0% found this document useful (0 votes)
100 views21 pages

Analytic Number Theory 1

This document provides an introduction to analytic number theory. It discusses counting arithmetic objects like primes up to a given value x. It presents three examples of how analytic techniques can be used to solve counting problems by relating them to evaluating integrals involving Dirichlet series and exponential sums. These techniques include using Poisson summation, inclusion-exclusion principles from sieve theory, and shifting contours in complex analysis. The goal is to introduce readers to analytic methods that can provide asymptotic formulas and error terms for solving enumerative number theory questions.

Uploaded by

Tamás Tornyi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
100 views21 pages

Analytic Number Theory 1

This document provides an introduction to analytic number theory. It discusses counting arithmetic objects like primes up to a given value x. It presents three examples of how analytic techniques can be used to solve counting problems by relating them to evaluating integrals involving Dirichlet series and exponential sums. These techniques include using Poisson summation, inclusion-exclusion principles from sieve theory, and shifting contours in complex analysis. The goal is to introduce readers to analytic methods that can provide asymptotic formulas and error terms for solving enumerative number theory questions.

Uploaded by

Tamás Tornyi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 21

Analytic Number Theory

Valentin Blomer
21. Oktober 2022

Only for personal use!


Warning: these lecture notes are likely to contain misprints.

1
Inhaltsverzeichnis
0 Introduction 3

1 Arithmetic functions and Dirichlet series 6

2 Fourier analysis and Poisson summation 16

2
0 Introduction
Recommended literature:
• J. Brüdern, Einführung in die analytische Zahlentheorie

• H. Davenport, Multiplicative Number Theory


• H. Iwaniec, E. Kowalski, Analytic number theory
Perhaps also useful:

• H. Montgomery, R. Vaughan, Multiplicative number theory I. Classical


theory
• G. Tenenbaum, Introduction to analytic and probabilistic number theory

Analytic number theory is the art of counting arithmetic objects. Typical


questions include:
• How many primes are there up to x?
• How many primes are there up to x in a given arithmetic progression a
(mod q)?

• How many primes are there in short interval (x, x + y] with y much less
than x?1
• in how many ways can one write a number as a sum of 2 (3, 4, . . .) primes?
• the same questions for squarefree numbers, sums of two squares, etc.

• how many numbers up to x are there with exactly k prime factors?


• how many numbers up to x are there with no prime factor ≥ y?
• what can one say about the distribution of quadratic non-residues, or
primitive roots modulo p?

• how many abelian groups are there of order less than x?


One can also study more complicated algebraic objects: class groups, elliptic
curves etc.

We will introduce analytic techniques to answer such questions. Often the


starting point for a counting problem is an analytic expression for a δ-function.
1 this is related to the previous question: two numbers are in the same progression modulo

q (a prime, say), if they are q-adically close, and they are in a short interval, if they are
archimedically close

3
Example 1
We use the notation e(x) = e2πix . For n ∈ Z we have
Z 1
e(αn)dα = δn=0 .
0

Hence the number of integral solutions to n = x21 + x22 + x23 + x24 (sums of four
squares) is
X Z 1 Z 1  X 4
e(α(x21 + . . . + x24 − n))dα = e(−αn) e(αx2 ) dα.
0 0 √
x1 ,...,x4 ∈Z |x|≤ n

We have now “reduced” the counting problem to evaluating P a single integral.


In order to do this, we observe that the exponential sum e(αx2 ) is typically
quite small due to cancellation unless α is close to a rational number a/q with
small denominator, in which case it is determined by the behaviour of squa-
res modulo q. Hence one could hope that the main contribution of the integral
comes from those α that are close to rational points with small denominator,
and the remaining parts are negligible. This gives indeed an asymptotic formula
with error term. It is also an instance of a local-global principle, since the glo-
bal asymptotic formula depends on the local behaviour in residue classes. This
method is called circle method.

Example 2
For natural numbers n we have
X
µ(d) = δn=1
d|n

where µ denotes the Möbius function


(
(−1)#{prime divisors of n} , n squarefree,
µ(n) =
0, otherwise,

see (1.4). This inclusion-exclusion formula√is essentially the sieve of Eratosthe-


nes: let P be the product of all primes ≤ x. Then we have
X X X X X X X hxi
1= 1= µ(d) = µ(d) 1= µ(d) .
√ d
x<p≤x n≤x n≤x d|(n,P ) d|P n≤x d|P
(n,P )=1 n≡0 (mod d)

Ignoring error terms, we could hope that this is approximately


X µ(d) Y  1

x =x 1− .
d √ p
d|P p≤ x

4

One can show that this is asymptotic to e−γ x/ log x, where γ = 0.577 . . .
is the Euler constant. Hence our heuristic argument gives the correct order of
magnitude, but the wrong constant. In particular, it is not allowed to ignore error
terms. Nevertheless, this method can be refined to prove interesting results (not
for counting primes, though), and leads to sieve theory.

Example 3
For c ∈ R let (c) be the vertical path c + it, −∞ < t < ∞. Then for c > 0 one
has Z (
1 −s ds 1, 0 < y < 1,
y =
2πi (c) s 0, y > 1.
(This is a non-trivial formula that we will prove later in (4.6)). The complex
contour integral is only conditionally convergent, but we ignore this for the
moment. Hence if an is any sequence of complex numbers, we have (ignoring
convergence)
Z  −s Z X
X X X 1 n ds 1 an s ds
an = an δn/x≤1 = an = s
x .
n n
2πi (c) x s 2πi (c) n n s
n≤x

Again our counting problem is “reduced” to evaluating one integral, which in


turn requires some
P understanding of the analytic properties of the associated
Dirichlet series n an n−s . As a toy problem, we have
Z
X 1 ds
1= ζ(s)xs , c > 1.
2πi (c) s
n≤x

By Cauchy’s integral theorem/residue theorem, this integral is independent of


the path, as long as it doesn’t cross a pole. If it does, we need to pick up the
residue. We shift the contour to the left to some 0 < c < 1. We pick up the pole
of the zeta-function at s = 1 getting
Z
X 1 ds ?
1=x+ ζ(s)xs = x + O(xc ).
2πi (c) s
n≤x

There are serious convergence issues, but they can be overcome in many appli-
cations. We will use this idea, for instance, to count primes up to x and prove
the prime number theorem.

5
1 Arithmetic functions and Dirichlet series
(1.1) Definition
An arithmetic function is a map f : N → C. An arithmetic function f 6= 0 is
multiplicative resp. completely multiplicative if f (nm) = f (n)f (m) for all
(m, n) ∈ N2 with (m, n) = 1 resp. for all (m, n) ∈ N2 . The set A of arithmetic
functions is a ring (an integral domain) with respect to pointwise addition and
Dirichlet convolution
X
(f ∗ g)(n) := f (d)g(n/d).
d|n

The identity element with respect to multiplication is the function η with η(1) =
1, η(n) = 0 for n > 1. (To show that there are no zero divisors let r, s minimal
with f (r) 6= 0 6= g(s) for 0 6= f, g ∈ A. Then (f ∗ g)(rs) = f (r)g(s) 6= 0.)

(1.2) Remarks
a) Since f (n) = f (1 · n) = f (1)f (n) for a multiplicative function f , we have
automatically f (1) = 1.
b) Multiplicative function are determined by their values on prime powers, com-
pletely multiplicative functions by their values on primes.
c) The units of A are precisely the function f with f (1) 6= 0: if f ∗ g = η, then
1 = η(1) = f (1)g(1). Conversely, if f (1) 6= 0, define its inverse g recursively by
g(1) = 1/f (1) and
1 X
g(n) = − f (d)g(n/d).
f (1)
d|n
d>1

(1.3) Lemma
The set of mutliplicative functions forms a subgroup of A∗ .

Proof. For (n, m) = 1 we have a bijection {d | nm} ↔ {d1 | n} × {d2 | m}. Let
f, g be multiplicative (hence invertible). Then
 nm  X X  
X nm
(f ∗ g)(nm) = f (d)g = f (d1 d2 )g
d d1 d2
d|nm d1 |n d2 |m
 X  
X n n
= f (d1 )g f (d2 )g = (f ∗ g)(n) · (f ∗ g)(m)
d1 d2
d1 |n d2 |m

for (n, m) = 1, so f ∗ g is multiplicative. To show that f −1 is multiplicative


define Y  Y
h pep := f −1 (pep ).
p p

6
Then h is multiplicative and h = f −1 on prime powers, so f ∗ h = η on prime
powers, and hence by multiplicativity everywhere. We conclude that f −1 =
f −1 ∗ η = f −1 ∗ f ∗ h = h is multiplicative.

(1.4) Examples
• the function 1(n) = 1 for all n ∈ N is completely multiplicative;
P
• the divisor function τ (n) = d|n 1 = (1 ∗ 1)(n) is not completely mul-
tiplicative. We have τ (pe ) = e + 1.
P
• the iterated divisor function τk (n) = (1 ∗ . . . ∗ 1)(n) = a1 ···ak =n 1 with
τk (pe ) = k+e−1

e ;
• Euler φ-function φ(n). We have (φ ∗ 1)(n) = n (check on prime powers);
P
• the functions ω(n) = #{p | n : p prim} and Ω(n) = pk kn k are not
multiplicative;
(
log p, n = pk
• the function Λ(n) = (von Mangoldt function) sa-
0, otherwise
Q ep
tisfies (Λ ∗ 1)(n) = log n: for n = p we have
X X X
Λ(d) = ep log p = log pep = log n;
d|n p p

• Let f ∈ Z[x] be a polynomial and ρf (n) = |{x ∈ Z/nZ | f (x) ≡


0 (mod n)}|. By the Chinese reminder theorem, ρf is multiplicative;
(
(−1)ω(n) , n squarefree
• the Möbius function µ(n) = satisfies µ∗1 = η
0, otherwise
P
(check on prime powers), so d|n µ(d) = δn=1 .

(1.5) Theorem (summation by parts)


Let y ∈ N, x ∈ R, x > y.
PLet g : [y, x] → C be continuously differentiable, and
let an ∈ C. Put A(t) = y≤n≤t an . Then
X Z x
an g(n) = A(x)g(x) − A(t)g 0 (t)dt.
y≤n≤x y

Proof. We have
X X
A(x)g(x) − an g(n) = an (g(x) − g(n))
y≤n≤x y≤n≤x
X Z x Z x X
= an g 0 (ξ)dξ = g 0 (ξ) an dξ.
y≤n≤x n y y≤n≤ξ

7
(1.6) Examples
a) The limit
X 1 
γ := lim − log M = 0.577 . . .
M →∞ m
m≤M

exists since writing t = [t] + {t} we have


Z x Z x
X 1 [x] [t] x − {x} t − {t}
= + 2
dt = + dt
n x 1 t x 1 t2
n≤x
Z ∞ Z ∞ Z ∞  
x − {x} {t} {t} {t} 1
= + log x − 2
dt + 2
dt = log x + 1 − 2
dt +O .
x 1 t x t 1 t x
| {z }
=:γ

b) We have X
τ (n) = x log x + (2γ − 1)x + O(x1/2 ),
n≤x

i.e. “on average” a number n has about log n divisors.

Proof: We have
X X X X X X √
τ (n) = 1= 1+ 1+ 1=2 1 + [ x]2 .
n≤x ab≤x ab≤x
√ ab≤x
√ ab≤x
√ ab≤x

a> x b> x a,b≤ x a> x
√ √
We have [ x]2 = x + O( x) and
X X X X x √ 
1= 1= − x + O(1)
√ √ √ b
ab≤x
√ b≤ x x<a≤x/b b≤ x
a> x
√ √ √ √ √
    
1 1
= x log x + γ + O √ − x[ x] + O( x) = x log x + γ − 1 + O( x).
x 2

(1.7) Lemma
For k ∈ N, ε > 0, there exists a constant C(k, ε) > 0 such that τk (n) ≤ C(k, ε)nε .

Proof. Let ε > 0. We have


k+α−1

τk (n) Y
α
= .
nε α
pαε
p kn

If p > C0 (k, ε) for some sufficiently large C0 (k, ε), then pαε > (α + k)k . We can
drop these factors and obtain
k+α−1
 k+α−1
 !C0 (k,ε)
τk (n) Y
α α
≤ ≤ max =: C(k, ε).
nε α
pαε α≥1 2αε
p kn
p≤C0 (k,ε)

8
(1.8) Definition and Lemma
Let f be an arithmetic
P∞ function. The Dirichlet series attached to f is the
−s
formal series n=1 f (n)n . If f, g are two functions whose Dirichlet series
converge absolutely for some s ∈ C, then
a)
∞ ∞ ∞
X (f ∗ g)(n) X f (n) X g(n)
= ;
n=1
ns n=1
ns n=1 ns
b)
∞ ∞
X f (n) Y X f (pk )
=
n=1
ns p
pks
k=0

if f is multiplicative (Euler product).

Proof. a) We have
∞ ∞ ∞ ∞
X (f ∗ g)(n) X 1 X X f (a) X g(b)
= f (a)g(b) = .
n=1
ns n=1
ns a=1
as bs
ab=n b=1

b) Multiply the right hand side and use uniqueness of prime factorization.

(1.9) Examples
a) For <s > 1 we have
∞ Y −1
X 1 1
ζ(s) := = 1− s 6= 0.
n=1
ns p
p

The Riemann zeta function has meromorphic continuation to <s > 0 with only
a simple pole at s = 1: for N ∈ N we have by partial summation we have
X Z N
n−s = [N ]N −s + s [x]x−s−1 dx.
n≤N 1

For <s > 1 we have


Z ∞ Z ∞
s
ζ(s) = s (x − {x})x−s−1 dx = −s {x}x−s−1 dx.
1 s−1 1

The right hand side exists for <s > 0 (except at s = 1) and yields the desired
continuation.
b) By (1.4) we have
∞ ∞
X τ (n) X φ(n) ζ(s − 1)
s
= ζ(s)2 , <s > 1, s
= , <s > 2,
n=1
n n=1
n ζ(s)
∞ ∞
ζ 0 (s) X Λ(n) X µ(n) 1
− = , <s > 1, = , <s > 1.
ζ(s) n=1
ns n=1
n s ζ(s)

9
(1.10) Theorem
P∞ −s
If F (s) = n=1 an n is convergent for some s0 ∈ C, then it converges for
all s ∈ C with <s > <s0 , uniformly on compacta, and defines a holomorphic
function in this region.

Proof. We show that the series converges uniformly in | arg(s − s0 )| ≤ π/2 − η


for any η > 0. By partial summation we have
N N N Z N X
X an X an s0 −s X an s0 −s an s0 −s−1
= n = N + (s − s0 ) t dt.
ns ns0 ns0 M M ≤n≤t n s0
n=M n=M n=M

For any ε > 0 there is M = M (ε) such that | M ≤n≤X an n−s0 | ≤ ε for all
P
X > M . Since <s ≥ <s0 we have
N Z N !    
X a
n <(s0 −s)−1 |s − s0 | 1
≤ ε 1 + |s − s0 | t dt ≤ ε 1 + ≤ε 1+ .

ns <(s − s0 ) sin η


n=M M

(1.11) Theorem
There is a number σ0 ∈ R∪{±∞} (abscissa of convergence) such that a Dirichlet
P for all s with <s > σ0 and diverges for all s with <s < σ0 . If
series converges
σ0 > 0, i.e. if n an diverges, it is given by
PN ( N
)
log | n=1 an | X
α
σ0 = lim sup = inf α | an = O(N ) .
N →∞ log N n=1

Remark: The assumption σ0 > 0 is no loss of generality and can always be


arranged by a suitable shift an 7→ an nα .

Proof. The existence of σ0 follows from (1.10). The second equality is obvious.
P call−σthe right hand side γ. It remains to show γ = σ0 . Let σ > σ0 . Then
We
an n is convergent, hence by partial summation
N N Z NX
X X an σ an σ−1
an = σ
N − σ σ
t dt = O(N σ ),
n=1 n=1
n 1 n
n≤t

so that γ ≤ σ. Since σ > σ0 was arbitrary, P


we obtain γ ≤ σ0 .
N
Now let σ > γ. Choose γ < α < σ so that n=1 an = O(N α ). We have
N N Z NX
X an X
−σ
σ
= an N + σ an t−σ−1 dt =: f (N ),
n=1
n n=1 1 n≤t

say, and limN →∞ f (N ) exists, since for N > M we have


Z N
f (N ) − f (M ) = O(M α−σ ) + O(tα ) · t−σ−1 dt = O(M α−σ ) → 0
M
for M → ∞. Hence σ ≥ σ0 and therefore also γ ≥ σ0 .

10
(1.12) Corollary
of convergence of n an n−s and σ1 the abscissa of con-
P
Let σ0 be the
Pabscissa
vergence of n |an |n−s . Then σ0 ≤ σ1 ≤ σ0 + 1.

Proof. Exercise

(1.13) Theorem (Landau)


Let F (s) = n an n−s be a Dirichlet series with nonnegative coefficients and let
P
σ0 be its abscissa of convergence. Then F has a singularity (i.e. is not holomor-
phic) at s = σ0 .

Proof. Wlog let σ0 = 0, and assume that F is holomorphic at s = 0. Then it is


holomorphic in a neighbourhood of 0, and hence we can expand it in a Taylor
series about 1 with radius of convergence > 1. Hence there is some δ > 0 such
the following series converges:
∞ ∞ ∞ ∞ ∞
X (−1 − δ)k X (1 + δ)k X (log n)k an X an X (1 + δ)k (log n)k
F (k) (1) = =
k! k! n=1
n n=1
n k!
k=0 k=0 k=0
∞ ∞
X an (1+δ) log n X
= e = an nδ .
n=1
n n=1

Here everything is absolutely convergent, but by assumption σ0 = 0 the right


hand side diverges, contradiction.

(1.14) Theorem (identity theorem)


Let F (s) = n an n−s and G(s) = n bn n−s be two Dirichlet series that con-
P P
verge in <s > c for some c ∈ R and that are identical in this region. Then
an = bn for all n.

Proof. If not, then let m be the smallest index with am 6= bm . For σ > c we
have
∞ ∞ ∞ ∞
!
X an X bn X  m σ X  m σ
σ
0=m − = am − bm + a n − bn
n=1
nσ n=1 nσ n=m+1
n n=m+1
n

For σ → ∞ this converges to am − bm 6= 0 (limit and summation may be


interchanged by uniform convergence), contradiction.

11
(1.15) Example
What is the “probability” that two numbers are coprime? By (1.4) we have
1 X 1 X X 1 X X X
1= µ(d) = µ(d) 1
x2 x2 x 2
n,m≤x n,m≤x d|(n,m) d≤x n≤x m≤x
(n,m)=1 d|n d|m
 
1 X x 2 X µ(d) 1 X1 1 X
= 2 µ(d) + O(1) = +O + 1
x d d2 x d x2
d≤x d≤x d≤x d≤x

!  
X µ(d) X 1 log x 1 1 log x 6
= 2
+O 2
+ + = +O → 2.
d d x x ζ(2) x π
d=1 d>x

(1.16) Notation (Vinogradov symbols)


For complex-valued functions f, g with g ≥ 0 we write f  g if f = O(g). If
in addition f ≥ 0, we write f  g if g = O(f ) and f  g if f = O(g) and
g = O(f ).

Appendix: characters
(1.17) Definition
Let q ∈ N. A homomorphism χ : (Z/qZ)∗ → S 1 is called a Dirichlet character.
We extend χ to a completely multiplicative q-periodic function Z → S 1 ∪ {0}
by putting χ(n) = 0 for (n, q) 6= 1 and call this again a Dirichlet character. The
function
X∞ Y
L(s, χ) := χ(n)n−s = (1 − χ(p)p−s )−1
n=1 p

is called Dirichlet L-function. We write χ (mod q) for the set of characters


\ ∗.
χ ∈ (Z/qZ)

(1.18) Remarks and Examples


a) Since (Z/qZ)∗ is a finite abelian group, it is isomorphic to its dual group2 ,
so there are φ(q) Dirichlet characters modulo q, and they form a group. We call
the trivial character χ0 , so χ0 (n) = δ(n,q)=1 .
b) If p is an odd prime, then the Legendre symbol (./p) is a quadratic (= order
2) Dirichlet character mod p (and the only one).
c) We have the orthogonality relations
( (
X φ(q), n ≡ 1 (mod q), X φ(q), χ = χ0 ,
χ(n) = χ(n) =
χ (mod q)
0, otherwise, n (mod q)
0, otherwise.

2 this is easy to see for cyclic groups and then follows in general

12
To show for instance the second formula for χ 6= χ0 , pick m with (m, q) = 1 and
χ(m) 6= 1, then
X X X
χ(n) = χ(nm) = χ(m) χ(n).
n (mod q) n (mod q) n (mod q)

d) Let r(n) = #{(a, b) ∈ Z2 | n = a2 + b2 } = #{α ∈ Z[i] | N α = n} = 4#{a ⊆


Z[i] | N a = n} since Z[i] is a principal ideal domain and has 4 units {±1, ±i}.
The discriminant is −4. The norm is multiplicative; primes p ≡ 1 (mod 4) are
split, p = 2 is ramified and primes p ≡ 3 (mod 4) are inert. Hence

e + 1,
 p ≡ 1 (mod 4)
1 e X
r(p ) = 1, p = 2, = χ−4 (pα )
4 1
 n 0≤α≤e
2 ((−1) + 1), p ≡ 3 (mod 4)

P−4 is the unique non-trivial character modulo 4. We conclude r = 4χ−4 ∗


where χ
1 and n r(n)n−s = 4ζ(s)L(s, χ−4 ).

(1.19) Definition
a) Let χ modulo q be a Dirichlet character. We say that χ has quasiperiod d
if χ(n) = χ(m) whenever n ≡ m (mod d) and (nm, q) = 1. The smallest quasi-
period is called conductor of χ.
b) If q ∗ | q, then χ mod q is induced by χ∗ mod q ∗ if χ(n) = χ∗ (n) for all
(n, q) = 1. A character χ mod q is called primitive if it is not induced by a
character χ∗ mod q ∗ with q ∗ < q.

Example: The nontrivial character χ mod 3 with values 1, −1, 0, 1, −1, 0 . . .


induces the character 1, 0, 0, 0, −1, 0, 1, 0, 0, 0, −1, 0, . . . mod 6.

Remark: If χ mod q is induced by χ∗ mod q ∗ , then


−1 Y  −1 Y 
Y χ∗ (p) χ∗ (p)

χ(p)
L(s, χ) = 1− s = 1− = 1− L(s, χ∗ ).
p ps ps
p-q p-q p|q

(1.20) Lemma
a) The conductor of a character χ mod q is a divisor of q.
b) Every character χ mod q is induced by a unique character χ∗ mod q ∗ where
q ∗ is the conductor of χ. This character χ∗ is primitive.

Proof. a) Let d be a quasiperiod of χ and let g = (d, q). Let n ≡ m (mod g)


and (nm, q) = 1. Then n − m = dx + qy for some x, y ∈ Z. Hence χ(m) =
χ(m + qy) = χ(n − dx) = χ(n) since n − dx = m + qy is coprime to q. Hence g
is a quasiperiod of χ, and in particular the smallest quasiperiod divides q.
b) Let q ∗ be the conductor of χ. For (n, q ∗ ) = 1 define χ∗ (n) = χ(n + kq ∗ ) for

13
any k ∈ Z such that (n + kq ∗ , q) = 1, and χ∗ (n) = 0 for (n, q ∗ ) > 1. Such a k
exists, for instance Y
k= p
p|q
p-q ∗ n

does the job3 . By definition of the conductor, this definition is independent of


the choice of k (as long as (n + kq ∗ , q) = 1), and defines q ∗ -periodic function χ∗
supported on (n, q ∗ ) = 1 that inherits complete multiplicativity from χ:
χ∗ (n)χ∗ (m) = χ((n+k1 q ∗ )(m+k2 q ∗ )) = χ(nm + (k1 m + k2 n + k1 k2 q ∗ )q ∗ ) = χ∗ (nm)
| {z }
coprime to q

for (nm, q ∗ ) = 1. Hence χ∗ is a character mod q ∗ , and it is primitive, since if it


had a smaller quasiperiod, then χ had the same quasiperiod contradicting the
minimality of conductor. Finally χ∗ is unique, since any χ1 mod q ∗ inducing χ
satisfies χ1 (n) = χ(n + kq ∗ ) = χ∗ (n) for (n, q ∗ ) = 1 and k as above.

(1.21) Lemma
If a, b ∈ Z and χ is a primitive character modulo q, then
1 X
χ(ac + b) = χ(b)δq|a .
q
c (mod q)

Proof. First we observe that we can replace a by (a, q) without changing the
sum: indeed, if we write d = (a, q), a0 = a/d, q 0 = q/d with (a0 , q 0 ) = 1, then
X X X
χ(ac + b) = χ(a0 dc + b) = d χ(d(a0 c) + b)
c (mod q) c (mod q) c (mod q 0 )
X X
=d χ(dc + b) = χ(dc + b).
c (mod q 0 ) c (mod q)

Hence we can assume wlog that a | q.


The claim of the lemma is obvious if q | a. Assume now q - a, i.e., a is a
proper divisor of q, and let S be the sum in question. For any x ∈ Z such that
(1 + ax, q) = 1 the map c 7→ c0 := c(1 + ax) + bx is a bijection of Z/qZ. Hence
1 X 1 X
χ(1 + ax)S = χ((ac + b)(1 + ax)) = χ(ac0 + b) = S.
q q
c (mod q) c0 (mod q)

Assume that χ(1 + ax) = 1 for all x ∈ Z with (1 + ax, q) = 1. Let (n, q) = 1 and
fix an integer n̄ such that n̄n ≡ 1 (mod q). Then χ(n + ay) = χ(n + nn̄ay) =
χ(n)χ(1 + an̄y) = χ(n) for all y ∈ Z with (1 + an̄y, q) = (n + ay, q) = 1, so that
by definition χ has quasiperiod a < q, contradiction. Hence there exists x with
χ(1 + ax) 6= 1, and so S = 0.
3 To check this, show that every p | q divides exactly one of n and kq ∗ , therefore it does

not divide n + kq ∗ .

14
(1.22) Definition
For a character χ modulo q we define the Gauß sum
X
τ (χ) = χ(h)e(h/q)
h (mod q)

where generally e(x) = e2πix .

(1.23) Theorem
Let χ be a character modulo q, a ∈ Z.
a) If (a, q) = 1, then4
X
χ(a)τ (χ̄) = χ̄(h)e(ha/q).
h (mod q)

b) If χ is primitive, this holds for all a ∈ Z.



c) If χ is primitive, then |τ (χ)| = q.

Proof. a) change variables h 7→ ha.


b) Let d = (a, q) > 1. Then
   
X ha X X (sq/d + r)a
χ̄(h)e = χ̄(sq/d + r)e
q q
h (mod q) r (mod q/d) s (mod d)
   
X ra X d X ra X
= e χ̄(sq/d + r) = e χ̄(sq/d + r) = 0
q q q
r (mod q/d) s (mod d) r (mod q/d) s (mod q)

by (1.21) since d > 1 and χ̄ is primitive.


c) We have
 
X X X n(a − b)
φ(q)|τ (χ)|2 = |χ(n)|2 |τ (χ̄)|2 = χ̄(a)χ(b)e
q
n (mod q) n (mod q) a,b (mod q)
X
=q |χ(a)|2 = qφ(q).
a (mod q)

Note in the first step that τ (χ) = χ(−1)τ (χ̄), so that |τ (χ)| = |τ (χ̄)|.

(1.24) Definition
A Dirichlet character is called even if χ(−1) = 1 and odd if χ(−1) = −1.

4 Note that χ̄ is a primitive character if χ is a primitive character and |τ (χ̄)| = |τ (χ)|.

15
2 Fourier analysis and Poisson summation
The main result of this section is the Poisson summation formula. We give two
arithmetic applications. More will follow later.

(2.1) Definition
A Schwartz class function on R is a C ∞ -function f : R → C such that
f (n) (x) n,A (1 + |x|)−A for all n ∈ N0 and all A > 0. We denote the vector
space of such functions by S(R).

(2.2) Definition
For f ∈ L1 (R) we define the Fourier transform fb ∈ L∞ (R) by
Z
F(f )(y) = f (y) :=
b f (x)e(−xy)dx.
R

(2.3) Lemma
Let f, g ∈ L1 (R), c ∈ R.
a) The Fourier transform is a linear operator.
b) If h(x) = f (x − c), then bh(y) = e(−cy)fb(y).
h(y) = |c|−1 fb(y/c).
c) If h(x) = f (cx) with c 6= 0, then b
d) If h(x) = (f ∗ g)(x) = R f (t)g(x − t)dt, then h ∈ L1 (R) and b
R
h(y) = fb(y)b
g (y).
R R
e) We have R f (x)b g (x)dx = R fb(x)g(x)dx.
2
f) The function f (x) = e−πx is self-dual with respect to the Fourier transform.

Proof. a) - e) Change of variables and Fubini.


f) We have
Z Z Z
−πx2 −πy 2 −π(x+iy)2 −πy 2 2
e e(−xy)dx = e e dx = e e−πx dx.
R R =x=y

In this complex contour integral we shift the line of integration to =x = 0 getting


Z Z
−πx2 −πy 2 2 2
e e(−xy)dx = e e−πx dx = e−πy .
R R

(2.4) Theorem
a) If f ∈ S(R), then fb ∈ S(R). More precisely, if f (r) (x)  (1 + |x|)−n−1−δ for
all r ≤ k and some δ > 0, then fb(n) (y)  (1 + |y|)−k .
b) If f ∈ C 1 (R) ∩ L1 (R) and fb ∈ L1 (R), in particular if f, f 0 , f 00  (1 + |x|)−2
(by part a) with n = 0, k = 2, δ = 1), the Fourier inversion formula holds:
Z
f (x) = fb(y)e(xy)dy = fb(−x).
b
R

16
c) If f ∈ S(R), Parseval’s identity holds:
Z Z
2
|f (x)| dx = |fb(y)|2 dy.
R R
Proof. a) differentiation under the integral sign and integration by parts:
Z
1 1
fb(n) (y)  r |(xn f (x))(r) |dx  r
|y| R |y|
for all r ≤ k.
b) By Lebesgue’s dominated convergence theorem we have
Z Z Z
2 2
fb(y)e(−xy)dy = lim e−πε y fb(y)e(−xy)dy = lim fb(y)gε (y)dy
R ε→0 R ε→0 R
−πε2 y 2
where gε (y) = gε (y; x) = e e(−xy). By (2.3b, c, f) we compute
1 − π(x−t)2

gbε (t) = e ε2
ε
so that by (2.3e)
Z Z Z
1 π(x−y)2 f ∈C 1 (R) 1 π(x−y)2
fb(y)e(−xy)dy = lim f (y) e− ε2 dy = lim (f (x) + O(|y − x|)) e− ε2 dy
R ε→0 R ε ε→0 R ε
= lim (f (x) + O(ε)) = f (x).
ε→0

c) Apply (2.3e) with f (x) = gb(x) = b̄


g(−x), so that fb(x) = ḡ(x) by part b).

(2.5) Remarks
a) Theorem 2.4b, c holds with less restrictive assumptions on f . A standard
condition in b) is f ∈ L1 (R) ∩ C(R) and fb ∈ L1 (R), and f ∈ L1 (R) ∩ L2 (R) in
c).
b) The results in (2.3), (2.4) generalize in an obvious way to functions on Rn .
c) The Fourier transform translates smoothness in decay conditions (and vice
versa), and it translates small support into large support. In particular, there is
an uncertainty principle: the support of f and fb cannot be small simultaneously.

(2.6) Theorem (Poisson summation)


Let f ∈ S(R). Then X X
f (n) = fb(n).
n∈Z n∈Z

P
Proof. Let F (x) = P n∈Z f (x + n) and expand this 1-periodic function into a
Fourier series: F (x) = m∈Z a(m)e(mx) where
Z 1 Z 1X Z ∞
a(m) = F (x)e(−mx)dx = f (x+n)e(−mx)dx = f (x)e(−mx)dx = fb(m).
0 0 n∈Z −∞

Now put x = 0.

17
(2.7) Corollary
Let f ∈ S(R), a, q ∈ N, χ a primitive Dirichlet character modulo q.
a) We have    
X 1 X b m am
f (m) = f e .
q q q
m∈Z m∈Z
m≡a (mod q)

b) We have  
X τ (χ) X b m
f (m)χ(m) = f χ̄(m).
q q
m∈Z m∈Z

Proof. Exercise

(2.8) Corollary
2
e−πn x
P
√ Theta function Θ(x) =
The n∈Z satisfies the functional equation Θ(1/x) =
xΘ(x).

(2.9) Lemma
Let X and 0 < Z < Y be three real numbers. There exists a smooth, non-
negative function f satisfying the following properties:
a) the support of f is contained in [X − Z, X + Y + Z]
b) f = 1 on [X, X + Y ];
c) kf (j) k1  Z 1−j for all j ∈ N.

Proof. Exercise.

(2.10) Theorem (Pólya-Vinogradov)


Let χ be a non-principal Dirichlet character modulo q, M ∈ Z, N ≥ 2. Then
X
χ(n)  q 1/2 log q.
M <n≤M +N

Proof. Let us first assume that χ is primitive (and q ≥ 2). We would like to
apply Poisson summation. To this end we need to smooth out the characteristic
function on the interval (M, M + N ]. Let w be a smooth bounded function with
support on [M −1, M +N +1] such that w = 1 on (M, M +N ] and kw(j) k1 j 1
for all j ∈ N. Then we have by (2.7b) that
Z  
X X τ (χ) X nx
χ(n) = χ(n)w(n)+O(1) = χ̄(n) w(x)e − dx+O(1).
q R q
M <n≤M +N n∈Z n∈Z

Note that, since q ≥ 2, χ(0) = 0. Integration by parts shows


Z    j
nx q
w(x)e − dx j
R q |n|

18
for all j ∈ N. Hence
 
X 1 X q X q2 
χ(n)  1 + √ +  q 1/2 log q.
q n n>q n2
M <n≤M +N n≤q

If χ is induced by χ1 modulo q1 with q = q1 r, then


X X X X
χ(n) = χ1 (n) = χ1 (n) µ(d)
M <n≤M +N M <n≤M +N M <n≤M +N d|(n,r)
(n,r)=1
X X X X
= µ(d) χ1 (n) = µ(d)χ1 (d) χ1 (n)
d|r M <n≤M +N d|r M/d<n≤(M +N )/d
d|n

By what we have proved, this is


1/2
 τ (r)q1 log q1  q 1/2 log q.

Remark. This is nontrivial for N  q 1/2 log q.

Application: let q be a prime. Then any interval of length ≥ cq 1/2 log q with c
sufficiently large contains a quadratic residue and a quadratic non-residue.

(2.11) Lemma
For k ∈ Z, x ∈ R define the Bessel function
Z π
1
Jk (x) = e−ikφ+ix sin φ dφ.
2π −π

It satisfies the following properties:


a) For θ ∈ [0, 2π], x ∈ R we have
X
eix sin θ = Jk (x)eikθ .
k∈Z

b) We have 2Jk0 (x) = Jk−1 (x) − Jk+1 (x) and 2k


x Jk (x) = Jk+1 (x) + Jk−1 (x).
c) We have (xk Jk (x))0 = xk Jk−1 (x).
d) If F : (0, ∞) → C is a smooth compactly supported function and α > 0, then
∞ j Z ∞
√ dj √
Z 
2
F (x)Jk (α x)dx = − (F (x)x−k/2 )x(k+j)/2 Jk+j (α x)dx
0 α 0 dxj

for all j ∈ N0 .
e) We have Jk (x) k min(1, |x|−1/2 ).

19
Proof. a) Follows directly from the theory of Fourier series.
b) Differentiate the generating series with respect to x:

eiθ − e−iθ ix sin θ X 0


e = Jk (x)eikθ
2
k∈Z

and θ:
eiθ + e−iθ ix sin θ X
x e = kJk (x)eikθ
2
k∈Z

and compare coefficients.


c) Add the two equation in part b) and multiply by xk /2 getting xk Jk0 (x) +
kxk−1 Jk (x) = xk Jk−1 (x).
d) Integrate by parts j times using the formula in c) in the form
d  r/2 √  d √ √  α √ √
x Jr (α x) = α−r (α x)r Jr (α x) = α−r √ (α x)r Jr−1 (α x)
dx dx 2 x
α (r−1)/2 √
= x Jr−1 (α x).
2
e) The bound |Jk (x)| ≤ 1 is obvious. To show the second bound, let 0 < α <
1/10 be a parameter. Let us assume that |x|α > 100k. Then integration by parts
shows
Z π/2−α Z π/2−α

e−ikφ+ix sin φ dφ = (−ik + ix cos φ)e−ikφ+ix sin φ
α α −ik + ix cos φ
 −ikφ+ix sin φ π/2−α Z π/2−α  
e d 1
= − e−ikφ+ix sin φ dφ
−ik + ix cos φ φ=α α dφ −ik + ix cos φ
Z π/2−α  
1 d 1 dφ  1

 +
|x|α α
dφ −ik + ix cos φ |x|α
since the integrand has no sign change in the range of integration. Hence
Z π/2
eix sin φ dφ  α + (α|x|)−1
0

for any 0 < α < 1/10 such that |x|α > 100k. We can p estimate the other regions
of the integral in the same way. Choosing α  1/ |x| and assuming wlog that
x is sufficiently large (i.e. x  k 2 ), we obtain the desired bound.

(2.12) Theorem (Hardy, Sierpiński)


The number of lattice points in a circle about the origin of radius R1/2 is
πR + Oε (R1/3+ε ) for every ε > 0.

Remark: This improves what one gets from the elementary Lipschitz principle:
πR + O(R1/2 ).

20
Proof. Let 1 ≤ T ≤ R1/2 be another parameter and let w be a smooth bounded
function with support on [0, R + T ] such that w = 1 on [T, R] and kw(j) k1 j
T 1−j for all j ∈ N. Let r(n) be as in (1.18d). We have
X
r(n) = 4 χ−4 (d)  τ (n)  nε .
d|n

Hence
X X X
r(n) = r(n)w(n) + O(T Rε ) = w(x2 + y 2 ) + O(T Rε )
n≤R n≤R x,y∈Z
X Z
= w(x2 + y 2 )e(−nx − my)dx dy + O(T Rε ).
n,m∈Z R2

The term with n = m = 0 is


Z Z ∞ Z T +R
w(x2 + y 2 )dx dy = 2π w(r2 )r dr = π w(r)dr = πR + O(T ).
R2 0 0
x x
√ change of variables ( y ) 7→ S ( y ) with
In the other terms we make an orthogonal
T 2 2
S ∈ SO(2) such that (n, m)S = (0, − n + m ). Then
X Z   
2 2 x
w(x + y )e −(n, m) dx dy
R2 y
(n,m)6=0
X Z ∞ Z 2π p
= w(r2 ) e( n2 + m2 r sin φ)dφ r dr
(n,m)6=0 0 0

X Z ∞ √
=π r(`) w(r)J0 (2π r`)dr.
`≥1 0

By Lemma 2.11d) and e) with k = 0 we obtain


Z ∞ √ Z T Z R+T !  
−j/2
w(r)J0 (2π r`)dr j ` + T −j rj/2 min 1, (r`)−1/4 dr
0 0 R
−j/2
` T 1−j
R j/2
(R`)−1/4
for any j ∈ N. Hence the sum over ` is at most
X r(`) X r(`) R3/4 R1/2+ε
1/4
R + 
2
`3/4 2
`5/4 T T 1/2
`≤R/T `≥R/T

1/3
Choosing T = R gives the result.

Remark: The proof has shown the summation formula


X Z ∞ X
r(n)w(n) = π w(x)dx + r(n)w̌(n)
n 0 n

with ∞ √
Z
w̌(n) = π w(x)J0 (2π xn)dx.
0

21

You might also like