• No results found

5.3 A Tauberian theorem for Dirichlet series

N/A
N/A
Protected

Academic year: 2021

Share "5.3 A Tauberian theorem for Dirichlet series"

Copied!
14
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Chapter 5

Tauberian theorems

5.1 Introduction

In 1826, Abel proved the following result for real power series. Let f (x) =P

n=0anxn be a power series with coefficients an ∈ R that converges on the real interval (−1, 1).

Assume that P

n=0an converges. Then limx↑1f (x) exists, and in fact, limx↑1f (x) =

X

n=0

an.

In general, the converse is not true, i.e., if limx↑1f (x) exists one can not conclude that P

n=0an converges. For instance, if f (x) = (1 + x)−1 = P

n=0(−1)nxn, then limx↑1f (x) = 12, but P

n=0(−1)n diverges.

In 1897, Tauber proved a converse to Abel’s Theorem, but under an additional hypothesis. Let again f (x) = P

n=0anxn be a power series with real coefficients converging on (−1, 1). Assume that

(5.1.1) lim

x↑1f (x) =: α exists, and moreover,

(5.1.2) lim

n→∞nan = 0.

Then (5.1.3)

X

n=0

an converges and is equal to α.

(2)

Tauber’s result led to various other “Tauberian theorems,” which are are all of the following shape:

- suppose that f (x) =P

n=0anxn converges for x ∈ R with |x| < 1;

- suppose one knows something about the behaviour of f (x) as x ↑ 1 (such as (5.1.1));

- further suppose one knows something about the growth of an as n → ∞ (such as (5.1.2));

- then one can conclude something about the convergence of P

n=0an (such as (5.1.3)).

There is now a very general “Tauberian theory,” which is about Tauberian the- orems for functions defined by integrals. These include as special cases Tauberian theorems for power series and Dirichlet series.

We will prove a Tauberian theorem for Laplace transforms G(z) :=

Z 0

F (t)e−ztdt,

where F : [0, ∞) → C is a ‘decent’ function and z is a complex variable. This Tauberian theorem has the following shape.

- Assume that the integral converges for Re z > 0;

- assume that one knows something about the limiting behaviour of G(z) as Re z ↓ 0;

- assume that one knows something about the growth order of F ; - then one can conclude something about the convergence of R

0 F (t)dt.

With some modifications, we may view power series as special cases of Laplace transforms. Let g(x) =P

n=0anxn be a power series converging for |x| < 1. Define the function F (t) on [0, ∞) by

F (t) := an if n 6 t < n + 1 (n ∈ Z>0).

Then if Re z > 0, Z

0

F (t)e−ztdt =

X

n=0

Z n+1 n

F (t)e−ztdt =

X

n=0

an Z n+1

n

e−ztdt

=

X

n=0

an1

z e−nz − e−(n+1)z

= 1 − e−z z

X

n=0

ane−nz.

(3)

Hence

g(e−z) = z 1 − e−z

Z 0

F (t)e−ztdt if Re z > 0.

Later, we show how a Dirichlet series can be expressed in terms of a Laplace trans- form.

Around 1930, Wiener developed a general Tauberian theory, which is now part of functional analysis. From this, in 1931, Ikehara deduced a Tauberian theorem for Dirichlet series (now known as the Wiener-Ikehara Theorem), with which one can give simple proofs of the Prime Number Theorem and various generalizations thereof. In 1980, Newman published a new method to derive Tauberian theorems, based on a clever contour integration and avoiding any functional analysis. This was developed further by Korevaar.

Using the ideas of Newman and Korevaar, we prove a Tauberian theorem for Laplace transforms, and deduce from this a weaker version of the Wiener-Ikehara theorem. This weaker version suffices for a proof of the Prime Number Theorem for arithmetic progressions. In Section 3.3 of Jameson’s book you may find proofs along the same lines of variations on the Tauberian theorems we are proving here.

Literature: G.J.O. Jameson, The Prime Number Theorem, Cambridge University Press 2003, London Mathematical Society, Student Texts 53,

J. Korevaar, Tauberian Theory, A century of developments, Springer Verlag 2004, Grundlehren der mathematischen Wissenschaften, vol. 329.

J. Korevaar, The Wiener-Ikehara Theorem by complex analysis, Proceedings of the American Mathematical Society, vol. 134, no. 4 (2005), 1107–1116.

5.2 A Tauberian theorem for Laplace transforms

Lemma 5.2.1. Let F : [0, ∞) → C be a measurable function. Further, assume there is a constant M such that

|F (t)| 6 M for t > 1.

Then

G(z) :=

Z 0

F (t)e−ztdt

converges, and defines an analytic function on {z ∈ C : Re z > 0}.

(4)

Proof. We apply Theorem 0.6.25. We check that the conditions of that theorem are satisfied. Let U := {z ∈ C : Re z > 0}. First, F (t)e−zt is measurable on [0, ∞) × U . Second, for every fixed t ∈ [0, ∞), the function z 7→ F (t)e−tz is analytic on U . Third, let K be a compact subset of U . Then there is δ > 0 such that Re z > δ for z ∈ K, and thus,

|F (t)e−zt| 6 Me−δt for z ∈ K.

The integral R

0 M · e−δtdt converges. So indeed, all conditions of Theorem 0.6.25 are satisfied and thus, by that Theorem, G(z) is analytic on U .

We are now ready to state our Tauberian theorem.

Theorem 5.2.2. Let F : [0, ∞) → C be a function with the following properties:

(i) F is measurable;

(ii) there is M > 0 such that |F (t)| 6 M for all t > 0;

(iii) there is an analytic function G(z) on an open set containing {z ∈ C : Re z > 0}, such that

Z 0

F (t)e−ztdt = G(z) for Re z > 0.

Then R

0 F (t)dt converges and is equal to G(0).

Remark. Theorem 5.2.2 may be rephrased as lim

z→0, Re z>0

Z 0

F (t)e−ztdt = Z

0

lim

z→0,Re z>0F (t)e−ztdt.

Although this seems plausible it is highly non-trivial. Indeed, it will imply the Prime Number Theorem!

Proof. The proof consists of several steps.

Step 1. Reduction to the case G(0) = 0.

We assume that Theorem 5.2.2 has been proved in the special case G(0) = 0 and deduce from this the general case.

Assume that G(0) 6= 0. Define new functions

F (t) := F (t) − G(0)ee −t, G(z) := G(z) −e G(0) z + 1.

(5)

Then eF satisfies (i),(ii), the function eG is analytic on an open set containing {z ∈ C : Re z > 0}, we have eG(0) = 0, and for Re z > 0 we have

Z 0

F (t)ee −ztdt = Z

0

F (t)e−ztdt − G(0) Z

0

e−(z+1)tdt = G(z) − G(0)

z + 1 = eG(z).

Hence eF satisfies (iii). Now if we have proved that R

0 F (t)dt = ee G(0) = 0, then it follows that

Z 0

F (t)dt = G(0) Z

0

e−tdt = G(0).

Henceforth we assume, in addition to the conditions (i)–(iii), that G(0) = 0.

Step 2. The function GT. For T > 0, define

GT(z) :=

Z T 0

F (t)e−ztdt.

We show that GT is analytic on C. We apply again Theorem 0.6.25 and verify the conditions of that theorem. First, F (t)e−zt is measurable on [0, T ] × C. Second, for every fixed t ∈ [0, T ], z 7→ F (t)e−ztis analytic on C. To verify the third property, let K be a compact subset of C. Then for z ∈ K, there is A > 0 such that Re z > −A for z ∈ K. Hence

|F (t)e−zt| 6 MeAt for 0 6 t 6 T, z ∈ K and clearly, RT

0 M · eAtdt < ∞ since we integrate over a bounded interval. So by Theorem 0.6.25, GT is indeed analytic on C.

We clearly have

GT(0) = Z T

0

F (t)dt.

So we have to prove:

(5.2.1) lim

T →∞GT(0) = G(0) = 0.

(6)

Step 3. An integral expression for GT(0).

We fix a parameter R > 0. It will be important in the proof that R can be chosen arbitrarily large. Let:

C+the semi-circle {z ∈ C : |z| = R, Re z > 0}, traversed counterclockwise;

Cthe semi-circle {z ∈ C : |z| = R, Re z 6 0}, traversed counterclockwise;

L the line segment from −iR to iR, traversed upwards.

Define the auxiliary function (invented by Newman):

JR,T(z) := eT z

 1 + z2

R2



· 1 z.

The function GT(z) · JR,T(z) is analytic for z 6= 0, and at z = 0 it has a simple pole with residue limz→0GT(z)eT z(1 + z2/R2) = GT(0) (or a removable singularity if GT(0) = 0). So by the Residue Theorem,

(A) 1

2πi I

C++C

GT(z)JR,T(z)dz = GT(0).

The function G(z) is analytic on an open set containing {Re z > 0}. Further, G(z)JR,T(z) is analytic on this open set. For it is clearly analytic if z 6= 0, and at z = 0 the simple pole of JR,T(z) is cancelled by the zero of G(z) at z = 0, thanks to our assumption G(0) = 0. So by Cauchy’s Theorem,

(B) 1

2πi I

C++(−L)

G(z)JR,T(z)dz = 0.

We derive an expression for GT(0) as a sum of three integrals by subtracting (B) from (A), splittingH

C++C in (A) into R

C++R

C andH

C++(−L) in (B) into R

C+−R

L

(7)

and lastly combining the two integrals over C+, more precisely, GT(0) = 1

2πi I

C++C

GT(z)JR,T(z)dz − 1 2πi

I

C++(−L)

G(z)JR,T(z)dz

= 1

2πi Z

C+

GT(z)JR,T(z)dz + 1 2πi

Z

C

GT(z)JR,T(z)dz

− 1 2πi

Z

C+

G(z)JR,T(z)dz + 1 2πi

Z

L

G(z)JR,T(z)dz

= 1

2πi Z

C+

(GT(z) − G(z))JR,T(z)dz + 1 2πi

Z

C

GT(z)JR,T(z)dz

+ 1 2πi

Z

L

G(z)JR,T(z)dz

=: I1 + I2+ I3,

where I1, I2, I3 denote the three integrals. To show that GT(0) → 0 as T → ∞, we estimate |I1|, |I2|, |I3|. Here we use

R

γf (z)dz

6length(γ) · supz∈γ|f (z)|.

Step 4. Estimation of |I1|.

We first estimate |(GT(z) − G(z))JR,T(z)| for z ∈ C+. First assume that z ∈ C+, Re z > 0. Using the condition |F (t)| 6 M for t > 0, we obtain

|GT(z) − G(z)| =

Z T

F (t)e−ztdt 6

Z T

|F (t)| · e−t·Re zdt

6 Z

T

M · e−t·Re zdt = M

Re z · e−T ·Re z.

The function JR,T(z) has been devised precisely for the purpose to get rid of the de- pendence on Re z in the estimate of |I1|, and to make the estimate for |I1| decreasing to 0 as R → ∞. Indeed, for z ∈ C with |z| = R we have z · z = R2, hence

(5.2.2) |JR,T(z)| = eT ·Re z



1 + z2 z · z

1 z

= eT ·Re z

z + z z · z

= 2eT ·Re z· |Re z|

R2 . This implies that for z ∈ C+ with Re z > 0 we have

|(GT(z) − G(z))JR,T(z)| 6 M

Re z · e−T ·Re z· 2eT ·Re z· Re z

R2 6 2M R2 .

(8)

By continuity, this is true also if Re z = 0. Hence

|I1| 6 1

2πlength(C+) · sup

z∈C+

|(GT(z) − G(z))JR,T(z)| 6 1

2π · πR · 2M R2 , i.e.,

|I1| 6 M R. Step 5. Estimation of |I2|.

The argument is similar to the estimation of |I1|. We start with estimating

|GT(z)JR,T(z)| for z ∈ C. Using again |F (t)| 6 M for t > 0, we have for z ∈ C with Re z < 0,

|GT(z)| =

Z T 0

F (t)e−ztdt 6

Z T 0

|F (t)| · e−t·Re zdt

6 Z T

0

M · e−t·Re zdt = − M

Re z e−T ·Re z− 1 6 M

|Re z|· e−T ·Re z, which together with (5.2.2) implies

|GT(z)JR,T(z)| 6 2M R2 . Again this holds true also if Re z = 0. So

|I2| 6 1

2πlength(C) · sup

z∈C

|GT(z)JR,T(z)| 6 1

2π · πR · 2M R2 , leading to

|I2| 6 M R. Step 6. Estimation of |I3|.

We choose for L the parametrization z = iy, −R 6 y 6 R. Thus, I3 = 1

2πi Z R

−R

G(iy)JR,T(iy)d(iy) = 1 2π

Z R

−R

HR(y)eiT ydt, where

HR(y) := G(iy) 1 − y2

R2

1 iy.

(9)

Since by assumption, G(0) = 0, the function G(z)/z is analytic on an open set containing {z ∈ C : Re z > 0}. Hence HR(y) is continuously differentiable on [−R, R]. Since HR is independent of T , there is a constant A(R) independent of T such that

|HR(y)| 6 A(R), |HR0 (y)| 6 A(R) for y ∈ [−R, R].

Using integration by parts, we get Z R

−R

HR(y)eiT ydy = 1 iT

Z R

−R

HR(y)deiT y

= 1

iT



HR(R)eiT R− HR(−R)e−iT R− Z R

−R

HR0 (y)eiT ydy

 .

Since |eiT y| = 1, we obtain

Z R

−R

HR(y)eiT ydy

6 1 T



A(R) + A(R) + Z R

−R

|HR0 (y)|dy



6 2A(R) + 2R · A(R)

T .

Hence

|I3| 6 C(R) T , where C(R) depends on R, but is independent of T . Step 7. Conclusion of the proof.

We have to prove that limT →∞GT(0) = G(0) = 0, in other words, for every ε > 0 there is T0 such that |GT(0)| < ε for all T > T0. Combining steps 3–6, we get, for every choice of R, T ,

|GT(0)| 6 |I1| + |I2| + |I3| 6 2M

R + C(R) T .

Let ε > 0. Then choose R such that 2M/R < ε/2, and subsequently T0 with C(R)/T0 < ε/2. For these choices, it follows that for T > T0,

|GT(0)| < 12ε +12ε = ε.

This completes our proof.

(10)

5.3 A Tauberian theorem for Dirichlet series

Let Lf(s) =P

n=1f (n)n−s be a Dirichlet series. Put A(x) := X

n6x

f (n).

We prove the following Tauberian theorem.

Theorem 5.3.1. Suppose Lf(s) satisfies the following conditions:

(i) f (n) > 0 for all n;

(ii) there are C > 0, σ > 0 such that |A(x)| 6 Cxσ for all x > 1;

(iii) Lf(s) converges for s ∈ C with Re s > σ;

(iv) There is an open subset U of C containing {s ∈ C : Re s > σ}, such that Lf(s) can be continued to a function that is analytic on U \ {σ} and for which lims→σ(s − σ)Lf(s) = α.

Then

x→∞lim A(x)

xσ = α σ.

Remarks. 1) Condition (iii) follows from (ii) (see Exercise 2.2a). Further, (iii) implies that Lf(s) is analytic for Re s > σ.

2) Condition (iv) means that Lf(s) has a simple pole with residue α at s = σ if α 6= 0, and a removable singularity at s = σ if α = 0.

3) The Wiener-Ikehara Theorem is the same as Theorem 5.3.1, except that only conditions (i),(iii),(iv) are required and (ii) can be dropped.

We start with some preparations. Notice that condition (iv) of Theorem 5.3.1 implies that there is an analytic function g(s) on U such that

(5.3.1) Lf(s) = α

s − σ + g(s) if Re s > 0.

Further, we need some lemmas.

Lemma 5.3.2. For s ∈ C with Re s > σ we have Lf(s) = s

Z 1

A(x)x−s−1dx.

(11)

Proof. Let Re s > σ. Then by partial summation we have for every integer N > 1,

N

X

n=1

f (n)n−s= A(N )N−s+ s Z N

1

A(x)x−s−1dx.

Since |A(N )| 6 PN

n=1f (n) 6 CNσ, we have A(N )N−s 6 CNσ · N−Re s → 0 as N → ∞. By letting N → ∞, the lemma follows.

Lemma 5.3.3.

Z 1

A(x) − (α/σ)xσ

xσ+1 · dx = σ−1g(σ) − σ−2α converges.

Proof. By substituting x = et, we see that the identity to be proved is equivalent to (5.3.2)

Z 0

e−σtA(et) − α/σdt = σ−1g(σ) − σ−2α.

We apply Theorem 5.2.2 to F (t) := e−σtA(et) − α/σ. We check that this F satisfies conditions (i),(ii),(iii) of Theorem 5.2.2.

First, F (t) is measurable (e.g., it has only countably many discontinuities). Sec- ond, by condition (ii) of Theorem 5.3.1,

|F (t)| 6 C + |α/σ| for t > 0.

Hence conditions (i),(ii) of Theorem 5.2.2 are satisfied. As for condition (iii), notice that for Re z > 0 we have

Z 0

F (t)e−ztdt = Z

0

e−σtA(et) − α/σe−ztdt

= Z

1

A(x)x−z−σ−1dx − (α/σ) Z

1

x−z−1dx (substitute x = et)

= 1

z + σLf(z + σ) − α

σz = 1 z + σ

z + g(z + σ)

− α

σz (by Lemma 5.3.2, (5.3.1)), that is,

(5.3.3)

Z 0

F (t)e−ztdt = 1

z + σ g(z + σ) − α/σ) if Re z > 0.

The right-hand side is analytic on an open set containing {z ∈ C : Re z > 0}, hence (iii) is satisfied as well. So by Theorem 5.2.2, identity (5.3.3) extends to z = 0, and this gives precisely (5.3.2).

(12)

By condition (i), we have f (n) > 0 for all n. Hence the function A(t) is non- decreasing. Now Theorem 5.3.1 follows by combining Lemma 5.3.3 with the lemma below.

Lemma 5.3.4. Let B : [1, ∞) → R be a non-decreasing function and let β ∈ R, σ > 0. Assume that

Z 1

B(x) − βxσ

xσ+1 · dx converges.

Then

x→∞lim B(x)

xσ = β.

Proof. We may assume without loss of generality that β = 1. Indeed, choose γ > 0 such that β + γ > 0, and replace B(x) by B(x) := β+γ1 B(x) + γxσ. Then B is non-decreasing and R

1 (B(x) − xσ)dx/xσ+1 converges. If we are able to prove that limx→∞B(x)/xσ = 1, then limx→∞B(x)/xσ = β follows.

So assume that β = 1. Assume that limx→∞B(x)/xσ does not exist or is not equal to 1. Then there are two possibilities:

(a) there are ε > 0 and an increasing sequence {xn}n=1 with xn → ∞ such that B(xn)/xσn > 1 + ε for all n;

(b) there are ε > 0 and an increasing sequence {xn}n=1 with xn → ∞ such that B(xn)/xσn 6 1 − ε for all n.

We consider only case (a); case (b) can be dealt with in the same manner. So assume (a). Then since R

1 (B(x) − xσ)dx/xσ+1 converges, we have

(5.3.4) lim

y1,y2→∞

Z y2

y1

B(x) − xσ

xσ+1 · dx = lim

y2→∞

Z y2

1

− lim

y1→∞

Z y1

1

= 0.

We choose y1, y2 appropriately and derive a contradiction. Notice that for x > xn we have, since B is non-decreasing,

B(x) − xσ

xσ+1 > B(xn) − xσ

xσ+1 > (1 + ε)xσn− xσ xσ+1 .

This is > 0 for xn 6 x 6 (1 + ε)1/σxn, so there is some hope that with the choice y1 = xn, y2 = (1 + ε)1/σxn the integral in (5.3.4) becomes strictly positive and does

(13)

not converge to 0 as n → ∞. Indeed we have Z (1+ε)1/σxn

xn

B(x) − xσ

xσ+1 · dx >

Z (1+ε)1/σxn

xn

(1 + ε)xσn− xσ xσ+1 · dx

=

Z (1+ε)1/σ 1

(1 + ε) − uσ

uσ+1 · du (u = x/xn)

= h

− (1 + ε)σ−1u−σ− log ui(1+ε)1/σ 1

= σ−1 ε − log(1 + ε).

This last number is independent of n and strictly positive, since σ > 0 and since log(1 + ε) < ε. This contradicts (5.3.4). Hence case (a) is impossible.

5.4 Exercises

Exercise 5.1. Work out in detail case (b) of the proof of Lemma 5.3.4.

Exercise 5.2. a) Prove that lim

x→∞

1 x

X

n6x

µ(n) = 0.

Hint. Apply Theorem 5.3.1 to ζ(s) + ζ(s)−1. Use Theorem 4.2, Corollary 4.4 and Theorem 4.5.

b) Prove that lim

x→∞

1 x

X

n6x

|µ(n)| = 6/π2.

Hint. Show first that L|µ|(s) = ζ(s)/ζ(2s) for Re s > 1.

c) Let A(x) denote the number of positive integers n 6 x with µ(n) = 1 and B(x) the number of positive integers n 6 x with µ(n) = −1. Prove that

x→∞lim A(x)

x = lim

x→∞

B(x)

x = 3/π2.

Exercise 5.3. Let f : Z>0 → R>0 be an arithmetic function satisfying conditions (i)–(iv) of Theorem 5.3.1, with α, σ ∈ R and σ > 0.

a) Let g : Z>0 → R be an arithmetic function, possibly assuming negative values, such that |g(n)| 6 f (n) for all n and such that there exist β ∈ R and an open

(14)

subset U of C containing {s ∈ C : Re s > σ} such that Lg(s) can be continued to a function which is analytic on U \ {σ} and for which lim

s→σ(s − σ)Lg(s) = β.

Prove that lim

x→∞

1 xσ

X

n6x

g(n) = β/σ.

b) Let now g be as in a) except that g may assume its values in C and not necessarily in R. Prove again that lim

x→∞

1 xσ

X

n6x

g(n) = β/σ.

Hint. Let g : n 7→ g(n) be the complex conjugate of g. Then Lg(s) = Lg(s).

In the proof of Corollary 0.6.24 from the Prerequisites it has been shown that if f is any analytic function on an open set V symmetric about the real axis (i.e., z ∈ V → z ∈ V , then z 7→ f (z) is also analytic on V . Use this to show that Lg(s) is also analytic on U \ {σ} and lims→σ(s − σ)Lg(s) = β and apply a) to Re g = 12(g + g) and Im g = 2i1(g − g).

Exercise 5.4. Let f : Z>0 → C be a multiplicative function such that |f(n)| 6 1 for all integers n > 1 and such that f (p) = −1 for every prime number p.

a) Prove that ζ(s)Lf(s) =Y

p



1 + f (p2)p−2s+ f (p3)p−3s+ · · · 1 − p−s



for s ∈ C with Re s > 1

and then show, using Corollary 0.6.29 from the Prerequisites, that the right- hand side defines an analytic function on {s ∈ C : Re s > 12}.

b) Prove that lim

x→∞

1 x

X

n6x

f (n) = 0.

c) Let A(x) denote the number of positive integers n 6 x such that ω(n) is even.

Prove that lim

x→∞

A(x) x = 12.

Referenties

GERELATEERDE DOCUMENTEN

The Theorem is a classic result within the theory of spaces of continuous maps on compact Hausdorff spaces and is named after the mathematicians Stefan Banach and Marshall Stone..

We need to think about convergence of series in Mat n ( C) before we start substituting matrices for the variables of the free associative algebra.. 4.1 Convergence of

If E is an elliptic curve over Q, we say that it is modular if a cusp form f as in the Modularity Theorem exists, and the theorem can be rephrased as: “all elliptic curves over Q

As a first remark, note that if a finite flat S-group scheme G is defined over a field k of characteristic zero, meaning that it is the pull-back to S of some finite group scheme

In a recent paper [12], Evertse and Schlickewei have proved a new absolute quantitative version of the Subspace Theorem which in turn makes use of the absolute Minkowski

In the proof of Theorem 2 we use a lower bound for resultants in terms of discriminants from [5] which has been proved by means of Schlickewei’s p-adic generalisation [10] of

We kunnen slechts raden naar zijn drijf- veren maar zijn analyse van de formulering van de stelling en zijn herformulering ervan doen sterk denken aan een wens de stelling en het

I thank Lawrence Forooghian from Cambridge University for drawing my attention to a mistake in a previous version of these notes and for providing a