed discussion of all of these topics.
80 Chapter 3. Elkies-McMullen theorem
Using LN we can write the union of the gaps of length < x/N as
Now let us talk about the preliminary reductions of these theorems. The
first simplification of the statement of Theorem 3.2.1 was annouced at the
fourth item of Elkies and McMullen scheme discussed above: namely, in
the statement of this theorem, we can assume that N is a square, i.e.,
N = s2 with s ∈ Z. More precisely, we have the following result:
Lemma 3.2.1 (Lemma 3.1 of Elkies and McMullen). Suppose that λs2
converges (uniformly on compact sets) to a continuous function λ∞ when
s → ∞. Then, λN also converges (uniformly on compact sets) to λ∞ when
N → ∞.
Proof.√ Observe that any integer N is far from some square s2 by a distance
2 2
√ O( N ). On the other hand, by replacing N by s , we change 3|N −s | .
of
N of the lengths of the gaps (at
p most) and we multiply the normalizing
factor 1/N by N/s2 = 1 + O( 1/N ). In particular, the desired lemma
follows.
The second simplification was described at the second and third items of
Elkies and McMullen program: during the study of λN (x) we can replace
the quadratic expressions by its linear approximations without changing
the asymptotic results. In order to formalize this fact, we will need more
2
notation. Recall
√ that
√ each√integer n can be uniquely written as a + b
where a = ⌊ n⌋ = n − { n}. Using Lemma 3.2.1, we can assume that
N is a perfect square, say N = s2 . In this situation, we see that
LN (t) = N (t2 − t1 )
Remark 3.2.1. For aj ∈ Z (j = 1, 2), we have (aj +tj )2 = a2j +2aj tj +t2j ∈
Z if and only if bj := 2aj tj + t2j ∈ Z. Furthermore, we have 0 ≤ bj ≤
(aj + 1)2 − a2j = 2aj + 1 when 0 ≤ tj ≤ 1.
Going back to the analysis of LN , we will apply the idea discussed in the
third item of the Elkies and McMullen program: in the definition of bj (see
remark 3.2.1), we replaced t2j by its linear approximation t2 + 2t(tj − t) =
2tj t − t2 nearby t. Consequently, bj is substituted by 2aj tj + (2tj t − t2 ) =
2(a + t)tj − t2 . By this reason, we will consider τj = (bj + t2 )/2(a + t) the
solution of the equation
2(a + t)τj − t2 = bj
b+t 2 a2 +b−(a+t)2
where ρt (a, b) := 2(a+t) −t = 2(a+t) and a, b vary over the set of
integers satisfying
As we told in the third item of the Elkies and McMullen program, re-
placing t2j by its linear approximation (or equivalently the replacement of
tj by τj ) in the definition of LN doesn’t affect the asymptotics. More
precisely, we have the following theorem:
3.2.3 eN
Geometrical interpretation of L
As we already told in the item 6 of the program, we are going to use a
convenient triangle T (whose specific form was described above) in our
arguments. In our current notation, T is the triangle of the (a, b)-plane
whose interior is determined by the inequalities
0 < a + t < s, 2c− (a + t)/s2 < b − 2ta − t2 < 2c+ (a + t)/s2 ,
84 Chapter 3. Elkies-McMullen theorem
for c− < 0 < c+ . Making a direct translation between our notations, the
e N is interpreted in terms of T as described
reader can easily check that L
in the following lemma:
Lemma 3.2.3. For each N = s2 and t ∈ [0, 1], we have the following
possibilities:
• if Le N (t) 6= 0, then L
e N (t) is the area c+ − c− of the biggest triangle
T as above whose interior doesn’t intersect Z2 − {(0, 0), (0, 1)};
e N (t) = 0, then all triangles T as above contains the point (a, b)
• if L
whose coordinates satisfy b − 2ta − t2 = 0 with 0 < a < s.
Now, we will “clean” the statement of this lemma by observing that the
e N (t) except for t ∈ [0, 1] in
possibilities (a, b) = (0, 0), (0, 1) don’t affect L
a subset of size O(1/s):
Lemma 3.2.4. The characterization of the quantity L e N (t) in lemma 3.2.3
isn’t affected by the inclusion of the extra cases (a, b) = (0, 0), (0, 1) except
when 0 ≤ t < 1/(s − 1) or t−1 − t < 1/(s − 1).
The proof of this lemma is a simple case-by-case verification (for further
details, see Lemma 3.7 of Elkies and McMullen). Using this lemma, it
follows that the inclusion of the extra cases (a, b) = (0, 0), (0, 1) doesn’t
affect the asymptotic of |IeN (t)| (because this minor modification alters the
values of Le N (t) only on a subset of size O(1/s)).
Now we will apply the idea discussed in the item 7 of the Elkies and
McMullen scheme: consider the affine transformation g of R2 defined by
g(a, b) = (w1 , w2 ) = (s(b − 2ta − t2 ), (a + t)/s).
Observe that g sends the vertex (−t, −t2 ) to the origin (0, 0), the triangle
T into the “standard” triangle
∆c− ,c+ := {(w1 , w2 ) ∈ R2 : 0 < w2 < 1, 2c− w2 < w1 < 2c+ w2 }
and the lattice Z2 to the lattice Λs2 (t) = g(Z2 ) given by
Λs2 (t) := {(s(b − 2ta − t2 ), (a + t)/s) : (a, b) ∈ Z2 )}.
Note that the “standard” triangle ∆c− ,c+ depends on c− , c+ but it doesn’t
depend on s, t (for the effect of comparison, this standard triangle corre-
sponds to the triangle St of item 7 of Elkies and McMullen program).
Closing our series of translations, we introduce the following definition:
Definition 3.2.1. Given an affine lattice Λ in the (w1 , w2 )-plane, we de-
note by L(Λ) the area c+ − c− of the biggest triangle of the form ∆c− ,c+
which is disjoint from Λ with the conventions L(Λ) = 0 when such a tri-
angle doesn’t exist and L(Λ) = ∞ when all of these triangles are disjoint
from Λ.
C. Matheus 85
Using this notation, we can apply Lemma 3.2.4 to resume the discussion
of this subsection into the following proposition:
In other words, Proposition 3.2.1 says that the issue of studying the
asymptotic of Ies2 is reduced to the study of the behavior of the function
L on the family of affine lattices Λs2 (t).
At this point, it remains “only” to complete the details of the items 9,
10, 11 of the program. This will be performed in the next two subsections.
Morally, these items essentially say that the study of L on the affine lattices
Λs2 (t) can be done through the Ergodic Theory of random affine lattices
(in particular, Ratner’s theorems will be useful during this task).
Proposition
√ 3.3.2. Suppose that the asymptotic gap distribution F (ξ) of
{ n} is continuous. Then,
Rx
0
ξF (ξ)dξ, so that in order to get a concrete formula for F in terms of
µE (L−1 ([0, x])) we should extract the two derivatives (with respect to the
x variable) of this function. However, at the present moment, it is not
clear even that µE (L−1 ([0, x])) is differentiable!
Hence, we should analyze more carefully the subsets L−1 ([0, x]). Keep-
ing this goal in mind, we introduce the subset Sc− ,c+ of E formed by
the affine lattices Λ with some point inside the triangle ∆c− ,c+ , where
c− < 0 < c+ . Observe that µE (Sc− ,c+ ) depends only on the area c+ −c− of
the triangle ∆c− ,c+ because any two triangles with the same area are equiv-
alent under some element of ASL2 (R) and the measure µE is ASL2 (R)-
invariant. In particular, we can define a function p : [0, ∞] → [0, 1] by
This completes the proofR xof the lemma in view of Proposition 3.3.2 and
the fact that |IN (x)| → 0 ξF (ξ)dξ.
88 Chapter 3. Elkies-McMullen theorem
∂2
F (x) = −p′′ (x) = − µE (Sc− ,c+ )
∂c− ∂c+
for any c− < 0 < c+ with c+ − c− = x. This gives the following geomet-
rical interpretation of F (x) in terms of µE : the value F (c+ − c− )dc− dc+
is the measure of the subset of affine lattices Λ ∈ E intersecting ∆c− ,c+
into exactly two points - one of them with coordinates (w1 , w2 ) verifying
w1 /2w2 ∈ (c− , c− + dc− ) and the other one with coordinates (w1 , w2 ) ver-
ifying w1 /2w2 ∈ (c+ − dc+ , c+ ).
√
From Lemma 3.3.1, the calculation of the gap distribution of F of { n}
is reduced to the explicit computation of the function p and the verification
of the fact p ∈ C 2 . In this direction, we will recall some known facts about
the theory of unimodular lattices.
We denote by B the space of unimodular lattices of R2 (i.e., discrete
subgroups Λ0 isomorphic to Z2 with covolume 1) and µB the Haar prob-
ability measure of B. A vector w ∈ Λ0 of a lattice Λ0 ∈ B is called
primitive whenever one can find w′ ∈ Λ0 such that {w, w′ } is a Z2 -basis
of Λ0 . Equivalently, w ∈ Λ0 is primitive when w/k ∈ / Λ0 for any k > 1. In
the sequel, we are going to use the following facts:
At this point, our objective is to combine the remark 3.3.1 with the desin-
tegration of µB in order to express F as a double integral. In this direction,
in view of the geometrical interpretation of F (see remark 3.3.1), we look
at the lattices Λ ∈ E intersecting the triangle ∆c− ,c+ into two points whose
(i) (i) (1) (1)
coordinates (w1 , w2 ) (i = 1, 2) verify w1 /2w2 ∈ (c− , c− + dc− ) and
(2) (2)
w1 /2w2 ∈ (c+ − dc+ , c+ ). Note that the difference between these two
points is a primitive vector of Λ: otherwise, Λ would contain a third point
C. Matheus 89
inside the line segment determined by these two points; since ∆c− ,c+ is
convex (since it is a triangle) it would follow that Λ intersects ∆c− ,c+ into
three points, a contradiction with our hypothesis. Using this primitive
vector, we apply the desintegration of µB to write F as an integral on
the w2 -coordinates v− , v+ of the vectors of Λ at the boundary of ∆c− ,c+ :
for v− , v+ ∈ (0, 1), we write w = (2c+ v+ , v+ ) − (2c− v− , v− ) and we recall
that Zw parametrize the subset of lattices containing w; next, we denote
by qx (v− , v+ ) ∈ [0, 1] the (µw )-measure of the subset of Zw formed by the
lattices do not touching the interior of ∆c− ,c+ . Observe that we write qx
instead of qc− ,c+ because this quantity depends only on x = c+ − c− . In
this notation, we can express F as a double integral:
and
{(w1 , v+ ) : 2(c+ − dc+ )v+ < w1 < 2c+ v+ }
where the vectors of the lattices reside, and the desintegration formula of
µB .
of area 1/2. Since the triangle ∆c− ,c+ has area c+ − c− , this affine trans-
formation multiplies the area by the factor
r := 1/2x.
Although the argument is not complicated, we will refer the curious reader
to Lemma 3.12 of Elkies and McMullen for a detailed proof of the following
fact:
Lemma 3.3.2. For any 0 < v, v ′ ≤ 1 and x > 0 it holds qx (v, v ′ ) =
qx (v ′ , v). Moreover, for v ≥ v ′ , we have
r
′ v(1 − v ′ ) − r
qx (v, v ) = max 0, min 1, ′ − max 0,
vv v(v − v ′ )
with r = 1/2x. Here we are using the following convention
(
v(1 − v ′ ) − r ∞ if v = v ′ and r < v(1 − v ′ )
max 0, =
v(v − v ′ ) 0 if v = v ′ and r ≥ v(1 − v ′ )
Once this fact is available, the task√of finding an explicit formula for F
(the asymptotic gap distribution of { n}) becomes a Calculus I exercise.
Indeed, combining Proposition 3.3.3 with Lemma 3.3.2 and computing
some integrals (as in the proof of Theorem 3.14 of Elkies and McMullen
paper), the reader will eventually prove the following result:
Theorem 3.3.2. It holds
2
6/π , t ∈ [0, 1/2],
F (t) = F2 (t), t ∈ [1/2, 2],
F3 (t), t ∈ [2, ∞),
Note that this group acts on R2 via the conservative affine transformations
X a b X x
7→ + .
Y c d Y y
This map is surjective and Λ(g) = Λ(h) if and only if h ∈ g · G(Z) (as the
reader can easily check), so that this map is an isomorphism between E
and G(R)/G(Z).
In the particular case of the affine lattices Λs2 (t), the corresponding
elements of G(R)/G(Z) under this isomorphism can be explicitly calculated
as follows: recall that
and
1 −2t −t2
U (t) := 0 1 t .
0 0 1
In other words, Λs2 (t) is identified with As U (t) via the isomorphism Λ. In
resume, we see that Theorem 3.3.1 is equivalent to:
Theorem 3.3.3. The circles {As U (t) : t ∈ [0, 1]} become equidistributed
in G(R)/G(Z) when s → ∞, i.e., for all f ∈ C0 (E), it holds
Z 1 Z
lim f (As U (t))dt = f dµE .
s→∞ 0 E
Finally, we concluded that these identifications reduce our task to the proof
of the following result:
Theorem 3.4.1. For all f ∈ C0 (E) it holds
Z 1 Z
f (As · U (t))dt → f dµE .
0 E
Remark 3.4.5. The main ingredients in this result are: the ”linear part”
of the horocycle is an unipotent matrix and the horocycle is non-linear.
Indeed, during the proof of Theorem 3.4.2, we will use the fact that the
linear part of the horocycle is unipotent to apply Ratner’s theorem in order
to reduce the list of candidates to the distribution law µ of the horocycle to
a countable quantity of possibilities (among them µE ). Then, we will use
the non-linearity to exclude all exotic possibilities.
Remark 3.4.6. The assumption of non-linearity of the horocycle is es-
sential: when it is linear, the conclusion of Theorem 3.4.2 is simply false!
We will come back to this point after the proof of this theorem.
After this considerations, we will dedicate the rest of this last section of
the last chapter of this book to the proof of Theorem 3.4.2. To do so, we
will use the following scheme:
• during the next subsection, we will revise some basic facts about
invariant measure and we will see some properties of the probability
measure µ associated to the distribution law of As · σ(t);
• in the sequel, we will use Ratner theorem to show that there are only
a countable quantity of possibilities to the distribution law µ;
• finally, in the last subsection, we will use the non-linearity of the
horocycle σ to show that µ = µE .
Now we start to formalize this program.
C. Matheus 95
when s → ∞.
Here the convergence means weak-* convergence. By the Banach-Alaoglu
theorem, we know that m(σs ) possesses a subsequence converging to a mea-
sure µ. In particular, our task consists into showing that µ = µE is the
unique possible limit of all convergent subsequences.
In this direction, consider the “derivative” map D from the space of
affine lattices E to the space of lattices B assigning to each element g ∈ E
its linear part D(g) ∈ B, i.e.,
a b x a b 0
D c d y := c d 0 .
0 0 1 0 0 1
Observe that, a priori, the projection of the Haar probability measure µE
of E by D isn’t necessarily equal to the Haar probability of µB of B. Thus,
as a preliminary work in the direction of the proof of Theorem 3.4.3, let
us verify that the projection of µ by D is correct:
Proposition 3.4.1. We have D∗ µ = µB .
Proof. The image H of D ◦ σ is a horocycle (in the usual sense) of the
space B. On the other hand, D sends the orbits of the ”Teichmuller
geodesic flow” As of E to the geodesics of B and D sends the measure
m(σ) to the Haar measure µH of H. Finally, a simple argument shows
that the geodesic flow of B pushes H far from the cusps of B so that H
becomes equidistributed (for further details see Theorem 2.4 of Elkies and
McMullen paper). Putting these facts together, it follows that
D∗ µ = lim(As )∗ µH = µB .
As we are going to see later, in order to fit the context of Ratner theorem,
we need to know that µ is invariant by an unipotent subgroup of SL2 (R).
In this direction, we introduce the subgroup
1 t 0
N (t) := 0 1 0 .
0 0 1
Note that this unipotent subgroup appears naturally in view of the formula
D ◦ σ(t) = N (t) whenever σ(t) is a horocycle. The following preparatory
result will put us in the setting of Ratner theorem:
Next, we use the fact that x(t) is Lipschitz, y(t) is bounded and u = τ /s2
to get
and
|y(t)/s − y(t − u)/s| ≤ (|y(t)| + |y(t − u)|)/s = O(1/s).
Thus, we see that d(ρs , σs ) → 0 when s → ∞. In particular, it follows that
lim m(ρs ) = lim m(σs ) = µ. Putting this information together with (3.3),
we get
(Nτ )∗ µ = µ
so that the proof is complete.
that is, J(ν) is the biggest subgroup of G(R) leaving ν invariant. Observe
that J(ν) is closed and N (R) ⊂ J(ν).
Proposition 3.4.3. For almost every ν of the ergodic decomposition of
µ, we have
D∗ ν = µB and D(J(ν)) = SL2 (R).
R
Proof. From Proposition 3.4.1, we know that µB = D∗ µ = D∗ νdP (ν).
Since the action of N (R) on (B, µB ) is ergodic (because it is the action of
the horocyclic flow on B), it follows that D∗ ν = µB for almost every ν.
On the other hand, by Ratner theorem, we know that ν is supported on
a closed orbit J(ν) · x ⊂ E. Thus,
Since D∗ ν = µB , we obtain
Proposition 3.4.4. Every affine action of SL2 (R) on Rk has fixed points.
Proof. By Weyl’s unitary trick, this action can be extended to a SL2 (C)-
action on Ck . On the other hand, a fixed point p ∈ Ck of the compact
subgroup SU2 (C) can be easily constructed (e.g., by averaging). Since
C · su2 (C) = sl2 (C), the point p is also fixed by the SL2 (C) and, a fortiori,
by the SL2 (R). Hence, the real part of p is a fixed point of SL2 (R) on
Rk .
Proof. Since D(H) = SL2 (R), the kernel K of the derivative map D :
H → SL2 (R) is a SL2 (R)-invariant subgroup of V2 (R) ≃ R2 , so that one
of the following two possibilities occurs:
At this stage, we can conclude this subsection with the proof of Theo-
rem 3.4.5:
100 Chapter 3. Elkies-McMullen theorem
R
Proof. We write the ergodic decomposition of µ as µ = νdP (ν). By the
proposition 3.4.6, almost every ergodic component ν of µ satisfy ν = µE
or supp(ν) ⊂ H(R) · E[n] for some n. Hence, we can write µ as:
∞
X
µ = a0 µE + an µn ,
n=1
P
∞
where an = 1 and supp(µn ) ⊂ H(R) · E[n]. In particular, if µ 6= µE
n=0
then an 6= 0 for some n ≥ 1, so that µ(H(R) · E[n]) > 0. This completes
the proof of the theorem.
U = H(r, ε) · E[n]
and
Ts = {t ∈ [0, p] : σs (t) ∈ U }.
We claim that
lim sup m(Ts ) = O(ε), (3.4)
s→∞
In particular, the points of G[n] belonging to the same fiber of σs (t) are
s st ni s + nj st
ρi,j
s (t) =
0 s−1 j −1 .
ns
0 0 1
Taking the Euclidean
S metric on the third column of the matrices above,
we see that Ts = Tsi,j where
i
i,j sx(t) n s + nj st
Ts = t : − ∈ H(r, ε) .
s−1 y(t) s−1 nj
On the other hand, we can use that x(t) is Lipschitz to estimate m(Xsi,j )
when j is large: more precisely, whenever |j| > M := 2n sup |x′ (t)|, the
0≤t≤p
subset Xsi,j is the pre-image of an interval of size 1/s by a map whose
derivative has order j/n. Hence,
m(Xsi,j ) = O(1/s|j|) for all |j| > M. (3.6)
Moreover, we note that
Ysi,j = ∅ when |j| ≥ Js := n(sε + sup |y(t)|). (3.7)
0≤t≤p
and
r j
Xsi,j = ∅ when |i| ≥ Is (j) := n( + | | + sup |x(t)|). (3.8)
s n 0≤t≤p
Next, we note that the equation (3.6) implies that the first part of this
sum is O(|Js |ε/s) = O(ε2 ) (because (3.9) says that |Is | = O(|j| + 1) and
|Js | = O(sε)) and the second part of this sum occurs over a finite set of
indices i, j so that (3.5) says that it tends to zero (when s increases). Thus,
putting these two estimates together with (3.10), we see that, for large s,
it holds
m(Ts ) = O(ε),
so that the desired estimate (3.4) follows.
Finally, we recall that ms (U ) = m(Ts )/p, so that the estimate (3.4)
implies µ(H(r, ε) · E[n]) = O(ε) for all r, ε > 0. Making ε → 0 first and
r → ∞ after, it follows that µ(H(R)·E[n]) = 0, so that the desired theorem
is proved.
m(As · σ) → µE
[2] V. Brun, La serie 1/5 + 1/7 + 1/11 + 1/13 + 1/17 + 1/19 + 1/29 +
1/31 + 1/41 + 1/43 + 1/59 + 1/61 + ..., les denominateurs sont nombres
premiers jumeaux est convergente ou finie. Bull. Sci. Math. 43, 124-
128, 1919.
[9] B. Green and T. Tao, The primes contain arbitrarily long arith-
metic progressions. Annals of Math. 167, 481–547, 2008.
103
104 Bibliography
[12] K. Roth, On certain sets of integers. J. London Math. Soc. 28, 245-
252, 1953.
[13] L. G. Schnirelman, Uspekhi Math. Nauk 6, 3-8, 1939.
[14] E. Szemerédi, On sets of integers containing no k elements in arith-
metic progression. Acta Arith. 27, 299-345, 1975.
[15] T. Tao, Arithmetic Progressions and the Primes (El Escorial Lec-
tures), Collect. Math., 37-88, 2006.
[16] T. Tao, The ergodic and combinatorial approaches to Szemerédi’s
theorem. CRM Proc. Lecture Notes 43, 145-193, 2007.
[17] T. Tao, The Gaussian primes contain arbitrarily shaped constella-
tions. J. d’Analyse Mathematique 99, 109-176, 2006.
[18] T. Tao and T. Ziegler, The primes contain arbitrarily long poly-
nomial progressions. Acta Math. 201, 213-305, 2008.
[19] J. Van der Corput, Uber Summen von Primzahlen und
Primzahlquadraten. Math. Ann. 116, 1-50, 1939.
[20] I. Vinogradov, Representation of an odd prime as a sum of three
primes. Comptes Rendus (Doklady) de l’Académie des Sciences de
PURSS 15, 291-294, 1937.
Alexander ARBIETO
UFRJ, Universidade Federal do Rio de Janeiro
Av. Athos da Silveira Ramos, 149, Ilha do Fundão, CEP 68530
Rio de Janeiro, RJ, Brazil
alexande@impa.br
http://www.im.ufrj.br/∼arbieto
Carlos MATHEUS
Collège de France
3, Rue d’Ulm, CEDEX 05
Paris, France
matheus@impa.br
http://www.impa.br/∼cmateus
Carlos Gustavo (Gugu) MOREIRA
IMPA, Instituto de Matemática Pura e Aplicada
Estrada D. Castorina, 110, CEP 22.460-320
Rio de Janeiro, RJ, Brazil
gugu@impa.br
http://www.impa.br/∼gugu