Analytic Number Theory
Daniel Naylor

Contents

1Estimating Primes
1.1Asymptotic Notation
1.2Partial Summation
1.3Arithmetic Functions and Dirichlet convolution
1.4Dirichlet Series
2Elementary Estimates for Primes
2.1Merten’s Theorems
2.2Sieve Methods
2.3Selberg Sieve
3The Riemann Zeta Function
3.1Partial fraction approximation of ζ
3.2Zero-free region
Index

What is analytic number theory?

What kind of problems are studied?

A variety of problems about integers, especially primes.

Key feature: To show that a set (of primes) is infinite, want to estimate the number of elements x.

Definition. Define

π(x)=|{primes x}|=px1.

Euclid showed: limxπ(x)=.

Theorem (Prime number theorem).

limxπ(x) log xx=1.

π(x)xlogx. (Conjectured: Legendre, Gauss. Proved: Hadamard, de la Vallée Poussin)

1 Estimating Primes

Theorem (Euler). p1p=.

Proof. Consider pN=pN(1+1p+1p2++1pN), for N={1,2,3,}. We have:

pNn=1N1nn=1N1nn+1dtt=1N1dtt=log(N1)

On the other hand, using 1+xex, so

pNpNexp(1p+1p2++1pN)=exp(pN(1p+1p2++1pN))exp(pN(1p+1p2p))exp(C+pN1p)

Comparing these two bounds gives

pN1p log log (N1)C.

Then letting N gives the desired result.

Theorem (Chebyshev’s Theorem).

π(x)cxlogx

(for x2, where c is an absolute constant).

Proof. Consider

SN=2NN=(2N)!(N!)2

for N. We have

SNj=02N2Nj=(1+1)2N=4N.

On the other hand,

SN=p2Npαp(N)

where αp(N) is the largest j such that pj|2NN. We have αp(N)=1 for p(N,2N]. So

(log4)NN<p2N log p.

Take N=x2, for x2. Hence

x<p2x log p( log 4)x2+ log x( log 4)x2+ log 4+ log x.

Then

px0j log x log 2(( log 4)x2j+2+ log x)(telescoping summation, take x2,x22,x23,)(log4)x+( log x)2+1

So for x2 and a suitable large enough constant c, we have

px( log 4)x+c( log x)2.

Hence

x( log x)2<px log p(log4)x+c( log x)2logx( log x)2(π(x)π(x( log x)2))(log4)x+c( log x)2π(x)x(logx)2+(log4)x log x( log x)2+c( log x)2 log x( log x)2(log4+𝜀)x log x

for any 𝜀, as long as xx(𝜀).

Take 𝜀=1. Choose c>0 large enough.

1.1 Asymptotic Notation

Definition 1.1 (Big O and little o notation). Let f,g,h:S, S.

Write f(x)=O(g(x)) if there is c>0 such that |f(x)|c|g(x)| for all xS.

Write f(x)=o(g(x)) if for any 𝜀>0 there is x𝜀>0 such that |f(x)|𝜀|g(x)| for xS, |x|x𝜀.

Write f(x)=g(x)+O(h(x)) if f(x)g(x)=O(h(x)) and write f(x)=g(x)+o(h(x)) if f(x)g(x)=o(h(x)).

Definition 1.2 (Vinogradov notation). Let f,g,h:S, S.

Write f(x)g(x) or g(x)f(x) if f(x)=O(g(x)).

Example.

Lemma. Let f,g,h,u:S.

Proof. Follows from the definition in a straightforward way. Example:

1.2 Partial Summation

Lemma 1.3 (Partial Summation). Assuming that:

  • (an)n are complex numbers

  • xy0

  • f:[y,x] is continuously differentiable

Then
y<nxanf(n)=A(x)f(x)A(y)f(y)yxA(t)f(t)dt,

where for t1, we define

A(t)=ntan=n=1tan.

Proof. It suffices to prove the y=0 case, since then

y<nxanf(n)=0<nxanf(n)0<nyanf(n).

Suppose y=0. By the fundamental theorem of calculus,

f(n)=f(x)nxf(t)dt=0xf(t)𝟙[n,x](t)dt.

Summing over nx, we get

nxanf(n)=A(x)f(x)0xf(t)(nx𝟙[n,x](t)an)dt=A(x)f(x)0xf(t)A(t)dt

Lemma. If x1, then

nx1n= log x+γ+O(1x),

where γ is Euler’s constant, which is given by γ=limNk=1N1k log N.

Proof. Apply Partial Summation with an=1, f(t)=1t, y=12. Clearly A(t)=t. Then,

nx1n=xx+1xtt2dt=x+O(1)x+1xt{t}t2dt=1+O(1x)+logx1x{t}t2dt=1+logx1{t}t2dt+O(1x)

The last equality is true since x{t}t2dtx1t2dt=1x.

Let γ=11{t}t2dt. Then we have the asymptotic equation as desired.

Taking x in the formula, we see that γ is equal to the formula for Euler’s constant, as desired.

Lemma. For x1,

px1p=1xπ(t)t2dt+O(1).

Proof. Apply Lemma 1.3 with an=𝟙(n) (where is the set of primes), f(t)=1t, and y=1.

We get A(t)=π(t), and then

px1p=π(x)x+1xπ(t)t2dt=1xπ(t)t2dt+O(1)

1.3 Arithmetic Functions and Dirichlet convolution

Definition (Arithmetic function). An arithmetic function is a function f:.

Definition (Multiplicative). An arithmetic function f is multiplicative if f(1)=1 and f(mn)=f(m)f(n) whenever m,n are coprime.

Moreover, f is completely multiplicative if f(mn)=f(m)f(n) for all m,n.

Example.

On the space of arithmetic functions, we have operations:

(f+g)(n)=f(n)+g(n)(fg)(n)=f(n)g(n)(fg)(n)=Dirichlet convolution

Definition (Dirichlet convolution). For f,g:, we define

(fg)(n)=d|nf(d)g(nd),

where d|n means sum over the divisors of n.

Lemma 1.4. The space of arithmetic functions with operations +, is a commutative ring.

Proof. Since arithmetic functions with + form an abelian group, it suffices to show:

Proofs:

Lemma. The set of arithmetic functions f with f(1)0 form an abelian group with operation .

Proof. Need g such that fg=I.

(fg)(1)=f(1)g(1)=1g(1)=1f(1).

Assume g(m) defined for m<n. We will defined g(n).

(fg)(n)=g(n)f(1)+d|nd1f(d)g(nd)g(n)=1f(1)d|nd1f(d)g(nd)<n

Lemma. Multiplicative arithmetic functions form an abelian group with . Moreover, for completely multiplicative f, the Dirichlet inverse is μf.

Proof. For the first part, suffices to show closedness. Let f,g: be multiplicative, and m,n coprime.

If mn=ab, we can write a=a1a2, b=b1b2 where a1=(a,m), a2=(a,n), b1=(b,m) and b2=(b,n). Therefore we have:

(fg)(mn)=mn=abf(a)g(b)=m=a1b1n=a2b2f(a1a2)g(b1b2)(above observation)=m=a1b1n=a2b2f(a1)f(a2)g(b1)g(b2)(by multiplicativity)=(fg)(m)(fg)(n)

(also need to check that inverses are multiplicative).

Now, remains to show that for completely multiplicative f, fμf=I.

Note that fμf is multiplicative by the first part. So enough to show (fμf)(pk)=I(pk) for prime powers pk. Calculate:

fμf(p)=f(p)+μf(p)=f(p)f(p)=0=I(p)

and for k2:

fμf(pk)=f(pk)+μf(p)f(pk1)=f(pk)f(p)f(pk1)=f(p)kf(p)f(p)k1=0=I(pk)

Example. τk(n)=n=n1nk1. Then

τk=111k times.

So τk is multiplicative by the previous result.

Definition (von Mangoldt function). The von Mangoldt function Λ: is

Λ(n)={logpif n=pk for some prime p0otherwise

Then log=1Λ since

logn= log p|npαp(n)=p|nαp(n) log p=1Λ(n).

Since μ is the inverse of 1, we have

logμ=1Λμ=Λ1μ=ΛI=Λ.

1.4 Dirichlet Series

For a sequence (an)n, we want to associate a generating function that gives information of (an)n. Might consider

(an)nn=1anxn.

If we do this, then pxp is hard to control. So this is not very useful.

The following series has nicer number-theoretic properties:

Definition (Formal series). For f:, define a (formal) series

Df(s)=n=1f(n)ns

for s.

Lemma. Assuming that:

  • f: satisfying |f(n)|no(1)

Then Df(s) converges absolutely for Re(s)>1 and defines an analytic function for Re(s)>1.

Proof. Let 𝜀>0 (fixed), n and Re(s)>1+2𝜀.

Then

n=N|f(n)ns|=n=N|f(n)|nRe(s)n=Nn𝜀Re(s)n=Nn1𝜀N1𝜀+n=N+1n1nt1𝜀dtN1𝜀+Nt1𝜀dtN𝜀

Hence we have absolute convergence for Re(s)>1+2𝜀. Also, Df(s) is a uniform limit of the functions n=1Nf(n)ns. From complex analysis, a uniform limit of analytic functions is analytic. Hence Df(s) is analytic for Re(s)>1+2𝜀.

Now let 𝜀0.

Theorem (Euler product). Assuming that:

Then
Df(s)=p(1+f(p)p2+f(p2)p2s+),

for Re(s)>1. Furthermore, if f is completely multiplicative, then

Df(s)=p(1f(p)ps)1,

for Re(s)>1.

Proof. Let N, Re(s)=σ>1. Let

Df(s,N)=pN(1+f(p)ps+f(p2)p2s+).

Note that the series defining the factors are absolutely convergent, since

k=1|f(pk)||pks|k=11pks<

(geometric series).

Therefore, multiplying out,

Df(s,N)=n=1a(n,N)f(n)ns

where

a(n,N)=#{ways to write n as a product of prime powers, where the primes are N}.

The fundamental theorem of arithmetic tells us that a(n,N){0,1} and a(n,N)=1 for nN.

Now,

|Df(S)Df(s,N)|n=N+1|f(n)||ns|n=N+1nσN0

(since n=1nσ<). Hence Df(s)=limNDf(s,N).

Finally, for f completely multiplicative, use geometric formula:

k=1f(p)kpks=11f(p)ps.

Lemma. Assuming that:

  • f,g:

  • |f(n)|,|g(n)|no(1)

Then
Dfg(s)=Df(s)Dg(s)

for Re(s)>1.

Proof. We know Df(s) and Dg(s) are absolutely convergent, so can expand out the product.

Df(s)Dg(s)=a,b=1f(a)g(b)(ab)s=n=1n=abf(a)g(b)ns=n=1(fg)(n)ns=Dfg(s)

Definition (Riemann zeta function). For Re(s)>1, define

ζ(s)=n=1ns.

Example.

Dirichlet hyperbola method

Problem: How many lattice points (a,b)2 satisfy abx?

PIC

Note that this number is nxτ(n)=abx1.

Dirichlet proved that for x2,

nxτ(n)=x log x+(2γ1)x+O(x12)

where γ is Euler’s constant.

We will see a proof of this shortly.

Conjecture: Can have O𝜀(x14+𝜀). Current best exponent is 0.314.

First, we prove a lemma:

Lemma (Dirichlet hyperbola method). Assuming that:

  • f,g:

  • xy1

Then
nx(fg)(n)=dyf(d)mxdg(m)+mxyg(m)y<dxmf(d).

Proof. nx(fg)(n)=dmxf(d)g(m). Split this sum into parts with dy and d>y to get the conclusion.

Theorem 1.5 (Dirichlet’s divisor problem). For x2,

nxτ(n)=x log x+(2γ1)x+O(x12)

where γ is Euler’s constant.

Proof. We use the Dirichlet hyperbola method, with y=x12. Note that τ=11. Then,

nxτ(n)=dx12mxd1+mx12x12<dxm1=xx12(xd+O(1))+mx12(xmx12+O(1))=xdx121d+xmx121mmx12x12+O(x12)

Recall from Lecture 2 that

dy1d= log y+γ+O(1y).

Taking y=x12, the previous expression becomes

2×(12logx+γ+O(1x12))x+O(x12)=xlogx+(2γ1)x+O(x12).

2 Elementary Estimates for Primes

Recall from Lecture 1:

2.1 Merten’s Theorems

Theorem 2.1 (Merten’s Theorem). Assuming that:

  • x3

Then
  • (i) px log pp= log x+O(1)
  • (ii) px1p= log log x+M+O(1 log x) (for some M)
  • (iii) px(11p)=c+o(1) log x (for some c>0)

Remark. Can show c=eγ.

Proof.

2.2 Sieve Methods

Definition (Sieve problem). Let P. Let

P(z)=pPpzp,

and let A. Denote

S(A,P,z)=|{nA:(n,P(z))=1}|.

Problem: Estimate S(A,P,z).

Note that if A[x2,x],

S(A,,x12)=|A|S(A,,x13)=|A({p1p2:p1p2>x13})|

Sieve hypothesis: There exists a multiplicative g:[0,1] and Rd such that

|{nA:n0(modd)}|=g(d)|A|+Rd

for all square-free d (no repeated prime factors).

Example.

Lemma. We have

S(A,P,z)=d|P(z)μ(d)|Ad|.

Proof. Recall that

𝟙n=1=(μ1)(n)

(since μ is the inverse of 1). Hence,

S(A,P,z)=nA𝟙(n,P(z))=1=nAd|P(z)d|nμ(d)=d|P(z)μ(a)|Ad|

Example. Let

π(x,z)=|{nx:(n,P(z))=1}|.

Let P=. Let A=[1,x]. Then

|Ad|=xd+O(1).

By the previous lemma,

π(x,z)=S(A,,z)=d|P(z)μ(d)(xd+O(1))=xd|P(z)μ(d)d+O(d|P(z)|μ(d)|1)=xd|P(z)μ(d)d+O(2π(z))P(z)=p1pπ(2)=xpz(1+μ(p)p1)+O(2π(2))fundamental theorem of arithmetic=c+o(1)logzx+O(2z)Merten’s theorem, for some c>0

For 2zlogx,

π(x,z)=c+o(1)logzx.

Theorem (Sieve of Erastothenes – Legendre). Assuming that:

  • A[1,x]

  • 2zx

  • Assume the Sieve Hypothesis

Then S(A,P,z)=|A|p2pP(1g(p))+O(x12( log x)122 log x4 log z(dxd|P(z)|Rd|2)12+|A|e log x log zpzpP(1+g(p))e)

Proof. Recall from previous lecture that

S(A,P,z)=d|P(z)dxμ(d)|Ad|(since Ad= for d>x)=d|P(z)dxμ(d)g(d)|A|+O(dxd|P(z)|Rd|)(sieve hypothesis)=|A|d|P(z)μ(d)g(d)+O(dxd|P(z)|Rd|+|A|d|P(z)d>xg(d))=|A|pzpP(1g(p))+O(dxd|P(z)|Rd|+|A|d|P(z)d>xg(d))

We estimate the first error term using Cauchy-Schwarz:

dxd|P(z)|Rd|(dxd|P(z)|Rd|2)12(dxd|P(z)1)12.

Note that if d|P(z), d>x12, then

ω(d)logx12logz,

(ω(d) – number of distinct prime factors of d), since zω(d)dx12.

Hence, 2ω(d)2logx2logz, d|P(z), d>x12.

Now we get

dxd|P(z)12logx2logzdxd|P(z)2ω(d)=τ(d)2logx2logzdxτ(d)2logx2logzxlogx

(the last step is by Dirichlet’s divisor problem: dxτ(d)=(1+o(1))x log x).

Substituting this in, the first error term becomes as desired.

Now we estimate the second error term. We have

d|P(z)d>xg(d)x1logzd|P(z)g(d)d1 log z(since d>x in the sum: Rankin’s trick)=x1logzpzpP(1+g(p)p1 log zz1 log z=e)x1logzpzpP(1+eg(p))x1 log z=e log x log zpzpP(1+g(p))e(1+ey(1+y)e)

Combining the error terms, the claim follows.

Example. Take A=[1,x], P=. Then g(d)=1d, Rd=O(1), so the sieve gives us

S(A,P,z)=π(x,z)=(x+O(1))pz(11p)+O(x12(logx)122 log x4 log zx12+(x+O(1))e log x log zpz(1+1p)e)

By Merten’s Theorem,

pz(11p)=C+o(1)logzpz(1+1p)pz(11p)1=(1C+o(1))logz

Hence,

π(x,z)=c+o(1)logz+O(x(logx)122 log x4 log z+xe log x log z( log z)e).

Hence, for 2zexp( log x10 log log x),

π(x,z)=c+o(1)logzx.

This asymptotic in fact holds for zxo(1).

In particular, the Erastothenes-Legendre sieve gives

π(x)π(x,z)+zxlogxlog log x

for z=exp( log x10 log log x). Not quite the Chebyshev bound π(x)xlogx.

2.3 Selberg Sieve

asymptotes good upper bound for primes
Erastothenes-Legendre
Selberg

Theorem 2.2 (Selberg sieve). Assuming that:

  • z2

  • A finite

  • P

  • Assume the sieve hypothesis

  • h:[0,) be the multiplicative function supported on square-free numbers, given on the primes by

    h(p)={g(p)1g(p)pP0pP

Then
S(A,P,z)|A|dzh(d)+dz2d|P(z)τ3(d)|Rd|.

Sieve hypothesis: There is a multiplicative g:[0,1] and Rd such that

|Ad|=g(d)|A|+Rd

for all square-free d1.

Proof. Let (ρd)d be real numbers with

ρ1=1,ρd=0,d>z(∗)

Then,

𝟙(n,P(z))=1(d|nd|P(z)ρd)2.

(If (n,P(z))=1, get 1ρ otherwise use 0x2).

Summing over nA,

S(A,P,z)=nA𝟙(n,P(z))=1nA(d|nd|P(z)ρd)2=d1d2|P(z)ρd1ρd2nA[d1,d2]|n1=|A|d1,d2|P(z)ρd1ρd2g([d1,d2])+d1,d2|P(z)ρd1ρd2R[d1,d2]E(sieve hypothesis)

([m,n] means lcm(m,n)).

We first estimate E:

Emaxk|ρk|2d1,d2|P(z)|R[d1,d2]|=maxk|ρk|2dzkd|P(z)d1,d2[d1,d2]=d|Rd|(d=[d1,d2])

We have

d1,d2d=[d1,d2]1=lzd1,d2d=ld1d2(d1,d2)=11(l=(d1,d2),d1=d1l,[d1,d2]=ld1d2)τ3(d)

Therefore

Edzzd|P(z)τ3(d)|Rd|maxk|ρk|2.

Now it suffices to prove that there is a choice of (ρd)d satisfying () such that

Claim 1: d1,d2|P(z)ρd1ρd2g([d1,d2])=1dzh(d).

Claim 2: |ρk|1 for all k.

Proof of claim 1:

We have, writing k=(d1,d2), di=dik,

d1,d2|P(z)ρd1ρd2g([d1,d2])=k|P(z)μ(k)2d1,d2|P(z)k(d1,d2)=1ρkd1ρkd2g(kd1d2)=k|P(z)μ(k)2g(k)d1,d2|P(z)k(d1,d2)=1ρkd1ρkd2g(d1)g(d2)(multiplicativity)

We have

𝟙(d1,d2)=1=c|d1c|d2μ(c)

(since μ1=I) so the previous expression becomes

k|P(z)μ(k)2g(k)c|P(z)kμ(c)d1,d2|P(z)d10(mod()c)d20(modc)ρkd1ρkd2g(d1)g(d2)=k|P(z)μ(k)2g(k)c|P(z)kμ(k)(d|P(z)kd0(modc))2=k|P(z)μ(k)2g(k)1c|P(z)kμ(c)(d|P(z)ckρckdg(ckd))2=nzh(m)1(d|P(z)d0(modn)ρdg(d))2

(multiplicativity, d=cd, m=ck, d=md) since 1h=μ2gμ (check on primes 1h(p)=1g(p)1).

Now we get

d1,d2|P(z)ρd1ρd2g([d1,d2])=mzh(m)1ζm2

where

ζm=d|P(z)d0(modm)ρdg(d).

Want to minimise this subject to ().

We need to translate the condition ρ1=1. Note that

mzm0(modc)μ(m)ζm=d|P(z)ρdg(d)m|dc|mμ(m)m|dcμ(c)μ(m)=μ(d)𝟙d=e=μ(e)ρeg(e)

Hence,

ρe=μ(e)g(e)m2m0(modc)μ(m)ζm.

Now,

ρ1=1=mzμ(m)ζm.

By Cauchy-Schwarz, we then get

(mzh(m)1ζm2)(mzμ(m)2h(m))(mzμ(m)ζm)2=1.

Hence

mzh(m)1ζm21mzμ(m)2h(m)=1mzh(m).

Equality holds for Tm=h(m)G(z), where G(z)=mzh(m). We now check that with these ζm, ρd=0 for d>z.

Note that

ρc=μ(c)g(c)G(z)mzm(modc)μ(m)h(m).

Hence, ρc=0 for c>2.

This proves Claim 1.

Now we prove Claim 2 (|ρc|1).

Note that any m has at most one representation as m=em, where e|d, (m,d)=1 (for any d).

Now,

G(z)c|dmzc(m,d)=1h(cm)=c|dh(c)mzc(m,d)=1h(m)c|dh(e)mzd(m,d)=1h(m)

Now,

ρd=μ(d)2h(d)g(d)G(z)mzd(m,d)=1μ(m)h(m).

Substituting the lower bound for G(z),

|ρd|h(d)g(d)c|dh(e)=1,

since 1h=hg.

Lemma 2.3. Assuming that:

  • z3

  • g:[0,1] multiplicative

  • for some K,A we have

    pzg(p) log pκ log z+A.

Then
1mzh(m)2pz1(eκ+1)(1g(p)),

where h is defined in terms of g as in Selberg’s sieve.

Proof. Note that for any c(0,1),

mzh(m)mzm|P(zc)h(m)=G(z,c).

Then

pzcpP(1g(p))1G(z,c)=pzcpP(1+h(p))mzm|P(zc)h(m)=m>zm|P(zc)h(m)

By Rankin’s trick,

1pzc(1g(p))G(z,c)=pzc(1g(p))m>zm|P(zc)h(m)pzc(1g(p))zλ log zm|P(zc)h(m)mλ log z(for any λ>0)=pzc(1g(p))eλpzc(1+h(p)pλ log z)=eλpzc(1g(p)+g(p)pλ log zeλexp(pzc(pλ log z1)λ log z( log p)pλ log z)exp(λ+λ log zpzcg(p)( log p)pλ log z)(1+tet)exp(λ+cecλλκ+λecλA log z)

Choose c=1λ and λ=eκ+1 to get the claim.

The Brun-Titchmarsh Theorem

Theorem 2.4 (Brun-Titchmarsh Theorem). Assuming that:

  • x0, y2

  • 𝜀>0 and y is large in terms of 𝜀

Then
π(x+y)π(x)(2+𝜀)ylogy.

Remark. We expect

π(x+y)π(x)=(n+o(1))ylogx

in a wide range of y (e.g. yx𝜀 for some 𝜀>0 fixed). The prime number theorem gives this for yx. The Brun-Titchmarsh Theorem gives an upper bound of the expected order for yx𝜀.

Proof. Apply the Selberg sieve with A=[x,x+y], P=. Note that for any d1,

|{aA:a0(modd)}=yd+O(1).

Hence, the sieve hypothesis holds with g(d)=1d, Rd=O(1).

Now, the function h in Selberg sieve is given on primes by

h(p)=g(p)1g(p)=1p1=1φ(p),

where φ is the Euler totient function. In general,

h(d)=μ(d)21φ(d)

(since φ is multiplicative). Now, for any z2, Selberg sieve yields

S(A,,z)ydzμ(d)2φ(d)+O(dz2τ3(d)).

By Problem 11 on Example Sheet 1, the error term is O(z2(logz)2).

Take z=y12𝜀10. Then

z2(logz)2(y12𝜀10)2( log y)2y1𝜀20

(for yy0(𝜀)). We estimate

dzμ(d)2φ(d)=dzμ(d)2ddφ(d)=dzμ(d)2dp|d(1+1p+1p2+)=pp1=pφ(p)nz1n

since any nz has at least one representation as

n=dp1a1pkak,

where dz is square-free and pi|d are primes and a10.

We have proved

nz1n= log z+O(1)(1110) log z

for zz0(𝜀).

Putting everything together gives us

π(x+y)π(x)S(A,,z)+zy(1𝜀10)logz+z+y1𝜀20(2+𝜀)ylogy

for yy0(𝜀).

3 The Riemann Zeta Function

Recall that the Riemann zeta function is defined by

ζ(s)=n=11ns

for Re(s)>1.

Remarkable properties:

Notation. f(x)g(x) means f(x)g(x)f(x). σ means that the constant in can depend on σ.

Lemma 3.1. Assuming that:

  • σ>1

  • t

Then |ζ(σ+it)|σ1.

Proof. By the Euler product formula,

ζ(σ+it)=p(1pσit)1,

hence

|ζ(σ+it)|=p|1pσit|1.

By the triangle inequality,

1pσ|1pσit|1+|pσit|=1+pσ.

Hence,

p(1pσ)1|ζ(σ+it)|p(1=pσ)1.

Note that these products converge if and only if ppσ converges, and this sum converges by the comparison test.

Lemma 3.2 (Polynomial growth of eta in half-planes).

  • (i)
    f(ζ) extends to a meromorphic function on , with the only pole being a simple pole which is at s=1.
  • (ii)
    Let k0 be an integer. Then, for Re(s)k and |s1|110 we have |ζ(s)|k|s|k+2+1.

Proof. Let k0 be an integer.

We claim that there exist polynomials Pk, Qk of degree k+1 and such that for Re(s)>1,

ζ(s)=1s1+Qk(s)+s(s+1)(s+k)1Pk({t})ts+k+1dt.(∗)

First assume () holds.

Then, since Pk({t})k1, for Re(s)>k12, |s1|110, () gives |ζ(s)|k|s|k+2+1. So (ii) follows.

For (i), using analytic continuation and (), it suffices to show that the RHS of () is meromorphic for Re(s)>k, with the only pole a simple one at s=1.

Suffices to show that

1Pk({t})ts+k+1dt

is analytic for Re(s)>k.

This follows from the following criterion: If U is open, f:U× is piecewise continuous, and if sf(s,t) is analytic in U for any t, then f(s,t)dt is analytic in U, provided that |f(s,t)|dt is bounded on compact subsets of U.

Applying this with f(s,t)=Pk({t})ts+k+1𝟙[1,) concludes the proof of (i) assuming ().

We are left with proving (). We use induction on k.

Case k=0: By Partial Summation,

ζ(s)=n=11ns=s1uus+1du(apply partial summation to nxns and let x)=s11usdu1{u}us+1du=1s1+1s1{u}us+1du

Take P0(u)=u, Q0(u)=1.

Case k+1 assuming k: Let

ck=01Pk({t})dt.

For Re(s)>1, we have

ζ(s)=1s1+Qk(s)+cks(s+1)(s+k1)+s(s+1)(s+k)1Pk({t})ckts+k+1dt.

Let

Pk+1(u)=0u(Pk(t)ck)dt.

This is a polynomial of degree k+2. By integration by parts,

1Pt({t})ckts+k+1dt=(s+k+1)1Pk+1({u})us+k+1du.

Substituting this into the previous equation, we get that case k+1 follows.

The Gamma function

Definition (Gamma function). For Re(s)>0, let

Γ(s)=0ts1etdt.

Lemma 3.3. Γ(s) is analytic for Re(s)>0.

Proof. Apply the same criterion for integral of analytic function being analytic as in the previous lemma, taking f(s,t)=ts1et𝟙[0,)(t).

Note that for 0<σ1Re(s)σ2,

|f(s,t)|dt01tσ11etdt+1tσ21etdt<.

Lemma (Functional equation for Γ). The Γ function extends meromorphically to , with the only poles being simple poles at s=0,1,2,.

Moreover:

Proof.

Theorem 3.4 (Functional equation for zeta). Assuming that:

  • ξ(s)=12s(s1)πs2Γ(ss)ζ(s) for s

Then ξ is an entire function and ξ(s)=ξ(1s) for s. Hence
πs2Γ(s2)ζ(s)=π1s2Γ(1s2)ζ(1s)

for s{0,1}.

Proof. Let Re(s)>1. Then,

Γ(s2)=0ts21etdt.

Make the change of variables f=πn2u to get

Γ(s2)=πs2ns0us21eπn2udu.

Hence

πs2nsΓ(s2)=0us21eπn2udu.

Summing over n and using Fubini,

πs2Γ(s2)ζ(s)=n=10us21eπn2udu=120us21(𝜃(u)1)du𝜃(u)=n=eπn2u=1201us21(𝜃(u)1)du+121us21(𝜃(u)1)du()

By the functional equation

𝜃(u)=1u𝜃(1u),

we have

01us21(𝜃(u)1)du=01us232𝜃(1u)du01us21du=1vs+12𝜃(v)dv01us21du(u=1v)=1vs+12(𝜃(v)1)dv+1vs+12dv1s1201us21du1s2

Plugging this into (), we get

πs2Γ(s2)ζ(s)=121(us+12+us21)(𝜃(u)1)du1s(s1).

Hence

ξ(s)=12+14s(s1)1(us+12+us21)(𝜃(u)1)du.(∗∗)

Since |𝜃(u)1|eπu, applying the criterion for integrals of analytic functions being analytic, we see that ξ(s) is entire. So by analytic continuation, () holds for all s.

Moreover, the expression for ζ(s) is symmetric with respect to s1s, so ξ(s)=ξ(1s), s.

Corollary 3.5 (Zeroes and poles of zeta). The ζ function extends to a meromorphic function in and it has

  • (i)
    Only one pole, which is a simple pole at s=1, residue 1.
  • (ii)
    Simple zeroes at s=2,4,6,.
  • (iii)
    Any other zeroes satisfy 0Re(s)1.

Proof.

3.1 Partial fraction approximation of ζ

This is a formula for ζ(s)ζ(s).

For the proof we need a lemma. We write, for z, r>0,

B(z0,r)={z:|zz0|<r}B(z0,r)¯={z:|zz0|r}B(z0,r)={z:|zz0|=r}

Lemma 3.6 (Borel-Caratheodory Theorem). Assuming that:

  • 0<r<R

  • f analytic in B(0,R)¯, with f(0)=0

Then
sup|z|r|f(z)|2rRrsup|z|=RRe(f(z)).

Proof. This is Exercise 10 on Example Sheet 2.

Lemma 3.7 (Landau). Assuming that:

  • z0 and r>0

  • f analytic in B(z0,r)

  • for some M>1 we have |f(z)|<eM|f(z0)| for all zB(z0,r)

Then for zB(z0,r4),
|f(z)f(z)ρZ1zρ|96Mr,

where Z is the set of zeroes of f in B(z0,r2)¯, counted with multiplicities.

Note. If f is a polynomial, we can factorise

f(z)=aρ(zρ),

and then

f(z)f(z)=(logf(z))=(loga+ρ log (zρ))=ρ1zρ

Proof. Let

g(z)=f(z)ρZ(zρ).

Then g is analytic and non-vanishing in B(z0,r2)¯. Note that

g(z)g(z)=f(z)f(z)ρZ1zρ

((f1fn)f1fn=i=1nf1f1).

Hence, it suffices to prove

|g(z)g(z)|96Mr

for zB(z0,r4). Write

h(z)=g(z0+z)g(z0).

Then h is analytic and non-vanishing in B(0,r2)¯, and h(0)=1. We want to show

|h(z)h(z)|96Mr

for zB(0,r4).

For all z0B(0,r) we have

|h(z)|=|f(z0+z)f(z0)ρZz0ρz0+zρ||f(z0+z)f(z0)|<eM

since

|z0ρ|r2=rr2|z0+zρ|

for zB(0,r).

By the maximum modulus principle, |h(z)|<eM for zB(0,r2)¯, so

Re log h(z)= log |h(z)|<M.

By the Borel-Caratheodory Theorem with radii 3r8, r4 we have for for zB(0,3r8),

|logh(z)|2r43r8r4M=4M.

Now, for zB(0,r4), Cauchy’s theorem gives us

|h(z)h(z)|=|12πiB(0,3r8) log h(w)(zw)2dw|12π2π3r84M(3r8r4)2=96Mr

Theorem 3.8 (Partial Fraction approximation of zetazeta).

  • (i) Let s=σ+it with |σ|10, s1 and ζ(s)0. Then
    ζ(s)ζ(s)=1s1+|ρs|1101sρ+O( log (|t|+2)).

    where the sum is over the zeroes ρ of ζ counted with multiplicity.

  • (ii) For any T0, there are log(T+2) many zeroes ρ of ζ (counted with multiplicity) with |Im(ρ)|[T,T+1].

Proof. We apply Landau with z0=2+it, r=50, with f(s)=(s1)ζ(s).

By the lemma on polynomial growth of ζ on vertical lines, for σ+itB(z0,50), we have

|f(s)|C(|t|+52)(|t+2)50Cexp(51 log (|t|+52))Cexp(51 log (|t|+52))|f(z)|

(|ζ(z+it)|1).

Let sB(z0,252). Then Landau gives us

f(s)f(s)=1s1+ζ(s)ζ(s)=|ρz0|2S1sρ+O( log (|t|+2))()

Since B(z0,252) contains all the points s=σ+it with |σ|10, it suffices to show

|ρz0|25|ρs|>1101sρ=O( log (|t|+2)).(∗∗)

Substituting s=z0 in (), we get

|ρz0|251z0ρ=O(|ζ(z0)ζ(z0)|+1+log(|t|+2))=O(log(|t|+2))

since

|ζ(2+it)ζ(2+it)|=|n=1Λ(n)n2+it|n=1Λ(n)n2=ζ(2)ζ(2)

Taking real parts,

|ρz0|252Re(ρ)|z0ρ|2=O( log (|t|+2)).

Since Re(ρ)1,

|ρz0|251|z0ρ|2=O( log (|t|+2)).

This proves part (ii). It gives also () since the sum there contains O(log(|t|+2)) zeros.

3.2 Zero-free region

Proposition 3.9. Assuming that:

  • σ>1 and t

Then
3ζζ(σ)+4Re(ζζ(σ+it))+Re(ζζ(σ+2it))0.(∗∗)

Proof. Recall that

ζζ(s)=n=1Λ(n)ns,

where Λ is the von Mangoldt function function, and Re(s)>1.

Taking linear combinations, the LHS of () becomes

n=1Λ(n)3+4Re(nit)+Re(n2it)nσ=n=1Λ(n)3+4cos(t log n)+cos(2+ log n)nσ

(Re(niu)=cos(u log n)).

We are done by the inequality:

3+4cosα+cos2α=2(1+cosα)20

for α.

Theorem 3.10 (Zero-free region). There is a constant c>0 such that ζ(σ+it)0 whenever σ>1clog(|t|+2). In particular, ζ(s)0 for Re(s)=1.

PIC

Proof. Let σ[1,2] and t. Suppose ζ(β+it). uwe know that β1. We know that ζ has no zeroes in some ball B(1,r) for some r>0 (otherwise the entire function (s1)ζ(s) would have an accumulation point for its zeros).

Choosing c>0 small enough, we can assume that |t|r. By the key inequality for ζζ,

3ζζ(σ)+4Re(ζζ(σ+it)).+Re(ζζ(σ+2it)).0.

Apply partial fraction decomposition of ζζ. So

ζζ(s)=1s1+|sρ|1101sρ+O( log (|t|+2)).(∗∗)

(t=Im(s)).

Since Re(ρ)1 for any zero ρ,

Re1σ+iuρ=σRe(ρ)|σ+iuρ|20

(σ>1).

Discarding terms, we get

3σ1+4σβClog(|t|+2).

Take σ=1+sclog(|t|+2), and assume β1clog(|t|+2) to get

3log(|t|+2)5c+4log(|t|+2)6cC(log|t|+2).

Take c=116C to get a contradiction.

Theorem 3.11 (Bounding zetazeta). Assuming that:

  • c>0 sufficiently small

  • T0

  • Re(s)10

  • |Im(s)|[T,T+1]

  • s is at least distance clog(T+2) away from any zero or pole

Then
|ζ(s)ζ(s)|c(log(T+2))2.

Proof. If s=σ+it with σ>10, then

|ζ(s)ζ(s)|=|n=1Λ(n)ns|n=1Λ(n)nσ1

Assume then that Re(s)[10,10]. Apply (). Each term satisfies

1|sρ|log(T+2)c1|s1|log(T+2)c

We know that there are O(log(T+2)) zeros with multiplicity having imaginary part [T2,T+2]. The claim follows from triangle inequality.

˙

Index

arithmetic function

Chebyshev’s Theorem

completely multiplicative

Euler’s constant

multiplicative

sieve hypothesis

Euler’s Theorem

von Mangoldt function